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

Sharp local Bernstein estimates for Laplace eigenfunctions on Riemannian manifolds

Kévin Le Balc’h111Inria, Sorbonne Université, CNRS, Laboratoire Jacques-Louis Lions, Paris, France. kevin.le-balc-h@inria.fr  –  The author is partially supported by the Project TRECOS ANR-20-CE40-0009 funded by the ANR (2021–2024) and by the Project CURED from Sorbonne Université (2024).
Abstract

In this paper we focus on local growth properties of Laplace eigenfunctions on a compact Riemannian manifold. The principal theme is that a Laplace eigenfunction behaves locally as a polynomial function of degree proportional to the square root of the eigenvalue. In this direction, we notably prove sharp local LL^{\infty}-Bernstein estimates, conjectured by Donnelly and Fefferman in 1990. As a byproduct we also obtain analogous inequalities for AA-harmonic functions where the square root of the eigenvalue is replaced by the doubling index of the solution. Our proof is based on a refinement of the original proof of L2L^{2}-Bernstein estimates by Donnelly and Fefferman, based on L2L^{2}-Carleman estimates, with a suitable bootstrap argument involving elliptic regularity estimates and Gagliardo–Nirenberg interpolation inequalities.

This actual version contains a bad mistake, located in Section 2.3 that invalidates the proofs of the main results.

Keywords: Bernstein estimates, Laplace eigenfunctions.

Classifications: 58J50, 35J05, 35P20, 35R01.

1 Introduction

Let MM be a CC^{\infty}-smooth, compact, connected, Riemannian manifold of dimension d2d\geqslant 2, without boundary, equipped with a Riemannian metric gg. In this article we are interested in growth properties of Laplace eigenfunctions φλC(M)\varphi_{\lambda}\in C^{\infty}(M) associated to the eigenvalue λ0\lambda\geqslant 0,

Δgφλ=λφλinM,-\Delta_{g}\varphi_{\lambda}=\lambda\varphi_{\lambda}\ \text{in}\ M, (1.1)

where Δg=divgg\Delta_{g}=\mathrm{div}_{g}\circ\nabla_{g} is the Laplace-Beltrami operator.

One may distinguish between global growth and local growth.

The famous classical Bernstein’s estimate on trigonometric polynomials is typically a global growth estimate on linear combination of Laplace eigenfunctions on the one-dimensional torus. Let n0n\geqslant 0 and T=k=nnakeikxT=\sum_{k=-n}^{n}a_{k}e^{ikx} for x(0,2π)x\in(0,2\pi), then

supx(0,2π)|T(x)|nsupx(0,2π)|T(x)|.\sup_{x\in(0,2\pi)}|T^{\prime}(x)|\leqslant n\sup_{x\in(0,2\pi)}|T(x)|. (1.2)

For a survey around (1.2), one can read [QZ19]. Global LL^{\infty}-Bernstein estimates also hold for a single Laplace eigenfunction on MM, i.e. there exist C>0C>0 depending only on MM such that for every Laplace eigenfunction φλ\varphi_{\lambda},

supM|φλ|CλsupM|φλ|.\sup_{M}|\nabla\varphi_{\lambda}|\leqslant C\sqrt{\lambda}\sup_{M}|\varphi_{\lambda}|. (1.3)

This estimate (1.3) is actually a consequence of standard elliptic estimates for harmonic functions, see [OCP13, Corollary 3.3] for a proof. Note that one can actually extend global Bernstein estimates for a single Laplace eigenfunction that is (1.3) to a linear combination of Laplace eigenfunctions, i.e. there exist C>0C>0 depending only on MM such that for ΦΛ=λkΛakφλk\Phi_{\Lambda}=\sum_{\lambda_{k}\leqslant\Lambda}a_{k}\varphi_{\lambda_{k}}, then

supM|ΦΛ|CΛsupM|ΦΛ|,\sup_{M}|\nabla\Phi_{\Lambda}|\leqslant C\sqrt{\Lambda}\sup_{M}|\Phi_{\Lambda}|, (1.4)

see for instance [FM10, Theorem 2.1] and [IO22, Theorem 1.2]. In this latter case, the proofs are considerably more involved.

Concerning local growth, from the breakthrough work of Donnelly, Fefferman [DF88], we know that φλ\varphi_{\lambda} also shares local growth properties with a polynomial function of degree proportional to λ\sqrt{\lambda}. One of the most celebrated result is the following bound on the doubling index of Laplace eigenfunctions, see [DF88, Theorem 4.2].

\bullet There exist r0,C>0r_{0},C>0 depending only on MM, such that for every Laplace eigenfunction φλC(M)\varphi_{\lambda}\in C^{\infty}(M), i.e. satisfying (1.1), for every xMx\in M, r(0,r0)r\in(0,r_{0}),

supBg(x,2r)|φλ|eCλsupBg(x,r)|φλ|.\sup_{B_{g}\left(x,2r\right)}|\varphi_{\lambda}|\leqslant e^{C\sqrt{\lambda}}\sup_{B_{g}\left(x,r\right)}|\varphi_{\lambda}|. (1.5)

Note that (1.5) is in perfect agreement to the previous heuristics because

supt(2r,2r)tλ=2λsupt(r,r)tλ.\sup_{t\in(-2r,2r)}t^{\sqrt{\lambda}}=2^{\sqrt{\lambda}}\sup_{t\in(-r,r)}t^{\sqrt{\lambda}}.

For fC(M)f\in C^{\infty}(M), xMx\in M, r>0r>0, the number

Nf(Bg(x,r)):=log(supBg(x,2r)|f|supBg(x,r)|f|),N_{f}(B_{g}(x,r)):=\log\left(\frac{\sup_{B_{g}\left(x,2r\right)}|f|}{\sup_{B_{g}\left(x,r\right)}|f|}\right),

is usually called the doubling index of ff in the ball Bg(x,r)B_{g}(x,r). Note that for xMx\in M,

limr0Nf(Bg(x,r))=vanishing order of fatx,\lim_{r\to 0}N_{f}(B_{g}(x,r))=\text{vanishing order of }f\ \text{at}\ x,

where the vanishing order of ff at xx is the smallest integer kk such that the derivatives of ff of order smaller than kk vanish while there is some non-zero derivative of order kk. As a consequence, the doubling index estimate (1.5) tells us that the vanishing order of φλ\varphi_{\lambda} is bounded by CλC\sqrt{\lambda}. This last result is sharp if we do not make extra assumptions on the Riemannian manifold because the vanishing order of spherical harmonics is comparable to λ\sqrt{\lambda}.

In [DF90a], the authors pursue the analogy between Laplace eigenfunctions and polynomial functions. They obtain the following local L2L^{2}-Bernstein estimates, see [DF90a, Theorem 1].

\bullet There exist r0,C>0r_{0},C>0 depending only on MM, such that for every Laplace eigenfunction φλC(M)\varphi_{\lambda}\in C^{\infty}(M), i.e. satisfying (1.1), for every xMx\in M, r(0,r0)r\in(0,r_{0}), λ1\lambda\geqslant 1,

φλL2(Bg(x,r(1+1λ)))CφλL2(Bg(x,r)),\|\varphi_{\lambda}\|_{L^{2}\left(B_{g}\left(x,r\left(1+\frac{1}{\sqrt{\lambda}}\right)\right)\right)}\leqslant C\|\varphi_{\lambda}\|_{L^{2}\left(B_{g}\left(x,r\right)\right)}, (1.6)

and

φλL2(Bg(x,r))CλrφλL2(Bg(x,r)).\|\nabla\varphi_{\lambda}\|_{L^{2}\left(B_{g}\left(x,r\right)\right)}\leqslant C\frac{\sqrt{\lambda}}{r}\|\varphi_{\lambda}\|_{L^{2}\left(B_{g}\left(x,r\right)\right)}. (1.7)

The inequality (1.7) is called a local L2L^{2}-Bernstein estimate due to the common feature with the standard global LL^{\infty}-Bernstein estimate (1.3).

\bullet In [DF90a], starting from (1.6) and an elementary elliptic regularity result, the authors also obtain the following local LL^{\infty}-Bernstein inequality

supBg(x,r)|φλ|Cλd+24rsupBg(x,r)|φλ|.\sup_{B_{g}\left(x,r\right)}|\nabla\varphi_{\lambda}|\leqslant C\frac{\lambda^{\frac{d+2}{4}}}{r}\sup_{B_{g}\left(x,r\right)}|\varphi_{\lambda}|. (1.8)

The authors also formulate the following conjecture.

Conjecture 1.1 ([DF90a]).

The Bernstein inequality (1.8) still holds replacing d+24\frac{d+2}{4} by 12\frac{1}{2}.

1.1 is again motivated by the heuristics that φλ\varphi_{\lambda} behaves as tλt^{\sqrt{\lambda}} because

supt(r,r)(ddttλ)=supt(r,r)λtλ1=λrsupt(r,r)tλ.\sup_{t\in(-r,r)}\left(\frac{d}{dt}t^{\sqrt{\lambda}}\right)=\sup_{t\in(-r,r)}\sqrt{\lambda}t^{\sqrt{\lambda}-1}=\frac{\sqrt{\lambda}}{r}\sup_{t\in(-r,r)}t^{\sqrt{\lambda}}.

\bullet In [Don95], Dong refines (1.8) for surfaces, i.e. when d=2d=2, using powerful geometric idea from [Don92], by obtaining

supBg(x,r)|φλ|Cmax(λr,λ34)supBg(x,r)|φλ|(d=2).\sup_{B_{g}\left(x,r\right)}|\nabla\varphi_{\lambda}|\leqslant C\max\left(\frac{\sqrt{\lambda}}{r},\lambda^{\frac{3}{4}}\right)\sup_{B_{g}\left(x,r\right)}|\varphi_{\lambda}|\qquad(d=2). (1.9)

\bullet In the very recent preprint [DM23], Decio and Malinnikova prove the following estimates

supBg(x,r)|φλ|Cmax(λr,λlog(λ))supBg(x,r)|φλ|(d=2),\sup_{B_{g}\left(x,r\right)}|\nabla\varphi_{\lambda}|\leqslant C\max\left(\frac{\sqrt{\lambda}}{r},\sqrt{\lambda}\log(\lambda)\right)\sup_{B_{g}\left(x,r\right)}|\varphi_{\lambda}|\qquad(d=2), (1.10)

and for δ>0\delta>0 arbitrary,

supBg(x,r)|φλ|C(δ)max(λrlog2+δ(λ),λlog2+δ(λ))supBg(x,r)|φλ|(d2).\sup_{B_{g}\left(x,r\right)}|\nabla\varphi_{\lambda}|\leqslant C(\delta)\max\left(\frac{\sqrt{\lambda}}{r}\log^{2+\delta}(\lambda),\lambda\log^{2+\delta}(\lambda)\right)\sup_{B_{g}\left(x,r\right)}|\varphi_{\lambda}|\qquad(d\geqslant 2). (1.11)

In particular, (1.10) is a strong refinement of (1.9) and gives 1.1 for surfaces up to a logarithm loss and (1.11) gives 1.1 up to a logarithm loss at the wavelength scale, i.e. r(0,r0λ1)r\in\left(0,r_{0}\sqrt{\lambda}^{-1}\right) while (1.11) resembles more to the Markov’s inequality for polynomials at larger scales in any dimension. Recall that for an algebraic polynomial of degree nn, the Markov inequality holds

supx(1,+1)|Pn(x)|n2supx(1,+1)|Pn(x)|.\sup_{x\in(-1,+1)}|P_{n}^{\prime}(x)|\leqslant n^{2}\sup_{x\in(-1,+1)}|P_{n}^{\prime}(x)|. (1.12)

Note that (1.12) is sharp because Chebychev polynomials are extremizers of this inequality.

The first main result of this paper is the establishment of 1.1 on LL^{\infty}-Bernstein estimates for Laplace eigenfunctions.

Theorem 1.2.

There exist r0,C>0r_{0},C>0 depending only on MM, such that for every λ1\lambda\geqslant 1, for every Laplace eigenfunction φλC(M)\varphi_{\lambda}\in C^{\infty}(M), i.e. satisfying (1.1), for every xMx\in M, r(0,r0)r\in(0,r_{0}),

supBg(x,r(1+1λ))|φλ|CsupBg(x,r)|φλ|,\sup_{B_{g}\left(x,r\left(1+\frac{1}{\sqrt{\lambda}}\right)\right)}|\varphi_{\lambda}|\leqslant C\sup_{B_{g}\left(x,r\right)}|\varphi_{\lambda}|, (1.13)

and

supBg(x,r)|φλ|CλrsupBg(x,r)|φλ|.\sup_{B_{g}\left(x,r\right)}|\nabla\varphi_{\lambda}|\leqslant C\frac{\sqrt{\lambda}}{r}\sup_{B_{g}\left(x,r\right)}|\varphi_{\lambda}|. (1.14)

The sharp doubling index estimate (1.5) is an easy consequence of (1.13), so we immediately deduce the sharpness of 1.2.

We actually prove a similar result for solutions to harmonic functions in the Euclidean space where the role of the square root of the eigenvalue is played by the doubling index, that serves as a local degree of the solution. More precisely, we look at LL^{\infty}-Bernstein estimates for AA-harmonic functions with a bounded doubling index.

The matrix A=(aij(x))1i,jdA=(a^{ij}(x))_{1\leqslant i,j\leqslant d} is supposed to be symmetric, uniformly elliptic, with Lipschitz entries

Λ11|ξ|2A(x)ξ,ξΛ1|ξ|2,|aij(x)aij(y)|Λ2|xy|,x,yB2,ξd,\Lambda_{1}^{-1}|\xi|^{2}\leqslant\langle A(x)\xi,\xi\rangle\leqslant\Lambda_{1}|\xi|^{2},\quad|a^{ij}(x)-a^{ij}(y)|\leqslant\Lambda_{2}|x-y|,\qquad x,y\in B_{2},\ \xi\in\mathbb{R}^{d}, (1.15)

for some Λ1,Λ2>0\Lambda_{1},\Lambda_{2}>0.

The second main result of this paper is the following one.

Theorem 1.3.

There exist r0,C>0r_{0},C>0 depending on AA such that for every uHloc1(B2)L(B2)u\in H_{\text{loc}}^{1}(B_{2})\cap L^{\infty}(B_{2}) satisfying

div(A(x)u)=0inB2,-\mathrm{div}(A(x)\nabla u)=0\ \text{in}\ B_{2}, (1.16)

and

Nu(B(0,1)):=log(supB(0,2)|u|supB(0,1)|u|)N,N1,N_{u}(B(0,1)):=\log\left(\frac{\sup_{B\left(0,2\right)}|u|}{\sup_{B\left(0,1\right)}|u|}\right)\leqslant N,\qquad N\geqslant 1, (1.17)

then for every r(0,r0)r\in(0,r_{0}),

supB(0,r(1+1N))|u|CsupB(0,r)|u|,\sup_{B\left(0,r\left(1+\frac{1}{N}\right)\right)}|u|\leqslant C\sup_{B\left(0,r\right)}|u|, (1.18)

and

supB(0,r)|u|CNrsupB(0,r)|u|.\sup_{B\left(0,r\right)}|\nabla u|\leqslant C\frac{N}{r}\sup_{B\left(0,r\right)}|u|. (1.19)

1.3 has to be compared to [DM23, Theorem 2] where the authors obtain a similar result with stronger regularity assumptions on the matrix AA and stronger smallness assumptions on the radius rr, that has to be small in function of the doubling index.

On the one hand, 1.2 and 1.3 are related by the standard lifting trick that allows to pass from Laplace eigenfunctions to harmonic functions. If φλ\varphi_{\lambda} satisfies (1.1) then the function

u(x,t)=φλ(x)eλt(x,t)M×,u(x,t)=\varphi_{\lambda}(x)e^{\sqrt{\lambda}t}\qquad(x,t)\in M\times\mathbb{R}, (1.20)

is harmonic on the product manifold M×M\times\mathbb{R} and by using (1.5) its doubling index is bounded by CλC\sqrt{\lambda}. This standard trick was first observed by [Lin91] in the study of the nodal volume for Laplace eigenfunctions on compact Riemannian manifolds, it has other applications like for instance the obtaining of the bound on the doubling index of Laplace eigenfunctions in (1.5), see [LM20, Proposition 2.4.1]. On the other hand, it is worth mentioning that 1.2 is not a direct consequence of 1.3, as it is the case in [DM23] where the authors deduce LL^{\infty}-Bernstein estimates for Laplace eigenfunctions from LL^{\infty}-Bernstein estimates for AA-harmonic functions because they are working at the wavelength scale, i.e. rCλ1r\leqslant C\sqrt{\lambda}^{-1}. In our case, the same phenomenon appears, we can only deduce 1.2 from 1.3 for rCλ1r\leqslant C\sqrt{\lambda}^{-1}. This is why we will actually prove 1.2 in an independent way by following the same strategy of the proof of 1.3 even if the some new technical difficulties appear.

New ingredient. The proof of 1.3, and therefore the one of 1.2, takes its source inside the proof of L2L^{2}-Bernstein estimates for Laplace eigenfunctions, i.e. (1.6) and (1.7) from [DF90a], that is based on an adequate L2L^{2}-Carleman estimate. Note that if we start directly from (1.6) and (1.7) then use elliptic regularity estimates and Sobolev embeddings, we cannot obtain better than the weak LL^{\infty}-Bernstein estimate from [DF90a] i.e. (1.8). The new ingredient consists in implementing a suitable bootstrap argument involving scaled versions of elliptic regularity estimates and Gagliardo-Nirenberg interpolations inequalities in the Carleman’s strategy.

Organization of the paper. In Section 2, we prove the main results i.e. 1.2 and 1.3. In Section 3, we state generalizations of 1.2, 1.3 and mention some related open problems.

2 Proof of the growth estimates

The goal of this part is to prove 1.2 and 1.3.

The first four parts are dedicated to the proof of 1.3 while the last part is devoted to the proof of 1.2. Recall that the proof of the growth estimates for Laplace eigenfunctions on Riemannian manifolds stated in 1.2 is a small adaptation of the one of the growth estimates for AA-harmonic functions on the Euclidean space stated in 1.3, this is why we will only insist on the new difficulties that appear in the fifth part. The first part consists in stating L2L^{2}-Carleman estimates for the operator div(A)\mathrm{div}(A\nabla\cdot), the second part proves vanishing order estimates for AA-harmonic functions with bounded doubling index, the third part establishes scaled versions of elliptic regularity estimates and Gagliardo-Nirenberg interpolations inequalities, the fourth part consists in applying the first three parts together with a suitable bootstrap argument to deduce the estimate (1.18) and then (1.19).

In the next four parts, the positive constants C>0C>0, c>0c>0 depend on AA and dd while in the last part, the positive constants C>0C>0, c>0c>0 are allowed to depend on MM, gg and dd. To insist on the dependence of a positive constant CC in function of some parameter ss, we will sometimes use the notation C=C(s)C=C(s). Moreover, the constants can vary from one line to another without explicitly mentioning it.

2.1 L2L^{2}-Carleman estimates

The goal of this part is to state L2L^{2}-Carleman estimates.

To simplify the notations in the next, we set

div(A(x)f)=ΔAf.\mathrm{div}(A(x)\nabla f)=\Delta_{A}f.

We also introduce the spherical coordinates of a point x=(x1,,xd)d{0}x=(x_{1},\dots,x_{d})\in\mathbb{R}^{d}\setminus\{0\} by

(ρ,θ)=(ρ,θ1,,θd1)(0,+)×𝕊d1,(\rho,\theta)=(\rho,\theta_{1},\dots,\theta_{d-1})\in(0,+\infty)\times\mathbb{S}^{d-1},

with

x1=ρcos(θ1),x2=ρsin(θ1)cos(θ2),x3=ρsin(θ1)sin(θ2)cos(θ3),,x_{1}=\rho\cos(\theta_{1}),\ x_{2}=\rho\sin(\theta_{1})\cos(\theta_{2}),\ x_{3}=\rho\sin(\theta_{1})\sin(\theta_{2})\cos(\theta_{3}),\ \dots\ ,
xd1=ρsin(θ1)sin(θd2)cos(θd1),xd=ρsin(θ1)sin(θd2)sin(θd1),x_{d-1}=\rho\sin(\theta_{1})\cdots\sin(\theta_{d-2})\cos(\theta_{d-1}),\ x_{d}=\rho\sin(\theta_{1})\cdots\sin(\theta_{d-2})\sin(\theta_{d-1}),

for

θ1,,θd2[0,π)andθd1[0,2π).\theta_{1},\dots,\theta_{d-2}\in[0,\pi)\ \text{and}\ \theta_{d-1}\in[0,2\pi).

The orthonormal spherical basis will be denoted by (eρ,eθ)(e_{\rho},e_{\theta}) and for a given vector field ξ\xi in d\mathbb{R}^{d}, its components with respect to this basis will be denoted respectively by (ξρ,ξθ)(\xi_{\rho},\xi_{\theta}).

First, we have the following standard L2L^{2}-Carleman estimate.

Lemma 2.1.

There exists a positive constant C=C(A)>0C=C(A)>0 such that for every αC\alpha\geqslant C, fCc(B2{0})f\in C_{c}^{\infty}(B_{2}\setminus\{0\}), the following estimate holds

α3B2ρ12α|f|2𝑑x+αB2ρ12α|f|2𝑑xCB2ρ22α|ΔAf|2𝑑x.\alpha^{3}\int_{B_{2}}\rho^{-1-2\alpha}|f|^{2}dx+\alpha\int_{B_{2}}\rho^{1-2\alpha}|\nabla f|^{2}dx\leqslant C\int_{B_{2}}\rho^{2-2\alpha}|\Delta_{A}f|^{2}dx. (2.1)

Note that 2.1 is a direct application of [EV03, Theorem 2], stated in the parabolic case.

The next result tells us how the Carleman estimate (2.1) from 2.1 translates when the function vanishes in a small ball centered at 0.

Lemma 2.2.

There exists a positive constant C=C(A)>0C=C(A)>0 and c=c(A)>0c=c(A)>0 such that for every r(0,c)r\in(0,c), αC\alpha\geqslant C, for all fCc(B2Br)f\in C_{c}^{\infty}(B_{2}\setminus B_{r}), the following estimate holds

α3B2ρ12α|f|2𝑑x+αB2ρ12α|f|2𝑑x+α4r2B(0,r(1+2α1))ρ2α|f|2𝑑xCB2ρ22α|ΔAf|2𝑑x.\alpha^{3}\int_{B_{2}}\rho^{-1-2\alpha}|f|^{2}dx+\alpha\int_{B_{2}}\rho^{1-2\alpha}|\nabla f|^{2}dx+\frac{\alpha^{4}}{r^{2}}\int_{B(0,r(1+2\alpha^{-1}))}\rho^{-2\alpha}|f|^{2}dx\\ \leqslant C\int_{B_{2}}\rho^{2-2\alpha}|\Delta_{A}f|^{2}dx. (2.2)

2.2 can be obtained by adapting the arguments of the proof of the Carleman estimate in [DF90a, Lemma A], stated for the operator Δguλu-\Delta_{g}u-\lambda u, in the Riemannian case.

For the sake of completeness and because we will need some small modifications of these Carleman estimates in the sequel of the paper, we decide to give a complete self-contained proof of 2.1 and 2.2 in Appendix A.

2.2 Vanishing order estimate

Before proving 1.3, one needs to prove a result on the vanishing order estimate for AA-harmonic functions with bounded doubling index.

First, we have the following modification of the Carleman estimate (2.1) from 2.1.

Lemma 2.3.

There exists a positive constant C=C(A)>0C=C(A)>0 such that for every x0B1/4x_{0}\in B_{1/4}, for every αC\alpha\geqslant C, fCc(B2{x0})f\in C_{c}^{\infty}(B_{2}\setminus\{x_{0}\}), the following estimate holds

α3B2|xx0|12α|f|2𝑑x+αB2|xx0|12α|f|2𝑑xCB2|xx0|22α|ΔAf|2𝑑x.\alpha^{3}\int_{B_{2}}|x-x_{0}|^{-1-2\alpha}|f|^{2}dx+\alpha\int_{B_{2}}|x-x_{0}|^{1-2\alpha}|\nabla f|^{2}dx\leqslant C\int_{B_{2}}|x-x_{0}|^{2-2\alpha}|\Delta_{A}f|^{2}dx. (2.3)

The proof of 2.3 is postponed in Appendix A, it is an adaptation of 2.1.

As a consequence of 2.3, we deduce the following result.

Lemma 2.4.

There exists C=C(A)>0C=C(A)>0 such that for every uHloc1(B2)L(B2)u\in H_{\text{loc}}^{1}(B_{2})\cap L^{\infty}(B_{2}) satisfying (1.16) and (1.17), the following estimate holds

supB(x,1/4)|u|exp(CN)supB2|u|xB1/2.\sup_{B(x,1/4)}|u|\geqslant\exp(-CN)\sup_{B_{2}}|u|\qquad\forall x\in B_{1/2}. (2.4)
Proof.

Let us fix x0B(0,1/2)x_{0}\in B(0,1/2).

Let xmaxB(0,1)x_{\max}\in B(0,1) be such that

|u(xmax)|=supB(0,1)|u|.|u(x_{\max})|=\sup_{B(0,1)}|u|.

We distinguish two cases.

First case: xmaxB(x0,1/4)x_{\max}\in B(x_{0},1/4). Then we have by (1.17),

supB(x0,1/4)|u|=supB(0,1)|u|exp(N)supB2|u|,\sup_{B(x_{0},1/4)}|u|=\sup_{B(0,1)}|u|\geqslant\exp(-N)\sup_{B_{2}}|u|,

so (2.4) holds.

Second case: xmaxB(x0,1/4)x_{\max}\notin B(x_{0},1/4). Let χCc(B2;[0,1])\chi\in C_{c}^{\infty}(B_{2};[0,1]) be a cut-off function such that

χ\displaystyle\chi =0inB(x0,1/16),\displaystyle=0\ \text{in}\ B(x_{0},1/16),
χ\displaystyle\chi =1inB(0,15/8)B(x0,1/8),\displaystyle=1\ \text{in}\ B(0,15/8)\setminus B(x_{0},1/8),
χ\displaystyle\chi =0inB(0,2)B(0,31/16).\displaystyle=0\ \text{in}\ B(0,2)\setminus B(0,31/16).

In particular χ1\chi\equiv 1 in B(xmax,1/8)B(x_{\max},1/8).

By local elliptic regularity, we have that uHloc2(B2)u\in H_{\text{loc}}^{2}(B_{2}) so by a straightforward density argument one can apply the modified version of the Carleman estimate (2.3) of 2.3 to f:=χuf:=\chi u. By denoting w(x)=|xx0|w(x)=|x-x_{0}|, we then obtain

α3B2w12α|f|2𝑑xCB2w22α|ΔAf|2𝑑x.\alpha^{3}\int_{B_{2}}w^{-1-2\alpha}|f|^{2}dx\leqslant C\int_{B_{2}}w^{2-2\alpha}|\Delta_{A}f|^{2}dx.

By using the equation (1.16), we deduce that

α3B2w12αχ2|u|2𝑑xCB2w22α(|χ|2|u|2+[|χ|2+|D2χ|2]|u|2)𝑑x.\alpha^{3}\int_{B_{2}}w^{-1-2\alpha}\chi^{2}|u|^{2}dx\leqslant C\int_{B_{2}}w^{2-2\alpha}(|\nabla\chi|^{2}|\nabla u|^{2}+[|\nabla\chi|^{2}+|D^{2}\chi|^{2}]|u|^{2})dx.

By using the definition of χ\chi and local elliptic regularity estimates, it is straightforward to get that there exist universal positive constants C3>C2>C1>0C_{3}>C_{2}>C_{1}>0 and a positive constant C=C(A)>0C=C(A)>0 such that

exp(C2α)|u(xmax)|2Cexp(C3α)(supB2|u|)2+Cexp(C1α)(supB(x0,1/8)|u|)2.\exp(-C_{2}\alpha)|u(x_{\max})|^{2}\leqslant C\exp(-C_{3}\alpha)(\sup_{B_{2}}|u|)^{2}+C\exp(-C_{1}\alpha)(\sup_{B(x_{0},1/8)}|u|)^{2}.

By using the definition of xmaxx_{\max} and the doubling index estimate (1.17), we therefore obtain that

exp(C2α)(supB1|u|)2Cexp(C3α)exp(2N)(supB1|u|)2+Cexp(C1α)(supB(x0,1/8)|u|)2.\exp(-C_{2}\alpha)(\sup_{B_{1}}|u|)^{2}\leqslant C\exp(-C_{3}\alpha)\exp(2N)(\sup_{B_{1}}|u|)^{2}+C\exp(-C_{1}\alpha)(\sup_{B(x_{0},1/8)}|u|)^{2}.

Now the punchline, by taking αC(C2,C3,A)N\alpha\geqslant C(C_{2},C_{3},A)N, the first right hand side term can be hidden in the left hand side term to deduce that

supB1|u|exp(CN)supB(x0,1/8)|u|.\sup_{B_{1}}|u|\leqslant\exp(CN)\sup_{B(x_{0},1/8)}|u|.

This concludes the proof of (2.4) recalling again (1.17). ∎

2.3 Elliptic regularity estimates and interpolation inequalities

There is a mistake in this part, because the scaling in α\alpha parameter is wrong, so the proof of the next lemma does not work. This point is crucial for the next.

The goal of this part is to establish scaled versions of local elliptic regularity estimates for the operator div(A)\mathrm{div}(A\nabla\cdot) and Gagliardo-Nirenberg interpolations inequalities.

We have the following relations for the divergence and gradient operators in spherical coordinates

div(ξ)=1ρn1ρ(ρn1ξρ)+1ρdivθ(ξθ),\mathrm{div}(\xi)=\frac{1}{\rho^{n-1}}\partial_{\rho}(\rho^{n-1}\xi_{\rho})+\frac{1}{\rho}\mathrm{div}_{\theta}(\xi_{\theta}), (2.5)

and

v=(ρv)eρ+1ρθv,\nabla v=(\partial_{\rho}v)e_{\rho}+\frac{1}{\rho}\nabla_{\theta}v, (2.6)

In the sequel, we need the following notation for the annulus

𝒞(r1,r2)=B(0,r2)B(0,r1)0<r1<r2.\mathcal{C}(r_{1},r_{2})=B(0,r_{2})\setminus B(0,r_{1})\qquad\forall 0<r_{1}<r_{2}. (2.7)
Lemma 2.5.

Let C3>C2>C1>C0>0C_{3}>C_{2}>C_{1}>C_{0}>0, r(0,c)r\in(0,c) and αC\alpha\geqslant C.

  1. 1.

    Let p(1,+)p\in(1,+\infty), then there exists a positive constant C=C(A,p,C0,C1,C2,C3)>0C=C(A,p,C_{0},C_{1},C_{2},C_{3})>0 such that for every uWloc2,p(B2)u\in W_{\text{loc}}^{2,p}(B_{2}), the following estimate holds

    r1αuLp(𝒞(r(1+C1α1),r(1+C2α1)))C(ΔAuLp(𝒞(r(1+C0α1),r(1+C3α1)))+r2α2uLp(𝒞(r(1+C0α1),r(1+C3α1)))).r^{-1}\alpha\|\nabla u\|_{L^{p}(\mathcal{C}(r(1+C_{1}\alpha^{-1}),r(1+C_{2}\alpha^{-1})))}\\ \leqslant C\Big{(}\|\Delta_{A}u\|_{L^{p}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}\\ +r^{-2}\alpha^{2}\|u\|_{L^{p}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}\Big{)}. (2.8)
  2. 2.

    Let p[2,+)p\in[2,+\infty), q(2,+]q\in(2,+\infty] be such that

    1q=12(1p1d)+12p.\frac{1}{q}=\frac{1}{2}\left(\frac{1}{p}-\frac{1}{d}\right)+\frac{1}{2p}. (2.9)

    Then there exists a positive constant C=C(A,p,C0,C3)>0C=C(A,p,C_{0},C_{3})>0 such that for every uWloc1,p(B2)u\in W_{\text{loc}}^{1,p}(B_{2}) then uLlocq(B2)u\in L_{loc}^{q}(B_{2}) and

    uLq(𝒞(r(1+C0α1),r(1+C3α1)))Cα1d2duLp(𝒞(r(1+C0α1),r(1+C3α1)))1/2uLp(𝒞(r(1+C0α1),r(1+C3α1)))1/2+C(rdα1)1/2duLp(𝒞(r(1+C0α1),r(1+C3α1))).\|u\|_{L^{q}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}\\ \leqslant C\alpha^{\frac{1-d}{2d}}\|\nabla u\|_{L^{p}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}^{1/2}\|u\|_{L^{p}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}^{1/2}\\ +C(r^{d}\alpha^{-1})^{-1/2d}\|u\|_{L^{p}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}. (2.10)
Proof.

We prove the first point. We set

u^(ρ,θ)=u(rρ,θ),A^(ρ,θ)=A(rρ,θ)(ρ,θ)(1+C0α1,1+C3α1)×𝕊d1.\hat{u}(\rho,\theta)=u(r\rho,\theta),\ \hat{A}(\rho,\theta)=A(r\rho,\theta)\qquad(\rho,\theta)\in(1+C_{0}\alpha^{-1},1+C_{3}\alpha^{-1})\times\mathbb{S}^{d-1}. (2.11)

Then, we find from (2.6)

u^(ρ,θ)=rρu(rρ,θ)eρ+1ρθu(rρ,θ)=r(ρu(rρ,θ)erρ+1rρθu(rρ,θ))=ru(rρ,θ),\nabla\hat{u}(\rho,\theta)=r\partial_{\rho}u(r\rho,\theta)e_{\rho}+\frac{1}{\rho}\nabla_{\theta}u(r\rho,\theta)=r(\partial_{\rho}u(r\rho,\theta)e_{r\rho}+\frac{1}{r\rho}\nabla_{\theta}u(r\rho,\theta))=r\nabla u(r\rho,\theta),

and then from (2.5)

div(A^u^)=1ρn1ρ(ρn1(A^u^)ρ)+1ρdivθ((A^u^)θ)=r2(1(rρ)n1ρ((rρ)n1(Au)ρ)+1rρdivθ((Au)θ))=r2div(A(rρ,θ)u(rρ,θ)).\mathrm{div}(\hat{A}\nabla\hat{u})=\frac{1}{\rho^{n-1}}\partial_{\rho}(\rho^{n-1}(\hat{A}\nabla\hat{u})_{\rho})+\frac{1}{\rho}\mathrm{div}_{\theta}((\hat{A}\nabla\hat{u})_{\theta})\\ =r^{2}(\frac{1}{(r\rho)^{n-1}}\partial_{\rho}((r\rho)^{n-1}(A\nabla u)_{\rho})+\frac{1}{r\rho}\mathrm{div}_{\theta}((A\nabla u)_{\theta}))=r^{2}\mathrm{div}(A(r\rho,\theta)\nabla u(r\rho,\theta)).

Then we set

u~(ρ,θ)=u^(1+α1ρ,θ),A~(ρ,θ)=A^(1+α1ρ,θ)(ρ,θ)(C0,C3)×𝕊d1.\tilde{u}(\rho,\theta)=\hat{u}(1+\alpha^{-1}\rho,\theta),\ \tilde{A}(\rho,\theta)=\hat{A}(1+\alpha^{-1}\rho,\theta)\qquad(\rho,\theta)\in(C_{0},C_{3})\times\mathbb{S}^{d-1}. (2.12)

In the same way, we find

u~(ρ,θ)=α1u^(1+α1ρ,θ),{\color[rgb]{1,0,0}\nabla\tilde{u}(\rho,\theta)=\alpha^{-1}\nabla\hat{u}(1+\alpha^{-1}\rho,\theta),}

and then

div(A~u~)(ρ,θ)=α2div(A^(1+α1ρ,θ)u^(1+α1ρ,θ)).{\color[rgb]{1,0,0}\mathrm{div}(\tilde{A}\nabla\tilde{u})(\rho,\theta)=\alpha^{-2}\mathrm{div}(\hat{A}(1+\alpha^{-1}\rho,\theta)\nabla\hat{u}(1+\alpha^{-1}\rho,\theta)).}

Therefore, we find successively by changes of variable

u~Lp(𝒞(C0,C3))(rdα)1/puLp(𝒞(r(1+C0α1),r(1+C3α1))),\|\tilde{u}\|_{L^{p}(\mathcal{C}(C_{0},C_{3}))}\approx(r^{-d}\alpha)^{1/p}\|u\|_{L^{p}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))},
u~Lp(𝒞(C1,C2))(rdα)1/p(rα1)uLp(𝒞(r(1+C1α1),r(1+C2α1))),\|\nabla\tilde{u}\|_{L^{p}(\mathcal{C}(C_{1},C_{2}))}\approx(r^{-d}\alpha)^{1/p}(r\alpha^{-1})\|\nabla u\|_{L^{p}(\mathcal{C}(r(1+C_{1}\alpha^{-1}),r(1+C_{2}\alpha^{-1})))},
ΔA~u~Lp(𝒞(C0,C3))(rdα)1/p(rα1)2ΔAuLp(𝒞(r(1+C0α1),r(1+C3α1))).\|\Delta_{\tilde{A}}\tilde{u}\|_{L^{p}(\mathcal{C}(C_{0},C_{3}))}\approx(r^{-d}\alpha)^{1/p}(r\alpha^{-1})^{2}\|\Delta_{A}u\|_{L^{p}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}.

In the three above estimates and in the sequel of the proof, for two positive quantities ff and gg, we denote by fgf\approx g the fact that there exist two positive constants c,C>0c,C>0 depending on A,p,C0,C1,C2,C3A,p,C_{0},C_{1},C_{2},C_{3} such that cfgCfcf\leqslant g\leqslant Cf.

By applying local elliptic regularity estimates, in particular [GT01, Theorem 9.11] and by deducing from (1.15),

Λ11|ξ|2A~(x)ξ,ξΛ1|ξ|2,|A~ij(x)A~ij(y)|CΛ2|xy|.\Lambda_{1}^{-1}|\xi|^{2}\leqslant\langle\tilde{A}(x)\xi,\xi\rangle\leqslant\Lambda_{1}|\xi|^{2},\quad|\tilde{A}^{ij}(x)-\tilde{A}^{ij}(y)|\leqslant C\Lambda_{2}|x-y|.

we obtain that there exists C=C(A,p,C0,C1,C2,C3)>0C=C(A,p,C_{0},C_{1},C_{2},C_{3})>0 such that

u~Lp(𝒞(C1,C2))C(ΔA~u~Lp(𝒞(1+C0,1+C3))+u~Lp(𝒞(1+C0,1+C3))).\|\nabla\tilde{u}\|_{L^{p}(\mathcal{C}(C_{1},C_{2}))}\leqslant C\Big{(}\|\Delta_{\tilde{A}}\tilde{u}\|_{L^{p}(\mathcal{C}(1+C_{0},1+C_{3}))}+\|\tilde{u}\|_{L^{p}(\mathcal{C}(1+C_{0},1+C_{3}))}\Big{)}.

Then we use the previous relations between the norms of u~\tilde{u} and uu and their derivatives to get

rα1uLp(𝒞(r(1+C1α1),r(1+C2α1)))C((rα1)2ΔAuLp(𝒞(r(1+C0α1),r(1+C3α1)))+uLp(𝒞(r(1+C0α1),r(1+C3α1)))).r\alpha^{-1}\|\nabla u\|_{L^{p}(\mathcal{C}(r(1+C_{1}\alpha^{-1}),r(1+C_{2}\alpha^{-1})))}\\ \leqslant C\Big{(}(r\alpha^{-1})^{2}\|\Delta_{A}u\|_{L^{p}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}\\ +\|u\|_{L^{p}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}\Big{)}.

We multiply the above estimate by (rα1)2(r\alpha^{-1})^{-2} to get the expected result (2.8).

We prove the second point, we apply Gagliardo-Nirenberg interpolation’s inequality, see [Nir66], to u~\tilde{u} defined by the relations (2.11) and (2.12) to get

u~Lq(𝒞(C0,C3))Cu~Lp(𝒞(C0,C3))1/2u~Lp(𝒞(C0,C3))1/2+Cu~Lp(𝒞(C0,C3)).\|\tilde{u}\|_{L^{q}(\mathcal{C}(C_{0},C_{3}))}\leqslant C\|\nabla\tilde{u}\|_{L^{p}(\mathcal{C}(C_{0},C_{3}))}^{1/2}\|\tilde{u}\|_{L^{p}(\mathcal{C}(C_{0},C_{3}))}^{1/2}+C\|\tilde{u}\|_{L^{p}(\mathcal{C}(C_{0},C_{3}))}.

Moreover, we have as before

u~Lq(𝒞(C0,C3))(rdα)1/quLq(𝒞(r(1+C0α1),r(1+C3α1))),\|\tilde{u}\|_{L^{q}(\mathcal{C}(C_{0},C_{3}))}\approx(r^{-d}\alpha)^{1/q}\|u\|_{L^{q}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))},
u~Lp(𝒞(C0,C3))(rdα)1/puLp(𝒞(r(1+C0α1),r(1+C3α1))),\|\tilde{u}\|_{L^{p}(\mathcal{C}(C_{0},C_{3}))}\approx(r^{-d}\alpha)^{1/p}\|u\|_{L^{p}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))},
u~Lp(𝒞(C0,C3))(rdα)1/p(rα1)uLp(𝒞(r(1+C0α1),r(1+C3α1))).\|\nabla\tilde{u}\|_{L^{p}(\mathcal{C}(C_{0},C_{3}))}\approx(r^{-d}\alpha)^{1/p}(r\alpha^{-1})\|\nabla u\|_{L^{p}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}.

By putting these relations in the previous Gagliardo-Nirenberg interpolation’s inequality, we obtain

uLq(𝒞(r(1+C0α1),r(1+C3α1)))(rdα)1/qCuLp(𝒞(r(1+C0α1),r(1+C3α1)))1/2(rdα)1/2p(rα1)1/2uLp(𝒞(r(1+C0α1),r(1+C3α1)))1/2(rdα)1/2p+CuLp(𝒞(r(1+C0α1),r(1+C3α1)))(rdα)1/p,\|u\|_{L^{q}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}(r^{-d}\alpha)^{1/q}\\ \leqslant C\|\nabla u\|_{L^{p}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}^{1/2}(r^{-d}\alpha)^{1/2p}(r\alpha^{-1})^{1/2}\\ \|u\|_{L^{p}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}^{1/2}(r^{-d}\alpha)^{1/2p}\\ +C\|u\|_{L^{p}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}(r^{-d}\alpha)^{1/p},

that simplifies into (2.10) by using (2.9). ∎

2.4 Proof of the growth estimate for AA-harmonic functions

The goal of this part is to prove 1.3.

The proof of (1.18) will be quite long and rather technical, while the proof of (1.19) will be a straightforward corollary of a generalization of (1.18) that we state below, see 2.6.

Proof of (1.18) of 1.3.

We split the proof into several steps. It is worth mentioning that the first three steps are strongly inspired by the original proof of L2L^{2}-Bernstein estimates for Laplace eigenfunctions from [DF90a].

Step 1: Carleman estimate to a truncated version of uu. Let us take χCc(B2;[0,1])\chi\in C_{c}^{\infty}(B_{2};[0,1]) be a cut-off function such that

χ\displaystyle\chi =0inB(0,r(1+(1/4)α1)),\displaystyle=0\ \text{in}\ B(0,r(1+(1/4)\alpha^{-1})),
χ\displaystyle\chi =1inB(0,15/8)B(0,r(1+(1/2)α1)),\displaystyle=1\ \text{in}\ B(0,15/8)\setminus B(0,r(1+(1/2)\alpha^{-1})),
χ\displaystyle\chi =0inB(0,2)B(0,31/16),\displaystyle=0\ \text{in}\ B(0,2)\setminus B(0,31/16),

satisfying the estimates

|χ|C,|D2χ|C,xB(0,31/16)B(0,15/8),|\nabla\chi|\leqslant C,\ |D^{2}\chi|\leqslant C,\qquad x\in B(0,31/16)\setminus B(0,15/8), (2.13)

and

|χ|Cr1α,|D2χ|Cr2α2,xB(0,r(1+(1/2)α1))B(0,r(1+(1/4)α1)).|\nabla\chi|\leqslant Cr^{-1}\alpha,\ |D^{2}\chi|\leqslant Cr^{-2}\alpha^{-2},\qquad x\in B(0,r(1+(1/2)\alpha^{-1}))\setminus B(0,r(1+(1/4)\alpha^{-1})). (2.14)

We define f=χuf=\chi u, note that fHloc2(B2)f\in H_{\text{loc}}^{2}(B_{2}) with supp(f)B2B(0,r)\text{supp}(f)\subset\subset B_{2}\setminus B(0,r) so we can apply the Carleman estimate (2.2) from 2.2 to ff to get

α3B2ρ12α|f|2𝑑x+α4r2B(0,r(1+2α1))ρ2α|f|2𝑑xCB2ρ22α|ΔAf|2𝑑x.\alpha^{3}\int_{B_{2}}\rho^{-1-2\alpha}|f|^{2}dx+\frac{\alpha^{4}}{r^{2}}\int_{B(0,r(1+2\alpha^{-1}))}\rho^{-2\alpha}|f|^{2}dx\leqslant C\int_{B_{2}}\rho^{2-2\alpha}|\Delta_{A}f|^{2}dx. (2.15)

We now use the elliptic equation (1.16) satisfied by uu to deduce from (2.15)

α3B2ρ12αχ2|u|2𝑑x+α4r2B(0,r(1+α1))ρ2αχ2|u|2𝑑xC(I1+I2),\alpha^{3}\int_{B_{2}}\rho^{-1-2\alpha}\chi^{2}|u|^{2}dx+\frac{\alpha^{4}}{r^{2}}\int_{B(0,r(1+\alpha^{-1}))}\rho^{-2\alpha}\chi^{2}|u|^{2}dx\leqslant C(I_{1}+I_{2}), (2.16)

where

I1=B(0,r(1+(1/2)α1))B(0,r(1+(1/4)α1))ρ22α(|χ|2|u|2+[|χ|2+|D2χ|2]|u|2)𝑑x,I_{1}=\int_{B(0,r(1+(1/2)\alpha^{-1}))\setminus B(0,r(1+(1/4)\alpha^{-1}))}\rho^{2-2\alpha}(|\nabla\chi|^{2}|\nabla u|^{2}+[|\nabla\chi|^{2}+|D^{2}\chi|^{2}]|u|^{2})dx, (2.17)

and

I2=B(0,31/16)B(0,15/8)ρ22α(|χ|2|u|2+[|χ|2+|D2χ|2]|u|2)𝑑x.I_{2}=\int_{B(0,31/16)\setminus B(0,15/8)}\rho^{2-2\alpha}(|\nabla\chi|^{2}|\nabla u|^{2}+[|\nabla\chi|^{2}+|D^{2}\chi|^{2}]|u|^{2})dx. (2.18)

Step 2: Absorption of the boundary terms. The goal of this step would be to absorb the cut-off terms from (2.16) that are located near the boundary of B(0,2)B(0,2) that is the term I2I_{2} defined in (2.18). By (2.13) and from local elliptic regularity estimates, it is straightforward to get that exist a universal positive constant C3>0C_{3}>0 and a positive constant C=C(A)>0C=C(A)>0 such that

I2Cexp(C3α)(supB2|u|)2.I_{2}\leqslant C\exp(-C_{3}\alpha)(\sup_{B_{2}}|u|)^{2}. (2.19)

Moreover, we can give a lower bound of the first term in the left hand side of (2.16) of the following form by using the definition of χ\chi and local elliptic regularity estimates

α3B2ρ12αχ2|u|2𝑑xC1exp(C2α)(supB(x0,1/4)|u|)2,\alpha^{3}\int_{B_{2}}\rho^{-1-2\alpha}\chi^{2}|u|^{2}dx\geqslant C^{-1}\exp(-C_{2}\alpha)(\sup_{B(x_{0},1/4)}|u|)^{2}, (2.20)

where 0<C2<C30<C_{2}<C_{3} and x0B2x_{0}\in B_{2} be such that |x0|=3/8|x_{0}|=3/8. Now by using the vanishing order estimate (2.4) from 2.4, note in particular that χ1\chi\equiv 1 in B(x0,1/4)B(x_{0},1/4) for r(0,c)r\in(0,c) with c>0c>0 small enough, we deduce from (2.20) that

α3B2ρ12αχ2|u|2𝑑xC1exp(C2α)exp(CN)(supB2|u|)2,\alpha^{3}\int_{B_{2}}\rho^{-1-2\alpha}\chi^{2}|u|^{2}dx\geqslant C^{-1}\exp(-C_{2}\alpha)\exp(-CN)(\sup_{B_{2}}|u|)^{2}, (2.21)

Now the punchline, by taking αC(C2,C3,A)N\alpha\geqslant C(C_{2},C_{3},A)N, we obtain from (2.16), (2.19) and (2.21) that I2I_{2} can be hidden in the left hand side of (2.16), that is there exists C=C(A)>0C=C(A)>0 such that

α3B2ρ12αχ2|u|2𝑑x+α4r2B(0,r(1+2α1))ρ2αχ2|u|2𝑑xCI1.\alpha^{3}\int_{B_{2}}\rho^{-1-2\alpha}\chi^{2}|u|^{2}dx+\frac{\alpha^{4}}{r^{2}}\int_{B(0,r(1+2\alpha^{-1}))}\rho^{-2\alpha}\chi^{2}|u|^{2}dx\leqslant CI_{1}. (2.22)

Step 3: Growth estimate at L2L^{2}-regularity. The goal of this step would be to compare the second left hand side term in (2.22) and the right hand side term in (2.22) to first obtain growth estimate at L2L^{2}-regularity. First note that ρ2α\rho^{-2\alpha} has the same amplitude in B(0,r(1+2α1))B(0,r)B(0,r(1+2\alpha^{-1}))\setminus B(0,r) so one can simplify (2.22) by using (2.17) into

α4r2B(0,r(1+2α1))χ2|u|2𝑑xCB(0,r(1+(1/2)α1))B(0,r(1+(1/4)α1))r2(|χ|2|u|2+[|χ|2+|D2χ|2]|u|2)𝑑x.\frac{\alpha^{4}}{r^{2}}\int_{B(0,r(1+2\alpha^{-1}))}\chi^{2}|u|^{2}dx\\ \leqslant C\int_{B(0,r(1+(1/2)\alpha^{-1}))\setminus B(0,r(1+(1/4)\alpha^{-1}))}r^{2}(|\nabla\chi|^{2}|\nabla u|^{2}+[|\nabla\chi|^{2}+|D^{2}\chi|^{2}]|u|^{2})dx. (2.23)

Now we use (2.14) to deduce from (2.23) that

α4r2B(0,r(1+2α1))χ2|u|2𝑑xCB(0,r(1+(1/2)α1))B(0,r(1+(1/4)α1))r2((r1α)2|u|2+(r2α2)2|u|2)𝑑x.\frac{\alpha^{4}}{r^{2}}\int_{B(0,r(1+2\alpha^{-1}))}\chi^{2}|u|^{2}dx\\ \leqslant C\int_{B(0,r(1+(1/2)\alpha^{-1}))\setminus B(0,r(1+(1/4)\alpha^{-1}))}r^{2}((r^{-1}\alpha)^{2}|\nabla u|^{2}+(r^{-2}\alpha^{2})^{2}|u|^{2})dx. (2.24)

By using the scaled local elliptic regularity estimate (2.8) for p=2p=2 from 2.5 to the right hand side of (2.24) we get for some C0,C0>0C_{0},C_{0}^{\prime}>0 such that C0<1/4<1/2<C0<3/4C_{0}<1/4<1/2<C_{0}^{\prime}<3/4,

α4r2B(0,r(1+2α1))χ2|u|2𝑑xCα4r2B(0,r(1+C0α))B(0,r(1+C0α))|u|2𝑑x.\frac{\alpha^{4}}{r^{2}}\int_{B(0,r(1+2\alpha^{-1}))}\chi^{2}|u|^{2}dx\leqslant C\frac{\alpha^{4}}{r^{2}}\int_{B(0,r(1+C_{0}^{\prime}\alpha))\setminus B(0,r(1+C_{0}\alpha))}|u|^{2}dx. (2.25)

In particular, (2.25) translates into

B(0,r(1+2α1))B(0,r(1+21α1))|u|2𝑑xCB(0,r(1+C0α))B(0,r(1+C0α))|u|2𝑑x.\int_{B(0,r(1+2\alpha^{-1}))\setminus B(0,r(1+2^{-1}\alpha^{-1}))}|u|^{2}dx\leqslant C\int_{B(0,r(1+C_{0}^{\prime}\alpha))\setminus B(0,r(1+C_{0}\alpha))}|u|^{2}dx. (2.26)

Note that this type of strategy already appears in [DF90a] to obtain growth estimate at L2L^{2}-regularity leading in particular to the Bernstein estimates at L2L^{2}-regularity for Laplace eigenfunctions, i.e. (1.6) and (1.7).

Step 4: Initialization of the bootstrap argument. The goal of this step would be to improve the growth estimate at L2L^{2}-regularity (2.26) by employing scaled versions of local elliptic regularity estimates and Gagliardo-Nirenberg interpolation inequalities from 2.5 by using the equation satisfied by uu, i.e. (1.16). First we have from (2.26) and (1.16) i.e. ΔAu=0\Delta_{A}u=0,

α4r2B(0,r(1+2α1))B(0,r(1+21α1))|u|2𝑑x+r2B(0,r(1+2α1))B(0,r(1+21α1))|ΔAu|2𝑑xCα4r2B(0,r(1+C0α))B(0,r(1+C0α))|u|2𝑑x.\alpha^{4}r^{-2}\int_{B(0,r(1+2\alpha^{-1}))\setminus B(0,r(1+2^{-1}\alpha^{-1}))}|u|^{2}dx+r^{2}\int_{B(0,r(1+2\alpha^{-1}))\setminus B(0,r(1+2^{-1}\alpha^{-1}))}|\Delta_{A}u|^{2}dx\\ \leqslant C\alpha^{4}r^{-2}\int_{B(0,r(1+C_{0}^{\prime}\alpha))\setminus B(0,r(1+C_{0}\alpha))}|u|^{2}dx. (2.27)

We now use the local elliptic regularity estimate (2.8) for p=2p=2 to obtain from (2.27) that

α2B(0,r(1+C3α1))B(0,r(1+C2α1))|u|2𝑑xCα4r2B(0,r(1+C0α))B(0,r(1+C0α))|u|2𝑑x,\alpha^{2}\int_{B(0,r(1+C_{3}\alpha^{-1}))\setminus B(0,r(1+C_{2}\alpha^{-1}))}|\nabla u|^{2}dx\\ \leqslant C\alpha^{4}r^{-2}\int_{B(0,r(1+C_{0}^{\prime}\alpha))\setminus B(0,r(1+C_{0}\alpha))}|u|^{2}dx, (2.28)

for some positive constants 1/2<C2<3/4<1<C3<21/2<C_{2}<3/4<1<C_{3}<2. We now use the interpolation inequality (2.10) for p=2p=2 with q>2q>2 defined by (2.9) to deduce from (2.27), (2.28) that

α212dr1/2uLq(𝒞(r(1+C2α1),r(1+C3α1)))Cα2r1uL2(𝒞(r(1+C0α1),r(1+C0α1))).\alpha^{2-\frac{1}{2d}}r^{-1/2}\|u\|_{L^{q}(\mathcal{C}(r(1+C_{2}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}\leqslant C\alpha^{2}r^{-1}\|u\|_{L^{2}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{0}^{\prime}\alpha^{-1})))}. (2.29)

In particular, by using Hölder’s estimate on the right hand side of (2.29) and the fact that the Lebesgue measure of 𝒞(r(1+C0α1),r(1+C0α1))\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{0}^{\prime}\alpha^{-1})) is bounded by Crdα1Cr^{d}\alpha^{-1},

α212dr1/2uLq(𝒞(r(1+C2α1),r(1+C3α1)))Cα2r1(rdα1)(1/21/q)uLq(C(r(1+C0α1),r(1+C0α1))).\alpha^{2-\frac{1}{2d}}r^{-1/2}\|u\|_{L^{q}(\mathcal{C}(r(1+C_{2}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}\\ \leqslant C\alpha^{2}r^{-1}(r^{d}\alpha^{-1})^{(1/2-1/q)}\|u\|_{L^{q}(C(r(1+C_{0}\alpha^{-1}),r(1+C_{0}^{\prime}\alpha^{-1})))}. (2.30)

By recalling (2.9) with p=2p=2, one can check that (2.30) simplifies into

uLq(𝒞(r(1+C2α1),r(1+C3α1)))CuLq(𝒞(r(1+C0α1),r(1+C0α1))).\|u\|_{L^{q}(\mathcal{C}(r(1+C_{2}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}\leqslant C\|u\|_{L^{q}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{0}^{\prime}\alpha^{-1})))}. (2.31)

In particular, (2.31) leads to a growth-estimate at LqL^{q}-regularity that is

uLq(B(0,r(1+C3α1)))CuLq(B(0,r(1+max(C0,C2)α1))),\|u\|_{L^{q}(B\left(0,r\left(1+C_{3}\alpha^{-1}\right)\right))}\leqslant C\|u\|_{L^{q}\left(B\left(0,r\left(1+\max(C_{0}^{\prime},C_{2})\alpha^{-1}\right)\right)\right)}, (2.32)

by adding uLq(B(0,r(1+max(C0,C2)α1)))\|u\|_{L^{q}(B(0,r(1+\max(C_{0}^{\prime},C_{2})\alpha^{-1})))} in both sides of (2.31). Note that C3>1>3/4>max(C0,C2)C_{3}>1>3/4>\max(C_{0}^{\prime},C_{2}). The estimate (2.32) will not be used in the next, it only furnishes a growth estimate at LqL^{q}-regularity leading in particular to the Bernstein estimates at LqL^{q}-regularity for AA-harmonic functions. We need to iterate such an argument to reach LL^{\infty}.

Step 5: Bootstrap argument. We iterate the previous argument starting from (2.29). Let us define by induction

p0=2,1pn+1=12(1pn1d)+12pnn{0,,d1}.p_{0}=2,\ \frac{1}{p_{n+1}}=\frac{1}{2}\left(\frac{1}{p_{n}}-\frac{1}{d}\right)+\frac{1}{2p_{n}}\qquad\forall n\in\{0,\dots,d-1\}. (2.33)

One can check that

1pn=12n2d,n{0,,d1}andpd=+.\frac{1}{p_{n}}=\frac{1}{2}-\frac{n}{2d},\ \forall n\in\{0,\dots,d-1\}\ \text{and}\ p_{d}=+\infty. (2.34)

In the same way, we define

β0=2,βn+1=βn12dn{0,,d1}βd=32,\beta_{0}=2,\ \beta_{n+1}=\beta_{n}-\frac{1}{2d}\qquad\forall n\in\{0,\dots,d-1\}\ \Rightarrow\beta_{d}=\frac{3}{2}, (2.35)
s0=1,sn+1=sn+12n{0,,d1}sd=1+d2.s_{0}=-1,\ s_{n+1}=s_{n}+\frac{1}{2}\qquad\forall n\in\{0,\dots,d-1\}\ \Rightarrow s_{d}=-1+\frac{d}{2}. (2.36)

From an easy induction applied to the previous step, conjugated with the scaled version of elliptic regularity estimates and Gagliardo-Nirenberg interpolation inequalities from 2.5 we obtain after dd iterations that

αβdrsduLpd(𝒞(r(1+Cdα1),r(1+Cdα1)))Cα2r1uL2(𝒞(r(1+C0α1),r(1+C0α1))).\alpha^{\beta_{d}}r^{s_{d}}\|u\|_{L^{p_{d}}(\mathcal{C}(r(1+C_{d}\alpha^{-1}),r(1+C_{d}^{\prime}\alpha^{-1})))}\leqslant C\alpha^{2}r^{-1}\|u\|_{L^{2}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{0}^{\prime}\alpha^{-1})))}. (2.37)

for some positive constants 21<Cd<3/4<1<Cd<22^{-1}<C_{d}<3/4<1<C_{d}^{\prime}<2. So (2.37) gives

α32r1+d2uL(𝒞(r(1+Cdα1),r(1+Cdα1)))Cα2r1uL2(𝒞(r(1+C0α1),r(1+C0α1))).\alpha^{\frac{3}{2}}r^{-1+\frac{d}{2}}\|u\|_{L^{\infty}(\mathcal{C}(r(1+C_{d}\alpha^{-1}),r(1+C_{d}^{\prime}\alpha^{-1})))}\leqslant C\alpha^{2}r^{-1}\|u\|_{L^{2}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{0}^{\prime}\alpha^{-1})))}. (2.38)

We now apply Hölder’s estimate to the right hand side of (2.38) and the fact that the Lebesgue measure of 𝒞(r(1+C0α1),r(1+C0α1))\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{0}^{\prime}\alpha^{-1})) is bounded by Crdα1Cr^{d}\alpha^{-1} to get

α32r1+d2uL(𝒞(r(1+Cdα1),r(1+Cdα1)))Cα2r1(rdα1)1/2uL(𝒞(r(1+C0α1),r(1+C0α1))),\alpha^{\frac{3}{2}}r^{-1+\frac{d}{2}}\|u\|_{L^{\infty}(\mathcal{C}(r(1+C_{d}\alpha^{-1}),r(1+C_{d}^{\prime}\alpha^{-1})))}\\ \leqslant C\alpha^{2}r^{-1}(r^{d}\alpha^{-1})^{1/2}\|u\|_{L^{\infty}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{0}^{\prime}\alpha^{-1})))},

that simplifies into

uL(𝒞(r(1+Cdα1),r(1+Cdα1)))CuL(𝒞(r(1+C0α1),r(1+C0α1))).\|u\|_{L^{\infty}(\mathcal{C}(r(1+C_{d}\alpha^{-1}),r(1+C_{d}^{\prime}\alpha^{-1})))}\leqslant C\|u\|_{L^{\infty}(\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{0}^{\prime}\alpha^{-1})))}. (2.39)

Recall that the constants are such that 1/2<Cd<3/4<1<Cd<21/2<C_{d}<3/4<1<C_{d}^{\prime}<2. In particular, this leads to a growth-estimate at LL^{\infty}-regularity that is

uL(B(0,r(1+Cdα1)))CuL(B(0,r(1+max(C0,Cd)α1))),\|u\|_{L^{\infty}\left(B\left(0,r\left(1+C_{d}^{\prime}\alpha^{-1}\right)\right)\right)}\leqslant C\|u\|_{L^{\infty}\left(B\left(0,r\left(1+\max(C_{0}^{\prime},C_{d})\alpha^{-1}\right)\right)\right)}, (2.40)

by adding uL(B(0,r(1+max(C0,Cd)α1)))\|u\|_{L^{\infty}(B(0,r(1+\max(C_{0}^{\prime},C_{d})\alpha^{-1})))} in both sides of (2.39). If we replace rr by r/((1+max(C0,Cd)α1))r/((1+\max(C_{0}^{\prime},C_{d})\alpha^{-1})), we then get from (2.40)

uL(B(0,r(1+Cdα1)))CuL(B(0,r)),\|u\|_{L^{\infty}\left(B\left(0,r\left(1+C_{d}^{\prime}\alpha^{-1}\right)\right)\right)}\leqslant C\|u\|_{L^{\infty}(B(0,r))}, (2.41)

for another constant Cd>0C_{d}^{\prime}>0. We then recall that αCN\alpha\geqslant CN so one has from (2.41)

uL(B(0,r(1+CdN1)))CuL(B(0,r)).\|u\|_{L^{\infty}\left(B\left(0,r\left(1+C_{d}^{\prime}N^{-1}\right)\right)\right)}\leqslant C\|u\|_{L^{\infty}(B(0,r))}. (2.42)

We then iterate (2.42) to finally deduce (1.18). ∎

From the proof of (1.18) of 1.3, one can also deduce the following result.

Lemma 2.6.

There exist r0,C>0r_{0},C>0 such that for every uHloc1(B2)L(B2)u\in H_{\text{loc}}^{1}(B_{2})\cap L^{\infty}(B_{2}) satisfying (1.16) and (1.17), then for every r(0,r0)r\in(0,r_{0}),

supB(x,r(1+1N))|u|CsupB(x,r)|u|,x(0,2r).\sup_{B\left(x,r\left(1+\frac{1}{N}\right)\right)}|u|\leqslant C\sup_{B\left(x,r\right)}|u|,\qquad\forall x\in(0,2r). (2.43)

Indeed, the proof of (2.43) consists in applying a modified version of the Carleman estimate (2.2) in the punctured domain B(0,2)B(x,r)B(0,2)\setminus B(x,r). Details are omitted.

We can now pass to the proof of (1.19) of 1.3.

Proof of (1.19) of 1.3.

Let us then take yB(0,r)y\in B(0,r) be such that

|u(y)|=supB(0,r)|u|.|\nabla u(y)|=\sup_{B(0,r)}|\nabla u|.

By a scaled local elliptic estimate applied to the equation (1.16) and by using (2.43), we have

|u(y)|CNrsupB(y,rN1)|u|CNrsupB(x,r(1+N1))|u|CNrsupB(x,r)|u|.|\nabla u(y)|\leqslant C\frac{N}{r}\sup_{B(y,rN^{-1})}|u|\leqslant C\frac{N}{r}\sup_{B(x,r(1+N^{-1}))}|u|\leqslant C\frac{N}{r}\sup_{B(x,r)}|u|.

This ends the proof of (1.19). ∎

2.5 Proof of the growth estimate for Laplace eigenfunctions

The goal of this part is to prove 1.2.

To simplify the notations, in all the following, the geodesic balls Bg(x,r)B_{g}(x,r) will be denoted by B(x,r)B(x,r) and the Riemannian volume of integration will be denoted by dxdx.

The following crucial L2L^{2}-Carleman estimate for the operator Δgλ-\Delta_{g}-\lambda will be one of the main ingredient of the proof.

Lemma 2.7.

There exists a positive constant C=C(M,g)>0C=C(M,g)>0 and c=c(M,g)>0c=c(M,g)>0 such that for every rM(0,c)r_{M}\in(0,c), r(0,c)r\in(0,c), x0Mx_{0}\in M, αC(λ+1)\alpha\geqslant C(\sqrt{\lambda}+1), for all fCc(BrMBr)f\in C_{c}^{\infty}(B_{r_{M}}\setminus B_{r}), the following estimate holds

α3B(x0,rM)ρ12α|f|2𝑑x+αB(x0,rM)ρ12α|f|2𝑑x+α4r2B(x0,r(1+2α1))ρ2α|f|2𝑑xCB(x0,rM)ρ22α|Δgf+λf|2𝑑x,\alpha^{3}\int_{B(x_{0},r_{M})}\rho^{-1-2\alpha}|f|^{2}dx+\alpha\int_{B(x_{0},r_{M})}\rho^{1-2\alpha}|\nabla f|^{2}dx+\frac{\alpha^{4}}{r^{2}}\int_{B(x_{0},r(1+2\alpha^{-1}))}\rho^{-2\alpha}|f|^{2}dx\\ \leqslant C\int_{B(x_{0},r_{M})}\rho^{2-2\alpha}|\Delta_{g}f+\lambda f|^{2}dx, (2.44)

where ρ(x)=dist(x,x0)\rho(x)=\mathrm{dist}(x,x_{0}).

The proof of 2.7 is similar to the one of 2.2, it is actually proved in [DF90a, Lemma A]. The crucial difference between the proof of 2.2 is the presence of the zero-order term λf\lambda f in the elliptic operator Δgfλf-\Delta_{g}f-\lambda f . This new term cannot be treated as a source term and absorbed by the Carleman parameter α\alpha because such a strategy would lead to (2.44) for αC(λ2/3+1)\alpha\geqslant C(\lambda^{2/3}+1). This is why it has to be directly included in the symmetric part of the conjugated operator in the Carleman’s strategy, note that here we crucially use the assumption that λ\lambda is constant then does not depend on the xx-variable.

The following classical vanishing order estimate holds for Laplace eigenfunctions.

Lemma 2.8.

There exist C=C(M,g)>0C=C(M,g)>0, c=c(M,g)>0c=c(M,g)>0 and rM,r0(0,c)r_{M},r_{0}\in(0,c), such that for every λ1\lambda\geqslant 1, for every φλC(M)\varphi_{\lambda}\in C^{\infty}(M) satisfying (1.1), the following estimate holds

supB(x0,r04)|φλ|exp(Cλ)supB(x0,rM)|φλ|xB(x0,r02).\sup_{B\left(x_{0},\frac{r_{0}}{4}\right)}|\varphi_{\lambda}|\geqslant\exp(-C\sqrt{\lambda})\sup_{B\left(x_{0},r_{M}\right)}|\varphi_{\lambda}|\qquad\forall x\in B\left(x_{0},\frac{r_{0}}{2}\right). (2.45)

The proof of 2.8 is similar to the one of 2.4 so we omit it.

With 2.7 and 2.8, we can now deduce the proof of the growth estimate (1.13) by adapting the arguments of the proof of the growth estimate (1.18).

Proof of (1.13) from 1.2.

The first three steps are the same. We apply the Carleman estimate (2.44) to a truncated version of φλ\varphi_{\lambda}, the boundary terms are then absorbed by the vanishing order estimate (2.45), we finally obtain a Bernstein estimate at L2L^{2}-regularity

B(x0,r(1+2α1))B(x0,r(1+21α1))|φλ|2𝑑xCB(x0,r(1+C0α))B(x0,r(1+C0α))|φλ|2𝑑x.\int_{B(x_{0},r(1+2\alpha^{-1}))\setminus B(x_{0},r(1+2^{-1}\alpha^{-1}))}|\varphi_{\lambda}|^{2}dx\leqslant C\int_{B(x_{0},r(1+C_{0}^{\prime}\alpha))\setminus B(x_{0},r(1+C_{0}\alpha))}|\varphi_{\lambda}|^{2}dx. (2.46)

Note that this type of strategy already appears in [DF90a] to obtain growth estimate at L2L^{2}-regularity leading in particular to the Bernstein estimates at L2L^{2}-regularity for Laplace eigenfunctions, i.e. (1.6) and (1.7).

Step 4: Initialization of the bootstrap argument. The goal of this step would be to improve the growth estimate at L2L^{2}-regularity (2.46) by employing scaled versions of local elliptic regularity estimates and Gagliardo-Nirenberg interpolation inequalities from 2.5 by using the equation satisfied by φλ\varphi_{\lambda}, i.e. (1.1). First we have from (2.46) and (1.1) i.e. Δgφλ+λφλ=0\Delta_{g}\varphi_{\lambda}+\lambda\varphi_{\lambda}=0,

α4r2B(x0,r(1+2α1))B(x0,r(1+21α1))|φλ|2𝑑x+r2B(x0,r(1+2α1))B(x0,r(1+21α1))|Δgφλ|2𝑑xCα4r2B(x0,r(1+C0α))B(x0,r(1+C0α))|φλ|2𝑑x+Cr2λ2B(x0,r(1+2α1))B(x0,r(1+21α1))|φλ|2𝑑x.\alpha^{4}r^{-2}\int_{B(x_{0},r(1+2\alpha^{-1}))\setminus B(x_{0},r(1+2^{-1}\alpha^{-1}))}|\varphi_{\lambda}|^{2}dx+r^{2}\int_{B(x_{0},r(1+2\alpha^{-1}))\setminus B(x_{0},r(1+2^{-1}\alpha^{-1}))}|\Delta_{g}\varphi_{\lambda}|^{2}dx\\ \leqslant C\alpha^{4}r^{-2}\int_{B(x_{0},r(1+C_{0}^{\prime}\alpha))\setminus B(x_{0},r(1+C_{0}\alpha))}|\varphi_{\lambda}|^{2}dx\\ +Cr^{2}\lambda^{2}\int_{B(x_{0},r(1+2\alpha^{-1}))\setminus B(x_{0},r(1+2^{-1}\alpha^{-1}))}|\varphi_{\lambda}|^{2}dx. (2.47)

The last right hand side term of (2.47) can be absorbed by the first left hand side term recalling that αCλ\alpha\geqslant C\sqrt{\lambda} to get

α4r2B(x0,r(1+2α1))B(x0,r(1+21α1))|φλ|2𝑑x+r2B(x0,r(1+2α1))B(x0,r(1+21α1))|Δgφλ|2𝑑xCα4r2B(x0,r(1+C0α))B(x0,r(1+C0α))|φλ|2𝑑x.\alpha^{4}r^{-2}\int_{B(x_{0},r(1+2\alpha^{-1}))\setminus B(x_{0},r(1+2^{-1}\alpha^{-1}))}|\varphi_{\lambda}|^{2}dx+r^{2}\int_{B(x_{0},r(1+2\alpha^{-1}))\setminus B(x_{0},r(1+2^{-1}\alpha^{-1}))}|\Delta_{g}\varphi_{\lambda}|^{2}dx\\ \leqslant C\alpha^{4}r^{-2}\int_{B(x_{0},r(1+C_{0}^{\prime}\alpha))\setminus B(x_{0},r(1+C_{0}\alpha))}|\varphi_{\lambda}|^{2}dx. (2.48)

We now use the local elliptic regularity estimate (2.8) for p=2p=2 to obtain from (2.48) that

α2B(x0,r(1+C3α1))B(x0,r(1+C2α1))|φλ|2𝑑xCα4r2B(x0,r(1+C0α))B(x0,r(1+C0α))|u|2𝑑x,\alpha^{2}\int_{B(x_{0},r(1+C_{3}\alpha^{-1}))\setminus B(x_{0},r(1+C_{2}\alpha^{-1}))}|\nabla\varphi_{\lambda}|^{2}dx\\ \leqslant C\alpha^{4}r^{-2}\int_{B(x_{0},r(1+C_{0}^{\prime}\alpha))\setminus B(x_{0},r(1+C_{0}\alpha))}|u|^{2}dx, (2.49)

for some positive constants 1/2<C2<3/4<1<C3<21/2<C_{2}<3/4<1<C_{3}<2. We now use the interpolation inequality (2.10) for p=2p=2 with q>2q>2 defined by (2.9) to deduce from (2.48), (2.49) that

α212dr1/2φλLq(𝒞(x0,r(1+C2α1),r(1+C3α1)))Cα2r1φλL2(𝒞(x0,r(1+C0α1),r(1+C0α1))).\alpha^{2-\frac{1}{2d}}r^{-1/2}\|\varphi_{\lambda}\|_{L^{q}(\mathcal{C}(x_{0},r(1+C_{2}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}\leqslant C\alpha^{2}r^{-1}\|\varphi_{\lambda}\|_{L^{2}(\mathcal{C}(x_{0},r(1+C_{0}\alpha^{-1}),r(1+C_{0}^{\prime}\alpha^{-1})))}. (2.50)

In particular, by using Hölder’s estimate on the right hand side of (2.50) and the fact that the Lebesgue measure of 𝒞(r(1+C0α1),r(1+C0α1))\mathcal{C}(r(1+C_{0}\alpha^{-1}),r(1+C_{0}^{\prime}\alpha^{-1})) is bounded by Crdα1Cr^{d}\alpha^{-1},

α212dr1/2φλLq(𝒞(x0,r(1+C2α1),r(1+C3α1)))Cα2r1(rdα1)(1/21/q)φλLq(C(x0,r(1+C0α1),r(1+C0α1))).\alpha^{2-\frac{1}{2d}}r^{-1/2}\|\varphi_{\lambda}\|_{L^{q}(\mathcal{C}(x_{0},r(1+C_{2}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}\\ \leqslant C\alpha^{2}r^{-1}(r^{d}\alpha^{-1})^{(1/2-1/q)}\|\varphi_{\lambda}\|_{L^{q}(C(x_{0},r(1+C_{0}\alpha^{-1}),r(1+C_{0}^{\prime}\alpha^{-1})))}. (2.51)

By recalling (2.9) with p=2p=2, one can check that (2.51) simplifies into

φλLq(𝒞(x0,r(1+C2α1),r(1+C3α1)))CφλLq(𝒞(x0,r(1+C0α1),r(1+C0α1))).\|\varphi_{\lambda}\|_{L^{q}(\mathcal{C}(x_{0},r(1+C_{2}\alpha^{-1}),r(1+C_{3}\alpha^{-1})))}\leqslant C\|\varphi_{\lambda}\|_{L^{q}(\mathcal{C}(x_{0},r(1+C_{0}\alpha^{-1}),r(1+C_{0}^{\prime}\alpha^{-1})))}. (2.52)

In particular, (2.31) leads to a growth-estimate at LqL^{q}-regularity that is

φλLq(B(x0,r(1+C3α1)))CφλLq(B(x0,r(1+max(C0,C2)α1))),\|\varphi_{\lambda}\|_{L^{q}(B\left(x_{0},r\left(1+C_{3}\alpha^{-1}\right)\right))}\leqslant C\|\varphi_{\lambda}\|_{L^{q}\left(B\left(x_{0},r\left(1+\max(C_{0}^{\prime},C_{2})\alpha^{-1}\right)\right)\right)}, (2.53)

by adding φλLq(B(x0,r(1+max(C0,C2)α1)))\|\varphi_{\lambda}\|_{L^{q}(B(x_{0},r(1+\max(C_{0}^{\prime},C_{2})\alpha^{-1})))} in both sides of (2.52). Note that C3>1>3/4>max(C0,C2)C_{3}>1>3/4>\max(C_{0}^{\prime},C_{2}). The estimate (2.53) will not be used in the next, it only furnishes a growth estimate at LqL^{q}-regularity leading in particular to the Bernstein estimates at LqL^{q}-regularity for Laplace eigenfunctions. We need to iterate such an argument to reach LL^{\infty}.

By simply following the fifth step of the proof of the growth estimate (1.18), consisting in implementing a bootstrap argument, we easily obtain the desired estimate (1.13). ∎

We then note that (1.14) is a consequence of (1.13).

Proof of (1.14) of 1.2.

Let us take yBg(x,r)y\in B_{g}(x,r) be such that

|φλ(y)|=supBg(x,r)|φλ|.|\nabla\varphi_{\lambda}(y)|=\sup_{B_{g}(x,r)}|\nabla\varphi_{\lambda}|.

By a scaled local elliptic estimate applied to the equation satisfied by φλ\varphi_{\lambda}, i.e. (1.1), and (1.13), we have

|φλ(y)|CλrsupBg(y,rλ1)|φλ|CλrsupBg(x,r(1+λ1))|φλ|CλrsupBg(x,r)|φλ|,|\nabla\varphi_{\lambda}(y)|\leqslant C\frac{\sqrt{\lambda}}{r}\sup_{B_{g}(y,r\sqrt{\lambda}^{-1})}|\varphi_{\lambda}|\leqslant C\frac{\sqrt{\lambda}}{r}\sup_{B_{g}(x,r(1+\sqrt{\lambda}^{-1}))}|\varphi_{\lambda}|\leqslant C\frac{\sqrt{\lambda}}{r}\sup_{B_{g}(x,r)}|\varphi_{\lambda}|,

that is exactly (1.14). ∎

3 Extensions

The goal of this part is to state several generalizations of our main results 1.2 and 1.3. For the sake of simplicity, we only give sketches of the proof, the details will be omitted. We also propose some open problems that can be investigated in the future.

3.1 Manifolds with boundaries

The treatment of CC^{\infty}-manifolds MM, possibly with boundaries i.e. M\partial M\neq\emptyset, are treated by the following result.

Theorem 3.1.

There exist r0,C>0r_{0},C>0 depending only on MM, such that for every Laplace eigenfunction φλC(M)\varphi_{\lambda}\in C^{\infty}(M), i.e. satisfying

Δgφλ=λφλinM,(φλ=0onM)or(νφλ=0onM),-\Delta_{g}\varphi_{\lambda}=\lambda\varphi_{\lambda}\ \text{in}\ M,\ (\varphi_{\lambda}=0\ \text{on}\ \partial M)\ \text{or}\ (\partial_{\nu}\varphi_{\lambda}=0\ \text{on}\ \partial M),

then, for every xMx\in M, r(0,r0)r\in(0,r_{0}), λ1\lambda\geqslant 1, (1.13) and (1.14) hold.

The proof of 3.1 can be obtained from 1.2 and the double manifold trick, see for instance [DF90b] or more precisely [BM23, Section 3], that consists in reducing the question to the case of a manifold without boundary by gluing two copies of MM along the boundary in such a way that the new double manifold M~\tilde{M} inherits a Lipschitz metric, which allows one to apply the previous results (without boundary) to this double manifold.

3.2 On elliptic differential inequalities

First, we have the following generalization of 1.2.

Theorem 3.2.

There exist r0,C>0r_{0},C>0 depending only on MM, such that for function φC(M)\varphi\in C^{\infty}(M), satisfying the elliptic differential inequality

|Δgφ|λ|φ|+μ|φ|inM,|-\Delta_{g}\varphi|\leqslant\lambda|\varphi|+\mu|\nabla\varphi|\ \text{in}\ M, (3.1)

where λ,μ1\lambda,\mu\geqslant 1, then for every xMx\in M, r(0,r0)r\in(0,r_{0}),

supBg(x,r(1+min(1λ2/3,1μ2)))|φ|CsupBg(x,r)|φ|,\sup_{B_{g}\left(x,r\left(1+\min\left(\frac{1}{\lambda^{2/3}},\frac{1}{\mu^{2}}\right)\right)\right)}|\varphi|\leqslant C\sup_{B_{g}\left(x,r\right)}|\varphi|, (3.2)

and

supBg(x,r)|φ|Cmax(λ2/3,μ2)rsupBg(x,r)|φ|.\sup_{B_{g}\left(x,r\right)}|\nabla\varphi|\leqslant C\frac{\max\left(\lambda^{2/3},\mu^{2}\right)}{r}\sup_{B_{g}\left(x,r\right)}|\varphi|. (3.3)

The proof of 3.2 cannot use the standard lifting trick from (1.20) so one needs to proceed differently.

First, by a standard L2L^{2}-Carleman estimate and the arguments of [DF88], the proof consists in establishing the following vanishing order estimate for φ\varphi.

\bullet There exist r0,C>0r_{0},C>0 depending only on MM, such that for every function φC(M)\varphi\in C^{\infty}(M) satisfying (3.1), for every xMx\in M, r(0,r0)r\in(0,r_{0}),

supBg(x,2r)|φ|eCmax(λ2/3,μ2)supBg(x,r)|φ|.\sup_{B_{g}\left(x,2r\right)}|\varphi|\leqslant e^{C\max\left(\lambda^{2/3},\mu^{2}\right)}\sup_{B_{g}\left(x,r\right)}|\varphi|. (3.4)

Secondly, by the application of a L2L^{2}-Carleman estimate on a punctured geodesic ball, taking αCmax(λ2/3,μ2)\alpha\geqslant C\max\left(\lambda^{2/3},\mu^{2}\right) to absorb the right hand side terms and by using the arguments of the proof of (1.18) together with (3.4), one is able to prove (3.2) then (3.3). An extra technical difficulty appears in Step 4 and consequently in Step 5, because the adding of terms involving LpL^{p}-norms of Δgu\Delta_{g}u in the left hand side introduces terms in function of uu and u\nabla u that needed to be absorbed. This is indeed the case by a bootstrap argument using scaled versions of local elliptic regularity estimates and Gagliardo-Nirenberg interpolations inequalities applied to uu and u\nabla u.

Note that (3.2) and (3.3) are crucially related to the vanishing order estimate (3.4) that is itself related to the possible rate of decay to the solutions of second order differential inequalities in the Euclidean space d\mathbb{R}^{d} inspired by the so-called Landis conjecture [KL88]. According to the Meshkov type counterxamples to the Landis conjecture for complex-valued functions [Mes91] and [Dav14], (3.4) is probably sharp so are (3.2) and (3.3) if we consider φC(M;)\varphi\in C^{\infty}(M;\mathbb{C}). But, one can probably sharpen (3.2) and (3.3) when d=2d=2, assuming that φ\varphi is real-valued by using the recent paper on vanishing order estimates of real-valued solutions to second order elliptic equations in the plane, [LMNN20] or [LBS23]. When d3d\geqslant 3, and assuming that φ\varphi is real-valued, we do not know if one can sharpen (3.2) and (3.3) but in that direction, one can read the very recent preprint [FK24] that constructs a real-valued counterexample to the quantitative Landis conjecture in d\mathbb{R}^{d} for d4d\geqslant 4.

3.3 Solutions to elliptic equations with bounded doubling index

Let us take a matrix A=(aij(x))1i,jdA=(a^{ij}(x))_{1\leqslant i,j\leqslant d} symmetric, uniformly elliptic, with Lipschitz entries, that is satisfying (1.15). The lower order terms are given by W=W(x)L(B2;d)W=W(x)\in L^{\infty}(B_{2};\mathbb{C}^{d}) and VL(B2;)V\in L^{\infty}(B_{2};\mathbb{C}) satisfying

|W(x)|+|V(x)|Λ3,xB2,|W(x)|+|V(x)|\leqslant\Lambda_{3},\qquad x\in B_{2},

for some Λ3>0\Lambda_{3}>0.

We have the following generalization of 1.3.

Theorem 3.3.

There exist r0,C>0r_{0},C>0 depending on A,W,VA,W,V such that for every uHloc1(B2)L(B2)u\in H_{\text{loc}}^{1}(B_{2})\cap L^{\infty}(B_{2}) satisfying

div(A(x)u)+Wu+Vu=0inB2,-\mathrm{div}(A(x)\nabla u)+W\cdot\nabla u+Vu=0\ \text{in}\ B_{2},

and

Nu(B(0,1))N,N1,N_{u}(B(0,1))\leqslant N,\qquad N\geqslant 1,

then for every r(0,r0)r\in(0,r_{0}),

supB(0,r(1+1N))|u|CsupB(0,r)|u|,\sup_{B\left(0,r\left(1+\frac{1}{N}\right)\right)}|u|\leqslant C\sup_{B\left(0,r\right)}|u|,

and

supB(0,r)|u|CNrsupB(0,r)|u|.\sup_{B\left(0,r\right)}|\nabla u|\leqslant C\frac{N}{r}\sup_{B\left(0,r\right)}|u|.

The proof of 3.3 follows the lines of the one of 1.3, the lower order terms Wu+VuW\cdot\nabla u+Vu are absorbed by the use of the Carleman parameter αCW2+CV2/3\alpha\geqslant C\|W\|_{\infty}^{2}+C\|V\|_{\infty}^{2/3}. Again Steps 4 and 5 are modified to include a bootstrap argument applied to u\nabla u.

3.4 Linear combination of eigenfunctions

An interesting open problem is the generalization of local LL^{\infty}-Bernstein estimates for linear combination of eigenfunctions. While global LL^{\infty}-Bernstein estimates (1.4) have been established, it seems that its local counterpart has not been investigated yet. Let us recall that from [JL99, Theorem 14.3], the following result holds.

\bullet There exist r0,C>0r_{0},C>0 depending only on MM, such that for every linear combination of Laplace eigenfunctions

ΦΛ=λkΛakφλk,ak,Λ1,withΔgφλk=λkφλkinM,\Phi_{\Lambda}=\sum_{\lambda_{k}\leqslant\Lambda}a_{k}\varphi_{\lambda_{k}},\qquad a_{k}\in\mathbb{C},\ \Lambda\geqslant 1,\ \text{with}\ -\Delta_{g}\varphi_{\lambda_{k}}=\lambda_{k}\varphi_{\lambda_{k}}\ \text{in}\ M, (3.5)

then for every xMx\in M, r(0,r0)r\in(0,r_{0}),

supBg(x,2r)|ΦΛ|eCΛsupBg(x,r)|ΦΛ|.\sup_{B_{g}\left(x,2r\right)}|\Phi_{\Lambda}|\leqslant e^{C\sqrt{\Lambda}}\sup_{B_{g}\left(x,r\right)}|\Phi_{\Lambda}|. (3.6)

It is then natural to conjecture the following result.

Conjecture 3.4.

There exist r0,C>0r_{0},C>0 depending only on MM such that for every linear combination of Laplace eigenfunctions ΦΛ\Phi_{\Lambda}, that is satisfying (3.5), the following Bernstein estimates hold

supBg(x,r(1+1Λ))|Φ|CsupBg(x,r)|Φ|,\sup_{B_{g}\left(x,r\left(1+\frac{1}{\sqrt{\Lambda}}\right)\right)}|\Phi|\leqslant C\sup_{B_{g}\left(x,r\right)}|\Phi|, (3.7)

and

supBg(x,r)|Φ|CΛrsupBg(x,r)|Φ|.\sup_{B_{g}\left(x,r\right)}|\nabla\Phi|\leqslant C\frac{\sqrt{\Lambda}}{r}\sup_{B_{g}\left(x,r\right)}|\Phi_{|}. (3.8)

The main difficulty for obtaining 3.4 comes from the fact that Φ\Phi does not satisfy an elliptic equation. The standard trick to remove this difficulty consists in setting

Φ~(x,t)=λkΛaksinh(λkt)λkφλk(x,t)M×,\tilde{\Phi}(x,t)=\sum_{\lambda_{k}\leqslant\Lambda}a_{k}\frac{\sinh(\sqrt{\lambda_{k}t})}{\sqrt{\lambda_{k}}}\varphi_{\lambda_{k}}\qquad(x,t)\in M\times\mathbb{R}, (3.9)

because ΔΦ~\Delta\tilde{\Phi} is harmonic in M×M\times\mathbb{R}, with respect to the metric g~=gdt\tilde{g}=g\otimes\mathrm{dt}. This transformation (3.9) is crucially used for proving (3.6). A first attempt for proving (3.7) would be to adapt the proof of (1.18) then (3.8) with a boundary L2L^{2}-Carleman-type inequality in the spirit of [JL99, Lemma 14.5]. Indeed, this boundary type estimate is useful for deducing an estimate of Φ\Phi from an estimate of Φ~\tilde{\Phi}.

Appendix A Proof of the Carleman estimates

In this part, we introduce standard notations inspired by Riemannian geometry. We do this because it simplifies the formulas appearing in the proof of the next lemmas. We set

if=fxi,xf=(1f,,df),f=Axf,div(ξ)=i=1diξiandΔf=div(f),\partial_{i}f=\frac{\partial f}{\partial x_{i}},\ \nabla_{x}f=(\partial_{1}f,\dots,\partial_{d}f),\ \nabla f=A\nabla_{x}f,\ \mathrm{div}(\xi)=\sum_{i=1}^{d}\partial_{i}\xi_{i}\ \text{and}\ \Delta f=\mathrm{div}(\nabla f),

Let A1=(aij(x))1i,jdA^{-1}=(a_{ij}(x))_{1\leqslant i,j\leqslant d} the inverse matrix of coefficients of AA.

For two vector fields ξ\xi and η\eta, we have

ξη=i,j=1daij(x)ξiηj,|ξ|2=ξξ,\xi\cdot\eta=\sum_{i,j=1}^{d}a_{ij}(x)\xi_{i}\eta_{j},\ |\xi|^{2}=\xi\cdot\xi, (A.1)

With these notations at hand, when ff, hh are smooth compactly supported functions, we have

div(A(x)f)=Δf,Δ(f2)=2fΔf+2|f|2,fΔh𝑑x=hΔf𝑑x=fhdx.\mathrm{div}(A(x)\nabla f)=\Delta f,\ \Delta(f^{2})=2f\Delta f+2|\nabla f|^{2},\ \int f\Delta hdx=\int h\Delta fdx=-\int\nabla f\cdot\nabla hdx.

The proof of 2.1 will follow the lines of the one of [EV03, Theorem 2], by keeping in the left-hand-side the whole anti-symmetric term of the conjugated operator, see A.1 for a precise formulation. Indeed, this term will then be exploited to give a more direct proof of 2.2 than the one of [DF90a, Lemma A].

A.1 The standard Carleman estimate

The main result of this section is the following Carleman estimate, that leads to 2.1.

Lemma A.1.

There exists a positive constant C=C(A)>0C=C(A)>0, a radial increasing function w=w(r)w=w(r) for 0<ρ<20<\rho<2 satisfying

C1w(ρ)ρC,C1|ρw(ρ)|Cρ(0,2),C^{-1}\leqslant\frac{w(\rho)}{\rho}\leqslant C,\ C^{-1}\leqslant|\partial_{\rho}w(\rho)|\leqslant C\ \qquad\forall\rho\in(0,2), (A.2)

such that for every αC\alpha\geqslant C, fCc(B2{0})f\in C_{c}^{\infty}(B_{2}\setminus\{0\}), the following estimate holds

α3B2w12α|f|2𝑑x+αB2w12α|xf|2𝑑x+α2B2|w|2w2|𝒜(g)|2𝑑xCB2w22α|Δf|2𝑑x,\alpha^{3}\int_{B_{2}}w^{-1-2\alpha}|f|^{2}dx+\alpha\int_{B_{2}}w^{1-2\alpha}|\nabla_{x}f|^{2}dx+\alpha^{2}\int_{B_{2}}\frac{|\nabla w|^{2}}{w^{2}}|\mathcal{A}(g)|^{2}dx\\ \leqslant C\int_{B_{2}}w^{2-2\alpha}|\Delta f|^{2}dx, (A.3)

where g=wαfg=w^{-\alpha}f and

𝒜(g)=wwg|w|2+12FwAg,FwA=wΔw|w|2|w|2,\mathcal{A}(g)=\frac{w\nabla w\cdot\nabla g}{|\nabla w|^{2}}+\frac{1}{2}F_{w}^{A}g,\qquad F_{w}^{A}=\frac{w\Delta w-|\nabla w|^{2}}{|\nabla w|^{2}}, (A.4)

together with the bound

|FwA|C.|F_{w}^{A}|\leqslant C. (A.5)
Proof.

We follow [EV03, Theorem 2], sometimes line by line.

Let g=wαfg=w^{-\alpha}f and we compute

wαΔf=Δg+α2|w|2w2g+2α|w|2w2𝒜(g),w^{-\alpha}\Delta f=\Delta g+\frac{\alpha^{2}|\nabla w|^{2}}{w^{2}}g+2\alpha\frac{|\nabla w|^{2}}{w^{2}}\mathcal{A}(g), (A.6)

where the definition of 𝒜(g)\mathcal{A}(g) is recalled in (A.4). We also set MwAM_{w}^{A} the d×dd\times d symmetric matrix

MwA=12(Mij+Mji)1i,jd,M_{w}^{A}=\frac{1}{2}(M_{ij}+M_{ji})_{1\leqslant i,j\leqslant d}, (A.7)

where, using the summation notation of repeated indices,

Mij=12div(ww|w|2)δijxj(waikxkw|w|2)+12aikwakllw|w|2kahi12FwAδij.M_{ij}=\frac{1}{2}\mathrm{div}\left(\frac{w\nabla w}{|\nabla w|^{2}}\right)\delta_{ij}-\partial_{x_{j}}\left(\frac{wa^{ik}\partial_{x_{k}}w}{|\nabla w|^{2}}\right)+\frac{1}{2}a_{ik}\frac{wa^{kl}\partial_{l}w}{|\nabla w|^{2}}\partial_{k}a^{hi}-\frac{1}{2}F_{w}^{A}\delta_{ij}. (A.8)

We split the proof in several steps.

Step 1: A first identity. The goal of this step is to prove that

w2|w|2(wαΔf)24αMwAgg+αFwAΔ(g2)+4α2|w|2w2𝒜(g)2.\int\frac{w^{2}}{|\nabla w|^{2}}(w^{-\alpha}\Delta f)^{2}\geqslant 4\alpha\int M_{w}^{A}\nabla g\cdot\nabla g+\alpha\int F_{w}^{A}\Delta(g^{2})+4\alpha^{2}\int\frac{|\nabla w|^{2}}{w^{2}}\mathcal{A}(g)^{2}. (A.9)

Let

J1=(2α|w|w𝒜(g))2,J2=2[2α𝒜(g)(Δg+α2|w|2w2g)].J_{1}=\int\left(2\alpha\frac{|\nabla w|}{w}\mathcal{A}(g)\right)^{2},\ J_{2}=2\int\left[2\alpha\mathcal{A}(g)\left(\Delta g+\alpha^{2}\frac{|\nabla w|^{2}}{w^{2}}g\right)\right]. (A.10)

Then, we have

w2|w|2(wαΔf)2J1+J2.\int\frac{w^{2}}{|\nabla w|^{2}}(w^{-\alpha}\Delta f)^{2}\geqslant J_{1}+J_{2}. (A.11)

First note that

|w|2w2𝒜(g)g=0,\int\frac{|\nabla w|^{2}}{w^{2}}\mathcal{A}(g)\cdot g=0, (A.12)

so that from (A.4) and the identity Δ(g2)=2gΔg+2|g|2\Delta(g^{2})=2g\Delta g+2|\nabla g|^{2},

J2=4α𝒜(g)Δg=2α(2wwg|w|2ΔgFwA|g|2)+αFwΔ(g2).J_{2}=4\alpha\int\mathcal{A}(g)\Delta g=2\alpha\int\left(\frac{2w\nabla w\cdot\nabla g}{|\nabla w|^{2}}\Delta g-F_{w}^{A}|\nabla g|^{2}\right)+\alpha\int F_{w}\Delta(g^{2}). (A.13)

We now have the following identity

2(βg)Δg=2div((βg)g)div(β|g|2)+div(β)|g|22iβkaijjgkg+βkkaijigjg,βC(B2{0};d),2(\beta\cdot\nabla g)\Delta g=2\mathrm{div}((\beta\cdot\nabla g)\nabla g)-\mathrm{div}(\beta|\nabla g|^{2})+\mathrm{div}(\beta)|\nabla g|^{2}\\ -2\partial_{i}\beta^{k}a^{ij}\partial_{j}g\partial_{k}g+\beta^{k}\partial_{k}a^{ij}\partial_{i}g\partial_{j}g,\qquad\forall\beta\in C^{\infty}(B_{2}\setminus\{0\};\mathbb{R}^{d}),

and we choose

β=ww|w|2,\beta=\frac{w\nabla w}{|\nabla w|^{2}},

to get from (A.13) and the divergence theorem

4α𝒜(g)Δg=4αMwAgg+αFwAΔ(g2),4\alpha\int\mathcal{A}(g)\Delta g=4\alpha\int M_{w}^{A}\nabla g\cdot\nabla g+\alpha\int F_{w}^{A}\Delta(g^{2}), (A.14)

where MwAM_{w}^{A} is defined in (A.7) and (A.8). By gathering (A.10), (A.11), (A.12) and (A.14) we obtain (A.9) so the conclusion of Step 1.

Step 2: Choice of ww. For μ1\mu\geqslant 1, a parameter to be chosen later, let

σ(x)=(i,j=1daij(0)xixj)1/2,φ(s)=sexp(0seμt1t𝑑t),ϕ(s)=φ(s)sφ(s)=eμs.\sigma(x)=\left(\sum_{i,j=1}^{d}a_{ij}(0)x_{i}x_{j}\right)^{1/2},\ \varphi(s)=s\exp\left(\int_{0}^{s}\frac{e^{-\mu t}-1}{t}dt\right),\ \phi(s)=\frac{\varphi(s)}{s\varphi^{\prime}(s)}=e^{\mu s}. (A.15)

We define

w(x)=φ(σ(x)).w(x)=\varphi(\sigma(x)).

With this definition, we can now compute

MσAσ=0,M_{\sigma}^{A}\nabla\sigma=0, (A.16)
FwA=FσAϕ(σ)σϕ(σ),MwA=ϕ(σ)MσA+σϕ(σ)(Iσσ|σ|2),F_{w}^{A}=F_{\sigma}^{A}\phi(\sigma)-\sigma\phi^{\prime}(\sigma),\ M_{w}^{A}=\phi(\sigma)M_{\sigma}^{A}+\sigma\phi^{\prime}(\sigma)\left(I-\frac{\nabla\sigma\otimes\nabla\sigma}{|\nabla\sigma|^{2}}\right), (A.17)

and

FσA(0)=n2,MσA(0)=0.F_{\sigma}^{A(0)}=n-2,\ M_{\sigma}^{A(0)}=0. (A.18)

Notice that the following properties hold

φ(r)>0,crφ(r)Cr,\varphi^{\prime}(r)>0,\ cr\leqslant\varphi(r)\leqslant Cr,

so we have

cσw(x)Cσ,c|w(x)|C,||w|2|C,|Δϕ|Cw1,|Fw|C.c\sigma\leqslant w(x)\leqslant C\sigma,\ c\leqslant|\nabla w(x)|\leqslant C,\ |\nabla|\nabla w|^{2}|\leqslant C,\ |\Delta\phi|\leqslant Cw^{-1},\ |F_{w}|\leqslant C.

Now we estimate the first two terms appearing in the right hand side of (A.9).

Let us treat the first term. From the second part of (A.17), we have

MwAgg=σϕ(|g|2(σg)2|σ|2),M_{w}^{A}\nabla g\cdot\nabla g=\sigma\phi^{\prime}\left(|\nabla g|^{2}-\frac{(\nabla\sigma\cdot\nabla g)^{2}}{|\nabla\sigma|^{2}}\right),

and denoting by ~g\tilde{\nabla}g the tangential components of the gradient of gg along the level sets of σ(x)\sigma(x) with respect to the metric i,j=1daij(x)dxidxj\sum_{i,j=1}^{d}a_{ij}(x)dx_{i}dx_{j}, we have

~g=gσg|σ|2σ=gwg|w|2w.\tilde{\nabla}g=\nabla g-\frac{\nabla\sigma\cdot\nabla g}{|\nabla\sigma|^{2}}\nabla\sigma=\nabla g-\frac{\nabla w\cdot\nabla g}{|\nabla w|^{2}}\nabla w. (A.19)

From (A.16), (A.18) and (A.19), we have that

MσAgg=(MσAMσA(0))~g~g.M_{\sigma}^{A}\nabla g\cdot\nabla g=(M_{\sigma}^{A}-M_{\sigma}^{A(0)})\tilde{\nabla}g\cdot\tilde{\nabla}g.

On the other hand, a computation and the Lipschitz condition on the matrix AA give that there exists C>0C>0 depending on on dd and Λ1,Λ2\Lambda_{1},\Lambda_{2} such that

|MσAMσA(0)|Cσ,|M_{\sigma}^{A}-M_{\sigma}^{A(0)}|\leqslant C\sigma,

so that

MσAggσ(ϕCϕ)|~g|2\int M_{\sigma}^{A}\nabla g\cdot\nabla g\geqslant\int\sigma(\phi^{\prime}-C\phi)|\tilde{\nabla}g|^{2} (A.20)

Now we estimate from below the second term appearing in (A.9). We observe that

FwA=(d2)ϕ+(Bϕσϕ),F_{w}^{A}=(d-2)\phi+(B\phi-\sigma\phi^{\prime}),

where B=FσAFσA(0)B=F_{\sigma}^{A}-F_{\sigma}^{A(0)} that satisfies by using the Lipschitz condition on the matrix AA,

|B(x)|Cσ.|B(x)|\leqslant C\sigma.

So we have from the identities

Δg2=2gΔg+2|g|2,|g|2=|~g|2+(wg)2|w|2,\Delta g^{2}=2g\Delta g+2|\nabla g|^{2},\ |\nabla g|^{2}=|\tilde{\nabla}g|^{2}+\frac{(\nabla w\cdot\nabla g)^{2}}{|\nabla w|^{2}},

that

FwAΔ(g2)\displaystyle\int F_{w}^{A}\Delta(g^{2}) =(d2)(Δϕ)g2+2(Bϕσϕ)gΔg\displaystyle=(d-2)\int(\Delta\phi)g^{2}+2\int(B\phi-\sigma\phi^{\prime})g\Delta g
+2(Bϕσϕ)|~g|2+2(Bϕσϕ)(σg)2|σ|2.\displaystyle\quad+2\int(B\phi-\sigma\phi^{\prime})|\tilde{\nabla}g|^{2}+2\int(B\phi-\sigma\phi^{\prime})\frac{(\nabla\sigma\cdot\nabla g)^{2}}{|\nabla\sigma|^{2}}. (A.21)

From (A.6), we then have that the second term of (A.1) writes as follows

2(Bϕσϕ)gΔg=2(Bϕσϕ)gΔfwα+2α2(σϕBϕ)|w|2w2g2+2(σϕBϕ)2α|w|2w2g𝒜(g).2\int(B\phi-\sigma\phi^{\prime})g\Delta g=2\int(B\phi-\sigma\phi^{\prime})g\Delta fw^{-\alpha}\\ +2\alpha^{2}\int(\sigma\phi^{\prime}-B\phi)\frac{|\nabla w|^{2}}{w^{2}}g^{2}+2\int(\sigma\phi^{\prime}-B\phi)2\alpha\frac{|\nabla w|^{2}}{w^{2}}g\mathcal{A}(g). (A.22)

Thus, from (A.20), (A.1) and (A.22) we have for α1\alpha\geqslant 1,

4αMwAgg+αFwΔ(g2)2α(σϕ2Cσϕ+Bϕ)|~g|2+2α3(σϕBϕ)|w|2w2g2R1,4\alpha\int M_{w}^{A}\nabla g\cdot\nabla g+\alpha\int F_{w}\Delta(g^{2})\\ \geqslant 2\alpha\int(\sigma\phi^{\prime}-2C\sigma\phi+B\phi)|\tilde{\nabla}g|^{2}+2\alpha^{3}\int(\sigma\phi^{\prime}-B\phi)\frac{|\nabla w|^{2}}{w^{2}}g^{2}-R_{1}, (A.23)

where for some constant C>0C>0 depending on dd, Λ1,Λ2\Lambda_{1},\Lambda_{2} and μ\mu,

R1C(αw1g2+αw1α|g||Δg|+α2w1|𝒜(g)||g|+αw|σg|2|σ|2).R_{1}\leqslant C\left(\alpha\int w^{-1}g^{2}+\alpha\int w^{1-\alpha}|g||\Delta g|+\alpha^{2}\int w^{-1}|\mathcal{A}(g)||g|+\alpha\int w\frac{|\nabla\sigma\cdot\nabla g|^{2}}{|\nabla\sigma|^{2}}\right). (A.24)

By choosing μ=C(d,Λ1,Λ2)1\mu=C(d,\Lambda_{1},\Lambda_{2})\geqslant 1 sufficiently large, we then get that for some c=c(d,Λ1,Λ2,μ)>0c=c(d,\Lambda_{1},\Lambda_{2},\mu)>0 by using (A.23),

4αMwAgg+αFwΔ(g2)c(ασ|~g|2+2α3σ|w|2w2g2)R14\alpha\int M_{w}^{A}\nabla g\cdot\nabla g+\alpha\int F_{w}\Delta(g^{2})\\ \geqslant c\left(\alpha\int\sigma|\tilde{\nabla}g|^{2}+2\alpha^{3}\int\sigma\frac{|\nabla w|^{2}}{w^{2}}g^{2}\right)-R_{1} (A.25)

Once we combine (A.9), (A.25) and (A.24), we obtain the following

4α2|w|2w2A(g)2+2ασϕ|~g|2+2α3σϕ|w|2w2g2w2|w|2(wαΔf)2+C(αw1g2+αw1α|g||Δg|+α2w1|𝒜(g)||g|+αw|σg|2|σ|2).4\alpha^{2}\int\frac{|\nabla w|^{2}}{w^{2}}A(g)^{2}+2\alpha\int\sigma\phi^{\prime}|\tilde{\nabla}g|^{2}+2\alpha^{3}\int\sigma\phi^{\prime}\frac{|\nabla w|^{2}}{w^{2}}g^{2}\leqslant\int\frac{w^{2}}{|\nabla w|^{2}}(w^{-\alpha}\Delta f)^{2}\\ +C\left(\alpha\int w^{-1}g^{2}+\alpha\int w^{1-\alpha}|g||\Delta g|+\alpha^{2}\int w^{-1}|\mathcal{A}(g)||g|+\alpha\int w\frac{|\nabla\sigma\cdot\nabla g|^{2}}{|\nabla\sigma|^{2}}\right). (A.26)

This ends Step 2.

Step 3: Absorption. To conclude the proof, recall the form of 𝒜(g)\mathcal{A}(g) and then

|FwA|C.|F_{w}^{A}|\leqslant C. (A.27)

Thus

w|w|2|wg|2|w|2Cw1|𝒜(g)|2+C|g|2w,\frac{w}{|\nabla w|^{2}}\frac{|\nabla w\cdot\nabla g|^{2}}{|\nabla w|^{2}}\leqslant Cw^{-1}|\mathcal{A}(g)|^{2}+C\frac{|g|^{2}}{w}, (A.28)

so that

αw|σg|2|σ|2=αw|wg|2|σ|2Cαw1|𝒜(g)|2+Cαw1|g|2Cα|w|2w|𝒜(g)|2+Cαw1|g|2.\alpha\int w\frac{|\nabla\sigma\cdot\nabla g|^{2}}{|\nabla\sigma|^{2}}=\alpha\int w\frac{|\nabla w\cdot\nabla g|^{2}}{|\nabla\sigma|^{2}}\leqslant C\alpha\int w^{-1}|\mathcal{A}(g)|^{2}+C\alpha\int w^{-1}|g|^{2}\\ \leqslant C\alpha\int\frac{|\nabla w|^{2}}{w}|\mathcal{A}(g)|^{2}+C\alpha\int w^{-1}|g|^{2}. (A.29)

Also,

Cα2w1|𝒜(g)||g|c2α2|w|2w|𝒜(g)|2+Cα2w1g2,C\alpha^{2}\int w^{-1}|\mathcal{A}(g)||g|\leqslant\frac{c}{2}\alpha^{2}\int\frac{|\nabla w|^{2}}{w}|\mathcal{A}(g)|^{2}+C\alpha^{2}\int w^{-1}g^{2}, (A.30)

and

αw(1α)|g||Δf|w2(wα|Δf|)2+Cα2w1g2.\alpha\int w^{(1-\alpha)}|g||\Delta f|\leqslant\int w^{2}(w^{-\alpha}|\Delta f|)^{2}+C\alpha^{2}\int w^{-1}g^{2}. (A.31)

From (A.26), (A.29), (A.30) and (A.31), conjugated with

α2|w|2w|A(g)|2cα2|w|2w|𝒜(g)|2,\alpha^{2}\int\frac{|\nabla w|^{2}}{w}|A(g)|^{2}\geqslant c\alpha^{2}\int\frac{|\nabla w|^{2}}{w}|\mathcal{A}(g)|^{2},

we obtain the expected inequality (A.3) and the bound (A.5) comes from (A.27). This concludes the proof. ∎

A.2 The L2L^{2}-Carleman estimate in a punctured domain

The goal of this part consists in proving 2.2. In the proof we assume that A(0)=IdA(0)=I_{d}. Note that this is not a restriction because by a local change of variables we can drop the assumption that A(0)=IdA(0)=I_{d}, replacing balls by ellipses, technical details are omitted here for an accurate argument.

Proof.

We start from (2.1) to obtain first

α2B(0,r(1+21α1))|w|2w2|𝒜(g)|2𝑑xCB2w22α|Δf|2𝑑x.\alpha^{2}\int_{B(0,r(1+2^{-1}\alpha^{-1}))}\frac{|\nabla w|^{2}}{w^{2}}|\mathcal{A}(g)|^{2}dx\leqslant C\int_{B_{2}}w^{2-2\alpha}|\Delta f|^{2}dx.

By using (A.2) and the assumption on ff that vanishes in B(0,r)B(0,r), this translates into

α2r2B(0,r(1+21α1))|𝒜(g)|2𝑑xCB2w22α|Δf|2𝑑x.\alpha^{2}r^{-2}\int_{B(0,r(1+2^{-1}\alpha^{-1}))}|\mathcal{A}(g)|^{2}dx\leqslant C\int_{B_{2}}w^{2-2\alpha}|\Delta f|^{2}dx.

Then we develop |𝒜(g)|2|\mathcal{A}(g)|^{2} by (A.4) to get that

α2r2B(0,r(1+21α1))(wwg|w|2)2CB2w22α|Δf|2𝑑x+Cα2r2B(0,r(1+21α1))|FwAg|2.\alpha^{2}r^{-2}\int_{B(0,r(1+2^{-1}\alpha^{-1}))}\left(\frac{w\nabla w\cdot\nabla g}{|\nabla w|^{2}}\right)^{2}\leqslant C\int_{B_{2}}w^{2-2\alpha}|\Delta f|^{2}dx+C\alpha^{2}r^{-2}\int_{B(0,r(1+2^{-1}\alpha^{-1}))}|F_{w}^{A}g|^{2}.

By using that A(0)=IdA(0)=I_{d}, the definition of σ\sigma in (A.15), the definition of the scalar product between two vector fields in (A.1) we have that

wg=A(x|ρ|)Axg=x|ρ|A2xg=ρg+x|ρ|(A2(x)Id)xg\nabla w\cdot\nabla g=A(\nabla_{x}|\rho|)\cdot A\nabla_{x}g=\nabla_{x}|\rho|\cdot A^{2}\nabla_{x}g=\partial_{\rho}g+\nabla_{x}|\rho|\cdot(A^{2}(x)-I_{d})\nabla_{x}g

By using the fact that AA is Lipschitz, i.e. the assumption (1.15), we deduce from the two previous estimates that

α2B(0,r(1+21α1))|ρg|2CB2w22α|Δf|2𝑑x+Cα2r2B(0,r(1+21α1))|FwAg|2+Cα2r2B(0,r(1+21α1))|g|2.\alpha^{2}\int_{B(0,r(1+2^{-1}\alpha^{-1}))}|\partial_{\rho}g|^{2}\\ \leqslant C\int_{B_{2}}w^{2-2\alpha}|\Delta f|^{2}dx+C\alpha^{2}r^{-2}\int_{B(0,r(1+2^{-1}\alpha^{-1}))}|F_{w}^{A}g|^{2}\\ +C\alpha^{2}r^{2}\int_{B(0,r(1+2^{-1}\alpha^{-1}))}|\nabla g|^{2}. (A.32)

By integrating along a radial line and by using that gg vanishes in B(0,r)B(0,r), we obtain that

α4r2B(0,r(1+2α1))|g|2Cα2B(0,r(1+21α1))|ρg|2,\alpha^{4}r^{-2}\int_{B(0,r(1+2\alpha^{-1}))}|g|^{2}\leqslant C\alpha^{2}\int_{B(0,r(1+2^{-1}\alpha^{-1}))}|\partial_{\rho}g|^{2},

so from the two previous estimates

α4r2B(0,r(1+2α1))|g|2CB2w22α|Δf|2𝑑x+Cα2r2B(0,r(1+21α1))|FwAg|2+Cα2r2B(0,r(1+21α1))|g|2.\alpha^{4}r^{-2}\int_{B(0,r(1+2\alpha^{-1}))}|g|^{2}\\ \leqslant C\int_{B_{2}}w^{2-2\alpha}|\Delta f|^{2}dx+C\alpha^{2}r^{-2}\int_{B(0,r(1+2^{-1}\alpha^{-1}))}|F_{w}^{A}g|^{2}+C\alpha^{2}r^{2}\int_{B(0,r(1+2^{-1}\alpha^{-1}))}|\nabla g|^{2}.

Then, for αC\alpha\geqslant C, using (A.5) one can absorb the second right hand side term to obtain

α4r2B(0,r(1+2α1))|g|2CB2w22α|Δf|2𝑑x+Cα2r2B(0,r(1+21α1))|g|2.\alpha^{4}r^{-2}\int_{B(0,r(1+2\alpha^{-1}))}|g|^{2}\leqslant C\int_{B_{2}}w^{2-2\alpha}|\Delta f|^{2}dx+C\alpha^{2}r^{2}\int_{B(0,r(1+2^{-1}\alpha^{-1}))}|\nabla g|^{2}.

We now come back to the variable ff to deduce

α4r2B(0,r(1+2α1))|f|2w2αCB2w22α|Δf|2𝑑x+Cα2r2B(0,r(1+21α1))w2α(|f|2+α2|f|2).\alpha^{4}r^{-2}\int_{B(0,r(1+2\alpha^{-1}))}|f|^{2}w^{-2\alpha}\\ \leqslant C\int_{B_{2}}w^{2-2\alpha}|\Delta f|^{2}dx+C\alpha^{2}r^{2}\int_{B(0,r(1+2^{-1}\alpha^{-1}))}w^{-2\alpha}(|\nabla f|^{2}+\alpha^{2}|f|^{2}).

The second term can be absorbed recalling that r(0,r0)r\in(0,r_{0}) with r0>0r_{0}>0 sufficiently small. Then, we add to the left hand side

α4r2B(0,r(1+2α1))|f|2w2α+B(0,r(1+2α1))w22α|Δf|2𝑑xCB2w22α|Δf|2𝑑x+Cα2r2B(0,r(1+21α1))w2α|f|2.\alpha^{4}r^{-2}\int_{B(0,r(1+2\alpha^{-1}))}|f|^{2}w^{-2\alpha}+\int_{B(0,r(1+2\alpha^{-1}))}w^{2-2\alpha}|\Delta f|^{2}dx\\ \leqslant C\int_{B_{2}}w^{2-2\alpha}|\Delta f|^{2}dx+C\alpha^{2}r^{2}\int_{B(0,r(1+2^{-1}\alpha^{-1}))}w^{-2\alpha}|\nabla f|^{2}.

So, by a L2L^{2} local elliptic regularity estimate, using that w2αw^{-2\alpha} has the same amplitude in the annulus B(0,r(1+2α1))B(0,r)B(0,r(1+2\alpha^{-1}))\setminus B(0,r), recalling that ff vanishes in B(0,r)B(0,r), we then obtain

α4r2B(0,r(1+2α1))|f|2w2α+α2B(0,r(1+α1))w2α|xf|2𝑑xCB2w22α|Δf|2𝑑x+Cα2r2B(0,r(1+21α1))w2α|f|2.\alpha^{4}r^{-2}\int_{B(0,r(1+2\alpha^{-1}))}|f|^{2}w^{-2\alpha}+\alpha^{2}\int_{B(0,r(1+\alpha^{-1}))}w^{-2\alpha}|\nabla_{x}f|^{2}dx\\ \leqslant C\int_{B_{2}}w^{2-2\alpha}|\Delta f|^{2}dx+C\alpha^{2}r^{2}\int_{B(0,r(1+2^{-1}\alpha^{-1}))}w^{-2\alpha}|\nabla f|^{2}.

Therefore, one can absorb the last right hand side term recalling that r(0,r0)r\in(0,r_{0}) with r0>0r_{0}>0 sufficiently small to get

α4r2B(0,r(1+2α1))|f|2w2α+α2B(0,r(1+α1))w2α|xf|2𝑑xCB2w22α|Δf|2𝑑x.\alpha^{4}r^{-2}\int_{B(0,r(1+2\alpha^{-1}))}|f|^{2}w^{-2\alpha}+\alpha^{2}\int_{B(0,r(1+\alpha^{-1}))}w^{-2\alpha}|\nabla_{x}f|^{2}dx\\ \leqslant C\int_{B_{2}}w^{2-2\alpha}|\Delta f|^{2}dx. (A.33)

By gathering (2.1) and (A.33) we get the expected result (2.2). ∎

A.3 The modified version of the standard Carleman estimate

We have the following Carleman estimate whose proof is an easy adaptation of the one of A.1.

Lemma A.2.

There exists a positive constant C=C(A)>0C=C(A)>0 such that for every x0B(0,1/4)x_{0}\in B(0,1/4), there exists an increasing function ww such that

C1w(|xx0|)|xx0|C,C^{-1}\leqslant\frac{w(|x-x_{0}|)}{|x-x_{0}|}\leqslant C,

such that for every αC\alpha\geqslant C, fCc(B2{x0})f\in C_{c}^{\infty}(B_{2}\setminus\{x_{0}\}), the following estimate holds

α3B2w12α|f|2𝑑x+αB2w12α|xf|2𝑑xCB2w22α|Δf|2𝑑x.\alpha^{3}\int_{B_{2}}w^{-1-2\alpha}|f|^{2}dx+\alpha\int_{B_{2}}w^{1-2\alpha}|\nabla_{x}f|^{2}dx\leqslant C\int_{B_{2}}w^{2-2\alpha}|\Delta f|^{2}dx.
Proof.

The proof exactly follows the lines of A.1. Only the choice of ww, coming from (A.15), is different. We instead set

σ(x)=(i,j=1daij(x0)(xix0,i)(xjx0,j))1/2.\sigma(x)=\left(\sum_{i,j=1}^{d}a_{ij}(x_{0})(x_{i}-x_{0,i})(x_{j}-x_{0,j})\right)^{1/2}.

Then the next estimates are established in function of ww that behaves as |xx0||x-x_{0}|. ∎

References

  • [BM23] Nicolas Burq and Iván Moyano. Propagation of smallness and control for heat equations. J. Eur. Math. Soc. (JEMS), 25(4):1349–1377, 2023.
  • [Dav14] Blair Davey. Some quantitative unique continuation results for eigenfunctions of the magnetic Schrödinger operator. Comm. Partial Differential Equations, 39(5):876–945, 2014.
  • [DF88] Harold Donnelly and Charles Fefferman. Nodal sets of eigenfunctions on Riemannian manifolds. Invent. Math., 93(1):161–183, 1988.
  • [DF90a] Harold Donnelly and Charles Fefferman. Growth and geometry of eigenfunctions of the Laplacian. In Analysis and partial differential equations, volume 122 of Lecture Notes in Pure and Appl. Math., pages 635–655. Dekker, New York, 1990.
  • [DF90b] Harold Donnelly and Charles Fefferman. Nodal sets of eigenfunctions: Riemannian manifolds with boundary. In Analysis, et cetera, pages 251–262. Academic Press, Boston, MA, 1990.
  • [DM23] Stefano Decio and Eugenia Malinnikova. On a Bernstein inequality for eigenfunctions. arXiv eprint: 2208.10541, 2023.
  • [Don92] Rui-Tao Dong. Nodal sets of eigenfunctions on Riemann surfaces. J. Differential Geom., 36(2):493–506, 1992.
  • [Don95] Rui-Tao Dong. A Bernstein type of inequality for eigenfunctions. J. Differential Geom., 42(1):23–29, 1995.
  • [EV03] Luis Escauriaza and Sergio Vessella. Optimal three cylinder inequalities for solutions to parabolic equations with Lipschitz leading coefficients. In Inverse problems: theory and applications (Cortona/Pisa, 2002), volume 333 of Contemp. Math., pages 79–87. Amer. Math. Soc., Providence, RI, 2003.
  • [FK24] Nikolai D. Filonov and Stanislav T. Krymskii. On the Landis conjecture in a cylinder, 2024.
  • [FM10] Frank Filbir and Hrushikesh Narhar Mhaskar. A quadrature formula for diffusion polynomials corresponding to a generalized heat kernel. J. Fourier Anal. Appl., 16(5):629–657, 2010.
  • [GT01] David Gilbarg and Neil S. Trudinger. Elliptic partial differential equations of second order. Classics in Mathematics. Springer-Verlag, Berlin, 2001. Reprint of the 1998 edition.
  • [IO22] Rafik Imekraz and El Maati Ouhabaz. Bernstein inequalities via the heat semigroup. Math. Ann., 382(1-2):783–819, 2022.
  • [JL99] David Jerison and Gilles Lebeau. Nodal sets of sums of eigenfunctions. In Harmonic analysis and partial differential equations (Chicago, IL, 1996), Chicago Lectures in Math., pages 223–239. Univ. Chicago Press, Chicago, IL, 1999.
  • [KL88] Vladimir A. Kondratev and Evgeni M Landis. Qualitative properties of the solutions of a second-order nonlinear equation. Mat. Sb. (N.S.), 135(177)(3):346–360, 415, 1988.
  • [LBS23] Kévin Le Balc’h and Diego A. Souza. Quantitative unique continuation for real-valued solutions to second order elliptic equations in the plane. arXiv eprint: 2401.00441, 2023.
  • [Lin91] Fang-Hua Lin. Nodal sets of solutions of elliptic and parabolic equations. Comm. Pure Appl. Math., 44(3):287–308, 1991.
  • [LM20] Alexander Logunov and Eugenia Malinnikova. Lecture notes on quantitative unique continuation for solutions of second order elliptic equations. In Harmonic analysis and applications, volume 27 of IAS/Park City Math. Ser., pages 1–33. Amer. Math. Soc., [Providence], RI, [2020] ©2020.
  • [LMNN20] Alexander Logunov, Eugenia Malinnikova, Nikolai Nadirashvili, and Fedor Nazarov. The Landis conjecture on exponential decay. arXiv eprint: 2007.07034, 2020.
  • [Mes91] Viktor Z. Meshkov. On the possible rate of decrease at infinity of the solutions of second-order partial differential equations. Mat. Sb., 182(3):364–383, 1991.
  • [Nir66] Louis Nirenberg. An extended interpolation inequality. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (3), 20:733–737, 1966.
  • [OCP13] Joaquim Ortega-Cerdà and Bharti Pridhnani. Carleson measures and Logvinenko-Sereda sets on compact manifolds. Forum Math., 25(1):151–172, 2013.
  • [QZ19] Hervé Queffélec and Rachid Zarouf. On Bernstein’s inequality for polynomials. Anal. Math. Phys., 9(3):1181–1207, 2019.