Uniform Calderón-Zygmund estimates
in multiscale elliptic homogenization
Abstract.
This paper is concerned with the elliptic equation in a bounded domain, where takes a form of , with being 1-periodic in each . We prove the uniform Calderón-Zygmund estimate, namely, the uniform boundedness of the linear map for any with a constant independent of small parameters . Our result includes the uniform Calderón-Zygmund estimate in quasiperiodic elliptic homogenization (even without the Diophantine condition), which was previously unknown. The proof novelly combines the Dirichlet’s theorem on the simultaneous Diophantine approximation from number theory, a technique of reperiodization, reiterated periodic homogenization and a large-scale real-variable argument. Using the idea of reperiodization, we also obtain some large-scale or mesoscopic-scale Lipschitz estimates.
Keywords: Multiscale homogenization, Diophantine approximation, Calderón-Zygmund estimate.
2020 Mathematics Subject Classification:
35B271. Introduction
In this paper, we study the uniform regularity for a family of linear elliptic equations or systems in divergence form in a bounded domain ,
(1.1) |
where the coefficient matrix takes a form of , and are small parameters describing different oscillating scales of the coefficients. We assume that is 1-periodic in each with . Due to the symmetry of ’s, it is harmless to assume . The full assumptions on will be given shortly.
The homogenization theorem (or quantitative convergence rates) and uniform regularity are the fundamental questions in homogenization theory. These two questions have been well answered for the equation (1.1) in the case of one oscillating scale (i.e., ); see recent monographs [She18, AK22] and references therein. For the case of multiple oscillating scales, we refer to our recent work [NZ23] for a brief survey. It is important to point out that before the work [NZ23], the homogenization theorem and uniform regularity were considered closely interconnected and always appear together, in the sense that the uniform regularity can be derived as a consequence of convergence rates either by a compactness method [AL87, KLS13] or a quantitative method (excess decay iteration [GNO20, AS16, She17] or a real-variable argument [CP98]). For the equation (1.1) with , we have the qualitative homogenization theorem or a quantitative convergence rate to a deterministic homogenized equation only if the scales are separated or well-separated [AB96], respectively. Recall that the scales are well-separated if for some . Consequently, under this scale-separation condition, the uniform regularity estimates for (1.1) were established in [NSX20] by a quantitative method. Yet, as pointed out by Avellaneda in [Ave96] (see also [GH03]) such an assumption is not adequate for treating the most general problems of transport and diffusion in self-similar (particularly random) media. The main contribution of [NZ23] is that, without any separation condition on , we obtained the uniform estimate for any by the compactness method combined with a novel scale-reduction process. This result was unexpected since it shows for the first time that the uniform regularity holds stably even without homogenization, and in some sense the averaging effect always takes places across all the mesoscopic scales.
This paper is a continuation of our previous work [NZ23]. It was conjectured in [NZ23] that with in (1.1) and without any scale-separation condition, the solution admits uniform Lipschitz estimate. In this paper, we partially solve the conjecture by proving a slightly weaker result, namely, the gradient is uniformly bounded in spaces for any . In particular, this recovers the uniform estimate in [NZ23] in view of the Sobolev embedding theorem. Moreover, our proof is quantitative (in contrast to the compactness method in [NZ23]) in the sense that the constant can be computed explicitly.
Before stating the main result, we list the assumptions on the coefficient matrix :
-
•
Strong ellipticity condition: there exists so that
(1.2) for every and for any
-
•
Periodicity: for any and ,
(1.3) -
•
Smoothness: there exists and such that
(1.4) for any .
We now state our main result.
Theorem 1.1.
The same uniform estimate holds if the Dirichlet boundary condition in (1.5) is replaced by the compatible Neumann boundary condition
(1.7) |
where is the unit outer normal vector of .
The new ingredient in the proof of Theorem 1.1 is the Dirichlet’s theorem (see Theorem 3.1) on simultaneous Diophantine approximation from number theory, which enters into a new technique of reperiodization and allow us to quantitatively separate one scale from the rest scales; see Section 3.1 for details. Precisely, for a given sequence in nonincreasing order and any number , we can find a new 1-periodic matrix , depending on and the ratios , such that can be rewritten as
(1.8) |
with some and the last scale is -separated from the rest ’s (i.e., for all ). Theoretically, can be chosen as large as with some , yielding a well-separation condition. However, as a payoff the regularity of in will become worse as increases, which leads to a small-scale estimate depending on in a blow up argument. Fortunately, a sufficiently large independent of will be just enough for our proof. With the -separation condition, we can perform a quantitative argument of the reiterated homogenization and approximate the solution of the original equation with oscillating scales by a solution of a new equation with at most oscillating scales. Hence, an inductive argument on the number of scales combined with a careful real-variable argument involving a double-averaging estimate will complete the proof.
It is crucial to point out that the proofs for estimate (a type of Schauder estimates) and gradient estimate (Calderón-Zygmund estimate related to singular integrals) are essentially different, for the latter is stronger and more informative. As in [NZ23], the estimate can be derived with a qualitative -convergence theorem and a compactness method (the correctors are not needed; also see [She15, Theorem 3.1] or [SZ18, Theorem 6.1] for similar situations). However, the estimate of gradient must involve a quantitative convergence rate or the fine properties of correctors. While our proof of Theorem 1.1 also uses a scale-reduction theorem and an inductive argument on the number scales, it is exactly the Dirichlet’s theorem that helps us quantify the error in the scale-reduction process and provides much stronger implications and more information about the behavior of the solutions. In particular, it possibly explains the reason behind the occurrence of the mesoscopic averaging effect in arbitrary multiscale media.
As a direct consequence, Theorem 1.1 implies the uniform Calderón-Zygmund estimate for the operator with arbitrary quasiperiodic coefficient . We recall that is said to be quasiperiodic if , where is -periodic in with and is an constant real matrix. Note that we do not have any restriction on the entries of and therefore is allowed to be periodic with a very degenerate periodic cell.
Corollary 1.2.
In the setting of quasiperiodic homogenization, the convergence rate was first obtained by Kozlov in [Koz78] under a Diophantine condition (a quantitative ergodicity condition): for each row of , for all and for some . It was observed in [AL91] that the uniform regularity is valid under Kozlov’s Diophantine condition. This condition was also required in recent work [She15, AGK16, BG19] applied to the case of quasiperiodic coefficients. In our Corollary 1.2, we do not need the Diophantine condition and the constant in the estimate is independent of the entries of the matrix (except for the dimension of ). This includes some typical degenerate cases, such as periodic coefficients with a degenerate periodic cell (e.g., a thin rectangle), quasiperiodic coefficients oscillating at two almost resonant frequencies (e.g., the ratio of the frequencies is a Liouville number), multiscale quasiperiodic coefficients, etc.
The proof of Corollary 1.2 based on Theorem 1.1 is simple and we include it here. In fact, it suffices to write into a periodic coefficient matrix with multiple oscillating scales and apply Theorem 1.1. To this end, let . Let with and be independent variables. Define a matrix with variables
Then is 1-periodic in each since is 1-periodic. Moreover, it is straightforward to verify
(1.9) |
This reduces the quasiperiodic coefficient matrix with one oscillating scale into a periodic coefficient matrix with oscillating scales. Clearly, in this case, all the scales are possibly not well-separated. Moreover, if is uniformly elliptic and Hölder continuous, so is . As a result, Corollary 1.2 follows readily from Theorem 1.1.
Finally, using a similar idea of scale separation, we can show the uniform large-scale or mesoscopic-scale Lipschitz estimates. In particular, in the case of two scales, i.e., , for any , we have for all ,
(1.10) |
where depends only on and ; see Theorem 5.1. For general cases , we have some mesoscopic-scale Lipschitz estimates near any given scale; see Theorem 5.2.
Organization. The rest of the paper is organized as follows. In Section 2, we recall some knowledge and known results from locally periodic homogenization and reiterated homogenization. In Section 3, we use the Dirichlet’s theorem and reperiodization technique to separate scales and reduce one scale by a uniform approximation. The main theorem is proved in Section 4 by a real-variable argument. In Section 5, we exploit the idea of scale separation to obtain the uniform large-scale or mesoscopic-scale Lipschitz estimates.
Acknowledgements. W. Niu is supported by NNSF of China (12371106, 11971031). J. Zhuge is partially supported by grants for Excellent Youth from the NSFC.
2. Preliminaries
2.1. Local periodic homogenization
In this subsection, we recall some results concerning the homogenization of the locally periodic operator with a scalar , which will be used in our quantitative scale reduction process in Section 3.
Suppose satisfies the ellipticity condition (1.2) and is 1-periodic in . The effective (homogenized) matrix is given by
(2.1) |
where is the corrector given by the cell problem
(2.2) |
for . Here denotes the th component of . Assume that
(2.3) |
The standard energy estimates for (2.2) imply that
(2.4) |
with depending only on This combined with the definition of yields
Likewise, if
(2.5) |
for some for any Then we have
(2.6) |
Let be supported in such that and . For functions of the form , we define a partial-smoothing operator
(2.7) |
The following theorem provides the convergence rate in for the locally periodic operator. The proof may be found in [NSX20].
Theorem 2.1.
Let be a bounded Lipschitz domain in . Let be the solution to in with on and the solution to in with on Then we have
(2.8) |
where we have extended to the whole space , and depends only on , and .
We emphasize that in the above theorem, has no smoothness assumption in .
2.2. Reiterated correctors and effective matrix
Consider the operator with scales , where satisfies the ellipticity and periodicity conditions (1.2) and (1.3). Assume that is separated from for By using the idea of reiterated homogenization, one can view as parameters and homogenize the finest scale to obtain an operator of the same type but with only scales.
Indeed rewrite
It is obvious that satisfies (1.2), and is locally 1-periodic in . Let be the corrector given by (2.2) with replaced by , and the effective matrix of given by (2.1). In view of the structure of , the corresponding corrector takes a form of
(2.9) |
where the reiterated corrector is defined by the cell problem
(2.10) |
for , which is also 1-periodic in , due to the periodicity of . Moreover, takes a form of
(2.11) |
with
(2.12) |
It is obvious that is 1-periodic in each with , and it is defined independent of . Similar to the case , satisfies the ellipticity condition (1.2). Furthermore, similar to (2.6), if is Hölder continuous in for , so is .
3. Quantitative scale separation and scale reduction
3.1. Scale separation
The key ingredient of this paper is the Dirichlet’s theorem on the simultaneous Diophantine approximation. The connection between the Dirichlet’s theorem and the regularity theory in homogenization is previously unknown. The only loosely related notion is the Diophantine condition imposed in the setting of quasiperiodic homogenization, as mentioned earlier. The Dirichlet’s theorem appears to be a powerful tool (instead of a condition) that eventually allows us to derive the uniform Calderón-Zygmund estimates in multiscale or quasiperiodic homogenization without any additional conditions.
Let us first recall the Dirichlet’s theorem; see [Sch91].
Theorem 3.1 (Dirichlet (1842)).
Suppose that are real numbers and . There exist integers such that and
(3.1) |
Roughly speaking, the Dirichlet’s theorem states that any real numbers can be approximated by good rational numbers with a quantitative small error. In the rest of this subsection, we apply the Dirichlet’s theorem to separate at least one scale in the coefficient matrix oscillating at multiple (possibly unseparated) scales.
Without loss of generality, hereafter we always assume
(3.2) |
Let be a large number. We emphasize that the fixed number will be chosen later independent of . We say is -separated from if . This scale separation condition is crucial in quantitative reiterated homogenization.
Given as above, by the Dirichlet’s theorem with for , we can find integers such that and
(3.3) |
Moreover, because of (3.2), we have . Note that it is possible that if is large enough that (i.e., is already -separated from ).
Define, for ,
(3.4) |
Then, we can write
(3.5) |
Now, we define
(3.6) |
The key insight here is that, since and are all integers, is 1-periodic in each . We thereby will call the transformation from to a technique of reperiodization. As a result of (3.5) and (3.6), we have
(3.7) |
In other words, the reperiodization allows us to rewrite the original periodic matrix oscillating at scales into a new periodic matrix oscillating at different scales . Note that if for some , then the corresponding scale for that (slower) variable does not change, i.e., . If for some , then we just do not have the scale and we have less oscillating scales after reperiodization (which is even better). Without loss of generality, we may assume this does not happen.
The new scales and the new 1-periodic matrix have the following crucial properties:
-
•
The smallest scale is -separated from for each . In fact, (3.3) implies
(3.8) This is the key property for our application.
-
•
If is Hölder continuous in for , i.e.,
(3.9) with for any , then is also Hölder continuous in for with the same constants . However, may lose good regularity in since many periods have been compressed into a single 1-periodic cell.
-
•
The construction of relies on the choice of as well as the ratios with . In other words, is scale-invariant. Also note that is a dimensionless parameter.
It is important to note that the above scale separation process can only separate the smallest scale from the rest larger scales, while the relationships among the rest larger scales cannot be determined. This allows us to perform a one-scale reduction by using the idea of reiterated homogenization.
3.2. Scale reduction and uniform approximation
In this subsection we use the scale separation technique in Section 3.1 to reduce at least one scale. Precisely, we shall prove that the solution of (1.1) can be uniformly approximated by a solution to the equation of the same type but with at most oscillating scales. The following is the main theorem.
Theorem 3.2.
Suppose that satisfies the assumptions (1.2), (1.3) and (3.9). Let be a weak solution to
(3.10) |
Then for any , there exists a coefficient matrix , scales with , and a weak solution to
such that
(3.11) |
and
(3.12) |
for some where is given by
(3.13) |
is the corrector given by (2.10) with replaced by , and is given by (3.18). Moreover, the constant depends only on and in (3.9), and the assumptions (1.2), (1.3) and (3.9) are preserved for .
The proof of Theorem 3.2 relies on the scale separation technique in Section 3.1 and the following approximation theorem that pertains to the quantitative reiterated homogenization.
Theorem 3.3.
Suppose satisfies the assumption (1.2), (1.3), and (3.9). Assume is -separated from . Let be a weak solution to
(3.14) |
Then there exist a 1-periodic matrix with scales and a weak solution to
such that
(3.15) |
and
(3.16) |
for some , where and are defined by (2.10) and (2.12), respectively, and
(3.17) |
Moreover, the constant depends only on and in (3.9).
Proof of Theorem 3.2.
We first consider the case . Given and , by the Dirichlet’s theorem, we can find integers such that , and (3.3)-(3.8) hold. Set
(3.18) | ||||
By (3.7) and (3.10), we know that satisfies
(3.19) |
By (3.8), is -separated from for all . Thus, by Theorem 3.3, there exist given by (2.12) with replaced by , and a weak solution to
(3.20) |
such that
(3.21) |
and
(3.22) | ||||
for some , where is the corrector defined in (2.10) with replaced by , and
We rename , , and derive the desired estimate in the case . Finally, note that and are scale-invariant. The estimate for general follows immediately by rescaling. The proof is complete.
∎
Proof of Theorem 3.3.
We consider the matrix
(3.23) |
Recall that is strongly elliptic and locally 1-periodic in . In view of (1.4), is Hölder continuous in , and
(3.24) |
In order to apply Theorem 2.1, we need to find an approximate matrix which is Lipschitz in the variable. In fact, we define
(3.25) |
with given as in Section 2.1. Then satisfies the ellipticity condition (1.2), is 1-periodic in , and
(3.26) |
where depends only on and .
Let be the weak solution to (3.14). Then by (3.23), it satisfies
(3.27) |
Let be the weak solution to
(3.28) |
Combining (3.27), (3.28) and the first inequality in (3.26), we can apply the energy estimate to the equation of
(3.29) |
to obtain
(3.30) |
Next, we apply Theorem 2.1 to the equation (3.28). Let be the corrector given by (2.2) with replaced by , and the corresponding effective coefficient matrix given by (2.1). Let be the solution to
(3.31) |
Thanks to Theorem 2.1,
(3.32) | ||||
By the Hölder’s inequality, the last term on the right-hand side of (3.32) is bounded by
(3.33) |
On the other hand, by the interior estimate for the equation (3.28), for any ,
Then it follows by a covering argument that
(3.34) | ||||
where the Hölder’s inequality has been used for the last integral. By Meyers’ estimate for the elliptic equations (3.27), (3.28) and (3.31), we have for some ,
(3.35) |
which implies that
(3.36) |
By taking (3.33) and (3.34) into (3.32), and using (3.36) we obtain that
(3.37) | ||||
Now, let be the corrector given by (2.2) with replaced by , and the effective matrix of given by (2.1) (recall that and take the forms of (2.9) and (2.11)). Note that satisfies
Thus, the standard energy estimate and the first inequality in (3.26) imply that
(3.38) |
where depends only on and . This together with the definitions of and gives
(3.39) |
Let be the weak solution to
(3.40) |
Similar to (3.30), we apply the standard energy estimate to the equation for ,
(3.41) |
and use (3.39) to obtain
(3.42) |
where the energy estimate for (or (3.36)) is also used in the last inequality.
Finally, combining (3.30), (3.38) and (3.42), we may replace the approximate solutions on the left-hand side of (3.37) by , respectively. In particular, by the property of the smoothing operator (see e.g., [She18, NSX20]), we have
(3.43) |
It follows that
(3.44) | ||||
Recall that is -separated from , i.e., for In view of (2.9) and (2.11), the desired estimate (3.16) follows immediately from (3.44), while the estimate (3.15) follows directly from (3.42). The proof is complete. ∎
By similar arguments as above, it is not difficult to prove the boundary version of the uniform approximation theorem.
Theorem 3.4.
Let be a bounded domain. Suppose that satisfies the assumptions (1.2), (1.3) and (3.9). Let with and be a weak solution to
(3.45) |
Then for any , there exist a coefficient matrix , scales with , and a weak solution to
such that
(3.46) |
and
(3.47) |
for some where is given by
(3.48) |
and is the corrector given by (2.2) with replaced by (we have extended by zero across the boundary). Moreover, the constant depends only on and in (3.9), and the assumptions (1.2), (1.3) and (1.4) are preserved for .
4. The uniform Calderón-Zygmund estimates
In this section we prove the uniform Calderón-Zygmund estimate for the equation (1.1). Let us first consider the interior estimate.
Theorem 4.1.
The proof of Theorem 4.1 relies on a real-variable argument involving a double-averaging estimate. We define the averaging operator as
(4.3) |
We need the averaging operator in the real-variable argument for two reasons. The first reason is that the approximation in (3.12) is meaningful only if for sufficiently large . This forces us to take the large-scale averaging estimate above the scale for sufficiently large and ignore the possible irregularity below that scale. The second reason to introduce another averaging is due to the special structure of the corrector in given by (3.13). Recall that is 1-periodic in each . But its gradient in does not have an estimate uniformly in before we prove it in case of oscillating scales (otherwise we will run into a circular reasoning), which causes a big problem in the real-variable argument. To resolve this issue, we take the averaging for another time at scale , a scale comparable to the size of the periodic cell (depending on as ). This allows us to use only the periodic structure of and the energy estimate (independent of ). Overall, the principle is that we have to make sure all the constants, except for , in the real-variable argument (Theorem 4.2 below) are independent of .
Theorem 4.2.
[She23, Theorem 2.1 and Remark 2.4] Let , and for some , where is a ball in . Let be a fixed number. Suppose that for each ball with the property that , there exists two measurable functions and on such that on , and
(4.4) | |||
(4.5) |
where , and . Then there exists , depending only on with the property that if , then
(4.6) |
where depends only on and . If , then (4.6) is replaced by
(4.7) |
Proof of Theorem 4.1.
Let us first consider the case . We use an inductive argument on the number of scales to prove the theorem. Note that the result is well-known for (see e.g., [She18]). Assume the theorem is true if the number of scales is strictly less than , which means that for any , the solution to
satisfies the uniform interior estimate
(4.8) |
for any and any .
Let to be determined later, and for some fixed . For each ball such that , we need to construct and , which satisfy (4.4) and (4.5), respectively. Since is a solution to in , by Theorem 3.2, there exist a 1-periodic matrix , scales , and a solution to
(4.9) |
such that
(4.10) |
and
(4.11) |
where is given by (3.13). Let
By the triangle inequality, we have
We recall a basic property of the averaging operator . If , then for any , we have
(4.12) |
This can be shown easily by the Fubini’s theorem.
Consequently, by the assumption , we have
(4.13) |
and
(4.14) |
Combining the last two inequalities with (4.11), we have
(4.15) |
We now choose large enough such that , and then take for some large enough such that where is given by Theorem 4.2 depending only on It follows that satisfies condition (4.5) for with . In the following, we will verify that satisfies (4.4) and the constant is independent of .
Note that
(4.16) |
Since satisfies (4.9), and the coefficient matrix involves at most scales and is periodic and Hölder continuous for . By the inductive assumption, we know that satisfies (4.8), which combined with (4.10) implies
for any . By the Hölder’s inequality and Fubini’s theorem,
(4.17) |
where we have used (4.12) in the last step.
To bound the second term on the right hand side of (4.16), we have to take advantage of the periodic structure of and it is right here that the averaging operator plays a key role. We observe that, for each ,
where we have used the energy estimate for (see the equation (2.10)) and depends only on and . This combined with (4.17) implies that
(4.18) |
for any . We point out that we cannot simply take the norm of directly since depends on and hence on , due to the loss of regularity of in caused by reperiodization.
By taking (4.17) and (4.18) into (4.16), we find that satisfies (4.4) for (without the part) with independent of . Thus the conditions of Theorem 4.2 are all satisfied. Thanks to (4.6), we obtain that
(4.19) |
where and and . This is a double-averaging estimate since we have taken average twice in the norm on the left-hand side. At this stage (after using Theorem 4.2), it is safe to remove one averaging in (4.19) by (4.12). In fact, since , (4.19) implies
(4.20) |
Next, we claim that for each , we have
(4.21) |
This pointwise Lipschitz estimate follows from a well-known blow-up argument at small scales and the fact . Inserting this into (4.20), we arrive at
(4.22) |
Since and are arbitrary with depending on , we actually have shown the uniform interior estimate for any in the case of .
Finally, we consider general case with . Let be a solution to (4.1). For each ball such that , let be the solution to
(4.23) |
and the solution to
(4.24) |
Thus, in . The energy estimate for (4.23) implies that
while the uniform estimates for proved above implies that
for some . Let , , and . We obtain (4.2) from Theorem 4.2 immediately (with ). ∎
Theorem 4.1 provides the uniform interior estimate for the equation (1.1). Based on Theorem 3.4 and a boundary version of Theorem 4.2 (See Theorem 4.1 and Remark 4.2 in [She23].), we can follow the same arguments to prove the following uniform boundary estimate.
Theorem 4.3.
We finally provide the proof of our main results.
5. Large-scale Lipschitz estimates
In this section, we investigate the large-scale Lipschitz estimates for (1.1). We distinguish between the cases of two scales and more scales. We recall that under the scale-separation condition: there exists a positive integer such that
(5.1) |
the large-scale Lipschitz estimate for the multiscale elliptic operators has been derived in [NSX20]. Indeed, let be the weak solution to
with for some . Suppose is strongly elliptic, periodic in for , and Hölder continuous in (no smoothness is needed for ). Then, under the condition (5.1), for any ,
(5.2) |
where is independent of .
However, as we have stated before, without any scale separation condition the full-scale uniform Lipschitz estimate for (1.1) seems difficult and still remains open. The following theorems provide suboptimal results in this direction using the idea of scale separation similar to Section 3.1.
Theorem 5.1.
Proof.
Applying the argument in Section 3.1 and by (3.7), for any we can find a 1-periodic matrix , Hölder continuous in , and write
(5.4) |
where and . Let . Then and . Thus and are well-separated. Moreover, satisfies the equation
(5.5) |
Thanks to (5.2), we have for ,
(5.6) |
This implies the desired estimate since . ∎
Theorem 5.2.
Proof.
Fix . Let with . By the Dirichlet’s theorem, there exist such that and
(5.8) |
where . As in Section 3.1, we set
and write
Define
(5.9) |
As a consequence, we have
(5.10) |
The last identity shows that we can rewrite original coefficient matrix with scales as a new 1-periodic matrix with scales. Moreover, the smallest scale is at least -separated from the remaining scales, i.e.,
(5.11) |
Hence, by a blow-up argument and (5.2) we have
(5.12) |
for any with . This is the desired estimate. To make sure , we require which gives .
References
- [AB96] G. Allaire and M. Briane. Multiscale convergence and reiterated homogenisation. Proc. Roy. Soc. Edinburgh Sect. A, 126(2):297–342, 1996.
- [AGK16] S. N. Armstrong, A. Gloria, and T. Kuusi. Bounded correctors in almost periodic homogenization. Arch. Ration. Mech. Anal., 222(1):393–426, 2016.
- [AK22] S. N. Armstrong and T. Kuusi. Elliptic homogenization from qualitative to quantitative. arXiv:2210.06488, 2022.
- [AL87] M. Avellaneda and F. Lin. Compactness methods in the theory of homogenization. Comm. Pure Appl. Math., 40(6):803–847, 1987.
- [AL91] M. Avellaneda and F. Lin. bounds on singular integrals in homogenization. Comm. Pure Appl. Math., 44(8-9):897–910, 1991.
- [AS16] S. N. Armstrong and C. K. Smart. Quantitative stochastic homogenization of convex integral functionals. Ann. Sci. Éc. Norm. Supér. (4), 49(2):423–481, 2016.
- [Ave96] M. Avellaneda. Homogenization and renormalization: the mathematics of multi-scale random media and turbulent diffusion. In Dynamical systems and probabilistic methods in partial differential equations (Berkeley, CA, 1994), volume 31 of Lectures in Appl. Math., pages 251–268. Amer. Math. Soc., Providence, RI, 1996.
- [BG19] A. Benoit and A. Gloria. Long-time homogenization and asymptotic ballistic transport of classical waves. Ann. Sci. Éc. Norm. Supér. (4), 52(3):703–759, 2019.
- [CP98] L. A. Caffarelli and I. Peral. On estimates for elliptic equations in divergence form. Comm. Pure Appl. Math., 51(1):1–21, 1998.
- [GH03] B. Gérard and O. Houman. Multiscale homogenization with bounded ratios and anomalous slow diffusion. Comm. Pure Appl. Math., 56(1):80–113, 2003.
- [GNO20] A. Gloria, S. Neukamm, and F. Otto. A regularity theory for random elliptic operators. Milan J. Math., 88(1):99–170, 2020.
- [KLS13] C. E. Kenig, F. Lin, and Z. Shen. Homogenization of elliptic systems with Neumann boundary conditions. J. Amer. Math. Soc., 26(4):901–937, 2013.
- [Koz78] S. M. Kozlov. Averaging of differential operators with almost periodic rapidly oscillating coefficients. Mat. Sb. (N.S.), 107(149)(2):199–217, 317, 1978.
- [NSX20] W. Niu, Z. Shen, and Y. Xu. Quantitative estimates in reiterated homogenization. J. Funct. Anal., 279(11):108759, 39, 2020.
- [NZ23] W. Niu and J. Zhuge. Compactness and stable regularity in multiscale homogenization. Math. Ann., 385(3-4):1431–1473, 2023.
- [Sch91] W. M. Schmidt. Diophantine approximations and Diophantine equations, volume 1467 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1991.
- [She15] Z. Shen. Convergence rates and Hölder estimates in almost-periodic homogenization of elliptic systems. Anal. PDE, 8(7):1565–1601, 2015.
- [She17] Z. Shen. Boundary estimates in elliptic homogenization. Anal. PDE, 10(3):653–694, 2017.
- [She18] Z. Shen. Periodic homogenization of elliptic systems, volume 269 of Operator Theory: Advances and Applications. Birkhäuser/Springer, Cham, 2018. Advances in Partial Differential Equations (Basel).
- [She23] Z. Shen. Weighted estimates for elliptic homogenization in Lipschitz domains. J. Geom. Anal., 33(1):Paper No. 3, 33, 2023.
- [SZ18] Z. Shen and J. Zhuge. Approximate correctors and convergence rates in almost-periodic homogenization. J. Math. Pures Appl. (9), 110:187–238, 2018.