Home Homogenization of oblique boundary value problems
Article Open Access

Homogenization of oblique boundary value problems

  • Sunhi Choi and Inwon C. Kim EMAIL logo
Published/Copyright: March 9, 2023

Abstract

We consider a nonlinear Neumann problem, with periodic oscillation in the elliptic operator and on the boundary condition. Our focus is on problems posed in half-spaces, but with general normal directions that may not be parallel to the directions of periodicity. As the frequency of the oscillation grows, quantitative homogenization results are derived. When the homogenized operator is rotation-invariant, we prove the Hölder continuity of the homogenized boundary data. While we follow the outline of Choi and Kim (Homogenization for nonlinear PDEs in general domains with oscillatory Neumann boundary data, Journal de Mathématiques Pures et Appliquées 102 (2014), no. 2, 419–448), new challenges arise due to the presence of tangential derivatives on the boundary condition in our problem. In addition, we improve and optimize the rate of convergence within our approach. Our results appear to be new even for the linear oblique problem.

MSC 2010: 35J66

1 Introduction

For given ε > 0 , ν S n 1 and τ R n , let u ε be a bounded solution of the following problem:

F D 2 u ε , x ε = 0 in Π { x R n : 1 < ( x τ ) ν < 0 } u ε = h ( x ) on H 1 { ( x τ ) ν = 1 } ν u ε = G D u ε , x ε on H 0 { ( x τ ) ν = 0 } . ( P ) ε

Here, F ( M , y ) and G ( p , y ) are Z n -periodic in the y variable. We also assume the boundary condition to be oblique and F to be uniformly elliptic: see Section 1.1 for precise assumptions on F and G .

The examples of boundary conditions we consider include the linear oblique problem:

(1) γ x ε D u + g x ε = 0 ,

where the vector field γ satisfies c x ε , ν γ x ε ν > 0 . In this case, one can write

G ( p , y ) = ( c ( y ) ) 1 [ γ ( y ) p T + g ( y ) ] ,

where p T p ( p ν ) is the tangential component of p R n on H 0 . A nonlinear example is capillarity-type conditions, for which G is given by

(2) G ( p , y ) = θ ( y ) 1 + p 2 ,

where θ ( x ) < 1 .

We are interested in the behavior of u ε as ε tends to zero. Our objective is to extend the results of [7,8], to establish a general framework to understand nonlinear elliptic problems with oscillatory Neumann boundary data. In particular, we have tried to carefully detail the double-scale averaging argument given in Section 5, which has been central in understanding continuity properties of the homogenized boundary condition in both Neumann and Dirichlet boundary problems: see [7,8,12,13]. We focus on problems posed on half-spaces here. To deal with domains with general geometry, the approach taken in [7] or [13] uses fundamental solutions as barriers to bound the potential singularity generated at points with rational normals. For our problem, while our result is likely to hold in general domains, we suspect that these singular solutions may cause new challenges in dealing with perturbative arguments, due to their singularity in tangential derivatives.

Note that, as first pointed out by Bensoussan et al. [5], if ν is a multiple of a vector in Z n (i.e., if ν is rational), then τ ν must be zero for u ε to converge, since otherwise the Neumann boundary condition changes drastically as ε changes, and thus, u ε would not have a limit. When ν is irrational, we expect u ε to average due to the ergodic property of its Neumann data. However, in this case, u ε is no longer periodic, and thus, interesting challenges arise in dealing with the inherent lack of compactness. Compared to [7] where the linear Neumann problem was considered, there is an additional challenge in our setting given by the presence of tangential derivatives on the boundary condition. We will discuss some of the relevant literature on this issue.

Let us state a convergence result on ( P ) ε to begin the discussion. Let F ¯ be the homogenized operator of F obtained by Evans [11].

Theorem 1.1

Let ν be irrational, or otherwise suppose that τ = 0 . Let us assume (F1)–(F3) and (G1)–(G3) (see Section 1.1). In addition, suppose that F ( , x ) is convex when G ( , x ) is nonlinear. Then there exists μ ( η , q ) : S n 1 × R n 1 R , where μ is independent of τ , such that u ε converges uniformly to the unique bounded solution u ¯ of the oblique boundary problem:

F ¯ ( D 2 u ¯ ) = 0 i n Π u ¯ = h ( x ) o n H 1 ν u ¯ = μ ( ν , D T u ¯ ) o n H 0 . ( P ¯ )

(here, D T u denotes the tangential derivative of u along the direction ν .) Moreover, μ is Lipschitz continuous with respect to q. Finally, if F ¯ ( M ) is rotation-invariant, then μ is also Hölder continuous over irrational directions ν with exponent α = 1 5 n .

The proof of Theorem 1.1 will be given later in this section, based on our main result (Theorem 1.2), which establishes rates of convergence for (approximate) cell problem solutions. Our work extends the previous results in [8] on linear Neumann problems, where G ( p , y ) = G ( y ) . For general, G ( p , y ) additional challenges arise due to the presence of tangential derivatives on the boundary condition, which necessitates Lipschitz regularity estimates for the solutions. As noted in [13], the continuity property of μ ( ν , q ) fails when F ¯ is not rotation-invariant, even when it is convex. When the continuity result holds for μ we expect to be able to address domains of general geometry, building on our result and proceeding as in [7].

It is unknown whether the form of the boundary condition such as (1) or (2) is preserved in the limit ε 0 . With the exception of linear problems, the interaction between the operator F and the boundary condition remains to be better understood to yield further characterizations of the homogenized problems.

Literature. Before proceeding further, let us briefly describe some of the relevant literature. In the classical article in [5], the following problem was considered:

(3) A x ε u ε = 0 in Ω , ν A x ε u ε ( x ) = g x ε on Ω .

For this co-normal boundary value problem, explicit integral formulas have been derived for the limiting operator as well as for the limiting boundary data, under the assumption that Ω does not contain any flat piece with a rational normal.

For linear elliptic systems with either Dirichlet or Neumann problem with co-normal derivatives, there has been a recent surge of development in quantitative homogenization relying on the integral representation of solutions: we refer to [2,15,20] and the references therein.

For nonlinear problems, or even for linear problems with non co-normal boundary data, until recently, the focus has been on half-space type domains with rational normal, with the origin on the boundary. In [21], Tanaka considered some model problems in half-space whose boundary is parallel to the axes of the periodicity by purely probabilistic methods. In [1], Arisawa studied specific problems in oscillatory domains near half spaces going through the origin. Generalizing the results of Arisawa [1] for nonlinear boundary conditions, Barles et al. [4] studied the problem for operators with oscillating coefficients, in half-space type domains whose boundary is parallel to the axes of periodicity. We also refer to [14], which adopts an integro-differential approach to study linear scalar problems with the specific Neumann problem G ( p , y ) = g ( y ) .

For the linear Neumann problem G ( p , y ) = g ( y ) in ( P ) ε , corresponding results to Theorems 1.1 and 1.2 have been recently shown in [8]. General domains has been considered in [7] based on the cell problem analysis in [8]. Corresponding results for the Dirichlet boundary data have been obtained in [12]. Finally, for general operator F , [13] discusses the generic nature of discontinuity for the homogenized boundary data, for either linear Neumann or Dirichlet problem.

Cell problem. By the formal expansion u ε = u ¯ ( x ) + ε v x , x ε + O ( ε 2 ) , the cell problem for v was derived in [4] for a rational ν and τ = 0 . There they find a unique constant μ = μ ( ν , q ) for q ν such that the boundary value problem

F ( D 2 v , y ) = 0 in { y ν 0 } , μ = G ( D v + p , y ) on H 0 , ( C )

with p = μ ν + q , has a bounded periodic solution v in { y ν 0 } . The existence of bounded v leads to the uniform convergence of u ε to u ¯ in the limit ε 0 with p = D u ¯ on H 0 .

For general ν and τ , an approximate cell problem needs to be derived, since v is no longer expected to be periodic and thus compactness is lost. In the context of ( C ) , our result shows that for irrational ν , there exists a unique constant μ = μ ( ν , q ) for q ν such that the problem

F ( D 2 v , y + τ ) = 0 in { y ν 0 } , μ = G ( D v + p , y + τ ) on H 0 ( C ˜ )

has a solution with sublinear growth at infinity, for any τ R n . To show this, we use the ergodicity of Neumann data in a scale depending on ν , and the stability of solutions under perturbation of boundary conditions. When the homogenized operator F ¯ is rotation-invariant, we show that v is stable as the normal direction of the domain ν varies. A quantitative version of this stability property yields the mode of continuity for μ as ν varies.

A discussion on assumptions on F and G . Our assumptions on F and G are mainly to obtain Lipschitz estimates for the solutions of ( C ˜ ) . The Lipschitz estimates ensure that the solution of the cell problem has the ergodic structure with respect to translations along the Neumann boundary (see Lemma 3.5), when ε changes in ( P ) ε and when τ is not the origin. In particular to guarantee the Lipschitz bound, available literature restricts F ( M , x ) to be convex with respect to M when G is a nonlinear function of D u . We refer to [3] for a detailed description of the regularity theory on nonlinear Neumann boundary problems. For the continuity properties of μ , we further need C 1 , α estimates for solutions of ( C ˜ ) ; however, this does not further restrict the class of problems we can address.

1.1 Assumptions and main results

Let T be the 1-periodic torus in R n , and let n be the space of real n × n symmetric matrices. Consider the functions F ( M , y ) : n × T R and G ( p , y ) : R n × T satisfying the following properties:

  1. (Uniform Ellipticity) There exist constants 0 < λ < Λ such that

    λ Tr ( N ) F ( M , y ) F ( M + N , y ) Λ Tr ( N )

    for all y T and M , N n with N 0 .

  2. (1-Homogeneity) F ( t M , y ) = t F ( M , y ) for all y T , t > 0 and M n .

  3. (Lipschitz continuity) There exists C > 0 such that for all y 1 , y 2 T and M , N n ,

    F ( M , y 1 ) F ( N , y 2 ) C ( y 1 y 2 ( 1 + M + N ) + M N ) .

  4. (At most linear growth) G ( p , x ) μ 0 ( 1 + p ) .

  5. (Lipschitz continuity) ( 1 + p ) G p , G y m ( 1 + p ) for some m > 0 .

  6. (Oblicity) G p ν c < 1 .

A typical example of an operator F satisfying (F1)–(F3) is the linear elliptic operator

(4) F ( D 2 u , x ) = Σ i , j a i j ( x ) x i x j u ,

where a i j : R n R is periodic and Lipschitz continuous. A nonlinear example is the Bellman-Isaacs operator arising from stochastic optimal control and differential games

(5) F ( D 2 u , x ) = inf β B sup α A { α , β u } ,

where α , β is a family of uniformly elliptic operators of the form (4). In fact, all operators satisfying (F1)–(F3) can be written as (5). As for G , the ones given in (1) and (2) with Lipschitz coefficients c 1 γ , c 1 g and θ satisfy (G1)–(G3).

For τ R n and ν S n 1 , let us define a strip domain

Π ( τ , ν ) { x R n : 1 ( x τ ) ν 0 }

and a hyperplane

H s ( τ , ν ) { ( x τ ) ν = s } .

We will denote H s ( τ , ν ) by H s throughout the article when it is unambiguous. For a given q ν , let u ε solve the following approximate cell problem:

F D 2 u ε , x ε = 0 in Π ( τ , ν ) ν u ε = G ( D u ε , x ε ) on H 0 u ε ( x ) = q x on H 1 ( P ) ε , ν , τ , q .

Now we are ready to state the main result.

Theorem 1.2

Let u ε solve ( P ) ε , ν , τ , q . Suppose that either ν is irrational or τ = 0 . Then the following holds:

  1. There exists μ = μ ( ν , q ) such that u ε converges uniformly to the linear profile

    u ( x ) μ ( ( x τ ) ν + 1 ) + q x .

    Here, μ ( ν , q ) is independent of τ and Lipschitz continuous with respect to q. Moreover, we have

    (6) u ε u C Λ ( ε , ν ) in Π ( τ , ν ) ,

    where Λ ( ε , ν ) (as given in (23)) is an increasing function of ε such that lim ε 0 Λ ( ε , ν ) = 0 .

  2. When F ¯ is rotation-invariant, there exists a continuous extension μ ¯ ( ν , q ) : S n 1 × R n R of μ ( ν , q ) over irrational directions ν S n 1 R Z n . Moreover, μ ¯ is Lipschitz in q and C α in ν , with α = 1 5 n .

The proof is given in Theorems 4.1, 5.2, and 5.1.

A discussion on the rate of convergence Λ ( ε , ν ) . Here, we briefly describe the geometric process used in Section 4 to obtain an upper bound for the rate function Λ in (6). Given δ > 0 , we are interested in finding ε 0 = ε 0 ( ν , δ ) such that u ε u C δ for ε ε 0 .

If ν is rational and τ = 0 , F and G are periodic along ν -direction with period T ν . Hence, we expect that ε 0 needs to be smaller than 1 / T ν for a fixed δ . In fact, Theorem 4.1 (d) yields that

Λ ( ε , ν ) δ for ε ε 0 = δ 2 / T ν

and thus yields a uniform bound

(7) Λ ( ε , ν ) C ( ν ) ε 1 / 2 .

If ν is irrational, for each δ , we choose a reference rational direction P as follows: choose a point P = P ( ν , δ ) Z n such that

(8) T ν P δ for some T = T ( ν , δ ) > 0 .

Then F and G are periodic along P -direction with period T + O ( δ ) . If we let θ = θ ( ν , δ ) be the angle between ν and P , then (8) can be written as θ < δ / T . If R < 1 / θ , then due to the proximity of ν to P direction, G ( p , ) takes only limited values of G on H 0 B R ( τ ) , even though ν is irrational. In other words, G ( p , ) exhibits ergodicity on H 0 only in a neighborhood of size R > 1 / θ . For this reason, u ε homogenizes only when ε O ( θ ) . Indeed Theorem 4.1 (c) yields that

Λ ( ε , ν ) δ for ε ε 0 = δ 2 θ .

Since θ depends on not only ν but also δ , we are not able to separate the dependence of the rate function on ε and ν , without further estimate of θ or T as δ varies. Such estimate would require better understanding of the discrepancy function discussed in in [7], [8] and [12].

Proof of Theorem 1.1

Once Theorem 1.2 (a) is obtained, one can derive our main theorem by the perturbed test function arguments introduced by Evans [10].

Let u ε solve ( P ) ε and define u and u as follows:

u = limsup u ε lim r 0 sup ( y , ε ) S r x u ε ( y ) ; u = liminf u ε lim r 0 inf ( y , ε ) S r x u ε ( y ) ,

where S r x = { ( y , ε ) : y Π , x y < r , 0 < ε < r } . First, observe that, by using a barrier of the form

φ M ( x ) M ( ( x τ ) ν + 1 ) + f ( x ) ,

where f is a C 2 -approximation of h that is larger than h , one can conclude that u ε φ M in Π for any large M , and thus, u h on H 1 . Similar arguments yield that u h on H 1 .

We claim that u and u are, respectively, a viscosity subsolution and supersolution of ( P ) . If the claim is true, then Corollary 3.4 applies to yield that u u . Since the opposite inequality is true from the definition, we conclude that u = u , which means that u ε uniformly converges in Ω ¯ .

Below we will only show that u is a subsolution of ( P ) , since the proof for u can be shown by parallel arguments. To this end, suppose that u ϕ has a local max in B r ( y 0 ) Π ¯ with a smooth test function ϕ . If y 0 is in the interior of Π , then F ¯ ( D 2 ϕ ) ( y 0 ) 0 due to standard interior homogenization (see, for instance, [10]). Hence, it remains to show that if y 0 is on the Neumann boundary, then ϕ satisfies

(9) ν ϕ μ ( ν , q D T ϕ ) at x = y 0 .

First, suppose that ν is rational and y 0 ν = 0 . We may assume for simplicity that u ( y 0 ) = ϕ ( y 0 ) = 0 and define P ( x ) D ϕ ( y 0 ) ( x y 0 ) . Since Π { x : x ν < 0 } , for any δ > 0 , we may choose r sufficiently small that l δ ( x ) P ( x ) δ ( x ν ) is strictly larger than u on B r ( 0 ) Π . Then for sufficiently small choice of ε , we have

(10) l δ > u ε on B r ( 0 ) H r δ , where H r δ = { x ν = r δ } .

Let ε ¯ ( r δ ) 1 ε and consider the re-scaled function v ε ( x ) ( r δ ) 1 u ε ( r δ x ) l δ ( x ) . Then v ε is a subsolution of ( P ) ε ¯ , ν , 0 , q , in the local domain Π B δ 1 ( 0 ) . Note that the corresponding Neumann boundary for v ε remains to be H 0 since y 0 ν = 0 : in general, it will be { ( x τ ) ν = 0 } with

(11) τ = ( ε ¯ ) 1 y 0 ,

and thus, the choice of τ must change as we vary ε ¯ . We will compare v ε with w ε ¯ , the unique bounded solution of ( P ) ε ¯ , ν , 0 , q in Π obtained in Lemma 3.3. Due to the localization lemma (Lemma 3.2), we have

(12) v ε w ε ¯ + M δ in Π B 1 ( 0 ) .

Due to Theorem 1.2, we have

w ε ¯ μ ( ν , q ) ( x ν + 1 ) + q x + Λ ( ε ¯ , ν ) in Π .

Since Λ ( ε , ν ) 0 as ε 0 , (10) and (12) yield that

(13) limsup ε 0 u ε ( r δ x ) r δ limsup ε 0 v ε ( x ) + l δ ( ν ) μ ( ν , q ) ( x ν + 1 ) + q x + l δ ( ν ) + M δ in Π B 1 ( 0 ) .

Now suppose that (9) is false, then there exists δ > 0 such that

(14) ν ϕ ( 0 ) = δ l δ ( ν ) > μ ( ν , q ) + ( M + 1 ) δ .

This means that the right-hand side of (13) is strictly negative at x = 0 , which contradicts the assumption that u ( 0 ) = 0 .

Next suppose that ν is irrational, we need to choose τ depending on ε ¯ so that (11) holds. Then we argue as earlier with a solution of ( P ) ε ¯ , ν , τ , q in Π . Here, we must use the fact that ν is irrational, and thus, Theorem 1.2 ensures the uniform convergence of w ε ¯ to the linear profile is regardless of the choice of τ .□

2 Preliminaries

We adopt the following definition of viscosity solutions, which is equivalent to the one given in [9]. Let Ω be domain in R n with Ω as a disjoint union of Γ 0 and Γ 1 . Let F satisfy (F1)–(F3) in the previous section, and let G satisfy (G3) with G ( p , x ) being uniformly continuous in p independent of the choice of x . For f C ( Γ 0 ) , consider the following problem:

F ( D 2 u , x ) = 0 in Ω u = f ( x ) on Γ 0 ν u = G ( D u , x ) on Γ 1 , ( P )

where ν = ν ( x ) is the outward unit normal at x Γ 1 . Here, we replace (G3) with

( G 3 ) (Oblicity) G p ν c < 1 on Ω , where ν = ν x is the outward normal at x Ω .

Definition 2.1

  1. An upper semi-continuous function u : Ω ¯ R is a viscosity subsolution of ( P ) if u cannot cross from below any C 2 function ϕ , which satisfies

    F ( D 2 ϕ , x ) > 0 in Ω , ϕ > f on Γ 0 , ν D ϕ > G ( D ϕ , x ) on Γ 1 .

  2. A lower semi-continuous function u : Ω ¯ R is a viscosity supersolution of ( P ) if u cannot cross from above any C 2 function φ , which satisfies

    F ( D 2 ϕ , x ) < 0 in Ω , ϕ < f on Γ 0 , ν D ϕ < G ( D ϕ , x ) on Γ 1 .

  3. u is a viscosity solution of ( P ) if its upper semi-continuous envelope u is a viscosity subsolution and its lower semi-continuous envelope u is a viscosity supersolution of ( P ) .

Existence and uniqueness of viscosity solutions of ( P ) are based on the comparison principle we state later. We refer to [9,16] for details on the proof of the following theorem as well as the well-posedness of the problem ( P ) .

Theorem 2.2

Let G and F satisfy the conditions (G1) and (G3) and (F1)–(F3) in the previous section, with G being uniformly continuous in p independent of the choice of x. Let u and v be, respectively, bounded viscosity subsolution and supersolution of ( P ) in a bounded domain Ω . Then u v in Ω .

For a symmetric n × n matrix M , we decompose M = M + M with M ± 0 and M + M = 0 . We define the Pucci operators as follows:

P + ( M ) = Λ tr ( M + ) + λ tr ( M )

and

P ( M ) = λ tr ( M + ) + Λ tr ( M )

where 0 < λ < Λ . Later this article, we will utilize the fact that the difference of two solutions of F ( D 2 u , x ) = 0 is both a subsolution of P + ( D 2 u ) 0 and a supersolution of P ( D 2 ( u ) ) 0 (see [6]).

Next we state some regularity results that will be used throughout this article.

Theorem 2.3

[Chapter 8, [6], modified for our setting] Let u be a viscosity solution of F ( D 2 u , x ) = 0 in a domain Ω . Then for any compact subset Ω of Ω , we have

D u L ( Ω ) C d 1 u L ( Ω ) ,

where d = d ( Ω , Ω ) and C > 0 depends on n, λ , and Λ .

As mentioned in Section 1, regularity results for nonlinear Neumann problems are rather limited. C 0 , α estimates have been obtained by Barles and Da Lio in the general framework [3]. While a priori results for the gradient bounds are available for general F and G in [19], their results are based on linearization and thus require existence of classical solutions. For G ( p , x ) that is linear in p , regularity estimates on D u were recently obtained by Li and Zhang [18].

Theorem 2.4

[18,19] Let u be a viscosity solution of ( P ) with u M .

B r + { x < r } { x e n 0 } a n d Γ { x e n = 0 } B 1 .

Let u be a viscosity solution of

F ( D 2 u , x ) = 0 i n B 1 + ν D u = G ( D u , x ) o n Γ .

For F and G satisfying (F1)–(F3) and (G1)–(G3), suppose that either (A) F ( M , x ) is convex with respect to M, or (B) G ( p , x ) is linear with respect to p. Then for any 0 < α < 1 , we have

(15) u C 0 , α ( B 1 / 2 + ) , D u C 0 , α ( B 1 / 2 + ) C ,

where C depends on α and M as well as the constants given in (F1)–(F3) and (G1)–(G3).

Our proof extends in general to the cases where estimate (15) holds for some α > 0 .

Below we mention interior homogenization result from [7], which is a modified version of homogenization results such as in [11].

Theorem 2.5

(Theorem 2.14, [7]) Let K be a positive constant and let f : R n R be bounded and Hölder continuous. Given ν S n 1 , let u N : { K x ν 0 } R be the unique bounded viscosity solution of

F ( D 2 u N , N x ) = 0 i n { K x ν 0 } ; ν D u N = f ( x ) o n { x ν = 0 } , u = 1 o n { x ν = K } . ( P N )

Then for any δ > 0 , there exists N 0 depending only on K, the bound of u N , and the Hölder exponent of f, such that

(16) u N u ¯ δ i n { x K } f o r N N 0 ,

where u ¯ is the unique bounded viscosity solution of

F ¯ ( D 2 u ¯ ) = 0 i n { K x ν 0 } ; ν D u ¯ = f ( x ) o n { x ν = 0 } , u = 1 o n { x ν = K } .

Next we state some consequences of ergodic property of irrational numbers in R mod Z . First, we state a version of Dirichlet’s approximation theorem, whose proof is based on the pigeon-hole principle.

Lemma 2.6

[Lemma 2.11 in [13]] For α 1 ,…, α n R and N N , there are integers p 1 , , p n , q Z with 1 q N such that

q α i p i N 1 / n .

Finally, we present a lemma that states ergodic property of hyperplanes with irrational normals in R n mod Z n .

Lemma 2.7

[Lemma 2.7 in [8], Lemma 2.3 in [12]]. For ν S n 1 and x 0 R n , let H ( x 0 ) { x R n : ( x x 0 ) ν = 0 } . Then the following holds:

  1. Suppose that ν is a rational direction. Then for any x H ( x 0 ) , there is y H ( x 0 ) , such that

    x y T ν ; y = x 0 mod Z n ,

    where T ν is the smallest positive number such that T ν ν Z n .

  2. Suppose that ν is an irrational direction, and let ω ν : N R + be defined as in (2.2) of [12]. Then there exists a dimensional constant C = C ( n ) > 0 such that the following is true: for any x H ( x 0 ) and N N , there is y R n such that

    x y C ( n ) N ; y = x 0 mod ε Z n

    and

    dist ( y , H d ) < ω ν ( N ) .

    We recall that ω ν ( N ) converges to 0 as N .

  3. If ν is an irrational direction, then for any z R n and δ > 0 , there is w H ( x 0 ) such that

    z w δ mod Z n .

3 Localization lemmas

In this section, we prove several lemmas on perturbing and localizing the solutions, which will be used frequently throughout the article. Below we prove a localization lemma, and as a corollary, we prove existence and uniqueness of solution u ε of ( P ) ε , ν , τ , q with Π = Π ( ν , τ ) for τ R n and ν S n 1 . Denote B R ( τ ) { x τ R } and recall H s { ( x τ ) ν = s } .

First, we state a basic lemma, which will be frequently used. The proof is a direct consequence of the oblicity assumption ( G 3 ) .

Lemma 3.1

There exists M = M ( q , c ) , such that q x ± M x ν are, respectively, super and subsolution of ( P ) ε , ν , τ , q .

Lemma 3.2

Let f C ( R n ) be bounded. Suppose w 1 and w 2 solve, in the viscosity sense,

  1. F D 2 w i , x ε = 0 in Σ R Π B R ( 0 ) for i = 1 , 2

  2. ν D w i = G D w i , x ε on H 0 for i = 1 , 2

  3. w 1 = w 2 on H 1

  4. 0 w 2 w 1 M on Π B R ( 0 ) .

Let L G p and 0 < c < 1 is the constant given in (G3). Then there exists a constant C Λ λ , c , L > 0 , such that

w 1 w 2 w 1 + C M ( 1 c ) R i n Π B 1 ( 0 ) .

Proof

Without loss of generality, let us set ν = e n and τ = 0 . The first inequality, w 1 w 2 , directly follows from Theorem 2.2. To show the second inequality, let

w w 1 + M ( h 1 + h 2 ) + C 1 h 3 ,

where

h 1 ( x ) = x 2 R 2 , h 2 ( x ) = C R 2 ( 1 ( x n ) 2 ) with C = n Λ λ , h 3 ( x ) = 1 + x n R ,

and C 1 > 0 is a large constant depending on n , Λ , λ , L , and c , which will be chosen below in the proof.

Note that in Σ R ,

F D 2 w , x ε = F D 2 w 1 + M ( D 2 h 1 + D 2 h 2 ) , x ε F D 2 w 1 , x ε P + ( M ( D 2 h 1 + D 2 h 2 ) ) = F D 2 w 1 , x ε = 0 .

Also w 2 = w 1 w on H 1 and w 2 w 1 + M w on B R ( 0 ) Π .

Hence, to show that w 2 w , it is enough to show that x n w G D w , x ε on H 0 . We will verify that this is true when C 1 is sufficiently large. Observe that in Σ R ,

(17) D ( h 1 + h 2 ) C 0 R for C 0 = C 0 ( n , Λ , λ ) .

Hence, on H 0 Σ R , we have

x n w x n w 1 + C 1 R C 0 R = G D w 1 , x ε + ( C 1 C 0 ) R G D w , x ε c C 1 R + C 0 L R + ( C 1 C 0 ) R ,

where the last inequality follows from the Lipschitz property of G with (17), if C 1 = C 1 ( n , Λ , λ , c ) is chosen sufficiently large. It follows from Theorem 2.2 that w 2 w in Σ R , and we obtain the lemma.□

As a corollary of Lemma 3.2, we prove existence and uniqueness of solutions in strip regions.

Lemma 3.3

There exists a unique solution u ε of ( P ) ε , ν , τ , q with the property u ε ( x ) q x L ( Π ) < , such that

u ε q x M .

Proof

  1. Let Σ R be as given in Lemma 3.2, and consider the viscosity solution w R ( x ) of ( P ) ε , ν , τ , q in Σ R with the lateral boundary data q x on B R ( τ ) Π . The existence and uniqueness of the viscosity solution w R is shown, for example, in [9, 16].

    From Lemma 3.1, q x ± M ( x τ + ν ) ν is a sub- and supersolution of ( P ) ε , ν , τ , q , and thus, by comparison principle, we obtain that

    w R ( x ) q x M for x Σ R .

    Due to Theorem 2.5 and the Arzela-Ascoli Theorem, w R locally uniformly converges to a continuous function u ε ( x ) . From the stability property of viscosity solutions, it follows that u ε ( x ) is a viscosity solution of ( P ) ε , ν , τ , q .

  2. To show uniqueness, suppose both u 1 and u 2 are viscosity solutions of ( P ) ε , ν , τ , q with u 1 q x , u 2 q x M . Then Lemma 3.2 yields that, for any point s H 0 ,

    u 1 u 2 O ( 1 / R ) in B 1 ( s ) Π .

    Hence, u 1 = u 2 .□

The following is immediate from Theorem 2.2 and the construction of u ε in the aforementioned lemma.

Corollary 3.4

Suppose u,v are bounded and continuous functions in Π ¯ = Π ( τ , ν ) ¯ . In addition, suppose they satisfy, for F satisfying (F1)–(F3) and G satisfying (G1)–(G2),

  1. F D 2 u , x ε 0 F D 2 v , x ε i n Π ;

  2. u v o n H 1 ;

  3. ν D u G ( D u , x / ε ) ; ν D v G ( D v , x / ε ) on H 0 .

Then u v in Π .

Lemma 3.5

There exists C > 0 such that the following holds: let u i for i = 1 , 2 solve

F ( D 2 u i ) = 0 i n Π B R ( 0 ) ν u i = G i ( D u i , x ) o n H 0 B R ( 0 ) u i = q x o n H 1 B R ( 0 ) ,

where Π = Π ( ν , 0 ) . Furthermore, suppose that G i satisfies the assumption in Theorem 2.4 and G 1 and G 2 satisfy

(18) G 1 ( p , x ) G 2 ( p , x ) δ ( 1 + p ) a n d u 1 u 2 M .

Let L denote the Lipschitz bound for u i and G s . Then there exists C = C ( Λ , λ , n ) , such that

u 1 u 2 δ ( L + 1 ) + C M / R i n Π B 1 ( 0 ) .

Proof

By our assumption, v ( u 1 u 2 ) / M satisfies v 1 in B R ( 0 ) with

P + ( D 2 v ) 0 in Π B R ( 0 ) v = 0 on H 1 B R ( 0 ) .

After a change of coordinates, we may assume ν = e n so that Π = { x : 1 x n 0 } , and we denote x = ( x , x n ) . Define

w ( x ) ( c 0 / M + c 1 / R ) ( x n + 1 ) + 2 x 2 Λ n λ ( x n 2 1 ) / R 2 ,

where c 0 and c 1 > 8 will be chosen later. Then w is a supersolution of the aforementioned problem with the Neumann boundary condition:

n w = ( c 0 / M + c 1 / R ) ( c 0 / M + 4 x / R 2 ) = ( c 0 / M + D T w ) on { x n = 0 } B R ( 0 ) .

Now suppose v w has positive maximum in Π B R ¯ ( 0 ) . Then the maximum would need to be achieved at a point τ H 0 B R ( 0 ) . At this point, we should have n ( v w ) 0 and D T v = D T w . Therefore,

(19) n v n w ( c 0 / M + D T w ) = ( c 0 / M + D T v ) . at x = τ ,

On the other hand,

G 1 ( D u 1 , x ) G 2 ( D u 2 , x ) = D G 1 ( p , x ) D ( M v ) + G 1 ( D u 2 , x ) G 2 ( D u 2 , x ) ,

and since D G 1 ( p , x ) e n c , we have, from (18) and the Lipschitz bound for u i given in Theorem 2.4,

( 1 c ) n v L D T v + 1 M G 1 ( D u 2 , x ) G 2 ( D u 2 , x ) L D T v + δ M ( L + 1 ) at x = τ .

Then using the fact that D T w = 4 x / R 2 4 / R in B R ( 0 ) , it follows that

(20) ( 1 c ) n v 4 L R + δ ( L + 1 ) M .

Hence, from (19), we obtain a contradiction if c 0 / M + c 1 / R is larger than the right-hand side of (31). This happens if we choose c 1 > 4 L and c 0 = δ ( L + 1 ) . Therefore, it follows that v w in Π B R ¯ . We can now conclude that

u 1 u 2 = M v c 0 + c 1 M / R + 2 M 1 + Λ n λ / R 2 in Π B 1 ( 0 ) .

The lower bound can be obtained with the aforementioned argument applied to u 2 u 1 .□

4 Homogenization in a strip domain

Let u ε solve ( P ) ε , ν , τ , q with linear boundary data l ( x ) on H 1 . We let v ε be the unique linear function on Π such that v ε coincides with u ε on H 1 and at a reference point τ ν / 2 . More precisely,

(21) v ε ( x ) = μ ε ( ( x τ ) ν + 1 ) + l ( x ) ,

where μ ε = 2 ( u ε ( τ ν / 2 ) u ε ( τ ν ) ) . Then we define the average slope μ ( u ε ) of u as follows:

(22) μ ( u ε ) ν v ε = μ ε .

Theorem 4.1

The followings hold for u ε solving ( P ) ε , ν , τ , q :

  1. For irrational directions ν , there exists a unique constant μ = μ ( ν , q ) , such that u ε converges uniformly to the linear profile

    u ( x ) μ ( ( x τ ) ν + 1 ) + l ( x ) ,

    where l ( x ) q x . The same holds for rational directions ν with τ = 0 .

  2. [Error estimate] There exists a constant C > 0 depending on λ , Λ , n , and the slope of l ( x ) such that the following holds: if ν is an irrational direction or ν is a rational direction with τ = 0 , then

    μ ( u ε ) μ C Λ ( ε , ν ) i n Π ,

    where

    (23) Λ ( ε , ν ) = inf 0 < k < 1 { ε k T ν + ε 1 k } i f ν i s a r a t i o n a l d i r e c t i o n inf 0 < k < 1 , N N { ε k N + ω ν ( N ) + ε 1 k } i f ν i s a n i r r a t i o n a l d i r e c t i o n .

    In (23), T ν and ω ν are as given in Lemma 2.7. T ν is the period of G ( P , y ) on the Neumann boundary H 0 and ω ν ( N ) 0 as N .

  3. Let ν be an irrational direction. For any δ > 0 , there exist T > 0 and P Z n such that

    T ν P δ .

    Let θ = θ ( δ , ν ) be the angle between ν and P, then

    Λ ( ε , ν ) 3 δ f o r ε < δ 2 θ .

  4. Let ν be a rational direction, and let δ > 0 . Then

    Λ ( ε , ν ) 2 δ f o r ε < δ 2 T ν .

To prove Theorem 4.1 we begin with a preliminary lemma. The following lemma states that u ε looks like a linear profile (almost flat) on each hyperplane normal to ν .

Lemma 4.2

Away from the Neumann boundary H 0 and u ε l ( x ) is almost a constant on hyperplanes parallel to H 0 . More precisely, for x 0 Π , we denote

d dist ( x 0 , H 0 ) > 0

and H d { ( x τ ) ν = d } = { ( x x 0 ) ν = 0 } . Then the following holds:

  1. If ν is a rational direction, there exists a constant C > 0 depending on α , λ , Λ , n , and the slope of l, such that for any x H d ,

    (24) ( u ε ( x ) l ( x ) ) ( u ε ( x 0 ) l ( x 0 ) ) C ( d 1 + 1 ) ( T ν ε ) ,

    where T ν is a constant depending on ν , given as in (a) of Lemma 2.7.

  2. If ν is an irrational direction, there exists a constant C > 0 depending on α , λ , Λ , n , and the slope of l, such that for any x H d ,

    (25) ( u ε ( x ) l ( x ) ) ( u ε ( x 0 ) l ( x 0 ) ) C ( d 1 ε ω ν ( N ) + ω ν ( N ) ) ,

    for any N N and ε > 0 with ε ω ν ( N ) < 1 , where ω ν ( N ) is given as in Lemma 2.7.

Proof

First, we consider a rational direction ν . By (a) of Lemma 2.7, for any x H d , there is y H d such that x y T ν ε and y = x 0 mod ε Z n . Then by comparison,

(26) u ε ( x ) = u ε ( x + ( y x 0 ) ) l ( y ) + l ( x 0 ) .

Hence, u ε ( x 0 ) = u ε ( y ) l ( y ) + l ( x 0 ) , and we obtain

( u ε ( x ) l ( x ) ) ( u ε ( x 0 ) l ( x 0 ) ) u ε ( x ) u ε ( y ) + l ( y ) l ( x ) u ε ( x ) u ε ( y ) + C T ν ε C d 1 T ν ε + C T ν ε ,

where the third inequality follows from Theorem 2.3.

Next, we consider an irrational direction ν and let x H d . By (b) of Lemma 2.7, for any N N , there exists y R n such that x y ε ω ν ( N ) , y = x 0 mod ε Z n and

(27) dist ( y , H d ) < ε ω ν ( N ) .

Observe that

( u ε ( x ) l ( x ) ) ( u ε ( x 0 ) l ( x 0 ) ) u ε ( x ) u ε ( y ) + ( u ε ( y ) l ( y ) ) ( u ε ( x 0 ) l ( x 0 ) ) + l ( y ) l ( x ) ,

where, from Theorem 2.3,

u ε ( x ) u ε ( y ) C d 1 ε ω ν ( N ) .

Next we project y to x 1 H d and use Lemma 3.5 for G 1 = G and G 2 ( p , x ) = G ( p , x + ( x 0 x 1 ) ) = G ( p , x + ( y x 1 ) ) with δ = ω ν ( N ) to conclude that

( u ε ( x 0 ) l ( x 0 ) ) ( u ε ( x 1 ) l ( x 1 ) ) C ω ν ( N ) .

Then by using Theorem 2.3 with (27) once again, we compare u ( y ) with u ( x 1 ) and conclude that

( u ε ( y ) l ( y ) ) ( u ε ( x 0 ) l ( x 0 ) ) C ( ω ν ( N ) + ε ) .

Finally,

l ( y ) l ( x ) C y x C ε ω ν ( N ) C d 1 ε ω ν ( N ) ,

where the last inequality follows since y x ε ω ν ( N ) and d 1 .□

Since u ε is flat on each hyperplanes located, a constant d -away from the Neumann boundary, u ε can be approximated well by a linear solution as in the following corollary. The proof of Corollary 4.3 follows from the comparison principle (Theorem 2.2) and Lemma 4.2 with d = ε 1 k .

Corollary 4.3

For a solution u ε of ( P ) ε , ν , τ , q , let v ε be the unique linear function given as in (21). Then there exists a constant C depending on λ , Λ , n, and the slope of l such that for any N N and 0 < k < 1 ,

u ε ( x ) v ε ( x ) C ( ε k T ν + ε 1 k ) i f ν i s a r a t i o n a l d i r e c t i o n C ( ε k N + ω ν ( N ) + ε 1 k ) i f ν i s a n i r r a t i o n a l d i r e c t i o n ,

and hence,

u ε ( x ) v ε ( x ) C Λ ( ε , ν ) .

Due to the uniform interior regularity of { u ε } (Theorem 2.3), along a subsequence, they locally uniformly converges to u in Π . Let us choose one of the convergent subsequence u ε j and denote it by u j , i.e., u j = u ε j . Let v j = v ε j and μ j = μ ( u ε j ) , both as given in (21) and (22). Corollary 4.3 implies that for any ν S n 1 , lim u j is linear. More precisely, the slope μ j converges as j (see Lemma 4.1 of [8]), and hence, by Corollary 4.3,

lim u j = lim v j = μ ( ( x τ ) ν + 1 ) + l ( x ) = u

for μ lim μ j .

Next, we prove that the subsequential limit is unique, i.e., μ does not depend on the subsequence { ε j } , when ν is irrational or ν is rational with τ = 0 . We will also obtain a mode of convergence of μ ε .

Proof of Theorem 4.1 (a) and (b) for irrational directions: Let ν be an irrational direction and let u be a subsequential limit of u ε . We claim that

u / ν = μ ( ν , q )

for a constant μ ( ν , q ) , which depends on ν and q , not on τ or the subsequence { ε j } . More precisely,

(28) μ ( u η ) μ ( u ε ) C ( Λ ( ε , ν ) + η ) .

where we let 0 < η < ε be sufficiently small.

For the proof of (28), let

w ε ( x ) = u ε ( ε x ) ε , w η ( x ) = u η ( η x ) η

and denote by H 1 and H 2 , the Neumann boundaries of w ε and w η , respectively. By (c) of Lemma 2.7, for τ R n , there exist s 1 H 1 and s 2 H 2 , such that

τ s 1 η mod Z n , and τ s 2 η mod Z n .

Hence, after translations by τ s 1 and τ s 2 , we may suppose that w ε ( x ) and w η ( x ) are defined on the extended strips

Ω ε x : 1 ε ( x τ ) ν 0 and Ω η x : 1 η ( x τ ) ν 0 ,

respectively, with

w ε = l ε ( x ) on ( x τ ) ν = 1 ε

and

w η = l η ( x ) on ( x τ ) ν = 1 η ,

where l ε and l η are linear functions with the same slope as l ( x ) . Moreover on H 0 , we have

w ε / ν = G ( D w ε , x z 1 ) and w η / ν = G ( D w η , x z 2 )

for some z 1 , z 2 η . Observe that by Lipschitz continuity of G , i.e., by (G2),

(29) G ( p , x z 1 ) G ( p , x z 2 ) < m ( 1 + p ) η .

Let v ε be given in (21). Then by Corollary 4.3 (after a translation),

(30) w ε ( x ) v ε ( ε x ) ε C Λ ( ε , ν ) ε .

Note that

v ε ( ε x ) ε = μ ε ( x τ ) ν + 1 ε + l ε ( x ) .

From (30) and the comparison principle, it follows that

(31) ( μ ε C Λ ( ε , ν ) ) ( x τ ) ν + 1 ε w ε ( x ) l ε ( x ) ( μ ε + C Λ ( ε , ν ) ) ( x τ ) ν + 1 ε .

Here, we denote by l 1 and l 2 , the following linear profiles

l 1 ( x ) = a 1 ( x τ ) ν + b 1 and l 2 ( x ) = a 2 ( x τ ) ν + b 2 ,

whose respective slopes are a 1 = μ ε + C Λ ( ε , ν ) and a 2 = μ ε C Λ ( ε , ν ) . b 1 and b 2 are chosen, so that

(32) l 1 ( x ) = l 2 ( x ) = w η ( x ) l η ( x ) = 0 on x : ( x τ ) ν = 1 η .

Now we define

w ¯ ( x ) l η ( x ) + l 1 ( x ) in { 1 / η ( x τ ) ν 1 / ε } w ε ( x ) l ε ( x ) + c 1 in { 1 / ε ( x τ ) ν 0 }

and

w ̲ ( x ) l η ( x ) + l 2 ( x ) in { 1 / η ( x τ ) ν 1 / ε } w ε ( x ) l ε ( x ) + c 2 in { 1 / ε ( x τ ) ν 0 } ,

where c 1 and c 2 are constants satisfying

l 1 = w ε l ε + c 1 = c 1 and l 2 = w ε l ε + c 2 = c 2

on { ( x τ ) ν = 1 / ε } . Note that by (32),

w ̲ = w ¯ = w η on x : ( x τ ) ν = 1 η ,

and also due to (31),

w ¯ ( x ) = l η ( x ) + min ( l 1 ( x ) , w ε ( x ) l ε ( x ) + c 1 )

and

w ̲ ( x ) = l η ( x ) + max ( l 2 ( x ) , w ε ( x ) l ε ( x ) + c 2 )

in 1 ε ( x τ ) ν 0 . Thus, it follows that w ¯ and w ̲ are, respectively, viscosity super- and subsolution of (P). Hence, we obtain

(33) w ̲ w ˜ η w ¯ ,

where w ˜ η is a solution of (P) in Ω η with w ˜ η = w η = l η ( x ) on { ( x τ ) ν = 1 / η } , and w ˜ η / ν = G ( D w ˜ η , x z 1 ) on H 0 . Then by (33) and Lemma 3.5 with (29),

μ η μ ε μ η μ ( w ˜ η ) + μ ( w ˜ η ) μ ε C ( Λ ( ε , ν ) + η ) ,

where μ ( w ˜ η ) is the slope of the linear approximation of w ˜ ε . The aforementioned inequality implies that the slope μ of a subsequential limit of u ε depends on neither the subsequence { ε j } nor τ . Also sending η 0 , we obtain an error estimate (d) when ν is irrational.

Proof of Theorem 4.1 (a) and (b) for rational directions: Let ν be a rational direction with τ = 0 . We claim that u / ν = μ ( ν , q ) for a constant μ ( ν , q ) , which depends on ν and q , not on the subsequence { ε j } . More precisely, if η ε , then

(34) μ ( u η ) μ ( u ε ) C Λ ( ε , ν ) .

The proof of (34) is parallel to that of (28). Let w ε and w η be as given in the proof of (28). Note that since Ω ε and Ω η have their Neumann boundaries passing through the origin, w ε / ν = G ( x ) = w η / ν without translation of the x variable, and thus, we do not need to use the properties of hyperplanes with an irrational normal (Lemma 2.7 (b)) to estimate the error between the shifted Neumann boundary datas. In other words, there exist q 1 H 1 and q 2 H 2 such that p = q 1 = q 2 mod Z n , and hence, G ( , x z 1 ) = G ( , x z 2 ) in the proof of (28). Following the proof of (28), we obtain an upper bound Λ ( ε , k ) of μ η μ ε . Note that we do not have the term η in (34) since G ( , x z 1 ) = G ( , x z 2 ) . By sending η 0 in (34), we obtain the error estimate (b) for rational directions with τ = 0 .

Proof of Theorem 4.1 (c) and (d): Let δ > 0 and let ν be an irrational direction. Lemma 2.6 implies that there is a positive number T ν ( δ ) δ ( n 1 ) such that T ν ( δ ) ν δ mod Z n . Then, for some P Z n and T = T ν ( δ ) + O ( δ ) ,

T ν P δ

and T ν P + P . Let θ = θ ( δ , ν ) > 0 be the angle between ν and P , then

(35) T ν P = T θ δ .

If we define q T ν P P , then q δ by (35). Then for 0 m 1 T θ , m T ν = m P + m q with 1 δ 1 T θ q 1 . Hence, we obtain

(36) ω ν ( N ) T θ when N = 1 θ .

Let ε ( δ , ν ) be a constant depending on δ and the direction ν such that

(37) ε ( δ , ν ) = δ 2 θ = δ 2 θ ( δ , ν ) .

Then for 0 < ε < ε ( δ , ν ) ,

Λ ( ε , ν ) = inf 0 < k < 1 , N N { ε k N + ω ν ( N ) + ε 1 k } inf 0 < k < 1 { ε k / θ + T θ + ε 1 k } inf 0 < k < 1 { ε k / θ + ε 1 k } + δ ,

where the first and last inequalities follow from (36) and (35), respectively. Then by (37),

inf 0 < k < 1 { ε k / θ + ε 1 k } inf 0 < k < 1 { ( δ 2 θ ) k / θ + ( δ 2 θ ) 1 k } .

The infimum is taken when 0 < k = ln ( θ δ ) / ln ( θ δ 2 ) < 1 and

inf 0 < k < 1 { ( δ 2 θ ) k / θ + ( δ 2 θ ) 1 k } = 2 δ .

Hence, we can conclude Λ ( ε , ν ) 3 δ for ε < ε ( δ , ν ) = δ 2 θ .

Next, we consider a rational direction ν . For δ > 0 , let ε < δ 2 / T ν . Then we can check

Λ ( ε , ν ) = inf 0 < k < 1 { ε k T ν + ε 1 k } inf 0 < k < 1 { δ 2 k T ν 1 k + δ 2 ( 1 k ) T ν k 1 } = 2 δ .

The following lemma will be used in the next section.

Lemma 4.4

Let ν = e n , τ = 0 , and let w solve

F ( D 2 w , x / ε ) = 0 i n { N ε x n 0 } ; w / x n = G ( D w , x / ε ) o n H 0 ; w = A o n H N ε ,

where N and A are constants. Then there is a constant C = C ( λ , Λ , n ) such that

w ( x ) w ( x 0 ) C ε f o r x , x 0 H s , N ε s N ε 2 .

Proof

For x 0 , x H s with s [ N ε , N ε 2 ] , choose y H s such that x y ε and y = x 0 mod ε Z n . Observe that w ( y ) = w ( x 0 ) , since G is 1-periodic on H 0 . Therefore,

w ( x ) w ( x 0 ) = w ( x ) w ( y ) C w A L x y N ε C ε ,

where the second inequality is from the interior Lipschitz regularity (Theorem 2.3) applied to w ( N ε x ) A .□

5 Continuity over normal directions

In the previous section, we have shown that for an irrational direction ν S n 1 R Z n , there is a unique homogenized slope μ ( ν , q ) for any solution u ε ν of ( P ) ε , ν , τ , q in Π ( ν , τ ) . In this section, we investigate the continuity properties of μ with respect to ν and q , as well as the mode of convergence for u ε ν as the normal direction ν of the domain varies.

We first show that μ is Lipschitz with respect to q , which directly follows from the 1-homogeneity of G .

Theorem 5.1

For ν S n 1 R Z n , μ ( ν , q ) is uniformly Lipschitz in q ν , independent of ν .

Proof

For q 1 , q 2 ν , let u ε i be the unique bounded solution of ( P ) ε , ν , τ , q i for i = 1 , 2 . Let m be the Lipschitz constant for G given in (G1) and c be as given in (G3). Then it follows that

w ± ( x ) u ε 1 ( x ) + ( q 2 q 1 ) x ± m 1 c q 1 q 2 ( x ν )

is, respectively, a super and subsolution of ( P ) ε , ν , τ , q 2 . Hence, by Corollary 3.4, we have

w u ε 2 w + in Π .

From here and Theorem 4.1, it follows that

μ ( ν , q 1 ) μ ( ν , q 2 ) m 1 c q 1 q 2 .

The dependence of μ on ν is a much more subtle matter due to the change of the domain and the resulting changes in boundary conditions on the Neumann boundary. From now on, we work with a fixed choice of q and denote μ = μ ( ν ) .

For s 0 , let T ν ( s ) be the smallest positive number 1 such that

T ν ( s ) ν s mod Z n .

Note T ν ( 0 ) is larger than all T ν ( s ) . In general, Lemma 2.6 yields

(38) T ν ( s ) n s ( n 1 ) .

Theorem 5.2

With fixed q, let us denote μ = μ ( , q ) : ( S n 1 R Z n ) R be as given in Theorem 4.1. Then μ has a continuous extension μ ¯ ( ν ) : S n 1 R . More precisely, let us fix a direction ν S n 1 and a constant δ > 0 . If ν 1 and ν 2 are irrational directions such that

(39) tan θ i < δ 5 / 2 T ν ( δ 5 / 2 ) f o r θ i ν i ν a n d i = 1 , 2 ,

then we have

  1. μ ( ν 1 ) μ ( ν 2 ) < C δ 1 / 2 for C = C ( ν ) .

  2. μ ¯ ( ν ) is Hölder continuous on S n 1 with a Hölder exponent of 1 5 n .

Remark 5.3

In the proof, we indeed show that, for any directions ν 1 and ν 2 satisfying (39), the range of { μ ( u ε ν i ) } ε , i fluctuates only by δ , if ε is sufficiently small. The fact that ν i ’s are irrational is only used to guarantee that there is only one subsequential limit for μ ( u ε ν i ) .

Remark 5.4

For notational simplicity and clarity in the proof, we will assume that n = 2 and ν = e 2 . We explain in Remark 5.6 how to modify the notations and proof for ν e 2 . For general dimension n , we refer to Remark 5.7.

For the rest of the article, we prove (a) of Theorem 5.2. Theorem 5.2 (b) follows from (38), (39), and Theorem 5.2 (a).

5.1 Basic settings and Sketch of the proof

We denote

Π Π ( e 2 , 0 ) and Π ν i Π ( ν i , 0 ) , for i = 1 , 2 .

We also denote

H 0 = H 0 ( e 2 ) , H 0 ν i H 0 ( ν i ) for i = 1 , 2 .

For given

m N and δ 1 / m > 0 ,

we divide the unit strip R × [ 0 , 1 ] by m numbers of small horizontal strips of width δ and define a family of functions { G k } k so that the value of G k at ( x 1 , x 2 ) is same as the value of G at ( x 1 , x ˜ 2 ) , where ( x 1 , x ˜ 2 ) is the projection of ( x 1 , x 2 ) onto the bottom of the k-th strip. More precisely, we define

(40) G k ( x 1 , x 2 ) G ( x 1 , δ ( k 1 ) ) for k = 1 , , m .

Then G k is a 1-periodic function with respect to x 1 .

Next we introduce the parameters

(41) θ 1 ν 1 e 2 , θ 2 ν 2 e 2

and

(42) N δ tan θ 1 , M δ tan θ 2 .

Without loss of generality, assume θ 2 θ 1 , and thus, N M .

If θ i ’s are sufficiently small, then we will be able to approximate G on both of the Neumann boundary H 0 ν 1 and H 0 ν 2 using the universal boundary data G k ’s, which depends only on δ , but not on the direction ν 1 nor ν 2 . In particular, in meso-scopic scale G can be approximated by many repeating pieces of G k ’s on H 0 ν i (approximately, N number of pieces of G k for ν 1 and M for ν 2 ). Thus, the problem already experiences averaging phenomena: we call this as the first or near-boundary homogenization. Note that in this step, the only difference in the averaging phenomena between the two directions ν 1 and ν 2 , besides the errors in terms of G and G k on H 0 ν i , is the number of repeating data G k for each k . This explains the proximity of μ ( ν 1 ) and μ ( ν 2 ) .

On the other hand, since ν s are irrational directions, the distribution of G k approximates the given G on H 0 ν i in large scale. Since ν 1 and ν 2 are close to the rational direction e 2 , the averaging behavior of a solution u ε ν i in Π ν i would appear in a very large scale, and in other words, only after ε obtains very small. We call this as the secondary homogenization.

The two-scale homogenization procedure has been introduced in [7,8]. It allows studying continuity properties of the homogenized boundary data as we approach the rational direction, which might be singular points as described in Section 1. This point of view was also employed in [12,13] to study homogenization for general operators, by studying the singularity of homogenized operator at rational directions. Let us also point out near the boundary the small-scale oscillation of the operator interacts with that of boundary data to create a meso-scale averaging phenomena. Due to this interaction, characterizing the homogenized boundary condition remains a challenging and interesting open problem. After the first homogenization, the boundary data change to periodic data in a meso-scale (which will be N ε below), and hence, the operator is well approximated by the homogenized operator F ¯ in the second homogenization in large scale.

Below we begin the analysis of the two-step homogenization as described earlier. We will work with small ε > 0 satisfying

(43) ε δ tan θ i T ν ( δ 5 / 2 ) for i = 1 , 2 ,

which can be stated as follows:

(44) 0 < ε δ tan θ i for i = 1 , 2

since T ν ( s ) 1 when ν = e 2 . It follows that

(45) m N ε m M ε δ .

After the near-boundary homogenization, u ε ν 1 will be approximated by a solution, which has periodic boundary data with period m N ε . With (45), it follows that u ε ν 1 fluctuates in order of δ in the interior of the strip domain.

On the other hand, (39) of Theorem 5.2 can be stated as follows:

(46) 0 < tan θ 1 δ 5 / 2 .

It follows then that

(47) 1 / N δ 3 / 2

which ensures u ε ν i to homogenize N ε -close to the Neumann boundary.

Next, we define vertical strips I k ¯ ’s so that in each I k ¯ , the Neumann boundary H 0 ν 1 is contained in the horizontal strip (parallel to H 0 )of width approximately δ ε . Let N 0 = 0 and

N k max N N j = 0 k 1 N j + N ε tan θ 1 < k δ ε for k N .

We define

I k ¯ = j = 0 k 1 N j ε , j = 0 k N j ε × R for k N j = 0 k + 1 N j ε , j = 0 k N j ε × R for k N { 0 } .

Then we can observe

(48) δ tan θ 1 1 N k δ tan θ 1 + 1

since the definition of N k implies

( N k 1 ) ε tan θ 1 δ ε ( N k + 1 ) ε tan θ 1 .

On the other hand, by the definitions of N k and I k ¯ , H 0 ν 1 I k ¯ is located within δ ε -distance from H 0 + δ ε ( k 1 ) e 2 , mod ε Z n , for each k Z . Thus, G is approximated well by G k on H 0 ν 1 I k ¯ , for 1 k m . Indeed, if we extend the definition of G k over k Z by letting G k = G k ¯ for k = k ¯ (mod m ), then we have

(49) G p , x ε G k p , x ε < C ( 1 + p ) δ on H 0 ν 1 I k ¯ for k Z .

Similarly for ν 2 , we define M k for k N { 0 } and the vertical strips J k ¯ for k Z .

Remark 5.5

Observe that (48) implies N k and M k are comparable, respectively, with N = δ tan θ 1 and M = δ tan θ 2 with N k N , M k M 1 . Thus, for simplicity of our proof, we assume

N k = N ; M k = M for k N

and

(50) I k = [ ( k 1 ) N ε , k N ε ] × R ; J k = [ ( k 1 ) M ε , k M ε ] × R for k Z .

Our simplification of N k does not affect our analysis in Section 5.2: For the first homogenization near the boundary, the estimate in Lemma 5.9 does not change since N k and N differ at most by 1, and the analysis is done in a local ball in the proof of Lemma 5.9. More precisely, Lemma 5.9 holds with μ N ( G k ) replaced by μ N k ( G k ) , where μ N ( G k ) μ N k ( G k ) is small enough by parallel arguments that show (75). For the second homogenization in the middle region, we can construct a periodic function Λ ( x ) with period ε / sin θ 1 similarly as in step 2 of Section 5.2, since there we view each m union of I ¯ k as a “block,” and

ε ( 1 tan θ 1 ) sin θ 1 k = 1 m I ¯ k H 0 ν 1 ε ( 1 + tan θ 1 ) sin θ 1

and also

ε ( L tan θ 1 ) sin θ 1 k = 1 L m I k ¯ H 0 ν 1 ε ( L + tan θ 1 ) sin θ 1

for any L N . This shows that the required period of Λ is ε / sin θ 1 , approximating the average period of k = 1 L m I ¯ k with the error ε tan θ 1 L sin θ 1 ε / L .

Remark 5.6

For ν e 2 in R 2 , there exists a rational direction ν ˜ such that for T = T ν ( δ 5 / 2 ) ,

T ν ˜ = 0 ( mod Z 2 ) ; ν ν ˜ δ 5 / 2 / T .

Observe that if Theorem 5.2 holds for the rational direction ν ˜ , it also holds for ν . For the proof of the theorem for ν ˜ , let x = x ( x ν ˜ ) ν ˜ and define

G k = G k ( x , x x ) = G ( x , δ ( k 1 ) ν ˜ ) for 1 k m .

Then G k is a periodic function on { x ν ˜ = 0 } with a period of T . The only difference between the case of ν ˜ and e 2 is in the periodicity of the function G k , and it does not make any essential difference in the proof. we point out that instead of the conditions (46), (47), and (45), we will need

1 T N δ 3 / 2 ; T tan θ 1 δ 5 / 2 ; m T M ε δ

since G k has a period of T . These conditions will be ensured if θ i and ε satisfy the assumptions as in Theorem 5.2.

Remark 5.7

For the dimension n > 2 and ν = e n , for a fixed m N and δ = 1 m , let us define

G i ( x 1 , , x n 1 , x n ) G ( x 1 , , x n 1 , δ ( i 1 ) ) for i = 0 , , m

and

I k 1 , k 2 , , k n 1 [ ( k 1 1 ) N ε , k 1 N ε ] × × [ ( k n 1 1 ) N ε , k n 1 N ε ] × R .

Then parallel arguments as in steps 1–9 in the next section would apply to yield the results in R n .

5.2 Proof of Theorem 5.2

In the first three steps, we follow the aforementioned heuristics and replace the Neumann condition with the locally projected boundary data G k . Then we go through the two-step homogenization procedures to obtain the first slope μ N ( G k ) on each I k near the boundary, and then the global slope μ ( ν 1 ) . While the actual first homogenization takes place in Π ν 1 , it turns out that its value has a small difference from μ N ( G k ) taken in Π (see Lemma 5.9). This fact is important in establishing a universal domain for both directions ν 1 and ν 2 . In fact, we rotate the middle and inner regions to compare the slopes in Π ν 1 and Π ν 2 . For this, we use the rotational invariance of the homogenized operator F ¯ . (See Lemmas 5.10 and 5.11.) The rest of steps are to verify that indeed μ ( ν 1 ) is the correct averaged slope for the problem ( P ) ε , ν 1 , τ , q .

Step 1. First homogenization near Boundary ( N ε -away from H 0 ν 1 )

We proceed to discuss the first homogenization. Denote x = ( x 1 , x 2 ) throughout this section. For a given linear function l ( x ) = l ( x 1 ) and k Z , let u = u N , ε and v k = v k N , ε solve the following problem with u = l ( x ) on H N ε ν 1 and v k = l ( x ) on H N ε :

(51) F ( D 2 u , x / ε ) = 0 in { N ε x ν 1 0 } ; u ν 1 ( x ) = G ( D u , x / ε ) on H 0 ν 1

and

(52) F ( D 2 v k , x / ε ) = 0 in { N ε x 2 0 } ; v k x 2 ( x ) = G k ( D v k , x / ε ) on H 0 .

Definition 5.8

For a given function u : { N ε x ν 0 } R and I k given as in (50), let a k and b k be the middle points of I k H N ε / 2 ν and I k H N ε ν , respectively, and consider the unique linear function h given by h = u at x = a k , b k and D T h ( b k ) = D T u ( b k ) . (Here, D T h denotes the tangential derivative of h along the direction ν .) Then μ k ( u ) is defined by

μ k ( u ) h / ν .

Note that the Neumann boundary data of v k are G k on each boundary piece H 0 I i ( i Z ), and hence, μ i ( v k ) = μ ( v k ) . (Here, μ ( v k ) is the average slope of v k given as in (22) with τ = ( N ε / 2 ) e 1 .) For N as given in (42), we denote

(53) μ N ( G k ) μ ( v k ) .

Lemma 5.9

For k Z and μ k ( u ) as given in Definition 5.8,

(54) μ k ( u ) μ N ( G k ) < C δ 1 / 2 .

Proof

We will prove the lemma for k = 1 , i.e., we will compare μ 1 ( u ) with μ ( v 1 ) . Let u ˜ and v ˜ 1 solve the following problem with u ˜ = l on H ε / δ ν 1 and v ˜ 1 = l on H ε / δ :

F ( D 2 u ˜ , x / ε ) = 0 in { ε / δ x ν 1 0 } ; u ˜ ν 1 ( x ) = G ( D u ˜ , x / ε ) on H 0 ν 1

and

F ( D 2 v ˜ 1 , x / ε ) = 0 in { ε / δ x 2 0 } v ˜ 1 x 2 ( x ) = G 1 ( D v ˜ 1 , x / ε ) on H 0 .

We will compare both of u ˜ ( x ) and v ˜ 1 ( x ) to w 1 ( x ) in the ball x δ 1 α 0 ε , where α 0 = 1 / 2 . For computational convenience, we will call this number as α 0 . Let w 1 ( x ) solve w 1 = l on H ε / δ ν 1 with

(55) F ( D 2 w 1 , x / ε ) = 0 in { ε / δ x ν 1 0 } ; w 1 ν 1 ( x ) = G 1 ( D w 1 , x / ε ) on H 0 ν 1 .

Here, observe that in the ball x δ 1 α 0 ε , the hyperplanes H 0 ν 1 and H 0 only differ by tan θ 1 δ 1 α 0 ε .

Below we derive some properties of w 1 . Consider

w ¯ ( x ) ε 1 w 1 ( ε x ) .

Then by Theorem 2.4, w ¯ is C 1 , 1 regular up to the Neumann boundary in a unit ball, if w ¯ has a bounded oscillation in the ball x 1 / δ . Observe that ( ε / δ ) 1 w 1 ( ε x / δ ) is defined in the strip { 1 x ν 1 0 } , and it has a periodic Neumann data G 1 ( , , x / δ ) with period δ . Since it has a periodic boundary data, it corresponds to the case of rational direction with Neumann boundary passing through the origin. Hence, we can use the error estimate Theorem 4.1 (b) for the rational direction passing through the origin, with T ν = 1 . Then we obtain

(56) ε δ 1 w 1 ε x δ h ( x ) inf 0 < k < 1 C ( δ k + δ 1 k ) = C δ 1 / 2 ,

where h is a linear solution approximating ( ε / δ ) 1 w 1 ( ε x / δ ) . Then by (56),

(57) w 1 ε x δ ε δ h ( x ) C δ 1 / 2 ε ,

and hence, the oscillation of w ¯ becomes less than C δ 1 / 2 in the ball x 1 / δ . Later in the proof, we will use C 1 , 1 regularity of w ¯ as well as the linear approximation (57) of w 1 .

First, we compare u ˜ to w 1 in B δ 1 α 0 ε ( 0 ) . For this, we compare the boundary data of u ˜ , that is G , to G 1 . Observe that if x H 0 ν 1 B δ 1 α 0 ε ( 0 ) , then x I k for some k δ 1 α 0 / N = δ 2 α 0 tan θ 1 . Hence, for x H 0 ν 1 B δ 1 α 0 ε ( 0 ) (i.e., for x H 0 ν 1 I k with k δ 2 α 0 tan θ 1 ),

(58) G ( p , x / ε ) G 1 ( p , x / ε ) G ( p , x / ε ) G k ( p , x / ε ) + G k ( p , x / ε ) G 1 ( p , x / ε ) C [ ( 1 + p ) δ + ( 1 + p ) k 1 δ ] C ( 1 + p ) ( δ + ( tan θ 1 δ ( 1 α 0 ) ) ) C ( 1 + p ) ( δ + δ ( 3 / 2 α 0 ) ) C ( 1 + p ) δ ,

where the second inequality follows from (49) and the construction of G k , third inequality follows from k δ 2 α 0 tan θ 1 , the fourth inequality follows from (46), and the last inequality follows since α 0 1 / 2 .

Note that u ˜ w 1 C ε δ in x 2 δ 1 α 0 ε . This implies that, by Lemma 3.5, u ˜ w 1 δ ( L + 1 ) + C δ α 0 ε δ in x δ 1 α 0 ε . Now we can compare u ˜ w 1 with linear profiles in the strip to obtain

(59) u ˜ ( x ) w 1 ( x ) C ( δ + δ α 0 ) x ν 1 + ε δ C δ α 0 x ν 1 + ε δ in x δ 1 α 0 ε .

Observe that (57) and (59) yield

u ˜ ( x ) L 1 ( x ) C ( δ α 0 + δ 1 / 2 ) x ν 1 + ε δ C δ α 0 x ν 1 + ε δ in x δ 1 α 0 ε ,

where L 1 ( x ) = l ( x ) + μ ( w 1 ) x ν 1 + ε δ , and μ ( w 1 ) is the average slope of w 1 . In other words, we obtain

(60) μ 1 ( u ˜ ) μ ( w 1 ) C δ α 0 .

Next, we compare v ˜ 1 and w 1 and prove that

μ ( v ˜ 1 ) μ ( w 1 ) C δ α 0 .

Recall that the oscillation of w ¯ is less than C δ 1 / 2 in the ball x 1 / δ (see (57)). If we consider w ˜ = δ 1 / 2 w ¯ , then this function solves the boundary condition:

w ˜ / ν = G ˜ ( D w ˜ , x ) = δ 1 / 2 G 1 ( δ 1 / 2 D w ˜ , x ) ,

which satisfies the assumptions for the C 1 , 1 regularity theory, Theorem 2.4. Thus, we have

w ¯ C 1 , 1 ( B 1 ) O ( δ 1 / 2 ) .

For x in the σ ε -neighborhood of H 0 ν 1 , choose x ˜ to be the closest point to x on H 0 . Then by (G1) and (G2) with the C 1 , 1 regularity of w ¯ given earlier, w 1 satisfies on H 0 ,

G D w 1 ( x ) , x ε G D w 1 ( x ˜ ) , x ˜ ε O ( δ 1 / 2 σ ) ( 1 + D w 1 ( x ) ) .

Recall that the Neumann boundaries of w 1 and v 1 ( H 0 ν 1 and H 0 ) only differ in the ball x δ 1 α 0 ε , by tan θ 1 δ 1 α 0 ε δ 3 / 2 α 0 ε (see (46)). So putting σ = δ 3 / 2 α 0 ,

G D w 1 ( x ) , x ε G D w 1 ( x ˜ ) , x ˜ ε O ( δ 1 α 0 ) ( 1 + D w 1 ( x ) ) on H 0 ,

and Lemma 3.5 yields that in x δ 1 α 0 ε ,

( v ˜ 1 w 1 ) ( x ) C ( δ 1 α 0 + δ α 0 ) x n + ε δ C δ α 0 x n + ε δ .

This and (57) yield that in x δ 1 α 0 ε ,

v ˜ 1 ( x ) L ( x ) C ( δ α 0 + δ 1 / 2 ) x n + ε δ C δ α 0 x n + ε δ ,

where L ( x ) = l ( x ) + μ ( w 1 ) x 2 + ε δ . In other words, we obtain

(61) μ ( w 1 ) μ ( v ˜ 1 ) C δ α 0 .

Recalling α 0 = 1 / 2 , we conclude from (60) and (61) that

(62) μ 1 ( u ˜ ) μ ( v ˜ 1 ) C δ 1 / 2 .

In the rest of proof, we will show

μ ( v 1 ) μ ( v ˜ 1 ) , μ 1 ( u ) μ 1 ( u ˜ ) C δ 1 / 2 .

Then the aforementioned inequalities and (62) would imply

μ 1 ( u ) μ ( v 1 ) μ 1 ( u ) μ 1 ( u ˜ ) + μ 1 ( u ˜ ) μ ( v ˜ 1 ) + μ ( v ˜ 1 ) μ ( v 1 ) C δ 1 / 2 .

First, observe that v 1 and v ˜ 1 have periodic Neumann data G 1 on H 0 . Hence, by similar arguments as in the proof of (28),

(63) μ ( v 1 ) μ ( v ˜ 1 ) C ( Λ ( δ , e 2 ) + N 1 ) C ( δ 1 / 2 + N 1 ) C δ 1 / 2 ,

where the last inequality follows from (47).

Next, recall that

μ 1 ( u ˜ ) μ ( w 1 ) C δ 1 / 2

for a solution w 1 of (55). (See (60).) Similarly, one can prove that

μ 1 ( u ) μ ( w ˜ 1 ) C N 1 / 2 C δ 1 / 2 ,

where w ˜ 1 solves similar equations as in (55) in the domain { N ε x ν 1 0 } , and the last inequality follows from (47). Then since w 1 and w ˜ 1 have periodic Neumann data G 1 on H 0 ν 1 , it corresponds to the case of ν = e 2 . Hence, by similar arguments as in (63),

μ ( w 1 ) μ ( w ˜ 1 ) C ( Λ ( δ , e 2 ) + N 1 ) C ( δ 1 / 2 + N 1 ) C δ 1 / 2 ,

and we can conclude

μ 1 ( u ) μ 1 ( u ˜ ) μ 1 ( u ) μ ( w ˜ 1 ) + μ ( w ˜ 1 ) μ ( w 1 ) + μ ( w 1 ) μ 1 ( u ˜ ) C δ 1 / 2 .

Step 2. Constructing middle region barrier ω ε (between H N ε / 2 and H K m N ε )

In step 1, we showed that N ε away from the boundary H 0 ν 1 , u ε ν 1 is homogenized with average slope approximated by μ N ( G k ) in each vertical strip I k . Now more than N ε away from H 0 ν 1 , we obtain the second homogenization of u ε ν 1 , whose slope is determined by μ N ( G k ) , k = 1 , , m . Since the width of I k = N ε , the homogenized slopes μ N ( G 1 ) ,…, μ N ( G m ) are repeated K times in a vertical strip of width K m N ε , N ε -away from H 0 ν 1 . We will specify

K 1 / δ ,

but for computational clarity, we will keep the symbol K .

We will construct middle region barrier ω ε in the region { K m N ε x 2 N ε / 2 } . To ensure that ω ε is regular near its Neumann boundary, we introduce a regularization of the original Neumann boundary data μ N ( G k ) as follows:

Consider a ball B δ α 0 / 2 N ε ( 0 ) . If I k H 0 , I j H 0 B δ α 0 / 2 N ε ( 0 ) , then k j δ α 0 / 2 and

(64) G k ( p , x / ε ) G j ( p , x / ε ) C ( 1 + p ) ( k j δ ) C ( 1 + p ) δ ( 1 α 0 / 2 ) .

By using this fact with Lemma 3.5, we can construct a C 1 function Λ ( x ) on H N ε / 2 , such that

  1. Λ C 1 ( H N ε / 2 ) with Λ C 1 δ ( N ε ) 1 ;

  2. μ N ( G k ) + δ α 0 Λ ( x ) μ N ( G k ) + δ α 0 + δ on each I k ;

  3. Λ ( x ) is periodic with period m N ε .

Note that when we patch the middle region barrier ω ε with the near-boundary barrier f ε in step 6, we will need that the average slope of ω ε is “sufficiently” larger than that of f ε . For this, we will make the average slope of ω ε to be μ N ( G k ) + O ( δ α 0 ) , i.e., ( b ) is to ensure that μ k ( ω ε ) is sufficiently larger than μ k ( f ε ) . Also when we show the flatness of barriers in steps 4 and 5, we will localize them in a “large” ball of size δ α 0 / 2 N ε .

Let Σ { K m N ε x 2 N ε / 2 } and ω ε solve the following Neumann boundary problem:

(65) F ( D 2 ω ε , x / ε ) = 0 in Σ ω ε x 2 = Λ ( x ) on H N ε / 2 ω ε = l ( x ) on H K m N ε .

Step 3. Homogenization of the operator in the middle region

Next we show, similar to Lemma 5.9, that the second homogenization does not change too much if the domain Π is replaced by Π ν 1 . More precisely, we will show that ω ε is close to ω ˜ ε solving

F ¯ ( D 2 ω ˜ ε ) = 0 in { K m N ε x ν 1 N ε / 2 } ω ˜ ε ν 1 = Λ ( x ) on H N ε / 2 ν 1 ω ˜ ε = l ( x ) on H K m N ε ν 1 .

Here, Λ ( x ) is a C 1 function constructed as in step 2, which approximates μ N ( G k ) on each I k , which is extended to R 2 so that Λ ( x ) = Λ ( p ( x ) ) for a projection p ( x ) onto H N ε / 2 .

To this end, we will first compare ω ε with ω ¯ ε , with the same Dirichlet data l on H k m N ε and solving

(66) F ¯ ( D 2 ω ¯ ε ) = 0 in Σ ω ¯ ε x 2 = Λ ( x ) on H N ε / 2 .

Lemma 5.10

For any σ > 0 , there exists N 0 such that for N 0 > N , we have

ω ε ( x ) ω ¯ ε ( x ) σ δ N ε i n Σ .

Proof

The proof follows from Theorem 2.5 applied to ( δ N ε ) 1 ω ε ( N ε x ) .□

Next we compare ω ¯ ε to ω ˜ ε to conclude. Here, we will use the rotational invariance of F ¯ .

Lemma 5.11

Let O be the rotation matrix that maps e 2 to ν 1 . Then

ω ˜ ε ( O x ) ω ¯ ε ( x ) δ 1 / 2 ( K m N ε )

in Σ { x δ 1 / 2 ( N ε ) } .

Proof

Observe that v ( x ) ω ˜ ( O x ) solves F ¯ ( D 2 v ) = 0 in Σ with Neumann boundary data Λ ( O x ) on H N ε / 2 and Dirichlet data l ( O x ) on H K m N ε . Note that due to (46) and the C 1 bound of Λ , we have

Λ ( K m N ε O x ) Λ ( K m N ε x ) tan θ 1 K m N ε x sup D Λ δ x .

and l ( K m N ε O x ) l ( K m N ε x ) K m N ε tan θ 1 x δ x .

Hence, one can apply Lemma 2.9 of [7] to τ 1 v ( τ x ) and τ 1 w ¯ ( τ x ) in τ 1 Σ , where τ = K m N ε and choose R δ 1 / 2 and ε = 2 to conclude.□

Step 4. Flatness of ω ε on H N ε , and the construction of near-boundary barrier f ε

Lemma 5.12

[Flatness of ω ε ] Let x 0 be any point on H N ε . Then for x H N ε B δ α 0 / 2 N ε ( x 0 ) ,

ω ε ( x ) ω ε ( x 0 ) 1 ω ε ( x 0 ) ( x x 0 ) 1 C δ 1 α 0 N ε .

Proof

Due to Lemma 5.10, it is enough to show aforementioned lemma for ω ¯ ε . Let ω 1 ( x ) ( K m N ε ) 1 ω ¯ ε ( ( K m N ε ) x ) , then it solves

F ¯ ( D 2 ω 1 ) = 0 in 1 x 2 1 2 K m ω 1 x 2 = Λ ( K m N ε x ) on H 1 2 K m ω 1 ( x ) = l ( x ) + C on H 1 .

We know that Λ C 1 δ ( N ε ) 1 , so the aforementioned Neumann boundary data has C 1 norm of δ K m . From Theorem 2.4, we have that

ω 1 C 1 , 1 C δ K m .

Hence,

(67) ω 1 ( x ) ω 1 ( x 0 ) 1 ω 1 ( x 0 ) ( x x 0 ) C δ K m x x 0 2 ,

which can be written in terms of ω ¯ ε ,

ω ¯ ε ( x ) ω ¯ ε ( x 0 ) 1 ω ¯ ε ( x 0 ) ( x x 0 ) C δ ( K m ) 2 ( N ε ) ( K m N ε ) 1 ( x x 0 ) 2 C δ ( δ α 0 / 2 ) 2 ( N ε ) = C δ 1 α 0 N ε

in δ α 0 / 2 N ε -neighborhood of x 0 .□

Now we construct the near-boundary barrier f ε using ω ε . Let f ε solve

F ( D 2 f ε , x / ε ) = 0 in { N ε x 2 0 } ; f ε = ω ε + δ 1 α 0 N ε on H N ε ; f ε x 2 = G D f ε , x ε on H 0 .

Step 5. Flatness of f ε

In this step, we compare μ N ( G k ) given in (53) with μ k ( f ε ) given in Definition 5.8. For simplicity, we put k = 1 . Note that Lemmas 3.2, 5.12, and 3.5 with (64) imply that

(68) μ N ( G 1 ) μ 1 ( f ε ) C ( δ 1 α 0 / 2 + δ + δ 1 α 0 ) C δ 1 α 0 .

Also from Lemma 5.12 and the definition of f ε , it follows that f ε is close to a linear function

(69) f ε ( x ) L 0 ( x ) C δ 1 α 0 N ε on H N ε B δ α 0 / 2 N ε ( 0 ) ,

where L 0 ( x ) f ε ( N ε e 2 ) + μ N ( G 1 ) ( x 2 + N ε ) + 1 f ε ( N ε e 2 ) x 1 . Then Lemma 4.4, (69), and Lemma 3.2 applied to the rescaled function ( N ε ) 1 f ε ( N ε x ) in the region { 1 x 2 1 / 2 } B δ α 0 / 2 yield that

(70) f ε L 0 C ( δ 1 α 0 + δ ( 1 α 0 / 2 ) ) N ε + C ε C δ 1 α 0 N ε

in { N ε x 2 N ε / 2 } B δ α 0 / 2 N ε ( 0 ) , where the last inequality follows from (47).

Before we proceed to the next step, observe that the C 1 regularity of Λ , Theorem 2.4, as well as Lemma 5.11 yield that

(71) ω ε ( x 1 , x 2 ) ω ε ( x 1 , N ε ) Λ ( x ) ( x 2 + N ε ) C δ 1 α 0 N ε on N ε x 2 N ε 2 .

Step 6. Patching up

Let h ( x ) l ( x ) + ( μ ( ω ε ) C δ 1 / 2 ) ( x 2 + K m N ε ) , where C > 0 is a constant given as in (b) of Theorem 4.1, and l ( x ) = l ( x 1 ) is a linear function chosen so that h ( x ) = q x on H 1 . We define

ρ ε h in { 1 x 2 K m N ε } , ω ε in { K m N ε x 2 N ε / 2 } .

Since Λ is m N ε -periodic, (b) of Theorem 4.1 implies that on { x 2 = K m N ε } ,

x 2 ω ε μ ( ω ε ) C Λ ( 1 / K , e 2 ) = μ ( ω ε ) C K 1 / 2 = μ ( ω ε ) C δ 1 / 2 = x 2 h .

Thus, it follows that F ( D 2 ρ ε , x ε ) 0 in { 1 x 2 N ε / 2 } .

Due to the flatness estimates (70) and (71), we can approximate f ε and ρ ε by linear functions, respectively, with normal derivatives of μ N ( G k ) and Λ ( x ) , with the error of O ( δ 1 α 0 N ε ) . Here, recall that Λ ( x ) was constructed so that Λ ( x ) μ N ( G k ) + δ α 0 , and α 0 is a constant satisfying α 0 1 / 2 . Then since f ε = ρ ε + δ 1 α 0 N ε on { x 2 = N ε } ,

(72) ρ ε > f ε on { x 2 = N ε / 2 } and f ε > ρ ε on { x 2 = N ε } .

Define ρ ̲ ε as follows:

ρ ̲ ε ρ ε in { 1 x 2 N ε } , min ( ρ ε , f ε ) in { N ε x 2 N ε / 2 } , f ε in { N ε / 2 x 2 0 } .

Then by (72), ρ ̲ ε is a viscosity supersolution of ( P ) ε , e 2 , 0 , q in { 1 x 2 0 } . Let us mention that, due to Lemmas 5.9, 5.10, and 5.11, a small perturbation of these barriers also yield a supersolution in { 1 x ν 1 0 } . Similarly, one can construct a subsolution ρ ¯ ε of ( P ) ε , e 2 , 0 , q by replacing Λ ( x ) given in the construction of ρ ε by Λ ˜ ( x ) μ N ( G k ) δ α 0 . Then by Lemmas 5.9 and 5.11,

(73) μ ( u ε ν 1 ) μ ( ρ ̲ ε ) μ ( ρ ¯ ε ) μ ( ρ ̲ ε ) + C δ 1 / 2 C ( δ 1 / 2 + δ α 0 ) C δ α 0 = C δ 1 / 2 ,

where the last inequality follows by choosing α 0 = 1 / 2 .

We denote ρ ¯ ε = ρ ¯ ε ν 1 and ρ ̲ ε = ρ ̲ ε ν 1 indicating that they are obtained from the direction ν 1 , i.e., with the scale N ε .

Step 7. Comparing the solutions u ε ν 1 and u ε ν 2 : Proof of Theorem 5.2 (a)

Parallel arguments as in the previous steps apply to the other direction ν 2 . Recall that

θ 2 = ν 2 e 2 < θ 1 , M = δ tan θ 2 > N .

Then similarly as in the direction ν 1 , we can construct barriers ρ ¯ ε ν 2 and ρ ̲ ε ν 2 , such that

(74) μ ( u ε ν 2 ) μ ( ρ ̲ ε ν 2 ) μ ( ρ ¯ ε ν 2 ) μ ( ρ ̲ ε ν 2 ) + C δ 1 / 2 C δ 1 / 2 .

Here, their corresponding Neumann boundary conditions satisfy

μ M ( G k ) δ α 0 δ x 2 ρ ¯ ε ν 2 ; x 2 ρ ̲ ε ν 2 μ M ( G k ) + δ α 0 + δ on H M ε I k ,

where α 0 = 1 / 2 , and the respective derivatives of ρ ¯ ε ν 2 and ρ ̲ ε ν 2 are taken as a limit from the region { 1 x 2 < M ε } .

Thus, to compare μ ( u ε ν 1 ) and μ ( u ε ν 2 ) , we compare μ N ( G k ) and μ M ( G k ) . Recall that we define μ M ( G k ) similarly as μ N ( G k ) . More precisely, μ M ( G k ) is the slope of the linear approximation of v k M , ε , where v k M , ε is defined similarly as in (52) in the region { M ε x 2 0 } with the boundary condition:

x 2 v k M , ε ( x ) = G k ( D v k M , ε , x / ε ) on H 0

and v k M , ε = l ( x ) on H M ε . Since G k is periodic on the Neumann boundary, it corresponds to the case of Neumman boundary with rational normal, passing through the origin. Hence, by applying arguments as in the proof of (34),

(75) μ N ( G k ) μ M ( G k ) C Λ ( 1 / N , e 2 ) = C inf 0 < k < 1 { 1 / N k + 1 / N 1 k } = C / N 1 / 2 .

Now we prove the following lemma using the estimate (75).

Lemma 5.13

For any ε satisfying (44),

μ ( u ε ν 1 ) μ ( u ε ν 2 ) C δ 1 / 2 .

Proof

By the construction of the viscosity supersolution ρ ̲ ε ν 1 and Lemma 5.10,

(76) μ ( ρ ̲ ε ν 1 ) μ ( ω ¯ ε ) C δ 1 / 2 ,

where ω ¯ ε is given as in (66). Similarly, we obtain

(77) μ ( ρ ̲ ε ν 2 ) μ ( ω ¯ ε ν 2 ) C δ 1 / 2 ,

where ω ¯ ε ν 2 solves

F ¯ ( D 2 ω ¯ ε ν 2 ) = 0 in { K m M ε x 2 M ε / 2 } ; ω ¯ ε ν 2 ν = Λ ν 2 ( x ) on H M ε / 2 ; ω ¯ ε ν 2 = l ( x ) on H K m M ε .

Here, Λ ν 2 ( x ) is constructed similarly as Λ ( x ) with N replaced by M , i.e., with μ N ( G k ) replaced by μ M ( G k ) . Then by (73), (74), (76), and (77), it suffices to prove

μ ( ω ¯ ε ) μ ( ω ¯ ε ν 2 ) C δ 1 / 2 .

Recall that Λ ( x ) μ N ( G k ) δ α 0 + δ on I k , and similarly, Λ ν 2 ( x ) μ M ( G k ) δ α 0 + δ on I k , with α 0 = 1 / 2 . Hence,

(78) μ ( ω ¯ ε ) μ ( h 1 ) , μ ( ω ¯ ε ν 2 ) μ ( h 2 ) C δ 1 / 2

for solutions h 1 and h 2 of

F ¯ ( D 2 h 1 ) = 0 in { K m N ε x 2 N ε / 2 } h 1 ν = μ N ( G k ) on H N ε / 2 I k h 1 = l ( x ) on H K m N ε

and

F ¯ ( D 2 h 2 ) = 0 in { K m M ε x 2 M ε / 2 } h 2 ν = μ M ( G k ) on H M ε / 2 I k h 2 = l ( x ) on H K m M ε .

Note that h 1 has a periodic Neumann condition on H N ε / 2 with period m N ε , and also h 2 has a periodic Neumann condition on H M ε / 2 with period m M ε . Hence, they correspond to the case of periodic Neumann boundary data, i.e., the case of Neumann boundary with a normal direction e 2 , and passing through the origin. Hence, by Theorem 4.1 with (75) and K = 1 / δ , we obtain

(79) μ ( h 1 ) μ ( h 2 ) Λ ( δ , e 2 ) + C / N 1 / 2 C ( δ 1 / 2 + ( 1 / N ) 1 / 2 ) C δ 1 / 2 ,

where the last inequality follows from (47). Then we can conclude from (78) and (79).□

  1. Funding information: Research supported by NSF DMS-1566578. I.K. is partially supported by NSF DMS-2153254.

  2. Conflict of interest: Authors state no conflict of interest.

References

[1] M. Arisawa, Long-time averaged reflection force and homogenization of oscillating Neumann boundary conditions, Ann. Inst. H. Poincare Anal. Lineaire 20 (2003), 293–332. 10.1016/s0294-1449(02)00025-2Search in Google Scholar

[2] S. Armstrong, T. Kuusi, and C. Prange, Quantitative analysis of boundary layers in periodic homogenization, ARMA 226 (2017), no. 2, 695–741. 10.1007/s00205-017-1142-zSearch in Google Scholar

[3] G. Barles and F. Da Lio, Local C0,α estimates for viscosity solutions of Neumann-type boundary value problems, J. Diff. Equ. 225 (2006), no. 1, 202–241. 10.1016/j.jde.2005.09.004Search in Google Scholar

[4] G. Barles, F. Da Lio, P. L. Lions, and P. E. Souganidis, Ergodic problems and periodic homogenization for fully nonlinear equations in half-type domains with Neumann boundary conditions, Indiana University Mathematics Journal 57 (2008), no. 5, 2355–2376. 10.1512/iumj.2008.57.3363Search in Google Scholar

[5] A. Bensoussan, J. L. Lions, and G. Papanicolaou, Asymptotic analysis for periodic structures, volume 5 of Studies in Mathematics and its Applications. North-Holland Publishing Co., Amsterdam, 1978. Search in Google Scholar

[6] L.A. Caffarelli and X. Cabre, Fully Nonlinear Elliptic Equations, Vol. 43, AMS Colloquium Publications, Providence, RI. Search in Google Scholar

[7] S. Choi and I. Kim. Homogenization for nonlinear PDEs in general domains with oscillatory Neumann boundary data, Journal de Mathématiques Pures et Appliquées 102 (2014), no. 2, 419–448. 10.1016/j.matpur.2013.11.015Search in Google Scholar

[8] S. Choi and I. Kim and Lee, Homogenization of Neumann boundary data with fully nonlinear PDEs, Analysis and PDE 6(2013), no. 4, 951–972. 10.2140/apde.2013.6.951Search in Google Scholar

[9] M. Crandall, H. Ishiii, and P. L. Lions, Users’ Guide to viscosity solutions of second order partial differential equations. Bull. Amer. Math. Soc. 27 (1992), no. 1, 1–67. 10.1090/S0273-0979-1992-00266-5Search in Google Scholar

[10] L. C. Evans, The perturbed test function method for viscosity solutions of nonlinear PDE, Proc. Roy. Soc. of Edinburgh: Section A. 111 (1989), no. 3-4, 359–375. 10.1017/S0308210500018631Search in Google Scholar

[11] L. C. Evans, Periodic homogenisation of certain fully nonlinear partial differential equations, Proc. Roy. Soc. of Edinburgh: Section A. 120(1992), no. 3–4, 245–265. 10.1017/S0308210500032121Search in Google Scholar

[12] W. M. Feldman, Homogenization of the Oscillating Dirichlet boundary condition in general domains, Journal de Mathématiques Pures et Appliquées 101 (2014), no. 5, 599–622. 10.1016/j.matpur.2013.07.003Search in Google Scholar

[13] W. M. Feldman and I. C. Kim, Continuity and discontinuity of the boundary layer tail, Ann. Sci. Éc. Norm. Supér. 50 (2017), no. 4, 1017–1064. 10.24033/asens.2338Search in Google Scholar

[14] N. Guillen and R. W. Schwab, Neumann homogenization via integro-differential operators, DCDS-A 36 (2016), no. 7, 3677–3703. 10.3934/dcds.2016.36.3677Search in Google Scholar

[15] D. Gerard-Varet and N. Masmoudi, Homogenization and boundary layers, Acta Math. 209 (2012), no. 1, 133–178. 10.1007/s11511-012-0083-5Search in Google Scholar

[16] H. Ishii and P. L. Lions, Viscosity solutions of fully nonlinear second-order elliptic partial differential equations, J. Differential Equations 83 (1990), no. 1, 26–78. 10.1016/0022-0396(90)90068-ZSearch in Google Scholar

[17] L. Kuipers and H. Niederreiter, Uniform Distribution of Sequences, Pure and Applied Mathematics, Wiley, New York, 1974. Search in Google Scholar

[18] D. Li and K. Zhang, Regularity for fully nonlinear elliptic equations with oblique boundary conditions, Archive for Rational Mechanics and Analysis 228 (2018), no. 3, 923–967. 10.1007/s00205-017-1209-xSearch in Google Scholar

[19] G. M. Lieberman and Neil S. Trudinger, Nonlinear Oblique boundary value problems for nonlinear elliptic equations, Trans. AMS 295 (1985), no. 2, 509–546. 10.1090/S0002-9947-1986-0833695-6Search in Google Scholar

[20] Z. Shen and J. Zhuge, Boundary Layers in Periodic Homogenization of Neumann Problems, Comm. Pure Appl. Math 71 (2018), 2163–2219. 10.1002/cpa.21740Search in Google Scholar

[21] H. Tanaka, Homogenization of diffusion processes with boundary conditions, Stochastic Analysis and Applications, Adv. Probab. Related Topics, vol. 7, Dekker, New York, 1984, pp. 411–437. Search in Google Scholar

[22] H. Weyl, UUUUber ein in der Theorie der säkutaren Störungen vorkommendes Problem, Rendiconti del Circolo Matematico di Palemo 330 (1910), 377–407. 10.1007/BF03014883Search in Google Scholar

Received: 2022-10-16
Revised: 2023-02-09
Accepted: 2023-02-13
Published Online: 2023-03-09

© 2023 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. Asymptotic properties of critical points for subcritical Trudinger-Moser functional
  3. The existence of positive solution for an elliptic problem with critical growth and logarithmic perturbation
  4. On some dense sets in the space of dynamical systems
  5. Sharp profiles for diffusive logistic equation with spatial heterogeneity
  6. Generic properties of the Rabinowitz unbounded continuum
  7. Global bifurcation of coexistence states for a prey-predator model with prey-taxis/predator-taxis
  8. Multiple solutions of p-fractional Schrödinger-Choquard-Kirchhoff equations with Hardy-Littlewood-Sobolev critical exponents
  9. Improved fractional Trudinger-Moser inequalities on bounded intervals and the existence of their extremals
  10. The existence of infinitely many boundary blow-up solutions to the p-k-Hessian equation
  11. A priori bounds, existence, and uniqueness of smooth solutions to an anisotropic Lp Minkowski problem for log-concave measure
  12. Existence of nonminimal solutions to an inhomogeneous elliptic equation with supercritical nonlinearity
  13. Non-degeneracy of multi-peak solutions for the Schrödinger-Poisson problem
  14. Gagliardo-Nirenberg-type inequalities using fractional Sobolev spaces and Besov spaces
  15. Ground states of Schrödinger systems with the Chern-Simons gauge fields
  16. Quasilinear problems with nonlinear boundary conditions in higher-dimensional thin domains with corrugated boundaries
  17. A system of equations involving the fractional p-Laplacian and doubly critical nonlinearities
  18. A modified Picone-type identity and the uniqueness of positive symmetric solutions for a prescribed mean curvature problem
  19. On a version of hybrid existence result for a system of nonlinear equations
  20. Special Issue: Geometric PDEs and applications
  21. Preface for the special issue on “Geometric Partial Differential Equations and Applications”
  22. Convex hypersurfaces with prescribed Musielak-Orlicz-Gauss image measure
  23. Total mean curvatures of Riemannian hypersurfaces
  24. On degenerate case of prescribed curvature measure problems
  25. A curvature flow to the Lp Minkowski-type problem of q-capacity
  26. Aleksandrov reflection for extrinsic geometric flows of Euclidean hypersurfaces
  27. A note on second derivative estimates for Monge-Ampère-type equations
  28. The Lp chord Minkowski problem
  29. Widths of balls and free boundary minimal submanifolds
  30. Smooth approximation of twisted Kähler-Einstein metrics
  31. The exterior Dirichlet problem for the homogeneous complex k-Hessian equation
  32. A Carleman inequality on product manifolds and applications to rigidity problems
  33. Asymptotic behavior of solutions to the Monge-Ampère equations with slow convergence rate at infinity
  34. Pinched hypersurfaces are compact
  35. The spinorial energy for asymptotically Euclidean Ricci flow
  36. Geometry of CMC surfaces of finite index
  37. Capillary Schwarz symmetrization in the half-space
  38. Regularity of optimal mapping between hypercubes
  39. Special Issue: In honor of David Jerison
  40. Preface for the special issue in honor of David Jerison
  41. Homogenization of oblique boundary value problems
  42. A proof of a trace formula by Richard Melrose
  43. Compactness estimates for minimizers of the Alt-Phillips functional of negative exponents
  44. Regularity properties of monotone measure-preserving maps
  45. Examples of non-Dini domains with large singular sets
  46. Sharp inequalities for coherent states and their optimizers
  47. Gradient estimates and the fundamental solution for higher-order elliptic systems with lower-order terms
  48. Propagation of symmetries for Ricci shrinkers
  49. Linear extension operators for Sobolev spaces on radially symmetric binary trees
  50. The Neumann problem on the domain in 𝕊3 bounded by the Clifford torus
  51. On an effective equation of the reduced Hartree-Fock theory
  52. Polynomial sequences in discrete nilpotent groups of step 2
  53. Integral inequalities with an extended Poisson kernel and the existence of the extremals
  54. On singular solutions of Lane-Emden equation on the Heisenberg group
Downloaded on 28.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2022-0051/html
Scroll to top button