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

Recovering a Riemannian Metric from Cherenkov Radiation in Inhomogeneous Anisotropic Medium

Antti T. P. Kujanpää [email protected] Department of Mathematics and Statistics, University of Helsinki
Abstract.

Although travelling faster than the speed of light in vacuum is not physically allowed, the analogous bound in medium can be exceeded by a moving particle. For an electron in dielectric material this leads to emission of photons which is usually referred to as Cherenkov radiation. In this article a related mathematical system for waves in inhomogeneous anisotropic medium with a maximum of three polarisation directions is studied. The waves are assumed to satisfy Pjkuk(x,t)=Sj(x,t)P^{k}_{j}u_{k}(x,t)=S_{j}(x,t), where PP is a vector-valued wave operator that depends on a Riemannian metric and SS is a point source that moves at speed β<c\beta<c in given direction θ𝕊2\theta\in\mathbb{S}^{2}. The phase velocity vphasev_{\text{phase}} is described by the metric and depends on both location and direction of motion. In regions where vphase(x,θ)<β<cv_{\text{phase}}(x,\theta)<\beta<c holds the source generates a cone-shaped front of singularities that propagate according to the underlying geometry. We introduce a model for a measurement setup that applies the mechanism and show that the Riemannian metric inside a bounded region can be reconstructed from partial boundary measurements. The result suggests that Cherenkov type radiation can be applied to detect internal geometric properties of an inhomogeneous anisotropic target from a distance.

1. Introduction

Figure 1. Cherenkov radiation wave front (in black) in three dimensional medium at a fixed time is a cone with apex at a moving charged particle (in red). The phenomenon is an electromagnetic equivalent of a sonic boom.

When a charge carrier moves in material faster than the phase velocity it radiates photons. This phenomenon, known in physics as Cherenkov radiation, plays an important role in particle detection systems such as the ring imaging Cherenkov detector (RICH). The mechanism has mathematical description (See Section 3) as a modification of propagation of singularities for real principal type operators. To introduce the concept, let us consider a system of the form

Pjkuk(x,t)=Sj(x,t),j=1,2,3,x3,t,P^{k}_{j}u_{k}(x,t)=S_{j}(x,t),\quad j=1,2,3,\quad x\in\mathbb{R}^{3},\quad t\in\mathbb{R},

where the Einstein summation convention is considered over k=1,2,3k=1,2,3, the operator PP with entries Pjk=Pjk(x,t,D)P^{k}_{j}=P^{k}_{j}(x,t,D), j,k=1,2,3j,k=1,2,3 is hyperbolic, and Sj(x,t)dxjS_{j}(x,t)dx^{j} is a singular source with each coordinate SjS_{j} conormal to the world line

K:={(z+tβθ,t)4:t},θ𝕊2,β(0,1)K:=\{(z+t\beta\theta,t)\in\mathbb{R}^{4}:t\in\mathbb{R}\},\quad\theta\in\mathbb{S}^{2},\quad\beta\in(0,1)

(i.e. trajectory in space-time) of the particle. An example of this is a scalar model, e.g. the scalar wave equation

(t2Δg)u=f,u𝒟(4),(\partial_{t}^{2}-\Delta_{g})u=f,\quad u\in\mathcal{D}^{\prime}(\mathbb{R}^{4}),

(identified as Pjk=δjk(t2Δg)P^{k}_{j}=\delta^{k}_{j}(\partial_{t}^{2}-\Delta_{g}) and Sj=fS_{j}=f) with a source f𝒟(4)f\in\mathcal{D}^{\prime}(\mathbb{R}^{4}) conormal to KK (e.g. f(x,t)=δ(xzβθ)f(x,t)=\delta(x-z-\beta\theta)). The source S=SjdxjS=S_{j}dx^{j} can be interpreted as a singularity that moves at a physically allowed (β<c=1\beta<c=1) constant velocity βθ\beta\theta, β(0,1)\beta\in(0,1) through z3z\in\mathbb{R}^{3}. If for PP, or more generally, for the product VPWVPW with non-degenerate VV and WW the leading part of the operator equals a scalar operator Q=QIQ=QI of real principal type, then the standard propagation of singularities extends to the case above. More precisely; the tensors VV and WW are neglectable in terms of singularities and the wave front set of u=ujdxju=u_{j}dx^{j} splits into a non-propagating part which does not leave j=13WF(Sj)NK\bigcup_{j=1}^{3}WF(S_{j})\subset N^{*}K, and a propagating part which spreads along the bicharacteristics of QQ into the surrounding space. We assume that the leading part QQ is of the form

Q=t2gjk(x)xjxkQ=\partial_{t}^{2}-g^{jk}(x)\frac{\partial}{\partial x^{j}}\frac{\partial}{\partial x^{k}}

where gg stands for a Riemannian metric. The characteristic manifold for QQ is the covector light-cone

L4={(x,t;ξ,ω)T4{0}:gjk(x)ξjξk=ω2},L\mathbb{R}^{4}=\{(x,t;\xi,\omega)\in T^{*}\mathbb{R}^{4}\setminus\{0\}:g^{jk}(x)\xi_{j}\xi_{k}=\omega^{2}\},

also known as the null-cone, and the speed of wave propagation at x3x\in\mathbb{R}^{3} in direction θ𝕊2\theta\in\mathbb{S}^{2} is vphase(x,θ):=(gjk(x)θjθk)1/2v_{\text{phase}}(x,\theta):=(g^{jk}(x)\theta_{j}\theta_{k})^{-1/2}. Provided that the source moves slower than waves (i.e. β<vphase\beta<v_{\text{phase}}) the normal bundle NKN^{*}K of the world line, and hence the wave front set of the source remains disjoint from the light-cone (the left picture in Figure 2) which implies that no propagating singularities are generated. This corresponds to absence of Cherenkov radiation in subluminal regimes. On the other hand, in regions where the point source moves faster than waves a non-empty intersection between the light-cone and the normal bundle appears (the right picture in Figure 2), thus enabling radiation of singularities from the source. The propagating singularities add up to a conical shock wave which is illustrated in Figure 1.

β<vphase(x,θ)\beta<v_{\text{phase}}(x,\theta) N(x,t)KN^{*}_{(x,t)}KTimeSpace
β>vphase(x,θ)\beta>v_{\text{phase}}(x,\theta) T(x,t)KT_{(x,t)}^{*}K Characteristic normals
Figure 2. The set of characteristic covectors of the wave operator in a single fibre T(x,t)4T_{(x,t)}^{*}\mathbb{R}^{4} is the light-cone L(x,t)4L_{(x,t)}\mathbb{R}^{4} which in some convenient spatial coordinates (e.g. normal coordinates at xx) is the Minkowsky light-cone (in black). The blue lines illustrate the cotangent space of the world-line KK through (x,t)K(x,t)\in K. The two pictures correspond to subluminal and superluminal cases, respectively. That is; the line KK is time-like at (x,t)(x,t) in the first picture, whereas at the second picture the source moves faster than null-geodesics. The normal plane N(x,t)KN_{(x,t)}^{*}K (in orange) of the line at (x,t)(x,t) does not intersect the cone in the first picture, whereas in the second picture there exist characteristic normals (in red), and hence bicharacteristics in L4L\mathbb{R}^{4} through them.

In this article it is shown that Cherenkov radiation can be applied to recover Riemannian structure of a target region within anisotropic inhomogeneous medium. The idea is to send singularities (particles or sharp pulses) into the region of interest and observe new singularities generated as the incident signals interact with the medium. These disruptions are carried by waves into the exterior of the target where they eventually hit a given hypersurface that surrounds the region of interest. On the surface the generated singularities are observed and data is collected by repeating the measurement with various initial parameters (z,θ,β)(z,\theta,\beta) of the source. We show that information collected in this way on a convenient part of the surface determines uniquely the metric gg inside the target for a large class of geometries and sources.

The structure of the article is as follows: In the first section the model and the main result together with two related examples are introduced. Some basic concepts are briefly discussed in Section 2. In Section 3 we build up the microlocal tools which are then used in the last section for proving the main result. Appendix A contains additional details for Section 3.

1.1. The Model

Let II stand for the identity matrix. Assume that the (1,1)-tensor P(x,t,D)P(x,t,D) on 3\mathbb{R}^{3} with differential operator entries Pjk(x,t,D)Ψcl2(4)P^{k}_{j}(x,t,D)\in\Psi^{2}_{cl}(\mathbb{R}^{4}) is similar to the vector wave-operator in the sense that the condition

(1) P(x,t,D)=(t2grl(x)xrxl)I+F(x,Dx),P(x,t,D)=\left(\partial_{t}^{2}-g^{rl}(x)\frac{\partial}{\partial x^{r}}\frac{\partial}{\partial x^{l}}\right)I+F(x,D_{x}),

that is,

(2) Pjk(x,t,D)=δjk(t2grl(x)xrxl)+Fjk(x,Dx),j,k=1,2,3,P^{k}_{j}(x,t,D)=\delta_{j}^{k}\left(\partial_{t}^{2}-g^{rl}(x)\frac{\partial}{\partial x^{r}}\frac{\partial}{\partial x^{l}}\right)+F^{k}_{j}(x,D_{x}),\quad j,k=1,2,3,

holds for some first order operators Fjk(x,Dx)F_{j}^{k}(x,D_{x}), j,k=1,2,3j,k=1,2,3. Consider a point-like object moving linearly in 3\mathbb{R}^{3} in direction θ𝕊2\theta\in\mathbb{S}^{2} at physically allowed speed β(0,c)=(0,1)\beta\in(0,c)=(0,1) (c=1c=1 in natural unit system). Let zz be the location at time t=0t=0. The associated world line in the space time 4\mathbb{R}^{4} is

K(z,θ,β):={(x,t)4:x=z+tβθ},K(z,\theta,\beta):=\{(x,t)\in\mathbb{R}^{4}:x=z+t\beta\theta\},

We study singularities of distributions vj𝒟(4)v_{j}\in\mathcal{D}^{\prime}(\mathbb{R}^{4}), j=1,2,3j=1,2,3, that obey

(3) Pjkvk(x,t)Sj(x,t)modC(4),j=1,2,3,P^{k}_{j}v_{k}(x,t)\equiv S_{j}(x,t)\mod C^{\infty}(\mathbb{R}^{4}),\quad j=1,2,3,

where the terms SjS_{j} are conormal distributions over K(z,θ,β)K(z,\theta,\beta), that is,

(4) Sj(x,t)=3ei(xztβθ)ξcj(x,t,ξ)𝑑ξ,j=1,2,3,cjSr(4×3).S_{j}(x,t)=\int_{\mathbb{R}^{3}}e^{i(x-z-t\beta\theta)\cdot\xi}c_{j}(x,t,\xi)d\xi,\quad j=1,2,3,\quad c_{j}\in S^{r}(\mathbb{R}^{4}\times\mathbb{R}^{3}).

E.g. S=δ(xβθt) dx1S=\delta(x-\beta\theta t) dx^{1} if cj=δj1c_{j}=\delta_{j1}. Above the relation abmodCa\equiv b\mod C^{\infty} refers to abCa-b\in C^{\infty}. The source S=SjdxjS=S_{j}dx^{j} can be interpreted as a point singularity that moves in 3\mathbb{R}^{3} along the trajectory tz+tβθt\mapsto z+t\beta\theta, θS2\theta\in S^{2}, β(0,1)\beta\in(0,1), whereas the field v=vjdxjv=v_{j}dx^{j} is a wave in anisotropic inhomogeneous material. The three components of each element correspond to different polarisations. Electromagnetic waves created by a moving charge disruption such as a charged particle in anisotropic medium is perhaps the most obvious application for the model. For materials with scalar wave impedance the electric field obeys (3) for a metric gg conformally equivalent to the anisotropic permeability tensor ε(x)\varepsilon(x). This model is explored in Section 1.6.

Remark 1.1.

Notice that an analogous scalar equation, e.g the scalar wave equation

(tΔg)v=f𝒟(4),f,v𝒟(4)(\partial_{t}-\Delta_{g})v=f\in\mathcal{D}^{\prime}(\mathbb{R}^{4}),\quad f,v\in\mathcal{D}^{\prime}(\mathbb{R}^{4})

with source ff conormal to K(x,θ,β)K(x,\theta,\beta), can also be expressed within the model by identifying any scalar operator PscalP_{scal} with the diagonal matrix Pjk:=δjkPscalP^{k}_{j}:=\delta^{k}_{j}P_{scal}, j,k=1,2,3j,k=1,2,3 and a scalar source SscalS_{scal} with Sj:=SscalS_{j}:=S_{scal}, j=1,2,3j=1,2,3. Similarly, the model also generalises waves with two polarisation directions.

Remark 1.2.

The model above is not as general as possible. It most likely suffices to assume that each entry FjkF^{k}_{j} of the tensor FF is an operator in the space Ψcl1(3)\Psi^{1}_{cl}(\mathbb{R}^{3}) of classical first order pseudo-differential operators. Moreover, the identity (1) can be replaced by the more general condition

(5) P(x,t,D)(t2grl(x)xrxl)I+F(x,Dx)P(x,t,D)\cong\left(\partial_{t}^{2}-g^{rl}(x)\frac{\partial}{\partial x^{r}}\frac{\partial}{\partial x^{l}}\right)I+F(x,D_{x})

where the relation PGP\cong G between two operators PP and GG refers to

Pjk=AjlGlrBrkP^{k}_{j}=A^{l}_{j}G_{l}^{r}B_{r}^{k}

for some operators AA and BB with pointwise non-degenerate principal symbol matrices [σm1(Ajl)][\sigma_{m_{1}}(A^{l}_{j})] and [σm2(Bjl)][\sigma_{m_{2}}(B^{l}_{j})] (see Lemma 3.3). Notice that the non-degeneracy is just a vector-valued equivalent of ellipticity and that a non-degenerate smooth time-dependent (1,1)(1,1)-tensor A:u(x,t)Ajk(x,t)uk(x,t)dxjA:u(x,t)\mapsto A^{k}_{j}(x,t)u_{k}(x,t)dx^{j} generalizes the concept of elliptic multiplication operator u(x,t)f(x,t)u(x,t)u(x,t)\mapsto f(x,t)u(x,t) for non-vanishing fC(4)f\in C^{\infty}(\mathbb{R}^{4}). Notice that in the vector valued setting ellipticity allows exchange of singularities between polarisations.

Moreover, the results of this work are expected to remain true even if the model is considered only microlocally near the covector light-cone

L4={(x,t,ξ,ω)T4:ω2=gjk(x)ξjξk}.L\mathbb{R}^{4}=\{(x,t,\xi,\omega)\in T^{*}\mathbb{R}^{4}:\omega^{2}=g^{jk}(x)\xi_{j}\xi_{k}\}.

This follows from the fact that methodologically the proofs are microlocal and based on propagation of singularities.

1.1.1. The Data

Due to physical background we require that waves propagate with speed less or equal to the speed of light in vacuum. This is equivalent to

(6) |θ|g(x)1,(x,θ)3×𝕊2,|\theta|_{g}(x)\geq 1,\quad\forall(x,\theta)\in\mathbb{R}^{3}\times\mathbb{S}^{2},

where we denote |θ|g(x):=θjθkgjk(x)|\theta|_{g}(x):=\sqrt{\theta^{j}\theta^{k}g_{jk}(x)}. Let W3W\subset\mathbb{R}^{3} be an open bounded set with smooth boundary W\partial W and fix open UWU\subset W. We consider UU as a region with unknown geometry inside the medium, whereas the boundary W\partial W is the hypersurface on which observations are made. The source SS is controlled by the variables z,θz,\theta and β\beta.

To avoid Cherenkov radiation from arbitrarily far we consider mediums that converge to vacuum at infinity. That is; |θ|g(x)1|\theta|_{g}(x)\longrightarrow 1 for every θ𝕊2\theta\in\mathbb{S}^{2} as xx\sqrt{x\cdot x}\longrightarrow\infty. Alternatively, one may assume that there is a relatively large velocity threshold β0(0,1)\beta_{0}\in(0,1) such that the preimage

(7) Rθ={x3:1β0<|θ|g(x)}R_{\theta}=\Big{\{}x\in\mathbb{R}^{3}:\frac{1}{\beta_{0}}<|\theta|_{g}(x)\Big{\}}

is bounded and contains UU for every θ𝕊2\theta\in\mathbb{S}^{2}, and then focus on velocities β(0,β0)\beta\in(0,\beta_{0}). In both cases the wave front set of the tail (1χ)S(1-\chi)S is disjoint from the characteristic set of PP for a test function χ=χg,βCc(3)\chi=\chi_{g,\beta}\in C_{c}^{\infty}(\mathbb{R}^{3}), 0χ10\leq\chi\leq 1 that equals 11 in a sufficiently large open set. Hence, for a convenient operator P1P^{-1} (see Lemma 3.3) that inverts PP microlocally near NKN^{*}K in {χ<1}\{\chi<1\} the element P1(1χ)SP^{-1}(1-\chi)S does not radiate singularities and u:=vP1(1χ)Su:=v-P^{-1}(1-\chi)S satisfies

(8) PuχS(4).Pu\equiv\chi S\in\mathcal{E}^{\prime}(\mathbb{R}^{4}).

The advantage in writing the system in this form is that the source on the right is compactly supported. Through this procedure we may pass on to a setup

PjkukSj,P^{k}_{j}u_{k}\equiv S_{j},

in which the source S=SjdxjS=S_{j}dx^{j} is not only conormal to K(z,θ,β)K(z,\theta,\beta) but also compactly supported.

Definition 1.1.

Let UW(3,g)U\subset W\subset(\mathbb{R}^{3},g) be as above. We say that a subset Υ=ΥU,gW\Upsilon=\Upsilon_{U,g}\subset\partial W is a stable part of the boundary with respect to UU if Υ\Upsilon is open in W\partial W and for every zUz\in U there is at least one x0=x0(z)Υx_{0}=x_{0}(z)\in\Upsilon that minimises distance to W\partial W, i.e.

distg(z,W)=distg(z,x0),\text{dist}_{g}(z,\partial W)=\text{dist}_{g}(z,x_{0}),

where distg(z,W):=minxWdistg(x,z)\text{dist}_{g}(z,\partial W):=\min_{x\in\partial W}\text{dist}_{g}(x,z).

Remark 1.3.

The whole boundary Υ=W\Upsilon=\partial W is always stable.

For ΛT4\Lambda\subset T^{*}\mathbb{R}^{4} we let Λ|T(Υ×)T(Υ×)\Lambda|_{T(\Upsilon\times\mathbb{R})}\subset T^{*}(\Upsilon\times\mathbb{R}) stand for the canonical projection Λ|T(Υ×):={(x,t,(ξ,ω)|T(x,t)(Υ×)):(x,t,ξ,ω)Λ,(x,t)Υ×}\Lambda|_{T(\Upsilon\times\mathbb{R})}:=\Big{\{}\big{(}x,t,(\xi,\omega)|_{T_{(x,t)}(\Upsilon\times\mathbb{R})}\big{)}:(x,t,\xi,\omega)\in\Lambda,(x,t)\in\Upsilon\times\mathbb{R}\Big{\}}. Given a family {Sj(;z,θ,β)dxj:(z,θ,β)U×𝕊2×(0,1)}\{S_{j}(\ \cdot\ ;z,\theta,\beta)dx^{j}:(z,\theta,\beta)\in U\times\mathbb{S}^{2}\times(0,1)\} of compactly supported sources of the form (4) and a stable ΥW\Upsilon\subset\partial W, the measurement data are defined as the following map:

(z,θ,β)j=1,2,3WF(uj)|T(Υ×),(z,θ,β)U×𝕊2×(0,1),(z,\theta,\beta)\mapsto\bigcup_{j=1,2,3}WF(u_{j})|_{T(\Upsilon\times\mathbb{R})},\quad(z,\theta,\beta)\in U\times\mathbb{S}^{2}\times(0,1),

where u=uj(;z,θ,β)dxju=u_{j}(\ \cdot\ ;z,\theta,\beta)dx^{j} solves for each j=1,2,3,j=1,2,3, the conditions

(9) PjkukSjmodC(4),\displaystyle P^{k}_{j}u_{k}\equiv S_{j}\mod C^{\infty}(\mathbb{R}^{4}),
(10) uj|3×(,T]C(3×(,T])\displaystyle u_{j}|_{\mathbb{R}^{3}\times(-\infty,-T]}\in C^{\infty}(\mathbb{R}^{3}\times(-\infty,-T])

for Sj=Sj(;z,θ,β)S_{j}=S_{j}(\ \cdot\ ;z,\theta,\beta) and some (sufficiently large) T>0T>0. Notice that each source is supported in 3×(T,)\mathbb{R}^{3}\times(-T,\infty) for large T>0T>0 (see Figure 4). As the high-energy limit β1\beta\longrightarrow 1 is unpractical, we are interested in smaller datasets where the velocities are restricted to a smaller subinterval z(0,1)\mathcal{I}_{z}\subset(0,1). The interval may depend on zUz\in U.

Remark 1.4.

Substituting PP in the definition with more general operator P~P\tilde{P}\cong P (see Remark 1.2) yields the same data. For example, the data are invariant in transformation u(x,t)Vjk(x,t)uk(x,t)dxju(x,t)\mapsto V^{k}_{j}(x,t)u_{k}(x,t)dx^{j} and S(x,t)Wjk(x,t)Sk(x,t)dxjS(x,t)\mapsto W^{k}_{j}(x,t)S_{k}(x,t)dx^{j}, where VV and WW are smooth non-degenerate (1,1)(1,1)-tensors.

UUWWΥ\Upsilon
Figure 3. A schematic 2D illustration of WW,UU and a stable part ΥW\Upsilon\subset\partial W. The ambient space represents 3\mathbb{R}^{3}.
zt=0t=0tTt\leq-T
Figure 4. A schematic illustration of the sets associated with (9-10). We consider waves uu that are smooth in the half-space 3×(,T]\mathbb{R}^{3}\times(-\infty,-T], indicated in gray. The sets U×(T,)U\times(-T,\infty) and Υ×(T,)\Upsilon\times(-T,\infty) are outlined in blue and red, respectively. Singularities of the source (the black line) lie in a compact subset of K(z,θ,β)(3×(T,))K(z,\theta,\beta)\cap(\mathbb{R}^{3}\times(-T,\infty)) for large T>0T>0.

1.2. The Main Result

The main outcome of this work is Theorem 1.1 below which states that the Riemannian metric gg in UU is uniquely determined by the data, provided that some natural conditions are satisfied. In fact, for each zUz\in U it suffices to consider data for velocities β\beta in a specific subinterval z(0,1)\mathcal{I}_{z}\subset(0,1) with supz\sup\mathcal{I}_{z} strictly less than c=1c=1, the speed of light in vacuum. If the maximum phase velocity in the closure U¯\overline{U} is strictly less than the speed of light in vacuum then the dependence on zUz\in U in the interval above can be dropped (see Remark 1.6). The requirements for the metric are as follows:

Definition 1.2.

Let gg be a Riemannian metric in 3\mathbb{R}^{3} and let U3U\subset\mathbb{R}^{3} be an open bounded set. We say that the metric gg is admissible with respect to UU if the following conditions hold:

  1. (i)

    Propagation speed of waves is everywhere less than or equal to the speed of light in vacuum:

    |θ|g(x)1,(x,θ)3×𝕊2.|\theta|_{g}(x)\geq 1,\quad\forall(x,\theta)\in\mathbb{R}^{3}\times\mathbb{S}^{2}.
  2. (ii)

    Propagation speed of waves in UU is strictly less than the speed of light in vacuum:

    |θ|g(z)>1,(z,θ)U×𝕊2.|\theta|_{g}(z)>1,\quad\forall(z,\theta)\in U\times\mathbb{S}^{2}.

To avoid technical difficulties, the following lower bound for speeds β\beta is considered in the main theorem.

Definition 1.3.

Let gg be admissible metric with respect to U3U\subset\mathbb{R}^{3}. For zUz\in U and a bounded set W3W\subset\mathbb{R}^{3} containing UU we define

Jz:=(maxθ𝕊21|θ|g(z),1),βW(z;g):=inf{βJz:βdistg(z,x)>|xz|eucl,xW},J_{z}:=\Big{(}\max_{\theta\in\mathbb{S}^{2}}\frac{1}{|\theta|_{g}(z)},1\Big{)},\quad\beta_{\partial W}(z;g):=\inf\big{\{}\beta\in J_{z}:\beta\text{dist}_{g}(z,x)>|x-z|_{eucl},\ \forall x\in\partial W\big{\}},

where ||eucl|\cdot|_{eucl} is the Euclidean norm.

Remark 1.5.

It follows from compactness of W\partial W that βW(z;g)<1\beta_{\partial W}(z;g)<1. Moreover, if propagation of waves in the closure U¯\overline{U} is strictly slower than in vacuum, we have supzUβW(z;g)<1\sup_{z\in U}\beta_{\partial W}(z;g)<1.

Below M3(X)M_{3}(X) refers to 3×33\times 3 matrices with entries in the class XX. In orthonormal Euclidean coordinates we identify a (1,1)-tensor TjkxkdxjT^{k}_{j}\frac{\partial}{\partial x^{k}}\otimes dx^{j}, TjkXT^{k}_{j}\in X, with the matrix [Tjk]M3(X)[T^{k}_{j}]\in M_{3}(X). The sources Sj=Sj(;z,θ,β)S_{j}=S_{j}(\ \cdot\ ;z,\theta,\beta) are required to satisfy for every (z,θ,β)U×𝕊2×(0,1)(z,\theta,\beta)\in U\times\mathbb{S}^{2}\times(0,1), j=1,2,3j=1,2,3 the following conditions:

(11) SjIclr+1/2(4;NK(z,θ,β)),r,i.e.Sj(x,t)=3ei(xztβθ)ξaj(x,t,ξ)𝑑ξ,aj=aj(;z,θ,β)Sclr(4×3),\begin{split}&S_{j}\in I_{cl}^{r+1/2}(\mathbb{R}^{4};N^{*}K(z,\theta,\beta)),\quad r\in\mathbb{R},\quad\text{i.e.}\\ &S_{j}(x,t)=\int_{\mathbb{R}^{3}}e^{i(x-z-t\beta\theta)\cdot\xi}a_{j}(x,t,\xi)d\xi,\quad a_{j}=a_{j}(\ \cdot\ ;z,\theta,\beta)\in S^{r}_{cl}(\mathbb{R}^{4}\times\mathbb{R}^{3}),\end{split}
(12) singsupp(Sj(;z,θ,β))B(0,Rβ)×,(cf. the cut-off in (8))\displaystyle\text{singsupp}(S_{j}(\ \cdot\ ;z,\theta,\beta))\subset B(0,R_{\beta})\times\mathbb{R},\quad\text{(cf. the cut-off in \eqref{cytor}) }
(13) (a1,r,a2,r,a3,r)(z,0,ξ)0,(ξ,ω)NK(z,θ,β)L(z,0)4(a_{1,r},a_{2,r},a_{3,r})(z,0,\xi)\neq 0,\quad\forall(\xi,\omega)\in N^{*}K(z,\theta,\beta)\cap L_{(z,0)}\mathbb{R}^{4}(ω=±|ξ|g)(\omega=\pm|\xi|_{g})

where aj,ra_{j,r} is the positively homogeneous principal part of the symbol ajaj,r+aj,r1+aj,r2+a_{j}\sim a_{j,r}+a_{j,r-1}+a_{j,r-2}+\cdots, L(z,0)4:={(ξ,ω)T(z,0)4{0}:|ξ|gλ2(z)=ω2}L_{(z,0)}\mathbb{R}^{4}:=\{(\xi,\omega)\in T^{*}_{(z,0)}\mathbb{R}^{4}\setminus\{0\}:|\xi|_{g_{\lambda}}^{2}(z)=\omega^{2}\} and the radius Rβ>0R_{\beta}>0 depends continuously on β\beta. The main result of the paper is the following theorem:

Theorem 1.1.

Let W3W\subset\mathbb{R}^{3} be a bounded open set with smooth boundary W\partial W and let gλ=gλ,jk(x)dxjdxkg_{\lambda}=g_{\lambda,jk}(x)dx^{j}dx^{k}, λ=1,2\lambda=1,2 be two Riemannian metrics in 3\mathbb{R}^{3}, both admissible ((Definition 1.2)) with respect to an open UWU\subset W. Let ΥW\Upsilon\subset\partial W be a stable part of the boudary ((Definition 1.1)) with respect to UU for both metrics gλg_{\lambda}, λ=1,2\lambda=1,2 and assume that

g1|x=g2|xxΥ.g_{1}|_{x}=g_{2}|_{x}\quad\forall x\in\Upsilon.

Consider two differential operators PλM3(Ψcl2(4))P_{\lambda}\in M_{3}(\Psi^{2}_{cl}(\mathbb{R}^{4})), λ=1,2\lambda=1,2 of the form

Pλ(x,t,D)=(t2gλjk(x)xjxk)I+Fλ(x,Dx),P_{\lambda}(x,t,D)=\left(\partial_{t}^{2}-g_{\lambda}^{jk}(x)\frac{\partial}{\partial x^{j}}\frac{\partial}{\partial x^{k}}\right)I+F_{\lambda}(x,D_{x}),

where the entries (Fλ)jk(F_{\lambda})^{k}_{j}, j,k=1,2,3j,k=1,2,3 of FλF_{\lambda} are first order operators on 3\mathbb{R}^{3}, and sources Sλ(x,t):=Sλ,j(x,t;z,θ,β)dxjS_{\lambda}(x,t):=S_{\lambda,j}(x,t;z,\theta,\beta)dx^{j} that satisfy (11-13). Let uλ(x,t):=uλ,j(x,t;z,θ,β)dxju_{\lambda}(x,t):=u_{\lambda,j}(x,t;z,\theta,\beta)dx^{j} solve

(14) PλuλSλC(4;3),\displaystyle P_{\lambda}u_{\lambda}-S_{\lambda}\in C^{\infty}(\mathbb{R}^{4};\mathbb{R}^{3}),
(15) uλ|3×(,T]C(3×(,T];3).\displaystyle u_{\lambda}|_{\mathbb{R}^{3}\times(-\infty,-T]}\in C^{\infty}(\mathbb{R}^{3}\times(-\infty,-T];\mathbb{R}^{3}).

for some T=Tz,β,θ>0T=T_{z,\beta,\theta}>0. Assume that for every zUz\in U there is an open nonempty subinterval

z(βW(z),1),βW(z):=maxλ=1,2βW(z;gλ),\mathcal{I}_{z}\subset\Big{(}\beta_{\partial W}(z),1\Big{)},\quad\beta_{\partial W}(z):=\max\limits_{\lambda=1,2}\beta_{\partial W}(z;g_{\lambda}),

such that the data on z\mathcal{I}_{z} coincide for both metrics, i.e.

j=1,2,3WF(uλ=1,j)|T(Υ×)=j=1,2,3WF(uλ=2,j)|T(Υ×),βz,\bigcup_{j=1,2,3}WF(u_{\lambda=1,j})|_{T(\Upsilon\times\mathbb{R})}=\bigcup_{j=1,2,3}WF(u_{\lambda=2,j})|_{T(\Upsilon\times\mathbb{R})},\quad\forall\beta\in\mathcal{I}_{z},

for every (z,θ)U×𝕊2(z,\theta)\in U\times\mathbb{S}^{2}. Then, g1|U=g2|Ug_{1}|_{U}=g_{2}|_{U}.

Remark 1.6.

Provided that (ii) in Definition 1.2 is substituted by the stronger condition

|θ|g(z)>1,(z,θ)U¯×𝕊2,|\theta|_{g}(z)>1,\quad\forall(z,\theta)\in\overline{U}\times\mathbb{S}^{2},

the compactness of U¯\overline{U} yields supzUβW(z)<1\sup_{z\in U}\beta_{\partial W}(z)<1 and we can choose z=\mathcal{I}_{z}=\mathcal{I} independently on zUz\in U for some (0,1)\mathcal{I}\subset(0,1) with sup<1\sup\mathcal{I}<1. This means that it suffices to make measurements with moderate speeds β\beta to obtain the result above.

1.3. Background and Previous Work

Cherenkov radiation is named after Pavel Cherenkov who shared the 1958 Nobel Price in Physics for its discovery. The historical background is briefly discussed in [Bol09]. The phenomenon is applied in particle detectors (see the review [YS94] and references therein) and in detection of biomolecules. Applications of Cherenkov luminescence in medical imaging and radiotherapy are studied extensively. [RHLG10], [ZQY+11], [DCB+18], [PK18], [KLA18] For Cherenkov radiation as a quantum effect in vacuum, see [BCK+]. Cherenkov radiation is analogous to so-called Askaryan radiation, first observed experimentally in 2000. [SGW+01] See [GSF+05], [BBB+06], [LGJRD04] for subsequent research.

Inverse problems related to Cherenkov or Askaryan radiation appears to be explored very little, if at all. There are, however, several studies on inverse problems for particle flows and analysis on wave fronts of propagating waves that are related to this paper or use analogous microlocal techniques. These include inverse problems related to interaction of waves [LUW18], [LUW18], [FLO20], [WZ19], [CLOP19], [CLOP20], [GLS+18], [HUZ21], single scattering inverse problems [KLU10], [dHILS20], [dHHI+15], [dHHI+14], and inverse problems for particle models [ST09], [CS96a], [CS96b], [LL20], [McD04], [McD05]. For studies with relativistic particles, see [BJ18], [Jol15], [Jol14], [Jol13]. Some of the methods applied in this work have points of resemblance in recovery of singularities in scattering theory which is a well-studied topic, especially in the Euclidean context. [GU93], [BFRV10], [PSS94], [PS91], [OPS01], [Jos98], [PS98], [RR12]

We refer to [H7̈1],[DH72], [GU81], [Gui75] [MU79], [GU90] for theory on FIOs. See also the preceding works [Lax57], [Lud60], [Ego69], and [Mas65, Mas88].

1.4. Acknowledgements

I would like to thank Academy Professor Matti Lassas for his help and guidance during the project. This work was financially supported by the ATMATH Collaboration and Academy of Finland (grants 336786 and 320113).

1.5. Illustrative Example

We give a short introduction to Cherenkov radiation in flat geometry which is the space-time that corresponds to homogeneous isotropic medium. The model in this example is trivial from a geometric point of view.

Let us consider the following forward propagating scalar wave:

(16) (t2k2Δ)u(x,t)=χ(x)S0(x,t),u𝒟(4),0<k<1,\displaystyle(\partial_{t}^{2}-k^{2}\Delta)u(x,t)=\chi(x)S_{0}(x,t),\quad u\in\mathcal{D}^{\prime}(\mathbb{R}^{4}),\quad 0<k<1,
(17) u|3×(,T]C(3×(,T]),\displaystyle u|_{\mathbb{R}^{3}\times(-\infty,T]}\in C^{\infty}(\mathbb{R}^{3}\times(-\infty,T]),
(18) S0:=δ(x1tβ)δ(x2)δ(x3).\displaystyle S_{0}:=\delta(x^{1}-t\beta)\delta(x^{2})\delta(x^{3}).

(z=0z=0, θ=(1,0,0)\theta=(1,0,0)) Here Δ=δjkxjxk\Delta=\delta^{jk}\frac{\partial}{\partial x^{j}}\frac{\partial}{\partial x^{k}} and χCc(3)\chi\in C^{\infty}_{c}(\mathbb{R}^{3}), χ|B(0,1)=1\chi|_{B(0,1)}=1. The characteristic manifold L4L\mathbb{R}^{4} for the wave operator is

Char(t2k2Δ)={(x,t,ξ,ω)4×(4{0}):ω2=k2|ξ|2}.\text{Char}(\partial_{t}^{2}-k^{2}\Delta)=\{(x,t,\xi,\omega)\in\mathbb{R}^{4}\times(\mathbb{R}^{4}\setminus\{0\}):\omega^{2}=k^{2}|\xi|^{2}\}.

The wave front set of the wave splits (see e.g. [Dui96]) into the static and propagating parts (cf. the discussion in Section 2.4). That is;

WF(u)WF(χS0)ΛfWF(χS0)\begin{split}WF(u)\subset WF(\chi S_{0})\ \cup\ \Lambda_{f}\circ WF(\chi S_{0})\end{split}

where Λf(T4×T4){0,0}\Lambda_{f}\subset(T^{*}\mathbb{R}^{4}\times T^{*}\mathbb{R}^{4})\setminus\{0,0\} stands for the forward propagating flow-out canonical relation generated by (t2k2Δ)(\partial_{t}^{2}-k^{2}\Delta), i.e.

Λf={(xrkv(ξ),t+r,ξ,±k|ξ| ;x,t,ξ,±k|ξ|):r0,(x,t,ξ)4×3},vj(ξ):=ξkδjk|ξ|.\Lambda_{f}=\bigg{\{}\Big{(}x\mp rkv(\xi),t+r,\xi,\pm k|\xi| \ ;\ x,t,\xi,\pm k|\xi|\Big{)}:r\geq 0,\ (x,t,\xi)\in\mathbb{R}^{4}\times\mathbb{R}^{3}\bigg{\}},\quad v^{j}(\xi):=\frac{\xi_{k}\delta^{jk}}{|\xi|}.

Here the composition ΛfWF(χS0)\Lambda_{f}\circ WF(\chi S_{0}) stands for

ΛfWF(χS0):={(x,t,ξ,ω):((x,t,ξ,ω),(y,s,η,ρ))Λf,(y,s,η,ρ)WF(χS0)}.\Lambda_{f}\circ WF(\chi S_{0}):=\{(x,t,\xi,\omega):\big{(}(x,t,\xi,\omega),(y,s,\eta,\rho)\big{)}\in\Lambda_{f},\ (y,s,\eta,\rho)\in WF(\chi S_{0})\}.

One checks that

WF(χS0)=NK(0,(1,0,0),β)(supp(χ)××4)={(tβ,0,0,t,ξ,βξ1):t,ξ3{0}}(supp(χ)××4).\begin{split}WF(\chi S_{0})&=N^{*}K(0,(1,0,0),\beta)\ \cap\ (\text{supp}(\chi)\times\mathbb{R}\times\mathbb{R}^{4})\\ &=\big{\{}(t\beta,0,0,t,\xi,-\beta\xi_{1}):t\in\mathbb{R},\ \xi\in\mathbb{R}^{3}\setminus\{0\}\big{\}}\ \cap\ (\text{supp}(\chi)\times\mathbb{R}\times\mathbb{R}^{4}).\end{split}

For β<k\beta<k we deduce

|βξ1|β|ξ|<k|ξ|βξ1±k|ξ|WF(χS0)Char(t2k2Δ)=.|\beta\xi_{1}|\leq\beta|\xi|<k|\xi|\quad\Rightarrow\quad-\beta\xi_{1}\neq\pm k|\xi|\quad\Rightarrow\quad WF(\chi S_{0})\cap\text{Char}(\partial_{t}^{2}-k^{2}\Delta)=\emptyset.

Thus, ΛfWF(χS0)=\Lambda_{f}\circ WF(\chi S_{0})=\emptyset and we obtain

WF(u)WF(χS0),forβ<k.WF(u)\subset WF(\chi S_{0}),\quad\text{for}\quad\beta<k.

That is; in the regime β<k\beta<k no singularities radiate from the source. If β>k\beta>k, then there exist bicharacteristics through WF(χS0)WF(\chi S_{0}) and the composition ΛfWF(χS0)\Lambda_{f}\circ WF(\chi S_{0}) is non-trivial. In fact, the projection of it to 4\mathbb{R}^{4} is a union of the sets

Qx,t={(x,t)+r(kθ~,1):θ~𝕊2,θ~1(β2k21)12=|(θ~2,θ~3)|,r>0}.\begin{split}Q_{x,t}=\bigg{\{}(x,t)+r(k\tilde{\theta},1):\tilde{\theta}\in\mathbb{S}^{2},\ \tilde{\theta}^{1}\left(\frac{\beta^{2}}{k^{2}}-1\right)^{\frac{1}{2}}=|(\tilde{\theta}^{2},\tilde{\theta}^{3})|,\ r>0\bigg{\}}.\end{split}

over points (x,t)(x,t) that satisfy xsupp(χ),x=(tβ,0,0)x\in\text{supp}(\chi),\ x=(t\beta,0,0). Each of these manifolds corresponds to propagating singularities generated at the associated point (x,t)(x,t) by the source. The equation θ~1(β2k21)12=|(θ~2,θ~3)|\tilde{\theta}^{1}\left(\frac{\beta^{2}}{k^{2}}-1\right)^{\frac{1}{2}}=|(\tilde{\theta}^{2},\tilde{\theta}^{3})| corresponds to the Frank-Tamm formula in physics. For β=k\beta=k the cone collapses into a line. This particular case is problematic in the microlocal framework as it fails to satisfy conditions required in the standard FIO calculuses.

1.6. Geometric Example: Electromagnetic Waves in Anisotropic Material with Scalar Wave Impedance

The objective of this example is to explicate how the equation

(t2gjkxjxk)Iv+FvSmodC(4;3),\left(\partial_{t}^{2}-g^{jk}\frac{\partial}{\partial x^{j}}\frac{\partial}{\partial x^{k}}\right)Iv+Fv\equiv S\mod C^{\infty}(\mathbb{R}^{4};\mathbb{R}^{3}),

is linked to electromagnetism. Consider a system of moving charges in dielectric anisotropic medium with smoothly varying inhomogeneities. The model is described by the Maxwell’s equations

(19) curlE(x,t)+tB(x,t)\displaystyle\text{curl}E(x,t)+\partial_{t}B(x,t) =0,\displaystyle=0,
(20) divB(x,t)\displaystyle\text{div}B(x,t) =0,\displaystyle=0,
(21) curlH(x,t)tD(x,t)\displaystyle\text{curl}H(x,t)-\partial_{t}D(x,t) =J(x,t),\displaystyle=J(x,t),
(22) divD(x,t)\displaystyle\text{div}D(x,t) =ρ(x,t),\displaystyle=\rho(x,t),

where E(x,t)=Ej(x,t)xjE(x,t)=E^{j}(x,t)\frac{\partial}{\partial x^{j}}, B(x,t)=Bj(x,t)xjB(x,t)=B^{j}(x,t)\frac{\partial}{\partial x^{j}}, D(x,t)=Dj(x,t)xjD(x,t)=D^{j}(x,t)\frac{\partial}{\partial x^{j}}, H(x,t)=Hj(x,t)xjH(x,t)=H^{j}(x,t)\frac{\partial}{\partial x^{j}}, ρ(x,t)\rho(x,t), and J(x,t)=Jj(x,t)xjJ(x,t)=J^{j}(x,t)\frac{\partial}{\partial x^{j}} are the associated electric field, magnetic field, electric displacement field, auxiliary magnetic field, charge density, and current density, respectively. Properties of the medium are encoded in the smooth (1,1)(1,1)-tensors

(23) ε(x)\displaystyle\varepsilon(x) =εjk(x)dxjxk,\displaystyle=\varepsilon_{j}^{k}(x)dx^{j}\otimes\frac{\partial}{\partial x^{k}},
(24) μ(x)\displaystyle\mu(x) =μjk(x)dxjxk\displaystyle=\mu_{j}^{k}(x)dx^{j}\otimes\frac{\partial}{\partial x^{k}}

which give connections between the fields via

(25) D(x,t)\displaystyle D(x,t) =ε(x)E(x,t),\displaystyle=\varepsilon(x)E(x,t),
(26) B(x,t)\displaystyle B(x,t) =μ(x)H(x,t).\displaystyle=\mu(x)H(x,t).

For lossless, optically inactive materials it is reasonable to require the tensors ε(x)\varepsilon(x) and μ(x)\mu(x) to be real valued, symmetric and non-degenerate. In addition, we assume that the material has scalar wave impedance. By definition, this means that there is a smooth scalar function α(x)\alpha(x) such that μ(x)=α2(x)ε(x)\mu(x)=\alpha^{2}(x)\varepsilon(x). Consider a charge carrier (e.g. an electron) or some other non-smooth charge perturbation moving along a straight line in 3\mathbb{R}^{3} through anisotropic dielectric medium with smoothly varying inhomogeneities. Let θ𝕊2\theta\in\mathbb{S}^{2}, β(0,c)=(0,1)\beta\in(0,c)=(0,1), and z3z\in\mathbb{R}^{3} be the direction of motion, speed, and location at t=0t=0 for the moving charge perturbation, respectively. Again, we work in a natural unit system that has c=1c=1 as the speed of light in vacuum. The perturbation signal is encoded in the charge density ρ(x,t)\rho(x,t) as an oscillatory superposition

(27) ρ(x,t)=3ei(xztβθ)ξa(x,ξ)𝑑ξ,\displaystyle\rho(x,t)=\int_{\mathbb{R}^{3}}e^{i(x-z-t\beta\theta)\cdot\xi}a(x,\xi)d\xi,

where the amplitude has an asymptotic development

(28) a(x,ξ)am(x,ξ)+am1(x,ξ)+a(x,\xi)\sim a_{m}(x,\xi)+a_{m-1}(x,\xi)+\cdots

into functions ajSj(3×(3{0}))a_{j}\in S^{j}(\mathbb{R}^{3}\times(\mathbb{R}^{3}\setminus\{0\})). For example, a single electron in vacuum corresponds to the density

ρe(x,t)=eγδ(xzβtθ)=eγ3ei(xztβθ)ξ𝑑ξ,\rho_{e^{-}}(x,t)=-e\gamma\delta(x-z-\beta t\theta)=-e\gamma\int_{\mathbb{R}^{3}}e^{i(x-z-t\beta\theta)\cdot\xi}d\xi,

where ee stands for the elementary charge and γ:=11β2\gamma:=\frac{1}{\sqrt{1-\beta^{2}}} is the Lorentz factor. It follows (see e.g. [Dui96]) that the distribution (27) is smooth outside the trajectory K(z,θ,β):={(x,t):x=z+tβθ}K(z,\theta,\beta):=\{(x,t):x=z+t\beta\theta\} and the order of singularity at each point corresponds to decay of aa with respect to ξ\xi. The signal is a moving point-like singularity even if all the charge is not necessarily concentrated at a single point. The current density obeys the continuity equation (conservation of charge)

divJ(x,t)+tρ(x,t)=0\text{div}J(x,t)+\partial_{t}\rho(x,t)=0

which follows from (21), (22) and the fact that divcurl=0\text{div}\ \text{curl}=0. By substitution one checks that the equation is solved by

J(x,t)=JI+JS(x,t),JI=(3ei(xztβθ)ξb(x,ξ)𝑑ξ)θJ(x,t)=J_{I}+J_{S}(x,t),\quad J_{I}=\left(\int_{\mathbb{R}^{3}}e^{i(x-z-t\beta\theta)\cdot\xi}b(x,\xi)d\xi\right)\theta

where b(x,ξ)j=mbj(x,ξ)b(x,\xi)\sim\sum_{j=-m}^{\infty}b_{-j}(x,\xi) is defined recursively by

(29) bm(x,ξ)=βam(x,ξ),\displaystyle b_{m}(x,\xi)=\beta a_{m}(x,\xi),
(30) iξθbj1(x,ξ)+θbj(x,ξ)=iβξθaj1(x,ξ),j=1,2,3,\displaystyle i\xi\cdot\theta b_{j-1}(x,\xi)+\theta\cdot\nabla b_{j}(x,\xi)=i\beta\xi\cdot\theta a_{j-1}(x,\xi),\quad j=1,2,3,\dots

and JS(x,t)J_{S}(x,t) is any solenoidal field. For simplicity, let us assume that JS(x,t)J_{S}(x,t) is smooth. Define a Riemannian metric gg by

gjk(x):=δjlεlk(x)α2(x)det(ε(x)).g^{jk}(x):=\frac{\delta^{jl}\varepsilon_{l}^{k}(x)}{\alpha^{2}(x)\det(\varepsilon(x))}.

and set δα:=(1)kαdα1\delta_{\alpha}:=(-1)^{k}*\alpha d\alpha^{-1}* where * is the Hodge star on kk-forms with respect to gg. Under the requirements above the electric field Ej=δjkEkE_{j}=\delta_{jk}E^{k} satisfies (See [KLS06])

(t2Δα)jkEk=Sj,j=1,2,3,(\partial_{t}^{2}-\Delta_{\alpha})^{k}_{j}E_{k}=S_{j},\quad j=1,2,3,

where Δα:=dδαδαd\Delta_{\alpha}:=-d\delta_{\alpha}-\delta_{\alpha}d and

(31) Sk=\displaystyle S_{k}= 1α2detε(x)3ei(xztβθ)ξck(x,ξ)𝑑ξ,\displaystyle\frac{1}{\alpha^{2}\det\varepsilon(x)}\int_{\mathbb{R}^{3}}e^{i(x-z-t\beta\theta)\cdot\xi}c_{k}(x,\xi)d\xi,
(32) ck(x,ξ)\displaystyle c_{k}(x,\xi)\equiv i(ξkβ2(θξ)gjkθj)am(x,ξ)modSm1.\displaystyle i(\xi_{k}-\beta^{2}(\theta\cdot\xi)g_{jk}\theta^{j})a_{m}(x,\xi)\mod S^{m-1}.

By applying Leibniz’ rule, we deduce that the standard codifferential δ=d\delta=-*d* differs from δα\delta_{\alpha} by only terms of order 0. For instance, for a 1-form ω\omega we have that

δαω=αdα1ω=dωdααω=δωgjkkααωj.\delta_{\alpha}\omega=-*\alpha d\alpha^{-1}*\omega=-*d*\omega-*\frac{d\alpha}{\alpha}\wedge*\omega=\delta\omega-g^{jk}\frac{\partial_{k}\alpha}{\alpha}\omega_{j}.

Thus, the leading terms in Δα\Delta_{\alpha} coincide with the ones in the Hodge Laplacian dδδd-d\delta-\delta d. As shown in [MMMT16, Lemma 2.8], the leading part in the Hodge Laplacian for 1-forms is (t2gjkjk)I(\partial_{t}^{2}-g^{jk}\partial_{j}\partial_{k})I. In conclusion, we have that the electric field satisfies the preferred equation:

(t2gjkjk000t2gjkjk000t2gjkjk)(E1E2E3)+(F11F21F31F12F22F32F13F23F33)(E1E2E3)=(S1S2S3),\begin{pmatrix}\partial_{t}^{2}-g^{jk}\partial_{j}\partial_{k}&0&0\\ 0&\partial_{t}^{2}-g^{jk}\partial_{j}\partial_{k}&0\\ 0&0&\partial_{t}^{2}-g^{jk}\partial_{j}\partial_{k}\end{pmatrix}\begin{pmatrix}E_{1}\\ E_{2}\\ E_{3}\\ \end{pmatrix}+\begin{pmatrix}F^{1}_{1}&F^{1}_{2}&F_{3}^{1}\\ F^{2}_{1}&F^{2}_{2}&F^{2}_{3}\\ F^{3}_{1}&F^{3}_{2}&F^{3}_{3}\end{pmatrix}\begin{pmatrix}E_{1}\\ E_{2}\\ E_{3}\\ \end{pmatrix}=\begin{pmatrix}S_{1}\\ S_{2}\\ S_{3}\\ \end{pmatrix},

where FjkΨcl1(3)F_{j}^{k}\in\Psi^{1}_{cl}(\mathbb{R}^{3}), j,k=1,2,3j,k=1,2,3.

2. Preliminaries

Henceforth the symbol j\partial_{j} with integer jj stands for the spatial partial derivative xj\frac{\partial}{\partial x^{j}}.

2.1. Symbols

Let XX be a smooth manifold of dimension nn. The class Sρ,δm(X×k) S^{m}_{\rho,\delta}(X\times\mathbb{R}^{k})  of symbols of order mm\in\mathbb{R} and type (ρ,δ)[0,1]×[0,1](\rho,\delta)\in[0,1]\times[0,1] is defined as the space of functions aC(X×k)a\in C^{\infty}(X\times\mathbb{R}^{k}) that satisfy the following: For each compact KXK\subset X, αn\alpha\in\mathbb{N}^{n} and βk\beta\in\mathbb{N}^{k} there exists C=CK,α,β,a0C=C_{K,\alpha,\beta,a}\geq 0 such that

|xαξβa(x,ξ)|C(1+|ξ|)mρ|β|+δ|α|,(x,ξ)K×k.|\partial_{x}^{\alpha}\partial_{\xi}^{\beta}a(x,\xi)|\leq C(1+|\xi|)^{m-\rho|\beta|+\delta|\alpha|},\quad\forall(x,\xi)\in K\times\mathbb{R}^{k}.

Only symbols of type (1,0)(1,0) are considered and hence the notational simplification Sm(X×k):=S1,0m(X×k)S^{m}(X\times\mathbb{R}^{k}):=S_{1,0}^{m}(X\times\mathbb{R}^{k}). It follows from the definition that Sρ,δm(X×k) Sρ,δm(X×k) S^{m}_{\rho,\delta}(X\times\mathbb{R}^{k}) \subset S^{m^{\prime}}_{\rho,\delta}(X\times\mathbb{R}^{k})  for mmm\leq m^{\prime} and we define Sρ,δ(X×k)S_{\rho,\delta}^{-\infty}(X\times\mathbb{R}^{k}) as the limit set Sρ,δ(X×k):=mSρ,δm(X×k)S^{-\infty}_{\rho,\delta}(X\times\mathbb{R}^{k}):=\bigcap_{m\in\mathbb{R}}S^{m}_{\rho,\delta}(X\times\mathbb{R}^{k}). It is common to treat symbols modulo a residual term in Sρ,δS^{-\infty}_{\rho,\delta}. For a sequence of symbols ajSρ,δmj(X×k) a_{j}\in S^{m_{j}}_{\rho,\delta}(X\times\mathbb{R}^{k}) , j=1,2,3,j=1,2,3,\dots, with mjm_{j}\searrow-\infty there exists aSρ,δm1(X×k) a\in S^{m_{1}}_{\rho,\delta}(X\times\mathbb{R}^{k}) , unique up to a term in S(X×k)S^{-\infty}(X\times\mathbb{R}^{k}), such that

aj=1lajSρ,δml+1(X×k) ,a-\sum_{j=1}^{l}a_{j}\in S^{m_{l+1}}_{\rho,\delta}(X\times\mathbb{R}^{k}) ,

for every l=1,2,3,4,l=1,2,3,4,\dots This is denoted by a(x,ξ)j=1aj(x,ξ)a(x,\xi)\sim\sum_{j=1}^{\infty}a_{j}(x,\xi) and referred to as an asymptotic expansion or asymptotic development of aa.

The set of classical symbols Sclm(X×k)Sm(X×k)S_{cl}^{m}(X\times\mathbb{R}^{k})\subset S^{m}(X\times\mathbb{R}^{k}) is defined as the elements aSm(X×k)a\in S^{m}(X\times\mathbb{R}^{k}) of the form

(33) a(x,ξ)j=1(1χ(ξ))amj(x,ξ),a(x,\xi)\sim\sum_{j=1}^{\infty}(1-\chi(\xi))a_{m-j}(x,\xi),

where χCc(k)\chi\in C^{\infty}_{c}(\mathbb{R}^{k}) takes value 11 in a neighbourhood of 0 and amjC(X×(k{0}))a_{m-j}\in C^{\infty}(X\times(\mathbb{R}^{k}\setminus\{0\})) is positively homogeneous of degree mjm-j with respect to the variable ξk{0}\xi\in\mathbb{R}^{k}\setminus\{0\}. Even though positively homogeneous functions are usually not smooth, they are treated as symbols by omitting a smoothing near the origin. The notation a(x,ξ)j=1amj(x,ξ)a(x,\xi)\sim\sum_{j=1}^{\infty}a_{m-j}(x,\xi) for positively homogeneous amj(x,ξ)a_{m-j}(x,\xi) is used also for referring to (33). The choice of the compactly supported cut-off function χ\chi around 0 is not significant as changing the function within contributes only by a term in S(X×k)S^{-\infty}(X\times\mathbb{R}^{k}). See e.g. [GS94] for a nice introduction to the topic.

2.2. Lagrangian Distributions and Fourier Integral Operators

A submanifold Λ\Lambda of a symplectic manifold (M,σ)(M,\sigma) of dimension 2n2n is called Lagrangian if dimΛ=n\dim\Lambda=n and σq(v,w)=0\sigma_{q}(v,w)=0, for every v,wTqΛv,w\in T_{q}\Lambda and qΛq\in\Lambda. The cotangent bundle TXT^{*}X of a smooth manifold XX is a symplectic manifold with the canonical symplectic structure (see [GS94, §5]). The space Ir(Λ)=Ir(X;Λ)I^{r}(\Lambda)=I^{r}(X;\Lambda) of Lagrangian distributions of order rr on XX associated with a conic Lagrangian manifold ΛTX{0}\Lambda\subset T^{*}X\setminus\{0\} consists of distributions u𝒟(X)u\in\mathcal{D}^{\prime}(X) that can be expressed as a locally finite sum of oscillatory integrals of the form

(34) I(φ,a)=keiφ(x,ξ)a(x,ξ)𝑑ξ,aSrk2+n4(X×k),I(\varphi,a)=\int_{\mathbb{R}^{k}}e^{i\varphi(x,\xi)}a(x,\xi)d\xi,\quad a\in S^{r-\frac{k}{2}+\frac{n}{4}}(X\times\mathbb{R}^{k}),

where φC(X×k)\varphi\in C^{\infty}(X\times\mathbb{R}^{k}) is a non-degenerate phase function, homogeneous of degree 1 with respect to ξ\xi, such that that the manifold Λ\Lambda coincides locally with the set Λφ={(x,dxφ(x,ξ)):dξφ(x,ξ)=0,ξ0,xX}\Lambda_{\varphi}=\{(x,d_{x}\varphi(x,\xi)):d_{\xi}\varphi(x,\xi)=0,\ \xi\neq 0,\ x\in X\} (see e.g. [Dui96] for details). For uIr(X;Λ)u\in I^{r}(X;\Lambda) the wave front set satisfies WF(u)ΛWF(u)\subset\Lambda and the asymptotic behaviour of the symbol defines the degree of regularity at given conic neighbourhood. For instance, if χaS(X×k)\chi a\in S^{-\infty}(X\times\mathbb{R}^{k}) where χC(X×k)\chi\in C^{\infty}(X\times\mathbb{R}^{k}) is positively homogeneous of degree 0 with respect to ξ\xi, then the wave front set of the oscillatory integral (34) does not meet a conic neighbourhood inside the support of χ\chi. As a special case, we have I(X;Λ)C(X)I^{-\infty}(X;\Lambda)\subset C^{\infty}(X). If two symbols aa and bb satisfy abS(X×k)a-b\in S^{-\infty}(X\times\mathbb{R}^{k}), i.e. abmodS(X×k)a\equiv b\mod S^{-\infty}(X\times\mathbb{R}^{k}), then I(φ,a)I(φ,b)=I(φ,ab)C(X)I(\varphi,a)-I(\varphi,b)=I(\varphi,a-b)\in C^{\infty}(X), that is, I(φ,a)I(φ,b)modC(X)I(\varphi,a)\equiv I(\varphi,b)\mod C^{\infty}(X) for the associated oscillatory integrals. We do not distinguish between distributions that differ by a smooth term.

A special class of Lagrangian distributions is obtained by assuming that Λ\Lambda equals NMN^{*}M, the conormal bundle of a submanifold MXM\subset X. Such elements are referred to as distributions conormal to MM. For example, the delta distribution u(x,t)=δ(tx)u(x,t)=\delta(t-x) on 2\mathbb{R}^{2} is conormal to the diagonal {(x,t)2:t=x}\{(x,t)\in\mathbb{R}^{2}:t=x\}. The class of distributions conormal to MM is often denoted by Im(M)I^{m}(M), where mm refers to the order of the symbol. Thus, Im(M)=Im+k2n4(X;NM)I^{m}(M)=I^{m+\frac{k}{2}-\frac{n}{4}}(X;N^{*}M), where kk is the codimension of MM in XX.

Considering a Lagrangian distribution ff on X×YX\times Y as a kernel gives an operator

F=Ff:C(Y)𝒟(X),Fϕ,ψ:=f,ψϕ,ϕC(Y),ψC(X).F=F_{f}:C^{\infty}(Y)\rightarrow\mathcal{D}^{\prime}(X),\quad\langle F\phi,\psi\rangle:=\langle f,\psi\otimes\phi\rangle,\quad\phi\in C^{\infty}(Y),\quad\psi\in C^{\infty}(X).

Functions of this form are called (standard) Fourier integral operators.111An operator and the kernel of it are often treated as the same object. For a closed cone ΓTY\Gamma\subset T^{*}Y that does not meet WFY(F):={(y,η):xXsuch that(x,0;y,η)WF(f)}WF_{Y}^{\prime}(F):=\{(y,\eta):\exists x\in X\ \text{such that}\ (x,0;y,\eta)\in WF(f)\} the definition can be extended to Γ(X)={v(X):WF(v)Γ}\mathcal{E}_{\Gamma}^{\prime}(X)=\{v\in\mathcal{E}^{\prime}(X):WF(v)\subset\Gamma\} and further to 𝒟Γ(X)={v𝒟(X):WF(v)Γ}\mathcal{D}_{\Gamma}^{\prime}(X)=\{v\in\mathcal{D}^{\prime}(X):WF(v)\subset\Gamma\} if the operator is properly supported. See e.g. [Dui96, Corollary 1.3.8] for details. If fIm(X×Y;Λ)f\in I^{m}(X\times Y;\Lambda) we say that the associated operator FfF_{f} lies in Im(X,Y;Λ)I^{m}(X,Y;\Lambda^{\prime}). The manifold Λ\Lambda^{\prime} refers to Λ:={(x,ξ;y,η):(x,ξ;y,η)Λ}\Lambda^{\prime}:=\{(x,\xi;y,\eta):(x,\xi;y,-\eta)\in\Lambda\} which is Lagrangian with respect to the symplectic form σXσY\sigma_{X}-\sigma_{Y}. Such a manifold is referred to as the canonical relation. Conversely, if a manifold ΛTX×TY\Lambda\subset T^{*}X\times T^{*}Y is a canonical relation, then Λ\Lambda^{\prime} is Lagrangian manifold for the standard symplectic form σX+σY\sigma_{X}+\sigma_{Y} and the kernel is a Lagrangian distribution on X×YX\times Y. A Fourier integral operator with a smooth kernel defines a regularizing map (X)C(X)\mathcal{E}^{\prime}(X)\rightarrow C^{\infty}(X) which is usually identified with the zero element.

An operator that admits the diagonal diag(TX{0})\text{diag}(T^{*}X\setminus\{0\}) as a canonical relation is called pseudo-differential operator. We denote by Ψm(X)\Psi^{m}(X) the class of pseudo-differential operators of order mm on XX. Pseudo-differential operators with classical symbols is denoted by Ψclm(X)\Psi^{m}_{cl}(X). All the pseudo-differential operators considered in this paper have X=nX=\mathbb{R}^{n}. By placing a smooth cut-off χ(xy)\chi(x-y) to the kernel (see [GS94, Remark 3.3]) a pseudo-differential operator can be identified with a properly supported one.

2.3. Transversal Intersection Calculus

A composition of two Fourier integral operators is not necessarily well defined as a Fourier integral operator. Sufficient conditions together with associated composition calculus, often referred to as transversal intersection calculus, was developed in [H7̈1], [DH72] by Hörmander and Duistermaat. Perhaps the most demanding one of the conditions for a composition FGF\circ G of two operators FIm(X,Y;Λ1)F\in I^{m}(X,Y;\Lambda_{1}) and GIm(Y,Z;Λ2)G\in I^{m^{\prime}}(Y,Z;\Lambda_{2}) with canonical relations Λ1\Lambda_{1} and Λ2\Lambda_{2} to be admissible in the framework is that the product Λ1×Λ2\Lambda_{1}\times\Lambda_{2} intersects the manifold TX×diagTY×TZT^{*}X\times\text{diag}T^{*}Y\times T^{*}Z transversally. Provided that the required conditions are satisfied, we have

(35) FGIm+m(X,Z;Λ1Λ2),\displaystyle F\circ G\in I^{m+m^{\prime}}(X,Z;\Lambda_{1}\circ\Lambda_{2}),
(36) Λ1Λ2:={(x,ξ;z,σ)TX{0}|(y,η):(x,ξ;y,η)Λ1,(y,η;z,σ)Λ2}.\displaystyle\Lambda_{1}\circ\Lambda_{2}:=\big{\{}(x,\xi;z,\sigma)\in T^{*}X\setminus\{0\}\ \big{|}\ \exists(y,\eta):(x,\xi;y,\eta)\in\Lambda_{1},\ (y,\eta;z,\sigma)\in\Lambda_{2}\big{\}}.

Moreover,

(37) σm+m(FG)(x,ξ;z,σ)=(y,η)Γσm(F)(x,ξ;y,η)σm(G)(y,η;z,σ),\sigma_{m+m^{\prime}}(F\circ G)(x,\xi;z,\sigma)=\sum_{(y,\eta)\in\Gamma}\sigma_{m}(F)(x,\xi;y,\eta)\sigma_{m^{\prime}}(G)(y,\eta;z,\sigma),

where σmSm/Sm1\sigma_{m}\in S^{m}/S^{m-1} refers to the principal symbol (i.e. the leading term in an asymptotic development) on the canonical relation and Γ=Γx,ξ;y,η:={(y,η):(x,ξ;y,η)Λ1,(y,η;z,σ)Λ2}\Gamma=\Gamma_{x,\xi;y,\eta}:=\{(y,\eta):(x,\xi;y,\eta)\in\Lambda_{1},\ (y,\eta;z,\sigma)\in\Lambda_{2}\}. The diagonal acts as an identity element on canonical relations:

diag(TX{0})Λ2=Λ2,Λ1diag(TY{0})=Λ1,\text{diag}(T^{*}X\setminus\{0\})\circ\Lambda_{2}=\Lambda_{2},\quad\Lambda_{1}\circ\text{diag}(T^{*}Y\setminus\{0\})=\Lambda_{1},

Considering Lagrangian distributions on YY as Fourier integral operators on Y×ZY\times Z with trivial ZZ one deduces

FuIm+r(X;ΛΛ~),forFIm(X,Y;Λ),uIr(Y;Λ~),Fu\in I^{m+r}(X;\Lambda\circ\tilde{\Lambda}),\quad\text{for}\quad F\in I^{m}(X,Y;\Lambda),\quad u\in I^{r}(Y;\tilde{\Lambda}),

provided that the required conditions hold. The element ΛΛ~\Lambda\circ\tilde{\Lambda} above refers to

ΛΛ~:={(x,ξ):(y,η)Λ~;(x,ξ;y,η)Λ}.\Lambda\circ\tilde{\Lambda}:=\{(x,\xi):\exists(y,\eta)\in\tilde{\Lambda};\ (x,\xi;y,\eta)\in\Lambda\}.

For a pseudo-differential operator AA we obtain

AuIm+r(X;diag(Tn{0})Λ~))=Im+r(X;Λ~)Au\in I^{m+r}(X;\text{diag}(T^{*}\mathbb{R}^{n}\setminus\{0\})\circ\tilde{\Lambda}))=I^{m+r}(X;\tilde{\Lambda})

(cf. microlocality WF(Au)WF(u)WF(Au)\subset WF(u)).

2.4. Distributions associated with a pair of Lagrangian manifolds.

We also consider the class Ir(X;Λ0,Λ1)I^{r}(X;\Lambda_{0},\Lambda_{1}) of Lagrangian distributions associated with a pair of cleanly intersecting Lagrangian manifolds Λ0,Λ1TX\Lambda_{0},\Lambda_{1}\subset T^{*}X instead of a single manifold. Calculus for operators with kernels of this form was developed by Melrose and Uhlmann in [MU79]. The theory is needed for describing parametrices of pseudo-differential operators. More precisely, a parametrix QQ for a pseudo-differential operator PP of real principal type on XX is associated with the pair

(diag(TX{0}),ΛP)(T(X)×T(X))×(T(X)×T(X)),(\text{diag}(T^{*}X\setminus\{0\}),\Lambda_{P})\subset(T^{*}(X)\times T^{*}(X))\times(T^{*}(X)\times T^{*}(X)),

of canonical relations where ΛP\Lambda_{P} consists of pairs ((x,ξ),(y,η))(TX×TX){0,0}((x,\xi),(y,\eta))\in(T^{*}X\times T^{*}X)\setminus\{0,0\} that lie in a same bicharacteristic. It is shown in [GU93, Proposition 2.1] that for a distribution uu in the class Im(X;Λ)I^{m}(X;\Lambda), where the Lagrangian manifold ΛTX\Lambda\subset T^{*}X intersects the characteristic manifold of PP transversally and each bicharacteristic of PP intersects the manifold a finite number of times, the composition QuQu is a Lagrangian distribution that corresponds to the pair

(diag(TX{0})Λ,ΛPΛ)=(Λ,ΛPΛ).(\text{diag}(T^{*}X\setminus\{0\})\circ\Lambda\ ,\ \Lambda_{P}\circ\Lambda)=(\Lambda,\Lambda_{P}\circ\Lambda).

Moreover, the identity (37) extends to (ΛPΛ)Λ(\Lambda_{P}\circ\Lambda)\setminus\Lambda.

2.5. Matrix Notation

We define the class M3(Ψm(4))M_{3}(\Psi^{m}(\mathbb{R}^{4})) as 3×33\times 3 matrices A=[Ajk]j,k=1,2,3A=[A_{j}^{k}]_{j,k=1,2,3} of pseudo-differential operators AjkΨm(4)A_{j}^{k}\in\Psi^{m}(\mathbb{R}^{4}), j,k=1,2,3j,k=1,2,3, of degree mm on 4\mathbb{R}^{4} as entries. The symbols of AjkA^{k}_{j}, j,k=1,2,3j,k=1,2,3 form an element in M3(Sm(4×4)):={[ajk]j,k=1,2,3:ajkSm(4×4)}M_{3}(S^{m}(\mathbb{R}^{4}\times\mathbb{R}^{4})):=\{[a^{k}_{j}]_{j,k=1,2,3}:a_{j}^{k}\in S^{m}(\mathbb{R}^{4}\times\mathbb{R}^{4})\} of matrices with entries in the symbol class Sm(4×4)S^{m}(\mathbb{R}^{4}\times\mathbb{R}^{4}). Each AM3(Ψm(4))A\in M_{3}(\Psi^{m}(\mathbb{R}^{4})) operates on a vector-valued distribution S=(S1,S2,S3)S=(S_{1},S_{2},S_{3}), Sj(4)S_{j}\in\mathcal{E}^{\prime}(\mathbb{R}^{4}), j=1,2,3j=1,2,3, in the obvious way:

(AS)j=AjkSk𝒟(4),j=1,2,3.(AS)_{j}=A^{k}_{j}S_{k}\in\mathcal{D}^{\prime}(\mathbb{R}^{4}),\quad j=1,2,3.

This is well defined also for Sk𝒟(4)S_{k}\in\mathcal{D}^{\prime}(\mathbb{R}^{4}) if the entries AjkA^{k}_{j} are properly supported. In addition, we define the composition ABM3(Ψm1+m2(4))AB\in M_{3}(\Psi^{m_{1}+m_{2}}(\mathbb{R}^{4})) of AM3(Ψm1(4))A\in M_{3}(\Psi^{m_{1}}(\mathbb{R}^{4})) and BM3(Ψm2(4))B\in M_{3}(\Psi^{m_{2}}(\mathbb{R}^{4})) by

(AB)jk:=AjlBlk,j,k=1,2,3.(AB)^{k}_{j}:=A^{l}_{j}B_{l}^{k},\quad j,k=1,2,3.

Any scalar operator LΨl(4)L\in\Psi^{l}(\mathbb{R}^{4}) is identified with the diagonal matrix [Lδjk]j,k=1,2,3[L\delta^{k}_{j}]_{j,k=1,2,3}, where δjk\delta^{k}_{j} stands for the Kronecker delta. That is;

(LS)j=LSj𝒟(4),j=1,2,3.(LS)_{j}=LS_{j}\in\mathcal{D}^{\prime}(\mathbb{R}^{4}),\quad j=1,2,3.

for S=(S1,S2,S3)S=(S_{1},S_{2},S_{3}), Sj𝒟(4)S_{j}\in\mathcal{D}^{\prime}(\mathbb{R}^{4}), and

(LA)jk=LAjkM3(Ψm+l(4)),(LA)^{k}_{j}=LA^{k}_{j}\in M_{3}(\Psi^{m+l}(\mathbb{R}^{4})),

for AM3(Ψm(4))A\in M_{3}(\Psi^{m}(\mathbb{R}^{4})).

3. Microlocal Methods

This section is based on the techniques developed in [H7̈1],[DH72], [MU79] and [GU93] for scalar operators of real principal type. The results here are modifications of the original ones. Some of the proofs are moved to Appendix A as they do not differ significantly from the ones for scalar-valued operators. The objective is to prove Proposition 3.1 which is the main tool in this paper.

We use the notation ±\pm to indicate a sign +/+/- that can be chosen to be either ++ or -. We also define :=±\mp:=-\pm. For a conic neighbourhood 𝒱T4\mathcal{V}\subset T^{*}\mathbb{R}^{4} and X,YM3(Ψm(4))X,Y\in M_{3}(\Psi^{m}(\mathbb{R}^{4})) we say that

XYmicrolocally in𝒱X\equiv Y\quad\text{microlocally in}\quad\mathcal{V}

if

𝒱WF(XjkYjk)=,j,k=1,2,3,\mathcal{V}\cap WF(X^{k}_{j}-Y^{k}_{j})=\emptyset,\quad\forall j,k=1,2,3,

where WF(A)WF(A) refers to the wave front set of a pseudo-differential operator, that is, WF(A)=diag(WF(A))WF^{\prime}(A)=\text{diag}(WF(A)) as a Fourier integral operator. Let us begin by factoring a vector-valued operator P(t2gjkkj)I+FP\equiv(\partial_{t}^{2}-g^{jk}\partial_{k}\partial_{j})I+F into first order operators. For a proof, see the appendix.

Lemma 3.1.

Consider a differential operator PM3(Ψcl2(4))P\in M_{3}(\Psi^{2}_{cl}(\mathbb{R}^{4})) of the form

P=(t2gjkjk)+F,P=(\partial_{t}^{2}-g^{jk}\partial_{j}\partial_{k})+F,

where gg is a Riemannian metric and F=Fjk(x,Dx)kdxjF=F^{k}_{j}(x,D_{x})\ \partial_{k}\otimes dx^{j} is a stationary first order operator. Then, for every ϵ>0\epsilon>0 there are operators W=W±M3(Ψcl1(4))W=W_{\pm}\in M_{3}(\Psi^{1}_{cl}(\mathbb{R}^{4})), Z=Z±M3(Ψcl1(4))Z=Z_{\pm}\in M_{3}(\Psi^{1}_{cl}(\mathbb{R}^{4})), and h=h±M3(Ψcl0(4))h=h_{\pm}\in M_{3}(\Psi^{0}_{cl}(\mathbb{R}^{4})) such that

P(DtW)Z,P\equiv(D_{t}-W)Z,

and

(38) W±|Dx|g+h,\displaystyle W\equiv\pm|D_{x}|_{g}+h,
(39) Z(Dt±|Dx|g)h,\displaystyle Z\equiv-(D_{t}\pm|D_{x}|_{g})-h,

microlocally in 𝒱={(x,t;ξ,ω)T4{0}:|ω|<ϵ|ξ|}\mathcal{V}=\{(x,t;\xi,\omega)\in T^{*}\mathbb{R}^{4}\setminus\{0\}:|\omega|<\epsilon|\xi|\}.

Define L±4={(x,t,ξ,ω)T4:ω=±|ξ|g}L^{\pm}\mathbb{R}^{4}=\{(x,t,\xi,\omega)\in T^{*}\mathbb{R}^{4}:\omega=\pm|\xi|_{g}\}, where gg is a Riemannian metric. Throughout this section we consider a non-zero covector (x0,t0,ξ0,ω0)(x_{0},t_{0},\xi_{0},\omega_{0})  in T4T^{*}\mathbb{R}^{4} with ξ00\xi_{0}\neq 0 and ω0|ξ0|g\omega_{0}\neq\mp|\xi_{0}|_{g} for a given sign ±{+,}\pm\in\{+,-\} and a fixed homogeneous canonical transformation

(40) (y,s,η,ρ)(y,s,η,ρ)T4,(y,s,\eta,\rho)\mapsto\mathcal{H}(y,s,\eta,\rho)\in T^{*}\mathbb{R}^{4},

defined on a conic neighbourhood of the covector (y0,s0,η0,ρ0)(y_{0},s_{0},\eta_{0},\rho_{0}) with

1(x0,t0,ξ0,ω0)=(y0,s0,η0,ρ0).\mathcal{H}^{-1}(x_{0},t_{0},\xi_{0},\omega_{0})=(y_{0},s_{0},\eta_{0},\rho_{0}).

We are only interested in (x0,t0,ξ0,ω0)L±4(x_{0},t_{0},\xi_{0},\omega_{0})\in L^{\pm}\mathbb{R}^{4} which shall be required. Moreover, we assume that the last coordinate of 1\mathcal{H}^{-1} equals ω|ξ|g\omega\mp|\xi|_{g}, i.e, ρ(x,t,ξ,ω)=ω|ξ|g\mathcal{H}_{*}\rho(x,t,\xi,\omega)=\omega\mp|\xi|_{g} (see Darboux’ theorem [Dui96, Theorem 3.5.6]). The graph of \mathcal{H} is denoted by GG_{\mathcal{H}}. The graph of 1\mathcal{H}^{-1}, denoted by G1G_{\mathcal{H}}^{-1}, is obtained by changing the order of coordinates in GG_{\mathcal{H}}.

In the next step we follow Egorov’s theorem [GS94, Theorem 10.1]. We fix Fourier integral operators 𝒜Icl0(4,4;G)\mathcal{A}\in I^{0}_{cl}(\mathbb{R}^{4},\mathbb{R}^{4};G_{\mathcal{H}}) and Icl0(4,4;G1)\mathcal{B}\in I^{0}_{cl}(\mathbb{R}^{4},\mathbb{R}^{4};G_{\mathcal{H}}^{-1}) with locally reciprocal principal symbols (positively homogeneous of degree 0) such that (x0,t0,ξ0,ω0,y0,s0,η0,ρ0)(x_{0},t_{0},\xi_{0},\omega_{0},y_{0},s_{0},\eta_{0},\rho_{0}) is noncharacteristic for 𝒜\mathcal{A}, (y0,s0,η0,ρ0,x0,t0,ξ0,ω0)(y_{0},s_{0},\eta_{0},\rho_{0},x_{0},t_{0},\xi_{0},\omega_{0}) is noncharacteristic for \mathcal{B} and there is a conic neighbourhood 𝒱1\mathcal{V}_{1} of (y0,s0,η0,ρ0)(y_{0},s_{0},\eta_{0},\rho_{0}) and a conic neighbourhood 𝒱2\mathcal{V}_{2} of (x0,t0,ξ0,ω0)(x_{0},t_{0},\xi_{0},\omega_{0}) such that

(41) WF(𝒜id)𝒱1=,WF(𝒜id)𝒱2=.\displaystyle WF(\mathcal{B}\mathcal{A}-id)\cap\mathcal{V}_{1}=\emptyset,\quad WF(\mathcal{A}\mathcal{B}-id)\cap\mathcal{V}_{2}=\emptyset.

We may assume that the Schwartz kernels of 𝒜\mathcal{A} and \mathcal{B} are compactly supported distributions. The principal symbol of (Dt|Dx|g)𝒜\mathcal{B}(D_{t}\mp|D_{x}|_{g})\mathcal{A} equals the coordinate ρ\rho in a conic neighbourhood of (y0,s0,η0,ρ0)(y_{0},s_{0},\eta_{0},\rho_{0}). Hence, microlocally near (y0,s0,η0,ρ0)(y_{0},s_{0},\eta_{0},\rho_{0}),

(DtW)𝒜(Dt|Dx|g+h)𝒜Ds+q,\mathcal{B}(D_{t}\mp W)\mathcal{A}\equiv\mathcal{B}(D_{t}\mp|D_{x}|_{g}+h)\mathcal{A}\equiv D_{s}+q,

modulo M3(Ψ(4))M_{3}(\Psi^{-\infty}(\mathbb{R}^{4})), where WW is as in Lemma 3.1 and qM3(Ψcl0(4))q\in M_{3}(\Psi^{0}_{cl}(\mathbb{R}^{4})).

Definition 3.1.

Let Shomm(4×4)S^{m}_{hom}(\mathbb{R}^{4}\times\mathbb{R}^{4}) stand for the symbols a(x,t,ξ,ω)a(x,t,\xi,\omega) in Sm(4×4)S^{m}(\mathbb{R}^{4}\times\mathbb{R}^{4}) that outside a bounded neighbourhood of the origin are positively homogeneous of degree mm in (ξ,ω)(\xi,\omega).

The following lemma provides a way to get rid of the extra term qq (cf. [DH72, Proposition 6.1.4], [GS94, Lemma 10.3]).

Lemma 3.2.

Let qM3(Ψcl0(4))q\in M_{3}(\Psi^{0}_{cl}(\mathbb{R}^{4})). There exists AM3(Ψcl0(4))A\in M_{3}(\Psi^{0}_{cl}(\mathbb{R}^{4})) such that

(Ds+q)AADsM3(Ψ(4)).(D_{s}+q)A-AD_{s}\in M_{3}(\Psi^{-\infty}(\mathbb{R}^{4})).

Moreover, there is a neighbourhood UU of (y0,s0)(y_{0},s_{0}) in 4\mathbb{R}^{4} and a0M3(Shom0(4×4))a_{0}\in M_{3}(S^{0}_{hom}(\mathbb{R}^{4}\times\mathbb{R}^{4})) that is invertible as a matrix at every point (y,s,η,ρ)U×4(y,s,\eta,\rho)\in U\times\mathbb{R}^{4} and satisfies

Aa0(y,s,Dy,Ds)M3(Ψcl1(4)),A-a_{0}(y,s,D_{y},D_{s})\in M_{3}(\Psi^{-1}_{cl}(\mathbb{R}^{4})),
Proof of Lemma 3.2.

Write each entry qkjq^{j}_{k} of qq in the asymptotic form

qkj(y,s,η,ρ)m=0qj,mk(y,s,η,ρ),q^{j}_{k}(y,s,\eta,\rho)\sim\sum_{m=0}^{\infty}q^{k}_{j,-m}(y,s,\eta,\rho),

where qj,mk(y,s,η,ρ)q^{k}_{j,-m}(y,s,\eta,\rho) is positively homogeneous of degree m-m as a function of η,ρ\eta,\rho. We substitute the ansatz Ajk=ajk(y,s,Dy,Ds)A^{k}_{j}=a^{k}_{j}(y,s,D_{y},D_{s}),

ajk(y,s,η,ρ)m=0aj,mk(y,s,η,ρ),j,k=1,2,3,a^{k}_{j}(y,s,\eta,\rho)\sim\sum_{m=0}^{\infty}a^{k}_{j,-m}(y,s,\eta,\rho),\quad j,k=1,2,3,

in

(42) (Ds+q)AADsM3(Ψ(4)),(D_{s}+q)A-AD_{s}\in M_{3}(\Psi^{-\infty}(\mathbb{R}^{4})),

apply the formula (70), and derive sufficient conditions by putting together the terms that are of the same degree. The equation (42) can be written as

(43) [Ds,Akj]+qklAljΨ(4),j,k=1,2,3.[D_{s},A^{j}_{k}]+q^{l}_{k}A^{j}_{l}\in\Psi^{-\infty}(\mathbb{R}^{4}),\quad j,k=1,2,3.

For every m=1,2,3,m=1,2,3,\dots, let ama_{-m} and qmq_{-m} be the matrices with entries aj,mka_{j,-m}^{k} and qj,mkq_{j,-m}^{k}, respectively. Setting the leading terms in the symbol of (43) equal zero implies

i{ρ,aj,0k}+qj,0lal,0k=0,-i\{\rho,a_{j,0}^{k}\}+q^{l}_{j,0}a^{k}_{l,0}=0,

that is,

sa0=iq0a0.\partial_{s}a_{0}=-iq_{0}a_{0}.

Instead of trying to solve the equation explicitly we refer to the basic theory of ordinary differential equations for existence of solutions. Provided an initial value at s=s0s=s_{0}, both uniqueness and existence of sa0(y,s,η,ρ)s\mapsto a_{0}(y,s,\eta,\rho) on a closed interval follows from [Rei71, Theorem 5.1]. We fix the initial value a0(y,s0,η,ρ)=Ia_{0}(y,s_{0},\eta,\rho)=I. Positive homogeneity of q0q_{0} in ρ,η\rho,\eta then imply that the solution really is positively homogeneous of degree 0 in ρ,η\rho,\eta. Moreover, regularity of the solution in (y,s,η,ρ)(y,s,\eta,\rho) is a consequence of [Rei71, Theorem 10.5]. As the initial value is an invertible matrix and the solution is positively homogeneous of degree 0, it follows by continuity that there is an open neighbourhood UU of (y0,s0)(y_{0},s_{0}) in 4\mathbb{R}^{4} such that a0(s,y,η,ρ)a_{0}(s,y,\eta,\rho) is invertible matrix for every (y,s,η,ρ)TU{0}(y,s,\eta,\rho)\in T^{*}U\setminus\{0\}. For the rest of the terms one derives

(44) sam=iq0amifm,m=1,2,3,\partial_{s}a_{-m}=-iq_{0}a_{-m}-if_{-m},\quad m=1,2,3,\dots

where fmf_{-m} is a matrix depending on entries of η,ραqwDy,sαav\partial_{\eta,\rho}^{\alpha}q_{-w}D_{y,s}^{\alpha}a_{-v}, where v=0,,m1v=0,\dots,m-1 and other indices satisfy |α|+m=v+w|\alpha|+m=v+w. Each ama_{-m} can be solved recursively from these equations. Proceeding inductively, similar arguments as in the case m=0m=0 apply for existence and regularity of solutions with smooth initial values. It is also straightforward to check that positive homogeneity of degree m-m in η,ρ\eta,\rho is consistent with the solutions.

The following lemma is the construction of a microlocal parametrix. The proof (see Appendix A) does not differ significantly from the scalar case (see [DH72]).

Lemma 3.3.

For LM3(Ψclm(4))L\in M_{3}(\Psi^{m}_{cl}(\mathbb{R}^{4})), mm\in\mathbb{R}, let amM3(Shomm(4×4))a_{m}\in M_{3}(S^{m}_{hom}(\mathbb{R}^{4}\times\mathbb{R}^{4})) be the principal symbol of LL in the sense that

Lam(x,t,Dx,Dt)M3(Ψclm1(4)).L-a_{m}(x,t,D_{x},D_{t})\in M_{3}(\Psi^{m-1}_{cl}(\mathbb{R}^{4})).

Assume that am(x,t,ξ,ω)a_{m}(x,t,\xi,\omega) is an invertible matrix at some (x0,t0,ξ0,ω0)T4(x_{0},t_{0},\xi_{0},\omega_{0})\in T^{*}\mathbb{R}^{4}. Then, there is RM3(Ψclm(4))R\in M_{3}(\Psi^{-m}_{cl}(\mathbb{R}^{4})) and a conic neighbourhood 𝒱\mathcal{V} of (x0,t0,ξ0,ω0)(x_{0},t_{0},\xi_{0},\omega_{0}) in T4T^{*}\mathbb{R}^{4} satisfying the conditions (i) and (ii) below.

  1. (i)

    RR is a (two sided) microlocal inverse of LL in 𝒱\mathcal{V}:

    WF(LjlRlkδjk)𝒱=,andWF(RjlLlkδjk)𝒱=.WF(L^{l}_{j}R^{k}_{l}-\delta^{k}_{j})\cap\mathcal{V}=\emptyset,\quad\text{and}\quad WF(R^{l}_{j}L^{k}_{l}-\delta^{k}_{j})\cap\mathcal{V}=\emptyset.
  2. (ii)

    The principal symbols of RR and LL are reciprocal in 𝒱\mathcal{V}: There is bmM3(Shomm(4×4))b_{-m}\in M_{3}(S^{-m}_{hom}(\mathbb{R}^{4}\times\mathbb{R}^{4})) satisfying

    Rbm(x,t,Dx,Dt)M3(Ψclm1(4))R-b_{-m}(x,t,D_{x},D_{t})\in M_{3}(\Psi^{-m-1}_{cl}(\mathbb{R}^{4}))

    and

    bm(x,t,ξ,ω)=(am)1(x,t,ξ,ω),(x,t,ξ,ω)𝒱.b_{-m}(x,t,\xi,\omega)=(a_{m})^{-1}(x,t,\xi,\omega),\quad\forall(x,t,\xi,\omega)\in\mathcal{V}.

Applying microlocality222WF(Au)WF(u)WF(Au)\subset WF(u) for AΨm(n)A\in\Psi^{m}(\mathbb{R}^{n}) to Lemma 3.3 yields the following:

Corollary 3.1.

Let LL, RR and 𝒱\mathcal{V} be as in Lemma 3.3. Then,

𝒱j=1,2,3WF(Ljkuk)=𝒱j=1,2,3WF(uj)=𝒱j=1,2,3WF(Rjkuk).\mathcal{V}\cap\bigcup_{j=1,2,3}WF(L_{j}^{k}u_{k})=\mathcal{V}\cap\bigcup_{j=1,2,3}WF(u_{j})=\mathcal{V}\cap\bigcup_{j=1,2,3}WF(R_{j}^{k}u_{k}).

for uk(4)u_{k}\in\mathcal{E}^{\prime}(\mathbb{R}^{4}), k=1,2,3k=1,2,3.

Let AA be as in Lemma 3.2 and let BB be the corresponding microlocal inverse of it, as in Lemma 3.3. Define 𝒳:=𝒜A\mathcal{X}:=\mathcal{A}A and 𝒴:=B\mathcal{Y}:=B\mathcal{B}. Applying the standard FIO calculus to Ajk,BjkΨcl0(4)A^{k}_{j},B^{k}_{j}\in\Psi_{cl}^{0}(\mathbb{R}^{4}), 𝒜Icl0(4,4;G)\mathcal{A}\in I^{0}_{cl}(\mathbb{R}^{4},\mathbb{R}^{4};G_{\mathcal{H}}), and Icl0(4,4;G1)\mathcal{B}\in I^{0}_{cl}(\mathbb{R}^{4},\mathbb{R}^{4};G_{\mathcal{H}}^{-1}) yields

(45) 𝒳jk\displaystyle\mathcal{X}^{k}_{j} =𝒜AjkIcl0+0(4,4;Gdiag(T4{0 }))=Icl0(4,4;G),\displaystyle=\mathcal{A}A^{k}_{j}\in I^{0+0}_{cl}(\mathbb{R}^{4},\mathbb{R}^{4};G_{\mathcal{H}}\circ\text{diag}(T^{*}\mathbb{R}^{4}\setminus\{0 \}))=I^{0}_{cl}(\mathbb{R}^{4},\mathbb{R}^{4};G_{\mathcal{H}}),
(46) 𝒴jk\displaystyle\mathcal{Y}^{k}_{j} =BjkIcl0+0(4,4;diag(T4{0})G1)=Icl0(4,4;G1),\displaystyle=B^{k}_{j}\mathcal{B}\in I^{0+0}_{cl}(\mathbb{R}^{4},\mathbb{R}^{4};\text{diag}(T^{*}\mathbb{R}^{4}\setminus\{0\})\circ G_{\mathcal{H}}^{-1})=I^{0}_{cl}(\mathbb{R}^{4},\mathbb{R}^{4};G_{\mathcal{H}}^{-1}),

together with the following identities for (y,s,η,ρ)(y,s,\eta,\rho) in the domain of \mathcal{H}:

(47) σ(𝒳jk)((y,s,η,ρ),y,s,η,ρ)=σ(𝒜)((y,s,η,ρ),y,s,η,ρ)σ(Ajk)(diag(y,s,η,ρ)),σ(𝒴jk)(y,s,η,ρ,(y,s,η,ρ))=σ(Bjk)(diag(y,s,η,ρ))σ()(y,s,η,ρ,(y,s,η,ρ)).\begin{split}\sigma(\mathcal{X}^{k}_{j})(\mathcal{H}(y,s,\eta,\rho),y,s,\eta,\rho)&=\sigma(\mathcal{A})(\mathcal{H}(y,s,\eta,\rho),y,s,\eta,\rho)\sigma(A^{k}_{j})(\text{diag}(y,s,\eta,\rho)),\\ \sigma(\mathcal{Y}^{k}_{j})(y,s,\eta,\rho,\mathcal{H}(y,s,\eta,\rho))&=\sigma(B^{k}_{j})(\text{diag}\mathcal{H}(y,s,\eta,\rho))\sigma(\mathcal{B})(y,s,\eta,\rho,\mathcal{H}(y,s,\eta,\rho)).\end{split}

Here σ=σ0S0/S1\sigma=\sigma_{0}\in S^{0}/S^{-1} refers to the principal symbol. As 𝒳𝒴I\mathcal{X}\mathcal{Y}\equiv I and 𝒴𝒳I\mathcal{Y}\mathcal{X}\equiv I near (x0,t0,ξ0,ω0)(x_{0},t_{0},\xi_{0},\omega_{0}) and (y0,s0,η0,ρ0)(y_{0},s_{0},\eta_{0},\rho_{0}), respectively, there are conic neighbourhoods 𝒱\mathcal{V} and 𝒲=1𝒱\mathcal{W}=\mathcal{H}^{-1}\mathcal{V} of (x0,t0,ξ0,ω0)(x_{0},t_{0},\xi_{0},\omega_{0}) and (y0,s0,η0,ρ0)(y_{0},s_{0},\eta_{0},\rho_{0}), respectively, such that

(48) 𝒲j=1,2,31WF(fj)=𝒲j=1,2,3WF(𝒴jkfk),\displaystyle\mathcal{W}\cap\bigcup_{j=1,2,3}\mathcal{H}^{-1}WF(f_{j})=\mathcal{W}\cap\bigcup_{j=1,2,3}WF(\mathcal{Y}^{k}_{j}f_{k}),
(49) 𝒱j=1,2,3WF(fj)=𝒱j=1,2,3WF(𝒳jkfk).\displaystyle\mathcal{V}\cap\bigcup_{j=1,2,3}\mathcal{H}WF(f_{j})=\mathcal{V}\cap\bigcup_{j=1,2,3}WF(\mathcal{X}^{k}_{j}f_{k}).

for any fj𝒟(4)f_{j}\in\mathcal{D}^{\prime}(\mathbb{R}^{4}), j=1,2,3j=1,2,3. See derivation of (48) in Appendix A. The other equation is computed similarly.

Definition 3.2.

For (x,t,ξ)4×3(x,t,\xi)\in\mathbb{R}^{4}\times\mathbb{R}^{3} we define a smooth curve

Σ=Σ±:T4,rΣ(r)=(X(r),T(r),Ξ(r),Ω(r))L±4,\Sigma=\Sigma_{\pm}:\mathbb{R}\rightarrow T^{*}\mathbb{R}^{4},\quad r\mapsto\Sigma(r)=(X(r),T(r),\Xi(r),\Omega(r))\in L^{\pm}\mathbb{R}^{4},

by

(50) X(r)=γx,v(r|ξ|g),T(r)=t+r,Ξ(r)=gγ˙x,v(r|ξ|g),Ω(r)=±|ξ|g,X(r)=\gamma_{x,v}\left(\frac{r}{|\xi|_{g}}\right),\quad T(r)=t+r,\quad\Xi(r)=\mp\flat_{g}\dot{\gamma}_{x,v}\left(\frac{r}{|\xi|_{g}}\right),\quad\Omega(r)=\pm|\xi|_{g},

where v:=gξ=gjkξkjv:=\mp\sharp_{g}\xi=\mp g^{jk}\xi_{k}\partial_{j}.

(x,t)(x,t)(ξ,|ξ|g)(\xi,|\xi|_{g})(X(r),T(r))(X(r),T(r))(Ξ(r),Ω(r))(\Xi(r),\Omega(r))
Figure 5. The curve (X,T,Ξ,Ω)(X,T,\Xi,\Omega) translates the initial covector (x,t,ξ,|ξ|g)(x,t,\xi,|\xi|_{g}) along the geodesic (X,T)(X,T) which is represented by the dashed line in the drawing. The covector (Ξ,Ω)(\Xi,\Omega) is normal to the geodesic (X,T)(X,T).

The lemmas below follow from standard geometric observations. For proofs, see Appendix A.

Lemma 3.4.

The curve (50) is the unique bicharacteristic for the operator Dt|Dx|gΨcl1(4)D_{t}\mp|D_{x}|_{g}\in\Psi^{1}_{cl}(\mathbb{R}^{4}) (i.e. an integral curve of the Hamilton vector field Hp±H_{p^{\pm}} for p±:=ω|ξ|gp^{\pm}:=\omega\mp|\xi|_{g} in char(p±)=L± 4\text{char}(p^{\pm})=L^{\pm} \mathbb{R}^{4}) with (x,t,ξ,±|ξ|g)(x,t,\xi,\pm|\xi|_{g}) as the initial value at r=0r=0.

Lemma 3.5.

Let (y,s,η)4×3(y,s,\eta)\in\mathbb{R}^{4}\times\mathbb{R}^{3} and r0>0r_{0}>0 such that the segment {(y,s+r,η,0)T4:r(r0,r0)}\{(y,s+r,\eta,0)\in T^{*}\mathbb{R}^{4}:r\in(-r_{0},r_{0})\} lies in the domain of the canonical transormation \mathcal{H}. Then, the curve

r(y,s+r,η,0),r(r0,r0)r\mapsto\mathcal{H}(y,s+r,\eta,0),\quad r\in(-r_{0},r_{0})

equals rΣ(r)r\mapsto\Sigma(r), r(r0,r0)r\in(-r_{0},r_{0}) with (x,t,ξ,±|ξ|g)=(y,s,η,0)(x,t,\xi,\pm|\xi|_{g})=\mathcal{H}(y,s,\eta,0) as the initial value. That is; the transformation \mathcal{H} takes a bicharacteristic for DsΨcl1(4)D_{s}\in\Psi^{1}_{cl}(\mathbb{R}^{4}) into a bicharacteristic for Dt|Dx|gΨcl1(4)D_{t}\mp|D_{x}|_{g}\in\Psi^{1}_{cl}(\mathbb{R}^{4}).

Definition 3.3.

We define

(51) Λf,±:={(Σ(r);x,t,ξ,±|ξ|g):r[0,),(x,t,ξ)4×3},\Lambda_{f,\pm}:=\Big{\{}\big{(}\Sigma(r)\ ;\ x,t,\xi,\pm|\xi|_{g}\big{)}:r\in[0,\infty),\ (x,t,\xi)\in\mathbb{R}^{4}\times\mathbb{R}^{3}\Big{\}},

where Σ(r)=Σ±(r)\Sigma(r)=\Sigma_{\pm}(r) is the curve (50) with the initial value Σ(0)=(x,t,ξ,±|ξ|g)L±4\Sigma(0)=(x,t,\xi,\pm|\xi|_{g})\in L^{\pm}\mathbb{R}^{4}. This is the forward propagating part of the characteristic flow-out canonical relation associated with Dt|Dx|gD_{t}\mp|D_{x}|_{g}. There is also the backwards propagating part Λb,±\Lambda_{b,\pm}, defined as

Λb,±:={(Σ(r);x,t,ξ,±|ξ|g):r(,0],(x,t,ξ)4×3}.\Lambda_{b,\pm}:=\Big{\{}\big{(}\Sigma(r)\ ;\ x,t,\xi,\pm|\xi|_{g}\big{)}:r\in(-\infty,0],\ (x,t,\xi)\in\mathbb{R}^{4}\times\mathbb{R}^{3}\Big{\}}.

The complete flow-out canonical relation associated with Dt|Dx|gD_{t}\mp|D_{x}|_{g} is the union

Λ±=Λf,±Λb,±.\Lambda_{\pm}=\Lambda_{f,\pm}\cup\Lambda_{b,\pm}.
Definition 3.4.

Let π3:T43\pi_{\mathbb{R}^{3}}:T^{*}\mathbb{R}^{4}\rightarrow\mathbb{R}^{3} be the projection π3(x,t,ξ,ω):=x\pi_{\mathbb{R}^{3}}(x,t,\xi,\omega):=x. For (z,θ,β)3×𝕊2×(0,1)(z,\theta,\beta)\in\mathbb{R}^{3}\times\mathbb{S}^{2}\times(0,1) we define

K(z,θ,β):={x3:|θ|g(x)=β1,x=z+rθ,r}.\begin{split}\mathcal{B}K(z,\theta,\beta):=\{x\in\mathbb{R}^{3}:|\theta|_{g}(x)=\beta^{-1},\ x=z+r\theta,\ r\in\mathbb{R}\}.\end{split}

The set K(z,θ,β)\mathcal{B}K(z,\theta,\beta) consists of all the spatial points where a particle that moves through the point zz at a constant velocity βθ\beta\theta breaks the “light-barrier”, that is, moves exactly at the speed of waves in the medium. The following proposition is the main tool in this paper.

Proposition 3.1.

Fix z3z\in\mathbb{R}^{3}, β(0,1)\beta\in(0,1), and θ𝕊2\theta\in\mathbb{S}^{2}. Let gg be a Riemannian metric in 3\mathbb{R}^{3} with ||g1|\cdot|_{g}\geq 1 on 𝕊2\mathbb{S}^{2} and assume that

(52) g(θθ,θ)(x)0,xK(z,θ,β),\displaystyle g(\nabla_{\theta}\theta,\theta)(x)\neq 0,\quad\forall x\in\mathcal{B}K(z,\theta,\beta),

((i.e. θ|θ|g(x)0\nabla_{\theta}|\theta|_{g}(x)\neq 0, xK(z,θ,β))\forall x\in\mathcal{B}K(z,\theta,\beta)). Let

SjIclm(4;NK(z,θ,β)),j=1,2,3,S_{j}\in I^{m}_{cl}(\mathbb{R}^{4};N^{*}K(z,\theta,\beta)),\quad j=1,2,3,

be compactly supported and PM3(Ψcl2(4))P\in M_{3}(\Psi^{2}_{cl}(\mathbb{R}^{4})) be of the form

P(t2gjkjk)IFM3(Ψ(4)),P-(\partial_{t}^{2}-g^{jk}\partial_{j}\partial_{k})I-F\in M_{3}(\Psi^{-\infty}(\mathbb{R}^{4})),

for some stationary operator F=[Fjk(x,Dx)]j,k=1,2,3F=[F^{k}_{j}(x,D_{x})]_{j,k=1,2,3} of order 11. Consider distributions uj𝒟(4)u_{j}\in\mathcal{D}^{\prime}(\mathbb{R}^{4}), j=1,2,3j=1,2,3, that for large T>0T>0 obey

(53) PjkukSjC(4),\displaystyle P^{k}_{j}u_{k}-S_{j}\in C^{\infty}(\mathbb{R}^{4}),
(54) uj|tTC(3×(,T]).\displaystyle u_{j}|_{t\leq-T}\in C^{\infty}(\mathbb{R}^{3}\times(-\infty,-T]).

Then,

j=1,2,3WF(uj)(k=1,2,3WF(Sk))j=1,2,3ΛfWF(Sj),Λf:=Λf,+Λf,.\bigcup_{j=1,2,3}WF(u_{j})\setminus\bigg{(}\bigcup_{k=1,2,3}WF(S_{k})\bigg{)}\subset\bigcup_{j=1,2,3}\Lambda_{f}\circ WF(S_{j}),\quad\Lambda_{f}:=\Lambda_{f,+}\cup\Lambda_{f,-}.

Moreover, if the initial value (x,t,ξ,±|ξ|g)=Σ(0)L±4(x,t,\xi,\pm|\xi|_{g})=\Sigma(0)\in L^{\pm}\mathbb{R}^{4} is the first intersection between a bicharacteristic Σ(r)\Sigma(r) of the form (50) and j=1,2,3WF(Sj)\bigcup_{j=1,2,3}WF(S_{j}), and the positively homogeneous principal symbol (of degree m1/2m-1/2) for at least one of the components SjS_{j}, j=1,2,3j=1,2,3 does not vanish at (x,t,ξ,±|ξ|g)(x,t,\xi,\pm|\xi|_{g}) (i.e. the 3-vector of these principal symbols is nonzero), then the open segment between the first and the second intersection lies in the wave front set. That is;

Σ(r)j=1,2,3WF(uj)\Sigma(r)\in\bigcup_{j=1,2,3}WF(u_{j})

for

0<r<sup{s>0:Σ(s)k=1,2,3WF(Sk)}.0<r<\sup\Big{\{}s>0:\Sigma(s)\notin\bigcup_{k=1,2,3}WF(S_{k})\Big{\}}.

For proving the proposition we need the following lemma:

Lemma 3.6.

Assume that the Riemannian metric gg on 3\mathbb{R}^{3} with ||g1|\cdot|_{g}\geq 1 satisfies

g(θθ,θ)(x)0,xK(z,θ,β).g(\nabla_{\theta}\theta,\theta)(x)\neq 0,\quad\forall x\in\mathcal{B}K(z,\theta,\beta).

Then the bundle L±4L^{\pm}\mathbb{R}^{4} is transversal to NK(z,θ,β)N^{*}K(z,\theta,\beta) and the intersection of NK(z,θ,β)N^{*}K(z,\theta,\beta) with a curve Σ(r)\Sigma(r) of the form (50) is a discrete set.

Proof of Lemma 3.6.

Denote K:=K(z,θ,β)K:=K(z,\theta,\beta). We begin by proving the first claim. As

dim(TvNK+TvL±4)=dim(NK)+dim(L±4)dim(TvNKTvL±4)=4+7dim(TvNKTvL±4)=dim(TvT4)+3dim(TvNKTvL±4),v=w\begin{split}&\text{dim}(T_{v}N^{*}K+T_{v}L^{\pm}\mathbb{R}^{4})\\ &=\text{dim}(N^{*}K)+\text{dim}(L^{\pm}\mathbb{R}^{4})-\text{dim}(T_{v}N^{*}K\cap T_{v}L^{\pm}\mathbb{R}^{4})\\ &=4+7-\text{dim}(T_{v}N^{*}K\cap T_{v}L^{\pm}\mathbb{R}^{4})\\ &=\text{dim}(T_{v}T^{*}\mathbb{R}^{4})+3-\text{dim}(T_{v}N^{*}K\cap T_{v}L^{\pm}\mathbb{R}^{4}),\quad v=\mathcal{H}w\\ \end{split}

we need to show that dim(TvNKTvL±4)3\text{dim}(T_{v}N^{*}K\cap T_{v}L^{\pm}\mathbb{R}^{4})\leq 3. Recall that

NK={(z+tβθ,t;ξ,βξθ):t,ξ3{0}}.N^{*}K=\{(z+t\beta\theta,t\ ;\ \xi,-\beta\xi\cdot\theta):t\in\mathbb{R},\ \xi\in\mathbb{R}^{3}\setminus\{0\}\}.

Thus, for vNKv\in N^{*}K we have the expression

TvNK={((δt)βθ,δt,δξ,β(δξ)θ):δt,δξ3}.T_{v}N^{*}K=\{\big{(}(\delta t)\beta\theta,\delta t,\delta\xi,-\beta(\delta\xi)\cdot\theta\big{)}:\delta t\in\mathbb{R},\ \delta\xi\in\mathbb{R}^{3}\}.

As L±4={(x,t,ξ,±|ξ|g):(x,t,ξ)4×(3{0})}L^{\pm}\mathbb{R}^{4}=\{(x,t,\xi,\pm|\xi|_{g}):(x,t,\xi)\in\mathbb{R}^{4}\times(\mathbb{R}^{3}\setminus\{0\})\} we have for v=(x,t,ξ,±|ξ|g)L±4v=(x,t,\xi,\pm|\xi|_{g})\in L^{\pm}\mathbb{R}^{4} that

TvL±4={(δx,δt,δξ,δω):δω=±g(ξ,δξ)+12dxg(ξ,ξ),δx|ξ|g(x)}.T_{v}L^{\pm}\mathbb{R}^{4}=\Big{\{}(\delta x,\delta t,\delta\xi,\delta\omega):\delta\omega=\pm\frac{g(\xi,\delta\xi)+\frac{1}{2}\langle d_{x}g(\xi,\xi),\delta x\rangle}{|\xi|_{g}(x)}\Big{\}}.

Consequently, for v=(x(t),t,ξ,ω(t,ξ))NKL±4v=(x(t),t,\xi,\omega(t,\xi))\in N^{*}K\cap L^{\pm}\mathbb{R}^{4} (x(t)=z+tβθx(t)=z+t\beta\theta, ω(t,ξ)=βξθ=±|ξ|g(z+tβθ)\omega(t,\xi)=-\beta\xi\cdot\theta=\pm|\xi|_{g}(z+t\beta\theta)) we have

dim(TvNKTvL+4)=dim(kerL).\text{dim}(T_{v}N^{*}K\cap T_{v}L^{+}\mathbb{R}^{4})=\text{dim}(\ker L).

where L=Lz,ξ,t,β,θ:4L=L_{z,\xi,t,\beta,\theta}:\mathbb{R}^{4}\mapsto\mathbb{R} is a linear map defined by

L:(δξ,δt)±g(ξ,δξ)+β2dxg(ξ,ξ),θδt|ξ|g|x=z+tβθ+β(δξ)θ.L:(\delta\xi,\delta t)\mapsto\pm\frac{g(\xi,\delta\xi)+\frac{\beta}{2}\langle d_{x}g(\xi,\xi),\theta\rangle\delta t}{|\xi|_{g}}\Big{|}_{x=z+t\beta\theta}+\beta(\delta\xi)\cdot\theta.

To prove that the dimension of the kernel is less or equal to 33 we need to show that the map LL is not identically zero. By setting L|3×{0}=0L|_{\mathbb{R}^{3}\times\{0\}}=0 one obtains

(55) ±ξ|ξ|g=βgθatx=z+tβθ\pm\frac{\xi}{|\xi|_{g}}=-\beta\flat_{g}\theta\quad\text{at}\quad x=z+t\beta\theta

which is possible only if |θ|g=1β|\theta|_{g}=\frac{1}{\beta} at x=z+tβθx=z+t\beta\theta. Let us study LL at such points. By substituting (55) in L(0,β3)L(0,\beta^{-3}) implies

L(0,β3)=±12dxg(θ,θ),θ|x=z+tβθ=±g(θθ,θ)|x=z+tβθL(0,\beta^{-3})=\pm\frac{1}{2}\langle d_{x}g(\theta,\theta),\theta\rangle|_{x=z+t\beta\theta}=\pm g(\nabla_{\theta}\theta,\theta)|_{x=z+t\beta\theta}

which is nonzero by assumptions. In conclusion, the dimension of kerL\text{ker}L is at most 33 which finishes the first part of the proof.

Let us now show that the intersection of NKN^{*}K and the curve Σ(r)=(X(r),T(r),Ξ(r),Ω(r))\Sigma(r)=(X(r),T(r),\Xi(r),\Omega(r)) of the form (50) is a collection of discrete points. To prove this we set Σ(0)NK\Sigma(0)\in N^{*}K and show that (X(r),T(r))(X(r),T(r)) is not in KK for r0r\neq 0 in a small neighbourhood of 0. By definition, X(r)=γv(r/|ξ|g)X(r)=\gamma_{v}(r/|\xi|_{g}), v=gξv=\mp\sharp_{g}\xi. If (X(r),T(r))(X(r),T(r)) is not tangent to KK at r=0r=0, then the claim clearly holds. Thus, we may assume that (X˙(0),T˙(0))TK(\dot{X}(0),\dot{T}(0))\in TK which implies the following approximation near r=0r=0:

X(r)=X(0)+βθrh(r),X(r)=X(0)+\beta\theta r-h(r),

where h(r)=O(r2)h(r)=O(r^{2}). In addition, (X˙(0),T˙(0))L±4(\dot{X}(0),\dot{T}(0))\in L^{\pm}\mathbb{R}^{4} so at X(0)X(0) we must have 1=|T˙(0)|=|X˙(0)|g=β|θ|g,1=|\dot{T}(0)|=|\dot{X}(0)|_{g}=\beta|\theta|_{g}, that is, |θ|g=1β|\theta|_{g}=\frac{1}{\beta}. In particular, g(θθ,θ)|x=X(0)0g(\nabla_{\theta}\theta,\theta)|_{x=X(0)}\neq 0 by assumptions. As rX(r)r\mapsto X(r) is a geodesic, it satisfies rX˙=0\nabla_{r}\dot{X}=0, where rV=V˙ll+ΓjklVjVkl\nabla_{r}V=\dot{V}^{l}\partial_{l}+\Gamma^{l}_{jk}V^{j}V^{k}\partial_{l} is the covariant derivative of a vector field along a curve (see e.g. [Lee97]). On the other hand,

g(rX˙,X˙)|r=0=β2g(θθ,θ)|x=X(0)βg(rh˙,θ)|r=0\begin{split}g(\nabla_{r}\dot{X},\dot{X})|_{r=0}=\beta^{2}g(\nabla_{\theta}\theta,\theta)|_{x=X(0)}-\beta g(\nabla_{r}\dot{h},\theta)|_{r=0}\\ \end{split}

Thus, g(rh˙,θ)|r=0=βg(θθ,θ)|x=X(0)0g(\nabla_{r}\dot{h},\theta)|_{r=0}=\beta g(\nabla_{\theta}\theta,\theta)|_{x=X(0)}\neq 0 which is possible only if h(r)=μr2+O(r3)h(r)=\mu r^{2}+O(r^{3}) around r=0r=0 for some nonzero vector μ\mu (i.e. the curve X(r)X(r) “bends” at r=0r=0). In conclusion, for r0r\neq 0 near the origin we have that (X(r),T(r))(X(r),T(r)) does not lie in the line KK. ∎

Proof of Proposition 3.1.

We relax the notation by omitting the parameters (z,θ,β)(z,\theta,\beta) whenever there is no danger of confusion. The proof follows the standard scheme of reducing the original vector-valued system microlocally along the characteristic flow into independent scalar transport equations. Solving such equations is a simple task and requires only extending the fundamental theorem of calculus.

For a fixed sign ±{+,}\pm\in\{+,-\} and a vector (x0,t0,ξ0,ω0)T4(x_{0},t_{0},\xi_{0},\omega_{0})\in T^{*}\mathbb{R}^{4}, ω0|ξ0|g\omega_{0}\neq\mp|\xi_{0}|_{g} we express the wave u=ujdxju=u_{j}dx^{j} microlocally in a suitable conic neighbourhood 𝒱=𝒱x0,t0,ξ0,ω0\mathcal{V}=\mathcal{V}_{x_{0},t_{0},\xi_{0},\omega_{0}} of (x0,t0,ξ0,ω0)(x_{0},t_{0},\xi_{0},\omega_{0}) as a sum

(56) uu±+u0modC,u\equiv u_{\pm}+u_{0}\mod C^{\infty},

of a term u±=u±,𝒱u_{\pm}=u_{\pm,\mathcal{V}} which creates (or annihilates) singularities that propagate forwards in time along the characteristic flowout from the source, and a residual term u0=u0,𝒱u_{0}=u_{0,\mathcal{V}} with a flow-invariant wave front set. That is;

u±,jIclr(4;Λf,±NK),j=1,2,3,microlocally away from NK,u_{\pm,j}\in I_{cl}^{r}(\mathbb{R}^{4};\Lambda_{f,\pm}\circ N^{*}K),\quad j=1,2,3,\quad\text{microlocally away from }N^{*}K,

which can be chosen such that u±0u_{\pm}\equiv 0 if (x0,t0,ξ0,ω0)j=1,2,3WF(Sj)(x_{0},t_{0},\xi_{0},\omega_{0})\notin\bigcup_{j=1,2,3}WF(S_{j}), and

Λ±WF𝒱(u0)=WF𝒱(u0)\Lambda_{\pm}\circ WF_{\mathcal{V}}(u_{0})=WF_{\mathcal{V}}(u_{0})

where

WF𝒱(f):=j=1,2,3WF(fj)𝒱,WF_{\mathcal{V}}(f):=\bigcup_{j=1,2,3}WF(f_{j})\cap\mathcal{V},

all in a suitable conic neighbourhood 𝒱\mathcal{V} of (x0,t0,ξ0,ω0)(x_{0},t_{0},\xi_{0},\omega_{0}). The residual term just moves existing singularities along bicharacteristics (50) in the sense that the wave front set j=1,2,3WF(u0,j)\bigcup_{j=1,2,3}WF(u_{0,j}) in one region 𝒱𝒱\mathcal{V}^{\prime}\subset\mathcal{V} defines it in the flowout Λ±𝒱\Lambda_{\pm}\circ\mathcal{V}^{\prime} within 𝒱\mathcal{V}. The wave front set of uu away from L4L^{\mp}\mathbb{R}^{4} is computed by solving the microlocal expressions u=u±+u0u=u_{\pm}+u_{0} along curves of the form (50). Let us briefly explain this before constructing the terms. First of all, early points in curves (50) have not yet hit the set 3×[T,)\mathbb{R}^{3}\times[-T,\infty) so regularity of the wave uu outside 3×(T,)\mathbb{R}^{3}\times(-T,\infty) implies that no singularities along the curve occur in uu until intersection with j=1,2,3WF(Sj)\bigcup_{j=1,2,3}WF(S_{j}) the first time in 3×(T,)\mathbb{R}^{3}\times(-T,\infty). Notice that the condition |θ|g1|\theta|_{g}\geq 1 on 𝕊2\mathbb{S}^{2} ensures that each bicharacteristic really goes through the Cauchy surface t=Tt=-T. We must therefore have uu±u\equiv u_{\pm} at the first intersection of the curve and j=1,2,3WF(Sj)\bigcup_{j=1,2,3}WF(S_{j}). Hence, a forward propagating singularity is created at the point if for some jj we can show that the positively homogeneous principal symbol of u±,ju_{\pm,j} does not vanish there. On the other hand, in a suitable conic neighbourhood around any point outside j=1,2,3WF(Sj)\bigcup_{j=1,2,3}WF(S_{j}) the expression u=u±+u0u=u_{\pm}+u_{0} leads to smooth u±u_{\pm} and hence flow-invariant j=1,2,3WF(uj)\bigcup_{j=1,2,3}WF(u_{j}). Thus, in a convenient sequence of conical neighbourhoods that cover a given bicharacteristic segment (50) emanating from the source each created singularity gets transported arbitrarily far until it possibly gets annihilated at the source.

Let us now derive the microlocal expression above. By Lemma 3.3 it suffices to study uu near (x0,t0,ξ0,ω0)(x_{0},t_{0},\xi_{0},\omega_{0}), ω0=±|ξ0|g\omega_{0}=\pm|\xi_{0}|_{g}. As before, we let \mathcal{H} be the homogeneous local transformation (40) on a conic neighbourhood of the point with ω|ξ|g\omega\mp|\xi|_{g} being the last coordinate. Applying Lemma 3.1 and Lemma 3.2 we derive

𝒴S=𝒴Pu𝒴(DtIW)Zu𝒴(DtIW)𝒳𝒴ZuB(DtIW)𝒜A𝒴ZuB(DtI+q)A𝒴ZuBADs𝒴ZuDs𝒴ZumodC\begin{split}\mathcal{Y}S=\mathcal{Y}Pu\equiv\mathcal{Y}(D_{t}I-W)Zu\equiv\mathcal{Y}(D_{t}I-W)\mathcal{X}\mathcal{Y}Zu\equiv B\mathcal{B}(D_{t}I-W)\mathcal{A}A\mathcal{Y}Zu\\ \equiv B(D_{t}I+q)A\mathcal{Y}Zu\equiv BAD_{s}\mathcal{Y}Zu\equiv D_{s}\mathcal{Y}Zu\mod C^{\infty}\end{split}

microlocally near (y0,s0,η0,ρ0)(y_{0},s_{0},\eta_{0},\rho_{0}). Thus, there are δ,ϵ>0\delta,\epsilon>0 such that in the open cylinder Wϵ,δ:=B(y0,δ)×(s0ϵ,s0+ϵ){W_{\epsilon,\delta}}:=B(y_{0},\delta)\times(s_{0}-\epsilon,s_{0}+\epsilon) we have modulo C(Wϵ,δ)C^{\infty}({W_{\epsilon,\delta}}) the local transport equations

(57) Dsu~j(y,s)S~j(y,s),j=1,2,3,(y,s)Wϵ,δ,D_{s}\tilde{u}_{j}(y,s)\equiv\tilde{S}_{j}(y,s),\quad j=1,2,3,\quad(y,s)\in{W_{\epsilon,\delta}},

for

(58) u~j:=\displaystyle\tilde{u}_{j}:= ϱ𝒴kjZlkul(4),\displaystyle\varrho\mathcal{Y}^{k}_{j}Z^{l}_{k}u_{l}\in\mathcal{E}^{\prime}(\mathbb{R}^{4}),
(59) S~j:=\displaystyle\tilde{S}_{j}:= ϱ𝒴kjSkImcl(1NK;4),\displaystyle\varrho\mathcal{Y}^{k}_{j}S_{k}\in I^{m}_{cl}(\mathcal{H}^{-1}N^{*}K;\mathbb{R}^{4}),
(60) ϱ:=\displaystyle\varrho:= χ(|(Dy,Ds)|(Dy,Ds)|(η0,ρ0)|(η0,ρ0)||)Ψ0cl(4),\displaystyle\chi\left(\bigg{|}\frac{(D_{y},D_{s})}{|(D_{y},D_{s})|}-\frac{(\eta_{0},\rho_{0})}{|(\eta_{0},\rho_{0})|}\bigg{|}\right)\in\Psi^{0}_{cl}(\mathbb{R}^{4}),

where χCc()\chi\in C_{c}^{\infty}(\mathbb{R}) is a smooth cut-off function that vanishes outside (δ,δ)(-\delta,\delta) and equals 1 in (δ/2,δ/2)(-\delta/2,\delta/2). Microlocally near (y0,s0,η0,ρ0)(y_{0},s_{0},\eta_{0},\rho_{0}) we have that u~𝒴Zu\tilde{u}\equiv\mathcal{Y}Zu and S~=𝒴S\tilde{S}=\mathcal{Y}S. In view of (48-49), Corollary 3.1 (applied to ZZ) and Lemma 3.5 it suffices to study singularities of each u~j\tilde{u}_{j} along bicharacteristics of DsD_{s}. The original wave uu is derived microlocally near (x0,t0,ξ0,ω0)(x_{0},t_{0},\xi_{0},\omega_{0}) by applying the inverse of 𝒴Z\mathcal{Y}Z to u~\tilde{u}. Formally,

(61) uR𝒳u~R𝒳EfS~+R𝒳vmodCu\equiv R\mathcal{X}\tilde{u}\equiv R\mathcal{X}E_{f}\tilde{S}+R\mathcal{X}v\mod C^{\infty}

where RR is the microlocal inverse of ZZ near (x0,t0,ξ0,ω0)(x_{0},t_{0},\xi_{0},\omega_{0}), EfE_{f} stands for the forward fundamental solution Ef(s,y;s~,y~)=iH(ss~)δ0(yy~)E_{f}(s,y;\tilde{s},\tilde{y})=iH(s-\tilde{s})\delta_{0}(y-\tilde{y}) (operating separately for each coordinate: (Efu)j:=Efuj(E_{f}u)_{j}:=E_{f}u_{j}), and vker(Ds)v\in\text{ker}(D_{s}) is a residual term with wave-front invariant in the characteristic flow of DsD_{s}:

(62) ΣjTB(y0,δ):WF(vj)={(y,s,η,0):(y,η)Σj},j=1,2,3,\exists\Sigma_{j}\subset T^{*}B(y_{0},\delta)\ :\ WF(v_{j})=\{(y,s,\eta,0):(y,\eta)\in\Sigma_{j}\},\quad j=1,2,3,

(see [DH72, Theorem 6.1.1]).333Choosing a different fundamental solution, say the backwards propagating solution Eb(s,y;s~,y~)=iH(s~s)δ0(yy~)E_{b}(s,y;\tilde{s},\tilde{y})=-iH(\tilde{s}-s)\delta_{0}(y-\tilde{y}) leads to a similar expression. In general, the difference E1S~E2S~E_{1}\tilde{S}-E_{2}\tilde{S} for two solution operators E1,E2E_{1},E_{2} solves Dsu=0D_{s}u=0, and therefore is invariant along the bicharacteristics. The creation term is u±:=R𝒳EfS~u_{\pm}:=R\mathcal{X}E_{f}\tilde{S}, whereas the residual term is given by u0:=R𝒳vu_{0}:=R\mathcal{X}v. For proving that EfS~E_{f}\tilde{S}, and hence the expression (61) is well defined and compatible with [GU93, Proposition 2.1] it suffices to show that the characteristic set {(y,s,η,0):(y,s)Wϵ,δ}\{(y,s,\eta,0):(y,s)\in{W_{\epsilon,\delta}}\} of Ds:𝒟(Wϵ,δ)𝒟(Wϵ,δ)D_{s}:\mathcal{D}^{\prime}({W_{\epsilon,\delta}})\rightarrow\mathcal{D}^{\prime}({W_{\epsilon,\delta}}) is transversal to 1NK\mathcal{H}^{-1}N^{*}K and each bicharacteristic {(y,s,η,0):s(s0ϵ,s0+ϵ)}\{(y,s,\eta,0):s\in(s_{0}-\epsilon,s_{0}+\epsilon)\}, yB(y0,δ)y\in B(y_{0},\delta) intersects 1NK\mathcal{H}^{-1}N^{*}K finite number of times. These properties follow for small ϵ,δ>0\epsilon,\delta>0 by combining Lemma 3.5 with Lemma 3.6 and hence the referred proposition applies and yields the expression for u±u_{\pm} as a vector of Lagrangian distributions associated with the pair (NK,Λf,±NK)(N^{*}K,\Lambda_{f,\pm}\circ N^{*}K). In particular, microlocally away from NKN^{*}K the term u±u_{\pm} is a Lagrangian distribution over Λf,±NK\Lambda_{f,\pm}\circ N^{*}K.

Finally we show that a propagating singularity in u±u_{\pm} is created at a point where the principal symbol of SS is non-vanishing. As u~EfS~Ef𝒴S\tilde{u}\equiv E_{f}\tilde{S}\equiv E_{f}\mathcal{Y}S microlocally near (x0,t0,ξ0,ω0)(x_{0},t_{0},\xi_{0},\omega_{0}), the standard principal symbol formula for compositions of FIOs (for EfE_{f} the formula is provided in the proposition referred above) implies that for (x0,t0,ξ0,ω0)L±4NK(x_{0},t_{0},\xi_{0},\omega_{0})\in L^{\pm}\mathbb{R}^{4}\cap N^{*}K, sufficiently small s>0s>0 and the positively homogeneous principal symbols p~±,j\tilde{p}_{\pm,j}, q~j\tilde{q}_{j}, qjq_{j} of u~±,j\tilde{u}_{\pm,j}, S~j\tilde{S}_{j}, SjS_{j} we have

p~±,j(y0,s0+s,η0,ρ0)q~j(y0,s0,η0,ρ0)=Mkjqk(x0,t0,ξ0,ω0),\tilde{p}_{\pm,j}(y_{0},s_{0}+s,\eta_{0},\rho_{0})\simeq\tilde{q}_{j}(y_{0},s_{0},\eta_{0},\rho_{0})=M^{k}_{j}q_{k}(x_{0},t_{0},\xi_{0},\omega_{0}),

where MM is an invertible matrix given as the principal symbol of 𝒴\mathcal{Y} at (y0,s0,η0,ρ0;x0,t0,ξ0,ω0)(y_{0},s_{0},\eta_{0},\rho_{0};x_{0},t_{0},\xi_{0},\omega_{0}). This implies that u~±\tilde{u}_{\pm} creates at (y0,s0,η0,ρ0)(y_{0},s_{0},\eta_{0},\rho_{0}) a forwards propagating singularity provided that the principal symbol qk(x0,t0,ξ0,ω0)q_{k}(x_{0},t_{0},\xi_{0},\omega_{0}) of SkS_{k} does not vanish for at least one k=1,2,3k=1,2,3. Applying this to (x0,t0,ξ0,ω0)=(x,t,ξ,|ξ|g)(x_{0},t_{0},\xi_{0},\omega_{0})=(x,t,\xi,|\xi|_{g}), where (x,t,ξ,|ξ|g)(x,t,\xi,|\xi|_{g}) is as in the assumptions, and using R𝒳R\mathcal{X} to transform u~±\tilde{u}_{\pm} microlocally back into u±u_{\pm} yields the creation of singularity for u±u_{\pm} at (x,t,ξ,|ξ|g)(x,t,\xi,|\xi|_{g}), hence finishing the proof.

4. Proof of Theorem 1.1

In this section we derive the claim of Theorem 1.1. The proof is constructed in three stages. The first and perhaps the most difficult step is to prove Lemma 4.1 below. The second step is to derive Proposition 4.1 from the lemma. The claim of Theorem 1.1 is deduced in the final step after that.

Lemma 4.1.

Assume that the conditions of Theorem 1.1 hold. Fix zUz\in U and let x0Υx_{0}\in\Upsilon be a 444There can be many of such points. nearest point to zz on the boundary with respect to g1g_{1}. Then there is a neighbourhood 𝒬Υ\mathcal{Q}\subset\Upsilon of x0x_{0} in Υ\Upsilon such that

distg2(z,x)distg1(z,x),x𝒬.\text{dist}_{g_{2}}(z,x)\leq\text{dist}_{g_{1}}(z,x),\quad\forall x\in\mathcal{Q}.

Moreover, if γ1:[0,l]3\gamma_{1}:[0,l]\rightarrow\mathbb{R}^{3} with l:=distg1(z,x)l:=\text{dist}_{g_{1}}(z,x) is a geodesic segment in (3,g1)(\mathbb{R}^{3},g_{1}) from γ1(0)=z\gamma_{1}(0)=z to γ1(l)=x\gamma_{1}(l)=x, then the unique geodesic γ2\gamma_{2} in (3,g2)(\mathbb{R}^{3},g_{2}) with (γ2(l),γ˙2(l))=(γ1(l),γ˙1(l))(\gamma_{2}(l),\dot{\gamma}_{2}(l))=(\gamma_{1}(l),\dot{\gamma}_{1}(l)) satisfies γ2(0)=z\gamma_{2}(0)=z.

Proof.

Fix zz and x0x_{0} as in the assumptions and a small neighbourhood 𝒬Υ\mathcal{Q}\subset\Upsilon of x0x_{0} in Υ\Upsilon. By making the neighbourhood small enough we may assume that every shortest geodesic segment in (3,g1)(\mathbb{R}^{3},g_{1}) from zz to a point in 𝒬\mathcal{Q} lies in WW and hits the boundary W\partial W transversally and for the first time at the endpoint. Fix arbitrary x𝒬x\in\mathcal{Q}. Let γ1:[0,l]3\gamma_{1}:[0,l]\rightarrow\mathbb{R}^{3} be an unit speed geodesic in (3,g1)(\mathbb{R}^{3},g_{1}) from zz to xx such that l=distg1(z,x)l=\text{dist}_{g_{1}}(z,x). Denote v:=γ˙1(0)v:=\dot{\gamma}_{1}(0). Let γ2\gamma_{2} be the unique geodesic in (3,g2)(\mathbb{R}^{3},g_{2}) such that (γ2(l),γ˙2(l))=(γ1(l),γ˙1(l))(\gamma_{2}(l),\dot{\gamma}_{2}(l))=(\gamma_{1}(l),\dot{\gamma}_{1}(l)). Since g1=g2g_{1}=g_{2} on Υ\Upsilon, also the geodesic γ2\gamma_{2} has unit speed. Define

M:={(θ,β)𝕊2×z:(z,0;g1v,1)NK(z,θ,β)L+(3;g1)},M:=\{(\theta,\beta)\in\mathbb{S}^{2}\times\mathcal{I}_{z}:(z,0;-\flat_{g_{1}}v,1)\in N^{*}K(z,\theta,\beta)\cap L^{+}(\mathbb{R}^{3};g_{1})\},

where z\mathcal{I}_{z} is as in the assumptions. The space MM is a trivial bundle over z\mathcal{I}_{z}, consisting of distinct loops Mβ=M(𝕊2×{β})M_{\beta}=M\cap(\mathbb{S}^{2}\times\{\beta\}) on the surface 𝕊2\mathbb{S}^{2} as fibers. In some coordinates the loops are just circles. For each θMβ\theta\in M_{\beta} there is the antipodal point θ~=θ~(θ)Mβ\tilde{\theta}=\tilde{\theta}(\theta)\in M_{\beta} in the circle MβM_{\beta} which satisfies

(63) NK(z,θ,β)NK(z,θ~,β)L+(3;g1)={(z,0;hg1v,h):h>0}.N^{*}K(z,\theta,\beta)\cap N^{*}K(z,\tilde{\theta},\beta)\cap L^{+}(\mathbb{R}^{3};g_{1})=\{(z,0;-h\flat_{g_{1}}v,h):h>0\}.

The construction is illustrated in Figure 6.

θ\thetaθ~\tilde{\theta}g1v{\color[rgb]{0,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,1}-\flat_{g_{1}}v}
Figure 6. A projection of the construction in (63) to the space 3\mathbb{R}^{3}. The loop MβM_{\beta} is illustrated in red. The intersections NK(z,θ,β)L+(3;g1)N^{*}K(z,\theta,\beta)\cap L^{+}(\mathbb{R}^{3};g_{1}) and NK(z,θ~,β)L+(3;g1)N^{*}K(z,\tilde{\theta},\beta)\cap L^{+}(\mathbb{R}^{3};g_{1}) appear as two cones (the dashed lines) that touch each other at the half-open segment {(z,0;hg1v,h):h>0}\{(z,0;-h\flat_{g_{1}}v,h):h>0\} (in blue).

Let γ3\gamma_{3} be the unique unit-speed geodesic in ( 3,g2)( \mathbb{R}^{3},g_{2}) through xx that is tangent to the mirror image of γ˙2(l)\dot{\gamma}_{2}(l) over the boundary TxWT_{x}\partial W at that point. Using the fact that MM is two dimensional as a manifold we fix (θ,β)M(\theta,\beta)\in M such that the lines K(z,θ,β)K(z,\theta,\beta) and K(z,θ~,β)K(z,\tilde{\theta},\beta) avoid the time-like geodesics

(64) {(γτ(r),r):r,τ=1,2,3}\{(\gamma_{\tau}(r),r):r\in\mathbb{R},\ \tau=1,2,3\}

inside (B×){(z,0)}(B\times\mathbb{R})\setminus\{(z,0)\} for a large closed Euclidean ball B3B\subset\mathbb{R}^{3} with radius R>supβzRβR>\sup_{\beta\in\mathcal{I}_{z}}R_{\beta}. By assumptions, B×B\times\mathbb{R} contains the singular supports of the sources. We would like to apply Proposition 3.1 to systems associated with the parameters (θ,β)(\theta,\beta) and (θ~,β)(\tilde{\theta},\beta). However, it is possible that the rays K(z,θ,β)K(z,\theta,\beta) and K(z,θ~,β)K(z,\tilde{\theta},\beta) do not break the velocity barriers K(z,θ,β)\mathcal{B}K(z,\theta,\beta) and K(z,θ~,β)\mathcal{B}K(z,\tilde{\theta},\beta) in the required way. To get around this we perturb the velocity β\beta slightly and then apply Sard’s theorem. If one makes the parameter β\beta slightly larger while keeping everything else fixed, the rays still avoid the geodesics (64) in (B×){(z,0)}(B\times\mathbb{R})\setminus\{(z,0)\} but the intersection (63) splits into two lines instead of one (See Figure 7).

Figure 7. Increasing β\beta makes the two touching cones in Figure 6 wider. Intersection of the wider cones consists of two lines, illustrated here by the blue and red arrows.

The splitting, however, happens continuously with respect to β\beta. Thus, for every conic neighbourhood 𝒱zTz3\mathcal{V}_{z}\subset T_{z}^{*}\mathbb{R}^{3} of g1v-\flat_{g_{1}}v in Tz3T_{z}^{*}\mathbb{R}^{3} there is δ>0\delta>0 such that

NK(z,θ,β+β)NK(z,θ~,β+β)L+(3;g1){(z,0,ξ,|ξ|g):ξ𝒱z},β[0,δ).N^{*}K(z,\theta,\beta+\beta^{\prime})\cap N^{*}K(z,\tilde{\theta},\beta+\beta^{\prime})\cap L^{+}(\mathbb{R}^{3};g_{1})\subset\{(z,0,\xi,|\xi|_{g}):\xi\in\mathcal{V}_{z}\},\quad\forall\beta^{\prime}\in[0,\delta).

That is; for fixed (β,θ)M(\beta,\theta)\in M, and associated θ~\tilde{\theta}, we may add a small positive perturbation to β\beta without changing the intersection (63)\eqref{sif} significantly. By applying Sard’s theorem to the functions r|θ|gλ(z+rθ)r\mapsto|\theta|_{g_{\lambda}}(z+r\theta) and r|θ~|gλ(z+rθ~)r\mapsto|\tilde{\theta}|_{g_{\lambda}}(z+r\tilde{\theta}) one constructs a sequence βj0\beta_{j}^{\prime}\longrightarrow 0 such that for every j=1,2,3,j=1,2,3,\dots and λ=1,2\lambda=1,2 both of the following conditions hold:

θ|θ|gλ(x)0,xK(z,θ,β+βj),θ~|θ~|gλ(x)0,xK(z,θ~,β+βj).\begin{split}\nabla_{\theta}|\theta|_{g_{\lambda}}(x)\neq 0,\quad\forall x\in\mathcal{B}K(z,\theta,\beta+\beta_{j}^{\prime}),\\ \nabla_{\tilde{\theta}}|\tilde{\theta}|_{g_{\lambda}}(x)\neq 0,\quad\forall x\in\mathcal{B}K(z,\tilde{\theta},\beta+\beta_{j}^{\prime}).\end{split}

This is equivalent to saying that gλ(θθ,θ)|x0g_{\lambda}(\nabla_{\theta}\theta,\theta)|_{x}\neq 0 and gλ(θ~θ~,θ~)|x0g_{\lambda}(\nabla_{\tilde{\theta}}\tilde{\theta},\tilde{\theta})|_{x}\neq 0 in the associated points and hence the phase velocity barrier is broken in the required way for the approximating rays. We may assume that β+βj\beta+\beta_{j}^{\prime} lies in z\mathcal{I}_{z} for all j=1,2,3,j=1,2,3,\dots by excluding a finite number of terms in the beginning of the sequence.

Let us fix vjTz3v_{j}\in T_{z}\mathbb{R}^{3}, j=1,2,3,j=1,2,3,\dots, such that

(z,g1vj,1)NK(z,θ,β+βj)NK(z,θ~,β+βj)L+(3;g1).(z,-\flat_{g_{1}}v_{j},1)\in N^{*}K(z,\theta,\beta+\beta_{j}^{\prime})\cap N^{*}K(z,\tilde{\theta},\beta+\beta_{j}^{\prime})\cap L^{+}(\mathbb{R}^{3};g_{1}).

Then, vjvv_{j}\longrightarrow v and the associated unit speed geodesics γ1,j:=γz,vj\gamma_{1,j}:=\gamma_{z,v_{j}} in (3,g1)(\mathbb{R}^{3},g_{1}) approximate the segment γ1\gamma_{1}. The segment γ1\gamma_{1} is then close to a shortest curve from zz to the boundary and therefore small perturbations of it hit the boundary transversally near the point γ1(l)=x\gamma_{1}(l)=x. Consequently, we may assume that there is a sequence of positive real numbers ljll_{j}\geq l such that ljll_{j}\longrightarrow l and γ1,j(lj)Υ\gamma_{1,j}(l_{j})\in\Upsilon. Let us denote by γ2,j\gamma_{2,j} the geodesics in (3,g2)(\mathbb{R}^{3},g_{2}) with the property (γ2,j(lj),γ˙2,j(lj))=(γ1,j(lj),γ˙1,j(lj))(\gamma_{2,j}(l_{j}),\dot{\gamma}_{2,j}(l_{j}))=(\gamma_{1,j}(l_{j}),\dot{\gamma}_{1,j}(l_{j})). These geodesics have unit speed and they approximate γ2\gamma_{2}.

We now apply Proposition 3.1 for g=g1g=g_{1} using (z,θ,β+βj)(z,\theta,\beta+\beta^{\prime}_{j}) and (z,θ~,β+βj)(z,\tilde{\theta},\beta+\beta^{\prime}_{j}), j=1,2,3,j=1,2,3,\dots as fixed parameters. The point is that there must be a singularity in uλ|λ=1u_{\lambda}|_{\lambda=1} that propagates along a bicharacteristic Σj(r)\Sigma_{j}(r) of the form (50) emanating from (z,0,ϵg1vj,ϵ)j=1,2,3WF(Sj)(z,0,-\epsilon\flat_{g_{1}}v_{j},\epsilon)\in\bigcup_{j=1,2,3}WF(S_{j}), where ϵ{1,1}\epsilon\in\{-1,1\}. Moreover, since β+βj>infz>βW(z)\beta+\beta_{j}^{\prime}>\inf\mathcal{I}_{z}>\beta_{\partial W}(z) by assumptions, the propagating singularity for large jj can not hit the source and get annihilated before hitting Υ\Upsilon as otherwise it would contradict the definition of βW(z)\beta_{\partial W}(z). Let us denote WF(u):=k=1,2,3WF(uk)WF(u):=\bigcup_{k=1,2,3}WF(u_{k}). By putting together the proposition and the construction above, we conclude

 Σj(lj)=(γ1,j(lj),lj,ϵg1γ˙1,j(lj),ϵ)WF(uλ(;z,θ,β+βj))|λ=1WF(uλ(;z,θ~,β+βj))|λ=1 \Sigma_{j}(l_{j})=(\gamma_{1,j}(l_{j}),l_{j},-\epsilon\flat_{g_{1}}\dot{\gamma}_{1,j}(l_{j}),\epsilon)\in WF(u_{\lambda}(\ \cdot\ ;z,\theta,\beta+\beta_{j}^{\prime}))\big{|}_{\lambda=1}\cap WF(u_{\lambda}(\ \cdot\ ;z,\tilde{\theta},\beta+\beta_{j}^{\prime}))\big{|}_{\lambda=1}

where ϵ{1,1}\epsilon\in\{-1,1\} and uλu_{\lambda} is as in the assumptions. Denote xj:=γ1,j(lj)=γ2,j(lj)x_{j}:=\gamma_{1,j}(l_{j})=\gamma_{2,j}(l_{j}). The covector (xj,lj,ϵg1γ˙1,j(lj),ϵ)=(xj,lj,ϵg2γ˙2,j(lj),ϵ)(x_{j},l_{j},-\epsilon\flat_{g_{1}}\dot{\gamma}_{1,j}(l_{j}),\epsilon)=(x_{j},l_{j},-\epsilon\flat_{g_{2}}\dot{\gamma}_{2,j}(l_{j}),\epsilon), has a non-zero correspondent in

WF(uλ=1(;z,θ,β+βj))|T(Υ×)WF(uλ=1(;z,θ~,β+βj))|T(Υ×).WF(u_{\lambda=1}(\ \cdot\ ;z,\theta,\beta+\beta_{j}^{\prime}))\big{|}_{T(\Upsilon\times\mathbb{R})}\cap WF(u_{\lambda=1}(\ \cdot\ ;z,\tilde{\theta},\beta+\beta_{j}^{\prime}))\big{|}_{T(\Upsilon\times\mathbb{R})}.

As the data for both metrics coincide, the set above equals

(65) WF(uλ=2(;z,θ,β+βj))|T(Υ×)WF(uλ=2(;z,θ~,β+βj))|T(Υ×)\begin{split}WF(u_{\lambda=2}(\ \cdot\ ;z,\theta,\beta+\beta_{j}^{\prime}))\big{|}_{T(\Upsilon\times\mathbb{R})}\cap WF(u_{\lambda=2}(\ \cdot\ ;z,\tilde{\theta},\beta+\beta_{j}^{\prime}))\big{|}_{T(\Upsilon\times\mathbb{R})}\end{split}

Applying Proposition 3.1, this time with respect to the metric g2g_{2}, yields that

WF(uλ=2(;z,θ,β+βj))NK(z,θ,β+βj)Λf,ϵWF(Sλ=2(z,θ,β+βj))Λf,ϵ(NK(z,θ,β+βj)T(B×))\begin{split}&WF(u_{\lambda=2}(\ \cdot\ ;z,\theta,\beta+\beta_{j}^{\prime}))\setminus N^{*}K(z,\theta,\beta+\beta_{j}^{\prime})\\ &\subset\Lambda_{f,\epsilon}\circ WF(S_{\lambda=2}(z,\theta,\beta+\beta_{j}^{\prime}))\\ &\subset\Lambda_{f,\epsilon}\circ\big{(}N^{*}K(z,\theta,\beta+\beta_{j}^{\prime})\cap T^{*}(B\times\mathbb{R})\big{)}\end{split}

where Λf,ϵ\Lambda_{f,\epsilon} refers to the forward flowout canonical relation (51) for g=g2g=g_{2}. Hence the projection of (xj,lj,ϵg2γ˙2,j(lj),ϵ)(x_{j},l_{j},-\epsilon\flat_{g_{2}}\dot{\gamma}_{2,j}(l_{j}),\epsilon) in (65) lifts into two possible covectors in the wave front set WF(uλ(;z,θ,β+βj))WF(u_{\lambda}(\ \cdot\ ;z,\theta,\beta+\beta_{j}^{\prime})), both lying in a curve of the form (50) emanating from (NK(z,θ,β+βj)T(B×))\big{(}N^{*}K(z,\theta,\beta+\beta_{j}^{\prime})\cap T^{*}(B\times\mathbb{R})\big{)}. These two covectors are (xj,lj,ϵg2γ˙2,j(lj),ϵ)(x_{j},l_{j},-\epsilon\flat_{g_{2}}\dot{\gamma}_{2,j}(l_{j}),\epsilon) and (xj,lj,ϵg2γ˙3,j(lj),ϵ)(x_{j},l_{j},-\epsilon\flat_{g_{2}}\dot{\gamma}_{3,j}(l_{j}),\epsilon) where, analogously to γ3\gamma_{3}, the γ3,j\gamma_{3,j} stands for the unit-speed geodesic in (3,g2)(\mathbb{R}^{3},g_{2}) through xjx_{j} with velocity tangent to the mirror image of γ˙2,j(lj)\dot{\gamma}_{2,j}(l_{j}) over TxjWT_{x_{j}}\partial W at that point. Thus at least one of these covectors lies in a bicharacteristic (50) (for g=g2g=g_{2}) that emanates from NK(z,θ,β+βj)T(B×)N^{*}K(z,\theta,\beta+\beta_{j}^{\prime})\cap T^{*}(B\times\mathbb{R}). In particular, one of the curves r(γ2,j(r),r)r\mapsto(\gamma_{2,j}(r),r) and r(γ3,j(r),r)r\mapsto(\gamma_{3,j}(r),r) intersects K(z,θ,β+βj)(B×)K(z,\theta,\beta+\beta_{j}^{\prime})\cap(B\times\mathbb{R}). These curves approximate the geodesics r(γ2(r),r)r\mapsto(\gamma_{2}(r),r) and r(γ3(r),r)r\mapsto(\gamma_{3}(r),r) arbitrarily well inside the bounded set (B×){(z,0)}(B\times\mathbb{R})\setminus\{(z,0)\} so the intersection for large jj can take place only at the point (z,0)(z,0) due to the way the parameters (θ,β)M(\theta,\beta)\in M were fixed. Thus, we deduce that γ2,j(0)=z\gamma_{2,j}(0)=z. In particular, each γ2,j\gamma_{2,j} is a geodesic from zz to xj=γ2,j(lj)Υx_{j}=\gamma_{2,j}(l_{j})\in\Upsilon in (3,g2)(\mathbb{R}^{3},g_{2}) with length ljl_{j}. Consequently, we have distg2(xj,z)lj\text{dist}_{g_{2}}(x_{j},z)\leq l_{j} which implies

distg2(x,z)l=distg1(x,z)\text{dist}_{g_{2}}(x,z)\leq l=\text{dist}_{g_{1}}(x,z)

at the limit jj\longrightarrow\infty. Moreover, as γ2,j\gamma_{2,j} converges to γ2\gamma_{2} we conclude that γ2\gamma_{2} connects zz to xx in (M2,g2)(M_{2},g_{2}) and has the distance above as length. ∎

We can now apply the lemma above to deduce the following proposition:

Proposition 4.1.

Assume that the conditions of Theorem 1.1 hold. Then

distg1(z,W)=distg2(z,W)\text{dist}_{g_{1}}(z,\partial W)=\text{dist}_{g_{2}}(z,\partial W)

for every zUz\in U. Moreover, a boundary point in Υ\Upsilon is the nearest to zz in (W,g1)(W,g_{1}) if and only if it is a nearest point in (W,g2)(W,g_{2}) and for any such point x0Υx_{0}\in\Upsilon there exists a neighbourhood 𝒬\mathcal{Q} of x0x_{0} in the space W\partial W such that

distg1(z,x)=distg2(z,x),x𝒬.\text{dist}_{g_{1}}(z,x)=\text{dist}_{g_{2}}(z,x),\quad\forall x\in\mathcal{Q}.
Proof.

As a direct consequence of the lemma above we get that

(66) distg2(z,W)distg2(z,x0)distg1(z,x0)=distg1(z,W).\text{dist}_{g_{2}}(z,\partial W)\leq\text{dist}_{g_{2}}(z,x_{0})\leq\text{dist}_{g_{1}}(z,x_{0})=\text{dist}_{g_{1}}(z,\partial W).

By exchanging the roles of metrics and repeating the proof one derives the opposite inequality:

distg1(z,W)distg2(z,W).\text{dist}_{g_{1}}(z,\partial W)\leq\text{dist}_{g_{2}}(z,\partial W).

Thus,

distg1(z,W)=distg2(z,W).\text{dist}_{g_{1}}(z,\partial W)=\text{dist}_{g_{2}}(z,\partial W).

Substituting the identity above in (66) one deduces that

distg2(z,W)=distg2(z,x0),\text{dist}_{g_{2}}(z,\partial W)=\text{dist}_{g_{2}}(z,x_{0}),

which implies that x0x_{0} is a nearest boundary point also in (W,g2)(W,g_{2}). Similarly one shows that a nearest boundary point in (W,g2)(W,g_{2}) is nearest also in (W,g1)(W,g_{1}). In view of Lemma 4.1 there is a neighbourhood 𝒬\mathcal{Q} of x0x_{0} such that

distg2(z,x)distg1(z,x),x𝒬.\text{dist}_{g_{2}}(z,x)\leq\text{dist}_{g_{1}}(z,x),\quad\forall x\in\mathcal{Q}.

Again, we may swap the metrics and repeat the argument to obtain the reversed inequality for a sufficiently small neighbourhood 𝒬\mathcal{Q}. In conlusion,

distg2(z,x)=distg1(z,x),x𝒬.\text{dist}_{g_{2}}(z,x)=\text{dist}_{g_{1}}(z,x),\quad\forall x\in\mathcal{Q}.

We finish the section by deriving the main result from the proposition above:

Proof of Theorem 1.1.

It suffices to show that g1|z=g2|zg_{1}|_{z}=g_{2}|_{z} at arbitrary zUz\in U. Let x0Υx_{0}\in\Upsilon be a nearest boundary point to zz. Let 𝒬\mathcal{Q} be a neighbourhood of x0x_{0} in W\partial W as in the proposition above. As shown in [KKL01, Lemma 2.15], the distance function xdistgλ(x,z)x\mapsto\text{dist}_{g_{\lambda}}(x,z) on a sufficiently small 𝒬\mathcal{Q} is smooth and the derivative ϕλ(x):=dwdistgλ(x,w)|w=z\phi_{\lambda}(x):=-d_{w}\text{dist}_{g_{\lambda}}(x,w)|_{w=z} defines a diffeomorphism from 𝒬\mathcal{Q} to an open set ϕλ𝒬\phi_{\lambda}\mathcal{Q} in {θTz3:|θ|gλ=1}\{\theta\in T_{z}\mathbb{R}^{3}:|\theta|_{g_{\lambda}}=1\}.555We unfortunately have the opposite roles for zz and xx compared to the notation in the reference. The derivative ϕλ(x)\phi_{\lambda}(x) points towards xx along an optimal geodesic segment between zz and xx. It follows from Proposition 4.1 that ϕ1=ϕ2\phi_{1}=\phi_{2} on 𝒬\mathcal{Q}. Thus, for every vv in the open cone Γz:=+ϕ1𝒬=+ϕ2𝒬Tz3\Gamma_{z}:=\mathbb{R}_{+}\phi_{1}\mathcal{Q}=\mathbb{R}_{+}\phi_{2}\mathcal{Q}\subset T_{z}\mathbb{R}^{3} we have

(67) |v|g1=|v|g2,vΓz.|v|_{g_{1}}=|v|_{g_{2}},\quad\forall v\in\Gamma_{z}.

Fix basis vectors e1,e2,e3Γze_{1},e_{2},e_{3}\in\Gamma_{z} for Tz3T_{z}\mathbb{R}^{3} so close to each other that ej+ekΓze_{j}+e_{k}\in\Gamma_{z} holds for every j,k=1,2,3j,k=1,2,3. The identity (67) then implies

g1(ej,ek)=12|ej+ek|g1212|ej|g1212|ek|g12=12|ej+ek|g2212|ej|g2212|ek|g22=g2(ej,ek).g_{1}(e_{j},e_{k})=\frac{1}{2}|e_{j}+e_{k}|_{g_{1}}^{2}-\frac{1}{2}|e_{j}|_{g_{1}}^{2}-\frac{1}{2}|e_{k}|_{g_{1}}^{2}=\frac{1}{2}|e_{j}+e_{k}|_{g_{2}}^{2}-\frac{1}{2}|e_{j}|_{g_{2}}^{2}-\frac{1}{2}|e_{k}|_{g_{2}}^{2}=g_{2}(e_{j},e_{k}).

In conclusion, for every v=vjejv=v^{j}e_{j} and w=wkekw=w^{k}e_{k} in Tz3T_{z}\mathbb{R}^{3} we have

g1(v,w)=vjwkg1(ej,ek)=vjwkg2(ej,ek)=g2(v,w),g_{1}(v,w)=v^{j}w^{k}g_{1}(e_{j},e_{k})=v^{j}w^{k}g_{2}(e_{j},e_{k})=g_{2}(v,w),

that is, g1|z=g2|zg_{1}|_{z}=g_{2}|_{z}. ∎

Appendix A Supplemental Material for Section 3

Proof of Lemma 3.1.

Let us prove the case ±=+\pm=+. The proof for the other sign is similar. By assumptions, Pkj=pkj(x,Dx,Dt)P^{k}_{j}=p^{k}_{j}(x,D_{x},D_{t}), where

pkj(x,ξ,ω)=(|ξ|g2ω2)δkj+rkj(x,ξ),j,k=1,2,3,p^{k}_{j}(x,\xi,\omega)=(|\xi|_{g}^{2}-\omega^{2})\delta^{k}_{j}+r^{k}_{j}(x,\xi),\quad j,k=1,2,3,

for some rjk(x,ξ)m=1aj,mk(x,ξ)r_{j}^{k}(x,\xi)\sim\sum_{m=-1}^{\infty}a_{j,-m}^{k}(x,\xi) (we may as well proceed with general FM3(Ψ1cl(3))F\in M_{3}(\Psi^{1}_{cl}(\mathbb{R}^{3}))) where aj,mk(x,ξ)a_{j,-m}^{k}(x,\xi) is positively homogeneous of degree m-m with respect to ξ\xi. Let us formally solve

(DtδjlWlj(x,Dx))ZklPkj0(D_{t}\delta_{j}^{l}-W^{l}_{j}(x,D_{x}))Z^{k}_{l}-P^{k}_{j}\equiv 0

by substituting an ansatz in the asymptotic form

(68) Wkj=wkj(x,Dx),wkj(x,ξ)m=1wkj,m(x,ξ),wkj,1:=|ξ|gδkj,\displaystyle W^{k}_{j}=w^{k}_{j}(x,D_{x}),\quad w^{k}_{j}(x,\xi)\sim\sum_{m=-1}^{\infty}w^{k}_{j,-m}(x,\xi),\quad w^{k}_{j,1}:=|\xi|_{g}\delta^{k}_{j},
(69) Zkj=zkj(x,Dx,Dt),zkj(x,ξ,ω)m=1zj,mk(x,ξ,ω),zj,1k(x,ξ,ω):=(ω+|ξ|g)δkj\displaystyle Z^{k}_{j}=z^{k}_{j}(x,D_{x},D_{t}),\quad z^{k}_{j}(x,\xi,\omega)\sim\sum_{m=-1}^{\infty}z_{j,-m}^{k}(x,\xi,\omega),\quad z_{j,1}^{k}(x,\xi,\omega):=-(\omega+|\xi|_{g})\delta^{k}_{j}

where wkj,mw^{k}_{j,-m} and zj,mz_{j,-m} are positively homogeneous of degree m-m. We apply the standard formula

(70) Symbol(AB)α1α!(ξαa)(Dαb),a:=Symbol(A),b:=Symbol(B),Symbol(AB)\sim\sum_{\alpha}\frac{1}{\alpha!}(\partial_{\xi}^{\alpha}a)(D^{\alpha}b),\quad a:=Symbol(A),\quad b:=Symbol(B),

(see e.g. [GS94, Theorem 3.6]) put together terms that correspond to the same degree of homogeneity and assume that they sum to zero. Multiplying the highest order terms of the ansatz gives the correct principal part, (|ξ|g2ω2)δkj(|\xi|_{g}^{2}-\omega^{2})\delta^{k}_{j}, whereas for the rest of the terms one deduces the following conditions:

(71) ωzkj,m1s+h+|α|=m1α!(ξαwj,hl)(Dxαzkl,s)=aj,mk,m=1,0,1,2,3,.\displaystyle\omega z^{k}_{j,-m-1}-\sum_{s+h+|\alpha|=m}\frac{1}{\alpha!}(\partial_{\xi}^{\alpha}w_{j,-h}^{l})(D_{x}^{\alpha}z^{k}_{l,-s})=a_{j,-m}^{k},\quad m=-1,0,1,2,3,\dots.

By extracting the terms corresponding to s=m+1s=m+1 and h=m+1h=m+1 we obtain

(72) (ω|ξ|g)zj,m1k+wkj,m1(ω+|ξ|g)Jm1α!(ξαwj,hl)(Dαzkl,s)=aj,mk,\displaystyle(\omega-|\xi|_{g})z_{j,-m-1}^{k}+w^{k}_{j,-m-1}(\omega+|\xi|_{g})-\sum_{J_{m}}\frac{1}{\alpha!}(\partial_{\xi}^{\alpha}w_{j,-h}^{l})(D^{\alpha}z^{k}_{l,-s})=a_{j,-m}^{k},
(73) Jm:={(α,h,s):s+h=m|α|,|α|1,s,h1}3×{1,,m}2\displaystyle J_{m}:=\{(\alpha,h,s):s+h=m-|\alpha|,\ |\alpha|\geq 1,\ s,h\geq-1\}\subset\mathbb{N}^{3}\times\{-1,\dots,m\}^{2}

for j,k=1,2,3j,k=1,2,3 and m=1,0,1,2,3,m=-1,0,1,2,3,\dots Fixing ω=|ξ|g\omega=|\xi|_{g} in the equation yields

wkj,m1(x,ξ)=12|ξ|g(Jm1α!(ξαwj,hl)(Dαzkl,s)+akj,m(x,ξ))w^{k}_{j,-m-1}(x,\xi)=\frac{1}{2|\xi|_{g}}\bigg{(}\sum_{J_{m}}\frac{1}{\alpha!}(\partial_{\xi}^{\alpha}w_{j,-h}^{l})(D^{\alpha}z^{k}_{l,-s})+a^{k}_{j,-m}(x,\xi)\bigg{)}

which solves wkj,m1w^{k}_{j,-m-1} recursively from the terms of degree m,,1-m,\dots,1. Analogously, one derives zj,m1k(x,ξ,ω)=wkj,m1(x,ξ)z_{j,-m-1}^{k}(x,\xi,\omega)=-w^{k}_{j,-m-1}(x,\xi) by setting ω=|ξ|g\omega=-|\xi|_{g} in (72) which also removes the terms with ω\omega. However, the elements obtained in this way have singularities at {ξ=0}\{\xi=0\}. To get around this one redefines the terms by multiplying with the cut-off χ(ω2ϵ|ξ|)\chi(\frac{\omega}{2\epsilon|\xi|}), where χCc()\chi\in C^{\infty}_{c}(\mathbb{R}) such that χ=1\chi=1 in (1/2,1/2)(-1/2,1/2) and 0 outside (1,1)(-1,1). The cut-off does not change the elements in 𝒱\mathcal{V} and hence the claim follows for hh given by hkjm=0wjmkh^{k}_{j}\sim\sum_{m=0}^{\infty}w_{j-m}^{k}. ∎

Proof of Lemma 3.3.

We begin by introducing an useful notation. A composition VWVW of two operators VM3(Ψl1(4))V\in M_{3}(\Psi^{l_{1}}(\mathbb{R}^{4})) and WM3(Ψl2(4))W\in M_{3}(\Psi^{l_{2}}(\mathbb{R}^{4})) with symbol matrices vM3(Sl1(4×4))v\in M_{3}(S^{l_{1}}(\mathbb{R}^{4}\times\mathbb{R}^{4})) and wM3(Sl2(4×4))w\in M_{3}(S^{l_{2}}(\mathbb{R}^{4}\times\mathbb{R}^{4})) admits as a symbol the matrix v#wM3(Sl1+l2(4×4))v\#w\in M_{3}(S^{l_{1}+l_{2}}(\mathbb{R}^{4}\times\mathbb{R}^{4})), given by

(v#w)kj(x,t,ξ,ω)s=0|α|=s1α!(αξvlj(x,t,ξ,ω))(Dxαwlk(x,t,ξ,ω)),j,k=1,2,3.(v\#w)^{k}_{j}(x,t,\xi,\omega)\sim\sum_{s=0}^{\infty}\sum_{|\alpha|=s}\frac{1}{\alpha!}(\partial^{\alpha}_{\xi}v^{l}_{j}(x,t,\xi,\omega))(D_{x}^{\alpha}w_{l}^{k}(x,t,\xi,\omega)),\quad j,k=1,2,3.

Further, if vv and ww have classical asymptotic developments vh=l1vhv\sim\sum_{h=-l_{1}}^{\infty}v_{-h} and wh=l2whw\sim\sum_{h=-l_{2}}^{\infty}w_{-h} for homogeneous wh,vhM3(Shhom(4×4))w_{-h},v_{-h}\in M_{3}(S^{-h}_{hom}(\mathbb{R}^{4}\times\mathbb{R}^{4})) of degree h-h, we have the classical asymptotics

v#wh=l2v#wh:=s=l1l2h=l2l1+svhs#whv\#w\sim\sum_{h=-l_{2}}v\#w_{-h}:=\sum_{s=-l_{1}-l_{2}}^{\infty}\sum_{h=-l_{2}}^{l_{1}+s}v_{h-s}\#w_{-h}

The proof of these formulas reduce to the scalar case (cf. [GS94, Theorem 3.6] or [Dui96, Theorem 2.5.2]) by considering each entry individually.

Let us now continue to the actual proof of the lemma. Set a:=(akj)j,k=1,2,3M3(Smcl(4×4))a:=(a^{k}_{j})_{j,k=1,2,3}\in M_{3}(S^{m}_{cl}(\mathbb{R}^{4}\times\mathbb{R}^{4})) where the entry akjSmcl(4×4)a^{k}_{j}\in S^{m}_{cl}(\mathbb{R}^{4}\times\mathbb{R}^{4}) stands for the symbol of LkjL^{k}_{j}. Each entry admits an asymptotic expansion ajkh=maj,hka_{j}^{k}\sim\sum_{h=-m}^{\infty}a_{j,-h}^{k}, where aj,hk(x,t,ξ,ω)a_{j,-h}^{k}(x,t,\xi,\omega) is positively homogeneous of degree h-h in ξ,ω\xi,\omega and we define ah=(aj,hk)j,k=1,2,3M3(Sh(4×4))a_{-h}=(a_{j,-h}^{k})_{j,k=1,2,3}\in M_{3}(S^{-h}(\mathbb{R}^{4}\times\mathbb{R}^{4})). Without loss of generality we may fix ϵ,δ>0\epsilon,\delta>0 such that ama_{m} is invertible in the conic set

(74) Σϵ,δ:={(x,t,ξ,ω)4×(40):|(x,t)(x0,t0)|ϵ,|(ξ,ω)|(ξ,ω)|(ξ0,ω0)|(ξ0,ω0)||δ}.\Sigma_{\epsilon,\delta}:=\left\{(x,t,\xi,\omega)\in\mathbb{R}^{4}\times(\mathbb{R}^{4}\setminus 0):|(x,t)-(x_{0},t_{0})|\leq\epsilon,\ \left|\frac{(\xi,\omega)}{|(\xi,\omega)|}-\frac{(\xi_{0},\omega_{0})}{|(\xi_{0},\omega_{0})|}\right|\leq\delta\right\}.

As the matrix ama_{m} is positively homogeneous of degree mm in ξ,ω\xi,\omega, the inverse matrix a1ma^{-1}_{m} on Σϵ,δ\Sigma_{\epsilon,\delta} is positively homogeneous of degree m-m. By proceeding as in the proof of [GS94, Theorem 4.1] we construct a suitable matrix b=(bkj)j,k=1,2,3b=(b^{k}_{j})_{j,k=1,2,3} where each bkjb^{k}_{j} defines the symbol of RkjR^{k}_{j}, j=1,2,3j=1,2,3. bh=mbhb\sim\sum_{h=m}^{\infty}b_{-h}. Let χC()\chi\in C^{\infty}(\mathbb{R}) be a cut-off function such that χ|(0,δ/2)=1\chi|_{(0,\delta/2)}=1 and supp(χ)(0,δ)\text{supp}(\chi)\subset(0,\delta). Set 𝒱:=Σϵ/2,δ/2\mathcal{V}:=\Sigma_{\epsilon/2,\delta/2}. We define the term bmb_{-m} by

bm(x,t,ξ,ω):={χ~(x,t,ξ,ω)(am)1(x,t,ξ,ω),(x,t,ξ,ω)Σϵ,δ0,(x,t,ξ,ω)T4Σϵ,δ,b_{-m}(x,t,\xi,\omega):=\begin{cases}\tilde{\chi}(x,t,\xi,\omega)(a_{m})^{-1}(x,t,\xi,\omega),&(x,t,\xi,\omega)\in\Sigma_{\epsilon,\delta}\\ 0,&(x,t,\xi,\omega)\in T^{*}\mathbb{R}^{4}\setminus\Sigma_{\epsilon,\delta},\end{cases}

where

(75) χ~(x,t,ξ,ω):=χ(|(x,t)(x0,t0)|)χ(|(ξ,ω)|(ξ,ω)|(ξ0,ω0)|(ξ0,ω0)||).\displaystyle\tilde{\chi}(x,t,\xi,\omega):=\chi(|(x,t)-(x_{0},t_{0})|)\chi\Big{(}\left|\frac{(\xi,\omega)}{|(\xi,\omega)|}-\frac{(\xi_{0},\omega_{0})}{|(\xi_{0},\omega_{0})|}\right|\Big{)}.

Then bmb_{-m} is positively homogeneous of degree m-m and equals (am)1(a_{m})^{-1} on 𝒱\mathcal{V}. The leading terms in the asymptotic expansions a#bmh=mah#bma\#b_{-m}\sim\sum_{h=-m}^{\infty}a_{-h}\#b_{-m} and bm#ah=mbm#ahb_{-m}\#a\sim\sum_{h=-m}^{\infty}b_{-m}\#a_{-h} are the compositions ambma_{m}b_{-m} and bmamb_{-m}a_{m} which in 𝒱\mathcal{V} reduce to the identity matrix by definition. Thus, we can find f,gM3(Scl1(4×4))f,g\in M_{3}(S_{cl}^{-1}(\mathbb{R}^{4}\times\mathbb{R}^{4})) such that for (x,t,ξ,ω)𝒱(x,t,\xi,\omega)\in\mathcal{V},

(76) (If)(x,t,ξ,ω)\displaystyle(I-f)(x,t,\xi,\omega) =a#bm(x,t,ξ,ω),\displaystyle=a\#b_{-m}(x,t,\xi,\omega),
(77) (Ig)(x,t,ξ,ω)\displaystyle(I-g)(x,t,\xi,\omega) =bm#a(x,t,ξ,ω),\displaystyle=b_{-m}\#a(x,t,\xi,\omega),

where II stands for the identity matrix. Define

(78) bfbm#(I+f+f#f+f#f#f+),\displaystyle b_{f}\sim b_{-m}\#(I+f+f\#f+f\#f\#f+\dots),
(79) bg(I+g+g#g+g#g#g+)#bm.\displaystyle b_{g}\sim(I+g+g\#g+g\#g\#g+\dots)\#b_{-m}.

In 𝒱\mathcal{V} we deduce a#bfIbg#aa\#b_{f}\equiv I\equiv b_{g}\#a modulo M3(S(4×4))M_{3}(S^{-\infty}(\mathbb{R}^{4}\times\mathbb{R}^{4})) by substituting the identities above. Consequently,

WF(ILRf)𝒱=,WF(IRgL)𝒱=,WF(I-LR_{f})\cap\mathcal{V}=\emptyset,\quad WF(I-R_{g}L)\cap\mathcal{V}=\emptyset,

where

(Rf)kj:=(bf)kj(x,t,Dx,Dt),(Rg)kj:=(bg)kj(x,t,Dx,Dt).\begin{split}(R_{f})^{k}_{j}:=(b_{f})^{k}_{j}(x,t,D_{x},D_{t}),\quad(R_{g})^{k}_{j}:=(b_{g})^{k}_{j}(x,t,D_{x},D_{t}).\end{split}

Further,

bgI#bg(bf#a)#bgbf#(a#bg)bf#Ibfin𝒱b_{g}\equiv I\#b_{g}\equiv(b_{f}\#a)\#b_{g}\equiv b_{f}\#(a\#b_{g})\equiv b_{f}\#I\equiv b_{f}\quad\text{in}\quad\mathcal{V}

which implies RfRgR_{f}\equiv R_{g}. In conclusion, we may define the terms with lower degree of homogeneity by

bmh:=bmh+1#f,h=1,2,3,,\begin{split}b_{-m-h}:=b_{-m-h+1}\#f,\quad h=1,2,3,\dots,\end{split}

(i.e. R:=Rf=RgR:=R_{f}=R_{g}) to obtain the result. ∎

Proof of Lemma 3.4.

Let us show that the bicharacteristics are of the form (50). Denote H:=Hp±H:=H_{p^{\pm}}, p±(x,t,ξ,ω):=ω|ξ|gp^{\pm}(x,t,\xi,\omega):=\omega\mp|\xi|_{g}. By definition,

H=t gjkξj|ξ|gxk± ξlξj2|ξ|ggjlxkξk.H=\frac{\partial}{\partial t}\mp \frac{g^{jk}\xi_{j}}{|\xi|_{g}}\frac{\partial}{\partial x^{k}}\pm \frac{\xi_{l}\xi_{j}}{2|\xi|_{g}}\frac{\partial g^{jl}}{\partial x^{k}}\frac{\partial}{\partial\xi_{k}}.

Taking a derivative of gjlglm=δjmg^{jl}g_{lm}=\delta^{j}_{m} yields glmkgjl+gjlkglm=0g_{lm}\partial_{k}g^{jl}+g^{jl}\partial_{k}g_{lm}=0. That is, kgjl=glmgjhkghm\partial_{k}g^{jl}=-g^{lm}g^{jh}\partial_{k}g_{hm}. Hence,

H=t gjkξj|ξ|gxk (ξlglm)(ξjgjh)2|ξ|ggmhxkξk.H=\frac{\partial}{\partial t}\mp \frac{g^{jk}\xi_{j}}{|\xi|_{g}}\frac{\partial}{\partial x^{k}}\mp \frac{(\xi_{l}g^{lm})(\xi_{j}g^{jh})}{2|\xi|_{g}}\frac{\partial g_{mh}}{\partial x^{k}}\frac{\partial}{\partial\xi_{k}}.

Thus, a bicharacteristic curve satisfies

(80) X˙k(r)=gjk(X(r))Ξj(r)|Ξ(r)|g,\displaystyle\dot{X}^{k}(r)=\mp\frac{g^{jk}(X(r))\Xi_{j}(r)}{|\Xi(r)|_{g}},
(81) T˙(r)=1,\displaystyle\dot{T}(r)=1,
(82) Ξ˙k(r)=12|Ξ(r)|Ξlglm(X(r))Ξjgjh(X(r))gmhxk|X(r),\displaystyle\dot{\Xi}_{k}(r)=\mp\frac{1}{2|\Xi(r)|}\Xi_{l}g^{lm}(X(r))\Xi_{j}g^{jh}(X(r))\frac{\partial g_{mh}}{\partial x^{k}}\Big{|}_{X(r)},
(83) Ω˙=0.\displaystyle\dot{\Omega}=0.

An initial value at r=0r=0 fixes the solution uniquely. The equations (80), (81) and (83) are solved by

X(r)=γx,v(r|v|g),Ξ(r)=gγ˙x,v(r|v|g),T(r)=t+r,Ω=±|v|g,X(r)=\gamma_{x,v}\left(\frac{r}{|v|_{g}}\right),\quad\Xi(r)=\mp\flat_{g}\dot{\gamma}_{x,v}\left(\frac{r}{|v|_{g}}\right),\quad T(r)=t+r,\quad\Omega=\pm|v|_{g},

where the initial value Γ(0)=(x,t,ξ,±|ξ|g)\Gamma(0)=(x,t,\xi,\pm|\xi|_{g}) is obtained by setting v:=gξv:=\mp\sharp_{g}\xi. Let us show that this solves also (82). It is suffices to show that

(84) r(gγ˙x,v(r|v|g))k=12|v|gγ˙mx,v(r|v|g)γ˙x,vh(r|v|g)kgmh|γx,v(r/|v|g).\partial_{r}\left(\flat_{g}\dot{\gamma}_{x,v}\left(\frac{r}{|v|_{g}}\right)\right)_{k}=\frac{1}{2|v|_{g}}\dot{\gamma}^{m}_{x,v}\left(\frac{r}{|v|_{g}}\right)\dot{\gamma}_{x,v}^{h}\left(\frac{r}{|v|_{g}}\right)\partial_{k}g_{mh}\big{|}_{\gamma_{x,v}(r/|v|_{g})}.

We compute

r(gγ˙x,v(r|v|g))=r(glk(γx,v(r|v|))γ˙x,vl(r|v|g))=1|v|gγ˙jx,v(r|v|g)γ˙x,vl(r|v|g)jglk|γx,v(r/|v|g)+1|v|gγ¨x,vh(r|v|g)ghk(γx,v(r|v|g)).\begin{split}&\partial_{r}\left(\flat_{g}\dot{\gamma}_{x,v}\left(\frac{r}{|v|_{g}}\right)\right)=\partial_{r}\left(g_{lk}\left(\gamma_{x,v}\left(\frac{r}{|v|}\right)\right)\dot{\gamma}_{x,v}^{l}\left(\frac{r}{|v|_{g}}\right)\right)\\ &=\frac{1}{|v|_{g}}\dot{\gamma}^{j}_{x,v}\left(\frac{r}{|v|_{g}}\right)\dot{\gamma}_{x,v}^{l}\left(\frac{r}{|v|_{g}}\right)\partial_{j}g_{lk}\big{|}_{\gamma_{x,v}(r/|v|_{g})}+\frac{1}{|v|_{g}}\ddot{\gamma}_{x,v}^{h}\left(\frac{r}{|v|_{g}}\right)g_{hk}\left(\gamma_{x,v}\left(\frac{r}{|v|_{g}}\right)\right).\end{split}

As γx,v\gamma_{x,v} is a geodesic, one derives

ghk(γ)γ¨x,vh=γ˙x,vjγ˙x,vlghk(γ)Γjlh(γ)=γ˙x,vjγ˙x,vljglk|γ+12kgjl|γγ˙x,vjγ˙x,vlg_{hk}(\gamma)\ddot{\gamma}_{x,v}^{h}=-\dot{\gamma}_{x,v}^{j}\dot{\gamma}_{x,v}^{l}g_{hk}(\gamma)\Gamma_{jl}^{h}(\gamma)=-\dot{\gamma}_{x,v}^{j}\dot{\gamma}_{x,v}^{l}\partial_{j}g_{lk}|_{\gamma}+\frac{1}{2}\partial_{k}g_{jl}|_{\gamma}\dot{\gamma}_{x,v}^{j}\dot{\gamma}_{x,v}^{l}

where Γjlh\Gamma_{jl}^{h}, j,l,h=1,2,3j,l,h=1,2,3, are the Christoffel symbols of gg. Substituting this in the previous equation yields (84). ∎

Proof of Lemma 3.5.

Since the Hamiltonian vector field corresponding to DsD_{s} is simply Hρ=sH_{\rho}=\partial_{s}, it follows immediately that the bicharacteristics of DsD_{s} (i.e. the characteristic integral curves of the Hamiltonian vector field) equal r(y,s+r,η,0)r\mapsto(y,s+r,\eta,0), for (y,s,η)4×3(y,s,\eta)\in\mathbb{R}^{4}\times\mathbb{R}^{3}. By Lemma 3.4 it suffices to show that the curve F:r(y,s+r,η,0)F:r\mapsto\mathcal{H}(y,s+r,\eta,0) is an integral curve of the Hamiltonian Hp±H_{p^{\pm}}. For fC(T4)f\in C^{\infty}(T^{*}\mathbb{R}^{4}) we compute (See [Duistermaat, Theorem 3.5.2.])

Hp±(F(r))f={p±,f}(y,s+r,η,0)={p±,f}(y,s+r,η,0)={ρ,f}(y,s+r,η,0)=F˙(r)f.\begin{split}H_{p^{\pm}}(F(r))f&=\{p^{\pm},f\}\circ\mathcal{H}(y,s+r,\eta,0)=\{p^{\pm}\circ\mathcal{H},f\circ\mathcal{H}\}(y,s+r,\eta,0)\\ &=\{\rho,f\circ\mathcal{H}\}(y,s+r,\eta,0)=\dot{F}(r)f.\\ \end{split}

This proves the claim. ∎

Derivation of (48).

Recall that 𝒳𝒴I\mathcal{X}\mathcal{Y}\equiv I in a small conic neighbourhood 𝒱\mathcal{V} of (x0,t0,ξ0,ω0)(x_{0},t_{0},\xi_{0},\omega_{0}) and that the operator wave-fronts of 𝒳\mathcal{X} and 𝒴\mathcal{Y} are the graphs GG_{\mathcal{H}} and G1G_{\mathcal{H}^{-1}}, respectively. Hence,

𝒱j=1,2,3WF(fj)=𝒱j=1,2,3WF(𝒳lj𝒴lkfk)𝒱j=1,2,3GWF(𝒴jkfk)=𝒱j=1,2,3WF(𝒴jkfk)𝒱j=1,2,3G1WF(fj)=𝒱j=1,2,3WF(fj),\begin{split}\mathcal{V}\cap\bigcup_{j=1,2,3}WF(f_{j})=\mathcal{V}\cap\bigcup_{j=1,2,3}WF(\mathcal{X}^{l}_{j}\mathcal{Y}_{l}^{k}f_{k})\\ \subset\mathcal{V}\cap\bigcup_{j=1,2,3}G_{\mathcal{H}}\circ WF(\mathcal{Y}_{j}^{k}f_{k})=\mathcal{V}\cap\bigcup_{j=1,2,3}\mathcal{H}WF(\mathcal{Y}_{j}^{k}f_{k})\\ \subset\mathcal{V}\cap\bigcup_{j=1,2,3}\mathcal{H}G_{\mathcal{H}^{-1}}\circ WF(f_{j})=\mathcal{V}\cap\bigcup_{j=1,2,3}WF(f_{j}),\end{split}
𝒱j=1,2,3WF(fj)=𝒱j=1,2,3WF(𝒴jkfk).\Rightarrow\mathcal{V}\cap\bigcup_{j=1,2,3}WF(f_{j})=\mathcal{V}\cap\bigcup_{j=1,2,3}\mathcal{H}WF(\mathcal{Y}_{j}^{k}f_{k}).

Define a conic neighbourhood 𝒲T4\mathcal{W}\subset T^{*}\mathbb{R}^{4} by 𝒲=1𝒱\mathcal{W}=\mathcal{H}^{-1}\mathcal{V}. Then,

𝒲1j=1,2,3WF(fj)=1(𝒱j=1,2,3WF(fj))=1(𝒱j=1,2,3WF(𝒴jkfk))=1𝒱j=1,2,31WF(𝒴jkfk)=𝒲j=1,2,3WF(𝒴jkfk).\begin{split}\Rightarrow\mathcal{W}\cap\mathcal{H}^{-1}\bigcup_{j=1,2,3}WF(f_{j})=\mathcal{H}^{-1}\bigg{(}\mathcal{V}\cap\bigcup_{j=1,2,3}WF(f_{j})\bigg{)}\\ =\mathcal{H}^{-1}\bigg{(}\mathcal{V}\cap\bigcup_{j=1,2,3}\mathcal{H}WF(\mathcal{Y}_{j}^{k}f_{k})\bigg{)}=\mathcal{H}^{-1}\mathcal{V}\cap\bigcup_{j=1,2,3}\mathcal{H}^{-1}\mathcal{H}WF(\mathcal{Y}_{j}^{k}f_{k})\\ =\mathcal{W}\cap\bigcup_{j=1,2,3}WF(\mathcal{Y}_{j}^{k}f_{k}).\end{split}

Hence,

𝒲1j=1,2,3WF(fj)=𝒲j=1,2,3WF(𝒴jkfk).\mathcal{W}\cap\mathcal{H}^{-1}\bigcup_{j=1,2,3}WF(f_{j})=\mathcal{W}\cap\bigcup_{j=1,2,3}WF(\mathcal{Y}_{j}^{k}f_{k}).

References

  • [BBB+06] S. W. Barwick, J. J. Beatty, D. Z. Besson, W. R. Binns, B. Cai, J. M. Clem, A. Connolly, D. F. Cowen, P. F. Dowkontt, M. A. DuVernois, P. A. Evenson, D. Goldstein, P. W. Gorham, C. L. Hebert, M. H. Israel, J. G. Learned, K. M. Liewer, J. T. Link, S. Matsuno, P. Miočinović, J. Nam, C. J. Naudet, R. Nichol, K. Palladino, M. Rosen, D. Saltzberg, D. Seckel, A. Silvestri, B. T. Stokes, G. S. Varner, and F. Wu. Constraints on cosmic neutrino fluxes from the antarctic impulsive transient antenna experiment. Phys. Rev. Lett., 96:171101, May 2006.
  • [BCK+] S P Bugaev, V A Cherepenin, V I Kanavets, A I Klimov, A D Kopenkin, V I Koshelev, V A Popov, and A I Slepkov. Relativistic multiwave cerenkov generators. IEEE Transactions on Plasma Science (Institute of Electrical and Electronics Engineers); (USA).
  • [BFRV10] J. A. Barceló, D. Faraco, A. Ruiz, and A. Vargas. Reconstruction of singularities from full scattering data by new estimates of bilinear Fourier multipliers. Math. Ann., 346(3):505–544, 2010.
  • [BJ18] Guillaume Bal and Alexandre Jollivet. Generalized stability estimates in inverse transport theory. Inverse Probl. Imaging, 12(1):59–90, 2018.
  • [Bol09] B. M. Bolotovskii. Vavilov-cherenkov radiation: its discovery and application. Phys. Usp., 52(11):1099–1110, 2009.
  • [CLOP19] Xi Chen, Matti Lassas, Lauri Oksanen, and Gabriel P. Paternain. Detection of hermitian connections in wave equations with cubic non-linearity, 2019.
  • [CLOP20] X. I. Chen, Matti Lassas, Lauri Oksanen, and Gabriel P. Paternain. Inverse problem for the Yang-Mills equations. 5 2020.
  • [CS96a] M. Choulli and P. Stefanov. Inverse scattering and inverse boundary value problems for the linear Boltzmann equation. Comm. Partial Differential Equations, 21(5-6):763–785, 1996.
  • [CS96b] M. Choulli and P. Stefanov. Reconstruction of the coefficients of the stationary transport equation from boundary measurements. Inverse Problems, 12(5):L19–L23, 1996.
  • [DCB+18] Emiko Desvaux, Alan Courteau, Pierre-Simon Bellaye, Mélanie Guillemin, Camille Drouet, Paul Walker, Bertrand Collin, and Richard Decréau. Cherenkov luminescence imaging is a fast and relevant preclinical tool to assess tumour hypoxia in vivo. EJNMMI Research, 8, 12 2018.
  • [DH72] J. J. Duistermaat and L. Hörmander. Fourier integral operators. II. Acta Math., 128(3-4):183–269, 1972.
  • [dHHI+14] Maarten V. de Hoop, Sean F. Holman, Einar Iversen, Matti Lassas, and Bjørn Ursin. Reconstruction of a conformally euclidean metric from local boundary diffraction travel times. SIAM Journal on Mathematical Analysis, 46(6):3705–3726, 2014.
  • [dHHI+15] Maarten V. de Hoop, Sean F. Holman, Einar Iversen, Matti Lassas, and Bjø rn Ursin. Recovering the isometry type of a Riemannian manifold from local boundary diffraction travel times. J. Math. Pures Appl. (9), 103(3):830–848, 2015.
  • [dHILS20] Maarten V. de Hoop, Joonas Ilmavirta, Matti Lassas, and Teemu Saksala. A foliated and reversible finsler manifold is determined by its broken scattering relation, 2020.
  • [Dui96] J. J. Duistermaat. Fourier integral operators, volume 130 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 1996.
  • [Ego69] Ju. V. Egorov. The canonical transformations of pseudodifferential operators. Uspehi Mat. Nauk, 24(5 (149)):235–236, 1969.
  • [FLO20] Ali Feizmohammadi, Matti Lassas, and Lauri Oksanen. Inverse problems for non-linear hyperbolic equations with disjoint sources and receivers, 2020.
  • [GLS+18] Allan Greenleaf, Matti Lassas, Matteo Santacesaria, Samuli Siltanen, and Gunther Uhlmann. Propagation and recovery of singularities in the inverse conductivity problem. Anal. PDE, 11(8):1901–1943, 2018.
  • [GS94] Alain Grigis and Johannes Sjöstrand. Microlocal analysis for differential operators, volume 196 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 1994. An introduction.
  • [GSF+05] P. W. Gorham, D. Saltzberg, R. C. Field, E. Guillian, R. Milinčić, P. Miočinović, D. Walz, and D. Williams. Accelerator measurements of the askaryan effect in rock salt: A roadmap toward teraton underground neutrino detectors. Phys. Rev. D, 72:023002, Jul 2005.
  • [GU81] V. Guillemin and G. Uhlmann. Oscillatory integrals with singular symbols. Duke Math. J., 48(1):251–267, 1981.
  • [GU90] Allan Greenleaf and Gunther Uhlmann. Estimates for singular Radon transforms and pseudodifferential operators with singular symbols. J. Funct. Anal., 89(1):202–232, 1990.
  • [GU93] Allan Greenleaf and Gunther Uhlmann. Recovering singularities of a potential from singularities of scattering data. Comm. Math. Phys., 157(3):549–572, 1993.
  • [Gui75] Victor Guillemin. Clean intersection theory and fourier integrals. In Jacques Chazarain, editor, Fourier Integral Operators and Partial Differential Equations, pages 23–35, Berlin, Heidelberg, 1975. Springer Berlin Heidelberg.
  • [H7̈1] Lars Hörmander. Fourier integral operators. I. Acta Math., 127(1-2):79–183, 1971.
  • [HUZ21] Peter Hintz, Gunther Uhlmann, and Jian Zhai. An inverse boundary value problem for a semilinear wave equation on lorentzian manifolds, 2021.
  • [Jol13] Alexandre Jollivet. On inverse scattering at fixed energy for the multidimensional Newton equation in a non-compactly supported field. J. Inverse Ill-Posed Probl., 21(6):713–734, 2013.
  • [Jol14] Alexandre Jollivet. Inverse scattering at high energies for the multidimensional Newton equation in a long range potential. Asymptot. Anal., 90(1-2):105–132, 2014.
  • [Jol15] Alexandre Jollivet. Inverse scattering at high energies for classical relativistic particles in a long-range electromagnetic field. Ann. Henri Poincaré, 16(11):2569–2602, 2015.
  • [Jos98] Mark S. Joshi. Recovering the total singularity of a conformal potential from backscattering data. Ann. Inst. Fourier (Grenoble), 48(5):1513–1532, 1998.
  • [KKL01] Alexander Katchalov, Yaroslav Kurylev, and Matti Lassas. Inverse boundary spectral problems, volume 123 of Chapman & Hall/CRC Monographs and Surveys in Pure and Applied Mathematics. Chapman & Hall/CRC, Boca Raton, FL, 2001.
  • [KLA18] Nalinikanth Kotagiri, Richard Laforest, and Samuel Achilefu. Reply to ’is cherenkov luminescence bright enough for photodynamic therapy?’. Nature Nanotechnology, 13(5):354–355, May 2018.
  • [KLS06] Yaroslav Kurylev, Matti Lassas, and Erkki Somersalo. Maxwell’s equations with a polarization independent wave velocity: direct and inverse problems. J. Math. Pures Appl. (9), 86(3):237–270, 2006.
  • [KLU10] Yaroslav Kurylev, Matti Lassas, and Gunther Uhlmann. Rigidity of broken geodesic flow and inverse problems. Amer. J. Math., 132(2):529–562, 2010.
  • [Lax57] Peter D. Lax. Asymptotic solutions of oscillatory initial value problems. Duke Math. J., 24:627–646, 1957.
  • [Lee97] John M. Lee. Riemannian manifolds, volume 176 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1997. An introduction to curvature.
  • [LGJRD04] Nikolai G. Lehtinen, Peter W. Gorham, Abram R. Jacobson, and Robert A. Roussel-Dupre. FORTE satellite constraints on ultra-high energy cosmic particle fluxes. Phys. Rev. D, 69:013008, 2004.
  • [LL20] Ru-Yu Lai and Qin Li. Parameter reconstruction for general transport equation. SIAM J. Math. Anal., 52(3):2734–2758, 2020.
  • [Lud60] Donald Ludwig. Exact and asymptotic solutions of the Cauchy problem. Comm. Pure Appl. Math., 13:473–508, 1960.
  • [LUW18] Matti Lassas, Gunther Uhlmann, and Yiran Wang. Inverse problems for semilinear wave equations on Lorentzian manifolds. Comm. Math. Phys., 360(2):555–609, 2018.
  • [Mas65] Viktor P. Maslov. Perturbation theory and asymptotic methods. 1965.
  • [Mas88] V. P. Maslov. Asimptoticheskie metody i teoriya vozmushchenii. “Nauka”, Moscow, 1988.
  • [McD04] Stephen R. McDowall. An inverse problem for the transport equation in the presence of a Riemannian metric. Pacific J. Math., 216(2):303–326, 2004.
  • [McD05] Stephen R. McDowall. Optical tomography on simple Riemannian surfaces. Comm. Partial Differential Equations, 30(7-9):1379–1400, 2005.
  • [MMMT16] Dorina Mitrea, Irina Mitrea, Marius Mitrea, and Michael Taylor. The Hodge-Laplacian, volume 64 of De Gruyter Studies in Mathematics. De Gruyter, Berlin, 2016. Boundary value problems on Riemannian manifolds.
  • [MU79] R. B. Melrose and G. A. Uhlmann. Lagrangian intersection and the Cauchy problem. Comm. Pure Appl. Math., 32(4):483–519, 1979.
  • [OPS01] Petri Ola, Lassi Päivärinta, and Valeri Serov. Recovering singularities from backscattering in two dimensions. Comm. Partial Differential Equations, 26(3-4):697–715, 2001.
  • [PK18] Guillem Pratx and Daniel S Kapp. Is cherenkov luminescence bright enough for photodynamic therapy? Nature nanotechnology, 13(5):354–354, 2018.
  • [PS91] Lassi Päivärinta and Erkki Somersalo. Inversion of discontinuities for the Schrödinger equation in three dimensions. SIAM J. Math. Anal., 22(2):480–499, 1991.
  • [PS98] Lassi Päivärinta and Valeri Serov. Recovery of singularities of a multidimensional scattering potential. SIAM J. Math. Anal., 29(3):697–711, 1998.
  • [PSS94] Lassi Päivärinta, Valeri S. Serov, and Erkki Somersalo. Reconstruction of singularities of a scattering potential in two dimensions. Adv. in Appl. Math., 15(1):97–113, 1994.
  • [Rei71] William T. Reid. Ordinary differential equations. John Wiley & Sons, Inc., New York-London-Sydney, 1971.
  • [RHLG10] Alessandro Ruggiero, Jason P Holland, Jason S Lewis, and Jan Grimm. Cerenkov luminescence imaging of medical isotopes. The Journal of nuclear medicine (1978), 51(7):1123–1130, 2010.
  • [RR12] Juan Manuel Reyes and Alberto Ruiz. Reconstruction of the singularities of a potential from backscattering data in 2D and 3D. Inverse Probl. Imaging, 6(2):321–355, 2012.
  • [SGW+01] David Saltzberg, Peter Gorham, Dieter Walz, Clive Field, Richard Iverson, Allen Odian, George Resch, Paul Schoessow, and Dawn Williams. Observation of the Askaryan effect: Coherent microwave Cherenkov emission from charge asymmetry in high-energy particle cascades. Phys. Rev. Lett., 86:2802–2805, 2001.
  • [ST09] Plamen Stefanov and Alexandru Tamasan. Uniqueness and non-uniqueness in inverse radiative transfer. Proc. Amer. Math. Soc., 137(7):2335–2344, 2009.
  • [WZ19] Yiran Wang and Ting Zhou. Inverse problems for quadratic derivative nonlinear wave equations. Comm. Partial Differential Equations, 44(11):1140–1158, 2019.
  • [YS94] T. Ypsilantis and J. Seguinot. Theory of ring imaging cherenkov counters. Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment, 343(1):30–51, 1994.
  • [ZQY+11] Jianghong Zhong, Chenghu Qin, Xin Yang, Shuping Zhu, Xin Zhang, and Jie Tian. Cerenkov luminescence tomography for in vivo radiopharmaceutical imaging. International journal of biomedical imaging, 2011:641618, 04 2011.