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

Uniform Calderón-Zygmund estimates
in multiscale elliptic homogenization

Weisheng Niu W. Niu: School of Mathematical Science, Anhui University, Hefei, China niuwsh@ahu.edu.cn  and  Jinping Zhuge J. Zhuge: Morningside Center of Mathematics, Academy of Mathematics and systems science, Chinese Academy of Sciences, Beijing, China. jpzhuge@amss.ac.cn
Abstract.

This paper is concerned with the elliptic equation div(Aεuε)=divf-\text{\rm div}(A_{\varepsilon}\nabla u_{\varepsilon})=\text{\rm div}f in a bounded C1C^{1} domain, where AεA_{\varepsilon} takes a form of Aε(x)=A(x/ε1,x/ε2,,x/εn)A_{\varepsilon}(x)=A(x/\varepsilon_{1},x/\varepsilon_{2},\cdots,x/\varepsilon_{n}), with A(y1,y2,,yn)A(y_{1},y_{2},\cdots,y_{n}) being 1-periodic in each yiy_{i}. We prove the uniform Calderón-Zygmund estimate, namely, the uniform LpL^{p} boundedness of the linear map fuεf\mapsto\nabla u_{\varepsilon} for any p(1,)p\in(1,\infty) with a constant independent of small parameters (ε1,ε2,,εn)(0,1]n(\varepsilon_{1},\varepsilon_{2},\cdots,\varepsilon_{n})\in(0,1]^{n}. Our result includes the uniform Calderón-Zygmund estimate in quasiperiodic elliptic homogenization (even without the Diophantine condition), which was previously unknown. The proof novelly combines the Dirichlet’s theorem on the simultaneous Diophantine approximation from number theory, a technique of reperiodization, reiterated periodic homogenization and a large-scale real-variable argument. Using the idea of reperiodization, we also obtain some large-scale or mesoscopic-scale Lipschitz estimates.

Keywords: Multiscale homogenization, Diophantine approximation, Calderón-Zygmund estimate.

2020 Mathematics Subject Classification:
35B27

1. Introduction

In this paper, we study the uniform regularity for a family of linear elliptic equations or systems in divergence form in a bounded domain Ωd\Omega\subset\mathbb{R}^{d},

div(Aε(x)uε)=divf,\displaystyle-\text{\rm div}(A_{\varepsilon}(x)\nabla u_{\varepsilon})=\text{\rm div}f, (1.1)

where the coefficient matrix takes a form of Aε(x)=A(x/ε1,x/ε2,,x/εn)A_{\varepsilon}(x)=A(x/\varepsilon_{1},x/\varepsilon_{2},\cdots,x/\varepsilon_{n}), and (εi)1in(0,1]n(\varepsilon_{i})_{1\leq i\leq n}\in(0,1]^{n} are small parameters describing nn different oscillating scales of the coefficients. We assume that A(y1,y2,,yn):d×nd×dA(y_{1},y_{2},\cdots,y_{n}):\mathbb{R}^{d\times n}\mapsto\mathbb{R}^{d\times d} is 1-periodic in each yidy_{i}\in\mathbb{R}^{d} with 1in1\leq i\leq n. Due to the symmetry of yiy_{i}’s, it is harmless to assume ε1ε2εn>0\varepsilon_{1}\geq\varepsilon_{2}\geq\cdots\geq\varepsilon_{n}>0. The full assumptions on AA will be given shortly.

The homogenization theorem (or quantitative convergence rates) and uniform regularity are the fundamental questions in homogenization theory. These two questions have been well answered for the equation (1.1) in the case of one oscillating scale (i.e., n=1n=1); see recent monographs [She18, AK22] and references therein. For the case of multiple oscillating scales, we refer to our recent work [NZ23] for a brief survey. It is important to point out that before the work [NZ23], the homogenization theorem and uniform regularity were considered closely interconnected and always appear together, in the sense that the uniform regularity can be derived as a consequence of convergence rates either by a compactness method [AL87, KLS13] or a quantitative method (excess decay iteration [GNO20, AS16, She17] or a real-variable argument [CP98]). For the equation (1.1) with n2n\geq 2, we have the qualitative homogenization theorem or a quantitative convergence rate to a deterministic homogenized equation only if the scales (ε1,ε2,,εn)(\varepsilon_{1},\varepsilon_{2},\cdots,\varepsilon_{n}) are separated or well-separated [AB96], respectively. Recall that the scales (ε1,ε2,,εn)(0,1]n(\varepsilon_{1},\varepsilon_{2},\cdots,\varepsilon_{n})\in(0,1]^{n} are well-separated if εi+1εiαi\varepsilon_{i+1}\lesssim\varepsilon_{i}^{\alpha_{i}} for some αi>1\alpha_{i}>1. Consequently, under this scale-separation condition, the uniform regularity estimates for (1.1) were established in [NSX20] by a quantitative method. Yet, as pointed out by Avellaneda in [Ave96] (see also [GH03]) such an assumption is not adequate for treating the most general problems of transport and diffusion in self-similar (particularly random) media. The main contribution of [NZ23] is that, without any separation condition on (εi)1in(\varepsilon_{i})_{1\leq i\leq n}, we obtained the uniform CαC^{\alpha} estimate for any α(0,1)\alpha\in(0,1) by the compactness method combined with a novel scale-reduction process. This result was unexpected since it shows for the first time that the uniform regularity holds stably even without homogenization, and in some sense the averaging effect always takes places across all the mesoscopic scales.

This paper is a continuation of our previous work [NZ23]. It was conjectured in [NZ23] that with f=0f=0 in (1.1) and without any scale-separation condition, the solution admits uniform Lipschitz estimate. In this paper, we partially solve the conjecture by proving a slightly weaker result, namely, the gradient uε\nabla u_{\varepsilon} is uniformly bounded in LpL^{p} spaces for any p<p<\infty. In particular, this recovers the uniform CαC^{\alpha} estimate in [NZ23] in view of the Sobolev embedding theorem. Moreover, our proof is quantitative (in contrast to the compactness method in [NZ23]) in the sense that the constant can be computed explicitly.

Before stating the main result, we list the assumptions on the coefficient matrix AA:

  • Strong ellipticity condition: there exists Λ(0,1]\Lambda\in(0,1] so that

    Λ|ξ|2ξA(y1,,yn)ξΛ1|ξ|2\Lambda|\xi|^{2}\leq\xi\cdot A(y_{1},\cdots,y_{n})\xi\leq\Lambda^{-1}|\xi|^{2} (1.2)

    for every ξd\xi\in\mathbb{R}^{d} and for any (y1,,yn)d×n.(y_{1},\cdots,y_{n})\in\mathbb{R}^{d\times n}.

  • Periodicity: for any (y1,,yn)d×n(y_{1},\cdots,y_{n})\in\mathbb{R}^{d\times n} and (z1,,zn)d×n(z_{1},\cdots,z_{n})\in\mathbb{Z}^{d\times n},

    A(y1+z1,,yn+zn)=A(y1,,yn).A(y_{1}+z_{1},\cdots,y_{n}+z_{n})=A(y_{1},\cdots,y_{n}). (1.3)
  • Smoothness: there exists τ(0,1]\tau\in(0,1] and L>0L>0 such that

    |A(y1,,yn)A(y1,,yn)|L{|y1y1|++|ynyn|}τ|A(y_{1},\cdots,y_{n})-A(y_{1}^{\prime},\cdots,y_{n}^{\prime})|\leq L\big{\{}|y_{1}^{\prime}-y_{1}|+\cdots+|y_{n}^{\prime}-y_{n}|\big{\}}^{\tau} (1.4)

    for any (y1,,yn),(y1,,yn)d×n(y_{1},\cdots,y_{n}),(y_{1}^{\prime},\cdots,y_{n}^{\prime})\in\mathbb{R}^{d\times n}.

We now state our main result.

Theorem 1.1.

Let Ω\Omega be a bounded C1C^{1} domain. Assume AA satisfies (1.2), (1.3) and (1.4) and (εi)1in(0,1]n(\varepsilon_{i})_{1\leq i\leq n}\in(0,1]^{n}. Assume 1<p<1<p<\infty. Let uεu_{\varepsilon} be the weak solution to the Dirichlet problem

div(Aεuε)=divf in Ω,uε=0 on Ω,\displaystyle-\text{\rm div}\big{(}A_{\varepsilon}\nabla u_{\varepsilon}\big{)}=\text{\rm div}f\,\text{ in }\Omega,\quad u_{\varepsilon}=0\,\text{ on }\partial\Omega, (1.5)

where Aε=A(x/ε1,x/ε2,,x/εn)A_{\varepsilon}=A(x/\varepsilon_{1},x/\varepsilon_{2},\cdots,x/\varepsilon_{n}) and fLp(Ω)df\in L^{p}(\Omega)^{d}. Then uεLp(Ω)d\nabla u_{\varepsilon}\in L^{p}(\Omega)^{d} and

uεLp(Ω)CfLp(Ω),\displaystyle\|\nabla u_{\varepsilon}\|_{L^{p}(\Omega)}\leq C\|f\|_{L^{p}(\Omega)}, (1.6)

where CC depends only on d,n,Λ,pd,n,\Lambda,p, (τ,L)(\tau,L) in (1.4), and Ω\Omega.

The same uniform estimate holds if the Dirichlet boundary condition in (1.5) is replaced by the compatible Neumann boundary condition

νAεuε=νf on Ω,\nu\cdot A_{\varepsilon}\nabla u_{\varepsilon}=-\nu\cdot f\,\text{ on }\partial\Omega, (1.7)

where ν\nu is the unit outer normal vector of Ω\partial\Omega.

The new ingredient in the proof of Theorem 1.1 is the Dirichlet’s theorem (see Theorem 3.1) on simultaneous Diophantine approximation from number theory, which enters into a new technique of reperiodization and allow us to quantitatively separate one scale from the rest n1n-1 scales; see Section 3.1 for details. Precisely, for a given sequence (ε1,ε2,,εn)(\varepsilon_{1},\varepsilon_{2},\cdots,\varepsilon_{n}) in nonincreasing order and any number Q>1Q>1, we can find a new 1-periodic matrix AA^{\sharp}, depending on QQ and the ratios εn/εi\varepsilon_{n}/\varepsilon_{i}, such that Aε(x)A_{\varepsilon}(x) can be rewritten as

A(x/ε1,x/ε2,,x/εn)=A(x/ε1,x/ε2,,x/εn),A(x/\varepsilon_{1},x/\varepsilon_{2},\cdots,x/\varepsilon_{n})=A^{\sharp}(x/\varepsilon^{\prime}_{1},x/\varepsilon^{\prime}_{2},\cdots,x/\varepsilon^{\prime}_{n}), (1.8)

with some (ε1,ε2,,εn)(0,)n(\varepsilon_{1}^{\prime},\varepsilon_{2}^{\prime},\cdots,\varepsilon_{n}^{\prime})\in(0,\infty)^{n} and the last scale εn\varepsilon_{n}^{\prime} is QQ-separated from the rest εi\varepsilon_{i}^{\prime}’s (i.e., εiQεn\varepsilon^{\prime}_{i}\geq Q\varepsilon_{n}^{\prime} for all 1in11\leq i\leq n-1). Theoretically, QQ can be chosen as large as (εn/r)σ(\varepsilon_{n}/r)^{-\sigma} with some σ>0\sigma>0, yielding a well-separation condition. However, as a payoff the regularity of A(y1,y2,,yn)A^{\sharp}(y_{1},y_{2},\cdots,y_{n}) in yny_{n} will become worse as QQ increases, which leads to a small-scale estimate depending on QQ in a blow up argument. Fortunately, a sufficiently large QQ independent of (εi)1in(\varepsilon_{i})_{1\leq i\leq n} will be just enough for our proof. With the QQ-separation condition, we can perform a quantitative argument of the reiterated homogenization and approximate the solution of the original equation with nn oscillating scales by a solution of a new equation with at most n1n-1 oscillating scales. Hence, an inductive argument on the number of scales combined with a careful real-variable argument involving a double-averaging estimate will complete the proof.

It is crucial to point out that the proofs for CαC^{\alpha} estimate (a type of Schauder estimates) and LpL^{p} gradient estimate (Calderón-Zygmund estimate related to singular integrals) are essentially different, for the latter is stronger and more informative. As in [NZ23], the CαC^{\alpha} estimate can be derived with a qualitative HH-convergence theorem and a compactness method (the correctors are not needed; also see [She15, Theorem 3.1] or [SZ18, Theorem 6.1] for similar situations). However, the estimate of gradient must involve a quantitative convergence rate or the fine properties of correctors. While our proof of Theorem 1.1 also uses a scale-reduction theorem and an inductive argument on the number scales, it is exactly the Dirichlet’s theorem that helps us quantify the error in the scale-reduction process and provides much stronger implications and more information about the behavior of the solutions. In particular, it possibly explains the reason behind the occurrence of the mesoscopic averaging effect in arbitrary multiscale media.

As a direct consequence, Theorem 1.1 implies the uniform Calderón-Zygmund estimate for the operator div(A(x/ε))-\text{\rm div}(A(x/\varepsilon)\nabla) with arbitrary quasiperiodic coefficient AA. We recall that A(y)A(y) is said to be quasiperiodic if A(y)=B(My)A(y)=B(My), where B(w)B(w) is 11-periodic in wNw\in\mathbb{R}^{N} with 1N1\leq N\in\mathbb{N} and MM is an N×dN\times d constant real matrix. Note that we do not have any restriction on the entries of MM and therefore AA is allowed to be periodic with a very degenerate periodic cell.

Corollary 1.2.

Let Ω\Omega be a bounded C1C^{1} domain. Assume that A(y)=B(My)A(y)=B(My) is quasi-periodic as above and BB satisfies (1.2), (1.3) and (1.4) (with n=1n=1). Let 1<p<1<p<\infty, fLp(Ω)df\in L^{p}(\Omega)^{d}, and uεu_{\varepsilon} be the weak solution to

div(A(x/ε)uε)=divf in Ω,uε=0 on Ω.\displaystyle-\text{\rm div}\big{(}A(x/\varepsilon)\nabla u_{\varepsilon}\big{)}=\text{\rm div}f\,\text{ in }\Omega,\quad u_{\varepsilon}=0\,\text{ on }\partial\Omega.

Then uεLp(Ω)d\nabla u_{\varepsilon}\in L^{p}(\Omega)^{d} and

uεLp(Ω)CfLp(Ω),\displaystyle\|\nabla u_{\varepsilon}\|_{L^{p}(\Omega)}\leq C\|f\|_{L^{p}(\Omega)},

where CC depends only on d,Λ,p,Ω,Nd,\Lambda,p,\Omega,N, and (τ,L)(\tau,L) in (1.4).

In the setting of quasiperiodic homogenization, the convergence rate was first obtained by Kozlov in [Koz78] under a Diophantine condition (a quantitative ergodicity condition): for each row MiM_{i} of MM, |Miz|c0|z|β|M_{i}\cdot z|\geq c_{0}|z|^{-\beta} for all zd{0}z\in\mathbb{Z}^{d}\setminus\{0\} and for some c0,β>0c_{0},\beta>0. It was observed in [AL91] that the uniform regularity is valid under Kozlov’s Diophantine condition. This condition was also required in recent work [She15, AGK16, BG19] applied to the case of quasiperiodic coefficients. In our Corollary 1.2, we do not need the Diophantine condition and the constant in the estimate is independent of the entries of the matrix MM (except for the dimension of MM). This includes some typical degenerate cases, such as periodic coefficients with a degenerate periodic cell (e.g., a thin rectangle), quasiperiodic coefficients oscillating at two almost resonant frequencies (e.g., the ratio of the frequencies is a Liouville number), multiscale quasiperiodic coefficients, etc.

The proof of Corollary 1.2 based on Theorem 1.1 is simple and we include it here. In fact, it suffices to write A(x/ε)A(x/\varepsilon) into a periodic coefficient matrix with multiple oscillating scales and apply Theorem 1.1. To this end, let M=(Mij)1iN,1jdM=(M_{ij})_{1\leq i\leq N,1\leq j\leq d}. Let yijdy_{ij}\in\mathbb{R}^{d} with 1iN1\leq i\leq N and 1jd1\leq j\leq d be NdNd independent variables. Define a matrix with variables yijy_{ij}

A+(y11,,yij,,yNd)=B(j=1dy1jej,,j=1dyNjej).A^{+}(y_{11},\cdots,y_{ij},\cdots,y_{Nd})=B\big{(}\sum_{j=1}^{d}y_{1j}\cdot e_{j},\cdots,\sum_{j=1}^{d}y_{Nj}\cdot e_{j}\big{)}.

Then A+A^{+} is 1-periodic in each yijy_{ij} since B(w)B(w) is 1-periodic. Moreover, it is straightforward to verify

A(x/ε)=B(Mx/ε)=A+(M11x/ε,,Mijx/ε,,MNdx/ε).A(x/\varepsilon)=B(Mx/\varepsilon)=A^{+}(M_{11}x/\varepsilon,\cdots,M_{ij}x/\varepsilon,\cdots,M_{Nd}x/\varepsilon). (1.9)

This reduces the quasiperiodic coefficient matrix with one oscillating scale into a periodic coefficient matrix with NdNd oscillating scales. Clearly, in this case, all the scales are possibly not well-separated. Moreover, if BB is uniformly elliptic and Hölder continuous, so is A+A^{+}. As a result, Corollary 1.2 follows readily from Theorem 1.1.

Finally, using a similar idea of scale separation, we can show the uniform large-scale or mesoscopic-scale Lipschitz estimates. In particular, in the case of two scales, i.e., n=2n=2, for any α(0,1)\alpha\in(0,1), we have for all ε21αr1\varepsilon_{2}^{1-\alpha}\leq r\leq 1,

(Br|uε|2)1/2Cα{(B1|uε|2)1/2+(B1|F|p)1/p},\bigg{(}\fint_{B_{r}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}\leq C_{\alpha}\bigg{\{}\bigg{(}\fint_{B_{1}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}+\bigg{(}\fint_{B_{1}}|F|^{p}\bigg{)}^{1/p}\bigg{\}}, (1.10)

where CαC_{\alpha} depends only on d,Λ,p,τ,Ld,\Lambda,p,\tau,L and α\alpha; see Theorem 5.1. For general cases n3n\geq 3, we have some mesoscopic-scale Lipschitz estimates near any given scale; see Theorem 5.2.

Organization. The rest of the paper is organized as follows. In Section 2, we recall some knowledge and known results from locally periodic homogenization and reiterated homogenization. In Section 3, we use the Dirichlet’s theorem and reperiodization technique to separate scales and reduce one scale by a uniform approximation. The main theorem is proved in Section 4 by a real-variable argument. In Section 5, we exploit the idea of scale separation to obtain the uniform large-scale or mesoscopic-scale Lipschitz estimates.

Acknowledgements. W. Niu is supported by NNSF of China (12371106, 11971031). J. Zhuge is partially supported by grants for Excellent Youth from the NSFC.

2. Preliminaries

2.1. Local periodic homogenization

In this subsection, we recall some results concerning the homogenization of the locally periodic operator div(A(x,x/ε))-\text{\rm div}(A(x,x/\varepsilon)\nabla) with a scalar ε(0,1)\varepsilon\in(0,1), which will be used in our quantitative scale reduction process in Section 3.

Suppose A(x,y)A(x,y) satisfies the ellipticity condition (1.2) and is 1-periodic in ydy\in\mathbb{R}^{d}. The effective (homogenized) matrix is given by

A^(x)=𝕋d(A(x,y)+A(x,y)yχ(x,y))𝑑y,\displaystyle\widehat{A}(x)=\fint_{\mathbb{T}^{d}}\Big{(}A(x,y)+A(x,y)\nabla_{y}\chi(x,y)\Big{)}dy, (2.1)

where χ(x,y)=(χ1(x,y),,χd(x,y))\chi(x,y)=(\chi_{1}(x,y),\dots,\chi_{d}(x,y)) is the corrector given by the cell problem

{divy(A(x,y)yχj)=divy(A(x,y)yyj) in 𝕋d:=d/d,χj=χj(x,y) is 1-periodic in y,𝕋dχj(x,y)𝑑y=0,\left\{\begin{aligned} &-\text{\rm div}_{y}\big{(}A(x,y)\nabla_{y}\chi_{j})=\text{\rm div}_{y}\big{(}A(x,y)\nabla_{y}y^{j}\big{)}\quad\text{ in }\mathbb{T}^{d}:=\mathbb{R}^{d}/\mathbb{Z}^{d},\\ &\chi_{j}=\chi_{j}(x,y)\text{ is 1-periodic in }y,\\ &\int_{\mathbb{T}^{d}}\chi_{j}(x,y)\,dy=0,\end{aligned}\right. (2.2)

for 1jd1\leq j\leq d. Here yjy^{j} denotes the jjth component of ydy\in\mathbb{R}^{d}. Assume that

xA=xAL(xd×yd)<.\|\nabla_{x}A\|_{\infty}=\|\nabla_{x}A\|_{L^{\infty}(\mathbb{R}_{x}^{d}\times\mathbb{R}^{d}_{y})}<\infty. (2.3)

The standard energy estimates for (2.2) imply that

yχL2(Y)C,xyχL(xd;L2(Y))CxA\|\nabla_{y}\chi\|_{L^{2}(Y)}\leq C,\quad\|\nabla_{x}\nabla_{y}\chi\|_{L^{\infty}(\mathbb{R}_{x}^{d};L^{2}(Y))}\leq C\|\nabla_{x}A\|_{\infty} (2.4)

with CC depending only on Λ,d.\Lambda,d. This combined with the definition of A^\widehat{A} yields

xA^CxA.\|\nabla_{x}\widehat{A}\|_{\infty}\leq C\|\nabla_{x}A\|_{\infty}.

Likewise, if

|A(x,y)A(x,y)|L|xx|τ|A(x,y)-A(x^{\prime},y)|\leq L|x-x^{\prime}|^{\tau} (2.5)

for some 0<τ10<\tau\leq 1 for any x,x,yd.x,x^{\prime},y\in\mathbb{R}^{d}. Then we have

|A^(x)A^(x)|CL|xx|τ.\displaystyle|\widehat{A}(x)-\widehat{A}(x^{\prime})|\leq CL|x-x^{\prime}|^{\tau}. (2.6)

Let φC0(d)\varphi\in C_{0}^{\infty}(\mathbb{R}^{d}) be supported in B1/2(0)B_{1/2}(0) such that φ0\varphi\geq 0 and dφ𝑑x=1\int_{\mathbb{R}^{d}}\varphi dx=1. For functions of the form gε(x)=g(x,x/ε)g^{\varepsilon}(x)=g(x,x/\varepsilon), we define a partial-smoothing operator

Sε(gε)(x)=εddg(z,x/ε)φ((xz)/ε)𝑑z.\displaystyle S_{\varepsilon}(g^{\varepsilon})(x)=\varepsilon^{-d}\int_{\mathbb{R}^{d}}g(z,x/\varepsilon)\varphi((x-z)/\varepsilon)dz. (2.7)

The following theorem provides the convergence rate in H1(Ω)H^{1}(\Omega) for the locally periodic operator. The proof may be found in [NSX20].

Theorem 2.1.

Let Ω\Omega be a bounded Lipschitz domain in d\mathbb{R}^{d}. Let uεu_{\varepsilon} be the solution to div(A(x,x/ε)uε)=F-\text{\rm div}(A(x,x/\varepsilon)\nabla u_{\varepsilon})=F in Ω\Omega with uε=gu_{\varepsilon}=g on Ω,\partial\Omega, and u0u_{0} the solution to div(A^(x)u0)=F-\text{\rm div}(\widehat{A}(x)\nabla u_{0})=F in Ω\Omega with u0=gu_{0}=g on Ω.\partial\Omega. Then we have

uεu0Sε((yχ)εu0)L2(Ω)Cε{(1+xA)u0L2(Ω)+2u0L2(ΩΩ3ε)}+Cu0L2(Ω4ε),\displaystyle\begin{split}&\|\nabla u_{\varepsilon}-\nabla u_{0}-S_{\varepsilon}((\nabla_{y}\chi)^{\varepsilon}\nabla u_{0})\|_{L^{2}(\Omega)}\\ &\leq C\varepsilon\big{\{}(1+\|\nabla_{x}A\|_{\infty})\|\nabla u_{0}\|_{L^{2}(\Omega)}+\|\nabla^{2}u_{0}\|_{L^{2}(\Omega\setminus\Omega_{3\varepsilon})}\big{\}}+C\|\nabla u_{0}\|_{L^{2}(\Omega_{4\varepsilon})},\end{split} (2.8)

where we have extended u0u_{0} to the whole space d,\mathbb{R}^{d}, Ωt={xΩ:dist(x,Ω)<t}\Omega_{t}=\big{\{}x\in\Omega:\text{\rm dist}(x,\partial\Omega)<t\big{\}}, and CC depends only on dd, Λ\Lambda and Ω\Omega.

We emphasize that in the above theorem, AA has no smoothness assumption in yy.

2.2. Reiterated correctors and effective matrix

Consider the operator with nn scales div(A(x/ε1,x/ε2,,x/εn))-\text{\rm div}(A(x/\varepsilon_{1},x/\varepsilon_{2},\cdots,x/\varepsilon_{n})\nabla), where AA satisfies the ellipticity and periodicity conditions (1.2) and (1.3). Assume that εn\varepsilon_{n} is separated from εi\varepsilon_{i} for 1in1.1\leq i\leq n-1. By using the idea of reiterated homogenization, one can view (εi)1in1(\varepsilon_{i})_{1\leq i\leq n-1} as parameters and homogenize the finest scale εn\varepsilon_{n} to obtain an operator of the same type but with only n1n-1 scales.

Indeed rewrite

𝒜ε(x,y):=A(x/ε1,x/ε2,,x/εn1,y).\mathcal{A}_{\varepsilon}(x,y):=A(x/\varepsilon_{1},x/\varepsilon_{2},\cdots,x/\varepsilon_{n-1},y).

It is obvious that 𝒜ε(x,y)\mathcal{A}_{\varepsilon}(x,y) satisfies (1.2), and is locally 1-periodic in yy. Let χε(x,y)\chi_{\varepsilon}(x,y) be the corrector given by (2.2) with A(x,y)A(x,y) replaced by 𝒜ε(x,y)\mathcal{A}_{\varepsilon}(x,y), and 𝒜^ε(x)\widehat{\mathcal{A}}_{\varepsilon}(x) the effective matrix of 𝒜ε\mathcal{A}_{\varepsilon} given by (2.1). In view of the structure of 𝒜ε(x,y)\mathcal{A}_{\varepsilon}(x,y), the corresponding corrector χε(x,y)\chi_{\varepsilon}(x,y) takes a form of

χε(x,y)=𝒳(x/ε1,,x/εn1,y),\chi_{\varepsilon}(x,y)=\mathcal{X}(x/\varepsilon_{1},\cdots,x/\varepsilon_{n-1},y), (2.9)

where the reiterated corrector 𝒳=(𝒳j)\mathcal{X}=(\mathcal{X}_{j}) is defined by the cell problem

{divy(A(y1,,yn1,y)y𝒳j)=divy(A(y1,,yn1,y)yyj) in 𝕋d,𝒳j=𝒳j(y1,,yn1,y) is 1-periodic in y,𝕋d𝒳j(y1,,yn1,y)𝑑y=0,\left\{\begin{aligned} &-\text{\rm div}_{y}\big{(}A(y_{1},\cdots,y_{n-1},y)\nabla_{y}\mathcal{X}_{j})=\text{\rm div}_{y}\big{(}A(y_{1},\cdots,y_{n-1},y)\nabla_{y}y^{j}\big{)}\quad\text{ in }\mathbb{T}^{d},\\ &\mathcal{X}_{j}=\mathcal{X}_{j}(y_{1},\cdots,y_{n-1},y)\text{ is 1-periodic in }y,\\ &\int_{\mathbb{T}^{d}}\mathcal{X}_{j}(y_{1},\cdots,y_{n-1},y)\,dy=0,\end{aligned}\right. (2.10)

for 1jd1\leq j\leq d, which is also 1-periodic in y1,,yn1y_{1},\cdots,y_{n-1}, due to the periodicity of AA. Moreover, 𝒜^ε(x)\widehat{\mathcal{A}}_{\varepsilon}(x) takes a form of

𝒜^ε(x)=𝒜^(x/ε1,,x/εn1)\widehat{\mathcal{A}}_{\varepsilon}(x)=\widehat{\mathcal{A}}(x/\varepsilon_{1},\cdots,x/\varepsilon_{n-1}) (2.11)

with

𝒜^(y1,y2,,yn1)=𝕋dA(y1,,yn1,y)(I+y𝒳(y1,,yn1,y))𝑑y.\widehat{\mathcal{A}}(y_{1},y_{2},\cdots,y_{n-1})=\fint_{\mathbb{T}^{d}}A(y_{1},\cdots,y_{n-1},y)\Big{(}I+\nabla_{y}\mathcal{X}(y_{1},\cdots,y_{n-1},y)\Big{)}dy. (2.12)

It is obvious that 𝒜^(y1,y2,,yn1)\widehat{\mathcal{A}}(y_{1},y_{2},\cdots,y_{n-1}) is 1-periodic in each yiy_{i} with 1in11\leq i\leq n-1, and it is defined independent of (ε1,,εn)(\varepsilon_{1},\cdots,\varepsilon_{n}). Similar to the case n=1n=1, 𝒜^\widehat{\mathcal{A}} satisfies the ellipticity condition (1.2). Furthermore, similar to (2.6), if AA is Hölder continuous in yiy_{i} for 1in11\leq i\leq n-1, so is 𝒜^\widehat{\mathcal{A}}.

3. Quantitative scale separation and scale reduction

3.1. Scale separation

The key ingredient of this paper is the Dirichlet’s theorem on the simultaneous Diophantine approximation. The connection between the Dirichlet’s theorem and the regularity theory in homogenization is previously unknown. The only loosely related notion is the Diophantine condition imposed in the setting of quasiperiodic homogenization, as mentioned earlier. The Dirichlet’s theorem appears to be a powerful tool (instead of a condition) that eventually allows us to derive the uniform Calderón-Zygmund estimates in multiscale or quasiperiodic homogenization without any additional conditions.

Let us first recall the Dirichlet’s theorem; see [Sch91].

Theorem 3.1 (Dirichlet (1842)).

Suppose that α1,α2,,αm\alpha_{1},\alpha_{2},\cdots,\alpha_{m} are mm real numbers and Q>1Q>1. There exist integers q,p1,p2,,pmq,p_{1},p_{2},\cdots,p_{m} such that 1q<Qm1\leq q<Q^{m} and

sup1im|αipiq|<1qQ.\sup_{1\leq i\leq m}\Big{|}\alpha_{i}-\frac{p_{i}}{q}\Big{|}<\frac{1}{qQ}. (3.1)

Roughly speaking, the Dirichlet’s theorem states that any real numbers can be approximated by good rational numbers with a quantitative small error. In the rest of this subsection, we apply the Dirichlet’s theorem to separate at least one scale in the coefficient matrix oscillating at multiple (possibly unseparated) scales.

Without loss of generality, hereafter we always assume

1ε1ε2εn>0.1\geq\varepsilon_{1}\geq\varepsilon_{2}\geq\cdots\geq\varepsilon_{n}>0. (3.2)

Let Q1Q\gg 1 be a large number. We emphasize that the fixed number QQ will be chosen later independent of {ε1,ε2,,εn}\{\varepsilon_{1},\varepsilon_{2},\cdots,\varepsilon_{n}\}. We say aa is QQ-separated from bb if bQab\geq Qa. This scale separation condition is crucial in quantitative reiterated homogenization.

Given Q>1Q>1 as above, by the Dirichlet’s theorem with αi=εn/εi(0,1]\alpha_{i}=\varepsilon_{n}/\varepsilon_{i}\in(0,1] for i=1,2,,n1i=1,2,\cdots,n-1, we can find integers q,p1,p2,,pn1q,p_{1},p_{2},\cdots,p_{n-1} such that 1q<Qn11\leq q<Q^{n-1} and

sup1in1|εnεipiq|<1qQ.\sup_{1\leq i\leq n-1}\Big{|}\frac{\varepsilon_{n}}{\varepsilon_{i}}-\frac{p_{i}}{q}\Big{|}<\frac{1}{qQ}. (3.3)

Moreover, because of (3.2), we have 0pipi+1q<Qn10\leq p_{i}\leq p_{i+1}\leq q<Q^{n-1}. Note that it is possible that pi=0p_{i}=0 if εi\varepsilon_{i} is large enough that εn/εi<1/Q\varepsilon_{n}/\varepsilon_{i}<1/Q (i.e., εn\varepsilon_{n} is already QQ-separated from εi\varepsilon_{i}).

Define, for 1in11\leq i\leq n-1,

γi:=|εnεipiq|andsi=sgn(εnεipiq).\gamma_{i}:=\Big{|}\frac{\varepsilon_{n}}{\varepsilon_{i}}-\frac{p_{i}}{q}\Big{|}\quad\text{and}\quad s_{i}=\text{sgn}\Big{(}\frac{\varepsilon_{n}}{\varepsilon_{i}}-\frac{p_{i}}{q}\Big{)}. (3.4)

Then, we can write

1εi=γisiεn+piq1εn.\frac{1}{\varepsilon_{i}}=\frac{\gamma_{i}s_{i}}{\varepsilon_{n}}+\frac{p_{i}}{q}\frac{1}{\varepsilon_{n}}. (3.5)

Now, we define

A(y1,y2,,yn)=A(s1y1+p1yn,s2y2+p2yn,,sn1yn1+pn1yn,qyn).A^{\sharp}(y_{1},y_{2},\cdots,y_{n})=A(s_{1}y_{1}+p_{1}y_{n},s_{2}y_{2}+p_{2}y_{n},\cdots,s_{n-1}y_{n-1}+p_{n-1}y_{n},qy_{n}). (3.6)

The key insight here is that, since si=±1s_{i}=\pm 1 and p1,p2,,pnp_{1},p_{2},\cdots,p_{n} are all integers, AA^{\sharp} is 1-periodic in each yiy_{i}. We thereby will call the transformation from AA to AA^{\sharp} a technique of reperiodization. As a result of (3.5) and (3.6), we have

A(xε1,,xεn)=A(γ1xεn,γ2xεn,,γn1xεn,xqεn).A\big{(}\frac{x}{\varepsilon_{1}},\cdots,\frac{x}{\varepsilon_{n}}\big{)}=A^{\sharp}\big{(}\frac{\gamma_{1}x}{\varepsilon_{n}},\frac{\gamma_{2}x}{\varepsilon_{n}},\cdots,\frac{\gamma_{n-1}x}{\varepsilon_{n}},\frac{x}{q\varepsilon_{n}}\big{)}. (3.7)

In other words, the reperiodization allows us to rewrite the original periodic matrix AA oscillating at scales {ε1,ε2,,εn}\{\varepsilon_{1},\varepsilon_{2},\cdots,\varepsilon_{n}\} into a new periodic matrix AA^{\sharp} oscillating at different scales {εn/γ1,εn/γ2,,εn/γn1,qεn}\{\varepsilon_{n}/\gamma_{1},\varepsilon_{n}/\gamma_{2},\cdots,\varepsilon_{n}/\gamma_{n-1},q\varepsilon_{n}\}. Note that if pi=0p_{i}=0 for some ii, then the corresponding scale for that (slower) variable does not change, i.e., εn/γi=εi\varepsilon_{n}/\gamma_{i}=\varepsilon_{i}. If γi=0\gamma_{i}=0 for some ii, then we just do not have the scale εn/γi\varepsilon_{n}/\gamma_{i} and we have less oscillating scales after reperiodization (which is even better). Without loss of generality, we may assume this does not happen.

The new scales and the new 1-periodic matrix AA^{\sharp} have the following crucial properties:

  • The smallest scale qεnq\varepsilon_{n} is QQ-separated from εn/γi\varepsilon_{n}/\gamma_{i} for each 1in11\leq i\leq n-1. In fact, (3.3) implies

    εn/γiqεn=1qγiQ.\frac{\varepsilon_{n}/\gamma_{i}}{q\varepsilon_{n}}=\frac{1}{q\gamma_{i}}\geq Q. (3.8)

    This is the key property for our application.

  • If AA is Hölder continuous in yiy_{i} for 1in11\leq i\leq n-1, i.e.,

    |A(y1,,yn1,yn)A(y1,,yn1,yn)|\displaystyle|A(y_{1},\cdots,y_{n-1},y_{n})-A(y_{1}^{\prime},\cdots,y_{n-1}^{\prime},y_{n})| (3.9)
    L{|y1y1|++|yn1yn1|}τ\displaystyle\leq L\big{\{}|y_{1}^{\prime}-y_{1}|+\cdots+|y_{n-1}^{\prime}-y_{n-1}|\big{\}}^{\tau}

    with 0<L,0<τ10<L,0<\tau\leq 1 for any (y1,,yn),(y1,,yn1,yn)d×n(y_{1},\cdots,y_{n}),(y_{1}^{\prime},\cdots,y_{n-1}^{\prime},y_{n})\in\mathbb{R}^{d\times n}, then AA^{\sharp} is also Hölder continuous in yiy_{i} for 1in11\leq i\leq n-1 with the same constants (τ,L)(\tau,L). However, AA^{\sharp} may lose good regularity in yny_{n} since many periods have been compressed into a single 1-periodic cell.

  • The construction of AA^{\sharp} relies on the choice of QQ as well as the ratios εn/εi\varepsilon_{n}/\varepsilon_{i} with 1in11\leq i\leq n-1. In other words, AA^{\sharp} is scale-invariant. Also note that QQ is a dimensionless parameter.

It is important to note that the above scale separation process can only separate the smallest scale from the rest n1n-1 larger scales, while the relationships among the rest n1n-1 larger scales cannot be determined. This allows us to perform a one-scale reduction by using the idea of reiterated homogenization.

3.2. Scale reduction and uniform approximation

In this subsection we use the scale separation technique in Section 3.1 to reduce at least one scale. Precisely, we shall prove that the solution of (1.1) can be uniformly approximated by a solution to the equation of the same type but with at most n1n-1 oscillating scales. The following is the main theorem.

Theorem 3.2.

Suppose that AA satisfies the assumptions (1.2), (1.3) and (3.9). Let uεu_{\varepsilon} be a weak solution to

div(A(x/ε1,x/ε2,,x/εn)uε)=Fin B2r:=B(0,2r).\displaystyle-\text{\rm div}\big{(}A(x/\varepsilon_{1},x/\varepsilon_{2},\cdots,x/\varepsilon_{n})\nabla u_{\varepsilon}\big{)}=F\quad\text{in }B_{2r}:=B(0,2r). (3.10)

Then for any Q>1Q>1, there exists a coefficient matrix A=A(y1,,yn1)A^{\flat}=A^{\flat}(y_{1},\cdots,y_{n-1}), scales εk(0,)\varepsilon_{k}^{\prime}\in(0,\infty) with k=1,,n1k=1,\cdots,n-1, and a weak solution uεu^{\flat}_{\varepsilon^{\prime}} to

div(A(x/ε1,x/ε2,,x/εn1)uε)=Fin Br,-\text{\rm div}\big{(}A^{\flat}(x/\varepsilon^{\prime}_{1},x/\varepsilon^{\prime}_{2},\cdots,x/\varepsilon^{\prime}_{n-1})\nabla u^{\flat}_{\varepsilon^{\prime}}\big{)}=F\quad\text{in }B_{r},

such that

uεL2(Br)C{uεL2(B2r)+rFL2(B2r)},\|\nabla u^{\flat}_{\varepsilon^{\prime}}\|_{L^{2}(B_{r})}\leq C\big{\{}\|\nabla u_{\varepsilon}\|_{L^{2}(B_{2r})}+r\|F\|_{L^{2}(B_{2r})}\big{\}}, (3.11)

and

uεuεUεL2(Br)C{(Qn1εnr)σ+Qτ}{uεL2(B2r)+rFL2(B2r)}\|\nabla u_{\varepsilon}-\nabla u^{\flat}_{\varepsilon^{\prime}}-U_{\varepsilon^{\prime}}\|_{L^{2}(B_{r})}\leq C\bigg{\{}\bigg{(}\frac{Q^{n-1}\varepsilon_{n}}{r}\bigg{)}^{\sigma}+Q^{-\tau}\bigg{\}}\big{\{}\|\nabla u_{\varepsilon}\|_{L^{2}(B_{2r})}+r\|F\|_{L^{2}(B_{2r})}\big{\}} (3.12)

for some σ>0,\sigma>0, where UεU_{\varepsilon^{\prime}} is given by

Uε(x)=(εn)ddφ(xzεn)(y𝒳)(zε1,,zεn1,xεn)uε(z)𝑑z,U_{\varepsilon^{\prime}}(x)=(\varepsilon_{n}^{\prime})^{-d}\int_{\mathbb{R}^{d}}\varphi\bigg{(}\frac{x-z}{\varepsilon_{n}^{\prime}}\bigg{)}(\nabla_{y}\mathcal{X}^{\sharp})\bigg{(}\frac{z}{\varepsilon_{1}^{\prime}},\cdots,\frac{z}{\varepsilon^{\prime}_{n-1}},\frac{x}{\varepsilon_{n}^{\prime}}\bigg{)}\nabla u^{\flat}_{\varepsilon^{\prime}}(z)dz, (3.13)

𝒳\mathcal{X}^{\sharp} is the corrector given by (2.10) with AA replaced by AA^{\sharp}, and εn\varepsilon^{\prime}_{n} is given by (3.18). Moreover, the constant CC depends only on d,n,Λd,n,\Lambda and (τ,L)(\tau,L) in (3.9), and the assumptions (1.2), (1.3) and (3.9) are preserved for AA^{\flat}.

The proof of Theorem 3.2 relies on the scale separation technique in Section 3.1 and the following approximation theorem that pertains to the quantitative reiterated homogenization.

Theorem 3.3.

Suppose AA satisfies the assumption (1.2), (1.3), and (3.9). Assume εn\varepsilon_{n} is QQ-separated from εk,k=1,,n1\varepsilon_{k},k=1,\cdots,n-1. Let uεu_{\varepsilon} be a weak solution to

div(A(x/ε1,x/ε2,,x/εn)uε)=F in B2.-\text{\rm div}\big{(}A(x/\varepsilon_{1},x/\varepsilon_{2},\cdots,x/\varepsilon_{n})\nabla u_{\varepsilon}\big{)}=F\quad\text{ in }B_{2}. (3.14)

Then there exist a 1-periodic matrix 𝒜^=𝒜^(y1,,yn1)\widehat{\mathcal{A}}=\widehat{\mathcal{A}}(y_{1},\cdots,y_{n-1}) with n1n-1 scales and a weak solution u^ε\widehat{u}_{\varepsilon} to

div(𝒜^(x/ε1,x/ε2,,x/εn1)u^ε)=F in B1,-\text{\rm div}\big{(}\widehat{\mathcal{A}}(x/\varepsilon_{1},x/\varepsilon_{2},\cdots,x/\varepsilon_{n-1})\nabla\widehat{u}_{\varepsilon}\big{)}=F\quad\text{ in }B_{1},

such that

u^εL2(B1)C{uεL2(B2)+FL2(B2)},\|\nabla\widehat{u}_{\varepsilon}\|_{L^{2}(B_{1})}\leq C\big{\{}\|\nabla u_{\varepsilon}\|_{L^{2}(B_{2})}+\|F\|_{L^{2}(B_{2})}\big{\}}, (3.15)

and

uεu^εSεn((y𝒳)εu^ε)L2(B1)C(Qτ+εnσ){uεL2(B2)+FL2(B2)},\|\nabla u_{\varepsilon}-\nabla\widehat{u}_{\varepsilon}-S_{\varepsilon_{n}}((\nabla_{y}\mathcal{X})^{\varepsilon}\nabla\widehat{u}_{\varepsilon})\|_{L^{2}(B_{1})}\leq C(Q^{-\tau}+\varepsilon_{n}^{\sigma})\big{\{}\|\nabla u_{\varepsilon}\|_{L^{2}(B_{2})}+\|F\|_{L^{2}(B_{2})}\big{\}}, (3.16)

for some σ>0\sigma>0, where 𝒳\mathcal{X} and 𝒜^\widehat{\mathcal{A}} are defined by (2.10) and (2.12), respectively, and

Sεn((y𝒳)εu^ε)(x)=(εn)ddφ(xzεn)(y𝒳)(zε1,,zεn1,xεn)u^ε(z)𝑑z.S_{\varepsilon_{n}}((\nabla_{y}\mathcal{X})^{\varepsilon}\nabla\widehat{u}_{\varepsilon})(x)=(\varepsilon_{n})^{-d}\int_{\mathbb{R}^{d}}\varphi\bigg{(}\frac{x-z}{\varepsilon_{n}}\bigg{)}(\nabla_{y}\mathcal{X})\bigg{(}\frac{z}{\varepsilon_{1}},\cdots,\frac{z}{\varepsilon_{n-1}},\frac{x}{\varepsilon_{n}}\bigg{)}\nabla\widehat{u}_{\varepsilon}(z)dz. (3.17)

Moreover, the constant CC depends only on d,n,Λd,n,\Lambda and (τ,L)(\tau,L) in (3.9).

We first prove Theorem 3.2, assuming Theorem 3.3.

Proof of Theorem 3.2.

We first consider the case r=1r=1. Given 1ε1ε2εn>01\geq\varepsilon_{1}\geq\varepsilon_{2}\geq\cdots\geq\varepsilon_{n}>0 and Q>1Q>1, by the Dirichlet’s theorem, we can find integers q,p1,p2,,pn1q,p_{1},p_{2},\cdots,p_{n-1} such that 1q<Qn11\leq q<Q^{n-1}, and (3.3)-(3.8) hold. Set

εi\displaystyle\varepsilon_{i}^{\prime} =εi/γi,for i=1,2,,n1,\displaystyle=\varepsilon_{i}/\gamma_{i},\quad\text{for }i=1,2,\cdots,n-1, (3.18)
εn\displaystyle\varepsilon_{n}^{\prime} =qεn.\displaystyle=q\varepsilon_{n}.

By (3.7) and (3.10), we know that uεu_{\varepsilon} satisfies

div(A(x/ε1,,x/εn)uε)=F in B2.\displaystyle-\text{\rm div}\big{(}A^{\sharp}(x/\varepsilon^{\prime}_{1},\cdots,x/\varepsilon^{\prime}_{n})\nabla u_{\varepsilon}\big{)}=F\,\text{ in }B_{2}. (3.19)

By (3.8), εn\varepsilon_{n}^{\prime} is QQ-separated from εi\varepsilon_{i}^{\prime} for all i=1,2,,n1i=1,2,\cdots,n-1. Thus, by Theorem 3.3, there exist 𝒜^=𝒜^(y1,,yn1)\widehat{\mathcal{A}}^{\sharp}=\widehat{\mathcal{A}}^{\sharp}(y_{1},\cdots,y_{n-1}) given by (2.12) with AA replaced by AA^{\sharp}, and a weak solution u^ε\widehat{u}_{\varepsilon^{\prime}} to

div(𝒜^(x/ε1,,x/εn1)u^ε)=Fin B1,-\text{\rm div}\big{(}\widehat{\mathcal{A}}^{\sharp}(x/\varepsilon^{\prime}_{1},\cdots,x/\varepsilon^{\prime}_{n-1})\nabla\widehat{u}_{\varepsilon^{\prime}}\big{)}=F\quad\text{in }B_{1}, (3.20)

such that

u^εL2(B1)C{uεL2(B2)+FL2(B2)},\|\nabla\widehat{u}_{\varepsilon^{\prime}}\|_{L^{2}(B_{1})}\leq C\{\|\nabla u_{\varepsilon}\|_{L^{2}(B_{2})}+\|F\|_{L^{2}(B_{2})}\}, (3.21)

and

uεu^εSεn((y𝒳)εu^ε)L2(B1)\displaystyle\|\nabla u_{\varepsilon}-\nabla\widehat{u}_{\varepsilon^{\prime}}-S_{\varepsilon_{n}^{\prime}}((\nabla_{y}\mathcal{X}^{\sharp})^{\varepsilon^{\prime}}\nabla\widehat{u}_{\varepsilon^{\prime}})\|_{L^{2}(B_{1})} (3.22)
C(Qτ+(εn)σ){uεL2(B2)+FL2(B2)}\displaystyle\leq C(Q^{-\tau}+(\varepsilon^{\prime}_{n})^{\sigma})\{\|\nabla u_{\varepsilon}\|_{L^{2}(B_{2})}+\|F\|_{L^{2}(B_{2})}\}
C(Qτ+(Qn1εn)σ){uεL2(B2)+FL2(B2)},\displaystyle\leq C(Q^{-\tau}+(Q^{n-1}\varepsilon_{n})^{\sigma})\{\|\nabla u_{\varepsilon}\|_{L^{2}(B_{2})}+\|F\|_{L^{2}(B_{2})}\},

for some σ>0\sigma>0, where 𝒳\mathcal{X}^{\sharp} is the corrector defined in (2.10) with AA replaced by AA^{\sharp}, and

Sεn((y𝒳)εu^ε)=(εn)ddφ(xzεn)(y𝒳)(zε1,,zεn1,xεn)u^ε(z)𝑑z.\displaystyle\begin{split}S_{\varepsilon_{n}^{\prime}}((\nabla_{y}\mathcal{X}^{\sharp})^{\varepsilon^{\prime}}\nabla\widehat{u}_{\varepsilon^{\prime}})&=(\varepsilon_{n}^{\prime})^{-d}\int_{\mathbb{R}^{d}}\varphi\bigg{(}\frac{x-z}{\varepsilon_{n}^{\prime}}\bigg{)}(\nabla_{y}\mathcal{X}^{\sharp})\left(\frac{z}{\varepsilon_{1}^{\prime}},\cdots,\frac{z}{\varepsilon^{\prime}_{n-1}},\frac{x}{\varepsilon_{n}^{\prime}}\right)\nabla\widehat{u}_{\varepsilon^{\prime}}(z)dz.\end{split}

We rename A=𝒜^,uε=u^εA^{\flat}=\widehat{\mathcal{A}}^{\sharp},u^{\flat}_{\varepsilon^{\prime}}=\widehat{u}_{\varepsilon^{\prime}}, Uε=Sεn((y𝒳)εu^ε)U_{\varepsilon^{\prime}}=S_{\varepsilon_{n}^{\prime}}((\nabla_{y}\mathcal{X}^{\sharp})^{\varepsilon^{\prime}}\nabla\widehat{u}_{\varepsilon^{\prime}}), and derive the desired estimate in the case r=1r=1. Finally, note that AA^{\sharp} and AA^{\flat} are scale-invariant. The estimate for general r>0r>0 follows immediately by rescaling. The proof is complete.

We now provide the proof of Theorem 3.3, following the idea of [NSX20].

Proof of Theorem 3.3.

We consider the matrix

𝒜ε(x,y):=A(x/ε1,x/ε2,,x/εn1,y).\mathcal{A}_{\varepsilon}(x,y):=A(x/\varepsilon_{1},x/\varepsilon_{2},\cdots,x/\varepsilon_{n-1},y). (3.23)

Recall that 𝒜ε(x,y)\mathcal{A}_{\varepsilon}(x,y) is strongly elliptic and locally 1-periodic in yy. In view of (1.4), 𝒜ε(x,y)\mathcal{A}_{\varepsilon}(x,y) is Hölder continuous in xx, and

|𝒜ε(x,y)𝒜ε(x,y)|Lk=1n1εkτ|xx|τ.\displaystyle|\mathcal{A}_{\varepsilon}(x^{\prime},y)-\mathcal{A}_{\varepsilon}(x,y)|\leq L\sum_{k=1}^{n-1}\varepsilon_{k}^{-\tau}|x-x^{\prime}|^{\tau}. (3.24)

In order to apply Theorem 2.1, we need to find an approximate matrix 𝒜εapp=𝒜εapp(x,y)\mathcal{A}^{\rm app}_{\varepsilon}=\mathcal{A}^{\rm app}_{\varepsilon}(x,y) which is Lipschitz in the xx variable. In fact, we define

𝒜εapp(x,y)=εnddφ(xzεn)𝒜ε(z,y)𝑑z,\mathcal{A}^{\rm app}_{\varepsilon}(x,y)=\varepsilon_{n}^{-d}\int_{\mathbb{R}^{d}}\varphi\Big{(}\frac{x-z}{\varepsilon_{n}}\Big{)}\mathcal{A}_{\varepsilon}(z,y)dz, (3.25)

with φ\varphi given as in Section 2.1. Then 𝒜εapp\mathcal{A}^{\rm app}_{\varepsilon} satisfies the ellipticity condition (1.2), is 1-periodic in yy, and

𝒜ε𝒜εappCLk=1n1εkτεnτ and x𝒜εappCLk=1n1εkτεnτ1,\|\mathcal{A}_{\varepsilon}-\mathcal{A}_{\varepsilon}^{\rm app}\|_{\infty}\leq CL\sum_{k=1}^{n-1}\varepsilon_{k}^{-\tau}\varepsilon_{n}^{\tau}\quad\text{ and }\quad\|\nabla_{x}\mathcal{A}_{\varepsilon}^{\rm app}\|_{\infty}\leq CL\sum_{k=1}^{n-1}\varepsilon_{k}^{-\tau}\varepsilon_{n}^{\tau-1}, (3.26)

where CC depends only on dd and τ\tau.

Let uεu_{\varepsilon} be the weak solution to (3.14). Then by (3.23), it satisfies

div(𝒜ε(x,x/εn)uε)=F in B2.-\text{\rm div}\big{(}\mathcal{A}_{\varepsilon}(x,x/\varepsilon_{n})\nabla u_{\varepsilon}\big{)}=F\quad\text{ in }B_{2}. (3.27)

Let uεappu_{\varepsilon}^{\rm app} be the weak solution to

div(𝒜εapp(x,x/εn)uεapp)=Fin B3/2,anduεapp=uεon B3/2.-\text{\rm div}\big{(}\mathcal{A}^{\rm app}_{\varepsilon}(x,x/\varepsilon_{n})\nabla u_{\varepsilon}^{\rm app}\big{)}=F\quad\text{in }B_{3/2},\quad\text{and}\quad u_{\varepsilon}^{\rm app}=u_{\varepsilon}\quad\text{on }\partial B_{3/2}. (3.28)

Combining (3.27), (3.28) and the first inequality in (3.26), we can apply the energy estimate to the equation of uεuεappu_{\varepsilon}-u_{\varepsilon}^{\rm app}

{div(𝒜εapp(x,x/εn)(uεappuε))=div((𝒜εapp(x,x/εn)𝒜ε(x,x/εn))uε)in B3/2,uεappuε=0on B3/2,\left\{\begin{aligned} &-\text{\rm div}\big{(}\mathcal{A}^{\rm app}_{\varepsilon}(x,x/\varepsilon_{n})\nabla(u_{\varepsilon}^{\rm app}-u_{\varepsilon})\big{)}=\text{\rm div}\big{(}(\mathcal{A}^{\rm app}_{\varepsilon}(x,x/\varepsilon_{n})-\mathcal{A}_{\varepsilon}(x,x/\varepsilon_{n}))\nabla u_{\varepsilon}\big{)}\quad\text{in }B_{3/2},\\ &u_{\varepsilon}^{\rm app}-u_{\varepsilon}=0\quad\text{on }\partial B_{3/2},\end{aligned}\right. (3.29)

to obtain

B3/2|(uεuεapp)|2\displaystyle\int_{B_{3/2}}|\nabla(u_{\varepsilon}-u_{\varepsilon}^{\rm app})|^{2} CL2k=1n1εk2τεn2τB3/2|uε|2.\displaystyle\leq CL^{2}\sum_{k=1}^{n-1}\varepsilon_{k}^{-2\tau}\varepsilon_{n}^{2\tau}\int_{B_{3/2}}|\nabla u_{\varepsilon}|^{2}. (3.30)

Next, we apply Theorem 2.1 to the equation (3.28). Let χεapp(x,y)\chi^{\rm app}_{\varepsilon}(x,y) be the corrector given by (2.2) with A(x,y)A(x,y) replaced by 𝒜εapp(x,y)\mathcal{A}^{\rm app}_{\varepsilon}(x,y), and 𝒜^εapp(x)\widehat{\mathcal{A}}^{\rm app}_{\varepsilon}(x) the corresponding effective coefficient matrix given by (2.1). Let u^εapp\widehat{u}_{\varepsilon}^{\rm app} be the solution to

div(𝒜^εapp(x)u^εapp)=Fin B5/4,andu^εapp=uεappon B5/4.\displaystyle-\text{\rm div}\big{(}\widehat{\mathcal{A}}^{\rm app}_{\varepsilon}(x)\nabla\widehat{u}_{\varepsilon}^{\rm app}\big{)}=F\quad\text{in }B_{5/4},\quad\text{and}\quad\widehat{u}_{\varepsilon}^{\rm app}=u_{\varepsilon}^{\rm app}\quad\text{on }\partial B_{5/4}. (3.31)

Thanks to Theorem 2.1,

B5/4|uεappu^εappSεn((yχεapp)εnuεapp)|2\displaystyle\int_{B_{5/4}}|\nabla u_{\varepsilon}^{\rm app}-\nabla\widehat{u}_{\varepsilon}^{\rm app}-S_{\varepsilon_{n}}((\nabla_{y}\chi^{\rm app}_{\varepsilon})^{\varepsilon_{n}}\nabla u_{\varepsilon}^{\rm app})|^{2} (3.32)
Cεn2(1+x𝒜εapp)2B5/4|u^εapp|2\displaystyle\leq C\varepsilon_{n}^{2}(1+\|\nabla_{x}\mathcal{A}_{\varepsilon}^{\rm app}\|_{\infty})^{2}\int_{B_{5/4}}|\nabla\widehat{u}_{\varepsilon}^{\rm app}|^{2}
+Cεn2B(5/4)3εn|2u^εapp|2+CB5/4B(5/4)4εn|u^εapp|2.\displaystyle\qquad+C\varepsilon_{n}^{2}\int_{B_{(5/4)-3\varepsilon_{n}}}|\nabla^{2}\widehat{u}_{\varepsilon}^{\rm app}|^{2}+C\int_{B_{5/4}\setminus B_{(5/4)-4\varepsilon_{n}}}|\nabla\widehat{u}_{\varepsilon}^{\rm app}|^{2}.

By the Hölder’s inequality, the last term on the right-hand side of (3.32) is bounded by

B5/4B(5/4)4εn|u^εapp|2Cεn12q(B5/4|u^εapp|q)2/q for q>2.\displaystyle\int_{B_{5/4}\setminus B_{(5/4)-4\varepsilon_{n}}}|\nabla\widehat{u}_{\varepsilon}^{\rm app}|^{2}\leq C\varepsilon_{n}^{1-\frac{2}{q}}\bigg{(}\int_{B_{5/4}}|\nabla\widehat{u}_{\varepsilon}^{\rm app}|^{q}\bigg{)}^{2/q}\quad\text{ for }q>2. (3.33)

On the other hand, by the interior H2H^{2} estimate for the equation (3.28), for any B(z,2ρ)B3/2B(z,2\rho)\subset B_{3/2},

B(z,ρ)|2u^εapp|2CB(z,2ρ)|F|2+C(x𝒜εapp2+ρ2)B(z,2ρ)|u^εapp|2.\int_{B(z,\rho)}|\nabla^{2}\widehat{u}_{\varepsilon}^{\rm app}|^{2}\leq C\int_{B(z,2\rho)}|F|^{2}+C\big{(}\|\nabla_{x}\mathcal{A}^{\rm app}_{\varepsilon}\|_{\infty}^{2}+\rho^{-2}\big{)}\int_{B(z,2\rho)}|\nabla\widehat{u}_{\varepsilon}^{\rm app}|^{2}.

Then it follows by a covering argument that

B(5/4)3εn|2u^εapp|2𝑑x\displaystyle\int_{B_{(5/4)-3\varepsilon_{n}}}|\nabla^{2}\widehat{u}_{\varepsilon}^{\rm app}|^{2}\,dx (3.34)
CB5/4|F|2+Cx𝒜εapp2B5/4|u^εapp|2+εn2q1(B5/4|u^εapp|q)2/q,\displaystyle\leq C\int_{B_{5/4}}|F|^{2}+C\|\nabla_{x}\mathcal{A}^{\rm app}_{\varepsilon}\|^{2}_{\infty}\int_{B_{5/4}}|\nabla\widehat{u}_{\varepsilon}^{\rm app}|^{2}+\varepsilon_{n}^{-\frac{2}{q}-1}\bigg{(}\int_{B_{5/4}}|\nabla\widehat{u}_{\varepsilon}^{\rm app}|^{q}\bigg{)}^{2/q},

where the Hölder’s inequality has been used for the last integral. By Meyers’ estimate for the elliptic equations (3.27), (3.28) and (3.31), we have for some q>2q>2,

(B3/2|uε|q)2/qCB2|F|2+CB2|uε|2,(B3/2|uεapp|q)2/qCB3/2|F|2+C(B3/2|uε|q)2/q,(B5/4|u^εapp|q)2/qCB5/4|F|2+C(B5/4|uεapp|q)2/q,\displaystyle\begin{split}&\bigg{(}\int_{B_{3/2}}|\nabla u_{\varepsilon}|^{q}\bigg{)}^{2/q}\leq C\int_{B_{2}}|F|^{2}+C\int_{B_{2}}|\nabla u_{\varepsilon}|^{2},\\ &\bigg{(}\int_{B_{3/2}}|\nabla u_{\varepsilon}^{\rm app}|^{q}\bigg{)}^{2/q}\leq C\int_{B_{3/2}}|F|^{2}+C\bigg{(}\int_{B_{3/2}}|\nabla u_{\varepsilon}|^{q}\bigg{)}^{2/q},\\ &\bigg{(}\int_{B_{5/4}}|\nabla\widehat{u}_{\varepsilon}^{\rm app}|^{q}\bigg{)}^{2/q}\leq C\int_{B_{5/4}}|F|^{2}+C\bigg{(}\int_{B_{5/4}}|\nabla u_{\varepsilon}^{\rm app}|^{q}\bigg{)}^{2/q},\end{split} (3.35)

which implies that

(B5/4|u^εapp|q)2/qCB2|F|2+CB2|uε|2.\displaystyle\bigg{(}\int_{B_{5/4}}|\nabla\widehat{u}_{\varepsilon}^{\rm app}|^{q}\bigg{)}^{2/q}\leq C\int_{B_{2}}|F|^{2}+C\int_{B_{2}}|\nabla u_{\varepsilon}|^{2}. (3.36)

By taking (3.33) and (3.34) into (3.32), and using (3.36) we obtain that

B5/4|uεappu^εappSεn((yχεapp)εnu^εapp)|2\displaystyle\int_{B_{5/4}}|\nabla u_{\varepsilon}^{\rm app}-\nabla\widehat{u}_{\varepsilon}^{\rm app}-S_{\varepsilon_{n}}((\nabla_{y}\chi_{\varepsilon}^{\rm app})^{\varepsilon_{n}}\nabla\widehat{u}_{\varepsilon}^{\rm app})|^{2} (3.37)
C{εn12q+L2k=1n1εk2τεn2τ}{B2|uε|2+B2|F|2}.\displaystyle\leq C\big{\{}\varepsilon_{n}^{1-\frac{2}{q}}+L^{2}\sum_{k=1}^{n-1}\varepsilon_{k}^{-2\tau}\varepsilon_{n}^{2\tau}\big{\}}\bigg{\{}\int_{B_{2}}|\nabla u_{\varepsilon}|^{2}+\int_{B_{2}}|F|^{2}\bigg{\}}.

Now, let χε(x,y)\chi_{\varepsilon}(x,y) be the corrector given by (2.2) with A(x,y)A(x,y) replaced by 𝒜ε(x,y)\mathcal{A}_{\varepsilon}(x,y), and 𝒜^ε(x)\widehat{\mathcal{A}}_{\varepsilon}(x) the effective matrix of 𝒜ε\mathcal{A}_{\varepsilon} given by (2.1) (recall that χε(x,y)\chi_{\varepsilon}(x,y) and 𝒜^ε(x)\widehat{\mathcal{A}}_{\varepsilon}(x) take the forms of (2.9) and (2.11)). Note that χεχεapp=(χε,jχε,japp)\chi_{\varepsilon}-\chi_{\varepsilon}^{\rm app}=(\chi_{\varepsilon,j}-\chi_{\varepsilon,j}^{\rm app}) satisfies

divy(𝒜ε(x,y)(yχε,jyχε,japp))\displaystyle-\text{\rm div}_{y}\big{(}\mathcal{A}_{\varepsilon}(x,y)(\nabla_{y}\chi_{\varepsilon,j}-\nabla_{y}\chi_{\varepsilon,j}^{\rm app})\big{)}
=divy((𝒜ε(x,y)𝒜εapp(x,y))yχε,japp)+divy((𝒜ε(x,y)𝒜εapp(x,y))ej) in 𝕋d.\displaystyle=\text{\rm div}_{y}\big{(}(\mathcal{A}_{\varepsilon}(x,y)-\mathcal{A}_{\varepsilon}^{\rm app}(x,y))\nabla_{y}\chi^{app}_{\varepsilon,j}\big{)}+\text{\rm div}_{y}\big{(}(\mathcal{A}_{\varepsilon}(x,y)-\mathcal{A}_{\varepsilon}^{\rm app}(x,y))e_{j}\big{)}\quad\text{ in }\mathbb{T}^{d}.

Thus, the standard energy estimate and the first inequality in (3.26) imply that

supxdyχε,j(x,)yχε,japp(x,)L2(𝕋d)CLk=1n1εkτεnτ,\displaystyle\sup_{x\in\mathbb{R}^{d}}\|\nabla_{y}\chi_{\varepsilon,j}(x,\cdot)-\nabla_{y}\chi^{\rm app}_{\varepsilon,j}(x,\cdot)\|_{L^{2}(\mathbb{T}^{d})}\leq CL\sum_{k=1}^{n-1}\varepsilon_{k}^{-\tau}\varepsilon_{n}^{\tau}, (3.38)

where CC depends only on d,Λd,\Lambda and τ\tau. This together with the definitions of 𝒜^εapp\widehat{\mathcal{A}}_{\varepsilon}^{\rm app} and 𝒜^ε\widehat{\mathcal{A}}_{\varepsilon} gives

𝒜^εapp𝒜^εCLk=1n1εkτεnτ.\|\widehat{\mathcal{A}}_{\varepsilon}^{\rm app}-\widehat{\mathcal{A}}_{\varepsilon}\|_{\infty}\leq CL\sum_{k=1}^{n-1}\varepsilon_{k}^{-\tau}\varepsilon_{n}^{\tau}. (3.39)

Let u^ε\widehat{u}_{\varepsilon} be the weak solution to

div(𝒜^ε(x)u^ε)=F in B5/4, and u^ε=u^εapp on B5/4.\displaystyle-\text{\rm div}\big{(}\widehat{\mathcal{A}}_{\varepsilon}(x)\nabla\widehat{u}_{\varepsilon}\big{)}=F\quad\text{ in }B_{5/4},\quad\text{ and }\quad\widehat{u}_{\varepsilon}=\widehat{u}^{\rm app}_{\varepsilon}\quad\text{ on }\partial B_{5/4}. (3.40)

Similar to (3.30), we apply the standard energy estimate to the equation for u^εappu^ε\widehat{u}_{\varepsilon}^{\rm app}-\widehat{u}_{\varepsilon},

{div(𝒜^ε(x,x/εn)(u^εu^εapp))=div((𝒜^ε(x,x/εn)𝒜^εapp(x,x/εn))u^εapp)in B5/4,u^εu^εapp=0on B5/4,\left\{\begin{aligned} &-\text{\rm div}\big{(}\widehat{\mathcal{A}}_{\varepsilon}(x,x/\varepsilon_{n})\nabla(\widehat{u}_{\varepsilon}-\widehat{u}_{\varepsilon}^{\rm app})\big{)}=\text{\rm div}\big{(}(\widehat{\mathcal{A}}_{\varepsilon}(x,x/\varepsilon_{n})-\widehat{\mathcal{A}}^{\rm app}_{\varepsilon}(x,x/\varepsilon_{n}))\nabla\widehat{u}_{\varepsilon}^{\rm app}\big{)}\quad\text{in }B_{5/4},\\ &\widehat{u}_{\varepsilon}-\widehat{u}_{\varepsilon}^{\rm app}=0\quad\text{on }\partial B_{5/4},\end{aligned}\right. (3.41)

and use (3.39) to obtain

B5/4|(u^εu^εapp)|2CL2k=1n1εk2τεn2τB5/4|u^εapp|2CL2k=1n1εk2τεn2τ{B2|uε|2+B2|F|2},\displaystyle\begin{split}\int_{B_{5/4}}|\nabla(\widehat{u}_{\varepsilon}-\widehat{u}_{\varepsilon}^{\rm app})|^{2}&\leq CL^{2}\sum_{k=1}^{n-1}\varepsilon_{k}^{-2\tau}\varepsilon_{n}^{2\tau}\int_{B_{5/4}}|\nabla\widehat{u}_{\varepsilon}^{\rm app}|^{2}\\ &\leq CL^{2}\sum_{k=1}^{n-1}\varepsilon_{k}^{-2\tau}\varepsilon_{n}^{2\tau}\bigg{\{}\int_{B_{2}}|\nabla u_{\varepsilon}|^{2}+\int_{B_{2}}|F|^{2}\bigg{\}},\end{split} (3.42)

where the energy estimate for u^εapp\widehat{u}_{\varepsilon}^{\rm app} (or (3.36)) is also used in the last inequality.

Finally, combining (3.30), (3.38) and (3.42), we may replace the approximate solutions uεapp,u^εapp,χεappu_{\varepsilon}^{\rm app},\widehat{u}_{\varepsilon}^{\rm app},\chi_{\varepsilon}^{\rm app} on the left-hand side of (3.37) by uε,u^ε,χεu_{\varepsilon},\widehat{u}_{\varepsilon},\chi_{\varepsilon}, respectively. In particular, by the property of the smoothing operator SεnS_{\varepsilon_{n}} (see e.g., [She18, NSX20]), we have

B1|Sεn((yχε)εnu^ε)Sεn((yχεapp)εnu^εapp)|22B1{|Sεn((yχε)εn(u^εu^εapp))|2+|Sεn(((yχεapp)εn(yχε)εn)u^εapp)|2}CB1+εn|u^εu^εapp|2+CL2k=1n1εk2τεn2τB1+εn|u^εapp|2CL2k=1n1εk2τεn2τ{B2|uε|2+B2|F|2}.\displaystyle\begin{split}&\int_{B_{1}}|S_{\varepsilon_{n}}\big{(}(\nabla_{y}\chi_{\varepsilon})^{\varepsilon_{n}}\nabla\widehat{u}_{\varepsilon}\big{)}-S_{\varepsilon_{n}}\big{(}(\nabla_{y}\chi_{\varepsilon}^{\rm app})^{\varepsilon_{n}}\nabla\widehat{u}_{\varepsilon}^{\rm app}\big{)}|^{2}\\ &\leq 2\int_{B_{1}}\big{\{}|S_{\varepsilon_{n}}\big{(}(\nabla_{y}\chi_{\varepsilon})^{\varepsilon_{n}}(\nabla\widehat{u}_{\varepsilon}-\nabla\widehat{u}_{\varepsilon}^{\rm app})\big{)}|^{2}+|S_{\varepsilon_{n}}\big{(}((\nabla_{y}\chi_{\varepsilon}^{\rm app})^{\varepsilon_{n}}-(\nabla_{y}\chi_{\varepsilon})^{\varepsilon_{n}})\nabla\widehat{u}_{\varepsilon}^{\rm app}\big{)}|^{2}\big{\}}\\ &\leq C\int_{B_{1+\varepsilon_{n}}}|\nabla\widehat{u}_{\varepsilon}-\nabla\widehat{u}_{\varepsilon}^{\rm app}|^{2}+CL^{2}\sum_{k=1}^{n-1}\varepsilon_{k}^{-2\tau}\varepsilon_{n}^{2\tau}\int_{B_{1+\varepsilon_{n}}}|\nabla\widehat{u}_{\varepsilon}^{\rm app}|^{2}\\ &\leq CL^{2}\sum_{k=1}^{n-1}\varepsilon_{k}^{-2\tau}\varepsilon_{n}^{2\tau}\bigg{\{}\int_{B_{2}}|\nabla u_{\varepsilon}|^{2}+\int_{B_{2}}|F|^{2}\bigg{\}}.\end{split} (3.43)

It follows that

B5/4|uεu^εSεn((yχ)εu^ε)|2\displaystyle\int_{B_{5/4}}|\nabla u_{\varepsilon}-\nabla\widehat{u}_{\varepsilon}-S_{\varepsilon_{n}}((\nabla_{y}\chi)^{\varepsilon}\nabla\widehat{u}_{\varepsilon})|^{2} (3.44)
C{εn12q+L2k=1n1εk2τεn2τ}{B2|uε|2+B2|F|2}.\displaystyle\leq C\big{\{}\varepsilon_{n}^{1-\frac{2}{q}}+L^{2}\sum_{k=1}^{n-1}\varepsilon_{k}^{-2\tau}\varepsilon_{n}^{2\tau}\big{\}}\bigg{\{}\int_{B_{2}}|\nabla u_{\varepsilon}|^{2}+\int_{B_{2}}|F|^{2}\bigg{\}}.

Recall that εn\varepsilon_{n} is QQ-separated from εk,k=1,,n1\varepsilon_{k},k=1,\cdots,n-1, i.e., εk/εnQ\varepsilon_{k}/\varepsilon_{n}\geq Q for k=1,,n1.k=1,\cdots,n-1. In view of (2.9) and (2.11), the desired estimate (3.16) follows immediately from (3.44), while the estimate (3.15) follows directly from (3.42). The proof is complete. ∎

By similar arguments as above, it is not difficult to prove the boundary version of the uniform approximation theorem.

Theorem 3.4.

Let Ω\Omega be a bounded C1C^{1} domain. Suppose that AA satisfies the assumptions (1.2), (1.3) and (3.9). Let x0Ω,Br=B(x0,r)x_{0}\in\partial\Omega,B_{r}=B(x_{0},r) with 0<r<r0=r0(Ω)0<r<r_{0}=r_{0}(\Omega) and uεu_{\varepsilon} be a weak solution to

div(A(x/ε1,x/ε2,,x/εn)uε)=F in B2rΩ,uε=0 on B2rΩ.\displaystyle-\text{\rm div}\big{(}A(x/\varepsilon_{1},x/\varepsilon_{2},\cdots,x/\varepsilon_{n})\nabla u_{\varepsilon}\big{)}=F\,\text{ in }B_{2r}\cap\Omega,\quad u_{\varepsilon}=0\,\text{ on }B_{2r}\cap\partial\Omega. (3.45)

Then for any Q>1Q>1, there exist a coefficient matrix A=A(y1,,yn1)A^{\flat}=A^{\flat}(y_{1},\cdots,y_{n-1}), scales εk(0,)\varepsilon_{k}^{\prime}\in(0,\infty) with k=1,,n1k=1,\cdots,n-1, and a weak solution uεu^{\flat}_{\varepsilon^{\prime}} to

div(A(x/ε1,x/ε2,,x/εn1)uε)=Fin Br,uε=0 on BrΩ,-\text{\rm div}\big{(}A^{\flat}(x/\varepsilon^{\prime}_{1},x/\varepsilon^{\prime}_{2},\cdots,x/\varepsilon^{\prime}_{n-1})\nabla u^{\flat}_{\varepsilon^{\prime}}\big{)}=F\quad\text{in }B_{r},\quad u^{\flat}_{\varepsilon^{\prime}}=0\text{ on }B_{r}\cap\partial\Omega,

such that

uεL2(BrΩ)C{uεL2(B2rΩ)+rFL2(B2rΩ)},\|\nabla u^{\flat}_{\varepsilon^{\prime}}\|_{L^{2}(B_{r}\cap\Omega)}\leq C\big{\{}\|\nabla u_{\varepsilon}\|_{L^{2}(B_{2r}\cap\Omega)}+r\|F\|_{L^{2}(B_{2r}\cap\Omega)}\big{\}}, (3.46)

and

uεuεUεL2(BrΩ)C{(Qn1εnr)σ+Qτ}{uεL2(B2rΩ)+rFL2(B2rΩ)}\|\nabla u_{\varepsilon}-\nabla u^{\flat}_{\varepsilon^{\prime}}-U_{\varepsilon^{\prime}}\|_{L^{2}(B_{r}\cap\Omega)}\leq C\bigg{\{}\bigg{(}\frac{Q^{n-1}\varepsilon_{n}}{r}\bigg{)}^{\sigma}+Q^{-\tau}\bigg{\}}\big{\{}\|\nabla u_{\varepsilon}\|_{L^{2}(B_{2r}\cap\Omega)}+r\|F\|_{L^{2}(B_{2r}\cap\Omega)}\big{\}} (3.47)

for some σ>0,\sigma>0, where UεU_{\varepsilon^{\prime}} is given by

Uε(x)=(εn)ddφ(xzεn)(y𝒳)(zε1,,zεn1,xεn)uε(z)𝑑z,U_{\varepsilon^{\prime}}(x)=(\varepsilon_{n}^{\prime})^{-d}\int_{\mathbb{R}^{d}}\varphi\bigg{(}\frac{x-z}{\varepsilon_{n}^{\prime}}\bigg{)}(\nabla_{y}\mathcal{X}^{\sharp})\bigg{(}\frac{z}{\varepsilon_{1}^{\prime}},\cdots,\frac{z}{\varepsilon^{\prime}_{n-1}},\frac{x}{\varepsilon_{n}^{\prime}}\bigg{)}\nabla u^{\flat}_{\varepsilon^{\prime}}(z)dz, (3.48)

and 𝒳\mathcal{X}^{\sharp} is the corrector given by (2.2) with AA replaced by AA^{\sharp} (we have extended uε(z)u^{\flat}_{\varepsilon^{\prime}}(z) by zero across the boundary). Moreover, the constant CC depends only on d,n,Λd,n,\Lambda and (τ,L)(\tau,L) in (3.9), and the assumptions (1.2), (1.3) and (1.4) are preserved for AA^{\flat}.

4. The uniform Calderón-Zygmund estimates

In this section we prove the uniform Calderón-Zygmund estimate for the equation (1.1). Let us first consider the interior estimate.

Theorem 4.1.

Assume AA satisfies the assumptions (1.2), (1.3) and (1.4). Let uεu_{\varepsilon} be a weak solution to

div(A(x/ε1,x/ε2,,x/εn)uε)=divfin 5B0,\displaystyle-\text{\rm div}\big{(}A(x/\varepsilon_{1},x/\varepsilon_{2},\cdots,x/\varepsilon_{n})\nabla u_{\varepsilon}\big{)}=\text{\rm div}f\quad\text{in }5B_{0}, (4.1)

with fLp(5B0)df\in L^{p}(5B_{0})^{d} for some 2<p<2<p<\infty, where B0=B(x0,r0)B_{0}=B(x_{0},r_{0}) for some x0d,r0>0x_{0}\in\mathbb{R}^{d},r_{0}>0. Then uεLp(B0)d\nabla u_{\varepsilon}\in L^{p}(B_{0})^{d} and

(B0|uε|p)1/pC{(5B0|uε|2)1/2+(5B0|f|p)1/p},\displaystyle\bigg{(}\fint_{B_{0}}|\nabla u_{\varepsilon}|^{p}\bigg{)}^{1/p}\leq C\bigg{\{}\bigg{(}\fint_{5B_{0}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}+\bigg{(}\fint_{5B_{0}}|f|^{p}\bigg{)}^{1/p}\bigg{\}}, (4.2)

where CC depends only on d,Λ,p,n,d,\Lambda,p,n, and (τ,L)(\tau,L) in (1.4).

The proof of Theorem 4.1 relies on a real-variable argument involving a double-averaging estimate. We define the averaging operator as

Mt[f](x):=(Bt(x)|f|2)1/2.M_{t}[f](x):=\bigg{(}\fint_{B_{t}(x)}|f|^{2}\bigg{)}^{1/2}. (4.3)

We need the averaging operator in the real-variable argument for two reasons. The first reason is that the approximation in (3.12) is meaningful only if rQn1εnr\gg Q^{n-1}\varepsilon_{n} for sufficiently large Q1Q\gg 1. This forces us to take the large-scale averaging estimate above the scale t=CQn1εnt=CQ^{n-1}\varepsilon_{n} for sufficiently large CC and ignore the possible irregularity below that scale. The second reason to introduce another averaging is due to the special structure of the corrector 𝒳\mathcal{X}^{\sharp} in UεU_{\varepsilon^{\prime}} given by (3.13). Recall that 𝒳=𝒳(y1,y2,,yn)\mathcal{X}^{\sharp}=\mathcal{X}^{\sharp}(y_{1},y_{2},\cdots,y_{n}) is 1-periodic in each yiy_{i}. But its gradient in yny_{n} does not have an LpL^{p} estimate uniformly in QQ before we prove it in case of nn oscillating scales (otherwise we will run into a circular reasoning), which causes a big problem in the real-variable argument. To resolve this issue, we take the averaging for another time at scale t=εn=qεnt=\varepsilon^{\prime}_{n}=q\varepsilon_{n}, a scale comparable to the size of the periodic cell (depending on QQ as 1<qQn11<q\leq Q^{n-1}). This allows us to use only the periodic structure of 𝒳\mathcal{X}^{\sharp} and the L2L^{2} energy estimate (independent of QQ). Overall, the principle is that we have to make sure all the constants, except for η\eta, in the real-variable argument (Theorem 4.2 below) are independent of QQ.

Theorem 4.2.

[She23, Theorem 2.1 and Remark 2.4] Let p0>2p_{0}>2, L2(4B0)\mathcal{F}\in L^{2}(4B_{0}) and GLp(4B0)G\in L^{p}(4B_{0}) for some 2<p<p02<p<p_{0}, where B0B_{0} is a ball in d\mathbb{R}^{d}. Let 0t<c0diam(B0)0\leq t<c_{0}\text{diam}(B_{0}) be a fixed number. Suppose that for each ball B2B0B\subset 2B_{0} with the property that t<diam(B)c0diam(B0)t<\text{diam}(B)\leq c_{0}\text{diam}(B_{0}), there exists two measurable functions B\mathcal{F}_{B} and B\mathcal{R}_{B} on 2B2B such that |||B|+|B||\mathcal{F}|\leq|\mathcal{F}_{B}|+|\mathcal{R}_{B}| on 2B2B, and

(2B|B|p0)1/p0C1{(4B||2)1/2+supBB4B0(B|G|2)1/2},\displaystyle\bigg{(}\fint_{2B}|\mathcal{R}_{B}|^{p_{0}}\bigg{)}^{1/p_{0}}\leq C_{1}\bigg{\{}\bigg{(}\fint_{4B}|\mathcal{F}|^{2}\bigg{)}^{1/2}+\sup_{B\subset B^{\prime}\subset 4B_{0}}\bigg{(}\fint_{B^{\prime}}|G|^{2}\bigg{)}^{1/2}\bigg{\}}, (4.4)
(2B|B|2)1/2C2supBB4B0(B|G|2)1/2+η(4B||2)1/2,\displaystyle\bigg{(}\fint_{2B}|\mathcal{F}_{B}|^{2}\bigg{)}^{1/2}\leq C_{2}\sup_{B\subset B^{\prime}\subset 4B_{0}}\bigg{(}\fint_{B^{\prime}}|G|^{2}\bigg{)}^{1/2}+\eta\bigg{(}\fint_{4B}|\mathcal{F}|^{2}\bigg{)}^{1/2}, (4.5)

where C1,C2>0,0<c0<1C_{1},C_{2}>0,0<c_{0}<1, and η0\eta\geq 0. Then there exists η0>0\eta_{0}>0, depending only on C1,C2,c0,p,p0,C_{1},C_{2},c_{0},p,p_{0}, with the property that if 0η<η00\leq\eta<\eta_{0}, then

(B0Mt[]p)1/pC{(4B0||2)1/2+(4B0|G|p)1/p},\displaystyle\bigg{(}\fint_{B_{0}}M_{t}[\mathcal{F}]^{p}\bigg{)}^{1/p}\leq C\bigg{\{}\bigg{(}\fint_{4B_{0}}|\mathcal{F}|^{2}\bigg{)}^{1/2}+\bigg{(}\fint_{4B_{0}}|G|^{p}\bigg{)}^{1/p}\bigg{\}}, (4.6)

where CC depends only on d,C1,C2,c0,pd,C_{1},C_{2},c_{0},p and p0p_{0}. If t=0t=0, then (4.6) is replaced by

(B0||p)1/pC{(4B0||2)1/2+(4B0|G|p)1/p}.\displaystyle\bigg{(}\fint_{B_{0}}|\mathcal{F}|^{p}\bigg{)}^{1/p}\leq C\bigg{\{}\bigg{(}\fint_{4B_{0}}|\mathcal{F}|^{2}\bigg{)}^{1/2}+\bigg{(}\fint_{4B_{0}}|G|^{p}\bigg{)}^{1/p}\bigg{\}}. (4.7)
Proof of Theorem 4.1.

Let us first consider the case f=0f=0. We use an inductive argument on the number of scales to prove the theorem. Note that the result is well-known for n=1n=1 (see e.g., [She18]). Assume the theorem is true if the number of scales is strictly less than nn, which means that for any >ε1ε2εn1>0\infty>\varepsilon_{1}\geq\varepsilon_{2}\geq\cdots\geq\varepsilon_{n-1}>0, the solution vεv_{\varepsilon} to

div(A(x/ε1,,x/εn1)vε)=0in B2r-\text{\rm div}\big{(}A(x/\varepsilon_{1},\cdots,x/\varepsilon_{n-1})\nabla v_{\varepsilon}\big{)}=0\quad\text{in }B_{2r}

satisfies the uniform interior W1,pW^{1,p} estimate

(B3r/2|vε|p)1/pCp(B2r|vε|2)1/2\displaystyle\bigg{(}\fint_{B_{3r/2}}|\nabla v_{\varepsilon}|^{p}\bigg{)}^{1/p}\leq C_{p}\bigg{(}\fint_{B_{2r}}|\nabla v_{\varepsilon}|^{2}\bigg{)}^{1/2} (4.8)

for any 0<r<0<r<\infty and any 2<p<2<p<\infty.

We next prove the theorem for nn scales by using Theorem 4.2 with :=Mεn[uε]\mathcal{F}:=M_{\varepsilon_{n}^{\prime}}[\nabla u_{\varepsilon}] with εn=qεn\varepsilon_{n}^{\prime}=q\varepsilon_{n} given in (3.18). Fix any p(2,)p\in(2,\infty) and p1(2,p)p_{1}\in(2,p).

Let Q>1Q>1 to be determined later, and εn=qεnQn1εn<r<c0r0\varepsilon_{n}^{\prime}=q\varepsilon_{n}\leq Q^{n-1}\varepsilon_{n}<r<c_{0}r_{0} for some fixed c0(0,1)c_{0}\in(0,1). For each ball Br=Br(x)B_{r}=B_{r}(x) such that Br2B0=B(x0,2r0)B_{r}\subset 2B_{0}=B(x_{0},2r_{0}), we need to construct B\mathcal{F}_{B} and B\mathcal{R}_{B}, which satisfy (4.4) and (4.5), respectively. Since uεu_{\varepsilon} is a solution to div(A(x/ε1,,x/εn)uε)=0-\text{\rm div}\big{(}A(x/\varepsilon_{1},\cdots,x/\varepsilon_{n})\nabla u_{\varepsilon}\big{)}=0 in B4r4B0B_{4r}\subset 4B_{0}, by Theorem 3.2, there exist a 1-periodic matrix A=A(y1,y2,,yn1)A^{\flat}=A^{\flat}(y_{1},y_{2},\cdots,y_{n-1}), scales ε=(ε1,,εn1)\varepsilon^{\prime}=(\varepsilon_{1}^{\prime},\cdots,\varepsilon_{n-1}^{\prime}), and a solution uε(x)u_{\varepsilon^{\prime}}^{\flat}(x) to

div(A(x/ε1,,x/εn1)uε)=0in B2r,\displaystyle-\text{\rm div}\big{(}A^{\flat}(x/\varepsilon^{\prime}_{1},\cdots,x/\varepsilon^{\prime}_{n-1})\nabla u_{\varepsilon^{\prime}}^{\flat}\big{)}=0\quad\text{in }B_{2r}, (4.9)

such that

uεL2(B2r)CuεL2(B4r),\displaystyle\|\nabla u_{\varepsilon^{\prime}}^{\flat}\|_{L^{2}(B_{2r})}\leq C\|\nabla u_{\varepsilon}\|_{L^{2}(B_{4r})}, (4.10)

and

uεuεUεL2(B2r)C{Qτ+(Qn1εnr)σ}uεL2(B4r),\displaystyle\begin{split}&\|\nabla u_{\varepsilon}-\nabla u_{\varepsilon^{\prime}}^{\flat}-U_{\varepsilon^{\prime}}\|_{L^{2}(B_{2r})}\leq C\bigg{\{}Q^{-\tau}+\bigg{(}\frac{Q^{n-1}\varepsilon_{n}}{r}\bigg{)}^{\sigma}\bigg{\}}\|\nabla u_{\varepsilon}\|_{L^{2}(B_{4r})},\end{split} (4.11)

where UεU_{\varepsilon^{\prime}} is given by (3.13). Let

B(x)=Mεn[uεuεUε](x) and B(x)=Mεn[uεUε](x).\displaystyle\mathcal{F}_{B}(x)=M_{\varepsilon_{n}^{\prime}}[\nabla u_{\varepsilon}-\nabla u_{\varepsilon^{\prime}}^{\flat}-U_{\varepsilon^{\prime}}](x)\,\,\text{ and }\,\,\mathcal{R}_{B}(x)=M_{\varepsilon_{n}^{\prime}}[\nabla u_{\varepsilon^{\prime}}^{\flat}-U_{\varepsilon^{\prime}}](x).

By the triangle inequality, we have |||B|+|B|.|\mathcal{F}|\leq|\mathcal{F}_{B}|+|\mathcal{R}_{B}|.

We recall a basic property of the averaging operator MtM_{t}. If FLloc2(d)F\in L^{2}_{\rm loc}(\mathbb{R}^{d}), then for any rt>0r\geq t>0, we have

BrMt[F]2CBr+t|F|2CB2r|F|2CB2rMt[F]2.\fint_{B_{r}}M_{t}[F]^{2}\leq C\fint_{B_{r+t}}|F|^{2}\leq C\fint_{B_{2r}}|F|^{2}\leq C\fint_{B_{2r}}M_{t}[F]^{2}. (4.12)

This can be shown easily by the Fubini’s theorem.

Consequently, by the assumption r>εnr>\varepsilon_{n}^{\prime}, we have

Br|B|2CB2r|uεuεUε|2,\fint_{B_{r}}|\mathcal{F}_{B}|^{2}\leq C\fint_{B_{2r}}|\nabla u_{\varepsilon}-\nabla u_{\varepsilon^{\prime}}^{\flat}-U_{\varepsilon^{\prime}}|^{2}, (4.13)

and

B4r|uε|2CB4r||2.\fint_{B_{4r}}|\nabla u_{\varepsilon}|^{2}\leq C\fint_{B_{4r}}|\mathcal{F}|^{2}. (4.14)

Combining the last two inequalities with (4.11), we have

Br|B|2C{Qτ+(Qn1εnr)σ}2B4r||2.\displaystyle\fint_{B_{r}}|\mathcal{F}_{B}|^{2}\leq C\bigg{\{}Q^{-\tau}+\bigg{(}\frac{Q^{n-1}\varepsilon_{n}}{r}\bigg{)}^{\sigma}\bigg{\}}^{2}\fint_{B_{4r}}|\mathcal{F}|^{2}. (4.15)

We now choose Q=Q0>1Q=Q_{0}>1 large enough such that CQ0τ<η0/2CQ_{0}^{-\tau}<\eta_{0}/2, and then take r>k0Q0n1εnr>k_{0}Q_{0}^{n-1}\varepsilon_{n} for some k0k_{0} large enough such that CQ0(n1)σ(εn/r)σ<η0/2,CQ_{0}^{(n-1)\sigma}(\varepsilon_{n}/r)^{\sigma}<\eta_{0}/2, where η0\eta_{0} is given by Theorem 4.2 depending only on C1,C2,c0,p,p1.C_{1},C_{2},c_{0},p,p_{1}. It follows that B\mathcal{F}_{B} satisfies condition (4.5) for r>k0Q0n1εnr>k_{0}Q_{0}^{n-1}\varepsilon_{n} with C2=0C_{2}=0. In the following, we will verify that B\mathcal{R}_{B} satisfies (4.4) and the constant C1C_{1} is independent of QQ.

Note that

Br|B|p=BrMεn[uεUε]p2p1Br|Mεn[uε](z)|p𝑑z+2p1Br|Mεn[Uε](z)|p𝑑z.\displaystyle\begin{split}\fint_{B_{r}}|\mathcal{R}_{B}|^{p}&=\fint_{B_{r}}M_{\varepsilon_{n}^{\prime}}[\nabla u_{\varepsilon^{\prime}}^{\flat}-U_{\varepsilon^{\prime}}]^{p}\\ &\leq 2^{p-1}\fint_{B_{r}}|M_{\varepsilon_{n}^{\prime}}[\nabla u_{\varepsilon^{\prime}}^{\flat}](z)|^{p}dz+2^{p-1}\fint_{B_{r}}|M_{\varepsilon_{n}^{\prime}}[U_{\varepsilon^{\prime}}](z)|^{p}dz.\end{split} (4.16)

Since uεu_{\varepsilon^{\prime}}^{\flat} satisfies (4.9), and the coefficient matrix AA^{\flat} involves at most n1n-1 scales and is periodic and Hölder continuous yi=x/εiy_{i}=x/\varepsilon_{i}^{\prime} for i=1,,n1i=1,\cdots,n-1. By the inductive assumption, we know that uεu_{\varepsilon^{\prime}}^{\flat} satisfies (4.8), which combined with (4.10) implies

(B3r/2|uε|p)1/pC(B4r|uε|2)1/2,\displaystyle\bigg{(}\fint_{B_{3r/2}}|\nabla u_{\varepsilon^{\prime}}^{\flat}|^{p}\bigg{)}^{1/p}\leq C\bigg{(}\fint_{B_{4r}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2},

for any k0εnk0Q0n1εnrc0r0k_{0}\varepsilon_{n}^{\prime}\leq k_{0}Q_{0}^{n-1}\varepsilon_{n}\leq r\leq c_{0}r_{0}. By the Hölder’s inequality and Fubini’s theorem,

Br|Mεn[uε](z)|p𝑑zBrBεn(z)|uε(y)|p𝑑y𝑑zCB3r/2|uε(y)|p𝑑yC(B4r|uε|2)p/2C(B4r||2)p/2,\displaystyle\begin{split}\fint_{B_{r}}|M_{\varepsilon_{n}^{\prime}}[\nabla u_{\varepsilon^{\prime}}^{\flat}](z)|^{p}dz&\leq\fint_{B_{r}}\fint_{B_{\varepsilon_{n}^{\prime}}(z)}|\nabla u_{\varepsilon^{\prime}}^{\flat}(y)|^{p}dydz\\ &\leq C\fint_{B_{3r/2}}|\nabla u_{\varepsilon^{\prime}}^{\flat}(y)|^{p}dy\\ &\leq C\bigg{(}\fint_{B_{4r}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{p/2}\\ &\leq C\bigg{(}\fint_{B_{4r}}|\mathcal{F}|^{2}\bigg{)}^{p/2},\end{split} (4.17)

where we have used (4.12) in the last step.

To bound the second term on the right hand side of (4.16), we have to take advantage of the periodic structure of 𝒳\mathcal{X}^{\sharp} and it is right here that the averaging operator MεnM_{\varepsilon_{n}^{\prime}} plays a key role. We observe that, for each yBry\in B_{r},

|Mεn[Uε](y)|2=Bεn(y)|(εn)ddφ(xzεn)(y𝒳)(zε1,,zεn1,xεn)uε(z)𝑑z|2𝑑xBεn(y)(εn)ddφ(xzεn)|(y𝒳)(zε1,,zεn1,xεn)uε(z)|2𝑑z𝑑xsupzdB1|(y𝒳)(zε1,,zεn1,y)|2𝑑yB2εn(y)|uε(z)|2𝑑zCB2εn(y)|uε(z)|2𝑑z=C|M2εn[uε](y)|2,\displaystyle\begin{split}|M_{\varepsilon_{n}^{\prime}}[U_{\varepsilon^{\prime}}](y)|^{2}&=\fint_{B_{\varepsilon_{n}^{\prime}}(y)}\bigg{|}(\varepsilon_{n}^{\prime})^{-d}\int_{\mathbb{R}^{d}}\varphi\bigg{(}\frac{x-z}{\varepsilon_{n}^{\prime}}\bigg{)}(\nabla_{y}\mathcal{X}^{\sharp})\bigg{(}\frac{z}{\varepsilon_{1}^{\prime}},\cdots,\frac{z}{\varepsilon^{\prime}_{n-1}},\frac{x}{\varepsilon_{n}^{\prime}}\bigg{)}\nabla u_{\varepsilon^{\prime}}^{\flat}(z)dz\bigg{|}^{2}dx\\ &\leq\fint_{B_{\varepsilon_{n}^{\prime}}(y)}(\varepsilon_{n}^{\prime})^{-d}\int_{\mathbb{R}^{d}}\varphi\bigg{(}\frac{x-z}{\varepsilon_{n}^{\prime}}\bigg{)}\bigg{|}(\nabla_{y}\mathcal{X}^{\sharp})\bigg{(}\frac{z}{\varepsilon_{1}^{\prime}},\cdots,\frac{z}{\varepsilon^{\prime}_{n-1}},\frac{x}{\varepsilon_{n}^{\prime}}\bigg{)}\nabla u_{\varepsilon^{\prime}}^{\flat}(z)\bigg{|}^{2}dzdx\\ &\leq\sup_{z\in\mathbb{R}^{d}}\fint_{B_{1}}\bigg{|}(\nabla_{y}\mathcal{X}^{\sharp})\bigg{(}\frac{z}{\varepsilon_{1}^{\prime}},\cdots,\frac{z}{\varepsilon^{\prime}_{n-1}},y\bigg{)}\bigg{|}^{2}dy\fint_{B_{2\varepsilon_{n}^{\prime}}(y)}|\nabla u_{\varepsilon^{\prime}}^{\flat}(z)|^{2}dz\\ &\leq C\fint_{B_{2\varepsilon_{n}^{\prime}}(y)}|\nabla u_{\varepsilon^{\prime}}^{\flat}(z)|^{2}dz=C|M_{2\varepsilon_{n}^{\prime}}[\nabla u_{\varepsilon^{\prime}}^{\flat}](y)|^{2},\end{split}

where we have used the L2L^{2} energy estimate for 𝒳\mathcal{X}^{\sharp} (see the equation (2.10)) and CC depends only on dd and Λ\Lambda. This combined with (4.17) implies that

Br|Mεn[Uε](z)|pCBr|M2εn[uε](y)|p𝑑yCB4r/3|Mεn[uε](y)|p𝑑yC(B4r||2)p/2,\displaystyle\begin{aligned} \fint_{B_{r}}|M_{\varepsilon_{n}^{\prime}}[U_{\varepsilon^{\prime}}](z)|^{p}&\leq C\fint_{B_{r}}|M_{2\varepsilon_{n}^{\prime}}[\nabla u_{\varepsilon^{\prime}}^{\flat}](y)|^{p}dy\\ &\leq C\fint_{B_{4r/3}}|M_{\varepsilon_{n}^{\prime}}[\nabla u_{\varepsilon^{\prime}}^{\flat}](y)|^{p}dy\leq C\bigg{(}\fint_{B_{4r}}|\mathcal{F}|^{2}\bigg{)}^{p/2},\end{aligned} (4.18)

for any k0Q0n1εnr<c0r0k_{0}Q_{0}^{n-1}\varepsilon_{n}\leq r<c_{0}r_{0}. We point out that we cannot simply take the LpL^{p} norm of UεU_{\varepsilon^{\prime}} directly since 𝒳Lp(B1,dyn)\|\nabla\mathcal{X}^{\sharp}\|_{L^{p}(B_{1},dy_{n})} depends on qq and hence on QQ, due to the loss of regularity of A(y1,y2,,yn)A^{\sharp}(y_{1},y_{2},\cdots,y_{n}) in yny_{n} caused by reperiodization.

By taking (4.17) and (4.18) into (4.16), we find that B\mathcal{R}_{B} satisfies (4.4) for k0Q0n1εn<r<c0r0k_{0}Q_{0}^{n-1}\varepsilon_{n}<r<c_{0}r_{0} (without the GG part) with C1C_{1} independent of QQ. Thus the conditions of Theorem 4.2 are all satisfied. Thanks to (4.6), we obtain that

(B0Mεn′′[Mεn[uε]]p1)1/p1C(4B0|Mεn[uε](x)|2)1/2,\displaystyle\bigg{(}\fint_{B_{0}}M_{\varepsilon_{n}^{\prime\prime}}[M_{\varepsilon_{n}^{\prime}}[\nabla u_{\varepsilon}]]^{p_{1}}\bigg{)}^{1/{p_{1}}}\leq C\bigg{(}\fint_{4B_{0}}|M_{\varepsilon_{n}^{\prime}}[\nabla u_{\varepsilon}](x)|^{2}\bigg{)}^{1/2}, (4.19)

where εn=qεn\varepsilon_{n}^{\prime}=q\varepsilon_{n} and εn′′=k0Q0n1εn\varepsilon_{n}^{\prime\prime}=k_{0}Q_{0}^{n-1}\varepsilon_{n} and p1(2,p)p_{1}\in(2,p). This is a double-averaging estimate since we have taken average twice in the LpL^{p} norm on the left-hand side. At this stage (after using Theorem 4.2), it is safe to remove one averaging in (4.19) by (4.12). In fact, since εn′′>εn\varepsilon_{n}^{\prime\prime}>\varepsilon_{n}^{\prime}, (4.19) implies

(B0M2εn′′[uε]p1)1/p1C(5B0|uε|2)1/2.\bigg{(}\fint_{B_{0}}M_{2\varepsilon_{n}^{\prime\prime}}[\nabla u_{\varepsilon}]^{p_{1}}\bigg{)}^{1/{p_{1}}}\leq C\bigg{(}\fint_{5B_{0}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}. (4.20)

Next, we claim that for each yB0y\in B_{0}, we have

|uε(y)|CM2εn′′[uε](y)=C(B2εn′′(y)|uε|2)1/2.|\nabla u_{\varepsilon}(y)|\leq CM_{2\varepsilon_{n}^{\prime\prime}}[\nabla u_{\varepsilon}](y)=C\bigg{(}\fint_{B_{2\varepsilon_{n}^{\prime\prime}}(y)}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}. (4.21)

This pointwise Lipschitz estimate follows from a well-known blow-up argument at small scales and the fact εn′′=k0Q0n1εn\varepsilon_{n}^{\prime\prime}=k_{0}Q_{0}^{n-1}\varepsilon_{n}. Inserting this into (4.20), we arrive at

(B0|uε|p1)1/p1C(5B0|uε|2)1/2.\bigg{(}\fint_{B_{0}}|\nabla u_{\varepsilon}|^{p_{1}}\bigg{)}^{1/p_{1}}\leq C\bigg{(}\fint_{5B_{0}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}. (4.22)

Since p1(2,p)p_{1}\in(2,p) and p(2,)p\in(2,\infty) are arbitrary with CC depending on p,p1p,p_{1}, we actually have shown the uniform interior W1,pW^{1,p} estimate for any p(0,)p\in(0,\infty) in the case of f=0f=0.

Finally, we consider general case with fLp(5B0)df\in L^{p}(5B_{0})^{d}. Let uεu_{\varepsilon} be a solution to (4.1). For each ball B=BrB=B_{r} such that 4B2B04B\subset 2B_{0}, let uε,1H01(4B)u_{\varepsilon,1}\in H^{1}_{0}(4B) be the solution to

div(A(x/ε1,x/ε2,,x/εn)uε,1)=div f in 4B,\displaystyle-\text{\rm div}\big{(}A(x/\varepsilon_{1},x/\varepsilon_{2},\cdots,x/\varepsilon_{n})\nabla u_{\varepsilon,1}\big{)}=\text{div }f\,\text{ in }4B, (4.23)

and uε,2u_{\varepsilon,2} the solution to

div(A(x/ε1,x/ε2,,x/εn)uε,2)=0 in 4B, and uε,2=uε on (4B).\displaystyle-\text{\rm div}\big{(}A(x/\varepsilon_{1},x/\varepsilon_{2},\cdots,x/\varepsilon_{n})\nabla u_{\varepsilon,2}\big{)}=0\,\text{ in }4B,\,\text{ and }\,u_{\varepsilon,2}=u_{\varepsilon}\,\text{ on }\partial(4B). (4.24)

Thus, uε=uε,1+uε,2u_{\varepsilon}=u_{\varepsilon,1}+u_{\varepsilon,2} in 4B4B. The energy estimate for (4.23) implies that

(4B|uε,1|2)1/2C(4B|f|2)1/2,\bigg{(}\fint_{4B}|\nabla u_{\varepsilon,1}|^{2}\bigg{)}^{1/2}\leq C\bigg{(}\fint_{4B}|f|^{2}\bigg{)}^{1/2},

while the uniform W1,pW^{1,p} estimates for uε,2u_{\varepsilon,2} proved above implies that

(2B|uε,2|p0)1/p0\displaystyle\bigg{(}\fint_{2B}|\nabla u_{\varepsilon,2}|^{p_{0}}\bigg{)}^{1/p_{0}} C(4B|uε,2|2)1/2\displaystyle\leq C\bigg{(}\fint_{4B}|\nabla u_{\varepsilon,2}|^{2}\bigg{)}^{1/2}
C(4B|uε|2)1/2+C(4B|f|2)1/2\displaystyle\leq C\bigg{(}\fint_{4B}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}+C\bigg{(}\fint_{4B}|f|^{2}\bigg{)}^{1/2}

for some p0(p,)p_{0}\in(p,\infty). Let =|uε|\mathcal{F}=|\nabla u_{\varepsilon}|, B=|uε,1|\mathcal{F}_{B}=|\nabla u_{\varepsilon,1}|, and B=|uε,2|\mathcal{R}_{B}=|\nabla u_{\varepsilon,2}|. We obtain (4.2) from Theorem 4.2 immediately (with t=0t=0). ∎

Theorem 4.1 provides the uniform interior W1,pW^{1,p} estimate for the equation (1.1). Based on Theorem 3.4 and a boundary version of Theorem 4.2 (See Theorem 4.1 and Remark 4.2 in [She23].), we can follow the same arguments to prove the following uniform boundary W1,pW^{1,p} estimate.

Theorem 4.3.

Let Ω\Omega be a bounded C1C^{1} domain. There exists r0=r0(Ω)r_{0}=r_{0}(\Omega) such that the following statement is true. Let B=B(x,r)B=B(x,r) with xΩx\in\partial\Omega, 0<r<r00<r<r_{0}. Assume AA satisfies the assumption (1.2), (1.3) and (1.4). Let uεu_{\varepsilon} be a weak solution to

div(A(x/ε1,,x/εn)uε)=divf in B2rΩ,uε=0 on B2rΩ,\displaystyle-\text{\rm div}\big{(}A(x/\varepsilon_{1},\cdots,x/\varepsilon_{n})\nabla u_{\varepsilon}\big{)}=\text{\rm div}f\,\text{ in }B_{2r}\cap\Omega,\quad u_{\varepsilon}=0\,\text{ on }B_{2r}\cap\partial\Omega, (4.25)

with fLp(B2rΩ)df\in L^{p}(B_{2r}\cap\Omega)^{d} for some 2<p<2<p<\infty. Then uεLp(BrΩ)d\nabla u_{\varepsilon}\in L^{p}(B_{r}\cap\Omega)^{d} and

(BrΩ|uε|p)1/pC{(B2rΩ|uε|2)1/2+(B2rΩ|f|p)1/p},\displaystyle\bigg{(}\fint_{B_{r}\cap\Omega}|\nabla u_{\varepsilon}|^{p}\bigg{)}^{1/p}\leq C\bigg{\{}\bigg{(}\fint_{B_{2r}\cap\Omega}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}+\bigg{(}\fint_{B_{2r}\cap\Omega}|f|^{p}\bigg{)}^{1/p}\bigg{\}}, (4.26)

where CC depends only on d,Λ,p,n,(τ,L)d,\Lambda,p,n,(\tau,L) in (1.4), and the C1C^{1} character of Ω\Omega.

We finally provide the proof of our main results.

Proof of Theorem 1.1.

Note that the case p=2p=2 follows from standard energy estimates, while the case p>2p>2 is a direct consequence of Theorems 4.1 and 4.3 together with a covering argument. Finally, the case 1<p<21<p<2 follows from the case p>2p>2 and a standard duality argument. ∎

5. Large-scale Lipschitz estimates

In this section, we investigate the large-scale Lipschitz estimates for (1.1). We distinguish between the cases of two scales and more scales. We recall that under the scale-separation condition: there exists a positive integer NN such that

(εi+1εi)Nεiεi1 for 1in1,\left(\frac{\varepsilon_{i+1}}{\varepsilon_{i}}\right)^{N}\leq\frac{\varepsilon_{i}}{\varepsilon_{i-1}}\quad\text{ for }1\leq i\leq n-1, (5.1)

the large-scale Lipschitz estimate for the multiscale elliptic operators has been derived in [NSX20]. Indeed, let uεu_{\varepsilon} be the weak solution to

div(A(x,x/ε1,,x/εn)uε)=F in B1-\text{\rm div}(A(x,x/\varepsilon_{1},\cdots,x/\varepsilon_{n})\nabla u_{\varepsilon})=F\quad\text{ in }B_{1}

with FLp(B1)F\in L^{p}(B_{1}) for some p>dp>d. Suppose A(x,y1,,yn)A(x,y_{1},\cdots,y_{n}) is strongly elliptic, periodic in yiy_{i} for 1in1\leq i\leq n, and Hölder continuous in x,y1,,yn1x,y_{1},\cdots,y_{n-1} (no smoothness is needed for yny_{n}). Then, under the condition (5.1), for any 0<εnr<10<\varepsilon_{n}\leq r<1,

(Br|uε|2)1/2C{(B1|uε|2)1/2+(B1|F|p)1/p},\displaystyle\bigg{(}\fint_{B_{r}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}\leq C\bigg{\{}\bigg{(}\fint_{B_{1}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}+\bigg{(}\fint_{B_{1}}|F|^{p}\bigg{)}^{1/p}\bigg{\}}, (5.2)

where CC is independent of ε\varepsilon.

However, as we have stated before, without any scale separation condition the full-scale uniform Lipschitz estimate for (1.1) seems difficult and still remains open. The following theorems provide suboptimal results in this direction using the idea of scale separation similar to Section 3.1.

Theorem 5.1.

Assume A(y1,y2)A(y_{1},y_{2}) satisfies (1.2), (1.3) and (3.9) with n=2n=2. For B1=B1(0)B_{1}=B_{1}(0), let uεu_{\varepsilon} be a weak solution of div(Aεuε)=F-\text{\rm div}(A_{\varepsilon}\nabla u_{\varepsilon})=F in B1B_{1} with FLp(B1)F\in L^{p}(B_{1}) for some p>dp>d. Then for any α>0\alpha>0, and any ε21αr1\varepsilon_{2}^{1-\alpha}\leq r\leq 1, we have

(Br|uε|2)1/2Cα{(B1|uε|2)1/2+(B1|F|p)1/p},\bigg{(}\fint_{B_{r}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}\leq C_{\alpha}\bigg{\{}\bigg{(}\fint_{B_{1}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}+\bigg{(}\fint_{B_{1}}|F|^{p}\bigg{)}^{1/p}\bigg{\}}, (5.3)

where CαC_{\alpha} depends only on d,Λ,p,α,d,\Lambda,p,\alpha, and (τ,L)(\tau,L) in (3.9).

Proof.

Applying the argument in Section 3.1 and by (3.7), for any Q>1Q>1 we can find a 1-periodic matrix A(y1,y2)A^{\sharp}(y_{1},y_{2}), Hölder continuous in y1y_{1}, and write

A(xε1,xε2)=A(γ1xε2,xqε2)=A(xε1,xε2),A\big{(}\frac{x}{\varepsilon_{1}},\frac{x}{\varepsilon_{2}}\big{)}=A^{\sharp}\big{(}\frac{\gamma_{1}x}{\varepsilon_{2}},\frac{x}{q\varepsilon_{2}}\big{)}=A^{\sharp}\big{(}\frac{x}{\varepsilon_{1}^{\prime}},\frac{x}{\varepsilon_{2}^{\prime}}\big{)}, (5.4)

where 1qQ,ε1=ε2/γ11\leq q\leq Q,\varepsilon_{1}^{\prime}=\varepsilon_{2}/\gamma_{1} and ε2=qε2\varepsilon_{2}^{\prime}=q\varepsilon_{2}. Let Q=ε2αQ=\varepsilon_{2}^{-\alpha}. Then ε2ε2=qε2ε21α\varepsilon_{2}\leq\varepsilon_{2}^{\prime}=q\varepsilon_{2}\leq\varepsilon_{2}^{1-\alpha} and ε1/ε2Q=ε2α\varepsilon_{1}^{\prime}/\varepsilon_{2}^{\prime}\geq Q=\varepsilon_{2}^{-\alpha}. Thus ε2\varepsilon_{2}^{\prime} and ε1\varepsilon_{1}^{\prime} are well-separated. Moreover, uεu_{\varepsilon} satisfies the equation

div(A(x/ε1,x/ε2)uε)=Fin B1.-\text{\rm div}(A^{\sharp}(x/\varepsilon_{1}^{\prime},x/\varepsilon_{2}^{\prime})\nabla u_{\varepsilon})=F\qquad\text{in }B_{1}. (5.5)

Thanks to (5.2), we have for ε2r1\varepsilon_{2}^{\prime}\leq r\leq 1,

(Br|uε|2)1/2Cα{(B1|uε|2)1/2+(B1|F|p)1/p}.\bigg{(}\fint_{B_{r}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}\leq C_{\alpha}\bigg{\{}\bigg{(}\fint_{B_{1}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}+\bigg{(}\fint_{B_{1}}|F|^{p}\bigg{)}^{1/p}\bigg{\}}. (5.6)

This implies the desired estimate since ε2ε21α\varepsilon_{2}^{\prime}\leq\varepsilon_{2}^{1-\alpha}. ∎

Theorem 5.2.

Assume A(y1,,yn)A(y_{1},\cdots,y_{n}) satisfies (1.2), (1.3) and (1.4) with n3n\geq 3. Let uεu_{\varepsilon} be a weak solution of div(Aεuε)=F-\text{\rm div}(A_{\varepsilon}\nabla u_{\varepsilon})=F in B1B_{1} with FLp(B1)F\in L^{p}(B_{1}) for some p>dp>d. Then for any α(0,1/n]\alpha\in(0,1/n] and δ[εn,1)\delta\in[\varepsilon_{n},1), there exists 1qδnα1\leq q\leq\delta^{-n\alpha}, such that for any qδrr1:=min{qδ1α,1}q\delta\leq r\leq r_{1}:=\min\{q\delta^{1-\alpha},1\}, we have

(Br|uε|2)1/2Cα{(Br1|uε|2)1/2+r11d/p(Br1|F|p)1/p},\bigg{(}\fint_{B_{r}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}\leq C_{\alpha}\bigg{\{}\bigg{(}\fint_{B_{r_{1}}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}+r_{1}^{1-d/p}\bigg{(}\fint_{B_{r_{1}}}|F|^{p}\bigg{)}^{1/p}\bigg{\}}, (5.7)

where CαC_{\alpha} depends only on d,n,Λ,p,αd,n,\Lambda,p,\alpha, and (τ,L)(\tau,L) in (1.4). Moreover, if δ=εj\delta=\varepsilon_{j} for some 1jn1\leq j\leq n, then α\alpha can be taken from (0,1/(n1)](0,1/(n-1)] and 1qδ(n1)α1\leq q\leq\delta^{-(n-1)\alpha}.

Proof.

Fix δ[εn,1)\delta\in[\varepsilon_{n},1). Let Q=δα>1Q=\delta^{-\alpha}>1 with α(0,1/n]\alpha\in(0,1/n]. By the Dirichlet’s theorem, there exist q,p1,p2,,pnq,p_{1},p_{2},\cdots,p_{n} such that 1<qQn=δnα1<q\leq Q^{n}=\delta^{-n\alpha} and

|δεipiq|1qQ,\big{|}\frac{\delta}{\varepsilon_{i}}-\frac{p_{i}}{q}\big{|}\leq\frac{1}{qQ}, (5.8)

where i=1,2,,ni=1,2,\cdots,n. As in Section 3.1, we set

γi:=|δεipiq|andsi=sgn(δεipiq),\gamma_{i}:=\Big{|}\frac{\delta}{\varepsilon_{i}}-\frac{p_{i}}{q}\Big{|}\quad\text{and}\quad s_{i}=\text{sgn}\Big{(}\frac{\delta}{\varepsilon_{i}}-\frac{p_{i}}{q}\Big{)},

and write

1εi=γisiδ+piq1δ.\frac{1}{\varepsilon_{i}}=\frac{\gamma_{i}s_{i}}{\delta}+\frac{p_{i}}{q}\frac{1}{\delta}.

Define

A1(y1,y2,,yn,yn+1)=A(s1y1+p1yn+1,,snyn+pnyn+1).A_{1}^{\sharp}(y_{1},y_{2},\cdots,y_{n},y_{n+1})=A(s_{1}y_{1}+p_{1}y_{n+1},\cdots,s_{n}y_{n}+p_{n}y_{n+1}). (5.9)

As a consequence, we have

A(xε1,,xεn)=A1(γ1xδ,,γnxδ,xqδ).A\big{(}\frac{x}{\varepsilon_{1}},\cdots,\frac{x}{\varepsilon_{n}}\big{)}=A_{1}^{\sharp}\big{(}\frac{\gamma_{1}x}{\delta},\cdots,\frac{\gamma_{n}x}{\delta},\frac{x}{q\delta}\big{)}. (5.10)

The last identity shows that we can rewrite original coefficient matrix AA with nn scales as a new 1-periodic matrix A1A_{1}^{\sharp} with n+1n+1 scales. Moreover, the smallest scale qδq\delta is at least QQ-separated from the remaining nn scales, i.e.,

δ/γiqδQ.\frac{\delta/\gamma_{i}}{q\delta}\geq Q. (5.11)

Hence, by a blow-up argument and (5.2) we have

(Br|uε|2)1/2Cα{(Br1|uε|2)1/2+r11d/p(Br1|F|p)1/p}\bigg{(}\fint_{B_{r}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}\leq C_{\alpha}\bigg{\{}\bigg{(}\fint_{B_{r_{1}}}|\nabla u_{\varepsilon}|^{2}\bigg{)}^{1/2}+r_{1}^{1-d/p}\bigg{(}\fint_{B_{r_{1}}}|F|^{p}\bigg{)}^{1/p}\bigg{\}} (5.12)

for any rr with qδrr1min{Qqδ,1}q\delta\leq r\leq r_{1}\leq\min\{Qq\delta,1\}. This is the desired estimate. To make sure qδ<1q\delta<1, we require qδQnδ=δ1nα<1q\delta\leq Q^{n}\delta=\delta^{1-n\alpha}<1 which gives α1/n\alpha\leq 1/n.

Finally, in the particular case δ=εj\delta=\varepsilon_{j} for some 1jn1\leq j\leq n, δ/εj=1\delta/\varepsilon_{j}=1 leads to a trivial approximation in (5.8). Thus we will only need (5.8) for iji\neq j and 1qQn11\leq q\leq Q^{n-1}. The rest of the proof follows similarly. ∎

References

  • [AB96] G. Allaire and M. Briane. Multiscale convergence and reiterated homogenisation. Proc. Roy. Soc. Edinburgh Sect. A, 126(2):297–342, 1996.
  • [AGK16] S. N. Armstrong, A. Gloria, and T. Kuusi. Bounded correctors in almost periodic homogenization. Arch. Ration. Mech. Anal., 222(1):393–426, 2016.
  • [AK22] S. N. Armstrong and T. Kuusi. Elliptic homogenization from qualitative to quantitative. arXiv:2210.06488, 2022.
  • [AL87] M. Avellaneda and F. Lin. Compactness methods in the theory of homogenization. Comm. Pure Appl. Math., 40(6):803–847, 1987.
  • [AL91] M. Avellaneda and F. Lin. LpL^{p} bounds on singular integrals in homogenization. Comm. Pure Appl. Math., 44(8-9):897–910, 1991.
  • [AS16] S. N. Armstrong and C. K. Smart. Quantitative stochastic homogenization of convex integral functionals. Ann. Sci. Éc. Norm. Supér. (4), 49(2):423–481, 2016.
  • [Ave96] M. Avellaneda. Homogenization and renormalization: the mathematics of multi-scale random media and turbulent diffusion. In Dynamical systems and probabilistic methods in partial differential equations (Berkeley, CA, 1994), volume 31 of Lectures in Appl. Math., pages 251–268. Amer. Math. Soc., Providence, RI, 1996.
  • [BG19] A. Benoit and A. Gloria. Long-time homogenization and asymptotic ballistic transport of classical waves. Ann. Sci. Éc. Norm. Supér. (4), 52(3):703–759, 2019.
  • [CP98] L. A. Caffarelli and I. Peral. On W1,pW^{1,p} estimates for elliptic equations in divergence form. Comm. Pure Appl. Math., 51(1):1–21, 1998.
  • [GH03] B. Gérard and O. Houman. Multiscale homogenization with bounded ratios and anomalous slow diffusion. Comm. Pure Appl. Math., 56(1):80–113, 2003.
  • [GNO20] A. Gloria, S. Neukamm, and F. Otto. A regularity theory for random elliptic operators. Milan J. Math., 88(1):99–170, 2020.
  • [KLS13] C. E. Kenig, F. Lin, and Z. Shen. Homogenization of elliptic systems with Neumann boundary conditions. J. Amer. Math. Soc., 26(4):901–937, 2013.
  • [Koz78] S. M. Kozlov. Averaging of differential operators with almost periodic rapidly oscillating coefficients. Mat. Sb. (N.S.), 107(149)(2):199–217, 317, 1978.
  • [NSX20] W. Niu, Z. Shen, and Y. Xu. Quantitative estimates in reiterated homogenization. J. Funct. Anal., 279(11):108759, 39, 2020.
  • [NZ23] W. Niu and J. Zhuge. Compactness and stable regularity in multiscale homogenization. Math. Ann., 385(3-4):1431–1473, 2023.
  • [Sch91] W. M. Schmidt. Diophantine approximations and Diophantine equations, volume 1467 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1991.
  • [She15] Z. Shen. Convergence rates and Hölder estimates in almost-periodic homogenization of elliptic systems. Anal. PDE, 8(7):1565–1601, 2015.
  • [She17] Z. Shen. Boundary estimates in elliptic homogenization. Anal. PDE, 10(3):653–694, 2017.
  • [She18] Z. Shen. Periodic homogenization of elliptic systems, volume 269 of Operator Theory: Advances and Applications. Birkhäuser/Springer, Cham, 2018. Advances in Partial Differential Equations (Basel).
  • [She23] Z. Shen. Weighted L2L^{2} estimates for elliptic homogenization in Lipschitz domains. J. Geom. Anal., 33(1):Paper No. 3, 33, 2023.
  • [SZ18] Z. Shen and J. Zhuge. Approximate correctors and convergence rates in almost-periodic homogenization. J. Math. Pures Appl. (9), 110:187–238, 2018.