1. Introduction
For some , let  be disjoint closed convex subsets of Euclidean 2-space , with each boundary  a  strictly convex Jordan curve. Let  be contained in the interior of the bounded component B of , where  is also a strictly convex Jordan curve.
By a 
geodesic or, equivalently, a 
ray in the closure 
M of 
, we mean a piecewise-affine constant-speed curve 
 whose junctions are points of reflection on 
 with equal angles of incidence and reflection. The restriction of 
 to an interval is also called a geodesic. Given 
, the set of all geodesics 
 with 
 is denoted by 
. Then, 
 is critical for the 
length functional
      over constant-speed piecewise-
 curves 
 satisfying 
. We write
      
	  Then, by Theorem 1.1 of [
1], 
K is uniquely determined by its 
travelling-time data
	  namely by the travelling times of geodesics joining pairs of given points on 
C. Similar results are proved in [
2] for obstacles in 
 where 
. Unfortunately, the proof of Theorem 1.1 in [
1] is not constructive: all that is shown is that 
 is different for different convex obstacles 
K. When 
, it is straightforward to calculate 
K from 
, and with a little more effort 
K can also be reconstructed when 
 (see 
Section 4 in [
3]). More interestingly, Theorem 1.1 of [
4] allows the area of 
K to be computed from 
. Indeed, ref. [
4] applies in a much more general setting, where the obstacles are not necessarily convex, and 
 is replaced by a Riemannian manifold of any finite dimension. Importantly, the application of the result in [
4] is made possible by the fact (proved in [
1]) that the set of points generating trapped trajectories in the exterior of obstacles 
K considered in this paper has a Lebesgue measure of zero. Constructing 
K is equivalent to constructing 
, but this seems difficult for 
. We refer to [
5,
6] for general definitions and information about geodesic (billiard) flows on Riemannian manifolds. The present paper shows how to construct 
 from 
 when no line meets more than two connected components of 
K. Equivalently, 
K is required to satisfy Ikawa’s no-eclipse condition [
7].
Inverse problems concerning metric rigidity have been studied for a long time in Riemannian geometry: we refer to [
8,
9,
10,
11] for more information. In the last 20 years or so similar problems have been considered for scattering by obstacles, where the task is to recover geometric information about an obstacle from its scattering length spectrum from travelling times of scattering rays in its exterior [
3].
In general, an obstacle in Euclidean space 
 (
) is a compact subset 
K of 
 with a smooth (e.g., 
) boundary 
 such that 
 is connected. The scattering rays in 
 are generalized geodesics (in the sense of Melrose and Sjöstrand [
12,
13,
14]) that are unbounded in both directions. Most of these scattering rays are billiard trajectories with finitely many reflection points at 
 (there are no reflections at 
C). When 
K is a finite disjoint union of strictly convex domains, then all scattering rays in 
 are billiard trajectories, namely geodesics of the type described above. We refer to [
15,
16,
17,
18] for general information about scattering theory and, in particular, for scattering by obstacles in Euclidean spaces.
It turns out that some kinds of obstacles are uniquely recoverable from their travelling-times spectra. For example, as mentioned above, this was proved in [
2] for obstacles 
K in 
 (
) that are finite disjoint unions of strictly convex bodies with 
 boundaries. The case 
 requires a different proof, given recently in [
1].
The set of the so-called trapped points (points that generate trajectories with infinitely many reflections) plays a rather important role in various inverse problems in scattering by obstacles, and also in problems on metric rigidity in Riemannian geometry. As an example shown by M. Livshits (see e.g., Figure 1 in [
19] or [
4]), in general, the set of trapped points may contain a non-trivial open set. In such a case, the obstacle cannot be recovered from travelling times, because of an argument given in [
4] due to Livshits based on the reflection properties of planar ellipses. In dimensions 
 examples similar to that of Livshits are given in [
19]. Other situations where geometric information cannot be recovered are studied in [
20,
21].
The layout of this paper is as folllows.
In 
Section 2, we collect some simple observations about linear (non-reflected) geodesics. This leads to the construction of 
, so-called 
vacuous arcs, 
 in 
, and then 
 initial arcs in 
. Our plan is to build on the initial arcs, using travelling-time data from reflected rays to construct 
incremental arcs in 
, until eventually the whole of 
 is found. (Note however that, unlike the initial arcs, there are countably many incremental arcs, yielding diminishing additional information from ever-increasing amounts of precisely known data. In practice, insufficient data and limited computing power makes it difficult to carry out more than a few inductive steps, and 
 is found only approximately.) In 
Section 6 we describe an 
inductive step for constructing incremental arcs from previously determined arcs and from observations of 
. To make the relevant observations, we need to understand some of the mathematical structure of 
.
The first step towards this understanding is made in 
Section 3, where some simple facts about (typically non-reflected) geodesics are recalled. These facts, including a known result for computing initial directions of geodesics, are applied in 
Section 4 to investigate the structure of travelling-time data of nowhere-tangent geodesics. In particular, cusps in so-called 
echographs of 
 correspond to geodesics that are tangent to 
.
The family of all such cusps is studied in 
Section 5, where the augmented travelling-time data 
 are shown to be the closure of a countable family of disjoint open 
 arcs 
. As described in 
Section 6, the property of 
extendibility can be checked for each 
. When 
 is extendible, it yields an incremental arc in 
. When 
 is not extendible, a trick using no-eclipse replaces 
 by an extendible 
 yielding an incremental arc as previously described.
Although our methods are elementary, the construction is intricate, requiring numerous steps and different notations. A glossary of terms and notation is given as 
Table 1 at the end of this paper.
  2. Linear Geodesics and Vacuous Arcs
From now on let  be an obstacle in , where  are disjoint closed convex subsets of  with boundaries that are  strictly convex Jordan curves. As before, assume that K is contained in the interior of the bounded component B of , where  is also a strictly convex Jordan curve. We also assume that K satisfies no-eclipse. For the following lemma, it is not enough to require that a geodesic may not be tangent to  at more than two points. Nor is it sufficient to assume that a line in  may not be tangent to  at more than 2 points. However, no-eclipse implies both these conditions, as is easily seen.
Lemma 1. Geodesics in Γ are not tangent to , except perhaps at the first or last points of contact with  (either or both).
 Proof.  If tangency was at an intermediate point of contact, the tangent line would have common points with at least 3 connected components of K, contradicting no-eclipse.    □
 We begin by investigating travelling-times of 
linear geodesics, namely geodesics in 
 that do not reflect at all. The travelling-time data from linear geodesics can be obtained as follows:
	  Excluding points of the form 
, we obtain
      
      where 
 is defined as the travelling-time data from linear geodesics meeting 
 exactly 
q times tangentially and nowhere else. By Lemma 1, 
 for 
; namely  
. In the simplest case where 
, 
 is empty, and 
 is constructed as the envelope of the smooth family 
 of line segments, where 
. Suppose 
 from now on.
Proposition 1.  is a union of  nonintersecting bounded open  arcs  whose boundaries in  comprise , which is finite of size .
 Proof.  For  there are 8 directed Euclidean line segments (linear bitangents) tangent to both  and . Each directed linear bitangent is an endpoint of two maximal open arcs of directed line segments that are singly tangent. The travelling-time data for the linear bitangents are . The travelling-time data for the open arcs  are the path components of .    □
 Therefore, n is found from .
Definition 1. For  the conjugate  is defined to be  or  according as  or .
 Evidently . Order the arcs  in  so that . The initial arcs in  are the  nonempty disjoint connected open subsets of  found as the envelopes of the , where  for . For each  there are  initial arcs in .
  3. Nonlinear Geodesics
Define  to be the space of geodesics  that are tangent exactly q times to . The corresponding travelling-time data are denoted by . Most nonlinear geodesics are nowhere-tangent; namely, they lie in . Their travelling-time data will be used to construct  travelling-time functions  for Lemma 3, as needed for Propositions 3, 4 of §4.
Whereas  is found by simple inspection of , some effort is required to isolate , which is needed to construct envelopes of nonlinear singly tangential geodesics. We recall some known results about directions of geodesics and travelling times.
For  let  be the geodesic satisfying  and . The endpoint map  is the continuous function given by . Then,  is the length of the restriction . Although  is continuous, it is not differentiable at points , where  is tangent to .
Lemma 2. For , suppose that  is nowhere tangent to , and that . Then  is smooth near , and the restriction of its derivative  to  is a linear isomorphism.
 Proof.  For variable perturbations 
 of 
 with 
 sufficiently small, the 
 are also nowhere-tangent to 
. Here we identify 
 with 
ℂ in the standard way. Considering the effects of repeated reflections of the 
, we find by routine calculation that the regular 
 parameterised curve
        
 meets the geodesic 
 transversally at 
; namely, 
 is not a multiple of 
. For variable perturbations 
 of 
 where 
 is small, the curve
        
 is a geodesic. Indeed, a reparameterisation of part of 
 with velocity 
 at 
. Therefore, 
 and 
 are linearly independent.    □
 Lemma 3. For  and with the hypotheses of Lemma 2, there exists an open neighbourhood  of  in , and a unique  function  whose gradient  is everywhere of unit length, satisfyingfor all . Here  with .  Proof.  By Lemma 2 and the implicit function theorem, there exist a unique 
 function 
 satisfying 
 for all 
. Because 
, 
X is never-zero for 
 sufficiently small. Then the geodesic 
 joining 
, has length
        
		Differentiating with respect to 
 in the direction of 
, we find 
, because geodesics are critical for 
J when variations have 
fixed endpoints.    □
 The 
order  of a geodesic 
 is the number of intersections with 
. Write
      
Let 
 be the minimum distance between obstacles. Writing 
 for the travelling time (length) of 
,
      
  4. Arcs and Generators for Nowhere-Tangent Geodesics
Recall that 
 is the space of geodesics that intersect 
 precisely 
r times, and that 
 is the space of geodesics 
 that are exactly 
q-times 
tangent to 
. Because of no-eclipse and Lemma 1, 
 is empty for 
. Set
      
	  Recall that 
 is the travelling-time data from geodesics meeting 
 exactly 
q times tangentially, and nowhere else. Then, 
 and 
 for 
. We have shown how to construct 
, but not yet 
 from 
 for 
.
For 
 define
      
	  Likewise 
 is the set of geodesics 
 with 
. Set
      
 Although 
 is found directly from the given travelling-time data 
, we have yet to show how 
 is found from 
 for 
 (this is done in the paragraph following Corollary 1 below).
It is easily seen that  is open and dense in , that  is open and dense in , and that  is discrete.
Proposition 2.  has at most  elements, and is empty for  in an open dense subset of C.
 Proof.  By Lemma 1, the first and last segments of any  are tangent to . Because  is strictly convex, there are at most  such geodesics. Therefore,  has a size at most of . A small perturbation of  causes a  perturbation of the last points of tangency of , destroying the first points of tangency, by strict convexity.    □
 Proposition 3. For , we have , where
- 1.
- the  are countably many pairwise-transversal  open bounded arcs in , 
- 2.
- , 
- 3.
- each  has a generator, namely a  function  with  open, such that -  is an open arc in C, and  is a diffeomorphism from  onto , 
- , where  for . 
 
 Proof.  For , there exists  with . Then, , where . By Lemma 3, for some open neighbourhood  of  in , there is a unique  function  with  and , such that  for all . In particular, the last equation holds for ; namely,  embeds  in . By continuation, the embedding extends uniquely in both directions around C, until just before  is tangent to some , which must eventually happen. Therefore,  is a countable union of  embedded arcs .
Pairwise transversality is proved by contradiction as follows. Suppose  meet tangentially at . Then  where , and  is tangent to C at . By Lemma 3,  and  point out from the bounded component B of  at . Therefore,  by Lemma 3, contradicting .
Because  is open and dense in , .    □
 By continuity, the orders 
 of the 
 are independent of 
. From (
1) we obtain, for all 
,
      
      and the arcs 
 are similarly bounded. For any 
i, the closures 
 and 
 in 
 are disjoint for all but finitely many 
, where 
. The generator 
 defines 
 for every 
. For 
, define
      
Then 
, and 
 where 
.
Proposition 4. If  then  for some unique  with . Then  and  are on the same side of  in C and, for , . We also have  Proof.  Write , where  has length t and . Suppose the last (respectively first) segment of  is not tangent to . Then, by no-eclipse, the first (last) segment is tangent. Perturbing the last segment while maintaining the endpoint  gives two arcs of nowhere-tangent geodesics, whose initial points  lie on the same side of  in C. Along one arc, the first (last) segment remains linear and the order decreases by 1. Along the other arc, the first (last) segment breaks into two linear segments, maintaining the order and increasing the travelling-time.
Write . For  near , the two arcs of geodesics define arcs ,  in  contained in maximal arcs , labelled so that . Then , and . We also have , and  .    □
 Because  is dense in , Propositions 3 and 4 have the
Corollary 1.  is the closure of a union  of locally finite pairwise-transverse  open bounded arcs, whose endpoints are cusps at points in .
 A  embedding  of  in  is given by , with  some constant-length nonzero outward-pointing normal field. Cusps in  are found by inspecting the echograph at , defined as . The open arcs joining cusps are the  for the  of Proposition 3. By Proposition 2, for  in an open dense subset of C, the cusps are at points in .
Example 1. Figure 1 displays the part of  corresponding to travelling-time data for 0 or 1 reflections by  obstacles, with  and C the circle of radius 4 and centre . For  starting at  and moving south-westerly along C, the blue line from  to  doesn’t intersect  at all; namely, it is a linear geodesic whose travelling-time data are in . As  moves downwards towards the endpoint of the dashed tangent, the echograph (shown outside B) traces out a blue curve from  to the endpoint . The other blue parts of the echograph also correspond to travelling-time data from linear geodesics (0 reflections). The green parts of the echograph in Figure 1 correspond to travelling-time data of geodesics for 1 reflection, such as the green geodesic starting north-westerly from  reflecting once on  to . More of the echograph is shown in Figure 2, where parts corresponding to travelling-time data for 2 reflections (olive) and 3 reflections (red) are also incorporated. For instance, the twice-reflecting red geodesic tangent to  defines the cusp  on the echograph. The other cusps also correspond to tangent geodesics, as indicated. The echograph is mainly smooth, but different smooth arcs (blue, green, olive and red) meet in cusps, and 6 transversal self-intersections are seen. The smooth arcs in Figure 2 are some of the  where the  are the open intervals of Proposition 3 (in no particular order). Cusps (labelled ) correspond to tangencies of geodesics ending at  to  or .  Our construction of K uses each cusp in an echograph to determine a pair of lines, one of which is tangent to . By Lemma 1, there is a tangent line which is the extension of either the initial segment of the geodesic from , or the terminal segment to . The pair of lines is determined from  using the echograph (additional work is carried out later to determine which of these is tangent). Then 1-parameter families of tangent lines are constructed by varying the echograph which depends on . Finally, the  are found as unions of envelopes of families of tangents.
Next we augment  and  to data sets  and  that include the initial velocities of geodesics. We first exclude points of intersection of the open arcs  in Proposition 3 (these points are reinserted later) by defining .
Remark 1. Any  intersects at most finitely many . Because intersections of  and  are transversal for ,  is dense in , and  Remark 2. The  partition . The generators  restrict to  functions on the open subsets of C.
 For 
, define 
 to be the unit vector 
 pointing inwards from 
C. Then set
      
To reinsert the excluded points, define 
 to be the closure of 
 in
      
      and 
.   Define 
 to be the closure of 
 in
      
      and 
. For 
 define
      
  5. Singly-Tangent Geodesics
Summarising so far, for :
-  is read directly from ; 
-  (respectively ) is the non-smooth (respectively smooth) part of ; 
- we have seen how to find arcs  and generators  for ; 
-  and  are obtained using the ; 
-  and  are found by varying . 
Proposition 5, below, is a structural result, analogous to Proposition 3, which will be used to distinguish  from . A geodesic  is said to be bitangent when it has two points of tangency to . Recall that  is linear when it has no other points of contact with .
Proposition 5. For a countable locally finite family  of disjoint bounded open  arcs in :
- 1.
- ; 
- 2.
- for ; , where the  are the vacuous arcs in , defined in Section 2; 
- 3.
- for every  (Including possibly ) there is a diffeomorphism  where  is an open arc in C, and  for all ; 
- 4.
- each  is an endpoint of four open arcs , where three of  are on one side of , and one is on the other side; 
- 5.
- . 
 Proof.  For , we have  and . Now,  is tangent to  at precisely one point. By Lemma 1, this is either the first or last point of contact with .
If the tangency is first, then perturbing the point of tangency in  gives a small open  arc around  contained in . Similarly, if the tangency is last, an open  arc in  is given by perturbing the point of tangency in . Therefore, the path components  of  in  are connected smooth 1-dimensional submanifolds of . They are bounded, nonclosed and, for , can be listed as augmentations of the . Thus, 1. and 2. hold.
For , the geodesic  is tangent to  at both the first and last points of contact, and nowhere else. Nearby geodesics in  are obtained by maintaining tangency either at a variable first point of contact, or at a variable last point of contact with . The tangencies at first (respectively last) points of contact generate arcs  (respectively ) in , separated by .
When the bitangent geodesic 
 is linear, there is an open arc 
 of initial points of perturbations initially tangent to 
, and another open arc 
 of initial points of perturbations initially tangent to 
, as in 
Figure 3, where 
 appear on the right of the illustration. Perturbations whose initial points are in 
 (green) and 
 (red) have no other points of contact with 
. There are also two unlabelled open arcs 
 bordered by 
, consisting of initial points of geodesics whose first points of contact are nontangent to 
, and whose second points of contact are tangent to 
 (green) or 
 (red), respectively. Similarly, the green and red arrows on the left of 
Figure 3 and 
Figure 4 indicate intervals of terminal points of perturbations.
Evidently, 
, because 
 and 
 are on opposite sides of 
, and similarly, 
 in 
Figure 3. Indeed, from the geometry of perturbations of 
, all of 
 are distinct.
In 
Figure 4, the nonlinear bitangent geodesic 
 is tangent to 
 and 
 at the first and last points of contact, respectively. It is not tangent anywhere else to 
, but is reflected at other points of contact, as suggested by the illustration. As before, the nonlinear bitangent is perturbed while maintaining tangency either with 
 (green) or with 
 (red), but now the first and last points of contact remain on 
 and 
, respectively. The initial points of perturbations tangent to 
 sweep out open arcs 
 (green) on either side of 
. Initial points of perturbations tangent to 
 give the other intervals 
 on one side of 
, as indicated by the two red arrows on the left of 
Figure 4. Again, 
 are distinct.
An element  of  corresponds precisely to the point of tangency (first or last contact) of  with . Because there is only one point of tangency it corresponds diffeomorphically to . This proves 3.
Thus, that  is an endpoint of precisely 4 open arcs and 4. are proved.
Because  is open in , . Because  is dense in , , proving 5.    □
 Corollary 2.  is the smooth part of the 1-dimensional space .
 Remark 3. For  the  for  depend only on . Therefore, we may write them as .
 We need the following definitions:
- The open arcs  where  are said to be vacuous; 
- For  denote the undirected line through  parallel to  by ; 
- For  define  by  where ; 
- Denote the envelope of  by . 
  6. Extendible Arcs and the Inductive Step
At the end of 
Section 2 the travelling-time data 
 are used to find 
 open arcs 
. Each of these is augmented, as described in Proposition 5, to a vacuous open arc 
. From the definition in 
Section 2 of the conjugate 
 of 
j, for 
,
      
	  We also obtain 
 parameterisations 
. More generally (inductively), suppose we have this kind of information where possibly 
.
In precise terms, suppose we are given a  parameterisation  of some possibly nonvacuous arc . Here,  is a maximal open arc with the property that, for all  and , the first segment of the geodesic  is tangent to . The inductive step extends the open arc  by adjoining another such arc to its clockwise endpoint, as follows.
For 
, the clockwise terminal limit of 
, set
      
	  By Proposition 5, there are three other open arcs 
 adjacent to 
 at 
, and the unordered set 
 is found by inspecting 
. In the proof of Proposition 5, the arcs 
 (respectively 
) are generated by geodesics whose first (respectively last) segments are tangent to 
. From the proof of Proposition 5, 
 has size 1 or 3. Construct
      
      where 
.
Definition 2.  is an extension of  when the closure  of  is a  strictly convex arc in . When an extension of  exists, the arc  is said to be extendible (otherwise nonextendible).
 Proposition 6. If  is extendible, the extension  is unique, and  is an arc in . If  is nonextendible, then  is vacuous and  is extendible.
 Proof.  By continuity of , the bitangent  is tangent to  at some  where , and q is a limit of points of first tangency and first contact with . By no-eclipse, q is either the first point of contact of the bitangent with  or the second point of contact.    □
 If q is the first point of contact, then  is extended by , whose associated geodesics maintain tangency to . Evidently,  is an arc in .
For 
 or 
, and 
 near 
, the last points of contact of 
 are tangent to 
 near 
 where 
. By the argument in §3 of [
1], the 
 are not all tangent to a 
 strictly convex arc; namely, 
 is not strictly convex, and 
 does not extend 
. Therefore, the extension 
 is unique.
If, alternatively, q is the second point of contact, then the first point of tangency is at  where  with . By Lemma 1,  is the first point of contact of the bitangent with . By Lemma 1, and because q is the second point of tangency, the bitangent is linear with , the only points of contact with . Therefore, , and  is the first point of contact of the linear bitangent  with . Then  is extended by requiring tangency to  of the associated geodesics.
The arc  in  is therefore extended by an incremental arc , where  is an extension either of  or of . This completes the inductive step.
Now the construction of  proceeds as follows. First,  is chosen with , and the inductive step is carried out repeatedly with  replacing  after each step, until the incremental arcs  in  are acceptably small. Countably many repetitions would be needed for perfect reconstruction. Then another vacuous arc is used to restart the iterative process. This is repeated until all the vacuous arcs are used. Finally,  is the union of the closures of all the arcs (initial and incremental) in .
  7. Conclusions
For unknown disjoint obstacles 
 contained in a compact 
m-manifold manifold with boundary 
B in Euclidean 
m-space 
, rays in 
B from 
 to itself are reflected in the usual way by 
 (but not by 
C). The problem is to find information about the obstacles from measurements of travelling times of reflected rays (geodesics). Even when 
B and the 
 are smoothly embedded 
m-dimensional balls, the travelling times sometimes do not determine the obstacles, as seen in the 2-dimensional examples of Livshits and those in [
19] where 
. These cases are excluded by requiring 
B and 
 to be smooth, strictly convex embeddings of the 
m-dimensional unit ball. Then the 
 are uniquely determined by the travelling times [
2] for 
. Unfortunately, the proof in  [
2] is not constructive and gives no clue about how the 
 might be found. This remains problematic, at least for 
, even when 
.
This most elementary case is studied in the present paper, assuming the no-eclipse hypothesis given in 
Section 1. Then, the number 
n of obstacles is found from travelling times using Proposition 1. We go on to reconstruct 
, as follows.
First, Lemma 3 uses travelling-time data of nowhere-tangent geodesics to construct 
 functions 
,  from which initial directions of geodesics at points on some subintervals of 
C can be calculated. Then 
C is the union of the closures of maximal subintervals. The maximal subintervals are found from the travelling-time functions 
, using Proposition 3. Corollary 1 identifies endpoints of maximal subintervals by finding cusps in sets 
 found from travelling times of geodesics that end on points 
. In 
Section 4, these sets and the cusps are displayed in echographs.
Cusps in echographs correspond to geodesics ending at 
 that are singly tangent to the 
. In 
Section 5, we keep track of singly tangent geodesics as 
 varies. Proposition 5 finds that the singly tangent geodesics define maximal open intervals in another computable set. The endpoints of the subintervals correspond to geodesics that are twice tangent to 
.
This is all put together in 
Section 6, where, using no-eclipse, Proposition 6 shows how to find countably many families of lines that are singly tangent to the 
. Then 
 is determined as the closure of the union of the envelopes of these countably many families. Therefore, in theory, the 
 are completely determined from travelling-times.
This construction is a complicated recipe that would be difficult to implement. In practice, we find only finitely many of the countable families of tangent lines, focusing first on families where lines correspond to geodesics that are once, twice, or maybe three times reflected. Even this requires extremely accurate measurements of travelling times, very large amounts of data and substantial computational effort to detect smooth families of cusps in echographs, then to accurately calculate envelopes. Then, large parts of the  are found, but not all. The scope of the present paper is necessarily limited: our ambient space  is flat with , and the  are smooth, strictly convex and satisfy no-eclipse.