1 Introduction
Let be a bounded Lipschitz domain in , , and . We consider the sharp convergence rate in the homogenization of the initial-boundary value problem with locally periodic coefficients
|
|
|
(1.1) |
where
|
|
|
(1.2) |
Note that the summation convention for repeated indices is used here and throughout the paper, and we may also omit the superscripts if it is clear to understand. The coefficient matrix defined on is assumed to be -periodic in , i.e.,
|
|
|
(1.3) |
and satisfy the boundedness and ellipticity conditions
|
|
|
(1.4) |
for any and a.e. , where and . We also assume that
|
|
|
(1.5) |
where is defined in Section 2.1.
System (1.1) is a simplified model describing physical processes or chemical reactions taking place in composite materials, such as thermal conduction, the sulfate corrosion of concrete (see [5, 10, 17] for more complicated models). It applies to processes in heterogeneous media, while the strictly periodic model
|
|
|
(1.6) |
is only suitable for homogeneous phenomena. From a micro perspective, the microscopic pattern of system (1.1) is allowed to differ at different times and positions. Since real media in biomechanics and engineering are almost never homogeneous, (1.1) covers better what happens in practical applications.
System (1.1) contains two levels of scales, the macroscopic scales and the microscopic scales . Usually, variables in are called macroscopic variables, denoting space-time positions, while are microscopic variables representing fast variations at the microscopic structure. In both (1.1) and (1.6), the scales of fast variations in space and time match naturally, that is, they are consistent with the intrinsic scaling of second order parabolic equations. Generally, one may consider models where the microscopic scales are with . In the periodic setting (1.6), when we say the scales are self-similar, and when they are non-self-similar [14]. Only in the case where , the spatial scale and the temporal scale are homogenized simultaneously. More generally, problems with multiple (matching or mismatching) scales in space and time have also been introduced in [5, 15, 22, 11, 7] and their references, where the qualitative homogenization of different types of matches was discussed widely. However, to the author’s knowledge, quite few quantitative results have been known for parabolic equations with multiple scales. For recent results on the quantitative homogenization of elliptic systems with multiple scales, we refer to [27, 18] and references therein.
Although the homogenized equation for (1.1) has been derived early in [5], there is not much progress in the quantitative homogenization. The only notable literature is [22, 21], where, for equation (1.1) with time-independent coefficients, the authors established the estimate of the operator exponential to its limit in for each . More recently, the rate in , as well as the full-scale Lipschitz estimates, for systems with non-self-similar scales is discussed widely under strong smoothness assumptions on the coefficients in [12]. We also refer to the series of work [2, 3, 4] for the qualitative pointwise convergence results of the same equation obtained by a probabilistic approach.
Under assumptions (1.4)–(1.5), the coefficient of system (1.1) is measurable on . We also assume that satisfy suitable conditions so that problem (1.1) admits a unique weak solution . As is well known, converges to a function weakly in as , where is the solution of the homogenized problem given by
|
|
|
(1.7) |
and is a divergence type elliptic operator with variable coefficients determined by solving unit cell problems at each point of the domain (see Section 2.3).
Our main goal is to establish the optimal convergence rate of to . Convergence rate is a core subject of homogenization and has aroused much interest in the past few years. So far, various results about convergence rates have been gained for parabolic systems with time-independent or time-dependent periodic coefficients. The reader may consult [28, 16, 13, 25, 20, 12] and their references. However, almost all these rates are in the sense of . In this paper, we establish a scale-invariant result for system (1.1) in with under quite general conditions.
Theorem 1.1.
Let be a bounded domain in , and . Assume that satisfies (1.4)–(1.5). Let and be the weak solutions to problems (1.1) and (1.7), respectively. Suppose further , with . Then
|
|
|
(1.8) |
where , depends only on and
|
|
|
Theorem 1.1 extends the result of [20], a similar error estimate for (1.6), to the locally periodic setting. It is remarkable that estimate (1.8) is scale-invariant under the parabolic rescaling and the constant in (1.8) is independent of the size of . This estimate is more elegant than that of [20] given as (with )
|
|
|
where the scale of does not coincide with the other terms when doing scaling. The earliest work on these kinds of scale-invariant results in homogenization should be attributed to Z. Shen who established the rate for elliptic systems with periodic coefficients in the noted book [23]. Later in [26, 27], the scale-invariant error estimates were extended to elliptic systems with stratified coefficients under rather general smoothness assumptions.
In one sense, the smoothness condition (1.5) means is -order differentiable in and -order differentiable in , which coincides with the regularity of general parabolic equations. Also the space has the same scale as . From the viewpoint of calculations, this smoothness condition is minimum to guarantee the error estimate.
Before describing the strategies and skills used in the paper, we introduce the notation
|
|
|
(1.9) |
which gives
|
|
|
(1.10) |
The main difficulties of the paper are essentially caused by the feature of two scales. As seen in the formal asymptotic expansion
|
|
|
the first-order term may not even be measurable on , as is not regular enough. To handle this problem, as in [27] for elliptic systems, we introduce the smoothing operator w.r.t. macroscopic variables . It makes smooth in for any which is -periodic in , thereby ensuring the measurability of . Moreover, by Fubini’s theorem this operator also helps us separate from in the coupled form . On the other hand, since is introduced to auxiliarily, it is necessary to control the difference (mostly in the case ). The idea is to write this difference into convolutions which act just like smoothing operators. In fact, it involves the differences in both space and time, namely,
|
|
|
where is the smoothing operator w.r.t. only (see (3.2)). The latter term is the difference in space and, by Poincaré’s inequality, it could be dominated by the “convolution”
|
|
|
The first term can be written formally into
|
|
|
where is a proper kernel. Both of these two terms can be controlled well by the critical estimates established in Section 3.1.
It should be pointed out that, due to the feature of the coefficient, the rate in involves many subtle inhomogeneous -type estimates of four-variable functions. This urges us to perform each step accurately in the right format, and by this reason, the process is much more delicate than that of elliptic problems in [27].
On the other hand, we introduce a new construction of flux correctors. In [13], flux correctors were constructed in an elliptic manner by lifting the function in both and , which results in high regularity in but low regularity in (especially for ). It does not work for higher-order parabolic systems, as more regularity in is required. Later, in [19] when dealing with higher-order systems, flux correctors were constructed by lifting the regularity w.r.t. space only. This provides enough regularity in , but no regularity in . Neither of them is applicable to our setting, as the microscopic regularities involved are very subtle. To this end, we construct flux correctors in a parabolic manner, in which way the regularities of flux correctors are the same as and even better than correctors. This construction seems more natural and is also valid for higher-order systems.
Another novelty of the paper is a new estimate of temporal boundary layers. Compared to the method in [20], it provides better estimates but requires no restriction on and . The estimate is based on a sharp embedding result for and it improves the estimate used in [14], where one half of the power of temporal layers was in fact lost.
We now describe the outline of the paper. Section 2 contains several parts of preliminaries. In Section 2.1, various vector-valued spaces of multi-variable functions are introduced, which are suitable tools to describe the properties of correctors and flux correctors. Section 2.2 provides some homogeneous Sobolev spaces. Next in Sections 2.3 and 2.4, correctors and flux correctors , together with their regularities, are studied. Section 3.1 is devoted to a series of critical estimates for the smoothing operator . Afterwards, a sharp embedding result for is built in Section 3.2, along with a corollary on the estimate of boundary layers.
These results are applied in Section 4 to establish the rate in and the rate in . More precisely, the estimate is established for the auxiliary function
|
|
|
(1.11) |
where is a cut-off function. Note that in this process those three in the last two terms of (1.11) play different roles: the first is mainly used to maintain the measurability on and control rapidly oscillating factors via Fubini’s theorem; the second operator helps us to reduce the smoothness assumption on in ; and the third one reduces the regularities of at the cost of the power of .
To derive the rate of to stated in Theorem 1.1, we adopt the classical duality argument, where the solution of the dual problem satisfies - estimates in . The main challenge in this process lies in the term
|
|
|
where is the corrector of the dual problem. For this term, the technique used in [20] is no longer in force, since is now a four-variable function and it does not have enough regularity to rearrange arbitrarily the order of integrals in mixed norms. Instead, the idea is, formally speaking, to transfer the gradient in to by integrating by parts, which is carried out by a regularity lifting argument (see Lemma 3.7).
Throughout this paper, unless otherwise stated, we will use to denote any positive constant which may depend on . It should be understood that may differ from each other even in the same line. We also use the notation for the integral average of over .
4 Convergence rates
This section is devoted to establishing the sharp convergence rate for problem (1.1). We always assume that henceforward.
Suppose , where and are two smooth cut-off functions on and , respectively, such that, and
|
|
|
(4.1) |
Let and be the solutions to problems (1.1) and (1.7) respectively. For the sake of simplicity, we extend onto , such that,
|
|
|
By equations (1.1) and (1.7), we calculate that
|
|
|
|
|
|
where with defined in Section 3.1 and notation (3.3) was used in . Note that we have regarded as functions on having value outside of , due to the cut-off effect of (see (4.1)). Recalling that is commutative with the smoothing operator , we deduce from (1.10) that
|
|
|
|
|
|
|
|
Thus, by setting , , where
|
|
|
it follows that
|
|
|
(4.2) |
where is defined by (2.8). We emphasize that
|
|
|
(4.3) |
where the main difference between and focuses on the smoothing acts of .
Noticing that , by Lemma 2.2, we write
|
|
|
|
|
|
|
|
|
|
|
|
where we have also used the fact that in the first equality, as well as the commutativity between and the partial derivatives w.r.t. in the second one. In view of (1.10), together with the skew-symmetry of in Lemma 2.2, this yields that
|
|
|
(4.4) |
Therefore, by defining
|
|
|
(4.5) |
and combining (4.2) and (4.4), we finally get
|
|
|
(4.6) |
Lemma 4.1.
Let be a bounded Lipschitz domain in and . Assume that satisfies (1.4)–(1.5). Let be defined by (4.5) with . Then for any and ,
|
|
|
|
|
|
|
|
|
(4.7) |
where and depends only on , , , .
Proof.
According to (4.6), we have
|
|
|
|
|
|
|
|
|
|
|
|
(4.8) |
where depends only on .
Firstly, by (4.3), we have
|
|
|
|
|
|
|
|
|
|
|
|
which, by the definition of , yields that
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
It is not hard to see that can be bounded by
|
|
|
Then we turn to , as can be handled in the same manner and has the same estimate. Precisely, can be dominated by
|
|
|
|
|
|
which, by applying estimates (3.13) and (3.14) to these two terms respectively, yields that
|
|
|
|
|
|
|
|
|
|
|
|
where Lemma 2.1 was also used and depends only on , , , . Therefore, can be bounded by
|
|
|
|
|
|
For , by the definitions of , and Lemmas 2.1, 3.2, we have
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
where we have used the fact that , and depends only on , , .
Similarly, has the same bound as .
To handle , we write
|
|
|
|
|
|
|
|
where we have used (1.10) and the notation . By using the arguments of together with Lemma 2.2 and Remark 3.1, it follows that
|
|
|
|
|
|
|
|
|
|
|
|
For the last term, we have
|
|
|
|
|
|
|
|
|
|
|
|
where we have used the fact
|
|
|
To bound , we apply Lemmas 2.2, 3.2 and 3.5 to obtain
|
|
|
where depends only on , , . Moreover, it follows from Lemmas 2.2 and 3.2 that
|
|
|
|
|
|
|
|
where and depends only on , . On the other hand,
|
|
|
|
|
|
|
|
where we have applied Lemma 3.2 in the second step. As a result, can be bounded by
|
|
|
|
|
|
By combining (4.8) and the estimates of –, we conclude the desired estimate.
∎
Theorem 4.1.
Suppose the assumptions of Lemma 4.1 hold. Then for ,
|
|
|
where depends only on and the Lipschitz character of . In particular,
|
|
|
Proof.
Note that for all and . By setting and in (4.7), we have
|
|
|
By Lemma 3.6, it is not hard to see that
|
|
|
Similarly, it follows from Lemmas 3.6 and 3.9 that for
|
|
|
On the other hand, thanks to Lemma 3.8 and Corollary 3.2, we have
|
|
|
(4.9) |
This completes the proof.
∎
Corollary 4.1.
Suppose the assumptions of Lemma 4.1 hold. Then for any ,
|
|
|
Now we prove the optimal -convergence rate in stated in Theorem 1.1. We first introduce the dual problem. For , let () be the weak solution to the following problem
|
|
|
(4.10) |
where is the adjoint operator of (). By setting , one can see that solves
|
|
|
(4.11) |
where is the operator given by (1.2) with replaced by and . Observe that satisfies the same conditions on as . Thus, the process and results discussed above for problem (1.1) remain valid for problem (4.11), thereby holding for problem (4.10). Especially, the correctors and flux correctors could be introduced for the operator (see also (2.6)).
Define
|
|
|
(4.12) |
Then has the same estimates as . As like Theorem 4.1, we have,
Corollary 4.2.
Let be defined by (4.12). Then
|
|
|
where depends only on and the Lipschitz character of .
Lemma 4.2.
Suppose is a bounded domain in and satisfies (1.4)–(1.5). Let be the solution to problem (4.10) with . Then
|
|
|
(4.13) |
where depends only on and the character of given by (4.15). Consequently,
|
|
|
(4.14) |
Proof.
Due to Corollary 2.1, . We claim that satisfies the so-called condition
|
|
|
where
|
|
|
(4.15) |
Indeed, let and we decompose into intervals , . Set
|
|
|
Then is finitely-valued w.r.t. and for each . By using inequality (3.9), together with Hölder’s inequality, we calculate that
|
|
|
|
|
|
|
|
which yields that is a VMO function on for each . Since is finitely-valued, we have
|
|
|
(4.16) |
Moreover, since , it holds that
|
|
|
(4.17) |
For , ,
|
|
|
|
|
|
|
|
|
|
|
|
|
|
|
which, together with (4.16)–(4.17), yields that .
Now, thanks to - estimates of non-divergence type parabolic systems with coefficients in cylinders (see [9] and references therein for the problems on the whole space and half space, from which one can deduce the estimates for bounded cylinders), we have
|
|
|
|
|
|
where we have also used - estimates of divergence type parabolic systems with coefficients [8] in the last inequality, and depends only on . This gives (4.13). (4.14) now follows from the same argument as (4.9).
∎
Armed with the previous results, we are now prepared to prove Theorem 1.1 by using the duality argument initiated in [24]. See also [13].
Proof of Theorem 1.1.
By interpolation, we deduce from Lemmas 2.1 and 2.2 that
|
|
|
Note that , as . Then Lemma 3.2, together with Remark 3.1, implies that
|
|
|
|
|
|
|
|
|
Thus, it is sufficient to bound . To do this, let be the solution of problem (4.10) and be given by (4.12). Then we have
|
|
|
|
|
|
|
|
|
(4.18) |
For , by Corollaries 4.1 and 4.2, we have
|
|
|
where we have used estimate (4.13) in the last inequality, and depends only on , , , and the character of (which can be reduced to the character of ).
To bound , we use (4.7) with to obtain
|
|
|
|
|
|
|
|
|
|
|
|
where we have used estimates (3.20), (4.9) and Lemma 3.9 in the first inequality, as well as (4.13)–(4.14) in the last one.
To deal with , we start from (4.8) with
|
|
|
Note that the latter term in can be written into
|
|
|
which has the same form as the former one, and is more regular than . This means the latter term can be handled in the same way as the former one. Similarly, taking the former term as the test function into (4.8), we find
|
|
|
where these two terms are in the same form and is more regular than . Thus, in view of in the proof of Lemma 4.1, the key is to bound
|
|
|
|
|
|
|
|
|
|
|
|
The rest terms can be handled in the same manner. Especially, the terms of boundary layers can be estimated by (4.9) and (4.14).
To bound , we write
|
|
|
|
|
|
|
|
where the second term can be handled by Lemma 3.7. For the first term, one can see from estimate (3.13) that it can be bounded by
|
|
|
|
|
|
|
|
|
where Lemma 3.9 was also used. Therefore, it follows from estimate (4.13) that
|
|
|
For , by using (3.15), we see that
|
|
|
|
|
|
|
|
|
|
|
|
where Lemmas 2.1, 3.9 and 4.2 were used and depends only on , , the character of .
To deal with , by Lemma 3.2, we have
|
|
|
|
|
|
|
|
|
|
|
|
where we have also used Lemmas 2.2, 3.5, 3.9 and 4.2. Consequently, we have
|
|
|
In view of the estimates of and (4.18), we get
|
|
|
which, by duality, yields the desired result. The proof is completed.
∎