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

The Calderón problem for the Fractional Dirac Operator

Hadrian Quan, Gunther Uhlmann
Abstract.

We show that knowledge of the source-to-solution map for the fractional Dirac operator acting over sections of a Hermitian vector bundle over a smooth closed connencted Riemannian manifold of dimension m2m\geq 2 determines uniquely the smooth structure, Riemannian metric, Hermitian bundle and connection, and its Clifford modulo up to a isometry. We also mention several potential applications in physics and other fields.

1. Introduction

The Calderón problem asks whether one can determine the electrical conductivity of a medium by making voltage and current measurements at the boundary. In the anisotropic case, that is when the conductivity depends on direction. It is modelled by a positive-definite symmetric matrix. It was shown in [24] that this problem is equivalent to determining a Riemannian metric from the associated Dirichlet to Neumann (DN) map associated with harmonic functions. Therefore this problem can be considered on a compact Riemannian manifold with boundary and arises also in the AdS/CFT correspondence [5]. See [33], [34] for more details and other results.

The study of the fractional Calderón problem was initiated in [15] where the unknown potential in the fractional Schrödinger equation on a bounded domain in the Euclidean space was determined from exterior measurements. An important generalization was in the work [13] where the Euclidean Laplacian is replaced by the Laplace-Beltrami operator of a Riemannian metric. Following these works, inverse problems of recovering lower order terms for fractional elliptic equations have been studied extensively, see for example [14], [13], [30], [29], [2], [4], [7], [6], [8], [9], [10], [25], [26], [28] for some of the contributions. In all of those papers, it is assumed that the leading order coefficients are known.

In the article [12], it was solved the fractional anisotropic Calderón problem on closed Riemannian manifolds of dimensions two and higher that the knowledge of the local source-to-solution map for the fractional Laplacian, given an arbitrary small open nonempty a priori known subsef a smooth closed connected Riemannian manifold, determines the Riemannian manifold up to an isometry. This can be viewed as a nonlocal analog of the anisotropic Calderón problem in the setting of closed Riemannian manifolds, which is open in dimensions three and higher.

We consider in this paper the anisotropic Calderon problem for the fractional Dirac operator acting on sections of a Hermitian vector bundle over a smooth closed connected Riemannian manifold MM of dimension m2m\geq 2. Just from local measurements confined to an arbitrary non-empty open set, we show that the fractional Dirac operator determines the smooth structure, Riemannian metric, Hermitian bundle and connection, and its Clifford module, up to a isometry which fixes the set in question. We will first briefly motivate the concept of a generalized Dirac operator, and its fractional powers, before explaining the types of measurements we allow, and stating our results.

In his attempts to quantize electromagnetism, Dirac defined the Dirac operator while trying to find a 1st order differential operator such that its square was the “Laplacian” on Minkowski space (in fact this was a factorization of the d’Alembertian). The data of the Clifford algebra (see §2) entered into his definition after he realized this was not possible for a scalar operator. On a vector bundle EE over a Riemannian manifold MM, the definition of a generalized Dirac operator \not{D} is motivated by the desire to capture this property that =ΔE\not{D}\circ\not{D}=\Delta_{E} is a generalized Laplace-type operator on EE, i.e. the Bochner Laplacian \nabla^{*}\nabla of some connection up to lower order terms. This factorization of ΔE\Delta_{E} implies \not{D} is elliptic and hence admits a spectral resolution {φk,λk}\{\varphi_{k},\lambda_{k}\} of L2(M;E)L^{2}(M;E)-orthonormal eigensections, with discrete eigenvalues accumulating only at ±\pm\infty. From the spectral theorem we can define for α(0,1)\alpha\in(0,1), the fractional Dirac operator by αφk=λkαφk\not{D}^{\alpha}\varphi_{k}=\lambda_{k}^{\alpha}\varphi_{k}. From unique solvability of the “Poisson equation”,

αu=f,\not{D}^{\alpha}u=f,

for data ff orthogonal to Ker()\operatorname{Ker}(\not{D}), we can consider the operator which maps a source function f𝒞0(𝒪;E|𝒪)Ker()f\in{\mathcal{C}}^{\infty}_{0}(\mathcal{O};E|_{\mathcal{O}})\cap\operatorname{Ker}(\not{D})^{\perp} to the solution,

L,𝒪:fu:=(αf)|𝒪,L_{\not{D},\mathcal{O}}:f\mapsto u:=(\not{D}^{-\alpha}f)|_{\mathcal{O}},

for fKer()f\perp\operatorname{Ker}(\not{D}). We refer to this operator as the source-to-solution map. We prove,

Theorem 1.1.

Let 𝒟j:=(Mj,Ej,gj,hEj,Ej)\mathscr{D}_{j}:=(M_{j},E_{j},g_{j},h_{E_{j}},\nabla^{E_{j}}) for j=1,2j=1,2 be two Dirac bundles (see §2) over closed Riemannian manifolds of dimension n2n\geq 2, and let 𝒪jMj\mathcal{O}_{j}\subset M_{j} be non-empty open sets. Assuming there exists a diffeomorphism ψ:𝒪1𝒪2\psi:\mathcal{O}_{1}\to\mathcal{O}_{2} satisfying

L1,𝒪1(ψf)=ψL2,𝒪2(f),L_{\not{D}_{1},\mathcal{O}_{1}}(\psi^{*}f)=\psi^{*}L_{\not{D}_{2},\mathcal{O}_{2}}(f),

for all f𝒞0(𝒪1;E1|𝒪1)f\in{\mathcal{C}}^{\infty}_{0}(\mathcal{O}_{1};E_{1}|_{\mathcal{O}_{1}}) orthogonal to Ker(1)\operatorname{Ker}(\not{D}_{1}). Then there is an isomorphism of Hermitian vector bundles Ψ:E1E2\Psi:E_{1}\to E_{2}, covering an isometry ψ:M1M2\psi:M_{1}\to M_{2} which restricts to Ψ|E1|𝒪1=ψ\Psi|_{E_{1}|_{\mathcal{O}_{1}}}=\psi.

Corollary 1.2.

Let 𝒟:=(M,E,g,hE,E)\mathscr{D}:=(M,E,g,h_{E},\nabla^{E}) be a Dirac bundle over a smooth closed manifold of dimension m2m\geq 2, with 𝒪M\mathcal{O}\subset M a non-empty open subset. Let {φk}k=1L2(M;E)\{\varphi_{k}\}_{k=1}^{\infty}\subset L^{2}(M;E) be the collection of positive eigensections with corresponding eigenvalues {λk}k=1+\{\lambda_{k}\}_{k=1}^{\infty}\subset\mathbb{R}^{+}. Then the partial spectral data plus clifford multiplication 𝒸𝓁:𝓁(𝒯,)End()\mathpzc{cl}:\mathbb{C}l(TM,g)\to\operatorname{End}(E),

(𝒪,φk|𝒪,λk,𝒸𝓁|𝒯𝒪)(\mathcal{O},\varphi_{k}|_{\mathcal{O}},\lambda_{k},\mathpzc{cl}|_{T^{*}\mathcal{O}})

determines the metric, Hermitian vector bundle, and connection, up to an isometry.

1.1. Applications

Already there has been some interest in studying generalizations of classical equations of physics with respect to a fractional time derivative, (see for example [23], [19]).

In this work we consider instead fractional differential operators, corresponding to fractional spatial derivatives. One place where there has been interest in applying such nonlocal operators is in the study of particle physics beyond the standard model, i.e. fields governed by fractional wave equations. The work of [38] considered nnth powers for n2n\geq 2 of the d’Alembert operator and demonstrated that for n>2n>2 the covariant wave equations generated by g1/n\Box_{g}^{1/n} generate a representation of SU(n)SU(n). Along similar lines [18] developed a form of local gauge invariance for such fractional fields and used this to deduce the Baryon mass spectrum via a fractional extension of the classical Zeeman effect [39]. For a more recent development of such fractional field theories see the work of [16] who suggest that anomalous power laws for the “strange metal” properties of Cuprate can be explained if the metal interacts with light via a gauge theory of fractional dimension; also the work of [22] who introduce a theory of fractional electromagnetism, which is motivated in part by a generalization of the Caffarelli-Silvestre extension to the case of the Hodge Laplacian.

The most surprising applications relate such nonlocal operators to questions coming from quantum information theory. A crucial first step in studying entanglement properties of algebraic quantum field theories is the Reeh-Schlieder theorem [37], which states all local fields in the field algebra of spacetime are entangled with fields localized to all other regions (c.f. [31] for a more mathematical exposition of this theorem in the language of von Neumann algebras). The work of [36] first gave a proof of the Reeh-Schlieder theorem for static spacetimes using only the strong anti-locality property of Δg1/2\Delta_{g}^{1/2}. In the work of [15] they demonstrate the comparable anti-locality property for Δgα\Delta_{g}^{\alpha} for α(0,1)\alpha\in(0,1), using standard techniques in inverse problems (Carleman estimates, etc.). That this entanglement property of the field algebra is equivalent to a strong unique continuation principle for certain nonlocal operators suggests other interesting connections between questions in inverse problems and quantum information theory.

2. Background on Generalized Dirac operators

Let (Mm,En,hE,g)(M^{m},E^{n},h_{E},g) be a Hermitian vector bundle of rank nn over smooth compact Riemannian manifold without boundary. We further call this Hermitian bundle a Clifford module if EME\to M is bundle over MM equipped with bundle morphism, known as Clifford multiplication,

𝒸𝓁:𝓁(𝒯,)End()\mathpzc{cl}:\mathbb{C}l(T^{*}M,g)\to\operatorname{End}(E)

from the unique Clifford module on TMT^{*}M induced by gg to End(E)\operatorname{End}(E). As a vector bundle, l(TM,g)\mathbb{C}l(T^{*}M,g) is isomorphic to ΛTM\Lambda^{\bullet}T^{*}M, and as a module has an algebra operation determined by the relation

𝒸𝓁(α)𝒸𝓁(β)+𝒸𝓁(β)𝒸𝓁(α)=2(α,β)Idα,βΛ𝒯.\mathpzc{cl}(\alpha)\mathpzc{cl}(\beta)+\mathpzc{cl}(\beta)\mathpzc{cl}(\alpha)=-2g(\alpha,\beta)\operatorname{Id}\quad\forall\alpha,\beta\in\Lambda^{\bullet}T^{*}M.

We say the Hermitian metric is compatible with Clifford multiplication if

hE(𝒸𝓁(θ)𝓋,𝓌)+𝒽(𝓋,𝒸𝓁(θ)𝓌)=0,𝓋,𝓌𝒞(;)h_{E}(\mathpzc{cl}(\theta)v,w)+h_{E}(v,\mathpzc{cl}(\theta)w)=0,\quad\forall v,w\in{\mathcal{C}}^{\infty}(M;E)

and a choice of Hermitian connection E\nabla^{E} on EE is is compatible with Clifford multiplication if

[E,𝒸𝓁(θ)]=𝒸𝓁(θ)[\nabla^{E},\mathpzc{cl}(\theta)]=\mathpzc{cl}(\nabla^{g}\theta)

where g\nabla^{g} is the Levi-Civita connection of gg.

Given a Hermitian bundle with connection both compatible with Clifford multiplication we can construct a generalized Dirac operator \not{D}, defined by

:𝒞(M;E)E𝒞(M;TME)i𝒸𝓁()𝒞(;),\not{D}:{\mathcal{C}}^{\infty}(M;E)\xrightarrow{\nabla^{E}}{\mathcal{C}}^{\infty}(M;T^{*}M\otimes E)\xrightarrow{i\mathpzc{cl}(-)}{\mathcal{C}}^{\infty}(M;E),

and call the collected data (M,E,g,hE,E)(M,E,g,h_{E},\nabla^{E}) a Dirac bundle. The generalized Dirac operator is a Dirac operator in the usual sense by arising as the ‘square root’ of a generalized Laplace operator on EE; compatibility of the Hermitian metric and connection implies

[,f]=[i𝒸𝓁,𝒻]=𝒾𝒸𝓁(𝒻),[\not{D},f]=[i\mathpzc{cl}\circ\nabla^{E},f]=i\mathpzc{cl}(f),

thus \not{D} has principal symbol

σ1()(x,ξ)=i𝒸𝓁(ξ),σ2(2)(𝓍,ξ)=|ξ|2Id,\sigma_{1}(\not{D})(x,\xi)=i\mathpzc{cl}(\xi),\implies\sigma_{2}(\not{D}^{2})(x,\xi)=|\xi|_{g}^{2}\operatorname{Id}_{E},

hence 2:=ΔE\not{D}^{2}:=\Delta_{E} is a principally scalar multiple of the metric gg. Further, we see from this computation that both \not{D} and 2\not{D}^{2} are principally scalar elliptic differential operators (of orders 1 and 2 respectively).

3. Fractional Dirac operator and determination of the Heat kernel

In this section we give two equivalent definitions of the fractional Dirac operator and use one to define the source-to-solution operator associated to an a priori known open set 𝒪M\mathcal{O}\subset M. Then, following [12], we show that knowledge of this source-to-solution operator determines the Heat kernel on 𝒪\mathcal{O}.

From the symbol calculation above we see that \not{D} is a symmetric operator. Because MM is a closed manifold, :𝒞(M;E)𝒞(M;E)\not{D}:{\mathcal{C}}^{\infty}(M;E)\to{\mathcal{C}}^{\infty}(M;E) is essentially self-adjoint on its core domain of smooth sections with a self-adjoint extension to H1(M;E)H^{1}(M;E). Unlike 2\not{D}^{2}, \not{D} fails to be a non-negative operator. On the other hand, its discrete spectrum (excluding the zero eigenvalue) is in correspondence with the spectrum of 2\not{D}^{2}; the discrete eigenvalues of \not{D} come in positive and negative pairs {±λk}\{\pm\lambda_{k}\} (corresponding to an eigenvalue λk2\lambda_{k}^{2} of 2\not{D}^{2}) which we index by their absolute values

0=λ0<|λ1||λ2|+0=\lambda_{0}<|\lambda_{1}|\leq|\lambda_{2}|\leq\cdots\nearrow+\infty

for the distinct eigenvalues of \not{D}, and denote dkd_{k} the multiplicity of λk\lambda_{k}. Let {φkj}j=1dk\{\varphi_{k_{j}}\}_{j=1}^{d_{k}} be an L2(M;E)L^{2}(M;E)-orthonormal basis for the eigenspace Ker(λk)\operatorname{Ker}(\not{D}-\lambda_{k}) corresponding to λk\lambda_{k}, and denote πk:L2(M;E)Ker(λk)\pi_{k}:L^{2}(M;E)\to\operatorname{Ker}(\not{D}-\lambda_{k}) for the orthogonal projection onto the corresponding eigenspace, written as

πk(f)=j=1dkf,φkjL2(M;E)φkj,\pi_{k}(f)=\sum_{j=1}^{d_{k}}\langle f,\varphi_{k_{j}}\rangle_{L^{2}(M;E)}\varphi_{k_{j}},

for all k=0,1,k=0,1,\ldots. Here ,L2(M;E)\langle\cdot,\cdot\rangle_{L^{2}(M;E)} is the L2L^{2}-inner product on sections of EE induced by our choice of hEh_{E}.

Fix α(0,1)\alpha\in(0,1). Given this spectral resolution of \not{D} we can define the Fractional Dirac operator α:𝒞(M;E)𝒞(M;E)\not{D}^{\alpha}:{\mathcal{C}}^{\infty}(M;E)\to{\mathcal{C}}^{\infty}(M;E)

αf=kλkαπk(f),\not{D}^{\alpha}f=\sum_{k\in\mathbb{Z}}\lambda_{k}^{\alpha}\pi_{k}(f),

which extends to an unbounded self-adjoint operator on L2(M;E)L^{2}(M;E) with domain 𝒟(α)=Hα(M;E)\mathcal{D}(\not{D}^{\alpha})=H^{\alpha}(M;E).

Unlike the scalar case, the nullspace of \not{D} may include more than just the constant functions. Say that dim Ker()=d0\operatorname{Ker}(\not{D})=d_{0} has orthonormal basis {φj0}j=1d0\{\varphi_{j}^{0}\}_{j=1}^{d_{0}}, then we can solve the equation

αu=f,\not{D}^{\alpha}u=f, (3.1)

for f𝒞0(𝒪;E)f\in{\mathcal{C}}^{\infty}_{0}(\mathcal{O};E), with 𝒪M\mathcal{O}\subset M open, whenever we impose

f,φ10L2==f,φd00L2=0\langle f,\varphi_{1}^{0}\rangle_{L^{2}}=\cdots=\langle f,\varphi_{d_{0}}^{0}\rangle_{L^{2}}=0

for a unique solution u=uf𝒞(M;E)u=u^{f}\in{\mathcal{C}}^{\infty}(M;E) defined by the condition that ufhEKer()u^{f}\perp_{h_{E}}\operatorname{Ker}(\not{D}). Associated to the equation (3.1) we can define the local source-to-source solution map LM,g,𝒪,L_{M,g,\mathcal{O},\ldots} by

L,𝒪(f):=uf|𝒪=(αf)|𝒪L_{\not{D},\mathcal{O}}(f):=u^{f}|_{\mathcal{O}}=(\not{D}^{-\alpha}f)|_{\mathcal{O}}

We can give an equivalent (spectral theoretic) definition for the Fractional Dirac operator via the heat kernel of the square of the Dirac operator. First note that we can define from the integral formula for the Gamma function,

Γ(α)=0ettα1dt,μα=1Γ(α)0etμtα1dt,\Gamma(\alpha)=\int_{0}^{\infty}e^{-t}t^{\alpha-1}dt,\implies\mu^{-\alpha}=\frac{1}{\Gamma(\alpha)}\int_{0}^{\infty}e^{-t\mu}t^{\alpha-1}dt,

thus we write

(ΔE)α=1Γ(α)0etΔEtα1𝑑t.(\Delta_{E})^{-\alpha}=\frac{1}{\Gamma(\alpha)}\int_{0}^{\infty}e^{-t\Delta_{E}}t^{\alpha-1}dt.

Using this formula, we have α=ΔEα12=ΔEα12\not{D}^{\alpha}=\not{D}\circ\Delta_{E}^{\frac{\alpha-1}{2}}=\Delta_{E}^{\frac{\alpha-1}{2}}\not{D}, i.e.

αu=1Γ(1α2)0t1α2etΔE(u)dtt\not{D}^{\alpha}u=\frac{1}{\Gamma(\frac{1-\alpha}{2})}\int_{0}^{\infty}t^{\frac{1-\alpha}{2}}e^{-t\Delta_{E}}(\not{D}u)\frac{dt}{t}

using the fact that etΔE=etΔE\not{D}e^{-t\Delta_{E}}=e^{-t\Delta_{E}}\not{D} (from uniqueness of the solution to the heat equation).

Using this definition of α\not{D}^{-\alpha} we can easily extend the proof of [12, Thm 1.5] to our setting:

Theorem 3.1.

Let 𝒟j:=(Mj,Ej,gj,hEj,Ej)\mathscr{D}_{j}:=(M_{j},E_{j},g_{j},h_{E_{j}},\nabla^{E_{j}}) for j=1,2j=1,2 be two Dirac bundles over closed Riemannian manifolds of dimension n2n\geq 2, then we denote by j\not{D}_{j} the generalized Dirac operator associated to 𝒟j\mathscr{D}_{j}. Let 𝒪jMj\mathcal{O}_{j}\subset M_{j} be non-empty open sets and assume that

(𝒪j,gj|𝒪j):=(𝒪,g)(\mathcal{O}_{j},g_{j}|_{\mathcal{O}_{j}}):=(\mathcal{O},g) (3.2)

and that there exists hermitian bundle isomorphism ϕ:E1|𝒪1E2|𝒪2\phi:E_{1}|_{\mathcal{O}_{1}}\to E_{2}|_{\mathcal{O}_{2}}. Assume furthermore that

L1,𝒪1(f)=L2,𝒪2(f)L_{\not{D}_{1},\mathcal{O}_{1}}(f)=L_{\not{D}_{2},\mathcal{O}_{2}}(f) (3.3)

for all f𝒞0(𝒪1,E1|𝒪)f\in{\mathcal{C}}^{\infty}_{0}(\mathcal{O}_{1},E_{1}|_{\mathcal{O}}) such that fL2(𝒪;E1|𝒪)Ker(1)f\perp_{L^{2}(\mathcal{O};E_{1}|_{\mathcal{O}})}\operatorname{Ker}(\not{D}_{1}), for any smooth extension of ff to M2M_{2}. Then

etΔE1(x,y)=etΔE2(x,y),x,y𝒪,t>0e^{-t\Delta_{E_{1}}}(x,y)=e^{-t\Delta_{E_{2}}}(x,y),\quad x,y\in\mathcal{O},\;t>0

Remark: The requirement for well-posed of the source-to-source solution maps that ff be orthogonal to the space Harmonic sections is subtle: the local bundle isomorphism ϕ\phi is already sufficient to ensure that dimKer(1)=dimKer(2)\text{dim}\;\operatorname{Ker}(\not{D}_{1})=\text{dim}\operatorname{Ker}(\not{D}_{2}). If fKer(2)f\in\operatorname{Ker}(\not{D}_{2}) then its restriction to 𝒪2\mathcal{O}_{2} pullbacks to a section of E1|𝒪1E_{1}|_{\mathcal{O}_{1}}, and in particular is also an element of Ker(1)\operatorname{Ker}(\not{D}_{1}) as the Dirac operator commutes with pullback. Similarly pullback by the inverse proves the reverse inclusion. The possibility that a non-trivial Harmonic section vanishes on the open sets in consideration is ruled out by the unique continuation principle for the Dirac operator: if f=0\not{D}f=0 and f|𝒪=0f|_{\mathcal{O}}=0 then f=0f=0 (see e.g. [3, Ch. 8]).

Proof.

Choosing ω1𝒪\omega_{1}\Subset\mathcal{O} non-empty and open, with ω2𝒪\omega_{2}\subset\mathcal{O} also non-empty open such that ω¯1ω¯2=\overline{\omega}_{1}\cap\overline{\omega}_{2}=\emptyset. For f𝒞0(𝒪;E|𝒪)f\in{\mathcal{C}}^{\infty}_{0}(\mathcal{O};E|_{\mathcal{O}}), due to (3.2), we have for all m=1,2,m=1,2,\ldots,

12m1f=22m1f=2m1fon ω1,\not{D}_{1}^{2m-1}f=\not{D}_{2}^{2m-1}f=\not{D}^{2m-1}f\quad\text{on $\omega_{1}$},

and 2m1f\not{D}^{2m-1}f is orthogonal to Ker(j)\operatorname{Ker}(\not{D}_{j}) for j=1,2j=1,2. Further from (3.3), for all m=1,2,3,m=1,2,3,\ldots that

(1α2m1f)|𝒪=(D2α2m1f)|𝒪,(\not{D}_{1}^{-\alpha}\not{D}^{2m-1}f)|_{\mathcal{O}}=(D_{2}^{-\alpha}\not{D}^{2m-1}f)|_{\mathcal{O}},

hence

0tα12((etΔE11\displaystyle\int_{0}^{\infty}t^{\frac{\alpha-1}{2}}((e^{-t\Delta_{E_{1}}}\not{D}_{1}- etΔE22)2m1f)(x)dtt=0tα12((etΔE1etΔE2)Δmf)(x)dtt=0\displaystyle e^{-t\Delta_{E_{2}}}\not{D}_{2})\not{D}^{2m-1}f)(x)\frac{dt}{t}=\int_{0}^{\infty}t^{\frac{\alpha-1}{2}}((e^{-t\Delta_{E_{1}}}-e^{-t\Delta_{E_{2}}})\Delta^{m}f)(x)\frac{dt}{t}=0
=0tα12tm((etΔE1etΔE2)f)(x)dtt\displaystyle=\int_{0}^{\infty}t^{\frac{\alpha-1}{2}}\partial_{t}^{m}((e^{-t\Delta_{E_{1}}}-e^{-t\Delta_{E_{2}}})f)(x)\frac{dt}{t}

for x𝒪,m=1,2,x\in\mathcal{O},\;m=1,2,\ldots, where we have again used (3.2) and that 2=ΔE\not{D}^{2}=\Delta_{E}. We next aim to integrate by parts mm-times, and observe that the boundary terms vanish due to a combination of classical Heat kernel estimates (both short and long time). The relevant Heat kernel estimates which hold for the Heat kernel of the Dirac Laplacian ΔEj\Delta_{E_{j}}, are

|etΔEj(x,y)|Ct(m+1)ecdgj(x,y)2t,0<t<1,x,yMj|e^{-t\Delta_{E_{j}}}(x,y)|\leq Ct^{-(m+1)}e^{-\frac{cdg_{j}(x,y)^{2}}{t}},\quad 0<t<1,\;x,y\in M_{j}

(has a slightly worse exponent in tt than what is known for the scalar case, but still suffices for our purposes), and

etΔEjL1LCtm/2,t>0||e^{-t\Delta_{E_{j}}}||_{L^{1}\to L^{\infty}}\leq Ct^{-m/2},\quad t>0

which is precisely the classical estimate of [35], equivalent to the scalar case. The first estimate can be found in [27, Thm 3.5]. Combining these two estimates, we can bound the integrand above, |t((etΔE1etΔE2)f)|(x)|\partial_{t}((e^{-t\Delta_{E_{1}}}-e^{-t\Delta_{E_{2}}})f)|(x), by functions vanishing at the endpoints of the integral. Proceeding to integrate by parts mm-times we have that

0tα12(m+1)((etΔE1etΔE2)f)(x)𝑑t=0\int_{0}^{\infty}t^{\frac{\alpha-1}{2}-(m+1)}((e^{-t\Delta_{E_{1}}}-e^{-t\Delta_{E_{2}}})f)(x)dt=0

and from here we can conclude that (χ[0,)φ)(s)\mathcal{F}(\chi_{[0,\infty)}\varphi)(s) is holomorphic with all derivatives vanishing at s=0s=0, for

φ(s)=((esΔE1esΔE2)f)(x)sα,\varphi(s)=\frac{((e^{s\Delta_{E_{1}}}-e^{s\Delta_{E_{2}}})f)(x)}{s^{\alpha}},

thus ((etΔE1etΔE2)f)(x)=0((e^{-t\Delta_{E_{1}}}-e^{-t\Delta_{E_{2}}})f)(x)=0 for all xω2x\in\omega_{2} and all t>0t>0. By unique continuation of the Heat equation this implies the same equality holds on all of 𝒪\mathcal{O}. Using that f𝒞0(ω1)f\in{\mathcal{C}}^{\infty}_{0}(\omega_{1}) was arbitrary we have

etΔE1(x,y)=etΔE2(x,y),x,y𝒪,t>0e^{-t\Delta_{E_{1}}}(x,y)=e^{-t\Delta_{E_{2}}}(x,y),\quad x,y\in\mathcal{O},t>0

as claimed. ∎

4. Reconstruction via the Wave equation

Have shown that the source-to-source solution operators determine the Heat kernel on the open neighborhood 𝒪\mathcal{O}. Using the Kannai transmutation formula

etΔEj(x,y)=1(4πt)1/2t0eτ4tsin(τ ΔEj )ΔEj 𝑑τt>0e^{-t\Delta_{E_{j}}}(x,y)=\frac{1}{(4\pi t)^{1/2}t}\int_{0}^{\infty}e^{-\frac{\tau}{4t}}\frac{\sin(\mathchoice{{\hbox{$\displaystyle\sqrt{\tau\,}$}\lower 0.4pt\hbox{\vrule height=4.30554pt,depth=-3.44446pt}}}{{\hbox{$\textstyle\sqrt{\tau\,}$}\lower 0.4pt\hbox{\vrule height=4.30554pt,depth=-3.44446pt}}}{{\hbox{$\scriptstyle\sqrt{\tau\,}$}\lower 0.4pt\hbox{\vrule height=3.01389pt,depth=-2.41113pt}}}{{\hbox{$\scriptscriptstyle\sqrt{\tau\,}$}\lower 0.4pt\hbox{\vrule height=2.15277pt,depth=-1.72223pt}}}\mathchoice{{\hbox{$\displaystyle\sqrt{\Delta_{E_{j}}\,}$}\lower 0.4pt\hbox{\vrule height=6.83331pt,depth=-5.46667pt}}}{{\hbox{$\textstyle\sqrt{\Delta_{E_{j}}\,}$}\lower 0.4pt\hbox{\vrule height=6.83331pt,depth=-5.46667pt}}}{{\hbox{$\scriptstyle\sqrt{\Delta_{E_{j}}\,}$}\lower 0.4pt\hbox{\vrule height=4.78333pt,depth=-3.82668pt}}}{{\hbox{$\scriptscriptstyle\sqrt{\Delta_{E_{j}}\,}$}\lower 0.4pt\hbox{\vrule height=3.41666pt,depth=-2.73334pt}}})}{\mathchoice{{\hbox{$\displaystyle\sqrt{\Delta_{E_{j}}\,}$}\lower 0.4pt\hbox{\vrule height=6.83331pt,depth=-5.46667pt}}}{{\hbox{$\textstyle\sqrt{\Delta_{E_{j}}\,}$}\lower 0.4pt\hbox{\vrule height=6.83331pt,depth=-5.46667pt}}}{{\hbox{$\scriptstyle\sqrt{\Delta_{E_{j}}\,}$}\lower 0.4pt\hbox{\vrule height=4.78333pt,depth=-3.82668pt}}}{{\hbox{$\scriptscriptstyle\sqrt{\Delta_{E_{j}}\,}$}\lower 0.4pt\hbox{\vrule height=3.41666pt,depth=-2.73334pt}}}}d\tau\quad t>0

the local determination of etΔEj(x,y)e^{-t\Delta_{E_{j}}}(x,y) and this equality imply

(sin(tΔE1 )ΔE1 f)(x)=(sin(tΔE2 )ΔE2 f)(x),\left(\frac{\sin(t\mathchoice{{\hbox{$\displaystyle\sqrt{\Delta_{E_{1}}\,}$}\lower 0.4pt\hbox{\vrule height=6.83331pt,depth=-5.46667pt}}}{{\hbox{$\textstyle\sqrt{\Delta_{E_{1}}\,}$}\lower 0.4pt\hbox{\vrule height=6.83331pt,depth=-5.46667pt}}}{{\hbox{$\scriptstyle\sqrt{\Delta_{E_{1}}\,}$}\lower 0.4pt\hbox{\vrule height=4.78333pt,depth=-3.82668pt}}}{{\hbox{$\scriptscriptstyle\sqrt{\Delta_{E_{1}}\,}$}\lower 0.4pt\hbox{\vrule height=3.41666pt,depth=-2.73334pt}}})}{\mathchoice{{\hbox{$\displaystyle\sqrt{\Delta_{E_{1}}\,}$}\lower 0.4pt\hbox{\vrule height=6.83331pt,depth=-5.46667pt}}}{{\hbox{$\textstyle\sqrt{\Delta_{E_{1}}\,}$}\lower 0.4pt\hbox{\vrule height=6.83331pt,depth=-5.46667pt}}}{{\hbox{$\scriptstyle\sqrt{\Delta_{E_{1}}\,}$}\lower 0.4pt\hbox{\vrule height=4.78333pt,depth=-3.82668pt}}}{{\hbox{$\scriptscriptstyle\sqrt{\Delta_{E_{1}}\,}$}\lower 0.4pt\hbox{\vrule height=3.41666pt,depth=-2.73334pt}}}}f\right)(x)=\left(\frac{\sin(t\mathchoice{{\hbox{$\displaystyle\sqrt{\Delta_{E_{2}}\,}$}\lower 0.4pt\hbox{\vrule height=6.83331pt,depth=-5.46667pt}}}{{\hbox{$\textstyle\sqrt{\Delta_{E_{2}}\,}$}\lower 0.4pt\hbox{\vrule height=6.83331pt,depth=-5.46667pt}}}{{\hbox{$\scriptstyle\sqrt{\Delta_{E_{2}}\,}$}\lower 0.4pt\hbox{\vrule height=4.78333pt,depth=-3.82668pt}}}{{\hbox{$\scriptscriptstyle\sqrt{\Delta_{E_{2}}\,}$}\lower 0.4pt\hbox{\vrule height=3.41666pt,depth=-2.73334pt}}})}{\mathchoice{{\hbox{$\displaystyle\sqrt{\Delta_{E_{2}}\,}$}\lower 0.4pt\hbox{\vrule height=6.83331pt,depth=-5.46667pt}}}{{\hbox{$\textstyle\sqrt{\Delta_{E_{2}}\,}$}\lower 0.4pt\hbox{\vrule height=6.83331pt,depth=-5.46667pt}}}{{\hbox{$\scriptstyle\sqrt{\Delta_{E_{2}}\,}$}\lower 0.4pt\hbox{\vrule height=4.78333pt,depth=-3.82668pt}}}{{\hbox{$\scriptscriptstyle\sqrt{\Delta_{E_{2}}\,}$}\lower 0.4pt\hbox{\vrule height=3.41666pt,depth=-2.73334pt}}}}f\right)(x), (4.1)

for all t>0t>0 and x𝒪x\in\mathcal{O}.

The benefit of this last equality is that the operator in (4.1) lets us represent the solution of the initial value problem,

{(t2ΔEj)uj(t,x)=f(t,x),(t,x)(0,)×Mjuj(0,x)=0,tuj(0,x)=0xMj\begin{cases}(\partial_{t}^{2}-\Delta_{E_{j}})u_{j}(t,x)=f(t,x),&(t,x)\in(0,\infty)\times M_{j}\\ u_{j}(0,x)=0,\partial_{t}u_{j}(0,x)=0&x\in M_{j}\\ \end{cases} (4.2)

for data f𝒞0(𝒪×Mj;Ej)f\in{\mathcal{C}}^{\infty}_{0}(\mathcal{O}\times M_{j};E_{j}) via the formula

uj(t,x)=0tsin(ts)ΔEj ΔEj f𝑑s.u_{j}(t,x)=\int_{0}^{t}\frac{\sin(t-s)\mathchoice{{\hbox{$\displaystyle\sqrt{\Delta_{E_{j}}\,}$}\lower 0.4pt\hbox{\vrule height=6.83331pt,depth=-5.46667pt}}}{{\hbox{$\textstyle\sqrt{\Delta_{E_{j}}\,}$}\lower 0.4pt\hbox{\vrule height=6.83331pt,depth=-5.46667pt}}}{{\hbox{$\scriptstyle\sqrt{\Delta_{E_{j}}\,}$}\lower 0.4pt\hbox{\vrule height=4.78333pt,depth=-3.82668pt}}}{{\hbox{$\scriptscriptstyle\sqrt{\Delta_{E_{j}}\,}$}\lower 0.4pt\hbox{\vrule height=3.41666pt,depth=-2.73334pt}}}}{\mathchoice{{\hbox{$\displaystyle\sqrt{\Delta_{E_{j}}\,}$}\lower 0.4pt\hbox{\vrule height=6.83331pt,depth=-5.46667pt}}}{{\hbox{$\textstyle\sqrt{\Delta_{E_{j}}\,}$}\lower 0.4pt\hbox{\vrule height=6.83331pt,depth=-5.46667pt}}}{{\hbox{$\scriptstyle\sqrt{\Delta_{E_{j}}\,}$}\lower 0.4pt\hbox{\vrule height=4.78333pt,depth=-3.82668pt}}}{{\hbox{$\scriptscriptstyle\sqrt{\Delta_{E_{j}}\,}$}\lower 0.4pt\hbox{\vrule height=3.41666pt,depth=-2.73334pt}}}}fds.

Associated to (4.2), we define the source-to-source solution map for the wave operator on 𝒪\mathcal{O} by

L𝒟j,𝒪jwave:𝒞0((0,)×𝒪)𝒞0([0,)×𝒪),L𝒟j,𝒪jwave(f)=uj|𝒪jL_{\mathscr{D}_{j},\mathcal{O}_{j}}^{\text{wave}}:{\mathcal{C}}^{\infty}_{0}((0,\infty)\times\mathcal{O})\to{\mathcal{C}}^{\infty}_{0}([0,\infty)\times\mathcal{O}),\quad L_{\mathscr{D}_{j},\mathcal{O}_{j}}^{\text{wave}}(f)=u_{j}|_{\mathcal{O}_{j}}

where uj(t,x)u_{j}(t,x) is the unique solution to (4.2). And by the equality in (4.1) we have equality of the source-to-solution maps for the wave operators for j=1,2j=1,2,

L𝒟1,𝒪1wave(f)=L𝒟2,𝒪2wave(f).L_{\mathscr{D}_{1},\mathcal{O}_{1}}^{\text{wave}}(f)=L_{\mathscr{D}_{2},\mathcal{O}_{2}}^{\text{wave}}(f).

Now we can move to the statement of our main theorem.

Theorem 4.1.

Let 𝒟j:=(Mj,Ej,gj,hEj,Ej)\mathscr{D}_{j}:=(M_{j},E_{j},g_{j},h_{E_{j}},\nabla^{E_{j}}) for j=1,2j=1,2 be two Dirac bundles over closed Riemannian manifolds of dimension n2n\geq 2, and let 𝒪jMj\mathcal{O}_{j}\subset M_{j} be non-empty open sets. Assuming there exists a diffeomorphism ψ:𝒪1𝒪2\psi:\mathcal{O}_{1}\to\mathcal{O}_{2} satisfying

L𝒟1,𝒪1wave(ψf)=ψ(L𝒟2,𝒪2wavef),f𝒞0((0,)×𝒪2;E).L_{\mathscr{D}_{1},\mathcal{O}_{1}}^{\text{wave}}(\psi^{*}f)=\psi^{*}(L_{\mathscr{D}_{2},\mathcal{O}_{2}}^{\text{wave}}f),\quad\forall f\in{\mathcal{C}}^{\infty}_{0}((0,\infty)\times\mathcal{O}_{2};E).

Then there is an isomorphism of Hermitian vector bundles Ψ:E1E2\Psi:E_{1}\to E_{2}, covering an isometry ψ:M1M2\psi:M_{1}\to M_{2} which coincides with Ψ|E1|𝒪1=ψ\Psi|_{E_{1}|_{\mathcal{O}_{1}}}=\psi.

We follow the proof of [17], (which itself was an application to closed manifolds of the boundary control method of Belishev [1] combined with the crucial unique continuation method of Tataru [32]) extended to the bundle case. The general structure of the proof is as follows

  1. (1)

    The source-to-source solution operator (𝒪,L𝒟,𝒪wave)(\mathcal{O},L_{\mathscr{D},\mathcal{O}}^{\text{wave}}) determines the distance function dgd_{g} on 𝒪×𝒪\mathcal{O}\times\mathcal{O}

  2. (2)

    The source-to-source solution data (𝒪,g|𝒪,hE|𝒪,L𝒟,𝒪wave)(\mathcal{O},g|_{\mathcal{O}},h_{E}|_{\mathcal{O}},L_{\mathscr{D},\mathcal{O}}^{\text{wave}}) determines the distance data (𝒪,g|𝒪,R(M))(\mathcal{O},g|_{\mathcal{O}},R(M))

  3. (3)

    The distance data (𝒪,g|𝒪,R(M))(\mathcal{O},g|_{\mathcal{O}},R(M)) determines the topology, smooth structure and Riemannian structure of (M,g)(M,g)

  4. (4)

    New: The distance data determines the isomorphism class of EME\to M and Hermitian metric hEh_{E}

  5. (5)

    New: The distance data determines the Hermitian connection E\nabla^{E}, and thus the homomorphism of clifford multiplication from the identity [E,𝒸𝓁(θ)]=𝒸𝓁(θ)[\nabla^{E},\mathpzc{cl}(\theta)]=\mathpzc{cl}(\nabla^{g}\theta)

This uses several facts about the solution to the linear wave equation on sections of EE: the first is its finite speed of propagation.

Theorem 4.2.

Let T>0T>0, and pMp\in M be open and define the open cone

CT,p:={(t,x)(0,T)×M:dg(x,p)<Tt}.C_{T,p}:=\{(t,x)\in(0,T)\times M:d_{g}(x,p)<T-t\}.

Let fL2(×M;E)f\in L^{2}(\mathbb{R}\times M;E) and suppose uu solves

{(t2ΔE)u=f(0,)×Mf|CT,p=0u|{t=0}×BT(p)=tu|{t=0}×BT(p)=0\begin{cases}(\partial_{t}^{2}-\Delta_{E})u=f&(0,\infty)\times M\\ f|_{C_{T,p}}=0\\ u|_{\{t=0\}\times B_{T}(p)}=\partial_{t}u|_{\{t=0\}\times B_{T}(p)}=0\end{cases}

then u|CT,p=0u|_{C_{T,p}}=0.

To give the statement of the relevant unique continuation principle we also define

M(T,𝒪)={xM:dg(x,𝒪)T}M(T,\mathcal{O})=\{x\in M:d_{g}(x,\mathcal{O})\leq T\}

for the domain of dependence of the wave equation. As a result of the Carleman estimates established in [11], the proof of the unique continuation theorem is as in the scalar case since the wave equation being considered is principally scalar (i.e. t2ΔE\partial_{t}^{2}-\Delta_{E}) see section 2.5 of [20],

Theorem 4.3.

Let T>0T>0, and 𝒪M\mathcal{O}\subset M be open and bounded. Let u𝒞0(×M;E|𝒪)u\in{\mathcal{C}}^{\infty}_{0}(\mathbb{R}\times M;E|_{\mathcal{O}}). Let

{(t2ΔE)u=0(0,2T)×M(T,𝒪)u|(0,2T)×𝒪0\begin{cases}(\partial_{t}^{2}-\Delta_{E})u=0&(0,2T)\times M(T,\mathcal{O})\\ u|_{(0,2T)\times\mathcal{O}}\equiv 0\end{cases}

Then u|C(T,𝒪)0u|_{C(T,\mathcal{O})}\equiv 0, for

C(T,𝒪)={(t,x)(0,2T)×N:dg(x,𝒪)min{t,2Tt}}C(T,\mathcal{O})=\{(t,x)\in(0,2T)\times N:d_{g}(x,\mathcal{O})\leq\min\{t,2T-t\}\}

Given the proper form of unique continuation we should be able to prove density of solutions with sources from the set,

(T,𝒪)={f𝒞0(×M;E)|supp(f)(0,T)×𝒪}\mathcal{F}(T,\mathcal{O})=\{f\in{\mathcal{C}}^{\infty}_{0}(\mathbb{R}\times M;E)|\operatorname{supp}(f)\subset(0,T)\times\mathcal{O}\}
Theorem 4.4 (Approximate controllability).

Let 𝒪M\mathcal{O}\subset M be open and bounded. For T>0T>0 the set

𝒲T={Wf(T,):f(T,𝒪)}\mathcal{W}_{T}=\{Wf(T,\cdot):f\in\mathcal{F}(T,\mathcal{O})\}

is dense in L2(M(T,𝒪);E)L^{2}(M(T,\mathcal{O});E). Further, by considering time reparametrization, we obtain that {Wf(T,):f𝒞0((Tr,T)×𝒪;E)\{Wf(T,\cdot):f\in{\mathcal{C}}^{\infty}_{0}((T-r,T)\times\mathcal{O};E) is dense in L2(M(r,𝒪);E)L^{2}(M(r,\mathcal{O});E) for all r>0r>0.

Proof.

From the finite speed of propagation we have that 𝒲TL2(M(T,𝒪);E)\mathcal{W}_{T}\subset L^{2}(M(T,\mathcal{O});E). Thus it suffices to show that the orthogonal complement of 𝒲T\mathcal{W}_{T} contains only the origin. Let ϕL2(M(T,𝒪);E)\phi\in L^{2}(M(T,\mathcal{O});E) satisfy Wf(T,),ϕL2(M;E)=0\langle Wf(T,\cdot),\phi\rangle_{L^{2}(M;E)}=0 for all f(T,𝒪)f\in\mathcal{F}(T,\mathcal{O}). Let u𝒞(×M)u\in{\mathcal{C}}^{\infty}(\mathbb{R}\times M) solve

{(t2ΔE)u=0,(0,T)×Mu|t=T=0tu|t=T=ϕ\begin{cases}(\partial_{t}^{2}-\Delta_{E})u=0,&(0,T)\times M\\ u|_{t=T}=0\;\partial_{t}u|_{t=T}=\phi\end{cases}

From Green’s identities we have

f,uL2((0,T)×M;E)=(t2ΔE)Wf,uL2((0,T)×M;E)Wf,(t2ΔE)uL2((0,T)×M;E)=0\langle f,u\rangle_{L^{2}((0,T)\times M;E)}=\langle(\partial_{t}^{2}-\Delta_{E})Wf,u\rangle_{L^{2}((0,T)\times M;E)}-\langle Wf,(\partial_{t}^{2}-\Delta_{E})u\rangle_{L^{2}((0,T)\times M;E)}=0

thus u0u\equiv 0 on (0,T]×𝒪(0,T]\times\mathcal{O} by density of (T,𝒪)L2((0,T)×𝒪;E)\mathcal{F}(T,\mathcal{O})\subset L^{2}((0,T)\times\mathcal{O};E). Extending uu across t=Tt=T by the odd reflection u(t,x)=u(2Tt,x)u(t,x)=-u(2T-t,x), and denoting the extension by UU we have that it satisfies

{(t2ΔE)U=0(0,2T)×MU|t=T=0,tU|t=T=tu|t=T=ϕ\begin{cases}(\partial_{t}^{2}-\Delta_{E})U=0&(0,2T)\times M\\ U|_{t=T}=0,\;\partial_{t}U|_{t=T}=\partial_{t}u|_{t=T}=\phi\end{cases}

by our odd reflection, thus U|(0,2T)×𝒪0U|_{(0,2T)\times\mathcal{O}}\equiv 0. Now by theorem 4.3 we conclude that U|C(T,𝒪)0U|_{C(T,\mathcal{O})}\equiv 0, in particular since {T}×M(T,𝒪)C(T,𝒪)\{T\}\times M(T,\mathcal{O})\subset C(T,\mathcal{O}) we have ϕ|M(T,𝒪)=tU|{T}×M(T,𝒪)0\phi|_{M(T,\mathcal{O})}=\partial_{t}U|_{\{T\}\times M(T,\mathcal{O})}\equiv 0 as claimed. ∎

Having proven this fact, the proofs of 1)  2) and 3) from [17] generalize immediately to the bundle-valued case. One lemma they depend on is the Blagovestchenskii identity:

Lemma 4.5 (Blagovestchenskii Identity).

Let (M,g)(M,g) be complete. Let T>0T>0 and 𝒪M\mathcal{O}\subset M open and bounded. Let f,h(2T,𝒪)f,h\in\mathcal{F}(2T,\mathcal{O}), then

Wf(T,),Wg(T,)L2(M;E)=f,(JL𝒟,𝒪wave(L𝒟,𝒪wave)J)gL2((0,T)×M;E)\langle Wf(T,\cdot),Wg(T,\cdot)\rangle_{L^{2}(M;E)}=\langle f,(JL_{\mathscr{D},\mathcal{O}}^{\text{wave}}-(L_{\mathscr{D},\mathcal{O}}^{\text{wave}})^{*}J)g\rangle_{L^{2}((0,T)\times M;E)}

where J:L2((0,2T);E)L2((0,T);E)J:L^{2}((0,2T);E)\to L^{2}((0,T);E) is the time averaging operator Jϕ(t)=12t2Ttϕ(s)𝑑sJ\phi(t)=\tfrac{1}{2}\int_{t}^{2T-t}\phi(s)ds.

Another such lemma is that the source-to-source solution operator determines the distance function:

Lemma 4.6.

Let (M,g)(M,g) and 𝒪M\mathcal{O}\subset M be open and bounded. Then (𝒪,L𝒟,𝒪wave)(\mathcal{O},L_{\mathscr{D},\mathcal{O}}^{\text{wave}}) determines the distance function dgd_{g} on 𝒪×𝒪\mathcal{O}\times\mathcal{O}.

Proof.

Set x,y𝒪x,y\in\mathcal{O}, and choose an auxiliary metric d0d_{0} on 𝒪\mathcal{O} which induces the same metric space topology on 𝒪\mathcal{O} as gg. Let ε>0\varepsilon>0 and consider x,ε:=(0,)×Bεd0(x)\mathcal{B}_{x,\varepsilon}:=(0,\infty)\times B_{\varepsilon}^{d_{0}}(x) and

tε:=inf{t>0:f𝒞0(x,ε;E),supp(L𝒟,𝒪wavef)(t,)Bεd0(y)}.t_{\varepsilon}:=\inf\{t>0:\exists f\in{\mathcal{C}}^{\infty}_{0}(\mathcal{B}_{x,\varepsilon};E),\;\operatorname{supp}{(L_{\mathscr{D},\mathcal{O}}^{\text{wave}}f)}(t,\cdot)\cap B_{\varepsilon}^{d_{0}}(y)\neq\emptyset\}.

From finite propagation speed and lemma (4.4) we have

tε=distg(Bεd0(x),Bεd0(y))t_{\varepsilon}=\text{dist}_{g}(B_{\varepsilon}^{d_{0}}(x),B_{\varepsilon}^{d_{0}}(y))

and thus limε0tε=dg(x,y)\lim_{\varepsilon\to 0}t_{\varepsilon}=d_{g}(x,y). ∎

Let γy,ξ(t)\gamma_{y,\xi}(t) be the time-tt flow of the unique unit speed geodesic starting at yy with initial velocity vector ξ\xi. Several facts about geometric determination in our setting will rely on the notion of cut time,

τ(y,ξ)=sup{t>0:dg(y,γy,ξ(t))=t}.\tau(y,\xi)=\sup\{t>0:d_{g}(y,\gamma_{y,\xi}(t))=t\}.

Note that we have

M={γy,ξ(t)M:y𝒪,ξSyM,t<τ(y,ξ)}M=\{\gamma_{y,\xi}(t)\in M:y\in\mathcal{O},\;\xi\in S_{y}M,\;t<\tau(y,\xi)\} (4.3)

for 𝒪M\mathcal{O}\subset M open and non-empty. In other words any point in MM can be reached by geodesics originating in 𝒪\mathcal{O} which do not meet the cut locus.

Further, the cut distance and distance between points can be determined using only the source-to-source solution data

(𝒪,g|𝒪,hE|𝒪,L𝒟,𝒪wave)(\mathcal{O},g|_{\mathcal{O}},h_{E}|_{\mathcal{O}},L_{\mathscr{D},\mathcal{O}}^{\text{wave}})

namely as shown by [17] in the scalar case,

Proposition 4.7.

For any y𝒪y\in\mathcal{O}, ξSyM\xi\in S_{y}M we can find the cut time τ(y,ξ)\tau(y,\xi) from the source-to-source solution data (𝒪,g|𝒪,hE|𝒪,L𝒟,𝒪wave)(\mathcal{O},g|_{\mathcal{O}},h_{E}|_{\mathcal{O}},L_{\mathscr{D},\mathcal{O}}^{\text{wave}}).

Further, given z,y𝒪z,y\in\mathcal{O}, ξTy𝒪\xi\in T_{y}\mathcal{O}, η=1||\eta||=1 and r<τ(y,η)r<\tau(y,\eta). Then (𝒪,g|𝒪,hE|𝒪,L𝒟,𝒪wave)(\mathcal{O},g|_{\mathcal{O}},h_{E}|_{\mathcal{O}},L_{\mathscr{D},\mathcal{O}}^{\text{wave}}) determines dg(p,z)d_{g}(p,z) where γy,ξ(r)=z\gamma_{y,\xi}(r)=z.

Proof.

For y,ξy,\xi fixed the geodesic segment γy,ξ([0,s])\gamma_{y,\xi}([0,s]) is determined by the given data for ss small. Choosing s>0s>0 sufficiently small that γy,ξ([0,s])𝒪\gamma_{y,\xi}([0,s])\subset\mathcal{O}, we observe that the condition

there exists ε>0\varepsilon>0 such that Br+ε(x)Bs+r(y)¯B_{r+\varepsilon}(x)\subset\overline{B_{s+r}(y)} (4.4)

determines the cut distance by the forumla,

τ(y,ξ)=inf{s+r>0:r,s>0,γy,ξ([0,s])𝒪, (4.4) holds}.\tau(y,\xi)=\inf\{s+r>0:r,s>0,\gamma_{y,\xi}([0,s])\subset\mathcal{O},\text{ \eqref{ball-cond} holds}\}.

This follows from its contrapositive, arguing via geodesic continuation of the arc γy,ξ([0,s])\gamma_{y,\xi}([0,s]) and the triangle inequality. So τ(y,ξ)\tau(y,\xi) is determined by relation (4.4).

Define Sε(x,r)=(T(rε),T)×Bε(x)S_{\varepsilon}(x,r)=(T-(r-\varepsilon),T)\times B_{\varepsilon}(x). Using finite propagation speed and lemma (4.4) we can show that for p,y,zMp,y,z\in M, ε>0\varepsilon>0, and p,y,z>ε\ell_{p},\ell_{y},\ell_{z}>\varepsilon the condition that

Bp(p)By(y)Bz(z)¯B_{\ell_{p}}(p)\subset\overline{B_{\ell_{y}}(y)\cup B_{\ell_{z}}(z)} (4.5)

is equivalent to the existence of {fj}𝒞0(Sε(y,y)Sε(z,z);E)\{f_{j}\}\subset{\mathcal{C}}^{\infty}_{0}(S_{\varepsilon}(y,\ell_{y})\cup S_{\varepsilon}(z,\ell_{z});E) for every f𝒞0(Sε(p,p);E)f\in{\mathcal{C}}^{\infty}_{0}(S_{\varepsilon}(p,\ell_{p});E) such that

Wf(T,)Wfj(T,)L2(M;E)j0.||Wf(T,\cdot)-Wf_{j}(T,\cdot)||_{L^{2}(M;E)}\xrightarrow{j\to\infty}0. (4.6)

Now choosing ε>0\varepsilon>0 sufficiently small that Bε(y)Bε(z)𝒪B_{\varepsilon}(y)\cup B_{\varepsilon}(z)\subset\mathcal{O}, and setting y=zy=z, y=z=s+r\ell_{y}=\ell_{z}=s+r, and p=r+ε\ell_{p}=r+\varepsilon, we see that relation (4.4) is equivalent to (4.6). However, using the Blagovestchenskii identity we see that (𝒪,g|𝒪,hE|𝒪,L𝒟,𝒪wave)(\mathcal{O},g|_{\mathcal{O}},h_{E}|_{\mathcal{O}},L_{\mathscr{D},\mathcal{O}}^{\text{wave}}) determines (4.6) and thus the cut time as claimed.

The proof of the second claim follows similarly, by considering s(0,r)s\in(0,r) such that γy,ξ([0,s])𝒪\gamma_{y,\xi}([0,s])\subset\mathcal{O}, and set γy,ξ(s)=p\gamma_{y,\xi}(s)=p. Then we can show that

dg(p,z)=infR>0{(4.5)holds with p=rs+εy=rz=R for some R and ε>0}d_{g}(p,z)=\inf_{R>0}\{\eqref{3ell}\text{holds with $\ell_{p}=r-s+\varepsilon$, $\ell_{y}=r$, $\ell_{z}=R$ for some $R$ and $\varepsilon>0$}\}

by first observing that dg(p,z)d_{g}(p,z) is less than or equal to this infimum (since r<τ(y,ξ)r<\tau(y,\xi)), and then showing this infimum is attained by a similar argument as above, using the Blagovestchenskii identity and lemma 4.4. From this formulation of dg(p,z)d_{g}(p,z) we have that it is determined by the source-to-solution data. ∎

Using this last proposition and the fact of (4.3)\eqref{exp-closure}, we have that (𝒪,g|𝒪,hE|𝒪,L𝒟,𝒪wave)(\mathcal{O},g|_{\mathcal{O}},h_{E}|_{\mathcal{O}},L_{\mathscr{D},\mathcal{O}}^{\text{wave}}) determines the family of distance functions,

R(M):={dg(x,)|𝒪:xM}𝒞0(𝒪¯).R(M):=\{d_{g}(x,\cdot)|_{\mathcal{O}}:x\in M\}\subset\mathcal{C}^{0}(\overline{\mathcal{O}}).

In Helin Lassas Oksanen Saksala they prove that this set can be topologized as a smooth Riemannian manifold, isometric to (M,g)(M,g), and serving as a background space depending only on the behavior of solutions in 𝒪\mathcal{O}, they show the two Riemannian manifolds are isometric. Their proof, depending only on the distance functions, has no dependence on whether or not the given wave equation is scalar or bundle-valued, and thus we obtain:

Proposition 4.8.

Let 𝒟j:=(Mj,Ej,gj,hEj,Ej)\mathscr{D}_{j}:=(M_{j},E_{j},g_{j},h_{E_{j}},\nabla^{E_{j}}) for j=1,2j=1,2 be two Dirac bundles over closed Riemannian manifolds of dimension n2n\geq 2, and let 𝒪jMj\mathcal{O}_{j}\subset M_{j} be non-empty open sets. Assuming there exists a diffeomorphism ψ:𝒪1𝒪2\psi:\mathcal{O}_{1}\to\mathcal{O}_{2} satisfying

L𝒟1,𝒪1wave(ψf)=ψ(L𝒟2,𝒪2wavef),f𝒞0((0,)×𝒪2;E).L_{\mathscr{D}_{1},\mathcal{O}_{1}}^{\text{wave}}(\psi^{*}f)=\psi^{*}(L_{\mathscr{D}_{2},\mathcal{O}_{2}}^{\text{wave}}f),\quad\forall f\in{\mathcal{C}}^{\infty}_{0}((0,\infty)\times\mathcal{O}_{2};E).

Then (M1,g1)(M_{1},g_{1}) and (M2,g2)(M_{2},g_{2}) are isometric Riemannian manifolds.

Having shown the two Riemannian manifolds are isometric, it remains to show that we can also recover the Hermitian bundle structure and connection.

Lemma 4.9.

Let f,h(2T,𝒪)f,h\in\mathcal{F}(2T,\mathcal{O}), for T>0,𝒪MT>0,\;\mathcal{O}\subset M open. Then L𝒟,𝒪waveL_{\mathscr{D},\mathcal{O}}^{\text{wave}} determines the inner products

Wf(T,),Wg(T,)L2(M;E).\langle Wf(T,\cdot),Wg(T,\cdot)\rangle_{L^{2}(M;E)}.

Further, for any xM(T,𝒪)x\in M(T,\mathcal{O}) there exists functions

{g}=1RkE(2T,𝒪)\{g_{\ell}\}_{\ell=1}^{\emph{Rk}\;E}\subset\mathcal{F}(2T,\mathcal{O})

such that {Wg(T,x))}=1RkE\{Wg_{\ell}(T,x))\}_{\ell=1}^{\emph{Rk}\;E} forms an orthonormal basis of ExE_{x}.

Proof.

The first claim follows immediately from the Blagovestchenskii identity above. For the second claim we note that it suffices to prove ExE_{x} is spanned by the vectors Wg(T,x)Wg(T,x), for g(T,𝒪)g\in\mathcal{F}(T,\mathcal{O}). Thus we are left to show that if eExe\in E_{x} and

e,Wg(T,x)hE=0,g(T,𝒪)\langle e,Wg(T,x)\rangle_{h_{E}}=0,\quad g\in\mathcal{F}(T,\mathcal{O})

then e=0e=0. By duality, applying W:Hs(M;E)Hs((0,T)×M;E)W^{*}:H^{-s}(M;E)\to H^{-s^{\prime}}((0,T)\times M;E) to eδxe\delta_{x} gives a solution to the initial value problem

{(t2ΔE)u=0u|t=T=0,tu|t=T=eδx,\begin{cases}(\partial_{t}^{2}-\Delta_{E})u=0\\ u|_{t=T}=0,\;\partial_{t}u|_{t=T}=e\delta_{x}\end{cases},

i.e. W(eδx)=uW^{*}(e\delta_{x})=u. Hence we have tu(t,x),g(t,x)hE=0t\mapsto\langle u(t,x),g(t,x)\rangle_{h_{E}}=0 for all t(0,T)t\in(0,T). Combining this with the initial condition on uu we see u|(0,T]×M0u|_{(0,T]\times M}\equiv 0. Extending uu by the odd reflection across t=Tt=T we obtain a section now satisfying (t2ΔE)u=0(\partial_{t}^{2}-\Delta_{E})u=0 on (0,2T)×M(0,2T)\times M, thus by 4.3 we have e=0e=0 as claimed. ∎

In particular this implies that hEh_{E} restricted to E|M(T,𝒪)E|_{M(T,\mathcal{O})} for T>0T>0 sufficiently small is determined by L𝒟,𝒪waveL_{\mathscr{D},\mathcal{O}}^{\text{wave}}.

Next we show how to construct sections of EE out of double sequences of functions supported in (Tr,T)×𝒪(T-r,T)\times\mathcal{O}. The following lemma is adapted from [21]:

Lemma 4.10.

Let 𝒪M\mathcal{O}\subset M be open and let xMx\in M. Then there exists y𝒪,ξSyMy\in\mathcal{O},\xi\in S_{y}M such that γy,ξ(t)=x\gamma_{y,\xi}(t)=x for t<τ(y,ξ)t<\tau(y,\xi). We define tk=t+1kt_{k}=t+\tfrac{1}{k}, and set

Yk=𝒪B1/k(y),Xk=Mθ<1/k(Yk,tk)M(𝒪,t).Y_{k}=\mathcal{O}\cap B_{1/k}(y),\quad X_{k}=M_{\theta<1/k}(Y_{k},t_{k})\setminus M(\mathcal{O},t).

where

Mθ<1/k(Yk,tk):={zM(Yk,tk):y~Yk,dg(z,y~)<t1k,ηTyM,γy,η(1k)=y,θ(ξ,η)<1k}M_{\theta<1/k}(Y_{k},t_{k}):=\{z\in M(Y_{k},t_{k}):\exists\widetilde{y}\in Y_{k},d_{g}(z,\widetilde{y})<t-\tfrac{1}{k},\;\exists\eta\in T_{y}M,\gamma_{y,\eta}(\tfrac{1}{k})=y,\;\theta(\xi,\eta)<\tfrac{1}{k}\}

with θ(,)\theta(\cdot,\cdot) equal to the angle between two vectors in TyMT_{y}M with respect to the inner product induced by gg. Suppose Φx={fjk}j,k=1𝒞((Ttk,T)×Yk;E)\Phi^{x}=\{f_{jk}\}_{j,k=1}^{\infty}\subset{\mathcal{C}}^{\infty}((T-t_{k},T)\times Y_{k};E) satisfies

  1. (1)

    For each k=1,2,k=1,2,\ldots the sequence {Wfjk(T,)}\{Wf_{jk}(T,\cdot)\} converges weakly in L2(M;E)L^{2}(M;E) to function supported in XkX_{k}.

  2. (2)

    There exists C>0C>0 such that

    Wfjk(T,)L2(M;E)CVol(Xk)1/2||Wf_{jk}(T,\cdot)||_{L^{2}(M;E)}\leq\frac{C}{\emph{Vol}(X_{k})^{1/2}}
  3. (3)

    The limit limj,kWfjk(T,),Wg(T,)L2(M;E)\lim_{j,k\to\infty}\langle Wf_{jk}(T,\cdot),Wg(T,\cdot)\rangle_{L^{2}(M;E)} exists for every g(2T,𝒪)g\in\mathcal{F}(2T,\mathcal{O}).

Then there is a vector e(x;Φx)Exe(x;\Phi^{x})\in E_{x} such that

limj,kWfjk(T,),ϕL2(M;E)=e(x;Φx),ϕ(x)hE,ϕ𝒞(M;E)\lim_{j,k\to\infty}\langle Wf_{jk}(T,\cdot),\phi\rangle_{L^{2}(M;E)}=\langle e(x;\Phi^{x}),\phi(x)\rangle_{h_{E}},\quad\phi\in{\mathcal{C}}^{\infty}(M;E)

This lemma is a consequence of lemma (4.9), and otherwise the proof is the same as given in [21], using the fact that from its construction diam(Xk)0\text{diam}(X_{k})\to 0 and XkxX_{k}\to x. Given this criterion for constructing vectors of eExe\in E_{x} we observe that such a double sequence Φx={fjk}j,k=1\Phi^{x}=\{f_{jk}\}_{j,k=1}^{\infty} satisfying the hypotheses of lemma (4.10) exists by applying lemma (4.4) to construct a sequence for each k=1,2,k=1,2,\ldots

Wfjkj1XkeVolXkWf_{jk}\xrightarrow{j\to\infty}\frac{1_{X_{k}}e}{\operatorname{Vol}{X_{k}}}

where 1Xk1_{X_{k}} is the indicator function of XkX_{k}. This sequence clearly satisfies the conclusions of lemma (4.10), as claimed. Further, if we choose Φx\Phi^{x} for xUx\in U such that

xe(x;Φx),Wg(T,)hE,g(2T,𝒪)x\mapsto\langle e(x;\Phi^{x}),Wg(T,\cdot)\rangle_{h_{E}},\quad g\in\mathcal{F}(2T,\mathcal{O})

is smooth, then the assignment e(x)=e(x;Φx)e(x)=e(x;\Phi^{x}) will be a smooth section of E|UE|_{U} as this implies ee depends smoothly in xx with respect to an orthonormal frame {Wg(T,)}=1RkE\{Wg_{\ell}(T,\cdot)\}_{\ell=1}^{\text{Rk}E} by lemma (4.9).

Theorem 4.11.

Let 𝒪M\mathcal{O}\subset M be open, non-empty, and bounded. Then the Dirac bundle 𝒟\mathscr{D} comprised of (M,g)(M,g), Hermitian vector bundle with connection (E,hE,E)(E,h_{E},\nabla^{E}) over MM, and clifford multiplication 𝒸𝓁:𝓁(𝒯,)End\mathpzc{cl}:\mathbb{C}l(T^{*}M,g)\to\operatorname{End}{E} are all determined by the source-to-source solution operator L𝒟,𝒪waveL_{\mathscr{D},\mathcal{O}}^{\text{wave}} and the Hermitian vector bundle E|𝒪E|_{\mathcal{O}}.

Proof.

Having already shown the Riemannian manifold (M,g)(M,g) is determined by L𝒟,𝒪waveL_{\mathscr{D},\mathcal{O}}^{\text{wave}}, so it remains to show the given data also determines the Hermitian bundle with connection. By making 𝒪\mathcal{O} smaller if needed, we may assume that 𝒪M\mathcal{O}\subset M is contractible and thus E|𝒪E|_{\mathcal{O}} is trivial.

For each x𝒪x\in\mathcal{O} we can choose a double sequence Φx={fjkx}j,k=1\Phi^{x}=\{f_{jk}^{x}\}_{j,k=1}^{\infty} satisfying the criteria of lemma 4.10 and hence converges to a prescribed eExe\in E_{x}. If we further require that this family (Φx)x𝒪(\Phi^{x})_{x\in\mathcal{O}} be chosen such that

xe(x;Φx),Wg(T,)hE,g(2T,𝒪),x\mapsto\langle e(x;\Phi^{x}),Wg(T,\cdot)\rangle_{h_{E}},\quad g\in\mathcal{F}(2T,\mathcal{O}),

is smooth, then e(x)=e(x;Φx)e(x)=e(x;\Phi^{x}) will be smooth in 𝒪\mathcal{O} as well.

Beginning with an orthonormal frame {e(x)}\{e_{\ell}(x)\} for E|𝒪E|_{\mathcal{O}}, we can similarly choose for each of the =1,,RkE\ell=1,\ldots,\text{Rk}\;E families of double sequences Φx\Phi_{\ell}^{x} representing e(x)e_{\ell}(x). Thus the coefficient functions hi,j(x)=ei(x),ej(x)hE=δij(x)h_{i,j}(x)=\langle e_{i}(x),e_{j}(x)\rangle_{h_{E}}=\delta_{ij}(x) are determined by L𝒟,𝒪waveL_{\mathscr{D},\mathcal{O}}^{\text{wave}}, and in this trivialization E|𝒪E|_{\mathcal{O}} we can express the Hermitian metric as v,whE=ve(x),wκeκ(x)hE=vw¯\langle v,w\rangle_{h_{E}}=\langle v^{\ell}e_{\ell}(x),w^{\kappa}e_{\kappa}(x)\rangle_{h_{E}}=\sum v^{\ell}\overline{w^{\ell}}.

To determine the bundle and metric globally we need only determine the bundle transition functions. From compactness injg>c\operatorname{inj}_{g}>c, and thus there exists a finite cover {Bc(pi)}i=1N\{B_{c}(p_{i})\}_{i=1}^{N} of MM. Further, for each xMx\in M there exists y𝒪y\in\mathcal{O} and ξSyM\xi\in S_{y}M such that γy,ξ(t)=x\gamma_{y,\xi}(t)=x with t<τ(y,ξ)t<\tau(y,\xi). Since each E|Bc(pi)Bc(pi)×EpiE|_{B_{c}(p_{i})}\simeq B_{c}(p_{i})\times E_{p_{i}} is trivial, it remains only to determine the transition functions Tij𝒞(Bc(pi)Bc(pj);End(Ep))T^{ij}\in{\mathcal{C}}^{\infty}(B_{c}(p_{i})\cap B_{c}(p_{j});\operatorname{End}(E_{p})). With respect to the chosen orthonormal frame e(x)e_{\ell}(x) over Bc(pi)Bc(pj)B_{c}(p_{i})\cap B_{c}(p_{j}) we can write this matrix valued function Tij(x)=(T,κij(x)),κ=1RkET^{ij}(x)=(T_{\ell,\kappa}^{ij}(x))_{\ell,\kappa=1}^{\text{Rk}E} in terms of the frame, i.e. by using the local expression of the metric hEh_{E} over Bc(pi)Bc(pj)B_{c}(p_{i})\cap B_{c}(p_{j}). The frame e(x)e_{\ell}(x), being generated by applying lemma 4.10 to points xBc(pi)Bc(pj)x\in B_{c}(p_{i})\cap B_{c}(p_{j}), is thus determined by L𝒟,𝒪waveL_{\mathscr{D},\mathcal{O}}^{\text{wave}} as above. Thus we have recovered the transition functions, hence the entire Hermitian bundle and metric.

Having determined the transition functions, to determine the connection it suffices to determine E\nabla^{E} on a trivialization E|UU×EpE|_{U}\simeq U\times E_{p} where E=d+A\nabla^{E}=d+A, for AΩ1(U;E)A\in\Omega^{1}(U;E). This portion of the argument follows that of Kurylev Oksanen Paternain and we include it for completeness. Choose T>0T>0 sufficiently large that UM(2T,𝒪)U\subset M(2T,\mathcal{O}). Since Wf(T,)Wf(T,\cdot) is determined by L𝒟,𝒪waveL_{\mathscr{D},\mathcal{O}}^{\text{wave}} for all f(2T,𝒪)f\in\mathcal{F}(2T,\mathcal{O}), and the wave equation (4.2) is time-translation invariant, we have also determined Wf(t,)Wf(t,\cdot) for t(0,T)t\in(0,T). Differentiating twice in tt gives ΔEWf(t,)\Delta_{E}Wf(t,\cdot) for t(0,T)t\in(0,T). Hence by duality

f(T,),ϕL2(M;E)=Wf(T,),ΔEϕL2(M;E)ϕ𝒞0(U;E),\langle f(T,\cdot),\phi\rangle_{L^{2}(M;E)}=\langle Wf(T,\cdot),\Delta_{E}\phi\rangle_{L^{2}(M;E)}\quad\phi\in{\mathcal{C}}^{\infty}_{0}(U;E),

thus by lemma (4.4) we can determine and densely approximate L2(U;E)L^{2}(U;E) by the functions Wf(T,)Wf(T,\cdot) thus determining ΔEϕ\Delta_{E}\phi. Now, because ΔE\Delta_{E} differs from (E)E(\nabla^{E})^{*}\nabla^{E} by a zeroth order term, and E=d+A\nabla^{E}=d+A over UU, we compute

(E)Eϕ=ddϕ2(A,dϕ)+(dA)u(A,Au),(\nabla^{E})^{*}\nabla^{E}\phi=d^{*}d\phi-2(A,d\phi)+(d^{*}A)u-(A,Au),

and conclude that the principal part of (E)E(\nabla^{E})^{*}\nabla^{E} is ddd^{*}d and its first order part is 2(A,dϕ)=2gjkAjkϕdxk2(A,d\phi)=2g^{jk}A_{j}\partial_{k}\phi dx^{k}. Because the principal part depends only on the metric gg, we can determine ddd^{*}d in UU. For xUx\in U we consider ϕ(x)=ϕk(x)\phi(x)=\phi_{\ell}^{k}(x) such that ϕ(x)=0\phi(x)=0 and jϕ(x)=δjke(x)\partial_{j}\phi(x)=\delta_{jk}e_{\ell}(x). Thus we can determine the endomorphism piece AA of the connection from the first order term in (E)Eϕ(x)(\nabla^{E})^{*}\nabla^{E}\phi(x), i.e. determine AΩ1(U;E)A\in\Omega^{1}(U;E) from the determination of

((E)Edd)ϕ=2(A,dϕ)=2gik(x)Aie(x)((\nabla^{E})^{*}\nabla^{E}-d^{*}d)\phi=-2(A,d\phi)=-2g^{ik}(x)A_{i}e_{\ell}(x)

for xUx\in U.

Finally, having determined every element of the Dirac bundle 𝒟\mathcal{D} except for the Clifford multiplication homomorphism 𝒸𝓁:𝓁(𝒯,)End()\mathpzc{cl}:\mathbb{C}l(T^{*}M,g)\to\operatorname{End}(E) we observe that having determined both E\nabla^{E} and g\nabla^{g} we have fixed 𝒸𝓁\mathpzc{cl} due to the Clifford compatibility condition that

[E,𝒸𝓁(θ)]=𝒸𝓁(θ),θ𝒯.[\nabla^{E},\mathpzc{cl}(\theta)]=\mathpzc{cl}(\nabla^{g}\theta),\quad\theta\in T^{*}M.

Proof of Corollary 1.2.

We follow the proof of [17, Cor 2]. To obtain the total spectral data on 𝒪\mathcal{O} we notice that not only are the remaining negative eigenvalues simply minus of the positive eigenvalues, but the unknown negative eigensections are simply the image of the positive eigensections under the Chirality operator,

γ=im2𝒸𝓁(1)𝒸𝓁(2)𝒸𝓁(𝓂)End(𝓍),\gamma=i^{\lceil\frac{m}{2}\rceil}\mathpzc{cl}(e_{1})\mathpzc{cl}(e_{2})\ldots\mathpzc{cl}(e_{m})\in\operatorname{End}(E_{x}),

where {ej}j=1m\{e_{j}\}_{j=1}^{m} is any choice of orthonormal basis of Tx𝒪T_{x}\mathcal{O}. Thus, because φ=λφ\not{D}\varphi=\lambda\varphi implies (γφ)=λγφ\not{D}(\gamma\varphi)=-\lambda\;\gamma\varphi we now have knowledge of all eigensections restricted to 𝒪\mathcal{O}. Writing the 2=ΔE\not{D}^{2}=\Delta_{E} eigensections as ψk(x)=φk(x)γφk(x)\psi_{k}(x)=\varphi_{k}(x)-\gamma\circ\varphi_{k}(x), we now have {λk2,ψk(x)}j=1\{\lambda_{k}^{2},\psi_{k}(x)\}_{j=1}^{\infty} as a spectral resolution of ΔE\Delta_{E} on L2(M;E)L^{2}(M;E). From here, as in [17], given any f𝒞0(+×𝒪)f\in{\mathcal{C}}^{\infty}_{0}(\mathbb{R}^{+}\times\mathcal{O}), we can calculate the approximate ‘Fourier coefficients’ of the wave source-to-solution operator k(t)=wf(t,),ψkL2\mathcal{I}_{k}(t)=\langle w^{f}(t,\cdot),\psi_{k}\rangle_{L^{2}}, and from there recover the approximate source-to-solution operator k=1~k(t)φk(x)=W[hf](t,x)\sum\limits_{k=1}^{\infty}\widetilde{\mathcal{I}}_{k}(t)\varphi_{k}(x)=W[h\cdot f](t,x) (i.e. the solution associated to a conformal change of source) with ~k(t)=0t𝒪sin(λkt)sin(λk)f(s,x)φj(x)h𝑑Vg(x)𝑑s\widetilde{\mathcal{I}}_{k}(t)=\int_{0}^{t}\int_{\mathcal{O}}\frac{\sin(\lambda_{k}t)}{\sin(\lambda_{k})}f(s,x)\varphi_{j}(x)\;h\cdot dV_{g}(x)ds. From here the result follows from 1.1 (after determining the conformal factor hh, as in [17, Thm 6]). ∎

5. acknowledgements

Travel funding in support of this work was provided to HQ by the Milliman committee of UW math. The research of GU was partly supported by NSF, a Simons Fellowship, a Walker Family Endowed Professorship at UW and a Si-Yuan Professorship at IAS, HKUST.

References

  • [1] M. I. Belishev. An approach to multidimensional inverse problems for the wave equation. Doklady Akademii Nauk SSR, 297(3):524–527, 1987.
  • [2] S. Bhattacharyya, T. Ghosh, and G. Uhlmann. Inverse problems for the fractional-Laplacian with lower order non-local perturbations. Transactions of the American Mathematical Society, 374(5):3053–3075, 2021.
  • [3] B. Booss, B. Booß-Bavnbek, K. P. Wojciechhowski, and K. Wojciechowski. Elliptic boundary problems for Dirac operators, volume 4. Springer Science & Business Media, 1993.
  • [4] M. Cekić, Y.-H. Lin, and A. Rüland. The Calderón problem for the fractional Schrödinger equation with drift. Calculus of Variations and Partial Differential Equations, 59(3):1–46, 2020.
  • [5] S.-Y. A. Chang and M. del Mar Gonzalez. Fractional Laplacian in conformal geometry. Advances in Mathematics, 226(2):1410–1432, 2011.
  • [6] G. Covi. An inverse problem for the fractional Schrödinger equation in a magnetic field. Inverse Problems, 36(4):045004, 2020.
  • [7] G. Covi. Inverse problems for a fractional conductivity equation. Nonlinear Analysis, 193:111418, 2020.
  • [8] G. Covi. Uniqueness for the fractional Calderón problem with quasilocal perturbations. arXiv preprint arXiv:2110.11063, 2021.
  • [9] G. Covi, M. Á. García-Ferrero, and A. Rüland. On the Calderón problem for nonlocal Schrödinger equations with homogeneous, directionally antilocal principal symbols. arXiv preprint arXiv:2109.14976, 2021.
  • [10] G. Covi, K. Mönkkönen, J. Railo, and G. Uhlmann. The higher order fractional Calderón problem for linear local operators: uniqueness. Advances in Mathematics, 399:108246, 2022.
  • [11] M. Eller, V. Isakov, G. Nakamura, and D. Tataru. Uniqueness and stability in the Cauchy problem for Maxwell and elasticity systems. In Nonlinear partial differential equations and their applications: Collège de France seminar. Volume XIV, pages 329–349. Elsevier, 2002.
  • [12] A. Feizmohammadi, T. Ghosh, K. Krupchyk, and G. Uhlmann. Fractional anisotropic Calder\\backslash’on problem on closed Riemannian manifolds. arXiv preprint arXiv:2112.03480, 2021.
  • [13] T. Ghosh, Y.-H. Lin, and J. Xiao. The Calderón problem for variable coefficients nonlocal elliptic operators. Communications in Partial Differential Equations, 42:1923 – 1961, 2017.
  • [14] T. Ghosh, A. Rüland, M. Salo, and G. Uhlmann. Uniqueness and reconstruction for the fractional Calderón problem with a single measurement. Journal of Functional Analysis, 279(1):108505, 2020.
  • [15] T. Ghosh, M. Salo, and G. Uhlmann. The Calderón problem for the fractional Schrödinger equation. Analysis & PDE, 13(2):455–475, 2020.
  • [16] S. A. Hartnoll and A. Karch. Scaling theory of the cuprate strange metals. Phys. Rev. B, 91:155126, Apr 2015.
  • [17] T. Helin, M. Lassas, L. Oksanen, and T. Saksala. Correlation based passive imaging with a white noise source. Journal de Mathématiques Pures et Appliquées, 116:132–160, 2018.
  • [18] R. Herrmann. Gauge invariance in fractional field theories. Physics Letters A, 372(34):5515–5522, 2008.
  • [19] B. Jin and W. Rundell. A tutorial on inverse problems for anomalous diffusion processes. Inverse problems, 31(3):035003, 2015.
  • [20] A. Kachalov, Y. Kurylev, and M. Lassas. Inverse boundary spectral problems. Chapman and Hall/CRC, 2001.
  • [21] Y. Kurylev, L. Oksanen, and G. P. Paternain. Inverse problems for the connection Laplacian. Journal of Differential Geometry, 110(3):457–494, 2018.
  • [22] G. La Nave, K. Limtragool, and P. W. Phillips. Colloquium: Fractional electromagnetism in quantum matter and high-energy physics. Rev. Mod. Phys., 91:021003, Jun 2019.
  • [23] N. Laskin. Fractional quantum mechanics. Phys. Rev. E, 62:3135–3145, Sep 2000.
  • [24] J. M. Lee and G. Uhlmann. Determining anisotropic real-analytic conductivities by boundary measurements. Communications on Pure and Applied Mathematics, 42(8):1097–1112, 1989.
  • [25] L. Li. The Calderón problem for the fractional magnetic operator. Inverse Problems, 36(7):075003, 2020.
  • [26] L. Li. Determining the magnetic potential in the fractional magnetic Calderón problem. Communications in Partial Differential Equations, 46(6):1017–1026, 2021.
  • [27] M. Ludewig. Strong short-time asymptotics and convolution approximation of the heat kernel. Annals of Global Analysis and Geometry, 55(2):371–394, 2019.
  • [28] A. Rüland. On single measurement stability for the fractional Calderón problem. SIAM Journal on Mathematical Analysis, 53(5):5094–5113, 2021.
  • [29] A. Rüland and M. Salo. Exponential instability in the fractional Calderón problem. Inverse Problems, 34(4):045003, 2018.
  • [30] A. Rüland and M. Salo. The fractional Calderón problem: low regularity and stability. Nonlinear Analysis, 193:111529, 2020.
  • [31] A. Strohmaier. The Reeh-Schlieder property for quantum qields on stationary spacetimes. Communications in Mathematical Physics, 215:105–118, 2000.
  • [32] D. Tataru. Unique continuation for solutions to pde’s; between Hörmander’s theorem and Holmgren’ theorem. Communications in Partial Differential Equations, 20(5-6):855–884, 1995.
  • [33] G. Uhlmann. Electrical impedance tomography and Calderón’s problem. Inverse problems, 25(12):123011, 2009.
  • [34] G. Uhlmann. Inverse problems: seeing the unseen. Bulletin of Mathematical Sciences, 4(2):209–279, 2014.
  • [35] N. T. Varopoulos. Hardy-Littlewood theory for semigroups. Journal of functional analysis, 63(2):240–260, 1985.
  • [36] R. Verch. Antilocality and a Reeh-Schlieder theorem on manifolds. Letters in Mathematical Physics, 28(2):143–154, June 1993.
  • [37] E. Witten. APS Medal for Exceptional Achievement in Research: Invited article on entanglement properties of quantum field theory. Rev. Mod. Phys., 90:045003, Oct 2018.
  • [38] P. Závada. Relativistic wave equations with fractional derivatives and pseudodifferential operators. Journal of Applied Mathematics, 2(4):163 – 197, 2002.
  • [39] P. Zeeman. The Effect of Magnetisation on the Nature of Light Emitted by a Substance. Nature, 55(347), 1897.