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

The higher order fractional Calderón problem for linear local operators: uniqueness

Giovanni Covi Department of Mathematics and Statistics, University of Jyväskylä, Jyväskylä, Finland Institut fur Angewandte Mathematik, Ruprecht-Karls-Universität Heidelberg, Im Neuenheimer Feld 205, 69120 Heidelberg, Germany [email protected] Keijo Mönkkönen Department of Mathematics and Statistics, University of Jyväskylä, Jyväskylä, Finland [email protected] Jesse Railo Seminar for Applied Mathematics, Department of Mathematics, ETH Zurich, Zürich, Switzerland Department of Pure Mathematics and Mathematical Statistics, University of Cambridge, Cambridge CB3 0WB, UK [email protected]  and  Gunther Uhlmann Department of Mathematics, University of Washington, Seattle, USA / Jockey Club Institute for Advanced Study, HKUST, Hong Kong [email protected]
Abstract.

We study an inverse problem for the fractional Schrödinger equation (FSE) with a local perturbation by a linear partial differential operator (PDO) of order smaller than the one of the fractional Laplacian. We show that one can uniquely recover the coefficients of the PDO from the exterior Dirichlet-to-Neumann (DN) map associated to the perturbed FSE. This is proved for two classes of coefficients: coefficients which belong to certain spaces of Sobolev multipliers and coefficients which belong to fractional Sobolev spaces with bounded derivatives. Our study generalizes recent results for the zeroth and first order perturbations to higher order perturbations.

Key words and phrases:
Inverse problems, fractional Calderón problem, fractional Schrödinger equation, Sobolev multipliers.

1. Introduction

Let s+s\in{\mathbb{R}}^{+}\setminus{\mathbb{Z}}, Ωn\Omega\subset{\mathbb{R}}^{n} a bounded open set where n1n\geq 1, Ωe=nΩ¯\Omega_{e}={\mathbb{R}}^{n}\setminus\overline{\Omega} its exterior and P(x,D)P(x,D) a linear partial differential operator (PDO) of order mm\in{\mathbb{N}}

(1) P(x,D)=|α|maα(x)DαP(x,D)=\sum_{\left\lvert\alpha\right\rvert\leq m}a_{\alpha}(x)D^{\alpha}

where the coefficients aα=aα(x)a_{\alpha}=a_{\alpha}(x) are functions defined in Ω\Omega. We study a nonlocal inverse problem for the perturbed fractional Schrödinger equation

(2) {(Δ)su+P(x,D)u=0inΩu=finΩe\begin{cases}(-\Delta)^{s}u+P(x,D)u=0\ \text{in}\ \Omega\\ u=f\ \text{in}\ \Omega_{e}\end{cases}

where (Δ)s(-\Delta)^{s} is a nonlocal pseudo-differential operator (Δ)su=1(||2su^)(-\Delta)^{s}u=\mathcal{F}^{-1}(\left\lvert\cdot\right\rvert^{2s}\hat{u}) in contrast to the local operator P(x,D)P(x,D). In the inverse problem, one aims to recover the local operator PP from the associated Dirichlet-to-Neumann map.

We always assume that 0 is not a Dirichlet eigenvalue of the operator ((Δ)s+P(x,D))((-\Delta)^{s}+P(x,D)), i.e.

(3) IfuHs(n)solves((Δ)s+P(x,D))u=0inΩandu|Ωe=0,thenu=0.\text{If}\ u\in H^{s}({\mathbb{R}}^{n})\ \text{solves}\ ((-\Delta)^{s}+P(x,D))u=0\ \text{in}\ \Omega\ \text{and}\ u|_{\Omega_{e}}=0,\ \text{then}\ u=0.

Our data for the inverse problem is the exterior Dirichlet-to-Neumann (DN) map ΛP:Hs(Ωe)(Hs(Ωe))\Lambda_{P}\colon H^{s}(\Omega_{e})\rightarrow(H^{s}(\Omega_{e}))^{*} which maps Dirichlet exterior values to a nonlocal version of the Neumann boundary value (see section 2 and 3.1). The main question that we study in this article is whether the exterior DN map ΛP\Lambda_{P} determines uniquely the coefficients aαa_{\alpha} in Ω\Omega. In other words, does ΛP1=ΛP2\Lambda_{P_{1}}=\Lambda_{P_{2}} imply that a1,α=a2,αa_{1,\alpha}=a_{2,\alpha} in Ω\Omega for all |α|m\left\lvert\alpha\right\rvert\leq m? We prove that the answer is positive under certain restrictions on the coefficients aαa_{\alpha} and the order of the PDOs.

This gives a positive answer to the uniqueness problem [12, Question 7.5] posed by the first three authors in a previous work. The precise statement in [12] asks to prove uniqueness for the higher order fractional Calderón problem in the case of a bounded domain with smooth boundary and PDOs with smooth coefficients (up to the boundary). The positive answer to this question follows from theorem 1.4. The study of the fractional Calderón problem was initiated by Ghosh, Salo and Uhlmann in the work [18] where the uniqueness for the associated inverse problem is proved when m=0m=0, s(0,1)s\in(0,1) and a0L(Ω)a_{0}\in L^{\infty}(\Omega).

We briefly note that by Peetre’s theorem any linear operator L:Cc(Ω)Cc(Ω)L\colon C_{c}^{\infty}(\Omega)\rightarrow C_{c}^{\infty}(\Omega) which does not increase supports, i.e. spt(Lf)spt(f)\operatorname{spt}(Lf)\subset\operatorname{spt}(f) for all fCc(Ω)f\in C_{c}^{\infty}(\Omega), is in fact a differential operator [38] (see also the original work [40]). Therefore our results apply to any local operator satisfying such properties and it is enough to study PDOs only. For a more general formulation of Peetre’s theorem on the level of vector bundles, see  [39].

1.1. Main results

We denote by M(Hs|α|Hs)M(H^{s-|\alpha|}\rightarrow H^{-s}) the space of all bounded Sobolev multipliers between the Sobolev spaces Hs|α|(n)H^{s-|\alpha|}({\mathbb{R}}^{n}) and Hs(n)H^{-s}({\mathbb{R}}^{n}). We denote by M0(Hs|α|Hs)M(Hs|α|Hs)M_{0}(H^{s-|\alpha|}\rightarrow H^{-s})\subset M(H^{s-|\alpha|}\rightarrow H^{-s}) the space of bounded Sobolev multipliers that can be approximated by smooth compactly supported functions in the multiplier norm of M(Hs|α|Hs)M(H^{s-|\alpha|}\rightarrow H^{-s}). We also write Hr,(Ω)H^{r,\infty}(\Omega) for the local Bessel potential space with bounded derivatives. See section 2 for more detailed definitions.

Our first theorem is a generalization of [44, Theorem 1.1] which considered the case m=0m=0 with s(0,1)s\in(0,1). It also generalizes [12, Theorem 1.5] which considered the higher order case s+s\in{\mathbb{R}}^{+}\setminus{\mathbb{Z}} when m=0m=0.

Theorem 1.1.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded open set where n1n\geq 1. Let s+s\in\mathbb{R}^{+}\setminus\mathbb{Z} and mm\in\mathbb{N} be such that 2s>m2s>m. Let

Pj=|α|maj,αDα,j=1,2,P_{j}=\sum_{|\alpha|\leq m}a_{j,\alpha}D^{\alpha},\quad j=1,2,

be linear PDOs of order mm with coefficients aj,αM0(Hs|α|Hs)a_{j,\alpha}\in M_{0}(H^{s-|\alpha|}\rightarrow H^{-s}). Given any two open sets W1,W2ΩeW_{1},W_{2}\subset\Omega_{e}, suppose that the exterior DN maps ΛPi\Lambda_{P_{i}} for the equations ((Δ)s+Pj)u=0((-\Delta)^{s}+P_{j})u=0 in Ω\Omega satisfy

ΛP1f|W2=ΛP2f|W2\Lambda_{P_{1}}f|_{W_{2}}=\Lambda_{P_{2}}f|_{W_{2}}

for all fCc(W1)f\in C^{\infty}_{c}(W_{1}). Then P1|Ω=P2|ΩP_{1}|_{\Omega}=P_{2}|_{\Omega}.

In theorem 1.1 one can pick the lower order coefficients (|α|<s\left\lvert\alpha\right\rvert<s) from Lp(Ω)L^{p}(\Omega) for high enough pp (especially from L(Ω)L^{\infty}(\Omega)) and higher order coefficients (s<|α|<2ss<\left\lvert\alpha\right\rvert<2s) from the closure of Cc(Ω)C^{\infty}_{c}(\Omega) in Hr,(Ω)H^{r,\infty}(\Omega) for certain values of rr\in{\mathbb{R}}. In the following propositions, which are proved in Section 2, we give more examples of Sobolev spaces which belong to the space of multipliers M0(Hs|α|Hs)M_{0}(H^{s-\left\lvert\alpha\right\rvert}\rightarrow H^{-s}):

Proposition 1.2.

Let Ωn\Omega\subset{\mathbb{R}}^{n} be an open set and let tt\in{\mathbb{R}} and rr\in{\mathbb{R}} be such that t>max{0,r}t>\max\{0,r\}. The following inclusions hold:

  1. (i)

    H~r,(Ω)M0(HrHt)\widetilde{H}^{r^{\prime},\infty}(\Omega)\subset M_{0}(H^{-r}\rightarrow H^{-t}) whenever rmax{0,r}r^{\prime}\geq\max\{0,r\}.

  2. (ii)

    H0r,(Ω)M0(HrHt)H^{r^{\prime},\infty}_{0}(\Omega)\subset M_{0}(H^{-r}\rightarrow H^{-t}) whenever rmax{0,r}r^{\prime}\geq\max\{0,r\} with r{12,32,52,}r^{\prime}\notin\{\frac{1}{2},\frac{3}{2},\frac{5}{2},\dotso\} and Ω\Omega is a Lipschitz domain.

  3. (iii)

    H~r(Ω)M0(HrHt)\widetilde{H}^{r^{\prime}}(\Omega)\subset M_{0}(H^{-r}\rightarrow H^{-t}) whenever rtr^{\prime}\geq t and r>n/2r^{\prime}>n/2. The same holds true for HΩ¯r(n)H^{r^{\prime}}_{\overline{\Omega}}({\mathbb{R}}^{n}) if Ω\Omega is a Lipschitz domain, and for H0r(Ω)H^{r^{\prime}}_{0}(\Omega) when Ω\Omega is a Lipschitz domain and r{12,32,52,}r^{\prime}\notin\{\frac{1}{2},\frac{3}{2},\frac{5}{2},\dotso\}.

Note that the assumptions in theorem 1.1 satisfy the conditions of proposition 1.2 since then r=|α|sr=\left\lvert\alpha\right\rvert-s and t=st=s. The next proposition gives examples of spaces of lower order coefficients (|α|s\left\lvert\alpha\right\rvert\leq s):

Proposition 1.3.

Let Ωn\Omega\subset{\mathbb{R}}^{n} be an open set and t>0t>0. The following inclusions hold:

  1. (i)

    Lp(Ω)M0(H0Ht)L^{p}(\Omega)\subset M_{0}(H^{0}\rightarrow H^{-t}) whenever 2p<2\leq p<\infty and p>n/tp>n/t. Especially, if Ω\Omega is bounded, then L(Ω)M0(H0Ht)L^{\infty}(\Omega)\subset M_{0}(H^{0}\rightarrow H^{-t}).

  2. (ii)

    H~r(Ω)M0(H0Ht)\widetilde{H}^{r}(\Omega)\subset M_{0}(H^{0}\rightarrow H^{-t}) whenever r0r\geq 0 and r>n/2tr>n/2-t. The same holds true for HΩ¯r(n)H^{r}_{\overline{\Omega}}({\mathbb{R}}^{n}) if Ω\Omega is a Lipschitz domain, and for H0r(Ω)H^{r}_{0}(\Omega) when Ω\Omega is Lipschitz domain and r{12,32,52,}r\notin\{\frac{1}{2},\frac{3}{2},\frac{5}{2},\dotso\}.

As mentioned above, we put t=s>0t=s>0 in theorem 1.1 and the condition in proposition 1.3 is satisfied. Note that under the assumption |α|s\left\lvert\alpha\right\rvert\leq s we have M0(H0Hs)M0(Hs|α|Hs)M_{0}(H^{0}\rightarrow H^{-s})\subset M_{0}(H^{s-\left\lvert\alpha\right\rvert}\rightarrow H^{-s}). Hence we can choose the lower order coefficients from a less regular space in theorem 1.1 (compare to proposition 1.2). We also note that when |α|=0\left\lvert\alpha\right\rvert=0 the space of multipliers M0(HsHs)M_{0}(H^{s}\rightarrow H^{-s}) coincides with the one studied in [44].

It follows from lemma 2.5 that the space of multipliers is trivial for higher order operators, i.e. M(Hs|α|Hs)={0}M(H^{s-|\alpha|}\rightarrow H^{-s})=\{0\} when s|α|<ss-\left\lvert\alpha\right\rvert<-s. It would be possible to state theorem 1.1 for higher order PDOs, but that forces aα=0a_{\alpha}=0 for all |α|>2s\left\lvert\alpha\right\rvert>2s. For this reason we only consider PDOs whose order is m<2sm<2s. See lemma 2.5 and the related remarks for more details.

Our second theorem generalizes [8, Theorem 1.1] and [18, Theorem 1.1], where similar results are proved when m=0,1m=0,1 and s(0,1)s\in(0,1). It also generalizes [12, Theorem 1.5] where the case m=0m=0 and s+s\in{\mathbb{R}}^{+}\setminus{\mathbb{Z}} was studied.

Theorem 1.4.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded Lipschitz domain where n1n\geq 1. Let s+s\in\mathbb{R}^{+}\setminus\mathbb{Z} and mm\in\mathbb{N} be such that 2s>m2s>m. Let

Pj(x,D)=|α|maj,α(x)Dα,j=1,2,P_{j}(x,D)=\sum_{|\alpha|\leq m}a_{j,\alpha}(x)D^{\alpha},\quad j=1,2,

be a linear PDOs of order mm with coefficients aj,αHrα,(Ω)a_{j,\alpha}\in H^{r_{\alpha},\infty}(\Omega) where

(4) rα:={0if|α|s<0,|α|s+δif|α|s{1/2,3/2,},|α|sifotherwise.\displaystyle r_{\alpha}:=\Bigg{\{}\begin{matrix}0&\mbox{if}&|\alpha|-s<0,\\ |\alpha|-s+\delta&\mbox{if}&|\alpha|-s\in\{1/2,3/2,...\},\\ |\alpha|-s&\mbox{if}&\mbox{otherwise}\\ \end{matrix}\Bigg{.}

for any fixed δ>0\delta>0. Given any two open sets W1,W2ΩeW_{1},W_{2}\subset\Omega_{e}, suppose that the exterior DN maps ΛPi\Lambda_{P_{i}} for the equations ((Δ)s+Pj(x,D))u=0((-\Delta)^{s}+P_{j}(x,D))u=0 in Ω\Omega satisfy

ΛP1f|W2=ΛP2f|W2\Lambda_{P_{1}}f|_{W_{2}}=\Lambda_{P_{2}}f|_{W_{2}}

for all fCc(W1)f\in C^{\infty}_{c}(W_{1}). Then P1(x,D)=P2(x,D)P_{1}(x,D)=P_{2}(x,D).

Our first theorem is formulated for general bounded open sets and the second theorem for Lipschitz domains. The difference arises in the proof of the well-posedness of the forward problem. We note that theorem 1.4 holds for coefficients aαa_{\alpha} which are smooth up to the boundary (aα=g|Ωa_{\alpha}=g|_{\Omega} where gC(n)g\in C^{\infty}({\mathbb{R}}^{n})). The conditions (4) imply that one can choose aαL(Ω)a_{\alpha}\in L^{\infty}(\Omega) for every α\alpha such that |α|<s|\alpha|<s. The case |α|=s|\alpha|=s never happens, as ss is assumed not to be an integer. If |α|>s|\alpha|>s, we have aαH|α|s,(Ω)a_{\alpha}\in H^{|\alpha|-s,\infty}(\Omega) when |α|s{1/2,3/2,}|\alpha|-s\not\in\{1/2,3/2,...\}. Thus the conditions (4) coincide with [8, 18] when m=0,1m=0,1 and s(0,1)s\in(0,1).

Our article is roughly divided into two parts. The first part of the article (theorem 1.1 and section 3) generalizes the study of the uniqueness problem for singular potentials in [44] and the second part (theorem 1.4 and section 4) generalizes the uniqueness problem for bounded first order perturbations in [8].

The approach to prove theorems 1.1 and 1.4 is the following. First one shows that the forward problem is well-posed and the corresponding bilinear forms are bounded. This leads to the boundedness of the exterior DN maps and an Alessandrini identity. By a unique continuation property of the higher order fractional Laplacian one obtains a Runge approximation property for equation (2). Using the Runge approximation and the Alessandrini identity for suitable test functions one proves the uniqueness of the inverse problem.

1.2. On the earlier literature

Equation (2) and theorems 1.1 and 1.4 are related to the Calderón problem for the fractional Schrödinger equation first introduced in [18]. There one tries to uniquely recover the potential qq in Ω\Omega by doing measurements in the exterior Ωe\Omega_{e}. This is a nonlocal (fractional) counterpart of the classical Calderón problem arising in electrical impedance tomography, where one obtains information about the electrical properties of some bounded domain by doing voltage and current measurements on the boundary [49, 50]. In [44] the study of the fractional Calderón problem is extended for “rough” potentials qq, i.e. potentials which are in general bounded Sobolev multipliers. First order perturbations were studied in [8] assuming that the fractional part dominates the equation, i.e. s(1/2,1)s\in(1/2,1), and that the perturbations have bounded fractional derivatives. A higher order version (s+s\in{\mathbb{R}}^{+}\setminus{\mathbb{Z}}) of the fractional Calderón problem was introduced and studied in [12]. These three articles [8, 12, 44] motivate the study of higher order (rough) perturbations to the fractional Laplacian (Δ)s(-\Delta)^{s} in equation (2). The natural restriction for the order of  P(x,D)P(x,D) in theorems  1.1 and 1.4 is then 2s>m2s>m, so that the fractional part governs the equation (2).

The fractional Calderón problem for s(0,1)s\in(0,1) has been studied in many settings. We refer to the survey  [46] for a more detailed treatment. In the work [44] stability was proved for singular potentials, and in [43] the related exponential instability was shown. The fractional Calderón problem has also been solved under single measurement [17]. The perturbed equation is related to the fractional magnetic Schrödinger equation which is studied in [10, 30, 31, 32]. See also [5] for a fractional Schrödinger equation with a lower order nonlocal perturbation. Other variants of the fractional Calderón problem include semilinear fractional (magnetic) Schrödinger equation  [24, 25, 30, 32], fractional heat equation [26, 45] and fractional conductivity equation [11] (see also [7, 16] for equations arising from a nonlocal Schrödinger-type elliptic operator). In the recent work [12], the first three authors of this article studied higher order versions (s+s\in{\mathbb{R}}^{+}\setminus{\mathbb{Z}}) of the fractional Calderón problem and proved uniqueness for the Calderón problem for the fractional magnetic Schrödinger equation (up to a gauge). This article continues these studies by showing uniqueness for the fractional Schrödinger equation with higher order perturbations and gives positive answer to a question posed in  [12, Question 7.5].

1.3. Examples of fractional models in the sciences

Equations involving fractional Laplacians like (2) have applications in mathematics and natural sciences. Fractional Laplacians appear in the study of anomalous and nonlocal diffusion, and these diffusion phenomena can be used in many areas such as continuum mechanics, graph theory and ecology just to mention a few [2, 6, 15, 34, 41]. Another place where the fractional counterpart of the classical Laplacian naturally shows up is the formulation of fractional quantum mechanics [27, 28, 29]. See [42] and references therein for possible applications of higher order fractional Laplacians. For more applications of fractional mathematical models, see  [6, 37] and the references therein.

1.4. The organization of the article

In section 2 we introduce the notation and give preliminaries on Sobolev spaces and fractional Laplacians. We also define the spaces of rough coefficients (Sobolev multipliers) and discuss some of the basic properties. In section 3 we prove theorem 1.1 in detail. Finally, in section 4 we prove theorem 1.4 but as the proofs of both theorems are very similar we do not repeat all identical steps and we keep our focus in the differences of the proofs.

Acknowledgements

G.C. was partially supported by the European Research Council under Horizon 2020 (ERC CoG 770924). K.M. and J.R. were supported by Academy of Finland (Centre of Excellence in Inverse Modelling and Imaging, grant numbers 284715 and 309963). G.U. was partly supported by NSF, a Walker Family Endowed Professorship at UW and a Si Yuan Professorship at IAS, HKUST. G.C. would like to thank Angkana Rüland for helpful discussions and her hospitality during his visit to Max Planck Institute for Mathematics in the Sciences. J.R. and G.U. wish to thank Maarten V. de Hoop, Rice University and Simons Foundation for providing the support to participate 2020 MATH + X Symposium on Inverse Problems and Deep Learning, Mitigating Natural Hazards, where this project was initiated.

2. Preliminaries

In this section we recall some basic theory of Sobolev spaces, Fourier analysis and fractional Laplacians on n{\mathbb{R}}^{n}. We also introduce the spaces of Sobolev multipliers and prove a few properties for them. Some auxiliary lemmas which are needed in the proofs of our main theorems are given as well. We follow the references [1, 18, 36, 35, 48, 51] (see also section 2 in [12]).

2.1. Sobolev spaces

The (inhomogeneous) fractional L2L^{2}-based Sobolev space of order rr\in{\mathbb{R}} is defined to be

(5) Hr(n)={u𝒮(n):1(ru^)L2(n)}H^{r}({\mathbb{R}}^{n})=\{u\in\mathscr{S}^{\prime}({\mathbb{R}}^{n}):\mathcal{F}^{-1}(\langle\cdot\rangle^{r}\hat{u})\in L^{2}({\mathbb{R}}^{n})\}

equipped with the norm

(6) uHr(n)=1(ru^)L2(n).\left\lVert u\right\rVert_{H^{r}({\mathbb{R}}^{n})}=\left\lVert\mathcal{F}^{-1}(\langle\cdot\rangle^{r}\hat{u})\right\rVert_{L^{2}({\mathbb{R}}^{n})}.

Here u^=(u)\hat{u}=\mathcal{F}(u) is the Fourier transform of a tempered distribution u𝒮(n)u\in\mathscr{S}^{\prime}({\mathbb{R}}^{n}), 1\mathcal{F}^{-1} is the inverse Fourier transform and x=(1+|x|2)1/2\langle x\rangle=(1+\left\lvert x\right\rvert^{2})^{1/2}. We define the fractional Laplacian of order s+s\in{\mathbb{R}}^{+}\setminus{\mathbb{Z}} as (Δ)sφ=1(||2sφ^)(-\Delta)^{s}\varphi=\mathcal{F}^{-1}(\left\lvert\cdot\right\rvert^{2s}\hat{\varphi}) where φ𝒮(n)\varphi\in\mathscr{S}({\mathbb{R}}^{n}) is a Schwartz function. Then (Δ)s(-\Delta)^{s} extends to a bounded operator (Δ)s:Hr(n)Hr2s(n)(-\Delta)^{s}\colon H^{r}({\mathbb{R}}^{n})\rightarrow H^{r-2s}({\mathbb{R}}^{n}) for all rr\in{\mathbb{R}} by density of 𝒮(n)\mathscr{S}({\mathbb{R}}^{n}) in Hr(n)H^{r}({\mathbb{R}}^{n}) [18, Lemma 2.1] (see also [12, Section 2.2]). It would be possible to define the fractional Laplacian in many other ways, according to the intended application (check e.g. [13, 23, 33]). In particular, our global definition of the fractional Laplacian using Fourier transform will be different from the spectral definition used in [21] (see for example [13, 47]).

Let Ωn\Omega\subset{\mathbb{R}}^{n} be an open set and FnF\subset{\mathbb{R}}^{n} a closed set. We define the following Sobolev spaces

(7) HFr(n)\displaystyle H_{F}^{r}({\mathbb{R}}^{n}) ={uHr(n):spt(u)F}\displaystyle=\{u\in H^{r}({\mathbb{R}}^{n}):\operatorname{spt}(u)\subset F\}
(8) H~r(Ω)\displaystyle\widetilde{H}^{r}(\Omega) =closure ofCc(Ω)inHr(n)\displaystyle=\ \text{closure of}\ C_{c}^{\infty}(\Omega)\ \text{in}\ H^{r}({\mathbb{R}}^{n})
(9) Hr(Ω)\displaystyle H^{r}(\Omega) ={u|Ω:uHr(n)}\displaystyle=\{u|_{\Omega}:u\in H^{r}({\mathbb{R}}^{n})\}
(10) H0r(Ω)\displaystyle H_{0}^{r}(\Omega) =closure ofCc(Ω)inHr(Ω).\displaystyle=\ \text{closure of}\ C_{c}^{\infty}(\Omega)\ \text{in}\ H^{r}(\Omega).

It trivially follows that H~r(Ω)H0r(Ω)\widetilde{H}^{r}(\Omega)\subset H_{0}^{r}(\Omega) and H~r(Ω)HΩ¯r(n)\widetilde{H}^{r}(\Omega)\subset H^{r}_{\overline{\Omega}}({\mathbb{R}}^{n}). Further, we have (H~r(Ω))=Hr(Ω)(\widetilde{H}^{r}(\Omega))^{*}=H^{-r}(\Omega) and (Hr(Ω))=H~r(Ω)(H^{r}(\Omega))^{*}=\widetilde{H}^{-r}(\Omega) for any open set Ω\Omega and rr\in{\mathbb{R}} [9, Theorem 3.3]. If Ω\Omega is in addition a Lipschitz domain, then we have H~r(Ω)=HΩ¯r(n)\widetilde{H}^{r}(\Omega)=H_{\overline{\Omega}}^{r}({\mathbb{R}}^{n}) for all rr\in{\mathbb{R}} and H0r(Ω)=HΩ¯r(n)H_{0}^{r}(\Omega)=H^{r}_{\overline{\Omega}}({\mathbb{R}}^{n}) when r0r\geq 0 such that r{12,32,52}r\notin\{\frac{1}{2},\frac{3}{2},\frac{5}{2}\dotso\} [36, Theorems 3.29 and 3.33].

More generally, let 1p1\leq p\leq\infty and rr\in{\mathbb{R}}. We define the Bessel potential space

(11) Hr,p(n)={u𝒮(n):1(ru^)Lp(n)}H^{r,p}({\mathbb{R}}^{n})=\{u\in\mathscr{S}^{\prime}({\mathbb{R}}^{n}):\mathcal{F}^{-1}(\langle\cdot\rangle^{r}\hat{u})\in L^{p}({\mathbb{R}}^{n})\}

equipped with the norm

(12) uHr,p(n)=1(ru^)Lp(n).\left\lVert u\right\rVert_{H^{r,p}({\mathbb{R}}^{n})}=\left\lVert\mathcal{F}^{-1}(\langle\cdot\rangle^{r}\hat{u})\right\rVert_{L^{p}({\mathbb{R}}^{n})}.

We also write 1(ru^)=:Jru\mathcal{F}^{-1}(\langle\cdot\rangle^{r}\hat{u})=:J^{r}u where the Fourier multiplier J=(IdΔ)1/2J=(\mathrm{Id}-\Delta)^{1/2} is called the Bessel potential. We have the continuous inclusions Hr,p(n)Ht,p(n)H^{r,p}({\mathbb{R}}^{n})\hookrightarrow H^{t,p}({\mathbb{R}}^{n}) whenever rtr\geq t [4, Theorem 6.2.3]. By the Mikhlin multiplier theorem one can show that (Δ)s:Hr,p(n)Hr2s,p(n)(-\Delta)^{s}\colon H^{r,p}({\mathbb{R}}^{n})\rightarrow H^{r-2s,p}({\mathbb{R}}^{n}) is continuous whenever s0s\geq 0 and 1<p<1<p<\infty (see [18, Remark 2.2] and [1, Theorem 7.2]). The local version of the space Hr,p(n)H^{r,p}({\mathbb{R}}^{n}) is defined as earlier by the restrictions

(13) Hr,p(Ω)={u|Ω:uHr,p(n)}H^{r,p}(\Omega)=\{u|_{\Omega}:u\in H^{r,p}({\mathbb{R}}^{n})\}

where Ωn\Omega\subset{\mathbb{R}}^{n} is any open set. This space is equipped with the quotient norm

(14) vHr,p(Ω)=inf{wHr,p(n):wHr,p(n),w|Ω=v}.\left\lVert v\right\rVert_{H^{r,p}(\Omega)}=\inf\{\left\lVert w\right\rVert_{H^{r,p}({\mathbb{R}}^{n})}:w\in H^{r,p}({\mathbb{R}}^{n}),\ w|_{\Omega}=v\}.

We have the continuous inclusions Hr,p(Ω)Ht,p(Ω)H^{r,p}(\Omega)\hookrightarrow H^{t,p}(\Omega) whenever rtr\geq t by the definition of the quotient norm.

We also define the spaces

(15) HFr,p(n)\displaystyle H_{F}^{r,p}({\mathbb{R}}^{n}) ={uHr,p(n):spt(u)F}\displaystyle=\{u\in H^{r,p}({\mathbb{R}}^{n}):\operatorname{spt}(u)\subset F\}
(16) H~r,p(Ω)\displaystyle\widetilde{H}^{r,p}(\Omega) =closure ofCc(Ω)inHr,p(n)\displaystyle=\ \text{closure of}\ C_{c}^{\infty}(\Omega)\ \text{in}\ H^{r,p}({\mathbb{R}}^{n})
(17) H0r,p(Ω)\displaystyle H_{0}^{r,p}(\Omega) =closure ofCc(Ω)inHr,p(Ω)\displaystyle=\ \text{closure of}\ C_{c}^{\infty}(\Omega)\ \text{in}\ H^{r,p}(\Omega)

where FnF\subset{\mathbb{R}}^{n} is a closed set. Note that H~r,p(Ω)H0r,p(Ω)\widetilde{H}^{r,p}(\Omega)\subset H_{0}^{r,p}(\Omega) since the restriction map |Ω:Hr,p(n)Hr,p(Ω)|_{\Omega}\colon H^{r,p}({\mathbb{R}}^{n})\rightarrow H^{r,p}(\Omega) is by definition continuous. One can also see that H~r,p(Ω)HΩ¯r,p(n).\widetilde{H}^{r,p}(\Omega)\subset H^{r,p}_{\overline{\Omega}}({\mathbb{R}}^{n}). If Ω\Omega is a bounded CC^{\infty}-domain and 1<p<1<p<\infty, then we have [48, Theorem 1 in section 4.3.2]

(18) H~r,p(Ω)\displaystyle\widetilde{H}^{r,p}(\Omega) =HΩ¯r,p(n),r\displaystyle=H^{r,p}_{\overline{\Omega}}({\mathbb{R}}^{n}),\quad r\in{\mathbb{R}}
(19) H0r,p(Ω)\displaystyle H^{r,p}_{0}(\Omega) =Hr,p(Ω),r1p.\displaystyle=H^{r,p}(\Omega),\quad r\leq\frac{1}{p}.

Some authors (especially in [8, 44]) use the notation Wr,p(Ω)W^{r,p}(\Omega) for Bessel potential spaces. We have decided to use the notation Hr,p(Ω)H^{r,p}(\Omega) so that these spaces are not confused with the Sobolev-Slobodeckij spaces which are in general different from the Bessel potential spaces [14, Remark 3.5].

The equation (2) we study is nonlocal. Instead of putting boundary conditions we impose exterior values for the equation. This can be done by saying that u=fu=f in Ωe\Omega_{e} if ufH~s(Ω)u-f\in\widetilde{H}^{s}(\Omega). Motivated by this we define the (abstract) trace space X=Hr(n)/H~r(Ω)X=H^{r}({\mathbb{R}}^{n})/\widetilde{H}^{r}(\Omega), i.e. functions in XX are the same (have the same trace) if they agree in Ωe\Omega_{e}. If Ω\Omega is a Lipschitz domain, then we have X=Hr(Ωe)X=H^{r}(\Omega_{e}) and X=HΩ¯er(n)X^{*}=H^{-r}_{\overline{\Omega}_{e}}({\mathbb{R}}^{n}) [18, p.463].

2.2. Properties of the fractional Laplacian

The fractional Laplacian admits two important properties which we need in our proofs. The first one is unique continuation property (UCP) which is used in proving the Runge approximation property.

Lemma 2.1 (UCP).

Let s+s\in{\mathbb{R}}^{+}\setminus{\mathbb{Z}}, rr\in{\mathbb{R}} and uHr(n)u\in H^{r}({\mathbb{R}}^{n}). If (Δ)su|V=0(-\Delta)^{s}u|_{V}=0 and u|V=0u|_{V}=0 for some nonempty open set VnV\subset{\mathbb{R}}^{n}, then u=0u=0.

Lemma 2.1 is proved in [12] for s>1s>1 by reducing the problem to the UCP result for s(0,1)s\in(0,1) in [18]. Note that such property is not true for local operators like the classical Laplacian (Δ)(-\Delta). The second property we need is the Poincaré inequality, which is used in showing that the forward problem for the perturbed fractional Schrödinger equation is well-posed.

Lemma 2.2 (Poincaré inequality).

Let s+s\in{\mathbb{R}}^{+}\setminus{\mathbb{Z}}, KnK\subset{\mathbb{R}}^{n} compact set and uHKs(n)u\in H^{s}_{K}({\mathbb{R}}^{n}). There exists a constant c=c(n,K,s)>0c=c(n,K,s)>0 such that

(20) uL2(n)c(Δ)s/2uL2(n).\left\lVert u\right\rVert_{L^{2}({\mathbb{R}}^{n})}\leq c\left\lVert(-\Delta)^{s/2}u\right\rVert_{L^{2}({\mathbb{R}}^{n})}.

Many different proofs for lemma 2.2 are given in [12]. We note that in the literature the fractional Poincaré inequality is typically considered only when s(0,1)s\in(0,1).

Finally, we recall the fractional Leibniz rule, also known as the Kato-Ponce inequality. It is used to show the boundedness of the bilinear forms associated to the perturbed fractional Schrödinger equation in the case when the coefficients of the PDO have bounded fractional derivatives.

Lemma 2.3 (Kato-Ponce inequality).

Let s0s\geq 0, 1<r<1<r<\infty, 1<q11<q_{1}\leq\infty and 1<p21<p_{2}\leq\infty such that 1r=1p1+1q1=1p2+1q2\frac{1}{r}=\frac{1}{p_{1}}+\frac{1}{q_{1}}=\frac{1}{p_{2}}+\frac{1}{q_{2}}. If fLp2(n)f\in L^{p_{2}}({\mathbb{R}}^{n}), JsfLp1(n)J^{s}f\in L^{p_{1}}({\mathbb{R}}^{n}), gLq1(n)g\in L^{q_{1}}({\mathbb{R}}^{n}) and JsgLq2(n)J^{s}g\in L^{q_{2}}({\mathbb{R}}^{n}), then Js(fg)Lr(n)J^{s}(fg)\in L^{r}({\mathbb{R}}^{n}) and

(21) Js(fg)Lr(n)C(JsfLp1(n)gLq1(n)+fLp2(n)JsgLq2(n))\left\lVert J^{s}(fg)\right\rVert_{L^{r}({\mathbb{R}}^{n})}\leq C(\left\lVert J^{s}f\right\rVert_{L^{p_{1}}({\mathbb{R}}^{n})}\left\lVert g\right\rVert_{L^{q_{1}}({\mathbb{R}}^{n})}+\left\lVert f\right\rVert_{L^{p_{2}}({\mathbb{R}}^{n})}\left\lVert J^{s}g\right\rVert_{L^{q_{2}}({\mathbb{R}}^{n})})

where JsJ^{s} is the Bessel potential of order ss and C=C(s,n,r,p1,p2,q1,q2)C=C(s,n,r,p_{1},p_{2},q_{1},q_{2}).

The proof of lemma 2.3 can be found in [20] (see also [19, 22]).

2.3. Spaces of rough coefficients

Following [35, Ch. 3], we introduce the space of multipliers M(HrHt)M(H^{r}\rightarrow H^{t}) between pairs of Sobolev spaces. Here we are assuming that r,tr,t\in\mathbb{R}. The coefficients of P(x,D)P(x,D) in theorem 1.1 will be picked from such spaces of multipliers.

If f𝒟(n)f\in\mathcal{D}^{\prime}({\mathbb{R}}^{n}) is a distribution, we say that fM(HrHt)f\in M(H^{r}\rightarrow H^{t}) whenever the norm

fr,t:=sup{|f,uv|;u,vCc(n),uHr(n)=vHt(n)=1}\|f\|_{r,t}:=\sup\{\left\lvert\left\langle f,uv\right\rangle\right\rvert\,;\,u,v\in C_{c}^{\infty}(\mathbb{R}^{n}),\left\lVert u\right\rVert_{H^{r}({\mathbb{R}}^{n})}=\left\lVert v\right\rVert_{H^{-t}({\mathbb{R}}^{n})}=1\}

is finite. Here uvuv indicates the pointwise product of functions, while ,\left\langle\cdot,\cdot\right\rangle is the duality pairing. If the distribution ff happens to be a function, the duality pairing can be defined as

f,uv=nf(x)u(x)v(x)dx.\langle f,uv\rangle=\int_{\mathbb{R}^{n}}f(x)u(x)v(x){\mathrm{d}}x.

By M0(HrHt)M_{0}(H^{r}\rightarrow H^{t}) we indicate the closure of Cc(n)C^{\infty}_{c}(\mathbb{R}^{n}) in M(HrHt)𝒟(n)M(H^{r}\rightarrow H^{t})\subset\mathcal{D}^{\prime}(\mathbb{R}^{n}). If fM(HrHt)f\in M(H^{r}\rightarrow H^{t}) and u,vCc(n)u,v\in C_{c}^{\infty}(\mathbb{R}^{n}) are both non-vanishing, we have the multiplier inequality

(22) |f,uv|=|f,uuHr(n)vvHt(n)|uHr(n)vHt(n)fr,tuHr(n)vHt(n).\left\lvert\left\langle f,uv\right\rangle\right\rvert=\left|\left\langle f,\frac{u}{\left\lVert u\right\rVert_{H^{r}({\mathbb{R}}^{n})}}\frac{v}{\left\lVert v\right\rVert_{H^{-t}({\mathbb{R}}^{n})}}\right\rangle\right|\left\lVert u\right\rVert_{H^{r}({\mathbb{R}}^{n})}\left\lVert v\right\rVert_{H^{-t}({\mathbb{R}}^{n})}\leq\|f\|_{r,t}\left\lVert u\right\rVert_{H^{r}({\mathbb{R}}^{n})}\left\lVert v\right\rVert_{H^{-t}({\mathbb{R}}^{n})}.

By the density of Cc(n)×Cc(n)C_{c}^{\infty}({\mathbb{R}}^{n})\times C_{c}^{\infty}({\mathbb{R}}^{n}) in Hr(n)×Ht(n)H^{r}({\mathbb{R}}^{n})\times H^{-t}({\mathbb{R}}^{n}) with respect to the product norm (u,v)=max{uHr(n),vHt(n)}\left\lVert(u,v)\right\rVert=\max\{\left\lVert u\right\rVert_{H^{r}({\mathbb{R}}^{n})},\left\lVert v\right\rVert_{H^{-t}({\mathbb{R}}^{n})}\} and estimate (22), there is a unique continuous extension of (u,v)f,uv(u,v)\mapsto\langle f,uv\rangle acting on (u,v)Hr(n)×Ht(n)(u,v)\in H^{r}({\mathbb{R}}^{n})\times H^{-t}({\mathbb{R}}^{n}). More precisely, each fM(HrHt)f\in M(H^{r}\rightarrow H^{t}) gives rise to a linear multiplication map mf:Hr(n)Ht(n)m_{f}:H^{r}({\mathbb{R}}^{n})\rightarrow H^{t}({\mathbb{R}}^{n}) defined by

(23) mf(u),v:=limif,uivifor all(u,v)Hr(n)×Ht(n),\displaystyle\langle m_{f}(u),v\rangle:=\lim_{i\to\infty}\langle f,u_{i}v_{i}\rangle\quad\mbox{for all}\quad(u,v)\in H^{r}({\mathbb{R}}^{n})\times H^{-t}({\mathbb{R}}^{n}),

where (ui,vi)Cc(n)×Cc(n)(u_{i},v_{i})\in C_{c}^{\infty}({\mathbb{R}}^{n})\times C_{c}^{\infty}({\mathbb{R}}^{n}) is any Cauchy sequence in Hr(n)×Ht(n)H^{r}({\mathbb{R}}^{n})\times H^{-t}({\mathbb{R}}^{n}) converging to (u,v)(u,v). The existence of the limit is granted by completeness and formula (22), which ensures that f,uivi\langle f,u_{i}v_{i}\rangle is also a Cauchy sequence. In fact, we have

|f,umvmf,unvn||f,um(vmvn)|+|f,(umun)vn|fr,t(umHr(n)vmvnHt(n)+umunHr(n)vnHt(n))fr,t(umHr(n)+vnHt(n))(um,vm)(un,vn),\begin{split}\left\lvert\langle f,u_{m}v_{m}\rangle-\langle f,u_{n}v_{n}\rangle\right\rvert&\leq\left\lvert\langle f,u_{m}(v_{m}-v_{n})\rangle\right\rvert+\left\lvert\langle f,(u_{m}-u_{n})v_{n}\rangle\right\rvert\\ &\leq\left\lVert f\right\rVert_{r,t}\left(\left\lVert u_{m}\right\rVert_{H^{r}({\mathbb{R}}^{n})}\left\lVert v_{m}-v_{n}\right\rVert_{H^{-t}({\mathbb{R}}^{n})}+\left\lVert u_{m}-u_{n}\right\rVert_{H^{r}({\mathbb{R}}^{n})}\left\lVert v_{n}\right\rVert_{H^{-t}({\mathbb{R}}^{n})}\right)\\ &\leq\left\lVert f\right\rVert_{r,t}\left(\left\lVert u_{m}\right\rVert_{H^{r}({\mathbb{R}}^{n})}+\left\lVert v_{n}\right\rVert_{H^{-t}({\mathbb{R}}^{n})}\right)\left\lVert(u_{m},v_{m})-(u_{n},v_{n})\right\rVert,\end{split}

where umHr(n)+vnHt(n)\left\lVert u_{m}\right\rVert_{H^{r}({\mathbb{R}}^{n})}+\left\lVert v_{n}\right\rVert_{H^{-t}({\mathbb{R}}^{n})} is bounded by a constant independent of mm and nn. The independence of the limit on the particular sequence (ui,vi)(u_{i},v_{i}) can be showed by a similar estimate.

We can analogously define the unique adjoint multiplication map mf:Ht(n)Hr(n)m_{f}^{*}:H^{-t}({\mathbb{R}}^{n})\to H^{-r}({\mathbb{R}}^{n}) such that

mf(v),u:=limif,uivifor all(u,v)Hr(n)×Ht(n).\left\langle m_{f}^{*}(v),u\right\rangle:=\lim_{i\to\infty}\langle f,u_{i}v_{i}\rangle\quad\mbox{for all}\quad(u,v)\in H^{r}({\mathbb{R}}^{n})\times H^{-t}({\mathbb{R}}^{n}).

Since one sees that the adjoint of mfm_{f} is mfm_{f}^{*}, the chosen notation is justified. For convenience, in the rest of the paper we will just write fufu for both mf(u)m_{f}(u) and mf(u)m_{f}^{*}(u).

Remark 2.4.

The spaces of rough coefficients we use are generalizations of the ones considered in [44]. In fact, the space Zs(n)Z^{-s}(\mathbb{R}^{n}) used there coincides with our space M(HsHs)M(H^{s}\rightarrow H^{-s}).

In the next lemma we state some elementary properties of the spaces of multipliers. Other interesting properties may be found in [35].

Lemma 2.5.

Let λ,μ0\lambda,\mu\geq 0 and r,tr,t\in\mathbb{R}. Then

  1. (i)

    M(HrHt)=M(HtHr)M(H^{r}\rightarrow H^{t})=M(H^{-t}\rightarrow H^{-r}), and the norms associated to the two spaces also coincide.

  2. (ii)

    M(HrλHt+μ)M(HrHt)M(H^{r-\lambda}\rightarrow H^{t+\mu})\hookrightarrow M(H^{r}\rightarrow H^{t}) continuously.

  3. (iii)

    M(HrHt)={0}M(H^{r}\rightarrow H^{t})=\{0\} whenever r<tr<t.

Proof.

i Let f𝒟(n)f\in\mathcal{D}^{\prime}({\mathbb{R}}^{n}) be a distribution. Then by just using the definition we see that

(24) fr,t\displaystyle\|f\|_{r,t} =sup{|f,uv|;u,vCc(n),uHr(n)=vHt(n)=1}\displaystyle=\sup\{\left\lvert\left\langle f,uv\right\rangle\right\rvert\,;\,u,v\in C_{c}^{\infty}(\mathbb{R}^{n}),\left\lVert u\right\rVert_{H^{r}({\mathbb{R}}^{n})}=\left\lVert v\right\rVert_{H^{-t}({\mathbb{R}}^{n})}=1\}
(25) =sup{|f,vu|;v,uCc(n),vHt(n)=uH(r)(n)=1}=ft,r.\displaystyle=\sup\{\left\lvert\left\langle f,vu\right\rangle\right\rvert\,;\,v,u\in C_{c}^{\infty}(\mathbb{R}^{n}),\left\lVert v\right\rVert_{H^{-t}({\mathbb{R}}^{n})}=\left\lVert u\right\rVert_{H^{-(-r)}({\mathbb{R}}^{n})}=1\}=\|f\|_{-t,-r}.

ii Observe that the given definition of fr,t\|f\|_{r,t} is equivalent to the following:

(26) fr,t=sup{|f,uv|;u,vCc(n),uHr(n)1,vHt(n)1}.\displaystyle\|f\|_{r,t}=\sup\{\left\lvert\left\langle f,uv\right\rangle\right\rvert\,;\,u,v\in C_{c}^{\infty}(\mathbb{R}^{n}),\left\lVert u\right\rVert_{H^{r}({\mathbb{R}}^{n})}\leq 1,\left\lVert v\right\rVert_{H^{-t}({\mathbb{R}}^{n})}\leq 1\}.

Since λ,μ0\lambda,\mu\geq 0, we also have

uHrλ(n)uHr(n),vH(t+μ)(n)vHt(n).\left\lVert u\right\rVert_{H^{r-\lambda}({\mathbb{R}}^{n})}\leq\left\lVert u\right\rVert_{H^{r}({\mathbb{R}}^{n})},\quad\left\lVert v\right\rVert_{H^{-(t+\mu)}({\mathbb{R}}^{n})}\leq\left\lVert v\right\rVert_{H^{-t}({\mathbb{R}}^{n})}.

This implies fr,tfrλ,t+μ\|f\|_{r,t}\leq\|f\|_{r-\lambda,t+\mu}, which in turn gives the wanted inclusion.

iii If 0r<t0\leq r<t, then this was considered in [35, Ch. 3]. The proof given there recalls the easier one for Sobolev spaces ([35, Sec. 2.1]), which is based on the explicit computation of derivatives of aptly chosen exponential functions.

If r<t0r<t\leq 0, then by point i we have M(HrHt)=M(HtHr)M(H^{r}\rightarrow H^{t})=M(H^{-t}\rightarrow H^{-r}). We need to show that M(HtHr)={0}M(H^{-t}\rightarrow H^{-r})=\{0\} whenever 0t<r0\leq-t<-r. This reduces the problem back to the case of non-negative Sobolev scales.

If r0<tr\leq 0<t, then r0-r\geq 0. Now by point ii, we have M(HrHt)M(Hr+(r)Ht)=M(L2Ht)M(H^{r}\rightarrow H^{t})\subseteq M(H^{r+(-r)}\rightarrow H^{t})=M(L^{2}\rightarrow H^{t}). It is therefore enough to show that this last space is trivial, which again immediately follows from the case of non-negative Sobolev scales.

If r<0tr<0\leq t, then the problem can be reduced again to the earlier cases. ∎

Remark 2.6.

We also have M0(HrλHt+μ)M0(HrHt)M_{0}(H^{r-\lambda}\rightarrow H^{t+\mu})\subseteq M_{0}(H^{r}\rightarrow H^{t}) whenever λ,μ0\lambda,\mu\geq 0, since the inclusion in ii is continuous.

Remark 2.7.

In light of lemma 2.5 ii we are only interested in M(HrHt)M(H^{r}\rightarrow H^{t}) in the case rtr\geq t, the case r<tr<t being trivial. For our theorem 1.1, this translates into the condition m2sm\leq 2s. We decided not to consider the limit case m=2sm=2s in this work, as our machinery (in particular, the coercivity estimate (99)) breaks down in this case. However, it should be noted that since by assumption we have mm\in\mathbb{Z} and ss\not\in\mathbb{Z}, the equality m=2sm=2s can only arise if mm is odd, which forces s=1/2+ks=1/2+k with kk\in\mathbb{Z}. This case was excluded in [8, 18] as well.

Propositions 1.2 and 1.3 relate our spaces of multipliers with some special Bessel potential spaces. This is interesting since in the coming section 3 we will consider the inverse problem for coefficients coming from such spaces. We now prove those propositions.

Proof of proposition 1.2.

Throughout the proof we assume that u,vCc(n)u,v\in C_{c}^{\infty}({\mathbb{R}}^{n}) such that uHr(n)=vHt(n)=1\left\lVert u\right\rVert_{H^{-r}({\mathbb{R}}^{n})}=\left\lVert v\right\rVert_{H^{t}({\mathbb{R}}^{n})}=1. In parts i and ii we can assume that r<tr^{\prime}<t since if rtr^{\prime}\geq t, then we have the continuous inclusion Hr,(Ω)Hr′′,(Ω)H^{r^{\prime},\infty}(\Omega)\hookrightarrow H^{r^{\prime\prime},\infty}(\Omega) where max{0,r}r′′<t\max\{0,r\}\leq r^{\prime\prime}<t (such r′′r^{\prime\prime} always exists since t>max{0,r}t>\max\{0,r\}).

i Let fH~r,(Ω)f\in\widetilde{H}^{r^{\prime},\infty}(\Omega). Now f=f1+f2f=f_{1}+f_{2} where f1Cc(Ω)f_{1}\in C^{\infty}_{c}(\Omega) and f2Hr,(n)ϵ\left\lVert f_{2}\right\rVert_{H^{r^{\prime},\infty}({\mathbb{R}}^{n})}\leq\epsilon. Then

(27) |f2,uv|f2vHr(n)uHr(n)\displaystyle\left\lvert\left\langle f_{2},uv\right\rangle\right\rvert\leq\left\lVert f_{2}v\right\rVert_{H^{r^{\prime}}({\mathbb{R}}^{n})}\left\lVert u\right\rVert_{H^{-r^{\prime}}({\mathbb{R}}^{n})} Cf2Hr,(n)vHr(n)uHr(n)\displaystyle\leq C\left\lVert f_{2}\right\rVert_{H^{r^{\prime},\infty}({\mathbb{R}}^{n})}\left\lVert v\right\rVert_{H^{r^{\prime}}({\mathbb{R}}^{n})}\left\lVert u\right\rVert_{H^{-r}({\mathbb{R}}^{n})}
(28) CϵvHt(n)=Cϵ.\displaystyle\leq C\epsilon\left\lVert v\right\rVert_{H^{t}({\mathbb{R}}^{n})}=C\epsilon.

Here we used the Kato-Ponce inequality (lemma 2.3)

(29) Jr(f2v)L2(n)\displaystyle\left\lVert J^{r^{\prime}}(f_{2}v)\right\rVert_{L^{2}({\mathbb{R}}^{n})} C(f2L(n)JrvL2(n)+Jrf2L(n)vL2(n))\displaystyle\leq C(\left\lVert f_{2}\right\rVert_{L^{\infty}({\mathbb{R}}^{n})}\left\lVert J^{r^{\prime}}v\right\rVert_{L^{2}({\mathbb{R}}^{n})}+\left\lVert J^{r^{\prime}}f_{2}\right\rVert_{L^{\infty}({\mathbb{R}}^{n})}\left\lVert v\right\rVert_{L^{2}({\mathbb{R}}^{n})})
(30) Cf2Hr,(n)vHr(n)\displaystyle\leq C\left\lVert f_{2}\right\rVert_{H^{r^{\prime},\infty}({\mathbb{R}}^{n})}\left\lVert v\right\rVert_{H^{r^{\prime}}({\mathbb{R}}^{n})}

and the assumption max{0,r}r<t\max\{0,r\}\leq r^{\prime}<t. Therefore ff1r,t=f2r,tCϵ\left\lVert f-f_{1}\right\rVert_{-r,-t}=\left\lVert f_{2}\right\rVert_{-r,-t}\leq C\epsilon which shows that fM0(HrHt)f\in M_{0}(H^{-r}\rightarrow H^{-t}).

ii Let fH0r,(Ω)f\in H^{r^{\prime},\infty}_{0}(\Omega). Now f=f1+f2f=f_{1}+f_{2} where f1Cc(Ω)f_{1}\in C^{\infty}_{c}(\Omega) and f2Hr,(Ω)ϵ\left\lVert f_{2}\right\rVert_{H^{r^{\prime},\infty}(\Omega)}\leq\epsilon. By the definition of the quotient norm Hr,(Ω)\left\lVert\cdot\right\rVert_{H^{r^{\prime},\infty}(\Omega)} we can take FHr,(n)F\in H^{r^{\prime},\infty}({\mathbb{R}}^{n}) such that F|Ω=f2F|_{\Omega}=f_{2} and FHr,(n)2f2Hr,(Ω)\left\lVert F\right\rVert_{H^{r^{\prime},\infty}({\mathbb{R}}^{n})}\leq 2\left\lVert f_{2}\right\rVert_{H^{r^{\prime},\infty}(\Omega)}. The assumptions imply the duality (Hr(Ω))=H0r(Ω)Hr(Ω)(H^{-r^{\prime}}(\Omega))^{*}=H^{r^{\prime}}_{0}(\Omega)\subset H^{r^{\prime}}(\Omega). Using the Kato-Ponce inequality for the extension FF we obtain as in the proof of part i that

(31) Jr(Fv)L2(n)\displaystyle\left\lVert J^{r^{\prime}}(Fv)\right\rVert_{L^{2}({\mathbb{R}}^{n})} CFHr,(n)vHr(n)2Cf2Hr,(Ω)vHt(n)2Cϵ\displaystyle\leq C\left\lVert F\right\rVert_{H^{r^{\prime},\infty}({\mathbb{R}}^{n})}\left\lVert v\right\rVert_{H^{r^{\prime}}({\mathbb{R}}^{n})}\leq 2C\left\lVert f_{2}\right\rVert_{H^{r^{\prime},\infty}(\Omega)}\left\lVert v\right\rVert_{H^{t}({\mathbb{R}}^{n})}\leq 2C\epsilon

and hence

(32) |f2,uv|\displaystyle\left\lvert\left\langle f_{2},uv\right\rangle\right\rvert f2v(Hr(Ω))uHr(Ω)f2vHr(Ω)uHr(n)\displaystyle\leq\left\lVert f_{2}v\right\rVert_{(H^{-r^{\prime}}(\Omega))^{*}}\left\lVert u\right\rVert_{H^{-r^{\prime}}(\Omega)}\leq\left\lVert f_{2}v\right\rVert_{H^{r^{\prime}}(\Omega)}\left\lVert u\right\rVert_{H^{-r}({\mathbb{R}}^{n})}
(33) Jr(Fv)L2(n)2Cϵ.\displaystyle\leq\left\lVert J^{r^{\prime}}(Fv)\right\rVert_{L^{2}({\mathbb{R}}^{n})}\leq 2C\epsilon.

This shows that fM0(HrHt)f\in M_{0}(H^{-r}\rightarrow H^{-t}).

iii Let fH~r(Ω)f\in\widetilde{H}^{r^{\prime}}(\Omega). Now f=f1+f2f=f_{1}+f_{2} where f1Cc(Ω)f_{1}\in C^{\infty}_{c}(\Omega) and f2Hr(n)ϵ\left\lVert f_{2}\right\rVert_{H^{r^{\prime}}({\mathbb{R}}^{n})}\leq\epsilon. Now [3, Theorem 7.3] implies the continuity of the multiplication Hr(n)×Ht(n)Ht(n)H^{r^{\prime}}({\mathbb{R}}^{n})\times H^{t}({\mathbb{R}}^{n})\hookrightarrow H^{t}({\mathbb{R}}^{n}) when rtr^{\prime}\geq t and r>n/2r^{\prime}>n/2. We obtain

(34) |f2,uv|f2vHt(n)uHt(n)\displaystyle\left\lvert\left\langle f_{2},uv\right\rangle\right\rvert\leq\left\lVert f_{2}v\right\rVert_{H^{t}({\mathbb{R}}^{n})}\left\lVert u\right\rVert_{H^{-t}({\mathbb{R}}^{n})} Cf2Hr(n)vHt(n)uHr(n)Cϵ.\displaystyle\leq C\left\lVert f_{2}\right\rVert_{H^{r^{\prime}}({\mathbb{R}}^{n})}\left\lVert v\right\rVert_{H^{t}({\mathbb{R}}^{n})}\left\lVert u\right\rVert_{H^{-r}({\mathbb{R}}^{n})}\leq C\epsilon.

Hence fM0(HrHt)f\in M_{0}(H^{-r}\rightarrow H^{-t}). If Ω\Omega is a Lipschitz domain, then HΩ¯r(n)=H~r(Ω)H^{r^{\prime}}_{\overline{\Omega}}({\mathbb{R}}^{n})=\widetilde{H}^{r^{\prime}}(\Omega). If in addition r{12,32,52,}r^{\prime}\notin\{\frac{1}{2},\frac{3}{2},\frac{5}{2},\dotso\}, we also have H0r(Ω)=H~r(Ω)H^{r^{\prime}}_{0}(\Omega)=\widetilde{H}^{r^{\prime}}(\Omega). ∎

Proof of proposition 1.3.

Throughout the proof we assume that u,vCc(n)u,v\in C_{c}^{\infty}({\mathbb{R}}^{n}) such that uL2(n)=vHt(n)=1\left\lVert u\right\rVert_{L^{2}({\mathbb{R}}^{n})}=\left\lVert v\right\rVert_{H^{t}({\mathbb{R}}^{n})}=1.

i Let fLp(Ω)f\in L^{p}(\Omega). By density of Cc(Ω)C^{\infty}_{c}(\Omega) in Lp(Ω)L^{p}(\Omega) we have f=f1+f2f=f_{1}+f_{2} where f1Cc(Ω)f_{1}\in C^{\infty}_{c}(\Omega) and f~2Lp(n)ϵ\left\lVert\widetilde{f}_{2}\right\rVert_{L^{p}({\mathbb{R}}^{n})}\leq\epsilon where f~2\widetilde{f}_{2} is the zero extension of f2Lp(Ω)f_{2}\in L^{p}(\Omega). The assumptions on pp imply the continuity of the multiplication Lp(n)×Ht(n)L2(n)L^{p}({\mathbb{R}}^{n})\times H^{t}({\mathbb{R}}^{n})\hookrightarrow L^{2}({\mathbb{R}}^{n}) ([3, Theorem 7.3]) and we have

(35) |f~2,uv|f~2vL2(n)uL2(n)\displaystyle\left\lvert\left\langle\widetilde{f}_{2},uv\right\rangle\right\rvert\leq\left\lVert\widetilde{f}_{2}v\right\rVert_{L^{2}({\mathbb{R}}^{n})}\left\lVert u\right\rVert_{L^{2}({\mathbb{R}}^{n})} Cf~2Lp(n)vHt(n)Cϵ.\displaystyle\leq C\left\lVert\widetilde{f}_{2}\right\rVert_{L^{p}({\mathbb{R}}^{n})}\left\lVert v\right\rVert_{H^{t}({\mathbb{R}}^{n})}\leq C\epsilon.

This gives that fM0(H0Ht)f\in M_{0}(H^{0}\rightarrow H^{-t}). If Ω\Omega is bounded, we have L(Ω)Lp(Ω)L^{\infty}(\Omega)\hookrightarrow L^{p}(\Omega) for all 1p<1\leq p<\infty, giving the second claim.

ii Let fH~r(Ω)f\in\widetilde{H}^{r}(\Omega). Now we have f=f1+f2f=f_{1}+f_{2} where f1Cc(Ω)f_{1}\in C^{\infty}_{c}(\Omega) and f2Hr(n)ϵ\left\lVert f_{2}\right\rVert_{H^{r}({\mathbb{R}}^{n})}\leq\epsilon. The assumptions on rr imply that the multiplication Hr(n)×Ht(n)L2(n)H^{r}({\mathbb{R}}^{n})\times H^{t}({\mathbb{R}}^{n})\hookrightarrow L^{2}({\mathbb{R}}^{n}) is continuous ([3, Theorem 7.3]). We obtain

(36) |f2,uv|f2vL2(n)uL2(n)\displaystyle\left\lvert\left\langle f_{2},uv\right\rangle\right\rvert\leq\left\lVert f_{2}v\right\rVert_{L^{2}({\mathbb{R}}^{n})}\left\lVert u\right\rVert_{L^{2}({\mathbb{R}}^{n})} Cf2Hr(n)vHt(n)Cϵ\displaystyle\leq C\left\lVert f_{2}\right\rVert_{H^{r}({\mathbb{R}}^{n})}\left\lVert v\right\rVert_{H^{t}({\mathbb{R}}^{n})}\leq C\epsilon

and therefore fM0(H0Ht)f\in M_{0}(H^{0}\rightarrow H^{-t}). The claims for HΩ¯r(n)H^{r}_{\overline{\Omega}}({\mathbb{R}}^{n}) and H0r(Ω)H^{r}_{0}(\Omega) follow as in the proof of part iii of lemma 1.2 from the usual identifications for Lipschtiz domains. ∎

3. Main theorem for singular coefficients

In this section, to shorten the notation, we will write Hs\left\lVert\cdot\right\rVert_{H^{s}}, L2\left\lVert\cdot\right\rVert_{L^{2}} and so on for the global norms in n{\mathbb{R}}^{n} when the base set is not written explicitly.

3.1. Well-posedness of the forward problem

Consider the problem

(37) (Δ)su+|α|maα(Dαu)\displaystyle(-\Delta)^{s}u+\sum_{|\alpha|\leq m}a_{\alpha}(D^{\alpha}u) =FinΩ,\displaystyle=F\quad\mbox{in}\;\;\Omega,
(38) u\displaystyle u =finΩe\displaystyle=f\quad\mbox{in}\;\;\Omega_{e}

and the corresponding adjoint-problem

(39) (Δ)su+|α|m(1)|α|Dα(aαu)\displaystyle(-\Delta)^{s}u^{*}+\sum_{|\alpha|\leq m}(-1)^{|\alpha|}D^{\alpha}(a_{\alpha}u^{*}) =FinΩ,\displaystyle=F^{*}\quad\mbox{in}\;\;\Omega,
(40) u\displaystyle u^{*} =finΩe.\displaystyle=f^{*}\quad\mbox{in}\;\;\Omega_{e}.

Note that if u,uHs(n)u,u^{*}\in H^{s}({\mathbb{R}}^{n}) and aαM(Hs|α|Hs)=M(HsH|α|s)a_{\alpha}\in M(H^{s-\left\lvert\alpha\right\rvert}\rightarrow H^{-s})=M(H^{s}\rightarrow H^{\left\lvert\alpha\right\rvert-s}), then aα(Dαu)Hs(n)a_{\alpha}(D^{\alpha}u)\in H^{-s}({\mathbb{R}}^{n}) and Dα(aαu)Hs(n)D^{\alpha}(a_{\alpha}u^{*})\in H^{-s}({\mathbb{R}}^{n}) matching with (Δ)su,(Δ)suHs(n)(-\Delta)^{s}u,(-\Delta)^{s}u^{*}\in H^{-s}({\mathbb{R}}^{n}).

The problems (37) and (39) are associated to the bilinear forms

(41) BP(v,w):=(Δ)s/2v,(Δ)s/2w+|α|maα,(Dαv)w\displaystyle B_{P}(v,w):=\langle(-\Delta)^{s/2}v,(-\Delta)^{s/2}w\rangle+\sum_{|\alpha|\leq m}\langle a_{\alpha},(D^{\alpha}v)w\rangle

and

(42) BP(v,w):=(Δ)s/2v,(Δ)s/2w+|α|maα,v(Dαw),\displaystyle B^{*}_{P}(v,w):=\langle(-\Delta)^{s/2}v,(-\Delta)^{s/2}w\rangle+\sum_{|\alpha|\leq m}\langle a_{\alpha},v(D^{\alpha}w)\rangle,

defined on v,wCc(n)v,w\in C^{\infty}_{c}({\mathbb{R}}^{n}).

Remark 3.1.

Observe that BPB_{P} is not symmetric, which motivates the introduction of the bilinear form BPB_{P}^{*}. Moreover, one sees by simple inspection that BP(v,w)=BP(w,v)B_{P}(v,w)=B_{P}^{*}(w,v) for all v,wCc(n)v,w\in C^{\infty}_{c}({\mathbb{R}}^{n}). This identity holds for v,wHs(n)v,w\in H^{s}({\mathbb{R}}^{n}) as well by density, thanks to the following boundedness lemma.

Lemma 3.2 (Boundedness of the bilinear forms).

Let s+s\in\mathbb{R}^{+}\setminus\mathbb{Z} and mm\in\mathbb{N} such that 2sm2s\geq m, and let aαM(Hs|α|Hs)a_{\alpha}\in M(H^{s-|\alpha|}\rightarrow H^{-s}). Then BPB_{P} and BPB_{P}^{*} extend as bounded bilinear forms on Hs(n)×Hs(n)H^{s}(\mathbb{R}^{n})\times H^{s}(\mathbb{R}^{n}).

Proof.

We only prove the boundedness of BPB_{P}, as for BPB_{P}^{*} one can proceed in the same way. The proof is a simple calculation following from inequality (22). Let u,vCc(n)u,v\in C_{c}^{\infty}({\mathbb{R}}^{n}). We can then estimate that

(43) |BP(v,w)|\displaystyle|B_{P}(v,w)| |(Δ)s/2v,(Δ)s/2w|+|α|m|aα,(Dαv)w|\displaystyle\leq|\langle(-\Delta)^{s/2}v,(-\Delta)^{s/2}w\rangle|+\sum_{|\alpha|\leq m}|\langle a_{\alpha},(D^{\alpha}v)w\rangle|
(44) wHs(n)vHs(n)+|α|maαs|α|,sDαvHs|α|(n)wHs(n)\displaystyle\leq\|w\|_{H^{s}(\mathbb{R}^{n})}\|v\|_{H^{s}(\mathbb{R}^{n})}+\sum_{|\alpha|\leq m}\|a_{\alpha}\|_{s-|\alpha|,-s}\|D^{\alpha}v\|_{H^{s-|\alpha|}(\mathbb{R}^{n})}\|w\|_{H^{s}(\mathbb{R}^{n})}
(45) (1+|α|maαs|α|,s)wHs(n)vHs(n).\displaystyle\leq\left(1+\sum_{|\alpha|\leq m}\|a_{\alpha}\|_{s-|\alpha|,-s}\right)\|w\|_{H^{s}(\mathbb{R}^{n})}\|v\|_{H^{s}(\mathbb{R}^{n})}.

Now the claim follows from the density of Cc(n)C_{c}^{\infty}({\mathbb{R}}^{n}) in Hs(n)H^{s}({\mathbb{R}}^{n}). ∎

Next we shall define the concept of weak solution to problems (37) and (39):

Definition 3.3 (Weak solutions).

Let f,fHs(n)f,f^{*}\in H^{s}(\mathbb{R}^{n}) and F,F(H~s(Ω))F,F^{*}\in(\widetilde{H}^{s}(\Omega))^{*}. We say that uHs(n)u\in H^{s}(\mathbb{R}^{n}) is a weak solution to (37) when ufH~s(Ω)u-f\in\widetilde{H}^{s}(\Omega) and BP(u,v)=F(v)B_{P}(u,v)=F(v) for all vH~s(Ω)v\in\widetilde{H}^{s}(\Omega). Similarly, we say that uHs(n)u^{*}\in H^{s}(\mathbb{R}^{n}) is a weak solution to (39) when ufH~s(Ω)u^{*}-f^{*}\in\widetilde{H}^{s}(\Omega) and BP(u,v)=F(v)B^{*}_{P}(u^{*},v)=F^{*}(v) for all vH~s(Ω)v\in\widetilde{H}^{s}(\Omega).

In order to prove the existence and uniqueness of weak solutions, we use the following form of Young’s inequality, which holds for all a,b,η+a,b,\eta\in\mathbb{R}^{+} and p,q(1,)p,q\in(1,\infty) such that 1/p+1/q=11/p+1/q=1:

(46) ab(qη)p/qpap+ηbq.\displaystyle ab\leq\frac{(q\eta)^{-p/q}}{p}a^{p}+\eta b^{q}.

The validity of (46) is easily proved by choosing a1=a(qη)1/qa_{1}=a(q\eta)^{-1/q} and b1=b(qη)1/qb_{1}=b(q\eta)^{1/q} in Young’s inequality a1b1a1p/p+b1q/qa_{1}b_{1}\leq a_{1}^{p}/p+b_{1}^{q}/q.

Lemma 3.4 (Well-posedness).

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded open set. Let s+s\in\mathbb{R}^{+}\setminus\mathbb{Z} and mm\in\mathbb{N} be such that 2s>m2s>m, and let aαM0(Hs|α|Hs)a_{\alpha}\in M_{0}(H^{s-|\alpha|}\rightarrow H^{-s}). There exist a real number μ>0\mu>0 and a countable set Σ(μ,)\Sigma\subset(-\mu,\infty) of eigenvalues λ1λ2\lambda_{1}\leq\lambda_{2}\leq...\rightarrow\infty such that if λΣ\lambda\in\mathbb{R}\setminus\Sigma, for any fHs(n)f\in H^{s}(\mathbb{R}^{n}) and F(H~s(Ω))F\in(\widetilde{H}^{s}(\Omega))^{*} there exists a unique uHs(n)u\in H^{s}(\mathbb{R}^{n}) such that ufH~s(Ω)u-f\in\widetilde{H}^{s}(\Omega) and

BP(u,v)λu,v=F(v)for allvH~s(Ω).B_{P}(u,v)-\lambda\langle u,v\rangle=F(v)\quad\mbox{for all}\quad v\in\widetilde{H}^{s}(\Omega).

One has the estimate

uHs(n)C(fHs(n)+F(H~s(Ω))).\|u\|_{H^{s}(\mathbb{R}^{n})}\leq C\left(\|f\|_{H^{s}(\mathbb{R}^{n})}+\|F\|_{(\widetilde{H}^{s}(\Omega))^{*}}\right).

The function uu is also the unique uHs(n)u\in H^{s}(\mathbb{R}^{n}) satisfying

rΩ((Δ)s+|α|maαDαλ)u=Fr_{\Omega}\left((-\Delta)^{s}+\sum_{|\alpha|\leq m}a_{\alpha}D^{\alpha}-\lambda\right)u=F

in the sense of distributions in Ω\Omega and ufH~s(Ω)u-f\in\widetilde{H}^{s}(\Omega). Moreover, if (63) holds then 0Σ0\notin\Sigma.

Proof.

Let u~:=uf\tilde{u}:=u-f. The above problem is reduced to finding a unique u~H~s(Ω)\tilde{u}\in\widetilde{H}^{s}(\Omega) such that BP(u~,v)λu~,v=F~(v)B_{P}(\tilde{u},v)-\lambda\langle\tilde{u},v\rangle=\tilde{F}(v), where F~:=FBP(f,)+λf,\tilde{F}:=F-B_{P}(f,\cdot)+\lambda\langle f,\cdot\rangle. Observe that the modified functional F~\tilde{F} belongs to (H~s(Ω))(\widetilde{H}^{s}(\Omega))^{*} as well, since by lemma 3.2 we have for all vH~s(Ω)v\in\widetilde{H}^{s}(\Omega)

|F~(v)||F(v)|+|BP(f,v)|+|λ||f,v|(F(H~s(Ω))+(C+|λ|)fHs(n))vHs(n).|\tilde{F}(v)|\leq|F(v)|+|B_{P}(f,v)|+|\lambda|\,|\langle f,v\rangle|\leq(\left\lVert F\right\rVert_{(\widetilde{H}^{s}(\Omega))^{*}}+(C+|\lambda|)\|f\|_{H^{s}(\mathbb{R}^{n})})\|v\|_{H^{s}(\mathbb{R}^{n})}.

Since aαM0(Hs|α|Hs)a_{\alpha}\in M_{0}(H^{s-|\alpha|}\rightarrow H^{-s}), for any ϵ>0\epsilon>0 we can write aα=aα,1+aα,2a_{\alpha}=a_{\alpha,1}+a_{\alpha,2}, where aα,1Cc(n)M(Hs|α|Hs)a_{\alpha,1}\in C^{\infty}_{c}(\mathbb{R}^{n})\cap M(H^{s-\left\lvert\alpha\right\rvert}\to H^{-s}) and aα,2s|α|,s<ϵ\|a_{\alpha,2}\|_{s-|\alpha|,-s}<\epsilon. Thus by formula (22), the continuity of the multiplication Hr(n)×Hs(n)Hs(n)H^{r}({\mathbb{R}}^{n})\times H^{s}({\mathbb{R}}^{n})\hookrightarrow H^{s}({\mathbb{R}}^{n}) for large enough rr\in{\mathbb{R}} (see [3, Theorem 7.3]) and the fact that aα,1Cc(n)Hr(n)a_{\alpha,1}\in C^{\infty}_{c}({\mathbb{R}}^{n})\subset H^{r}({\mathbb{R}}^{n}) for all rr\in{\mathbb{R}} we obtain

(47) |aα,(Dαv)w|\displaystyle|\langle a_{\alpha},(D^{\alpha}v)w\rangle| |aα,1,(Dαv)w|+|aα,2,(Dαv)w|\displaystyle\leq|\langle a_{\alpha,1},(D^{\alpha}v)w\rangle|+|\langle a_{\alpha,2},(D^{\alpha}v)w\rangle|
(48) aα,1Hr(n)DαvHs(n)wHs(n)+aα,2s|α|,sDαvHs|α|(n)wHs(n)\displaystyle\leq\|a_{\alpha,1}\|_{H^{r}({\mathbb{R}}^{n})}\|D^{\alpha}v\|_{H^{-s}({\mathbb{R}}^{n})}\|w\|_{H^{s}({\mathbb{R}}^{n})}+\|a_{\alpha,2}\|_{s-|\alpha|,-s}\|D^{\alpha}v\|_{H^{s-|\alpha|}({\mathbb{R}}^{n})}\|w\|_{H^{s}({\mathbb{R}}^{n})}
(49) cwHs(n)(aα,1Hr(n)vH|α|s(n)+ϵvHs(n))\displaystyle\leq c\|w\|_{H^{s}({\mathbb{R}}^{n})}\left(\|a_{\alpha,1}\|_{H^{r}({\mathbb{R}}^{n})}\|v\|_{H^{|\alpha|-s}({\mathbb{R}}^{n})}+\epsilon\|v\|_{H^{s}({\mathbb{R}}^{n})}\right)

where rr\in{\mathbb{R}} is large enough (r>max{s,n/2}r>\max\{s,n/2\} is sufficient). If |α|<s|\alpha|<s, from formulas (47) and (46) with p=q=2p=q=2 we get directly

(50) |aα,(Dαv)v|\displaystyle|\langle a_{\alpha},(D^{\alpha}v)v\rangle| C(vHs(n)vL2(n)+ϵvHs(n)2)\displaystyle\leq C\left(\|v\|_{H^{s}({\mathbb{R}}^{n})}\|v\|_{L^{2}({\mathbb{R}}^{n})}+\epsilon\|v\|_{H^{s}({\mathbb{R}}^{n})}^{2}\right)
(51) C(ϵ1vL2(n)2+ϵvHs(n)2)\displaystyle\leq C(\epsilon^{-1}\|v\|_{L^{2}({\mathbb{R}}^{n})}^{2}+\epsilon\|v\|_{H^{s}({\mathbb{R}}^{n})}^{2})

for a constant CC independent of v,w,ϵv,w,\epsilon. If instead |α|>s|\alpha|>s (observe that we can not have |α|=s|\alpha|=s, because ss can not be an integer), we use the interpolation inequality

vH|α|s(n)CvL2(n)1(|α|s)/svHs(n)(|α|s)/s=CvL2(n)2|α|/svHs(n)|α|/s1\|v\|_{H^{|\alpha|-s}(\mathbb{R}^{n})}\leq C\|v\|_{L^{2}(\mathbb{R}^{n})}^{1-(|\alpha|-s)/s}\|v\|_{H^{s}(\mathbb{R}^{n})}^{(|\alpha|-s)/s}=C\|v\|_{L^{2}(\mathbb{R}^{n})}^{2-|\alpha|/s}\|v\|_{H^{s}(\mathbb{R}^{n})}^{|\alpha|/s-1}

in order to get

(52) |aα,(Dαv)w|\displaystyle|\langle a_{\alpha},(D^{\alpha}v)w\rangle| CwHs(n)(vL2(n)2|α|/svHs(n)|α|/s1+ϵvHs(n)).\displaystyle\leq C\|w\|_{H^{s}({\mathbb{R}}^{n})}\left(\|v\|_{L^{2}({\mathbb{R}}^{n})}^{2-|\alpha|/s}\|v\|_{H^{s}({\mathbb{R}}^{n})}^{|\alpha|/s-1}+\epsilon\|v\|_{H^{s}({\mathbb{R}}^{n})}\right).

Then by formula (46) with

a=vL2(n)2|α|/s,b=vHs(n)|α|/s1,p=s2s|α|,q=s|α|s,η=ϵa=\|v\|_{L^{2}({\mathbb{R}}^{n})}^{2-|\alpha|/s},\quad b=\|v\|_{H^{s}({\mathbb{R}}^{n})}^{|\alpha|/s-1},\quad p=\frac{s}{2s-|\alpha|},\quad q=\frac{s}{|\alpha|-s},\quad\eta=\epsilon

we obtain

(53) |aα,(Dαv)w|\displaystyle|\langle a_{\alpha},(D^{\alpha}v)w\rangle| CwHs(n)(ϵs|α|2s|α|vL2(n)+ϵvHs(n))\displaystyle\leq C\|w\|_{H^{s}({\mathbb{R}}^{n})}\left(\epsilon^{\frac{s-|\alpha|}{2s-|\alpha|}}\|v\|_{L^{2}({\mathbb{R}}^{n})}+\epsilon\|v\|_{H^{s}({\mathbb{R}}^{n})}\right)

for a constant CC independent of v,w,ϵv,w,\epsilon. Now we use formula (46) again, but this time we choose

a=vL2(n),b=vHs(n),q=p=2,η=ϵs/(2s|α|).a=\|v\|_{L^{2}({\mathbb{R}}^{n})},\quad b=\|v\|_{H^{s}({\mathbb{R}}^{n})},\quad q=p=2,\quad\eta=\epsilon^{s/(2s-|\alpha|)}.

This leads to

(54) |aα,(Dαv)v|\displaystyle|\langle a_{\alpha},(D^{\alpha}v)v\rangle| C(ϵs|α|2s|α|vL2(n)vHs(n)+ϵvHs(n)2)\displaystyle\leq C\left(\epsilon^{\frac{s-|\alpha|}{2s-|\alpha|}}\|v\|_{L^{2}({\mathbb{R}}^{n})}\|v\|_{H^{s}({\mathbb{R}}^{n})}+\epsilon\|v\|_{H^{s}({\mathbb{R}}^{n})}^{2}\right)
(55) C(ϵ|α|2s|α|vL2(n)2+2ϵvHs(n)2)\displaystyle\leq C\left(\epsilon^{\frac{-|\alpha|}{2s-|\alpha|}}\|v\|_{L^{2}({\mathbb{R}}^{n})}^{2}+2\epsilon\|v\|_{H^{s}({\mathbb{R}}^{n})}^{2}\right)
(56) C(ϵ|α|2s|α|vL2(n)2+ϵvHs(n)2)\displaystyle\leq C\left(\epsilon^{\frac{-|\alpha|}{2s-|\alpha|}}\|v\|_{L^{2}({\mathbb{R}}^{n})}^{2}+\epsilon\|v\|_{H^{s}({\mathbb{R}}^{n})}^{2}\right)
(57) C(ϵm2smvL2(n)2+ϵvHs(n)2)\displaystyle\leq C^{\prime}\left(\epsilon^{\frac{-m}{2s-m}}\|v\|_{L^{2}({\mathbb{R}}^{n})}^{2}+\epsilon\|v\|_{H^{s}({\mathbb{R}}^{n})}^{2}\right)

where C,CC,C^{\prime} are constants changing from line to line. Observe that CC^{\prime} can be taken independent of α\alpha. Eventually, using (50) and (54) we get

(58) BP(v,v)\displaystyle B_{P}(v,v) (Δ)s/2vL2(n)2|α|m|aα,(Dαv)v|\displaystyle\geq\|(-\Delta)^{s/2}v\|^{2}_{L^{2}({\mathbb{R}}^{n})}-\sum_{|\alpha|\leq m}|\langle a_{\alpha},(D^{\alpha}v)v\rangle|
(59) (Δ)s/2vL2(n)2C((ϵm2sm+ϵ1)vL2(n)2+ϵvHs(n)2).\displaystyle\geq\|(-\Delta)^{s/2}v\|^{2}_{L^{2}({\mathbb{R}}^{n})}-C^{\prime}\left((\epsilon^{\frac{-m}{2s-m}}+\epsilon^{-1})\|v\|_{L^{2}({\mathbb{R}}^{n})}^{2}+\epsilon\|v\|_{H^{s}({\mathbb{R}}^{n})}^{2}\right).

By the higher order Poincaré inequality (lemma 2.2) (94) turns into

(60) BP(v,v)\displaystyle B_{P}(v,v) c((Δ)s/2vL2(n)2+vL2(n)2)C((ϵm2sm+ϵ1)vL2(n)2+ϵvHs(n)2)\displaystyle\geq c\left(\|(-\Delta)^{s/2}v\|^{2}_{L^{2}({\mathbb{R}}^{n})}+\|v\|_{L^{2}({\mathbb{R}}^{n})}^{2}\right)-C^{\prime}\left((\epsilon^{\frac{-m}{2s-m}}+\epsilon^{-1})\|v\|_{L^{2}({\mathbb{R}}^{n})}^{2}+\epsilon\|v\|_{H^{s}({\mathbb{R}}^{n})}^{2}\right)
(61) cvHs(n)2C((ϵm2sm+ϵ1)vL2(n)2+ϵvHs(n)2)\displaystyle\geq c\|v\|_{H^{s}({\mathbb{R}}^{n})}^{2}-C^{\prime}\left((\epsilon^{\frac{-m}{2s-m}}+\epsilon^{-1})\|v\|_{L^{2}({\mathbb{R}}^{n})}^{2}+\epsilon\|v\|_{H^{s}({\mathbb{R}}^{n})}^{2}\right)

for some constant c=c(Ω,n,s)c=c(\Omega,n,s) changing from line to line. For ϵ\epsilon small enough, this eventually gives the coercivity estimate

(62) BP(v,v)c0vHs(n)2μvL2(n)2B_{P}(v,v)\geq c_{0}\|v\|_{H^{s}(\mathbb{R}^{n})}^{2}-\mu\|v\|_{L^{2}(\mathbb{R}^{n})}^{2}

for some constants c0,μ>0c_{0},\mu>0 independent of vv.

As a consequence of the coercivity estimate, the bilinear form BP(,)+μ,L2(n)B_{P}(\cdot,\cdot)+\mu\langle\cdot,\cdot\rangle_{L^{2}(\mathbb{R}^{n})} satisfies the assumptions of the Lax–Milgram theorem, and there exists a bounded linear operator Gμ:(H~s(Ω))H~s(Ω)G_{\mu}:(\widetilde{H}^{s}(\Omega))^{*}\rightarrow\widetilde{H}^{s}(\Omega) associating each functional in (H~s(Ω))(\widetilde{H}^{s}(\Omega))^{*} to its unique representative in the bilinear form BP(,)+μ,L2(n)B_{P}(\cdot,\cdot)+\mu\langle\cdot,\cdot\rangle_{L^{2}(\mathbb{R}^{n})} on H~s(Ω)\widetilde{H}^{s}(\Omega). Thus u~:=GμF~\tilde{u}:=G_{\mu}\tilde{F} verifies

BP(u~,v)+μu~,vL2(n)=F~(v)for allvH~s(Ω)B_{P}(\tilde{u},v)+\mu\langle\tilde{u},v\rangle_{L^{2}(\mathbb{R}^{n})}=\tilde{F}(v)\quad\mbox{for all}\quad v\in\widetilde{H}^{s}(\Omega)

and it is the required unique solution u~H~s(Ω)\tilde{u}\in\widetilde{H}^{s}(\Omega). Moreover, GμG_{\mu} induces a compact operator G~μ:L2(Ω)L2(Ω)\tilde{G}_{\mu}:L^{2}(\Omega)\rightarrow L^{2}(\Omega) by the compact Sobolev embedding theorem. The remaining claims follow from the spectral theorem of compact operators for G~μ\tilde{G}_{\mu} and from the Fredholm alternative as in [18]. ∎

By the above lemma 3.4, both problems (37) and (39) have a countable set of Dirichlet eigenvalues. Throughout the paper we will assume that the coefficients aαa_{\alpha} are such that 0 is not a Dirichlet eigenvalue for either of the problems. That is, we assume that

(63) {if uHs(n) solves (Δ)su+|α|maαDαu=0 in Ω and u|Ωe=0,then u0.\displaystyle\Bigg{\{}\begin{matrix}\mbox{if $u\in H^{s}(\mathbb{R}^{n})$ solves $(-\Delta)^{s}u+\sum_{|\alpha|\leq m}a_{\alpha}D^{\alpha}u=0$ in $\Omega$ and $u|_{\Omega_{e}}=0$,}\\ \mbox{then $u\equiv 0$}\end{matrix}\Bigg{.}

and

(64) {if uHs(n) solves (Δ)su+|α|m(1)|α|Dα(aαu)=0 in Ω and u|Ωe=0,then u0..\displaystyle\Bigg{\{}\begin{matrix}\mbox{if $u^{*}\in H^{s}(\mathbb{R}^{n})$ solves $(-\Delta)^{s}u^{*}+\sum_{|\alpha|\leq m}(-1)^{|\alpha|}D^{\alpha}(a_{\alpha}u^{*})=0$ in $\Omega$ and $u^{*}|_{\Omega_{e}}=0$,}\\ \mbox{then $u^{*}\equiv 0$.}\end{matrix}\Bigg{.}

With this in mind, we shall define the exterior DN maps associated to the problems (37) and (39). Consider the abstract trace space X:=Hs(n)/H~s(Ω)X:=H^{s}(\mathbb{R}^{n})/\widetilde{H}^{s}(\Omega) equipped with the quotient norm

[f]X:=infϕH~s(Ω)fϕHs(n),fHs(n)\|[f]\|_{X}:=\inf_{\phi\in\widetilde{H}^{s}(\Omega)}\|f-\phi\|_{H^{s}(\mathbb{R}^{n})},\quad f\in H^{s}(\mathbb{R}^{n})

and its dual space XX^{*}.

Definition 3.5.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded open set. Let s+s\in\mathbb{R}^{+}\setminus\mathbb{Z} and mm\in\mathbb{N} such that 2s>m2s>m, and let aαM0(Hs|α|Hs)a_{\alpha}\in M_{0}(H^{s-|\alpha|}\rightarrow H^{-s}). The exterior DN maps ΛP\Lambda_{P} and ΛP\Lambda_{P}^{*} are

ΛP:XXdefined byΛP[f],[g]:=BP(uf,g)\Lambda_{P}:X\rightarrow X^{*}\quad\mbox{defined by}\quad\langle\Lambda_{P}[f],[g]\rangle:=B_{P}(u_{f},g)

and

ΛP:XXdefined byΛP[f],[g]:=BP(uf,g)\Lambda^{*}_{P}:X\rightarrow X^{*}\quad\mbox{defined by}\quad\langle\Lambda^{*}_{P}[f],[g]\rangle:=B^{*}_{P}(u^{*}_{f},g)

where uf,ufu_{f},u^{*}_{f} are the unique solutions to the equations

(Δ)su+|α|maαDαu\displaystyle(-\Delta)^{s}u+\sum_{|\alpha|\leq m}a_{\alpha}D^{\alpha}u =0inΩ,ufH~s(Ω)\displaystyle=0\quad\mbox{in}\;\;\Omega,\quad u-f\in\widetilde{H}^{s}(\Omega)

and

(Δ)su+|α|m(1)|α|Dα(aαu)\displaystyle(-\Delta)^{s}u^{*}+\sum_{|\alpha|\leq m}(-1)^{|\alpha|}D^{\alpha}(a_{\alpha}u^{*}) =0inΩ,ufH~s(Ω)\displaystyle=0\quad\mbox{in}\;\;\Omega,\quad u^{*}-f\in\widetilde{H}^{s}(\Omega)

with f,gHs(n)f,g\in H^{s}(\mathbb{R}^{n}).

The next lemma proves that the exterior DN maps ΛP\Lambda_{P} and ΛP\Lambda_{P}^{*} are well-defined and have some expected properties:

Lemma 3.6 (Exterior DN maps).

The exterior DN maps ΛP\Lambda_{P} and ΛP\Lambda_{P}^{*} are well-defined, linear and continuous. Moreover, the identity ΛP[f],[g]=[f],ΛP[g]\langle\Lambda_{P}[f],[g]\rangle=\langle[f],\Lambda^{*}_{P}[g]\rangle holds.

Proof.

We show well-definedness and continuity only for ΛP\Lambda_{P}, the proof being similar for ΛP\Lambda_{P}^{*}. We note that the required unique solutions exist by lemma 3.4.

If ϕH~s(Ω)\phi\in\widetilde{H}^{s}(\Omega), then uf|Ωe=f=uf+ϕ|Ωeu_{f}|_{\Omega_{e}}=f=u_{f+\phi}|_{\Omega_{e}}, and also ufu_{f}, uf+ϕu_{f+\phi} both solve (Δ)su+Pu=0(-\Delta)^{s}u+Pu=0 in Ω\Omega. By unicity of solutions, we must then have that ufu_{f} and uf+ϕu_{f+\phi} coincide. On the other hand, if ψH~s(Ω)\psi\in\widetilde{H}^{s}(\Omega), then ψ|Ωe=0\psi|_{\Omega_{e}}=0. These two facts imply the well-definedness of ΛP\Lambda_{P}, since

BP(uf+ϕ,g+ψ)=BP(uf,g)+BP(uf,ψ)=BP(uf,g).B_{P}(u_{f+\phi},g+\psi)=B_{P}(u_{f},g)+B_{P}(u_{f},\psi)=B_{P}(u_{f},g).

The continuity of ΛP\Lambda_{P} is an easy consequence of lemma 3.2 and the estimate in lemma 3.4. If f,gHs(n)f,g\in H^{s}(\mathbb{R}^{n}) and ϕ,ψH~s(Ω)\phi,\psi\in\widetilde{H}^{s}(\Omega), then

|ΛP[f],[g]|\displaystyle|\langle\Lambda_{P}[f],[g]\rangle| =|BP(ufϕ,gψ)|CufϕHsgψHsCfϕHsgψHs.\displaystyle=|B_{P}(u_{f-\phi},g-\psi)|\leq C\|u_{f-\phi}\|_{H^{s}}\|g-\psi\|_{H^{s}}\leq C\|f-\phi\|_{H^{s}}\|g-\psi\|_{H^{s}}.

By taking the infimum on both sides with respect to ϕ\phi and ψ\psi, we end up with

|ΛP[f],[g]|CinfϕH~s(Ω)fϕHsinfψH~s(Ω)gψHs=C[f]X[g]X.|\langle\Lambda_{P}[f],[g]\rangle|\leq C\inf_{\phi\in\widetilde{H}^{s}(\Omega)}\|f-\phi\|_{H^{s}}\inf_{\psi\in\widetilde{H}^{s}(\Omega)}\|g-\psi\|_{H^{s}}=C\|[f]\|_{X}\|[g]\|_{X}.

The well-posedness result proved above implies that for all f,gHs(n)f,g\in H^{s}(\mathbb{R}^{n}) we have ΛP[f],[g]=BP(uf,eg)\langle\Lambda_{P}[f],[g]\rangle=B_{P}(u_{f},e_{g}), where ege_{g} is a generic extension of g|Ωeg|_{\Omega_{e}} from Ωe\Omega_{e} to n\mathbb{R}^{n}. In particular, ΛP[f],[g]=BP(uf,ug)\langle\Lambda_{P}[f],[g]\rangle=B_{P}(u_{f},u_{g}^{*}). By lemma 3.2 this leads to

ΛP[f],[g]=BP(uf,ug)=BP(ug,uf)=ΛP[g],[f],\langle\Lambda_{P}[f],[g]\rangle=B_{P}(u_{f},u_{g}^{*})=B_{P}^{*}(u_{g}^{*},u_{f})=\langle\Lambda^{*}_{P}[g],[f]\rangle,

which conlcudes the proof. ∎

Remark 3.7.

We should observe at this point that a priori ΛP\Lambda_{P}^{*} has no reason to be the adjoint of ΛP\Lambda_{P}, as the symbols would suggest. However, the identity we proved in lemma 3.6 shows that this is in fact true, and thus there is no abuse of notation.

3.2. Proof of injectivity

The proof of injectivity is based on an Alessandrini identity and the Runge approximation property for our operator, following the scheme developed in [18].

Lemma 3.8 (Alessandrini identity).

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded open set. Let s+s\in\mathbb{R}^{+}\setminus\mathbb{Z} and mm\in\mathbb{N} such that 2s>m2s>m. For j=1,2j=1,2, let aj,αM0(Hs|α|Hs)a_{j,\alpha}\in M_{0}(H^{s-|\alpha|}\rightarrow H^{-s}). For any f1,f2Hs(n)f_{1},f_{2}\in H^{s}(\mathbb{R}^{n}), let u1,u2Hs(n)u_{1},u_{2}^{*}\in H^{s}(\mathbb{R}^{n}) respectively solve

(Δ)su1+|α|ma1,αDαu1\displaystyle(-\Delta)^{s}u_{1}+\sum_{|\alpha|\leq m}a_{1,\alpha}D^{\alpha}u_{1} =0inΩ,u1f1H~s(Ω)\displaystyle=0\quad\mbox{in}\;\;\Omega,\quad u_{1}-f_{1}\in\widetilde{H}^{s}(\Omega)

and

(Δ)su2+|α|m(1)|α|Dα(a2,αu2)\displaystyle(-\Delta)^{s}u_{2}^{*}+\sum_{|\alpha|\leq m}(-1)^{|\alpha|}D^{\alpha}(a_{2,\alpha}u_{2}^{*}) =0inΩ,u2f2H~s(Ω).\displaystyle=0\quad\mbox{in}\;\;\Omega,\quad u_{2}^{*}-f_{2}\in\widetilde{H}^{s}(\Omega).

Then we have the integral identity

(ΛP1ΛP2)[f1],[f2]=|α|m(a1,αa2,α),(Dαu1)u2.\langle(\Lambda_{P_{1}}-\Lambda_{P_{2}})[f_{1}],[f_{2}]\rangle=\sum_{|\alpha|\leq m}\langle(a_{1,\alpha}-a_{2,\alpha}),(D^{\alpha}u_{1})u_{2}^{*}\rangle.
Proof.

The proof is a simple computation following from lemma 3.6:

(ΛP1ΛP2)[f1],[f2]\displaystyle\langle(\Lambda_{P_{1}}-\Lambda_{P_{2}})[f_{1}],[f_{2}]\rangle =ΛP1[f1],[f2]ΛP2[f1],[f2]=ΛP1[f1],[f2][f1],ΛP2[f2]\displaystyle=\langle\Lambda_{P_{1}}[f_{1}],[f_{2}]\rangle-\langle\Lambda_{P_{2}}[f_{1}],[f_{2}]\rangle=\langle\Lambda_{P_{1}}[f_{1}],[f_{2}]\rangle-\langle[f_{1}],\Lambda_{P_{2}}^{*}[f_{2}]\rangle
=BP1(u1,u2)BP2(u2,u1)=|α|m(a1,αa2,α),(Dαu1)u2.\displaystyle=B_{P_{1}}(u_{1},u_{2}^{*})-B_{P_{2}}^{*}(u_{2}^{*},u_{1})=\sum_{|\alpha|\leq m}\langle(a_{1,\alpha}-a_{2,\alpha}),(D^{\alpha}u_{1})u_{2}^{*}\rangle.\qed
Lemma 3.9 (Runge approximation property).

Let Ω,Wn\Omega,W\subset\mathbb{R}^{n} respectively be a bounded open set and a non-empty open set such that W¯Ω¯=\overline{W}\cap\overline{\Omega}=\emptyset. Let s+s\in\mathbb{R}^{+}\setminus\mathbb{Z} and mm\in\mathbb{N} be such that 2s>m2s>m, and let aαM0(Hs|α|Hs)a_{\alpha}\in M_{0}(H^{s-|\alpha|}\rightarrow H^{-s}). Moreover, let :={uff:fCc(W)}H~s(Ω)\mathcal{R}:=\{\,u_{f}-f:f\in C^{\infty}_{c}(W)\,\}\subset\widetilde{H}^{s}(\Omega) where ufu_{f} solves

(Δ)suf+|α|maαDαuf\displaystyle(-\Delta)^{s}u_{f}+\sum_{|\alpha|\leq m}a_{\alpha}D^{\alpha}u_{f} =0inΩ,uffH~s(Ω)\displaystyle=0\quad\mbox{in}\;\;\Omega,\quad u_{f}-f\in\widetilde{H}^{s}(\Omega)

and :={uff:fCc(W)}H~s(Ω)\mathcal{R}^{*}:=\{\,u^{*}_{f}-f:f\in C^{\infty}_{c}(W)\,\}\subset\widetilde{H}^{s}(\Omega) where ufu^{*}_{f} solves

(Δ)suf+|α|m(1)|α|Dα(aαuf)\displaystyle(-\Delta)^{s}u_{f}^{*}+\sum_{|\alpha|\leq m}(-1)^{|\alpha|}D^{\alpha}(a_{\alpha}u_{f}^{*}) =0inΩ,uffH~s(Ω).\displaystyle=0\quad\mbox{in}\;\;\Omega,\quad u_{f}^{*}-f\in\widetilde{H}^{s}(\Omega).

Then \mathcal{R} and \mathcal{R}^{*} are dense in H~s(Ω)\widetilde{H}^{s}(\Omega).

Proof.

The proofs of the two statements are similar, so we show only the density of \mathcal{R} in H~s(Ω)\widetilde{H}^{s}(\Omega). By the Hahn-Banach theorem, it is enough to prove that any functional FF acting on H~s(Ω)\widetilde{H}^{s}(\Omega) that vanishes on \mathcal{R} must be identically 0. Thus, let F(H~s(Ω))F\in(\widetilde{H}^{s}(\Omega))^{*} and assume F(uff)=0F(u_{f}-f)=0 for all fCc(W)f\in C^{\infty}_{c}(W). Let ϕ\phi be the unique solution of

(65) (Δ)sϕ+|α|m(1)|α|Dα(aαϕ)=FinΩ,ϕH~s(Ω).(-\Delta)^{s}\phi+\sum_{|\alpha|\leq m}(-1)^{|\alpha|}D^{\alpha}(a_{\alpha}\phi)=-F\quad\mbox{in}\;\;\Omega,\quad\phi\in\widetilde{H}^{s}(\Omega).

In other words, ϕ\phi is the unique function in H~s(Ω)\widetilde{H}^{s}(\Omega) such that BP(ϕ,w)=F(w)B_{P}^{*}(\phi,w)=-F(w) for all wH~s(Ω)w\in\widetilde{H}^{s}(\Omega). Then we can compute

(66) 0\displaystyle 0 =F(uff)=BP(ϕ,uff)=BP(ϕ,f)\displaystyle=F(u_{f}-f)=-B_{P}^{*}(\phi,u_{f}-f)=B_{P}^{*}(\phi,f)
(67) =(Δ)s/2f,(Δ)s/2ϕ+|α|maα,Dαfϕ\displaystyle=\langle(-\Delta)^{s/2}f,(-\Delta)^{s/2}\phi\rangle+\sum_{|\alpha|\leq m}\langle a_{\alpha},D^{\alpha}f\phi\rangle
(68) =f,(Δ)sϕ.\displaystyle=\langle f,(-\Delta)^{s}\phi\rangle.

On the first line of (66) we used that ϕH~s(Ω)\phi\in\widetilde{H}^{s}(\Omega) and ufu_{f} solves the equation in Ω\Omega, and on the last line we used the support condition for ff. By the arbitrariety of fCc(W)f\in C^{\infty}_{c}(W) we have obtained that (Δ)sϕ=0(-\Delta)^{s}\phi=0 in WW, and on the same set we also have ϕ=0\phi=0. Using the unique continuation result for the higher order fractional Laplacian given in lemma 2.1 we deduce ϕ0\phi\equiv 0 on all of n\mathbb{R}^{n}. The vanishing of the functional FF now follows easily from the definition of  ϕ\phi. ∎

Remark 3.10.

We remark that using the same proof one can show that rΩL2(Ω)r_{\Omega}\mathcal{R}\subset L^{2}(\Omega) and rΩL2(Ω)r_{\Omega}\mathcal{R}^{*}\subset L^{2}(\Omega) are dense in L2(Ω)L^{2}(\Omega), where rΩr_{\Omega} is the restriction to Ω\Omega. If FL2(Ω)F\in L^{2}(\Omega), then FF induces an element in (H~s(Ω))(\widetilde{H}^{s}(\Omega))^{*} via the integral F(w):=F,rΩwL2(Ω)F(w):=\left\langle F,r_{\Omega}w\right\rangle_{L^{2}(\Omega)}, where wH~s(Ω)w\in\widetilde{H}^{s}(\Omega). Hence one can choose the solution ϕ\phi in equation (65) with FF as a source term and complete the proof as in equation (66) showing that (rΩ)={0}(r_{\Omega}\mathcal{R})^{\perp}=\{0\} in L2(Ω)L^{2}(\Omega) (similarly (rΩ)={0}(r_{\Omega}\mathcal{R}^{*})^{\perp}=\{0\}).

We are ready to prove the main result of the paper.

Proof of theorem 1.1.

Step 1. Since one can always shrink the sets W1W_{1} and W2W_{2} if necessary, we can assume without loss of generality that W1¯W2¯=\overline{W_{1}}\cap\overline{W_{2}}=\emptyset. Let v1,v2Cc(Ω)v_{1},v_{2}\in C^{\infty}_{c}(\Omega). By the Runge approximation property proved in lemma 3.9 we can find two sequences of functions {fj,k}kCc(Wj)\{f_{j,k}\}_{k\in\mathbb{N}}\subset C^{\infty}_{c}(W_{j}), j=1,2j=1,2, such that

u1,k=f1,k+v1+r1,k,u2,k=f2,k+v2+r2,k\displaystyle u_{1,k}=f_{1,k}+v_{1}+r_{1,k},\quad u^{*}_{2,k}=f_{2,k}+v_{2}+r_{2,k}

where u1,k,u2,kH~s(Ω)u_{1,k},\,u_{2,k}^{*}\in\widetilde{H}^{s}(\Omega) respectively solve

(Δ)su1,k+|α|ma1,αDαu1,k\displaystyle(-\Delta)^{s}u_{1,k}+\sum_{|\alpha|\leq m}a_{1,\alpha}D^{\alpha}u_{1,k} =0inΩ,u1,kf1,kH~s(Ω)\displaystyle=0\quad\mbox{in}\;\;\Omega,\quad u_{1,k}-f_{1,k}\in\widetilde{H}^{s}(\Omega)

and

(Δ)su2,k+|α|m(1)|α|Dα(a2,αu2,k)\displaystyle(-\Delta)^{s}u_{2,k}^{*}+\sum_{|\alpha|\leq m}(-1)^{|\alpha|}D^{\alpha}(a_{2,\alpha}u_{2,k}^{*}) =0inΩ,u2,kf2,kH~s(Ω)\displaystyle=0\quad\mbox{in}\;\;\Omega,\quad u_{2,k}^{*}-f_{2,k}\in\widetilde{H}^{s}(\Omega)

and r1,k,r2,k0r_{1,k},\,r_{2,k}\rightarrow 0 in H~s(Ω)\widetilde{H}^{s}(\Omega) as kk\rightarrow\infty. By the assumption on the exterior DN maps and the Alessandrini identity from lemma 3.8 we have

(69) 0\displaystyle 0 =(ΛP1ΛP2)[f1,k],[f2,k]=|α|m(a1,αa2,α),(Dαu1,k)u2,k.\displaystyle=\langle(\Lambda_{P_{1}}-\Lambda_{P_{2}})[f_{1,k}],[f_{2,k}]\rangle=\sum_{|\alpha|\leq m}\langle(a_{1,\alpha}-a_{2,\alpha}),(D^{\alpha}u_{1,k})u_{2,k}^{*}\rangle.

On the other hand, the support conditions imply that

(70) |α|m(a1,αa2,α),(Dαu1,k)u2,k\displaystyle\sum_{|\alpha|\leq m}\langle(a_{1,\alpha}-a_{2,\alpha}),(D^{\alpha}u_{1,k})u_{2,k}^{*}\rangle =|α|m(a1,αa2,α),(Dα(u1,kf1,k))(u2,kf2,k)\displaystyle=\sum_{|\alpha|\leq m}\langle(a_{1,\alpha}-a_{2,\alpha}),(D^{\alpha}(u_{1,k}-f_{1,k}))(u_{2,k}^{*}-f_{2,k})\rangle
(71) =|α|m(a1,αa2,α),(Dα(v1+r1,k))(v2+r2,k).\displaystyle=\sum_{|\alpha|\leq m}\langle(a_{1,\alpha}-a_{2,\alpha}),(D^{\alpha}(v_{1}+r_{1,k}))(v_{2}+r_{2,k})\rangle.

Thus by taking the limit kk\to\infty and using lemma 3.2, we obtain

(72) |α|m(a1,αa2,α),(Dαv1)v2=0for allv1,v2Cc(Ω)\sum_{|\alpha|\leq m}\langle(a_{1,\alpha}-a_{2,\alpha}),(D^{\alpha}v_{1})v_{2}\rangle=0\quad\mbox{for all}\quad v_{1},v_{2}\in C^{\infty}_{c}(\Omega)

by formula (69).

Step 2. Assume that we have a1,α|Ω=a2,α|Ωa_{1,\alpha}|_{\Omega}=a_{2,\alpha}|_{\Omega} for all α\alpha such that |α|<N|\alpha|<N for some NN\in\mathbb{N}. We show that the equality of the coefficients also holds for α\alpha for which |α|=N|\alpha|=N, and this will prove the theorem by the principle of complete induction.

To this end, consider v2Cc(Ω)v_{2}\in C^{\infty}_{c}(\Omega), and then take v1Cc(Ω)v_{1}\in C^{\infty}_{c}(\Omega) such that v1(x)=xαv_{1}(x)=x^{\alpha} on supp(v2)Ω(v_{2})\Subset\Omega. Recall that since α=(α1,α2,,αn)n\alpha=(\alpha_{1},\alpha_{2},...,\alpha_{n})\in\mathbb{N}^{n} is a multi-index and x=(x1,x2,,xn)nx=(x_{1},x_{2},...,x_{n})\in\mathbb{R}^{n}, the symbol xαx^{\alpha} is intended to mean x1α1x2α2xnαnx_{1}^{\alpha_{1}}x_{2}^{\alpha_{2}}...\,x_{n}^{\alpha_{n}}. With this choice of v1,v2v_{1},v_{2}, equation (72) becomes

(73) 0\displaystyle 0 =|β|m(a1,βa2,β),(Dβv1)v2=N|β|m(a1,βa2,β),(Dβxα)v2\displaystyle=\sum_{|\beta|\leq m}\langle(a_{1,\beta}-a_{2,\beta}),(D^{\beta}v_{1})v_{2}\rangle=\sum_{N\leq|\beta|\leq m}\langle(a_{1,\beta}-a_{2,\beta}),(D^{\beta}x^{\alpha})v_{2}\rangle
(74) =N<|β|m(a1,βa2,β),(Dβxα)v2+|β|=N,βα(a1,βa2,β),(Dβxα)v2\displaystyle=\sum_{N<|\beta|\leq m}\langle(a_{1,\beta}-a_{2,\beta}),(D^{\beta}x^{\alpha})v_{2}\rangle+\sum_{|\beta|=N,\,\beta\neq\alpha}\langle(a_{1,\beta}-a_{2,\beta}),(D^{\beta}x^{\alpha})v_{2}\rangle
(75) +(a1,αa2,α),(Dαxα)v2.\displaystyle\quad+\langle(a_{1,\alpha}-a_{2,\alpha}),(D^{\alpha}x^{\alpha})v_{2}\rangle.

If |β|>N=|α||\beta|>N=|\alpha|, then there must exist k{1,2,,n}k\in\{1,2,...,n\} such that βk>αk\beta_{k}>\alpha_{k}. This is true also if |β|=N|\beta|=N with βα\beta\neq\alpha. In both cases we can compute

Dβ(xα)=(x1β1x1α1)(x2β2x2α2)(xnβnxnαn)=0D^{\beta}(x^{\alpha})=(\partial_{x_{1}}^{\beta_{1}}x_{1}^{\alpha_{1}})\;(\partial_{x_{2}}^{\beta_{2}}x_{2}^{\alpha_{2}})\;...\;(\partial_{x_{n}}^{\beta_{n}}x_{n}^{\alpha_{n}})=0

because xkβkxkαk=0\partial_{x_{k}}^{\beta_{k}}x_{k}^{\alpha_{k}}=0. Therefore formula (73) becomes

0=(a1,αa2,α),(Dαxα)v2=α!a1,αa2,α,v20=\langle(a_{1,\alpha}-a_{2,\alpha}),(D^{\alpha}x^{\alpha})v_{2}\rangle=\alpha!\langle a_{1,\alpha}-a_{2,\alpha},v_{2}\rangle

which by the arbitrariety of v2Cc(Ω)v_{2}\in C^{\infty}_{c}(\Omega) implies a1,α|Ω=a2,α|Ωa_{1,\alpha}|_{\Omega}=a_{2,\alpha}|_{\Omega} also for α\alpha for which |α|=N|\alpha|=N.

Step 3. We have proved that a1,α|Ω=a2,α|Ωa_{1,\alpha}|_{\Omega}=a_{2,\alpha}|_{\Omega} for all α\alpha of order |α|m|\alpha|\leq m. Since this entails P1|Ω=P2|ΩP_{1}|_{\Omega}=P_{2}|_{\Omega}, the proof is complete. ∎

4. Main theorem for bounded coefficients

We shall now study the case when the coefficients of PDOs are from the bounded spaces Hrα,(Ω)H^{r_{\alpha},\infty}(\Omega). It should be noted, however, that most of the considerations of the previous section still apply identically.

4.1. Well-posedness of the forward problem

We shall define the bilinear forms for the problems (37) and (39) respectively by (41) and (42), just as in the case of singular coefficients. These will turn out to be bounded in Hs(n)×Hs(n)H^{s}({\mathbb{R}}^{n})\times H^{s}({\mathbb{R}}^{n}) as well, but the proof we give of this fact is a fortiori different. Since now we assume that aαHrα,(Ω)L(Ω)a_{\alpha}\in H^{r_{\alpha},\infty}(\Omega)\subset L^{\infty}(\Omega) for rα0r_{\alpha}\geq 0, the duality pairing aα,(Dαv)w\left\langle a_{\alpha},(D^{\alpha}v)w\right\rangle becomes an integral over Ω\Omega.

Lemma 4.1 (Boundedness of the bilinear forms).

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded Lipschitz domain and s+s\in\mathbb{R}^{+}\setminus\mathbb{Z}, mm\in\mathbb{N} such that 2s>m2s>m. Let aαHrα,(Ω)a_{\alpha}\in H^{r_{\alpha},\infty}(\Omega), with rαr_{\alpha} defined as in (4). Then BPB_{P} and BPB_{P}^{*} extend as bounded bilinear forms on Hs(n)×Hs(n)H^{s}(\mathbb{R}^{n})\times H^{s}(\mathbb{R}^{n}).

Remark 4.2.

Since s+s\in\mathbb{R}^{+}\setminus\mathbb{Z} and |α|m<2s|\alpha|\leq m<2s, we also have that max(0,|α|s)rα<s\max(0,|\alpha|-s)\leq r_{\alpha}<s for δ>0\delta>0 small (see formula (4)).

Proof of lemma 4.1.

We only prove the boundedness of BPB_{P}, as for BPB_{P}^{*} one can proceed in the same way. If v,wCc(n)v,w\in C^{\infty}_{c}(\mathbb{R}^{n}), then

(76) |aα(x)Dαv,w|\displaystyle|\langle a_{\alpha}(x)D^{\alpha}v,w\rangle| =|Ωaαw(Dαv)𝑑x|aαw(Hrα(Ω))DαvHrα(Ω).\displaystyle=\left|\int_{\Omega}a_{\alpha}w(D^{\alpha}v)\,dx\right|\leq\|a_{\alpha}w\|_{(H^{-r_{\alpha}}(\Omega))^{*}}\|D^{\alpha}v\|_{H^{-r_{\alpha}}(\Omega)}.

Since Ω\Omega is a Lipschitz domain and rα0r_{\alpha}\geq 0, rα{12,32,52}r_{\alpha}\not\in\left\{\frac{1}{2},\frac{3}{2},\frac{5}{2}...\right\}, we have (Hrα(Ω))=H0rα(Ω)Hrα(Ω)(H^{-r_{\alpha}}(\Omega))^{*}=H^{r_{\alpha}}_{0}(\Omega)\subset H^{r_{\alpha}}(\Omega). Therefore

(77) |aα(x)Dαv,w|\displaystyle|\langle a_{\alpha}(x)D^{\alpha}v,w\rangle| CaαwHrα(Ω)DαvHrα(Ω)CAαwHrα(n)DαvHrα(Ω)\displaystyle\leq C\|a_{\alpha}w\|_{H^{r_{\alpha}}(\Omega)}\|D^{\alpha}v\|_{H^{-r_{\alpha}}(\Omega)}\leq C\|A_{\alpha}w\|_{H^{r_{\alpha}}(\mathbb{R}^{n})}\|D^{\alpha}v\|_{H^{-r_{\alpha}}(\Omega)}
(78) CJrα(Aαw)L2(n)vH|α|rα(Ω)\displaystyle\leq C\|J^{r_{\alpha}}(A_{\alpha}w)\|_{L^{2}(\mathbb{R}^{n})}\|v\|_{H^{|\alpha|-r_{\alpha}}(\Omega)}

where J=(IdΔ)1/2J=(\mathrm{Id}-\Delta)^{1/2} is the Bessel potential and AαA_{\alpha} is an extension of aαa_{\alpha} from Ω\Omega to n\mathbb{R}^{n} such that Aα|Ω=aαA_{\alpha}|_{\Omega}=a_{\alpha} and AαHrα,(n)2aαHrα,(Ω)\|A_{\alpha}\|_{H^{r_{\alpha},\infty}(\mathbb{R}^{n})}\leq 2\|a_{\alpha}\|_{H^{r_{\alpha},\infty}(\Omega)}. Since rα0r_{\alpha}\geq 0, we may estimate the last term of (77) by the Kato-Ponce inequality given in lemma 2.3

(79) Jrα(Aαw)L2(n)\displaystyle\|J^{r_{\alpha}}(A_{\alpha}w)\|_{L^{2}(\mathbb{R}^{n})} C(AαL(n)JrαwL2(n)+JrαAαL(n)wL2(n))\displaystyle\leq C\left(\|A_{\alpha}\|_{L^{\infty}(\mathbb{R}^{n})}\|J^{r_{\alpha}}w\|_{L^{2}(\mathbb{R}^{n})}+\|J^{r_{\alpha}}A_{\alpha}\|_{L^{\infty}(\mathbb{R}^{n})}\|w\|_{L^{2}(\mathbb{R}^{n})}\right)
(80) CAαHrα,(n)wHrα(n)CaαHrα,(Ω)wHrα(n).\displaystyle\leq C\|A_{\alpha}\|_{H^{r_{\alpha},\infty}(\mathbb{R}^{n})}\|w\|_{H^{r_{\alpha}}(\mathbb{R}^{n})}\leq C\|a_{\alpha}\|_{H^{r_{\alpha},\infty}(\Omega)}\|w\|_{H^{r_{\alpha}}(\mathbb{R}^{n})}.

Substituting this into (77) gives

(81) |aα(x)Dαv,w|\displaystyle|\langle a_{\alpha}(x)D^{\alpha}v,w\rangle| CaαHrα,(Ω)wHrα(n)vH|α|rα(Ω)\displaystyle\leq C\|a_{\alpha}\|_{H^{r_{\alpha},\infty}(\Omega)}\|w\|_{H^{r_{\alpha}}(\mathbb{R}^{n})}\|v\|_{H^{|\alpha|-r_{\alpha}}(\Omega)}
(82) CaαHrα,(Ω)wHs(n)vHs(n)\displaystyle\leq C\|a_{\alpha}\|_{H^{r_{\alpha},\infty}(\Omega)}\|w\|_{H^{s}(\mathbb{R}^{n})}\|v\|_{H^{s}(\mathbb{R}^{n})}

given that both rα<sr_{\alpha}<s and |α|rαs|\alpha|-r_{\alpha}\leq s hold by remark 4.2. Eventually we obtain

(83) |BP(v,w)|\displaystyle|B_{P}(v,w)| |(Δ)s/2v,(Δ)s/2w|+|α|m|aαDαv,w|\displaystyle\leq|\langle(-\Delta)^{s/2}v,(-\Delta)^{s/2}w\rangle|+\sum_{|\alpha|\leq m}|\langle a_{\alpha}D^{\alpha}v,w\rangle|
(84) wHs(n)vHs(n)+|α|mCaαHrα,(Ω)wHs(n)vHs(n)\displaystyle\leq\|w\|_{H^{s}(\mathbb{R}^{n})}\|v\|_{H^{s}(\mathbb{R}^{n})}+\sum_{|\alpha|\leq m}C\|a_{\alpha}\|_{H^{r_{\alpha},\infty}(\Omega)}\|w\|_{H^{s}(\mathbb{R}^{n})}\|v\|_{H^{s}(\mathbb{R}^{n})}
(85) CwHs(n)vHs(n).\displaystyle\leq C\|w\|_{H^{s}(\mathbb{R}^{n})}\|v\|_{H^{s}(\mathbb{R}^{n})}.\qed

Next we shall prove existence and uniqueness of solutions for the problems (37) and (39). The reasoning is similar to the one for the proof of lemma 3.4, but the details of the computations are quite different.

Lemma 4.3 (Well-posedness).

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded Lipschitz domain and s+s\in\mathbb{R}^{+}\setminus\mathbb{Z}, mm\in\mathbb{N} such that 2s>m2s>m. Let aαHrα,(Ω)a_{\alpha}\in H^{r_{\alpha},\infty}(\Omega), with rαr_{\alpha} defined as in (4). There exist a real number μ>0\mu>0 and a countable set Σ(μ,)\Sigma\subset(-\mu,\infty) of eigenvalues λ1λ2\lambda_{1}\leq\lambda_{2}\leq...\rightarrow\infty such that if λΣ\lambda\in\mathbb{R}\setminus\Sigma, for any fHs(n)f\in H^{s}(\mathbb{R}^{n}) and F(H~s(Ω))F\in(\widetilde{H}^{s}(\Omega))^{*} there exists a unique uHs(n)u\in H^{s}(\mathbb{R}^{n}) such that ufH~s(Ω)u-f\in\widetilde{H}^{s}(\Omega) and

BP(u,v)λu,v=F(v)for allvH~s(Ω).B_{P}(u,v)-\lambda\langle u,v\rangle=F(v)\quad\mbox{for all}\quad v\in\widetilde{H}^{s}(\Omega).

One has the estimate

uHs(n)C(fHs(n)+F(H~s(Ω))).\|u\|_{H^{s}(\mathbb{R}^{n})}\leq C\left(\|f\|_{H^{s}(\mathbb{R}^{n})}+\|F\|_{(\widetilde{H}^{s}(\Omega))^{*}}\right).

The function uu is also the unique uHs(n)u\in H^{s}(\mathbb{R}^{n}) satisfying

rΩ((Δ)s+|α|maα(x)Dαλ)u=Fr_{\Omega}\left((-\Delta)^{s}+\sum_{|\alpha|\leq m}a_{\alpha}(x)D^{\alpha}-\lambda\right)u=F

in the sense of distributions in Ω\Omega and ufH~s(Ω)u-f\in\widetilde{H}^{s}(\Omega). Moreover, if (63) holds then 0Σ0\notin\Sigma.

Proof.

Again it is enough to find unique u~H~s(Ω)\tilde{u}\in\widetilde{H}^{s}(\Omega) such that BP(u~,v)λu~,v=F~(v)B_{P}(\tilde{u},v)-\lambda\langle\tilde{u},v\rangle=\tilde{F}(v), where F~:=FBP(f,)+λf,\tilde{F}:=F-B_{P}(f,\cdot)+\lambda\langle f,\cdot\rangle. Consider v,wCc(Ω)v,w\in C^{\infty}_{c}(\Omega) and rα0r_{\alpha}\neq 0. Since 0<rα<s0<r_{\alpha}<s, the interpolation inequality

wHrα(n)CwL2(n)1rα/swHs(n)rα/s\|w\|_{H^{r_{\alpha}}(\mathbb{R}^{n})}\leq C\|w\|_{L^{2}(\mathbb{R}^{n})}^{1-r_{\alpha}/s}\|w\|_{H^{s}(\mathbb{R}^{n})}^{r_{\alpha}/s}

holds. Using this and formula (81) we get, for a constant C=C(Ω,n,s,rα)C=C(\Omega,n,s,r_{\alpha}) which may change from line to line,

(86) |aα(x)Dαv,w|\displaystyle|\langle a_{\alpha}(x)D^{\alpha}v,w\rangle| CaαHrα,(Ω)vHs(n)wHrα(n)\displaystyle\leq C\|a_{\alpha}\|_{H^{r_{\alpha},\infty}(\Omega)}\|v\|_{H^{s}(\mathbb{R}^{n})}\|w\|_{H^{r_{\alpha}}(\mathbb{R}^{n})}
(87) CaαHrα,(Ω)vHs(n)wL2(n)1rα/swHs(n)rα/s\displaystyle\leq C\|a_{\alpha}\|_{H^{r_{\alpha},\infty}(\Omega)}\|v\|_{H^{s}(\mathbb{R}^{n})}\|w\|_{L^{2}(\mathbb{R}^{n})}^{1-r_{\alpha}/s}\|w\|_{H^{s}(\mathbb{R}^{n})}^{r_{\alpha}/s}
(88) aαHrα,(Ω)vHs(n)(Cϵrα/(rαs)wL2(n)+ϵwHs(n)).\displaystyle\leq\|a_{\alpha}\|_{H^{r_{\alpha},\infty}(\Omega)}\|v\|_{H^{s}(\mathbb{R}^{n})}\left(C\epsilon^{r_{\alpha}/(r_{\alpha}-s)}\|w\|_{L^{2}(\mathbb{R}^{n})}+\epsilon\|w\|_{H^{s}(\mathbb{R}^{n})}\right).

In the last step of (86) we used formula (46) with

q=srα,p=ssrα,b=wHs(n)rα/s,a=CwL2(n)1rα/s,η=ϵ.q=\frac{s}{r_{\alpha}},\quad p=\frac{s}{s-r_{\alpha}},\quad b=\|w\|_{H^{s}(\mathbb{R}^{n})}^{r_{\alpha}/s},\quad a=C\|w\|_{L^{2}(\mathbb{R}^{n})}^{1-r_{\alpha}/s},\quad\eta=\epsilon.

If instead rα=0r_{\alpha}=0, just by formula (81) we already have

|aα(x)Dαv,w|CaαL(Ω)vHs(n)wL2(n).|\langle a_{\alpha}(x)D^{\alpha}v,w\rangle|\leq C\|a_{\alpha}\|_{L^{\infty}(\Omega)}\|v\|_{H^{s}(\mathbb{R}^{n})}\|w\|_{L^{2}(\mathbb{R}^{n})}.

Moreover, the two estimates above also hold for v,wH~s(Ω)v,w\in\widetilde{H}^{s}(\Omega) by the density of Cc(Ω)C^{\infty}_{c}(\Omega) in H~s(Ω)\widetilde{H}^{s}(\Omega). Now we use formula (46) again, but this time we choose

q=p=2,b=vHs(n),a=vL2(n),η=ϵs/(srα).q=p=2,\quad b=\|v\|_{H^{s}(\mathbb{R}^{n})},\quad a=\|v\|_{L^{2}(\mathbb{R}^{n})},\quad\eta=\epsilon^{s/(s-r_{\alpha})}.

This leads to

(89) |aα(x)Dαv,v|\displaystyle|\langle a_{\alpha}(x)D^{\alpha}v,v\rangle| aαHrα,(Ω)vHs(n)(Cϵrα/(rαs)vL2(n)+ϵvHs(n))\displaystyle\leq\|a_{\alpha}\|_{H^{r_{\alpha},\infty}(\Omega)}\|v\|_{H^{s}(\mathbb{R}^{n})}\left(C\epsilon^{r_{\alpha}/(r_{\alpha}-s)}\|v\|_{L^{2}(\mathbb{R}^{n})}+\epsilon\|v\|_{H^{s}(\mathbb{R}^{n})}\right)
(90) =aαHrα,(Ω)(Cϵrα/(rαs)vL2(n)vHs(n)+ϵvHs(n)2)\displaystyle=\|a_{\alpha}\|_{H^{r_{\alpha},\infty}(\Omega)}\left(C\epsilon^{r_{\alpha}/(r_{\alpha}-s)}\|v\|_{L^{2}(\mathbb{R}^{n})}\|v\|_{H^{s}(\mathbb{R}^{n})}+\epsilon\|v\|_{H^{s}(\mathbb{R}^{n})}^{2}\right)
(91) aαHrα,(Ω)(Cϵrα+srαsvL2(n)2+ϵ(C+1)vHs(n)2)\displaystyle\leq\|a_{\alpha}\|_{H^{r_{\alpha},\infty}(\Omega)}\left(C\epsilon^{\frac{r_{\alpha}+s}{r_{\alpha}-s}}\|v\|_{L^{2}(\mathbb{R}^{n})}^{2}+\epsilon(C+1)\|v\|_{H^{s}(\mathbb{R}^{n})}^{2}\right)
(92) CaαHrα,(Ω)(ϵrα+srαsvL2(n)2+ϵvHs(n)2)\displaystyle\leq C\|a_{\alpha}\|_{H^{r_{\alpha},\infty}(\Omega)}\left(\epsilon^{\frac{r_{\alpha}+s}{r_{\alpha}-s}}\|v\|_{L^{2}(\mathbb{R}^{n})}^{2}+\epsilon\|v\|_{H^{s}(\mathbb{R}^{n})}^{2}\right)
(93) CaαHrα,(Ω)(ϵM+sMsvL2(n)2+ϵvHs(n)2)\displaystyle\leq C^{\prime}\|a_{\alpha}\|_{H^{r_{\alpha},\infty}(\Omega)}\left(\epsilon^{\frac{M+s}{M-s}}\|v\|_{L^{2}(\mathbb{R}^{n})}^{2}+\epsilon\|v\|_{H^{s}(\mathbb{R}^{n})}^{2}\right)

where C=C(Ω,n,s,rα)C=C(\Omega,n,s,r_{\alpha}) and C=C(Ω,n,s)C^{\prime}=C^{\prime}(\Omega,n,s) are constants changing from line to line and M[0,s)M\in[0,s) is defined by M:=max|α|mrαM:=\max_{|\alpha|\leq m}r_{\alpha}. Eventually

(94) BP(v,v)\displaystyle B_{P}(v,v) (Δ)s/2vL2(n)2|α|m|aα(x)Dαv,v|\displaystyle\geq\|(-\Delta)^{s/2}v\|^{2}_{L^{2}(\mathbb{R}^{n})}-\sum_{|\alpha|\leq m}|\langle a_{\alpha}(x)D^{\alpha}v,v\rangle|
(95) (Δ)s/2vL2(n)2C(ϵM+sMsvL2(n)2+ϵvHs(n)2)|α|maαHrα,(Ω)\displaystyle\geq\|(-\Delta)^{s/2}v\|^{2}_{L^{2}(\mathbb{R}^{n})}-C^{\prime}\left(\epsilon^{\frac{M+s}{M-s}}\|v\|_{L^{2}(\mathbb{R}^{n})}^{2}+\epsilon\|v\|_{H^{s}(\mathbb{R}^{n})}^{2}\right)\sum_{|\alpha|\leq m}\|a_{\alpha}\|_{H^{r_{\alpha},\infty}(\Omega)}
(96) =(Δ)s/2vL2(n)2CC′′(ϵM+sMsvL2(n)2+ϵvHs(n)2)\displaystyle=\|(-\Delta)^{s/2}v\|^{2}_{L^{2}(\mathbb{R}^{n})}-C^{\prime}C^{\prime\prime}\left(\epsilon^{\frac{M+s}{M-s}}\|v\|_{L^{2}(\mathbb{R}^{n})}^{2}+\epsilon\|v\|_{H^{s}(\mathbb{R}^{n})}^{2}\right)

where C′′:=|α|maαHrα,(Ω)C^{\prime\prime}:=\sum_{|\alpha|\leq m}\|a_{\alpha}\|_{H^{r_{\alpha},\infty}(\Omega)} is a constant independent of ϵ\epsilon and vv. By the higher order Poincaré inequality (lemma 2.2) (94) turns into

(97) BP(v,v)\displaystyle B_{P}(v,v) c((Δ)s/2vL2(n)2+vL2(n)2)CC′′(ϵM+sMsvL2(n)2+ϵvHs(n)2)\displaystyle\geq c\left(\|(-\Delta)^{s/2}v\|^{2}_{L^{2}(\mathbb{R}^{n})}+\|v\|_{L^{2}(\mathbb{R}^{n})}^{2}\right)-C^{\prime}C^{\prime\prime}\left(\epsilon^{\frac{M+s}{M-s}}\|v\|_{L^{2}(\mathbb{R}^{n})}^{2}+\epsilon\|v\|_{H^{s}(\mathbb{R}^{n})}^{2}\right)
(98) cvHs(n)2CC′′(ϵM+sMsvL2(n)2+ϵvHs(n)2)\displaystyle\geq c\|v\|_{H^{s}(\mathbb{R}^{n})}^{2}-C^{\prime}C^{\prime\prime}\left(\epsilon^{\frac{M+s}{M-s}}\|v\|_{L^{2}(\mathbb{R}^{n})}^{2}+\epsilon\|v\|_{H^{s}(\mathbb{R}^{n})}^{2}\right)

for some constant c=c(Ω,n,s)c=c(\Omega,n,s) changing from line to line. For ϵ\epsilon small enough (notice that Ms<0M-s<0), this eventually gives the coercivity estimate

(99) BP(v,v)c0vHs(n)2μvL2(n)2B_{P}(v,v)\geq c_{0}\|v\|_{H^{s}(\mathbb{R}^{n})}^{2}-\mu\|v\|_{L^{2}(\mathbb{R}^{n})}^{2}

for some constants c0,μ>0c_{0},\mu>0 independent of vv. The proof is now concluded as in lemma 3.4. ∎

Assuming as in Section 3 that both (63) and (64) hold, by means of the above lemma 4.3 we can define the DN-maps ΛP,ΛP\Lambda_{P},\Lambda_{P}^{*} just as in lemma 3.6.

Definition 4.4.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded open set. Let s+s\in\mathbb{R}^{+}\setminus\mathbb{Z} and mm\in\mathbb{N} such that 2s>m2s>m, and let aαHrα,(Ω)a_{\alpha}\in H^{r_{\alpha},\infty}(\Omega), with rαr_{\alpha} defined as in (4). The exterior DN maps ΛP\Lambda_{P} and ΛP\Lambda_{P}^{*} are

ΛP:XXdefined byΛP[f],[g]:=BP(uf,g)\Lambda_{P}:X\rightarrow X^{*}\quad\mbox{defined by}\quad\langle\Lambda_{P}[f],[g]\rangle:=B_{P}(u_{f},g)

and

ΛP:XXdefined byΛP[f],[g]:=BP(uf,g)\Lambda^{*}_{P}:X\rightarrow X^{*}\quad\mbox{defined by}\quad\langle\Lambda^{*}_{P}[f],[g]\rangle:=B^{*}_{P}(u^{*}_{f},g)

where uf,ufu_{f},u^{*}_{f} are the unique solutions to the equations

(Δ)su+|α|maαDαu\displaystyle(-\Delta)^{s}u+\sum_{|\alpha|\leq m}a_{\alpha}D^{\alpha}u =0inΩ,ufH~s(Ω)\displaystyle=0\quad\mbox{in}\;\;\Omega,\quad u-f\in\widetilde{H}^{s}(\Omega)

and

(Δ)su+|α|m(1)|α|Dα(aαu)\displaystyle(-\Delta)^{s}u^{*}+\sum_{|\alpha|\leq m}(-1)^{|\alpha|}D^{\alpha}(a_{\alpha}u^{*}) =0inΩ,ufH~s(Ω)\displaystyle=0\quad\mbox{in}\;\;\Omega,\quad u^{*}-f\in\widetilde{H}^{s}(\Omega)

with f,gHs(n)f,g\in H^{s}(\mathbb{R}^{n}).

4.2. Proof of injectivity

We also arrive at the same Alessandrini identity and Runge approximation property which we get in lemmas 3.8 and 3.9.

Lemma 4.5 (Alessandrini identity).

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded Lipschitz domain and s+s\in\mathbb{R}^{+}\setminus\mathbb{Z}, mm\in\mathbb{N} such that 2s>m2s>m. Let aαHrα,(Ω)a_{\alpha}\in H^{r_{\alpha},\infty}(\Omega), with rαr_{\alpha} defined as in (4). For any f1,f2Hs(n)f_{1},f_{2}\in H^{s}(\mathbb{R}^{n}), let u1,u2Hs(n)u_{1},u_{2}^{*}\in H^{s}(\mathbb{R}^{n}) respectively solve

(Δ)su1+|α|ma1,α(x)Dαu1\displaystyle(-\Delta)^{s}u_{1}+\sum_{|\alpha|\leq m}a_{1,\alpha}(x)D^{\alpha}u_{1} =0inΩ,u1f1H~s(Ω)\displaystyle=0\quad\mbox{in}\;\;\Omega,\quad u_{1}-f_{1}\in\widetilde{H}^{s}(\Omega)

and

(Δ)su2+|α|m(1)|α|Dα(a2,α(x)u2)\displaystyle(-\Delta)^{s}u_{2}^{*}+\sum_{|\alpha|\leq m}(-1)^{|\alpha|}D^{\alpha}(a_{2,\alpha}(x)u_{2}^{*}) =0inΩ,u2f2H~s(Ω).\displaystyle=0\quad\mbox{in}\;\;\Omega,\quad u_{2}^{*}-f_{2}\in\widetilde{H}^{s}(\Omega).

Then we have the integral identity

(ΛP1ΛP2)[f1],[f2]=|α|m(a1,αa2,α)Dαu1,u2.\langle(\Lambda_{P_{1}}-\Lambda_{P_{2}})[f_{1}],[f_{2}]\rangle=\sum_{|\alpha|\leq m}\langle(a_{1,\alpha}-a_{2,\alpha})D^{\alpha}u_{1},u_{2}^{*}\rangle.
Lemma 4.6 (Runge approximation property).

Let Ω,Wn\Omega,W\subset\mathbb{R}^{n} respectively be a bounded Lipschitz domain and a non-empty open set such that W¯Ω¯=\overline{W}\cap\overline{\Omega}=\emptyset. Let s+s\in\mathbb{R}^{+}\setminus\mathbb{Z}, mm\in\mathbb{N} such that 2s>m2s>m. Let aαHrα,(Ω)a_{\alpha}\in H^{r_{\alpha},\infty}(\Omega), with rαr_{\alpha} defined as in (4). Moreover, let :={uff:fCc(W)}H~s(Ω)\mathcal{R}:=\{\,u_{f}-f:f\in C^{\infty}_{c}(W)\,\}\subset\widetilde{H}^{s}(\Omega), where ufu_{f} solves

(Δ)suf+|α|maα(x)Dαuf\displaystyle(-\Delta)^{s}u_{f}+\sum_{|\alpha|\leq m}a_{\alpha}(x)D^{\alpha}u_{f} =0inΩ,uffH~s(Ω)\displaystyle=0\quad\mbox{in}\;\;\Omega,\quad u_{f}-f\in\widetilde{H}^{s}(\Omega)

and :={uff:fCc(W)}H~s(Ω)\mathcal{R}^{*}:=\{\,u^{*}_{f}-f:f\in C^{\infty}_{c}(W)\,\}\subset\widetilde{H}^{s}(\Omega), where ufu^{*}_{f} solves

(Δ)suf+|α|m(1)|α|Dα(aα(x)uf)\displaystyle(-\Delta)^{s}u_{f}^{*}+\sum_{|\alpha|\leq m}(-1)^{|\alpha|}D^{\alpha}(a_{\alpha}(x)u_{f}^{*}) =0inΩ,uffH~s(Ω).\displaystyle=0\quad\mbox{in}\;\;\Omega,\quad u_{f}^{*}-f\in\widetilde{H}^{s}(\Omega).

Then \mathcal{R} and \mathcal{R}^{*} are dense in H~s(Ω)\widetilde{H}^{s}(\Omega).

With this at hand, we can prove the main theorem for bounded coefficients.

Proof of theorem 1.4.

The proof is virtually identical to the one of theorem 1.1, the unique difference being in the way the error terms of the Runge approximation are estimated. We make use of (81), which relied on the Kato-Ponce inequality instead of multiplier space estimates. The proof is otherwise completed as the proof of theorem 1.1. ∎

References

  • [1] H. Abels. Pseudodifferential and Singular Integral Operators. De Gruyter Graduate Lectures. De Gruyter, Berlin, 2012. An introduction with applications.
  • [2] F. Andreu-Vaillo, J. M. Mazón, J. D. Rossi, and J. J. Toledo-Melero. Nonlocal Diffusion Problems, volume 165 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI; Real Sociedad Matemática Española, Madrid, 2010.
  • [3] A. Behzadan and M. Holst. Multiplication in Sobolev spaces, revisited. 2017. arXiv:1512.07379.
  • [4] J. Bergh and J. Löfström. Interpolation Spaces. An Introduction. Springer-Verlag, Berlin-New York, 1976. Grundlehren der Mathematischen Wissenschaften, No. 223.
  • [5] S. Bhattacharyya, T. Ghosh, and G. Uhlmann. Inverse problem for fractional-Laplacian with lower order non-local perturbations. Trans. Amer. Math. Soc., 374(5):3053–3075, 2021.
  • [6] C. Bucur and E. Valdinoci. Nonlocal Diffusion and Applications, volume 20 of Lecture Notes of the Unione Matematica Italiana. Springer, [Cham]; Unione Matematica Italiana, Bologna, 2016.
  • [7] X. Cao, Y.-H. Lin, and H. Liu. Simultaneously recovering potentials and embedded obstacles for anisotropic fractional Schrödinger operators. Inverse Probl. Imaging, 13(1):197–210, 2019.
  • [8] M. Cekić, Y.-H. Lin, and A. Rüland. The Calderón problem for the fractional Schrödinger equation with drift. Calc. Var. Partial Differential Equations, 59(3):Paper No. 91, 46, 2020.
  • [9] S. N. Chandler-Wilde, D. P. Hewett, and A. Moiola. Sobolev spaces on non-Lipschitz subsets of n\mathbb{R}^{n} with application to boundary integral equations on fractal screens. Integral Equations Operator Theory, 87(2):179–224, 2017.
  • [10] G. Covi. An inverse problem for the fractional Schrödinger equation in a magnetic field. Inverse Problems, 36(4):045004, 24, 2020.
  • [11] G. Covi. Inverse problems for a fractional conductivity equation. Nonlinear Anal., 193:111418, 18, 2020.
  • [12] G. Covi, K. Mönkkönen, and J. Railo. Unique continuation property and Poincaré inequality for higher order fractional Laplacians with applications in inverse problems. Inverse Probl. Imaging, 15(4):641–681, 2021.
  • [13] M. Daoud and E. H. Laamri. Fractional Laplacians: A short survey. Discrete Contin. Dyn. Syst. Ser. S, 2021. Published online.
  • [14] E. Di Nezza, G. Palatucci, and E. Valdinoci. Hitchhiker’s guide to the fractional Sobolev spaces. Bull. Sci. Math., 136(5):521–573, 2012.
  • [15] Q. Du, M. Gunzburger, R. B. Lehoucq, and K. Zhou. Analysis and Approximation of Nonlocal Diffusion Problems with Volume Constraints. SIAM Rev., 54, No. 4:667–696, 2012.
  • [16] T. Ghosh, Y.-H. Lin, and J. Xiao. The Calderón problem for variable coefficients nonlocal elliptic operators. Comm. Partial Differential Equations, 42(12):1923–1961, 2017.
  • [17] T. Ghosh, A. Rüland, M. Salo, and G. Uhlmann. Uniqueness and reconstruction for the fractional Calderón problem with a single measurement. J. Funct. Anal., 279(1):108505, 42, 2020.
  • [18] T. Ghosh, M. Salo, and G. Uhlmann. The Calderón problem for the fractional Schrödinger equation. Anal. PDE, 13(2):455–475, 2020.
  • [19] L. Grafakos and S. Oh. The Kato-Ponce inequality. Comm. Partial Differential Equations, 39(6):1128–1157, 2014.
  • [20] A. Gulisashvili and M. A. Kon. Exact Smoothing Properties of Schrödinger Semigroups. Amer. J. Math., 118(6):1215–1248, 1996.
  • [21] T. Helin, M. Lassas, L. Ylinen, and Z. Zhang. Inverse problems for heat equation and space–time fractional diffusion equation with one measurement. J. Differential Equations, 269(9):7498–7528, 2020.
  • [22] T. Kato and G. Ponce. Commutator Estimates and the Euler and Navier-Stokes Equations. Comm. Pure Appl. Math, 41(7):891–907, 1988.
  • [23] M. Kwaśnicki. Ten equivalent definitions of the fractional Laplace operator. Fract. Calc. Appl. Anal., 20(1):7–51, 2017.
  • [24] R.-Y. Lai and Y.-H. Lin. Global uniqueness for the fractional semilinear Schrödinger equation. Proc. Amer. Math. Soc., 147(3):1189–1199, 2019.
  • [25] R.-Y. Lai and Y.-H. Lin. Inverse problems for fractional semilinear elliptic equations. 2020. arXiv:2004.00549.
  • [26] R.-Y. Lai, Y.-H. Lin, and A. Rüland. The Calderón problem for a space-time fractional parabolic equation. SIAM J. Math. Anal., 52(3):2655–2688, 2020.
  • [27] N. Laskin. Fractional quantum mechanics and Lévy path integrals. Phys. Lett. A, 268(4-6):298–305, 2000.
  • [28] N. Laskin. Fractional Schrödinger equation. Phys. Rev. E, 66:056108, 2002.
  • [29] N. Laskin. Fractional Quantum Mechanics. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2018.
  • [30] L. Li. A Semilinear Inverse Problem For The Fractional Magnetic Laplacian. 2020. arXiv:2005.06714.
  • [31] L. Li. The Calderón problem for the fractional magnetic operator. Inverse Problems, 36(7):075003, 2020.
  • [32] L. Li. Determining the magnetic potential in the fractional magnetic Calderón problem. Comm. Partial Differential Equations, 46(6):1017–1026, 2021.
  • [33] A. Lischke, G. Pang, M. Gulian, F. Song, C. Glusa, X. Zheng, Z. Mao, W. Cai, M. M. Meerschaert, M. Ainsworth, and G. E. Karniadakis. What is the fractional Laplacian? A comparative review with new results. J. Comput. Phys., 404:109009, 2020.
  • [34] A. Massaccesi and E. Valdinoci. Is a nonlocal diffusion strategy convenient for biological populations in competition? J. Math. Biol., 74(1):113–147, 2017.
  • [35] V. G. Maz’ya and T. O. Shaposhnikova. Theory of Sobolev Multipliers, volume 337 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 2009. With applications to differential and integral operators.
  • [36] W. McLean. Strongly Elliptic Systems and Boundary Integral Equations. Cambridge University Press, Cambridge, 2000.
  • [37] R. Metzler and J. Klafter. The random walk’s guide to anomalous diffusion: a fractional dynamics approach. Phys. Rep., 339(1):1–77, 2000.
  • [38] M. Mišur. A Refinement of Peetre’s Theorem. Results Math., 74(4):199, 2019.
  • [39] J. Navarro and J. B. Sancho. Peetre-Slovák’s theorem revisited. 2014. arXiv:1411.7499.
  • [40] J. Peetre. Une caractérisation abstraite des opérateurs différentiels. Math. Scand., 7:211–218, 1959.
  • [41] X. Ros-Oton. Nonlocal equations in bounded domains: a survey. Publ. Mat., 60(1):3–26, 2016.
  • [42] X. Ros-Oton and J. Serra. Local integration by parts and Pohozaev identities for higher order fractional Laplacians. Discrete Contin. Dyn. Syst., 35(5):2131–2150, 2015.
  • [43] A. Rüland and M. Salo. Exponential instability in the fractional Calderón problem. Inverse Problems, 34(4):045003, 21, 2018.
  • [44] A. Rüland and M. Salo. The fractional Calderón problem: low regularity and stability. Nonlinear Anal., 193:111529, 56, 2020.
  • [45] A. Rüland and M. Salo. Quantitative approximation properties for the fractional heat equation. Math. Control Relat. Fields, 10(1):1–26, 2020.
  • [46] M. Salo. The fractional Calderón problem. Journées équations aux dérivées partielles, Exp. No.(7), 2017.
  • [47] R. Servadei and E. Valdinoci. On the spectrum of two different fractional operators. Proc. Roy. Soc. Edinburgh Sect. A, 144(4):831–855, 2014.
  • [48] H. Triebel. Interpolation Theory, Function Spaces, Differential Operators. North-Holland Publishing Company, 1978.
  • [49] G. Uhlmann. Electrical impedance tomography and Calderón’s problem. Inverse Problems, 25(12):123011, 2009.
  • [50] G. Uhlmann. Inverse problems: seeing the unseen. Bull. Math. Sci., 4(2):209–279, 2014.
  • [51] M. W. Wong. An Introduction to Pseudo-Differential Operators. World Scientific, Third edition, 2014.