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

On inverse problems arising in fractional elasticity

Li Li Institute for Pure and Applied Mathematics, University of California,
Los Angeles, CA 90095, USA

ABSTRACT.  We first formulate an inverse problem for a linear fractional Lamé system. We determine the Lamé parameters from exterior partial measurements of the Dirichlet-to-Neumann map. We further study an inverse obstacle problem as well as an inverse problem for a nonlinear fractional Lamé system. Our arguments are based on the unique continuation property for the fractional operator as well as the associated Runge approximation property.

1 Introduction

The classical Lamé operator Lλ,μL_{\lambda,\mu} for a three-dimensional isotropic elastic body is given by

(Lλ,μu)i:=j=13i(λuj,j)+j=13j(μ(ui,j+uj,i)),(1i3)(L_{\lambda,\mu}u)_{i}:=\sum^{3}_{j=1}\partial_{i}(\lambda u_{j,j})+\sum^{3}_{j=1}\partial_{j}(\mu(u_{i,j}+u_{j,i})),\quad(1\leq i\leq 3)

where uiu_{i} denotes the ii-th component of the vector-valued displacement function uu and ui,j:=juiu_{i,j}:=\partial_{j}u_{i}. The associated inverse problem has been studied in [9, 20] where the authors considered the Dirichlet problem

Lλ,μu=0inΩ,u=gonΩ.L_{\lambda,\mu}u=0\,\,\,\,\text{in}\,\,\Omega,\qquad u=g\,\,\,\,\text{on}\,\,\partial\Omega.

They determined the variable Lamé parameters λ,μ\lambda,\mu from the Dirichlet-to-Neumann (displacement-to-traction) map

(Λg)i:=j=13(λuj,j)νi+j=13μ(ui,j+uj,i)νj,(1i3)(\Lambda g)_{i}:=\sum^{3}_{j=1}(\lambda u_{j,j})\nu_{i}+\sum^{3}_{j=1}\mu(u_{i,j}+u_{j,i})\nu_{j},\quad(1\leq i\leq 3)

where ν\nu is the unit outer normal to Ω\partial\Omega under certain assumptions on λ,μ\lambda,\mu.

In this paper, we study a fractional analogue of Lλ,μL_{\lambda,\mu} and its associated inverse problems.

First we recall that the classical elastic model is based on the constitutive relation

σij=k,l=13Cijklekl\sigma_{ij}=\sum_{k,l=1}^{3}C_{ijkl}e_{kl}

where the fourth-order elastic stiffness tensor is

Cijkl:=λδijδkl+μ(δikδjl+δilδjk),C_{ijkl}:=\lambda\delta_{ij}\delta_{kl}+\mu(\delta_{ik}\delta_{jl}+\delta_{il}\delta_{jk}),

the linearized strain tensor is

eij:=12(ui,j+uj,i)e_{ij}:=\frac{1}{2}(u_{i,j}+u_{j,i})

and σij\sigma_{ij} denotes the stress tensor. Here δij\delta_{ij} is the standard Kronecker delta.

Recently, the theory of nonlocal elasticity has attracted much attention. The integral linear constitutive relation

σij=k,l=13Cijkl(Kekl)\sigma_{ij}=\sum_{k,l=1}^{3}C_{ijkl}(K*e_{kl})

has been introduced to describe complex materials characterized by nonlocality. Then the fractional Taylor series approximation for the Fourier transform of the interaction kernel KK, which is given by

K^(|ξ|)K^(0)+cs|ξ|2s,(0<s<1)\hat{K}(|\xi|)\approx\hat{K}(0)+c_{s}|\xi|^{2s},\quad(0<s<1)

leads to the definition of the fractional Lamé operator

:=Lλ,μ+(Δ)sLλ0,μ0.\mathcal{L}:=L_{\lambda,\mu}+(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}. (1)

See [24] and the references there for more background information. In [24], λ,μ\lambda,\mu are constants proportional to λ0,μ0\lambda_{0},\mu_{0} but in this paper we allow λ,μ\lambda,\mu to be variable functions.

We consider the exterior Dirichlet problem

u=0inΩ,u=ginΩe\mathcal{L}u=0\,\,\,\,\text{in}\,\,\Omega,\qquad u=g\,\,\,\,\text{in}\,\,\Omega_{e} (2)

where Ω\Omega is a bounded Lipschitz domain and Ωe:=3Ω¯\Omega_{e}:=\mathbb{R}^{3}\setminus\bar{\Omega}. Under appropriate assumptions on λ0,μ0,λ,μ\lambda_{0},\mu_{0},\lambda,\mu, we can show its well-posedness so we will be able to define the associated Dirichlet-to-Neumann map Λ\Lambda, which is formally given by

Λg:=(Δ)sLλ0,μ0ug|Ωe.\Lambda g:=(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}u_{g}|_{\Omega_{e}}. (3)

Our goal here is to determine both λ\lambda and μ\mu from exterior partial measurements of Λ\Lambda.

We remark that our problem can be viewed as a variant of the fractional Calderón problem first introduced in [13] where the authors considered the exterior Dirichlet problem

((Δ)s+q)u=0inΩ,u=ginΩe((-\Delta)^{s}+q)u=0\,\,\,\text{in}\,\,\Omega,\qquad u=g\,\,\,\,\text{in}\,\,\Omega_{e}

and they proved the fundamental uniqueness theorem that the potential qq in Ω\Omega can be determined from exterior partial measurements of the map

Λq:g(Δ)sug|Ωe.\Lambda_{q}:g\to(-\Delta)^{s}u_{g}|_{\Omega_{e}}.

It has been shown that the knowledge of Λqg\Lambda_{q}g is equivalent to the knowledge of the nonlocal Neumann derivative of ugu_{g} (see [13] for more details). Hence our problem can also be viewed as a nonlocal analogue of the inverse problem for the classical Lamé system.

We mention that inverse problems for fractional operators have been extensively studied so far. See [23] for low regularity and stability results for the fractional Calderón problem. See [12] for reconstruction and single measurement results for the fractional Calderón problem. See [11] for inverse problems for variable coefficients fractional elliptic operators. See [1, 6] for inverse problems for fractional Schrödinger operators with local and non-local perturbations. See [5, 16, 17] for inverse problems for fractional magnetic operators. See [15, 18, 19] for inverse problems for fractional parabolic operators.

The following theorem is our first main result in this paper.

Theorem 1.1.

Let 0<s<10<s<1. Let μ0>0\mu_{0}>0 and λ0+μ00\lambda_{0}+\mu_{0}\geq 0. Let 0λ(j),μ(j)C1(Ω¯)0\leq\lambda^{(j)},\mu^{(j)}\in C^{1}(\bar{\Omega}) and let WjΩeW_{j}\subset\Omega_{e} be nonempty and open (j=1,2j=1,2). Suppose

Λ(1)g|W2=Λ(2)g|W2\Lambda^{(1)}g|_{W_{2}}=\Lambda^{(2)}g|_{W_{2}}

for all gCc(W1)g\in C^{\infty}_{c}(W_{1}). Then λ(1)=λ(2),μ(1)=μ(2)\lambda^{(1)}=\lambda^{(2)},\mu^{(1)}=\mu^{(2)} in Ω\Omega.

We remark that our problem provides an example which suggests that the inverse problem for the fractional operator is more manageable than its classical counterpart.

Recall that to solve the classical inverse problem, we first reduce the Lamé system to a first order system perturbation of the Laplacian. Then we construct complex geometrical optics (CGO) solutions and apply the integral identity to obtain the uniqueness of λ,μ\lambda,\mu. For some technical reasons, the uniqueness result is only proved provided that μ\mu is close to a constant. The full classical problem remains open. See [9, 21] for details.

Here such a priori knowledge of μ\mu is not required for solving the fractional problem. Instead of constructing CGO solutions, we will use the unique continuation property and the Runge approximation property associated with our fractional operator to prove the strong uniqueness result. This scheme was first introduced in [13] for solving the fractional Calderón problem.

We further study an inverse obstacle problem associated with our fractional operator.

We consider the following obstacle problem

u=0inΩD¯,u=0inD,u=ginΩe\mathcal{L}u=0\,\,\text{in}\,\,\Omega\setminus\bar{D},\quad u=0\,\,\text{in}\,\,D,\quad u=g\,\,\text{in}\,\,\Omega_{e} (4)

where DΩD\subset\Omega is a nonempty open set satisfying that ΩD¯\Omega\setminus\bar{D} is a bounded Lipschitz domain.

As we did for the exterior problem (2), we can similarly show the well-posedness of (4) and define the Dirichlet-to-Neumann map ΛD\Lambda_{D} by

ΛDg:=(Δ)sLλ0,μ0ug|Ωe.\Lambda_{D}g:=(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}u_{g}|_{\Omega_{e}}. (5)

Our next goal is to determine D,λ,μD,\lambda,\mu from the knowledge of ΛD\Lambda_{D}.

The following theorem is our second main result.

Theorem 1.2.

Let 0<s<10<s<1. Let μ0>0\mu_{0}>0 and λ0+μ00\lambda_{0}+\mu_{0}\geq 0. Let DjΩD_{j}\subset\Omega be nonempty and open s.t. ΩDj¯\Omega\setminus\bar{D_{j}} is a bounded Lipschitz domain, 0λ(j),μ(j)C1(Ω¯Dj)0\leq\lambda^{(j)},\mu^{(j)}\in C^{1}(\bar{\Omega}\setminus D_{j}) and let WjΩeW_{j}\subset\Omega_{e} be nonempty and open (j=1,2j=1,2). Suppose

ΛD1(1)g|W2=ΛD2(2)g|W2\Lambda^{(1)}_{D_{1}}g|_{W_{2}}=\Lambda^{(2)}_{D_{2}}g|_{W_{2}}

for a nonzero gCc(W1)g\in C^{\infty}_{c}(W_{1}). Then D1=D2=:DD_{1}=D_{2}=:D. Further assume the identity holds for all gCc(W1)g\in C^{\infty}_{c}(W_{1}). Then λ(1)=λ(2)\lambda^{(1)}=\lambda^{(2)} and μ(1)=μ(2)\mu^{(1)}=\mu^{(2)} in ΩD¯\Omega\setminus\bar{D}.

We also study an inverse problem for a nonlinear fractional Lamé system.

We consider the following nonlinear exterior problem

u+𝒩u=0inΩ,u=ginΩe\mathcal{L}u+\mathcal{N}u=0\,\,\,\,\text{in}\,\,\Omega,\qquad u=g\,\,\,\,\text{in}\,\,\Omega_{e} (6)

where the nonlinear operator 𝒩\mathcal{N} is given by

(𝒩u)i:=j=13jNiju,(1i3)(\mathcal{N}u)_{i}:=\sum^{3}_{j=1}\partial_{j}N_{ij}u,\quad(1\leq i\leq 3) (7)
Niju:=λ+2m,num,n2δij+𝒞(mum,m)2δij+2m,num,nun,mδijN_{ij}u:=\frac{\lambda+\mathcal{B}}{2}\sum_{m,n}u^{2}_{m,n}\delta_{ij}+\mathcal{C}(\sum_{m}u_{m,m})^{2}\delta_{ij}+\frac{\mathcal{B}}{2}\sum_{m,n}u_{m,n}u_{n,m}\delta_{ij}
+mum,muj,i+𝒜4muj,mum,i+(λ+)mum,mui,j+\mathcal{B}\sum_{m}u_{m,m}u_{j,i}+\frac{\mathcal{A}}{4}\sum_{m}u_{j,m}u_{m,i}+(\lambda+\mathcal{B})\sum_{m}u_{m,m}u_{i,j}
+(μ+𝒜4)m(um,ium,j+ui,muj,m+ui,mum,j).+(\mu+\frac{\mathcal{A}}{4})\sum_{m}(u_{m,i}u_{m,j}+u_{i,m}u_{j,m}+u_{i,m}u_{m,j}).

This nonlinearity comes from the higher order expansion of the energy density as well as the nonlinear term in the strain tensor

eij:=12(ui,j+uj,i+mum,ium,j).e_{ij}:=\frac{1}{2}(u_{i,j}+u_{j,i}+\sum_{m}u_{m,i}u_{m,j}).

We remark that our 𝒩\mathcal{N} is the static version of the nonlinearity considered in [7, 25], where the inverse problem for the associated nonlinear elastic wave equation was studied. See the references there for more background information on 𝒩\mathcal{N}.

Under certain assumptions, we can show the well-posedness of (6) for small gCc(Ωe)g\in C^{\infty}_{c}(\Omega_{e}) and then for such gg we can define the associated Dirichlet-to-Neumann map ΛN\Lambda_{N} formally given by

ΛNg:=(Δ)sLλ0,μ0ug|Ωe.\Lambda_{N}g:=(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}u_{g}|_{\Omega_{e}}. (8)

Our last goal is to determine λ,μ,𝒜,𝒞\lambda,\mu,\mathcal{A},\mathcal{C} from the knowledge of ΛN\Lambda_{N}.

The following theorem is our third main result.

Theorem 1.3.

Let 12s<1\frac{1}{2}\leq s<1. Let μ0>0\mu_{0}>0 and λ0+μ00\lambda_{0}+\mu_{0}\geq 0 and let C1(Ω¯)\mathcal{B}\in C^{1}(\bar{\Omega}). Let 0λ(j),μ(j)C1(Ω¯)0\leq\lambda^{(j)},\mu^{(j)}\in C^{1}(\bar{\Omega}), let 𝒜(j),𝒞(j)C1(Ω¯)\mathcal{A}^{(j)},\mathcal{C}^{(j)}\in C^{1}(\bar{\Omega}) and let WjΩeW_{j}\subset\Omega_{e} be nonempty and open (j=1,2j=1,2). Suppose

ΛN(1)g|W2=ΛN(2)g|W2\Lambda^{(1)}_{N}g|_{W_{2}}=\Lambda^{(2)}_{N}g|_{W_{2}}

for small gCc(W1)g\in C^{\infty}_{c}(W_{1}). Then λ(1)=λ(2),μ(1)=μ(2),𝒜(1)=𝒜(2),𝒞(1)=𝒞(2)\lambda^{(1)}=\lambda^{(2)},\mu^{(1)}=\mu^{(2)},\mathcal{A}^{(1)}=\mathcal{A}^{(2)},\mathcal{C}^{(1)}=\mathcal{C}^{(2)} in Ω\Omega.

Note that here we only claim that 𝒜,𝒞\mathcal{A},\mathcal{C} can be determined for a fixed \mathcal{B}. The question whether we can simultaneously determine 𝒜,,𝒞\mathcal{A},\mathcal{B},\mathcal{C} is still open.

The rest of this paper is organized in the following way. In Section 2, we summarize the background knowledge. In Section 3, we show the well-posedness of the linear exterior problem; We prove the unique continuation property and the Runge approximation property associated with our fractional operator; Then we prove Theorem 1.1 and Theorem 1.2. In Section 4, we show the well-posedness of the nonlinear exterior problem for small exterior data; We combine linearization arguments with the Runge approximation property to prove Theorem 1.3.

Acknowledgements. The author would like to thank Professor Gunther Uhlmann for suggesting the problem and for helpful discussions.

2 Preliminaries

Throughout this paper we use the following notations.

  • We fix the space dimension n=3n=3 and the fractional power 0<s<10<s<1.

  • x=(x1,x2,x3)x=(x_{1},x_{2},x_{3}) denotes the spatial variable.

  • For vector-valued function uu, uiu_{i} denotes the ii-th component of uu and ui,j:=juiu_{i,j}:=\partial_{j}u_{i}.

  • Ω\Omega denotes a bounded Lipschitz domain and Ωe:=nΩ¯\Omega_{e}:=\mathbb{R}^{n}\setminus\bar{\Omega}.

  • δij\delta_{ij} denotes the standard Kronecker delta.

  • ,\langle\cdot,\cdot\rangle denotes the distributional pairing so formally, f,g=fg\langle f,g\rangle=\int fg.

Throughout this paper we refer all function spaces to real-valued function spaces. For convenience, we use the same notation for the scalar-valued function space and the vector-valued one. For instance, Cc(Ω)C^{\infty}_{c}(\Omega) can be either Cc(Ω;)C^{\infty}_{c}(\Omega;\mathbb{R}) or Cc(Ω;3)C^{\infty}_{c}(\Omega;\mathbb{R}^{3}).

2.1 Sobolev spaces

For rr\in\mathbb{R}, we have the Sobolev space

Hr(n):={f𝒮(n):n(1+|ξ|2)r|f(ξ)|2𝑑ξ<}H^{r}(\mathbb{R}^{n}):=\{f\in\mathcal{S}^{\prime}(\mathbb{R}^{n}):\int_{\mathbb{R}^{n}}(1+|\xi|^{2})^{r}|\mathcal{F}f(\xi)|^{2}d\xi<\infty\}

where \mathcal{F} is the Fourier transform and 𝒮(n)\mathcal{S}^{\prime}(\mathbb{R}^{n}) is the space of temperate distributions.

We have the natural identification

Hr(n)=Hr(n).H^{-r}(\mathbb{R}^{n})=H^{r}(\mathbb{R}^{n})^{*}.

Let UU be an open set in n\mathbb{R}^{n}. Let FF be a closed set in n\mathbb{R}^{n}. Then

Hr(U):={u|U:uHr(n)},HFr(n):={uHr(n):suppuF},H^{r}(U):=\{u|_{U}:u\in H^{r}(\mathbb{R}^{n})\},\qquad H^{r}_{F}(\mathbb{R}^{n}):=\{u\in H^{r}(\mathbb{R}^{n}):\mathrm{supp}\,u\subset F\},
H~r(U):=theclosureofCc(U)inHr(n).\tilde{H}^{r}(U):=\mathrm{the\,\,closure\,\,of}\,\,C^{\infty}_{c}(U)\,\,\mathrm{in}\,\,H^{r}(\mathbb{R}^{n}).

Since Ω\Omega is a bounded Lipschitz domain, we also have the identifications

H~r(Ω)=HΩ¯r(n),Hr(Ω)=H~r(Ω).\tilde{H}^{r}(\Omega)=H^{r}_{\bar{\Omega}}(\mathbb{R}^{n}),\quad H^{-r}(\Omega)=\tilde{H}^{r}(\Omega)^{*}.

Let 0<s<10<s<1. It is well known that we have the following continuous embedding

H~s(Ω)L2nn2s(Ω).\tilde{H}^{s}(\Omega)\hookrightarrow L^{\frac{2n}{n-2s}}(\Omega).

See for instance, Section 1.5 in [2]. It has also been proved that we have the continuous embedding

Ln2s(Ω)M(HsHs)L^{\frac{n}{2s}}(\Omega)\hookrightarrow M(H^{s}\to H^{-s})

where M(HsHs)M(H^{s}\to H^{-s}) is the space of pointwise multipliers from Hs(n)H^{s}(\mathbb{R}^{n}) to Hs(n)H^{-s}(\mathbb{R}^{n}) equipped with the norm

||f||s,s=sup{|f,ϕψ|:ϕ,ψCc(Ω),||ϕ||Hs(n)=||ψ||Hs(n)=1}.||f||_{s,-s}=\sup\{|\langle f,\phi\psi\rangle|:\,\phi,\psi\in C^{\infty}_{c}(\Omega),\,\,||\phi||_{H^{s}(\mathbb{R}^{n})}=||\psi||_{H^{s}(\mathbb{R}^{n})}=1\}.

See for instance, Section 2 in [6, 23].

2.2 Fractional Laplacian

Let 0<s<10<s<1. The fractional Laplacian (Δ)s(-\Delta)^{s} is formally given by the pointwise definition

(Δ)su(x):=cn,slimϵ0+nBϵ(x)u(x)u(y)|xy|n+2s𝑑y(-\Delta)^{s}u(x):=c_{n,s}\lim_{\epsilon\to 0^{+}}\int_{\mathbb{R}^{n}\setminus B_{\epsilon}(x)}\frac{u(x)-u(y)}{|x-y|^{n+2s}}\,dy

as well as the the equivalent Fourier transform definition

(Δ)su(x):=1(|ξ|2su(ξ))(x).(-\Delta)^{s}u(x):=\mathcal{F}^{-1}(|\xi|^{2s}\mathcal{F}u(\xi))(x).

It is well-known that one of the equivalent forms of the HsH^{s}-norm is given by

fHs2:=fL22+nn|f(x)f(y)|2|xy|n+2s,fHs(n)||f||^{2}_{H^{s}}:=||f||^{2}_{L^{2}}+\int_{\mathbb{R}^{n}}\int_{\mathbb{R}^{n}}\frac{|f(x)-f(y)|^{2}}{|x-y|^{n+2s}},\quad f\in H^{s}(\mathbb{R}^{n})

and we have the following bilinear form formula

(Δ)su,v=cn,snn(u(x)u(y))(v(x)v(y))|xy|n+2s𝑑x𝑑y,u,vHs(n).\langle(-\Delta)^{s}u,v\rangle=c^{\prime}_{n,s}\int_{\mathbb{R}^{n}}\int_{\mathbb{R}^{n}}\frac{(u(x)-u(y))(v(x)-v(y))}{|x-y|^{n+2s}}\,dxdy,\quad u,v\in H^{s}(\mathbb{R}^{n}).

It is also well-known that one of the equivalent forms of the H1+sH^{1+s}-norm is given by

fH1+s2:=fH12+jnn|jf(x)jf(y)|2|xy|n+2s,fH1+s(n).||f||^{2}_{H^{1+s}}:=||f||^{2}_{H^{1}}+\sum_{j}\int_{\mathbb{R}^{n}}\int_{\mathbb{R}^{n}}\frac{|\partial_{j}f(x)-\partial_{j}f(y)|^{2}}{|x-y|^{n+2s}},\quad f\in H^{1+s}(\mathbb{R}^{n}).

By the classical and fractional Poincaré inequalities, we have the following norm equivalence

||f||H1+s2j(||jf||L22+(Δ)sjf,jf)j(Δ)sjf,jf),fH~1+s(Ω).||f||^{2}_{H^{1+s}}\sim\sum_{j}(||\partial_{j}f||^{2}_{L^{2}}+\langle(-\Delta)^{s}\partial_{j}f,\partial_{j}f\rangle)\sim\sum_{j}\langle(-\Delta)^{s}\partial_{j}f,\partial_{j}f\rangle),\quad f\in\tilde{H}^{1+s}(\Omega).

The following unique continuation property of (Δ)s(-\Delta)^{s} was first proved in [13].

Proposition 2.1.

Suppose uHr(n)u\in H^{r}(\mathbb{R}^{n}) for some rr\in\mathbb{R}. Let WnW\subset\mathbb{R}^{n} be open and non-empty. If

(Δ)su=u=0inW,(-\Delta)^{s}u=u=0\quad\mathrm{in}\,\,W,

then u=0u=0 in n\mathbb{R}^{n}.

3 Linear fractional elasticity

3.1 Well-posedness

We first study the equation

u=finΩ.\mathcal{L}u=f\,\,\,\,\text{in}\,\,\Omega. (9)

The bilinear form associated with -\mathcal{L} (see (1) in Section 1 for its definition) is

B[u,v]=j(Δ)sλ0uj,j,jvj,j+Ωλ(juj,j)(jvj,j)B[u,v]=\langle\sum_{j}(-\Delta)^{s}\lambda_{0}u_{j,j},\sum_{j}v_{j,j}\rangle+\int_{\Omega}\lambda(\sum_{j}u_{j,j})(\sum_{j}v_{j,j})
+12i,j((Δ)sμ0(ui,j+uj,i),vi,j+vj,i+Ωμ(ui,j+uj,i)(vi,j+vj,i)).+\frac{1}{2}\sum_{i,j}(\langle(-\Delta)^{s}\mu_{0}(u_{i,j}+u_{j,i}),v_{i,j}+v_{j,i}\rangle+\int_{\Omega}\mu(u_{i,j}+u_{j,i})(v_{i,j}+v_{j,i})). (10)

Assume λ,μL(Ω)\lambda,\mu\in L^{\infty}(\Omega). Then it is clear that BB is bounded over H1+s(3)×H1+s(3)H^{1+s}(\mathbb{R}^{3})\times H^{1+s}(\mathbb{R}^{3}) since the bilinear form (Δ)sϕ,ψ\langle(-\Delta)^{s}\phi,\psi\rangle is bounded over Hs(3)×Hs(3)H^{s}(\mathbb{R}^{3})\times H^{s}(\mathbb{R}^{3}). Now we show that BB is coercive over H~1+s(Ω)×H~1+s(Ω)\tilde{H}^{1+s}(\Omega)\times\tilde{H}^{1+s}(\Omega) if we further assume λ,μ0\lambda,\mu\geq 0, μ0>0\mu_{0}>0 and λ0+μ00\lambda_{0}+\mu_{0}\geq 0.

In fact, we note that

(Δ)sui,j,uj,i=(Δ)sui,i,uj,j\langle(-\Delta)^{s}u_{i,j},u_{j,i}\rangle=\langle(-\Delta)^{s}u_{i,i},u_{j,j}\rangle

for uH~1+s(Ω)u\in\tilde{H}^{1+s}(\Omega) so we have

j(Δ)sλ0uj,j,juj,j+12i,j(Δ)sμ0(ui,j+uj,i),ui,j+uj,i\langle\sum_{j}(-\Delta)^{s}\lambda_{0}u_{j,j},\sum_{j}u_{j,j}\rangle+\frac{1}{2}\sum_{i,j}\langle(-\Delta)^{s}\mu_{0}(u_{i,j}+u_{j,i}),u_{i,j}+u_{j,i}\rangle
=j(Δ)sλ0uj,j,juj,j+i,j(Δ)sμ0ui,j,ui,j+i,j(Δ)sμ0ui,j,uj,i=\langle\sum_{j}(-\Delta)^{s}\lambda_{0}u_{j,j},\sum_{j}u_{j,j}\rangle+\sum_{i,j}\langle(-\Delta)^{s}\mu_{0}u_{i,j},u_{i,j}\rangle+\sum_{i,j}\langle(-\Delta)^{s}\mu_{0}u_{i,j},u_{j,i}\rangle
=j(Δ)sλ0uj,j,juj,j+i,j(Δ)sμ0ui,j,ui,j+i,j(Δ)sμ0ui,i,uj,j=\langle\sum_{j}(-\Delta)^{s}\lambda_{0}u_{j,j},\sum_{j}u_{j,j}\rangle+\sum_{i,j}\langle(-\Delta)^{s}\mu_{0}u_{i,j},u_{i,j}\rangle+\sum_{i,j}\langle(-\Delta)^{s}\mu_{0}u_{i,i},u_{j,j}\rangle
=j(Δ)s(λ0+μ0)uj,j,juj,j+i,j(Δ)sμ0ui,j,ui,j=\langle\sum_{j}(-\Delta)^{s}(\lambda_{0}+\mu_{0})u_{j,j},\sum_{j}u_{j,j}\rangle+\sum_{i,j}\langle(-\Delta)^{s}\mu_{0}u_{i,j},u_{i,j}\rangle
i,j(Δ)sμ0ui,j,ui,j.\geq\sum_{i,j}\langle(-\Delta)^{s}\mu_{0}u_{i,j},u_{i,j}\rangle.

Then the coerciveness immediately follows from the H1+sH^{1+s}-norm equivalence (see Subsection 2.2).

Now the Lax-Milgram theorem implies that the solution operator fuff\to u_{f} associated with (9) is well-defined, which is a homeomorphism from H1s(Ω)H^{-1-s}(\Omega) to H~s+1(Ω)\tilde{H}^{s+1}(\Omega).

From now on we will always assume 0λ,μL(Ω)0\leq\lambda,\mu\in L^{\infty}(\Omega), μ0>0\mu_{0}>0 and λ0+μ00\lambda_{0}+\mu_{0}\geq 0.

Let f:=(Δ)sLλ0,μ0g|Ωf:=-(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}g|_{\Omega} in (9). Then we have the well-posedness of the exterior problem (2).

Proposition 3.1.

For each gH1+s(3)g\in H^{1+s}(\mathbb{R}^{3}), there exists a unique solution ugH1+s(3)u_{g}\in H^{1+s}(\mathbb{R}^{3}) of (2) s.t. uggH~1+s(Ω)u_{g}-g\in\tilde{H}^{1+s}(\Omega). Moreover, the solution operator gugg\to u_{g} is bounded on H1+s(3)H^{1+s}(\mathbb{R}^{3}).

3.2 Dirichlet-to-Neumann map and integral identity

Let X:=H1+s(3)/H~1+s(Ω)=H1+s(Ωe)X:=H^{1+s}(\mathbb{R}^{3})/\tilde{H}^{1+s}(\Omega)=H^{1+s}(\Omega_{e}) and g~:=\tilde{g}:= the natural image of gH1+s(3)g\in H^{1+s}(\mathbb{R}^{3}) in XX.

We define the Dirichlet-to-Neumann map Λ\Lambda by

Λg~,h~:=B[ug,h],g,hH1+s(3)\langle\Lambda\tilde{g},\tilde{h}\rangle:=-B[u_{g},h],\quad g,h\in H^{1+s}(\mathbb{R}^{3})

where ugu_{g} is the solution corresponding to the exterior data gg in (2).

It is easy to verify that Λ:XX\Lambda:X\to X^{*} is well-defined and this bilinear form definition coincides with the one given by (3) for gCc(Ωe)g\in C^{\infty}_{c}(\Omega_{e}).

For convenience, we will write Λg\Lambda g and Λg,h\langle\Lambda g,h\rangle instead of Λg~\Lambda\tilde{g} and Λg~,h~\langle\Lambda\tilde{g},\tilde{h}\rangle. Note that

Λg,h=B[ug,uh]=B[uh,ug]=Λh,g\langle\Lambda g,h\rangle=-B[u_{g},u_{h}]=-B[u_{h},u_{g}]=\langle\Lambda h,g\rangle

so we have the integral identity

Λ(1)g(1),g(2)Λ(2)g(1),g(2)=B(1)[u(1),u(2)]+B(2)[u(2),u(1)]\langle\Lambda^{(1)}g^{(1)},g^{(2)}\rangle-\langle\Lambda^{(2)}g^{(1)},g^{(2)}\rangle=-B^{(1)}[u^{(1)},u^{(2)}]+B^{(2)}[u^{(2)},u^{(1)}]
=Ω(λ(2)λ(1))(juj,j(1))(juj,j(2))+12Ω(μ(2)μ(1))k,j(uk,j(1)+uj,k(1))(uk,j(2)+uj,k(2))=\int_{\Omega}(\lambda^{(2)}-\lambda^{(1)})(\sum_{j}u^{(1)}_{j,j})(\sum_{j}u^{(2)}_{j,j})+\frac{1}{2}\int_{\Omega}(\mu^{(2)}-\mu^{(1)})\sum_{k,j}(u^{(1)}_{k,j}+u^{(1)}_{j,k})(u^{(2)}_{k,j}+u^{(2)}_{j,k}) (11)

where (j)\mathcal{L}^{(j)},B(j)B^{(j)},λ(j)\lambda^{(j)} correspond to Lamé parameters λ(j),μ(j)\lambda^{(j)},\mu^{(j)}; u(j)u^{(j)} denotes the solution of

(j)u=0inΩ,u=g(j)inΩe.\mathcal{L}^{(j)}u=0\,\,\,\,\text{in}\,\,\Omega,\qquad u=g^{(j)}\,\,\,\,\text{in}\,\,\Omega_{e}.

We remark that this integral identity has the same form as its classical counterpart (see [9]).

3.3 Unique continuation property and Runge approximation property

Recall that a classical operator LL possesses the unique continuation property in a domain UU if

u=0inU,u=0inV\mathcal{L}u=0\,\,\,\,\text{in}\,\,U,\quad u=0\,\,\,\,\text{in}\,\,V

where VV is a nonempty open subset of UU imply that u=0u=0 in UU.

It is well-known that the classical constant coefficients Lamé operator Lλ0,μ0L_{\lambda_{0},\mu_{0}} possesses the unique continuation property in this sense. (This property even holds true for the general variable coefficients Lamé operator Lλ,μL_{\lambda,\mu}. See for instance, the main theorem in [8].)

The following proposition is the unique continuation property of (Δ)sLλ0,μ0(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}.

Proposition 3.2.

Let uH1+s(3)u\in H^{1+s}(\mathbb{R}^{3}). Let WW be open. If

(Δ)sLλ0,μ0u=u=0inW,(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}u=u=0\quad\text{in}\,\,W,

then u=0u=0 in 3\mathbb{R}^{3}.

Proof.

By the unique continuation property of (Δ)s(-\Delta)^{s} (Proposition 2.1), we have Lλ0,μ0ϕ=0L_{\lambda_{0},\mu_{0}}\phi=0 in 3\mathbb{R}^{3}. Then by the unique continuation property of Lλ0,μ0L_{\lambda_{0},\mu_{0}} we have u=0u=0 in 3\mathbb{R}^{3}. ∎

Based on the unique continuation property above, we can prove the following Runge approximation property.

Proposition 3.3.

Let WΩeW\subset\Omega_{e} be nonempty and open. Then

S:={ug|Ω:gCc(W)}S:=\{u_{g}|_{\Omega}:g\in C^{\infty}_{c}(W)\}

is dense in H~1+s(Ω)\tilde{H}^{1+s}(\Omega) where ugu_{g} is the solution corresponding to the exterior data gg in (2).

Proof.

By the Hahn-Banach Theorem, it suffices to show that:

If fH1s(Ω)f\in H^{-1-s}(\Omega) and f,w=0\langle f,w\rangle=0 for all wSw\in S, then f=0f=0.

In fact, we can choose vH~1+s(Ω)v\in\tilde{H}^{1+s}(\Omega) to be the solution of v=f\mathcal{L}v=f in Ω\Omega. Then for any gCc(W)g\in C^{\infty}_{c}(W), by the assumption f,w=0\langle f,w\rangle=0 for wSw\in S we have

0=f,ugg=B[v,ugg]=B[v,g]0=\langle f,u_{g}-g\rangle=-B[v,u_{g}-g]=B[v,g]

since B[ug,v]=0B[u_{g},v]=0. This implies (Δ)sLλ0,μ0v=0(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}v=0 in WW. By the unique continuation property above we have v=0v=0 in 3\mathbb{R}^{3}. Hence f=0f=0. ∎

3.4 Proof of Theorem 1.1

Now we are ready to prove Theorem 1.1. The key point is to approximate certain carefully chosen functions by solutions of the linear exterior problems based on the Runge approximation property. This will enable us to exploit the integral identity (11) to determine λ,μ\lambda,\mu.

Proof.

Let u(j)u^{(j)} denote the solution of

(j)u=0inΩ,u=g(j)inΩe.\mathcal{L}^{(j)}u=0\,\,\,\,\text{in}\,\,\Omega,\qquad u=g^{(j)}\,\,\,\,\text{in}\,\,\Omega_{e}.

For any given ϵ>0\epsilon>0 and f(j)Cc(Ω)f^{(j)}\in C^{\infty}_{c}(\Omega), by the Runge approximation property (Proposition 3.3), we can choose g(1)Cc(W1)g^{(1)}\in C^{\infty}_{c}(W_{1}) s.t.

u(1)f(1)H~1+s(Ω)ϵ||u^{(1)}-f^{(1)}||_{\tilde{H}^{1+s}(\Omega)}\leq\epsilon

and for this chosen g(1)g^{(1)}, we can choose g(2)Cc(W2)g^{(2)}\in C^{\infty}_{c}(W_{2}) s.t.

u(1)H~1+s(Ω)u(2)f(2)H~1+s(Ω)ϵ.||u^{(1)}||_{\tilde{H}^{1+s}(\Omega)}||u^{(2)}-f^{(2)}||_{\tilde{H}^{1+s}(\Omega)}\leq\epsilon.

(Actually we only need the H1H^{1}-norm approximation.) By the assumption

Λ(1)g|W2=Λ(2)g|W2\Lambda^{(1)}g|_{W_{2}}=\Lambda^{(2)}g|_{W_{2}}

for gCc(W1)g\in C^{\infty}_{c}(W_{1}) and the integral identity (11) we get

Ω(λ(2)λ(1))(juj,j(1))(juj,j(2))+12Ω(μ(2)μ(1))k,j(uk,j(1)+uj,k(1))(uk,j(2)+uj,k(2))=0.\int_{\Omega}(\lambda^{(2)}-\lambda^{(1)})(\sum_{j}u^{(1)}_{j,j})(\sum_{j}u^{(2)}_{j,j})+\frac{1}{2}\int_{\Omega}(\mu^{(2)}-\mu^{(1)})\sum_{k,j}(u^{(1)}_{k,j}+u^{(1)}_{j,k})(u^{(2)}_{k,j}+u^{(2)}_{j,k})=0.

Based on our choice for g(j)g^{(j)}, we get

|Ω(λ(2)λ(1))(jfj,j(1))(jfj,j(2))+12Ω(μ(2)μ(1))k,j(fk,j(1)+fj,k(1))(fk,j(2)+fj,k(2))|Cϵ|\int_{\Omega}(\lambda^{(2)}-\lambda^{(1)})(\sum_{j}f^{(1)}_{j,j})(\sum_{j}f^{(2)}_{j,j})+\frac{1}{2}\int_{\Omega}(\mu^{(2)}-\mu^{(1)})\sum_{k,j}(f^{(1)}_{k,j}+f^{(1)}_{j,k})(f^{(2)}_{k,j}+f^{(2)}_{j,k})|\leq C\epsilon

where CC is a constant depending on λ(j),μ(j)\lambda^{(j)},\mu^{(j)} and f(j)f^{(j)}. Hence we conclude that

Ω(λ(2)λ(1))(jfj,j(1))(jfj,j(2))+12Ω(μ(2)μ(1))k,j(fk,j(1)+fj,k(1))(fk,j(2)+fj,k(2))=0\int_{\Omega}(\lambda^{(2)}-\lambda^{(1)})(\sum_{j}f^{(1)}_{j,j})(\sum_{j}f^{(2)}_{j,j})+\frac{1}{2}\int_{\Omega}(\mu^{(2)}-\mu^{(1)})\sum_{k,j}(f^{(1)}_{k,j}+f^{(1)}_{j,k})(f^{(2)}_{k,j}+f^{(2)}_{j,k})=0 (12)

since ϵ\epsilon is arbitrary.

We will appropriately choose f(j)f^{(j)} in (12) to determine λ,μ\lambda,\mu.

In fact, for any given ψCc(Ω)\psi\in C^{\infty}_{c}(\Omega), we can choose ϕ\phi s.t. ϕCc(Ω)\phi\in C^{\infty}_{c}(\Omega) and ϕ=x1\phi=x_{1} on suppψ\mathrm{supp}\,\psi.

We can show that

Ω(μ(2)μ(1))jψ=0,(1j3).\int_{\Omega}(\mu^{(2)}-\mu^{(1)})\partial_{j}\psi=0,\quad(1\leq j\leq 3).

For instance, to obtain the equality above for j=2j=2, we can choose

f(1)=(0,ϕ,0),f(2)=(ψ,0,0)f^{(1)}=(0,\phi,0),\qquad f^{(2)}=(\psi,0,0)

in (12). We can also show that

Ω(λ(2)λ(1))jψ=0,(1j3).\int_{\Omega}(\lambda^{(2)}-\lambda^{(1)})\partial_{j}\psi=0,\quad(1\leq j\leq 3).

For instance, to obtain the equality above for j=2j=2, we can choose

f(1)=(ϕ,0,0),f(2)=(0,ψ,0)f^{(1)}=(\phi,0,0),\qquad f^{(2)}=(0,\psi,0)

in (12). Hence we get

(μ(2)μ(1))=(λ(2)λ(1))=0.\nabla(\mu^{(2)}-\mu^{(1)})=\nabla(\lambda^{(2)}-\lambda^{(1)})=0.

Now we show that the constants cμ:=μ(2)μ(1)c_{\mu}:=\mu^{(2)}-\mu^{(1)} and cλ:=λ(2)λ(1)c_{\lambda}:=\lambda^{(2)}-\lambda^{(1)} are zeros.

In fact, we can choose ψCc(Ω)\psi\in C^{\infty}_{c}(\Omega) s.t. 1ψL22ψL2||\partial_{1}\psi||_{L^{2}}\neq||\partial_{2}\psi||_{L^{2}}. Then we can show that

(cλ+2cμ)1ψL22+cμ2ψL22+cμ3ψL22=0,(c_{\lambda}+2c_{\mu})||\partial_{1}\psi||^{2}_{L^{2}}+c_{\mu}||\partial_{2}\psi||^{2}_{L^{2}}+c_{\mu}||\partial_{3}\psi||^{2}_{L^{2}}=0,
(cλ+2cμ)2ψL22+cμ3ψL22+cμ1ψL22=0,(c_{\lambda}+2c_{\mu})||\partial_{2}\psi||^{2}_{L^{2}}+c_{\mu}||\partial_{3}\psi||^{2}_{L^{2}}+c_{\mu}||\partial_{1}\psi||^{2}_{L^{2}}=0,
(cλ+2cμ)3ψL22+cμ1ψL22+cμ2ψL22=0.(c_{\lambda}+2c_{\mu})||\partial_{3}\psi||^{2}_{L^{2}}+c_{\mu}||\partial_{1}\psi||^{2}_{L^{2}}+c_{\mu}||\partial_{2}\psi||^{2}_{L^{2}}=0.

For instance, to obtain the first identity, we can choose

f(1)=f(2)=(ψ,0,0)f^{(1)}=f^{(2)}=(\psi,0,0)

in (12). Now we combine the three identities to obtain cλ+4cμ=0c_{\lambda}+4c_{\mu}=0. Then we combine the first two identities to obtain cμ1ψL22=cμ2ψL22c_{\mu}||\partial_{1}\psi||^{2}_{L^{2}}=c_{\mu}||\partial_{2}\psi||^{2}_{L^{2}}, which implies cμ=0c_{\mu}=0 and thus cλ=0c_{\lambda}=0. ∎

Remark.

Similar strategy has been applied to solve inverse problems for fractional Schrödinger operators with local perturbations. See [4, 6] for details.

3.5 Proof of Theorem 1.2

We can identically apply the considerations for the exterior problem (2) in previous subsections to the obstacle problem (4). (We just replace Ω\Omega by ΩD¯\Omega\setminus\bar{D} in earlier arguments.) Our main task in this subsection is to prove the first part of Theorem 1.2. We will see that this part is an immediate consequence of the unique continuation property (Proposition 3.2). Once we have determined the obstacle DD, we can use the same arguments as in the proof of Theorem 1.1 to determine λ,μ\lambda,\mu.

Proof.

For the fixed nonzero gCc(W1)g\in C^{\infty}_{c}(W_{1}), let u(j)u^{(j)} denote the solution of the obstacle problem

(j)u=0inΩDj¯,u=0inDj,u=ginΩe.\mathcal{L}^{(j)}u=0\,\,\text{in}\,\,\Omega\setminus\bar{D_{j}},\quad u=0\,\,\text{in}\,\,D_{j},\quad u=g\,\,\text{in}\,\,\Omega_{e}.

Since we have the assumption

ΛD1(1)g|W2=ΛD2(2)g|W2\Lambda^{(1)}_{D_{1}}g|_{W_{2}}=\Lambda^{(2)}_{D_{2}}g|_{W_{2}}

and u(1)=u(2)=gu^{(1)}=u^{(2)}=g in Ωe\Omega_{e}, we get

(Δ)sLλ0,μ0(u(1)u(2))=u(1)u(2)=0inW2.(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}(u^{(1)}-u^{(2)})=u^{(1)}-u^{(2)}=0\quad\text{in}\,\,W_{2}.

Then Proposition 3.2 implies that u(1)=u(2)u^{(1)}=u^{(2)} in 3\mathbb{R}^{3}.

Suppose D1D2D_{1}\neq D_{2}. Without loss of generality we can assume V:=D2D¯1V:=D_{2}\setminus\bar{D}_{1} is nonempty. Note that u(1)=u(2)=0u^{(1)}=u^{(2)}=0 in VV. Then the equation for u(1)u^{(1)} in VV implies (Δ)sLλ0,μ0u(1)=0(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}u^{(1)}=0 in VV. But now Proposition 3.2 implies that u(1)=0u^{(1)}=0 in 3\mathbb{R}^{3}, which contradicts that gg is nonzero. ∎

Remark.

Similar inverse obstacle problems have been studied for fractional elliptic operators. See [3] for details.

4 Nonlinear fractional elasticity

4.1 Well-posedness and Dirichlet-to-Neumann map

We first study the nonlinear equation

u+𝒩u=finΩ\mathcal{L}u+\mathcal{N}u=f\,\,\text{in}\,\,\Omega (13)

See (7) in Section 1 for the definition of 𝒩\mathcal{N}.

From now on we will always assume 𝒜,,𝒞C1(Ω¯)\mathcal{A},\mathcal{B},\mathcal{C}\in C^{1}(\bar{\Omega}) in (7). Let 12s<1\frac{1}{2}\leq s<1. Then 2nn2sn2s\frac{2n}{n-2s}\geq\frac{n}{2s} for n=3n=3 so we have the continuous embeddings (see Subsection 2.1)

H~s(Ω)L2nn2s(Ω)Ln2s(Ω)M(HsHs).\tilde{H}^{s}(\Omega)\hookrightarrow L^{\frac{2n}{n-2s}}(\Omega)\hookrightarrow L^{\frac{n}{2s}}(\Omega)\hookrightarrow M(H^{s}\to H^{-s}).

Note that for uH~s+1(Ω)u\in\tilde{H}^{s+1}(\Omega), each component of NuNu is a sum of terms which have the form j(aϕψ)\partial_{j}(a\phi\psi) where aC1(Ω¯)a\in C^{1}(\bar{\Omega}) and ϕ,ψH~s(Ω)\phi,\psi\in\tilde{H}^{s}(\Omega) so NuH1s(3)Nu\in H^{-1-s}(\mathbb{R}^{3}). Hence the map FF defined by

F(f,u):=(u+𝒩u)|ΩfF(f,u):=(\mathcal{L}u+\mathcal{N}u)|_{\Omega}-f

maps from H1s(Ω)×H~s+1(Ω)H^{-1-s}(\Omega)\times\tilde{H}^{s+1}(\Omega) to H1s(Ω)H^{-1-s}(\Omega).

Note that F(0,0)=0F(0,0)=0 and it is easy to verify that the Fréchet derivative

DuF|(0,0)(v)=v|Ω,D_{u}F|_{(0,0)}(v)=\mathcal{L}v|_{\Omega},

which is a homeomorphism from H~s+1(Ω)\tilde{H}^{s+1}(\Omega) to H1s(Ω)H^{-1-s}(\Omega). By Implicit function theorem (see for instance, Theorem 10.6 in [22]), there exists δ>0\delta>0 s.t. whenever fH1s(Ω)δ||f||_{H^{-1-s}(\Omega)}\leq\delta, we have both existence and uniqueness of small solutions of (13), and ufu_{f} smoothly depends on ff.

Let f:=(Δ)sLλ0,μ0g|Ωf:=-(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}g|_{\Omega} in (13). Then we have the well-posedness of (6).

Proposition 4.1.

For each sufficiently small gCc(Ωe)g\in C^{\infty}_{c}(\Omega_{e}), there exists a unique small solution ugu_{g} of the exterior problem (6) s.t. uggH~1+s(Ω)u_{g}-g\in\tilde{H}^{1+s}(\Omega) and ugu_{g} smoothly depends on gg.

Now we can conclude that the associated Dirichlet-to-Neumann map ΛN\Lambda_{N} given by (8) is well-defined at least for small gCc(Ωe)g\in C^{\infty}_{c}(\Omega_{e}).

4.2 Proof of Theorem 1.3

We are ready to prove Theorem 1.3. We will first apply the first order linearization and the linear result (Theorem 1.1) to determine λ,μ\lambda,\mu. Then we will apply the second order linearization and the Runge approximation property (Proposition 3.3) to determine 𝒜,𝒞\mathcal{A},\mathcal{C}.

Proof.

Determine λ,μ\lambda,\mu: Let uϵ,g(j)u^{(j)}_{\epsilon,g} be the solution of the exterior problem

(j)u+𝒩(j)u=0inΩ,u=ϵginΩe\mathcal{L}^{(j)}u+\mathcal{N}^{(j)}u=0\,\,\,\,\text{in}\,\,\Omega,\qquad u=\epsilon g\,\,\,\,\text{in}\,\,\Omega_{e} (14)

for gCc(W1)g\in C^{\infty}_{c}(W_{1}) and small ϵ\epsilon. Applying ϵ|ϵ=0\frac{\partial}{\partial\epsilon}|_{\epsilon=0} to (14), we obtain that

ug(j):=ϵ|ϵ=0uϵ,g(j)u^{(j)}_{g}:=\frac{\partial}{\partial\epsilon}|_{\epsilon=0}u^{(j)}_{\epsilon,g}

is the solution of

(j)u=0inΩ,u=ginΩe.\mathcal{L}^{(j)}u=0\,\,\,\,\text{in}\,\,\Omega,\qquad u=g\,\,\,\,\text{in}\,\,\Omega_{e}.

Since we have the assumption

ΛN(1)g=ΛN(2)ginW2,\Lambda^{(1)}_{N}g=\Lambda^{(2)}_{N}g\qquad\text{in}\,\,W_{2},
i.e.(Δ)sLλ0,μ0uϵ,g(1)=(Δ)sLλ0,μ0uϵ,g(2)inW2,i.e.\quad(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}u^{(1)}_{\epsilon,g}=(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}u^{(2)}_{\epsilon,g}\qquad\text{in}\,\,W_{2},

we can apply ϵ|ϵ=0\frac{\partial}{\partial\epsilon}|_{\epsilon=0} to the identity to obtain that

(Δ)sLλ0,μ0ug(1)=(Δ)sLλ0,μ0ug(2)inW2,(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}u^{(1)}_{g}=(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}u^{(2)}_{g}\qquad\text{in}\,\,W_{2},
i.e.Λ(1)g=Λ(2)ginW2.i.e.\quad\Lambda^{(1)}g=\Lambda^{(2)}g\qquad\text{in}\,\,W_{2}.

Hence we conclude that λ(1)=λ(2):=λ\lambda^{(1)}=\lambda^{(2)}:=\lambda, μ(1)=μ(2):=μ\mu^{(1)}=\mu^{(2)}:=\mu based on Theorem 1.1.

Now we use \mathcal{L} to denote both (1)\mathcal{L}^{(1)} and (2)\mathcal{L}^{(2)}.

Determine 𝒜,𝒞\mathcal{A},\mathcal{C}: Let uϵ,g(j)u^{(j)}_{\epsilon,g} be the solution of the exterior problem

u+𝒩(j)u=0inΩ,u=ϵ1g(1)+ϵ2g(2)inΩe\mathcal{L}u+\mathcal{N}^{(j)}u=0\,\,\,\,\text{in}\,\,\Omega,\quad u=\epsilon_{1}g^{(1)}+\epsilon_{2}g^{(2)}\,\,\,\,\text{in}\,\,\Omega_{e} (15)

for g(j)Cc(W1)g^{(j)}\in C^{\infty}_{c}(W_{1}) and small ϵj\epsilon_{j}. First note that

ϵj|ϵj=0uϵ,g(1)=ϵj|ϵj=0uϵ,g(2)\frac{\partial}{\partial\epsilon_{j}}|_{\epsilon_{j}=0}u^{(1)}_{\epsilon,g}=\frac{\partial}{\partial\epsilon_{j}}|_{\epsilon_{j}=0}u^{(2)}_{\epsilon,g}

since both of them are the solution of

u=0inΩ,u=g(j)inΩe.\mathcal{L}u=0\,\,\,\ \text{in}\,\,\Omega,\qquad u=g^{(j)}\,\,\,\,\text{in}\,\,\Omega_{e}.

Hence we can denote both of them by v(j)v^{(j)}. Next we apply 2ϵ1ϵ2|ϵ1=ϵ2=0\frac{\partial^{2}}{\partial\epsilon_{1}\partial\epsilon_{2}}|_{\epsilon_{1}=\epsilon_{2}=0} to (15). Then

w(j):=2ϵ1ϵ2|ϵ1=ϵ2=0uϵ,g(j)w^{(j)}:=\frac{\partial^{2}}{\partial\epsilon_{1}\partial\epsilon_{2}}|_{\epsilon_{1}=\epsilon_{2}=0}u^{(j)}_{\epsilon,g}

satisfies

w(j)+N~(j)(v(1),v(2))=0inΩ,w(j)=0inΩe.\mathcal{L}w^{(j)}+\tilde{N}^{(j)}(v^{(1)},v^{(2)})=0\,\,\,\,\text{in}\,\,\Omega,\qquad w^{(j)}=0\,\,\,\,\text{in}\,\,\Omega_{e}. (16)

Here (based on (7)) we can compute that

N~i(l)=j=13jN~ij(l),(1l2; 1i3)\tilde{N}^{(l)}_{i}=\sum^{3}_{j=1}\partial_{j}\tilde{N}^{(l)}_{ij},\quad(1\leq l\leq 2;\,1\leq i\leq 3)
N~ij(l)(v(1),v(2))=(λ+)m,nvm,n(1)vm,n(2)δij\tilde{N}^{(l)}_{ij}(v^{(1)},v^{(2)})=(\lambda+\mathcal{B})\sum_{m,n}v^{(1)}_{m,n}v^{(2)}_{m,n}\delta_{ij}
+2𝒞(l)(mvm,m(1))(mvm,m(2))δij+m,nvm,n(1)vm,n(2)δij+m(vm,m(1)vj,i(2)+vm,m(2)vj,i(1))+2\mathcal{C}^{(l)}(\sum_{m}v^{(1)}_{m,m})(\sum_{m}v^{(2)}_{m,m})\delta_{ij}+\mathcal{B}\sum_{m,n}v^{(1)}_{m,n}v^{(2)}_{m,n}\delta_{ij}+\mathcal{B}\sum_{m}(v^{(1)}_{m,m}v^{(2)}_{j,i}+v^{(2)}_{m,m}v^{(1)}_{j,i})
+𝒜(l)4m(vj,m(1)vm,i(2)+vj,m(2)vm,i(1))+(λ+)m(vm,m(1)vi,j(2)+vm,m(2)vi,j(1))+\frac{\mathcal{A}^{(l)}}{4}\sum_{m}(v^{(1)}_{j,m}v^{(2)}_{m,i}+v^{(2)}_{j,m}v^{(1)}_{m,i})+(\lambda+\mathcal{B})\sum_{m}(v^{(1)}_{m,m}v^{(2)}_{i,j}+v^{(2)}_{m,m}v^{(1)}_{i,j})
+(μ+𝒜(l)4)m(vm,i(1)vm,j(2)+vm,i(2)vm,j(1)+vi,m(1)vj,m(2)+vi,m(2)vj,m(1)+vi,m(1)vm,j(2)+vi,m(2)vm,j(1)).+(\mu+\frac{\mathcal{A}^{(l)}}{4})\sum_{m}(v^{(1)}_{m,i}v^{(2)}_{m,j}+v^{(2)}_{m,i}v^{(1)}_{m,j}+v^{(1)}_{i,m}v^{(2)}_{j,m}+v^{(2)}_{i,m}v^{(1)}_{j,m}+v^{(1)}_{i,m}v^{(2)}_{m,j}+v^{(2)}_{i,m}v^{(1)}_{m,j}).

We also apply 2ϵ1ϵ2|ϵ1=ϵ2=0\frac{\partial^{2}}{\partial\epsilon_{1}\partial\epsilon_{2}}|_{\epsilon_{1}=\epsilon_{2}=0} to the Dirichlet-to-Neumann map assumption

(Δ)sLλ0,μ0uϵ,g(1)=(Δ)sLλ0,μ0uϵ,g(2)inW2.(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}u^{(1)}_{\epsilon,g}=(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}u^{(2)}_{\epsilon,g}\qquad\text{in}\,\,W_{2}.

Then we get

(Δ)sLλ0,μ0w(1)=(Δ)sLλ0,μ0w(2)inW2.(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}w^{(1)}=(-\Delta)^{s}L_{\lambda_{0},\mu_{0}}w^{(2)}\qquad\text{in}\,\,W_{2}.

Since w(1)=w(2)=0w^{(1)}=w^{(2)}=0 in Ωe\Omega_{e}, the unique continuation property (Proposition 3.2) implies that w(1)=w(2)w^{(1)}=w^{(2)} in 3\mathbb{R}^{3}.

Now we combine the two equations (j=1,2j=1,2) in (16) to obtain

j=13jG~ij(v(1),v(2))=0,(1i3)\sum^{3}_{j=1}\partial_{j}\tilde{G}_{ij}(v^{(1)},v^{(2)})=0,\quad(1\leq i\leq 3) (17)

where

G~ij(v(1),v(2))=N~ij(2)(v(1),v(2))N~ij(1)(v(1),v(2))\tilde{G}_{ij}(v^{(1)},v^{(2)})=\tilde{N}^{(2)}_{ij}(v^{(1)},v^{(2)})-\tilde{N}^{(1)}_{ij}(v^{(1)},v^{(2)})
=𝒜(2)𝒜(1)4(m(vj,m(1)+vm,j(1))(vi,m(2)+vm,i(2))+m(vj,m(2)+vm,j(2))(vi,m(1)+vm,i(1)))=\frac{\mathcal{A}^{(2)}-\mathcal{A}^{(1)}}{4}(\sum_{m}(v^{(1)}_{j,m}+v^{(1)}_{m,j})(v^{(2)}_{i,m}+v^{(2)}_{m,i})+\sum_{m}(v^{(2)}_{j,m}+v^{(2)}_{m,j})(v^{(1)}_{i,m}+v^{(1)}_{m,i}))
+2(𝒞(2)𝒞(1))(mvm,m(1))(mvm,m(2))δij.+2(\mathcal{C}^{(2)}-\mathcal{C}^{(1)})(\sum_{m}v^{(1)}_{m,m})(\sum_{m}v^{(2)}_{m,m})\delta_{ij}.

Let both sides of (17) act on ψCc(Ω)\psi\in C^{\infty}_{c}(\Omega). Then Proposition 3.3 implies that

j=13G~ij(f(1),f(2))jψ=0,(1i3)\sum^{3}_{j=1}\int\tilde{G}_{ij}(f^{(1)},f^{(2)})\partial_{j}\psi=0,\quad(1\leq i\leq 3)

for any f(j)Cc(Ω)f^{(j)}\in C^{\infty}_{c}(\Omega). We can appropriately choose f(j)f^{(j)} to show that

(𝒜(2)𝒜(1))jψ=0,(1j3).\int(\mathcal{A}^{(2)}-\mathcal{A}^{(1)})\partial_{j}\psi=0,\quad(1\leq j\leq 3).

For instance, if we choose i=1i=1, f(1)=(0,ϕ,0)f^{(1)}=(0,\phi,0) and f(2)=(ϕ,0,0)f^{(2)}=(\phi,0,0) where ϕCc(Ω)\phi\in C^{\infty}_{c}(\Omega) satisfies ϕ=x1\phi=x_{1} on suppψ\mathrm{supp}\,\psi for a chosen ψCc(Ω)\psi\in C^{\infty}_{c}(\Omega), a direct computation shows that

G~11=0,G~12=𝒜(2)𝒜(1)2,G~13=0,\tilde{G}_{11}=0,\quad\tilde{G}_{12}=\frac{\mathcal{A}^{(2)}-\mathcal{A}^{(1)}}{2},\quad\tilde{G}_{13}=0,

which verifies the equality for j=2j=2. Hence we conclude that 𝒜(2)𝒜(1):=c𝒜\mathcal{A}^{(2)}-\mathcal{A}^{(1)}:=c_{\mathcal{A}} is a constant.

We can also appropriately choose f(j)f^{(j)} to show that

(𝒞(2)𝒞(1))jψ=0,(1j3).\int(\mathcal{C}^{(2)}-\mathcal{C}^{(1)})\partial_{j}\psi=0,\quad(1\leq j\leq 3).

For instance, if we choose i=1i=1, f(1)=(0,φ,0)f^{(1)}=(0,\varphi,0) and f(2)=(ϕ,0,0)f^{(2)}=(\phi,0,0) where ϕ\phi is defined as before and φCc(Ω)\varphi\in C^{\infty}_{c}(\Omega) satisfies φ=x2\varphi=x_{2} on suppψ\mathrm{supp}\,\psi for a chosen ψCc(Ω)\psi\in C^{\infty}_{c}(\Omega), a direct computation shows that

G~11=2(𝒞(2)𝒞(1)),G~12=0,G~13=0,\tilde{G}_{11}=2(\mathcal{C}^{(2)}-\mathcal{C}^{(1)}),\quad\tilde{G}_{12}=0,\quad\tilde{G}_{13}=0,

which verifies the equality for j=1j=1. Hence we conclude that 𝒞(2)𝒞(1):=c𝒞\mathcal{C}^{(2)}-\mathcal{C}^{(1)}:=c_{\mathcal{C}} is a constant.

Now we show that c𝒜c_{\mathcal{A}} and c𝒞c_{\mathcal{C}} are zeros.

In fact, we can choose ψCc(Ω)\psi\in C^{\infty}_{c}(\Omega) s.t. 1ψL22ψL2||\partial_{1}\psi||_{L^{2}}\neq||\partial_{2}\psi||_{L^{2}}.

Then we can appropriately choose f(j)f^{(j)} to obtain

(4c𝒜+4c𝒞)1ψL22+c𝒜2ψL22+c𝒜3ψL22=0,(4c_{\mathcal{A}}+4c_{\mathcal{C}})||\partial_{1}\psi||^{2}_{L^{2}}+c_{\mathcal{A}}||\partial_{2}\psi||^{2}_{L^{2}}+c_{\mathcal{A}}||\partial_{3}\psi||^{2}_{L^{2}}=0,
(4c𝒜+4c𝒞)2ψL22+c𝒜3ψL22+c𝒜1ψL22=0,(4c_{\mathcal{A}}+4c_{\mathcal{C}})||\partial_{2}\psi||^{2}_{L^{2}}+c_{\mathcal{A}}||\partial_{3}\psi||^{2}_{L^{2}}+c_{\mathcal{A}}||\partial_{1}\psi||^{2}_{L^{2}}=0,
(4c𝒜+4c𝒞)3ψL22+c𝒜1ψL22+c𝒜2ψL22=0.(4c_{\mathcal{A}}+4c_{\mathcal{C}})||\partial_{3}\psi||^{2}_{L^{2}}+c_{\mathcal{A}}||\partial_{1}\psi||^{2}_{L^{2}}+c_{\mathcal{A}}||\partial_{2}\psi||^{2}_{L^{2}}=0.

For instance, if we choose i=1i=1, f(1)=(ψ,0,0)f^{(1)}=(\psi,0,0) and f(2)=(ϕ,0,0)f^{(2)}=(\phi,0,0) where ϕ\phi is defined as before, a direct computation shows that

G~11=2c𝒜1ψ+2c𝒞1ψ,G~12=c𝒜22ψ,G~13=c𝒜23ψ,\tilde{G}_{11}=2c_{\mathcal{A}}\partial_{1}\psi+2c_{\mathcal{C}}\partial_{1}\psi,\quad\tilde{G}_{12}=\frac{c_{\mathcal{A}}}{2}\partial_{2}\psi,\quad\tilde{G}_{13}=\frac{c_{\mathcal{A}}}{2}\partial_{3}\psi,

which verifies the first identity. Now we combine the three identities to obtain 6c𝒜+4c𝒞=06c_{\mathcal{A}}+4c_{\mathcal{C}}=0. Then we combine the first two identities to obtain c𝒜1ψL22=c𝒜2ψL22c_{\mathcal{A}}||\partial_{1}\psi||^{2}_{L^{2}}=c_{\mathcal{A}}||\partial_{2}\psi||^{2}_{L^{2}}, which implies c𝒜=0c_{\mathcal{A}}=0 and thus c𝒞=0c_{\mathcal{C}}=0.

Remark.

The multiple-fold linearization procedure performed in the proof has been widely applied in solving inverse problems. For instance, see [10, 14] for this approach for inverse problems for semilinear elliptic operators. Also see [25] for this approach for an inverse problem for a nonlinear elastic wave operator.

References

  • [1] Sombuddha Bhattacharyya, Tuhin Ghosh, and Gunther Uhlmann. Inverse problems for the fractional-laplacian with lower order non-local perturbations. Transactions of the American Mathematical Society, 2021.
  • [2] Giovanni Molica Bisci, Vicentiu Radulescu, and Raffaella Servadei. Variational methods for nonlocal fractional problems, volume 162. Cambridge University Press, 2016.
  • [3] Xinlin Cao, Yi-Hsuan Lin, and Hongyu Liu. Simultaneously recovering potentials and embedded obstacles for anisotropic fractional Schrödinger operators. Inverse Problems and Imaging, 13:197–210, 2019.
  • [4] Mihajlo Cekić, Yi-Hsuan Lin, and Angkana Rüland. The Calderón problem for the fractional Schrödinger equation with drift. Calculus of Variations and Partial Differential Equations, 59(91), 2020.
  • [5] Giovanni Covi. An inverse problem for the fractional Schrödinger equation in a magnetic field. Inverse Problems, 36(4):045004, 2020.
  • [6] Giovanni Covi, Keijo Mönkkönen, Jesse Railo, and Gunther Uhlmann. The higher order fractional Calderón problem for linear local operators: uniqueness. arXiv preprint arXiv:2008.10227, 2020.
  • [7] Maarten De Hoop, Gunther Uhlmann, and Yiran Wang. Nonlinear interaction of waves in elastodynamics and an inverse problem. Mathematische Annalen, 376(1):765–795, 2020.
  • [8] Dang Ding Ang, Dang Duc Trong And, Masaru Ikehata, and Masahiro Yamamoto. Unique continuation for a stationary isotropic Lamé system with variable coefficients. Communications in partial differential equations, 23(1-2):599–617, 1998.
  • [9] Gregory Eskin and James Ralston. On the inverse boundary value problem for linear isotropic elasticity. Inverse Problems, 18(3):907–921, 2002.
  • [10] 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.
  • [11] Tuhin Ghosh, Yi-Hsuan Lin, and Jingni Xiao. The Calderón problem for variable coefficients nonlocal elliptic operators. Communications in Partial Differential Equations, 42(12):1923–1961, 2017.
  • [12] Tuhin Ghosh, Angkana Rüland, Mikko Salo, and Gunther Uhlmann. Uniqueness and reconstruction for the fractional Calderón problem with a single measurement. Journal of Functional Analysis, page 108505, 2020.
  • [13] Tuhin Ghosh, Mikko Salo, and Gunther Uhlmann. The Calderón problem for the fractional Schrödinger equation. Analysis & PDE, 13(2):455–475, 2020.
  • [14] Katya Krupchyk and Gunther Uhlmann. A remark on partial data inverse problems for semilinear elliptic equations. Proceedings of the American Mathematical Society, 148(2):681–685, 2020.
  • [15] Ru-Yu Lai, Yi-Hsuan Lin, and Angkana Rüland. The Calderón problem for a space-time fractional parabolic equation. SIAM Journal on Mathematical Analysis, 52(3):2655–2688, 2020.
  • [16] Ru-Yu Lai and Ting Zhou. An inverse problem for non-linear fractional magnetic schrodinger equation. arXiv preprint arXiv:2103.08180, 2021.
  • [17] Li Li. Determining the magnetic potential in the fractional magnetic Calderón problem. Communications in Partial Differential Equations, in press, 2020.
  • [18] Li Li. A fractional parabolic inverse problem involving a time-dependent magnetic potential. SIAM Journal on Mathematical Analysis, 53(1):435–452, 2021.
  • [19] Li Li. An inverse problem for a fractional diffusion equation with fractional power type nonlinearities. arXiv preprint arXiv:2104.00132, 2021.
  • [20] Gen Nakamura and Gunther Uhlmann. Global uniqueness for an inverse boundary problem arising in elasticity. Inventiones mathematicae, 118(1):457–474, 1994.
  • [21] Gen Nakamura and Gunther Uhlmann. Global uniqueness for an inverse boundary value problem arising in elasticity (Erratum). Inventiones mathematicae, 152(1):205–207, 2003.
  • [22] Michael Renardy and Robert Rogers. An introduction to partial differential equations, volume 13. Springer Science & Business Media, 2006.
  • [23] Angkana Rüland and Mikko Salo. The fractional Calderón problem: low regularity and stability. Nonlinear Analysis, 193:111529, 2020.
  • [24] Vasily Tarasov and Elias Aifantis. On fractional and fractal formulations of gradient linear and nonlinear elasticity. Acta Mechanica, 230(6):2043–2070, 2019.
  • [25] Gunther Uhlmann and Jian Zhai. On an inverse boundary value problem for a nonlinear elastic wave equation. Journal de Mathématiques Pures et Appliquées, 153:114–136, 2021.