This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Doubling inequalities and nodal sets in periodic elliptic homogenization

Carlos E. Kenig Department of Mathematics
University of Chicago
Chicago, IL 60637, USA
Email: [email protected]
Jiuyi Zhu Department of Mathematics
Louisiana State University
Baton Rouge, LA 70803, USA
Email: [email protected]
 and  Jinping Zhuge Department of Mathematics
University of Chicago
Chicago, IL 60637, USA
Email: [email protected]
Abstract.

We prove explicit doubling inequalities and obtain uniform upper bounds (under (d1)(d-1)-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 sets
2010 Mathematics Subject Classification:
35A02, 35B27, 35J15.
Kenig is supported in part by NSF grant DMS-1800082; Zhu is supported in part by NSF grant OIA-1832961

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

ε=(A(x/ε)),\displaystyle\mathcal{L}_{\varepsilon}=-\nabla\cdot\big{(}A(x/\varepsilon)\nabla\big{)}, (1.1)

where ε>0\varepsilon>0, and A(y)=(aij(y))A(y)=\big{(}a_{ij}(y)\big{)} is a symmetric d×dd\times d matrix-valued function in d\mathbb{R}^{d} with dimension d2d\geq 2. Assume that A(y)A(y) satisfies the following assumptions:

  • Strong ellipticity: there is Λ>0\Lambda>0 such that

    Λ|ξ|2A(y)ξ,ξ|ξ|2,for any yd,ξd.\displaystyle\Lambda|\xi|^{2}\leq\langle A(y)\xi,\ \xi\rangle\leq|\xi|^{2},\quad\quad\text{for any }y\in\mathbb{R}^{d},\xi\in\mathbb{R}^{d}. (1.2)
  • Periodicity:

    A(y+z)=A(y)for any ydandzd.A(y+z)=A(y)\quad\quad\mbox{for any }\ y\in\mathbb{R}^{d}\ \mbox{and}\ z\in\mathbb{Z}^{d}. (1.3)
  • Lipschitz continuity: There exists a constant γ0\gamma\geq 0 such that

    |A(x)A(y)|γ|xy|,for any x,yd.\displaystyle|A(x)-A(y)|\leq\gamma|x-y|,\qquad\text{for any }x,y\in\mathbb{R}^{d}. (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 uεu_{\varepsilon} is a weak solution of ε(uε)=0\mathcal{L}_{\varepsilon}(u_{\varepsilon})=0 in B2=B2(0)B_{2}=B_{2}(0) and

B2uε2NBΛuε2,\int_{B_{2}}u_{\varepsilon}^{2}\leq N\int_{B_{\Lambda}}u_{\varepsilon}^{2}, (1.5)

then for any r(0,1)r\in(0,1),

Bruε2C(N)Br/2uε2,\int_{B_{r}}u_{\varepsilon}^{2}\leq C(N)\int_{B_{r/2}}u_{\varepsilon}^{2}, (1.6)

where C(N)C(N) depends only on d,Λ,γd,\Lambda,\gamma and NN. The point here is that the constant C(N)C(N) is independent of the small parameter ε\varepsilon. This cannot be derived directly from the classical doubling inequality as the Lipschitz constant of the coefficients blows up as ε\varepsilon approaches zero. However, it is not known that how the constant C(N)C(N) in (1.6) depends on NN, because (1.6) was proved by a compactness argument. We mention that if ε=1\varepsilon=1, the classical doubling inequality shows that C(N)=CNKC(N)=CN^{K} for some C,K1C,K\geq 1; 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 C(N)C(N) 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 A=A(y)A=A(y) satisfies the conditions (1.2), (1.3) and (1.4). Let uεu_{\varepsilon} be a weak solution of ε(uε)=0\mathcal{L}_{\varepsilon}(u_{\varepsilon})=0 in B1B_{1}.

  • (i)

    For d3d\geq 3 and every τ>0\tau>0, there exist θ(0,1/2)\theta\in(0,1/2) and C>1C>1, depending only on d,τ,Λd,\tau,\Lambda and γ\gamma, such that if uεu_{\varepsilon} satisfies

    B1uε2NBθuε2,\int_{B_{1}}u_{\varepsilon}^{2}\leq N\int_{B_{\theta}}u_{\varepsilon}^{2}, (1.7)

    then for every r(0,1)r\in(0,1),

    Bruε2exp(exp(CNτ))Bθruε2.\int_{B_{r}}u_{\varepsilon}^{2}\leq\exp(\exp(CN^{\tau}))\int_{B_{\theta r}}u_{\varepsilon}^{2}. (1.8)
  • (ii)

    For d=2d=2, there exists a constant C>1C>1 depending only on Λ\Lambda and γ\gamma such that if uεu_{\varepsilon} satisfies

    B1uε2NBΛ2uε2,\int_{B_{1}}u_{\varepsilon}^{2}\leq N\int_{B_{\frac{\Lambda}{2}}}u_{\varepsilon}^{2}, (1.9)

    then for every r(0,1)r\in(0,1),

    Bruε2exp(C(lnN)2)Br2uε2.\int_{B_{r}}u_{\varepsilon}^{2}\leq\exp(C(\ln N)^{2})\int_{B_{\frac{r}{2}}}u_{\varepsilon}^{2}. (1.10)

The double exponential growth exp(exp(CNτ))\exp(\exp(CN^{\tau})) for d3d\geq 3 in (1.8) and sub-exponential growth exp(C(lnN)2)=NClnN\exp(C(\ln N)^{2})=N^{C\ln N} for d=2d=2 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 C(N)=NC(lnlnN)pC(N)=N^{{C(\ln\ln N)}^{p}}, for some p>0p>0 depending on d,Λd,\Lambda and γ\gamma. 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 uεu_{\varepsilon} at the origin does not exceed exp(CNτ)\exp(CN^{\tau}) for d3d\geq 3 and C(lnN)2C(\ln N)^{2} for d=2d=2. 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 d3d\geq 3),

Bθruε2exp(exp(CNτ))(Bθ2ruε2)τ1(Bruε2)1τ1\displaystyle\int_{B_{\theta r}}u_{\varepsilon}^{2}\leq\exp(\exp(CN^{\tau}))\bigg{(}\int_{B_{\theta^{2}r}}u_{\varepsilon}^{2}\bigg{)}^{\tau_{1}}\bigg{(}\int_{B_{r}}u_{\varepsilon}^{2}\bigg{)}^{1-\tau_{1}} (1.11)

for any 0<τ1<10<\tau_{1}<1.

The proof of Theorem 1.1 breaks down into three steps:

  • Step 1: ε/rN5\varepsilon/r\lesssim N^{-5}. 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 d3d\geq 3 and d=2d=2. For d3d\geq 3, we let N5ε/rN12τN^{-5}\lesssim\varepsilon/r\lesssim N^{-\frac{1}{2}\tau}, 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 τ>0\tau>0 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 d=2d=2, we let N5ε/r1N^{-5}\lesssim\varepsilon/r\lesssim 1, 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: ε/rN12τ\varepsilon/r\gtrsim N^{-\frac{1}{2}\tau} for d3d\geq 3 or ε/r1\varepsilon/r\gtrsim 1 for d=2d=2. 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 d3d\geq 3, the Lipschitz constant of the coefficients turns out to be O(N12τ)O(N^{\frac{1}{2}\tau}) after rescaling. A careful calculation shows that the constant in (1.8) is at least exp(exp(CNτ))\exp(\exp(CN^{\tau})), if the periodicity is not used. If d=2d=2, the Lipschitz constant of the coefficients after rescaling is bounded by CC, independent of ε\varepsilon and NN. This allows us to obtain a much better estimate in two dimensions.

For d3d\geq 3, 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 ε/rN12τ\varepsilon/r\gtrsim N^{-\frac{1}{2}\tau}? To gain some intuition, consider a typical harmonic function wk=rkcos(kα)w_{k}=r^{k}\cos(k\alpha) in 2\mathbb{R}^{2} (see [6] or [7]). Note that

B1wk2Bθwk2=Cθ2k.\frac{\int_{B_{1}}w_{k}^{2}}{\int_{B_{\theta}}w_{k}^{2}}=C\theta^{-2k}.

By setting N=Cθ2kN=C\theta^{-2k}, we see that the intrinsic frequency of wkw_{k} (i.e., the number of times that wkw_{k} changes signs) is approximately lnN/(lnθ)\ln N/(-\ln\theta). Now, let uεu_{\varepsilon} be a weak solution of ε(uε)=0\mathcal{L}_{\varepsilon}(u_{\varepsilon})=0 whose limit is wkw_{k} (the homogenized solution) as ε0\varepsilon\to 0. In view of the interior first-order approximation uεwk+εχ(x/ε)wku_{\varepsilon}\sim w_{k}+\varepsilon\chi(x/\varepsilon)\nabla w_{k}, the intrinsic frequency of wkw_{k} will interact with the frequency of oscillation of the corrector χ(x/ε)\chi(x/\varepsilon). Particularly, under rescaling, if r/εlnNr/\varepsilon\approx\ln N, the frequency of oscillation of the rescaled coefficients A(rx/ε)A(rx/\varepsilon) (or correctors) is comparable to the intrinsic frequency of wkw_{k}. 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 ε/r(lnN)1N12τ{\varepsilon/r}\approx(\ln N)^{-1}\gtrsim N^{-\frac{1}{2}\tau} (note that τ\tau can be arbitrarily small and thus N12τN^{-\frac{1}{2}\tau} is close to the resonant situation), and we do not have a tool to handle this situation (except for d=2d=2). 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 (d1)(d-1)-dimensional Hausdorff measure of nodal sets centers around Yau’s conjecture for Laplace eigenfunctions on smooth manifolds:

Δgϕλ=λϕλ,on ,-\Delta_{g}\phi_{\lambda}=\lambda\phi_{\lambda},\qquad\text{on }\mathcal{M}, (1.12)

where \mathcal{M} 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

cλHd1({x|ϕλ(x)=0})Cλc\sqrt{\lambda}\leq H^{d-1}(\{x\in\mathcal{M}|\phi_{\lambda}(x)=0\})\leq C\sqrt{\lambda} (1.13)

where C,cC,c depend only on the manifold \mathcal{M} and Hd1H^{d-1} denotes the (d1)(d-1)-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 ε(uε)=0\mathcal{L}_{\varepsilon}(u_{\varepsilon})=0 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.

Assume that A=A(y)A=A(y) satisfies the conditions (1.2), (1.3), and (1.4). Let uεu_{\varepsilon} be a nonzero weak solution of ε(uε)=0\mathcal{L}_{\varepsilon}(u_{\varepsilon})=0 satisfying (1.5).

  • (i)

    If d3d\geq 3, then for any α>8\alpha>8, it holds that

    Hd1({xBΛ16|uε(x)=0})exp(CNα),\displaystyle H^{d-1}(\{x\in B_{\frac{\Lambda}{16}}|u_{\varepsilon}(x)=0\})\leq\exp(CN^{\alpha}), (1.14)

    where CC depends only on dd, Λ,γ\Lambda,\gamma and α\alpha.

  • (ii)

    If d=2d=2, then it holds that

    H1({xBΛ16|uε(x)=0})exp(C(lnN)2),\displaystyle H^{1}(\{x\in B_{\frac{\Lambda}{16}}|u_{\varepsilon}(x)=0\})\leq\exp(C(\ln N)^{2}), (1.15)

    where CC depends only on Λ\Lambda and γ\gamma.

The strategy of the proof is as follows. For relatively large ε\varepsilon, we adapt a blow-up argument to obtain the upper bounds of nodal sets. For small ε\varepsilon, the solution uεu_{\varepsilon} can be approximated by a harmonic function u0u_{0}, and thus the nodal set of uεu_{\varepsilon} is a small perturbation of the nodal set of u0u_{0}. We then derive a quantitative estimate for the nodal set of uεu_{\varepsilon} by carefully studying the small perturbations near the nodal set and critical set of u0u_{0}, 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 uεu_{\varepsilon}. The restriction α>8\alpha>8 for d3d\geq 3 arises from the doubling inequality (5.2) for β(34,1)\beta\in(\frac{3}{4},1). If we consider NN to be exp(CM)\exp(CM) for some large constant MM, which is the case for the doubling inequality of eigenfunctions, the upper bounds of nodal sets are double exponential functions exp(exp(CM))\exp(\exp(CM)). In this sense, the restriction α>8\alpha>8 only affects the constant CC in such upper bounds, which does not play an important role. For d=2d=2, we point out that there is no misprint in the exponential (compared to d3d\geq 3). 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 cc, CC, C^\hat{C}, C~\tilde{C}, CiC_{i}, cic_{i} denote positive constants that do not depend on ε\varepsilon or uεu_{\varepsilon}, 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 ε/rN1β34\varepsilon/r\gtrsim N^{-\frac{1}{\beta-\frac{3}{4}}} for all dimensions. Indeed, we will prove a quantitative version of [15, Theorem 3.1].

Let 0=(A^)\mathcal{L}_{0}=-\nabla\cdot(\widehat{A}\nabla) be the homogenized operator and A^\widehat{A} be the homogenized coefficient matrix of AA (see, e.g., [24] for the general theory of periodic elliptic homogenization). Define the ellipsoid

Er={xd:(A^)1x,x<r2}.E_{r}=\big{\{}x\in\mathbb{R}^{d}:\langle(\widehat{A})^{-1}x,x\rangle<r^{2}\big{\}}.

The following is the main theorem of this section.

Theorem 2.1.

Let θ(0,1/2]\theta\in(0,1/2] and AA satisfy conditions (1.2), (1.3) and (1.4). There exists C>0C>0 depending only on dd and Λ\Lambda such that if ε(uε)=0\mathcal{L}_{\varepsilon}(u_{\varepsilon})=0 in E1E_{1} and

E1uε2NEθuε2,\int_{E_{1}}u_{\varepsilon}^{2}\leq N\int_{E_{\theta}}u_{\varepsilon}^{2},

then for any CN1β34ε<r<1εCN^{\frac{1}{\beta-\frac{3}{4}}}\varepsilon<r<1-\sqrt{\varepsilon}, we have

Eruε22NEθruε2.\int_{E_{r}}u_{\varepsilon}^{2}\leq 2N\int_{E_{\theta r}}u_{\varepsilon}^{2}.

This follows from Lemma 2.2 and Lemma 2.3.

Lemma 2.2.

Let θ(0,1/2]\theta\in(0,1/2]. Suppose uεu_{\varepsilon} is a solution of ε(uε)=0\mathcal{L}_{\varepsilon}(u_{\varepsilon})=0 in E1E_{1} satisfying

E1uε2NEθuε2.\int_{E_{1}}u_{\varepsilon}^{2}\leq N\int_{E_{\theta}}u_{\varepsilon}^{2}.

For any β(3/4,1)\beta\in(3/4,1), there exist c,C>0c,C>0, depending only on d,Λd,\Lambda and β\beta, such that if ε<cN1β34\varepsilon<cN^{-\frac{1}{\beta-\frac{3}{4}}}, then for any r[θ,1ε]r\in[\theta,1-\sqrt{\varepsilon}]

Eruε2N(1+CNεβ34)Eθruε2.\int_{E_{r}}u_{\varepsilon}^{2}\leq N(1+CN\varepsilon^{\beta-\frac{3}{4}})\int_{E_{\theta r}}u_{\varepsilon}^{2}. (2.1)
Proof.

Let t>0t>0, to be determined. Since dist(E1t,E1)Ct\text{dist}(\partial E_{1-t},\partial E_{1})\leq Ct, by the Caccioppoli inequality, we have

E1t|uε|2+E1tuε2Ct2E1uε2CNt2Eθuε2.\int_{E_{1-t}}|\nabla u_{\varepsilon}|^{2}+\int_{E_{1-t}}u_{\varepsilon}^{2}\leq\frac{C}{t^{2}}\int_{E_{1}}u_{\varepsilon}^{2}\leq\frac{CN}{t^{2}}\int_{E_{\theta}}u_{\varepsilon}^{2}.

By the co-area formula, we can find some c0(1,2)c_{0}\in(1,2) so that

E1c0t|uε|2+E1c0tuε2CNt3Eθuε2.\int_{\partial E_{1-c_{0}t}}|\nabla u_{\varepsilon}|^{2}+\int_{\partial E_{1-c_{0}t}}u_{\varepsilon}^{2}\leq\frac{CN}{t^{3}}\int_{E_{\theta}}u_{\varepsilon}^{2}.

Without loss of generality, let us simply assume c0=1c_{0}=1. Hence uε|E1tH1(E1t)u_{\varepsilon}|_{\partial E_{1-t}}\in H^{1}(\partial E_{1-t}). By [10, Theorem 1.1],

E1t(uεu0)2Cε2βuεH1(E1t)2CNε2βt3Eθuε2,\int_{E_{1-t}}(u_{\varepsilon}-u_{0})^{2}\leq C\varepsilon^{2\beta}\lVert u_{\varepsilon}\rVert_{H^{1}(\partial E_{1-t})}^{2}\leq\frac{CN\varepsilon^{2\beta}}{t^{3}}\int_{E_{\theta}}u_{\varepsilon}^{2}, (2.2)

where u0u_{0} is the solution of 0(u0)=0\mathcal{L}_{0}(u_{0})=0 and u0=uεu_{0}=u_{\varepsilon} on E1t\partial E_{1-t} and β(0,1)\beta\in(0,1) is arbitrary.

As a result, we have

u0L2(E1t)\displaystyle\lVert u_{0}\rVert_{L^{2}(E_{1-t})} uεL2(E1)+uεu0L2(E1t)\displaystyle\leq\lVert u_{\varepsilon}\rVert_{L^{2}(E_{1})}+\lVert u_{\varepsilon}-u_{0}\rVert_{L^{2}(E_{1-t})} (2.3)
N(1+Cεβt3/2)uεL2(Eθ).\displaystyle\leq\sqrt{N}(1+C\varepsilon^{\beta}t^{-3/2})\lVert u_{\varepsilon}\rVert_{L^{2}(E_{\theta})}.

Also,

uεL2(Eθ)u0L2(Eθ)+CNεβt3/2uεL2(Eθ).\lVert u_{\varepsilon}\rVert_{L^{2}(E_{\theta})}\leq\lVert u_{0}\rVert_{L^{2}(E_{\theta})}+C\sqrt{N}\varepsilon^{\beta}t^{-3/2}\lVert u_{\varepsilon}\rVert_{L^{2}(E_{\theta})}.

We will choose t<1t<1 so that CNεβt3/2<1/2C\sqrt{N}\varepsilon^{\beta}t^{-3/2}<1/2. Consequently,

uεL2(Eθ)(1CNεβt3/2)1u0L2(Eθ)(1+CNεβt3/2)u0L2(Eθ).\lVert u_{\varepsilon}\rVert_{L^{2}(E_{\theta})}\leq\big{(}1-C\sqrt{N}\varepsilon^{\beta}t^{-3/2}\big{)}^{-1}\lVert u_{0}\rVert_{L^{2}(E_{\theta})}\leq\big{(}1+C\sqrt{N}\varepsilon^{\beta}t^{-3/2}\big{)}\lVert u_{0}\rVert_{L^{2}(E_{\theta})}. (2.4)

Inserting this into (2.3), we have

u0L2(E1t)\displaystyle\lVert u_{0}\rVert_{L^{2}(E_{1-t})} N(1+CNεβt3/2)(1+Cεβt3/2)u0L2(Eθ)\displaystyle\leq\sqrt{N}\big{(}1+C\sqrt{N}\varepsilon^{\beta}t^{-3/2}\big{)}(1+C\varepsilon^{\beta}t^{-3/2})\lVert u_{0}\rVert_{L^{2}(E_{\theta})} (2.5)
N(1+CNεβt3/2)u0L2(Eθ),\displaystyle\leq\sqrt{N}\big{(}1+C\sqrt{N}\varepsilon^{\beta}t^{-3/2}\big{)}\lVert u_{0}\rVert_{L^{2}(E_{\theta})},

where we have used the simple fact that (1+a)21+3a(1+a)^{2}\leq 1+3a for a[0,1]a\in[0,1] and enlarged the constant CC in the last inequality.

Next, by the interior LL^{\infty} estimate for A^\widehat{A}-harmonic functions, we have

u0L2(Eθ)\displaystyle\lVert u_{0}\rVert_{L^{2}(E_{\theta})} u0L2(Eθ(1t))+u0L2(EθEθ(1t))\displaystyle\leq\lVert u_{0}\rVert_{L^{2}(E_{\theta(1-t)})}+\lVert u_{0}\rVert_{L^{2}(E_{\theta}\setminus E_{\theta(1-t)})}
u0L2(Eθ(1t))+Cθtu0L2(E1t)\displaystyle\leq\lVert u_{0}\rVert_{L^{2}(E_{\theta(1-t)})}+C\sqrt{\theta t}\lVert u_{0}\rVert_{L^{2}(E_{1-t})}

Inserting this into (2.5) and choosing tt sufficiently small so that CNθt<1/2C\sqrt{N}\sqrt{\theta t}<1/2, we obtain

u0L2(E1t)N(1+NCεβt3/2)(1+CNθt)u0L2(Eθ(1t)).\lVert u_{0}\rVert_{L^{2}(E_{1-t})}\leq\sqrt{N}\big{(}1+\sqrt{N}C\varepsilon^{\beta}t^{-3/2}\big{)}\big{(}1+C\sqrt{N}\sqrt{\theta t}\big{)}\lVert u_{0}\rVert_{L^{2}(E_{\theta(1-t)})}.

Choose t=εt=\sqrt{\varepsilon}. We arrive at

u0L2(E1t)N(1+CNεβ3/4)u0L2(Eθ(1t)).\lVert u_{0}\rVert_{L^{2}(E_{1-t})}\leq\sqrt{N}(1+C\sqrt{N}\varepsilon^{\beta-3/4})\lVert u_{0}\rVert_{L^{2}(E_{\theta(1-t)})}. (2.6)

Note that the above calculation goes through only if NCεβt3/2<1/2\sqrt{N}C\varepsilon^{\beta}t^{-3/2}<1/2 and CNt<1/2C\sqrt{N}\sqrt{t}<1/2. This implies that we require

εC1N12(β3/4),\varepsilon\leq C^{-1}N^{\frac{-1}{2(\beta-3/4)}},

for some large constant CC.

Recall that u0u_{0} is a weak solution of 0(u0)=0\mathcal{L}_{0}(u_{0})=0 in E(1ε)E(1-\sqrt{\varepsilon}). Let w0(x)=u0(A^12x)w_{0}(x)=u_{0}(\widehat{A}^{\frac{1}{2}}x). Then Δw0=0\Delta w_{0}=0 in B1εB_{1-\sqrt{\varepsilon}} and (2.6) is equivalent to

w0L2(B1ε)N(1+CNεβ3/4)w0L2(Bθ(1ε)).\lVert w_{0}\rVert_{L^{2}(B_{1-\sqrt{\varepsilon}})}\leq\sqrt{N}(1+C\sqrt{N}\varepsilon^{\beta-3/4})\lVert w_{0}\rVert_{L^{2}(B_{\theta(1-\sqrt{\varepsilon})})}. (2.7)

Now, as a consequence of the well-known three-sphere theorem for harmonic functions,

φ(r)=log2B2rw02\varphi(r)=\log_{2}\fint_{B_{2^{r}}}w_{0}^{2} (2.8)

is a convex function in (,0](-\infty,0] and therefore φ(t)φ(tc)\varphi(t)-\varphi(t-c) is a nondecreasing function in tt, for any fixed c>0c>0. Hence, we obtain from (2.7) that for any r(0,1ε)r\in(0,1-\sqrt{\varepsilon}),

w0L2(Br)N(1+CNεβ3/4)w0L2(Bθr).\lVert w_{0}\rVert_{L^{2}(B_{r})}\leq\sqrt{N}(1+C\sqrt{N}\varepsilon^{\beta-3/4})\lVert w_{0}\rVert_{L^{2}(B_{\theta r})}.

(The doubling index with θ\theta is an increasing function of radius.) Again, this is equivalent to

u0L2(Er)N(1+CNεβ3/4)u0L2(Eθr)\lVert u_{0}\rVert_{L^{2}(E_{r})}\leq\sqrt{N}(1+C\sqrt{N}\varepsilon^{\beta-3/4})\lVert u_{0}\rVert_{L^{2}(E_{\theta r})} (2.9)

for any r(0,1ε)r\in(0,1-\sqrt{\varepsilon}).

Now, let r[θ,1ε)r\in[\theta,1-\sqrt{\varepsilon}). It follows by (2.9) that

uεL2(Er)\displaystyle\lVert u_{\varepsilon}\rVert_{L^{2}(E_{r})} uεu0L2(Er)+u0L2(Er)\displaystyle\leq\lVert u_{\varepsilon}-u_{0}\rVert_{L^{2}(E_{r})}+\lVert u_{0}\rVert_{L^{2}(E_{r})}
CNεβ3/4uεL2(Eθ)+N(1+CNεβ3/4)u0L2(Eθr)\displaystyle\leq C\sqrt{N}\varepsilon^{\beta-3/4}\lVert u_{\varepsilon}\rVert_{L^{2}(E_{\theta})}+\sqrt{N}(1+C\sqrt{N}\varepsilon^{\beta-3/4})\lVert u_{0}\rVert_{L^{2}(E_{\theta r})}
CNεβ3/4uεL2(Er)+N(1+CNεβ3/4)u0uεL2(Eθr)\displaystyle\leq C\sqrt{N}\varepsilon^{\beta-3/4}\lVert u_{\varepsilon}\rVert_{L^{2}(E_{r})}+\sqrt{N}(1+C\sqrt{N}\varepsilon^{\beta-3/4})\lVert u_{0}-u_{\varepsilon}\rVert_{L^{2}(E_{\theta r})}
+N(1+CNεβ3/4)uεL2(Eθr)\displaystyle\qquad+\sqrt{N}(1+C\sqrt{N}\varepsilon^{\beta-3/4})\lVert u_{\varepsilon}\rVert_{L^{2}(E_{\theta r})}
CNεβ3/4(1+CNεβ3/4)uεL2(Er)+N(1+CNεβ3/4)uεL2(Eθr),\displaystyle\leq CN\varepsilon^{\beta-3/4}(1+C\sqrt{N}\varepsilon^{\beta-3/4})\lVert u_{\varepsilon}\rVert_{L^{2}(E_{r})}+\sqrt{N}(1+C\sqrt{N}\varepsilon^{\beta-3/4})\lVert u_{\varepsilon}\rVert_{L^{2}(E_{\theta r})},

where we have used the fact EθErE_{\theta}\subset E_{r} in the third inequality and (2.2) in the second and last inequalities. Assume further that εcN1/(β3/4)\varepsilon\leq cN^{-1/(\beta-3/4)}. Then

uεL2(Er)\displaystyle\lVert u_{\varepsilon}\rVert_{L^{2}(E_{r})} N(1+CNεβ3/4)1CNεβ3/4uεL2(Eθr)\displaystyle\leq\frac{\sqrt{N}(1+C\sqrt{N}\varepsilon^{\beta-3/4})}{1-CN\varepsilon^{\beta-3/4}}\lVert u_{\varepsilon}\rVert_{L^{2}(E_{\theta r})}
N(1+CNεβ3/4)uεL2(Eθr).\displaystyle\leq\sqrt{N}(1+CN\varepsilon^{\beta-3/4})\lVert u_{\varepsilon}\rVert_{L^{2}(E_{\theta r})}.

This proves the lemma. ∎

Now, if ε<cN1/(β3/4)\varepsilon<cN^{-1/(\beta-3/4)}, the above lemma allows us to iterate (2.1) down to r=c1N1/(β3/4)εr=c^{-1}N^{1/(\beta-3/4)}\varepsilon. Precisely, if r=θk>CN1/(β3/4)εr=\theta^{k}>CN^{1/(\beta-3/4)}\varepsilon and

Eruε2AkEθruε2,\int_{E_{r}}u_{\varepsilon}^{2}\leq A_{k}\int_{E_{\theta r}}u_{\varepsilon}^{2},

with A0=NA_{0}=N, then

Eθruε2Ak+1Eθ2ruε2,\int_{E_{\theta r}}u_{\varepsilon}^{2}\leq A_{k+1}\int_{E_{\theta^{2}r}}u_{\varepsilon}^{2},

where

Ak+1=Ak(1+CAk(θkε)β3/4),A_{k+1}=A_{k}(1+CA_{k}(\theta^{-k}\varepsilon)^{\beta-3/4}),

provided Ak(θkε)β3/4<cA_{k}(\theta^{-k}\varepsilon)^{\beta-3/4}<c.

Lemma 2.3.

For all kk0k\leq k_{0} with θk0εc1N1/(β3/4)\theta^{-k_{0}}\varepsilon\simeq c_{1}N^{-1/(\beta-3/4)} and c1>0c_{1}>0 sufficiently small, one has

Ak2N.A_{k}\leq 2N.
Proof.

Define Bk=Ak/NB_{k}=A_{k}/N. Then B0=1B_{0}=1 and

Bk+1=Bk(1+CBkθk(β3/4)εβ3/4N)Bk(1+c1CBkθ(k0k)(β3/4)).B_{k+1}=B_{k}(1+CB_{k}\theta^{-k(\beta-3/4)}\varepsilon^{\beta-3/4}N)\leq B_{k}(1+c_{1}CB_{k}\theta^{(k_{0}-k)(\beta-3/4)}).

It follows that

Bk+1Bkc1CBk2δk0k,B_{k+1}-B_{k}\leq c_{1}CB_{k}^{2}\delta^{k_{0}-k},

where δ=θβ3/4(1/2)β3/4<1\delta=\theta^{\beta-3/4}\leq(1/2)^{\beta-3/4}<1. The above inequality yields

Bk+1B0+j=0kc1CBj2δk0j.B_{k+1}\leq B_{0}+\sum_{j=0}^{k}c_{1}CB_{j}^{2}\delta^{k_{0}-j}. (2.10)

We prove by induction that if c1c_{1} is sufficiently small, then Bk2B_{k}\leq 2 and Ak(θkε)β3/42c1A_{k}(\theta^{-k}\varepsilon)^{\beta-3/4}\leq 2c_{1} for all kk0k\leq k_{0}. Actually, if

c1(4Cj=0δj)1=1δ4C,c_{1}\leq(4C\sum_{j=0}^{\infty}\delta^{j})^{-1}=\frac{1-\delta}{4C},

and Bj2B_{j}\leq 2 for all 1jk1\leq j\leq k, then it is easy to see from (2.10) that Bk+12B_{k+1}\leq 2 and Ak(θkε)β3/42N(θk0ε)β3/42c1A_{k}(\theta^{-k}\varepsilon)^{\beta-3/4}\leq 2N(\theta^{-k_{0}}\varepsilon)^{\beta-3/4}\leq 2c_{1}. This proves the desired estimate. ∎

Remark 2.4.

Observe that in the above proof, the smoothness of the coefficients has not been used explicitly, except for (2.2) by [10, Theorem 1.1]. But this actually can be replaced by, e.g., [23, Theorem 1.4] with m=1m=1, which does not require any smoothness.

Remark 2.5.

It is not difficult to see that (2.1) implies the following three-ball inequality with an error term

uεL2(Eθr)uεL2(Er)12uεL2(Eθ2r)12+Cε12(β34)uεL2(Er),\lVert u_{\varepsilon}\rVert_{L^{2}(E_{\theta r})}\leq\lVert u_{\varepsilon}\rVert_{L^{2}(E_{r})}^{\frac{1}{2}}\lVert u_{\varepsilon}\rVert_{L^{2}(E_{\theta^{2}r})}^{\frac{1}{2}}+C\varepsilon^{\frac{1}{2}(\beta-\frac{3}{4})}\lVert u_{\varepsilon}\rVert_{L^{2}(E_{r})}, (2.11)

for any θ(0,12]\theta\in(0,\frac{1}{2}] and β(34,1)\beta\in(\frac{3}{4},1). 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 ε0\varepsilon\to 0, (2.11) recovers precisely the three-ball inequality for A^\widehat{A}-harmonic functions.

Theorem 2.6.

Given arbitrary θ(0,Λ/2]\theta\in(0,\Lambda/2], there exists C>0C>0 depending only on dd and Λ\Lambda such that if ε(uε)=0\mathcal{L}_{\varepsilon}(u_{\varepsilon})=0 in B1B_{1} and

B1uε2NBθuε2,\int_{B_{1}}u_{\varepsilon}^{2}\leq N\int_{B_{\theta}}u_{\varepsilon}^{2}, (2.12)

then for any CN1β34ε<r<1CN^{\frac{1}{\beta-\frac{3}{4}}}\varepsilon<r<1, we have

Bruε28N3Bθruε2.\int_{B_{r}}u_{\varepsilon}^{2}\leq 8N^{3}\int_{B_{\theta r}}u_{\varepsilon}^{2}.
Proof.

This is deduced from Theorem 2.1 and the fact

BΛrErBr.B_{\sqrt{\Lambda}r}\subset E_{r}\subset B_{r}. (2.13)

Indeed, (2.12) and (2.13) imply

E1uε2NBθuε2NEθ/Λuε2.\int_{E_{1}}u_{\varepsilon}^{2}\leq N\int_{{B_{\theta}}}u_{\varepsilon}^{2}\leq N\int_{{E_{\theta/\sqrt{\Lambda}}}}u_{\varepsilon}^{2}.

Note that θ(0,Λ/2]\theta\in(0,\Lambda/2] implies θ/Λ(0,Λ/2](0,1/2]\theta/\sqrt{\Lambda}\in(0,\sqrt{\Lambda}/2]\subset(0,1/2]. Now, if r(CN1β34ε,1ε)r\in(CN^{\frac{1}{\beta-\frac{3}{4}}}\varepsilon,1-\sqrt{\varepsilon}), we may apply Theorem 2.1 (three times) with θ=θ/Λ\theta^{\prime}=\theta/\sqrt{\Lambda} and obtain

Bruε2Er/Λuε2(2N)3Eθ3Λ2ruε28N3Eθruε28N3Bθruε2.\int_{B_{r}}u_{\varepsilon}^{2}\leq\int_{E_{r/\sqrt{\Lambda}}}u_{\varepsilon}^{2}\leq(2N)^{3}\int_{E_{\theta^{3}\Lambda^{-2}r}}u_{\varepsilon}^{2}\leq 8N^{3}\int_{E_{\theta r}}u_{\varepsilon}^{2}\leq 8N^{3}\int_{B_{\theta r}}u_{\varepsilon}^{2}.

For r[1ε,1]r\in[1-\sqrt{\varepsilon},1] (without loss of generality, assume ε<1/4\varepsilon<1/4), we may apply Theorem 2.1 once to obtain

Bruε2B1uε2NBθuε22N2Bθ2Λ1uε22N2Bθruε2.\int_{B_{r}}u_{\varepsilon}^{2}\leq\int_{B_{1}}u_{\varepsilon}^{2}\leq N\int_{B_{\theta}}u_{\varepsilon}^{2}\leq 2N^{2}\int_{B_{\theta^{2}\Lambda^{-1}}}u_{\varepsilon}^{2}\leq 2N^{2}\int_{B_{\theta r}}u_{\varepsilon}^{2}.

This ends the proof. ∎

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

1(u)=(A(x)u)=0,\displaystyle\mathcal{L}_{1}(u)=-\nabla\cdot\big{(}A(x)\nabla u\big{)}=0, (3.1)

where A(x)A(x) satisfies (1.2) and

|A(x)A(y)|L|xy|\displaystyle|A({x})-A({y})|\leq L|x-y| (3.2)

for some large positive constant L>1L>1. We emphasize that throughout this section, the constant CC will never depend on LL. Since the LL^{\infty} norm and the L2L^{2} norm of uu are comparable, parallel to the assumption (1.7), we may assume the following

uL(B1)MuL(Bθ)\displaystyle\|u\|_{L^{\infty}(B_{1})}\leq M\|u\|_{L^{\infty}(B_{\theta})} (3.3)

for some large constant M>1M>1.

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 LL. For d3d\geq 3, we define the Lipschitz metric g^=g^ij(x)dxidxj\hat{g}=\hat{g}_{ij}(x)dx_{i}\otimes dx_{j} as follows

g^ij(x)=aij(x)det(A(x))1d2,\displaystyle\hat{g}_{ij}({x})=a^{ij}({x})\det(A({x}))^{\frac{1}{d-2}}, (3.4)

where aij(x)a^{ij}({x}) is the entry of A1(x)A^{-1}({x}). The case d=2d=2 will be discussed in Remark 3.3. Note that g^\hat{g} is Lipsthitz continuous and satisfies

|g^(x)g^(y)|CL|xy|.\displaystyle|\hat{g}({x})-\hat{g}({y})|\leq CL|x-y|. (3.5)

Define

r2=r2(x)=g^ij(0)xixj\displaystyle r^{2}=r^{2}(x)=\hat{g}_{ij}(0)x_{i}x_{j} (3.6)

and

ψ(x)=g^kl(x)rxkrxl.\displaystyle\psi(x)=\hat{g}^{kl}({x})\frac{\partial r}{\partial x_{k}}\frac{\partial r}{\partial x_{l}}.

From (3.6), we can also write

ψ(x)=1r2g^kl(x)g^ik(0)g^jl(0)xixj.\displaystyle\psi(x)=\frac{1}{r^{2}}\hat{g}^{kl}({x})\hat{g}_{ik}(0)\hat{g}_{jl}(0)x_{i}x_{j}.

Thus, we can check that ψ(x)\psi(x) is a non-negative Lipschitz function satisfying

|ψ(x)ψ(y)|CL|xy|,\displaystyle|\psi(x)-\psi(y)|\leq CL|x-y|, (3.7)

where CC depends only on dd and Λ\Lambda. We introduce a new metric g=gij(x)dxidxj{g}={g}_{ij}(x)dx_{i}\otimes dx_{j} by setting

gij(x)=ψ(x)g^ij(x).\displaystyle{g}_{ij}(x)=\psi(x)\hat{g}_{ij}({x}). (3.8)

We can write the metric gg in terms of the intrinsic geodesic polar coordinates (r,σ1,,σd1)(r,\sigma_{1},\cdots,\sigma_{d-1}),

g=drdr+r2bij(r,σ)dσidσj,\displaystyle g=dr\otimes dr+r^{2}b_{ij}(r,\sigma)d\sigma_{i}\otimes d\sigma_{j}, (3.9)

where bijb_{ij} satisfies

|rbij(r,σ)|CL,fori,j=1,,d1,\displaystyle|\partial_{r}b_{ij}(r,\sigma)|\leq CL,\quad\mbox{for}\ i,j=1,\cdots,d-1, (3.10)

and CC depends only on dd and Λ\Lambda.

The existence of the geodesic polar coordinates (r,σ)(r,\ \sigma) allows us to consider geodesic balls. Denote by 𝔹r\mathbb{B}_{r} the geodesic ball in the metric gg of radius rr and centred at the origin. In particular, from (3.6) and (3.9), r(x)=g^ij(0)xixjr(x)=\sqrt{\hat{g}_{ij}(0)x_{i}x_{j}} is the geodesic distance from xx to the origin in the new metric gg. 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., g^ij(0)=δij\hat{g}_{ij}(0)=\delta_{ij}.

Let

η(x)=ψd22.\displaystyle\eta(x)=\psi^{-\frac{d-2}{2}}. (3.11)

Obviously, η(x)\eta(x) is a Lipschitz function satisfying

C1η(x)C2,\displaystyle C_{1}\leq\eta(x)\leq C_{2}, (3.12)

where C1C_{1} and C2C_{2} depend on dd and Λ\Lambda. In the polar coordinates,

|rη(r,σ)|CL.\displaystyle|\partial_{r}\eta(r,\sigma)|\leq{C}L. (3.13)

In this new metric gg, the equation (3.1) can be written as

g(η(x)gu(x))=0inB1.\displaystyle-\nabla_{g}\cdot(\eta(x)\nabla_{g}u(x))=0\quad\mbox{in}\ B_{1}. (3.14)

Let

D(r)=Brη|gu|2𝑑Vg\displaystyle D(r)=\int_{B_{r}}\eta|\nabla_{g}u|^{2}dV_{g} (3.15)

and

H(r)=Brηu2𝑑Sg,\displaystyle H(r)=\int_{\partial B_{r}}\eta u^{2}dS_{g}, (3.16)

where dSgdS_{g} represents the area element of Br\partial B_{r} under the metric gg. We define the frequency function by

𝒩(r)=rD(r)H(r).\displaystyle\mathcal{N}(r)=\frac{rD(r)}{H(r)}. (3.17)

For future application, we will also use the notation 𝒩(p,r)\mathcal{N}(p,r) to specify the center of the ball Br(p)B_{r}(p) in the definition of frequency function.

Lemma 3.1.

Let uH1(B1)u\in H^{1}(B_{1}) be a nontrivial solution of (3.1). There exists a positive constant CC depending on dd and Λ\Lambda such that

𝒩¯(r)=exp(CLr)𝒩(r)\displaystyle\overline{\mathcal{N}}(r)=\exp({CLr})\mathcal{N}(r) (3.18)

is a non-decreasing function of r(0,1)r\in(0,1).

Proof.

The proof of the lemma is essentially contained in [6]. Since we want to show the explicit dependence of the Lipschtiz constant LL in the estimates, we sketch the proof by considering the role of LL. Taking derivative with respect to rr for 𝒩\mathcal{N}, we have

𝒩(r)𝒩(r)=(1r+D(r)D(r)H(r)H(r)).\displaystyle\frac{\mathcal{N}^{\prime}(r)}{\mathcal{N}(r)}=\bigg{(}\frac{1}{r}+\frac{D^{\prime}(r)}{D(r)}-\frac{H^{\prime}(r)}{H(r)}\bigg{)}. (3.19)

In order to prove the lemma, it suffices to show

1r+D(r)D(r)H(r)H(r)CL.\displaystyle\frac{1}{r}+\frac{D^{\prime}(r)}{D(r)}-\frac{H^{\prime}(r)}{H(r)}\geq-CL. (3.20)

Thus, we consider the derivatives of H(r)H(r) and D(r)D(r), respectively. Setting b(r,σ)=|det(bij(r,σ))|b(r,\sigma)=|\det(b_{ij}(r,\sigma))|. Note that dSg=rd1b(r,σ)dσdS_{g}=r^{d-1}\sqrt{b(r,\sigma)}d\sigma. We write H(r)H(r) as

H(r)=rd1B1η(r,σ)u2(r,σ)b(r,σ)𝑑σ.\displaystyle H(r)=r^{d-1}\int_{\partial B_{1}}\eta(r,\sigma)u^{2}(r,\sigma)\sqrt{b(r,\sigma)}d\sigma. (3.21)

Taking derivative with respect to rr, one has

H(r)=d1rH(r)+Br1br(ηb)u2dSg+2BrηurudSg,\displaystyle H^{\prime}(r)=\frac{d-1}{r}H(r)+\int_{\partial B_{r}}\frac{1}{\sqrt{b}}\partial_{r}(\eta\sqrt{b})u^{2}dS_{g}+2\int_{\partial B_{r}}\eta u\partial_{r}udS_{g}, (3.22)

where ru=gu,xr\partial_{r}u=\langle\nabla_{g}u,\frac{x}{r}\rangle on Br\partial B_{r}. By (3.10), (3.12) and (3.13), we have

H(r)=(d1r+O(L))H(r)+2BrηurudSg.\displaystyle H^{\prime}(r)=\Big{(}\frac{d-1}{r}+{O(L)}\Big{)}H(r)+2\int_{\partial B_{r}}\eta u\partial_{r}udS_{g}. (3.23)

Multiplying both sides of (3.14) by uu and performing the integration by parts give that

D(r)=Brη|gu|2𝑑Vg=BrηurudSg.\displaystyle D(r)=\int_{B_{r}}\eta|\nabla_{g}u|^{2}dV_{g}=\int_{\partial B_{r}}\eta u\partial_{r}udS_{g}. (3.24)

It follows that

H(r)=(d1r+O(L))H(r)+2D(r).\displaystyle H^{\prime}(r)=\Big{(}\frac{d-1}{r}+{O(L)}\Big{)}H(r)+2D(r). (3.25)

Similarly, we may compute the derivative of D(r)D(r) as in [6] and obtain

D(r)=(d2r+O(L))D(r)+2Brη(ru)2𝑑Sg.\displaystyle D^{\prime}(r)=\Big{(}\frac{d-2}{r}+{O(L)}\Big{)}D(r)+2\int_{\partial B_{r}}\eta(\partial_{r}u)^{2}dS_{g}. (3.26)

Combining the estimates (3.25) and (3.26), and using the Cauchy-Schwarz inequality, we obtain

1r+D(r)D(r)H(r)H(r)\displaystyle\frac{1}{r}+\frac{D^{\prime}(r)}{D(r)}-\frac{H^{\prime}(r)}{H(r)} =O(L)+2Brη(ru)2𝑑SgBrηurudSg2BrηurudSgBrηu2𝑑Sg\displaystyle={O(L)}+2\frac{\int_{\partial B_{r}}\eta(\partial_{r}u)^{2}dS_{g}}{\int_{\partial B_{r}}\eta u\partial_{r}udS_{g}}-2\frac{\int_{\partial B_{r}}\eta u\partial_{r}udS_{g}}{\int_{\partial B_{r}}\eta u^{2}dS_{g}}
O(L).\displaystyle\geq{O(L)}.

This proves (3.20) and thus the lemma. ∎

Next we derive the doubling inequality with an explicit dependence on LL.

Lemma 3.2.

Let uu be a solution of (3.1) satisfying (3.2) and (3.3). For a fixed constant 0<θ120<\theta\leq\frac{1}{2}, we have

uL2(Br)MC1eC2LuL2(Bθr)\displaystyle\|u\|_{L^{2}(B_{r})}\leq M^{C_{1}e^{{C_{2}L}}}\|u\|_{L^{2}(B_{\theta r})} (3.27)

for 0<r120<r\leq\frac{1}{2}, where C1C_{1} depends on θ\theta, and C2C_{2} depends on dd, Λ\Lambda.

Proof.

From (3.25) and the definition of 𝒩¯(r)\overline{\mathcal{N}}(r), we have

(lnH(r)rd1)=O(L)+2r𝒩¯(r)exp(CLr).\displaystyle\bigg{(}\ln\frac{H(r)}{r^{d-1}}\bigg{)}^{\prime}={O(L)}+\frac{2}{r}\overline{\mathcal{N}}(r)\exp(-{CL}r). (3.28)

Note that here O(L)O(L) is a function in rr satisfying CLO(L)CL-CL\leq O(L)\leq CL. We would like to obtain an upper bound and a lower bound for the quotient H(r2)/H(r1)H(r_{2})/H(r_{1}) with 0<r1<r20<r_{1}<r_{2}. To find the upper bound, we integrate the equality (3.28) from r1r_{1} to r2r_{2} and use the monotonicity of 𝒩¯(r)\overline{\mathcal{N}}(r) to obtain

lnH(r2)r2d1lnH(r1)r1d1CL(r2r1)+2𝒩¯(r2)ln(r2r1)exp(CLr1).\displaystyle\ln\frac{H(r_{2})}{r^{d-1}_{2}}-\ln\frac{H(r_{1})}{r^{d-1}_{1}}\leq{CL(r_{2}-r_{1})}+2\overline{\mathcal{N}}(r_{2})\ln\Big{(}\frac{r_{2}}{r_{1}}\Big{)}\exp(-CLr_{1}). (3.29)

Taking the exponential of both sides gives the upper bound

H(r2)H(r1)eCL(r2r1)(r2r1)2𝒩¯(r2)exp(CLr1)+d1.\displaystyle\frac{H(r_{2})}{H(r_{1})}\leq e^{{CL(r_{2}-r_{1})}}\Big{(}\frac{r_{2}}{r_{1}}\Big{)}^{2\overline{\mathcal{N}}(r_{2})\exp(-CLr_{1})+d-1}. (3.30)

To see the lower bound, we integrate (3.28) from r1r_{1} to r2r_{2} and apply the monotonicity of 𝒩¯(r)\overline{\mathcal{N}}(r) again to obtain

lnH(r2)r2d1lnH(r1)r1d1CL(r2r1)+2𝒩¯(r1)exp(CLr2)ln(r2r1).\displaystyle\ln\frac{H(r_{2})}{r^{d-1}_{2}}-\ln\frac{H(r_{1})}{r^{d-1}_{1}}\geq-{CL}(r_{2}-r_{1})+2\overline{\mathcal{N}}(r_{1})\exp(-{CLr_{2}})\ln\Big{(}\frac{r_{2}}{r_{1}}\Big{)}. (3.31)

Raising to the exponential form, we have

H(r2)H(r1)eCL(r2r1)(r2r1)2𝒩¯(r1)exp(CLr2)+d1.\displaystyle\frac{H(r_{2})}{H(r_{1})}\geq e^{{-CL(r_{2}-r_{1})}}\Big{(}\frac{r_{2}}{r_{1}}\Big{)}^{2\overline{\mathcal{N}}(r_{1})\exp(-{CLr_{2}})+d-1}. (3.32)

Combining (3.30) and (3.32), we arrive at

eCL(r2r1)(r2r1)2𝒩¯(r1)exp(CLr2)+d1H(r2)H(r1)eCL(r2r1)(r2r1)2𝒩¯(r2)exp(CLr1)+d1.\displaystyle e^{{-CL(r_{2}-r_{1})}}\Big{(}\frac{r_{2}}{r_{1}}\Big{)}^{2\overline{\mathcal{N}}(r_{1})\exp(-{CLr_{2}})+d-1}\leq\frac{H(r_{2})}{H(r_{1})}\leq e^{{CL(r_{2}-r_{1})}}\Big{(}\frac{r_{2}}{r_{1}}\Big{)}^{2\overline{\mathcal{N}}(r_{2})\exp(-CLr_{1})+d-1}. (3.33)

Next we want to show an upper bound for 𝒩¯(34)\overline{\mathcal{N}}(\frac{3}{4}). Let r2=34r_{2}=\frac{3}{4} and 0<r1=r<340<r_{1}=r<\frac{3}{4}. From the estimate (3.32), we have

eCL(34r)(34r)d1H(34)H(r).\displaystyle e^{{-CL(\frac{3}{4}-r)}}\bigg{(}\frac{\frac{3}{4}}{r}\bigg{)}^{d-1}\leq\frac{H(\frac{3}{4})}{H(r)}. (3.34)

Using the fact that 0<θ120<\theta\leq\frac{1}{2}, we have

uL(Bθ)2\displaystyle\|u\|^{2}_{L^{\infty}(B_{\theta})} CB34u2𝑑VgC034H(r)𝑑r\displaystyle\leq C\int_{B_{\frac{3}{4}}}u^{2}dV_{g}\leq C\int^{\frac{3}{4}}_{0}H(r)dr
C034rd1eCL(34r)H(34)𝑑r\displaystyle\leq C\int^{\frac{3}{4}}_{0}r^{d-1}e^{{CL(\frac{3}{4}-r)}}H\Big{(}\frac{3}{4}\Big{)}dr
CeCLH(34),\displaystyle\leq Ce^{{CL}}H\Big{(}\frac{3}{4}\Big{)}, (3.35)

where CC depends on dd and Λ\Lambda. Obviously,

uL(B1)2\displaystyle\|u\|^{2}_{L^{\infty}(B_{1})} CB1u2𝑑Sg.\displaystyle\geq C\int_{\mathbb{\partial}B_{{1}}}u^{2}dS_{g}. (3.36)

Therefore, from (3.3), (3.32) and (3.35), we have

M2\displaystyle M^{2} uL(B1)2uL(Bθ)2CH(1)CeCLH(34)\displaystyle\geq\frac{\|u\|^{2}_{L^{\infty}(B_{1})}}{\|u\|^{2}_{L^{\infty}(B_{\theta})}}\geq\frac{CH(1)}{Ce^{{C}L}H(\frac{3}{4})}
eCL(43)d1+2𝒩¯(34)eCL.\displaystyle\geq e^{{-CL}}\Big{(}\frac{4}{3}\Big{)}^{d-1+2\overline{\mathcal{N}}(\frac{3}{4})e^{{-CL}}}. (3.37)

Thus, we can get an upper bound for 𝒩¯(34)\overline{\mathcal{N}}(\frac{3}{4}) as

𝒩¯(34)CeCLlnM,\displaystyle\overline{\mathcal{N}}\Big{(}\frac{3}{4}\Big{)}\leq Ce^{{CL}}\ln M, (3.38)

where M>1M>1 is a large constant. Choosing any r12r\leq\frac{1}{2}, we integrate (3.28) from θr\theta r to rr, by the monotonicity of 𝒩¯\overline{\mathcal{N}}, we derive that

lnH(r)rd1lnH(θr)(θr)d1\displaystyle\ln\frac{H(r)}{r^{d-1}}-\ln\frac{H(\theta r)}{(\theta r)^{d-1}} CLr+2𝒩¯(34)ln1θ\displaystyle\leq{CLr}+2\overline{\mathcal{N}}(\frac{3}{4})\ln\frac{1}{\theta}
CLr+CeCLlnMln1θ.\displaystyle\leq{CLr}+Ce^{{CL}}\ln M\ln\frac{1}{\theta}. (3.39)

Thus, we obtain that

H(r)\displaystyle H(r) exp(CLr+eCLlnMln1θ)H(θr)\displaystyle\leq\exp({CLr}+e^{{CL}}\ln M\ln\frac{1}{\theta})H(\theta r)
θ1dM(lnθ)eCLH(θr),\displaystyle\leq\theta^{1-d}M^{-(\ln\theta)e^{{CL}}}H(\theta r), (3.40)

where M>1M>1 is large. By further integrations, we can also obtain that

uL2(Br)θd2M(lnθ)eCLuL2(Bθr)\displaystyle\|u\|_{L^{2}(B_{r})}\leq\theta^{\frac{-d}{2}}M^{-(\ln\theta)e^{{CL}}}\|u\|_{L^{2}(B_{\theta r})} (3.41)

for 0<r120<r\leq\frac{1}{2}, where CC depends only on dd and Λ\Lambda. ∎

Remark 3.3.

For the case d=2d=2, we introduce a new variable to apply a lifting argument. Let v(x,t)=etu(x)v(x,t)=e^{t}u(x). Then the new function v(x,t)v(x,t) satisfies the equation

(A~(x,t)v)+v=0inB^1,\displaystyle-\nabla\cdot\big{(}\widetilde{A}(x,t)\nabla v\big{)}+v=0\quad\quad\mbox{in}\ \hat{B}_{1}, (3.42)

where

A~(x,t)=(A(x)001)\widetilde{A}(x,t)=\begin{pmatrix}A(x)&0\\ 0&1\end{pmatrix}

and B^1\hat{B}_{1} is the ball with radius 11 in 3\mathbb{R}^{3}. It is easy to see that A~\widetilde{A} satisfies the conditions (1.2) and (3.2). Following the procedure performed as d3d\geq 3, we are able to introduce the new metric gg and geodesic polar coordinates. Thus, in the metric gg as (3.8) and η\eta as (3.13), we have

g(η(x)gv)+cgv=0inB^1,\displaystyle-\nabla_{g}\cdot(\eta(x)\nabla_{g}v)+c_{g}v=0\quad\mbox{in}\ \hat{B}_{1}, (3.43)

where cg=1detgc_{g}=\frac{1}{\sqrt{\det{g}}}. As before, we could make use of the monotonicity of the frequency function to obtain the doubling inequality. Precisely, we may define

D(r)=B^rη|gv|2+cgv2dVg\displaystyle D(r)=\int_{\hat{B}_{r}}\eta|\nabla_{g}v|^{2}+c_{g}v^{2}dV_{g} (3.44)

and

H(r)=B^rηv2𝑑Sg.\displaystyle H(r)=\int_{\partial\hat{B}_{r}}\eta v^{2}dS_{g}. (3.45)

Then the frequency function is defined as

𝒩(r)=rD(r)H(r).\displaystyle\mathcal{N}(r)=\frac{rD(r)}{H(r)}. (3.46)

Following the proof of Lemma 3.1 and [7, Theorem 3.2.1], we can obtain the almost monotonicity of 𝒩(r)\mathcal{N}(r). That is, for any r0(0,1)r_{0}\in(0,1), it holds that

exp(CLr)𝒩(r)exp(CLr0)+exp(CLr0)𝒩(r0)\displaystyle\exp{(CLr)}\mathcal{N}(r)\leq\exp{(CLr_{0})}+\exp{(CLr_{0})}\mathcal{N}(r_{0}) (3.47)

for any r(0,r0)r\in(0,\ r_{0}) where CC depends on Λ\Lambda. By mimicking the argument in the proof of Lemma 3.2, we can obtain the doubling inequality for vv in B^r\hat{B}_{r}. This also leads to the doubling inequality for uu as (3.27) in BrB_{r}.

Remark 3.4.

For a better estimate when d=2d=2, 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 β\beta such that 1β34=5\frac{1}{\beta-\frac{3}{4}}=5. Our argument works for any β(34,1)\beta\in(\frac{3}{4},1). To handle the case N5ε/rN12τN^{-5}\lesssim\varepsilon/r\lesssim N^{-\frac{1}{2}\tau} 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 d3d\geq 3, we will employ the three-ball inequality with a sharp exponential error term obtained in [2]; for d=2d=2, we will use quasi-regular mappings [1] which provides a much better doubling estimate.

We first introduce a three-ball inequality for all dimensions d2d\geq 2. For convenience, we define the normalized L2L^{2} norm by

uL¯2(Bt)=(Btu2)1/2.\lVert u\rVert_{\underline{L}^{2}(B_{t})}=\bigg{(}\fint_{B_{t}}u^{2}\bigg{)}^{1/2}.

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 τ^(0,1/2)\hat{\tau}\in(0,1/2), there exist c=c(d,Λ)>0c=c(d,\Lambda)>0 and θ=θ(τ^,d,Λ)(0,1/2]\theta=\theta(\hat{\tau},d,\Lambda)\in(0,1/2] such that if uu is a weak solution of 1(u)=0\mathcal{L}_{1}(u)=0 in BRB_{R} with θ2R>2\theta^{2}R>2, then

uL¯2(BθR)uL¯2(Bθ2R)τ^uL¯2(BR)1τ^+exp(cθ2R)uL¯2(BR).\lVert u\rVert_{\underline{L}^{2}(B_{\theta R})}\leq\lVert u\rVert_{\underline{L}^{2}(B_{\theta^{2}R})}^{\hat{\tau}}\lVert u\rVert_{\underline{L}^{2}(B_{R})}^{1-\hat{\tau}}+\exp(-c\theta^{2}R)\lVert u\rVert_{\underline{L}^{2}(B_{R})}. (4.1)

As a simple corollary, we have

Corollary 4.2.

Let uu be a weak solution of 1(u)=0\mathcal{L}_{1}(u)=0 in BRB_{R}. For every α1>0\alpha_{1}>0, there exist C>0C>0 and θ(0,1/2)\theta\in(0,1/2) such that if

N>C and θ2RClnN,N>C\quad\text{ and }\quad\theta^{2}R\geq C\ln N, (4.2)

and

uL¯2(BR)NuL¯2(BθR),\lVert u\rVert_{\underline{L}^{2}(B_{R})}\leq N\lVert u\rVert_{\underline{L}^{2}(B_{\theta R})}, (4.3)

then

uL¯2(BθR)CN1+α1uL¯2(Bθ2R).\lVert u\rVert_{\underline{L}^{2}(B_{\theta R})}\leq CN^{1+\alpha_{1}}\lVert u\rVert_{\underline{L}^{2}(B_{\theta^{2}R})}. (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 lnN\ln N 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 d=2d=2, 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 uu be a weak solution of the equation 1(u)=0\mathcal{L}_{1}(u)=0 in BRB_{R} with only bounded measurable coefficients satisfying (1.2). Let z=x+iyz=x+iy for x,yx,y\in\mathbb{R}. Define

z¯f=12(xf+iyf),zf=12(xfiyf).\displaystyle\partial_{\bar{z}}f=\frac{1}{2}(\partial_{x}f+i\partial_{y}f),\quad\partial_{z}f=\frac{1}{2}(\partial_{x}f-i\partial_{y}f). (4.5)

We introduce a stream function (the generalized harmonic conjugate) associated with uu as

v=JAu,\displaystyle\nabla v=JA\nabla u,

where JJ is the rotation matrix in the plane

J=(0110).\displaystyle J=\begin{pmatrix}0&-1\vskip 6.0pt plus 2.0pt minus 2.0pt\\ 1&0\end{pmatrix}.

Let f=u+ivf=u+iv. Then we have fHloc1(BR)f\in H^{1}_{\text{loc}}(B_{R}) and satisfies

z¯f=μzf+νzf¯,\displaystyle\partial_{\bar{z}}f=\mu\partial_{z}f+\nu\overline{\partial_{z}f},

where the complex valued function μ\mu and ν\nu can be explicitly written in term of AA and

|μ|+|ν|1Λ1+Λ<1.\displaystyle|\mu|+|\nu|\leq\frac{1-\Lambda}{1+\Lambda}<1.

Hence, f:BRf:B_{R}\to\mathbb{C} is a 1Λ\frac{1}{\Lambda}-quasi-regular mapping. Moreover, it can be written as f=Fχ^f=F\circ\hat{\chi}, where FF is holomorphic and χ^:BRBR\hat{\chi}:B_{R}\to B_{R} is a 1Λ\frac{1}{\Lambda}-quasiconformal homeomorphism satisfying χ^(0)=0\hat{\chi}(0)=0 and χ^(1)=1\hat{\chi}(1)=1. Define

r^={zBR:|χ^(z)|<r^}.\displaystyle\mathcal{B}_{\hat{r}}=\{z\in B_{R}:|\hat{\chi}(z)|<{\hat{r}}\}.

The quasi-balls r^\mathcal{B}_{\hat{r}} are comparable to the standard Euclidean balls in the sense

R=BR,andBR(r^CR)1αr^BR(Cr^R)α,forr<R,\displaystyle\mathcal{B}_{R}=B_{R},\quad\text{and}\quad B_{R(\frac{{\hat{r}}}{CR})^{\frac{1}{\alpha}}}\subset\mathcal{B}_{\hat{r}}\subset B_{R(\frac{C{\hat{r}}}{R})^{{\alpha}}},\quad\mbox{for}\ r<R, (4.6)

where C1C\geq 1 and 0<α<10<\alpha<1 depend only on Λ\Lambda. Observe that r^\mathcal{B}_{\hat{r}} tends to be singular if r^R\hat{r}\ll R, which fortunately is not too restrictive as we only use it in the transition at intermediate scales.

From the fact that FF is a holomorphic function, the following doubling property holds [1].

Lemma 4.3.

If uHloc1(BR)u\in H^{1}_{{\rm loc}}(B_{R}) is a nonzero weak solution of 1(u)=0\mathcal{L}_{1}(u)=0 in BRB_{R}, then

uL(r^)uL(r^2)CuL()uL(R4),for 0<r^R.\displaystyle\frac{\|u\|_{L^{\infty}(\mathcal{B}_{\hat{r}})}}{\|u\|_{L^{\infty}(\mathcal{B}_{\frac{{\hat{r}}}{2}})}}\leq C\frac{\|u\|_{L^{\infty}(\mathcal{B_{R}})}}{\|u\|_{L^{\infty}(\mathcal{B}_{\frac{R}{4}})}},\quad\mbox{for}\ 0<{\hat{r}}\leq R. (4.7)
Remark 4.4.

Note that Lemma 4.3 does not use periodicity, and it is also true for solutions of ε(uε)=0\mathcal{L}_{\varepsilon}(u_{\varepsilon})=0, with a constant independent of ε\varepsilon. 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 rr goes to zero. Because of this, we still need to use the periodicity assumption and Step 1 below, when d=2d=2, to cover the range rCN5εr\geq CN^{5}\varepsilon. We then apply Lemma 4.3 in the range Cε<r<CN5εC\varepsilon<r<CN^{5}\varepsilon, using (4.6) since rr is not too small. Finally, the case 0<r<Cε0<r<C\varepsilon is handled by scaling and Lemma 3.2. The details are below.

Equipped with Corollary 4.2 and Lemma 4.3, we are ready to prove Theorem 1.1.

Proof of Theorem 1.1.

According to the relationship between ε\varepsilon and NN, one needs to consider three cases based on the comparison of ε\varepsilon with N5N^{-5} and Nτ2N^{-\frac{\tau}{2}} (or 11 for d=2d=2). Without loss of generality, we may just consider the most complicated case εN5\varepsilon\lesssim N^{-5}, since all the three steps listed in the introduction will be involved as rr approaches 0. Hence, we fix ε\varepsilon and NN so that CN5ε<1CN^{5}\varepsilon<1, and then discuss the different ranges of rr.

Step 1: CN5ε<r<1CN^{5}\varepsilon<r<1. Under either (1.7) or (1.9), Theorem 2.6 implies

Bruε28N3Bθruε2,\int_{B_{r}}u_{\varepsilon}^{2}\leq 8N^{3}\int_{B_{\theta r}}u_{\varepsilon}^{2}, (4.8)

for any given θ(0,Λ2]\theta\in(0,\frac{\Lambda}{2}]. This estimate holds for all dimensions d2d\geq 2.

Step 2: In this step, we need to treat the cases d3d\geq 3 and d=2d=2 separately.

Case 1: d3d\geq 3 and CεNτ2<r<CN5εC\varepsilon N^{\frac{\tau}{2}}<r<CN^{5}\varepsilon for any fixed τ>0\tau>0. Let mm be the smallest integer so that θmr>CN5ε\theta^{-m}r>CN^{5}\varepsilon. If NN is bounded by some absolute constant, then Step 2 is not needed. Since r>CεNτ2r>C\varepsilon N^{\frac{\tau}{2}}, for sufficiently large NN, mm satisfies

m6lnNlnθ.m\leq\frac{6\ln N}{-\ln\theta}. (4.9)

Because of (4.8), we have

Bθmruε28N3Bθm+1ruε2.\int_{B_{\theta^{-m}r}}u_{\varepsilon}^{2}\leq 8N^{3}\int_{B_{\theta^{-m+1}r}}u_{\varepsilon}^{2}. (4.10)

Let M0=8N3M_{0}=8N^{3} and MjM_{j} be the constant such that

Bθm+jruε2MjBθm+j+1ruε2.\int_{B_{\theta^{-m+j}r}}u_{\varepsilon}^{2}\leq M_{j}\int_{B_{\theta^{-m+j+1}r}}u_{\varepsilon}^{2}. (4.11)

The goal is to estimate MmM_{m} with mm comparable to the bound in (4.9).

Thanks to Corollary 4.2, and by rescaling, we know that for a given α1>0\alpha_{1}>0 with θ\theta small enough, we have

Mj=CMj11+α1.M_{j}=CM_{j-1}^{1+\alpha_{1}}. (4.12)

Note that the left-end restriction r>CεlnMjr>C\varepsilon\ln M_{j} is needed in order to apply Corollary 4.2, due to (4.2). This can be guaranteed if we eventually show MjMm<Cexp(Nτ2)M_{j}\leq M_{m}<C\exp(N^{\frac{\tau}{2}}).

We now proceed to estimate MjM_{j}. Using the initial condition M0=8N3M_{0}=8N^{3}, one can show explicitly that

Mj=exp(lnC/α1)exp[(1+α1)j(3lnN+ln(8C1/α1))].M_{j}=\exp(-\ln C/\alpha_{1})\exp\Big{[}(1+\alpha_{1})^{j}\big{(}3\ln N+\ln(8C^{1/\alpha_{1}})\big{)}\Big{]}. (4.13)

It follows from (4.9) that

MmCexp[exp(ln(1+α1)(lnθ)16lnN)(3lnN+ln(8C1/α1))].M_{m}\leq C\exp\Big{[}\exp\big{(}\ln(1+\alpha_{1})(-\ln\theta)^{-1}6\ln N\big{)}\cdot\big{(}3\ln N+\ln(8C^{1/\alpha_{1}})\big{)}\Big{]}. (4.14)

Note that τ\tau is any given positive constant. Then, we may choose α1\alpha_{1} small enough (hence θ\theta is also small), so that

τ36ln(1+α1)(lnθ)1.\frac{\tau}{3}\geq 6\ln(1+\alpha_{1})(-\ln\theta)^{-1}. (4.15)

Thus, if NN is large enough,

MmCexp(N12τ).M_{m}\leq C\exp(N^{\frac{1}{2}\tau}). (4.16)

This implies that for any CN12τε<r<CN5εCN^{\frac{1}{2}\tau}\varepsilon<r<CN^{5}\varepsilon, we have

Bruε2Cexp(N12τ)Bθruε2.\int_{B_{r}}u_{\varepsilon}^{2}\leq C\exp(N^{\frac{1}{2}\tau})\int_{B_{\theta r}}u_{\varepsilon}^{2}. (4.17)

Case 2: d=2d=2 and Cε<r<CN5εC\varepsilon<r<CN^{5}\varepsilon. From (4.8) with θ=Λ2\theta=\frac{\Lambda}{2} in Step 1, for RCεN5R\simeq C\varepsilon N^{5},

BRuε28N3BΛ2Ruε2.\displaystyle\int_{B_{R}}u_{\varepsilon}^{2}\leq 8N^{3}\int_{B_{\frac{\Lambda}{2}R}}u_{\varepsilon}^{2}.

By the LL^{\infty} norm estimates, it follows that

uεL(BR)CN32uεL(BR2).\displaystyle\|u_{\varepsilon}\|_{L^{\infty}(B_{R})}\leq CN^{\frac{3}{2}}\|u_{\varepsilon}\|_{L^{\infty}(B_{\frac{R}{2}})}. (4.18)

We would like to to apply Lemma 4.3 to uεu_{\varepsilon}. From the relation (4.6) of quasi-balls r^\mathcal{B}_{\hat{r}} and the standard balls, as well as the iteration of the doubling inequality (4.18), we have

uεL(R)uεL(R4)uεL(BR)uεL(BR(4C)1α)CNk,\displaystyle\frac{\|u_{\varepsilon}\|_{L^{\infty}(\mathcal{B}_{R})}}{\|u_{\varepsilon}\|_{L^{\infty}(\mathcal{B}_{\frac{R}{4}})}}\leq\frac{\|u_{\varepsilon}\|_{L^{\infty}({B_{R}})}}{\|u_{\varepsilon}\|_{L^{\infty}({B_{R(4C)^{-\frac{1}{\alpha}}}})}}\leq CN^{k}, (4.19)

where kk depends only on Λ\Lambda. Thus, (4.7) implies that for any 0<r^<R0<\hat{r}<R

uεL(r^)uεL(r^2)CNk.\displaystyle\frac{\|u_{\varepsilon}\|_{L^{\infty}(\mathcal{B}_{\hat{r}})}}{\|u_{\varepsilon}\|_{L^{\infty}(\mathcal{B}_{\frac{{\hat{r}}}{2}})}}\leq CN^{k}. (4.20)

In order to establish a doubling inequality at small scale on standard Euclidean balls, we iterate the above doubling inequality mm times to obtain

uεL(r^)(CNk)muεL(r^2m).\displaystyle{\|u_{\varepsilon}\|_{L^{\infty}(\mathcal{B}_{\hat{r}})}}\leq(CN^{k})^{m}{\|u_{\varepsilon}\|_{L^{\infty}(\mathcal{B}_{\frac{{\hat{r}}}{2^{m}}})}}. (4.21)

By the relation (4.6),

uεL(BR(r^CR)1α)(CNk)muεL(BR(Cr^2mR)α).\displaystyle\|u_{\varepsilon}\|_{L^{\infty}(B_{R(\frac{{\hat{r}}}{CR})^{\frac{1}{\alpha}}})}\leq(CN^{k})^{m}\|u_{\varepsilon}\|_{L^{\infty}(B_{R(\frac{C{\hat{r}}}{2^{m}R})^{{\alpha}}})}.

We choose m>0m>0 to be the smallest integer so that

R(Cr^2mR)α12R(r^CR)1α.R(\frac{C{\hat{r}}}{2^{m}R})^{{\alpha}}\leq\frac{1}{2}R(\frac{{\hat{r}}}{CR})^{\frac{1}{\alpha}}. (4.22)

Consequently,

uεL(BR(r^CR)1α)(CNk)muεL(B12R(r^CR)1α).\|u_{\varepsilon}\|_{L^{\infty}(B_{R(\frac{{\hat{r}}}{CR})^{\frac{1}{\alpha}}})}\leq(CN^{k})^{m}\|u_{\varepsilon}\|_{L^{\infty}(B_{\frac{1}{2}R(\frac{{\hat{r}}}{CR})^{\frac{1}{\alpha}}})}. (4.23)

Note that r^\hat{r}, satisfying 0<r^<RCεN50<\hat{r}<R\simeq C\varepsilon N^{5}, is arbitrary and mm is chosen depending on r^\hat{r}. We now assume r^CεαR1αCεN5(1α){\hat{r}}\geq C\varepsilon^{\alpha}R^{1-\alpha}\simeq C\varepsilon N^{5(1-\alpha)}. Hence, r:=R(r^CR)1αCεr:=R(\frac{{\hat{r}}}{CR})^{\frac{1}{\alpha}}\geq C\varepsilon. Moreover, from (4.22), we have mClnNm\leq C\ln N, where CC depends only on Λ\Lambda. Thus, it follows from (4.23) that

uεL(Br)\displaystyle\|u_{\varepsilon}\|_{L^{\infty}(B_{r})} CNClnNuεL(Br2)\displaystyle\leq CN^{C\ln N}\|u_{\varepsilon}\|_{L^{\infty}(B_{\frac{r}{2}})} (4.24)

for all Cε<r<RCεN5C\varepsilon<r<R\simeq C\varepsilon N^{5}. Since LL^{\infty} norm can be replaced by L2L^{2} norm in the above inequality, we derive the desired estimate for the case d=2d=2.

Step 3: For r<CεN12τr<C\varepsilon N^{\frac{1}{2}\tau} (or r<Cεr<C\varepsilon for d=2d=2), by rescaling, the equation may be reduced to the case in which the Lipschitz constant of coefficients is bounded by CNτ2CN^{\frac{\tau}{2}} (bounded by CC for d=2d=2). It follows from (3.27) and (4.17) that for d3d\geq 3 and any 0<r<CεN12τ0<r<C\varepsilon N^{\frac{1}{2}\tau},

Bruε2\displaystyle\int_{B_{r}}u_{\varepsilon}^{2} C[exp(N12τ)]Cexp(N12τ)Bθruε2\displaystyle\leq C\Big{[}\exp(N^{\frac{1}{2}\tau})\Big{]}^{C\exp(N^{\frac{1}{2}\tau})}\int_{B_{\theta r}}u_{\varepsilon}^{2}
exp(exp(CNτ))Bθruε2.\displaystyle\leq\exp(\exp(CN^{\tau}))\int_{B_{\theta r}}u_{\varepsilon}^{2}.

For d=2d=2, it follows from (3.27) and (4.24) that for any 0<r<Cε0<r<C\varepsilon,

Bruε2C[NClnN]CBr2uε2CNClnNBr2uε2.\int_{B_{r}}u_{\varepsilon}^{2}\leq C\big{[}N^{C\ln N}\big{]}^{C}\int_{B_{\frac{r}{2}}}u_{\varepsilon}^{2}\leq CN^{C\ln N}\int_{B_{\frac{r}{2}}}u_{\varepsilon}^{2}. (4.25)

Note that NClnN=exp(C(lnN)2)N^{C\ln N}=\exp(C(\ln N)^{2}). 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 τ^=12\hat{\tau}=\frac{1}{2}), without any smoothness assumption on the coefficients. If the critical τ^=1/2\hat{\tau}=1/2 in (4.1) can also be achieved (which seems like a very difficult task), then Corollary 4.2 with α1=0\alpha_{1}=0 would follow. By the argument in Step 2, this would yield the estimate

Bruε2CNkBθruε2\displaystyle\int_{B_{r}}u_{\varepsilon}^{2}\leq CN^{k}\int_{B_{\theta r}}u_{\varepsilon}^{2} (4.26)

for CεlnNrCεN5C\varepsilon\ln N\leq r\leq C\varepsilon N^{5}. If we then apply (3.27) as in Step 3, with Lipschitz constant ClnNC\ln N, we would obtain the bound C(N)=exp(CNC)C(N)=\exp{(CN^{C})} for 0<r120<r\leq\frac{1}{2} (for the range 0<r<Cε0<r<C\varepsilon, (3.27) does give the optimal bound). On the other hand, the estimate (3.27) in term of the large Lipschitz constant LL 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 CεrCεlnNC\varepsilon\leq r\leq C\varepsilon\ln N, one could try to use a method taking advantage of both periodicity and smoothness. No such method is available at the moment.

For d=2d=2, if the critical case τ^=12\hat{\tau}=\frac{1}{2} 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 d=2d=2, 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 r>CεlnNr>C\varepsilon\ln N. By reproducing the argument in Step 2 (Case 2), we can then show mClnlnNm\leq C\ln\ln N and therefore

uεL(Br)CNClnlnNuεL(Br2)\|u_{\varepsilon}\|_{L^{\infty}(B_{r})}\leq CN^{C\ln\ln N}\|u_{\varepsilon}\|_{L^{\infty}(B_{\frac{r}{2}})} (4.27)

for all r>Cεr>C\varepsilon. Then a blow-up argument gives the same estimate for 0<r<Cε0<r<C\varepsilon. Observe that (4.27) is exactly the ultimate estimate we expect, as mentioned in the introduction.

Remark 4.6.

If we consider, when d=2d=2, elliptic operators with Lipschitz coefficients, with Lipschitz constant L>1L>1 (and no periodicity assumption), we can obtain the improved bound MC1lnLM^{C_{1}\ln L} in Lemma 3.2. To show this, we break down the scales into 1Lr<1\frac{1}{L}\leq r<1 and 0<r<1L0<r<\frac{1}{L}. For the case 1Lr<1\frac{1}{L}\leq r<1, we use Lemma 4.3, (4.6) and the argument from (4.19) to (4.23). For the case 0<r<1L0<r<\frac{1}{L}, we scale to reduce to the case L=1L=1 and then apply Lemma 3.2 as it stands. This may suggest that the bound in Lemma 3.2 is not optimal, also for d3d\geq 3.

Remark 4.7.

The disadvantage of Theorem 4.1 for d3d\geq 3 is that θ\theta may be very small. If we do not apply Theorem 4.1 to improve the exponent τ\tau, Step 1 and 3 in the proof of Theorem 1.1 allows θ\theta to be any number in (0,Λ/2](0,\Lambda/2]. In particular, under (1.5), for any β(34,1)\beta\in(\frac{3}{4},1), we have

Bruε2exp(exp(CN1β34))BΛr/2uε2.\int_{B_{r}}u_{\varepsilon}^{2}\leq\exp(\exp(CN^{\frac{1}{\beta-\frac{3}{4}}}))\int_{B_{\Lambda r/2}}u_{\varepsilon}^{2}. (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 α\alpha has to be larger than 88 in (1.14).

5. Upper bounds of Nodal sets

In this section, we study of the upper bounds of nodal sets for uεu_{\varepsilon}, where uεu_{\varepsilon} is a nonzero solution of ε(uε)=0\mathcal{L}_{\varepsilon}(u_{\varepsilon})=0 satisfying (1.5). We will focus on the general treatment for all dimensions d2d\geq 2 and with an eye towards d=2d=2 in the end. Throughout this section, up to a change of variable, we assume 0=Δ\mathcal{L}_{0}=-\Delta. Note that in this case, ErE_{r}’s are just balls, and in view of Theorem 2.1, the assumption (1.5) can be replaced by

B2uε2NB1uε2,\int_{B_{2}}u_{\varepsilon}^{2}\leq N\int_{B_{1}}u_{\varepsilon}^{2}, (5.1)

and (4.28) holds with Λ=1\Lambda=1.

5.1. Small scales

We first show that a doubling inequality centered at 0 implies the doubling inequality with shifted centers.

Lemma 5.1.

Let uεu_{\varepsilon} be a weak solution of ε(uε)=0\mathcal{L}_{\varepsilon}(u_{\varepsilon})=0 in B2B_{2} satisfying (5.1). Then for any xB1/3x\in B_{1/3} and B2r(x)B2B_{2r}(x)\subset B_{2}, we have

B2r(x)uε2exp(exp(CN2β34))Br(x)uε2.\int_{B_{2r}(x)}u_{\varepsilon}^{2}\leq\exp(\exp(CN^{\frac{2}{\beta-\frac{3}{4}}}))\int_{B_{r}(x)}u_{\varepsilon}^{2}. (5.2)
Proof.

Let us first assume CεN1β34<1C\varepsilon N^{\frac{1}{\beta-\frac{3}{4}}}<1 for some large CC. In this case, by Theorem 2.1 with θ=1/2\theta=1/2, we have

B2uε2NB1uε22N2B1/2uε2.\int_{B_{2}}u_{\varepsilon}^{2}\leq N\int_{B_{1}}u_{\varepsilon}^{2}\leq 2N^{2}\int_{B_{1/2}}u_{\varepsilon}^{2}. (5.3)

Now, for any xB1/3x\in B_{1/3}, note that B1/2B5/6(x)B_{1/2}\subset B_{5/6}(x) and B5/3(x)B2B_{5/3}(x)\subset B_{2}. It follows from (5.3) that

B5/3(x)uε2B2uε22N2B1/2uε22N2B5/6(x)uε2.\int_{B_{5/3}(x)}u_{\varepsilon}^{2}\leq\int_{B_{2}}u_{\varepsilon}^{2}\leq 2N^{2}\int_{B_{1/2}}u_{\varepsilon}^{2}\leq 2N^{2}\int_{B_{5/6}(x)}u_{\varepsilon}^{2}. (5.4)

Since Theorem 2.1 and (4.28) are invariant under translation, we can apply them in B5/3(x)B_{5/3}(x) with NN replaced by 2N22N^{2}. Thus, for all r(0,5/6)r\in(0,5/6),

B2r(x)uε2exp(exp(CN2β34))Br(x)uε2.\int_{B_{2r}(x)}u_{\varepsilon}^{2}\leq\exp(\exp(CN^{\frac{2}{\beta-\frac{3}{4}}}))\int_{B_{r}(x)}u_{\varepsilon}^{2}.

To handle the case CεN1β341C\varepsilon N^{\frac{1}{\beta-\frac{3}{4}}}\geq 1, we use (4.28) directly and obtain

B2uε2NB1uε2exp(exp(CN1β34))B1/2uε2.\int_{B_{2}}u_{\varepsilon}^{2}\leq N\int_{B_{1}}u_{\varepsilon}^{2}\leq\exp(\exp(CN^{\frac{1}{\beta-\frac{3}{4}}}))\int_{B_{1/2}}u_{\varepsilon}^{2}.

Then the desired estimate follows from the same idea as the first case and a blow up argument as in Step 3 in the proof of Theorem 1.1. ∎

Let us define the nodal sets as

Z(uε)={xB2|uε=0}\displaystyle Z(u_{\varepsilon})=\{x\in B_{2}|u_{\varepsilon}=0\} (5.5)

and the density function of nodal sets as

Eε(y,r)=Hd1(Z(uε)Br(y))rd1.\displaystyle E_{\varepsilon}(y,r)=\frac{H^{d-1}(Z(u_{\varepsilon})\cap B_{r}(y))}{r^{d-1}}. (5.6)

Based on Lemma 5.1 and a blow up argument, we can estimate the Hausdoff measure of the nodal set of uεu_{\varepsilon} in small balls.

Lemma 5.2.

For any 0<r<1/30<r<1/3 and x0B1/3x_{0}\in B_{1/3} such that Br(x0)B1/3B_{r}(x_{0})\subset B_{1/3},

Eε(x0,r)(1+rε)exp(CN2β34),\displaystyle E_{\varepsilon}(x_{0},r)\leq\Big{(}1+\frac{r}{\varepsilon}\Big{)}\exp(CN^{\frac{2}{\beta-\frac{3}{4}}}), (5.7)

where CC depends on d,Λ,βd,\Lambda,\beta and γ\gamma.

Proof.

First of all, we consider the case 0<rε0<r\leq\varepsilon and Br(x0)B1/3B_{r}(x_{0})\subset B_{1/3}. Let v(x)=uε(x0+rx)v(x)=u_{\varepsilon}(x_{0}+rx) and Ax0ε,r(x)=A(ε1(x0+rx))A^{\varepsilon,r}_{x_{0}}(x)=A(\varepsilon^{-1}(x_{0}+rx)). Then

(Ax0ε,r(x)v(x))=0.\displaystyle\nabla(A^{\varepsilon,r}_{x_{0}}(x)\nabla v(x))=0. (5.8)

By (1.4),

|Ax0ε,r(x)Ax0ε,r(y)|γrε1|xy|γ|xy|\displaystyle|A^{\varepsilon,r}_{x_{0}}(x)-A^{\varepsilon,r}_{x_{0}}(y)|\leq\gamma r\varepsilon^{-1}|x-y|\leq\gamma|x-y| (5.9)

for x,yB2x,y\in B_{2}. Therefore, in this case, the coefficient matrix has a uniform Lipschitz constant independent of ε\varepsilon and NN. Then, a change of variable and the doubling inequality in Lemma 5.1 give that

B2v2𝑑x\displaystyle\fint_{B_{2}}v^{2}\,dx =B2r(x0)uε2𝑑x\displaystyle=\fint_{B_{2r}(x_{0})}u_{\varepsilon}^{2}\,dx
exp(exp(CN2β34))Br(x0)uε2𝑑x\displaystyle\leq\exp(\exp(CN^{\frac{2}{\beta-\frac{3}{4}}}))\fint_{B_{r}(x_{0})}u_{\varepsilon}^{2}\,dx
exp(exp(CN2β34))B1v2𝑑x.\displaystyle\leq\exp(\exp(CN^{\frac{2}{\beta-\frac{3}{4}}}))\fint_{B_{1}}v^{2}\,dx. (5.10)

By the upper bound of nodal sets in [16], there exists a constant β0>12\beta_{0}>\frac{1}{2} so that

Hd1(Z(v)B1)[exp(CN2β34)]β0exp(C1N2β34),\displaystyle H^{d-1}(Z(v)\cap B_{1})\leq\big{[}\exp(CN^{\frac{2}{\beta-\frac{3}{4}}})\big{]}^{\beta_{0}}\leq\exp(C_{1}N^{\frac{2}{\beta-\frac{3}{4}}}), (5.11)

which implies, by rescaling,

Hd1(Z(uε)Br(x0))exp(C1N2β34)rd1H^{d-1}(Z(u_{\varepsilon})\cap B_{r}(x_{0}))\leq\exp(C_{1}N^{\frac{2}{\beta-\frac{3}{4}}})\;r^{d-1}

for any r(0,ε]r\in(0,\varepsilon] and Br(x0)B1/3B_{r}(x_{0})\subset B_{1/3}.

Next, to deal with the case r>εr>\varepsilon, we simply use a covering argument. Let x0B1/3x_{0}\in B_{1/3} and r>εr>\varepsilon. There there exists a family of balls Bε(xi),i=1,2,,M,B_{\varepsilon}(x_{i}),i=1,2,\cdots,M, that covers Br(x0)B_{r}(x_{0}) with a finite number of overlaps depending only on dd. Note that M(r/ε)dM\approx(r/\varepsilon)^{d}. Consequently,

Hd1(Z(uε)Br(x0))\displaystyle H^{d-1}(Z(u_{\varepsilon})\cap B_{r}(x_{0})) i=1MHd1(Z(uε)Bε(xi))\displaystyle\leq\sum_{i=1}^{M}H^{d-1}(Z(u_{\varepsilon})\cap B_{\varepsilon}(x_{i}))
Mexp(C1N2β34)εd1\displaystyle\leq M\exp(C_{1}N^{\frac{2}{\beta-\frac{3}{4}}})\;\varepsilon^{d-1}
Crdε1exp(C1N2β34).\displaystyle\leq Cr^{d}\varepsilon^{-1}\exp(C_{1}N^{\frac{2}{\beta-\frac{3}{4}}}).

We obtain the desired estimate by enlarging the constant C1C_{1}. ∎

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 vv is a solution of (A(x)v)=0\nabla\cdot(A(x)\nabla v)=0 in B2B_{2}. In addition to the ellipticity condition (1.2), we assume

|A(x)A(y)|L|xy|.\displaystyle|A(x)-A(y)|\leq L|x-y|. (5.12)

Then

Eε(x0,r)C(1+Lr)N(v,Q)β0,E_{\varepsilon}(x_{0},r)\leq C(1+Lr)N(v,Q)^{\beta_{0}},

for Br(x0)QB_{r}(x_{0})\subset Q, where the definition of N(v,Q)N(v,Q) 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 u0u_{0}, close to uεu_{\varepsilon} under LL^{\infty} norm, and satisfying a doubling inequality.

Lemma 5.4.

Suppose r>3CNεr>3C\sqrt{N}\varepsilon for some large CC. Let uεu_{\varepsilon} be a solution of ε(uε)=0\mathcal{L}_{\varepsilon}(u_{\varepsilon})=0 in B2rB_{2r} satisfying

B2ruε2NBruε2.\int_{B_{2r}}u_{\varepsilon}^{2}\leq N\int_{B_{r}}u_{\varepsilon}^{2}. (5.13)

Then there exists u0u_{0} satisfying 0(u0)=0\mathcal{L}_{0}(u_{0})=0 in B7r4B_{\frac{7r}{4}} such that

uεu0L(B3r2)CεruεL¯2(B2r),\displaystyle\|u_{\varepsilon}-u_{0}\|_{L^{\infty}(B_{\frac{3r}{2}})}\leq\frac{C\varepsilon}{r}\|u_{\varepsilon}\|_{\underline{L}^{2}(B_{2r})}, (5.14)

and

Bru0216N2Br/2u02,\int_{B_{r}}u_{0}^{2}\leq 16N^{2}\int_{B_{r/2}}u_{0}^{2}, (5.15)

where CC depends on d,Λd,\Lambda and γ\gamma.

Proof.

By rescaling, we may assume r=1r=1. The construction of such locally homogenized solution u0u_{0} and the estimate (5.14) can be found in [15, Theorem 2.3]. Note that it is not necessary that uε=u0u_{\varepsilon}=u_{0} on B74\partial B_{\frac{7}{4}}. Then, it suffices to show (5.15). By (5.13) and (5.14), we have

uεu0L(B32)2Cε2B2uε2Cε2NB1uε2.\displaystyle\|u_{\varepsilon}-u_{0}\|^{2}_{L^{\infty}(B_{\frac{3}{2}})}\leq C\varepsilon^{2}\int_{B_{2}}u_{\varepsilon}^{2}\leq{C\varepsilon^{2}N}\int_{B_{1}}u_{\varepsilon}^{2}. (5.16)

We now establish estimates to compare the norms of uεu_{\varepsilon} and u0u_{0}. Thanks to (5.16),

u0L2(B32)\displaystyle\|u_{0}\|_{L^{2}(B_{\frac{3}{2}})} uεL2(B2)+uεu0L2(B32)\displaystyle\leq\|u_{\varepsilon}\|_{L^{2}(B_{2})}+\|u_{\varepsilon}-u_{0}\|_{L^{2}(B_{\frac{3}{2}})}
NuεL2(B1)+CNεuεL2(B1)\displaystyle\leq\sqrt{N}\|u_{\varepsilon}\|_{L^{2}(B_{1})}+{C\sqrt{N}\varepsilon}\|u_{\varepsilon}\|_{L^{2}(B_{1})}
=N(1+Cε)uεL2(B1).\displaystyle=\sqrt{N}(1+C\varepsilon)\|u_{\varepsilon}\|_{L^{2}(B_{1})}. (5.17)

By the same strategy, using (5.16), we obtain that

uεL2(B1)\displaystyle\|u_{\varepsilon}\|_{L^{2}(B_{1})} uεu0L2(B1)+u0L2(B1)\displaystyle\leq\|u_{\varepsilon}-u_{0}\|_{L^{2}(B_{1})}+\|u_{0}\|_{L^{2}(B_{1})}
CNεuεL2(B1)+u0L2(B1).\displaystyle\leq C\sqrt{N}\varepsilon\|u_{\varepsilon}\|_{L^{2}(B_{1})}+\|u_{0}\|_{L^{2}(B_{1})}. (5.18)

Since CNε13C\sqrt{N}\varepsilon\leq\frac{1}{3}, the above estimate yields

uεL2(B1)(1+CNε)u0L2(B1).\displaystyle\|u_{\varepsilon}\|_{L^{2}(B_{1})}\leq(1+C\sqrt{N}\varepsilon)\|u_{0}\|_{L^{2}(B_{1})}. (5.19)

Combining (5.17) and (5.19) together yields that

u0L2(B32)\displaystyle\|u_{0}\|_{L^{2}(B_{\frac{3}{2}})} N(1+CNε)u0L2(B1)\displaystyle\leq\sqrt{N}(1+C\sqrt{N}\varepsilon)\|u_{0}\|_{L^{2}(B_{1})}
2Nu0L2(B1).\displaystyle\leq 2\sqrt{N}\|u_{0}\|_{L^{2}(B_{1})}. (5.20)

Now, we use the fact that

φ(s)=log2B2su02\varphi(s)=\log_{2}\fint_{B_{2^{s}}}u_{0}^{2} (5.21)

is a convex function with respect to ss. Then f(s)=φ(s)φ(sc)f(s)=\varphi(s)-\varphi(s-c) is nondecreasing for any c>0c>0. This implies

u0L2(B1)\displaystyle\|u_{0}\|_{L^{2}(B_{1})} 2Nu0L2(B23)\displaystyle\leq 2\sqrt{N}\|u_{0}\|_{L^{2}(B_{\frac{2}{3}})}
4Nu0L2(B49)\displaystyle\leq 4{N}\|u_{0}\|_{L^{2}(B_{\frac{4}{9}})}
4Nu0L2(B12).\displaystyle\leq 4{N}\|u_{0}\|_{L^{2}(B_{\frac{1}{2}})}. (5.22)

This proves (5.15) and the lemma. ∎

Remark 5.5.

We would like to point out that the advantage of Lemma 5.4, compared to (2.2), is that it provides an LL^{\infty} (or pointwise) error estimate which is much stronger than the L2L^{2} error estimate in (2.2)(\ref{est.ue-u0}). This LL^{\infty} estimate will play an essential role in the estimation of nodal sets.

Let BB be a ball and u0u_{0} be a C1C^{1} function in 2B2B. In order to show some quantitative stratification results for u0u_{0} and u0\nabla u_{0}, we introduce the doubling index:

N(u0,B)=log2sup2B|u0|supB|u0|\displaystyle N(u_{0},B)=\log_{2}\frac{\sup_{2B}|u_{0}|}{\sup_{B}|u_{0}|} (5.23)

and

N(u0,B)=log2sup2B|u0|supB|u0|.\displaystyle N(\nabla u_{0},B)=\log_{2}\frac{\sup_{2B}|\nabla u_{0}|}{\sup_{B}|\nabla u_{0}|}. (5.24)

If u0u_{0} is a weak solution of the equation 0(u0)=0\mathcal{L}_{0}(u_{0})=0, the doubling index for |u0||u_{0}| and |u0||\nabla u_{0}| are monotonic in the sense that

N(u0,tB)CN(u0,B)\displaystyle N(u_{0},tB)\leq CN(u_{0},B) (5.25)

and

N(u0,tB)CN(u0,B),\displaystyle N(\nabla u_{0},tB)\leq CN(\nabla u_{0},B), (5.26)

for t12t\leq\frac{1}{2} and CC depending only on dd. 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 QQ, denote by s(Q)s(Q) the side length of QQ. Define the doubling index in the cube QQ by

N(u0,Q)=supxQ,rs(Q)log2supB2r(x)|u0|supBr(x)|u0|\displaystyle N(u_{0},Q)=\sup_{x\in Q,r\leq s(Q)}\log_{2}\frac{\sup_{B_{2r}(x)}|u_{0}|}{\sup_{B_{r}(x)}|u_{0}|} (5.27)

and

N(u0,Q)=supxQ,rs(Q)log2supB2r(x)|u0|supBr(x)|u0|.\displaystyle N(\nabla u_{0},Q)=\sup_{x\in Q,r\leq s(Q)}\log_{2}\frac{\sup_{B_{2r}(x)}|\nabla u_{0}|}{\sup_{B_{r}(x)}|\nabla u_{0}|}. (5.28)

The doubling index defined in cubes is convenient in the sense that if a cube qq is a subset of QQ, then N(u0,q)N(u0,Q)N(u_{0},q)\leq N(u_{0},Q). Let qq be a subcube of QQ and K=s(Q)s(q)2K=\frac{s(Q)}{s(q)}\geq 2. Then

supq|u0|KCN(u0,Q)supQ|u0|,\displaystyle\sup_{q}|u_{0}|\geq K^{-CN(u_{0},Q)}\sup_{Q}|u_{0}|, (5.29)

where CC depends only on dd. Similarly, it also holds

supq|u0|KCN(u0,Q)supQ|u0|.\displaystyle\sup_{q}|\nabla u_{0}|\geq K^{-CN(\nabla u_{0},Q)}\sup_{Q}|\nabla u_{0}|. (5.30)

The following quantitative stratification for u0u_{0} and u0\nabla u_{0} 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 u0u_{0} is harmonic in 5Q5Q.

  1. (1)

    Suppose N(u0,Q)N0N(u_{0},Q)\leq{N_{0}}. If 0<δ<exp(CN0)0<\delta<\exp(-C^{*}N_{0}) for some C>0C^{*}>0, there exists a finite sequence of balls {Bti(xi)}i=1m\{B_{t_{i}}(x_{i})\}_{i=1}^{m} such that

    Gδ={x12Q:|u0(x)|<δsupQ|u0(x)|}i=1mBti(xi)\displaystyle G_{\delta}=\Big{\{}x\in\frac{1}{2}Q:|u_{0}(x)|<\delta\sup_{Q}|u_{0}(x)|\Big{\}}\subset\bigcup^{m}_{i=1}B_{t_{i}}(x_{i}) (5.31)

    and

    i=1mtid1CN0Cs(Q)d1,\sum^{m}_{i=1}t_{i}^{d-1}\leq C{N_{0}}^{C}s(Q)^{d-1}, (5.32)

    where CC and CC^{*} depend only on dd.

  2. (2)

    Suppose N(u0,Q)N^0N(\nabla u_{0},Q)\leq{\hat{N}_{0}}. If 0<δ^eCN^030<\hat{\delta}\leq e^{-C\hat{N}_{0}^{3}} for some C>0C>0 depending on dd, there exists a finite sequence of balls {Bt^j(xj)}j=1m^\{B_{\hat{t}_{j}}(x_{j})\}_{j=1}^{\hat{m}} such that

    G^δ^={x12Q:|u0(x)|<δ^supQ|u0(x)|}j=1m^Bt^j(xj)\displaystyle\hat{G}_{\hat{\delta}}=\Big{\{}x\in\frac{1}{2}Q:|\nabla u_{0}(x)|<\hat{\delta}\sup_{Q}|\nabla u_{0}(x)|\Big{\}}\subset\bigcup^{\hat{m}}_{j=1}B_{\hat{t}_{j}}(x_{j}) (5.33)

    and

    j=1m^t^jd114(s(Q)4)d1.\sum^{\hat{m}}_{j=1}\hat{t}_{j}^{d-1}\leq\frac{1}{4}(\frac{s(Q)}{4})^{d-1}.\quad (5.34)
Proof.

In the following proof, all the constants C,C,C0,C1,,C14C,C^{*},C_{0},C_{1},\cdots,C_{14} depend only on dd, and N0N_{0}, N^0\hat{N}_{0} are large constants.

(1) Let K=δτN0K=\delta^{-\frac{\tau}{{N_{0}}}} and δeC0N0τ\delta\leq e^{-\frac{C_{0}N_{0}}{\tau}}, where τ\tau is small to be specified later. We can assume that KK is an integer and K8K\geq 8. We divide the cube 12Q\frac{1}{2}Q into KdK^{d} equal subcubes qiq_{i}. Then 4qiQ4q_{i}\subset Q. We would like to estimate the number of cubes qiq_{i} that intersect GδG_{\delta}.

Let qiq_{i} be a cube with qiGδq_{i}\cap G_{\delta}\not=\emptyset. Thus, we have infqi|u0|<δsupQ|u0|\inf_{q_{i}}|u_{0}|<\delta\sup_{Q}|u_{0}|. We claim that if δ<eCN0\delta<e^{-C^{*}N_{0}} for some large C>1C^{*}>1, then u0u_{0} changes sign in 2qi2q_{i}. Assume that u0u_{0} does not change sign in 2qi2q_{i}. By the Harnack inequality,

supqi|u0|C1infqi|u0|C1δsupQ|u0|.\displaystyle\sup_{q_{i}}|u_{0}|\leq C_{1}\inf_{q_{i}}|u_{0}|\leq C_{1}\delta\sup_{Q}|u_{0}|. (5.35)

On the other hand, by the monotonicity of the doubling index in cubes (5.29),

supqi|u0|C3KC2N0supQ|u0|=C3δC2τsupQ|u0|.\displaystyle\sup_{q_{i}}|u_{0}|\geq C_{3}K^{-C_{2}{N_{0}}}\sup_{Q}|u_{0}|=C_{3}\delta^{C_{2}\tau}\sup_{Q}|u_{0}|. (5.36)

Choosing τ=12C2\tau=\frac{1}{2C_{2}}, we reach a contradiction if

δ12min{C3C1,eC0C2N0}.\displaystyle\delta^{\frac{1}{2}}\leq\min\bigg{\{}\frac{C_{3}}{C_{1}},\ e^{-{C_{0}C_{2}N_{0}}}\bigg{\}}.

Since N01N_{0}\geq 1, the last inequality holds if we choose δ<eCN0\delta<e^{-{C^{\ast}N_{0}}}. This proves the claim. Hence, there are zeros in each 2qi2q_{i} and qiGδq_{i}\cap G_{\delta}\not=\emptyset with

Gδi=1mqi.\displaystyle G_{\delta}\subset\bigcup^{m}_{i=1}q_{i}. (5.37)

This implies (5.31) as we may replace qiq_{i} by Bti(xi)B_{t_{i}}(x_{i}) with the same center and ti=s(qi)d2t_{i}=\frac{s(q_{i})\sqrt{d}}{2}.

Next, to show the first part of (5.32), we need to estimate the number mm of the cubes qiq_{i}. Recall that 4qiQ4q_{i}\subset Q and each point in QQ may be covered by at most a finite number of 4qi4q_{i}. By the lower bound estimate of nodal sets in [17], we have

Hd1({u0=0}Q)Ci=1mHd1({u0=0}4qi)C4m(s(Q)K)d1.\displaystyle H^{d-1}(\{u_{0}=0\}\cap Q)\geq C\sum_{i=1}^{m}H^{d-1}(\{u_{0}=0\}\cap 4q_{i})\geq C_{4}m\Big{(}\frac{s(Q)}{K}\Big{)}^{d-1}. (5.38)

On the other hand, by the upper bound estimate of nodal sets in [16], it holds that

Hd1({u0=0}Q)C5N0C6s(Q)d1,\displaystyle H^{d-1}(\{u_{0}=0\}\cap Q)\leq C_{5}N_{0}^{C_{6}}s(Q)^{d-1}, (5.39)

where C6>12C_{6}>\frac{1}{2}. Combining (5.38) and (5.39), we arrive at

C4m(s(Q)K)d1C5N0C6s(Q)d1,\displaystyle C_{4}m\Big{(}\frac{s(Q)}{K}\Big{)}^{d-1}\leq C_{5}{N}_{0}^{C_{6}}s(Q)^{d-1}, (5.40)

which yields

mC5C4N0C6Kd1.\displaystyle m\leq\frac{C_{5}}{C_{4}}{N}_{0}^{C_{6}}K^{d-1}. (5.41)

Thus,

i=1mtid1=Cdms(qi)d1CN0C6Kd1(s(Q)K)d1CN0C6s(Q)d1.\displaystyle\sum_{i=1}^{m}t_{i}^{d-1}=C_{d}m\cdot s(q_{i})^{d-1}\leq C{N}_{0}^{C_{6}}K^{d-1}\Big{(}\frac{s(Q)}{K}\Big{)}^{d-1}\leq C{N}_{0}^{C_{6}}s(Q)^{d-1}. (5.42)

This proves (1).

(2) Next, we establish the estimates (5.33) and (5.34). We divide the cube 12Q\frac{1}{2}Q into K1dK_{1}^{d} subcubes with side length s(Q)2K1\frac{s(Q)}{2K_{1}}. The size of K1K_{1}, depending on δ^\hat{\delta}, will be chosen later. The cube qjq_{j} is called bad if

infqj|u0|csup2qj|u0|\displaystyle\inf_{q_{j}}|\nabla u_{0}|\leq c\sup_{2q_{j}}|\nabla u_{0}| (5.43)

for some small cc depending only on dd. We claim that the number of bad cubes qjq_{j} is not greater than eCdN^02K1d2e^{C_{d}\hat{N}_{0}^{2}}K_{1}^{d-2}, where CdC_{d} depends on dd.

To show the above claim, we need to use [22, Theorem 1.1]. Recall the effective critical set is defined as

𝒞r(u0)={xQ:infBr(x)r2|u0|2d16B2r(x)(uu(x))2}.\mathcal{C}_{r}(u_{0})=\bigg{\{}x\in Q:\inf_{B_{r}(x)}r^{2}|\nabla u_{0}|^{2}\leq\frac{d}{16}\fint_{\partial B_{2r}(x)}(u-u(x))^{2}\bigg{\}}.

Let Br(𝒞r(u0))B_{r}(\mathcal{C}_{r}(u_{0})) be the rr-neighborhood of 𝒞r(u0)\mathcal{C}_{r}(u_{0}), namely, Br(𝒞r(u0))={xQ:dist(x,𝒞r(u0))<r}B_{r}(\mathcal{C}_{r}(u_{0}))=\{x\in Q:\text{dist}(x,\mathcal{C}_{r}(u_{0}))<r\}. Then [22, Theorem 1.1] implies

|Br(𝒞r(u0))Bs|C(N~(u0,B2s))2(rs)2|Bs|,|B_{r}(\mathcal{C}_{r}(u_{0}))\cap B_{s}|\leq C^{\big{(}\widetilde{N}(u_{0},B_{2s})\big{)}^{2}}\Big{(}\frac{r}{s}\Big{)}^{2}|B_{s}|, (5.44)

where Bs,B2sB_{s},B_{2s} are concentric balls such that B4sQB_{4s}\subset Q and N~\widetilde{N} is the modified frequency function defined by

N~(u0,B2s):=2sB2s|u0|2B2s(u0u0(z))2,\widetilde{N}(u_{0},B_{2s}):=\frac{2s\int_{B_{2s}}|\nabla u_{0}|^{2}}{\int_{\partial B_{2s}}(u_{0}-{u_{0}(z)})^{2}},

where zz is the center of BsB_{s}. By [7, Corollary 2.2.6] and the mean value property of harmonic functions, we have

N~(u0,B2s)Clog2B4s(u0u0(z))2B2s(u0u0(z))2Clog2supB4s|u0|supBs|u0|CN(u0,Q)CN^0,\widetilde{N}(u_{0},B_{2s})\leq C\log_{2}\frac{\int_{B_{4s}}(u_{0}-{u_{0}(z)})^{2}}{\int_{B_{2s}}(u_{0}-u_{0}(z))^{2}}\leq C\log_{2}\frac{\sup_{B_{4s}}|\nabla u_{0}|}{\sup_{B_{s}}|\nabla u_{0}|}\leq CN(\nabla u_{0},Q)\leq C\hat{N}_{0},

where we have also used a gradient estimate for harmonic functions in the second inequality. Hence, (5.44) implies

|Br(𝒞r(u0))Bs|CN^02(rs)2|Bs|.|B_{r}(\mathcal{C}_{r}(u_{0}))\cap B_{s}|\leq C^{\hat{N}_{0}^{2}}\Big{(}\frac{r}{s}\Big{)}^{2}|B_{s}|. (5.45)

Next, we show that if qjq_{j} is a bad cube with sufficiently small cc, then qj𝒞r(u0).q_{j}\cap\mathcal{C}_{r}(u_{0})\neq\emptyset. Actually if qjq_{j} is bad and xjx_{j} is the point in qjq_{j} so that |u0(xj)|=infqj|u0||\nabla u_{0}(x_{j})|=\inf_{{q_{j}}}|\nabla u_{0}|, then

infBr(xj)|u0||u0(xj)|=infqj|u0|csup2qj|u0|,\inf_{B_{r}(x_{j})}|\nabla u_{0}|\leq|\nabla u_{0}(x_{j})|=\inf_{q_{j}}|\nabla u_{0}|\leq c\sup_{2q_{j}}|\nabla u_{0}|,

where we used the condition (5.43) in the last inequality. Fix r=2ds(qj)r=2\sqrt{d}s(q_{j}). Then 2qjBr(xj)2q_{j}\subset B_{r}(x_{j}). It follows from the gradient estimate and the Caccioppoli inequality that

infBr(xj)r|u0|crsupBr(xj)|u0|\displaystyle\inf_{B_{r}(x_{j})}r|\nabla u_{0}|\leq cr\sup_{B_{r}(x_{j})}|\nabla u_{0}| cCdr(B32r(xj)|u0|2)1/2\displaystyle\leq cC_{d}r\bigg{(}\fint_{B_{\frac{3}{2}r}(x_{j})}|\nabla u_{0}|^{2}\bigg{)}^{1/2}
cCd2(B2r(xj)|u0u0(xj)|2)1/2.\displaystyle\leq cC_{d}^{2}\bigg{(}\fint_{B_{2r}(x_{j})}|u_{0}-u_{0}(x_{j})|^{2}\bigg{)}^{1/2}.

In view of [7, Corollary 2.2.7], we have

infBr(xj)r|u0|\displaystyle\inf_{B_{r}(x_{j})}r|\nabla u_{0}| cCd2d(B2r(xj)|u0u0(xj)|2)1/2\displaystyle\leq\frac{cC_{d}^{2}}{d}\bigg{(}\fint_{\partial B_{2r}(x_{j})}|u_{0}-u_{0}(x_{j})|^{2}\bigg{)}^{1/2}
d16(B2r(xj)|u0u0(xj)|2)1/2,\displaystyle\leq\sqrt{\frac{d}{16}}\bigg{(}\fint_{\partial B_{2r}(x_{j})}|u_{0}-u_{0}(x_{j})|^{2}\bigg{)}^{1/2},

where in the last inequality, we choose cc small so that cCd2/d<d/16cC_{d}^{2}/d<\sqrt{d/16}. This implies that xj𝒞r(u0)x_{j}\in\mathcal{C}_{r}(u_{0}) and qj𝒞r(u0)q_{j}\cap\mathcal{C}_{r}(u_{0})\neq\emptyset. Because r=2ds(qj)r=2\sqrt{d}s(q_{j}), we have qjBr(𝒞r(u0))q_{j}\subset B_{r}(\mathcal{C}_{r}(u_{0})). This means that all the bad cubes qjq_{j} are contained in Br(𝒞r(u0))B_{r}(\mathcal{C}_{r}(u_{0})). Finally, let ss be comparable to s(Q)s(Q) and note that 12Q\frac{1}{2}Q can be covered by finitely many, depending only on dd, BsB_{s} with B4sQB_{4s}\subset Q. Then, by (5.45), the total volume of bad cubes in 12Q\frac{1}{2}Q is bounded by CN^02(s(qj)/s(Q))2|Q|CN^02K12|Q|C^{\hat{N}_{0}^{2}}\big{(}s(q_{j})/s(Q)\big{)}^{2}|Q|\leq C^{\hat{N}_{0}^{2}}K_{1}^{-2}|Q|. Hence, the number of bad cubes is not greater than CN^02K1d2C^{\hat{N}_{0}^{2}}K_{1}^{d-2}. The claim has been proved.

Now, for any qjq_{j}, the monotonicity of the doubling index of u0\nabla u_{0} in cubes in (5.30) shows that

supqj|u0|C8K1C7N^0supQ|u0|.\displaystyle\sup_{q_{j}}|\nabla u_{0}|\geq C_{8}K_{1}^{-C_{7}\hat{N}_{0}}\sup_{Q}|\nabla u_{0}|. (5.46)

If qjq_{j} is not bad, the reverse inequality of (5.43) yields

infqj|u0|>C9K1C7N^0supQ|u0|.\displaystyle\inf_{q_{j}}|\nabla u_{0}|>C_{9}K_{1}^{-C_{7}\hat{N}_{0}}\sup_{Q}|\nabla u_{0}|. (5.47)

Given δ^\hat{\delta}, small enough (to be quantified later), we want to estimate the set G^δ^\hat{G}_{\hat{\delta}} defined in (5.33). If qjq_{j} is not bad and we choose K1K_{1} to be the smallest integer such that

C9(K1+1)C7N^0<δ^,\displaystyle C_{9}(K_{1}+1)^{-C_{7}\hat{N}_{0}}<\hat{\delta}, (5.48)

then (5.47) gives

infqj|u0|C9K1C7N^0supQ|u0|>δ^supQ|u0|.\inf_{q_{j}}|\nabla u_{0}|\geq C_{9}K_{1}^{-C_{7}\hat{N}_{0}}\sup_{Q}|\nabla u_{0}|>\hat{\delta}\sup_{Q}|\nabla u_{0}|.

This implies that qjq_{j} does not intersect G^δ\hat{G}_{\delta}. It also shows that K1δ^1C10N^0K_{1}\approx\hat{\delta}^{\frac{-1}{C_{10}\hat{N}_{0}}}. Thus, the set G^δ^\hat{G}_{\hat{\delta}} is covered by the union of bad cubes of size s(Q)2K1\frac{s(Q)}{2K_{1}}. Again, we may now replace bad qjq_{j} by Bt^j(xj)B_{\hat{t}_{j}}(x_{j}) with the same center and t^j=s(qj)d2\hat{t}_{j}=\frac{s(q_{j})\sqrt{d}}{2}. Let m^\hat{m} be the number of bad cubes and recall that eCdN^02K1d2m^e^{C_{d}\hat{N}_{0}^{2}}K_{1}^{d-2}\geq\hat{m}. It follows that

j=1m^t^jd1=Cm^s(qj)d1\displaystyle\sum^{\hat{m}}_{j=1}\hat{t}_{j}^{d-1}=C\hat{m}\cdot s(q_{j})^{d-1} CeCdN^02K1d2(s(Q)2K1)d1\displaystyle\leq Ce^{C_{d}\hat{N}_{0}^{2}}K_{1}^{d-2}\bigg{(}\frac{s(Q)}{2K_{1}}\bigg{)}^{d-1}
(12)d1sd1(Q)CeCdN^02K11\displaystyle\leq\Big{(}\frac{1}{2}\Big{)}^{d-1}s^{d-1}(Q)Ce^{C_{d}\hat{N}_{0}^{2}}K_{1}^{-1}
(12)d1sd1(Q)CeCdN^02δ^1C10N^0\displaystyle\leq\Big{(}\frac{1}{2}\Big{)}^{d-1}s^{d-1}(Q)Ce^{C_{d}\hat{N}_{0}^{2}}{\hat{\delta}}^{\frac{1}{C_{10}\hat{N}_{0}}}
(14)dsd1(Q)\displaystyle\leq\Big{(}\frac{1}{4}\Big{)}^{d}s^{d-1}(Q) (5.49)

where we have chosen δ^eC11N^03\hat{\delta}\leq e^{-C_{11}\hat{N}_{0}^{3}} in the last inequality. This completes the proof. ∎

Next we estimate the density function Eε(y,r)E_{\varepsilon}(y,r) 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 QQ in Lemma 5.6 by B18B_{\frac{1}{8}}.

Lemma 5.7.

Let β(34,1)\beta\in(\frac{3}{4},1) and (5.1) hold. If εexp(C(lnN)3)\varepsilon\leq\exp(-C(\ln N)^{3}), then there exists a finite sequence of balls {Bt^j(yj):j=1,m^}\{B_{\hat{t}_{j}}(y_{j}):j=1\cdots,\hat{m}\} such that yjB116y_{j}\in B_{\frac{1}{16}}, t^j(0,1128)\hat{t}_{j}\in(0,\frac{1}{128}) and

Eε(0,116)exp(CN2β34)+14sup1jm^Eε(yj,t^j),\displaystyle E_{\varepsilon}(0,\frac{1}{16})\leq\exp(CN^{\frac{2}{\beta-\frac{3}{4}}})+\frac{1}{4}\sup_{1\leq j\leq\hat{m}}E_{\varepsilon}(y_{j},\hat{t}_{j}), (5.50)

where cc and CC depend on dd, Λ\Lambda and γ\gamma.

Proof.

We assume Eε(0,116)>0E_{\varepsilon}(0,\frac{1}{16})>0. 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 r=1r=1, we have

uεu0L(B18)C^εuεL2(B2).\displaystyle\|u_{\varepsilon}-u_{0}\|_{L^{\infty}(B_{\frac{1}{8}})}\leq\hat{C}\varepsilon\|u_{\varepsilon}\|_{L^{2}(B_{2})}. (5.51)

By normalization, we may assume that B2uε2=1\int_{B_{2}}u_{\varepsilon}^{2}=1. We would like to estimate the doubling index for u0u_{0} and u0\nabla u_{0}. From (5.20) and convexity of φ\varphi in (5.21), we have

u0L2(B3t2)\displaystyle\|u_{0}\|_{L^{2}(B_{\frac{3t}{2}})} 2Nu0L2(Bt)\displaystyle\leq 2\sqrt{N}\|u_{0}\|_{L^{2}(B_{{t}})} (5.52)

for all 0<t<10<t<1. By elliptic estimates and using (5.52) twice,

u0L(B1)+u0L(B1)\displaystyle\|u_{0}\|_{L^{\infty}(B_{1})}+\|\nabla u_{0}\|_{L^{\infty}(B_{1})} C(B98u02)12\displaystyle\leq C\bigg{(}\int_{B_{\frac{9}{8}}}u^{2}_{0}\bigg{)}^{\frac{1}{2}} (5.53)
CN(B34u02)12CN(B12u02)12CNu0L(B12).\displaystyle\leq C\sqrt{N}\bigg{(}\int_{B_{\frac{3}{4}}}u^{2}_{0}\bigg{)}^{\frac{1}{2}}\leq C{N}\bigg{(}\int_{B_{\frac{1}{2}}}u^{2}_{0}\bigg{)}^{\frac{1}{2}}\leq CN\|u_{0}\|_{L^{\infty}(B_{\frac{1}{2}})}.

The above estimate includes

u0L(B1)CNu0L(B12).\|u_{0}\|_{L^{\infty}({B_{1}})}\leq CN\|u_{0}\|_{L^{\infty}({B_{\frac{1}{2}}})}. (5.54)

Thus, N(u0,B12)log2(CN)N(u_{0},B_{\frac{1}{2}})\leq\log_{2}{(CN)}. Similarly to the frequency function in Section 3, one can define the frequency function introduced for the harmonic function u0u_{0} as

𝒩(x,r)=rBr(x)|u0|2Br(x)u02.\displaystyle\mathcal{N}(x,r)=\frac{r\int_{B_{r}(x)}|\nabla u_{0}|^{2}}{\int_{\partial B_{r}(x)}u_{0}^{2}}.

By Theorem 2.2.8 in [7], it holds that

𝒩(x,12(12R))C𝒩(0,12)\displaystyle\mathcal{N}(x,\ \frac{1}{2}(\frac{1}{2}-R))\leq C\mathcal{N}(0,\frac{1}{2}) (5.55)

for any xBRx\in B_{R} and 0<R<120<R<\frac{1}{2}. It is known that the doubling index and the frequency function are comparable. It follows from Lemma 7.1 in [16] that

𝒩(0,12)CN(u0,B12)Clog2(CN),\displaystyle\mathcal{N}(0,\frac{1}{2})\leq CN(u_{0},B_{\frac{1}{2}})\leq C\log_{2}{(CN)}, (5.56)

where CC depends on dd. From (5.55), let R=18R=\frac{1}{8}, we obtain that

𝒩(x,316)Clog2(CN).\displaystyle\mathcal{N}(x,\frac{3}{16})\leq C\log_{2}{(CN)}.

Using Lemma 7.1 in [16] again, we see that

N(u0,Br(x))C𝒩(x,316)Clog2(CN)\displaystyle N(u_{0},B_{r}(x))\leq C\mathcal{N}(x,\frac{3}{16})\leq C\log_{2}{(CN)} (5.57)

for any 0<r18.0<r\leq\frac{1}{8}. Thus, we obtain that

u0L(B2r(x))CNC0u0L(Br(x))\displaystyle\|u_{0}\|_{L^{\infty}({B_{2r}(x)})}\leq CN^{C_{0}}\|u_{0}\|_{L^{\infty}({B_{r}(x)})} (5.58)

for any 0<r180<r\leq\frac{1}{8} and any xB18x\in B_{\frac{1}{8}}, where C0C_{0} depends on dd.

Next, we estimate the doubling index of u0\nabla u_{0}. We first claim that there are zeros for u0u_{0} in B18B_{\frac{1}{8}}. In fact, from (5.51) and Theorem 2.1 with θ=1/2\theta=1/2, we have

uεu0L2(B18)CN2εuεL2(B18).\displaystyle\|u_{\varepsilon}-u_{0}\|_{L^{2}(B_{\frac{1}{8}})}\leq{C}{N^{2}}\varepsilon\|u_{\varepsilon}\|_{L^{2}(B_{\frac{1}{8}})}. (5.59)

Then

u0L2(B18)\displaystyle\|u_{0}\|_{L^{2}(B_{\frac{1}{8}})} (1CN2ε)uεL2(B18)\displaystyle\geq(1-{C}{N^{2}}\varepsilon)\|u_{\varepsilon}\|_{L^{2}(B_{\frac{1}{8}})}
(1CN2ε)CN2uεL2(B2).\displaystyle\geq\frac{(1-{C}{N^{2}}\varepsilon)}{C{N^{2}}}\|u_{\varepsilon}\|_{L^{2}(B_{2})}. (5.60)

It follows that u0L2(B18)1CN2\|u_{0}\|_{L^{2}(B_{\frac{1}{8}})}\geq\frac{1}{CN^{2}} if ε<cN2\varepsilon<cN^{-{2}}. Hence, (5.15) implies

1CN2u0L2(B18)CNu0L2(B116).\displaystyle\frac{1}{CN^{2}}\leq\|u_{0}\|_{L^{2}(B_{\frac{1}{8}})}\leq CN\|u_{0}\|_{L^{2}(B_{\frac{1}{16}})}. (5.61)

Now, let us assume that u0u_{0} has no zeros in B18B_{\frac{1}{8}} and therefore does not change signs in B18B_{\frac{1}{8}}. Without loss of generality, we may assume that u0u_{0} is positive. By the Harnack inequality and (5.61),

infB116|u0|CsupB116|u0|1CN3.\displaystyle\inf_{B_{\frac{1}{16}}}|u_{0}|\geq C\sup_{B_{\frac{1}{16}}}|u_{0}|\geq\frac{1}{CN^{3}}. (5.62)

From (5.51), for xB116x\in B_{\frac{1}{16}}, we get

1CN3C^εuε(x).\displaystyle\frac{1}{CN^{3}}-\hat{C}\varepsilon\leq u_{\varepsilon}(x). (5.63)

Since ε<CN3\varepsilon<CN^{-3}, then uε(x)>0u_{\varepsilon}(x)>0 for xB116x\in B_{\frac{1}{16}}. This contradicts our assumption that Eε(0,116)>0E_{\varepsilon}(0,\frac{1}{16})>0. Thus, the claim has been shown.

Now, since u0u_{0} has zeros in B18B_{\frac{1}{8}} (and hence in B12B_{\frac{1}{2}}), we obtain from (5.53) and the mean value theorem that

supB1|u0|CNsupB12|u0|.\displaystyle\sup_{B_{1}}|\nabla u_{0}|\leq CN\sup_{B_{\frac{1}{2}}}|\nabla u_{0}|. (5.64)

Again, by the relation of the frequency function for |u0||\nabla u_{0}| and the doubling index, we can argue as the derivation of (5.58) that (5.64) implies that

supB2r(x)|u0|CNC0supBr(x)|u0|\displaystyle\sup_{B_{2r}(x)}|\nabla u_{0}|\leq CN^{C_{0}}\sup_{B_{r}(x)}|\nabla u_{0}| (5.65)

for any 0<r180<r\leq\frac{1}{8} and any xB18x\in B_{\frac{1}{8}}.

Thanks to the monotonicity of the doubling index for u0u_{0} and u0\nabla u_{0}, from the definition of N(u0,Q)N(u_{0},Q) and N(u0,Q)N(\nabla u_{0},Q), we know that N(u0,Q)N0:=ClogNN(u_{0},Q)\leq{N}_{0}:=C\log N and N(u0,Q)N^0:=ClogNN(\nabla u_{0},Q)\leq\hat{N}_{0}:=C\log N. In order to apply Lemma 5.6, we assume that C^εδeCClogN\hat{C}\varepsilon\leq\delta\leq e^{-C^{\ast}C\log N} and δ^212eC(logN)3\frac{\hat{\delta}}{2}\approx\frac{1}{2}e^{-C(\log N)^{3}}. Thus, we require εCNα\varepsilon\leq CN^{-\alpha} for some α\alpha depending on dd, which is satisfied by the assumption of ε\varepsilon in the lemma. With the aid of (5.51), we have

Z(uε)B116\displaystyle Z(u_{\varepsilon})\cap B_{\frac{1}{16}} Z(uε){xB116:|u0(x)|C^ε}\displaystyle\subset Z(u_{\varepsilon})\cap\{x\in B_{\frac{1}{16}}:|u_{0}(x)|\leq\hat{C}\varepsilon\}
Z(uε){xB116:|u0(x)|C^εand|u0(x)|δ^2supB18|u0|}\displaystyle\subset Z(u_{\varepsilon})\cap\{x\in{B_{\frac{1}{16}}}:|u_{0}(x)|\leq\hat{C}\varepsilon\ \mbox{and}\ |\nabla u_{0}(x)|\geq\frac{\hat{\delta}}{2}\sup_{{B_{\frac{1}{8}}}}|\nabla u_{0}|\}
Z(uε){xB116:|u0(x)|C^εand|u0(x)|δ^2supB18|u0|}\displaystyle\bigcup Z(u_{\varepsilon})\cap\{x\in{B_{\frac{1}{16}}}:|u_{0}(x)|\leq\hat{C}\varepsilon\ \mbox{and}\ |\nabla u_{0}(x)|\leq\frac{\hat{\delta}}{2}\sup_{B_{\frac{1}{8}}}|\nabla u_{0}|\}
(i=1mZ(uε)Gi)(j=1m^Z(uε)Bt^j(yj)),\displaystyle\subset\big{(}\cup^{m}_{i=1}Z(u_{\varepsilon})\cap G_{i}\big{)}\bigcup\big{(}\cup^{\hat{m}}_{j=1}Z(u_{\varepsilon})\cap B_{\hat{t}_{j}}(y_{j})\big{)}, (5.66)

where

Gi={xBti(xi)||u0(x)|C^ε and |u0(x)|δ^2supB18|u0|},\displaystyle G_{i}=\bigg{\{}x\in B_{{t}_{i}}(x_{i})|\ |u_{0}(x)|\leq\hat{C}\varepsilon\ \mbox{ and }\ |\nabla u_{0}(x)|\geq\frac{\hat{\delta}}{2}\sup_{B_{\frac{1}{8}}}|\nabla u_{0}|\bigg{\}}, (5.67)

and Bt^j(yj)B_{\hat{t}_{j}}(y_{j}) and Bti(xi)B_{t_{i}}(x_{i}) are given by Lemma 5.6. Thus, it follows from Lemma 5.6 that

Hd1(Z(uε)B116)\displaystyle H^{d-1}(Z(u_{\varepsilon})\cap{B_{\frac{1}{16}}}) i=1mHd1(Z(uε)Gi)+j=1m^Hd1(Z(uε)Bt^j(yj))\displaystyle\leq\sum^{m}_{i=1}H^{d-1}(Z(u_{\varepsilon})\cap G_{i})+\sum^{\hat{m}}_{j=1}H^{d-1}(Z(u_{\varepsilon})\cap B_{\hat{t}_{j}}(y_{j}))
(supiHd1(Z(uε)Gi)tid1)i=1mtid1+supjEε(yj,t^j)j=1m^t^jd1\displaystyle\leq\bigg{(}\sup_{i}\frac{H^{d-1}(Z(u_{\varepsilon})\cap G_{i})}{t_{i}^{d-1}}\bigg{)}\sum^{m}_{i=1}t_{i}^{d-1}+\sup_{j}E_{\varepsilon}(y_{j},\hat{t}_{j})\sum^{\hat{m}}_{j=1}\hat{t}_{j}^{d-1}
C(116)d1(logN)α^supiHd1(Z(uε)Gi)tid1+14(116)d1supjEε(yj,t^j),\displaystyle\leq C(\frac{1}{16})^{d-1}(\log N)^{\hat{\alpha}}\sup_{i}\frac{H^{d-1}(Z(u_{\varepsilon})\cap G_{i})}{t_{i}^{d-1}}+\frac{1}{4}(\frac{1}{16})^{d-1}\sup_{j}E_{\varepsilon}(y_{j},\hat{t}_{j}), (5.68)

where α^\hat{\alpha} depends on dd. Since NN is large, by the decomposition in Lemma 5.6, we may assume 0<t^j<11280<\hat{t}_{j}<\frac{1}{128}.

Next we estimate the upper bound for Hd1(Z(uε)Gi)H^{d-1}(Z(u_{\varepsilon})\cap G_{i}) for each ii. We will discuss two cases tiC^εt_{i}\leq\hat{C}\varepsilon and tiC^εt_{i}\geq\hat{C}\varepsilon. If tiC^εt_{i}\leq\hat{C}\varepsilon, by Lemma 5.2,

Hd1(Z(uε)Gi)\displaystyle H^{d-1}(Z(u_{\varepsilon})\cap G_{i}) Hd1(Z(uε)Bti(xi))\displaystyle\leq H^{d-1}(Z(u_{\varepsilon})\cap B_{t_{i}}(x_{i}))
Cexp(CN2β34)tid1.\displaystyle\leq C\exp(CN^{\frac{2}{\beta-\frac{3}{4}}})t_{i}^{d-1}. (5.69)

Now, we consider the case tiC^εt_{i}\geq\hat{C}\varepsilon. Note that (5.61), the fact that u0u_{0} has zeros in B18B_{\frac{1}{8}} and the definition of GiG_{i} imply

|u0(x)|δ^C~2N2,and|u0(x)|C^ε|\nabla u_{0}(x)|\geq\frac{\hat{\delta}\tilde{C}}{2N^{2}},\quad\text{and}\quad|u_{0}(x)|\leq\hat{C}\varepsilon (5.70)

for any point xGix\in G_{i}. Fix ii. For k=1,2,,dk=1,2,\cdots,d, define

Fk±={xBti(xi)||u0(x)|C^ε,±u0xk(x)δ^c2dN2}.F_{k}^{\pm}=\Big{\{}x\in B_{t_{i}}(x_{i})\Big{|}\ |u_{0}(x)|\leq\hat{C}\varepsilon,\ \pm\frac{\partial u_{0}}{\partial x_{k}}(x)\geq\frac{\hat{\delta}c}{2dN^{2}}\Big{\}}.

Then (5.70) implies that GiG_{i} is contained in k=1d(Fk+Fk)\cup_{k=1}^{d}(F_{k}^{+}\cup F_{k}^{-}). Without loss of generality, it suffices to estimate Fk+F_{k}^{+}. By the C2C^{2} regularity of u0u_{0}, for any x0Fk+x_{0}\in F_{k}^{+}, there exists a cylinder 𝒞(x0)\mathcal{C}(x_{0}) centered at x0x_{0}, whose base is a square perpendicular to eke_{k} with side length CεC\varepsilon, such that the height of 𝒞(x0)\mathcal{C}(x_{0}) is δ^c12dN2\frac{\hat{\delta}c_{1}}{2dN^{2}} and

u0xk(x)δ^c12dN2,for any x𝒞(x0),\frac{\partial u_{0}}{\partial x_{k}}(x)\geq\frac{\hat{\delta}c_{1}}{2dN^{2}},\quad\text{for any }x\in\mathcal{C}(x_{0}), (5.71)

where c1>0c_{1}>0 is a constant smaller than cc.

Refer to caption
Figure 1. The cylinder 𝒞(x0)\mathcal{C}(x_{0})

We would like to show that 𝒞(x0)Fk+\mathcal{C}(x_{0})\cap F_{k}^{+} can be covered by m1m_{1} balls with radius CεC\varepsilon, where m1C22dN2δ^m_{1}\leq\frac{C_{2}2dN^{2}}{\hat{\delta}}. Let 𝒮(x0)\mathcal{S}(x_{0}) be the cross section containing x0x_{0} of the cylinder 𝒞(x0)\mathcal{C}(x_{0}) which is perpendicular to eke_{k}. Since |u0|C|\nabla u_{0}|\leq C and |u0(x0)|C^ε|u_{0}(x_{0})|\leq\hat{C}\varepsilon, we see that |u0(y)|C1ε|u_{0}(y)|\leq C_{1}\varepsilon for any y𝒮(x0)y\in\mathcal{S}(x_{0}). Next, because of (5.71), for any y𝒮(x0)y\in\mathcal{S}(x_{0}) and t>0t>0,

u0(y+tek)tδ^c12dN2C1ε.u_{0}(y+te_{k})\geq t\frac{\hat{\delta}c_{1}}{2dN^{2}}-C_{1}\varepsilon.

This implies that y+tekFk+y+te_{k}\notin F_{k}^{+} if t>(C1+C^)ε2dN2δ^c1t>\frac{(C_{1}+\hat{C})\varepsilon 2dN^{2}}{\hat{\delta}c_{1}}. Similarly, y+tekFk+y+te_{k}\notin F_{k}^{+} if t<(C1+C^)ε2dN2δ^c1t<-\frac{(C_{1}+\hat{C})\varepsilon 2dN^{2}}{\hat{\delta}c_{1}}. This implies that

Fk+𝒞(x0){y+tek|y𝒮(x0),|t|(C1+C^)ε2dN2δ^c1}.F_{k}^{+}\cap\mathcal{C}(x_{0})\subset\Big{\{}y+te_{k}|\ y\in\mathcal{S}(x_{0}),\ |t|\leq\frac{(C_{1}+\hat{C})\varepsilon 2dN^{2}}{\hat{\delta}c_{1}}\Big{\}}.

Consequently, Fk+𝒞(x0)F_{k}^{+}\cap\mathcal{C}(x_{0}) can be covered by m1m_{1} balls with radius CεC\varepsilon and m1C22dN2δ^m_{1}\leq\frac{C_{2}2dN^{2}}{\hat{\delta}}.
Now, because Fk+Bti(xi)F_{k}^{+}\subset B_{t_{i}}(x_{i}) can be covered by m2m_{2} cylinders, with

m2{C|Bti||𝒞(x0)|=CdtidN2δ^εd1,if tiL=δ^c12dN2,Ctid1(Cε)d1=Ctid1εd1,if C^εtiL=δ^c12dN2,m_{2}\leq\left\{\begin{aligned} &\frac{C|B_{t_{i}}|}{|\mathcal{C}(x_{0})|}=\frac{Cdt_{i}^{d}N^{2}}{\hat{\delta}\varepsilon^{d-1}},\qquad\text{if }t_{i}\geq L=\frac{\hat{\delta}c_{1}}{2dN^{2}},\\ &\frac{Ct_{i}^{d-1}}{(C\varepsilon)^{d-1}}=\frac{Ct_{i}^{d-1}}{\varepsilon^{d-1}},\qquad\text{if }\hat{C}\varepsilon\leq t_{i}\leq L=\frac{\hat{\delta}c_{1}}{2dN^{2}},\end{aligned}\right.

such that the 𝒞(x0)\mathcal{C}(x_{0})’s have finite overlaps, then Fk+F_{k}^{+} can be covered by mm balls with radius CεC\varepsilon (denoted by {BCε(zk,+):=1,2,,m}\{B_{C\varepsilon}(z_{k,\ell}^{+}):\ell=1,2,\cdots,m\}) , where

m=m1m2Ctid1d2N4εd1δ^2.m=m_{1}m_{2}\leq\frac{Ct_{i}^{d-1}d^{2}N^{4}}{\varepsilon^{d-1}\hat{\delta}^{2}}.

Note that the same estimate holds for FkF_{k}^{-} as well for each k=1,2,,dk=1,2,\cdots,d.

Hence, by Lemma 5.2, we derive that

Hd1(Z(uε){xBti(xi):|u0|C~δ^2N})\displaystyle H^{d-1}(Z(u_{\varepsilon})\cap\{x\in B_{t_{i}}(x_{i}):|\nabla u_{0}|\geq\frac{\tilde{C}\hat{\delta}}{2N}\}) k=1dl=1mHd1(Z(uε)BCε(zk,l±))\displaystyle\leq\sum_{k=1}^{d}\sum^{m}_{l=1}H^{d-1}(Z(u_{\varepsilon})\cap B_{C\varepsilon}(z_{k,l}^{\pm}))
Cexp(CN2β34)mεd1\displaystyle\leq C\exp(CN^{\frac{2}{\beta-\frac{3}{4}}})m\varepsilon^{d-1}
CN4δ^2exp(CN2β34)tid1\displaystyle\leq CN^{4}\hat{\delta}^{-2}\exp(CN^{\frac{2}{\beta-\frac{3}{4}}})t_{i}^{d-1}
exp(CN2β34)tid1,\displaystyle\leq\exp(CN^{\frac{2}{\beta-\frac{3}{4}}})t_{i}^{d-1}, (5.72)

where in the last inequality, we have used the fact δ^1exp(C(lnN)3)\hat{\delta}^{-1}\approx\exp(C(\ln N)^{3}) and enlarged the constant CC. Note that here β(34,1)\beta\in(\frac{3}{4},1) can be arbitrary. Thus, together with (5.69), we obtain that

Hd1(Z(uε)Gi)exp(CN2β34)tid1.\displaystyle H^{d-1}(Z(u_{\varepsilon})\cap G_{i})\leq\exp(CN^{\frac{2}{\beta-\frac{3}{4}}})t_{i}^{d-1}. (5.73)

Taking (5.68) into account, we arrive at the conclusion (5.50). ∎

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 uεu_{\varepsilon} in the interior domain.

Proof of Theorem 1.2 (d3d\geq 3).

We first consider the case εexp(C(lnN)3)\varepsilon\leq\exp(-C(\ln N)^{3}). Recall from (5.4) that

B5/3(x)uε22N2B5/6(x)uε2\int_{B_{5/3}(x)}u_{\varepsilon}^{2}\leq 2N^{2}\int_{B_{5/6}(x)}u_{\varepsilon}^{2} (5.74)

for any xB1/3x\in B_{1/3}. By Theorem 2.1, it follows that

B2r(x)uε24N2Br(x)uε2\int_{B_{2r}(x)}u_{\varepsilon}^{2}\leq 4N^{2}\int_{B_{r}(x)}u_{\varepsilon}^{2} (5.75)

for CN1β34ε<r<56CN^{\frac{1}{\beta-\frac{3}{4}}}\varepsilon<r<\frac{5}{6} and any xB1/3x\in B_{1/3}. By (5.75) and Lemma 5.4, we derive that

Br(x)u02CN4Br2(x)u02,\displaystyle\int_{B_{r}(x)}u_{0}^{2}\leq CN^{4}\int_{B_{\frac{r}{2}}(x)}u_{0}^{2}\ , (5.76)

as in (5.15) for xB13x\in B_{\frac{1}{3}} and CN1β34ε<r<56CN^{\frac{1}{\beta-\frac{3}{4}}}\varepsilon<r<\frac{5}{6}. 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 E(x0,s)E(x_{0},s) for x0B18x_{0}\in B_{\frac{1}{8}} and Cεexp(C(lnN)3)<sr16C\varepsilon\exp(C(\ln N)^{3})<s\leq{\frac{r}{16}}.

Let v(x)=uε(x0+tx)v(x)=u_{\varepsilon}(x_{0}+tx) for any tt satisfying CN1β34εCεexp(C(lnN)3)<t<56CN^{\frac{1}{\beta-\frac{3}{4}}}\varepsilon\leq C\varepsilon\exp(C(\ln N)^{3})<t<\frac{5}{6} and x0B1/8x_{0}\in B_{1/8}. Then v(x)v(x) satisfies

(Ax0ε,t(x)v(x))=0inB2,\displaystyle\nabla\cdot({A}^{\varepsilon,t}_{x_{0}}(x)\nabla v(x))=0\quad\mbox{in}\ B_{2}, (5.77)

where Ax0ε,t(x)=A(ε1(x0+tx)){A}^{\varepsilon,t}_{x_{0}}(x)=A(\varepsilon^{-1}(x_{0}+tx)). By Lemma 5.7, we have

Hd1(Z(v)B116(0))(116)d1\displaystyle\frac{H^{d-1}(Z(v)\cap B_{\frac{1}{16}}(0))}{(\frac{1}{16})^{d-1}} exp(CN2β34)+14supjHd1(Z(v)Bs~j(yj))(s~j)d1,\displaystyle\leq\exp(CN^{\frac{2}{\beta-\frac{3}{4}}})+\frac{1}{4}\sup_{j}\frac{H^{d-1}(Z(v)\cap B_{\tilde{s}_{j}}(y_{j}))}{(\tilde{s}_{j})^{d-1}}, (5.78)

where s~j(0,116×8)\tilde{s}_{j}\in(0,\frac{1}{16\times 8}) and yjB116(0)y_{j}\in B_{\frac{1}{16}}(0). By rescaling, we reduce the estimate to uεu_{\varepsilon} and obtain that

Hd1(Z(uε)Bt16(x0))(t16)d1\displaystyle\frac{H^{d-1}(Z(u_{\varepsilon})\cap B_{\frac{t}{16}}(x_{0}))}{(\frac{t}{16})^{d-1}} exp(CN2β34)+14supjHd1(Z(uε)Bts~j(x0+tyj))(ts~j)d1.\displaystyle\leq\exp(CN^{\frac{2}{\beta-\frac{3}{4}}})+\frac{1}{4}\sup_{j}\frac{H^{d-1}(Z(u_{\varepsilon})\cap B_{t\tilde{s}_{j}}(x_{0}+ty_{j}))}{(t\tilde{s}_{j})^{d-1}}. (5.79)

Let τ=t16\tau=\frac{t}{16}, then Cεexp(C(lnN)3)<τ<596C\varepsilon\exp(C(\ln N)^{3})<\tau<\frac{5}{96}. Thus,

Eε(x0,τ)exp(CN2β34)+14supjEε(y^j,s^j).\displaystyle E_{\varepsilon}(x_{0},\tau)\leq\exp(CN^{\frac{2}{\beta-\frac{3}{4}}})+\frac{1}{4}\sup_{j}E_{\varepsilon}(\hat{y}_{j},\hat{s}_{j}). (5.80)

where y^j=x0+tyjBτ(x0)\hat{y}_{j}=x_{0}+ty_{j}\in B_{\tau}(x_{0}), s^j=ts~j(0,τ8)\hat{s}_{j}=t\tilde{s}_{j}\in(0,\frac{\tau}{8}). Note that Bs^j(y^j)B_{\hat{s}_{j}}(\hat{y}_{j}) may not be fully contained in Bτ(x0)B_{\tau}(x_{0}), since y^j\hat{y}_{j} may be the centers of subcubes which intersect the boundary of Bτ(x0)B_{\tau}(x_{0}) (we identify the ball Bτ(x0)B_{\tau}(x_{0}) as a cube when we perform the subcubes decomposition). However, Bs^j(y^j)Bτ+τ8(x0)B_{\hat{s}_{j}}(\hat{y}_{j})\subset B_{\tau+\frac{\tau}{8}}(x_{0}) since s^j(0,τ8)\hat{s}_{j}\in(0,\frac{\tau}{8}). If we iterate (5.80), Bs^j(y^j)B_{\hat{s}_{j}}(\hat{y}_{j}) still stays close to Bτ(x0)B_{\tau}(x_{0}). Actually, Bs^j(yj)Bτ^(x0)B_{\hat{s}_{j}}(y_{j})\subset B_{\hat{\tau}}(x_{0}) for any large jj, where τ^=j=1τ8j1=8τ7\hat{\tau}=\sum^{\infty}_{j=1}\frac{\tau}{8^{j-1}}=\frac{8\tau}{7}.

Now, we iterate (5.80) to obtain the desired estimate. The estimate (5.50) yields the initial step of the iteration,

Eε(0,116)exp(CN2β34)+14sup1jm^Eε(yj,t^j).\displaystyle E_{\varepsilon}(0,\frac{1}{16})\leq\exp(CN^{\frac{2}{\beta-\frac{3}{4}}})+\frac{1}{4}\sup_{1\leq j\leq\hat{m}}E_{\varepsilon}(y_{j},\hat{t}_{j}). (5.81)

Assume that sup1jm^Eε(yj,t^j)\sup_{1\leq j\leq\hat{m}}E_{\varepsilon}(y_{j},\hat{t}_{j}) is achieved at some Eε(yj0,t^j0)E_{\varepsilon}(y_{j_{0}},\hat{t}_{j_{0}}) with |yj0|<116|y_{j_{0}}|<\frac{1}{16} and |t^j0|<1128|\hat{t}_{j_{0}}|<\frac{1}{128}. Let x0=yj0x_{0}=y_{j_{0}} and t^j0=τ\hat{t}_{j_{0}}=\tau. Since s^j<τ8\hat{s}_{j}<\frac{\tau}{8}, we apply (5.80) to Eε(yj0,t^j0)E_{\varepsilon}(y_{j_{0}},\hat{t}_{j_{0}}) to get to the estimates of nodal sets at a smaller scale, that is,

Eε(0,116)(1+14)exp(CN2β34)+14supjEε(y^j,s^j).\displaystyle E_{\varepsilon}(0,\frac{1}{16})\leq(1+\frac{1}{4})\exp(CN^{\frac{2}{\beta-\frac{3}{4}}})+\frac{1}{4}\sup_{j}E_{\varepsilon}(\hat{y}_{j},\hat{s}_{j}). (5.82)

We apply (5.80) repeatedly down to the case rC^εexp(C(lnN)3)r\approx\hat{C}\varepsilon\exp(C(\ln N)^{3}) or the case that Eε(y,r)E_{\varepsilon}(y,r) is empty. Note that Bs^j(y^j)B116+87×128(0)B112(0)B_{\hat{s}_{j}}(\hat{y}_{j})\subset B_{\frac{1}{16}+\frac{8}{7\times 128}}(0)\subset B_{\frac{1}{12}}(0). Thus, we derive that

Eε(0,116)\displaystyle E_{\varepsilon}(0,\frac{1}{16}) i=04iexp(CN2β34)+supyB112(0){Eε(y,r): 0<rC^εexp(C(lnN)3)}\displaystyle\leq\sum^{\infty}_{i=0}4^{-i}\exp(CN^{\frac{2}{\beta-\frac{3}{4}}})+\sup_{y\in B_{\frac{1}{12}}(0)}\big{\{}E_{\varepsilon}(y,r):\ 0<r\leq\hat{C}\varepsilon\exp(C(\ln N)^{3})\big{\}}
exp(CN2β34)+(1+C^exp(C(lnN)3))exp(CN2β34)\displaystyle\leq\exp(CN^{\frac{2}{\beta-\frac{3}{4}}})+(1+\hat{C}\exp(C(\ln N)^{3}))\exp(CN^{\frac{2}{\beta-\frac{3}{4}}})
exp(CN2β34),\displaystyle\leq\exp(CN^{\frac{2}{\beta-\frac{3}{4}}}), (5.83)

where we have used (5.7) in the second inequality. This proves the desired estimate for the case εexp(C(lnN)3)\varepsilon\leq\exp(-C(\ln N)^{3}).

Finally, for the case εexp(C(lnN)3)\varepsilon\geq\exp(-C(\ln N)^{3}), the desired estimate follows directly from (5.7). Since β(34,1)\beta\in(\frac{3}{4},1) is arbitrary, so (1.14) holds for any α>8\alpha>8. This ends the proof of the theorem. ∎

Following the above proof of Theorem 1.2 for d3d\geq 3, we sketch the proof of upper bounds of nodal sets in d=2d=2.

Proof of Theorem 1.2 (d=2d=2).

Since the proof is parallel to d3d\geq 3, we only present the changes for d=2d=2. Thus, we only present the changes for d=2d=2. By the argument in Lemma 5.2 and the doubling inequality (1.10), we can obtain for 0<r<1/30<r<1/3,

Eε(x0,r)C(1+rε)(lnN)2.\displaystyle E_{\varepsilon}(x_{0},r)\leq C\Big{(}1+\frac{r}{\varepsilon}\Big{)}(\ln N)^{2}. (5.84)

On the other hand, the statements of (5.31) and (5.32) still hold for u0u_{0}; while for u0\nabla u_{0}, we have a better bound for δ^\hat{\delta}. Precisely for d=2d=2, [22, Theorem B.1] implies

|Br(𝒞r(u0))Bs|C(N~(u0,B2s))(rs)2|Bs|.|B_{r}(\mathcal{C}_{r}(u_{0}))\cap B_{s}|\leq C^{\big{(}\widetilde{N}(u_{0},B_{2s})\big{)}}\Big{(}\frac{r}{s}\Big{)}^{2}|B_{s}|. (5.85)

By mimicking the argument in part (2) of Lemma 5.6, we can show (5.33) and (5.34) with 0<δ^<eCN^020<\hat{\delta}<e^{-C\hat{N}^{2}_{0}}. Following the proof of Lemma 5.7, we may show that if 0<ε<exp(C(lnN)2)0<\varepsilon<\exp(-C(\ln N)^{2}), then

Eε(0,116)exp(C(lnN)2)+14sup1jm^Eε(yj,t^j),\displaystyle E_{\varepsilon}(0,\frac{1}{16})\leq\exp(C(\ln N)^{2})+\frac{1}{4}\sup_{1\leq j\leq\hat{m}}E_{\varepsilon}(y_{j},\hat{t}_{j}), (5.86)

with yjB116y_{j}\in B_{\frac{1}{16}}, t^j(0,1128)\hat{t}_{j}\in(0,\frac{1}{128}). Observe that the quantitative stratification of critical sets u0\nabla u_{0}, instead of the doubling inequality, plays the dominant role in the estimate (5.86). We iterate (5.86), as in the proof for d3d\geq 3, to get

Eε(0,116)\displaystyle E_{\varepsilon}(0,\frac{1}{16}) i=04iexp(C(lnN)2)+supyB112(0){Eε(y,r): 0<rC^εexp(C(lnN)2)}\displaystyle\leq\sum^{\infty}_{i=0}4^{-i}\exp(C(\ln N)^{2})+\sup_{y\in B_{\frac{1}{12}}(0)}\big{\{}E_{\varepsilon}(y,r):\ 0<r\leq\hat{C}\varepsilon\exp(C(\ln N)^{2})\big{\}}
exp(C(lnN)2)+C(1+C^exp(C(lnN)2))(lnN)2\displaystyle\leq\exp(C(\ln N)^{2})+C(1+\hat{C}\exp(C(\ln N)^{2}))(\ln N)^{2}
exp(C(lnN)2).\displaystyle\leq\exp(C(\ln N)^{2}).

This provides the desired estimate (1.15) for 0<ε<exp(C(lnN)2)0<\varepsilon<\exp(-C(\ln N)^{2}). For εexp(C(lnN)2)\varepsilon\geq\exp(-C(\ln N)^{2}), (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, ApA_{p} 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 L2L^{2} 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