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

Uniqueness of the partial travel time representation of a compact Riemannian manifold with strictly convex boundary

Ella Pavlechko Department of Mathematics, North Carolina State University, Raleigh, NC 27695, USA ([email protected])  and  Teemu Saksala Department of Mathematics, North Carolina State University, Raleigh, NC 27695, USA ([email protected])
Abstract.

In this paper a compact Riemannian manifold with strictly convex boundary is reconstructed from its partial travel time data. This data assumes that an open measurement region on the boundary is given, and that for every point in the manifold, the respective distance function to the points on the measurement region is known. This geometric inverse problem has many connections to seismology, in particular to microseismicity. The reconstruction is based on embedding the manifold in a function space. This requires the differentiation of the distance functions. Therefore this paper also studies some global regularity properties of the distance function on a compact Riemannian manifold with strictly convex boundary.

Key words and phrases:
Inverse problem, Riemannian geometry, Distance function, Geodesics
2020 Mathematics Subject Classification:
53C21, 53C24, 53C80, 86A22

1. Introduction

This paper is devoted to an inverse problem for smooth compact Riemannian manifolds with smooth boundaries. Suppose that there is a dense but unknown set of point sources going off in an unknown Riemannian manifold. For each of these point sources we measure the travel time of the respective wave on some small known open subset of the boundary of the manifold. The point source can be natural (e.g. an earthquake as a source of seismic waves) or artificial (e.g. produced by focusing of waves or by a wave sent in scattering from a point scatterer). We show that this partial travel time data determines the unknown Riemannian manifold up to a Riemannian isometry under some geometric constraints. Namely, we assume that the unknown manifold has a strictly convex boundary. We also provide an example demonstrating that the convexity is a necessary assumption. To prove our result, we embed a Riemannian manifold with boundary into a function space and use smooth boundary distance functions to give a coordinate and the Riemannian structures.

The inverse problem introduced above can be rephrased as the following problem in seismology. Imagine that earthquakes occur at known times but unknown locations within Earth’s interior and arrival times are measured on some small area on the surface. Are such partial travel time measurements sufficient to determine the possibly anisotropic elastic wave speed everywhere in the interior and pinpoint the locations of the earthquakes? While earthquake times are not known in practice, this is a fundamental mathematical problem that underlies more elaborate geophysical scenarios.

An elastic body — e.g. a planet — can be modeled as a manifold, where distance is measured in travel time: The distance between two points is the shortest time it takes for a wave to go from one point to the other. If the material is isotropic or elliptically anisotropic, then this elastic geometry is Riemannian. However this sets a very stringent assumption on the stiffness tensor describing the elastic system, one which is not physically observed in the Earth, and Riemannian geometry is therefore insufficient to describe the propagation of seismic waves in the Earth. If no structural assumptions on the stiffness tensor apart from the physically necessary symmetry and positivity properties are made, this leads necessarily to modeling the planet by a Finsler manifold as was explained in [13].

An isotropically elastic medium carries pressure (P) and shear (S) wave speeds that are conformally Euclidean metrics. Of these two the P-waves are faster [8]. In order to be true to the isotropic elasticity we should measure both P- and S-wave arrivals. In this paper we simplify this aspect of the problem by disregarding polarizations and considering only one type of isotropic waves.

Acknowledgements

EP was funded by the AWM and NSF-DMS grant # 1953892 for travel to the 2022 Joint Mathematics Meetings (JMM), where this paper is to be presented.

1.1. Main theorem and the geometric assumptions

We consider a compact nn-dimen-sional smooth manifold MM with smooth boundary M\partial M, equipped with a smooth Riemannian metric gg. For points p,qMp,q\in M the Riemannian distance between them is denoted by d(p,q)d(p,q). Then for pMp\in M we define the boundary distance function r^p:M𝐑\hat{r}_{p}:\partial M\to{\bf R} given by r^p(z)=d(p,z)\hat{r}_{p}(z)=d(p,z). Let Γ\Gamma be a non-empty open subset of the boundary M\partial M. We denote the restriction of the boundary distance function on this set as rp:=r^p|Γr_{p}:=\hat{r}_{p}\big{|}_{\Gamma}. The collection

(1) Γand{rp:Γ𝐑:rp(z)=d(p,z)},\Gamma\qquad\text{and}\qquad\{r_{p}:\Gamma\to{\bf R}~{}:~{}r_{p}(z)=d(p,z)\},

are called the partial travel time data of ΓM\Gamma\subset\partial M. With these data we seek to recover the Riemannian manifold (M,g)(M,g) up to a Riemannian isometry. The following definition explains when two Riemannian manifolds have the same partial travel time data (1).

Definition 1.

Let (M1,g1)(M_{1},g_{1}) and (M2,g2)(M_{2},g_{2}) be compact, connected, and oriented Riemannian manifolds of dimension n𝐍,n2n\in\mathbf{N},\>n\geq 2 with smooth boundaries M1\partial M_{1} and M2\partial M_{2} and open non-empty regions ΓiMi\Gamma_{i}\subset\partial M_{i} respectively. We say that the partial travel time data of (M1,g1)(M_{1},g_{1}) and (M2,g2)(M_{2},g_{2}) coincide if there exists a diffeomorphism ϕ:Γ1Γ2\phi:\Gamma_{1}\to\Gamma_{2} such that

(2) {rpϕ1:pM1}={rq:qM2}.\{r_{p}\circ\phi^{-1}~{}:~{}p\in M_{1}\}=\{r_{q}~{}:~{}q\in M_{2}\}.

We want to emphasize that the equality (2) is for the non-indexed sets of travel time functions. Thus, for any pM1p\in M_{1} there exists a point qM2q\in M_{2} such that rp(ϕ1(z))=rq(z)r_{p}(\phi^{-1}(z))=r_{q}(z) for every zΓ2z\in\Gamma_{2}. We do not know a priori where the point pM1p\in M_{1} is or if there are several points qM2q\in M_{2} that satisfy this equation.

We use the notations TMTM and SMSM for the tangent and unit sphere bundles of MM. Their respective fibers, for each point pMp\in M, are denoted by TpMT_{p}M and SpMS_{p}M. In order to show that the data (1) determine (M,g)(M,g), up to an isometry or in other words that the Riemannian manifolds (M1,g1)(M_{1},g_{1}) and (M2,g2)(M_{2},g_{2}) of Definition 1 are Riemanian isometric, we need to place an additional geometric restriction. We assume that (M,g)(M,g) has a strictly convex boundary M\partial M which means that the shape operator S:TMTMS\colon T\partial M\to T\partial M as a linear operator on each tangent space TxMT_{x}\partial M of the boundary M\partial M for a point xMx\in\partial M is negative definite. This means that the second fundamental form

Πx(X,Y)=S(x)X,Yg,\Pi_{x}(X,Y)=\langle S(x)X,Y\rangle_{g},

is strictly negative whenever X,YTxMX,Y\in T_{x}\partial M agree, but do not vanish. It has been shown in [5, 3] that the strict convexity of the boundary implies the geodesic convexity of (M,g)(M,g). That is any pair of points p,qMp,q\in M can be connected by a distance minimizing geodesic (not necessarily unique) which is contained in the interior MintM^{int} of MM modulo the terminal points. In particular any geodesic of MM that hits the boundary exits immediately.

The main theorem of this paper is the following:

Theorem 2.

Let (M1,g1)(M_{1},g_{1}) and (M2,g2)(M_{2},g_{2}) be compact, connected, and oriented Riemannian manifolds of dimension n𝐍,n2n\in\mathbf{N},\>n\geq 2 with smooth and strictly convex boundaries M1\partial M_{1} and M2\partial M_{2} and open non-empty measurement regions ΓiMi\Gamma_{i}\subset\partial M_{i} respectively. If the travel time data of (M1,g1)(M_{1},g_{1}) and (M2,g2)(M_{2},g_{2}) coincide, in the sense of Definition 1, then the Riemannian manifolds (M1,g1)(M_{1},g_{1}) and (M2,g2)(M_{2},g_{2}) are Riemannian isometric.

Remark 3.

Our assumptions in Theorem 2 do not prevent the existence of the conjugate points. Actually quite a lot of work in this paper is needed to handle their existence. We also allow the manifolds to have trapped geodesics.

1.2. Outline of the proof of Theorem 2

The main tool of proving Theorem 2 is to differentiate the travel time functions given in equation (1). As these functions are defined only on a small open subset of the boundary we need to develop some regularity theorem for them. For this reason in Section 3 we study the regularity properties of the distance function on Riemannian manifolds satisfying the geometric constraints of Theorem 2. Section 3 has two main results. Theorem 4 is the aforementioned regularity result and the key of the proof of Theorem 2. In order to prove Theorem 4 we need to study, for each point in our manifold, the properties of its cut locus. This is the set past which the geodesics shot from the chosen point are not anymore distance minimizers. Theorem 12 collects the needed properties of these sets. Up to the best of our knowledge the material presented in Section 3 does not exist or is not easily accessible in the literature. Nevertheless, the corresponding results for manifolds without boundaries are well known.

In Section 4 we apply Theorem 4 to reconstruct the Riemannian manifold from its partial travel time data (1). This is done in five parts. Firstly we recover the geometry of the measurement region. As the second step we recover the topological structure by embedding the unknown manifold into a function space. Then we determine the boundary. The fourth step is to find local coordinates. Since our manifold has a boundary, we need different types of local coordinates for the interior and boundary points. Lastly we reconstruct the Riemannian metric. All the steps in Section 4 are fully data driven. Finally, in Section 5 we show that if two Riemannian manifolds, as in Theorem 2, have coinciding partial travel time data, in the sense of the Definition 1, then they are isometric.

1.3. The convexity of the domain in Theorem 2 is necessary

Let us construct an explicit example of a surface MM and a subset ΓM\Gamma\subset\partial M so that our results fail with data recorded only on Γ\Gamma (this example was originally presented in [11]). We recall that every pair of points on a smooth compact Riemannian manifold with boundary is always connected by a C1C^{1}-smooth distance minimizing curve [1]. We choose our a manifold to be the horseshoe-shaped domain of Figure 1. We split the domain MM into two pieces M1M_{1} and M2M_{2} with respect to the line (red dotted line) that is normal to M\partial M at x0Mx_{0}\in\partial M (blue dot). Then we choose a domain ΓM1\Gamma\subset\partial M_{1} (red arch) so that any minimizing curve joining a point on Γ\Gamma and a point in M2M_{2} touches the boundary near x0x_{0}. The curve PMP\subset M is any involute of the boundary, meaning that the distance from all points on PP to x0x_{0} is the same. Because d(z,p)=d(z,q)d(z,p)=d(z,q) for any zΓz\in\Gamma and p,qPp,q\in P, from the point of view of our data (1), the set PP appears to collapse to a point.

Refer to captionPPM2M_{2}x0x_{0}Γ\GammaM1M_{1}
Figure 1. A domain where partial data is insufficient.

2. Some geometric inverse problems arising from seismology

In the following subsections we review some seismologically relevant geometric inverse problems on Riemannian and Finsler manifolds.

2.1. Related geometric inverse problems on Riemannian manifolds

Let (M,g)(M,g) be as in Theorem 2 and denote the Laplace-Beltrami operator of the metric gg by Δg\Delta_{g}. Let (p,t0)Mint×𝐑(p,t_{0})\in M^{int}\times{\bf R}. In this paper we consider an inverse problem related to the wave equation

(3) {(t2Δg)u(x,t)=δp(x)δt0(t),u(x,t)=0,t<t0,xM,\begin{cases}(\partial_{t}^{2}-\Delta_{g})u(x,t)=\delta_{p}(x)\delta_{t_{0}}(t),\\ u(x,t)=0,\qquad t<t_{0},~{}x\in M,\end{cases}

where the solution u(x,t)u(x,t) is a spherical wave produced by an interior point source given by the delta function δp(x)δt0(t)\delta_{p}(x)\delta_{t_{0}}(t) of the space time M×𝐑M\times{\bf R} at (p,t0)(p,t_{0}). We define the arrival time function 𝒯p,t0:M𝐑\mathcal{T}_{p,t_{0}}\colon\partial M\to{\bf R}, of the wave uu by the formula

𝒯p,t0(z)=sup{t𝐑: the point (z,t)M×𝐑 has a neighborhoodUM×𝐑 such that u(,)|U=0}.\begin{split}\mathcal{T}_{p,t_{0}}(z)=\sup\{&t\in{\bf R}:\hbox{ the point $(z,t)\in\partial M\times{\bf R}$ has a neighborhood}\\ &\text{$U\subset M\times{\bf R}$ such that $u(\cdot,\cdot)\big{|}_{U}=0$}\}.\end{split}

We recall that it was shown in [32, Proposition 3.1], by applying the results of [16, 18], that 𝒯p,t0(z)=d(p,z)+t0.\mathcal{T}_{p,t_{0}}(z)=d(p,z)+t_{0}. If the initial time t0t_{0} is zero (or given), the knowledge of the arrival time function 𝒯p,t0\mathcal{T}_{p,t_{0}} on the open set ΓM\Gamma\subset\partial M is equivalent to the corresponding travel time function rp()r_{p}(\cdot) as in (1).

The problem of determining the isometry type of a compact Riemannian manifold from its boundary distance data

{r^p:M(0,);r^p(z)=d(p,x),pMint}\{\hat{r}_{p}\colon\partial M\to(0,\infty);\>\hat{r}_{p}(z)=d(p,x),\>p\in M^{int}\}

was introduced for the first time in [26]. The reconstructions of the smooth atlas on the manifold and the metric tensor in these coordinates was originally considered in [23]. In contrast to the paper at hand, these earlier results do not need any extra assumption for the geometry, but have complete data in the sense that the measurement area Γ\Gamma is the whole boundary. Their counterpart of our Theorem 4, is [23, Lemma 2.15], which says that the boundary distance function r^p\hat{r}_{p} is smooth near its minimizers (the set of closest boundary points to the source point pMp\in M). This lemma is a key of their proof. However, the same technique is not available to us as it requires the access to the whole boundary.

The problem of boundary distance data is related to many other geometric inverse problems. For instance, it is a crucial step in proving uniqueness for Gel’fand’s inverse boundary spectral problem [23]. Gel’fand’s problem concerns the question whether the data

(M,(λj,νϕj|M)j=1)(\partial M,(\lambda_{j},\partial_{\nu}\phi_{j}|_{\partial M})_{j=1}^{\infty})

determine (M,g)(M,g) up to isometry, when (λj,ϕj)(\lambda_{j},\phi_{j}) are the Dirichlet eigenvalues and the corresponding L2L^{2}-orthonormal eigenfunctions of the Laplace–Beltrami operator. Belishev and Kurylev provide an affirmative answer to this problem in [4].

In [24] the authors studied a question of approximating a Riemannian manifold under the assumption: For a finite set of receivers RMR\subset\partial M one can measure the travel times d(p,)|Rd(p,\cdot)|_{R} for finitely many pPMintp\in P\subset M^{int} under the a priori assumption that RMR\subset\partial M is ε\varepsilon-dense and that {d(p,)|R:pP}{d(p,)|R:pMint}\{d(p,\cdot)|_{R}:\>p\in P\}\subset\{d(p,\cdot)|_{R}:\>p\in M^{int}\} is also ε\varepsilon-dense. Thus {d(p,)|R:pP}\{d(p,\cdot)|_{R}:\>p\in P\} is a finite measurement. The authors construct an approximate finite metric space MεM_{\varepsilon} and show that the Gromov-Hausdorff distance of MM and MεM_{\varepsilon} is proportional to some positive power of ε\varepsilon. In [24] an independent travel time measurement is made for each interior source point in PP, whereas in [11] the authors studied the approximate reconstruction of a simple Riemannian manifold (a compact Riemannian manifold with strictly convex boundary where each pair of points is connected by the unique smoothly varying distance minimizing geodesic) by measuring the arrival times of wave fronts produced by several point sources, that go off at unknown times, and moreover, the signals from the different point sources are mixed together. To describe the similarity of two metric spaces ‘with the same boundary’ the authors defined a labeled Gromov-Hausdorff distance. This is an extension of the classical Gromov-Hausdorff distance which compares both the similarity of the metric spaces and the sameness of the boundaries — with a fixed model space for the boundary. In addition to reconstructing a discrete metric space approximation of (M,g)(M,g), the authors in [11] estimated the density of the point sources and established an explicit error bound for the reconstruction in the labeled Gromov-Hausdorff sense.

If we do not know the initial time t0t_{0} in (3), but we can recover the arrival times 𝒯p,t0(z)\mathcal{T}_{p,t_{0}}(z) for each zMz\in\partial M, then taking the difference of the arrival times one obtains a boundary distance difference function

Dp(z1,z2):=d(p,z1)d(p,z2)D_{p}(z_{1},z_{2}):=d(p,z_{1})-d(p,z_{2})

for all z1,z2Mz_{1},z_{2}\in\partial M, which is independent from the initial time t0t_{0}. In [32] it is shown that if UNU\subset N is a compact subset of a closed Riemannian manifold (N,g)(N,g) with a non-empty interior, then distance difference data

((U,g|U),{Dp:U×U𝐑|pN})((U,g|_{U}),\{D_{p}\colon U\times U\to{\bf R}\>|\>p\in N\})

determine (N,g)(N,g) up to an isometry. This result was generalized for complete Riemannian manifolds [22] and for compact Riemannian manifolds with boundary [14, 21]. These results require the full boundary measurement in the sense of Γ=M\Gamma=\partial M, unlike Theorem 2 in the present paper.

If the sign in the definition of the distance difference functions is changed, we arrive at the distance sum functions,

Dp+(z1,z2)=d(z1,p)+d(z2,p)D^{+}_{p}(z_{1},z_{2})=d(z_{1},p)+d(z_{2},p)

for all pMp\in M and z1,z2Mz_{1},z_{2}\in\partial M. These functions give the lengths of the broken geodesics, that is, the union of the shortest geodesics connecting z1z_{1} to pp and the shortest geodesics connecting pp to z2z_{2}. Also, the gradients of Dp+(z1,z2)D^{+}_{p}(z_{1},z_{2}) with respect to z1z_{1} and z2z_{2} give the velocity vectors of these geodesics. The inverse problem of determining the manifold (M,g)(M,g) from the broken geodesic data, consisting of the initial and the final points and directions, and the total length of the broken geodesics, has been considered in [27]. The authors show that broken geodesic data determine the boundary distance data of any compact smooth manifold of dimension three and higher. Finally they use the results of [23, 26] to prove that the broken geodesic data determine the Riemannian manifold up to an isometry. A different variant of broken geodesic data was recently considered in [38].

The Riemannian wave operator is a globally hyperbolic linear partial differential operator of real principal type. Therefore, the Riemannian distance function and the propagation of a singularity initiated by a point source in space time are related to one another. We let uu be the solution of the Riemannian wave equation as in (3). In [17, 18] it is shown that the image, Λ\Lambda, of the wave front set of uu, under the musical isomorphism TM(x,ξ)(x,gij(x)ξi)TMT^{\ast}M\ni(x,\xi)\mapsto(x,g^{ij}(x)\xi_{i})\in TM, coincides with the image of the tangent space  TpMT_{p}M at pMintp\in M^{int} under the geodesic flow of gg. Thus Λ(SM)\Lambda\cap\partial(SM), where SMSM is the unit sphere bundle of (M,g)(M,g), coincides with the exit directions of geodesics emitted from pp. In [33] the authors show that if (M,g)(M,g) is a compact smooth non-trapping Riemannian manifold with smooth strictly convex boundary, then generically the scattering data of point sources (M,RM(M))(\partial M,R_{\partial M}(M)) determine (M,g)(M,g) up to an isometry. Here, RM(p)RM(M)R_{\partial M}(p)\in R_{\partial M}(M) for pMp\in M stands for the collection of tangential components to the boundary of exit directions of geodesics from pp to M\partial M.

A classical geometric inverse problem, that is closely related to the distance functions, asks: Does the Dirichlet-to-Neumann mapping of a Riemannian wave operator determine a Riemannian manifold up to an isometry? For the full boundary data this problem was solved originally in [4] using the Boundary control method. Partial boundary data questions have been studied for instance in [31, 40]. Recently [29] extended these results for connection Laplacians. Lately also inverse problems related to non-linear hyperbolic equations have been studied extensively [28, 34, 51]. For a review of inverse boundary value problems for partial differential equations see [30, 50].

Maybe the most extensively studied geometric inverse problem formulated with the distance functions is the Boundary rigidity problem. This problem asks: Does the boundary distance function, that gives a distance between any two boundary points, determine (M,g)(M,g) up to an isometry? In an affirmative case (M,g)(M,g) is said to be boundary rigid. For a general Riemannian manifold the problem is false: Suppose the manifold contains a domain with very slow wave speed, such that all the geodesics starting and ending at the boundary avoid this domain. Then in this domain one can perturb the metric in such a way that the boundary distance function does not change. It was conjectured in [39] that for all simple Riemannian manifolds the answer is affirmative. In two dimensions this was verified in [44]. For higher dimensional cases the problem is still open, but different variations of it has been considered for instance in [7, 9, 48, 49].

2.2. Related geometric inverse problems on Finsler manifolds

In [13] the authors studied the recovery of a compact Finsler manifold from its boundary distance data. In contrast to earlier Riemannian results [23, 26] the data only determines the topological and smooth structures, but not the global geometry. However the Finsler function F:TM[0,)F\colon TM\to[0,\infty) can be recovered in a closure of the set G(M,F)TMG(M,F)\subset TM, which consists of points (p,v)TM(p,v)\in TM such that the corresponding geodesic γp,v\gamma_{p,v} is distance minimizing to the terminal boundary point. They also showed that if the set TMG(M,F)TM\setminus G(M,F) is non-empty then any small perturbation of FF in this set leads to a Finsler metric whose boundary distance data agrees with the one of FF. If G(M,F)=TMG(M,F)=TM, then the boundary distance data determines (M,F)(M,F) up to a Finsler isometry. For instance the isometry class of any simple Finsler manifold is determined by this data. The same is not true if only the boundary distance function is known [20]. Thus a simple Finsler manifold is never boundary rigid. In [12] the main result of [13] was utilized to generalize the result of [27], about the broken geodesic data, on reversible Finsler manifolds, satisfying a convex foliation condition.

Although, simple Finsler manifolds are not boundary rigid there are results considering their rigidity questions for some special Finsler metrics. For instance it was shown in [41] that Randers metrics Fk=Gk+βkF_{k}=G_{k}+\beta_{k} indexed with k{1,2}k\in\{1,2\} with simple and boundary rigid Riemannian norm Gk(x,v)=gij(x)vivjG_{k}(x,v)=\sqrt{g_{ij}(x)v^{i}v^{j}} and closed one-form βk\beta_{k}, have the same boundary distance function if and only if G1=ΨG2G_{1}=\Psi^{\ast}G_{2} for some boundary fixing diffeomorphism Ψ:MM\Psi\colon M\to M and β1β2=dϕ\beta_{1}-\beta_{2}=\mathrm{d}\phi for some smooth function ϕ\phi vanishing on M\partial M. It is worth mentioning that analogous results have been presented earlier on a Riemannian manifold in the presence of a magnetic field [2, 10].

3. Distance functions on compact manifolds with strictly convex boundary

The aim of this section is to prove the following regularity result for the Riemannian distance function.

Theorem 4.

Let (M,g)(M,g) be a smooth, compact, connected, and oriented Riemannian manifold of dimension n𝐍,n2n\in\mathbf{N},\>n\geq 2 with smooth and strictly convex boundary. For any p0Mp_{0}\in M there exists an open and dense set Wp0MW_{p_{0}}\subset\partial M such that for every z0Wp0z_{0}\in W_{p_{0}} there are neighborhoods Up0MU_{p_{0}}\subset M of p0p_{0} and Vp0MV_{p_{0}}\subset M of z0z_{0} such that the distance function d(,)d(\cdot,\cdot) is smooth in the product set Up0×Vp0U_{p_{0}}\times V_{p_{0}}.

This result is the key of the proof of Theorem 2.

3.1. Critical distances, extensions and the cut locus

In this section we consider a Riemannian manifold (M,g)(M,g) as in Theorem 4, and study the properties of several critical distance functions. We define the exit time function

τexit:SM𝐑{},τexit(p,v)=sup{t>0:γp,v(t)Mint},{\tau_{\text{exit}}}\colon SM\to{\bf R}\cup\{\infty\},\quad{\tau_{\text{exit}}}(p,v)=\sup\{t>0:\gamma_{p,v}(t)\in M^{int}\},

where γp,v\gamma_{p,v} is the geodesic of (M,g)(M,g) with the initial conditions (p,v)SM(p,v)\in SM. Since the boundary of MM is strictly convex, τexit(p,v){\tau_{\text{exit}}}(p,v) is the first time when the geodesic γp,v\gamma_{p,v} hits the boundary, and (τexit(p,v),τexit(p,v))(-{\tau_{\text{exit}}}(p,-v),{\tau_{\text{exit}}}(p,v)) is the maximal interval where the geodesic is defined. We do not assume that τexit(p,v)<{\tau_{\text{exit}}}(p,v)<\infty for all (p,v)SM(p,v)\in SM. That is, (M,g)(M,g) may have trapped geodesics. Here we denote by JSMJ\subset SM the set of all non-trapped directions, that are those (p,v)SM(p,v)\in SM for which τexit(p,v)<{\tau_{\text{exit}}}(p,v)<\infty. It is well known that on compact Riemannian manifolds with strictly convex boundary the set JJ is open in SMSM, the exit time function τexit{\tau_{\text{exit}}} is continuous in JJ, and smooth on JTMJ\setminus T\partial M (See for instance [47, Chapter 4]).

For any pMp\in M we define a star shaped set

(4) Mp:={vTpM:v=0, or vgτexit(p,vvg)}.M_{p}:=\left\{v\in T_{p}M:\>v=0,\>\text{ or }\|v\|_{g}\leq{\tau_{\text{exit}}}\left(p,\frac{v}{\|v\|_{g}}\right)\right\}.

Thus MpM_{p} is the largest subset of TpMT_{p}M where the exponential map of pp

expp:MpM,expp(v)=γp,v(1)\exp_{p}\colon M_{p}\to M,\quad\exp_{p}(v)=\gamma_{p,v}(1)

is defined. Since M\partial M is strictly convex this map is onto, but it does not need to be one-to-one, since there can be several geodesics of the same length connecting pp to some common point. This leads to the following definition of the cut distance function:

(5) τcut:SM𝐑,τcut(p,v)=sup{t(0,τexit(p,v)]:d(p,γp,v(t))=t}.{\tau_{\text{cut}}}\colon SM\to{\bf R},\quad{\tau_{\text{cut}}}(p,v)=\sup\{t\in(0,{\tau_{\text{exit}}}(p,v)]:\>d(p,\gamma_{p,v}(t))=t\}.

Thus the geodesic segment γp,v:[0,t]M\gamma_{p,v}\colon[0,t]\to M is a distance minimizing curve for any t[0,τcut(p,v)]t\in[0,{\tau_{\text{cut}}}(p,v)].

Traditionally on a closed Riemannian manifold (N,g)(N,g) the set

(6) cutN(p):={γp,v(τcut(p,v))N:vSpN}{\text{cut}}_{N}(p):=\{\gamma_{p,v}({\tau_{\text{cut}}}(p,v))\in N:\>v\in S_{p}N\}

is known as the cut locus of the point pNp\in N and each point in this set is called a cut point of pp. Moreover, the cut locus of pp coincides with the closure of the set of those points qNq\in N such that there is more than one distance minimizing geodesic from pp to qq (see for instance [25, Theorem 2.1.14]). It has been also shown in [45, Section 9.1] that d(p,)d(p,\cdot) is smooth in N({p}cutN(p))N\setminus(\{p\}\cup{\text{cut}}_{N}(p)) but not at any q({p}cutN(p))q\in(\{p\}\cup{\text{cut}}_{N}(p)). In order to understand the smoothness properties of the distance function on a Riemannian manifold (M,g)(M,g) with a strictly convex boundary, our aim is to define the set analogous to the one in (6) in this context.

If NN is a closed manifold and (p,v)SN(p,v)\in SN then by Klingenberg’s lemma [36, Proposition 10.32] either there is a second distance minimizing geodesic from pp to γp,v(τcut(p,v))\gamma_{p,v}({\tau_{\text{cut}}}(p,v)) or these points are conjugate to each other along γp,v\gamma_{p,v}. In particular, the geodesic γp,v\gamma_{p,v} is not a distance minimizer beyond the interval [0,τcut(p,v)][0,{\tau_{\text{cut}}}(p,v)]. The following lemma extends this result in our case.

Lemma 5.

Let Riemannian manifold (M,g)(M,g) be as in Theorem 4 and (p,v)SM(p,v)\in SM. If

τcut(p,v)<τexit(p,v){\tau_{\text{cut}}}(p,v)<{\tau_{\text{exit}}}(p,v)

then at least one of the following holds for q:=γp,v(τcut(p,v))q:=\gamma_{p,v}({\tau_{\text{cut}}}(p,v)):

  • There exists another distance minimizing geodesic from pp to qq.

  • qq is the first conjugate point to pp along γp,v\gamma_{p,v}.

Moreover, for any t0(0,τcut(p,v))t_{0}\in(0,{\tau_{\text{cut}}}(p,v)) the geodesic segment γp,v:[0,t0]M\gamma_{p,v}\colon[0,t_{0}]\to M has no conjugate points and is the unique unit-speed distance minimizing curve between its endpoints.

Proof.

Since the exit time function is continuous on the non-trapping part of SMSM and qq is an interior point of MM, the proof is identical to the proof of the analogous claim in [36, Proposition 10.32]. ∎

Since the manifold MM has a non-empty boundary M\partial M it holds that both the tangent bundle TMTM and the unit sphere bundle SMSM are manifolds with boundaries TM\partial TM and SM\partial SM respectively.

(p,v)TM,((p,v)SM) if and only if pM.(p,v)\in\partial TM,\>((p,v)\in\partial SM)\quad\hbox{ if and only if }\quad p\in\partial M.

We equip TMTM with the Sasaki metric gSg_{S}. Thus we can consider TMTM, and its submanifold SMSM, as Riemannian manifolds. In the following the convergence and other metric properties in TMTM or SMSM will be considered with respect to this metric.

Lemma 6.

Let Riemannian manifold (M,g)(M,g) be as in Theorem 4. The cut distance function τcut{\tau_{\text{cut}}} is continuous in SMSM.

Proof.

Since the exit time function is continuous on the non-trapping part of SMSM, the proof of this claim is almost identical to the proof of the analogous claims in [36, Theorem 10.33] and [25, Lemma 2.1.5]. ∎

As MM has a boundary, the definition of the cut time function τcut{\tau_{\text{cut}}}, in the equation (5), has an issue. Namely if τcut(p,v)=τexit(p,v){\tau_{\text{cut}}}(p,v)={\tau_{\text{exit}}}(p,v) for some (p,v)SM(p,v)\in SM we do not know a priori if the geodesic γp,v\gamma_{p,v} just hits the boundary at γp,v(τcut(p,v))\gamma_{p,v}({\tau_{\text{cut}}}(p,v)) or if it is possible to find an extension of (M,g)(M,g) such that γp,v\gamma_{p,v} also extends as a distance minimizer.

To address this question, from here onwards we assume that (M,g)(M,g) has been isometrically embedded in some closed Riemannian manifold (N,g)(N,g). This can be done for instance by constructing the double of the manifold MM as explained in [35, Example 9.32] and extending the metric gg smoothly across the boundary M\partial M. The issue with this extension is that it might create ‘short cuts’ in the sense that there can be a curve in NN, connecting some points of MM, which is shorter than any curve entirely contained in MM. Therefore we always have

dM(p,q)dN(p,q), for all p,qM,d_{M}(p,q)\geq d_{N}(p,q),\quad\hbox{ for all }p,q\in M,

where dM(,)d_{M}(\cdot,\cdot) and dN(,)d_{N}(\cdot,\cdot) are the distance functions of MM and NN respectively. The following proposition shows that while we stay close enough to MM we do not need to worry about these short cuts.

Proposition 7.

Let (N,g)(N,g) be a smooth, connected, orientable, and closed Riemannian manifold and MNM\subset N an open set whose boundary M\partial M is a smooth strictly convex hyper-surface of (N,g)(N,g). There exists an open subset M^\hat{M} of NN, that contains the closure of M,M, and whose boundary is a smooth, strictly convex hyper-surface of NN.

Moreover

(7) dM^(p,q)=dM(p,q), for all p,qM¯.d_{\hat{M}}(p,q)=d_{M}(p,q),\quad\text{ for all }p,q\in\bar{M}.
Proof.

Since M\partial M is a smooth hyper-surface of NN there exists a smooth function s:N𝐑s\colon N\to{\bf R} and a neighborhood UU of M\partial M such that

|s(x)|=dist(x,M):=inf{dN(x,z):zM},andgrads(x)g1,|s(x)|=\text{dist}(x,\partial M):=\inf\{d_{N}(x,z):\>z\in\partial M\},\quad\text{and}\quad\|\text{grad}\;s(x)\|_{g}\equiv 1,

for every xUx\in U. Moreover for each xUx\in U there exists a unique zMz\in\partial M such that dN(x,z)=|s(x)|.d_{N}(x,z)=|s(x)|. We choose the sign convention of ss such that s(x)0s(x)\geq 0 for xUMx\in U\setminus M. Then on M\partial M the gradient of the function s()s(\cdot) agrees with the outward pointing unit normal vector field of M\partial M. The existence of this function is explained for instance in [36, Example 6.43].

By this construction, each pUp\in U can be written uniquely as

p=(z(p),s(p))M×𝐑,p=(z(p),s(p))\in\partial M\times{\bf R},

where z(p)z(p) is the closest point of M\partial M to pp. Thus on UU we write the Riemannian metric as a function of (z,ε)M×𝐑(z,\varepsilon)\in\partial M\times{\bf R} in the form ds2+g~(ε,z),\mathrm{ds}^{2}+\tilde{g}(\varepsilon,z), where g~(ε,z)\tilde{g}(\varepsilon,z) is the first fundamental form of the smooth hyper-surface Ω(ε):=s1{ε}.\Omega(\varepsilon):=s^{-1}\{\varepsilon\}. By [36, Proposition 8.18] we can then write the second fundamental form of Ω(ε)\Omega(\varepsilon) as a bi-linear form

Π(z,ε)(X,Y)=12εg~αβ(ε,z)XαYβ𝐑\Pi_{(z,\varepsilon)}(X,Y)=-\frac{1}{2}\frac{\partial}{\partial\varepsilon}\tilde{g}_{\alpha\beta}(\varepsilon,z)X^{\alpha}Y^{\beta}\in{\bf R}

on TΩ(s)T\Omega(s). Thus the eigenvalues λ1(z,ε),,λn1(z,ε)\lambda_{1}(z,\varepsilon),\ldots,\lambda_{n-1}(z,\varepsilon) of Π(z,ε)\Pi_{(z,\varepsilon)} are continuous functions of (z,ε)(z,\varepsilon) [52, Appendix V, Section 4, Theorem 4A]. Since Ω(0)\Omega(0) coincides with M\partial M, which is strictly convex, we have that λα(z,0)<0\lambda_{\alpha}(z,0)<0 for every α{1,,n1}\alpha\in\{1,\ldots,n-1\}. Thus there exists ε0>0\varepsilon_{0}>0 so that

λα(z,ε)<0,for every α{1,,n1} and |ε|<ε0.\lambda_{\alpha}(z,\varepsilon)<0,\quad\text{for every $\alpha\in\{1,\ldots,n-1\}$ and $|\varepsilon|<\varepsilon_{0}$}.

Therefore, for small enough ε>0\varepsilon>0, we have that

M(ε):=s1(,ε)MUM(\varepsilon):=s^{-1}(-\infty,\varepsilon)\subset M\cup U

is an open set of NN that contains M¯\bar{M}, and whose boundary M(ε)=Ω(ε)\partial M(\varepsilon)=\Omega(\varepsilon) is a smooth strictly convex hyper-surface of NN. We choose ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}) and set M^=M(ε)\hat{M}=M(\varepsilon).

Let p,qM¯p,q\in\bar{M} and choose a distance minimizing unit speed geodesic γ:[0,dM^(p,q)]M^\gamma\colon[0,d_{\hat{M}}(p,q)]\to\hat{M} that connects these points. Now without loss of generality we may assume that γ(t~)U\gamma(\tilde{t})\in U for some t~[0,dM^(p,q)]\tilde{t}\in[0,d_{\hat{M}}(p,q)]. If this is not true then the trace of γ\gamma is contained in MM and we are done.

Since UU is open and γ(t~)U\gamma(\tilde{t})\in U we can choose an interval [a,b][0,dM^(p,q)][a,b]\subset[0,d_{\hat{M}}(p,q)] such that γ([a,b])U\gamma([a,b])\subset U and define a smooth function

s~:[a,b]𝐑,s~(t):=s(γ(t)).\tilde{s}\colon[a,b]\to{\bf R},\quad\tilde{s}(t):=s(\gamma(t)).

Since pp and qq are in M¯\bar{M} we may without loss of generality assume that s~(a),s~(b)0.\tilde{s}(a),\tilde{s}(b)\leq 0.

We aim to verify that s~\tilde{s} is always non-positive. To establish this we show that the maximum value m𝐑m\in{\bf R} of s~\tilde{s} is attained at the endpoints of the domain interval. So suppose that m=s~(t0)m=\tilde{s}(t_{0}) is attained in some interior point t0(a,b)t_{0}\in(a,b). As t0t_{0} is a maximum point of s~\tilde{s}, laying in the interior of the domain interval, it must hold that s~˙(t0)=0\dot{\tilde{s}}(t_{0})=0 and s~¨(t0)0\ddot{\tilde{s}}(t_{0})\leq 0. On the other hand since γ\gamma is a geodesic, we have by Weingarten equation [36, Theorem 8.13 (c)] that

(8) s~˙(t0)=grads(γ(t0)),γ˙(t0)g, and s~¨(t0)=Dtgrads(γ(t0)),γ˙(t0)g=Πγ(t0)(γ˙(t0),γ˙(t0)).\dot{\tilde{s}}(t_{0})=\langle\operatorname{grad}s(\gamma(t_{0})),\dot{\gamma}(t_{0})\rangle_{g},\>\text{ and }\>\ddot{\tilde{s}}(t_{0})=\langle D_{t}\operatorname{grad}s(\gamma(t_{0})),\dot{\gamma}(t_{0})\rangle_{g}=-\Pi_{\gamma(t_{0})}(\dot{\gamma}(t_{0}),\dot{\gamma}(t_{0})).

Here DtD_{t} stands for the covariant differentiation along the curve γ\gamma. Therefore γ˙(t0)\dot{\gamma}(t_{0}) is tangential to the strictly convex hyper-surface Ω(m)\Omega(m) which implies that Πγ(t0)(γ˙(t0),γ˙(t0))<0.\Pi_{\gamma(t_{0})}(\dot{\gamma}(t_{0}),\dot{\gamma}(t_{0}))<0. This in conjunction with (8) leads into a contradiction with s~¨(t0)0\ddot{\tilde{s}}(t_{0})\leq 0. We have verified that for all p,qM¯p,q\in\bar{M} any distance minimizing geodesic in M^\hat{M}, between these points, is contained in M¯\bar{M}. Therefore the equation (7) is true. ∎

By Proposition 7 we may assume that MM is contained in the interior of some compact, Riemannian manifold (M^,g)(\hat{M},g) with a smooth strictly convex boundary. Moreover the distance function of M^\hat{M} restricts to the one of MM. Thus for every (p,v)SM^(p,v)\in S\hat{M} where pp is in MM we always have that

(9) τcut(p,v)τcut^(p,v), and τexit(p,v)<τexit^(p,v),{\tau_{\text{cut}}}(p,v)\leq\hat{\tau_{\text{cut}}}(p,v),\quad\text{ and }\quad{\tau_{\text{exit}}}(p,v)<\hat{\tau_{\text{exit}}}(p,v),

where τcut^\hat{\tau_{\text{cut}}} and τexit^\hat{\tau_{\text{exit}}} are the cut distance and the exit time functions of (M^,g)(\hat{M},g) respectively. Motivated by this observation we define the cut locus of a point pMp\in M as

(10) cut(p):={γp,v(τcut(p,v))M:vSpM,τcut(p,v)=τcut^(p,v)}.\begin{split}{\text{cut}}(p):=&\{\gamma_{p,v}({\tau_{\text{cut}}}(p,v))\in M:\>v\in S_{p}M,\>{\tau_{\text{cut}}}(p,v)=\hat{\tau_{\text{cut}}}(p,v)\}.\end{split}

The following result summarizes the basic properties of these sets.

Proposition 8.

Let Riemannian manifold (M,g)(M,g) be as in Theorem 4. Let pMp\in M.

  • The cut locus cut(p){\text{cut}}(p) of the point pp is a closed set of measure zero.

  • If qcut(p)q\in{\text{cut}}(p) and γ\gamma is a unit speed distance minimizing geodesic of MM between pp and qq then at least one of the following holds:

    1. (1)

      There exists another distance minimizing geodesic from pp to qq.

    2. (2)

      qq is the first conjugate point to pp along γ\gamma.

Proof.
  • The proof of the first claim is identical to the proof of [36, Theorem 10.34 (a)], thus omitted here.

  • If qcut(p)q\in{\text{cut}}(p) then by (9) there is vSpMv\in S_{p}M such that

    q=γp,v(τcut(p,v)),τcut(p,v)=τcut^(p,v)τexit(p,v)<τexit^(p,v),q=\gamma_{p,v}({\tau_{\text{cut}}}(p,v)),\quad{\tau_{\text{cut}}}(p,v)=\hat{\tau_{\text{cut}}}(p,v)\leq{\tau_{\text{exit}}}(p,v)<\hat{\tau_{\text{exit}}}(p,v),

    Thus Lemma 5 and Proposition 7 yield the second claim.

The following result introduces an open and dense subset of MM where the distance function of an interior point is smooth.

Lemma 9.

Let Riemannian manifold (M,g)(M,g) be as in Theorem 4. Let pMp\in M. The distance function d(p,):M𝐑,d(p,\cdot)\colon M\to{\bf R}, is smooth precisely in the open and dense set M({p}cut(p))M\setminus(\{p\}\cup{\text{cut}}(p)).

Proof.

The proof is identical to the proof of the analogous claim of [45, Chapter 5, Section 9, Corollary 7] thus omitted here. ∎

Proposition 10.

Let Riemannian manifold (M,g)(M,g) be as in Theorem 4. Let pMp\in M and qM({p}cut(p))q\in M\setminus(\{p\}\cup{\text{cut}}(p)). There exist neighborhoods UMU\subset M of pp and VMV\subset M of qq such that the distance function d(,)d(\cdot,\cdot) is smooth in the product set U×VU\times V.

Proof.

Let (N,g)(N,g) be a closed extension of (M,g)(M,g) as in Proposition 7. We define a smooth map

F:(x,v)TN(x,expx(v))N×N.F:(x,v)\in TN\mapsto(x,\exp_{x}(v))\in N\times N.

Then the differential of this map can be written as

DF(x,v)=[Id0Dexpx(v)].\mathrm{D}F(x,v)=\begin{bmatrix}\text{Id}&0\\ *&\mathrm{D}\exp_{x}(v)\end{bmatrix}.

Since qq is not in the cut locus of pp there is a v0TpNv_{0}\in T_{p}N such that v0g=dM(p,q)\|v_{0}\|_{g}=d_{M}(p,q), expp(v0)=q\exp_{p}(v_{0})=q and Dexpp(v0)\mathrm{D}\exp_{p}(v_{0}) is not singular. Therefore det(DF(p,v0))=det(Dexpp(v0))\det(\mathrm{D}F(p,v_{0}))=\det(\mathrm{D}\exp_{p}(v_{0})) does not vanish. Thus the Inverse function theorem implies that there are neighborhoods W~TN\tilde{W}\subset TN of (p,v0)(p,v_{0}) and WN×NW\subset N\times N of (p,q)(p,q) such that the local inverse function of FF,

F1:WW~,F1(x,y)=(x,expx1(y)),F^{-1}\colon W\to\tilde{W},\quad F^{-1}(x,y)=(x,\exp_{x}^{-1}(y)),

is a diffeomorphism.

Let (M^,g)(\hat{M},g) be the extension of (M,g)(M,g) as in Proposition 7. Since qq is not in the cut locus of pp we have F1(p,q)g<τcut^(p,v0v0g).\|F^{-1}(p,q)\|_{g}<\hat{\tau_{\text{cut}}}\left(p,\frac{v_{0}}{\|v_{0}\|_{g}}\right). Thus by the continuity of the cut distance function τcut^(,)\hat{\tau_{\text{cut}}}\left(\cdot,\cdot\right) of (M^,g)(\hat{M},g) we can choose a neighborhood W1WW_{1}\subset W of (p,q)(p,q) such that

F1(x,y)g<τcut^(x,F1(x,y)F1(x,y)g),for all (x,y)W1.\|F^{-1}(x,y)\|_{g}<\hat{\tau_{\text{cut}}}\left(x,\frac{F^{-1}(x,y)}{\|F^{-1}(x,y)\|_{g}}\right),\quad\text{for all }(x,y)\in W_{1}.

This gives dM^(x,y)=F1(x,y)g, for all (x,y)W1.d_{\hat{M}}(x,y)=\|F^{-1}(x,y)\|_{g},\text{ for all }(x,y)\in W_{1}.

Finally we choose disjoint neighborhoods UMU\subset M of pp and VMV\subset M of qq such that U×VU\times V is contained in W1W_{1}. Let (x,y)U×V(x,y)\in U\times V then γ(t):=expx(tF1(x,y))\gamma(t):=\exp_{x}(tF^{-1}(x,y)) for t[0,1]t\in[0,1] is a geodesic of M^\hat{M} that connects xx to yy having the length of dM^(x,y)d_{\hat{M}}(x,y). Since both xx and yy are in MM, we get from the proof of Proposition 7 that γ(t)\gamma(t) is contained in MM. This yields

(11) dM(x,y)=F1(x,y)g, for all (x,y)U×V.d_{M}(x,y)=\|F^{-1}(x,y)\|_{g},\text{ for all }(x,y)\in U\times V.

Since the sets UU and VV are disjoint we have that F1F^{-1} does not vanish in U×VU\times V. Hence equation (11) gives the smoothness of dM(,)d_{M}(\cdot,\cdot) on U×VU\times V. ∎

Recall that we have assumed that MM is isometrically embedded in a closed Riemannian manifold (N,g)(N,g). Thus any geodesic starting from MM can be extended to the entire 𝐑{\bf R}. Let pNp\in N. We define the conjugate distance function τcon:SpN𝐑{\tau_{\text{con}}}\colon S_{p}N\to{\bf R}\cup{\infty} by the formula:

τcon(p,v)=inf{t>0:γp,v(t) is a conjugate point to p}.{\tau_{\text{con}}}(p,v)=\inf\{t>0:\>\gamma_{p,v}(t)\text{ is a conjugate point to $p$}\}.

As the infimum of the empty set is positive infinity we set τcon(p,v)={\tau_{\text{con}}}(p,v)=\infty in the case when the geodesic γp,v\gamma_{p,v} does not have any conjugate points to pp. Since geodesics do not minimize the distance beyond the first conjugate point it holds that

τcut(p,v)τcon(p,v), if (p,v)SM.{\tau_{\text{cut}}}(p,v)\leq{\tau_{\text{con}}}(p,v),\quad\text{ if }(p,v)\in SM.

The following result is well known, but we could not find its proof in the existing literature, so we provide one below.

Lemma 11.

Let (N,g)(N,g) be a closed Riemannian manifold and pNp\in N. The conjugate distance function is continuous on SpNS_{p}N.

Proof.

Let viSpNv_{i}\in S_{p}N for i𝐍i\in\mathbf{N} converge to vv.

We set

C=τcon(p,v),B=lim infiτcon(p,vi), and A=lim supiτcon(p,vi).C={\tau_{\text{con}}}(p,v),\quad B=\liminf_{i\to\infty}{\tau_{\text{con}}}(p,v_{i}),\quad\hbox{ and }\quad A=\limsup_{i\to\infty}{\tau_{\text{con}}}(p,v_{i}).

It suffices to show that ACBA\leq C\leq B.

We assume first that C=C=\infty. If A<A<\infty we choose a sub-sequence vikv_{i_{k}} of viv_{i} such that τcon(p,vik){\tau_{\text{con}}}(p,v_{i_{k}}) converges to AA. Then det(Dexpp(τcon(p,vik)vik))=0,\det(\mathrm{D}\exp_{p}({\tau_{\text{con}}}(p,v_{i_{k}})v_{i_{k}}))=0, and the smoothness of the exponential map gives det(Dexpp(Av))=0\det(\mathrm{D}\exp_{p}(Av))=0 yielding that expp(Av)\exp_{p}(Av) is conjugate to pp along γp,v\gamma_{p,v}. This implies that τcon(p,v)A{\tau_{\text{con}}}(p,v)\leq A, which is impossible. By the same argument we see that B=B=\infty.

Let C<C<\infty, and by the same limiting argument as above we get CBC\leq B. Then we show that ACA\leq C. Choose a sub-sequence vikv_{i_{k}} such that τcon(p,vik)A{\tau_{\text{con}}}(p,v_{i_{k}})\to A as kk\to\infty. For the sake of contradiction we suppose that A>CA>C. By the definition of the conjugate distance function we have that q=γp,v(C)q=\gamma_{p,v}(C) is the first conjugate to pp along γp,v\gamma_{p,v}. By [36, Theorem 10.26] for any ε(0,AC)\varepsilon\in(0,A-C) there exists a piecewise smooth vector field XX on the geodesic segment γp,v:[0,C+ε]N\gamma_{p,v}\colon{[0,C+\varepsilon]}\to N, that vanishes on 0 and C+εC+\varepsilon such that the index form of γp,v\gamma_{p,v} over XX is strictly negative. That is

(12) Iv(X,X):=0C+εDtX,DtXg+R(γ˙p,v,X)γ˙p,v,Xgdt<0.I_{v}(X,X):=\int_{0}^{C+\varepsilon}\langle\mathrm{D}_{t}X,\mathrm{D}_{t}X\rangle_{g}+\langle R(\dot{\gamma}_{p,v},X)\dot{\gamma}_{p,v},X\rangle_{g}\;\mathrm{d}t<0.

Here we used the notation Dt\mathrm{D}_{t} for the covariant differentiation along γp,v\gamma_{p,v}. The capital RR stands for the Riemannian curvature tensor.

We choose vectors E1,,EnE_{1},\ldots,E_{n} of TpNT_{p}N that form a basis of TpNT_{p}N and extend them on γp,v(t)\gamma_{p,v}(t) for t[0,C+ε]t\in[0,C+\varepsilon] via the parallel transport. Since parallel transport is an isomorphism the vector fields {E1(t),,En(t)}\{E_{1}(t),\ldots,E_{n}(t)\} constitute a basis of Tγp,v(t)MT_{\gamma_{p,v}(t)}M. We write X(t)=Xj(t)Ej(t)X(t)=X^{j}(t)E_{j}(t). Since XX is piecewise smooth it holds that the component functions Xj(t)X^{j}(t) are piecewise smooth. This lets us ‘extend’ XX on γp,vik\gamma_{p,v_{i_{k}}} by the formula

(13) Xk(t)=Xj(t)Ejk(t),X_{k}(t)=X^{j}(t)E_{j}^{k}(t),

where the vector field Ejk(t)E_{j}^{k}(t) is the parallel transport of EjE_{j} along γp,vik\gamma_{p,v_{i_{k}}}. Thus XkX_{k} is a piecewise smooth vector field on γp,vik\gamma_{p,v_{i_{k}}} that vanishes at t=0t=0 and t=C+εt=C+\varepsilon.

Since vikv, as k,v_{i_{k}}\to v,\hbox{ as $k\to\infty$}, it holds that

γp,vik(t)γp,v(t), and Ejk(t)Ej(t),uniformly in t[0,C+ε] as k\gamma_{p,v_{i_{k}}}(t)\to\gamma_{p,v}(t),\quad\hbox{ and }\quad E^{k}_{j}(t)\to E_{j}(t),\quad\hbox{uniformly in $t\in[0,C+\varepsilon]$ as $k\to\infty$. }

Therefore by (12), (13), the continuity of the Levi-Civita connection, and the Riemannian curvature tensors we have

Ivik(Xk,Xk)<0, for large enough k𝐍.I_{v_{i_{k}}}(X_{k},X_{k})<0,\quad\hbox{ for large enough $k\in\mathbf{N}$.}

By [36, Theorem 10.28] there exists sk(0,C+ε]s_{k}\in(0,C+\varepsilon] so that γp,vik(0)\gamma_{p,v_{i_{k}}}(0) and γp,vik(sk)\gamma_{p,v_{i_{k}}}(s_{k}) are conjugate points. Therefore we must have skτcon(p,vik)s_{k}\geq{\tau_{\text{con}}}(p,v_{i_{k}}) and we arrive at a contradiction τcon(p,vik)<A.{\tau_{\text{con}}}(p,v_{i_{k}})<A. This ends the proof. ∎

3.2. The Hausdorff dimension of the cut locus

We fix a point pMp\in M for this sub-section. By Lemma 9 we know that for any pMp\in M the distance function d(p,)d(p,\cdot) is smooth in the open set M(cut(p){p})M\setminus({\text{cut}}(p)\cup\{p\}). Moreover by Proposition 10 for each qq in this open set there are neighborhoods UU of pp and VV of qq such that the distance function is smooth in the product set U×VU\times V. As we are interested in the inverse problem where we study distance function d(p,)d(p,\cdot) restricted on some open subset Γ\Gamma of the boundary, we do not know a priori if this function is smooth on Γ\Gamma. In particular we do not know the size of the set cut(p)M{\text{cut}}(p)\cap\partial M yet. In this sub-section we show that the set Mcut(p)\partial M\setminus{\text{cut}}(p), where d(p,)d(p,\cdot) is smooth, is always an open and dense subset of M\partial M.

Proposition 8 yields that cut(p){\text{cut}}(p) can be written as a disjoint union of

  • Conjugate cut points:

    Q(p):={γp,v(t)M:vSpM,t=τcut(p,v)=τcon(p,v)}cut(p),Q(p):=\{\gamma_{p,v}(t)\in M:v\in S_{p}M,\>t={\tau_{\text{cut}}}(p,v)={\tau_{\text{con}}}(p,v)\}\subset{\text{cut}}(p),

    that are those points qcut(p)q\in{\text{cut}}(p) such that there exits a distance minimizing geodesic from pp to qq along which these points are conjugate to each other. By Proposition 8 and Lemma 11 the set Q(p)Q(p) is closed in MM.

  • Typical cut points: T(p)(cut(p)Q(p))T(p)\subset(cut(p)\setminus Q(p)) that can be connected to pp with exactly two distance minimizing geodesics.

  • A-typical cut points: L(p)(cut(p)Q(p))L(p)\subset(cut(p)\setminus Q(p)) that can be connected to pp with more than two distance minimizing geodesics. Thus an a-typical cut point is both non-conjugate and non-typical.

It was proven in [19] that the Hausdorff dimension of the cut locus on a closed Riemannian manifold (N,g)(N,g) is locally an integer that does not exceed dimN1\dim N-1. Moreover T(p)T(p) is a smooth hyper-surface of NN and the Hausdorff dimension of Q(p)L(p)Q(p)\cup L(p) does not exceed dimN2\dim N-2. In this paper we will extended these results for manifolds with strictly convex boundary. The main result of this section is as follows:

Theorem 12.

Let (M,g)(M,g) be a smooth, compact, connected, and oriented Riemannian manifold of dimension n𝐍,n2n\in\mathbf{N},\>n\geq 2 with smooth and strictly convex boundary. If pMp\in M then

  1. (1)

    The set T(p)T(p) of typical cut points is a smooth hyper-surface of MM that is transverse to M\partial M.

  2. (2)

    The Hausdorff dimension of Q(p)L(p)Q(p)\cup L(p) does not exceed n2n-2.

  3. (3)

    The Hausdorff dimension of cut(p){\text{cut}}(p) does not exceed n1n-1.

  4. (4)

    The set Mcut(p)\partial M\setminus{\text{cut}}(p) is open and dense in M\partial M.

For the readers who want to learn more about Hausdorff measure and dimension we suggest to have look at [6, 37]. Some basic properties of the Hausdorff dimension dim\dim_{\mathcal{H}} are collected in the following lemma.

Lemma 13.

Basic properties of the Hausdorff dimension are:

  • If XX is a metric space and AXA\subset X then dim(A)dim(X)\dim_{\mathcal{H}}(A)\leq\dim_{\mathcal{H}}(X).

  • If XX is a metric space and 𝒳\mathcal{X} is a countable cover of XX then dim(X)=supA𝒳dim(A)\dim_{\mathcal{H}}(X)=\sup_{A\in\mathcal{X}}\dim_{\mathcal{H}}(A).

  • If X,YX,Y are metric spaces and f:XYf\colon X\to Y is a bi-Lipschitz map then dim(A)=dim(f(A))\dim_{\mathcal{H}}(A)=\dim_{\mathcal{H}}(f(A)) for any AXA\subset X.

  • If UMU\subset M is open and MM is a Riemannian manifold of dimension n𝐍n\in\mathbf{N} then dim(U)=n\dim_{\mathcal{H}}(U)=n.

From here onwards we follow the steps of [19, 42] and develop machinery needed for the proof of Theorem 12. We recall that we have isometrically embedded MM into the closed Riemannian manifold (N,g)(N,g). The maximal subset MpTpMM_{p}\subset T_{p}M where the exponential map expp:MpM\exp_{p}\colon M_{p}\to M of (M,g)(M,g) is well defined was given in (4) as

Mp={vTpM:v=0 or vgτexit(p,vvg)}.M_{p}=\left\{v\in T_{p}M:\>v=0\hbox{ or }\|v\|_{g}\leq{\tau_{\text{exit}}}\left(p,\frac{v}{\|v\|_{g}}\right)\right\}.

Thus the exponential function expp:TpNN\exp_{p}\colon T_{p}N\to N of NN agrees with that of MM in MpM_{p}.

Let v0Mpv_{0}\in M_{p} be such that the exponential map expp\exp_{p} is not singular at v0v_{0}. The Inverse function theorem yields that there are neighborhoods UTpNU\subset T_{p}N of v0v_{0} and V~N\tilde{V}\subset N of x0=expp(v0)Mx_{0}=\exp_{p}(v_{0})\in M such that expp:UV~\exp_{p}\colon U\to\tilde{V} is a diffeomorphism. We want to emphasize that even when v0Mpv_{0}\in M_{p} the set UTpNU\subset T_{p}N does not need to be contained in MpM_{p}. If we equate TpNT_{p}N and Tv0(TpN)T_{v_{0}}(T_{p}N) we note that the formula

Y(x)=Dexpp|expp1(x)expp1(x)Y(x)=\mathrm{D}\exp_{p}\bigg{|}_{\exp_{p}^{-1}(x)}\exp_{p}^{-1}(x)

defines a smooth vector field on V~\tilde{V} that satisfies the following properties

(14) Y(x)=γ˙p,expp1(x)(1), and Y(x)g=expp1(x)g.Y(x)=\dot{\gamma}_{p,\exp_{p}^{-1}(x)}(1),\quad\hbox{ and }\quad\|Y(x)\|_{g}=\|\exp_{p}^{-1}(x)\|_{g}.

The vector field YY is called a distance vector field related to pp and UU. Let xV~x\in\tilde{V} and XTxNX\in T_{x}N. It holds by a similar proof to [42, Lemma 2.2.] that

(15) XY(x)g=X,Y(x)gY(x)g.X\|Y(x)\|_{g}=\frac{\langle X,Y(x)\rangle_{g}}{\|Y(x)\|_{g}}.

In what follows we will always consider cut(p){\text{cut}}(p) as defined for (M,g)(M,g) in equation (10). Let qcut(p)Q(p)q\in{\text{cut}}(p)\setminus Q(p) and λ=d(p,q)\lambda=d(p,q). By Proposition 8 it holds that there are at least two MM-distance minimizing geodesics from pp to qq. Thus the set

Ep,q:=expp1{q}SλM, where SλM={wTpM:wg=λ},E_{p,q}:=\exp_{p}^{-1}\{q\}\cap S_{\lambda}M,\quad\hbox{ where }S_{\lambda}M=\{w\in T_{p}M:\>\|w\|_{g}=\lambda\},

contains at least two points.

By the compactness of SλMS_{\lambda}M and the continuity of the exit time function on the non-trapping part of SMSM we get from the assumption that qQ(p)q\notin Q(p) that the set Ep,qE_{p,q} is finite. We write

Ep,q={wi:i{1,,kp(q)}},E_{p,q}=\{w_{i}:\>i\in\{1,\ldots,k_{p}(q)\}\},

where kp(q)𝐍k_{p}(q)\in\mathbf{N} is the number of distance minimizing geodesics from pp to qq. Since the set Q(p)Q(p) is closed in MM, the complement cut(p)Q(p){\text{cut}}(p)\setminus Q(p) is relatively open in cut(p){\text{cut}}(p), and there exists an open neighborhood WMW\subset M of qq such that Q(p)W=.Q(p)\cap W=\emptyset. Thus by the previous discussion for any xcut(p)Wx\in{\text{cut}}(p)\cap W there are only kp(x)𝐍k_{p}(x)\in\mathbf{N} many distance minimizing geodesics connecting pp to xx. Moreover, as the following lemma shows, qq is a local maximum of the function kpk_{p} defined on cut(p)W.{\text{cut}}(p)\cap W. This statement is an adaptation of the analogous result given in [42].

Lemma 14.

Let (M,g)(M,g) be a Riemannian manifold as in Theorem 12. Let pMp\in M and qcut(p)Q(p)q\in{\text{cut}}(p)\setminus Q(p). Let the closed manifold (N,g)(N,g) be as in Proposition 7. Then there is a neighborhood VV of qq in MM such that

(16) kp(x)kp(q), for every xcut(p)V.k_{p}(x)\leq k_{p}(q),\quad\hbox{ for every }x\in{\text{cut}}(p)\cap V.
Proof.

Since the set Ep,qE_{p,q} is finite we can choose disjoint neighborhoods UiTpNU_{i}\subset T_{p}N for each wiEp,qw_{i}\in E_{p,q}, so that for each i{1,,kp(q)}i\in\{1,\ldots,k_{p}(q)\} the map expp:UiV~\exp_{p}\colon U_{i}\to\tilde{V} is a diffeomorphism on some open set V~N\tilde{V}\subset N that contains qq. We want to show that there is a neighborhood VMV\subset M of qq such that for every xVx\in V and for any MM-distance minimizing unit speed geodesic γ\gamma from pp to xx there is i{1,,kp(q)}i\in\{1,\ldots,k_{p}(q)\} such that

γ(t)=expp(tXXg), for some XMpUi.\gamma(t)=\exp_{p}\left(t\frac{X}{\|X\|_{g}}\right),\quad\hbox{ for some }X\in M_{p}\cap U_{i}.

Clearly this implies the inequality (16).

If the former is not true then there exist a sequence qkMq_{k}\in M that converges to qq and XkMpX_{k}\in M_{p} so that for each k𝐍k\in\mathbf{N} we have

  • expp(Xk)=qk\exp_{p}(X_{k})=q_{k}

  • Xkg=dM(p,qk)τexit(p,XkXkg)\|X_{k}\|_{g}=d_{M}(p,q_{k})\leq{\tau_{\text{exit}}}(p,\frac{X_{k}}{\|X_{k}\|_{g}})

  • expp(tXkXkg)\exp_{p}\left(t\frac{X_{k}}{\|X_{k}\|_{g}}\right) for t[0,dM(p,qk)]t\in[0,d_{M}(p,q_{k})] is a unit speed distance minimizing geodesic from pp to qkq_{k}.

  • XkU1Ukp(q)X_{k}\notin U_{1}\cup\ldots\cup U_{k_{p}(q)}.

These imply that

limkXkg=limkdM(p,qk)=dM(p,q).\lim_{k\to\infty}\|X_{k}\|_{g}=\lim_{k\to\infty}d_{M}(p,q_{k})=d_{M}(p,q).

Moreover, the sequence XkTpMX_{k}\in T_{p}M is contained in some compact subset KK of TpMT_{p}M. After passing to a sub-sequence we may assume XkXTpMX_{k}\to X\in T_{p}M and the continuity of the exit time function on the non-trapping part of SMSM gives Xgτexit(p,XXg)\|X\|_{g}\leq{\tau_{\text{exit}}}(p,\frac{X}{\|X\|_{g}}). Thus XMpX\in M_{p} and by the continuity of expp\exp_{p} we get expp(X)=q, and Xg=dM(p,q).\exp_{p}(X)=q,\hbox{ and }\|X\|_{g}=d_{M}(p,q).

Therefore texpp(tXXg)t\mapsto\exp_{p}\left(t\frac{X}{\|X\|_{g}}\right) is a MM-distance minimizing geodesic from pp to qq and XX must coincide with wiw_{i} for some i{1,,kp(q)}i\in\{1,\ldots,k_{p}(q)\}. Therefore XkUiX_{k}\in U_{i} for large enough k𝐍k\in\mathbf{N}. This contradicts the choice of XkX_{k}, and possibly after choosing a smaller V~\tilde{V}, we can set V=MV~V=M\cap\tilde{V}. ∎

Suppose now that qT(p)q\in T(p) is a typical cut point, and VMV\subset M is a neighborhood of qq as in Lemma 14. Then by (16) it holds that

(17) cut(p)VT(p),{\text{cut}}(p)\cap V\subset T(p),

and Ep,q={w1,w2}TpNE_{p,q}=\{w_{1},w_{2}\}\subset T_{p}N are the directions that give the two distance minimizing geodesics expp(twi),t[0,1]\exp_{p}(tw_{i}),\>t\in[0,1] from pp to qq. Let U1,U2TpNU_{1},U_{2}\subset T_{p}N be the neighborhoods of w1w_{1} and w2w_{2} and V~N\tilde{V}\subset N a neighborhood of qq as in the proof of Lemma 14. Finally we consider the distance vector fields Y1,Y2Y_{1},Y_{2} related to pp and U1U_{1} and U2U_{2}. Since these vector fields do not vanish on V~\tilde{V} the function

ρ:V~𝐑,ρ(x)=Y1(x)gY2(x)g,\rho\colon\tilde{V}\to{\bf R},\quad\rho(x)=\|Y_{1}(x)\|_{g}-\|Y_{2}(x)\|_{g},

is smooth. The following result is an adaptation of [42, Propositions 2.3 & 2.4].

Lemma 15.

Let Riemannian manifold (M,g)(M,g) be as in Theorem 12 and pMp\in M. Let qT(p)Mq\in T(p)\subset M and define the closed manifold NN as in Proposition 7. Let the neighborhood V~N\tilde{V}\subset N of qq and function ρ:V~𝐑\rho\colon\tilde{V}\to{\bf R} be as above. Then possibly after choosing a small enough V~\tilde{V} we have

(18) ρ1{0}M=V~cut(p).\rho^{-1}\{0\}\cap M=\tilde{V}\cap{\text{cut}}(p).

Moreover, the set ρ1{0}\rho^{-1}\{0\} is a smooth hyper-surface of NN whose tangent bundle is given by the orthogonal complement of the vector field Y1Y2Y_{1}-Y_{2}.

Proof.

We prove first the equation (18).

  • Let xρ1{0}Mx\in\rho^{-1}\{0\}\cap M. By the proof of Lemma 14 we can assume that a MM-distance minimizing unit speed geodesic from pp to xx is given by expp(tX1),t[0,1], for some X1MpU1.\exp_{p}\left(tX_{1}\right),\>t\in[0,1],\>\hbox{ for some }X_{1}\in M_{p}\cap U_{1}. Also x=expp(X2)x=\exp_{p}(X_{2}) for some X2U2X_{2}\in U_{2}, but we do not know a priori if expp(tX2)M\exp_{p}(tX_{2})\in M for all t[0,1]t\in[0,1] or equivalently if X2MpX_{2}\in M_{p}. However, by the definition of the distance vector fields and the assumption xρ1{0}x\in\rho^{-1}\{0\} we have that

    (19) dM(p,x)=X1g=Y1(x)g=Y2(x)g=X2g.d_{M}(p,x)=\|X_{1}\|_{g}=\|Y_{1}(x)\|_{g}=\|Y_{2}(x)\|_{g}=\|X_{2}\|_{g}.

    Let M^\hat{M} be as in Proposition 7. Thus we can assume that V~M^\tilde{V}\subset\hat{M}. Since qT(p)q\in T(p) there is w2U2w_{2}\in U_{2} so that expp(w2)=q\exp_{p}(w_{2})=q and expp(tw2)MM^,\exp_{p}(tw_{2})\in M\subset\hat{M}, for every t[0,1]t\in[0,1]. Since qq is an interior point of M^\hat{M} we can again choose smaller V~\tilde{V} so that expp(tX)M^,\exp_{p}(tX)\in\hat{M}, for every t[0,1]t\in[0,1] and XUiX\in U_{i}, for i{1,2}i\in\{1,2\}. Since xMx\in M and expp(tX2)\exp_{p}(tX_{2}) is a geodesic of M^\hat{M} that connects pp to xx having the length of X2g\|X_{2}\|_{g}, the equation (19) and Proposition 7 imply that expp(tX2)M\exp_{p}(tX_{2})\in M for every t[0,1]t\in[0,1]. Therefore equation (19) gives xV~cut(p)x\in\tilde{V}\cap{\text{cut}}(p).

  • Let xV~cut(p)T(p)x\in\tilde{V}\cap{\text{cut}}(p)\subset T(p). Thus there are exactly two distance minimizing geodesics of MM from pp to xx. Since xV~x\in\tilde{V}, it holds by the proof of Lemma 14 that one of these geodesics has the initial velocity in U1U_{1} and the other in U2U_{2}. Therefore ρ(x)\rho(x) is zero by the definition of the distance vector fields.

Then we prove that the set ρ1{0}\rho^{-1}\{0\} is a smooth hyper-surface whose tangent bundle is orthogonal to the vector field Y1Y2Y_{1}-Y_{2}. By (15) we get

(20) Xρ(x)=X,Y1(x)gY1(x)gX,Y2(x)gY2(x)g=X,Y1(x)Y2(x)gY1(x)g, for every xρ1{0}.X\rho(x)=\frac{\langle X,Y_{1}(x)\rangle_{g}}{\|Y_{1}(x)\|_{g}}-\frac{\langle X,Y_{2}(x)\rangle_{g}}{\|Y_{2}(x)\|_{g}}=\frac{\langle X,Y_{1}(x)-Y_{2}(x)\rangle_{g}}{\|Y_{1}(x)\|_{g}},\quad\hbox{ for every }x\in\rho^{-1}\{0\}.

Moreover the vector field Y1Y2Y_{1}-Y_{2} does not vanish on V~\tilde{V}, since the geodesics related to these two vector fields are different. This implies that the differential of the map ρ\rho does not vanish in V~\tilde{V}. Thus the set ρ1{0}\rho^{-1}\{0\} is a smooth hyper-surface of NN, and by (20) its tangent bundle is given by those vectors that are orthogonal to Y1Y2Y_{1}-Y_{2}. ∎

Now we consider the set of conjugate cut points Q(p)Q(p). First we define a function

δ:SpN{0,1,,n1},δ(v) is the dimension of the kernel of Dexpp at τcon(p,v)v.\delta\colon S_{p}N\to\{0,1,\ldots,n-1\},\quad\delta(v)\text{ is the dimension of the kernel of }\mathrm{D}\exp_{p}\text{ at }{\tau_{\text{con}}}(p,v)v.

If τcut(p,v)={\tau_{\text{cut}}}(p,v)=\infty we set δ(v)=0\delta(v)=0.

Lemma 16.

Let (N,g)(N,g) be a closed Riemannian manifold. Let pNp\in N and v0SpNv_{0}\in S_{p}N be such that δ(v0)=1\delta(v_{0})=1. There exists a neighborhood USpNU\subset S_{p}N of v0v_{0} such that the δ()\delta(\cdot) is the constant function one in UU.

Before proving this lemma we recall one auxiliary result from linear algebra.

Lemma 17.

Let L:𝐑n𝐑nL\colon{\bf R}^{n}\to{\bf R}^{n} be a self-adjoint bijective linear operator. Then the index i(L)i(L) of LL, the dimension of the largest vector subspace of 𝐑n{\bf R}^{n} where LL is negative definite, equals the amount of the negative eigenvalues of the operator LL counted up to a multiplicity.

Proof.

The proof follows directly from the spectral theorem and thus we omit it here. ∎

Proof of Lemma 16.

In this proof we adopt the definitions and results of [15, Chapter 11] appearing in the proof of the Morse index theorem. For each vSpMv\in S_{p}M and t>0t>0 we use the notation 𝒱(t,v)\mathcal{V}(t,v) for the vector space of all piecewise smooth vector fields that are normal to the geodesic γp,v\gamma_{p,v} in the interval [0,t][0,t] and vanish at the endpoints. Then we define the function i:[0,)×SpN𝐍,i\colon[0,\infty)\times S_{p}N\to\mathbf{N}, to be the index of the symmetric bilinear form

It,v(X,Y):=0tDsX,DsYg+R(γ˙p,v,X)γ˙p,v,Ygds,X,Y𝒱(t,v).I_{t,v}(X,Y):=\int_{0}^{t}\langle\mathrm{D}_{s}X,\mathrm{D}_{s}Y\rangle_{g}+\langle R(\dot{\gamma}_{p,v},X)\dot{\gamma}_{p,v},Y\rangle_{g}\;\mathrm{d}s,\quad X,Y\in\mathcal{V}(t,v).

Hence,

{δ(v)>i(t,v)=0,tτcon(p,v)δ(v)=i(t,v),t(τcon(p,v),ε(v))δ(v)<i(t,v),t>ε(v)\left\{\begin{array}[]{ll}\delta(v)>i(t,v)=0,&t\leq{\tau_{\text{con}}}(p,v)\\ \delta(v)=i(t,v),&t\in({\tau_{\text{con}}}(p,v),\varepsilon(v))\\ \delta(v)<i(t,v),&t>\varepsilon(v)\end{array}\right.

where ε(v)>0\varepsilon(v)>0 depends on vSpNv\in S_{p}N. Moreover, no γp,v(t)\gamma_{p,v}(t), for t(τcon(p,v),ε(v))t\in({\tau_{\text{con}}}(p,v),\varepsilon(v)) is conjugate to pp along γp,v\gamma_{p,v}. We choose t(τcon(p,v0),ε(v0))t\in({\tau_{\text{con}}}(p,v_{0}),\varepsilon(v_{0})). Thus Lemma 11, gives δ(v)i(t,v)\delta(v)\leq i(t,v) for vSpNv\in S_{p}N close enough to v0v_{0}. Our aim is to find a neighborhood USpNU\subset S_{p}N of v0v_{0} for which

(21) 1δ(v)i(t,v)=i(t,v0)=δ(v0)=1,for vU.1\leq\delta(v)\leq i(t,v)=i(t,v_{0})=\delta(v_{0})=1,\quad\text{for $v\in U$}.

Clearly this gives the result of the lemma.

The space 𝒱(t,v)\mathcal{V}(t,v) can be written as a direct sum of two of its vector sub-spaces 𝒱+(t,v)\mathcal{V}_{+}(t,v) and 𝒱(t,v)\mathcal{V}_{-}(t,v), defined so that index form It,vI_{t,v} is positive definite on 𝒱+(t,v)\mathcal{V}_{+}(t,v) and the space 𝒱(t,v)\mathcal{V}_{-}(t,v) is finite dimensional. Moreover, these vector spaces are It,vI_{t,v} orthogonal. Thus the index of It,vI_{t,v} coincides with the index of its restriction on 𝒱(t,v)\mathcal{V}_{-}(t,v). Since the dimension of 𝒱(t,v)\mathcal{V}_{-}(t,v) is independent of a direction vSpNv\in S_{p}N, that is close to v0v_{0}, we can identify all the spaces 𝒱(t,v)\mathcal{V}_{-}(t,v) with 𝒱(t,v0)\mathcal{V}_{-}(t,v_{0}) and consider the bilinear forms It,vI_{t,v} as a family of operators on the finite dimensional vector space 𝒱(t,v0)\mathcal{V}_{-}(t,v_{0}), depending continuously on the parameter vSpNv\in S_{p}N.

For each vSpNv\in S_{p}N, we consider the linear operator Lt,v:𝒱(t,v)𝒱(t,v)L_{t,v}\colon\mathcal{V}_{-}(t,v)\to\mathcal{V}_{-}(t,v), corresponding to the bilinear form It,vI_{t,v}. Since γp,v0(t)\gamma_{p,v_{0}}(t) is not a conjugate point to pp along γp,v0\gamma_{p,v_{0}}, zero is not an eigenvalue of the linear operator Lt,v0L_{t,v_{0}}. Thus Lemma 17 implies that the operator Lt,v0L_{t,v_{0}} has i(t,v0)i(t,v_{0}) negative eigenvalues. Since the eigenvalues of the operator Lt,vL_{t,v} depend continuously on the initial direction vSpNv\in S_{p}N, that are near v0v_{0}, we can find a neighborhood USpNU\subset S_{p}N of v0v_{0} such that the linear operator Lt,vL_{t,v} is invertible and has i(t,v0)i(t,v_{0}) negative eigenvalues for every vUv\in U. Hence, by Lemma 17 we have again that i(t,v)=i(t,v0)i(t,v)=i(t,v_{0}) for vUv\in U. We have verified the equation (21). ∎

Lemma 18.

Let (N,g)(N,g) be a closed Riemannian manifold, pNp\in N and suppose that δ\delta is constant in some open set USpMU\subset S_{p}M. Then τcon(p,){\tau_{\text{con}}}(p,\cdot) is smooth in UU.

Proof.

If δ\delta is zero in UU then τcon(p,){\tau_{\text{con}}}(p,\cdot) is infinite and we are done. So we suppose that δ\delta equals to k{1,,n1}k\in\{1,\ldots,n-1\} in UU and get τcon(p,v)<{\tau_{\text{con}}}(p,v)<\infty for every vUv\in U.

Let ξ1,,ξn1\xi_{1},\ldots,\xi_{n-1} be a base of Tv0SpMT_{v_{0}}S_{p}M and use the formula

Jv0,β(t)=Dexpp|tv0tξβ, for β{1,,n1},J_{v_{0},\beta}(t)=\mathrm{D}\exp_{p}\bigg{|}_{tv_{0}}t\xi_{\beta},\quad\text{ for }\beta\in\{1,\ldots,n-1\},

from [36, Proposition 10.10], to define (n1)(n-1)-Jacobi fields Jv0,1(t),,Jv0,n1(t)J_{v_{0},1}(t),\ldots,J_{v_{0},n-1}(t) along the geodesic γp,v0\gamma_{p,v_{0}}. They span the vector space of all Jacobi fields along γp,v0(t)\gamma_{p,v_{0}}(t) that vanish at t=0t=0 and are normal to γ˙p,v0(t)\dot{\gamma}_{p,v_{0}}(t). As δ(v0)=k\delta(v_{0})=k we may assume that Jv0,β(τcon(p,v0))=0J_{v_{0},\beta}({\tau_{\text{con}}}(p,v_{0}))=0 for β{1,,k}\beta\in\{1,\ldots,k\}, implying DtJv0,β(τcon(p,v0))0D_{t}J_{v_{0},\beta}({\tau_{\text{con}}}(p,v_{0}))\neq 0 and Jv0,α(τcon(p,v0))0J_{v_{0},\alpha}({\tau_{\text{con}}}(p,v_{0}))\neq 0 for β{1,,k}\beta\in\{1,\ldots,k\} and α{k+1,,n1}\alpha\in\{k+1,\ldots,n-1\}. Moreover, the vectors

(22) DtJv0,1(τcon(p,v0)),,DtJv0,k(τcon(p,v0)),Jv0,k+1(τcon(p,v0)),Jv0,n1(τcon(p,v0))D_{t}J_{v_{0},1}({\tau_{\text{con}}}(p,v_{0})),\ldots,D_{t}J_{v_{0},k}({\tau_{\text{con}}}(p,v_{0})),J_{v_{0},k+1}({\tau_{\text{con}}}(p,v_{0})),\ldots J_{v_{0},n-1}({\tau_{\text{con}}}(p,v_{0}))

are linearly independent due to [25, Proposition 2.5.8 (ii)] and the properties of the geodesic flow on SNSN presented in [43, Lemma 1.40].

Since Jacobi fields are solutions of the second order ODE they depend smoothly on the coefficients of the respective equation. In particular, after choosing smaller UU if necessary, we can construct the family of Jacobi fields Jv,1(t),,Jv,n1(t)J_{v,1}(t),\ldots,J_{v,n-1}(t) along the geodesic γp,v(t)\gamma_{p,v}(t) that depend smoothly on vUv\in U, and span the vector space of all Jacobi fields along γp,v(t)\gamma_{p,v}(t) that vanishes at t=0t=0 and are normal to γ˙p,v(t)\dot{\gamma}_{p,v}(t). Therefore the function

f:U×[0,τcon(p,v0)+1]𝐑,f(v,t)=det(Jv,1(t),,Jv,n1(t)),f\colon U\times[0,{\tau_{\text{con}}}(p,v_{0})+1]\to{\bf R},\quad f(v,t)=\det(J_{v,1}(t),\ldots,J_{v,n-1}(t)),

is smooth and vanishes at (v,t)(v,t) if and only if γp,v(t)\gamma_{p,v}(t) is conjugate to pp.

We choose a parallel frame E1(t),,En1(t)E_{1}(t),\ldots,E_{n-1}(t) along γp,v0(t)\gamma_{p,v_{0}}(t) that is orthogonal to γ˙p,v0(t)\dot{\gamma}_{p,v_{0}}(t). With respect to this frame we write

Jv0,β(t)=jβα(t)Eα(t), for α,β1,,n1,J_{v_{0},\beta}(t)=j_{\beta}^{\alpha}(t)E_{\alpha}(t),\quad\hbox{ for }\alpha,\beta\in{1,\ldots,n-1},

for some some smooth functions jβα(t)j_{\beta}^{\alpha}(t). From here we get

jtjf(v0,τcon(p,v0))=jtj(σsign(σ)j1σ(1)(t)j2σ(2)(t)jn1σ(n1)(t))|t=τcon(p,v0)=0,\frac{\partial^{j}}{\partial t^{j}}f(v_{0},{\tau_{\text{con}}}(p,v_{0}))=\frac{\partial^{j}}{\partial t^{j}}\left(\sum_{\sigma}\text{sign}(\sigma)j_{1}^{\sigma(1)}(t)j_{2}^{\sigma(2)}(t)\cdots j_{n-1}^{\sigma(n-1)}(t)\right)\bigg{|}_{t={\tau_{\text{con}}}(p,v_{0})}=0,

for every j{0,,k1}\hbox{for every }j\in\{0,\ldots,k-1\}. Above, σ\sigma is a permutation of the set {1,,n1}\{1,\ldots,n-1\}. Moreover the covariant derivative of Jv0,βJ_{v_{0},\beta} along γp,v0\gamma_{p,v_{0}} is written as DtJv0,β(t)=(ddtjβα(t))Eα(t).D_{t}J_{v_{0},\beta}(t)=\left(\frac{\mathrm{d}}{\mathrm{d}t}j_{\beta}^{\alpha}(t)\right)E_{\alpha}(t).

If AA is the square matrix whose column vectors are given in the formula (22) we have

ktkf(v0,τcon(p,v0))=±k!det(A)0.\begin{split}&\frac{\partial^{k}}{\partial t^{k}}f(v_{0},{\tau_{\text{con}}}(p,v_{0}))=\pm k!\det(A)\neq 0.\end{split}

Since δ()\delta(\cdot) is constant kk in the set UU we have that k1tk1f(v,τcon(p,v))=0\frac{\partial^{k-1}}{\partial t^{k-1}}f(v,{\tau_{\text{con}}}(p,v))=0 for every vUv\in U. Therefore the Implicit function theorem gives that the conjugate distance function is smooth in some neighborhood VUV\subset U of v0v_{0}. Since v0Uv_{0}\in U was chosen arbitrarily the claim follows. ∎

Let v0SpNv_{0}\in S_{p}N be such that τcon(p,v0)<{\tau_{\text{con}}}(p,v_{0})<\infty. Then by lemmas 6 and 11 the function eq(v)=expp(τcon(p,v)v)Ne_{q}(v)=\exp_{p}({\tau_{\text{con}}}(p,v)v)\in N is well defined and continuous on some neighborhood USpNU\subset S_{p}N of v0v_{0}. Moreover, we have that

(23) Q(p)={eq(v)M:τcut(p,v)=τcon(p,v)}.Q(p)=\{e_{q}(v)\in M:\>{\tau_{\text{cut}}}(p,v)={\tau_{\text{con}}}(p,v)\}.

The following result is an adaptation of [19, Lemma 2].

Proposition 19.

Let Riemannian manifold (M,g)(M,g) be as in Theorem 12 and pMp\in M. The Hausdorff dimension of Q(p)Q(p) does not exceed n2n-2.

Proof.

By (23) we can write the conjugate cut locus Q(p)Q(p) as a disjoint union of the sets

A1={eq(v)M:τcut(p,v)=τcon(p,v),δ(v)=1}A_{1}=\{e_{q}(v)\in M:\>{\tau_{\text{cut}}}(p,v)={\tau_{\text{con}}}(p,v),\>\delta(v)=1\}

and

A2={eq(v)M:τcut(p,v)=τcon(p,v),δ(v)2}.A_{2}=\{e_{q}(v)\in M:\>{\tau_{\text{cut}}}(p,v)={\tau_{\text{con}}}(p,v),\>\delta(v)\geq 2\}.

To prove the claim of this proposition it suffices to show that

(24) A1{eq(v)N:dim(Deq(TvSpM))n2},A_{1}\subset\{e_{q}(v)\in N:\>\dim\left(\mathrm{D}e_{q}(T_{v}S_{p}M)\right)\leq n-2\},

since clearly we have that

A2{expp(w)N:wTpN,dim(Dexpp(Tw(TpN)))n2},A_{2}\subset\{\exp_{p}(w)\in N:\>w\in T_{p}N,\>\dim\left(\mathrm{D}\exp_{p}(T_{w}(T_{p}N))\right)\leq n-2\},

and therefore by the generalization of the classical Sard’s theorem [46] the Hausdorff dimension of Q(p)=A1A2Q(p)=A_{1}\cup A_{2} is at most n2n-2.

We choose v0SpNv_{0}\in S_{p}N such that eq(v0)A1e_{q}(v_{0})\in A_{1}. By the properties of the Jacobi fields normal to γp,v0\gamma_{p,v_{0}}, we can identify the kernel Dexpp(τcon(p,v0)v0)\mathrm{D}\exp_{p}({\tau_{\text{con}}}(p,v_{0})v_{0}) with some vector sub-space of Tv0SpM.T_{v_{0}}S_{p}M.

Since dimTv0SpM=n1\dim T_{v_{0}}S_{p}M=n-1 we can verify the inclusion (24) if we show that

(25) kerDexpp(τcon(p,v0)v0)kerDeq(v0).\ker\mathrm{D}\exp_{p}({\tau_{\text{con}}}(p,v_{0})v_{0})\subset\ker\mathrm{D}e_{q}(v_{0}).

Since δ(v0)=1\delta(v_{0})=1 we get by lemmas 16 and 18 that there exists a neighborhood USpNU\subset S_{p}N of v0v_{0} where the conjugate distance τcon(p,){\tau_{\text{con}}}(p,\cdot) and the map eq(v)=expp(τcon(p,v)v)e_{q}(v)=\exp_{p}({\tau_{\text{con}}}(p,v)v) are smooth. Let ξTv0SpM\xi\in T_{v_{0}}S_{p}M be in the kernel of the differential of the exponential map. Then by the chain and Leibniz rules we get

Deq(v0)ξ=γ˙p,v0(τcon(p,v0))Dτcon(p,v0)ξ.\begin{split}\mathrm{D}e_{q}(v_{0})\xi=\dot{\gamma}_{p,v_{0}}({\tau_{\text{con}}}(p,v_{0}))\mathrm{D}{\tau_{\text{con}}}(p,v_{0})\xi.\end{split}

Therefore Deq(v0)ξ=0\mathrm{D}e_{q}(v_{0})\xi=0 if and only if Dτcon(p,v0)ξ=0\mathrm{D}{\tau_{\text{con}}}(p,v_{0})\xi=0. So we suppose that Deq(v0)ξ0\mathrm{D}e_{q}(v_{0})\xi\neq 0.

By the Rank theorem [35, Theorem 4.12] we get that the subset of TpNT_{p}N, near τcon(p,v0)v0{\tau_{\text{con}}}(p,v_{0})v_{0}, where Dexpp\mathrm{D}\exp_{p} vanishes is diffeomorphic to a smooth sub-bundle of TUTSpNTU\subset TS_{p}N. Then we use the existence of the ODE theorem to choose a smooth curve v():(1,1)USpNv(\cdot)\colon(-1,1)\to U\subset S_{p}N such that v(0)=v0v(0)=v_{0}, v˙(0)=ξ\dot{v}(0)=\xi and v˙(t)kerDexpp(τcon(p,v(t))v(t))\dot{v}(t)\in\ker\mathrm{D}\exp_{p}({\tau_{\text{con}}}(p,v(t))v(t)) for every t(1,1)t\in(-1,1). Thus c(t):=eq(v(t))c(t):=e_{q}(v(t)) is a smooth curve in M^\hat{M} that satisfies

(26) c˙(t)=γ˙p,v(t)(τcon(p,v(t)))ddt(τcon(p,v(t))).\dot{c}(t)=\dot{\gamma}_{p,v(t)}({\tau_{\text{con}}}(p,v(t)))\frac{d}{dt}({\tau_{\text{con}}}(p,v(t))).

Since ddt(τcon(p,v(t0)))=Dτcon(p,v0)ξ\frac{d}{dt}({\tau_{\text{con}}}(p,v(t_{0})))=\mathrm{D}{\tau_{\text{con}}}(p,v_{0})\xi we can assume that ddt(τcon(p,v(t)))>0\frac{d}{dt}({\tau_{\text{con}}}(p,v(t)))>0 on some interval (ε,ε)(-\varepsilon,\varepsilon) for 0<ε<10<\varepsilon<1. Thus by equation (26) and the Fundamental theorem of calculus we get that the length of c(t)c(t) on [ε,0][-\varepsilon,0] is τcon(p,v0)τcon(p,v(ε)).{\tau_{\text{con}}}(p,v_{0})-{\tau_{\text{con}}}(p,v(-\varepsilon)). Below we denote the length of c(t)c(t) as (c)\mathcal{L}(c). From here by the assumption τcon(p,v0)=τcut(p,v0){\tau_{\text{con}}}(p,v_{0})={\tau_{\text{cut}}}(p,v_{0}) and the triangle inequality we get

τcon(p,v(ε))dM^(p,eq(v(ε)))dM^(p,eq(v0))(c)τcon(p,v(ε)),\begin{split}{\tau_{\text{con}}}(p,v(-\varepsilon))\geq&d_{\hat{M}}(p,e_{q}(v(-\varepsilon)))\geq d_{\hat{M}}(p,e_{q}(v_{0}))-\mathcal{L}(c)\geq{\tau_{\text{con}}}(p,v(-\varepsilon)),\end{split}

and the inequality above must hold as an equality. Therefore

dM^(p,eq(v(ε)))+(c)=dM^(p,eq(v0)),d_{\hat{M}}(p,e_{q}(v(-\varepsilon)))+\mathcal{L}(c)=d_{\hat{M}}(p,e_{q}(v_{0})),

and the curve c():[ε,0]M^c(\cdot)\colon[-\varepsilon,0]\to\hat{M} is part of some distance minimizing geodesic γ\gamma of M^{\hat{M}} from pp to eq(v(0))e_{q}(v(0)) that contains eq(v(ε))e_{q}(v(-\varepsilon)). Thus we have after some reparametrization t=t(s)t=t(s) that

γ(s)=eq(v(t(s)))=c(t(s))\gamma(s)=e_{q}(v(t(s)))=c(t(s))

for every t(s)(ε,0)t(s)\in(-\varepsilon,0). By (26) we get that γ\gamma is a parallel to γp,v(t)\gamma_{p,v(t)} for every t(ε,0)t\in(-\varepsilon,0). This is not possible unless the geodesics γp,v(t)\gamma_{p,v(t)} are all the same for every t(ε,0)t\in(-\varepsilon,0). Hence v(t)v(t) and c(t)c(t) are constant curves. This leads to a contradiction. The inclusion (25) is confirmed and the proof is complete. ∎

We are ready to present the proof of Theorem 12.

Proof of Theorem 12.

Let pMp\in M. In this proof we combine the observations made earlier in this section. The proofs of the four sub-claims are given below.

  • (1)

    By Lemma 15 we know that T(p)T(p) is a smooth hyper-surface of MM whose tangent space is normal to the vector field ν(q)=Y1(q)Y2(q)\nu(q)=Y_{1}(q)-Y_{2}(q) for qT(p)q\in T(p). Since Y1(q)Y2(q)Y_{1}(q)\neq Y_{2}(q) and Y1(q)g=Y2(q)g\|Y_{1}(q)\|_{g}=\|Y_{2}(q)\|_{g} we get from Cauchy-Schwarz inequality that

    Y1(q),ν(q)>0 and Y2(q),ν(q)<0.\langle Y_{1}(q),\nu(q)\rangle>0\quad\text{ and }\quad\langle Y_{2}(q),\nu(q)\rangle\rangle<0.

    Thus Y1(q)Y_{1}(q) and Y2(q)Y_{2}(q) hit T(p)T(p) from different sides. If qT(p)Mq\in T(p)\cap\partial M and these surfaces are tangential to each other at qq we arrive in a contradiction: Since ν(q)\nu(q) is normal to both T(T(p))T(T(p)) and TMT\partial M we can without loss of generality assume that Y2(q)Y_{2}(q) is inward pointing at qq. This is not possible since the geodesic related to Y2(q)Y_{2}(q), that connects pp to qq, is contained in MM. Thus by equation (14) Y2(q)Y_{2}(q) is also outward pointing which is not possible.

  • (2)

    By Proposition 19 we know that the Hausdorff dimension of the conjugate cut locus Q(p)Q(p) does not exceed n2n-2. If we can prove the same for the set L(p)L(p) of a-typical cut points the claim (2) follows from Lemma 13.

    Recall that L(p)(cut(p)Q(p))L(p)\subset({\text{cut}}(p)\setminus Q(p)) is the set of points in MM that can be connected to pp with more than two distance minimizing geodesics of MM. Let qL(p)q\in L(p) and define kp(q)𝐍k_{p}(q)\in\mathbf{N} to be the number of distance minimizing geodesics from pp to qq. Then we choose vectors w1,,wkp(q)Mpw_{1},\ldots,w_{k_{p}(q)}\in M_{p} and their respective neighborhoods UiTpNU_{i}\in T_{p}N such that for each i{1,,kp(q)}i\in\{1,\ldots,k_{p}(q)\}

    expp(wi)=q, and expp:UiV~\exp_{p}(w_{i})=q,\quad\hbox{ and }\exp_{p}\colon U_{i}\to\tilde{V}

    is a diffeomorphism on some open set V~N\tilde{V}\subset N. Let YiY_{i} for i{1,,kp(q)}i\in\{1,\ldots,k_{p}(q)\} be the distance vector fields related to the UiU_{i} and pp. Then we define a collection of smooth functions

    ρij:V~𝐑,ρij(x)=Yi(x)gYj(x)g,i,j{1,,kp(q)}.\rho_{ij}\colon\tilde{V}\to{\bf R},\quad\rho_{ij}(x)=\|Y_{i}(x)\|_{g}-\|Y_{j}(x)\|_{g},\quad i,j\in\{1,\ldots,k_{p}(q)\}.

    By the proof of Lemma 15 it holds that the sets Kij:=ρij1{0}, for i<jK_{ij}:=\rho_{ij}^{-1}\{0\},\>\hbox{ for }i<j are smooth hyper-surfaces of NN that contain qq. Also by [42, Proposition 2.6] it holds that the sets

    Ki,j,k:=KikKjk, for i<j<kK_{i,j,k}:=K_{ik}\cap K_{jk},\quad\hbox{ for }i<j<k

    are smooth submanifolds of NN of co-dimension two. Next we set K(q):=i<j<kKi,j,kK(q):=\bigcup_{i<j<k}K_{i,j,k} and claim that

    (27) L(p)V~=K(q)M.L(p)\cap\tilde{V}=K(q)\cap M.

    Since the sets Ki,j,kK_{i,j,k} are smooth sub-manifolds of dimension n2n-2 their Hausdorff dimension is also n2n-2. Thus the equation (27) and Lemma 13 imply that Hausdorff dimension of L(p)L(p) does not exceed n2n-2.

    Finally we verify the equation (27). If xL(p)V~x\in L(p)\cap\tilde{V} it holds there are at least three distance minimizing geodesics of MM connecting pp to xx. Thus there are 1i<j<kkp(q)1\leq i<j<k\leq k_{p}(q) so that

    Yi(x)g=Yj(x)g=Yk(x)g=dM(p,x),\|Y_{i}(x)\|_{g}=\|Y_{j}(x)\|_{g}=\|Y_{k}(x)\|_{g}=d_{M}(p,x),

    which yields

    ρik(x)=ρjk(x)=0,and xKi,j,kK(q).\rho_{ik}(x)=\rho_{jk}(x)=0,\quad\hbox{and $x\in K_{i,j,k}\subset K(q)$}.

    If xK(q)Mx\in K(q)\cap M then xKi,j,kMx\in K_{i,j,k}\cap M for some i<j<ki<j<k. Thus by the proof of Lemma 15 it holds that there are at least three distance minimizing geodesics of MM connecting pp to xx. Therefore xL(p)V~x\in L(p)\cap\tilde{V}.

  • (3)

    Since we can write the cut locus of pp as a disjoint union cut(p)=T(p)L(p)Q(p){\text{cut}}(p)=T(p)\cup L(p)\cup Q(p) the parts (1) and (2) in conjunction with Lemma 13 yield the claim of part (3).

  • (4)

    Since M\partial M is a smooth hyper-surface of a nn-dimensional Riemannian manifold MM we have by part (1) that T(p)MT(p)\cap\partial M is a smooth sub-manifold of dimension n2n-2, thus it has the Hausdorff-dimension n2n-2. Also by part (3) we know that the Hausdorff dimension of L(p)Q(p)L(p)\cap Q(p) does not exceed n2n-2. We have proven that the Hausdorff dimension of the closed set cut(p)M{\text{cut}}(p)\cap\partial M does not exceed n2n-2. Since the boundary of MM has the Hausdorff-dimension n1n-1 it follows that Mcut(p)\partial M\setminus{\text{cut}}(p) is open and dense in M\partial M. The density claim follows from the observation that by Lemma 13 the set cut(p)M{\text{cut}}(p)\cap\partial M cannot contain any open subsets of M\partial M as their Hausdorff dimension is n1n-1.

We are ready to prove Theorem 4.

Proof of Theorem 4.

The proof follows from Proposition 10 and Theorem 12. ∎

4. Reconstruction of the manifold

4.1. Geometry of the measurement region

In this section we consider only one Riemannian manifold (M,g)(M,g) that satisfies the assumptions of Theorem 4 and whose partial travel time data (1) is known. Let ν(z)\nu(z) be the outward pointing unit normal vector field at zMz\in\partial M. The inward pointing bundle at the boundary is the set

inTM={(z,v)TM|zM,v,ν(z)g<0}.\partial_{in}TM=\{(z,v)\in TM~{}|~{}z\in\partial M,\langle v,\nu(z)\rangle_{g}<0\}.

We restrict our attention to the vectors that are inward pointing and of unit length: inSM={(z,v)inTM:vg=1}\partial_{in}SM=\{(z,v)\in\partial_{in}TM~{}:~{}\|v\|_{g}=1\}. We emphasize that this set or its restriction on the open measurement region ΓM\Gamma\subset\partial M is not a priori given by the data (1). Our first task is to recover a diffeomorphic copy of this set. We consider the orthogonal projection

(28) h:inSMTM,h(z,v)=vv,ν(z)gν(z),h:\partial_{in}SM\to T\partial M,\qquad h(z,v)=v-\langle v,\nu(z)\rangle_{g}\nu(z),

and denote the set, that contains the image of hh, as P(M)={(z,w)TM:wg<1}P(\partial M)=\{(z,w)\in T\partial M~{}:~{}\|w\|_{g}<1\}. It is straightforward to show that the map hh is a diffeomorphism onto P(M)P(\partial M).

For the rest of this section we will be considering the vectors in P(M)P(\partial M), and with a slight abuse of notation, each vector (z,v)P(M)(z,v)\in P(\partial M) represents an inward-pointing unit vector at zz. In the next lemma we show that the data (1) determines the restriction of P(M)P(\partial M) on Γ\Gamma.

Lemma 20.

Let Riemannian manifold (M,g)(M,g) be as in Theorem 4. The first fundamental form g|Γg|_{\Gamma} of Γ\Gamma and the set

P(Γ)={(z,v)TM:zΓ,vg<1}P(\Gamma)=\{(z,v)\in T\partial M:\>z\in\Gamma,\>\|v\|_{g}<1\}

can be recovered from data (1).

Proof.

Let (z,v)TΓ(z,v)\in T\Gamma. We choose a smooth curve c:(1,1)Γc\colon(-1,1)\to\Gamma for which c(0)=z,c(0)=z, and c˙(0)=v\dot{c}(0)=v. Since the boundary M\partial M of MM is strictly convex the inverse function of the exponential map expz\exp_{z} is smooth and well defined near zz on MM. In addition, we have that

rz(c(t))=d(z,c(t))=expz1(c(t))g.r_{z}(c(t))=d(z,c(t))=\|\exp^{-1}_{z}(c(t))\|_{g}.

We set c~(t)=expz1(c(t))TzM.\tilde{c}(t)=\exp^{-1}_{z}(c(t))\in T_{z}M. As the differential of the exponential map at the origin is an identity operator we get c~(0)=0, and c~˙(0)=vTzM.\tilde{c}(0)=0,\text{ and }\dot{\tilde{c}}(0)=v\in T_{z}M. From here the continuity of the norm yields

(29) limt0rz(c(t))|t|=limt0c~(0)c~(t)tg=c~˙(0)g=vg.\lim_{t\to 0}\frac{r_{z}(c(t))}{|t|}=\lim_{t\to 0}\left\|\frac{\tilde{c}(0)-\tilde{c}(t)}{t}\right\|_{g}=\left\|\dot{\tilde{c}}(0)\right\|_{g}=\|v\|_{g}.

By the data (1) and the choice of the path c(t)Γc(t)\in\Gamma we know the left hand side of equation (29). Therefore we have recovered the length of an arbitrary vector (z,v)TΓ(z,v)\in T\Gamma. Moreover, the set P(Γ)P(\Gamma) is recovered.

Since we know the unit sphere {vTzM:vg=1}\{v\in T_{z}\partial M:\>\|v\|_{g}=1\} for each zΓz\in\Gamma the reconstruction of the first fundamental form of Γ\Gamma can be carried out as explained in the next lemma. ∎

Lemma 21.

Let (X,g)(X,g) be a finite dimensional inner product space. Let a>0a>0 and S(a):={vX:vg=a}.S(a):=\{v\in X:\>\|v\|_{g}=a\}. Then any open subset UU of S(a)S(a) determines the inner product gg on XX.

Proof.

This proof is the same as the one in [23, Lemma 3.33] and thus omitted here. ∎

Let p0Mp_{0}\in M. By Theorem 4 we can find a boundary point z0Γz_{0}\in\Gamma and neighborhoods Up0U_{p_{0}} and Vp0V_{p_{0}} for p0p_{0} and z0z_{0} respectively such that the distance function d(,)d(\cdot,\cdot) is smooth in the product set Up0×Vp0U_{p_{0}}\times V_{p_{0}}. For each zΓVp0z\in\Gamma\cap V_{p_{0}} we let γz\gamma_{z} be the unique distance minimizing unit speed geodesic from p0p_{0} to zz. If we decompose the velocity of the geodesic γz\gamma_{z} at rp0(z)r_{p_{0}}(z) into its tangential and normal components to the boundary, then the tangential component coincides with the boundary gradient of the travel time function rp0r_{p_{0}} at zz. For this vector field we use the notation gradMrp0(z)Pz(Γ).\operatorname{grad}_{\partial M}r_{p_{0}}(z)\in P_{z}(\Gamma). Furthermore, by Lemma 20 we have recovered the metric tensor of the measurement domain ΓM\Gamma\subset\partial M. Thus we can compute gradMrp0(z)\operatorname{grad}_{\partial M}r_{p_{0}}(z) whenever the respective travel time function rp0r_{p_{0}} is differentiable on Γ\Gamma.

4.2. Topological reconstruction

We first show that the data (1) separates the points in the manifold MM.

Lemma 22.

Let (M,g)(M,g) be as in Theorem 4. Let ΓM\Gamma\subset\partial M be open and p1,p2Mp_{1},p_{2}\in M be such that rp1(z)=rp2(z)r_{p_{1}}(z)=r_{p_{2}}(z) for all zΓz\in\Gamma, then p1=p2p_{1}=p_{2}.

Proof.

First we choose open and dense subsets W1,W2MW_{1},W_{2}\subset\partial M for the points p1p_{1} and p2p_{2} as we have for the point p0p_{0} in Theorem 4. Then we choose any point z0Wp1Wp2Γz_{0}\in W_{p_{1}}\cap W_{p_{2}}\cap\Gamma, neighborhoods Up1U_{p_{1}} of p1p_{1}, Up2U_{p_{2}} of p2p_{2} and Vp1V_{p_{1}}, Vp2V_{p_{2}} of z0z_{0} as we have for p0p_{0} in Theorem 4. Thus the distance function d(,)d(\cdot,\cdot) is smooth in the product sets Up1×VU_{p_{1}}\times V and Up2×VU_{p_{2}}\times V, where V=Vp1Vp2V=V_{p_{1}}\cap V_{p_{2}} is an open neighborhood of z0z_{0}. Moreover for each (p,z)Upi×V,i{1,2}(p,z)\in U_{p_{i}}\times V,\>i\in\{1,2\} there exists a unique distance minimizing geodesic of MM connecting pp to zz.

If γi\gamma_{i} is the distance minimizing geodesic from pip_{i} to z0z_{0} for i={1,2}i=\{1,2\} then by the discussion preceding this lemma we have that gradMrpi(z0)\operatorname{grad}_{\partial M}r_{p_{i}}(z_{0}) represents the tangential component of γ˙i\dot{\gamma}_{i} at rpi(z0)r_{p_{i}}(z_{0}). Since rp1=rp2r_{p_{1}}=r_{p_{2}} the tangential components of γ˙1\dot{\gamma}_{1} and γ˙2\dot{\gamma}_{2} are the same. Since the velocity vectors of γ˙i\dot{\gamma}_{i} at rpi(z0)r_{p_{i}}(z_{0}) have unit length, they must also coincide. We get

z0=γ1(rp1(z0))=γ2(rp2(z0)) and γ˙1(rp1(z0))=γ˙2(rp2(z0)).z_{0}=\gamma_{1}(r_{p_{1}}(z_{0}))=\gamma_{2}(r_{p_{2}}(z_{0}))\quad\text{ and }\quad\dot{\gamma}_{1}(r_{p_{1}}(z_{0}))=\dot{\gamma}_{2}(r_{p_{2}}(z_{0})).

Thus the geodesics γ1\gamma_{1} and γ2\gamma_{2} agree and we have p1=p2p_{1}=p_{2}. ∎

We are now ready to reconstruct the topological structure of (M,g)(M,g) from the partial travel time data (1). Let B(Γ)B(\Gamma) be the collection of all bounded functions f:Γ𝐑f\colon\Gamma\to{\bf R} and \|\cdot\|_{\infty} the supremum norm of B(Γ)B(\Gamma). Thus (B(Γ),)(B(\Gamma),\|\cdot\|_{\infty}) is a Banach space. Since (M,g)(M,g) is a compact Riemannian manifold each travel time function rpr_{p}, for pMp\in M, is bounded by the diameter of MM, which is finite. Thus

{rp=d(p,):Γ[0,)|pM}B(Γ),\{r_{p}=d(p,\cdot)\colon\Gamma\to[0,\infty)|~{}p\in M\}\subset B(\Gamma),

and the map

(30) R:(M,g)(B(Γ),),R(p)=rpR:(M,g)\to(B(\Gamma),\|\cdot\|_{\infty}),\quad R(p)=r_{p}

is well defined.

Proposition 23.

Let Riemannian manifold (M,g)(M,g) be as in Theorem 4. The map RR as in (30) is a topological embedding.

Proof.

By Lemma 22, we know that the map RR is injective, and by the triangle inequality we get that it is also continuous. Let KK be a closed set in MM. Since MM is a compact Hausdorff space the set KK is compact. Since the image of a compact set under a continuous mapping is compact, it follows that R(K)R(K) is closed. This makes RR a closed map and thus a topological embedding. ∎

4.3. Boundary Determination

We recall that the data (1) only gives us the subset Γ\Gamma of the boundary, and we do not know yet if the travel time function rR(M)r\in R(M) is related to an interior or a boundary point of MM. In this subsection we will use the data (1) to determine the boundary of the unknown manifold MM as a point set. However, due to Proposition 23 we may assume without loss of generality that the topology of MM is known. Also the set P(Γ)P(\Gamma), as in Lemma 20, is known to us.

Let (z,v)P(Γ)(z,v)\in P(\Gamma), and define the set,

(31) σ(z,v)={pM|the point p has a neighborhood UM such that,rq:Γ𝐑 is smooth near z for every qU,qgradMrq(z) is continuous in U,gradMrp(z)=v}{z},\begin{split}\sigma(z,v)=&\{p\in M~{}|~{}\text{the point $p$ has a neighborhood $U\subset M$ such that,}\\ &\>r_{q}\colon\Gamma\to{\bf R}\text{ is smooth near $z$ for every $q\in U$},\\ &q\mapsto\operatorname{grad}_{\partial M}r_{q}(z)\text{ is continuous in $U$},\\ &\operatorname{grad}_{\partial M}r_{p}(z)=-v\}\cup\{z\},\end{split}

where gradMrp(z)\operatorname{grad}_{\partial M}r_{p}(z) is the boundary gradient of rpr_{p} at zΓz\in\Gamma. We recall that by Proposition 23 we know the topology of MM, and by Lemma 20 we know the geometry of Γ\Gamma. These in conjunction with the data (1) imply that we can recover the set σ(z,v)\sigma(z,v) for every (z,v)P(Γ)(z,v)\in P(\Gamma). In the next lemma we generalize the result [32, Lemma 2.9] and relate σ(z,v)\sigma(z,v) to the maximal distance minimizing segment of the geodesic γz,v\gamma_{z,v}.

Lemma 24.

Let (z,v)P(Γ)(z,v)\in P(\Gamma) then σ(z,v)¯=γz,v([0,τcut(z,v)])\overline{\sigma(z,v)}=\gamma_{z,v}([0,{\tau_{\text{cut}}}(z,v)]).

Proof.

Let t[0,τcut(z,v))t\in[0,{\tau_{\text{cut}}}(z,v)) and define y:=γz,v(t)y:=\gamma_{z,v}(t). Thus yy is not in the cut locus of zz (see equation (10)). By Proposition 10 there exist neighborhoods U,VMU,V\subset M of yy and zz respectively, having the property that the distance function d(,)d(\cdot,\cdot) is smooth in the product set U×VU\times V. Therefore the function rq()=d(q,)|Γr_{q}(\cdot)=d(q,\cdot)|_{\Gamma} is smooth near zz for any qUq\in U. Furthermore, gradMrp(z)=v\operatorname{grad}_{\partial M}r_{p}(z)=-v, and the function pgradMrp(z)p\mapsto\operatorname{grad}_{\partial M}r_{p}(z) is continuous in UU. Therefore yy is in σ(z,v)\sigma(z,v) and the inclusion γz,v([0,τcut(z,v)))σ(z,v)\gamma_{z,v}([0,{\tau_{\text{cut}}}(z,v)))\subseteq\sigma(z,v) is true. This gives γz,v([0,τcut(z,v)])σ(z,v)¯\gamma_{z,v}([0,{\tau_{\text{cut}}}(z,v)])\subseteq\overline{\sigma(z,v)}.

Let pσ(z,v)p\in\sigma(z,v), then rp(z)r_{p}(z) is smooth in a neighborhood of zz and gradMrp(z)=v\operatorname{grad}_{\partial M}r_{p}(z)=-v. Thus γz,v\gamma_{z,v} is the unique distance minimizing geodesic connecting zz to pp. Since the geodesic γz,v\gamma_{z,v} is not distance minimizing beyond the interval [0,τcut(z,v)][0,{\tau_{\text{cut}}}(z,v)] we have pγz,v([0,τcut(z,v)])p\in\gamma_{z,v}([0,{\tau_{\text{cut}}}(z,v)]) and therefore σ(z,v)¯γz,v([0,τcut(z,v)]).\overline{\sigma(z,v)}\subset\gamma_{z,v}([0,{\tau_{\text{cut}}}(z,v)]).

We set

(32) Tz,v:=suppσ(z,v)rp(z)=suppσ(z,v)d(p,z), for (z,v)P(Γ).T_{z,v}:=\sup_{p\in\sigma(z,v)}r_{p}(z)=\sup_{p\in\sigma(z,v)}d(p,z),\qquad\text{ for }(z,v)\in P(\Gamma).

Notice that this number is determined entirely by the data (1), as opposed to τcut(z,v){\tau_{\text{cut}}}(z,v) which requires our knowledge of when the geodesics were distance minimizing. By the following corollary, whose proof is evident, these two numbers are the same.

Corollary 25.

For any (z,v)P(Γ)(z,v)\in P(\Gamma) we have that Tz,v=τcut(z,v)T_{z,v}={\tau_{\text{cut}}}(z,v).

We will use the sets σ(z,v)\sigma(z,v), for (z,v)P(Γ)(z,v)\in P(\Gamma) to determine the boundary M\partial M of MM. Since the topology of MM is known by Proposition 23, we can determine the topology of these σ\sigma sets from the data. The next lemma shows if σ(z,v)\sigma(z,v) is closed then γz,v(Tz,v)\gamma_{z,v}(T_{z,v}) is on the boundary of MM.

Lemma 26.

Let (z,v)P(Γ)(z,v)\in P(\Gamma). If σ(z,v)\sigma(z,v) is closed then Tz,v=τexit(z,v)T_{z,v}={\tau_{\text{exit}}}(z,v).

Proof.

By the definition of Tz,vT_{z,v} we must have Tz,vτexit(z,v)T_{z,v}\leq{\tau_{\text{exit}}}(z,v).

Suppose that Tz,v<τexit(z,v)T_{z,v}<{\tau_{\text{exit}}}(z,v). From Corollary 25 then we also know

(33) τcut(z,v)=Tz,v<τexit(z,v){\tau_{\text{cut}}}(z,v)=T_{z,v}<{\tau_{\text{exit}}}(z,v)

Let p=γz,v(Tz,v)p=\gamma_{z,v}(T_{z,v}), and by Lemma 24 it holds that pσ(z,v)¯=σ(z,v)p\in\overline{\sigma(z,v)}=\sigma(z,v). Since pcut(z)p\in{\text{cut}}(z) we have by Lemma 5, that there either exists a second distance minimizing geodesic from zz to pp or pp is a conjugate point to zz along γz,v\gamma_{z,v}. In the first case let wP(Γ)w\in P(\Gamma) such that γz,w\gamma_{z,w} is another unit-speed distance minimizing geodesic from zz to pp. We note that Tz,v=τcut(z,w)T_{z,v}={\tau_{\text{cut}}}(z,w).

Let UU be a neighborhood of pp as in (31). We consider a sequence ti[0,Tz,v],i𝐍t_{i}\in[0,T_{z,v}],\>i\in\mathbf{N} such that tiTz,vt_{i}\to T_{z,v} as ii\to\infty. Then for sufficiently large ii the points pi=γz,v(ti)p_{i}=\gamma_{z,v}(t_{i}) and qi=γz,w(ti)q_{i}=\gamma_{z,w}(t_{i}) are in UU and converge to pp. By the continuity of the boundary gradient in UU we have gradMrpi(z)gradMrp(z)\operatorname{grad}_{\partial M}r_{p_{i}}(z)\to\operatorname{grad}_{\partial M}r_{p}(z) and gradMrqi(z)gradMrp(z)\operatorname{grad}_{\partial M}r_{q_{i}}(z)\to\operatorname{grad}_{\partial M}r_{p}(z), when ii\to\infty. However, by construction gradMrpi(z)=v\operatorname{grad}_{\partial M}r_{p_{i}}(z)=-v while gradMrqi(z)=w\operatorname{grad}_{\partial M}r_{q_{i}}(z)=-w for all i𝐍i\in\mathbf{N}. Thus gradMrp(z)\operatorname{grad}_{\partial M}r_{p}(z) has multiple values, and rpr_{p} is not differentiable at zz, contradicting that pσ(z,v)p\in\sigma(z,v).

If the second case is valid, and since pMintp\in M^{int}, we get by a similar proof as in [25, Theorem 2.1.12] that the exponential map expz\exp_{z} is not a local injection at Tz,vvTzMT_{z,v}v\in T_{z}M. From here [25, Theorem 2.1.14] implies that there is a sequence of points (pi)i=1(p_{i})_{i=1}^{\infty} in MintM^{int}, that converges to pp and can be connected to zz by at least two distance minimizing geodesics. By the same argument as in the previous case, rpir_{p_{i}} is not differentiable at zz for any i𝐍i\in\mathbf{N}, which contradicts the fact that pσ(z,v)p\in\sigma(z,v). Thus inequality (33) cannot occur and we must have Tz,v=τexit(z,v)T_{z,v}={\tau_{\text{exit}}}(z,v). ∎

Lemma 27.

Let p0Mp_{0}\in\partial M and z0Γz_{0}\in\Gamma, Up0U_{p_{0}}, and Vp0V_{p_{0}} be as in Theorem 4. For every pUp0p\in U_{p_{0}} we denote η(p)=gradMrp(z0)\eta(p)=-\operatorname{grad}_{\partial M}r_{p}(z_{0}). There exists a neighborhood Up0Up0U_{p_{0}}^{\prime}\subseteq U_{p_{0}} of p0p_{0} such that for all pUp0p\in U_{p_{0}}^{\prime} we have that pp is in the closed set σ(z0,η(p))\sigma(z_{0},\eta(p)).

Proof.

By these assumptions, d(,)d(\cdot,\cdot) in Up0×Vp0U_{p_{0}}\times V_{p_{0}} is smooth. Define v0=η(p0)v_{0}=\eta(p_{0}) and t0=τexit(z0,v0)t_{0}={\tau_{\text{exit}}}(z_{0},v_{0}), then p0=expz0(t0v0)p_{0}=\exp_{z_{0}}(t_{0}v_{0}). Since z0z_{0} was chosen to be a point outside the cut locus of p0p_{0}, these points are not conjugate to each other along the geodesic γz0,v0\gamma_{z_{0},v_{0}} connecting them. Therefore the differential Dexpz0\mathrm{D}\exp_{z_{0}} of the exponential map is invertible at t0v0Tz0Mt_{0}v_{0}\in T_{z_{0}}M. From here the claim follows from the Inverse function theorem for expz0\exp_{z_{0}} near t0v0t_{0}v_{0}, the continuity of the exit time function on the non-trapping part of SMSM, and the inequality

rp(z0)=expz01(p)qτexit(z0,η(p)),for pUp0.r_{p}(z_{0})=\|\exp_{z_{0}}^{-1}(p)\|_{q}\leq{\tau_{\text{exit}}}(z_{0},\eta(p)),\quad\text{for }p\in U_{p_{0}}.

We omit the further details.

Corollary 28.

Let p0Mp_{0}\in\partial M, z0Γz_{0}\in\Gamma and Up0U_{p_{0}}^{\prime} be as in Lemma 27. If we denote η(p)=gradMrp(z0)\eta(p)=-\operatorname{grad}_{\partial M}r_{p}(z_{0}) then Tz0,η(p)T_{z_{0},\eta(p)} is smooth for all pUp0p\in U_{p_{0}}^{\prime}.

Proof.

Since the exit time function is smooth on those (z,v)inSM(z,v)\in\partial_{in}SM that satisfy τexit(z,v)<{\tau_{\text{exit}}}(z,v)<\infty we only need to show that Tz,η(p)=τexit(z0,η(p)).T_{z,\eta(p)}={\tau_{\text{exit}}}(z_{0},\eta(p)). This equation follows from lemmas 26 and 27. ∎

We are now ready to determine the boundary of MM from the data (1).

Proposition 29.

Let (M,g)(M,g) be as in Theorem 4 and p0Mp_{0}\in M. Then p0Mp_{0}\in\partial M if and only if there exists (z,v)P(Γ)(z,v)\in P(\Gamma) such that p0σ(z,v)p_{0}\in\sigma(z,v) and rp0(z)=Tz,vr_{p_{0}}(z)=T_{z,v}.

Proof.

If p0Mp_{0}\in\partial M then we get from Lemma 27 that there exists (z0,v)P(Γ)(z_{0},v)\in P(\Gamma) such that p0p_{0} is in the closed set σ(z0,v)\sigma(z_{0},v). By Lemma 26 we have Tz0,v=τexit(z0,v)T_{z_{0},v}={\tau_{\text{exit}}}(z_{0},v). Firstly the strict convexity of M\partial M implies that each geodesic has at most two boundary points. Secondly since p0z0p_{0}\neq z_{0} are both boundary points contained in σ(z0,v)\sigma(z_{0},v), which is a trace of a distance minimising geodesic, it follows that Tz0,v=rp0(z0).T_{z_{0},v}=r_{p_{0}}(z_{0}).

To show the reverse direction, let (z,v)P(Γ)(z,v)\in P(\Gamma) be such that p0σ(z,v)p_{0}\in\sigma(z,v) and Tz,v=rp0(z)T_{z,v}=r_{p_{0}}(z). Thus γz,v([0,rp0(z)])σ(z,v),\gamma_{z,v}([0,r_{p_{0}}(z)])\subseteq\sigma(z,v), and it follows from Lemma 24 and Corollary 25 that σ(z,v)\sigma(z,v) is closed. By Lemma 26, the closedness of σ(z,v)\sigma(z,v) implies Tz,v=τexit(z,v)T_{z,v}={\tau_{\text{exit}}}(z,v). Thus, rp0(z)=τexit(z,v)r_{p_{0}}(z)={\tau_{\text{exit}}}(z,v), making p0Mp_{0}\in\partial M. ∎

4.4. Local Coordinates

By Proposition 29 we have reconstructed the boundary M\partial M of the smooth manifold MM. In this section we use the partial travel time data (1) to construct two local coordinate systems for p0Mp_{0}\in M. Since MM has a boundary, we need different coordinates systems based on whether p0Mintp_{0}\in M^{int} or p0Mp_{0}\in\partial M.

Proposition 30.

Let (M,g)(M,g) be as in Theorem 4. Let p0Mintp_{0}\in M^{int}, and choose z0Γz_{0}\in\Gamma, Up0U_{p_{0}}, and Vp0V_{p_{0}} as in Theorem 4. Let the map α:Up0Pz0(Γ)×𝐑\alpha:U_{p_{0}}\to P_{z_{0}}(\Gamma)\times{\bf R} be defined as

(34) α(p)=(gradMrp(z0),rp(z0)).\alpha(p)=(-\operatorname{grad}_{\partial M}r_{p}(z_{0}),r_{p}(z_{0})).

This map is a diffeomorphism onto its image α(Up0)Pz0(Γ)×𝐑\alpha(U_{p_{0}})\subset P_{z_{0}}(\Gamma)\times{\bf R}.

Proof.

Since the distance function d(,)d(\cdot,\cdot) is smooth in Up0×Vp0U_{p_{0}}\times V_{p_{0}} also the function α\alpha is smooth on Up0U_{p_{0}}. By a direct computation we see that the inverse function of α\alpha, is given as,

α1(v,t)=expz0(th1(v)), for (v,t)Pz0(Γ)×𝐑.\alpha^{-1}(v,t)=\exp_{z_{0}}\left(th^{-1}(v)\right),\qquad\text{ for }(v,t)\in P_{z_{0}}(\Gamma)\times{\bf R}.

where h:inSz0MPz0(Γ)h\colon\partial_{in}S_{z_{0}}M\to P_{z_{0}}(\Gamma), is the orthogonal projection given in (28). By the smoothness of h1h^{-1} and the exponential map, it follows that α1\alpha^{-1} is smooth. Thus, α\alpha is a diffeomorphism onto its image, which is open in Pz0(Γ)×𝐑P_{z_{0}}(\Gamma)\times{\bf R}. ∎

In particular, the function α\alpha, in (34), gives a local coordinate system near the interior point p0p_{0}. In order to define a coordinate system for a point at the boundary we will adjust the last coordinate function of α\alpha to be a boundary defining function.

Proposition 31.

Let (M,g)(M,g) be as in Theorem 4. Let p0Mp_{0}\in\partial M and choose z0Γz_{0}\in\Gamma, and Up0U_{p_{0}}^{\prime} as in Lemma 27. Let η(p):=gradMrp(z0)\eta(p):=-\operatorname{grad}_{\partial M}r_{p}(z_{0}) and βz0:Up0Pz0(Γ)×[0,)\beta_{z_{0}}:U_{p_{0}}^{\prime}\to P_{z_{0}}(\Gamma)\times[0,\infty) be defined as

(35) β(p)=(η(p),Tz0,η(p)rp(z0)).\beta(p)=(\eta(p),T_{z_{0},\eta(p)}-r_{p}(z_{0})).

This map is a diffeomorphism onto its image β(Up0)Pz0(Γ)×[0,)\beta(U_{p_{0}}^{\prime})\subset P_{z_{0}}(\Gamma)\times[0,\infty).

Proof.

Since the distance function d(,)d(\cdot,\cdot) is smooth in Up0×Vp0U_{p_{0}}^{\prime}\times V_{p_{0}} and p0Mp_{0}\in\partial M we have by Corollary 28, that the map Tz0,η(p)T_{z_{0},\eta(p)} is smooth for all pUp0p\in U_{p_{0}}^{\prime}. Thus β\beta is smooth in Up0U_{p_{0}}^{\prime}. Again by a direct computation we get that the inverse function of β\beta is given as

β1(v,t)=expz0((Tz0,vt)h1(v)) for (v,t)Pz0(Γ)×[0,).\beta^{-1}(v,t)=\exp_{z_{0}}\left(\left(T_{z_{0},v}-t\right)h^{-1}(v)\right)\qquad\text{ for }(v,t)\in P_{z_{0}}(\Gamma)\times[0,\infty).

By the local invertibility of the exponential map expz0\exp_{z_{0}} at rp0(z0)h1(η(p0))Tz0Mr_{p_{0}}(z_{0})h^{-1}(\eta(p_{0}))\in T_{z_{0}}M and the equation rp(z0)=expz01(p)gr_{p}(z_{0})=\|\exp_{z_{0}}^{-1}(p)\|_{g} for pUp0p\in U_{p_{0}}^{\prime}, the set η(Up0)Pz0(Γ)\eta(U_{p_{0}}^{\prime})\subset P_{z_{0}}(\Gamma) is open and the function vTz0,vv\mapsto T_{z_{0},v}, in this set is smooth, making β1\beta^{-1} smooth. Thus, β\beta is a diffeomorphism onto its image, which is open in Pz0(Γ)×[0,)P_{z_{0}}(\Gamma)\times[0,\infty).

Finally by Proposition 26 we get that Tz0,η(p)rp(z0)=0T_{z_{0},\eta(p)}-r_{p}(z_{0})=0 if and only if pUp0Mp\in U_{p_{0}}^{\prime}\cap\partial M. Thus this function defines the boundary. ∎

Combining the results of Propositions 30 and 31, we know that for p0Mp_{0}\in M, either the function α\alpha as in (34) or the function β\beta as in (35), gives a smooth local coordinate system. Moreover these maps can be recovered fully from the data (1). As these two types of coordinate charts cover MM the smooth structure on MM is then the same as the maximal smooth atlas determined by these coordinate charts [36, Proposition 1.17].

4.5. Reconstruction of the Riemannian Metric

So far we recovered both the topological and smooth structures of the Riemannian manifold (M,g)(M,g) from the data (1). In this section we recover the Riemannian metric gg. We recall that by Lemma 20 we know the first fundamental form of Γ\Gamma.

In order to recover the metric on MM we consider the distance function

d(p,z)=rp(z),for (p,z)M×Γ,d(p,z)=r_{p}(z),\quad\text{for }(p,z)\in M\times\Gamma,

which we have recovered by Proposition 23. Let p0Mp_{0}\in M. By Theorem 4 we can choose z0Γz_{0}\in\Gamma and neighborhoods Up0U_{p_{0}} and Vp0V_{p_{0}} for p0p_{0} and z0z_{0} respectively such that the distance function d(p,z)d(p,z) for (p,z)Up0×Vp0(p,z)\in U_{p_{0}}\times V_{p_{0}} is smooth. Thus the map

(36) Hp0:Vp0ΓTp0M,Hp0(z)=Dd(p0,z)H_{p_{0}}\colon V_{p_{0}}\cap\Gamma\to T^{\ast}_{p_{0}}M,\quad H_{p_{0}}(z)=\mathrm{D}d(p_{0},z)

is well defined and smooth. Here D\mathrm{D} stands for the differential of the distance function d(p,z)d(p,z) with respect to the pp variable in the open set Up0MU_{p_{0}}\subset M and Tp0MT^{\ast}_{p_{0}}M is the cotangent space at p0p_{0}. As we have recovered the smooth structure of MM we can find Hp0H_{p_{0}}.

For zVp0z\in V_{p_{0}} the gradient gradpd(p,z)\text{grad}_{p}d(p,z) for pUp0p\in U_{p_{0}} is the velocity of the distance minimizing unit speed geodesic from zz to pp (see for instance [36, theorems 6.31, 6.32]). In particular the map

H~p0:(Vp0Γ)zgradpd(p0,z)Sp0M\tilde{H}_{p_{0}}\colon(V_{p_{0}}\cap\Gamma)\ni z\to\text{grad}_{p}d(p_{0},z)\in S_{p_{0}}M

is well defined and satisfies H~p0(z)=Hp0(z),\tilde{H}_{p_{0}}(z)=H_{p_{0}}(z)^{\sharp}, where :Tp0MTp0M\sharp\colon T^{\ast}_{p_{0}}M\to T_{p_{0}}M is the musical isomorphism, raising the indices, given in any local coordinates near p0p_{0} as (ξ)i=gij(p0)ξj.(\xi^{\sharp})^{i}=g^{ij}(p_{0})\xi_{j}. Note that the inverse of \sharp is given by :Tp0MTp0M\flat:T_{p_{0}}M\to T_{p_{0}}^{*}M, that lowers the indices. Although we know the map Hp0H_{p_{0}}, we do not know its sister map H~p0\tilde{H}_{p_{0}}.

Lemma 32.

Let p0Mp_{0}\in M. Let z0Γz_{0}\in\Gamma, Up0U_{p_{0}} and Vp0V_{p_{0}} be as in Theorem 4. Then the image of the map Hp0H_{p_{0}}, as in (36), is contained the unit co-sphere

Sp0M:={ξTp0M:ξg1=1},S^{\ast}_{p_{0}}M:=\{\xi\in T_{p_{0}}^{\ast}M:\>\|\xi\|_{g^{-1}}=1\},

and has a nonempty interior.

Proof.

Let vTp0Mv\in T_{p_{0}}M such that expp0(v)=z0\exp_{p_{0}}(v)=z_{0}. Since :Tp0MTp0M\flat\colon T_{p_{0}}M\to T^{\ast}_{p_{0}}M is a linear isomorphism that preserves the inner product, the claim holds due to the local invertibility of the exponential map expp\exp_{p} near vv, the equality d(p0,z)=expp01(z)gd(p_{0},z)=\|\exp_{p_{0}}^{-1}(z)\|_{g}, which is true for all zVp0z\in V_{p_{0}}, and the continuity of the exit time function near vvgSp0M\frac{v}{\|v\|_{g}}\in S_{p_{0}}M. We omit the further details. ∎

Finally Lemma 21 in conjunction with the previous lemma lets us recover the inverse metric g1(p0)g^{-1}(p_{0}) and thus the metric gp0g_{p_{0}}. This is formalized in the proposition below.

Proposition 33.

Let (M,g)(M,g) be as in Theorem 4 and p0Mp_{0}\in M. The data (1) determines the metric tensor gg near p0p_{0} in the local coordinates given in Propositions 30 and 31.

Proof.

Let z0Γz_{0}\in\Gamma, Up0U_{p_{0}} and Vp0V_{p_{0}} be as in Theorem 4. By Proposition 29 we can tell whether p0p_{0} is an interior or a boundary point. Based on this we choose local coordinates of p0p_{0} as in Proposition 30 or as in Proposition 31. Then we consider the function Hp0H_{p_{0}} given in the equation (36). By Lemma 32 we know that image of the function Hp0H_{p_{0}} contains an open subset of Sp0MS_{p_{0}}^{*}M.

From here, by applying Lemma 21 we determine the inverse metric gij(p0)g^{ij}(p_{0}) in the aforementioned coordinates. Finally taking the inverse of gij(p0)g^{ij}(p_{0}) determines gij(p0)g_{ij}(p_{0}). As this procedure can be done for any point pMp\in M, which is close enough to p0p_{0}, we have recovered the metric gg near p0p_{0} in the appropriate local coordinates. ∎

5. The proof of Theorem 2

Let Riemannian manifolds (M1,g1)(M_{1},g_{1}) and (M2,g2)(M_{2},g_{2}) be as in Theorem 2. We recall that the partial travel time data of these manifolds coincide in the sense of Definition 1. Let (B(Γi),)(B(\Gamma_{i}),\|\cdot\|_{\infty}), for i{1,2}i\in\{1,2\}, be the Banach space of bounded real valued functions on Γi\Gamma_{i}. We set a mapping

(37) F:B(Γ1)B(Γ2),F(f)=fϕ1,F:B(\Gamma_{1})\to B(\Gamma_{2}),\qquad F(f)=f\circ\phi^{-1},

where ϕ\phi is the diffeomorphism from Γ1\Gamma_{1} to Γ2\Gamma_{2}. By the triangle inequality we have that FF is a metric isometry whose inverse mapping is given by F1(h)=hϕ.F^{-1}(h)=h\circ\phi. Taking Ri:(Mi,gi)(B(Γi),)R_{i}:(M_{i},g_{i})\to(B(\Gamma_{i}),\|\cdot\|_{\infty}), as in the equation (30), we have by the equation (2) in Definition 1 that

F(R1(M1))=R2(M2).F(R_{1}(M_{1}))=R_{2}(M_{2}).

Therefore we get from Proposition 23 that the map

(38) Ψ:(M1,g1)R1(B(Γ1),)𝐹(B(Γ2),)R21(M2,g2),\Psi:(M_{1},g_{1})\xrightarrow{R_{1}}(B(\Gamma_{1}),\|\cdot\|_{\infty})\xrightarrow{F}(B(\Gamma_{2}),\|\cdot\|_{\infty})\xrightarrow{R_{2}^{-1}}(M_{2},g_{2}),

is a well defined homeomorphism, that satisfies the equation

(39) d2(Ψ(x),ϕ(z))=F(d1(x,))(ϕ(z))=d1(x,z), for all (x,z)M1×Γ1.d_{2}(\Psi(x),\phi(z))=F(d_{1}(x,\cdot))(\phi(z))=d_{1}(x,z),\quad\text{ for all }(x,z)\in M_{1}\times\Gamma_{1}.

Here di(,)d_{i}(\cdot,\cdot) is the distance function of (Mi,gi)(M_{i},g_{i}). The goal of this section is to show that Ψ\Psi is a Riemannian isometry. In the following lemma we show first that ϕ\phi preserves the Riemannian structure of the measurement regions.

Lemma 34.

Let Riemannian manifolds (M1,g1)(M_{1},g_{1}) and (M2,g2)(M_{2},g_{2}) be as in Theorem 2. Then Ψ|Γ1=ϕ\Psi|_{\Gamma_{1}}=\phi and ϕ:(Γ1,g1)(Γ2,g2)\phi\colon(\Gamma_{1},g_{1})\to(\Gamma_{2},g_{2}) is a Riemannian isometry.

Proof.

Let z1z_{1} be in Γ1\Gamma_{1}. From equation (39) we get d2(Ψ(z1),ϕ(z1))=0.d_{2}(\Psi(z_{1}),\phi(z_{1}))=0. Thus Ψ(z1)=ϕ(z1)\Psi(z_{1})=\phi(z_{1}) and we have verified the first claim Ψ|Γ1=ϕ\Psi|_{\Gamma_{1}}=\phi. It follows from the proof of Lemma 20 and equation (39) that Dϕvg2=vg1, for all vTΓ1.\|\mathrm{D}\phi v\|_{g_{2}}=\|v\|_{g_{1}},\text{ for all }v\in T\Gamma_{1}. Then the polarization identity implies that the differential Dϕ\mathrm{D}\phi of ϕ\phi also preserves the first fundamental forms:

Dϕv1,Dϕv2g2=v1,v2g1, for all v1,v2TΓ1,\langle\mathrm{D}\phi v_{1},\mathrm{D}\phi v_{2}\rangle_{g_{2}}=\langle v_{1},v_{2}\rangle_{g_{1}},\quad\text{ for all }v_{1},v_{2}\in T\Gamma_{1},

making ϕ\phi a Riemannian isometry. ∎

In particular we get from this lemma that Dϕ(P(Γ1))=P(Γ2).\mathrm{D}\phi(P(\Gamma_{1}))=P(\Gamma_{2}). Next we show that the mapping Ψ\Psi takes the boundary of M1M_{1} onto the boundary of M2M_{2}. In light of Proposition 29 we need to understand how this map carries over the sets σ(z,v)\sigma(z,v), as in (31). The following lemma gives an answer to this question.

Lemma 35.

Let Riemannian manifolds (M1,g1)(M_{1},g_{1}) and (M2,g2)(M_{2},g_{2}) be as in Theorem 2. If (z0,v)P(Γ1)(z_{0},v)\in P(\Gamma_{1}) then Ψ(σ(z0,v))=σ(ϕ(z0),Dϕv)\Psi(\sigma(z_{0},v))=\sigma(\phi(z_{0}),\mathrm{D}\phi v).

Proof.

Clearly we have that Ψ(z0)=ϕ(z0)σ(ϕ(z0),Dϕv)\Psi(z_{0})=\phi(z_{0})\in\sigma(\phi(z_{0}),\mathrm{D}\phi v). So suppose that p0σ(z0,v){z0}p_{0}\in\sigma(z_{0},v)\setminus\{z_{0}\}. Hence, by the same argument as in the proof of Lemma 26, we get that z0z_{0} is not in the cut-locus of p0p_{0}. Thus by Proposition 10 we can choose a neighborhood U×VM1×M1U\times V\subset M_{1}\times M_{1} of (p0,z0)(p_{0},z_{0}) where the distance function d1(,)d_{1}(\cdot,\cdot) is smooth. Since the map Ψ\Psi is a homeomorphism the set Ψ(U)M2\Psi(U)\subset M_{2} is open, and we have by (39) that for each qΨ(U)q\in\Psi(U), the function rq()=d1(Ψ1(q),ϕ1())r_{q}(\cdot)=d_{1}(\Psi^{-1}(q),\phi^{-1}(\cdot)) is smooth on the open set ϕ(VΓ1)Γ2\phi(V\cap\Gamma_{1})\subset\Gamma_{2}.

Since ϕ:Γ1Γ2\phi\colon\Gamma_{1}\to\Gamma_{2} is a Riemannian isometry we have that

(40) gradM2rΨ(p)(ϕ(z0))=Dϕ(z0)gradM1rp(z0),for pU.\operatorname{grad}_{\partial M_{2}}r_{\Psi(p)}(\phi(z_{0}))=\mathrm{D}\phi(z_{0})\operatorname{grad}_{\partial M_{1}}r_{p}(z_{0}),\quad\text{for }p\in U.

Here D\mathrm{D} stands for the differential, gradM1\operatorname{grad}_{\partial M_{1}} for the boundary gradient of Γ1\Gamma_{1} and gradM2\operatorname{grad}_{\partial M_{2}} for that of Γ2\Gamma_{2}. Since the right hand side of equation (40) is continuous in pp, the function

q=Ψ(p)gradM2rq(ϕ(z0))q=\Psi(p)\mapsto\operatorname{grad}_{\partial M_{2}}r_{q}(\phi(z_{0}))

is continuous in Ψ(U)\Psi(U). Finally

gradM2rΨ(p0)(ϕ(z0))=Dϕ(z0)gradM1rp0(z0)=Dϕ(z0)v\operatorname{grad}_{\partial M_{2}}r_{\Psi(p_{0})}(\phi(z_{0}))=\mathrm{D}\phi(z_{0})\operatorname{grad}_{\partial M_{1}}r_{p_{0}}(z_{0})=-\mathrm{D}\phi(z_{0})v

implies Ψ(p0)σ(ϕ(z0),Dϕv).\Psi(p_{0})\in\sigma(\phi(z_{0}),\mathrm{D}\phi v).

On the other hand after reversing the roles of M1M_{1} and M2M_{2} we can use the same proof to show σ(z0,v)Ψ1(σ(ϕ(z0),Dϕv))\sigma(z_{0},v)\supset\Psi^{-1}(\sigma(\phi(z_{0}),\mathrm{D}\phi v)), implying Ψ(σ(z0,v))=σ(ϕ(z0),Dϕv).\Psi(\sigma(z_{0},v))=\sigma(\phi(z_{0}),\mathrm{D}\phi v). This ends the proof. ∎

Lemma 36.

Let Riemannian manifolds (M1,g1)(M_{1},g_{1}) and (M2,g2)(M_{2},g_{2}) be as in Theorem 2. Then Ψ(M1)=M2\Psi(\partial M_{1})=\partial M_{2}. Moreover, Ψ(M1int)=M2int\Psi(M_{1}^{int})=M_{2}^{int}.

Proof.

Let pM1p\in\partial M_{1}. Due to Proposition 29 there is a (z,v)P(Γ1)(z,v)\in P(\Gamma_{1}) such that pp is in the closed set σ(z,v)\sigma(z,v) and rp1(z)=Tz,vr_{p_{1}}(z)=T_{z,v}. Thus Lemma 35 gives Ψ(σ(z,v))=σ(ϕ(z),Dϕv),\Psi(\sigma(z,v))=\sigma(\phi(z),\mathrm{D}\phi v), and since Ψ\Psi is a homeomorphism, also the set σ(ϕ(z),Dϕv)\sigma(\phi(z),\mathrm{D}\phi v) is closed and contains Ψ(p)\Psi(p). Furthermore, by equation (39) we have that rΨ(q)(ϕ(z))=rq(z),for all qσ(z,v).r_{\Psi(q)}(\phi(z))=r_{q}(z),\>\text{for all }\>q\in\sigma(z,v). Therefore

Tϕ(z),Dϕv=Tz,v=rp(z)=rΨ(p)(ϕ(z)).T_{\phi(z),\mathrm{D}\phi v}=T_{z,v}=r_{p}(z)=r_{\Psi(p)}(\phi(z)).

From here Proposition 29 implies that Ψ(p)\Psi(p) is in M2\partial M_{2}. Thus Ψ(M1)M2\Psi(\partial M_{1})\subset\partial M_{2} and by using the same argument for Ψ1\Psi^{-1} it follows that Ψ(M1)=M2\Psi(\partial M_{1})=\partial M_{2}. Since M1intM^{int}_{1} and M1\partial M_{1} are disjoint and Ψ\Psi is a bijection we also have that Ψ(M1int)=M2int\Psi(M^{int}_{1})=M^{int}_{2}. ∎

Lemma 37.

Let Riemannian manifolds (M1,g1)(M_{1},g_{1}) and (M2,g2)(M_{2},g_{2}) be as in Theorem 2. The mapping Ψ:M1M2\Psi:M_{1}\to M_{2}, given in formula (38), is a diffeomorphism.

Proof.

Let p0M1p_{0}\in M_{1}, and choose Wp0M1W_{p_{0}}\subset\partial M_{1} as in Theorem 4. Since ϕ:Γ1Γ2\phi\colon\Gamma_{1}\to\Gamma_{2} is a diffeomorphism, the set ϕ(Wp0Γ1)\phi(W_{p_{0}}\cap\Gamma_{1}) is open and dense in Γ2\Gamma_{2}. Then for Ψ(p0)M2\Psi(p_{0})\in M_{2} we choose WΨ(p0)M2W_{\Psi(p_{0})}\subset\partial M_{2} as in Theorem 4 and consider the non-empty open set WΨ(p0)ϕ(Wp0Γ1)Γ2W_{\Psi(p_{0})}\cap\phi(W_{p_{0}}\cap\Gamma_{1})\subset\Gamma_{2}. We pick z0Wp0Γ1z_{0}\in W_{p_{0}}\cap\Gamma_{1} such that ϕ(z0)WΨ(p0)ϕ(Wp0Γ1)\phi(z_{0})\in W_{\Psi(p_{0})}\cap\phi(W_{p_{0}}\cap\Gamma_{1}).

Let neighborhoods Up0M1U_{p_{0}}\subset M_{1} of p0p_{0} and Vp0M1V_{p_{0}}\subset M_{1} of z0z_{0} be such that the distance function d1(,)d_{1}(\cdot,\cdot) is smooth in the product set Up0×Vp0U_{p_{0}}\times V_{p_{0}}. We also choose neighborhoods UΨ(p0)M2U_{\Psi(p_{0})}\subset M_{2} of Ψ(p0)\Psi(p_{0}) and VΨ(p0)M2V_{\Psi(p_{0})}\subset M_{2} of ϕ(z0)=Ψ(z0)\phi(z_{0})=\Psi(z_{0}) to be such that the distance function d2(,)d_{2}(\cdot,\cdot) is smooth in the product set UΨ(p0)×VΨ(p0)U_{\Psi(p_{0})}\times V_{\Psi(p_{0})}. Since Ψ:M1M2\Psi\colon M_{1}\to M_{2} is a homeomorphism we may choose these four sets in such a way that they satisfy

Ψ(Up0)=UΨ(p0), and Ψ(Vp0)=VΨ(p0).\Psi(U_{p_{0}})=U_{\Psi(p_{0})},\quad\text{ and }\quad\Psi(V_{p_{0}})=V_{\Psi(p_{0})}.

By Lemma 36 we know that Ψ(p0)M2int\Psi(p_{0})\in M_{2}^{int} if and only if p0M1intp_{0}\in M_{1}^{int}, and Ψ(p0)M2\Psi(p_{0})\in\partial M_{2} if and only if p0M1p_{0}\in\partial M_{1}. Next we consider the interior and boundary cases separately.

Suppose first that p0p_{0} is an interior point of M1M_{1}. The functions

Up0pα1(p)=(gradM1rp(z0),rp(z0))Pz0(Γ1)×𝐑,U_{p_{0}}\ni p\mapsto\alpha_{1}(p)=(-\operatorname{grad}_{\partial M_{1}}r_{p}(z_{0}),r_{p}(z_{0}))\in P_{z_{0}}(\Gamma_{1})\times{\bf R},

and

UΨ(p0)qα2(q)=(gradM2rq(ϕ(z0)),rq(ϕ(z0)))Pϕ(z0)(Γ2)×𝐑,U_{\Psi(p_{0})}\ni q\mapsto\alpha_{2}(q)=(-\operatorname{grad}_{\partial M_{2}}r_{q}(\phi(z_{0})),r_{q}(\phi(z_{0})))\in P_{\phi(z_{0})}(\Gamma_{2})\times{\bf R},

as in Proposition 30, are smooth local coordinate maps of M1M_{1} and M2M_{2} respectively. Moreover, by the computations done in the proof of Lemma 35 we get for every pUp0p\in U_{p_{0}} that

rp(z0)=rΨ(p)(ϕ(z0)),andDϕ(z0)gradM1rp(z0)=gradM2rΨ(p)(ϕ(z0)).r_{p}(z_{0})=r_{\Psi(p)}(\phi(z_{0})),\quad\text{and}\quad\mathrm{D}\phi(z_{0})\operatorname{grad}_{\partial M_{1}}r_{p}(z_{0})=\operatorname{grad}_{\partial M_{2}}r_{\Psi(p)}(\phi(z_{0})).

Therefore for any (v,t)α1(Up0)(v,t)\in\alpha_{1}(U_{p_{0}}) we have that

(α2Ψα11)(v,t)=(Dϕ(z0)v,t).(\alpha_{2}\circ\Psi\circ\alpha^{-1}_{1})(v,t)=(\mathrm{D}\phi(z_{0})v,t).

Thus we have proven that the map α2Ψα11:α1(Up0)α2(UΨ(p0))\alpha_{2}\circ\Psi\circ\alpha^{-1}_{1}\colon\alpha_{1}(U_{p_{0}})\to\alpha_{2}(U_{\Psi(p_{0})}) is smooth.

Then we let p0p_{0} be a boundary point of M1M_{1}. Let η1(p):=gradM1rp(z0)\eta_{1}(p):=-\operatorname{grad}_{\partial M_{1}}r_{p}(z_{0}) for pUp0p\in U_{p_{0}} and choose Up0Up0U_{p_{0}}^{\prime}\subset U_{p_{0}} as in Lemma 27 to be such that the set σ(z0,η1(p))\sigma(z_{0},\eta_{1}(p)) is closed and the function pTz0,η1(p)p\mapsto T_{z_{0},\eta_{1}(p)} is smooth for every pUp0p\in U_{p_{0}}^{\prime}. Let UΨ(p0):=Ψ(Up0)UΨ(p0)U^{\prime}_{\Psi(p_{0})}:=\Psi(U_{p_{0}}^{\prime})\subset U_{\Psi(p_{0})} and denote η2(q):=gradM2rq(ϕ(z0))\eta_{2}(q):=-\operatorname{grad}_{\partial M_{2}}r_{q}(\phi(z_{0})) for qUΨ(p0)q\in U^{\prime}_{\Psi(p_{0})}. Since we have that Dϕ(z0)η1(p)=η2(Ψ(p))\mathrm{D}\phi(z_{0})\eta_{1}(p)=\eta_{2}(\Psi(p)) it holds by Lemma 35 that the set σ(ϕ(z0),η2(Ψ(p)))=Ψ(σ(z0,η1(p))),\sigma(\phi(z_{0}),\eta_{2}(\Psi(p)))=\Psi(\sigma(z_{0},\eta_{1}(p))), is closed for every pUp0p\in U_{p_{0}}^{\prime}, and thus the function UΨ(p0)qTϕ(z0),η2(q)U^{\prime}_{\Psi(p_{0})}\ni q\to T_{\phi(z_{0}),\eta_{2}(q)} is smooth by Corollary 25. Moreover, we have Tz0,η1(p)=Tϕ(z0),η2(Ψ(p))T_{z_{0},\eta_{1}(p)}=T_{\phi(z_{0}),\eta_{2}(\Psi(p))} for every pUp0p\in U_{p_{0}}^{\prime}.

Then we consider local coordinate maps

Up0pβ1(p)=(η1(p),Tz0,η1(p)rp(z0))Pz0(Γ1)×[0,),U_{p_{0}}^{\prime}\ni p\mapsto\beta_{1}(p)=(\eta_{1}(p),T_{z_{0},\eta_{1}(p)}-r_{p}(z_{0}))\in P_{z_{0}}(\Gamma_{1})\times[0,\infty),

of M1M_{1} and

UΨ(p0)qβ2(q)=(η2(q),Tϕ(z0),η2(q)rq(ϕ(z0)))Pϕ(z0)(Γ2)×[0,),U^{\prime}_{\Psi(p_{0})}\ni q\mapsto\beta_{2}(q)=(\eta_{2}(q),T_{\phi(z_{0}),\eta_{2}(q)}-r_{q}(\phi(z_{0})))\in P_{\phi(z_{0})}(\Gamma_{2})\times[0,\infty),

of M2M_{2}, as in Proposition 31. By the discussion above we have for any (v,t)β1(Up0)(v,t)\in\beta_{1}(U_{p_{0}}^{\prime}) that

(β2Ψβ11)(v,t)=(Dϕ(z0)v,t),(\beta_{2}\circ\Psi\circ\beta_{1}^{-1})(v,t)=(\mathrm{D}\phi(z_{0})v,t),

which implies that the map (β2Ψβ11):β1(Up0)β2(UΨ(p0))(\beta_{2}\circ\Psi\circ\beta_{1}^{-1})\colon\beta_{1}(U_{p_{0}}^{\prime})\to\beta_{2}(U^{\prime}_{\Psi(p_{0})}) is smooth.

By combining these two cases we have proved that for every p0Mp_{0}\in M a local representation of the map Ψ\Psi is smooth, making Ψ:M1M2\Psi\colon M_{1}\to M_{2} smooth. Finally by an analogous argument for Ψ1\Psi^{-1} we can show that this map is also smooth. Thus Ψ:M1M2\Psi\colon M_{1}\to M_{2} is a diffeomorphism as claimed. ∎

We are ready to present the proof of our main inverse problem:

Proof of Theorem 2.

By Lemma 37 we know that the map Ψ:M1M2\Psi\colon M_{1}\to M_{2} is a diffeomorphism. We define a metric tensor g~2\tilde{g}_{2} on M1M_{1} as the pull back of the metric g2g_{2} with respect to map Ψ\Psi. Thus it suffices to consider a smooth manifold M=M1M=M_{1} with an open measurement region Γ=Γ1M\Gamma=\Gamma_{1}\subset\partial M and two Riemannian metrics g1g_{1} and g2~\tilde{g_{2}}. Moreover M\partial M is strictly convex with respect to both of these metrics.

Let d~2(,)\tilde{d}_{2}(\cdot,\cdot) be the distance function of g2~\tilde{g_{2}}. We note that due to equation (39) we have d1(p,z)=d~2(p,z),for all (p,z)M×Γ.d_{1}(p,z)=\tilde{d}_{2}(p,z),\>\text{for all }(p,z)\in M\times\Gamma. By Lemma 34 we get that g1(p)=g~2(p)g_{1}(p)=\tilde{g}_{2}(p) for all pΓp\in\Gamma. Let p0Mp_{0}\in M. Thus the map Hp0H_{p_{0}} given by (36) is the same for both metrics. From here Lemma 32 and Proposition 33 imply that g1(p0)=g~2(p0)g_{1}(p_{0})=\tilde{g}_{2}(p_{0}). Therefore map Ψ\Psi is a Riemannian isometry as claimed. ∎

References

  • [1] R. Alexander and S. Alexander. Geodesics in Riemannian manifolds-with-boundary. Indiana University Mathematics Journal, 30(4):481–488, 1981.
  • [2] Y. M. Assylbekov and H. Zhou. Boundary and scattering rigidity problems in the presence of a magnetic field and a potential. Inverse Problems & Imaging, 9(4):935–950, 2015.
  • [3] R. Bartolo, E. Caponio, A. V. Germinario, and M. Sánchez. Convex domains of Finsler and Riemannian manifolds. Calculus of Variations and Partial Differential Equations, 40(3-4):335–356, 2011.
  • [4] M. I. Belishev and Y. V. Kurylev. To the reconstruction of a Riemannian manifold via its spectral data (Bc–Method). Communications in partial differential equations, 17(5–6):767–804, 1992.
  • [5] R. L. Bishop. Infinitesimal convexity implies local convexity. Indiana Univ. Math. J, 24(169-172):75, 1974.
  • [6] D. Burago, Y. Burago, and S. Ivanov. A course in metric geometry, volume 33. American Mathematical Soc., 2001.
  • [7] D. Burago and S. Ivanov. Boundary rigidity and filling volume minimality of metrics close to a flat one. Annals of mathematics, pages 1183–1211, 2010.
  • [8] V. Cerveny. Seismic ray theory. Cambridge university press, 2005.
  • [9] C. B. Croke et al. Rigidity and the distance between boundary points. Journal of Differential Geometry, 33(2):445–464, 1991.
  • [10] N. S. Dairbekov, G. P. Paternain, P. Stefanov, and G. Uhlmann. The boundary rigidity problem in the presence of a magnetic field. Advances in mathematics, 216(2):535–609, 2007.
  • [11] M. V. de Hoop, J. Ilmavirta, M. Lassas, and T. Saksala. Stable reconstruction of simple Riemannian manifolds from unknown interior sources. arXiv preprint arXiv:2102.11799, 2021.
  • [12] M. V. de Hoop, J. Ilmavirta, M. Lassas, and T. Saksala. A foliated and reversible Finsler manifold is determined by its broken scattering relation. Pure and Applied Analysis, 3(4):789–811, 2022.
  • [13] M. V. de Hoop, J. Ilmavirta, M. Lassas, and T. Saksala. Determination of a compact Finsler manifold from its boundary distance map and an inverse problem in elasticity. Communications in Analysis and Geometry, arXiv:1901.03902, to be appear, arXiv preprint 2019.
  • [14] M. V. de Hoop and T. Saksala. Inverse problem of travel time difference functions on a compact Riemannian manifold with boundary. The Journal of Geometric Analysis, 29(4):3308–3327, 2019.
  • [15] M. P. do Carmo. Riemannian geometry. Birkhäuser, 1992.
  • [16] J. J. Duistermaat and L. Hörmander. Fourier integral operators. ii. Acta mathematica, 128:183–269, 1972.
  • [17] J. J. Duistermaat and L. Hörmander. Fourier integral operators, volume 2. Springer, 1996.
  • [18] A. Greenleaf and G. Uhlmann. Recovering singularities of a potential from singularities of scattering data. Communications in mathematical physics, 157(3):549–572, 1993.
  • [19] J.-I. Itoh and M. Tanaka. The dimension of a cut locus on a smooth Riemannian manifold. Tohoku Mathematical Journal, Second Series, 50(4):571–575, 1998.
  • [20] S. Ivanov. Local monotonicity of Riemannian and Finsler volume with respect to boundary distances. Geometriae Dedicata, 164(1):83–96, 2013.
  • [21] S. Ivanov. Distance difference functions on non-convex boundaries of Riemannian manifolds, 2020. arXiv:2008.13153.
  • [22] S. Ivanov. Distance difference representations of Riemannian manifolds. Geometriae Dedicata, 207(1):167–192, 2020.
  • [23] A. Katchalov, Y. Kurylev, and M. Lassas. Inverse boundary spectral problems, volume 123 of Monographs and Surveys in Pure and Applied Mathematics. Chapman & Hall/CRC, Boca Raton, FL, 2001.
  • [24] A. Katsuda, Y. Kurylev, and M. Lassas. Stability of boundary distance representation and reconstruction of Riemannian manifolds. Inverse Problems & Imaging, 1(1):135, 2007.
  • [25] W. Klingenberg. Riemannian geometry, volume 1. Walter de Gruyter, 1982.
  • [26] Y. Kurylev. Multidimensional Gel’fand inverse problem and boundary distance map. Inverse Problems Related with Geometry (ed. H. Soga), pages 1–15, 1997.
  • [27] Y. Kurylev, M. Lassas, and G. Uhlmann. Rigidity of broken geodesic flow and inverse problems. American journal of mathematics, 132(2):529–562, 2010.
  • [28] Y. Kurylev, M. Lassas, and G. Uhlmann. Inverse problems for Lorentzian manifolds and non-linear hyperbolic equations. Inventiones mathematicae, 212(3):781–857, 2018.
  • [29] Y. Kurylev, L. Oksanen, and G. P. Paternain. Inverse problems for the connection Laplacian. Journal of Differential Geometry, 110(3):457–494, 2018.
  • [30] M. Lassas. Inverse problems for linear and non-linear hyperbolic equations. Proceedings of International Congress of Mathematicians – 2018 Rio de Janeiro, 3:3739–3760, 2018.
  • [31] M. Lassas and L. Oksanen. Inverse problem for the Riemannian wave equation with Dirichlet data and Neumann data on disjoint sets. Duke Mathematical Journal, 163(6):1071–1103, 2014.
  • [32] M. Lassas and T. Saksala. Determination of a Riemannian manifold from the distance difference functions. Asian journal of mathematics, 23(2):173–200, 2019.
  • [33] M. Lassas, T. Saksala, and H. Zhou. Reconstruction of a compact manifold from the scattering data of internal sources. Inverse Problems & Imaging, 12(4), 2018.
  • [34] M. Lassas, G. Uhlmann, and Y. Wang. Inverse problems for semilinear wave equations on Lorentzian manifolds. Communications in Mathematical Physics, pages 1–55, 2018.
  • [35] J. M. Lee. Introduction to Smooth Manifolds. Springer, 2013.
  • [36] J. M. Lee. Introduction to Riemannian manifolds. Springer, 2018.
  • [37] P. Mattila. Geometry of sets and measures in Euclidean spaces: fractals and rectifiability. Cambridge university press, 1999.
  • [38] R. Meyerson. Stitching data: Recovering a manifold’s geometry from geodesic intersections. The Journal of Geometric Analysis, 32(3):1–22, 2022.
  • [39] R. Michel. Sur la rigidité imposée par la longueur des géodésiques. Inventiones mathematicae, 65(1):71–83, 1981.
  • [40] T. Milne and A.-R. Mansouri. Codomain rigidity of the Dirichlet to Neumann operator for the Riemannian wave equation. Trans. Amer. Math. Soc., 371:8781–8810, 2019.
  • [41] K. Mönkkönen. Boundary rigidity for Randers metrics. arXiv:2010.11484, 2020.
  • [42] V. Ozols. Cut loci in Riemannian manifolds. Tohoku Mathematical Journal, Second Series, 26(2):219–227, 1974.
  • [43] G. P. Paternain. Geodesic flows, volume 180. Springer Science & Business Media, 2012.
  • [44] L. Pestov and G. Uhlmann. Two dimensional compact simple Riemannian manifolds are boundary distance rigid. Annals of mathematics, pages 1093–1110, 2005.
  • [45] P. Petersen. Riemannian geometry, volume 171. Springer, 2006.
  • [46] A. Sard. Hausdorff measure of critical images on Banach manifolds. American Journal of Mathematics, 87(1):158–174, 1965.
  • [47] V. A. Sharafutdinov. Integral geometry of tensor fields, volume 1. Walter de Gruyter, 1994.
  • [48] P. Stefanov, G. Uhlmann, and A. Vasy. Boundary rigidity with partial data. Journal of the American Mathematical Society, 29(2):299–332, 2016.
  • [49] P. Stefanov, G. Uhlmann, and A. Vasy. Local and global boundary rigidity and the geodesic x-ray transform in the normal gauge. Annals of Mathematics, 194(1):1–95, 2021.
  • [50] G. Uhlmann. Inverse boundary value problems for partial differential equations. In Proceedings of the International Congress of Mathematicians, Berlin, pages 77–86, 1998.
  • [51] Y. Wang and T. Zhou. Inverse problems for quadratic derivative nonlinear wave equations. Communications in Partial Differential Equations, 44(11):1140–1158, 2019.
  • [52] H. Whitney. Complex analytic varieties, volume 131. Addison-Wesley Reading, 1972.