Doubling inequalities and nodal sets in periodic elliptic homogenization
Abstract.
We prove explicit doubling inequalities and obtain uniform upper bounds (under -dimensional Hausdorff measure) of nodal sets of weak solutions for a family of linear elliptic equations with rapidly oscillating periodic coefficients. The doubling inequalities, explicitly depending on the doubling index, are proved at different scales by a combination of convergence rates, a three-ball inequality from certain “analyticity”, and a monotonicity formula of a frequency function. The upper bounds of nodal sets are shown by using the doubling inequalities, approximations by harmonic functions and an iteration argument.
Key words and phrases:
Periodic homogenization, Doubling inequalities, Nodal sets2010 Mathematics Subject Classification:
35A02, 35B27, 35J15.1. Introduction
The paper is concerned with doubling inequalities and upper bounds of nodal sets of solutions in periodic elliptic homogenization. We consider a family of elliptic operators in divergence form with rapidly oscillating periodic coefficients
(1.1) |
where , and is a symmetric matrix-valued function in with dimension . Assume that satisfies the following assumptions:
-
•
Strong ellipticity: there is such that
(1.2) -
•
Periodicity:
(1.3) -
•
Lipschitz continuity: There exists a constant such that
(1.4)
The doubling inequality describes quantitative behavior to characterize the strong unique continuation property, which has important applications in inverse problems, control theory and the study of nodal sets of eigenfunctions. For harmonic functions or solutions of general elliptic equations in divergence form with Lipschitz coefficients, the doubling inequality is a consequence of a monotonicity formula or Carleman estimates; see [6, 8, 9, 26]. In periodic elliptic homogenization, the first doubling inequality was obtained recently by Lin and Shen [15] with an implicit dependence on the doubling index. Precisely, they proved that if is a weak solution of in and
(1.5) |
then for any ,
(1.6) |
where depends only on and . The point here is that the constant is independent of the small parameter . This cannot be derived directly from the classical doubling inequality as the Lipschitz constant of the coefficients blows up as approaches zero. However, it is not known that how the constant in (1.6) depends on , because (1.6) was proved by a compactness argument. We mention that if , the classical doubling inequality shows that for some ; also see Lemma 3.2.
On the other hand, the Hadamard three-ball inequality also describes the quantitative unique continuation property. In periodic elliptic homogenization, two different versions of the three-ball inequality with error terms were discovered in [2] and [11]. In general, the three-ball inequalities with errors are weaker than the doubling inequalities, as they alone do not imply the strong unique continuation.
Our first goal of this paper is to find an explicit estimate for the constant in the doubling inequality in periodic elliptic homogenization. The explicit doubling inequality not only provides more clear quantitative information for the solutions (such as the vanishing order), but also has more applications. We state the result as follows.
Theorem 1.1.
Assume that satisfies the conditions (1.2), (1.3) and (1.4). Let be a weak solution of in .
-
(i)
For and every , there exist and , depending only on and , such that if satisfies
(1.7) then for every ,
(1.8) -
(ii)
For , there exists a constant depending only on and such that if satisfies
(1.9) then for every ,
(1.10)
The double exponential growth for in (1.8) and sub-exponential growth for in (1.10) seem to be the best we can obtain from the existing tools and results; see Remark 4.5. Our ultimate hope is for an estimate of the form , for some depending on and . Such an estimate would have very important consequences for the study of long-standing open problems regarding the spectral properties of second order elliptic operators with periodic coefficients and their quantitative unique continuation properties (see for instance Conjecture 6.13, Theorem 6.15 and Conjecture 6.16 in [12]). This connection between the conjectured optimal doubling estimates and Conjecture 6.13 in [12] was observed by the first author, D. Mendelson and C. Smart in the fall of 2019. This motivated the current work.
As a straightforward corollary, Theorem 1.1 implies that the vanishing order of at the origin does not exceed for and for . Theorem 1.1 also implies a three-ball inequality without an error term, in contrast to the results in [2] and [11], namely (e.g., for ),
(1.11) |
for any .
The proof of Theorem 1.1 breaks down into three steps:
-
•
Step 1: . In this case, we take advantage of the convergence rate in homogenization theory and use the precise three-ball inequality of harmonic functions. The smoothness of the coefficients is not needed in this step.
-
•
Step 2: In this step, we need to use “analyticity”, which distinguishes between and . For , we let , and use a three-ball inequality with a sharp exponential error term proved recently in [2] by Armstrong, Kuusi and Smart, which is a consequence of the “large-scale analyticity” from periodic homogenization. This will lead to a nontrivial improvement on the exponent so that in Theorem 1.1 can be arbitrarily small. Again in this case, the periodic structure will play a role; but the smoothness of coefficients is still not required. For , we let , and apply a doubling inequality derived from quasi-regular mappings [1] (related to complex analyticity), which requires no smoothness or periodicity on the coefficients. Unfortunately, this method works only in two dimensions.
-
•
Step 3: for or for . In this case, the classical doubling inequality for elliptic operators with Lipschitz coefficients can be handled by a monotonicity formula for the frequency function. If , the Lipschitz constant of the coefficients turns out to be after rescaling. A careful calculation shows that the constant in (1.8) is at least , if the periodicity is not used. If , the Lipschitz constant of the coefficients after rescaling is bounded by , independent of and . This allows us to obtain a much better estimate in two dimensions.
For , one will see in the proof that the estimate in Step 3 leads to the double-exponential growth of the constant in (1.8). What happens when ? To gain some intuition, consider a typical harmonic function in (see [6] or [7]). Note that
By setting , we see that the intrinsic frequency of (i.e., the number of times that changes signs) is approximately . Now, let be a weak solution of whose limit is (the homogenized solution) as . In view of the interior first-order approximation , the intrinsic frequency of will interact with the frequency of oscillation of the corrector . Particularly, under rescaling, if , the frequency of oscillation of the rescaled coefficients (or correctors) is comparable to the intrinsic frequency of . Note that the intrinsic frequency does not change under rescaling. It seems that the resonance between these two frequencies causes the failure of the arguments in Step 1 and Step 2 when (note that can be arbitrarily small and thus is close to the resonant situation), and we do not have a tool to handle this situation (except for ). We believe that an effective argument should take advantage of both the periodicity and the Lipschitz continuity of the coefficients.
Our second goal is to obtain an upper bound for the nodal sets of solutions in periodic elliptic homogenization. The study of the -dimensional Hausdorff measure of nodal sets centers around Yau’s conjecture for Laplace eigenfunctions on smooth manifolds:
(1.12) |
where is a compact smooth Riemannian manifold without boundary. It was conjectured in [25] that the bounds of nodal sets of eigenfunctions in (1.12) are controlled by
(1.13) |
where depend only on the manifold and denotes the -dimensional Hausdorff measure. The conjecture (1.13) was shown for real analytic manifolds by Donnelly-Fefferman in [5]. Lin [14] also proved the upper bound for the analytic case, using an approach by frequency functions. We should mention that, by a lifting argument, Yau’s conjecture can be reduced to studying the nodal sets of harmonic functions on smooth manifolds. In recent years, there was an important breakthrough made by Logunov and Malinnikova [18], [16] and [17]. A polynomial upper bound was given in [16] and the sharp lower bound in the conjecture was shown in [17]. We are interested in the upper bound of nodal sets for with rapidly oscillating periodic coefficients. The study of nodal sets in homogenization was initiated by Lin and Shen [15], where an implicit upper bound depending on the doubling index was shown. We are able to provide an explicit upper bound.
Theorem 1.2.
The strategy of the proof is as follows. For relatively large , we adapt a blow-up argument to obtain the upper bounds of nodal sets. For small , the solution can be approximated by a harmonic function , and thus the nodal set of is a small perturbation of the nodal set of . We then derive a quantitative estimate for the nodal set of by carefully studying the small perturbations near the nodal set and critical set of , which has its root in the analogous qualitative estimates obtained in [15]. By iterating such quantitative estimate, we are able to show the upper bound for the nodal sets of . The restriction for arises from the doubling inequality (5.2) for . If we consider to be for some large constant , which is the case for the doubling inequality of eigenfunctions, the upper bounds of nodal sets are double exponential functions . In this sense, the restriction only affects the constant in such upper bounds, which does not play an important role. For , we point out that there is no misprint in the exponential (compared to ). We still have the exponential, because in this situation, instead of the doubling inequality in (1.10), the suboptimal quantitative stratification of the critical set of harmonic functions [22] dominates the upper bound.
The paper is organized as follows. Section 2 is devoted to a doubling inequality at relatively large scales by the homogenization theory. In section 3, we derive the doubling inequality, using frequency functions, and show how it depends on the large Lipschitz constant of the coefficients. Then, Theorem 1.1 is proved in section 4 and Theorem 1.2 is proved in section 5. Throughout the paper, the letters , , , , , denote positive constants that do not depend on or , and they may vary from line to line.
Acknowledgements: Parts of this work were carried out during the second author’s visit to the Department of Mathematics at the University of Chicago during January-March 2020. The first author would like to thank D. Mendelson and C. Smart for many insightful discussions on doubling inequalities in periodic homogenization. The second author would like to thank the Department of Mathematics at Chicago for the warm hospitality and the wonderful academic atmosphere. The second author also would like to thank F. Lin for helpful discussions on three-ball inequalities in periodic homogenization. The third author would like to thank D. Mendelson for insightful discussions during the early stages of this work.
2. Homogenization
In this section, we deal with the case for all dimensions. Indeed, we will prove a quantitative version of [15, Theorem 3.1].
Let be the homogenized operator and be the homogenized coefficient matrix of (see, e.g., [24] for the general theory of periodic elliptic homogenization). Define the ellipsoid
The following is the main theorem of this section.
Theorem 2.1.
Lemma 2.2.
Let . Suppose is a solution of in satisfying
For any , there exist , depending only on and , such that if , then for any
(2.1) |
Proof.
Let , to be determined. Since , by the Caccioppoli inequality, we have
By the co-area formula, we can find some so that
Without loss of generality, let us simply assume . Hence . By [10, Theorem 1.1],
(2.2) |
where is the solution of and on and is arbitrary.
As a result, we have
(2.3) | ||||
Also,
We will choose so that . Consequently,
(2.4) |
Inserting this into (2.3), we have
(2.5) | ||||
where we have used the simple fact that for and enlarged the constant in the last inequality.
Next, by the interior estimate for -harmonic functions, we have
Inserting this into (2.5) and choosing sufficiently small so that , we obtain
Choose . We arrive at
(2.6) |
Note that the above calculation goes through only if and . This implies that we require
for some large constant .
Recall that is a weak solution of in . Let . Then in and (2.6) is equivalent to
(2.7) |
Now, as a consequence of the well-known three-sphere theorem for harmonic functions,
(2.8) |
is a convex function in and therefore is a nondecreasing function in , for any fixed . Hence, we obtain from (2.7) that for any ,
(The doubling index with is an increasing function of radius.) Again, this is equivalent to
(2.9) |
for any .
Now, if , the above lemma allows us to iterate (2.1) down to . Precisely, if and
with , then
where
provided .
Lemma 2.3.
For all with and sufficiently small, one has
Proof.
Define . Then and
It follows that
where . The above inequality yields
(2.10) |
We prove by induction that if is sufficiently small, then and for all . Actually, if
and for all , then it is easy to see from (2.10) that and . This proves the desired estimate. ∎
Remark 2.4.
Remark 2.5.
It is not difficult to see that (2.1) implies the following three-ball inequality with an error term
(2.11) |
for any and . Compared to the three-ball inequalities in [2] (see Theorem 4.1 below) and [11], our major term on the right-hand side of (2.11) is sharp. In particular, if , (2.11) recovers precisely the three-ball inequality for -harmonic functions.
Theorem 2.6.
Given arbitrary , there exists depending only on and such that if in and
(2.12) |
then for any , we have
3. Dependence on the Lipschitz constant
In this section, we derive the doubling inequality with a large Lipschitz constant, which will be used in the Step 3 of the proof of Theorem 1.1. We aim to show how the Lipschitz character of the coefficients plays a role in quantitative unique continuation, which seems to be largely unexplored. Assume that
(3.1) |
where satisfies (1.2) and
(3.2) |
for some large positive constant . We emphasize that throughout this section, the constant will never depend on . Since the norm and the norm of are comparable, parallel to the assumption (1.7), we may assume the following
(3.3) |
for some large constant .
In order to define the frequency function later, we need to construct the geodesic polar coordinates. The construction of polar coordinates has been obtained in [3]. We adopt a slightly different construction of the metric from [7, Chapter 3.1]. We follow the construction with an eye on the explicit dependence of the Lipschtiz constant . For , we define the Lipschitz metric as follows
(3.4) |
where is the entry of . The case will be discussed in Remark 3.3. Note that is Lipsthitz continuous and satisfies
(3.5) |
Define
(3.6) |
and
From (3.6), we can also write
Thus, we can check that is a non-negative Lipschitz function satisfying
(3.7) |
where depends only on and . We introduce a new metric by setting
(3.8) |
We can write the metric in terms of the intrinsic geodesic polar coordinates ,
(3.9) |
where satisfies
(3.10) |
and depends only on and .
The existence of the geodesic polar coordinates allows us to consider geodesic balls. Denote by the geodesic ball in the metric of radius and centred at the origin. In particular, from (3.6) and (3.9), is the geodesic distance from to the origin in the new metric . Thus, it is conformal to the usual Euclidean ball. For convenience of presentation, we may assume that the geodesic balls coincide with the Euclidean balls, i.e., .
Let
(3.11) |
Obviously, is a Lipschitz function satisfying
(3.12) |
where and depend on and . In the polar coordinates,
(3.13) |
In this new metric , the equation (3.1) can be written as
(3.14) |
Let
(3.15) |
and
(3.16) |
where represents the area element of under the metric . We define the frequency function by
(3.17) |
For future application, we will also use the notation to specify the center of the ball in the definition of frequency function.
Lemma 3.1.
Let be a nontrivial solution of (3.1). There exists a positive constant depending on and such that
(3.18) |
is a non-decreasing function of .
Proof.
The proof of the lemma is essentially contained in [6]. Since we want to show the explicit dependence of the Lipschtiz constant in the estimates, we sketch the proof by considering the role of . Taking derivative with respect to for , we have
(3.19) |
In order to prove the lemma, it suffices to show
(3.20) |
Thus, we consider the derivatives of and , respectively. Setting . Note that . We write as
(3.21) |
Taking derivative with respect to , one has
(3.22) |
where on . By (3.10), (3.12) and (3.13), we have
(3.23) |
Multiplying both sides of (3.14) by and performing the integration by parts give that
(3.24) |
It follows that
(3.25) |
Next we derive the doubling inequality with an explicit dependence on .
Lemma 3.2.
Proof.
From (3.25) and the definition of , we have
(3.28) |
Note that here is a function in satisfying . We would like to obtain an upper bound and a lower bound for the quotient with . To find the upper bound, we integrate the equality (3.28) from to and use the monotonicity of to obtain
(3.29) |
Taking the exponential of both sides gives the upper bound
(3.30) |
To see the lower bound, we integrate (3.28) from to and apply the monotonicity of again to obtain
(3.31) |
Raising to the exponential form, we have
(3.32) |
Combining (3.30) and (3.32), we arrive at
(3.33) |
Next we want to show an upper bound for . Let and . From the estimate (3.32), we have
(3.34) |
Using the fact that , we have
(3.35) |
where depends on and . Obviously,
(3.36) |
Therefore, from (3.3), (3.32) and (3.35), we have
(3.37) |
Thus, we can get an upper bound for as
(3.38) |
where is a large constant. Choosing any , we integrate (3.28) from to , by the monotonicity of , we derive that
(3.39) |
Thus, we obtain that
(3.40) |
where is large. By further integrations, we can also obtain that
(3.41) |
for , where depends only on and . ∎
Remark 3.3.
For the case , we introduce a new variable to apply a lifting argument. Let . Then the new function satisfies the equation
(3.42) |
where
and is the ball with radius in . It is easy to see that satisfies the conditions (1.2) and (3.2). Following the procedure performed as , we are able to introduce the new metric and geodesic polar coordinates. Thus, in the metric as (3.8) and as (3.13), we have
(3.43) |
where . As before, we could make use of the monotonicity of the frequency function to obtain the doubling inequality. Precisely, we may define
(3.44) |
and
(3.45) |
Then the frequency function is defined as
(3.46) |
Following the proof of Lemma 3.1 and [7, Theorem 3.2.1], we can obtain the almost monotonicity of . That is, for any , it holds that
(3.47) |
for any where depends on . By mimicking the argument in the proof of Lemma 3.2, we can obtain the doubling inequality for in . This also leads to the doubling inequality for as (3.27) in .
Remark 3.4.
For a better estimate when , see Remark 4.6.
4. Proof of Theorem 1.1
This section is devoted to the proof of Theorem 1.1. Step 1 and Step 3 of the proof have been handled in Section 2 and Section 3, respectively. For convenience of presentation, we choose such that . Our argument works for any . To handle the case in Step 2, we will use particular doubling properties, obtained from some sort of “analyticity”, as a transition in order to improve our estimates. Indeed, for , we will employ the three-ball inequality with a sharp exponential error term obtained in [2]; for , we will use quasi-regular mappings [1] which provides a much better doubling estimate.
We first introduce a three-ball inequality for all dimensions . For convenience, we define the normalized norm by
The following theorem is essentially taken from [2, Theorem 1.4], which is a corollary of the “large-scale analyticity” in periodic homogenization. This result relies on the periodic structure of the coefficients, but does not depend on the smoothness of coefficients.
Theorem 4.1.
For each , there exist and such that if is a weak solution of in with , then
(4.1) |
As a simple corollary, we have
Corollary 4.2.
Let be a weak solution of in . For every , there exist and such that if
(4.2) |
and
(4.3) |
then
(4.4) |
The sharp exponential tail in (4.1) is crucial for our purpose which is related to the condition (4.2). The lower bound in (4.2) allows us to iterate the estimate down to a scale at which the classical theory in Section 3 may apply.
Next, for the case , we introduce a stronger doubling property using quasi-regular mappings (related to complex analyticity). We briefly give some background on quasi-regular mappings. For a detailed account of this topic, please refer to the presentation in [1], [4, Chapter II.6] and references therein.
Let be a weak solution of the equation in with only bounded measurable coefficients satisfying (1.2). Let for . Define
(4.5) |
We introduce a stream function (the generalized harmonic conjugate) associated with as
where is the rotation matrix in the plane
Let . Then we have and satisfies
where the complex valued function and can be explicitly written in term of and
Hence, is a -quasi-regular mapping. Moreover, it can be written as , where is holomorphic and is a -quasiconformal homeomorphism satisfying and . Define
The quasi-balls are comparable to the standard Euclidean balls in the sense
(4.6) |
where and depend only on . Observe that tends to be singular if , which fortunately is not too restrictive as we only use it in the transition at intermediate scales.
From the fact that is a holomorphic function, the following doubling property holds [1].
Lemma 4.3.
If is a nonzero weak solution of in , then
(4.7) |
Remark 4.4.
Note that Lemma 4.3 does not use periodicity, and it is also true for solutions of , with a constant independent of . This gives “almost monotonicity of the doubling constant”, a statement stronger than that of Theorem 1.1, but for quasi-balls, as opposed to the usual balls. As we pointed out above, quasi-balls are difficult to manage as the radius goes to zero. Because of this, we still need to use the periodicity assumption and Step 1 below, when , to cover the range . We then apply Lemma 4.3 in the range , using (4.6) since is not too small. Finally, the case is handled by scaling and Lemma 3.2. The details are below.
Proof of Theorem 1.1.
According to the relationship between and , one needs to consider three cases based on the comparison of with and (or for ). Without loss of generality, we may just consider the most complicated case , since all the three steps listed in the introduction will be involved as approaches 0. Hence, we fix and so that , and then discuss the different ranges of .
Step 1: . Under either (1.7) or (1.9), Theorem 2.6 implies
(4.8) |
for any given . This estimate holds for all dimensions .
Step 2: In this step, we need to treat the cases and separately.
Case 1: and for any fixed . Let be the smallest integer so that . If is bounded by some absolute constant, then Step 2 is not needed. Since , for sufficiently large , satisfies
(4.9) |
Because of (4.8), we have
(4.10) |
Let and be the constant such that
(4.11) |
The goal is to estimate with comparable to the bound in (4.9).
Thanks to Corollary 4.2, and by rescaling, we know that for a given with small enough, we have
(4.12) |
Note that the left-end restriction is needed in order to apply Corollary 4.2, due to (4.2). This can be guaranteed if we eventually show .
We now proceed to estimate . Using the initial condition , one can show explicitly that
(4.13) |
It follows from (4.9) that
(4.14) |
Note that is any given positive constant. Then, we may choose small enough (hence is also small), so that
(4.15) |
Thus, if is large enough,
(4.16) |
This implies that for any , we have
(4.17) |
Case 2: and . From (4.8) with in Step 1, for ,
By the norm estimates, it follows that
(4.18) |
We would like to to apply Lemma 4.3 to . From the relation (4.6) of quasi-balls and the standard balls, as well as the iteration of the doubling inequality (4.18), we have
(4.19) |
where depends only on . Thus, (4.7) implies that for any
(4.20) |
In order to establish a doubling inequality at small scale on standard Euclidean balls, we iterate the above doubling inequality times to obtain
(4.21) |
By the relation (4.6),
We choose to be the smallest integer so that
(4.22) |
Consequently,
(4.23) |
Note that , satisfying , is arbitrary and is chosen depending on . We now assume . Hence, . Moreover, from (4.22), we have , where depends only on . Thus, it follows from (4.23) that
(4.24) |
for all . Since norm can be replaced by norm in the above inequality, we derive the desired estimate for the case .
Step 3: For (or for ), by rescaling, the equation may be reduced to the case in which the Lipschitz constant of coefficients is bounded by (bounded by for ). It follows from (3.27) and (4.17) that for and any ,
For , it follows from (3.27) and (4.24) that for any ,
(4.25) |
Note that . This completes the proof of Theorem 1.1. ∎
Remark 4.5.
It was shown in [2] that the exponential tail in (4.1) is sharp (up to the end point ), without any smoothness assumption on the coefficients. If the critical in (4.1) can also be achieved (which seems like a very difficult task), then Corollary 4.2 with would follow. By the argument in Step 2, this would yield the estimate
(4.26) |
for . If we then apply (3.27) as in Step 3, with Lipschitz constant , we would obtain the bound for (for the range , (3.27) does give the optimal bound). On the other hand, the estimate (3.27) in term of the large Lipschitz constant may not be sharp. This is a well-known difficult issue in quantitative unique continuation, for which none of the currently known methods apply. Any improvement here would have many consequences. Alternatively, in the range , one could try to use a method taking advantage of both periodicity and smoothness. No such method is available at the moment.
For , if the critical case in (4.1) is true, then we can actually show our expected estimate which has very important consequences for the study of long-standing open problems regarding the spectral properties of second order elliptic operators with periodic coefficients. Note though that when , these problems have already been solved in [20, 21] (also see [13, subsection 7.3, 7.4]). However, the approach just outlined would require lower regularity on the coefficients than [20, 21, 13]. To obtain the expected estimate assuming this critical case holds, note that (4.26) holds for . By reproducing the argument in Step 2 (Case 2), we can then show and therefore
(4.27) |
for all . Then a blow-up argument gives the same estimate for . Observe that (4.27) is exactly the ultimate estimate we expect, as mentioned in the introduction.
Remark 4.6.
If we consider, when , elliptic operators with Lipschitz coefficients, with Lipschitz constant (and no periodicity assumption), we can obtain the improved bound in Lemma 3.2. To show this, we break down the scales into and . For the case , we use Lemma 4.3, (4.6) and the argument from (4.19) to (4.23). For the case , we scale to reduce to the case and then apply Lemma 3.2 as it stands. This may suggest that the bound in Lemma 3.2 is not optimal, also for .
Remark 4.7.
The disadvantage of Theorem 4.1 for is that may be very small. If we do not apply Theorem 4.1 to improve the exponent , Step 1 and 3 in the proof of Theorem 1.1 allows to be any number in . In particular, under (1.5), for any , we have
(4.28) |
For convenience, we will use this doubling inequality (4.28), instead of (1.8), in estimating the upper bound of nodal sets in the next section. The price is that has to be larger than in (1.14).
5. Upper bounds of Nodal sets
In this section, we study of the upper bounds of nodal sets for , where is a nonzero solution of satisfying (1.5). We will focus on the general treatment for all dimensions and with an eye towards in the end. Throughout this section, up to a change of variable, we assume . Note that in this case, ’s are just balls, and in view of Theorem 2.1, the assumption (1.5) can be replaced by
(5.1) |
and (4.28) holds with .
5.1. Small scales
We first show that a doubling inequality centered at implies the doubling inequality with shifted centers.
Lemma 5.1.
Let be a weak solution of in satisfying (5.1). Then for any and , we have
(5.2) |
Proof.
Let us define the nodal sets as
(5.5) |
and the density function of nodal sets as
(5.6) |
Based on Lemma 5.1 and a blow up argument, we can estimate the Hausdoff measure of the nodal set of in small balls.
Lemma 5.2.
For any and such that ,
(5.7) |
where depends on and .
Proof.
First of all, we consider the case and . Let and . Then
(5.8) |
By (1.4),
(5.9) |
for . Therefore, in this case, the coefficient matrix has a uniform Lipschitz constant independent of and . Then, a change of variable and the doubling inequality in Lemma 5.1 give that
(5.10) |
By the upper bound of nodal sets in [16], there exists a constant so that
(5.11) |
which implies, by rescaling,
for any and .
Next, to deal with the case , we simply use a covering argument. Let and . There there exists a family of balls that covers with a finite number of overlaps depending only on . Note that . Consequently,
We obtain the desired estimate by enlarging the constant . ∎
Remark 5.3.
The above lemma does not rely on the periodicity of the coefficients. Actually, its proof also gives how the estimate depends on the Lipschitz constant of the coefficients. Precisely, if is a solution of in . In addition to the ellipticity condition (1.2), we assume
(5.12) |
Then
for , where the definition of is given below in (5.27).
5.2. Large scales
To deal with the nodal sets at large scales, we need to use the homogenization theory. Precisely, in the following, we find an approximate solution , close to under norm, and satisfying a doubling inequality.
Lemma 5.4.
Suppose for some large . Let be a solution of in satisfying
(5.13) |
Then there exists satisfying in such that
(5.14) |
and
(5.15) |
where depends on and .
Proof.
By rescaling, we may assume . The construction of such locally homogenized solution and the estimate (5.14) can be found in [15, Theorem 2.3]. Note that it is not necessary that on . Then, it suffices to show (5.15). By (5.13) and (5.14), we have
(5.16) |
We now establish estimates to compare the norms of and . Thanks to (5.16),
(5.17) |
By the same strategy, using (5.16), we obtain that
(5.18) |
Since , the above estimate yields
(5.19) |
Combining (5.17) and (5.19) together yields that
(5.20) |
Now, we use the fact that
(5.21) |
is a convex function with respect to . Then is nondecreasing for any . This implies
(5.22) |
This proves (5.15) and the lemma. ∎
Remark 5.5.
Let be a ball and be a function in . In order to show some quantitative stratification results for and , we introduce the doubling index:
(5.23) |
and
(5.24) |
If is a weak solution of the equation , the doubling index for and are monotonic in the sense that
(5.25) |
and
(5.26) |
for and depending only on . This follows from (2.8) and the line after it.
We also define a variant of the above doubling index for cubes. For a cube , denote by the side length of . Define the doubling index in the cube by
(5.27) |
and
(5.28) |
The doubling index defined in cubes is convenient in the sense that if a cube is a subset of , then . Let be a subcube of and . Then
(5.29) |
where depends only on . Similarly, it also holds
(5.30) |
The following quantitative stratification for and is the key ingredient of this section. The idea of the proof originates from Lemmas 3.5 and 5.2 in [19].
Lemma 5.6.
Assume that is harmonic in .
-
(1)
Suppose . If for some , there exists a finite sequence of balls such that
(5.31) and
(5.32) where and depend only on .
-
(2)
Suppose . If for some depending on , there exists a finite sequence of balls such that
(5.33) and
(5.34)
Proof.
In the following proof, all the constants depend only on , and , are large constants.
(1) Let and , where is small to be specified later. We can assume that is an integer and . We divide the cube into equal subcubes . Then . We would like to estimate the number of cubes that intersect .
Let be a cube with . Thus, we have . We claim that if for some large , then changes sign in . Assume that does not change sign in . By the Harnack inequality,
(5.35) |
On the other hand, by the monotonicity of the doubling index in cubes (5.29),
(5.36) |
Choosing , we reach a contradiction if
Since , the last inequality holds if we choose . This proves the claim. Hence, there are zeros in each and with
(5.37) |
This implies (5.31) as we may replace by with the same center and .
Next, to show the first part of (5.32), we need to estimate the number of the cubes . Recall that and each point in may be covered by at most a finite number of . By the lower bound estimate of nodal sets in [17], we have
(5.38) |
On the other hand, by the upper bound estimate of nodal sets in [16], it holds that
(5.39) |
where . Combining (5.38) and (5.39), we arrive at
(5.40) |
which yields
(5.41) |
Thus,
(5.42) |
This proves (1).
(2) Next, we establish the estimates (5.33) and (5.34). We divide the cube into subcubes with side length . The size of , depending on , will be chosen later. The cube is called bad if
(5.43) |
for some small depending only on . We claim that the number of bad cubes is not greater than , where depends on .
To show the above claim, we need to use [22, Theorem 1.1]. Recall the effective critical set is defined as
Let be the -neighborhood of , namely, . Then [22, Theorem 1.1] implies
(5.44) |
where are concentric balls such that and is the modified frequency function defined by
where is the center of . By [7, Corollary 2.2.6] and the mean value property of harmonic functions, we have
where we have also used a gradient estimate for harmonic functions in the second inequality. Hence, (5.44) implies
(5.45) |
Next, we show that if is a bad cube with sufficiently small , then Actually if is bad and is the point in so that , then
where we used the condition (5.43) in the last inequality. Fix . Then . It follows from the gradient estimate and the Caccioppoli inequality that
In view of [7, Corollary 2.2.7], we have
where in the last inequality, we choose small so that . This implies that and . Because , we have . This means that all the bad cubes are contained in . Finally, let be comparable to and note that can be covered by finitely many, depending only on , with . Then, by (5.45), the total volume of bad cubes in is bounded by . Hence, the number of bad cubes is not greater than . The claim has been proved.
Now, for any , the monotonicity of the doubling index of in cubes in (5.30) shows that
(5.46) |
If is not bad, the reverse inequality of (5.43) yields
(5.47) |
Given , small enough (to be quantified later), we want to estimate the set defined in (5.33). If is not bad and we choose to be the smallest integer such that
(5.48) |
then (5.47) gives
This implies that does not intersect . It also shows that . Thus, the set is covered by the union of bad cubes of size . Again, we may now replace bad by with the same center and . Let be the number of bad cubes and recall that . It follows that
(5.49) |
where we have chosen in the last inequality. This completes the proof. ∎
Next we estimate the density function of nodal sets, which is the initial step for an iterative argument to obtain Theorem 1.2. The following lemma is a quantitative version of [15, Lemma 4.5]. Without loss of generality, we may identify in Lemma 5.6 by .
Lemma 5.7.
Let and (5.1) hold. If , then there exists a finite sequence of balls such that , and
(5.50) |
where and depend on , and .
Proof.
We assume . Otherwise, (5.50) is trivial. We will make use of the approximation estimate by a harmonic function in Lemma 5.4. Using (5.14) with , we have
(5.51) |
By normalization, we may assume that . We would like to estimate the doubling index for and . From (5.20) and convexity of in (5.21), we have
(5.52) |
for all . By elliptic estimates and using (5.52) twice,
(5.53) | ||||
The above estimate includes
(5.54) |
Thus, . Similarly to the frequency function in Section 3, one can define the frequency function introduced for the harmonic function as
By Theorem 2.2.8 in [7], it holds that
(5.55) |
for any and . It is known that the doubling index and the frequency function are comparable. It follows from Lemma 7.1 in [16] that
(5.56) |
where depends on . From (5.55), let , we obtain that
Using Lemma 7.1 in [16] again, we see that
(5.57) |
for any Thus, we obtain that
(5.58) |
for any and any , where depends on .
Next, we estimate the doubling index of . We first claim that there are zeros for in . In fact, from (5.51) and Theorem 2.1 with , we have
(5.59) |
Then
(5.60) |
It follows that if . Hence, (5.15) implies
(5.61) |
Now, let us assume that has no zeros in and therefore does not change signs in . Without loss of generality, we may assume that is positive. By the Harnack inequality and (5.61),
(5.62) |
From (5.51), for , we get
(5.63) |
Since , then for . This contradicts our assumption that . Thus, the claim has been shown.
Now, since has zeros in (and hence in ), we obtain from (5.53) and the mean value theorem that
(5.64) |
Again, by the relation of the frequency function for and the doubling index, we can argue as the derivation of (5.58) that (5.64) implies that
(5.65) |
for any and any .
Thanks to the monotonicity of the doubling index for and , from the definition of and , we know that and . In order to apply Lemma 5.6, we assume that and . Thus, we require for some depending on , which is satisfied by the assumption of in the lemma. With the aid of (5.51), we have
(5.66) |
where
(5.67) |
and and are given by Lemma 5.6. Thus, it follows from Lemma 5.6 that
(5.68) |
where depends on . Since is large, by the decomposition in Lemma 5.6, we may assume .
Next we estimate the upper bound for for each . We will discuss two cases and . If , by Lemma 5.2,
(5.69) |
Now, we consider the case . Note that (5.61), the fact that has zeros in and the definition of imply
(5.70) |
for any point . Fix . For , define
Then (5.70) implies that is contained in . Without loss of generality, it suffices to estimate . By the regularity of , for any , there exists a cylinder centered at , whose base is a square perpendicular to with side length , such that the height of is and
(5.71) |
where is a constant smaller than .

We would like to show that can be covered by balls with radius , where . Let be the cross section containing of the cylinder which is perpendicular to . Since and , we see that for any . Next, because of (5.71), for any and ,
This implies that if . Similarly, if . This implies that
Consequently, can be covered by balls with radius and .
Now, because can be covered by cylinders, with
such that the ’s have finite overlaps, then can be covered by balls with radius (denoted by ) , where
Note that the same estimate holds for as well for each .
5.3. Proof of Theorem 1.2
Thanks to Lemma 5.7, we are able to show the upper bound of the nodal sets of in the interior domain.
Proof of Theorem 1.2 ().
We first consider the case . Recall from (5.4) that
(5.74) |
for any . By Theorem 2.1, it follows that
(5.75) |
for and any . By (5.75) and Lemma 5.4, we derive that
(5.76) |
as in (5.15) for and . By examining the proof of Lemma 5.7, the estimates (5.75) and (5.76) guarantee that the arguments in the Lemma 5.7 hold for for and .
Let for any satisfying and . Then satisfies
(5.77) |
where . By Lemma 5.7, we have
(5.78) |
where and . By rescaling, we reduce the estimate to and obtain that
(5.79) |
Let , then . Thus,
(5.80) |
where , . Note that may not be fully contained in , since may be the centers of subcubes which intersect the boundary of (we identify the ball as a cube when we perform the subcubes decomposition). However, since . If we iterate (5.80), still stays close to . Actually, for any large , where .
Now, we iterate (5.80) to obtain the desired estimate. The estimate (5.50) yields the initial step of the iteration,
(5.81) |
Assume that is achieved at some with and . Let and . Since , we apply (5.80) to to get to the estimates of nodal sets at a smaller scale, that is,
(5.82) |
We apply (5.80) repeatedly down to the case or the case that is empty. Note that . Thus, we derive that
(5.83) |
where we have used (5.7) in the second inequality. This proves the desired estimate for the case .
Following the above proof of Theorem 1.2 for , we sketch the proof of upper bounds of nodal sets in .
Proof of Theorem 1.2 ().
Since the proof is parallel to , we only present the changes for . Thus, we only present the changes for . By the argument in Lemma 5.2 and the doubling inequality (1.10), we can obtain for ,
(5.84) |
On the other hand, the statements of (5.31) and (5.32) still hold for ; while for , we have a better bound for . Precisely for , [22, Theorem B.1] implies
(5.85) |
By mimicking the argument in part (2) of Lemma 5.6, we can show (5.33) and (5.34) with . Following the proof of Lemma 5.7, we may show that if , then
(5.86) |
with , . Observe that the quantitative stratification of critical sets , instead of the doubling inequality, plays the dominant role in the estimate (5.86). We iterate (5.86), as in the proof for , to get
This provides the desired estimate (1.15) for . For , (5.84) yields the desired estimate directly. This ends the proof. ∎
References
- [1] Giovanni Alessandrini and Luis Escauriaza, Null-controllability of one-dimensional parabolic equations, ESAIM Control Optim. Calc. Var. 14 (2008), no. 2, 284–293. MR 22394511
- [2] Scott Armstrong, Tuomo Kuusi, and Charles Smart, Large-scale analyticity and unique continuation for periodic elliptic equations, To appear in Comm. Pure Appl. Math., arXiv:2005.01199 (2020).
- [3] Nachman Aronszajn, Andrzej Krzywicki, and Jacek Szarski, A unique continuation theorem for exterior differential forms on Riemannian manifolds, Ark. Mat. 4 (1962), 417–453. MR 140031
- [4] Lipman Bers, Fritz John and Martin Schechter, Partial differential equations, with special lectures by Lars Gårding and A. N. Milgram, Interscience, New York, 1964.
- [5] Harold Donnelly and Charles Fefferman, Nodal sets of eigenfunctions on Riemannian manifolds, Invent. Math. 93 (1988), no. 1, 161-183. MR 943927
- [6] Nicola Garofalo and Fanghua Lin, Monotonicity properties of variational integrals, weights and unique continuation, Indiana Univ. Math. J. 35 (1986), no. 2, 245–268. MR 833393
- [7] Qing Han and Fanghua Lin, Nodal sets of solutions of elliptic differential equations, Book available on Han’s homepage (2013).
- [8] David Jerison and Gilles Lebeau, Nodal sets of sums of eigenfunctions, Harmonic analysis and partial differential equations (Chicago, IL, 1996) Chicago Lectures in Math., Univ. Chicago Press, Chicago, IL, 1999, pp. 223-239. MR 1743865
- [9] Carlos E. Kenig, Some recent applications of unique continuation, Recent developments in nonlinear partial differential equations, Contemp. Math., vol. 439, Amer. Math. Soc., Providence, RI, 2007, pp. 25–56. MR 2359019
- [10] Carlos E. Kenig, Fanghua Lin, and Zhongwei Shen, Convergence rates in for elliptic homogenization problems, Arch. Ration. Mech. Anal. 203 (2012), no. 3, 1009–1036. MR 2928140
- [11] Carlos E. Kenig and Jiuyi Zhu, Propagation of smallness in elliptic periodic homogenization, SIAM J. Math. Anal. 53 (2021), no. 1, 111–132. MR 4194318
- [12] Peter Kuchment, An overview of periodic elliptic operators, Bull. Amer. Math. Soc. (N.S.) 53 (2016), no. 3, 343–414. MR 3501794
- [13] Peter Kuchment and Sergei Levendorskiî, On the structure of spectra of periodic elliptic operators, Trans. Amer. Math. Soc. 354 (2002), no. 2, 537–569. MR 1862558
- [14] Fanghua Lin, Nodal sets of solutions of elliptic equations of elliptic and parabolic equations, Comm. Pure Appl Math. 44 (1991), no. 3, 287-308. MR 1090434
- [15] Fanghua Lin and Zhongwei Shen, Nodal sets and doubling conditions in elliptic homogenization, Acta Math. Sin. (Engl. Ser.) 35 (2019), no. 6, 815–831. MR 3952693
- [16] Alexander Logunov, Nodal sets of Laplace eigenfunctions: polynomial upper estimates of the Hausdorff measure, Ann. of Math. (2) 187 (2018), no. 1, 221–239. MR 3739231
- [17] by same author, Nodal sets of Laplace eigenfunctions: proof of Nadirashvili’s conjecture and of the lower bound in Yau’s conjecture, Ann. of Math. (2) 187 (2018), no. 1, 241–262. MR 3739232
- [18] Alexander Logunov and Eugenia Malinnikova, Nodal sets of Laplace eigenfunctions: estimates of the Hausdorff measure in dimension two and three, 50 years with Hardy spaces, 333–344, Oper. Theory Adv. Appl., 261, Birkhäuser/Springer, Cham, 2018. MR 3792104
- [19] by same author, Quantitative propagation of smallness for solutions of elliptic equations, Proceedings of the International Congress of Mathematicians—Rio de Janeiro 2018. Vol. III. Invited lectures, World Sci. Publ., Hackensack, NJ, 2018, pp. 2391–2411. MR 3966855
- [20] Abderemane Morame, Absence of singular spectrum for a perturbation of a two-dimensional Laplace-Beltrami operator with periodic electromagnetic potential, J. Phys. A, 31 (1998), no. 37, 7593–7601. MR 1652918
- [21] by same author, The absolute continuity of the spectrum of Maxwell operator in a periodic media, J. Math. Phys. 41 (2000), no. 10, 7099–7108. MR 1781426
- [22] Aaron Naber and Daniele Valtorta, Volume estimates on the critical sets of solutions to elliptic PDEs, Comm. Pure Appl. Math. 70 (2017), no. 10, 1835–1897. MR 3688031
- [23] Weisheng Niu and Yao Xu, Uniform boundary estimates in homogenization of higher-order elliptic systems, Ann. Mat. Pura Appl. (4) 198 (2019), no. 1, 97–128. MR 3918621
- [24] Zhongwei Shen, Periodic homogenization of elliptic systems, Oper. Theory Adv. Appl., 269, Birkhäuser/Springer, Cham, 2018. MR 3838419
- [25] Shing-Tung Yau, Problem section, Seminar on Differential Geometry, Annals of Mathematical Studies 102, Princeton, 1982, 669–706.
- [26] Jiuyi Zhu, Quantitative uniqueness of elliptic equations, Amer. J. Math. 138 (2016), no. 3, 733–762. MR 3506384