Recovering a Riemannian Metric from Cherenkov Radiation in Inhomogeneous Anisotropic Medium
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 , where is a vector-valued wave operator that depends on a Riemannian metric and is a point source that moves at speed in given direction . The phase velocity is described by the metric and depends on both location and direction of motion. In regions where 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
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
where the Einstein summation convention is considered over , the operator with entries , is hyperbolic, and is a singular source with each coordinate conormal to the world line
(i.e. trajectory in space-time) of the particle. An example of this is a scalar model, e.g. the scalar wave equation
(identified as and ) with a source conormal to (e.g. ). The source can be interpreted as a singularity that moves at a physically allowed () constant velocity , through . If for , or more generally, for the product with non-degenerate and the leading part of the operator equals a scalar operator of real principal type, then the standard propagation of singularities extends to the case above. More precisely; the tensors and are neglectable in terms of singularities and the wave front set of splits into a non-propagating part which does not leave , and a propagating part which spreads along the bicharacteristics of into the surrounding space. We assume that the leading part is of the form
where stands for a Riemannian metric. The characteristic manifold for is the covector light-cone
also known as the null-cone, and the speed of wave propagation at in direction is . Provided that the source moves slower than waves (i.e. ) the normal bundle 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.
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 of the source. We show that information collected in this way on a convenient part of the surface determines uniquely the metric 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 stand for the identity matrix. Assume that the (1,1)-tensor on with differential operator entries is similar to the vector wave-operator in the sense that the condition
(1) |
that is,
(2) |
holds for some first order operators , . Consider a point-like object moving linearly in in direction at physically allowed speed ( in natural unit system). Let be the location at time . The associated world line in the space time is
We study singularities of distributions , , that obey
(3) |
where the terms are conormal distributions over , that is,
(4) |
E.g. if . Above the relation refers to . The source can be interpreted as a point singularity that moves in along the trajectory , , , whereas the field 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 conformally equivalent to the anisotropic permeability tensor . This model is explored in Section 1.6.
Remark 1.1.
Notice that an analogous scalar equation, e.g the scalar wave equation
with source conormal to , can also be expressed within the model by identifying any scalar operator with the diagonal matrix , and a scalar source with , . 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 of the tensor is an operator in the space of classical first order pseudo-differential operators. Moreover, the identity (1) can be replaced by the more general condition
(5) |
where the relation between two operators and refers to
for some operators and with pointwise non-degenerate principal symbol matrices and (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 -tensor generalizes the concept of elliptic multiplication operator for non-vanishing . 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
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) |
where we denote . Let be an open bounded set with smooth boundary and fix open . We consider as a region with unknown geometry inside the medium, whereas the boundary is the hypersurface on which observations are made. The source is controlled by the variables and .
To avoid Cherenkov radiation from arbitrarily far we consider mediums that converge to vacuum at infinity. That is; for every as . Alternatively, one may assume that there is a relatively large velocity threshold such that the preimage
(7) |
is bounded and contains for every , and then focus on velocities . In both cases the wave front set of the tail is disjoint from the characteristic set of for a test function , that equals in a sufficiently large open set. Hence, for a convenient operator (see Lemma 3.3) that inverts microlocally near in the element does not radiate singularities and satisfies
(8) |
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
in which the source is not only conormal to but also compactly supported.
Definition 1.1.
Let be as above. We say that a subset is a stable part of the boundary with respect to if is open in and for every there is at least one that minimises distance to , i.e.
where .
Remark 1.3.
The whole boundary is always stable.
For we let stand for the canonical projection . Given a family of compactly supported sources of the form (4) and a stable , the measurement data are defined as the following map:
where solves for each the conditions
(9) | |||
(10) |
for and some (sufficiently large) . Notice that each source is supported in for large (see Figure 4). As the high-energy limit is unpractical, we are interested in smaller datasets where the velocities are restricted to a smaller subinterval . The interval may depend on .
Remark 1.4.
Substituting in the definition with more general operator (see Remark 1.2) yields the same data. For example, the data are invariant in transformation and , where and are smooth non-degenerate -tensors.
1.2. The Main Result
The main outcome of this work is Theorem 1.1 below which states that the Riemannian metric in is uniquely determined by the data, provided that some natural conditions are satisfied. In fact, for each it suffices to consider data for velocities in a specific subinterval with strictly less than , the speed of light in vacuum. If the maximum phase velocity in the closure is strictly less than the speed of light in vacuum then the dependence on in the interval above can be dropped (see Remark 1.6). The requirements for the metric are as follows:
Definition 1.2.
Let be a Riemannian metric in and let be an open bounded set. We say that the metric is admissible with respect to if the following conditions hold:
-
(i)
Propagation speed of waves is everywhere less than or equal to the speed of light in vacuum:
-
(ii)
Propagation speed of waves in is strictly less than the speed of light in vacuum:
To avoid technical difficulties, the following lower bound for speeds is considered in the main theorem.
Definition 1.3.
Let be admissible metric with respect to . For and a bounded set containing we define
where is the Euclidean norm.
Remark 1.5.
It follows from compactness of that . Moreover, if propagation of waves in the closure is strictly slower than in vacuum, we have .
Below refers to matrices with entries in the class . In orthonormal Euclidean coordinates we identify a (1,1)-tensor , , with the matrix . The sources are required to satisfy for every , the following conditions:
(11) |
(12) | |||
(13) |
where is the positively homogeneous principal part of the symbol , and the radius depends continuously on . The main result of the paper is the following theorem:
Theorem 1.1.
Let be a bounded open set with smooth boundary and let , be two Riemannian metrics in , both admissible Definition 1.2 with respect to an open . Let be a stable part of the boudary Definition 1.1 with respect to for both metrics , and assume that
Consider two differential operators , of the form
where the entries , of are first order operators on , and sources that satisfy (11-13). Let solve
(14) | |||
(15) |
for some . Assume that for every there is an open nonempty subinterval
such that the data on coincide for both metrics, i.e.
for every . Then, .
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]
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) | |||
(17) | |||
(18) |
(, ) Here and , . The characteristic manifold for the wave operator is
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;
where stands for the forward propagating flow-out canonical relation generated by , i.e.
Here the composition stands for
One checks that
For we deduce
Thus, and we obtain
That is; in the regime no singularities radiate from the source. If , then there exist bicharacteristics through and the composition is non-trivial. In fact, the projection of it to is a union of the sets
over points that satisfy . Each of these manifolds corresponds to propagating singularities generated at the associated point by the source. The equation corresponds to the Frank-Tamm formula in physics. For 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
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) | ||||
(20) | ||||
(21) | ||||
(22) |
where , , , , , and 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 -tensors
(23) | ||||
(24) |
which give connections between the fields via
(25) | ||||
(26) |
For lossless, optically inactive materials it is reasonable to require the tensors and 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 such that . Consider a charge carrier (e.g. an electron) or some other non-smooth charge perturbation moving along a straight line in through anisotropic dielectric medium with smoothly varying inhomogeneities. Let , , and be the direction of motion, speed, and location at for the moving charge perturbation, respectively. Again, we work in a natural unit system that has as the speed of light in vacuum. The perturbation signal is encoded in the charge density as an oscillatory superposition
(27) |
where the amplitude has an asymptotic development
(28) |
into functions . For example, a single electron in vacuum corresponds to the density
where stands for the elementary charge and is the Lorentz factor. It follows (see e.g. [Dui96]) that the distribution (27) is smooth outside the trajectory and the order of singularity at each point corresponds to decay of with respect to . 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)
which follows from (21), (22) and the fact that . By substitution one checks that the equation is solved by
where is defined recursively by
(29) | |||
(30) |
and is any solenoidal field. For simplicity, let us assume that is smooth. Define a Riemannian metric by
and set where is the Hodge star on -forms with respect to . Under the requirements above the electric field satisfies (See [KLS06])
where and
(31) | ||||
(32) |
By applying Leibniz’ rule, we deduce that the standard codifferential differs from by only terms of order . For instance, for a 1-form we have that
Thus, the leading terms in coincide with the ones in the Hodge Laplacian . As shown in [MMMT16, Lemma 2.8], the leading part in the Hodge Laplacian for 1-forms is . In conclusion, we have that the electric field satisfies the preferred equation:
where , .
2. Preliminaries
Henceforth the symbol with integer stands for the spatial partial derivative .
2.1. Symbols
Let be a smooth manifold of dimension . The class of symbols of order and type is defined as the space of functions that satisfy the following: For each compact , and there exists such that
Only symbols of type are considered and hence the notational simplification . It follows from the definition that for and we define as the limit set . It is common to treat symbols modulo a residual term in . For a sequence of symbols , , with there exists , unique up to a term in , such that
for every This is denoted by and referred to as an asymptotic expansion or asymptotic development of .
The set of classical symbols is defined as the elements of the form
(33) |
where takes value in a neighbourhood of and is positively homogeneous of degree with respect to the variable . Even though positively homogeneous functions are usually not smooth, they are treated as symbols by omitting a smoothing near the origin. The notation for positively homogeneous is used also for referring to (33). The choice of the compactly supported cut-off function around is not significant as changing the function within contributes only by a term in . See e.g. [GS94] for a nice introduction to the topic.
2.2. Lagrangian Distributions and Fourier Integral Operators
A submanifold of a symplectic manifold of dimension is called Lagrangian if and , for every and . The cotangent bundle of a smooth manifold is a symplectic manifold with the canonical symplectic structure (see [GS94, §5]). The space of Lagrangian distributions of order on associated with a conic Lagrangian manifold consists of distributions that can be expressed as a locally finite sum of oscillatory integrals of the form
(34) |
where is a non-degenerate phase function, homogeneous of degree 1 with respect to , such that that the manifold coincides locally with the set (see e.g. [Dui96] for details). For the wave front set satisfies and the asymptotic behaviour of the symbol defines the degree of regularity at given conic neighbourhood. For instance, if where is positively homogeneous of degree with respect to , then the wave front set of the oscillatory integral (34) does not meet a conic neighbourhood inside the support of . As a special case, we have . If two symbols and satisfy , i.e. , then , that is, 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 equals , the conormal bundle of a submanifold . Such elements are referred to as distributions conormal to . For example, the delta distribution on is conormal to the diagonal . The class of distributions conormal to is often denoted by , where refers to the order of the symbol. Thus, , where is the codimension of in .
Considering a Lagrangian distribution on as a kernel gives an operator
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 that does not meet the definition can be extended to and further to if the operator is properly supported. See e.g. [Dui96, Corollary 1.3.8] for details. If we say that the associated operator lies in . The manifold refers to which is Lagrangian with respect to the symplectic form . Such a manifold is referred to as the canonical relation. Conversely, if a manifold is a canonical relation, then is Lagrangian manifold for the standard symplectic form and the kernel is a Lagrangian distribution on . A Fourier integral operator with a smooth kernel defines a regularizing map which is usually identified with the zero element.
An operator that admits the diagonal as a canonical relation is called pseudo-differential operator. We denote by the class of pseudo-differential operators of order on . Pseudo-differential operators with classical symbols is denoted by . All the pseudo-differential operators considered in this paper have . By placing a smooth cut-off 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 of two operators and with canonical relations and to be admissible in the framework is that the product intersects the manifold transversally. Provided that the required conditions are satisfied, we have
(35) | |||
(36) |
Moreover,
(37) |
where refers to the principal symbol (i.e. the leading term in an asymptotic development) on the canonical relation and . The diagonal acts as an identity element on canonical relations:
Considering Lagrangian distributions on as Fourier integral operators on with trivial one deduces
provided that the required conditions hold. The element above refers to
For a pseudo-differential operator we obtain
(cf. microlocality ).
2.4. Distributions associated with a pair of Lagrangian manifolds.
We also consider the class of Lagrangian distributions associated with a pair of cleanly intersecting Lagrangian manifolds 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 for a pseudo-differential operator of real principal type on is associated with the pair
of canonical relations where consists of pairs that lie in a same bicharacteristic. It is shown in [GU93, Proposition 2.1] that for a distribution in the class , where the Lagrangian manifold intersects the characteristic manifold of transversally and each bicharacteristic of intersects the manifold a finite number of times, the composition is a Lagrangian distribution that corresponds to the pair
Moreover, the identity (37) extends to .
2.5. Matrix Notation
We define the class as matrices of pseudo-differential operators , , of degree on as entries. The symbols of , form an element in of matrices with entries in the symbol class . Each operates on a vector-valued distribution , , , in the obvious way:
This is well defined also for if the entries are properly supported. In addition, we define the composition of and by
Any scalar operator is identified with the diagonal matrix , where stands for the Kronecker delta. That is;
for , , and
for .
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 to indicate a sign that can be chosen to be either or . We also define . For a conic neighbourhood and we say that
if
where refers to the wave front set of a pseudo-differential operator, that is, as a Fourier integral operator. Let us begin by factoring a vector-valued operator into first order operators. For a proof, see the appendix.
Lemma 3.1.
Consider a differential operator of the form
where is a Riemannian metric and is a stationary first order operator. Then, for every there are operators , , and such that
and
(38) | |||
(39) |
microlocally in .
Define , where is a Riemannian metric. Throughout this section we consider a non-zero covector in with and for a given sign and a fixed homogeneous canonical transformation
(40) |
defined on a conic neighbourhood of the covector with
We are only interested in which shall be required. Moreover, we assume that the last coordinate of equals , i.e, (see Darboux’ theorem [Dui96, Theorem 3.5.6]). The graph of is denoted by . The graph of , denoted by , is obtained by changing the order of coordinates in .
In the next step we follow Egorov’s theorem [GS94, Theorem 10.1]. We fix Fourier integral operators and with locally reciprocal principal symbols (positively homogeneous of degree ) such that is noncharacteristic for , is noncharacteristic for and there is a conic neighbourhood of and a conic neighbourhood of such that
(41) |
We may assume that the Schwartz kernels of and are compactly supported distributions. The principal symbol of equals the coordinate in a conic neighbourhood of . Hence, microlocally near ,
modulo , where is as in Lemma 3.1 and .
Definition 3.1.
Let stand for the symbols in that outside a bounded neighbourhood of the origin are positively homogeneous of degree in .
The following lemma provides a way to get rid of the extra term (cf. [DH72, Proposition 6.1.4], [GS94, Lemma 10.3]).
Lemma 3.2.
Let . There exists such that
Moreover, there is a neighbourhood of in and that is invertible as a matrix at every point and satisfies
Proof of Lemma 3.2.
Write each entry of in the asymptotic form
where is positively homogeneous of degree as a function of . We substitute the ansatz ,
in
(42) |
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) |
For every , let and be the matrices with entries and , respectively. Setting the leading terms in the symbol of (43) equal zero implies
that is,
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 , both uniqueness and existence of on a closed interval follows from [Rei71, Theorem 5.1]. We fix the initial value . Positive homogeneity of in then imply that the solution really is positively homogeneous of degree in . Moreover, regularity of the solution in is a consequence of [Rei71, Theorem 10.5]. As the initial value is an invertible matrix and the solution is positively homogeneous of degree , it follows by continuity that there is an open neighbourhood of in such that is invertible matrix for every . For the rest of the terms one derives
(44) |
where is a matrix depending on entries of , where and other indices satisfy . Each can be solved recursively from these equations. Proceeding inductively, similar arguments as in the case apply for existence and regularity of solutions with smooth initial values. It is also straightforward to check that positive homogeneity of degree in 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 , , let be the principal symbol of in the sense that
Assume that is an invertible matrix at some . Then, there is and a conic neighbourhood of in satisfying the conditions (i) and (ii) below.
-
(i)
is a (two sided) microlocal inverse of in :
-
(ii)
The principal symbols of and are reciprocal in : There is satisfying
and
Applying microlocality222 for to Lemma 3.3 yields the following:
Corollary 3.1.
Let be as in Lemma 3.2 and let be the corresponding microlocal inverse of it, as in Lemma 3.3. Define and . Applying the standard FIO calculus to , , and yields
(45) | ||||
(46) |
together with the following identities for in the domain of :
(47) |
Here refers to the principal symbol. As and near and , respectively, there are conic neighbourhoods and of and , respectively, such that
(48) | |||
(49) |
for any , . See derivation of (48) in Appendix A. The other equation is computed similarly.
Definition 3.2.
For we define a smooth curve
by
(50) |
where .
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 (i.e. an integral curve of the Hamilton vector field for in ) with as the initial value at .
Lemma 3.5.
Let and such that the segment lies in the domain of the canonical transormation . Then, the curve
equals , with as the initial value. That is; the transformation takes a bicharacteristic for into a bicharacteristic for .
Definition 3.3.
We define
(51) |
where is the curve (50) with the initial value . This is the forward propagating part of the characteristic flow-out canonical relation associated with . There is also the backwards propagating part , defined as
The complete flow-out canonical relation associated with is the union
Definition 3.4.
Let be the projection . For we define
The set consists of all the spatial points where a particle that moves through the point at a constant velocity 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 , , and . Let be a Riemannian metric in with on and assume that
(52) |
i.e. , . Let
be compactly supported and be of the form
for some stationary operator of order . Consider distributions , , that for large obey
(53) | |||
(54) |
Then,
Moreover, if the initial value is the first intersection between a bicharacteristic of the form (50) and , and the positively homogeneous principal symbol (of degree ) for at least one of the components , does not vanish at (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;
for
For proving the proposition we need the following lemma:
Lemma 3.6.
Assume that the Riemannian metric on with satisfies
Then the bundle is transversal to and the intersection of with a curve of the form (50) is a discrete set.
Proof of Lemma 3.6.
Denote . We begin by proving the first claim. As
we need to show that . Recall that
Thus, for we have the expression
As we have for that
Consequently, for (, ) we have
where is a linear map defined by
To prove that the dimension of the kernel is less or equal to we need to show that the map is not identically zero. By setting one obtains
(55) |
which is possible only if at . Let us study at such points. By substituting (55) in implies
which is nonzero by assumptions. In conclusion, the dimension of is at most which finishes the first part of the proof.
Let us now show that the intersection of and the curve of the form (50) is a collection of discrete points. To prove this we set and show that is not in for in a small neighbourhood of . By definition, , . If is not tangent to at , then the claim clearly holds. Thus, we may assume that which implies the following approximation near :
where . In addition, so at we must have that is, . In particular, by assumptions. As is a geodesic, it satisfies , where is the covariant derivative of a vector field along a curve (see e.g. [Lee97]). On the other hand,
Thus, which is possible only if around for some nonzero vector (i.e. the curve “bends” at ). In conclusion, for near the origin we have that does not lie in the line . ∎
Proof of Proposition 3.1.
We relax the notation by omitting the parameters 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 and a vector , we express the wave microlocally in a suitable conic neighbourhood of as a sum
(56) |
of a term which creates (or annihilates) singularities that propagate forwards in time along the characteristic flowout from the source, and a residual term with a flow-invariant wave front set. That is;
which can be chosen such that if , and
where
all in a suitable conic neighbourhood of . The residual term just moves existing singularities along bicharacteristics (50) in the sense that the wave front set in one region defines it in the flowout within . The wave front set of away from is computed by solving the microlocal expressions 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 so regularity of the wave outside implies that no singularities along the curve occur in until intersection with the first time in . Notice that the condition on ensures that each bicharacteristic really goes through the Cauchy surface . We must therefore have at the first intersection of the curve and . Hence, a forward propagating singularity is created at the point if for some we can show that the positively homogeneous principal symbol of does not vanish there. On the other hand, in a suitable conic neighbourhood around any point outside the expression leads to smooth and hence flow-invariant . 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 near , . As before, we let be the homogeneous local transformation (40) on a conic neighbourhood of the point with being the last coordinate. Applying Lemma 3.1 and Lemma 3.2 we derive
microlocally near . Thus, there are such that in the open cylinder we have modulo the local transport equations
(57) |
for
(58) | ||||
(59) | ||||
(60) |
where is a smooth cut-off function that vanishes outside and equals 1 in . Microlocally near we have that and . In view of (48-49), Corollary 3.1 (applied to ) and Lemma 3.5 it suffices to study singularities of each along bicharacteristics of . The original wave is derived microlocally near by applying the inverse of to . Formally,
(61) |
where is the microlocal inverse of near , stands for the forward fundamental solution (operating separately for each coordinate: ), and is a residual term with wave-front invariant in the characteristic flow of :
(62) |
(see [DH72, Theorem 6.1.1]).333Choosing a different fundamental solution, say the backwards propagating solution leads to a similar expression. In general, the difference for two solution operators solves , and therefore is invariant along the bicharacteristics. The creation term is , whereas the residual term is given by . For proving that , and hence the expression (61) is well defined and compatible with [GU93, Proposition 2.1] it suffices to show that the characteristic set of is transversal to and each bicharacteristic , intersects finite number of times. These properties follow for small by combining Lemma 3.5 with Lemma 3.6 and hence the referred proposition applies and yields the expression for as a vector of Lagrangian distributions associated with the pair . In particular, microlocally away from the term is a Lagrangian distribution over .
Finally we show that a propagating singularity in is created at a point where the principal symbol of is non-vanishing. As microlocally near , the standard principal symbol formula for compositions of FIOs (for the formula is provided in the proposition referred above) implies that for , sufficiently small and the positively homogeneous principal symbols , , of , , we have
where is an invertible matrix given as the principal symbol of at . This implies that creates at a forwards propagating singularity provided that the principal symbol of does not vanish for at least one . Applying this to , where is as in the assumptions, and using to transform microlocally back into yields the creation of singularity for at , 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 and let be a 444There can be many of such points. nearest point to on the boundary with respect to . Then there is a neighbourhood of in such that
Moreover, if with is a geodesic segment in from to , then the unique geodesic in with satisfies .
Proof.
Fix and as in the assumptions and a small neighbourhood of in . By making the neighbourhood small enough we may assume that every shortest geodesic segment in from to a point in lies in and hits the boundary transversally and for the first time at the endpoint. Fix arbitrary . Let be an unit speed geodesic in from to such that . Denote . Let be the unique geodesic in such that . Since on , also the geodesic has unit speed. Define
where is as in the assumptions. The space is a trivial bundle over , consisting of distinct loops on the surface as fibers. In some coordinates the loops are just circles. For each there is the antipodal point in the circle which satisfies
(63) |
The construction is illustrated in Figure 6.
Let be the unique unit-speed geodesic in through that is tangent to the mirror image of over the boundary at that point. Using the fact that is two dimensional as a manifold we fix such that the lines and avoid the time-like geodesics
(64) |
inside for a large closed Euclidean ball with radius . By assumptions, contains the singular supports of the sources. We would like to apply Proposition 3.1 to systems associated with the parameters and . However, it is possible that the rays and do not break the velocity barriers and in the required way. To get around this we perturb the velocity slightly and then apply Sard’s theorem. If one makes the parameter slightly larger while keeping everything else fixed, the rays still avoid the geodesics (64) in but the intersection (63) splits into two lines instead of one (See Figure 7).
The splitting, however, happens continuously with respect to . Thus, for every conic neighbourhood of in there is such that
That is; for fixed , and associated , we may add a small positive perturbation to without changing the intersection significantly. By applying Sard’s theorem to the functions and one constructs a sequence such that for every and both of the following conditions hold:
This is equivalent to saying that and in the associated points and hence the phase velocity barrier is broken in the required way for the approximating rays. We may assume that lies in for all by excluding a finite number of terms in the beginning of the sequence.
Let us fix , , such that
Then, and the associated unit speed geodesics in approximate the segment . The segment is then close to a shortest curve from to the boundary and therefore small perturbations of it hit the boundary transversally near the point . Consequently, we may assume that there is a sequence of positive real numbers such that and . Let us denote by the geodesics in with the property . These geodesics have unit speed and they approximate .
We now apply Proposition 3.1 for using and , as fixed parameters. The point is that there must be a singularity in that propagates along a bicharacteristic of the form (50) emanating from , where . Moreover, since by assumptions, the propagating singularity for large can not hit the source and get annihilated before hitting as otherwise it would contradict the definition of . Let us denote . By putting together the proposition and the construction above, we conclude
where and is as in the assumptions. Denote . The covector , has a non-zero correspondent in
As the data for both metrics coincide, the set above equals
(65) |
Applying Proposition 3.1, this time with respect to the metric , yields that
where refers to the forward flowout canonical relation (51) for . Hence the projection of in (65) lifts into two possible covectors in the wave front set , both lying in a curve of the form (50) emanating from . These two covectors are and where, analogously to , the stands for the unit-speed geodesic in through with velocity tangent to the mirror image of over at that point. Thus at least one of these covectors lies in a bicharacteristic (50) (for ) that emanates from . In particular, one of the curves and intersects . These curves approximate the geodesics and arbitrarily well inside the bounded set so the intersection for large can take place only at the point due to the way the parameters were fixed. Thus, we deduce that . In particular, each is a geodesic from to in with length . Consequently, we have which implies
at the limit . Moreover, as converges to we conclude that connects to in 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
for every . Moreover, a boundary point in is the nearest to in if and only if it is a nearest point in and for any such point there exists a neighbourhood of in the space such that
Proof.
As a direct consequence of the lemma above we get that
(66) |
By exchanging the roles of metrics and repeating the proof one derives the opposite inequality:
Thus,
Substituting the identity above in (66) one deduces that
which implies that is a nearest boundary point also in . Similarly one shows that a nearest boundary point in is nearest also in . In view of Lemma 4.1 there is a neighbourhood of such that
Again, we may swap the metrics and repeat the argument to obtain the reversed inequality for a sufficiently small neighbourhood . In conlusion,
∎
We finish the section by deriving the main result from the proposition above:
Proof of Theorem 1.1.
It suffices to show that at arbitrary . Let be a nearest boundary point to . Let be a neighbourhood of in as in the proposition above. As shown in [KKL01, Lemma 2.15], the distance function on a sufficiently small is smooth and the derivative defines a diffeomorphism from to an open set in .555We unfortunately have the opposite roles for and compared to the notation in the reference. The derivative points towards along an optimal geodesic segment between and . It follows from Proposition 4.1 that on . Thus, for every in the open cone we have
(67) |
Fix basis vectors for so close to each other that holds for every . The identity (67) then implies
In conclusion, for every and in we have
that is, . ∎
Appendix A Supplemental Material for Section 3
Proof of Lemma 3.1.
Let us prove the case . The proof for the other sign is similar. By assumptions, , where
for some (we may as well proceed with general ) where is positively homogeneous of degree with respect to . Let us formally solve
by substituting an ansatz in the asymptotic form
(68) | |||
(69) |
where and are positively homogeneous of degree . We apply the standard formula
(70) |
(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, , whereas for the rest of the terms one deduces the following conditions:
(71) |
By extracting the terms corresponding to and we obtain
(72) | |||
(73) |
for and Fixing in the equation yields
which solves recursively from the terms of degree . Analogously, one derives by setting in (72) which also removes the terms with . However, the elements obtained in this way have singularities at . To get around this one redefines the terms by multiplying with the cut-off , where such that in and outside . The cut-off does not change the elements in and hence the claim follows for given by . ∎
Proof of Lemma 3.3.
We begin by introducing an useful notation. A composition of two operators and with symbol matrices and admits as a symbol the matrix , given by
Further, if and have classical asymptotic developments and for homogeneous of degree , we have the classical asymptotics
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 where the entry stands for the symbol of . Each entry admits an asymptotic expansion , where is positively homogeneous of degree in and we define . Without loss of generality we may fix such that is invertible in the conic set
(74) |
As the matrix is positively homogeneous of degree in , the inverse matrix on is positively homogeneous of degree . By proceeding as in the proof of [GS94, Theorem 4.1] we construct a suitable matrix where each defines the symbol of , . . Let be a cut-off function such that and . Set . We define the term by
where
(75) |
Then is positively homogeneous of degree and equals on . The leading terms in the asymptotic expansions and are the compositions and which in reduce to the identity matrix by definition. Thus, we can find such that for ,
(76) | ||||
(77) |
where stands for the identity matrix. Define
(78) | |||
(79) |
In we deduce modulo by substituting the identities above. Consequently,
where
Further,
which implies . In conclusion, we may define the terms with lower degree of homogeneity by
(i.e. ) to obtain the result. ∎
Proof of Lemma 3.4.
Let us show that the bicharacteristics are of the form (50). Denote , . By definition,
Taking a derivative of yields . That is, . Hence,
Thus, a bicharacteristic curve satisfies
(80) | |||
(81) | |||
(82) | |||
(83) |
An initial value at fixes the solution uniquely. The equations (80), (81) and (83) are solved by
where the initial value is obtained by setting . Let us show that this solves also (82). It is suffices to show that
(84) |
We compute
As is a geodesic, one derives
where , , are the Christoffel symbols of . Substituting this in the previous equation yields (84). ∎
Proof of Lemma 3.5.
Since the Hamiltonian vector field corresponding to is simply , it follows immediately that the bicharacteristics of (i.e. the characteristic integral curves of the Hamiltonian vector field) equal , for . By Lemma 3.4 it suffices to show that the curve is an integral curve of the Hamiltonian . For we compute (See [Duistermaat, Theorem 3.5.2.])
This proves the claim. ∎
Derivation of (48).
Recall that in a small conic neighbourhood of and that the operator wave-fronts of and are the graphs and , respectively. Hence,
Define a conic neighbourhood by . Then,
Hence,
∎
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.