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

Calderón problem for the quasilinear conductivity equation in dimension 22

Tony Liimatainen    Ruirui Wu
Abstract

In this paper we prove a uniqueness result for the Calderón problem for the quasilinear conductivity equation on a bounded domain 2\mathbb{R}^{2}. The proof of the result is based on the higher order linearization method, which reduces the problem to showing density of products of solutions to the linearized equation and their gradients. In contrast to the higher dimensional case, the proof involves delicate analysis of the correction terms of Bukhgeim type complex geometric solutions (CGOs), which have only limited decay. To prove our results, we construct suitable families of CGOs whose phase functions have and do not have critical points. We also combine stationary phase analysis with LpL^{p} estimates for the correction terms of the CGOs.

1 Introduction

Let Ω2\Omega\subset\mathbb{R}^{2} be a bounded open domain with smooth boundary Ω\partial\Omega. In this paper we consider a quasilinear conductivity equation of the form

div(γ(x,u,u)u)=\displaystyle\operatorname{div}(\gamma(x,u,\nabla u)\nabla u)= 0,\displaystyle 0, (1.1)
u|Ω=\displaystyle u|_{\partial\Omega}= f,\displaystyle f, (1.2)

where

γ:Ω¯××2\gamma:\overline{\Omega}\times\mathbb{C}\times\mathbb{C}^{2}\rightarrow\mathbb{C}

is CC^{\infty} smooth function. We also assume that the quasilinear conductivity γ\gamma satisfies

  • (a)

    0<γ(,0,0)C(Ω¯)0<\gamma(\,\cdot\,,0,0)\in C^{\infty}(\bar{\Omega})

  • (b)

    The map ×2(ρ,μ)γ(,ρ,μ)\mathbb{C}\times\mathbb{C}^{2}\ni(\rho,\mu)\rightarrow\gamma(\,\cdot\,,\rho,\mu) is holomorphic with values in the Hölder space C1,α(Ω¯)C^{1,\alpha}(\bar{\Omega}) for some α(0,1)\alpha\in(0,1).

With the above assumptions, the boundary value problem (1.1) is well-posed in the following sense. There exists δ>0\delta>0 and C>0C>0 such that for all

fBδ(Ω)={fC2,α(Ω):fC2,α(Ω)}<δf\in B_{\delta}(\partial\Omega)=\{f\in C^{2,\alpha}(\partial\Omega)\ :\ \lVert f\rVert_{C^{2,\alpha}(\partial\Omega)}\}<\delta

there exists a unique solution u=ufC2,α(Ω)u=u_{f}\in C^{2,\alpha}(\Omega) satisfying uC2,α(Ω¯)<Cδ\lVert u\rVert_{C^{2,\alpha}(\overline{\Omega})}<C\delta. The proof of the above fact follows from the Banach implicit function theorem and the fact that the linearization of (1.1) is injective at the constant solution 0. See for example [LLLS21] or [KKU22] for similar proofs. We then define the Dirichlet-to-Neumann map Λγ:Bδ(Ω)C2,α(Ω)\Lambda_{\gamma}:B_{\delta}(\partial\Omega)\to C^{2,\alpha}(\partial\Omega) by

Λγ(f)=(γ(x,u,u)νu)|Ω,\Lambda_{\gamma}(f)=\left.\left(\gamma(x,u,\nabla u)\partial_{\nu}u\right)\right|_{\partial\Omega},

where fBδ(Ω)f\in B_{\delta}(\partial\Omega), u=ufu=u_{f}, and ν\nu is the unit outer normal to Ω\partial\Omega.

We prove the following uniqueness theorem.

Theorem 1.

Let Ω2\Omega\subset\mathbb{R}^{2} be a bounded open set with CC^{\infty} boundary. Assume that γ1,γ2:Ω¯××2\gamma_{1},\gamma_{2}:\bar{\Omega}\times\mathbb{C}\times\mathbb{C}^{2}\rightarrow\mathbb{C} satisfy the assumptions (a) and (b). Assume also that γ1\gamma_{1} and γ2\gamma_{2} agree up to infinite order on the boundary Ω\partial\Omega. Suppose that we have

Λγ1(f)=Λγ2(f),fBδ(Ω).\Lambda_{\gamma_{1}}(f)=\Lambda_{\gamma_{2}}(f),\quad\forall f\in B_{\delta}(\partial\Omega).

Then

γ1=γ2 in Ω¯××2.\gamma_{1}=\gamma_{2}\quad\text{ in }\overline{\Omega}\times\mathbb{C}\times\mathbb{C}^{2}.

The assumption that the unknown quantities in the theorem are known on the boundary, is to avoid proving a standard-like boundary determination results. We refer to [CLLO24, Theorem 3.2] for an example of a related boundary determination result.

To prove Theorem (1), we use the higher order linearization method originating in the elliptic setting from [LLLS21, FO20]. In our case the higher order linearization argument is the same as the one derived in [CFK+21] and is as follows. Denote γ(,0,0)\gamma(\,\cdot\,,0,0) by γ(,0)\gamma(\,\cdot\,,0). By linearization and the uniqueness result for linear conductivity equations, we can first conclude that

γ0():=γ1(,0)=γ2(,0).\displaystyle\gamma_{0}(\,\cdot\,):=\gamma_{1}(\,\cdot\,,0)=\gamma_{2}(\,\cdot\,,0). (1.3)

Then the following can be proven by induction, see [CFK+21]¸: Assume for k=0,1,,m2k=0,1,\ldots,m-2, that

γ1(k)(x,0)=γ2(k)(x,0),xΩ.\gamma_{1}^{(k)}(x,0)=\gamma_{2}^{(k)}(x,0),\quad x\in\Omega.

Then, linearizing the equation (1.1) several times, one obtains

(l1,,lm)π(m)j1,,jm1=02Ω((λj1λjm1γ1)(x,0)(λj1λjm1γ2)(x,0))(v(l1),v(l1))j1(v(lm1),v(lm1))jm1v(lm)v(m+1)dx=0\sum_{\left(l_{1},\ldots,l_{m}\right)\in\pi(m)}\sum_{j_{1},\ldots,j_{m-1}=0}^{2}\int_{\Omega}\left(\left(\partial_{\lambda_{j_{1}}}\ldots\partial_{\lambda_{j_{m-1}}}\gamma_{1}\right)(x,0)-\left(\partial_{\lambda_{j_{1}}}\ldots\partial_{\lambda_{j_{m-1}}}\gamma_{2}\right)(x,0)\right)\\ \left(v^{\left(l_{1}\right)},\nabla v^{\left(l_{1}\right)}\right)_{j_{1}}\ldots\left(v^{\left(l_{m-1}\right)},\nabla v^{\left(l_{m-1}\right)}\right)_{j_{m-1}}\nabla v^{\left(l_{m}\right)}\cdot\nabla v^{(m+1)}dx=0 (1.4)

for all v(l)C(Ω¯)v^{(l)}\in C^{\infty}(\bar{\Omega}) solving (γ0v(l))=0\nabla\cdot\left(\gamma_{0}\nabla v^{(l)}\right)=0 in Ω,l=1,,m+1\Omega,l=1,\ldots,m+1.

Therefore, by also recalling the assumption (b) on γ1\gamma_{1} and γ2\gamma_{2}, the proof of Theorem 1 reduces to the following completeness result:

Proposition 1.

Let Ω2\Omega\subset\mathbb{R}^{2}, be a bounded open domain with CC^{\infty} boundary. Let γ0C(Ω¯)\gamma_{0}\in C^{\infty}(\bar{\Omega}) and assume that γ0\gamma_{0} satisfies the assumption (a). Let mm\in\mathbb{N} and let TT be CC^{\infty} smooth function with values in the space of symmetric tensors of rank mm. Assume also that TT vanishes to infinite order on Ω\partial\Omega. Suppose that

(l1,,lm+1)π(m+1)j1,,jm=02ΩTj1jm(x)(ul1,ul1)j1(ulm,ulm)jm×ulm+1um+2dx=0.\sum_{\left(l_{1},\ldots,l_{m+1}\right)\in\pi(m+1)}\sum_{j_{1},\ldots,j_{m}=0}^{2}\int_{\Omega}T^{j_{1}\ldots j_{m}}(x)\left(u_{l_{1}},\nabla u_{l_{1}}\right)_{j_{1}}\ldots\left(u_{l_{m}},\nabla u_{l_{m}}\right)_{j_{m}}\\ \times\nabla u_{l_{m+1}}\cdot\nabla u_{m+2}\hskip 0.5ptdx=0. (1.5)

for all ulC(Ω¯)u_{l}\in C^{\infty}(\bar{\Omega}) solving (γ0ul)=0\nabla\cdot\left(\gamma_{0}\nabla u_{l}\right)=0 in Ω,l=1,,m+2\Omega,l=1,\ldots,m+2. Then TT vanishes identically on Ω\Omega. Here (ul,ul)j,j=0,1,2\left(u_{l},\nabla u_{l}\right)_{j},j=0,1,2 stands for the jjth component of the vector (ul,x1ul,,xnul)\left(u_{l},\partial_{x_{1}}u_{l},\ldots,\partial_{x_{n}}u_{l}\right), and π(m+1)\pi(m+1) stands for the set of all distinct permutations of {1,,m+1}\{1,\ldots,m+1\}.

To prove the above proposition, we use the reduction of the conductivity equation to the Schrödinger equation and Bukhgeim type CGO solutions constructed for the Schrödinger equation in [Buk08]. We also use modifications of the Bukhgeim solutions introduced very recently in [CLT24]. These modifications include CGOs whose phase functions do not have critical points. This yields better estimates for the corresponding correction terms of the solutions, which are needed when considering inverse problems for nonlinear equations. The CGO solutions we use may have phases without critical points, or have at most one critical point.

1.0.1 Earlier results

Before going into the proof of Proposition 1, we review earlier related works. For the linear conductivity equation Sylvester and Uhlmann [SU87] and Novikov [Nov88] proved the uniqueness result for smooth conductivities in dimensions 3 and higher. In dimension 2, Nachman [Nac96] proved a uniqueness result for C2C^{2} conductivities. The regularity assumptions have since been relaxed by several authors. In dimensions three and higher the uniqueness is known for W1,(Ω)W^{1,\infty}(\Omega) conductivities by the work of Haberman and Tataru [HT13], and in dimension 22 for LL^{\infty} conductivities by the work of Astala and Päivärinta [PA06]. In the work by Bukhgeim [Buk08] the potential of a Shrödinger equation was recover from the corresponding DN map in dimension 22. In the case the conductivity is matrix valued, the best results are in dimension 22. In the work [IUY12] by Imanuvilov-Uhlmann-Yamamoto, a matrix valued conductivity was recover on bounded domains in 2\mathbb{R}^{2}. The very recent related work [CLT24] by Cârstea, Liimatainen and Tzou recovered the conformal structure of Riemannian surface from the Dirichlet-to-Neumann map (DN map) of the associated Shrödinger equation.

The approach in the study of inverse problems for nonlinear elliptic equations was proposed in [Isa93]. There the author linearized the nonlinear DN map, which reduced the inverse problem for a nonlinear equation to an inverse problem of a linear equation, which the author was able to solve by using methods for linear equations. For the quasilinear conductivity equation, where γ(x,u)\gamma(x,u) depends on uu, [Sun96] proves a uniqueness result for C1,1C^{1,1} regular conductivities by linearizing of the nonlinear DN map. The linearization technique was further examined in [IN95, IS94] for elliptic equations. Later, second order linearizations, where data depends on two independent parameters, were used to solve inverse problems for example in [AZ21, CNV19, KN02, Sun10, SU97].

Inverse problems for semilinear elliptic equations were also recently considered in the works by Feizmohammadi and Oksanen [FO20] and Lassas, Liimatainen, Lin and Salo [LLLS21]. These works realized how to use higher order linearizations in inverse problems for elliptic equations. The method is by now called the higher order linearization method and it was motivated by the seminal work [KLU18] by Lassas and Uhlmann that considered nonlinear hyperbolic equations.

After the works [KLU18, FO20, LLLS21], the literature on the research of inverse problems for nonlinear equations based on the higher order linearization method, has grown rapidly. Earlier inverse problems for quasilinear conductivity equations have been considered in [HS02, EPS14, MU20, Sha20]. The recent works [LLLS20, LLST22, KU20a, KU20b, FLL23, LL24] investigated inverse problems for semilinear elliptic equations with general nonlinearities and in the case of partial data. For quasilinear conductivity equation where γ(x,u,u)\gamma(x,u,\nabla u), the uniqueness result we prove in this paper in dimension 22 was obtained in [CFK+21] in dimensions 33 and higher, and the corresponding partial data result was proven in [KKU22]. The works [CLLO24, Nur23a, Nur23b, CLT24] studied inverse problems for the minimal surface equation (which is quasilinear) on Riemannian surfaces and in Euclidean domains.

Acknowledgements R.W. would like to thank Gunther Uhlmann for proposing this project and helpful discussions throughout the progress.

T.L. was partially supported by PDE-Inverse project of the European Research Council of the European Union, and the grant 336786 of the Research Council of Finland. Views and opinions expressed are those of the authors only and do not necessarily reflect those of the European Union or the other funding organizations. Neither the European Union nor the other funding organizations can be held responsible for them.

2 CGO Solutions

In this section we construct the CGO solutions that we will use in the proof of Proposition 1. In what follows we write x=(x1,x2)2x=(x_{1},x_{2})\in\mathbb{R}^{2} with z=x1+ix2z=x_{1}+ix_{2}\in\mathbb{C}, and =12(1i2)\partial=\frac{1}{2}(\partial_{1}-i\partial_{2}) and ¯=12(1+i2)\overline{\partial}=\frac{1}{2}(\partial_{1}+i\partial_{2}). By the identity

γ(γ1/2u)=γ1/2(Δ+q)u,-\nabla\cdot\gamma\nabla\left(\gamma^{-1/2}u\right)=\gamma^{1/2}(-\Delta+q)u, (2.1)

where

q=Δγγ,q=\frac{\Delta\sqrt{\gamma}}{\sqrt{\gamma}},

we obtain solutions to the conductivity equation from solutions to the Schrödinger equation.

We start by recalling the construction of CGO solutions of Bukhgeim in [Buk08]. For this, let ψ\psi be a holomorphic function such that 2(ψ)=φ2\Im(\psi)=\varphi is a real valued function with one non-degenerate critical point at z0Ωz_{0}\in\Omega. For h>0h>0 and a function ff, we write

f±h=e±iφ/hf,f_{\pm h}=e^{\pm\mathrm{i}\varphi/h}f,

and define the operators SS and StS^{t} by

Su=14¯1(bh1(ahu)),Su=\frac{1}{4}\overline{\partial}^{-1}(b_{-h}\partial^{-1}(a_{h}u)),
Stu=141(bh¯1(ahu)),S^{t}u=\frac{1}{4}\partial^{-1}(b_{-h}\overline{\partial}^{-1}(a_{h}u)),

where aLpa\in L^{p} and bW1,b\in W^{1,\infty}, and ¯1\overline{\partial}^{-1} is the usual Cauchy operator

(¯1u)(z)=1πΩu(w)zwdw.(\overline{\partial}^{-1}u)(z)=\frac{1}{\pi}\int_{\Omega}\frac{u(w)}{z-w}\mathrm{~{}d}w.

Here also 1\partial^{-1} is an operator with the conjugate kernel π1(zw)1¯\pi^{-1}\overline{(z-w)^{-1}}. By the proof of in [GT11, Lemma 3.1], we have

SLrLr=O(h1/r)andSL2L2=O(h1/2ϵ)||S||_{L^{r}\to L^{r}}=O(h^{1/r})\quad\text{and}\quad||S||_{L^{2}\to L^{2}}=O(h^{1/2-\epsilon})

for any r>2r>2 and 0<ϵ<120<\epsilon<\frac{1}{2}.

Let us then choose a=Δγγa=\frac{\Delta\gamma}{\sqrt{\gamma}} and b=1b=-1. By Theorem 3.5 in [Buk08] and identity (2.1),

u\displaystyle u =1γeψ/h(1+rh),\displaystyle=\frac{1}{\sqrt{\gamma}}e^{\psi/h}(1+r_{h}), (2.2)
rh\displaystyle r_{h} =n=1Sn1\displaystyle=\sum_{n=1}^{\infty}S^{n}1 (2.3)

is a solution to (γu)=0\nabla\cdot(\gamma\nabla u)=0. Note that the correction term rhr_{h} depends on both hh and ψ\psi. For the convenience of the reader, we also denote

¯φ1f:=¯1eiφ/hf,φ1f:=1eiφ/hf,Th:=φ1q¯φ1\bar{\partial}_{\varphi}^{-1}f:=\bar{\partial}^{-1}e^{-i\varphi/h}f,\quad\partial_{\varphi}^{-1}f:=\partial^{-1}e^{i\varphi/h}f,\quad T_{h}:=-\partial_{\varphi}^{-1}q\bar{\partial}_{\varphi}^{-1}

for easier comparison to the work [CLT24] from which we next borrow material from. With this notation, (2.3) reads

rh=¯φ1sh,sh=n=0Thnφ1q.r_{h}=\bar{\partial}_{\varphi}^{-1}s_{h},\quad s_{h}=-\sum_{n=0}^{\infty}T_{h}^{n}\partial_{\varphi}^{-1}q.

We have similarly for antiholomorphic phase function ψ\psi.

In particular, after possible translation of coordinates, we can choose z2z^{2} and z¯2-\bar{z}^{2} as the phases for the CGO solutions we use in our proofs. In these cases, the CGOs are

u\displaystyle{u} =1γez2/h(1+rh),\displaystyle=\frac{1}{\sqrt{\gamma}}e^{z^{2}/h}(1+{r}_{h}), (2.4)
rh\displaystyle{r}_{h} =n=1Sn1\displaystyle=\sum_{n=1}^{\infty}S^{n}1 (2.5)
u\displaystyle{u} =1γez¯2/h(1+r~h),\displaystyle=\frac{1}{\sqrt{\gamma}}e^{-\overline{z}^{2}/h}(1+\tilde{r}_{h}), (2.6)
r~h\displaystyle\tilde{r}_{h} =n=1(St)n1\displaystyle=\sum_{n=1}^{\infty}(S^{t})^{n}1 (2.7)

By [CLT24, Section 4.1], we have that the remainder above satisfy the estimates

r~hLr,r~hLr,¯r~hLr,s~hLr=O(h1/r+ϵr)\displaystyle\left\|\tilde{r}_{h}\right\|_{L^{r}},\left\|\partial\tilde{r}_{h}\right\|_{L^{r}},\left\|\bar{\partial}\tilde{r}_{h}\right\|_{L^{r}},||\tilde{s}_{h}||_{L^{r}}=O\left(h^{1/r+\epsilon_{r}}\right) (2.8)

for any r[2,)r\in[2,\infty) and ϵr>0\epsilon_{r}>0 depending on rr. We also recall from [GT11, Lemma 2.2] that for smooth ff and all ϵ>0\epsilon>0 small enough

¯φ1(f)=OL2(h1/2+ϵ) and φ1(f)=OL2(h1/2+ϵ).\bar{\partial}_{\varphi}^{-1}(f)=O_{L^{2}}\left(h^{1/2+\epsilon}\right)\text{ and }\partial_{\varphi}^{-1}(f)=O_{L^{2}}\left(h^{1/2+\epsilon}\right). (2.9)

In addition to the above CGOs, we will also use their modification introduced recently in [CLT24]. The phase functions ψ\psi of the modifications are holomorphic or antiholomorphic, but they have no critical points in Ω\Omega. After possible scaling and translation of coordinates, we may assume on our domain that z+12z2z+\frac{1}{2}z^{2} and z¯12z¯2-\bar{z}-\frac{1}{2}\bar{z}^{2} are such phase functions. In this case we have the solutions

u\displaystyle{u} =1γe(z+12z2)/h(1+rh),\displaystyle=\frac{1}{\sqrt{\gamma}}e^{(z+\frac{1}{2}z^{2})/h}(1+r_{h}), (2.10)
rh\displaystyle r_{h} =n=1Sn1\displaystyle=\sum_{n=1}^{\infty}S^{n}1 (2.11)

and

u\displaystyle{u} =1γe(z¯12z¯2)/h(1+r~h),\displaystyle=\frac{1}{\sqrt{\gamma}}e^{(-\bar{z}-\frac{1}{2}\bar{z}^{2})/h}(1+\tilde{r}_{h}), (2.12)
r~h\displaystyle\tilde{r}_{h} =n=1(St)n1\displaystyle=\sum_{n=1}^{\infty}(S^{t})^{n}1 (2.13)

to the Shr̈odinger equation (Δ+q)u=0(-\Delta+q)u=0. By integration by parts we have

¯1eiφ/hf=ih2[eiφ/hf¯φ+ih2¯1(eiφ/h¯(f¯φ))],\bar{\partial}^{-1}e^{\mathrm{i}\varphi/h}f=\frac{ih}{2}\left[e^{\mathrm{i}\varphi/h}\frac{f}{\bar{\partial}\varphi}+\frac{ih}{2}\bar{\partial}^{-1}\left(e^{\mathrm{i}\varphi/h}\bar{\partial}\left(\frac{f}{\bar{\partial}\varphi}\right)\right)\right],

which holds for any fC01(Ω¯)f\in C_{0}^{1}(\bar{\Omega}) and φ\varphi having no critical points in Ω\Omega. Therefore, using Calderón-Zygmund estimate (see e.g. [GT01]), we have

¯1eiφ/hfLrChfW1,r||\bar{\partial}^{-1}e^{\mathrm{i}\varphi/h}f||_{L^{r}}\leq Ch||f||_{W^{1,r}}

for all r(1,)r\in(1,\infty). This leads to better estimates

rhLr,rhLr,¯rhLr,shLr=O(h)\displaystyle||r_{h}||_{L^{r}},||\partial r_{h}||_{L^{r}},||\bar{\partial}r_{h}||_{L^{r}},||s_{h}||_{L^{r}}=O(h) (2.14)

for the correction terms in the above solutions. We refer to [CLT24, Section 4] for more details about CGOs with phases without critical points.

3 Proof of Proposition 1

We now prove Proposition 1, which consequently proves also Theorem 1 by the discussion in the introduction of this paper. We will choose the solutions in the integral identity (1.5) to be CGO solutions that we introduced in Section 2. The proof is somewhat complicated and also technical. Most of the complications come from the fact that we are recovering the components of the tensor TT by using stationary phase, while on the other hand the correction terms of the CGOs we use satisfy LpL^{p} estimates with only limited decay. It can be seen that the LpL^{p} estimates are not enough to show that the correction terms correspond to negligible terms in the asymptotic analysis of the integral identity (1.5). Roughly speaking, stationary phase analysis is not well compatible with LpL^{p} estimates.

To overcome the above difficulty, we will use also CGOs whose phase functions do not have critical points, which we will see to lead to sufficiently improved decay for the correction terms. We will also use the explicit forms of the correction terms in the analysis. Since the integral identity (1.5) is also increasingly complicated in mm, the proof will also be somewhat technical. For this reason, we split the proof into cases with respect to mm. The case m=1m=1 contains the most important arguments and constructions of the proof.

3.1 The case m=1m=1:

By assumption, the entries Tj1jmT^{j_{1}\cdots j_{m}} in the proposition vanish to infinite order on the boundary. For m=1m=1, the integral identity (1.5) reads

0\displaystyle 0 =(l1,l2)π(2)j=0nTj(x)(ul1,ul1)jul2u3\displaystyle=\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\sum_{j=0}^{n}\int T^{j}(x)\left(u_{l_{1}},\nabla u_{l_{1}}\right)_{j}\nabla u_{l_{2}}\cdot\nabla u_{3}
=j=0nTj(x)(u1,u1)ju2u3+j=0nTj(x)(u2,u2)ju1u3,\displaystyle=\sum_{j=0}^{n}\int T^{j}(x)\left(u_{1},\nabla u_{1}\right)_{j}\nabla u_{2}\cdot\nabla u_{3}+\sum_{j=0}^{n}\int T^{j}(x)\left(u_{2},\nabla u_{2}\right)_{j}\nabla u_{1}\cdot\nabla u_{3},

Where (u1,u1)j(u_{1},\nabla u_{1})_{j}, j=0,1,,nj=0,1,\ldots,n, stands for the jthj^{\textrm{th}} component of (u1,x1u1,,xnu1)(u_{1},\partial_{x_{1}}u_{1},...,\partial_{x_{n}}u_{1}), and similarly for (u2,u2)j(u_{2},\nabla u_{2})_{j}. By setting u2=1u_{2}=1 in the integral identity, the terms involving u2\nabla u_{2} all vanishes and so we have

T0(x)uv=0\int T^{0}(x)\nabla u\cdot\nabla v=0 (3.1)

for any functions uu and vv solving

(γ0u)=0 in Ω.\nabla\cdot\left(\gamma_{0}\nabla u\right)=0\text{ in }\Omega. (3.2)

We rewrite (3.1) as

(T0γ0)γ0uv=0\int\left(\frac{T^{0}}{\gamma_{0}}\right)\gamma_{0}\nabla u\cdot\nabla v=0

We integrate by parts to move the gradient on vv to the other terms. By using also (3.2), we get

γ0v(T0γ0)u=0.\displaystyle\int\gamma_{0}v\nabla\left(\frac{T^{0}}{\gamma_{0}}\right)\cdot\nabla u=0. (3.3)

There is no boundary term since T0T^{0} vanishes on the boundary. By integrating by parts again, we obtain

0=γ0v(T0γ0)u=uv(γ0(T0γ0))+γ0u(T0γ0)v\displaystyle 0=\int\gamma_{0}v\nabla\left(\frac{T^{0}}{\gamma_{0}}\right)\cdot\nabla u=\int uv\nabla\cdot\left(\gamma_{0}\nabla\left(\frac{T^{0}}{\gamma_{0}}\right)\right)+\int\gamma_{0}u\nabla\left(\frac{T^{0}}{\gamma_{0}}\right)\cdot\nabla v
=uv(γ0(T0γ0)).\displaystyle=\int uv\nabla\cdot\left(\gamma_{0}\nabla\left(\frac{T^{0}}{\gamma_{0}}\right)\right). (3.4)

Here in the last identity we used (3.3) with uu in place of vv.

Let us denote

A:=(γ0(T0γ0)),A:=\nabla\cdot\left(\gamma_{0}\nabla\left(\frac{T^{0}}{\gamma_{0}}\right)\right),

and let uu and vv be the CGO solutions

u=\displaystyle u= 1γ0ez2/h(1+rh)\displaystyle\frac{1}{\sqrt{\gamma_{0}}}e^{z^{2}/h}(1+r_{h})
v=\displaystyle v= 1γ0ez¯2/h(1+r~h)\displaystyle\frac{1}{\sqrt{\gamma_{0}}}e^{-\bar{z}^{2}/h}(1+\tilde{r}_{h})

as in 2.4 and 2.6. We have by (3.1) that

e(z2z¯2)/hAγ0(1+rh+r~h+rhr~h)=0.\displaystyle\int e^{(z^{2}-\bar{z}^{2})/h}\frac{A}{\gamma_{0}}(1+r_{h}+\tilde{r}_{h}+r_{h}\tilde{r}_{h})=0. (3.5)

By using stationary phase, we have the expansion

e(z2z¯2)/hAγ0=h(Aγ0)(0)+o(h).\int e^{(z^{2}-\bar{z}^{2})/h}\frac{A}{\gamma_{0}}=h\left(\frac{A}{\gamma_{0}}\right)(0)+o(h).

(as well as for 1{\partial}^{-1})

At this point, we mention that integration by parts also works for ¯1\bar{\partial}^{-1} as it does for ¯\bar{\partial}. By [Vek62, Theorem 1.13] the following holds: Let fL1f\in L^{1}, φLp\varphi\in L^{p} for some p>2p>2, and assume that ff and φ\varphi both vanish on the boundary. Then by Fubini’s theorem

Ω(¯1f)φ𝑑z=1πΩ(Ωf(ξ)zξ𝑑ξ)φ(z)𝑑z=1πΩf(ξ)(Ωφ(z)zξ𝑑z)𝑑ξ=Ωf(¯1φ)𝑑z.\int_{\Omega}(\bar{\partial}^{-1}f)\varphi dz=\frac{1}{\pi}\int_{\Omega}\left(\int_{\Omega}\frac{f(\xi)}{z-\xi}d\xi\right)\varphi(z)dz\\ =\frac{1}{\pi}\int_{\Omega}f(\xi)\left(\int_{\Omega}\frac{\varphi(z)}{z-\xi}dz\right)d\xi=-\int_{\Omega}f(\bar{\partial}^{-1}\varphi)dz. (3.6)

We use this observation for the terms in (3.5) involving the remainder terms rhr_{h} and r~h\tilde{r}_{h}. An integration by parts gives

e(z2z¯2)/hAγ0rh=¯1(e(z2z¯2)/hAγ0)e(z¯2z2)/hsh.\int e^{(z^{2}-\bar{z}^{2})/h}\frac{A}{\gamma_{0}}r_{h}=-\int\bar{\partial}^{-1}\left(e^{(z^{2}-\bar{z}^{2})/h}\frac{A}{\gamma_{0}}\right)e^{(\bar{z}^{2}-{z}^{2})/h}s_{h}. (3.7)

Using (2.9) and the estimates (2.8),

r~hLr,r~hLr,¯r~hLr,s~hLr=O(h1/r+ϵr),r2,ϵr>0,\displaystyle\left\|\tilde{r}_{h}\right\|_{L^{r}},\left\|\partial\tilde{r}_{h}\right\|_{L^{r}},\left\|\bar{\partial}\tilde{r}_{h}\right\|_{L^{r}},||\tilde{s}_{h}||_{L^{r}}=O\left(h^{1/r+\epsilon_{r}}\right),\quad r\geq 2,\quad\epsilon_{r}>0,

we get by Cauchy’s inequality that the right hand side of (3.7) is O(h1+ϵ)O(h^{1+\epsilon}) for some ϵ>0\epsilon>0. We have similarly that the terms in (3.5) involving r~h\tilde{r}_{h} and rhr~hr_{h}\tilde{r}_{h} are O(h1+ϵ)O(h^{1+\epsilon}). Hence, by dividing (3.5) by hh and letting h0h\to 0, we obtain A(0)γ01(0)=0A(0)\gamma_{0}^{-1}(0)=0. This shows that A=0A=0 at z=0z=0.

By translation we can vary the critical point of the phases of the CGOs to show that A=0A=0 in Ω\Omega,

(γ0(T0γ0))=0in Ω.\nabla\cdot\left(\gamma_{0}\nabla\left(\frac{T^{0}}{\gamma_{0}}\right)\right)=0\quad\text{in $\Omega$}.

Thus T0/γ0T^{0}/\gamma_{0} is a solution to an elliptic equation in Ω\Omega. Since T0T^{0} is identically zero on the boundary, uniqueness of solutions to the Dirichlet problem of the above equation shows that

T0=0 in Ω.T^{0}=0\text{ in }\Omega.

By using that T00T^{0}\equiv 0, the integral identity (1.5) we started from now reduces to

(l1,l2)π(2)j=12Tj(x)xjul1ul2u3dx=0,\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\sum_{j=1}^{2}\int T^{j}(x)\partial_{x_{j}}u_{l_{1}}\nabla u_{l_{2}}\cdot\nabla u_{3}dx=0, (3.8)

in the current case m=1m=1 holding for all ulC(Ω¯)u_{l}\in{C}^{\infty}(\overline{\Omega}), l=1,2,3l=1,2,3, solving 3.2. We continue by letting ulu_{l}, l=1,2,3l=1,2,3, to be solutions as in (2.10) and (2.12):

u1\displaystyle{u_{1}} =1γ0e(z+12z2)/h(1+r1),\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(z+\frac{1}{2}z^{2})/h}(1+r_{1}),
u2\displaystyle{u_{2}} =1γ0e(z+12z2)/h(1+r2),\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(-z+\frac{1}{2}z^{2})/h}(1+r_{2}),
u3\displaystyle{u_{3}} =1γ0ez¯2/h(1+r~)\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{-\bar{z}^{2}/h}(1+\tilde{r})

Notice that we can rewrite these solutions as

u1=1γ0v1,u2=1γ0v2,u3=1γ0v3,\displaystyle{u_{1}}=\frac{1}{\sqrt{\gamma_{0}}}v_{1},\quad{u_{2}}=\frac{1}{\sqrt{\gamma_{0}}}v_{2},\quad{u_{3}}=\frac{1}{\sqrt{\gamma_{0}}}v_{3}, (3.9)

where the functions vlv_{l} solve

Δv+qv=0\Delta v+qv=0

with q=Δγ0/γ0q=\Delta\gamma_{0}/\sqrt{\gamma_{0}}. Since

ul2u3=12(Δ(ul2u3)ul2Δu3(Δul2)u3),\nabla u_{l_{2}}\cdot\nabla u_{3}=\frac{1}{2}(\Delta(u_{l_{2}}u_{3})-u_{l_{2}}\Delta u_{3}-(\Delta u_{l_{2}})u_{3}),

we can rewrite (3.8) as

(l1,l2)π(2)j=12(Tjxjul1)(Δ(ul2u3)ul2Δu3(Δul2)u3)=0\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\sum_{j=1}^{2}\int(T^{j}\partial_{x_{j}}u_{l_{1}})(\Delta(u_{l_{2}}u_{3})-u_{l_{2}}\Delta u_{3}-(\Delta u_{l_{2}})u_{3})=0

By using (3.9), we also have

Δul=Δ(1γ0)vl+2(1γ0)vlqvlγ0.\displaystyle\Delta u_{l}=\Delta\left(\frac{1}{\sqrt{\gamma_{0}}}\right)v_{l}+2\nabla\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\cdot\nabla v_{l}-\frac{qv_{l}}{\sqrt{\gamma_{0}}}. (3.10)

Let us write Δ=4¯\Delta=4\partial\bar{\partial} and integrate by parts to obtain

0\displaystyle 0 =(l1,l2)π(2)j=12(Tjxjul1)(Δ(ul2u3)ul2Δu3(Δul2)u3)\displaystyle=\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\sum_{j=1}^{2}\int(T^{j}\partial_{x_{j}}u_{l_{1}})(\Delta(u_{l_{2}}u_{3})-u_{l_{2}}\Delta u_{3}-(\Delta u_{l_{2}})u_{3})
=(l1,l2)π(2)j=12{4¯(Tjxjul1)(ul2u3)\displaystyle=\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\sum_{j=1}^{2}\int\Bigg{\{}4\partial\bar{\partial}(T^{j}\partial_{x_{j}}u_{l_{1}})(u_{l_{2}}u_{3}) (3.11)
(Tjxjul1)(2ul2(1γ0)v3+2u3(1γ0)vl2)\displaystyle\qquad\qquad-(T^{j}\partial_{x_{j}}u_{l_{1}})(2u_{l_{2}}\nabla(\frac{1}{\sqrt{\gamma_{0}}})\cdot\nabla v_{3}+2u_{3}\nabla(\frac{1}{\sqrt{\gamma_{0}}})\cdot\nabla v_{l_{2}})
+(Tjxjul1)FT,γ0(ul2v3+vl2u3)},\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad+(T^{j}\partial_{x_{j}}u_{l_{1}})F_{T,\gamma_{0}}(u_{l_{2}}v_{3}+v_{l_{2}}u_{3})\Bigg{\}},

where FT,γ0F_{T,\gamma_{0}} is a general term that depends smoothly only on TlT^{l} and γ0\gamma_{0}, and their derivatives. As before, by estimates (2.14), (2.8) and stationary phase, we have

(Tjxjul1FT,γ0)(ul2v3+vl2u3)=O(1).\int(T^{j}\partial_{x_{j}}u_{l_{1}}F_{T,\gamma_{0}})(u_{l_{2}}v_{3}+v_{l_{2}}u_{3})=O(1). (3.12)

Using the above and by writing uv=2(u¯v+¯uv)\nabla u\cdot\nabla v=2(\partial u\bar{\partial}v+\bar{\partial}u\partial v), the integral in (3.1) reads

4¯(Tjxjul1)(ul2u3)4(Tjxjul1)[ul2((1γ0)¯v3+¯(1γ0)v3)\displaystyle\int 4\partial\bar{\partial}(T^{j}\partial_{x_{j}}u_{l_{1}})(u_{l_{2}}u_{3})-\int 4(T^{j}\partial_{x_{j}}u_{l_{1}})\left[u_{l_{2}}\left(\partial\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\bar{\partial}v_{3}+\bar{\partial}\left(\frac{1}{\sqrt{\gamma_{0}}}\right){\partial}v_{3}\right)\right.
+u3((1γ0)¯vl2+¯(1γ0)vl2)]+O(1).\displaystyle\qquad\quad\qquad\qquad\qquad\qquad\qquad\qquad\left.+u_{3}\left(\partial\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\bar{\partial}v_{l_{2}}+\bar{\partial}\left(\frac{1}{\sqrt{\gamma_{0}}}\right){\partial}v_{l_{2}}\right)\right]+O(1).

Here we have

4¯(Tjxjul1)(ul2u3)=4(Tj¯xjul1+4¯Tjxjul1)(ul2u3)+4Tjxj((1γ0)¯vl1+¯(1γ0)vl1)(ul2u3)+O(1)\int 4\partial\bar{\partial}(T^{j}\partial_{x_{j}}u_{l_{1}})(u_{l_{2}}u_{3})=\int 4(\partial T^{j}\bar{\partial}\partial_{x_{j}}u_{l_{1}}+4\bar{\partial}T^{j}{\partial}\partial_{x_{j}}u_{l_{1}})(u_{l_{2}}u_{3})\\ +\int 4T^{j}\partial_{x_{j}}\left(\partial\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\bar{\partial}v_{l_{1}}+\bar{\partial}\left(\frac{1}{\sqrt{\gamma_{0}}}\right){\partial}v_{l_{1}}\right)(u_{l_{2}}u_{3})+O(1) (3.13)

since in the case both \partial and ¯\bar{\partial} hit TjT^{j} in the term 4¯(Tjxjul1)4\partial\bar{\partial}(T^{j}\partial_{x_{j}}u_{l_{1}}), then the resulting integral term will be O(1)O(1) by the argument after (3.7). We also used (3.10) and (3.12) again.

Note next that

Tj¯xjul1=O(1),\int\partial T^{j}\bar{\partial}\partial_{x_{j}}u_{l_{1}}=O(1),

because the phase function of ul1u_{l_{1}} is holomorphic. Indeed, when ¯\bar{\partial} hits the exponential factor exp((±z+12z2)/h)\exp{((\pm z+\frac{1}{2}z^{2})/h)} of ul1u_{l_{1}} the result vanishes and the situation is then similar to the case where there are only first order derivatives of ul1u_{l_{1}}. Regarding this, we note that when ¯\bar{\partial}\partial hits the correction term, the corresponding integral is also O(1)O(1). This is due to the Calderón-Zygmund estimate explained below.

By the Calderón-Zygmund inequality (see for example [GT01, Corollary 9.10]), the L2L^{2} norm of any second order derivative of rl1r_{l_{1}} multiplied by a compactly supported CC^{\infty} function, say HH, can be estimated as

H2rl1L22(Hrl1)L2+2Hrl1L2+rl12HL2¯(Hrl1)L2+O(h1/2+ϵ)(eiφ/hsh)L2+O(h1/2+ϵ)(1hshL2+shL2)+O(h1/2+ϵ)=O(1).||H\nabla^{2}r_{l_{1}}||_{L^{2}}\leq||\nabla^{2}(Hr_{l_{1}})||_{L^{2}}+2||\nabla H\otimes\nabla r_{l_{1}}||_{L^{2}}+||r_{l_{1}}\nabla^{2}H||_{L^{2}}\\ \lesssim||\partial\bar{\partial}(Hr_{l_{1}})||_{L^{2}}+O(h^{1/2+\epsilon})\lesssim||\partial(e^{i\varphi/h}s_{h})||_{L^{2}}+O(h^{1/2+\epsilon})\\ \lesssim\left(\frac{1}{h}||s_{h}||_{L^{2}}+||\partial s_{h}||_{L^{2}}\right)+O(h^{1/2+\epsilon})=O(1). (3.14)

Here we used the estimate shL2=O(h)||s_{h}||_{L^{2}}=O(h) from 2.14 and

sh=k=0Thkφ1qs_{h}=-\sum_{k=0}^{\infty}T_{h}^{\hskip 0.5ptk}\partial_{\varphi}^{-1}q

to have that shL2=O(1)||\partial s_{h}||_{L^{2}}=O(1). So the third identity holds.

Returning to the main line of the proof, combining our computations so far shows that (3.1) equals

(l1,l2)π(2)j=12{4¯Tj(xjul1)ul2u3\displaystyle\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\sum_{j=1}^{2}\int\Bigg{\{}4\bar{\partial}T^{j}({\partial}\partial_{x_{j}}u_{l_{1}})u_{l_{2}}u_{3}
+4Tjxj((1γ0)¯vl1+¯(1γ0)vl1)ul2u3\displaystyle+4T^{j}\partial_{x_{j}}\left(\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\bar{\partial}v_{l_{1}}+\bar{\partial}\left(\frac{1}{\sqrt{\gamma_{0}}}\right){\partial}v_{l_{1}}\right)u_{l_{2}}u_{3}
4(Tjxjul1)[ul2((1γ0)¯v3+¯(1γ0)v3)\displaystyle-4(T^{j}\partial_{x_{j}}u_{l_{1}})\left[u_{l_{2}}\left(\partial\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\bar{\partial}v_{3}+\bar{\partial}\left(\frac{1}{\sqrt{\gamma_{0}}}\right){\partial}v_{3}\right)\right. (3.15)
+u3((1γ0)¯vl2+¯(1γ0)vl2)]}+O(1).\displaystyle\qquad\quad\qquad\qquad\qquad\qquad\qquad\qquad\left.+u_{3}\left(\partial\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\bar{\partial}v_{l_{2}}+\bar{\partial}\left(\frac{1}{\sqrt{\gamma_{0}}}\right){\partial}v_{l_{2}}\right)\right]\Bigg{\}}+O(1).

Let us then recall the notations x1=1=+¯\partial_{x_{1}}=\partial_{1}=\partial+\bar{\partial} and x2=2=i(¯)\partial_{x_{2}}=\partial_{2}=i(\partial-\bar{\partial}). So we have

1=2+¯,2=i(2¯).\partial\partial_{1}=\partial^{2}+\partial\bar{\partial},\quad\partial\partial_{2}=i(\partial^{2}-\partial\bar{\partial}). (3.16)

Let us then consider the first term in (3.1). We have

(l1,l2)π(2)j=124¯Tj(xjul1)ul2u3=(l1,l2)π(2)4¯(T1+iT2)(2ul1)(ul2u3)+O(1)\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\sum_{j=1}^{2}\int 4\bar{\partial}T^{j}({\partial}\partial_{x_{j}}u_{l_{1}})u_{l_{2}}u_{3}=\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\int 4\bar{\partial}(T^{1}+iT^{2})(\partial^{2}u_{l_{1}})(u_{l_{2}}u_{3})+O(1)

by arguing as above. Consequently

(l1,l2)π(2)j=124¯Tj(xjul1)ul2u3=(l1,l2)π(2)4¯(T1+iT2)(2ul1)ul2u3+O(1)\displaystyle\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\sum_{j=1}^{2}\int 4\bar{\partial}T^{j}({\partial}\partial_{x_{j}}u_{l_{1}})u_{l_{2}}u_{3}=\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\int 4\bar{\partial}(T^{1}+iT^{2})({\partial}^{2}u_{l_{1}})u_{l_{2}}u_{3}+O(1)
=4h2γ03/2¯(T1+iT2)e(z2z¯2)/h((1+z)2+(1+z)2)(1+r1)(1+r2)(1+r3)+O(1)\displaystyle=\int\frac{4}{h^{2}\gamma_{0}^{3/2}}\bar{\partial}(T^{1}+iT^{2})e^{(z^{2}-\bar{z}^{2})/h}\left((1+z)^{2}+(-1+z)^{2}\right)(1+r_{1})(1+r_{2})(1+r_{3})+O(1)
=8h(γ0(0))3/2¯(T1+iT2)(0)+o(h1)\displaystyle=\frac{8}{h}(\gamma_{0}(0))^{-3/2}\hskip 0.5pt\bar{\partial}(T^{1}+iT^{2})(0)+o(h^{-1}) (3.17)

by the stationary phase. Here we also used that the integral corresponding to 2rl1\partial^{2}r_{l_{1}} is O(1)O(1) by the Calderón-Zygmund estimate.

For the second term of (3.1), we similarly have

(l1,l2)π(2)j=124Tjxj[(1γ0)¯vl1+¯(1γ0)vl1]ul2u3\displaystyle\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\sum_{j=1}^{2}\int 4T^{j}\partial_{x_{j}}\left[\partial\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\bar{\partial}v_{l_{1}}+\bar{\partial}\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\partial v_{l_{1}}\right]u_{l_{2}}u_{3}
=(l1,l2)π(2)4(T1+iT2)¯(1γ0)(2vl1)ul2u3+O(1)\displaystyle=\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\int 4(T^{1}+iT^{2})\bar{\partial}\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\left(\partial^{2}v_{l_{1}}\right)u_{l_{2}}u_{3}+O(1)
=4h2γ0(T1+iT2)¯(1γ0)[e(z2z¯2)/h((1+z)2+(1+z)2)(1+r1)(1+r2)(1+r3)]\displaystyle=\int\frac{4}{h^{2}\gamma_{0}}(T^{1}+iT^{2})\bar{\partial}\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\left[e^{(z^{2}-\bar{z}^{2})/h}((1+z)^{2}+(-1+z)^{2})(1+r_{1})(1+r_{2})(1+r_{3})\right]
+O(1)=8hγ0(0)1¯(γ0(0)1/2)(T1+iT2)(0)+o(h1).\displaystyle+O(1)=\frac{8}{h}\gamma_{0}(0)^{-1}\bar{\partial}(\gamma_{0}(0)^{-1/2})\hskip 0.5pt(T^{1}+iT^{2})(0)+o(h^{-1}). (3.18)

For the last term of (3.1), we first notice that

(l1,l2)π(2)j=124Tjxjul1{ul2[(1γ0)¯v3+¯(1γ0)v3]\displaystyle\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\sum_{j=1}^{2}\int-4T^{j}\partial_{x_{j}}u_{l_{1}}\Bigg{\{}u_{l_{2}}\left[\partial\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\bar{\partial}v_{3}+\bar{\partial}\left(\frac{1}{\sqrt{\gamma_{0}}}\right){\partial}v_{3}\right]
+u3[(1γ0)¯vl2+¯(1γ0)vl2]}\displaystyle\qquad\quad\qquad\qquad\qquad\quad\qquad\qquad+u_{3}\left[\partial\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\bar{\partial}v_{l_{2}}+\bar{\partial}\left(\frac{1}{\sqrt{\gamma_{0}}}\right){\partial}v_{l_{2}}\right]\Bigg{\}}
=(l1,l2)π(2)j=124Tjxjul1[ul2(1γ0)¯v3+u3¯(1γ0)vl2]+o(h1)\displaystyle=\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\sum_{j=1}^{2}\int-4T^{j}\partial_{x_{j}}u_{l_{1}}\left[u_{l_{2}}\partial\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\bar{\partial}v_{3}+u_{3}\bar{\partial}\left(\frac{1}{\sqrt{\gamma_{0}}}\right){\partial}v_{l_{2}}\right]+o(h^{-1})

by arguing similarly as before. Then we compute

(l1,l2)π(2)j=124(Tjxjul1)ul2(1γ0)¯v3\displaystyle\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\sum_{j=1}^{2}\int-4(T^{j}\partial_{x_{j}}u_{l_{1}})u_{l_{2}}\partial\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\bar{\partial}v_{3}
=4(T1+iT2)h2γ0(1γ0)[e(z2+z¯2)/h((1+z)+(1+z))(2z¯)(1+r1)(1+r2)(1+r3)]\displaystyle=\int-\frac{4(T^{1}+iT^{2})}{h^{2}\gamma_{0}}\partial\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\left[e^{(z^{2}+\bar{z}^{2})/h}((1+z)+(-1+z))(-2\bar{z})(1+r_{1})(1+r_{2})(1+r_{3})\right]
+o(h1)=o(h1).\displaystyle+o(h^{-1})=o(h^{-1}).

We also obtain

(l1,l2)π(2)j=124(Tjxjul1)u3¯(1γ0)vl2\displaystyle\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\sum_{j=1}^{2}\int-4(T^{j}\partial_{x_{j}}u_{l_{1}})u_{3}\bar{\partial}\left(\frac{1}{\sqrt{\gamma_{0}}}\right){\partial}v_{l_{2}}
=(l1,l2)π(2)Ω4(T1+iT2)¯(1γ0)(ul1)(vl2)u3+O(1)\displaystyle=\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\int_{\Omega}-4(T^{1}+iT^{2})\bar{\partial}\left(\frac{1}{\sqrt{\gamma_{0}}}\right)(\partial u_{l_{1}})(\partial v_{l_{2}})u_{3}+O(1)
=Ω4h2γ0(T1+iT2)¯(1γ0)[e(z2+z¯2)/h2(1+z)(1+z)(1+r1)(1+r2)(1+r3)]\displaystyle=\int_{\Omega}\frac{-4}{h^{2}\gamma_{0}}(T^{1}+iT^{2})\bar{\partial}\left(\frac{1}{\sqrt{\gamma_{0}}}\right)\left[e^{(z^{2}+\bar{z}^{2})/h}2(1+z)(-1+z)(1+r_{1})(1+r_{2})(1+r_{3})\right]
+O(1)=8hγ0(0)1¯(γ0(0)1/2)¯(T1+iT2)(0)+o(h1).\displaystyle+O(1)=\frac{8}{h}\gamma_{0}(0)^{-1}\bar{\partial}(\gamma_{0}(0)^{-1/2})\hskip 0.5pt\bar{\partial}(T^{1}+iT^{2})(0)+o(h^{-1}). (3.19)

Combining the above the results in (3.1), (3.1) and (3.1), multiply the right hand side of (3.1) by hh and letting hh\to\infty shows that

¯(T1+iT2)¯γ0γ0(T1+iT2)=0{\bar{\partial}(T^{1}+iT^{2})}-\frac{\bar{\partial}{\gamma_{0}}}{{\gamma_{0}}}(T^{1}+iT^{2})=0

in Ω\Omega. By applying \partial to both sides of the above equation, we get a second order elliptic equation for T1+iT2T^{1}+iT^{2}. Since T1T^{1} and T2T^{2} are real and vanish to first order on Ω\partial\Omega by boundary determination, we conclude that T1+iT2=0T^{1}+iT^{2}=0 in Ω\Omega by unique continuation.

So far we have shown that T0=T1+iT2=0T^{0}=T^{1}+iT^{2}=0. Let us next choose

u1\displaystyle{u_{1}} =1γ0e(z¯12z¯2)/h(1+r1),\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(-\bar{z}-\frac{1}{2}\bar{z}^{2})/h}(1+r_{1}),
u2\displaystyle{u_{2}} =1γ0e(z¯12z¯2)/h(1+r2),\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(\bar{z}-\frac{1}{2}\bar{z}^{2})/h}(1+r_{2}),
u3\displaystyle{u_{3}} =1γ0ez2/h(1+r~)\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{{z}^{2}/h}(1+\tilde{r})

as solutions to (γ0u)=0\nabla\cdot\left(\gamma_{0}\nabla u\right)=0. By using these solutions, and arguing in a similar manner as before, we obtain T1iT2=0T^{1}-iT^{2}=0 in Ω\Omega. Combining everything, we thus have shown T0=T1=T2=0T^{0}=T^{1}=T^{2}=0 in Ω\Omega. This concludes the proof of the case m=1m=1.

3.2 The case m=2m=2:

For m=2m=2, the integral identity (1.5) reads

(l1,l2,l3)π(3)j1,j2=02Tj1j2(x)(ul1,ul1)j1(ul2,ul2)j2ul3u4=0.\sum_{\left(l_{1},l_{2},l_{3}\right)\in\pi(3)}\sum_{j_{1},j_{2}=0}^{2}\int T^{j_{1}j_{2}}(x)\left(u_{l_{1}},\nabla u_{l_{1}}\right)_{j_{1}}\left(u_{l_{2}},\nabla u_{l_{2}}\right)_{j_{2}}\nabla u_{l_{3}}\cdot\nabla u_{4}\hskip 0.5pt=0. (3.20)

If we let two of the functions u1,u2,u3u_{1},u_{2},u_{3} be constant functions equal to 11, then we get

T00(x)uv=0,\int T^{00}(x)\nabla u\cdot\nabla v=0,

which is the same as the identity (3.1) in the m=1m=1 case for T0T^{0}. This proves T00=0T^{00}=0 in Ω\Omega. Next, we let one of u1,u2,u3u_{1},u_{2},u_{3} be the constant function 11. This yields

(l1,l2)π(2)j=12T0j(x)xjul1ul2u3=0,\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\sum_{j=1}^{2}\int T^{0j}(x)\partial_{x_{j}}u_{l_{1}}\nabla u_{l_{2}}\cdot\nabla u_{3}=0,

which is the same identity (3.8) as in the m=1m=1 case for T1T^{1} and T2T^{2}. Thus we obtain T01T^{01} and T02=0T^{02}=0 in Ω\Omega.

By using T00=T01=T02=0T^{00}=T^{01}=T^{02}=0, the identity (3.20) becomes

(l1,l2,l3)π(3)j,k=12Tjk(x)xjul1xkul2ul3u4=0.\displaystyle\sum_{\left(l_{1},l_{2},l_{3}\right)\in\pi(3)}\sum_{j,k=1}^{2}\int T^{jk}(x)\partial_{x_{j}}u_{l_{1}}\partial_{x_{k}}u_{l_{2}}\nabla u_{l_{3}}\cdot\nabla u_{4}=0. (3.21)

Next we choose solutions ulu_{l}, l=1,,4l=1,\ldots,4, as in (2.10) and (2.6), as we did when proving T1+iT2=0T^{1}+iT^{2}=0 in m=1m=1 case. That is, we choose

u1\displaystyle{u_{1}} =1γ0e(z+13z2)/h(1+r1),\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(z+\frac{1}{3}z^{2})/h}(1+r_{1}),
u2\displaystyle{u_{2}} =1γ0e(z+13z2)/h(1+r2),\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(z+\frac{1}{3}z^{2})/h}(1+r_{2}),
u3\displaystyle{u_{3}} =1γ0e(2z+13z2)/h(1+r3),\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(-2{z}+\frac{1}{3}{z}^{2})/h}(1+r_{3}),
u4\displaystyle{u_{4}} =1γ0ez¯2/h(1+r4).\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{-\bar{z}^{2}/h}(1+r_{4}).

Rewrite (3.21) as

(l1,l2)π(2)j,k=12(Tjkxjul1xkul2)(Δ(ul3u4)ul3Δu4(Δul3)u4)=0\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\sum_{j,k=1}^{2}\int(T^{jk}\partial_{x_{j}}u_{l_{1}}\partial_{x_{k}}u_{l_{2}})(\Delta(u_{l_{3}}u_{4})-u_{l_{3}}\Delta u_{4}-(\Delta u_{l_{3}})u_{4})=0

and we can write solutions as

ul=1γ0vl,l=1,2,3,4.\displaystyle{u_{l}}=\frac{1}{\sqrt{\gamma_{0}}}v_{l},\qquad l=1,2,3,4.

We obtain by substituting the solutions

0\displaystyle 0 =(l1,l2,l3)π(3)j,k=12(Tjkxjul1xkul2)(Δ(ul3u4)ul3Δu4(Δul3)u4)\displaystyle=\sum_{\left(l_{1},l_{2},l_{3}\right)\in\pi(3)}\sum_{j,k=1}^{2}\int(T^{jk}\partial_{x_{j}}u_{l_{1}}\partial_{x_{k}}u_{l_{2}})(\Delta(u_{l_{3}}u_{4})-u_{l_{3}}\Delta u_{4}-(\Delta u_{l_{3}})u_{4})
=(l1,l2,l3)π(3)j=12{4¯(Tjkxjul1xkul2)(ul3u4)\displaystyle=\sum_{\left(l_{1},l_{2},l_{3}\right)\in\pi(3)}\sum_{j=1}^{2}\int\Bigg{\{}4\partial\bar{\partial}(T^{jk}\partial_{x_{j}}u_{l_{1}}\partial_{x_{k}}u_{l_{2}})(u_{l_{3}}u_{4}) (3.22)
(Tjkxjul1xkul2)(2ul3(1γ0)v4+2u4(1γ0)vl3)\displaystyle\qquad\qquad-(T^{jk}\partial_{x_{j}}u_{l_{1}}\partial_{x_{k}}u_{l_{2}})(2u_{l_{3}}\nabla(\frac{1}{\sqrt{\gamma_{0}}})\cdot\nabla v_{4}+2u_{4}\nabla(\frac{1}{\sqrt{\gamma_{0}}})\cdot\nabla v_{l_{3}})
+(Tjkxjul1xkul2)FT,γ0(ul3v4+vl3u4)},\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad+(T^{jk}\partial_{x_{j}}u_{l_{1}}\partial_{x_{k}}u_{l_{2}})F_{T,\gamma_{0}}(u_{l_{3}}v_{4}+v_{l_{3}}u_{4})\Bigg{\}},

where FT,γ0F_{T,\gamma_{0}} is a general term that depends smoothly only on TjkT^{jk} and γ0\gamma_{0}, and their derivatives. Following the same argument as we did when proving T1+iT2=0T^{1}+iT^{2}=0 after (3.9) in the m=1m=1 case, we can prove the leading order term is O(h2)O(h^{-2}) while other terms is o(h2)o(h^{-2}). Hence we obtain

T11+2iT12T22=0.T^{11}+2iT^{12}-T^{22}=0. (3.23)

We similarly get

T112iT12T22=0\displaystyle T^{11}-2iT^{12}-T^{22}=0 (3.24)

by choosing u1u_{1}, u2u_{2}, u3u_{3} as in (2.12) and u4u_{4} as in (2.4)\eqref{csolu1}.

Now we let the solutions ulu_{l} be as in (2.10) and (2.12):

u1\displaystyle{u_{1}} =1γ0e(z+12z2)/h(1+r1),\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(z+\frac{1}{2}z^{2})/h}(1+r_{1}),
u2\displaystyle{u_{2}} =1γ0e(z+12z2)/h(1+r2),\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(-z+\frac{1}{2}z^{2})/h}(1+r_{2}),
u3\displaystyle{u_{3}} =1γ0e(z¯12z¯2)/h(1+r3),\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(-\bar{z}-\frac{1}{2}\bar{z}^{2})/h}(1+r_{3}),
u4\displaystyle{u_{4}} =1γ0e(z¯12z¯2)/h(1+r4).\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(\bar{z}-\frac{1}{2}\bar{z}^{2})/h}(1+r_{4}).

Again we write 1=+¯\partial_{1}=\partial+\bar{\partial}, 2=i(¯)\partial_{2}=i(\partial-\bar{\partial}) and uv=2(u¯v+¯uv)\nabla u\cdot\nabla v=2(\partial u\bar{\partial}v+\bar{\partial}u\partial v) so that (3.21) becomes

(l1,l2,l3)π(3)[T11((+¯)ul1)((+¯)ul2)+T12((+¯)ul1)(i(¯)ul2)+T21(i(¯)ul1)((+¯)ul2)T22((¯)ul1)((¯)ul2)](ul3¯u4+¯ul3u4)=0.\sum_{\left(l_{1},l_{2},l_{3}\right)\in\pi(3)}\int\Big{[}T^{11}((\partial+\bar{\partial})u_{l_{1}})((\partial+\bar{\partial})u_{l_{2}})+T^{12}((\partial+\bar{\partial})u_{l_{1}})(i(\partial-\bar{\partial}){u_{l_{2}}})\\ +T^{21}(i(\partial-\bar{\partial})u_{l_{1}})((\partial+\bar{\partial}){u_{l_{2}}})-T^{22}((\partial-\bar{\partial})u_{l_{1}})((\partial-\bar{\partial})u_{l_{2}})\Big{]}(\partial u_{l_{3}}\bar{\partial}u_{4}+\bar{\partial}u_{l_{3}}\partial u_{4})=0. (3.25)

Consider first l3=1l_{3}=1 in the above equation. Let us first focus on the terms in (3.25) where \partial hits u1u_{1} and u2u_{2}, and ¯\bar{\partial} hits u3u_{3} and u4u_{4}. The resulting term is

2(T11+T22)(u1¯u3)(u2¯u4).\displaystyle 2(T^{11}+T^{22})(\partial u_{1}\bar{\partial}u_{3})(\partial u_{2}\bar{\partial}u_{4}). (3.26)

Now if the derivatives in (3.26) all hit the exponential factor of the solutions, we get

21h4γ02(T11+T22)e(z2z¯2)/h(1+z)(1+z)(1z¯)(1z¯)×(1+r1)(1+r2)(1+r3)(1+r4)=2h3((T11+T22)/γ0)(0)+o(h3),2\int\frac{1}{h^{4}\gamma_{0}^{2}}(T^{11}+T^{22})e^{(z^{2}-\bar{z}^{2})/h}(1+z)(-1+z)(1-\bar{z})(-1-\bar{z})\\ \times(1+r_{1})(1+r_{2})(1+r_{3})(1+r_{4})=\frac{2}{h^{3}}((T^{11}+T^{22})/\gamma_{0})(0)+o(h^{-3}),

where we used estimate (2.14) and the same argument, that begun from (3.7), involving the remainders in the case m=1m=1. Therefore, to show T11+T22=0T^{11}+T^{22}=0, it remains to show that in (3.25) all the other terms are o(h3)o(h^{-3}).

We now consider in the above case l3=1l_{3}=1 that one derivative hits the γ0\gamma_{0} term. For example, consider the following term

(1/γ0)h3γ03Te(z2z¯2)/h(1+z)(1z¯)(1z¯)(1+r1)(1+r2)(1+r3)(1+r4).\int\frac{\partial(1/\sqrt{\gamma_{0}})}{h^{3}\sqrt{\gamma_{0}}^{3}}Te^{(z^{2}-\bar{z}^{2})/h}(-1+z)(1-\bar{z})(-1-\bar{z})(1+r_{1})(1+r_{2})(1+r_{3})(1+r_{4}).

In the expansion of the above of product, the term with no remainder term is of order O(h2)O(h^{-2}) by stationary phase, and the terms with remainder term is also O(h2)O(h^{-2}) by estimate (2.14). The other cases where one or more derivatives hit γ0\gamma_{0} term instead of the exponential term are similar.

Now if one derivative hits the (1+r)(1+r) term, we consider

1h3γ02Te(z2z¯2)/h(1+z)(1z¯)(1z¯)(r1)(1+r2)(1+r3)(1+r4).\int\frac{1}{h^{3}\gamma_{0}^{2}}Te^{(z^{2}-\bar{z}^{2})/h}(-1+z)(1-\bar{z})(-1-\bar{z})(\partial r_{1})(1+r_{2})(1+r_{3})(1+r_{4}).

By estimate (2.14), we conclude that every term also in the above integral is O(h2)O(h^{-2}). Similarly for all the other cases where derivatives hit 1+r1+r term instead of the exponential term, we have the same conclusion.

Let us then consider the remaining terms in (3.25) where \partial does not hit both u1u_{1} and u2u_{2} if ¯\bar{\partial} hits both u3u_{3} and u4u_{4}, and vice versa. For example, consider terms of form

Tu1u3u2¯u4.\displaystyle T\partial u_{1}\partial u_{3}\partial u_{2}\bar{\partial}u_{4}. (3.27)

Then the term where the derivative for u3u_{3} hitting on the exponential will vanish. Therefore, we only need to consider

Te(z2z¯2)/h(1/γ0)u1u2¯u4\int Te^{(z^{2}-\bar{z}^{2})/h}\partial(1/\sqrt{\gamma}_{0})\partial u_{1}\partial u_{2}\bar{\partial}u_{4}

and

Te(z2z¯2)/h(r3)u1u2¯u4.\int Te^{(z^{2}-\bar{z}^{2})/h}(\partial r_{3})\partial u_{1}\partial u_{2}\bar{\partial}u_{4}.

By stationary phase, both integrals are O(h2)O(h^{-2}). The case l3=2l_{3}=2 is similar to l3=1l_{3}=1, thus we omit its proof.

Finally, we consider the case l3=3l_{3}=3. In this case, (ul3¯u4+¯ul3u4)(\partial u_{l_{3}}\bar{\partial}u_{4}+\bar{\partial}u_{l_{3}}\partial u_{4}) becomes (u3¯u4+¯u3u4)(\partial u_{3}\bar{\partial}u_{4}+\bar{\partial}u_{3}\partial u_{4}), so we only need to consider terms of the form

Tu1u3u2¯u4orTu1¯u3u2u4.\displaystyle T\partial u_{1}\partial u_{3}\partial u_{2}\bar{\partial}u_{4}\qquad\text{or}\qquad T\partial u_{1}\bar{\partial}u_{3}\partial u_{2}{\partial}u_{4}. (3.28)

Similar argument for (3.27) shows the above terms are also O(h2)O(h^{-2}). Therefore, we have shown all the other terms except (3.26) are o(h3)o(h^{-3}), and we get T11+T22=0T^{11}+T^{22}=0 in Ω\Omega. Combining with (3.23) and (3.24), we conclude that T11=T12=T22=0T^{11}=T^{12}=T^{22}=0 in Ω\Omega. This finishes the proof for m=2m=2.

3.3 The case m3m\geq 3:

Let us recall the integral identity for general mm: (1.5)

(l1,,lm+1)π(m+1)j1,,jm=02Tj1jm(x)(ul1,ul1)j1(ulm,ulm)jm×ulm+1um+2=0.\sum_{\left(l_{1},\ldots,l_{m+1}\right)\in\pi(m+1)}\sum_{j_{1},\ldots,j_{m}=0}^{2}\int T^{j_{1}\cdots j_{m}}(x)\left(u_{l_{1}},\nabla u_{l_{1}}\right)_{j_{1}}\ldots\left(u_{l_{m}},\nabla u_{l_{m}}\right)_{j_{m}}\\ \times\nabla u_{l_{m+1}}\cdot\nabla u_{m+2}\hskip 0.5pt=0. (3.29)

We first prove T0j1jm1=0T^{\hskip 0.5pt0\hskip 0.5ptj_{1}\hskip 0.5pt\cdots\hskip 0.5ptj_{m-1}}=0 where jk{0,1,2}j_{k}\in\{0,1,2\} and k=1,,m1k=1,\ldots,m-1. Firstly, let mm of the functions ulu_{l}, l=1,,m+1l=1,\ldots,m+1, in the identity (3.29) to be the constants functions 11. This yields

T00(x)uv=0,\int T^{\hskip 0.5pt0\hskip 0.5pt\cdots\hskip 0.5pt0}(x)\nabla u\cdot\nabla v=0, (3.30)

which is of the form (3.1) we had in the case m=1m=1. Thus we obtain T00=0T^{\hskip 0.5pt0\hskip 0.5pt\cdots\hskip 0.5pt0}=0 in Ω\Omega.

Next, we let m1m-1 of solutions ulu_{l} to be the constant functions 11. This yields

(l1,l2)π(2)j=12T00j(x)xjul1ul2u3=0,\sum_{\left(l_{1},l_{2}\right)\in\pi(2)}\sum_{j=1}^{2}\int T^{0\ldots 0j}(x)\partial_{x_{j}}u_{l_{1}}\nabla u_{l_{2}}\cdot\nabla u_{3}=0, (3.31)

which is of the form (3.8) we also had in the case m=1m=1. Thus we have T00j=0T^{\hskip 0.5pt0\hskip 0.5pt\cdots\hskip 0.5pt0\hskip 0.5ptj}=0 in Ω\Omega for j=1,2j=1,2. Continuing in similar fashion, we let m2m-2 of the solutions uiu_{i} to be the constant functions 11. This yields an integral identity similar to (3.21) we had in the case m=2m=2. The same argument used in that case proves T00j1j2=0T^{\hskip 0.5pt0\hskip 0.5pt\cdots\hskip 0.5pt0\hskip 0.5ptj_{1}\hskip 0.5ptj_{2}}=0 in Ω\Omega for j1,j2{1,2}j_{1},j_{2}\in\{1,2\}. Proceeding in this manner, by induction we obtain

T0j1jm1=0,j1,,jm1{0,1,2}T^{\hskip 0.5pt0\hskip 0.5ptj_{1}\hskip 0.5pt\cdots\hskip 0.5ptj_{m-1}}=0,\quad j_{1},\ldots,j_{m-1}\in\{0,1,2\} (3.32)

in Ω\Omega.

It remains to prove Tj1jm=0T^{j_{1}\cdots j_{m}}=0 where all the indices jkj_{k}, k=1,,mk=1,\ldots,m, are nonzero. By (3.32), the integral identity (3.29) is reduced to

(l1,,lm+1)π(m+1)j1,jm=12Tj1jm(x)xj1ul1xj2ul2xjmulmulm+1um+2=0\displaystyle\sum_{\left(l_{1},\cdots,l_{m+1}\right)\in\pi(m+1)}\sum_{j_{1},\ldots j_{m}=1}^{2}\int T^{j_{1}\ldots j_{m}}(x)\partial_{x_{j_{1}}}u_{l_{1}}\partial_{x_{j_{2}}}u_{l_{2}}\ldots\partial_{x_{j_{m}}}u_{l_{m}}\nabla u_{l_{m+1}}\cdot\nabla u_{m+2}=0 (3.33)

for all uu solving (3.2). Since Tj1jmT^{j_{1}\cdots j_{m}} is symmetric in exchange of any of its two indices, it has (d+m1)!/(m!(d1)!)(d+m-1)!/(m!(d-1)!) independent components, where d=2d=2. Thus we have m+1m+1 unknown entries Tj1jmT^{j_{1}\cdots j_{m}} in (3.33). To recover these entries we will find m+1m+1 linearly independent equations for the entries. Firstly, we choose CGO solutions such that exactly one of them has an antiholomorphic phase:

u1=u2==um\displaystyle{u_{1}}=u_{2}=\cdots=u_{m} =1γ0e(z+1m+1z2)/h(1+r1),\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(z+\frac{1}{m+1}z^{2})/h}(1+r_{1}),
um+1\displaystyle{u_{m+1}} =1γ0e(mz+1m+1z2)/h(1+r2),\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(-mz+\frac{1}{m+1}z^{2})/h}(1+r_{2}),
um+2\displaystyle{u_{m+2}} =1γ0ez¯2/h(1+r3).\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{-\bar{z}^{2}/h}(1+r_{3}).

Following how we proved T1+iT2=0T^{1}+iT^{2}=0 in m=1m=1 case, we can show that the principal order term in (3.33) is O(hm)O(h^{-m}), and the integrals involving correction terms r1,r2,r3r_{1},r_{2},r_{3} are o(hm)o(h^{-m}). Stationary phase shows that the principal order term of (3.33) gives the linear equation

(m0)T11+i(m1)T112+i2(m2)T1122++im(mm)T22\displaystyle\binom{m}{0}T^{1\ldots 1}+i\binom{m}{1}T^{1\ldots 12}+i^{2}\binom{m}{2}T^{1\ldots 122}+\ldots+i^{m}\binom{m}{m}T^{2\ldots 2} =0,\displaystyle=0, (3.34)

after dividing by a nonzero constant. To see how (3.34) is obtained, we first note that the coefficients come from expanding the integral

j1,jm=12¯Tj1jm(x)(xj1ul1)(xj2ul2)(xjmulm)(ulm+1um+2)\sum_{j_{1},\ldots j_{m}=1}^{2}\int\bar{\partial}T^{j_{1}\cdots j_{m}}(x)(\partial\partial_{x_{j_{1}}}u_{l_{1}})(\partial_{x_{j_{2}}}u_{l_{2}})\ldots(\partial_{x_{j_{m}}}u_{l_{m}})(u_{l_{m+1}}u_{m+2}) (3.35)

by stationary phase. Here the principal order term results from the solutions \partial hitting u1,u2,,um+1u_{1},u_{2},\ldots,u_{m+1} and ¯\bar{\partial} hitting the solution um+2u_{m+2}, which have holomorphic and antiholomorphic phases respectively. This is similar to what we had in the proof for the case m=1m=1. We also used

1=2+¯,2=i(2¯).\partial\partial_{1}=\partial^{2}+\partial\bar{\partial},\quad\partial\partial_{2}=i(\partial^{2}-\partial\bar{\partial}). (3.36)

to compute the exact coefficients.

Similarly, if we choose CGO solutions such that exactly one of them has holomorphic phase, we get the following linear equation up to a scalar multiple:

(m0)T11+(i)(m1)T112+(i)2(m2)T1122++(i)m(mm)T22\displaystyle\binom{m}{0}T^{\hskip 0.5pt1\hskip 0.5pt\cdots\hskip 0.5pt1}+(-i)\binom{m}{1}T^{\hskip 0.5pt1\hskip 0.5pt\cdots\hskip 0.5pt1\hskip 0.5pt2}+(-i)^{2}\binom{m}{2}T^{\hskip 0.5pt1\hskip 0.5pt\cdots\hskip 0.5pt1\hskip 0.5pt2\hskip 0.5pt2}+\cdots+(-i)^{m}\binom{m}{m}T^{\hskip 0.5pt2\hskip 0.5pt\cdots\hskip 0.5pt2} =0.\displaystyle=0. (3.37)

Next we choose CGO solutions so that more than one of the solutions have holomorphic phases and also that more than one solution have antiholomorphic phases. Therefore, we can choose every solution to be of form (2.10) or (2.12) whose phase functions have no critical points. For general mm, we choose CGOs to that all their phases add up to z2z¯2z^{2}-\bar{z}^{2}. The explicit formula for general mm is complicated to write down. Therefore, we only consider the case m=3m=3 as an example. In this case, we choose two solutions with holomorphic phases and three with antiholomorphic phases:

u1\displaystyle{u_{1}} =1γ0e(z+12z2)/h(1+r1),\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(z+\frac{1}{2}z^{2})/h}(1+r_{1}),
u2\displaystyle{u_{2}} =1γ0e(z+12z2)/h(1+r2),\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(-z+\frac{1}{2}z^{2})/h}(1+r_{2}),
u3\displaystyle{u_{3}} =1γ0e(z¯13z¯2)/h(1+r3~),\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(-\bar{z}-\frac{1}{3}\bar{z}^{2})/h}(1+\tilde{r_{3}}),
u4\displaystyle{u_{4}} =1γ0e(z¯13z¯2)/h(1+r4~),\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(-\bar{z}-\frac{1}{3}\bar{z}^{2})/h}(1+\tilde{r_{4}}),
u5\displaystyle{u_{5}} =1γ0e(2z¯13z¯2)/h(1+r5~).\displaystyle=\frac{1}{\sqrt{\gamma_{0}}}e^{(2\bar{z}-\frac{1}{3}\bar{z}^{2})/h}(1+\tilde{r_{5}}).

Note that here all the solutions have phases without critical points. Consequently, their correction terms satisfy the better estimates (2.14), which simplifies the asymptotic analysis.

By arguing similarly as we did after (3.25), we obtain

T111iT112+T122iT222\displaystyle T^{111}-iT^{112}+T^{122}-iT^{222} =0.\displaystyle=0. (3.38)

Note that the coefficients in (3.38) agree with those in the expansion of the polynomial (a+ib)(aib)2(a+ib)(a-ib)^{2} of variables aa and bb. This is true in general: Choose tt, 1tm+11\leq t\leq m+1, solutions to have holomorphic phases and m+2tm+2-t solutions to have antiholomorphic phases in (3.29). Then, by stationary phase, we may compute the coefficient of T1122T^{\hskip 0.5pt1\hskip 0.5pt\cdots\hskip 0.5pt1\hskip 0.5pt2\hskip 0.5pt\cdots 2}, where the number of indices with index 11 is ss and the number of indices with index 22 is (ms)(m-s). The coefficient will agree with the coefficient of asbmsa^{s}b^{m-s} in the expansion of (a+ib)t1(aib)m+2t1(a+ib)^{t-1}(a-ib)^{m+2-t-1}. We explain next why the above holds.

The reason why the above holds is the following: The principal order term of the integral (3.35) for the chosen solutions corresponds to \partial hitting solutions with holomorphic phases and ¯\bar{\partial} hitting solutions with antiholomorphic phases. (See the part of the proof after (3.25).) Then, since

1=+¯,2=i(¯),\partial_{1}=\partial+\bar{\partial},\quad\partial_{2}=i(\partial-\bar{\partial}),

we know that 1\partial_{1} acting on a holomorphic phase gives the coefficient 11, 2\partial_{2} acting on holomorphic phase gives a coefficient of ii, 1\partial_{1} acting on antiholomorphic phase gives a coefficient of 11 and 2\partial_{2} acting on antiholomorphic phase gives a coefficient of i-i.

To compute the coefficient of T1122T^{\hskip 0.5pt1\hskip 0.5pt\cdots\hskip 0.5pt1\hskip 0.5pt2\hskip 0.5pt\cdots\hskip 0.5pt2}, where the number of indices 11 is ss, and the number of indices 22 is (ms)(m-s), we note the following. In the integral (3.33), the coefficient T1122T^{\hskip 0.5pt1\hskip 0.5pt\cdots\hskip 0.5pt1\hskip 0.5pt2\hskip 0.5pt\cdots\hskip 0.5pt2} appears together with ss instances of 1\partial_{1} and msm-s instances of 2\partial_{2}. In the principal order term, if uu is a solution with holomorphic phase, we may only consider the terms where we have u\partial u. So for the term 1u=(+¯)u\partial_{1}u=(\partial+\bar{\partial})u following T1122T^{\hskip 0.5pt1\hskip 0.5pt\cdots\hskip 0.5pt1\hskip 0.5pt2\hskip 0.5pt\cdots\hskip 0.5pt2} in (3.33), we may consider only u\partial u, while in the term 2u=i(¯)u\partial_{2}u=i(\partial-\bar{\partial})u, we may consider only iui\partial u.

We have similarly for uu with antiholomorphic phase: For 1u=(+¯)u\partial_{1}u=(\partial+\bar{\partial})u, we consider only ¯u\bar{\partial}u, while in the term 2u=i(¯)u\partial_{2}u=i(\partial-\bar{\partial})u, we consider only i¯u-i\bar{\partial}u. This implies each of 1\partial_{1} corresponds to a factor of 11 for both holomorphic phase and antiholomorphic phase, and each 2\partial_{2} corresponds to a factor of ii for holomorphic phase, and a factor of i-i for antiholomorphic phase. From the proof after (3.25)\eqref{eqT11+T22}, we see there are t1t-1 solutions having holomorphic phases and m+2t1m+2-t-1 solutions having antiholomorphic phases among ul1,,ulmu_{l_{1}},\ldots,u_{l_{m}}. Therefore, the coefficient that we are considering should be equal to that of asbmsa^{s}b^{m-s} in the polynomial (a+ib)t1(aib)m+2t1(a+ib)^{t-1}(a-ib)^{m+2-t-1}.

Note that in the case of mm, we have m+1m+1 many choices of solutions. Each choice gives a linear equation of the form (3.34). Finally, we show that the m+1m+1 linear equations we have now obtained for the coefficients Tj1j2j3T^{j_{1}j_{2}j_{3}} are linearly independent. This implies that the coefficients are uniquely determined. Let us inspect the linear system we obtain for m=3m=3. This is

T111+3iT1123T122iT222\displaystyle T^{111}+3iT^{112}-3T^{122}-iT^{222} =0\displaystyle=0
T111iT112T122iT222\displaystyle-T^{111}-iT^{112}-T^{122}-iT^{222} =0\displaystyle=0
T111iT112+T122iT222\displaystyle T^{111}-iT^{112}+T^{122}-iT^{222} =0\displaystyle=0
T111+3iT112+3T122iT222\displaystyle-T^{111}+3iT^{112}+3T^{122}-iT^{222} =0.\displaystyle=0.

As we have shown above, the coefficients in each row agree with those of (a+ib)3(a+ib)^{3}, (a+ib)2(aib)(a+ib)^{2}(a-ib), (a+ib)(aib)2(a+ib)(a-ib)^{2} and (aib)3(a-ib)^{3} respectively. Since these polynomials are linearly independent, so is the coefficient matrix of the above linear system. The proof for general mm is similar.

References

  • [AZ21] Yernat M Assylbekov and Ting Zhou. Direct and inverse problems for the nonlinear time-harmonic Maxwell equations in Kerr-type media. J. Spectr. Theory, 11:1–38, 2021.
  • [Buk08] A. Bukhgeim. Recovering a potential from cauchy data in the two-dimensional case. J.Inv. Ill-Posed Problems, 16:19–33, 2008.
  • [CFK+21] C. Cârstea, A. Feizmohammadi, Y. Kian, K. Krupchyk, and G. Uhlmann. The caldeón inverse problem for isotropic quasilinear conductivities. Advances in Mathematics, 391, 2021.
  • [CLLO24] Cătălin I Cârstea, Matti Lassas, Tony Liimatainen, and Lauri Oksanen. An inverse problem for the riemannian minimal surface equation. Journal of Differential Equations, 379:626–648, 2024.
  • [CLT24] Cătălin I Cârstea, Tony Liimatainen, and Leo Tzou. The calder\\backslash’on problem on riemannian surfaces and of minimal surfaces. arXiv preprint arXiv:2406.16944, 2024.
  • [CNV19] Cătălin I Cârstea, Gen Nakamura, and Manmohan Vashisth. Reconstruction for the coefficients of a quasilinear elliptic partial differential equation. Applied Mathematics Letters, 98:121–127, 2019.
  • [EPS14] Herbert Egger, Jan-Frederik Pietschmann, and Matthias Schlottbom. Simultaneous identification of diffusion and absorption coefficients in a quasilinear elliptic problem. Inverse Problems, 30(3):035009, 2014.
  • [FLL23] Ali Feizmohammadi, Tony Liimatainen, and Yi-Hsuan Lin. An inverse problem for a semilinear elliptic equation on conformally transversally anisotropic manifolds. Annals of PDE, 9(2):12, 2023.
  • [FO20] Ali Feizmohammadi and Lauri Oksanen. An inverse problem for a semi-linear elliptic equation in riemannian geometries. Journal of Differential Equations, 269(6):4683–4719, 2020.
  • [GT01] David Gilbarg and Neil S. Trudinger. Elliptic Partial Differential Equations of Second Order. Springer Berlin, Heidelberg, 2001.
  • [GT11] C. Guillarmou and L. Tzou. Identification of a connection from cauchy data on a riemann surface with boundary. Geometric and Functional Analysis, 21(2):393–418, 2011.
  • [HS02] David Hervas and Ziqi Sun. An inverse boundary value problem for quasilinear elliptic equations. 2002.
  • [HT13] B. Haberman and D. Tataru. Uniqueness in calderón’s problem with lipschitz conductivities. Duke Math. J, 162:497–516, 2013.
  • [IN95] V. Isakov and A. Nachman. Global uniqueness for a two-dimensional semilinear elliptic inverse problem. Trans. Am. Math. Soc., 347:3375–3390, 1995.
  • [IS94] V. Isakov and J. Sylvester. Global uniqueness for a semilinear elliptic inverse problem. Commun. Pure Appl. Math., 47:1403–1410, 1994.
  • [Isa93] Victor Isakov. On uniqueness in inverse problems for semilinear parabolic equations. Archive for Rational Mechanics and Analysis, 124(1):1–12, 1993.
  • [IUY12] Oleg Yu Imanuvilov, Gunther Uhlmann, and Masahiro Yamamoto. Partial cauchy data for general second order elliptic operators in two dimensions. Publications of the Research Institute for Mathematical Sciences, 48(4):971–1055, 2012.
  • [KKU22] Yavar Kian, Katya Krupchyk, and Gunther Uhlmann. Partial data inverse problems for quasilinear conductivity equations. Mathematische Annalen, pages 1–28, 2022.
  • [KLU18] Yaroslav Kurylev, Matti Lassas, and G. Uhlmann. Inverse problems for Lorentzian manifolds and non-linear hyperbolic equations. Invent. Math., 212(3):781–857, 2018.
  • [KN02] Hyeonbae Kang and Gen Nakamura. Identification of nonlinearity in a conductivity equation via the Dirichlet-to-Neumann map. Inverse Problems, 18:1079–1088, 2002.
  • [KU20a] Katya Krupchyk and Gunther Uhlmann. Partial data inverse problems for semilinear elliptic equations with gradient nonlinearities. Mathematical Research Letters, 27(6):1801–1824, 2020.
  • [KU20b] Katya Krupchyk and Gunther Uhlmann. A remark on partial data inverse problems for semilinear elliptic equations. Proc. Amer. Math. Soc., 148:681–685, 2020.
  • [LL24] Tony Liimatainen and Yi-Hsuan Lin. Uniqueness results for inverse source problems for semilinear elliptic equations. Inverse Problems, 40(4):045030, 2024.
  • [LLLS20] Matti Lassas, Tony Liimatainen, Yi-Hsuan Lin, and Mikko Salo. Partial data inverse problems and simultaneous recovery of boundary and coefficients for semilinear elliptic equations. Revista Matemática Iberoamericana, 37(4):1553–1580, 2020.
  • [LLLS21] Matti Lassas, Tony Liimatainen, Yi-Hsuan Lin, and Mikko Salo. Inverse problems for elliptic equations with power type nonlinearities. Journal de Mathématiques Pures et Appliquées, 145:44–82, 2021.
  • [LLST22] Tony Liimatainen, Yi-Hsuan Lin, Mikko Salo, and Teemu Tyni. Inverse problems for elliptic equations with fractional power type nonlinearities. Journal of Differential Equations, 306:189–219, 2022.
  • [MU20] Claudio Munoz and Gunther Uhlmann. The calderón problem for quasilinear elliptic equations. In Annales de l’Institut Henri Poincaré C, Analyse non linéaire, volume 37, pages 1143–1166. Elsevier, 2020.
  • [Nac96] A. Nachman. Global uniqueness theorem for a two-dimensional inverse boundary value problem. Ann. of Math., 143(1):71–96, 1996.
  • [Nov88] R. G. Novikov. A multidimensional inverse spectral problem for the equation Δψ+(v(x)Eu(x))ψ=0-\Delta\psi+(v(x)-Eu(x))\psi=0. (Russian) Funktsional. Anal. i Prilozhen. Translation in Funct. Anal. Appl. 22 (1988), no. 4, 263–272 (1989)., 22(4):11–22, 1988.
  • [Nur23a] J. Nurminen. An inverse problem for the minimal surface equation. Nonlinear Analysis, 227:113163, 2023.
  • [Nur23b] J. Nurminen. An inverse problem for the minimal surface equation in the presence of a riemannian metric. arXiv preprint arXiv:2304.05808, 2023.
  • [PA06] L. Päivärinta and K. Astala. Calderón’s inverse conductivity problem in the plane. Ann. of Math., 163(1):265–299, 2006.
  • [Sha20] Ravi Shankar. Recovering a quasilinear conductivity from boundary measurements. Inverse problems, 37(1):015014, 2020.
  • [SU87] J. Sylvester and G. Uhlmann. A global uniqueness theorem for an inverse boundary value problem. Ann. of Math., 125(1):153–169, 1987.
  • [SU97] Ziqi Sun and Gunther Uhlmann. Inverse problems in quasilinear anisotropic media. American Journal of Mathematics, 119(4):771–797, 1997.
  • [Sun96] Z. Sun. On a quasilinear inverse boundary value problem. Mathematische Zeitschrift, 2(221):293–305, 1996.
  • [Sun10] Ziqi Sun. An inverse boundary-value problem for semilinear elliptic equations. Electronic Journal of Differential Equations (EJDE)[electronic only], 2010:Paper–No, 2010.
  • [Vek62] I. Vekua. Generalized Analytic Functions. Pergamon Press, 1962.