Home Viscosity method for random homogenization of fully nonlinear elliptic equations with highly oscillating obstacles
Article Open Access

Viscosity method for random homogenization of fully nonlinear elliptic equations with highly oscillating obstacles

  • Ki-Ahm Lee and Se-Chan Lee EMAIL logo
Published/Copyright: September 8, 2022

Abstract

In this article, we establish a viscosity method for random homogenization of an obstacle problem with nondivergence structure. We study the asymptotic behavior of the viscosity solution u ε of fully nonlinear equations in a perforated domain with the stationary ergodic condition. By capturing the behavior of the homogeneous solution, analyzing the characters of the corresponding obstacle problem, and finding the capacity-like quantity through the construction of appropriate barriers, we prove that the limit profile u of u ε satisfies a homogenized equation without obstacles.

MSC 2010: 35B05; 35B27; 74Q10

1 Introduction

This article is devoted to the random homogenization of fully nonlinear equations with highly oscillating random obstacles, via a viscosity method. Indeed, a variety of physical and biological phenomena can be modeled by partial differential equations (PDEs) on the media with periodic structure (or oscillating obstacles); see [3,18]. Then the solutions, u ε , of these equations are expected to possess periodic oscillation in microscopic scale (often denoted by ε ), which is much smaller than the size of the domain with macroscopic scale. The homogenization process is interested in describing the asymptotic behavior of u ε when ε 0 and determining the effective model which is satisfied by the limit solution u = lim ε 0 u ε .

To state our main theorem, let ( Ω , , P ) be a given probability space. For every ω Ω and every ε > 0 , we consider a domain D ε ( ω ) obtained by perforating holes from an open, bounded domain D of R n . We denote, by T ε ( ω ) , the set of holes (i.e., D ε ( ω ) = D T ε ( ω ) ) and impose two assumptions on T ε ( ω ) , namely Assumptions 1.3 and 1.4, which will be stated later. Moreover, let us consider a special smooth function φ ( x ) in D such that φ 0 on D and φ > 0 in some region of D . Then we are going to consider highly oscillating obstacles φ ε ( x ) which are zero in D ε ( ω ) and φ ( x ) on holes T ε ( ω ) , i.e.,

φ ε ( x ) φ ( x ) if x T ε ( ω ) , 0 otherwise .

Now we consider the standard obstacle problem asking the least viscosity supersolution of Laplacian operator above the given oscillating obstacle:

Δ u ε 0 in D , u ε ( x ) = 0 on D , u ε ( x ) φ ε ( x ) in D . ( L ε )

The concept of viscosity solution and its regularity can be found in [6]. Then our main theorem concerning the Laplacian operator is the following:

Theorem 1.1

Let u ε be the least viscosity supersolution of ( L ε ) .

  1. There is a continuous function u such that u ε u in D with respect to L p -norm, for p > 0 , and for any δ > 0 , there is a subset D δ D and ε 0 such that for 0 < ε < ε 0 , u ε u uniformly in D δ as ε 0 and D D δ < δ .

  2. There exists a critical value β 0 > 0 such that u is a viscosity solution of

    Δ u + β 0 ( φ u ) + = 0 i n D , u = 0 o n D . ( L ¯ )

Here the critical value β 0 can be interpreted as a capacity-like quantity; see [7] for details. Moreover, a viscosity method for the Laplacian case can be extended to a general class of fully nonlinear operators. More precisely, we will consider a fully nonlinear operator F , which satisfies the following two properties:

  1. F is uniformly elliptic: there exist positive constants 0 < λ Λ such that for any symmetric n × n -matrix M ,

    λ N F ( M + N ) F ( M ) Λ N , N 0 .

    Here we write N 0 whenever N is a nonnegative definite symmetric matrix;

  2. F is positively homogeneous of degree one: F ( t M ) = t F ( M ) for any t > 0 and any M S n .

Typical examples that satisfy (F1) and (F2) are the Pucci extremal operators

P + ( D 2 u ( x ) ) λ μ i < 0 μ i + Λ μ i 0 μ i , P ( D 2 u ( x ) ) Λ μ i < 0 μ i + λ μ i 0 μ i ,

where μ 1 , , μ n are eigenvalues of D 2 u ( x ) . Then we will deal with the fully nonlinear version of equation ( L ε ) : find the least viscosity supersolution u ε such that

F ( D 2 u ε ( x ) ) 0 in D , u ε = 0 on D , u ε ( x ) φ ε ( x ) in D . ( F ε )

Then our main theorem concerning the fully nonlinear operator is the following:

Theorem 1.2

Let u ε be the least viscosity supersolution of ( F ε ) .

  1. There is a continuous function u such that u ε u in D with respect to L p -norm, for p > 0 , and for any δ > 0 , there is a subset D δ D and ε 0 such that for 0 < ε < ε 0 , u ε u uniformly in D δ as ε 0 and D D δ < δ .

  2. There exists a fully nonlinear, uniformly elliptic operator F ¯ such that u is a viscosity solution of

    F ¯ ( D 2 u , ( φ u ) + ) = 0 i n D , u = 0 o n D . ( F ¯ )

1.1 History

There has been a large body of literature on the periodic homogenization of linear and nonlinear PDEs; for classical results, see [3,5,13,14,18] and references therein. Here we concentrate on summarizing the homogenization results which are closely related to our circumstances.

Cioranescu and Murat employed an energy method to analyze the asymptotic behavior of u ε in their paper [10], entitled “A strange term coming from nowhere.” To be precise, they proved that the solution u ε of Laplace equation ( Δ u ε = f ) in a perforated domain with critical hole size converges to the solution u of Laplace equation with an additional term depending on the capacity of holes ( Δ u + μ u = f ). The proof relies on the construction of appropriate correctors with desired properties under abstract framework. Note that in their periodic setting, all holes have the identical size and u ε = 0 on T ε rather than u ε 0 on T ε .

The homogenization result in [10] was extended to the stationary ergodic setting for the Laplace equations with obstacles by Caffarelli and Mellet [8]. Here the hypothesis of stationary ergodicity is an extension of the notion of periodicity or almost periodicity, and it requires a random variable to have self-averaging behavior. They overcame the difficulty coming from randomness by exploiting the subadditive ergodic theorem: we refer to [1,9,12] for details. Tang [27] and Lee and Lee [22] generalized this result for p -Laplacian operator and φ -Laplacian operator, respectively, where φ is a general N-function.

Caffarelli and one of the authors [7] developed a viscosity method for periodic homogenization of Laplacian and fully nonlinear operator with highly oscillating obstacles. They considered a viscosity solution satisfying a uniformly elliptic equation with nondivergence structure and established a viscosity method to find an effective equation satisfied by the limit function. The main ingredient for proof is the construction of viscosity supersolutions and subsolutions (i.e., barriers) adopting the homogeneous solution. See also [19] and [23] for an application of a viscosity method for periodic homogenization of nonlinear parabolic equations and semilinear equations, respectively.

1.2 Main steps

We summarize the main strategies of this article and explain related key features briefly.

  1. (The critical value β 0 ) In the stationary ergodic environment, the determination of the critical value β 0 is performed by an application of the subadditive ergodic theorem. For this purpose, we define a proper contact set (often with zero obstacle) of some auxiliary functions so that the measure of a contact set has a subadditive property. For the equations with divergence structure [8, 22,27], this process has been done by considering the Dirac-delta distribution δ 0 . Unfortunately, the inherent lack of divergence structure (i.e., integration by parts) prevents us from employing similar techniques. To overcome this obstruction, our idea is to approximate the homogeneous solution in the sense of “shape,” which enables us to define auxiliary functions without the notion of δ 0 .

    We now denote two auxiliary functions, namely, free solutions w β , σ , A and obstacle solutions v β , σ , A (see Section 4 for precise definitions). To find the critical value β 0 , we further have to check the convergence of these functions when σ 0 . Unlike the linear case, there is no monotone property for the fully nonlinear case; however, such difficulty could be overcome by the isolated singularity theorem, Theorem 4.11. In short, this theorem guarantees that a singular solution must behave like the corresponding homogeneous solution, near an isolated singularity. With the help of Theorem 4.11 and Arzela-Ascoli theorem, we derive the existence and uniqueness of such limit function.

  2. (Properties of a corrector w ε ) After determining the critical value β 0 , we define a corrector w ε (see Section 5 for precise definitions). Here we require two properties for w ε to finish the proof of our main theorem:

    1. lim ε 0 w ε = 0 away from each hole;

    2. w ε = 1 (or w ε 1 , see Section 5) on the boundary of each hole.

    Note that (P2) is trivial by the definition of w ε . Our strategy is to check these properties for the auxiliary functions w β , A ε lim ε 0 w β , σ , A ε and v β , A ε lim ε 0 v β , σ , A ε first, and then transport the convergence property (P1) to the corrector w ε via the comparison principle. Indeed, we show that the auxiliary functions satisfy (P1) and (P2) by studying the theory for obstacle problems and singular solutions, together with the criticality of β 0 . More precisely, we discover the “spreading effect” of obstacle solutions using the quadratic growth of obstacle problems and construct appropriate barriers using the behavior of (approximated) homogeneous solutions. Again, although the linear case is fairly straightforward, an additional challenge occurs for the fully nonlinear case; the Alexandrov-Backelman-Pucci estimate (for viscosity solutions) and the stability of coincidence sets (for obstacle problems) will help us.

1.3 Assumptions on the holes

Let us make precise assumptions on the holes

T ε ( ω ) = ( k Z n B a ε ( r ( k , ω ) ) ( ε k ) ) D ,

where the size of hole is determined randomly, but the center of hole is periodically distributed.

Assumption 1.3

For all k Z n and a.e. ω Ω , there exists γ ( k , ω ) (independent of ε ) such that

a ε ( r ( k , ω ) ) α = ε α + 2 γ ( k , ω ) ,

where α denotes the scaling exponent of F (see Remark 3.2 (i)) and a ε ( r ) = r ε α + 2 α . Moreover, we assume that there exists a constant γ ¯ > 0 :

γ ( k , ω ) γ ¯ for all k Z n and a.e. ω Ω .

Assumption 1.4

The process γ : Z n × Ω [ 0 , ) is stationary ergodic: there exists a family of measure-preserving transformations τ k : Ω Ω such that

  1. (Stationary) γ ( k + k , ω ) = γ ( k , τ k ω ) for all k , k Z n and ω Ω ;

  2. (ergodic) if A Ω and τ k A = A for all k Z n , then P ( A ) = 0 or P ( A ) = 1 . (In other words, the only invariant set of positive measure is the whole set.)

Remark 1.5

In this article, we will concentrate on the nontrivial case with critical hole size a ε ε α + 2 α so that the limit solution satisfies an effective equation without obstacles. In fact, the behavior of limit solution u can be different (but trivial) if the radius of holes a ε is not critical. First, if the decay rate of the hole size is higher than the critical one, then the obstacles rarely influence the behavior of limit solution. Thus, the limit solution u becomes a viscosity solution of

F ( D 2 u ) = 0 in D , u = 0 on D .

Second, on the contrary, if the decay rate of hole size is lower than the critical one, then the obstacles completely enforce the behavior of limit solution. Thus, the limit solution u becomes a least viscosity supersolution of

F ( D 2 u ) 0 in D , u φ in D , u = 0 on D .

See [7] for details.

Remark 1.6

For a uniformly elliptic operator with divergence structure, the assumption on the hole size is usually stated in terms of capacity. For example, when we consider the Laplacian case with n 3 (see [8]), the assumption corresponding to Assumption 1.3 is: for all k Z n and a.e. ω Ω , there exists γ ( k , ω ) (independent of ε ) such that

cap ( B a ε ( r ( k , ω ) ) ( ε k ) ) = ε n γ ( k , ω ) ,

where cap ( A ) denotes a (variational) capacity of a subset A of R n and a ε ( r ) = r ε n n 2 .

Here since capa ( B r ) = c n r n 2 , we have

r ( k , ω ) = γ ( k , ω ) c n 1 n 2 .

Nevertheless, for the fully nonlinear case, the notion of capacity is unclear and so we imposed different assumptions on the hole size. Note that in Assumption 1.3, we can write

r ( k , ω ) = γ ( k , ω ) 1 α ,

which is analogous to the Laplacian case where α = n 2 .

1.4 Outline

This article is organized as follows. In Section 2, we investigate the behavior of u ε away from holes, and as a consequence, we derive the convergence of u ε to the homogenized solution u . Section 3 is devoted to the explanation of a homogeneous solution and its C 1 , 1 -approximation in the sense of “shape.” In Section 4, we define free solutions w β , σ , A and obstacle solutions v β , σ , A and prove the convergence of these auxiliary functions when σ 0 . Then we conclude that the critical value β 0 is well-defined by the subadditive ergodic theorem. In Section 5, we justify two properties of w β , σ , A and transport the information to the corrector w ε , which enables us to finish the proof for our main theorem. Note that to clarify the difficulties coming from nonlinearity, we deal with the Laplacian case and the fully nonlinear case in consecutive order within each section.

2 Estimates and convergence

We begin with the estimate for the oscillation of u ε on B b ε ( k ) where b ε is chosen to have an intermediate growth rate between ε and a ε . We first consider the Laplacian case.

Lemma 2.1

Set b ε ( k , ω ) = ( ε a ε ( k , ω ) ) 1 / 2 where a ε ( k , ω ) ε n n 2 is the critical rate. Then

osc B b ε ( k ) u ε = o ( ε γ )

for k ε Z n supp φ and for some 0 < γ < 1 .

Proof

See Lemma 3.4. in [7] or Lemma 2.9. in [19] for proof.□

Next, we control the behavior of u ε in D ( k Z n B b ε ( k ) ) by constructing appropriate barrier functions h ε ± and applying the comparison principle with u ε . In this method, we do not require the size of perforating holes to be identical. Note that in the periodic setting, similar results follow from the discrete gradient estimate [7,19]; if we define

Δ e i u ε u ε ( x + ε e i ) u ε ( x ) ε for unit vector e i R n ,

then there exists a uniform constant C > 0 such that

(2.1) Δ e u ε < C .

Lemma 2.2

For ε ( 0 , 1 ) , let h ε ± be the solutions of the Dirichlet problem

Δ h ε ± = 0 i n D ( k Z n B b ε ( k ) ) , h ε ± = 0 o n D , h ε + ( x ) = sup B b ε ( k ) u ε h ε ( x ) = inf B b ε ( k ) u ε f o r x B b ε ( k ) w h e r e k Z n .

Then h ε ± have the following properties:

  1. 0 h ε u ε h ε + .

  2. h ε ± C 2 , α ; in particular, we have

    h ε ± ( x ) h ε ± ( y ) C x y α ,

    for any α ( 0 , 1 ) and any x , y D ( k Z n B b ε ( k ) ) .

  3. h ε + h ε max k Z n osc B b ε ( k ) u ε .

Proof

  1. It follows directly from the construction of h ε ± and the comparison principle with u ε .

  2. Since the boundary data for h ε ± are clearly in C α for any α ( 0 , 1 ) , the desired result follows from the boundary C 2 , α -estimate; for example, see [15].

  3. The maximum principle for h ε + h ε yields the inequality.□

We also need the following version of Arzela-Ascoli theorem, whose proof is a simple modification of the original one. In short, the equicontinuous assumption in Arzela-Ascoli theorem can be relaxed to “almost equicontinuity.”

Lemma 2.3

[Arzela-Ascoli theorem] Let A be a compact subset of R n . Suppose that a sequence of functions { f l } l N defined on A satisfies

  1. (Uniformly bounded) There exists a constant M > 0 such that

    f l ( x ) M ,

    for any l N and x A .

  2. (Almost equicontinuous) There exists a constant α ( 0 , 1 ) , C > 0 and a function g : N R 0 such that lim l g ( l ) = 0 and

    f l ( x ) f l ( y ) C x y α + g ( l ) ,

    for any x , y A .

Then there exists a subsequence { f l k } k N which converges uniformly on A. Moreover, if we denote the limit function by f, then f C α ( A ) .

Theorem 2.4

(Uniform convergence) There is a continuous function u such that u ε u weakly in D with respect to L p -norm for any p > 0 . Also for any δ > 0 , there is a subset D δ D and a sequence { ε l } l N such that ε l > 0 , lim l ε l = 0 , and u ε l u uniformly in D δ as l and D D δ < δ .

Proof

See [7,19] for detailed proof. Here the only difference occurs when applying the discrete gradient estimate in the references. More precisely, the absence of periodicity in stationary ergodic setting prevents us from achieving the discrete gradient estimate (2.1). Nevertheless, Lemmas 2.1 and 2.2 provide the “almost equicontinuity” of u ε , i.e.,

u ε ( x ) u ε ( y ) = u ε ( x ) u ε ( y ) h ε + ( x ) h ε ( y ) h ε + ( x ) h ε + ( y ) + h ε + ( y ) h ε ( y ) C x y α + max k Z n osc B b ε ( k ) u ε C x y α + o ( ε γ ) ,

for any x , y D ( k Z n B b ε ( k ) ) where we assumed u ε ( x ) u ε ( y ) without loss of generality. Then we can apply the modified Arzela-Ascoli theorem, Lemma 2.3, for u ε and finish the proof following the previous references.□

Remark 2.5

Note that the argument for the Laplacian operator in this section can be repeated for the uniformly elliptic, fully nonlinear operator F . Indeed, we only used the comparison principle, boundary C α -estimate, Harnack inequality, and oscillation lemma which still hold for F ; for example, see [6,11].

In conclusion, we presented the proof for the first part of Theorems 1.1 and 1.2, which concerns the convergence of u ε to the limit function u . In the remaining of this article, we will concentrate on the second part of our main theorems by constructing a proper corrector and investigating its properties.

3 Approximation of a homogeneous solution

To determine the critical value β 0 , the essential step is to define the corresponding subadditive quantity since the size of hole is not identical, but random. In the articles [8] (for Laplacian case), [27] (for p -Laplacian case), and [22] (for φ -Laplacian case), they described the subadditive quantity in terms of Dirac-delta distribution δ 0 and proved the properties of correctors using an energy method. However, this approach is not suitable for our case, because the operators that we consider do not have the divergence structure. Hence, we concentrate on the nondivergence structure of F ; in particular, we will capture its “shape” and employ a viscosity method to verify the properties of correctors.

3.1 A homogeneous solution

The starting point is a homogeneous solution for F , a uniformly elliptic fully nonlinear operator being homogeneous of degree one. The following lemma shows the existence and uniqueness, and classifies the behavior of a homogeneous solution for F .

Lemma 3.1

(A homogeneous solution; [2,7]) There exists a nonconstant solution of F ( D 2 u ) = 0 in R n { 0 } that is bounded below in B 1 and bounded above in R n B 1 . Moreover, the set of all such solutions is of the form { a Φ + b a > 0 , b R } , where Φ C loc 1 , γ ( R n { 0 } ) can be chosen to satisfy one of the following homogeneity relations: for all t > 0

(3.1) Φ ( x ) = Φ ( t x ) + log t i n R n { 0 } w h e r e α = 0

or

(3.2) Φ ( x ) = t α Φ ( t x ) , α Φ > 0 i n R n { 0 } ,

for some number α ( 1 , ) { 0 } that depends only on F and n.

Remark 3.2

  1. We call the number α = α ( F ) the scaling exponent of F .

  2. In the remaining of this article, we concentrate on the case α > 0 (which corresponds to n 3 in the Laplacian case) to simplify the statement. Indeed, the same argument can be applied to α = 0 (which corresponds to n = 2 in the Laplacian case) and α < 0 (which corresponds to n = 1 in the Laplacian case).

  3. We can easily calculate the exact scaling exponent of the Pucci extremal operators. For example, we have

    α ( P λ , Λ + ) = ( n 1 ) λ Λ 1 ( > 0 ) if 1 Λ λ < n 1 , 0 if Λ λ = n 1 , ( n 1 ) λ Λ 1 ( < 0 ) if n 1 < Λ λ .

    In a similar way, α ( P λ , Λ ) can be computed; see [20].

  4. Since F is positively homogeneous of degree one, we have a Φ + b is again a homogeneous solution for a > 0 , b R and a homogeneous solution Φ . In the remaining of this article, we fix the “normalized” homogeneous solution Φ by

    Φ ( x ) = x α Φ x x x α ϕ ( θ ) , for θ = x x S n 1 ,

    where ϕ is chosen so that min θ S n 1 ϕ ( θ ) = 1 . Note that here we normalize a homogeneous solution in the sense of the “height” (at x = 1 ) while in the Laplacian case, we typically normalize in the sense of “mass” (i.e., measure):

    Δ Φ = δ 0 .

3.2 Approximation of a homogeneous solution

In a divergence case, it is natural to approximate the Dirac-delta measure δ 0 by measurable functions { f σ } . More precisely, we define

f σ ( x ) = 1 B σ χ B σ ( x ) ,

for any σ > 0 and let a regularized homogeneous solution Φ σ by the solution of T Φ σ = f σ in R n , where T is a uniformly elliptic operator with divergence structure. Then Φ σ Φ in L 1 ( R n ) and f σ δ 0 in distribution sense as σ 0 + .

In nondivergence case, we do not have the corresponding measure such as the Dirac-delta δ 0 . In other words, it is difficult to define F ( D 2 Φ ) in the whole space R n , while we know that F ( D 2 Φ ) = 0 in R n { 0 } ; see [20]. Thus, instead of measure-sense, we focus on the “shape” of Φ ; we define an approximated homogeneous solution Φ σ for σ > 0 by

Φ σ = Φ in R n B a ¯ σ , W σ in B a ¯ σ ,

where Φ is the normalized homogeneous solution and W σ , a ¯ σ will be determined later. (Note that a radius a ¯ σ must converge to zero when σ 0 .) Then we define a corresponding function ν σ by

ν σ F ( D 2 Φ σ ) in R n .

Since Φ σ = Φ in B a ¯ σ c and Φ is a homogeneous solution, we immediately have that ν σ 0 in B a ¯ σ c and so supp ν σ B a ¯ σ .

3.2.1 Laplacian case

Continuing to the aforementioned argument, we can define a radius a ¯ σ and an approximated homogeneous solution W σ . Note that for the Laplacian case, we have α = n 2 and so the normalized homogeneous solution is given by Φ ( x ) = x 2 n . On the other hand, we see that the radius of hole a ε is assumed to be comparable to ε n / ( n 2 ) . Since the corrector w ε will be constructed so that w ε 1 near T ε (see Section 5), we require the homogeneous solution Φ ( x ) ε 2 near x = a ¯ ε = a ε / ε . Here we need to distinguish scale ε and scale 1.

Therefore, we let a ¯ σ σ 2 n 2 and determine a quadratic polynomial W σ which is rotationally symmetric and satisfies

W σ ( x ) = Φ ( x ) and W σ ( x ) = Φ ( x ) ,

if x = a ¯ σ . Indeed, for σ > 0 , we set

Φ σ ( x ) Φ ( x ) = x 2 n x a ¯ σ , W σ ( x ) = m σ x 2 + k σ x < a ¯ σ ,

where m σ = n 2 2 σ 2 n n 2 and k σ = n 2 σ 2 . Then Φ σ C 1 , 1 ( R n ) and it follows that

Δ Φ σ ( x ) = ν σ ( x ) 0 x a ¯ σ , 2 n m σ x < a ¯ σ .

On the other hand, by its construction, we immediately have that

Φ σ Φ as σ 0 ,

locally uniformly on R n { 0 } . Moreover, for 0 < σ 1 σ 2 ,

(3.3) Φ σ 1 = Φ σ 2 in B a ¯ σ 2 c ,

and

(3.4) Φ σ 1 Φ σ 2 in R n .

Note that this approximation is related to the Dirac-delta measure:

ν σ n ( n 2 ) ω n δ 0 as σ 0 ,

in distribution sense, i.e.,

B a ¯ σ ν σ ( x ) η ( x ) d x n ( n 2 ) ω n η ( 0 ) as σ 0 ,

for any η C c ( R n ) .

3.2.2 Fully nonlinear case

In this case, the radius of hole a ε is comparable to ε α + 2 α . Since the normalized homogeneous solution is given by

Φ ( x ) = r α ϕ ( θ ) ,

in spherical coordinates, we let a ¯ σ = σ 2 α . Moreover, we consider a (strict) superlevel set of Φ :

C a ¯ σ { x = ( r , θ ) R n : r α ϕ ( θ ) > a ¯ σ α } .

Note that we have C a ¯ σ = B a ¯ σ as before, if we let ϕ ( θ ) 1 , i.e., F is rotationally symmetric. Then for σ > 0 , we set

Φ σ ( x ) Φ ( x ) = r α ϕ ( θ ) x C a ¯ σ c , W σ ( x ) = m σ ( r α ϕ ( θ ) 1 ) s + k σ x C a ¯ σ ,

where s (which is independent of σ > 0 ), m σ , and k σ will be determined.

  1. ( Φ σ C 1 , 1 ) We only need to check this property on C a ¯ σ . In fact, for ( r , θ ) C a ¯ σ , we have

    Φ ( r , θ ) = a ¯ σ α , W σ ( r , θ ) = m σ a ¯ σ α s + k σ , r Φ ( r , θ ) = α r a ¯ σ α , r W σ ( r , θ ) = m σ α s r a ¯ σ α s ,

    and

    θ Φ ( r , θ ) = r α θ ϕ , θ W σ ( r , θ ) = m σ s r α s ϕ ( θ ) s 1 θ ϕ = m σ s r α a ¯ σ α ( s + 1 ) θ ϕ .

    Therefore, we conclude W σ C 1 , 1 provided that

    m σ = 1 s σ 2 ( s + 1 ) and k σ = 1 + 1 s σ 2 ,

    for some constant s > 0 .

  2. ( F ( D 2 W σ ) ν σ 0 ) To verify this property, it is enough to show that there exists a sufficiently large s such that

    P ( D 2 W σ ) 0 .

    For this purpose, we claim that for sufficiently large s = s ( λ , Λ , f ) > 0 , we have

    P ( D 2 w ) 0 ,

    where w ( r , θ ) r s f ( θ ) for a positive function f C 2 ( S n 1 ) . Indeed, one can calculate the Hessian of w as follows:

    Hess ( w ) ( a i j ( s , θ ) ) r s 2 ,

    where

    a i j ( s , θ ) = s ( s 1 ) f ( θ ) if ( i , j ) = ( 1 , 1 ) , o ( s 2 ) g i j ( θ ) otherwise ,

    for g i j C ( S n 1 ) , 1 i , j n . See Appendix in [24] for the computation of the Hessian matrix in spherical coordinates. In short, we have the dominant s 2 -order only in ( 1 , 1 ) -component of Hess ( w ) , since the power of s is added if and only if we take a radial derivative with respect to w .

    Moreover, since det ( t I A ) = ( t λ 1 ( A ) ) ( t λ n ( A ) ) and the determinant function is smooth, one can easily check that the eigenvalues of Hess ( w ) are given by

    ( s 2 f ( θ ) + o ( s 2 ) b 1 ( θ ) , o ( s 2 ) b 2 ( θ ) , , o ( s 2 ) b n ( θ ) ) r s 2 ,

    for some functions b i C ( S n 1 ) . Hence, for sufficiently large s = s ( λ , Λ , f ) > 0 , we have

    P ( D 2 w ) [ λ s 2 f ( θ ) Λ b ( θ ) o ( s 2 ) ] r s 2 0 ,

    as claimed.

After choosing s , m σ , and k σ in this way, we immediately have that

Φ σ Φ as σ 0 ,

locally uniformly on R n { 0 } . Moreover, by its construction, we have

Φ σ 1 = Φ σ 2 in C a ¯ σ 2 c

and

Φ σ 1 Φ σ 2 in R n ,

for 0 < σ 1 σ 2 .

4 Convergence of free solutions and obstacle solutions

To apply the subadditive ergodic theorem (see [9,12]) and determine the critical value β 0 , we first consider an obstacle problem and its solution as an auxiliary function for a corrector w ε . In view of [8,22,27], the forcing term of an obstacle problem was presented by the Dirac-delta measure. In contrast to those operators of divergence form, we cannot exploit this energy-type method in fully nonlinear operator of nondivergence form. Instead, to capture the behavior of a corrector w ε , we are going to adopt the approximation of a homogeneous solution which was obtained in the previous section. Moreover, to connect the properties between an obstacle solution and a corrector, we need one more auxiliary function, namely a “free” solution.

While the argument concerning these auxiliary functions is relatively straightforward in the Laplacian case, there arises several challenges in the fully nonlinear case. Hence, we will first investigate nice properties of obstacle solutions and free solutions in the Laplacian case, and then justify the validity of those properties in the fully nonlinear case.

4.1 Laplacian operator

We start with the definition of obstacle solutions and free solutions in the Laplacian operator:

Definition 4.1

Let A be an open and bounded subset of R n .

  1. For β R , we define an “obstacle” solution

    v β , σ , A ( x , ω ) inf v ( x ) : Δ v β k Z n A γ ( k , ω ) ν σ ( x k ) in A , v 0 in A , v = 0 on A ,

    and its rescaled function

    v ¯ β , σ ε ( y , ω ) ε 2 v β , σ , ε 1 A ( y / ε , ω ) , in A .

  2. For β R , we define a “free” solution

    w β , σ , A ( x , ω ) inf w ( x ) : Δ w β k Z n A γ ( k , ω ) ν σ ( x k ) in A , w = 0 on A ,

    and its rescaled function

    w ¯ β , σ ε ( y , ω ) ε 2 w β , σ , ε 1 A ( y / ε , ω ) , in A .

Lemma 4.2

(Multiple sources; Δ ). Let 0 < σ 1 σ 2 1 . For a nonnegative function γ : Z n R 0 , we consider the solutions w i , defined by

Δ w i ( x ) = k Z n γ ( k ) ν σ i ( x k ) i n A , w i ( x ) = 0 o n A .

Then we have

w 1 w 2 i n A .

Proof

First, we let w ˜ i ( x ) = k Z n γ ( k ) Φ σ i ( x k ) . Then we have Δ w ˜ i ( x ) = k Z n γ ( k ) ν σ i ( x k ) and

w ˜ 1 w ˜ 2 in A

by (3.4). Recalling (3.3), (for almost every x A ), we also have that

w ˜ 1 = w ˜ 2 on A .

Thus, if we let g be the solution of the Dirichlet problem

Δ g = 0 in A , g = w ˜ 1 ( = w ˜ 2 ) on A ,

then we conclude

w i = w ˜ i + g ,

which completes the proof.□

Lemma 4.3

(Additional source; Δ ). Suppose that a function w i defined by the Dirichlet problem

Δ w i = f i in A , w i = 0 on A ,

satisfy

w 1 w 2 in A .

Moreover, for a constant β 0 , we define a function w β , i by

Δ w β , i = f i + β in A , w β , i = 0 on A .

Then we have

w β , 1 w β , 2 in A .

Proof

Let g β be the solution of the Dirichlet problem

Δ g β = β in A , g β = 0 on A .

Then the result follows immediately from

w β , i = w i + g β .

Remark 4.4

(Existence of limit free solutions; Δ ). Recall that a free solution is defined by

w β , σ , A ( x , ω ) inf w ( x ) : Δ w β k Z n A γ ( k , ω ) ν σ ( x k ) in A , w = 0 on A .

For 0 < σ 1 σ 2 1 , applying Lemmas 4.2, 4.3, and their proofs, we have

(4.1) w β , σ 1 , A w β , σ 2 , A in A ,

and furthermore,

(4.2) w β , σ 1 , A ( x ) = w β , σ 2 , A ( x ) if x k Z n ( B a ¯ σ 2 ( k ) ) c .

In particular, (4.1), the monotonicity of { w β , σ , A } σ > 0 yields the convergence of free solutions w β , σ , A when σ 0 + . We denote the limit function by w β , A .

Lemma 4.5

(Obstacle solution; Δ ). Let 0 < σ 1 σ 2 1 . For a constant β 0 and a nonnegative function γ : Z n R 0 , we define (an obstacle solution) v β , i by

(4.3) v β , i ( x ) inf v ( x ) : Δ v β k Z n A γ ( k ) ν σ i ( x k ) i n A , v 0 i n A , v = 0 o n A .

Then we have

v β , 1 v β , 2 i n A .

Proof

Note that we have the equivalent definition of an obstacle problem (4.3):

Δ v β , i β k Z n A γ ( k ) ν σ i ( x k ) , v β , i 0 in A and Δ v β , i = β k Z n A γ ( k ) ν σ i ( x k ) if v β , i > 0 .

Moreover, for sufficiently small σ i > 0 and k Z n with γ ( k ) 0 , we have β γ ( k ) ν σ i ( x k ) < 0 for x k < a ¯ σ i . Since β γ ( k ) ν σ i ( x k ) < 0 in B a ¯ σ i ( k ) for any k Z n A with γ ( k ) 0 , we have

(4.4) v β , i > 0 in k Z n A B a ¯ σ i ( k ) .

Now, if we let v ˜ ( x ) v β , 2 ( x ) + k Z n A γ ( k ) ( Φ σ 1 ( x k ) Φ σ 2 ( x k ) ) , then by (3.4), we have v ˜ v β , 2 in A . Thus, the proof will be completed if we prove v β , 1 = v ˜ . Indeed, v ˜ 0 in A . We split two cases:

  1. ( v ˜ ( x ) = 0 ) By the definition of v ˜ , for each k Z n , we have either γ ( k ) = 0 or Φ σ 1 ( x k ) = Φ σ 2 ( x k ) . In the latter case, the construction of Φ σ yields x k > a ¯ σ 2 a ¯ σ 1 and so

    ν σ i ( x k ) = 0 .

    Thus, in both cases, we have

    γ ( k ) ν σ i ( x k ) = 0

    and so

    0 = Δ v ˜ ( x ) β k Z n A γ ( k ) ν σ i ( x k ) = β .

  2. ( v ˜ ( x ) > 0 ) By the definition of v ˜ , we have either v β , 2 ( x ) > 0 or x k a ¯ σ 2 for some k Z n with γ ( k ) 0 . Recalling (4.4), in both cases, we have v β , 2 ( x ) > 0 which yields that

    Δ v β , 2 ( x ) = β k Z n A γ ( k ) ν σ 2 ( x k ) .

    Therefore, we conclude that

    Δ v ˜ ( x ) = Δ v β , 2 ( x ) + k Z n A γ ( k ) ( Φ σ 1 ( x k ) Φ σ 2 ( x k ) ) = β k Z n A γ ( k ) ν σ 1 ( x k ) .

Remark 4.6

(Existence of limit obstacle solutions; Δ ). We know that the assumption in Lemma 4.5 holds by Lemmas 4.2 and 4.3, i.e., we have (4.1) and further (4.2). Thus, by applying Lemma 4.5 and its proof, we conclude that

(4.5) v β , σ 1 , A v β , σ 2 , A in A ,

and furthermore,

(4.6) v β , σ 1 , A ( x ) = v β , σ 2 , A ( x ) if x k Z n ( B a ¯ σ 2 ( k ) ) c .

In particular, (4.5), the monotonicity of { v β , σ , A } σ > 0 yields the convergence of obstacle solutions v β , σ , A when σ 0 + .

Now we are ready to define the measure of contact set for an obstacle problem and to determine the critical value β 0 . Indeed, we define a random variable m β , A by

m β , A { x A : v β , A = 0 } .

Lemma 4.7

(A subadditive quantity).

  1. The random variable m β , A is subadditive: in other words, for the finite family of sets ( A i ) i I such that

    A i A for all i I , A i A j = for all i j , A i I A i = 0 ,

    then

    m β , A i I m β , A i .

  2. The process

    T k m β , A = m β , k + A

    has the same distribution for all k Z n .

Proof

  1. Since v β , A is admissible for v β , A i for each i , we have

    v β , A i v β , A in A i .

    Thus, we have the desired result.

  2. It follows immediately from our assumptions on the process γ ( k , ω ) .□

Due to the previous lemma, we can apply a subadditive ergodic theorem (see [9,12]). More precisely, we have

l ( β ) = lim t m β , B t B t ,

and a scaled version:

l ( β ) = lim ε 0 { y : v ¯ β ε ( y , ω ) = 0 } B 1 a.s. ,

where v ¯ β ε ( y , ω ) ε 2 v β , ε 1 B ( y / ε , ω ) .

Lemma 4.8

(Properties of l ( β ) )

  1. l ( β ) is nondecreasing function with respect to β .

  2. If β < 0 , then l ( β ) = 0 .

  3. If β > 0 is large enough, then l ( β ) > 0 .

Proof

  1. For β 1 β 2 , we have v β 2 , A v β 1 , A , which implies that m β 1 , A m β 2 , A .

  2. Since v β , σ , B t is a solution of an obstacle problem, we have

    Δ v β , σ , B t β k Z n A γ ( k , ω ) ν σ ( k , ω ) β ,

    in B t . Thus, by the comparison principle,

    v β , σ , A β 2 n ( x 2 t 2 ) > 0

    in B t . Letting σ 0 , we have v β , A > 0 in A , which implies that l ( β ) = 0 for β < 0 .

  3. For k Z n , we define

    h k ( x ) = β 2 n x k 2 + γ ( k , ω ) Φ σ ( x k ) .

    Then we have h k C 1 , 1 and Δ h k = β γ ( k , ω ) ν σ ( x k ) . Moreover, a direct calculation yields that a rotationally symmetric function h k attains its minimum at

    x k = r k = n ( n 2 ) γ ( k , ω ) β 1 / n .

    Since γ ( k , ω ) γ ¯ , we can choose β > 0 large enough so that r k < 1 / 2 for any k Z n . Moreover, we can choose a constant D k so that the minimum of h k ( x ) D k is exactly 0. Now if we define

    h ˜ k ( x ) = h k ( x ) D k if x < r k , 0 if x r k ,

    then h ˜ k is well-defined and it belongs to C 1 , 1 . Moreover, since Δ h ˜ k = β γ ( k , ω ) ν σ ( x k ) in B r k ,

    k Z n t B h ˜ k

    is admissible for v β , σ , t B . Therefore, we conclude that

    l σ ( β ) lim t { x t B : v β , σ , t B = 0 } t B C 1 B 1 / 2 C 1 1 ω n 2 n > 0 ,

    which ensures that l ( β ) > 0 for large enough β .□

Finally, we let the (nonnegative) critical value

β 0 sup { β : l ( β ) = 0 } ,

which is well-defined by the previous lemma.

4.2 Fully nonlinear operator

Definition 4.9

  1. For M S n and β R , we define an “obstacle” solution

    v β , σ , A ; M ( x , ω ) inf v ( x ) : F ( M + D 2 v ( x ) ) β + F ( M ) k Z n A γ ( k , ω ) ν σ ( x k ) in A , v 0 in A , v = 0 on A } ,

    and its rescaled function

    v ¯ β , σ ; M ε ( y , ω ) ε 2 v β , σ , ε 1 A ; M ( y / ε , ω ) , in A .

  2. For M S n and β R , we define a “free” solution

    w β , σ , A ; M ( x , ω ) inf w ( x ) : F ( M + D 2 w ( x ) ) β + F ( M ) k Z n A γ ( k , ω ) ν σ ( x k ) in A , w = 0 on A ,

    and its rescaled function

    w ¯ β , σ ; M ε ( y , ω ) ε 2 w β , σ , ε 1 A ; M ( y / ε , ω ) , in A .

Before showing the convergence of these functions when σ 0 + , we first describe the local behavior of a singular solution with an isolated singularity at some point x 0 . Roughly speaking, we will demonstrate that the growth rate of a singular solution near the singularity point is the same as the growth rate of the corresponding homogeneous solution Φ . This type of result was first proved by Bôcher [4] for the Laplacian operator in 1903. Similar results can be found in [25,26] for quasilinear divergence-type equations, [20] for Pucci operators, [2,7] for fully nonlinear operators with homogeneous degree one, and [16,17] for a class of subequations. Note that they considered the local behavior of solutions for equations with zero-forcing term; in the following lemmas, we present generalized results by choosing a general forcing term.

Lemma 4.10

Let u C ( B 1 { 0 } ) be a viscosity solution of

F ( D 2 u ) = g ( x ) in B 1 { 0 } ,

where g L ( B 1 ) , u is bounded on B 1 , and lim x 0 u ( x ) = . Then there exist positive constants δ 0 and C 0 such that

δ 0 Φ ( x ) C 0 u ( x ) 1 δ 0 Φ ( x ) + C 0 .

Proof

We may assume u is positive in B 1 { 0 } by adding a constant on u , if necessary. To show the lower bound, suppose that there exist sequences δ i 0 , ε i 0 , and x i B 1 { 0 } such that

(4.7) u ( x i ) δ i Φ ( x i ) for x i = ε i .

Note that from [6], we have the Harnack inequality

sup B 1 / 2 u C ( inf B 1 / 2 u + g L n ( B 1 ) ) .

Recalling that F is positively homogeneous of degree one and considering the scaled function u r ( x ) u ( r x ) for small r > 0 , we deduce that

(4.8) sup B r / 2 u C ( inf B r / 2 u + r g L n ( B 1 ) ) .

Thus, (4.7), (4.8), and the homogeneity of Φ imply that

u ( x ) C ( u ( x i ) + ε i g L n ( B 1 ) ) C ( δ i Φ ( x i ) + ε i g L n ( B 1 ) ) C ˜ δ i Φ ( x ) + C ε i g L n ( B 1 ) ,

for x = ε i . Since F ( D 2 ( Φ c x 2 ) ) = F ( D 2 Φ 2 c I ) F ( D 2 Φ ) 2 c λ g ( x ) for sufficiently large c > 0 , the comparison principle yields that

u ( x ) C ˜ δ i Φ ( x ) + c 1 c 2 x 2 for ε i x 1 ,

for some c 1 , c 2 > 0 . Letting i , we have u is bounded above in B 1 { 0 } , which contradicts to the assumption lim x 0 u ( x ) = . Therefore, we obtain the lower bound and from the similar argument, we finish the proof.□

Theorem 4.11

(An isolated singularity) Let u C ( B 1 { 0 } ) be a solution of

F ( D 2 u ) = g ( x ) i n B 1 { 0 } ,

where g L ( B 1 ) , u is bounded on B 1 , and lim x 0 u ( x ) = . Then there exists a positive constant such that

(4.9) lim x 0 u ( x ) Φ ( x ) = a .

Proof

Set u ˜ ε ( x ) ε α u ( ε x ) . Then the homogeneity of the homogeneous solution Φ gives

u ( ε x ) Φ ( ε x ) = ε α u ( ε x ) ε α Φ ( ε x ) = u ˜ ε ( x ) Φ ( x ) .

For a compact set K R n { 0 } , an application of Lemma 4.10 leads to

sup x K u ˜ ε ( x ) Φ ( x ) = sup x ε K u ( ε x ) Φ ( ε x ) 1 δ 0 + C 1 ,

for some constant C 1 > 0 , which is independent of ε > 0 . Employing a similar argument for the lower bound, we conclude that

sup 0 < ε < ε 0 u ˜ ε L ( K ) C K .

Since

F ( D 2 u ˜ ε ( x ) ) = ε α + 2 g ( ε x )

holds for any x B 1 { 0 } , we also obtain the uniform Hölder estimates for the sequence { u ˜ ε } ε > 0 in K . Therefore, Arzela-Ascoli theorem implies that there exist a subsequence ε j 0 and a function v C ( R n { 0 } ) such that u ˜ ε j v locally uniformly in R n { 0 } . Then for any x R n { 0 } , the homogeneity of Φ yields

v ( x ) Φ ( x ) = lim ε 0 u ˜ ε ( x ) Φ ( x ) = lim ε 0 u ( ε x ) Φ ( ε x ) [ a ̲ , a ¯ ] ,

where

a ̲ liminf ε 0 inf x = ε u ( x ) Φ ( x ) , a ¯ limsup ε 0 sup x = ε u ( x ) Φ ( x ) .

Here a ̲ , a ¯ ( 0 , ) by Lemma 4.10. Moreover, since u ˜ ε v and ε α + 2 g ( ε x ) 0 uniformly on every compact subset K , we have F ( D 2 v ) = 0 in R n { 0 } . Finally, choose x ε B ε so that u ( x ε ) = inf x = ε u ( x ) Φ ( x ) Φ ( x ε ) . Then there exist a (further) subsequence ε j 0 and y B 1 such that x ε j ε j y . Since

v ( y ) = lim j u ˜ ε j ( ε j 1 x ε j ) = lim j ε j α u ( x ε j ) = lim j inf x = ε j u ( x ) Φ ( x ) Φ x ε j ε j = a ̲ Φ ( y ) ,

we conclude that v a ̲ Φ in R n { 0 } by the strong maximum principle. Hence, we have

limsup x 0 u ( x ) Φ ( x ) = limsup ε 0 max x B 1 u ˜ ε ( x ) Φ ( x ) = max x B 1 v ( x ) Φ ( x ) = a ̲ ,

which implies the desired result

lim x 0 u ( x ) Φ ( x ) = a ( a ̲ = a ¯ ) .

4.2.1 Free solutions

For notational simplicity, we write a free solution

w σ ( x , ω ) w β , σ , A ; M ( x , ω ) ,

where β 0 , A R n , M S n are fixed and σ > 0 . Recall that we defined

F ( M + D 2 w β , σ , A ; M ( x , ω ) ) = β + F ( M ) k Z n A γ ( k , ω ) ν σ ( x k ) in A , w β , σ , A ; M = 0 , on A .

Moreover, we denote

z σ k Z n A γ ( k , ω ) Φ σ ( x k ) .

For 0 < σ 1 σ 2 1 , we have

z σ 1 ( x ) = z σ 2 ( x ) ,

for x A satisfying x k a ¯ σ 2 ( k ) for any k Z n .

We will estimate F ( M + D 2 z σ ) : we may expect

F ( M + D 2 z σ ) F ( M ) k Z n A γ ( k , ω ) ν σ ( x k )

in A , by heuristic computation. Indeed, recall that F is positively homogeneous of degree one, F ( D 2 Φ σ ) = ν σ by the construction of approximated homogeneous solution Φ σ and Φ σ ( x k ) is “flat” away from k Z n (i.e., D 2 Φ σ ( x k ) 0 , away from k Z n ). We prove this observation rigorously in the following lemma:

Lemma 4.12

  1. There exists a constant C > 0 which is independent of σ > 0 such that

    (4.10) F ( M + D 2 z σ ( x ) ) k Z n A γ ( k , ω ) ν σ ( x k ) C ,

    in A.

  2. There exists a constant C > 0 which is independent of σ > 0 such that

    F ( M + D 2 z σ ) F ( M + D 2 w σ ) C ,

    in A.

  3. There exists a constant C > 0 which is independent of σ > 0 such that

    (4.11) z σ w σ L ( A ) C .

  4. There exists a subsequence { w σ m } m = 1 and a limit function w C ( A k Z n { k } ) such that w σ m w when σ 0 + uniformly on every compact subset of A k Z n { k } .

Proof

  1. Since F is uniformly elliptic and positively homogeneous of degree one, we have

    F ( M + D 2 z σ ) γ ( k 0 ) ν σ ( x k ) + P + ( M ) + k k 0 γ ( k ) P + ( D 2 Φ ( x k ) ) ,

    where x k 0 1 / 3 for k 0 Z n A . Since there exists a positive constant γ ¯ > 0 such that γ ( k ) γ ¯ , there exists a constant C = C ( F , A , M , γ ¯ ) such that

    P + ( M ) + k k 0 γ ( k ) P + ( D 2 Φ ( x k ) ) C ,

    where x k 0 1 / 3 for k 0 Z n A . Similarly, we also have

    F ( M + D 2 z σ ) P + ( M ) + k Z n γ ( k ) P + ( D 2 Φ ( x k ) ) C ,

    where x k > 1 / 3 for any k Z n A . We can apply the same argument for finding the lower bound of F ( M + D 2 z σ ) , and thus, we conclude the desired result (4.10).

  2. It follows directly from the part (i) and the definition of w σ .

  3. We may assume x 0 = 0 A and A B l ( x 0 ) for some l > 0 . Note that

    F D 2 z σ + 1 2 x T M x + C 2 n λ ( x 2 l 2 ) F ( M + D 2 z σ ) + P C n λ I = F ( M + D 2 z σ ) + C F ( M + D 2 w σ ) = F D 2 w σ + 1 2 x T M x .

    Moreover, by the construction of Φ σ , z σ 1 = z σ 2 on A for any σ 1 , σ 2 > 0 . Thus, there exists a constant C independent of σ such that z σ C on A . Therefore, the comparison principle leads to

    z σ C + 1 2 x T M x + C 2 n λ ( x 2 l 2 ) w σ + 1 2 x T M x ,

    which implies that

    z σ w σ C .

    From the same argument, we derive (4.11).

  4. Let K be a compact subset of A k Z n { k } . Again by the construction of Φ σ , we have z σ ( x ) = k Z n γ ( k ) Φ ( x k ) in K for any sufficiently small σ > 0 . In other words, the function z σ in K is independent of σ > 0 (if it is sufficiently small). Due to part (iii), we have a uniform L -bound for w σ : w σ L ( K ) C , and F ( D 2 w σ ) = β in K . Hence, application of Arzela-Ascoli theorem together with the interior C α -estimate and standard diagonal process ensures the existence of a convergent subsequence and corresponding limit function.□

Theorem 4.13

[Convergence of free solutions] There exists a unique limit function w C ( A k Z n { k } ) such that w σ w when σ 0 + uniformly on every compact subset of A k Z n { k } . Moreover, w satisfies

F ( M + D 2 w ) = β + F ( M ) i n A k Z n { k } , w = 0 , o n A ,

and

lim x k w ( x , ω ) Φ ( x k ) = γ ( k , ω ) ,

for any k Z n A .

Proof

According to Lemma 4.12 (iv), there exists a limit function w C ( A k Z n { k } ) such that w σ m w when m uniformly on every compact subset of A k Z n { k } . Recalling Proposition 2.9. in [6] (the stability of viscosity solutions), we deduce that w is a viscosity solution of

F ( M + D 2 w ) = β + F ( M ) in A k Z n { k } , w = 0 , on A .

Moreover, since z σ ( k ) when σ 0 + for k Z n A with γ ( k ) > 0 , Lemma 4.12 (iii) yields that w has an isolated singularity at each k Z n A whenever γ ( k ) > 0 . Thus, applying Theorem 4.11 for an isolated singularity k Z n A , we have

(4.12) lim x k w ( x ) Φ ( x k ) = a ,

for some positive constant a > 0 .

Now we claim that

a = γ ( k ) .

Indeed, for ε > 0 , there exists δ > 0 such that

0 < x k < δ ( a ε ) Φ ( x k ) w ( x ) ( a + ε ) Φ ( x k ) ,

by (4.12). Let { w σ m } m = 1 be a subsequence such that w σ m w uniformly on every compact subset of A k Z n { k } . Then for any m N large enough, we have

w ( x ) w σ m ( x ) 1 , w σ m ( x ) z σ m ( x ) C ,

for min ( δ / 2 , ε ) < x k < δ . Note that C is independent of m . (see Lemma 4.12 (iii)). Recalling the definition of z σ , for m large enough, we have

γ ( k ) Φ ( x k ) z σ m ( x ) γ ( k ) Φ ( x k ) + C ,

for min ( δ / 2 , ε ) < x k < δ and C > 0 , which is independent of m . Combining these estimates together, we conclude that

( a ε ) Φ ( x k ) C γ ( k ) Φ ( x k ) ( a + ε ) Φ ( x k ) + C .

Dividing by Φ ( x k ) and letting ε 0 lead to a = γ ( k ) , as desired.

Finally, it only remains to prove the uniqueness of limit functions. For this purpose, let w ¯ , w ̲ be two limit functions of { w σ } σ > 0 . Then both w ¯ and w ̲ are viscosity solutions of

F ( M + D 2 w ) = β + F ( M ) in A k Z n { k } , w = 0 , on A .

Moreover, the aforementioned argument allows us to capture the behavior of limit functions near an isolated singularity, namely:

lim x k w ¯ ( x ) Φ ( x k ) = γ ( k ) = lim x k w ̲ ( x ) Φ ( x k ) .

Fix ε > 0 . Then there exists small enough δ > 0 such that

( 1 + ε ) w ¯ w ̲ ,

for x k = δ with k Z n A . Employing a similar argument as in the proof of Lemma 4.10 and Theorem 4.11, there exist constants c 1 , c 2 > 0 such that

F ( M + D 2 ( 1 + ε ) w ¯ + β ε ( c 1 c 2 x 2 ) ) β + F ( M ) = F ( M + D 2 w ̲ ) in A k B δ ( k )

and

( 1 + ε ) w ¯ + β ε ( c 1 c 2 x 2 ) w ̲ on ( A k B δ ( k ) ) .

Applying the comparison principle and letting ε 0 , we have w ¯ w ̲ and by the symmetry, we conclude that w ¯ = w ̲ .□

4.2.2 Obstacle solutions

For notational simplicity, we write an obstacle solution

v σ ( x , ω ) v β , σ , A ; M ( x , ω ) ,

where β 0 , A R n , M S n are fixed and σ > 0 . Recall that we defined

v β , σ , A ; M ( x , ω ) inf v ( x ) : F ( M + D 2 v ( x ) ) β + F ( M ) k Z n A γ ( k , ω ) ν σ ( x k ) in A , v 0 in A , v = 0 on A } .

The convergence of obstacle solutions { v σ } σ > 0 can be achieved if we exploit the result for free solutions.

Lemma 4.14

  1. We have

    0 F ( M + D 2 w σ ) F ( M + D 2 v σ ) β ,

    in A .

  2. There exists a constant C > 0 which is independent of σ > 0 such that

    z σ v σ L ( A ) C .

  3. There exists a subsequence { v σ m } m = 1 and a limit function v C ( A k Z n { k } ) such that v σ m v when σ 0 + uniformly on every compact subset of A k Z n { k } .

Proof

  1. Since

    F ( M + D 2 v σ ) = F ( M ) + β k Z n A γ ( k , ω ) ν σ ( x k ) χ { v σ > 0 } ,

    we have

    F ( M + D 2 w σ ) F ( M + D 2 v σ ) = β k Z n A γ ( k , ω ) ν σ ( x k ) χ { v σ = 0 } = β χ { v σ = 0 } .

    Here we used that ν σ > 0 near k Z n with γ ( k , ω ) > 0 , recalling the proof for Lemma 4.5.

  2. It follows from the comparison principle (similarly as in the proof of Lemma 4.12 (iii)) and (4.11).

  3. See the proof of Lemma 4.12 (iv).□

Theorem 4.15

(Convergence of obstacle solutions). There exists a unique limit function v C ( A k Z n { k } ) such that v σ v when σ 0 + uniformly on every compact subset of A k Z n { k } . Moreover, v satisfies

(4.13) F ( M + D 2 v ) = F ( M ) + β χ { v > 0 } i n A k Z n { k } , v 0 i n A k Z n { k } , v = 0 , o n A ,

and

lim x k v ( x , ω ) Φ ( x k ) = γ ( k , ω ) ,

for any k Z n A .

Proof

The most part of the proof is the same as the proof of Theorem 4.13, which is an application of the uniform convergence obtained in the previous lemma, the stability of obstacle problems, and the isolated singularity theorem, Theorem 4.11.

Again it only remains to prove the uniqueness part. Let v ¯ , v ̲ be two limit functions of { v σ } σ > 0 . Since v σ behaves like z σ (or Φ σ ) near k Z n , Theorem 4.11 implies that

lim x k v ¯ ( x ) Φ ( x k ) = γ ( k ) = lim x k v ̲ ( x ) Φ ( x k ) .

Fix ε > 0 . Then there exists small enough δ > 0 such that

( 1 + ε ) v ¯ v ̲ ,

for x k = δ with k Z n A . Similarly as in the proof of Theorem 4.13, there exist constants c 1 , c 2 > 0 such that

F ( M + D 2 ( ( 1 + ε ) v ¯ + β ε ( c 1 c 2 x 2 ) ) ) β + F ( M ) in A k B δ ( k ) ,

( 1 + ε ) v ¯ + β ε ( c 1 c 2 x 2 ) v ̲ on ( A k B δ ( k ) ) ,

and

( 1 + ε ) v ¯ + β ε ( c 1 c 2 x 2 ) 0 in A k B δ ( k ) .

Note that v ̲ can be written as the unique solution of the following obstacle problem:

inf { v ( x ) : F ( M + D 2 v ( x ) ) β + F ( M ) in A k B δ ( k ) , v 0 in A k B δ ( k ) , v v ̲ on k B δ ( k ) , v 0 on A . }

Since the function ( 1 + ε ) v ¯ + β ε ( c 1 c 2 x 2 ) is admissible for the obstacle problem above, we have

( 1 + ε ) v ¯ + β ε ( c 1 c 2 x 2 ) v ̲ .

Letting ε 0 , we have v ¯ v ̲ and by the symmetry, we conclude that v ¯ = v ̲ .□

Now as we have done in the Laplacian case, we are able to define the measure of contact set for an obstacle problem and to determine the critical value β 0 . Indeed, for any M S n , we define a random variable m β , A ; M by

m β , A ; M { x A : v β , A ; M = 0 } .

Lemma 4.16

(A subadditive quantity)

  1. The random variable m β , A ; M is subadditive.

  2. The process

    T k m β , A ; M = m β , k + A ; M

    has the same distribution for all k Z n .

Proof

See the proof of Lemma 4.7.□

Due to the previous lemma, applying a subadditive ergodic theorem, we have:

l ( β ; M ) = lim t m β , B t ; M B t a.s.

and a scaled version:

l ( β ; M ) = lim ε 0 { y : v ¯ β ; M ε ( y , ω ) = 0 } B 1 a.s. ,

where v ¯ β ; M ε ( y , ω ) ε 2 v β , ε 1 B ; M ( y / ε , ω ) .

Lemma 4.17

(Properties of l ( β ; M ) ) Let M S n .

  1. l ( β ; M ) is nondecreasing function with respect to β .

  2. If β 0 , then l ( β ; M ) = 0 .

  3. If β > 0 is large enough, then l ( β ; M ) > 0 .

Proof

See the proof of Lemma 4.8.□

Finally, we let the (nonnegative) critical value

β 0 ( M ) sup { β : l ( β ; M ) = 0 } ,

which is well-defined by the previous lemma.

5 Properties of free solutions and obstacle solutions

In short, the argument in the previous section enables us to show the convergence of obstacle functions { v β , σ , A } σ > 0 and free functions { w β , σ , A } σ > 0 when σ 0 ; so we could define the critical value β 0 . Note that, in the Laplacian case, we have a further information such that w β , σ 1 , A = w β , σ 2 , A and v β , σ 1 , A = v β , σ 2 , A if x k Z n ( B a ¯ σ 2 ( k ) ) c and 0 < σ 1 σ 2 .

In this section, we first extract useful properties (namely, (P1) and (P2) in Section 1) of limit obstacle solutions and free solutions by investigating the behaviors of approximated solutions (whose parameter is given by σ ). Here we should carefully check whether the auxiliary functions are rescaled or not. Then we define the corrector in terms of the critical value β 0 and transport the desired properties for correctors by comparing to the auxiliary functions. Finally, we end up with our main homogenization result employing the correctors.

Again we justify each step above with respect to the Laplacian operator Δ first, and then to the general fully nonlinear operator F .

5.1 Laplacian operator

We begin with the step which illustrates the behavior of an obstacle solution and a free solution away from perforated holes, when ε 0 . In other words, we are going to show that for the critical value β 0 , we have

lim ε 0 w ¯ β 0 ε = 0 ,

away from holes (which will be precisely stated later). Note that

  1. we split two cases depending on the value l ( β ) , more precisely,

    • if l ( β ) = 0 , i.e., v ¯ β ε never meet the (zero) obstacle, then we expect that v ¯ β ε > 0 in D ;

    • if l ( β ) > 0 , i.e., v ¯ β ε meets the (zero) obstacle in some region, then we expect that v ¯ β ε = 0 occurs throughout the whole domain (we will prove the “spreading effect” of contact point);

  2. we first prove for an obstacle solution v ¯ β ε and transport this information to a free solution w ¯ β ε .

Lemma 5.1

If l ( β ) = 0 , then

liminf ε 0 w ¯ β ε 0 ,

in D.

Proof

We may assume D = B 1 and write v ¯ β , σ ε ( x , ω ) = ε 2 v β , σ , ε 1 B 1 ( x / ε , ω ) and w ¯ β , σ ε ( x , ω ) = ε 2 w β , σ , ε 1 B 1 ( x / ε , ω ) . First recall that for each fixed ε > 0 , we have v β , σ , ε 1 D v β , ε 1 B 1 when σ 0 + . Moreover, v β , σ , ε 1 B 1 v β , ε 1 B 1 only can occur in k Z n B a ¯ σ ( k ) , where a ¯ σ = σ 2 n 2 . Thus, for any sufficiently small σ > 0 , the coincidence sets are identical:

(5.1) { v ¯ β ε = 0 } = { v ¯ β , σ ε = 0 } .

On the other hand, recalling (4.4), we obtain

v β , σ , ε 1 B 1 ( x ) > 0 if x k Z n B a ¯ σ ( k ) ,

which yields

Δ ( w β , σ , ε 1 B 1 v β , σ , ε 1 B 1 ) = β k Z n γ ( k ) ν σ ( x k ) χ { v β , σ , ε 1 B 1 = 0 } = β χ { v β , σ , ε 1 B 1 = 0 } .

Applying the Alexandrov-Backelman-Pucci estimate (e.g., [6]), we obtain that

sup B ε 1 ( v β , σ , ε 1 B 1 w β , σ , ε 1 B 1 ) C ε 1 B ε 1 ( β χ { v β , σ , ε 1 B 1 = 0 } ) n 1 / n = C β ε 1 { v β , σ , ε 1 B 1 = 0 } B ε 1 1 / n .

By rescaling, we have

sup B 1 ( v ¯ β , σ ε w ¯ β , σ ε ) C β { v ¯ β , σ ε = 0 } B 1 1 / n ,

or equivalently,

w ¯ β , σ ε v ¯ β , σ ε C β { v ¯ β , σ ε = 0 } 1 n C β { v ¯ β , σ ε = 0 } 1 n ,

in B 1 . Letting σ 0 + and applying (5.1), we have

w ¯ β ε C β { v ¯ β ε = 0 } 1 n ,

in B 1 . Finally, since

0 = l ( β ) = lim ε 0 { v ¯ β ε = 0 } B 1 ,

we conclude that

liminf ε 0 w ¯ β ε 0 ,

in B 1 .□

Next, to estimate the upper bound for w ¯ β ε when l ( β ) > 0 , we study the quadratic growth property of an obstacle problem. For this purpose, let u L ( D ) be a nonnegative solution of

(5.2) Δ u = f ( x ) χ { u > 0 } in D ,

for f L ( D ) . For an open set D ( u ) = { u > 0 } , we define the free boundary

Γ ( u ) D ( u ) D .

The following lemma explains the quadratic growth of the solution for an obstacle problem near the free boundary. In other words, the solution has the optimal C 1 , 1 -regularity.

Lemma 5.2

(Quadratic growth) Let u L ( D ) , u 0 , satisfy (5.2), x 0 Γ ( u ) , and B 2 r ( x 0 ) D . Then there exists a constant C = C ( n ) > 0 such that

sup B r ( x 0 ) u C f L ( D ) r 2 .

Proof

See the proof of Lemma 5.14, which deals with the same result for the fully nonlinear operator.□

Lemma 5.3

If l ( β ) > 0 , then we have

lim ε 0 v ¯ β , σ ε = 0 in D ,

for each sufficiently small σ > 0 .

Proof

We may assume D = Q 1 , where Q r [ r / 2 , r / 2 ] n is a cube of width r > 0 . By (5.1), we have

(5.3) l σ ( β ) lim ε 0 { x Q 1 : v ¯ β , σ ε = 0 } Q 1 = l ( β ) > 0

for any sufficiently small σ > 0 . Here we fix a sufficiently small σ > 0 , and for any m N , we split Q 1 into 2 m n smaller cubes of equal size, whose width is exactly 1 / 2 m . For Q being any of these cubes, we have v ¯ β , σ ε 0 on Q . Thus, applying the comparison principle in Q , we have

(5.4) lim ε 0 { x Q : v ¯ β , σ ε = 0 } Q l σ ( β ) .

Then from (5.3) and (5.4), we deduce that for sufficiently small ε > 0 , the set { x Q : v ¯ β , σ ε = 0 } is nonempty for any smaller cubes Q . In other words, we have shown the contact set { v ¯ β , σ ε = 0 } “spreads” all over the Q 1 .

For a smaller cube Q (with width 1 / 2 m ) such that Q Q 1 = , the above result yields that there exists a point x 0 Q such that v ¯ β , σ ε ( x 0 ) = 0 for any sufficiently small ε > 0 . Recalling the definition of the obstacle solution v ¯ β , σ ε , we have

Δ v ¯ β , σ ε = β k Z n γ ( k ) ν σ ( x / ε k ) χ { v ¯ β , σ ε > 0 } in B 1 .

Thus, by applying Lemma 5.2, we have

sup Q v ¯ β , σ ε sup B n 2 m ( x 0 ) v ¯ β , σ ε C n ( β + c γ ¯ σ 2 n n 2 ) n 2 2 m ,

for any smaller cubes Q such that Q Q 1 = . Thus, letting m and then choosing sufficiently small ε = ε ( m ) > 0 , we conclude that

lim ε 0 v ¯ β , σ ε = 0 ,

in Q 1 .□

Since w ¯ β ε v ¯ β ε by their definitions and v ¯ β , σ ε = v ¯ β ε in D k Z n B ε a ¯ σ ( ε k ) , we deduce the following corollary:

Corollary 5.4

Let q ( 0 , 1 / 2 ) . If l ( β ) > 0 , then

lim ε 0 sup D k Z n B q ε ( ε k ) w ¯ β ε 0 ,

in D.

Applying Lemma 5.1, Corollary 5.4, and the fact that β 0 is the critical value, we have:

Corollary 5.5

Let B η ( x 0 ) D . Then there exists a sequence { ε j } j N such that ε j 0 + and

lim j w ¯ β 0 ε j ( x ) = 0 ,

for any x B η ( x 0 ) { x 0 } .

Next, we study the asymptotic behavior of auxiliary functions near the boundary of holes T ε . More precisely, we need to show that

w ¯ β 0 ε = 1 + o ( 1 ) , on T ε .

In short, this can be done by comparing the auxiliary functions with the normalized homogeneous solution.

Lemma 5.6

For k Z n , we denote

h β , σ , k ( x ) β 2 n x k 2 + γ ( k , ω ) Φ σ ( x k ) .

  1. For every β and for every k Z n , we have

    v β , σ , ε 1 D ( x ) h β , σ , k ( x ) β 2 n r ( k , ω ) n 2 ,

    for all x B 1 ( k ) and almost every ω Ω .

  2. For every β > β 0 , we have

    v β , σ , ε 1 D ( x ) h β , σ , k ( x ) + o ( ε 2 ) ,

    for all x B 1 / 2 ( k ) and almost every ω Ω .

Proof

We refer to the proof of Lemma 4.3 in [8]. The only difference arises from that we are considering functions with σ -dependence, which does not change the proof. Otherwise, it can also be shown by applying the comparison principle and Lemma 5.3; indeed, see the proof of Lemma 5.18 later in the fully nonlinear case.□

We denote

h β , k ( x ) lim σ 0 h β , σ , k ( x ) = β 2 n x k 2 + γ ( k , ω ) Φ ( x k )

and

h β , k ε ( x ) ε 2 h β , k ( x / ε ) .

Then since h β , k B a ¯ ε ( r ( k , ω ) ) ( k ) = ε 2 + β 2 n a ¯ ε ( r ( k , ω ) ) 2 , we deduce the following corollary by letting σ 0 + in Lemma 5.6:

Corollary 5.7

  1. For every β and k Z n such that r ( k , ω ) > 0 , we have

    v β , ε 1 D ( x ) ε 2 + O ( 1 ) on B a ¯ ε ( r ( k , ω ) ) ( k ) a.e. ω Ω ,

    and so

    v ¯ β ε ( x ) 1 + o ( 1 ) on T ε ( ω ) a.e. ω Ω ,

    for all β .

  2. For every β > β 0 and every k Z n , we have

    v β , ε 1 D ( x ) ε 2 + o ( ε 2 ) on B a ¯ ε ( r ( k , ω ) ) ( k ) a.e. ω Ω ,

    and so

    v ¯ β ε ( x ) 1 + o ( 1 ) on T ε ( ω ) a.e. ω Ω ,

    for all β > β 0 .

Lemma 5.8

For every k Z n , w ¯ β 0 ε satisfies

h β 0 , k ε ( x ) o ( 1 ) w ¯ β 0 ε ( x ) h β 0 , k ε ( x ) + o ( 1 ) x B ε / 2 ( ε k ) D a.e. ω Ω .

In particular,

w ¯ β 0 ε = 1 + o ( 1 ) on T ε D .

Proof

Recall that for every β , we denote

v ¯ β ε ( x ) = ε 2 v β , ε 1 D ( x / ε ) ,

which is defined in D and v ¯ β ε = 0 on D .

  1. Let β > β 0 . Note that

    Δ w β 0 , σ , ε 1 D = β 0 k Z n ε 1 D γ ( k , ω ) ν σ ( x k )

    and

    Δ v β , σ , ε 1 D = β k Z n ε 1 D γ ( k , ω ) ν σ ( x k ) χ { v β , σ , ε 1 D > 0 } ,

    in ε 1 D . Thus, we have

    Δ ( w β 0 , σ , ε 1 D v β , σ , ε 1 D ) β 0 β

    and w β 0 , σ , ε 1 D v β , σ , ε 1 D = 0 on ( ε 1 D ) . By rescaling, we obtain

    Δ ( w ¯ β 0 , σ ε v ¯ β , σ ε ) β 0 β

    and w ¯ β 0 , σ ε v ¯ β , σ ε = 0 on D . The Green representation formula yields that

    w ¯ β 0 , σ ε ( x 0 ) v ¯ β , σ ε ( x 0 ) D G ( x , x 0 ) ( β 0 β ) d x ( β β 0 ) D Φ ( x x 0 ) d x O ( β β 0 ) ,

    where G ( , ) is the Green function on D . ( Δ G ( , x 0 ) = δ x 0 and G = 0 on D .) Applying Lemma 5.6 (ii), we conclude that

    w ¯ β 0 , σ ε ( x ) ε 2 h β , σ , k ( x / ε ) + o ( 1 ) + O ( β β 0 )

    for all x B ε / 2 ( ε k ) . Letting σ 0 and then β β 0 , we obtain the desired upper bound.

  2. Arguing as in (i), for every β β 0 , we have

    Δ ( v β , σ , ε 1 D w β 0 , σ , ε 1 D ) = β β 0 β χ { v β , σ , ε 1 D = 0 } .

    By rescaling, we obtain that

    Δ ( v ¯ β , σ ε w ¯ β 0 , σ ε ) = β β 0 β χ { v ¯ β , σ ε = 0 } .

    Again, the Green representation formula leads to

    v ¯ β , σ ε w ¯ β 0 , σ ε O ( β 0 β ) + C β { v ¯ β , σ ε = 0 } 1 / ( n 1 ) in D .

    Applying Lemma 5.6 (i), we conclude that

    w ¯ β 0 , σ ε ( x ) ε 2 h β , σ , k ( x / ε ) o ( 1 ) O ( β 0 β ) C β { v ¯ β , σ ε = 0 } 1 / ( n 1 ) .

    Letting σ 0 (see (5.1))

    w ¯ β 0 ε ( x ) h β , k ε ( x ) o ( 1 ) O ( β 0 β ) C β { v ¯ β ε = 0 } 1 / ( n 1 ) .

    Since lim ε 0 { v ¯ β ε = 0 } = l ( β ) D = 0 for any β β 0 , we obtain the desired lower bound.□

Now we define a corrector:

(5.5) Δ w ε ( x , ω ) = β 0 for x D T ε , w ε ( x , ω ) = 1 for x T ε , w ε ( x , ω ) = 0 for x D T ε .

Lemma 5.9

Let B η ( x 0 ) D . Then there exists a sequence { ε j } j N such that ε j 0 + and

lim j w ε j ( x ) = 0 ,

for any x B η ( x 0 ) { x 0 } .

Proof

Since

Δ w ¯ β 0 , σ ε = β 0 in D k B ε a ¯ σ ( ε k ) ,

letting σ = ε yields that

Δ w ¯ β 0 ε = Δ w ¯ β 0 , ε ε = β 0 in D k B a ε ( ε k ) = D T ε .

Recalling Lemma 5.8 and applying the comparison principle for w ε and w ¯ β 0 ε in D T ε , we have

w ¯ β 0 ε ( x ) o ( 1 ) w ε ( x ) w ¯ β 0 ε ( x ) + o ( 1 ) in D T ε .

Therefore, the desired result follows from Corollary 5.5.□

Finally, we are ready to finish the proof of our main theorem for the Laplacian case:

Proof of Theorem 1.1 (ii)

We are going to show that u is a subsolution. Let us assume that there is a parabola P touching u from above at x 0 and

Δ P + β 0 ( φ ( x 0 ) P ( x 0 ) ) + < 2 μ 0 < 0 .

In a small neighborhood of x 0 , B η ( x 0 ) , there exists another parabola Q such that

D 2 P < D 2 Q in B η ( x 0 ) , P ( x 0 ) > Q ( x 0 ) + δ 0 , P ( x ) < Q ( x ) on B η ( x 0 ) .

In addition, we can choose a sufficiently small ε 0 > 0 so that Q satisfies

Δ Q + β 0 ( φ ( x 0 ) u ( x 0 ) + 3 ε 0 ) Δ Q + β 0 ξ 0 < μ 0 < 0

and

Q ( x ) Q ( x 0 ) + φ ( x ) φ ( x 0 ) < ε 0 ,

for x B η ( x 0 ) . Let us consider

Q ε ( x ) Q ( x ) + w ε ( x ) ξ 0 .

Then by the definition of w ε , we have

Δ Q ε < μ 0 < 0

and

Q ε ( x ) = Q ( x ) + ξ 0 = Q ( x ) + φ ( x 0 ) u ( x 0 ) + 3 ε 0 > φ ( x ) ,

on T a ε B η ( x 0 ) . Therefore, the maximum principle yields that Q ε φ ε in B η ( x 0 ) .

We now define the function

v ε min ( u ε , Q ε ) in B η ( x 0 ) , u ε in D B η ( x 0 ) .

Applying Lemma 5.9, for sufficiently small ε > 0 (at least for a subsequence { ε j } ), we have Q ε > u ε on B η ( x 0 ) . Thus, the function v ε is well-defined and will be a viscosity supersolution of ( L ε ) . Since u ε is the least viscosity supersolution, we have

u ε v ε Q ε in B η ( x 0 ) .

Letting ε 0 , we have

u ( x 0 ) Q ( x 0 ) < P ( x 0 ) = u ( x 0 ) ,

which is a contradiction. By an argument similar to the proof of Lemma 4.1 in [7], we can show that u is also a viscosity supersolution of ( L ¯ ) (or see the proof of Theorem 1.2 (ii) which deals with the argument for a viscosity supersolution of ( F ¯ ) ).□

5.2 Fully nonlinear operator

In the Laplacian case, the strong properties of the obstacle solutions (see Lemma 4.5 and its proof) immediately led to the equality between coincidence sets, (5.1): for any sufficiently small σ > 0 ,

{ v ¯ β , σ ε = 0 } = { v ¯ β ε = 0 } .

On the other hand, in the fully nonlinear case, we only have the uniform convergence of the obstacle solutions and we require some auxiliary lemmas to derive the stability of coincidence sets, which is a weaker consequence compared to (5.1).

We begin with a simple lemma:

Lemma 5.10

For an open set D R n , let { u m } m = 1 , and u 0 be continuous functions on D. If u m u 0 uniformly on every compact subset of D as m , then

limsup m { u m = 0 } { u 0 = 0 } .

Proof

Suppose that

limsup m { u m = 0 } > { u 0 = 0 } .

Then there exist an open neighborhood A of { u 0 = 0 } and a compact set K D such that ( { u m = 0 } K ) A is nonempty (up to subsequence, if necessary). In other words, there exists a sequence of points x m K satisfying x m { u m = 0 } A . Again, up to subsequence, there exists a point x 0 K A such that x m x 0 . Then for any ε > 0 , we have

u m ( x m ) u 0 ( x 0 ) u m ( x m ) u 0 ( x m ) + u 0 ( x m ) u 0 ( x 0 ) < ε ,

for sufficiently large m , which yields that

u 0 ( x 0 ) = lim m u m ( x m ) = 0 .

This contradicts to x 0 { u 0 = 0 } and so we have the desired inequality.□

Next, for the other direction of the inequality obtained from Lemma 5.10, we need an additional work in terms of obstacle problem theory. Let u be a nonnegative solution of an obstacle problem

F ( D 2 u ) = f ( x ) χ { u > 0 } in D ,

for f L ( D ) . By Theorem 1.2.1 in [21], we have u C loc 1 , 1 ( D ) C ( D ¯ ) . We set

D ( u ) { x D : u ( x ) > 0 } , C ( u ) { x D : u ( x ) = 0 } , Γ ( u ) D ( u ) D .

Lemma 5.11

(Nondegeneracy; Lemma 3.4, [21]) Suppose that f M > 0 in D and let x 0 be any point in D ( u ) ¯ . Then for any ball B r ( x 0 ) D ,

sup B r ( x 0 ) [ u ( x ) u ( x 0 ) ] M r 2 2 n Λ .

Lemma 5.12

Suppose that f M > 0 in D and u 0 , u m , m N are nonnegative solutions of

F ( D 2 v ) = f ( x ) χ { v > 0 } in D .

If u m u 0 uniformly in every compact subset of D, then we have

  1. limsup m D ( u m ) D ( u 0 ) ¯ ;

  2. limsup m D ( u m ) D ( u 0 ) ¯ ;

  3. liminf m C ( u m ) C ( u 0 ) ,

where limsup m A m means the set of all limit points of sequences { x m } , x m A m .

Proof

  1. It is a consequence of Lemma 5.11; see Corollary 3.5 in [21].

  2. Its proof is similar to the one of Lemma 5.10. Indeed, we suppose that

    limsup m D ( u m ) > D ( u 0 ) ¯ .

    Then there exists an open neighborhood A of D ( u 0 ) ¯ such that there is a sequence of points x m D (up to subsequence, if necessary) satisfying x m D ( u m ) A . Then there is a point x 0 D such that x m x 0 , which implies that

    x 0 limsup m D ( u m ) A limsup m D ( u m ) D ( u 0 ) ¯ .

    This contradicts to (i).

  3. Since D ( u m ) = D C ( u m ) and D ( u 0 ) ¯ = ( D C ( u 0 ) ) Γ ( u 0 ) ,

    liminf m C ( u m ) C ( u 0 ) Γ ( u 0 )

    follows from (ii). Note that when F is positively homogeneous of degree one, the free boundary Γ ( u 0 ) is a C 1 , α -graph; see Theorem 3.3 in [21]. In particular, Γ ( u 0 ) = 0 , and thus, the desired inequality follows.□

Combining the results from Lemmas 5.10 and 5.12 (iii), we have the stability of coincidence sets; i.e., since v β , σ ε converges to v β ε uniformly on every compact subset of D as σ 0 , we have

(5.6) lim σ 0 { v β , σ ε = 0 } = { v β ε = 0 } .

Now, we proceed similarly as in the Laplacian case:

Lemma 5.13

If l ( β ) = 0 , then

liminf ε 0 w ¯ β ε 0 ,

in D.

Proof

We may assume D = B 1 and write v ¯ β , σ ε ( x , ω ) = ε 2 v β , σ , ε 1 B 1 ( x / ε , ω ) and w ¯ β , σ ε ( x , ω ) = ε 2 w β , σ , ε 1 B 1 ( x / ε , ω ) . Recalling the proof for Lemma 4.5, we obtain

v β , σ , ε 1 B 1 ( x ) > 0 if x k Z n B a ¯ σ ( k ) ,

which yields

F ( M + D 2 w β , σ , ε 1 B 1 ) F ( M + D 2 v β , σ , ε 1 B 1 ) = β k Z n γ ( k ) ν σ ( x k ) χ { v β , σ , ε 1 B 1 = 0 } = β χ { v β , σ , ε 1 B 1 = 0 } .

Now let h be the solution of

P + ( D 2 h ) = β χ { v β , σ , ε 1 B 1 = 0 } in ε 1 B 1 , h = 0 on ( ε 1 B 1 ) .

Then the uniformly ellipticity of F yields that

F ( M + D 2 ( w β , σ , ε 1 B 1 + h ) ) F ( M + D 2 w β , σ , ε 1 B 1 ) + P + ( D 2 h ) = F ( M + D 2 v β , σ , ε 1 B 1 ) ,

and thus, we have w β , σ , ε 1 B 1 + h v β , σ , ε 1 B 1 in ε 1 B 1 by the comparison principle (e.g., see [11]). Applying the Alexandrov-Backelman-Pucci estimate for h , we obtain that

sup B ε 1 h C ε 1 B ε 1 ( β χ { v β , σ , ε 1 B 1 = 0 } ) n 1 / n = C β ε 1 { v β , σ , ε 1 B 1 = 0 } B ε 1 1 / n .

Now the remaining part is the same as in the proof of Lemma 5.1. The only difference arises when applying the stability of coincidence sets, (5.6).□

Similarly as in the Laplacian case (Lemma 5.2), one can prove the quadratic growth property in the fully nonlinear case:

Lemma 5.14

(Quadratic growth; fully nonlinear operator) Let u be a nonnegative solution of an obstacle problem, i.e., u satisfies

F ( D 2 u ) = f ( x ) χ { u > 0 } in D ,

for f L ( D ) . If x 0 Γ ( u ) , and B 2 r ( x 0 ) D , then

sup B r ( x 0 ) u C ( n , λ , Λ ) f L ( D ) r 2 .

Proof

For simplicity, we may assume x 0 = 0 and write B R = B R ( 0 ) for any R > 0 . Then we split u into the sum u 1 + u 2 in B 2 r , where

F ( D 2 u 1 ) = F ( D 2 u ) , P + ( D 2 u 2 ) = 0 in B 2 r ; u 1 = 0 , u 2 = u on B 2 r .

We estimate these functions u 1 and u 2 separately.

  1. To estimate u 1 , we consider a barrier function

    g ( x ) 1 2 n ( 4 r 2 x 2 ) .

    Then we immediately obtain

    F ( D 2 g ) = F 1 n I λ

    in B 2 r and g = 0 on B 2 r . Thus, the comparison principle yields that for x B 2 r ,

    u 1 ( x ) M λ g ( x ) C ( n , λ ) M r 2 ,

    where M f L ( D ) . Considering g + ( x ) 1 2 n ( x 2 4 r 2 ) ( = g ( x ) ) , we can conclude that

    u 1 ( x ) C ( n , λ ) M r 2 .

  2. To estimate u 2 , note that u 2 0 in B 2 r since F ( D 2 u 2 ) = 0 in B 2 r and u 2 = u 0 on B 2 r . Moreover, since 0 Γ ( u ) , we have

    u 2 ( 0 ) = u 1 ( 0 ) C ( n , λ ) M r 2 ,

    by the previous result. Thus, applying the Harnack inequality to a nonnegative function u 2 in B 2 r , we conclude that for x B r ,

    u 2 ( x ) C ( n , λ , Λ ) u 2 ( 0 ) C ( n , λ , Λ ) M r 2 .

    Finally, combining the estimates for u 1 and u 2 , we obtain the desired estimates for u .□

Lemma 5.15

If l ( β ) > 0 , then we have

lim ε 0 v ¯ β , σ ε = 0 in D ,

for each sufficiently small σ > 0 .

Proof

Its proof can be done by following the proof of Lemma 5.3. Note that here we use the stability of coincidence sets and the quadratic separation occurred at the contact point.□

Since w ¯ β ε v ¯ β ε by their definitions and v ¯ β , σ ε uniformly converges to v ¯ β ε on every compact subset of D k Z n ( ε k ) , we deduce the following corollary:

Corollary 5.16

Let q ( 0 , 1 / 2 ) . If l ( β ) > 0 , then

lim ε 0 sup D k Z n B q ε ( ε k ) w ¯ β ε 0 ,

in D.

Applying Lemma 5.13, Corollary 5.16, and the fact that β 0 is the critical value, we have:

Corollary 5.17

Let B η ( x 0 ) D . Then there exists a sequence { ε j } j N such that ε j 0 + and

lim j w ¯ β 0 ε j ( x ) = 0 ,

for any x B η ( x 0 ) { x 0 } .

Lemma 5.18

For k Z n and M S n , we denote

h β , σ , k ; M ( x ) β + F ( M ) 2 n λ x k 2 + γ ( k , ω ) Φ σ ( x k ) 1 2 ( x k ) T M ( x k )

and

h σ , k ; M + ( x ) F ( M ) 2 n Λ x k 2 + γ ( k , ω ) Φ σ ( x k ) 1 2 ( x k ) T M ( x k ) .

  1. For every β , we have

    v β , σ , ε 1 D ; M ( x ) h β , σ , k ; M ( x ) β 8 n λ ( 2 r ( k , ω ) ) α 1 8 M ,

    for all x B 1 / 2 ( k ) and almost every ω Ω .

  2. For every β > β 0 , we have

    v β , σ , ε 1 D ; M ( x ) h σ , k ; M + ( x ) + o ( ε 2 ) ,

    for all x B 1 / 2 ( k ) and almost every ω Ω .

Proof

  1. By a direct calculation, for x B 1 / 2 ( k ) , we have

    F ( M + D 2 h β , σ , k ; M ) = F β + F ( M ) n λ I + γ ( k , ω ) D 2 Φ σ ( x k ) β + F ( M ) γ ( k , ω ) ν σ ( x k ) F ( M + D 2 v β , σ , ε 1 D ; M ) .

    Thus, the comparison principle yields the desired inequality.

  2. Combining Lemma 4.15 (uniform convergence on every compact subset of D k Z n { ε k } ) and Lemma 5.15, we have (after rescaling)

    v β , σ , ε 1 D ; M = o ( ε 2 ) ,

    if x ε 1 D k Z n B 1 / 2 ( k ) . On the other hand, in B 1 / 2 ( k ) ,

    F ( M + D 2 h σ , k ; M + ) = F F ( M ) n Λ I + γ ( k , ω ) D 2 Φ σ ( x k ) F ( M ) γ ( k , ω ) ν σ ( x k ) F ( M + D 2 v β , σ , ε 1 D ; M ) .

    Moreover, there exists a constant L = L ( n , F , M ) 0 (independent of k ) such that

    h β , σ , k ; M + + L 0

    on B 1 / 2 ( k ) . Thus, again by the comparison principle, we conclude the desired inequality.□

We denote

h β , k ; M ( x ) lim σ 0 h β , σ , k ; M ( x ) , h k ; M + ( x ) lim σ 0 h σ , k ; M + ( x ) ,

and

h β , k ; M , ε ( x ) ε 2 h β , k ; M ( x / ε ) h k ; M + , ε ( x ) ε 2 h k ; M + ( x / ε ) .

Recalling Assumption 1.3, we have

γ ( k , ω ) ( a ¯ ε ) α = ε 2 ,

where a ¯ ε = a ε ( r ( k , ω ) ) / ε . Thus, we deduce

h β , k ; M B a ¯ ε ( r ( k , ω ) ) ( k ) = ε 2 ϕ ( θ ) + O ( 1 ) , h k ; M + B a ¯ ε ( r ( k , ω ) ) ( k ) = ε 2 ϕ ( θ ) + O ( 1 ) ,

which yields the following corollary after letting σ 0 + in Lemma 5.18:

Corollary 5.19

  1. For every β and k Z n such that r ( k , ω ) > 0 , we have

    v β , ε 1 D ; M ( x ) ε 2 ϕ ( θ ) + O ( 1 ) o n B a ¯ ε ( r ( k , ω ) ) ( k ) a . e . ω Ω

    and so

    v ¯ β ; M ε ( x ) ϕ ( θ ) + o ( 1 ) o n T ε ( ω ) a . e . ω Ω ,

    for all β .

  2. For every β > β 0 and every k Z n , we have

    v β , ε 1 D ; M ( x ) ε 2 ϕ ( θ ) + o ( ε 2 ) o n B a ¯ ε ( r ( k , ω ) ) ( k ) a.e. ω Ω ,

    and so

    v ¯ β ; M ε ( x ) ϕ ( θ ) + o ( 1 ) on T ε ( ω ) a.e. ω Ω ,

    for all β > β 0 .

Lemma 5.20

For every k Z n , w ¯ β 0 ε satisfies

h β 0 , k ; M , ε ( x ) o ( 1 ) w ¯ β 0 ; M ε ( x ) h k ; M + , ε ( x ) + o ( 1 ) x B ε / 2 ( ε k ) D a.e. ω Ω .

In particular,

w ¯ β 0 ; M ε = ϕ ( θ ) + o ( 1 ) o n T ε D .

Proof

For simplicity, we drop the subscript M in this proof. Recall that for every β , we denote

v ¯ β ε ( x ) = ε 2 v β , ε 1 D ( x / ε ) ,

which is defined in D and v ¯ β ε = 0 on D . Compared to the Laplacian case (Lemma 5.8), we cannot exploit the Green representation formula when estimating an auxiliary function. Instead, we will use the Alexandrov-Backelman-Pucci estimate.

  1. Let β > β 0 . Note that

    F ( M + D 2 w β 0 , σ , ε 1 D ) = F ( M ) + β 0 k Z n ε 1 D γ ( k , ω ) ν σ ( x k )

    and

    F ( M + D 2 v β , σ , ε 1 D ) = F ( M ) + β k Z n ε 1 D γ ( k , ω ) ν σ ( x k ) χ { v β , σ , ε 1 D > 0 } = F ( M ) + β χ { v β , σ , ε 1 D > 0 } k Z n ε 1 D γ ( k , ω ) ν σ ( x k ) F ( M ) + β k Z n ε 1 D γ ( k , ω ) ν σ ( x k ) .

    Now let h be the solution of

    P + ( D 2 h ) = β 0 β in ε 1 D , h = 0 on ( ε 1 D ) .

    Since

    F ( M + D 2 v β , σ , ε 1 D + D 2 h ) F ( M + D 2 v β , σ , ε 1 D ) + P + ( D 2 h ) F ( M + D 2 w β 0 , σ , ε 1 D ) ,

    the comparison principle leads to

    v β , σ , ε 1 D + h w β 0 , σ , ε 1 D .

    Then an application of Alexandrov-Backelman-Pucci estimate for h indicates that

    sup ε 1 D h C ε 1 diam D β β 0 L n ( ε 1 D ) C ( ε 1 diam D ) 2 ( β β 0 ) .

    Thus, by rescaling, we conclude that

    w ¯ β 0 , σ ε v ¯ β , σ ε + C ( diam D ) 2 ( β β 0 ) .

    Applying Lemma 5.18 (ii), we conclude that

    w ¯ β 0 , σ ε ( x ) ε 2 h σ , k + ( x / ε ) + o ( 1 ) + C ( diam D ) 2 ( β β 0 )

    for all x B ε / 2 ( ε k ) . Letting σ 0 and then β β 0 , we obtain the desired upper bound.

  2. Arguing as in (i), for every β β 0 , let h be the solution of

    P + ( D 2 h ) = β β 0 β χ { v β , σ , ε 1 D = 0 } in ε 1 D , h = 0 on ( ε 1 D ) .

    Then since

    F ( M + D 2 w β 0 , σ , ε 1 D + D 2 h ) F ( M + D 2 v β , σ , ε 1 D ) ,

    we have

    w β 0 , σ , ε 1 D + h v β , σ , ε 1 D .

    Again, the Alexandrov-Backelman-Pucci estimate yields that

    sup ε 1 D h C ( ε 1 diam D ) 2 ( β 0 β ) + { v β , σ , ε 1 D = 0 } ε 1 D 1 / n .

    By rescaling, we obtain that

    w ¯ β 0 , σ ε v ¯ β , σ ε C ( diam D ) 2 ( β 0 β ) + { v ¯ β , σ ε = 0 } D 1 / n .

    Applying Lemma 5.18 (i) and letting σ 0 (recall the stability of coincidence sets)

    w ¯ β 0 ε ( x ) ε 2 h β , k ( x / ε ) o ( 1 ) C ( diam D ) 2 ( β 0 β ) + { v ¯ β ε = 0 } D 1 / n .

    Since lim ε 0 { v ¯ β ε = 0 } D = l ( β ) = 0 for any β β 0 , we obtain the desired lower bound.□

For each symmetric matrix M S n , we define a corrector w M ε by

F ( M + D 2 w M ε ) = β 0 ( M ) + F ( M ) in D ε , w M ε ( x ) = ϕ ( θ ) on T ε , w M ε = 0 on D ,

where β 0 ( M ) is the critical value. Note that we impose the boundary condition w ε = ϕ ( θ ) on T ε instead of w ε = 1 , which we wrote w ε 1 in Section 1.

Lemma 5.21

Let B η ( x 0 ) D . Then there exists a sequence { ε j } j N such that ε j 0 + and

lim j w ε j ( x ) = 0 ,

for any x B η ( x 0 ) { x 0 } .

Proof

First of all, by Theorem 4.13, the free solution w ¯ β 0 ; M ε satisfies

F ( M + D 2 w ¯ β 0 ; M ε ) = β 0 ( M ) + F ( M ) in D k Z n { ε k } ,

while the corrector w M ε satisfies

F ( M + D 2 w M ε ) = β 0 ( M ) + F ( M ) .

Applying Lemma 5.20 together with the comparison principle for w M ε and w ¯ β 0 ; M ε in D T ε , it holds that

w ¯ β 0 ; M ε ( x ) o ( 1 ) w ε ( x ) w ¯ β 0 ; M ε ( x ) + o ( 1 ) in D T ε .

Therefore, the desired result follows from Corollary 5.17.□

Before finding the effective equation satisfied by the limit profile u , we show the uniform ellipticity of the homogenized operator.

Lemma 5.22

For M S n and c 0 , set

F ¯ ( M , c ) F ( M + c D 2 w M ε ) = c F ( M / c + D 2 w M ε ) = c β 0 ( M / c ) + F ( M ) i f c > 0 , F ( M ) i f c = 0 .

Then

F ¯ ( M , c ) + λ N F ¯ ( M + N , c ) F ¯ ( M , c ) + Λ N ,

for any N 0 .

Proof

If c = 0 , then the result follows from the uniform ellipticity of F . For c = 1 , we have

F ¯ ( M + N , 1 ) = F ( M + N + D 2 w M + N ε ) F ( M + D 2 w M + N ε ) + Λ N .

Let w ˜ be the solution of

F ( M + D 2 w ˜ ) = F ¯ ( M + N , 1 ) Λ N in D ε , w ˜ ( x ) = ϕ ( θ ) on T ε , w ˜ = 0 on D .

Then the comparison principle yields that w ˜ w M + N ε and so w ˜ 0 letting ε 0 . Recalling the definition of w M ε , we have

F ( M + D 2 w ˜ ) F ( M + D 2 w M ε ) = β 0 ( M ) + F ( M ) = F ¯ ( M , 1 ) ,

and so

F ¯ ( M + N , 1 ) F ¯ ( M , 1 ) + Λ N .

The lower bound can be proved similarly. Moreover, considering M / c and N / c instead of M and N , we can finish the proof for general c > 0 .□

Finally, we are ready to finish the proof of our main theorem for the fully nonlinear case:

Proof of Theorem 1.2 (ii)

Recall that F ¯ is uniformly elliptic from Lemma 5.22. We are going to show that u is a supersolution. Let us assume that there is a parabola P touching u from below at x 0 and

F ¯ ( D 2 P , ( φ ( x 0 ) P ( x 0 ) ) + ) > 2 μ 0 > 0 .

In a small neighborhood of x 0 , B η ( x 0 ) , there exists another parabola Q such that

D 2 P > D 2 Q in B η ( x 0 ) , P ( x 0 ) + δ 0 < Q ( x 0 ) , P ( x ) > Q ( x ) on B η ( x 0 ) .

In addition, for ξ 0 ( φ ( x 0 ) u ( x 0 ) ) + ,

F ¯ ( D 2 Q , ξ 0 ) > μ 0 > 0 .

Then for a corrector w M ε with M = D 2 Q , we have

F ( D 2 Q ε ) = F ( D 2 Q + ξ 0 D 2 w M ε ) = F ¯ ( D 2 Q , ξ 0 ) > μ 0 > 0 ,

where Q ε ( x ) = Q ( x ) + ξ 0 w M ε ( x ) . Since Q ε < u ε on B ρ ( x 0 ) for sufficiently small ε > 0 , the comparison principle yields that Q ( x 0 ) u ( x 0 ) . It contradicts to the fact that Q ( x 0 ) > P ( x 0 ) + δ 0 = u ( x 0 ) + δ 0 . A similar argument tells us u is also a subsolution.□

Acknowledgments

Ki-Ahm Lee was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT): NRF-2021R1A4A1027378.

  1. Conflict of interest: The authors state no conflict of interest.

References

[1] M. A. Akcoglu and U. Krengel, Ergodic theorems for superadditive processes, J. Reine Angew. Math. 323 (1981), 53–67. 10.1007/BF01013192Search in Google Scholar

[2] S. N. Armstrong, B. Sirakov, and C. K. Smart, Fundamental solutions of homogeneous fully nonlinear elliptic equations, Comm. Pure Appl. Math. 64 (2011), no. 6, 737–777. 10.1002/cpa.20360Search in Google Scholar

[3] A. Bensoussan, J.-L. Lions, and G. Papanicolaou, Asymptotic analysis for periodic structures, Studies in Mathematics and its Applications, vol. 5, North-Holland Publishing Co., Amsterdam-New York, 1978. 10.1090/chel/374Search in Google Scholar

[4] M. Bôcher, Singular points of functions which satisfy partial differential equations of the elliptic type, Bull. Amer. Math. Soc. 9 (1903), no. 9, 455–465. 10.1090/S0002-9904-1903-01017-9Search in Google Scholar

[5] L. A. Caffarelli, A note on nonlinear homogenization, Comm. Pure Appl. Math. 52 (1999), no. 7, 829–838. 10.1002/(SICI)1097-0312(199907)52:7<829::AID-CPA3>3.0.CO;2-MSearch in Google Scholar

[6] L. A. Caffarelli and X. Cabré, Fully Nonlinear Elliptic Equations, American Mathematical Society Colloquium Publications, vol. 43, American Mathematical Society, Providence, RI, 1995. 10.1090/coll/043Search in Google Scholar

[7] L. A. Caffarelli and K.-A. Lee, Viscosity method for homogenization of highly oscillating obstacles, Indiana Univ. Math. J. 57 (2008), no. 4, 1715–1741. 10.1512/iumj.2008.57.3261Search in Google Scholar

[8] L. A. Caffarelli and A. Mellet, Random homogenization of an obstacle problem, Ann. Inst. H. Poincaré Anal. Non Linéaire 26 (2009), no. 2, 375–395. 10.1016/j.anihpc.2007.09.001Search in Google Scholar

[9] L. A. Caffarelli, P. E. Souganidis, and L. Wang, Homogenization of fully nonlinear, uniformly elliptic and parabolic partial differential equations in stationary ergodic media, Comm. Pure Appl. Math. 58 (2005), no. 3, 319–361. 10.1002/cpa.20069Search in Google Scholar

[10] D. Cioranescu and F. Murat, Un terme étrange venu daailleurs. II, In: Nonlinear Partial Differential Equations and Their Applications. Collège de France Seminar, Vol. III (Paris, 1980/1981), vol. 70 of Research Notes in Mathematics, Pitman, Boston, Mass.-London, 1982, 425–426. Search in Google Scholar

[11] M. G. Crandall, H. Ishii, and P.-L. Lions, User’s guide to viscosity solutions of second order partial differential equations, Bull. Amer. Math. Soc. (N.S.) 27 (1992), no. 1, 1–67. 10.1090/S0273-0979-1992-00266-5Search in Google Scholar

[12] G. Dal Maso and L. Modica, Nonlinear stochastic homogenization and ergodic theory, J. Reine Angew. Math. 368 (1986), 28–42. 10.1515/crll.1986.368.28Search in Google Scholar

[13] G. Dal Maso and F. Murat, Asymptotic behaviour and correctors for Dirichlet problems in perforated domains with homogeneous monotone operators, Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4) 24 (1997), no. 2, 239–290. Search in Google Scholar

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

[15] D. Gilbarg and N. S. Trudinger, Elliptic partial differential equations of second order, vol. 224 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], second edition, Springer-Verlag, Berlin, 1983. Search in Google Scholar

[16] F. R. Harvey and H. B. Lawson Jr, The Dirichlet problem with prescribed interior singularities, Adv. Math. 303 (2016), 1319–1357. 10.1016/j.aim.2016.08.024Search in Google Scholar

[17] F. R. Harvey and H. B. Lawson Jr., Tangents to subsolutions: existence and uniqueness, Part I, Ann. Fac. Sci. Toulouse Math. (6) 27 (2018), no. 4, 777–848. 10.5802/afst.1583Search in Google Scholar

[18] V. V. Jikov, S. M. Kozlov, and O. A. Oleĭnik, Homogenization of Differential Operators and Integral Functionals, Springer-Verlag, Berlin, 1994. Translated Russian by G. A. Yosifian. 10.1007/978-3-642-84659-5Search in Google Scholar

[19] S. Kim and K.-A. Lee, Viscosity method for homogenization of parabolic nonlinear equations in perforated domains, J. Differ. Equ. 251 (2011), no. 8, 2296–2326. 10.1016/j.jde.2011.07.014Search in Google Scholar

[20] D. A. Labutin, Isolated singularities for fully nonlinear elliptic equations, J. Differ. Equ. 177 (2001), no. 1, 49–76. 10.1006/jdeq.2001.3998Search in Google Scholar

[21] K.-A. Lee, Obstacle Problems for the Fully Nonlinear Elliptic Operators, Ph.D. thesis, New York University, New York, 1998. Search in Google Scholar

[22] K.-A. Lee and S.-C. Lee, Random homogenization of φ-Laplace equations with highly oscillating obstacles, Indiana Univ. Math. J. http://www.iumj.indiana.edu/IUMJ/Preprints/9220.pdf.Search in Google Scholar

[23] K.-A. Lee and M. Yoo, The viscosity method for the homogenization of soft inclusions, Arch. Ration. Mech. Anal. 206 (2012), no. 1, 297–332. 10.1007/s00205-012-0533-4Search in Google Scholar

[24] M. Masi, On compressive radial tidal forces, Am J Phys. 75 (2007), no. 2, 116–124. 10.1119/1.2366736Search in Google Scholar

[25] J. Serrin, Local behavior of solutions of quasi-linear equations, Acta Math. 111 (1964), 247–302. 10.1007/BF02391014Search in Google Scholar

[26] J. Serrin, Isolated singularities of solutions of quasi-linear equations, Acta Math. 113 (1965), 219–240. 10.1007/BF02391778Search in Google Scholar

[27] L. Tang, Random homogenization of p-Laplacian with obstacles in perforated domain, Comm. Partial Differ. Equ. 37 (2012), no. 3, 538–559. 10.1080/03605302.2011.648038Search in Google Scholar

Received: 2022-01-27
Revised: 2022-05-03
Accepted: 2022-07-31
Published Online: 2022-09-08

© 2023 Ki-Ahm Lee and Se-Chan Lee, published by De Gruyter

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

Articles in the same Issue

  1. Regular Articles
  2. On sufficient “local” conditions for existence results to generalized p(.)-Laplace equations involving critical growth
  3. On the critical Choquard-Kirchhoff problem on the Heisenberg group
  4. On the local behavior of local weak solutions to some singular anisotropic elliptic equations
  5. Viscosity method for random homogenization of fully nonlinear elliptic equations with highly oscillating obstacles
  6. Double-phase parabolic equations with variable growth and nonlinear sources
  7. Logistic damping effect in chemotaxis models with density-suppressed motility
  8. Bifurcation diagrams of one-dimensional Kirchhoff-type equations
  9. Standing wave solution for the generalized Jackiw-Pi model
  10. Weak-strong uniqueness and energy-variational solutions for a class of viscoelastoplastic fluid models
  11. Eigenvalue inequalities for the buckling problem of the drifting Laplacian of arbitrary order
  12. Homoclinic solutions for a differential inclusion system involving the p(t)-Laplacian
  13. Equivalence between a time-fractional and an integer-order gradient flow: The memory effect reflected in the energy
  14. Bautin bifurcation with additive noise
  15. Small solitons and multisolitons in the generalized Davey-Stewartson system
  16. Nonstationary Poiseuille flow of a non-Newtonian fluid with the shear rate-dependent viscosity
  17. A global compactness result with applications to a Hardy-Sobolev critical elliptic system involving coupled perturbation terms
  18. On a strongly damped semilinear wave equation with time-varying source and singular dissipation
  19. Singularity for a nonlinear degenerate hyperbolic-parabolic coupled system arising from nematic liquid crystals
  20. Stability of stationary solutions to the three-dimensional Navier-Stokes equations with surface tension
  21. Uniform decay estimates for the semi-linear wave equation with locally distributed mixed-type damping via arbitrary local viscoelastic versus frictional dissipative effects
  22. Symmetry and nonsymmetry of minimal action sign-changing solutions for the Choquard system
  23. Periodic and quasi-periodic solutions of a four-dimensional singular differential system describing the motion of vortices
  24. Multiple nontrivial solutions of superlinear fractional Laplace equations without (AR) condition
  25. Existence and blow-up of solutions in Hénon-type heat equation with exponential nonlinearity
  26. Existence and concentration behavior of positive solutions to Schrödinger-Poisson-Slater equations
  27. On the fractional Korn inequality in bounded domains: Counterexamples to the case ps < 1
  28. Gradient estimates for nonlinear elliptic equations involving the Witten Laplacian on smooth metric measure spaces and implications
  29. Dynamical analysis of a reaction–diffusion mosquito-borne model in a spatially heterogeneous environment
  30. Existence and multiplicity of solutions for a quasilinear system with locally superlinear condition
  31. Nontrivial solution for Klein-Gordon equation coupled with Born-Infeld theory with critical growth
  32. Modeling Wolbachia infection frequency in mosquito populations via a continuous periodic switching model
  33. Infinitely many localized semiclassical states for nonlinear Kirchhoff-type equation
  34. Fujita-type theorems for a quasilinear parabolic differential inequality with weighted nonlocal source term
  35. Approximations of center manifolds for delay stochastic differential equations with additive noise
  36. Periodic solutions to a class of distributed delay differential equations via variational methods
  37. Groundstate for the Schrödinger-Poisson-Slater equation involving the Coulomb-Sobolev critical exponent
  38. Existence of nontrivial solutions for the Klein-Gordon-Maxwell system with Berestycki-Lions conditions
  39. Global Sobolev regular solution for Boussinesq system
  40. Normalized solutions for the p-Laplacian equation with a trapping potential
  41. Nonlinear elliptic–parabolic problem involving p-Dirichlet-to-Neumann operator with critical exponent
  42. Blow-up for compressible Euler system with space-dependent damping in 1-D
  43. High energy solutions of general Kirchhoff type equations without the Ambrosetti-Rabinowitz type condition
  44. On the dynamics of grounded shallow ice sheets: Modeling and analysis
  45. A survey on some vanishing viscosity limit results
  46. Blow-up for logarithmic viscoelastic equations with delay and acoustic boundary conditions
  47. Generalized Liouville theorem for viscosity solutions to a singular Monge-Ampère equation
  48. Front propagation in a double degenerate equation with delay
  49. Positive solutions for a class of singular (pq)-equations
  50. Higher integrability for anisotropic parabolic systems of p-Laplace type
  51. The Poincaré map of degenerate monodromic singularities with Puiseux inverse integrating factor
  52. On a system of multi-component Ginzburg-Landau vortices
  53. Existence and concentration of solutions to Kirchhoff-type equations in ℝ2 with steep potential well vanishing at infinity and exponential critical nonlinearities
  54. Multiplicity results for fractional Schrödinger-Kirchhoff systems involving critical nonlinearities
  55. On double phase Kirchhoff problems with singular nonlinearity
  56. Estimates for eigenvalues of the Neumann and Steklov problems
  57. Nodal solutions with a prescribed number of nodes for the Kirchhoff-type problem with an asymptotically cubic term
  58. Dirichlet problems involving the Hardy-Leray operators with multiple polars
  59. Incompressible limit for compressible viscoelastic flows with large velocity
  60. Characterizations for the genuine Calderón-Zygmund operators and commutators on generalized Orlicz-Morrey spaces
  61. Non-local gradients in bounded domains motivated by continuum mechanics: Fundamental theorem of calculus and embeddings
  62. Noncoercive parabolic obstacle problems
  63. Touchdown solutions in general MEMS models
  64. Existence and nonexistence of nontrivial solutions for the Schrödinger-Poisson system with zero mass potential
  65. Short-time existence of a quasi-stationary fluid–structure interaction problem for plaque growth
  66. Fine bounds for best constants of fractional subcritical Sobolev embeddings and applications to nonlocal PDEs
  67. Symmetries of Ricci flows
  68. Asymptotic behavior of solutions of a second-order nonlinear discrete equation of Emden-Fowler type
  69. On the topological gradient method for an inverse problem resolution
  70. Supersolutions to nonautonomous Choquard equations in general domains
  71. Uniform complex time heat Kernel estimates without Gaussian bounds
  72. Global existence for time-dependent damped wave equations with nonlinear memory
  73. Normalized solutions for a critical fractional Choquard equation with a nonlocal perturbation
  74. Reversed Hardy-Littlewood-Sobolev inequalities with weights on the Heisenberg group
  75. Lamé system with weak damping and nonlinear time-varying delay
  76. Global well posedness for the semilinear edge-degenerate parabolic equations on singular manifolds
  77. Blowup in L1(Ω)-norm and global existence for time-fractional diffusion equations with polynomial semilinear terms
  78. Boundary regularity results for minimisers of convex functionals with (p, q)-growth
  79. Parametric singular double phase Dirichlet problems
  80. Special Issue on Nonlinear analysis: Perspectives and synergies
  81. Editorial to Special issue “Nonlinear analysis: Perspectives and synergies”
  82. Identification of discontinuous parameters in double phase obstacle problems
  83. Global boundedness to a 3D chemotaxis-Stokes system with porous medium cell diffusion and general sensitivity
  84. On regular solutions to compressible radiation hydrodynamic equations with far field vacuum
  85. On Cauchy problem for fractional parabolic-elliptic Keller-Segel model
  86. The evolution of immersed locally convex plane curves driven by anisotropic curvature flow
  87. Stability of combination of rarefaction waves with viscous contact wave for compressible Navier-Stokes equations with temperature-dependent transport coefficients and large data
  88. On concave perturbations of a periodic elliptic problem in R2 involving critical exponential growth
  89. Existence of solutions for a double-phase variable exponent equation without the Ambrosetti-Rabinowitz condition
Downloaded on 17.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/anona-2022-0273/html
Scroll to top button