Skip to main content
Springer logoLink to Springer
. 2025 Apr 10;11(2):24. doi: 10.1007/s40879-025-00816-x

Coarse extrinsic curvature of Riemannian submanifolds

Marc Arnaudon 1, Xue-Mei Li 2,3, Benedikt Petko 2,
PMCID: PMC11985601  PMID: 40225791

Abstract

We introduce a novel concept of coarse extrinsic curvature for Riemannian submanifolds, inspired by Ollivier’s notion of coarse Ricci curvature. This curvature is derived from the Wasserstein 1-distance between probability measures supported in the tubular neighborhood of a submanifold, providing new insights into the extrinsic curvature of isometrically embedded manifolds in Euclidean spaces. The framework also offers a method to approximate the mean curvature from statistical data, such as point clouds generated by a Poisson point process. This approach has potential applications in manifold learning and the study of metric embeddings, enabling the inference of geometric information from empirical data.

Keywords: Coarse curvature, Extrinsic curvature, Optimal transport

Introduction

Synthetic lower bounds on Ricci curvature is a powerful tool in the study of classical geometric analysis and metric measure spaces. Ollivier’s notion of Coarse Ricci Curvature is distinct in that it approximates the curvature itself, rather than merely providing a lower bound. By selecting as test measures weighted localized volume measures, supported on a ball of radius ε, the Wasserstein 1-distance between two such measures reveals the generalized Ricci tensor; applying to random geometric graphs sampled from a Poisson point process with non-uniform intensity leads to similar conclusions [2].

Inspired by the concept of coarse Ricci curvature, in this article we seek a suitable notion of extrinsic curvature for embedded manifolds. With manifold learning applications in mind, we initially work with curves and surfaces and subsequently define a concept of coarse extrinsic curvature for general embedded manifolds. This notion captures the inner product between the mean curvature and the second fundamental form in a principal curvature direction. It may prove useful for studying embedded metric spaces and could be relevant in manifold learning contexts.

Let M be a smooth manifold isometrically embedded in another Riemannian manifold. We propose a family of test measures Inline graphic, where σ,ε are small parameters, whose ‘derivative’ in the 1-Wasserstein distance with respect to variation of the point x describes some kind of curvature.

This consideration leads to a novel concept of coarse curvature in the setting of Riemannian submanifolds. Within the applicable range of the parameters, we have an approximation of the mean curvature and the second fundamental form, providing a valuable tool for evaluating these extrinsic curvatures. In more practical applications, we can take test measures built from statistical data and simulations; for instance through the empirical measures of point cloud samples. There is scope for extending to metric embeddings of metric spaces.

In contrast to the intrinsic Riemannian curvature, which characterizes the geometry of a manifold independently of its embedding, the second fundamental form of submanifolds is an extrinsic concept. It provides a means for describing the shape of a submanifold in relation to its ambient space, offering views into its bending properties. For instance, a surface embedded in R3 is locally isometric to a plane if and only if its second fundamental form vanishes.

The extrinsic curvature of M, isometrically embedded in N, is expressed by the second fundamental form, which we recall to be defined as the bilinear form

graphic file with name 40879_2025_816_Equ1_HTML.gif 1.1

where W is an arbitrary vector field on M with W(x)=w. Letting m denote the dimension of M, the mean curvature is defined as the vector field

H(x):=i=1miNei(x)-iMei(x).

Here (ei)i=1m is an arbitrary local orthonormal frame on a neighbourhood of x in M. Note that we omit the factor of 1/m that usually appears in this definition in the literature in order to simplify the statement of our results. It is a standard fact that both Inline graphic and H(x) are vectors which are perpendicular to the submanifold M. We refer to e.g. [19, Chapter 5] for a detailed treatment of these objects. For instance, one of the examples we consider below is that of a planar curve γ with radius of osculating circle R(α). A simple computation shows that in this case

graphic file with name 40879_2025_816_Equ180_HTML.gif

where · is the Euclidean magnitude.

There exists a considerable body of literature on description of submanifold properties by tubular volume, of which we name a few representatives. The early work of Weyl [32] proved the classical tube formula for submanifolds embedded in Euclidean spaces, which is an expansion with respect to the width of the tubular volume and its coefficients are geometric invariants of the submanifold. Federer [10] introduced the notion of boundary measures, which lead to generalization of the tube formula to compact subsets of Euclidean spaces. More recent works of Chazal et al. [5, 6] studied geometric inference via point cloud approximations to boundary measures using Monte Carlo methods. For a comprehensive treatment on properties of tubular neighbourhoods, we refer to the monograph [16]. The approach in our present work differs from the above in that it gives a local and directional information about the second fundamental form, and also the mean curvature.

Notions of synthetic Ricci curvature were motivated by the study of geometry of metric measure spaces and were pioneered by the seminal works [4, 23, 30], see also the survey [22]. In a metric measure space, a global lower bound on the synthetic Ricci curvature leads to properties of the metric measure space which are analogous to the Riemannian setting, such as the Poincaré and log-Sobolev inequalities, the concentration of measure phenomenon, and closure under measured Gromov–Hausdorff convergence [1, 8, 28]. We note also the related direction of the works [4, 31].

To our understanding, there has not been a notion of a synthetic extrinsic curvature. Our notion of coarse extrinsic curvature is inspired by coarse Ricci curvature of Ollivier [25], which is defined in the Riemannian setting through the expansion of the 1-Wasserstein distance of two uniform measures supported on geodesic balls of a small radius, the radius being the variable of expansion [25, Example 7], see also the survey [26]. This is different from the above mentioned synthetic Ricci curvature lower bounds in that it puts a precise number on the value of curvature at a point. Moreover, it can be applied to general metric spaces by choosing a family of measures indexed by points in the space for the evaluation of the 1-Wasserstein distance. Coarse Ricci curvature can be computed explicitly for a number of examples on graphs, where the measures are provided by a Markov chain. We adopt and modify Ollivier’s approach to the submanifold setting by choosing suitable measures for the expansion of the 1-Wasserstein distance, showing that this yields a geometrically meaningful information.

As an immediate application of our result, we venture into the setting of [2, 18] to explore retrieval of curvature information from point clouds generated by a Poisson point process. In the first of the mentioned works, Hoorn et al. proved that Ollivier’s coarse Ricci curvature of random geometric graphs sampled from a Poisson point process with increasing intensity on a Riemannian manifold converges in expectation at every point to the classical Ricci curvature of the manifold. This was extended in the second mentioned work to weighted Riemannian manifolds. In the present work, we show that coarse extrinsic curvature can recover the mean curvature in expectation at a point. In this case, it is not necessary to impose a graph structure to connect points of the sample.

In the context of deep learning, it is noteworthy that computational algorithms for effectively computing optimal transport maps have been proposed, as discussed in [17, 29]. Additionally, relevant work in the fields of manifold learning and inverse problems is worth mentioning. One particularly interesting inverse problem is whether an embedded manifold can be learned from a set of samples xj+ξj where xj belongs to a submanifold MRm+p and ξj are independent Gaussian random variables on Rm+p. The reconstruction of embedded manifolds has been studied in [9, 11, 12, 27]. An algorithm for constructing an embedded submanifold is provided in [13]. Although manifold learning is still in its early stages, manifold approximation and reconstruction have a longer history, we point out some more recent publications on this topic [3, 7, 7, 14, 15].

Main results

In our setting, M is an m-dimensional compact Riemannian manifold embedded isometrically in a Euclidean space Rm+k and Mσ is the local σ-tubular neighbourhood of M in Rm+k, defined for σ sufficiently small as

Mσ={x+v:xM,vTxM,vσ}.

For any compact subset UM, the projection mapping from its tubular neighbourhood

π:UσU,π(z):=argminxUz-x

is well-defined for all σ>0 sufficiently small, with the same notation for Uσ as above.

Denote by expM,x:TxMM the exponential mapping in M with base point x. Fix a point x0M, a unit tangent vector vTx0M and denote y:=expM,x0(δv) for δ>0. Fix a constant ε0>0 smaller than the uniform injectivity radius of some fixed compact neighbourhood of x0 in M. Assume δ,ε<ε0/3 so that Bε(x0)Bε(y) lie within the uniform injectivity radius away from x0 and assume σ is small enough so that the projection π is well-defined on the σ-tubular neighbourhood of the ε0-geodesic ball at x0 in M. These requirements on the parameters δ,ε,σ will henceforth be encapsulated in the assumption that they are “sufficiently small”. This ensures that all locally defined maps are well-defined, in particular the projection map (smallness of σ) and the Fermi coordinates (smallness of δ and ε) used later on.

As our test measures, we choose the probability measures

graphic file with name 40879_2025_816_Equ181_HTML.gif

where BεM(x) denotes the ε-geodesic ball at x in M. Note that these measures are supported on compact subsets of Mσ. We seek to obtain the expansion of W1(μx0σ,ε,μyσ,ε) with respect to the parameters δ,σ and ε.

To relate the Wasserstein distance to the second fundamental form, we first localize to a tubular neighbourhood of a fixed open set on the submanifold. We expand the densities of the test measures in Fermi coordinates, and for the subsequent computations we rely on a crucial observation developed in Sect. 2.2: if T is an approximate transport map from μx0σ,ε to μyσ,ε, in a sense defined later, then W1(μx0σ,ε,μyσ,ε) is close to W1(μx0σ,ε,Tμx0σ,ε). The remaining task involves proposing a concrete approximate transport map, which is at the same time close enough to optimal.

When dealing with test measures on an embedded manifold, accounting for the effect of the bending of the submanifold in the ambient space becomes crucial. The proposed transport map is thus formulated in terms of the Fermi frame along γ, adapted to the submanifold M in a way that separates tangent and normal coordinate directions at every point.

We give a rough outline of the proposed transport map, made precise in Sect. 2.4. In terms of Fermi coordinates, if α=(α1,,αm) represent submanifold tangent directions with α1 being associated with the direction of γ, and if β=(β1,,βk) represent the normal directions, an initial proposal informed by the circle example (Sect. 3.1) was

(α,β)(δ-α1,α2,,αm,β1,,βk).

This can be construed as translation by δ in the direction of the first coordinate, together with reflection in the first coordinate. From studying the planar curve example (Sect. 3.2), it turned out that an additional bending correction needs to be put on top of the β components of the transport by adding terms involving the derivative of the mean curvature. Favourably, such a correction contributes to the final estimate of the Wasserstein distance only at the fourth order and higher, and hence does not interfere with the mean curvature term, which will appear at third order of the expansion. The test measures are first expressed in Fermi coordinates in Sect. 2.3. The proposed transport map is then presented in Sect. 2.4, where we prove that it is indeed an approximate transport map of degree 3, i.e.

d(Tμx0σ,ε)dμyσ,ε(ϕ(α,β))=1+O(δ3).

This precision is sufficient for obtaining the 1-Wasserstein distance approximation (see Sect. 2.2):

W1(μx0σ,ε,μyσ,ε)=W1(μx0σ,ε,Tμx0σ,ε)+O(δ4).

From here the strategy is to construct a test function Inline graphic with Lipschitz norm approximately 1 and satisfying the estimate

f(Tz)-f(z)=Tz-z+O(δ4)=O(δ),

which allows us to estimate the distance between the original measure and its transport by means of the relation

W1(μx0σ,ε,Tμx0σ,ε)=(f(Tz)-f(z))dμx0σ,ε(z)+O(δ4).

On the whole, we find that the Wasserstein distance between the initial measure μx0σ,ε and the target measure μyσ,ε is approximated by MTz-zdμx0σ,ε up to O(δ4) (see Lemma 2.25), which is explicitly computable as an expansion in δ,σ and ε with geometric quantities as coefficients.

Using the above tools, in Sect. 3 we thus compute the expansion of W1(μx0σ,ε,μyσ,ε), beginning with the case of a planar curve:

Proposition 1.1

Let γ be a smooth unit speed curve in R2 such that γ(0)=x0 and γ(δ)=y. For all δ,ε,σ>0 sufficiently small with σεδ/4, it holds that

W1(μx0σ,ε,μyσ,ε)=x0-y(1-ε26R2+σ23R2)+O(δ4)

where R is the radius of the osculating circle of the curve at x0.

This expansion can be rearranged as

1-W1(μx0σ,ε,μyσ,ε)x0-y=ε26R2-σ23R2+O(δ3).

We refer to the quantity on the left as the coarse extrinsic curvature of γ between x0 and y at scales σ,ε. A version of this result for spatial curves is presented in Theorem 3.10. In Theorem 3.16, we then proceed to study the case of coarse extrinsic curvature along a geodesic on a surface embedded in R3.

This work culminates with the most general form:

[Style2 Style3 Style3]Theorem 4.1

Let M be an isometrically embedded submanifold of Rm+k, and γ a unit speed geodesic in M such that γ(0)=x0 and γ(δ)=y. Let (ej)j=1m be an orthonormal basis of Tx0M with e1=γ˙(0) and assume that Inline graphic for all j=2,,m. Then for every σ,ε,δ>0 sufficiently small with σεδ/4 it holds that

graphic file with name 40879_2025_816_Equ182_HTML.gif

The assumption on the second fundamental form is necessary for optimality of our proposed transport map up to sufficient order and can always be satisfied for submanifolds of codimension 1, in particular surfaces embedded in R3, by choosing the basis of principal curvature directions. Further commentary is provided in Remark 4.2.

To interpret such expansions in terms of mean curvature, we can remove the directionality of the above result caused by transport in the direction of γ. Denoting the square norm of the mean curvature vector as

H(x0)2=i=1kH(x0),ni(x0)2

for an arbitrary orthonormal basis (ni(x0))i=1k of the normal space Tx0MTx0N, we deduce the following:

Corollary 1.2

Let (ej)j=1m be an orthonormal basis of Tx0M, and for j=1,,m, let yj=expM,x0(δej). Assume that Inline graphic for ij. Then for all σ,ε,δ>0 sufficiently small with σεδ/4 it holds that

j=1m(1-W1(μx0σ,ε,μyjσ,ε)x0-yj)=(ε22(m+2)-σ2k+2)H(x0)2+O(δ3).

Observe that the left side of the equation is independent of the choice of orthonormal basis (ej)j=1m because the norm on the right side is basis-invariant. Moreover, the assumption on the second fundamental form always holds for submanifolds of codimension 1 (see Remark 4.2).

In Proposition 5.4, we deduce that the coarse extrinsic curvature of suitable test measures on Poisson point clouds sampled from the tubular neighbourhood retrieves the same extrinsic geometric information consistent with Theorem 4.1.

One key ingredient in the proofs of the above theorems is the geometric approximate transport map introduced in Definition 2.18, defined by means of Fermi coordinates (as per Definition 2.14) adapted to the submanifold. Test measures in these coordinates encode information about the second fundamental form of the submanifold. The proposed map is verified to be an approximate transport map between the test measures with sufficient order of accuracy, as specified and motivated in Sect. 2.2. The optimality up to fourth order is proved by choosing a concrete test function for the Wasserstein lower bound by the Kantorovich–Rubinstein duality.

In the resulting expansion of the Wasserstein distance, the second fundamental form at the fixed point x0 appears at third order, and its derivatives appear at fourth and higher orders. As a consequence, information about the second fundamental form at a point can be retrieved in a suitably scaled limit of coarse curvature. Please see the discussion below for an example.

Discussion

We illustrate this work using the following prototypical example. Let γ:(-δ0,δ0)R2 be a smooth, unit speed planar curve, and n:(-δ0,δ0)R2 a unit normal vector field along γ, unique up to sign. Denote by R(α):=1γ¨(α) the radius of the osculating circle at the point γ(α). To detect the extrinsic curvature at x0:=γ(0), captured here by R(0), we define test probability measures centered at nearby points y:=γ(δ) indexed by δ>0.

Denote μ the Lebesgue measure on R2, M:=γ((-δ0,δ0)) as the image of the curve, and Mσ0 as a small enough tubular neighbourhood of M so that the orthogonal projection π:Mσ0M is well-defined. Denote

Bσ,ε(y):={zR2:z-π(z)<σ,dγ(y,π(z))<ε}

where dγ is the distance along γ. Define for σ,ε>0 with σεδ/4, the Borel measure on R2,

μyσ,ε(A):=μ(ABσ,ε(y))μ(Bσ,ε(y)).

We compare these in 1-Wasserstein distance to the initial measure, i.e. when δ=0 and is denoted μx0σ,ε. The Wasserstein distance has the form:

W1(μx0σ,ε,μyσ,ε)=x0-y(1-ε26R(0)2+σ23R(0)2)+O(δ4).

Rearranging this expansion yields

1-W1(μx0σ,ε,μyσ,ε)x0-y=1R(0)2(ε26-σ23)+O(δ3). 1.2

From this point, depending on the application, we may consider three different regimes for the parameters ε and σ as y converges to x0. We recall the asymptotic notation σ=Θ(δ) means there exist c,C,δ0>0 such that for all δ<δ0,

cδ<σ(δ)<Cδ,

and σ=o(δ) means limδ0σ(δ)δ=0.

  • (i)
    limδ0ε(δ)σ(δ)=C2 for some known constant C>0, i.e. the decay of both σ and ε is controlled. In this case,
    1R(0)2=limδ0-6(C2-2)σ2(1-W1(μx0σ,ε,μyσ,ε)x0-y),
  • (ii)
    σ=Θ(δ) and ε=o(δ), i.e. the decay of σ is controlled, while the parameter of support size ε vanishes fast. In this case,
    1R(0)2=limε=o(σ),σ=Θ(δ)δ0-3σ2(1-W1(μx0σ,ε,μyσ,ε)x0-y),
  • (iii)
    ε=Θ(δ) and σ=o(δ), i.e. the decay of ε is controlled, while the size of the tubular neighbourhood σ vanishes fast. In this case,
    1R(0)2=limσ=o(ε),ε=Θ(δ)δ06ε2(1-W1(μx0σ,ε,μyσ,ε)x0-y).

The requirements σ=Θ(δ) and ε=Θ(δ) in the respective cases are in place to ensure the remainder term O(δ3) in (1.2) does not explode upon division by σ2 (resp. ε2) in the limit as δ0.

In light of the above discussion, we may define the coarse extrinsic curvature between x0 and y at scales ε,σ>0 as:

κσ,ε(x0,y):=1-W1(μx0σ,ε,μyσ,ε)x0-y. 1.3

This quantity can be estimated from point cloud data and used for geometric inference.

In summary, this work focuses on Riemannian submanifolds embedded isometrically in Euclidean spaces with the aim of producing a reasonable measurement for the bending energy. This bending energy can also be estimated from point clouds obtained from sampling. One of the novel ingredients is the construction of a test function for using the Kantorovich–Rubinstein duality to obtain a lower bound for the Wasserstein distance in this setting.

The outline of this work is as follows. In Sect. 2, we establish geometric preliminaries pertaining to the volumes of tubular neighbourhoods and present approximate transport maps as a novel tool for approximating the 1-Wasserstein distance. In Sect. 3, we give description of coarse extrinsic curvature for a planar curve, space curve and a 2-surface embedded in R3. The coarse extrinsic curvature of a general submanifold of arbitrary codimension is studied in Sect. 4. We present several immediate corollaries to our results with practical applications in Sect. 5. Although the cases of curves and surfaces in Sect. 3 are just instances of the general result in Sect. 4, they provide value in understanding this general case. Sections 3 and 4 can be read separately after reading Sect. 2, which contains all preliminaries.

Preliminaries

We prove a formula for volume growth of tubular neighbourhoods of submanifolds, leading to a disintegration of the ambient volume measure adapted to the submanifold. This formula is subsequently utilized to derive explicit formulas for such disintegration in Fermi coordinates, considering cases such as a planar curve, space curve, and a surface in Sect. 3, and general Riemannian submanifolds in Sect. 4.

Following the geometric preliminaries, we introduce the notion of an approximate transport map, enabling the computation of Wasserstein distances up to a sufficiently high degree of error. Subsequently, we define the test measures to be transported and their representation in Fermi coordinates. Finally, we propose a transport map to evaluate the Wasserstein distance of these test measures.

Ambient volume disintegration

We begin with a simple lemma on evolution of probability densities. We denote Inline graphic as the space of probability measures on a measurable space Inline graphic. The notation μν denotes the fact that the measure μ is absolutely continuous with respect to the measure ν.

Lemma 2.1

Consider Inline graphic such that μtμs for all st. Let Inline graphic, t0, be a family of functions with tht(x) locally integrable, and such that dds|s=0dμt+sdμt(x)=ht(x) for every t0. Then

dμtdμ0(x)=e0ths(x)ds.

Proof

The change of density at any t0 satisfies

dds|s=0dμt+sdμt(x)=dds|s=0dμt+sdμ0(x)dμtdμ0(x)=ht(x)

implying

dds|s=0dμt+sdμ0(x)=ht(x)dμtdμ0(x)

which has the unique solution dμtdμ0(x)=e0ths(x)ds by standard ODE theory.

Notation 2.2

Throughout this article, M is a compact Riemannian manifold of dimension m, isometrically immersed in a Riemannian manifold N of dimension n. Set k:=n-m. Let σ0>0 be a fixed number smaller than half the reach of M in N. The reach is defined as the maximal number r such that each point within a distance r from M has a unique orthogonal projection to M, Inline graphic. The projection map is well-defined within the ‘reach’.

Let UM be a sufficiently small open neighbourhood such that there exists an orthonormal frame of unit normal vector fields (n1,,nk) on U and a one-parameter family of vector fields {(ei(s))i=1m:s(-σ0,σ0)} such that (ei(s))i=1m is an orthonormal frame on ψs(U) for every s(-σ0,σ0), and sei(s) is smooth for every i=1,,m. The latter can be constructed by taking the pushforward of an arbitrary initial orthonormal frame by ψs, denoted by (Dei(0)ψs)i=1m, and applying the Gram–Schmidt orthonormalization procedure.

Definition 2.3

Let nΓ(TU) be a unit normal vector field, and define the normal flow ψ:M×(-2σ0,2σ0)N by

ψt(x):=expN,x(tn(x))

where expN,x:TxNN denotes the exponential mapping on N. Denote by Inline graphic the parallel transport with respect to the Levi-Civita connection N along tψt(x) for a fixed xM, and note that Inline graphic.

For every t(-2σ0,2σ0), ψt is a diffeomorphism onto its image, and tψt(x) is smooth with non-vanishing derivative for every xM. Equip every ψt(M) with the Riemannian metric inherited from the ambient space. The mean curvature of the leaf ψs(U) is then given by

H(ψs(x))=i=1miNei(ψs(x))-iMei(ψs(x)).

In particular, for any unit normal vector field n on U,

graphic file with name 40879_2025_816_Equ183_HTML.gif

The following lemma shows Inline graphic stays normal to the leaves ψt(U) as t changes.

Lemma 2.4

The vector field Inline graphic is normal to ψt(U) for every t(-σ0,σ0), i.e. Inline graphic for any local tangent frame (ei)i=1m on M.

Proof

For every t(-σ0,σ0), ψt being a diffeomorphism implies that if (ei)i=1m is a frame on U, then (Deiψt)i=1m is a frame on ψt(U), not necessarily orthonormal. Then

graphic file with name 40879_2025_816_Equ184_HTML.gif

where on the second line we used that Inline graphic. The initial condition ψ0=id gives Deiψ0,n=ei,n=0, so we may conclude that Inline graphic is normal to all tangent directions on ψt(U) for all t.

The action of push-forwards of volume forms on any orthonormal basis of tangent vectors is characterized by the determinant of the mapping which we make precise below. Let M1,M2 be Riemannian manifolds of the same dimension m, ψ:M1M2 a diffeomorphism, (ei)i=1m an orthonormal frame on an open set U1M1 and (e~i)i=1m an orthonormal frame on an open set U2M2, and (ei)i=1m,(e~i)i=1m the corresponding coframes characterized by ei(ej)=δji,e~i(e~j)=δji. Below, by the determinant of Dψ-1(x):TxM2TΨ-1(x)M1 we mean that of the matrix representing the map in these bases:

detDψ-1=σSmi=1msign(σ)ei(ψ-1e~σ(i)).

By the rules of differential forms acting on tangent vectors, for all xU2,

ψ(e1em)(x)(e~1(x),,e~m(x))=(e1(x)em(x))(ψ-1e~1(x),,ψ-1e~m(x))=σSmi=1msign(σ)ei(x)(ψ-1e~σ(i)(x))=detDψ-1.

Since linear maps are determined by their values on basis vectors, we may deduce

ψ(e1em)(x)=detDψ-1(x)e~1e~m(x). 2.1

With the above notation we return to the exponential map ψt(x):=expN,x(tn(x)).

Proposition 2.5

(Change of volume) For every t(-σ0,σ0),

graphic file with name 40879_2025_816_Equ5_HTML.gif 2.2

and hence the volume of the image of any Borel measurable AU can be expressed as

graphic file with name 40879_2025_816_Equ6_HTML.gif 2.3

where volψt(M) is the Riemannian volume on ψt(M).

Proof

First, we extend the map ψ:(-σ0,σ0)×UN to ψ~:(-σ0,σ0)×Uσ0N by the flow condition

ψs~(ψt(x)):=ψt+s(x)

for all s and t in (-σ0,σ0). This determines ψ~ uniquely because {ψt(U)}t(-σ0,σ0) is a foliation of the tubular neighbourhood Uσ0. Then on every leaf ψt(U) of the foliation, we have ψ~0=id. If (ei)i=1m and (e~i)i=1m are orthonormal coframes on ψt+s(U) and ψt(U) respectively, the change of variable formula for volume forms (2.1) states that

(ψ~-s)e1en=(detDψ~s)e~1e~n.

Then for every AU Borel measurable and s(-σ0,σ0),

volψt+s(M)(ψt+s(A))=volψt+s(M)(ψ~s(ψt(A)))=((ψ~-s)volψt+s(M))(ψt(A))=ψt(A)detDψ~s(x)dvolψt(M)(x)

using respectively the flow property, definition of the push-forward of a measure, and the change of variable formula with volψt+s(M)=e1em and volψt(M)=e~1e~m.

Denoting by Ds the covariant derivative along sψ~s, the Jacobi formula for the derivative of determinants gives

graphic file with name 40879_2025_816_Equ185_HTML.gif

From the second to third line, the other term coming from the product rule applied on the bracket does not contribute to the trace, because for i=j,

Dei(t)ψ~0,Ddtei(t)=ei(t),Ddtei(t)=12tei(t),ei(t)=0

using that ψ~0(x)=x so Dψ~0=id. From the fourth to fifth line, we used normality of the flow s|s=0ψ~s(x),ei(t)(x)=0, and on the last line applied sψ~s(ψt(x))=sψt+s(x) from definition of the extension ψ~, before pulling the integral from ψt(A) back to A.

Hence the evolved volume measure pulled back to U satisfies the dynamics

graphic file with name 40879_2025_816_Equ7_HTML.gif 2.4

which together with the initial condition d(ψ0)-1volψ0(M)=dvolM implies

graphic file with name 40879_2025_816_Equ186_HTML.gif

by Lemma 2.1, setting ht to be the right-hand side of (2.4). Equation (2.3) is then simply the change of variable formula for the map ψt.

Remark 2.6

The formula of Proposition 2.5 can be extended from the neighbourhood U to all of M by a partition of unity argument, nonetheless the local formulation is sufficient for our purpose.

We proceed to derive a disintegration of the ambient volume measure adapted to a submanifold of arbitrary codimension, at the cost of specialising to the case N=Rn. Note that the covariant derivative n then becomes the plain derivative denoted by D.

Notation 2.7

Let (nj)j=1k be a local orthonormal frame for TM on U and denote Inline graphic the centered Euclidean ball of radius σ0. Define the map

graphic file with name 40879_2025_816_Equ8_HTML.gif 2.5

which gives the k-dimensional foliation {ψ(U,β):βBσ0k} of π-1(U) with leaves of dimension m. Extend (nj)j=1k and (ei)i=1m smoothly to π-1(U) so that the restrictions to the submanifold ψ(U,β) are an orthonormal frame in the tangent space and the normal space, respectively, for every βB~σ0k.

Denote n(x,β)=j=1kβjnj(x)β2 which was shown in Lemma 2.4 to be normal to each leaf ψ(U,β). Then the mean curvature of ψ(U,β) in the direction n(x,β) is

H(ψ(x,β)),n(x,β)=n(x,β),i=1mim+kei(ψ(x,β)).

Denote also the components of mean curvature in each of the directions of the normal frame,

Hj(ψ(x,β)):=nj(x),H(ψ(x,β))=nj(x),i=1mim+kei(ψ(x,β))

so that

βH(ψ(x,β)),n(x,β)=j=1kβjHj(ψ(x,β)).

Remark 2.8

The collection of submanifolds {ψ(U,β):βB~σ0k} is indeed a foliation of π-1(U)Rm+k (see e.g. the definition of foliation in [21]). The leaves are disjoint submanifolds of dimension m. Defining

F:(p1,,pm,β1,,βk)ξ(p1,,pm)+j=1kβjnj(ξ(p1,,pm))

where ξ:ORmU is an arbitrary chart on U, we have by definition that

ψ(U,β)=F({β}),

so each leaf is a level set of F and thus F is a flat chart for the foliation.

Proposition 2.9

(Disintegration) The ambient volume measure on π-1(U)Rn disintegrates with respect to the submanifold and the normal frame (nj)j=1k as

graphic file with name 40879_2025_816_Equ9_HTML.gif 2.6

for any Borel measurable set Inline graphic in the tubular neighbourhood.

Proof

We apply the change of coordinates by the map defined by (2.5), which at every (x,β) has block-triangular derivative with respect to the orthonormal bases

(e1(x),,em(x),β1(x),,βk(x))

and

(e1(ψ(x,β)),,em(ψ(x,β)),n1(x),,nk(x))

in the domain and codomain respectively, since

βiψ(x,β),ej(x,β)=ni(x),ej(ψ(x,β))=0.

Hence the determinant can be computed as

detDψ=det(Deiψ,ej)det(βiψ,nj)=det(Deiψ,ej)det(δji)=det(Deiψ,ej)

for which we have the right-hand side of (2.2).

Denoting (ei)i=1m, (ni)i=1k the coframes characterized by ei(ej)=δji and ni(nj)=δji,

graphic file with name 40879_2025_816_Equ187_HTML.gif

on the second line using the change of variable formula and on the third line plugging in the determinant expression (2.2) with n(x,β)=j=1kβjnj(x)β and t=β2. The final expression is obtained by the substitution s=sβ so that

graphic file with name 40879_2025_816_Equ188_HTML.gif

Corollary 2.10

(Codimension 1) If M has codimension 1 then the ambient volume measure on π-1(U) can be written in terms of the disintegration

volRn(A)=U-σσ1A(ψ(x,β))e-0βH(ψ(x,β),n(x,β)dβdβvolM(dx), 2.7

for all Inline graphic, where volM(dx) is the volume measure of the submanifold M and H(ψ(·,β)) is the mean curvature on the Riemannian submanifold ψ(U,β).

Approximate transport maps

In the sequel, we work with transport maps which are only optimal up to sufficiently high degree for asymptotically small diameter of support of the test measures. We present a result which justifies the use of such transport maps.

Let Inline graphic be a Polish space, Inline graphic the set of probability measures, define

graphic file with name 40879_2025_816_Equ189_HTML.gif

and consider two families of probability measures Inline graphic, Inline graphic.

Lemma 2.11

(W1 distance approximation) If diamsuppμ2δ=O(δ) and μ1δμ2δ for every δ0 with the density satisfying supxsuppμ2dμ1δdμ2δ(x)=1+O(δk), then

graphic file with name 40879_2025_816_Equ190_HTML.gif

Proof

By the reverse triangle inequality and Kantorovich–Rubinstein duality, for all Inline graphic:

graphic file with name 40879_2025_816_Equ191_HTML.gif

where x0suppμ2δ is arbitrary. On the last line, we introduced the term

graphic file with name 40879_2025_816_Equ192_HTML.gif

because f(x0) is constant and the density integrates to 1, and then used the 1-Lipschitz property of f together with the O(δ) bound on the diameter of the support of μ2.

Let Inline graphic be another family of probability measures.

Definition 2.12

(Approximate transport map) A measurable map Inline graphic is said to be an approximate transport from μδ to μ2δ with degree k if Tδμδμ2 and the density satisfies

graphic file with name 40879_2025_816_Equ193_HTML.gif

Corollary 2.13

If Inline graphic is an approximate transport map from μδ to μ2δ with degree k and diamsuppμ2δ=O(δ) then

graphic file with name 40879_2025_816_Equ194_HTML.gif

Proof

Set μ1δ:=Tδμδ and apply the previous lemma.

Test measures in Fermi coordinates

Let (Mg) be a Riemannian submanifold of codimension k in Rm+k and UM an open neighbourhood of a point x0M as in Notation 2.2. The Fermi coordinates are a suitable tool for explicit computations and will be used throughout the rest of this work. The following is a modification of classical Fermi coordinates to the submanifold setting.

Definition 2.14

(Fermi coordinates) Let γ:(-δ0,δ0)M be a unit speed geodesic with δ0>0 small enough for γ to be contained in U, ε0 the uniform injectivity radius in M along γ and σ0 smaller than half the reach of U in Rm+k. Let (ei)i=1m be an orthonormal frame for the fibres of TM along γ such that e1(α1)=γ˙(α1) and γ˙Mei(α1)=0 for i=1,,m and every α1(-δ0,δ0). Also let (ni)i=1k be a local orthonormal frame for fibres of the normal bundle TM along γ.

Denote by B~εm-1 the centered ball of radius ε>0 in Rm-1 and by B~σk the centered ball of radius σ>0 in Rk. Denote α=(α1,,αm),β=(β1,,βk) and define

ϕ:(-δ0,δ0)×B~ε0m-1×B~σ0kRm+k,ϕ(α,β):=expM,γ(α1)(i=2mαiei(α1))+j=1kβjnj(α),

which is a diffeomorphism provided that δ0,ε0,σ0>0 are sufficiently small. This is referred to as the Fermi chart along γ adapted to the submanifold M. See Fig. 3 for an illustration on a 2-surface in R3.

Fig. 3.

Fig. 3

Fermi coordinates along γ adapted to the surface M embedded in R3

The Riemannian metric is expressed in the Fermi coordinates as

gij(α)=αiϕ(α,0),αjϕ(α,0). 2.8

Remark 2.15

The advantage of ϕ over a generic ψ as given in Notation 2.7 is that ϕ is adapted to the geodesic γ in a way that simplifies computations of distances relevant to our optimal transport problem. The chart ϕ yields again a foliation {ϕ(U,β):βB~σ0k} of Mσ0.

Definition 2.16

(Test measures) Denote the cylinder-like segment in Rn of height σ and radius ε centered at xM as

Bσ,ε(x):={zMσ:dM(π(z),x)<ε}

and let μ be the Lebesgue measure on Rn. For any xM define the family of test probability measures

graphic file with name 40879_2025_816_Equ195_HTML.gif

indexed by ε,σ>0.

Denote α^=(α2,,αm) so that α=(α1,α^). The main purpose of the expansion in the following lemma is twofold. First, we use it to design the third order corrections in the approximate transport map of Definition 2.18 so that density matching occurs in Proposition 2.23. Second, the first order term of the expansion interacts with first order term of pointwise distance when integrating to get the Wasserstein upper bound in the proofs of Sects. 3 and 4.

Lemma 2.17

(Test measures in Fermi coordinates) For any y=γ(δ), the expansion of the density of test measures in Fermi coordinates is

(ϕ-1μyσ,ε)(dα,dβ)=1Z1B~σ,ε(δ+α1,α^,β)(1-i=1kβiHi(ϕ(0))-i=1kj=1mαjβiαj(Hiϕ)(0)-i,j=1kβiβjβj(Hiϕ)(0)+12i,j=1kβiβjHi(ϕ(0))Hj(ϕ(0))+14q,=2mi=1mαqααqαgii(0)+O(δ3))dαdβ 2.9

where Z is the probability normalization constant and g=(gij) is the Riemannian metric of M in Fermi coordinates given by (2.8).

Proof

First, note the pull-back of the test measure μyσ,ε(dα,dβ) to Fermi coordinates is

graphic file with name 40879_2025_816_Equ196_HTML.gif

The Riemannian metric in Fermi coordinates expands as

graphic file with name 40879_2025_816_Equ13_HTML.gif 2.10

Indeed, αigj(α1,0)=0 for all α1(-δ0,δ0), and hence also α1αigj(α1,0)=0. We show this by cases:

  • for all i,j,=2,,m: igj(α1,0)=0 because for every fixed α1(-δ0,δ0), ϕ(α1,·) are normal coordinates within the injectivity radius of expγ(α1)({γ˙(α1)})M at γ(α1) (see e.g. [19, Section 1.4] for a proof),

  • for all i,j=1,,m: 1gij(α1,0)=0 for any α1(-δ0,δ0) as the orthonormal frame along γ used to define the Fermi chart is parallel translated along γ,

  • for all i=2,,m and j=1,,m and any α1(-δ0,δ0):
    graphic file with name 40879_2025_816_Equ197_HTML.gif
    using that α1ϕnαiϕ(α1,0)-α1ϕMαiϕ(α1,0)M and αiϕMαjϕ(α1,0)=0. The latter vanishes for j1 again by normality of the chart on expγ(α1)({γ˙(α1)}), and for j=1 because αiϕ is given by parallel translation along γ.

The Riemannian volume expanded in the Fermi coordinates then simplifies to

graphic file with name 40879_2025_816_Equ198_HTML.gif

Moreover, expand the exponent in the normal part of the disintegration as

graphic file with name 40879_2025_816_Equ199_HTML.gif

and apply the approximation up to second order ex=1+x+x22+O(x3) to obtain

graphic file with name 40879_2025_816_Equ14_HTML.gif 2.11

We conclude the result by taking the product of the two factors (2.10) and (2.11), merging third order terms in α,β into O(δ3) by the assumption εσδ/4.

The probability normalization constant can be deduced by integration of (2.11) with respect to μyσ,ε as

Z=1+12(k+2)σ2i=1kHi(ϕ(0))2+O(δ3),

and so

1Z=1-12(k+2)σ2i=1kHi(ϕ(0))2+O(δ3).

Proposed transport map

As mentioned in the introduction, when considering an embedded manifold, it is crucial to include the mean curvature in the transport map. We will show that the transport map proposed below is an approximate transport map with degree 3. We then present a criterion for optimality of the proposed map in Lemma 2.25.

Definition 2.18

Define T:Bσ,ε(x0)Bσ,ε(y) in Fermi coordinates as

T(ϕ(α,β)):=ϕ(δ-α1,α2,,αm,β1-12(σ2-β12)(δ-2α1)α1(H1ϕ)(0),,βk-12(σ2-βk2)(δ-2α1)α1(Hkϕ)(0)).

Denote by α^=(α2,,αm) and similarly for α^, and denote the input vector on the right in the above definition as Inline graphic. Note that

α1=δ-α1,α^=α^,β=β+O(δ3).

Remark 2.19

Observe that T is a local diffeomorphism and

|detD(ϕ-1Tϕ)(α,β)|=1-i=1kβi(δ-2α1)α1(Hiϕ)(0)+O(δ3)

and deduce

graphic file with name 40879_2025_816_Equ15_HTML.gif 2.12

Remark 2.20

The third order terms in the definition of T are adjustments to cancel out second order terms in the proof of Proposition 2.23 below, obtaining an approximate transport of degree 3 as a result. In fact, the form of T is tailored precisely for this to occur. It turns out these third order adjustment terms do not influence the Wasserstein distance computation up to order 4.

We need two general lemmas to show that T is an approximate transport of degree 3.

Lemma 2.21

(Density under pushforward) Let Inline graphic be measurable spaces, Inline graphic a measurable bijection with measurable inverse, and μ,ν two measures on Inline graphic with μν. Then the push-forward measures are also absolutely continuous with density

d(ψμ)d(ψν)(x)=dμdν(ψ-1(x)).

Proof

For all measurable sets Inline graphic,

ψμ(A)=μ(ψ-1(A))=ψ-1(A)dμdν(x)dν(x)=ψ-1(A)dμdν(ψ-1ψ(x))dν(x)=Adμdν(ψ-1(x))d(ψν)(x).

Noting that the representations (2.6) and (2.7) are decompositions into skew-products of two measures, the following will be used for density comparisons.

Let Inline graphic be measurable spaces. Given measures Inline graphic on Inline graphic and a measure μ2 on Inline graphic, the skew-product is defined as follows: For all bounded measurable real valued functions f on Inline graphic,

graphic file with name 40879_2025_816_Equ200_HTML.gif

Lemma 2.22

(Skew-product density factorization) Consider two families of measures Inline graphic and Inline graphic on Inline graphic such that μ1yν1y for every Inline graphic and the map (x,y)dμ1ydν1y(x) is measurable. Furthermore, let μ2 and ν2 be measures on Inline graphic with μ2ν2. Consider the skew products of Inline graphic with μ2 and that of Inline graphic with ν2. Then μ1μ2ν1ν2 and

d(μ1μ2)d(ν1ν2)(x,y)=dμ1ydν1y(x)dμ1dν2(y).

Proof

Plugging in the densities, for all Inline graphic bounded:

graphic file with name 40879_2025_816_Equ201_HTML.gif

We verify that the density of Tμx0σ,ε via T matches that of μyσ,ε up to O(δ3).

Proposition 2.23

The proposed map is an approximate transport map of degree 3, i.e.

d(Tμx0σ,ε)dμyσ,ε(ϕ(α,β))=1+O(δ3).

Proof

First, combining the elementary change of variable formula with the Fermi coordinate representation of μx0σ,ε, with notation of Definition 2.18 we have

graphic file with name 40879_2025_816_Equ16_HTML.gif 2.13

using the expansions (2.12) for the determinant of ϕ-1Tϕ and (2.9) for the coordinate representation of ϕ-1μx0σ,ε.

We use Lemma 2.21 to push the density into Fermi coordinates, and then Lemma 2.22 allows us to take the ratio of the densities of (2.13) and (2.9), obtaining

d(Tμx0σ,ε)dμyσ,ε(ϕ(α,β))=d(ϕ-1Tμx0σ,ε)d(ϕ-1μyσ,ε)(α,β)=1B~σ,ε(0)(α1-δ,α^,β)(1+O(δ3))

because the second order terms cancel out. Here we also used that T is a diffeomorphism from B~σ,ε(0m+k) to B~σ,ε(δ,0m+k-1), hence

graphic file with name 40879_2025_816_Equ202_HTML.gif

Remark 2.24

Building upon the preceding proposition and leveraging Corollary 2.13, we readily deduce that the proposed transport map satisfies:

W1(μx0σ,ε,μyσ,ε)=W1(μx0σ,ε,Tμx0σ,ε)+O(δ4),

taking also into account that suppTμx0σ,ε=suppμyσ,ε leading to diamsuppTμx0σ,ε=O(δ) when σεδ/4. Thus, when computing the coarse curvature, we may use W1(μx0σ,ε,Tμx0σ,ε). This is justified as terms involving the second fundamental form at the point x0 emerge only at the third order in the expansion of W1(μx0σ,ε,μyσ,ε), making precision up to O(δ4) sufficient.

The following will allow us to deduce a Wasserstein lower bound from an upper bound provided by an approximate transport map of degree 3, and merging these into a both-sided estimate up to O(δ4).

Lemma 2.25

If Inline graphic is smooth and takes the form

f(Tz)-f(z)=Tz-z+O(δ4)=O(δ) 2.14

and the magnitude of its gradient satisfies

graphic file with name 40879_2025_816_Equ203_HTML.gif

then for all σ,ε,δ sufficiently small with σεδ/4,

W1(μx0σ,ε,Tμx0σ,ε)=Tz-zdμx0σ,ε(z)+O(δ4)=(f(Tz)-f(z))dμx0σ,ε(z)+O(δ4).

Proof

Using the expansion 1/(1+a)=1-a+O(a2), we deduce that

graphic file with name 40879_2025_816_Equ204_HTML.gif

By the mean value theorem, a differentiable function divided by the supremum of its gradient is 1-Lipschitz. Then by Kantorovich–Rubinstein duality

graphic file with name 40879_2025_816_Equ205_HTML.gif

using the assumption (2.14) on the third and fourth line.

Finally, in order to integrate over the correct range of Fermi coordinates to cover precisely Bσ,ε(x0) as the support of μx0σ,ε, we need to find the range parameter ε(α) such that if

Bσ,ε(0):={(α,β):|α1|ε,j=2mαj2ε(α)2,i=1kβi2σ2}Rm+k

then

ϕ(Bσ,ε(0))=Bε(x0).

This is a necessary consideration, because in general non-flat spaces

ϕ(B~σ,ε(0))Bσ,ε(x0).

The following is a classical result of Toponogov, which is a generalization of the Pythagoras theorem for Riemannian manifolds and gives a characterisation of sectional curvature. See e.g. [24] for a proof.

Lemma 2.26

For any point x0M and any w1,w2Tx0M sufficiently small, the Riemannian distance between expx0(w1) and expx0(w2) has the expansion

d(expx0(w1),expx0(w2))=w1-w2-13R(w1,w2)w2,w1+O(max(w1,w2)5).

As a consequence, we deduce that given a coordinate α1(-ε,ε), the range parameter ε(α) is characterized by the relation

ε2=α12+ε(α)2+O(max(α12,ε(α)2))

where the coefficient in the remainder term only depends on a fixed neighbourhood of x0. This implies ε(α)=O(ε) and

ε(α)=(1+O(ε2))ε2-α12=ε2-α12+O(ε3).

We shall label the remainder term r(α)=O(ε3) for the purpose of the following proof. The next corollary will allow us to ignore the distinction between B~σ,ε(0) and Bσ,ε(0) up to O(δ4) whenever we integrate with respect to the test measure μx0σ,ε in Fermi coordinates.

Corollary 2.27

If Inline graphic is a polynomial with no constant term and max(σ,ε)δ then

graphic file with name 40879_2025_816_Equ206_HTML.gif

Proof

We split the domain of integral on the right so that one part matches the domain on the left and the integral of the other part is O(δ4):

graphic file with name 40879_2025_816_Equ207_HTML.gif

on the last line we using that P(α,β) has no constant term and r(α)=O(ε3)=O(δ3).

Curves and surfaces

We establish explicit formulas for the coarse extrinsic curvature defined by (1.3) in four practically relevant cases: a circle, a planar curve, a space curve, and a surface. We begin by presenting the common setup shared among all these cases.

The circle example

Our motivating example is the circle SR1 with a fixed radius R>0, which avoids technicalities arising from varying radius in the osculating circle, an issue that will be addressed in Sect. 3.2 in the case of planar curves.

Notation 3.1

Denote the polar coordinates

ϕ(α,β):=(R-β)cos(α/R)(R-β)sin(α/R), 3.1

where α(-πR,πR) parametrizes arc-length distance from the point (R, 0) along the circle and β(-σ,σ) parametrizes the direction normal to the circle.

Denote x0:=ϕ(0,0)=(R,0) and for every δ>0 denote y:=(Rcos(δ/R)Rsin(δ/R)).

Lemma 3.2

The test measures in polar coordinates take the form

(ϕ-1μyσ,ε)(dα,dβ)=14σε1(δ-ε,δ+ε)×(-σ,σ)(α,β)(1-βR)dαdβ.

Proof

At any (α,β), the radial coordinate is R-β, the radial length element is dβ and the angular element is dαR, giving the volume element (R-β)dβdαR=(1-βR)dαdβ, with 14σε as the probability normalization factor for the support (δ-ε,δ+ε)×(-σ,σ).

This is consistent with the formula of Proposition 2.9, as the mean curvature at (α,β) is 1R-β, which gives the density

e-01βR-sβds=elog(R-β)-logR=1-βR

on (δ-ε,δ+ε)×(-σ,σ).

The transport map of Definition 2.18 boils down to

T(ϕ(α,β))=ϕ(δ-α,β)=(R-β)cos((δ-α)/R)(R-β)sin((δ-α)/R), 3.2

and note that y=Tx0=T(ϕ(0,0)). See also Fig. 1 below.

Fig. 1.

Fig. 1

Planar curve case: test measures in red with some transport pairs of T in blue (Color figure online)

Remark 3.3

In this case the transport map T is precise in the sense that Tμx0σ,ε=μyσ,ε. Indeed, for any f:R2R Borel measurable,

graphic file with name 40879_2025_816_Equ208_HTML.gif

Proposition 3.4

For all δ,ε,σ>0 sufficiently small with σεδ/2, it holds that

W1(μx0σ,ε,μyσ,ε)=2R2sin(δ2R)1εsin(εR)(1+σ23R2)=x0-yRεsin(εR)(1+σ23R2).

Proof

For every point z=ϕ(α,β),

Tz-z=2(R-β)sin(δ-2α2R),

which is the Euclidean distance of two points on the circle at angle (δ-2α)/R apart. Integrating with respect to the test measure yields

W1(μx0σ,ε,μyσ,ε)Tz-zdμx0σ,ε(z)=14σε-σσdβ-εεdα(1-βR)2(R-β)sin(δ-2α2R)=14σε-σσdβ(1-βR)2(R-β)×-εεdα(sin(δ2R)cos(αR)-sin(αR)cos(δ2R))=2R2sin(δ2R)1εsin(εR)(1+σ23R2).

For the lower bound, we test against the 1-Lipschitz function

f(z):=z-x0,y-x0y-x0.

We have

y-x0=ϕ(δ,0)-ϕ(0,0)=Rcos(δ/R)Rsin(δ/R)-R0=R(cos(δ/R)-1)Rsin(δ/R),

and so y-x0=R2(1-cos(δ/R)=2Rsin(δ/(2R)), giving

y-x0y-x0=12sin(δ/(2R))cos(δ/R)-1sin(δ/R). 3.3

Then we compute using (3.1), (3.2) and (3.3):

f(Tz)-f(z)=ϕ(δ-α,β)-ϕ(α,β),y-x0y-x0=(R-β)2sin(δ/(2R))cos((δ-α)/R)-cos(α/R)sin((δ-α)/R)-sin(α/R)·cos(δ/R)-1sin(δ/R)=R-βsin(δ/(2R))(cos(α/R)-cos((δ-α)/R))=2(R-β)sin((δ-2α)/(2R))=Tz-z

by trigonometric identities. Therefore

W1(μx0σ,ε,μyσ,ε)f(z)(dμyσ,ε(z)-dμ0σ,ε(z))=(f(Tz)-f(z))dμ0σ,ε(z)=Tz-zdμ0σ,ε(z),

which shows the lower bound agrees exactly with the upper bound.

Planar curve

Let γ:(-δ0,δ0)R2 be a smooth unit speed curve. As before, let x0:=γ(0), y:=γ(δ) where δ(-δ0,δ0).

The normal vector field along γ is given by n(α):=γ¨(α)γ¨(α), the radius of the osculating circle is R(α):=1γ¨(α) and we have the relationships

γ¨(α)=n(α)R(α),γ(α)=-1R(α)2γ˙(α)-R˙(α)R(α)2n(α),n˙(α)=-γ˙(α)R(α),n¨(α)=R˙(α)R(α)2γ˙(α)-1R(α)2n(α). 3.4

Let ϕ:(-δ0,δ0)×(-σ0,σ0)R2 be given as follows:

ϕ(α,β):=γ(α)+βn(α).

This is the Fermi chart along γ. While we have the general Fermi coordinate representation in terms of the expansion in Lemma 2.17, in this case we arrive at a precise form:

Lemma 3.5

The test measures at y=γ(δ) are

(ϕ-1μyσ,ε)(dα,dβ)=14σε1(δ-ε,δ+ε)×(-σ,σ)(α,β)(1-βR(α))dαdβ.

Proof

To evaluate H(ϕ(α,β)) in applying Proposition 2.9, normalize the vector field tangent to the curve αϕ(α,β) and compute the second derivative in R2 as

ααϕ(α,β)(αϕ(α,β)αϕ(α,β))=α2ϕ(α,β)αϕ(α,β)2-α2ϕ(α,β),αϕ(α,β)αϕ(α,β)3αϕ(α,β).

The second term is tangential to the curve, so may be ignored for the computation of H. Moreover,

graphic file with name 40879_2025_816_Equ209_HTML.gif

Note that n(α) is normal to αϕ(α,β) for every β since

n(α),αϕ(α,β)=n(α),γ˙(α)+βn˙(α)=0,

therefore the mean curvature is

graphic file with name 40879_2025_816_Equ210_HTML.gif

Finally,

graphic file with name 40879_2025_816_Equ211_HTML.gif

and the Lebesgue measure of the support 14σε is the normalization factor because the β term vanishes when integrating over β(-σ,σ).

In this case the proposed transport map of Definition 2.18 reduces to

T(ϕ(α,β))=ϕ(δ-α,β-12R˙(0)R(0)2(σ2-β2)(δ-2α)).

As a consequence of Corollary 2.13,

Lemma 3.6

For all δ,ε,σ>0 sufficiently small with σεδ/4, it holds that

W1(μx0σ,ε,μyσ,ε)=W1(μ0σ,ε,Tμ0σ,ε)+O(δ4).

For notational ease we shall from here onwards denote R:=R(0) and R˙:=R˙(0).

Proposition 3.7

Let γ be a smooth unit speed curve in R2 such that γ(0)=x0 and γ(δ)=y. For all δ,ε,σ>0 sufficiently small with σεδ/4, it holds that

W1(μx0σ,ε,μyσ,ε)=x0-y(1-ε26R2+σ23R2)+O(δ4)

where R is the radius of the osculating circle of the curve at x0.

Proof

Lemma 3.6 allows computing W1(μ0σ,ε,Tμ0σ,ε) instead. Throughout the proof, terms of order δ4 and higher are absorbed into O(δ4). For the upper bound, we compute by expansion with respect to the orthonormal basis (γ˙(0),n(0)) at x0,

ϕ(α,β)=γ(0)+βn(0)+α(γ˙(0)+βn˙(0))+α22(γ¨(0)+βn¨(0))+α36γ(0)+O(δ4)=x0+(α-αβR-α36R2+βα2R˙2R2)γ˙(0)+(β+α22R-α3R˙6R2-βα22R2)n(0)+O(δ4)

having inserted for the derivatives at 0 using the list (3.4). Then the distance of the transport pairs up to order 4 is

T(ϕ(α,β))-ϕ(α,β)=ϕ(δ-α,β)-ϕ(α,β)=(δ-2α)(1-βR-16R2(δ2-δα+α2)+βR˙2R2δ+O(δ3))γ˙(0)+(δ2R+O(δ2))n(0) 3.5

having used the factorizations (δ-α)3-α3=(δ-2α)(δ2-δα+α2) and (δ-α)2-α2=δ(δ-α). By orthonormality of (γ˙(0),n(0)), we compute this norm as

graphic file with name 40879_2025_816_Equ23_HTML.gif 3.6

by the expansion 1+x=1+12x-18x2+O(x3) for the square root on the last line.

Moreover, expanding the volume distortion factor as

(1-βR(α))=1-βR+R˙R2αβ+O(δ3)

and multiplying the expression for T(ϕ(α,β))-ϕ(α,β) by this factor, we integrate and note that only terms of even order in both α and β contribute, yielding

W1(μx0σ,ε,μyσ,ε)14σε-εεdα-σσdβ(1-βR+R˙R2αβ+O(δ3))T(ϕ(α,β))-ϕ(α,β)=14σε-εεdα-σσdβ(1-βR+R˙R2αβ+O(δ3))×(δ-2α)(1-βR-δ224R2+δα6R2-α26R2+R˙2R2βδ)+O(δ4)=δ(1-δ224R2-ε26R2+σ23R2)+O(δ4)=y-x0(1-ε26R2+σ23R2)+O(δ4).

To obtain the factor y-x0 on the last line, we applied that

y-x0=δ(1-δ224R2)+O(δ4)

which can be deduced by plugging in for α=β=0 in the previous computation of T(ϕ(α,β))-ϕ(α,β). The σ2 coefficient came from integrating the β2 term of the integrand, σ23R2=12σ-σσ(-βR)×(-βR)dβ. The terms with odd power in α or β such as δαβ vanished as they are mean zero.

We proceed with showing the lower bound, using again the 1-Lipschitz test function

f(z):=z-x0,y-x0y-x0.

Express the vector between the centres of the two test measures, recalling γ(0)=x0,

γ(δ)-γ(0)=T(ϕ(0,0))-ϕ(0,0)=δ(1-δ26R2)γ˙(0)-δ(δ2R+O(δ2))n(0)+O(δ4).

This vector has magnitude

γ(δ)-γ(0)=δ(1-δ224R2)+O(δ4),

and so we deduce that

y-x0y-x0=γ(δ)-γ(0)γ(δ)-γ(0)=(1-δ28R2+O(δ3))γ˙(0)+(δ2R+O(δ2))n(0).

Then we compute, using the expression (3.5) for T(ϕ(α,β))-ϕ(α,β) obtained above,

f(Tz)-f(z)=T(ϕ(α,β))-ϕ(α,β),y-x0y-x0=(δ-2α)(1-βR-16R2(δ2-δα+α2)+βδR˙2R2)(1-δ28R2+O(δ3))+(δ-2α)(δ2R+O(δ2))(δ2R+O(δ2))+O(δ4)=(δ-2α)(1-βR-δ224R2+δα6R2-α26R2+βδR˙2R2)+O(δ4).

We see that this agrees with the pairwise transport distance (3.6) up to O(δ4), hence Lemma 2.25 applies and the upper and lower bounds agree up to an O(δ4) term.

Space curve

Let γ:(-δ0,δ0)R3 be a smooth, unit speed curve with velocity γ˙. Define the unit normal and binormal vector fields along γ as

n(α):=γ¨(α)γ¨(α),b(α):=γ˙(α)×n(α)γ˙(α)×n(α).

This yields the so-called Frenet–Serret frame (γ˙(α),n(α),b(α)) of R3 along γ. Writing R(α):=1γ¨(α) for the radius of the osculating circle and τ(α):=b˙(α) for the torsion, the Frenet–Serret formulas give relationships between the vector fields of the frame,

γ¨(α)=n(α)R(α),n˙(α)=-γ˙(α)R(α)+τ(α)b(α),b˙(α)=-τ(α)n(α). 3.7

From these, we deduce the higher order derivatives

γ(α)=-1R(α)2γ˙(α)-R˙(α)R(α)2n(α)+τ(α)R(α)b(α),n¨(α)=R˙(α)R(α)2γ˙(α)-(τ(α)2+1R(α)2)n(α)+τ˙(α)b(α),b¨(α)=τ(α)R(α)γ˙(α)-τ˙(α)n(α)-τ(α)2b(α). 3.8

We will employ the Frenet–Serret frame for explicit computations of distances between points in the tubular neighborhood of a space curve. Additionally, we will employ it in formulating a sufficiently accurate approximate transport map between test measures, represented through an expansion in Fermi coordinates.

Definition 2.14 for Fermi coordinates requires a choice of a local orthonormal frame of the normal bundle along γ. We choose (n1,n2) as follows:

n1(α):=n(α)-ατ(α)b(α)1+α2τ(α)2=(1+O(α2))n(α)-(ατ(α)+O(α2))b(α)),n2(α):=b(α)+ατ(α)n(α)1+α2τ(α)2=(1+O(α2))b(α)+(ατ(α)+O(α2))n(α)

where n,b come from the Frenet–Serret frame.

Definition 3.8

Define the Fermi coordinates ϕ:(-δ0,δ0)3R3, adapted to γ, by the formula:

ϕ(α,β1,β2):=γ(α)+β1n1(α)+β2n2(α)=γ(α)+(β1+αβ2τ(α)+O(δ3))n(α)+(β2-αβ1τ(α)+O(δ3))b(α).

Denote R=R(0), R˙=R˙(0), τ=τ(0), τ˙=τ˙(0).

Consider the family of curves {αϕ(α,β1,β2):(β1,β2)Bσ}. Denote the particular unit normal vector fields

n~(α,β1,β2):=1β12+β22(β1n1(α)+β2n2(α)).

The mean curvature of each curve in the direction n~ is expressed as

H(ϕ(α,β1,β2)),n~(α,β1,β2)=n~(α,β1,β2),ααϕ(α,β1,β2)(αϕ(α,β1,β2)αϕ(α,β1,β2))=n~(α,β1,β2),α2ϕ(α,β1,β2)αϕ(α,β1,β2)2.

where the second equality holds because n~ is normal to αϕ by Lemma 2.4.

Remark 3.9

We perform computations in terms of the Frenet–Serret frame as it can be interpreted in terms of the radius of the osculating circle and torsion of the curve.

  1. As a special case of Lemma 2.17, using that the mean curvature components are
    H1(ϕ(0))=n(0),γ¨(0)=n(0),n(0)R(0)=1R(0),H2(ϕ(0))=b(0),γ¨(0)=b(0),n(0)R(0)=0,
    the test measures at every y=γ(δ) in Fermi coordinates along γ are
    (ϕ-1μyσ,ε)(dα,dβ1,dβ2)=1B~σ,ε(α,β1,β2)B~σ,ε(1+r(α,β1,β2))d(ϕ-1μyσ,ε)(α,β1,β2)×(1-β1R(0)+r(α,β1,β2))dαdβ1dβ2
    where r(α,β1,β2)=O(δ2) is the remainder.
  2. The proposed transport map of Definition 2.18 reduces, in this case, to
    T(ϕ(α,β1,β2))=ϕ(δ-α,β1+O(δ3),β2+O(δ3)).
    The transport map in this case is depicted in Fig. 2. These expressions will be used in the proof of the next theorem.
Fig. 2.

Fig. 2

Space curve case: test measures in red with some transport pairs of T in blue (Color figure online)

Theorem 3.10

Let γ:(-δ0,δ0)R3 be a space curve with x0=γ(0),y=γ(δ) and μx0σ,ε,μyσ,ε the test measures defined in Definition 2.16 with coordinate representation of Remark 3.9. For all δ,σ,ε>0 sufficiently small and with σεδ/4, it holds that

W1(μx0σ,ε,μyσ,ε)=x0-y(1+σ24R2-ε26R2)+O(δ4).

where R=1γ¨(0) is the radius of the osculating circle.

Proof

Due to Lemma 3.6, it is sufficient to work with the distance W1(μx0σ,ε,Tμx0σ,ε) as it approximates W1(μx0σ,ε,μyσ,ε). The computation of the pairwise distances is similar to the planar curve case, (3.5), with additional terms due to the component b(α). Concretely, since

b(α)=b(0)+αb˙(0)+12α2b¨(0)+O(α3)=α2τ2Rγ˙(0)-(ατ+12α2τ˙)n(0)+(1-12α2τ2)b(0)+O(α3),

and using the derivatives (3.7) and (3.8), compute

ϕ(α,β1,β2)=x0+(α-αβ1R-α36R2+β1α2R˙2R2+β2α2τ2R+O(δ4))γ˙(0)+(β1+α22R+O(δ3))n(0)+(β2+O(δ3))b(0).

Then similarly to (3.5) we obtain

T(ϕ(α,β1,β2))-ϕ(α,β1,β2)=γ(δ-α)+β1n(δ-α)+β2b(δ-α)-γ(α)-β1n(α)-β2b(α)=(δ-2α)(1-β1R-16R2(δ2-δα+α2)+β1R˙2R2δ+β2δτ2R+O(δ3))γ˙(0)+(12Rδ+O(δ2))n(0)+O(δ2)b(0)=(δ-2α)(1-β1R-δ224R2+δα6R2-α26R2+β1R˙2R2δ+β2δτ2R+O(δ3)). 3.9

The Wasserstein distance upper bound is then computed by integration with respect to μx0σ,ε using the coordinate representation of Remark 3.9 as

graphic file with name 40879_2025_816_Equ212_HTML.gif

applying on the last line that x0-y=δ(1-δ224R2)+O(δ4). In the integral on the first line, terms of odd order vanish upon integration, and the remaining terms amount to integration of quadratic polynomials.

We now address the lower bound. Analogously to the plane curve case, define the test function for the Kantorovich–Rubinstein duality as

f(ϕ(α,β1,β2)):=ϕ(α,β1,β2)-x0,y-x0y-x0

which is again clearly 1-Lipschitz in R3. We wish to apply Lemma 2.25 to show the lower bound and upper bound coincide up to O(δ4). Noting that y-x0=ϕ(δ,0,0)-ϕ(0,0,0), we deduce from (3.9) that

y-x0=δ(1-δ26R2+O(δ3))γ˙(0)+δ(δ2R+O(δ2))n(0)+O(δ3)b(0),y-x0=δ(1-δ224R2+O(δ3)),

and so

y-x0y-x0=(1-δ28R2+O(δ3))γ˙(0)+(δ2R+O(δ2))n(0)+O(δ2)b(0).

Therefore

f(ϕ(T(α,β1,β2)))-f(α,β1,β2)=T(ϕ(α,β1,β2))-ϕ(α,β1,β2),y-x0y-x0=(δ-2α)(1-β1R-δ224R2+δα6R2-α26R2+β1δR˙2R2+β2δτ2R+O(δ3)).

This is the same expression as for T(ϕ(α,β1,β2))-ϕ(α,β1,β2), hence Lemma 2.25 applies and the lower and upper bounds agree up to O(δ4).

Surface

We now consider a smooth 2-surface MR3 and γ:(-1,1)M a unit speed geodesic in M, denoting again x0:=γ(0),y:=γ(δ) for δ>0 sufficiently small. Let nΓ(TM) be the unit normal vector field and mΓ(TM|γ) the unit vector field along γ orthogonal to the velocity γ˙. Both n and m are unique up to sign.

Definition 3.11

  • Define the Fermi coordinates ψ:(-δ0,δ0)×(-ε0,ε0)M along γ in M as
    ψ(α1,α2)=expM,γ(α1)(α2m(α1)).
  • Define the Fermi coordinates ϕ:(-δ0,δ0)×(-ε0,ε0)×(-σ0,σ0)R3 along γ in R3 adapted to the surface M as
    ϕ(α1,α2,β)=ψ(α1,α2)+βn(α1,α2).
    See Fig. 3 for a graphical representation of ψ and ϕ.
  • For i,j{1,2} denote the components of the second fundamental form in the Fermi coordinates
    graphic file with name 40879_2025_816_Equ213_HTML.gif

Remark 3.12

We point out that we overload the second fundamental form symbol Inline graphic depending on the context of use. In the notation (1.1) and in the statements of Theorem 3.16 and Theorem 4.1, the subscript is the point x0 on the manifold and the bracket arguments are tangent vectors. On the other hand, in coordinate computations taking place in the proofs, the subscripts will represent components with respect to the Fermi frame at Fermi coordinates α,β in brackets.

Similarly to the Frenet–Serret frame in the case of a planar curve, we now consider the orthonormal frame (γ˙,m,n) with the intent to expand at x0, i.e. α1=α2=β=0.

Lemma 3.13

The first derivatives of the normal vector field at (α1,α2)=0 are

graphic file with name 40879_2025_816_Equ214_HTML.gif

Hence the derivatives of ϕ at (α1,α2,β)=0 up to third order are

graphic file with name 40879_2025_816_Equ215_HTML.gif

Proof

The derivatives involving β are clear, recalling the definition

ϕ(α1,α2,β):=ψ(α1,α2)+βn(α1,α2),

and the first derivatives in α1,α2 follow from the definition of ψ(α1,α2).

For α1n(0) we check its components with respect to the frame (γ˙,m,n),

graphic file with name 40879_2025_816_Equ216_HTML.gif

and similarly for α2n(0).

For the second derivatives in α1,α2 at (α1,0,0) for any α1(-δ0,δ0) and j=1,2,

graphic file with name 40879_2025_816_Equ217_HTML.gif

having introduced the term α1Mαjψ(α1,0), which vanishes for j=1,2, because α1ψ(α1,0) is a geodesic on M and m(α1) is the parallel translation of m(0) along γ. By the same argument, for any (α1,α2)(-δ0,δ0)×(-ε0,ε0),

graphic file with name 40879_2025_816_Equ218_HTML.gif

because α2ψ(α1,α2) is a geodesic for every α1(-δ0,δ0).

For the second derivatives in β and one of α1 and α2, deduce βαiϕ(0)=αin(0) and plug in for αin(0).

For the third derivatives at (α1,α2,β)=0, write by the chain rule

graphic file with name 40879_2025_816_Equ219_HTML.gif

and plug in for αin(0) in each.

Denote DuV the plain derivative in the direction uRn of a vector field V as a smooth map from an open subset of Rn to Rn.

Notation 3.14

Denote B~σ,ε:={(α1,α2,β):α12+α22<ε2,|β|<σ}R3.

Consider the family of surfaces {ϕ(U,β):β(-σ0,σ0)}. For any β(-σ0,σ0), we denote the unit normal vector field of the surface as n(α1,α2), which is unique up to sign. The corresponding mean curvature is:

H(ϕ(α1,α2,β))=n(α1,α2),i=12i3ei(ϕ(α1,α2,β))

where (e1,e2) is an orthonormal frame on each ϕ(U,β).

Remark 3.15

  1. As a special case of Lemma 2.17, the test measures at y=γ(δ) in these Fermi coordinates are
    graphic file with name 40879_2025_816_Equ27_HTML.gif 3.10
    where r(α,β)=O(δ2) is a second order remainder.
  2. The proposed transport map of Definition 2.18 reduces, in this case, to
    T(ϕ(α1,α2,β))=ϕ(δ-α1,α2+O(δ3),β+O(δ3)).
    See Figs. 4, 5 and 6 for a pictorial representation of this map.
Fig. 4.

Fig. 4

Test measures in red with some transport pairs of T in blue (Color figure online)

Fig. 5.

Fig. 5

Top-down perspective for the transport map T

Fig. 6.

Fig. 6

Cross-sectional perspective for the transport map T

Theorem 3.16

Let M be an isometrically embedded surface in R3, let x0 be a point and (e1,e2) an orthonormal basis of principal curvature directions at x0. Let γ be a unit speed geodesic in M with γ(0)=x0, γ˙(0)=e1 and denote y=γ(δ). For all δ,ε,σ>0 sufficiently small with σεδ/4, it holds that

graphic file with name 40879_2025_816_Equ220_HTML.gif

Remark 3.17

If we set ε=223σ, we note that the bracket on the right reduces to 1. This is due to the effects of second fundamental form and the curvature of the submanifold cancelling out, so it would appear in such special case that the coarse extrinsic curvature is flat, even though the second fundamental form may be non-vanishing. Such a special case is due to having an additional degree of freedom because of the additional σ parameter and the sign of the σ2 term happens to oppose that of the ε2 term. The extrinsic curvature should thus be seen as encapsulated by varying both σ and ε in W1(μx0σ,ε,μyσ,ε).

Proof

The conclusion of Corollary 2.13 holds, so we may compute W1(μx0σ,ε,Tμx0σ,ε) instead. For every point ϕ(α1,α2,β), expanding up to third order and using the list of derivatives of Lemma 3.13, we collect terms as components of the frame (γ˙,m,n) at 0,

graphic file with name 40879_2025_816_Equ28_HTML.gif 3.11

While the terms βαiαjβαiαjϕ(0) are only of order 3, they are linear in β, and hence will not influence the integral with respect to μx0σ,ε up to O(δ4). In the same way an expression for the proposed transport

T(ϕ(α1,α2,β))=ϕ(δ-α1,α2+O(δ3),β+O(δ3))

can be obtained by making corresponding substitutions for the components in the above expression for ϕ(α1,α2,β). Then the pointwise transport vector is

graphic file with name 40879_2025_816_Equ29_HTML.gif 3.12

and its magnitude is

graphic file with name 40879_2025_816_Equ30_HTML.gif 3.13

Using the density of the test measure μx0σ,ε in Fermi coordinates given by (3.10), the upper bound is

graphic file with name 40879_2025_816_Equ221_HTML.gif

In the third equality, we plugged in for T(ϕ(α,β))-ϕ(α,β) as computed above and used that terms of odd order in one of α1,α2,β integrate to 0 and again absorbed higher order terms into O(δ4). On the last line, we used that

graphic file with name 40879_2025_816_Equ222_HTML.gif

We proceed with showing the lower bound. Define

p(α2,β):=ϕ(δ,α2,β)-ϕ(0,α2,β)ϕ(δ,α2,β)-ϕ(0,α2,β)

and the test function

f(ϕ(α1,α2,β)):=ϕ(α1,α2,β)-x0,p(α2,β) 3.14

for the Kantorovich–Rubinstein duality, with the intention of applying Lemma 2.25 to conclude. We first expand

graphic file with name 40879_2025_816_Equ223_HTML.gif

Deduce

graphic file with name 40879_2025_816_Equ224_HTML.gif

and therefore

graphic file with name 40879_2025_816_Equ32_HTML.gif 3.15

Then it can be verified using expansions (3.12) and (3.15) to compute the inner product that

f(Tz)-f(z)=T(ϕ(α1,α2,β))-ϕ(α1,α2,β),p(α2,β)=T(ϕ(α1,α2,β))-ϕ(α1,α2,β)+O(δ4)

by comparison with (3.13).

It remains to show that the magnitude of the gradient of f satisfies

supzB2δ(x0)f(z)=1+O(δ3).

For this we need to expand the inverse matrix of the metric in Fermi coordinates. Using the expansion (3.11), compute

graphic file with name 40879_2025_816_Equ225_HTML.gif

We shall label the term

r(α):=13α1α12n(0),γ˙(0)+13α2α1α2n(0),γ˙(0).

Then the metric matrix has the shape

G=g11g120g21g220001

with

graphic file with name 40879_2025_816_Equ226_HTML.gif

Note that the matrix is of the form

G=I+A

with A=O(δ), which means the expansion of its inverse is

G-1=I-A+A2+O(δ3).

We compute

graphic file with name 40879_2025_816_Equ227_HTML.gif

and thus

graphic file with name 40879_2025_816_Equ228_HTML.gif

From (3.15) we deduce the derivatives of the projection vector field in coordinates are

α2p(α2,β)=O(δ2)γ˙(0)+O(δ)m(0)+O(δ)n(0),βp(α2,β)=O(δ2)γ˙(0)+O(δ)m(0)+O(δ)n(0).

Then the first derivatives of the test function defined in (3.14) are

graphic file with name 40879_2025_816_Equ229_HTML.gif

Then the magnitude of the gradient is

f(ϕ(α1,α2,β)2=(g11ϕ)(α1(fϕ))2+2(g12ϕ)α1(fϕ)α2(fϕ)+(g22ϕ)(α2(fϕ))2+(β(fϕ))2,

and we find the individual summands

graphic file with name 40879_2025_816_Equ230_HTML.gif

which indeed gives

f(ϕ(α1,α2,β)=1+O(δ3)

as the first and second order terms cancel out. Hence Lemma 2.25 applies and we conclude the lower bound coincides up to O(δ4) with the upper bound.

General Riemannian submanifolds

We now consider a Riemannian submanifold M of arbitrary dimension m and codimension k embedded isometrically in Rm+k. Theorems 3.10 and 3.16 are thus special cases of Theorem 4.1 below. We begin by defining an orthonormal frame of Rm+k-valued vector fields on a sufficiently small open domain U in the submanifold M, which is used to define the Fermi coordinates on U in this general setting.

Frame extension

We take the ambient manifold to be Rm+k. Recall the second fundamental form at a point xM is

graphic file with name 40879_2025_816_Equ231_HTML.gif

where W is an arbitrary local vector field on M with W(x)=w2. The mean curvature at x is

graphic file with name 40879_2025_816_Equ232_HTML.gif

for an arbitrary orthonormal basis (ej)j=1m of TxM. Both Inline graphic and H(x) are normal to the submanifold, i.e.

graphic file with name 40879_2025_816_Equ233_HTML.gif

Recall from Definition 2.14 that the Fermi coordinates in M along γ are given by

ψ(α)=expM,γ(α1)(j=2mαjej(α1)),

where (ej(α1)j=1m is the parallel transport along γ of an orthonormal basis (ej(0))j=1m of Tx0M with e1(0)=γ˙(0). We refer back to Sect. 2 for properties of the Fermi chart.

Denote α^=(α2,,αm) so that α=(α1,α^). Extend the frame (ej(α1))j=1m defined along α1γ(α1) to UM by imposing

Ddsej(α1,sα^)=0,

i.e. by parallel translating in M along the geodesic sψ(α1,sα^).

Given an initial orthonormal basis (ni)i=1k of Tx0M, first extend it to a frame along γ by requiring that

α1ni(α1),ej(α1)=0and(α1ni(α1))=0for allα1(-δ0,δ0).

The first requirement implies

graphic file with name 40879_2025_816_Equ234_HTML.gif

which together with the second requirement implies the first order ODE

graphic file with name 40879_2025_816_Equ33_HTML.gif 4.1

The solution exists and is unique by standard ODE theory. Having defined the frame (ni(α1))i=1k along the geodesic α1γ(α1), we may also extend it to the submanifold by requiring that for every α1(-δ0,δ0) and α^B~0m-1,

graphic file with name 40879_2025_816_Equ235_HTML.gif

Similarly to the above, the first requirement implies that for all j=1,,m,

ddsni(α1,sα^),ej(α1,sα^)=-ni(α1,sα^),ddsej(α1,sα^),

and from the second requirement we conclude the frame satisfies the first order ODE

ddsni(α1,sα^)=-j=1mni(α1,sα^),ddsej(α1,sα^)ej(α1,sα^)

along each geodesic sϕ(α1,sα^) in M.

With these concrete vector fields, recall the Fermi coordinates in Rm+k along γ adapted to the submanifold M were defined in Definition 2.14 as

ϕ(α,β)=ψ(α)+i=1kβini(α)

and note that ϕ(α,0)=ψ(α).

For every α1(-δ0,δ0), the map ψ(α1,·):B~0m-1M is the exponential chart on its image. It is known that the Christoffel symbols vanish at the centre for such charts, i.e.

αiψMαjψ(α1,0)=0for alli,j=2,,m.

Moreover, since αjψ(α1,0)=ej(α1) for j=1,,m is parallel transport of ej(0) along γ, also

α1ψMαjψ(α1,0)=0for allj=1,,m,

noting that γ˙(α1)=α1ψ(α1,0).

Denote the components of the second fundamental form with respect to the Fermi coordinates as

graphic file with name 40879_2025_816_Equ236_HTML.gif

Note that the first index represents the normal direction and the latter two represent manifold directions. Then we can write for every j,=1,,m,

graphic file with name 40879_2025_816_Equ34_HTML.gif 4.2

In addition, (4.1) can be written as

graphic file with name 40879_2025_816_Equ237_HTML.gif

Thus the third derivatives with at least one in α1 are

graphic file with name 40879_2025_816_Equ35_HTML.gif 4.3

Main theorem

In the statement of the theorem, Inline graphic is the vector of second fundamental form. In the proof exclusively, Inline graphic denotes the ij-component of the second fundamental form with respect to the Fermi frame at Fermi coordinates α,β.

[Style2 Style3 Style3]Theorem 4.1

Let M be an isometrically embedded submanifold of Rm+k, and γ a unit speed geodesic in M such that γ(0)=x0 and γ(δ)=y. Let (ej)j=1m be an orthonormal basis of Tx0M with e1=γ˙(0) and assume that Inline graphic for all j=2,,m. Then for every σ,ε,δ>0 sufficiently small with σεδ/4 it holds that

graphic file with name 40879_2025_816_Equ238_HTML.gif

Remark 4.2

We point out two special cases:

  • If the submanifold has dimension 1 then the condition on the second fundamental form is trivially satisfied as there are no submanifold directions other than that of the curve itself. In this case
    graphic file with name 40879_2025_816_Equ239_HTML.gif
    and hence Inline graphic. This is the square curvature of the curve and for m=1, k=2 agrees with Theorem 3.10.
  • If the submanifold has codimension 1 with a normal vector field n on the submanifold, then the orthonormal eigenbasis of Inline graphic satisfies the condition Inline graphic for ij. Such a basis always exists as Inline graphic is symmetric and consists of the so-called principal curvature directions. Thus for m=2, k=1, we obtain Theorem 3.16 as a special case.

  • In general codimension, however, such a basis may not exist for a general submanifold, hence the assumption on the second fundamental form needs to be made and is highly restrictive. If this assumption was dropped, the upper bound for the Wasserstein distance via the proposed transport map would still apply. However, the computation of the lower bound using a projection plane, as done in the proof of Theorem 3.16 and applied again in the proof below, would yield additional lower order terms not agreeing with the upper bound. This is symptomatic of the non-optimality of the transport map up to third order. The more general computation including the off-diagonal terms to show this is straightforward but rather lengthy and is thus omitted. Qualitatively, the issue is that the off-diagonal terms of the second fundamental form introduce a deformation of the supports of the test measures which is not easily remedied and leaves the fully general case open. The deformation arises because the principal curvature directions above the reference point x0 for each leaf of the foliation of the tubular neighbourhood change their vertical alignment as we consider leaves further away from the base submanifold M. On the other hand, the diagonal assumption on the second fundamental form ensures an aligned stacking of principal curvature directions of leaves above x0, leading to the favourable cylinder-like support of the test measures.

  • For the interpretation of the special case of the parameters ε=2(m+2)k+2σ, we refer back to Remark 3.17.

Proof of Theorem 4.1

Expand the Fermi chart up to and including third order as

graphic file with name 40879_2025_816_Equ240_HTML.gif

From the definition of the Fermi chart and (4.2), (4.3), we have the derivatives at the origin on the right-hand side:

graphic file with name 40879_2025_816_Equ241_HTML.gif

With these we obtain:

graphic file with name 40879_2025_816_Equ242_HTML.gif

In the above, the sum of third derivative terms in α was split into those that involve at least one power in α1, for which we have a formula, and those that don’t. The other third derivatives αiαrαϕ(0) for i,r,2 are not easily written in Fermi coordinates, but will not be needed for our computations. Rearranging the terms, we write ϕ in terms of the basis (e1(0),,em(0), n1(0),,nk(0)) and apply the assumption Inline graphic:

graphic file with name 40879_2025_816_Equ36_HTML.gif 4.4

We will henceforth denote

ri(α):=12α1α12ni(0),e1(0)+=2mααα1ni(0),e1(0).

Let T be the transport map defined in Definition 2.18. With asymptotic notation for the third order terms,

T(ϕ(α1,α^,β))=ϕ(δ-α1,α^,β+O(δ3)).

In the expansion of ϕ above, from the third derivatives in α we only needed to specify those involving α1, because the transport map T changes only the first coordinate up to O(δ3). These derivatives were given by (4.3). Then the pointwise transport vector is

graphic file with name 40879_2025_816_Equ37_HTML.gif 4.5

Therefore, using the expansion 1+x=1+12x-18x2+O(x3), the pointwise transport distance is

graphic file with name 40879_2025_816_Equ38_HTML.gif 4.6

Lemma 2.17 expressed the density of the test measure μx0σ,ε in Fermi coordinates up to second order. Denoting the second order remainder of the density as r(α,β), the density simplifies to give

graphic file with name 40879_2025_816_Equ39_HTML.gif 4.7

where the form of the normalizing factor in the denominator is deduced from the two facts

graphic file with name 40879_2025_816_Equ243_HTML.gif

We deduce the upper bound in the statement of Theorem 4.1 by computing the integral on the right side of the inequality

graphic file with name 40879_2025_816_Equ244_HTML.gif

up to and including third order terms. Using the product of expressions (4.7) and (4.6), this amounts to integrating a quadratic polynomial in α,β. First, as terms with odd power in one of the coordinates vanish, we simplify the integral to

graphic file with name 40879_2025_816_Equ245_HTML.gif

We now use the fact that the average integral of the square of any coordinate over a d-dimensional ball of arbitrary radius r>0 is

graphic file with name 40879_2025_816_Equ246_HTML.gif

where Inline graphic denotes the integral normalised by the volume of the ball and using that

|Brd|=πd/2Γ(d/2+1)rd,|Bsd|=2πd/2Γ(d/2)sd-1.

This in particular gives

graphic file with name 40879_2025_816_Equ247_HTML.gif

Then

graphic file with name 40879_2025_816_Equ248_HTML.gif

Furthermore, from (4.6) for α=0,β=0 we deduce

graphic file with name 40879_2025_816_Equ249_HTML.gif

Therefore, we can rewrite in terms of the Euclidean distance:

graphic file with name 40879_2025_816_Equ250_HTML.gif

We now address the lower bound. Denoting

p(α^,β):=ϕ(δ,α^,β)-ϕ(0,α^,β)ϕ(δ,α^,β)-ϕ(0,α^,β),

emphasizing that this vector does not depend on α1, we propose

f(ϕ(α,β)):=ϕ(α,β)-x0,p(α^,β)

as the test function for Kantorovich–Rubinstein duality, with the intention of applying Lemma 2.25 to conclude the upper bound is also a lower bound up to O(δ4). We deduce from (4.5) that

graphic file with name 40879_2025_816_Equ40_HTML.gif 4.8

Then it can be verified, using the expansions (4.5) and (4.8) to compute the inner product up to and including third order terms, that

f(T(ϕ(α,β)))-f(ϕ(α,β))=T(ϕ(α,β))-ϕ(α,β)+O(δ4),p(α^,β)=T(ϕ(α,β))-ϕ(α,β)+O(δ4)

by comparison with (4.6).

Finally, we wish to compute the square magnitude of the gradient of the test function in order to verify that its supremum over B2δ(x0) is 1+O(δ3) for Lemma 2.25 to apply. For this we need to establish the Riemannian metric in Fermi coordinates gij=αiϕ,αjϕ. The first derivatives of the Fermi chart are deduced by differentiating (4.4) as

graphic file with name 40879_2025_816_Equ251_HTML.gif

Then the entries of the inverse metric matrix are computed from these to be

graphic file with name 40879_2025_816_Equ252_HTML.gif

Note that α1ϕ and g11 needed to be expanded up to second order due to the particular role of the first coordinate. For the rest, expansion up to first order is sufficient. The above means the metric matrix has the block structure

G=(gj)j,mO(δ2)O(δ2)Ik.

In particular, denoting

graphic file with name 40879_2025_816_Equ253_HTML.gif

and the matrix

graphic file with name 40879_2025_816_Equ254_HTML.gif

having used that a1j=0 for j=2,,m as Inline graphic by assumption, we can write

G=Im+k+AO(δ2)O(δ2)0.

Noting that the second matrix is O(δ), the expansion of its inverse is

G-1=Im-A+A2+O(δ3)O(δ2)O(δ2)Ik+O(δ4)

due to the block structure. Computing

graphic file with name 40879_2025_816_Equ255_HTML.gif

we deduce

graphic file with name 40879_2025_816_Equ256_HTML.gif

by plugging in for aj and b, and also

graphic file with name 40879_2025_816_Equ257_HTML.gif

We remark that for j,2 the expansion of gj up to the linear term suffices for the computations to follow, while the expansion of g11 up to second order is necessary.

We now compute the expansions of the derivatives of the test function. The first derivatives of the projection vector field in coordinates can be computed from (4.8) as

αjp(α^,β)=O(δ)for all2jm,βip(α^,β)=O(δ)for all1ik.

Then computing the inner products, using (4.4) for the derivatives of the chart,

graphic file with name 40879_2025_816_Equ258_HTML.gif

and for 2jm,

αj(fϕ)(α,β)=αjϕ(α,β),p(α^,β)+ϕ(α,β)-x0,αjp(α^,β)=O(δ2),

and for ik,

graphic file with name 40879_2025_816_Equ259_HTML.gif

We wish to compute

f(ϕ(α,β))2=j,=1mgj(ϕ(α,β))αj(fϕ)(α,β)α(fϕ)(α,β)+2i=1kj=1mgm+i,j(ϕ(α,β))βi(fϕ)(α,β)αj(fϕ)(α,β)+i=1k(βi(fϕ)(α,β))2.

The individual summands are

graphic file with name 40879_2025_816_Equ260_HTML.gif

All first and second order terms vanish upon summation, hence we may conclude that f(ϕ(α,β))2=1+O(δ3) as required.

Applications

Poisson point processes on manifolds

In applications one may wish to recover curvature information from coarse curvature of a random point cloud represented by a Poisson point process. Such an approach has already been investigated in [18] and [2] for the Ricci curvature and generalised Ricci curvature, respectively.

We first recall the definition of a Poisson point process. Let Inline graphic be a σ-finite measure space, Inline graphic the set of measures on Inline graphic and Inline graphic a probability space.

Definition 5.1

A Poisson point process on Inline graphic with intensity measure μ is a random measure Inline graphic (equivalently Inline graphic) such that the following three properties hold:

  • For all μ-finite measurable sets Inline graphic: Inline graphic is a Poisson(μ(A)) random variable.

  • For all disjoint, measurable μ-finite sets Inline graphic: Inline graphic and Inline graphic are independent random variables.

  • For all ωΩ: Inline graphic is a measure on Inline graphic.

It turns out (see [20, Chapter 6]) that all Poisson point processes with a finite intensity measure take the form of a random empirical measure, i.e.

graphic file with name 40879_2025_816_Equ261_HTML.gif

where N is a Inline graphic random variable, (Xi)iN are independent μ-distributed random variables on Inline graphic and (Xi)iN,N are independent. Denote the random set of points thus generated by Inline graphic as

graphic file with name 40879_2025_816_Equ262_HTML.gif

Notation 5.2

Let Inline graphic be a sequence of Poisson point processes on the ambient space Rm+k with uniform intensity measure nvolRm+k(dz). Denote by Inline graphic the discrete random set of points generated by Inline graphic. Let x0M, (δn)nN, (σn)nN,(εn)nN sequences of positive reals and yn:=expx0(δnv) for a fixed unit vector vTx0M. As the discrete counterpart to the test measures μxσ,ε, for any point xM denote the random empirical measures adapted to the submanifold,

graphic file with name 40879_2025_816_Equ263_HTML.gif

If σnεnδn/4 then Bσn,εn(x0)Bσn,εn(yn)x0+[-2δn,2δn]m+k.

Using the following result proved in [18, Corollary 3], it is possible to quantify the approximation of the test measures by the empirical measures in the Wasserstein metric:

Lemma 5.3

For all nN, it holds that

graphic file with name 40879_2025_816_Equ41_HTML.gif 5.1

We may then deduce that coarse curvature of point clouds with the empirical measures as test measures has the same limit as coarse extrinsic curvature if the intensity of the point process increases fast enough relative to the parameter δn. Denote

graphic file with name 40879_2025_816_Equ264_HTML.gif

This leads immediately to a corollary of Theorem 4.1:

Proposition 5.4

Under the assumptions of Theorem 4.1, if the sequences (δn)nN, (σn)nN and (εn)nN satisfy σnεnδn/4 and log(n)n-1m+k=o(δn3), then

graphic file with name 40879_2025_816_Equ265_HTML.gif

Proof

By the triangle inequality and (5.1),

graphic file with name 40879_2025_816_Equ266_HTML.gif

At the same time, from Theorem 4.1 we have

graphic file with name 40879_2025_816_Equ267_HTML.gif

which gives the final result upon substitution and taking the limit as n.

Retrieving mean curvature

Theorem 4.1 could in practice be exploited in the two settings already alluded to in the introduction, which considered the planar curve case for illustrative purposes. In the scope of generality of Theorem 4.1, we have

graphic file with name 40879_2025_816_Equ268_HTML.gif

In particular, we may distinguish two limit regimes:

  • Assuming σ=Θ(δ) and ε=o(σ),
    graphic file with name 40879_2025_816_Equ269_HTML.gif
    This represents a situation where one can obtain a sample from the ambient measure in a tubular neighbourhood of the surface. Decreasing ε corresponds to localization of the geometric information thus retrieved.
  • Assuming ε=Θ(δ) and σ=o(ε),
    graphic file with name 40879_2025_816_Equ270_HTML.gif
    In this case, we have a noisy sample from the surface and obtain convergence of the coarse extrinsic curvature under attenuation of the noise as σ decreases.

Note that these expressions depend on the vector v with yδ=expM,x0(δv). We can remove this directionality by adding up coarse curvatures in all directions of an orthonormal frame at x0, thus obtaining an expression involving the mean curvature.

Denote the square norm of the mean curvature vector as

H(x0)2=i=1kH(x0),ni(x0)2

for an arbitrary orthonormal basis (ni(x0))i=1k of the normal space Inline graphic.

Corollary 5.5

Let (ej)j=1m be an orthonormal basis of Tx0M, and for j=1,,m, let yj=expM,x0(δej). Assume that Inline graphic for ij. Then for all σ,ε,δ>0 sufficiently small with σεδ/4 it holds that

j=1m(1-W1(μx0σ,ε,μyjσ,ε)x0-yj)=(ε22(m+2)-σ2k+2)H(x0)2+O(δ3).

Proof

We express the coarse curvatures using the expansion of Theorem 4.1 and sum up, noting that j=1,,m indexing each direction plays the role of the first coordinate,

graphic file with name 40879_2025_816_Equ271_HTML.gif

completing the proof.

This implies that given the family of coarse curvatures

{1-W1(μx0σ,ε,μyjσ,ε)x0-yj:σ,ε,δ>0,j=1,,m},

one can retrieve the square magnitude of the mean curvature vector of the surface at x0 as

graphic file with name 40879_2025_816_Equ272_HTML.gif

In conclusion, we introduced the notion of coarse extrinsic curvature of Riemannian submanifolds embedded isometrically in a Euclidean space and verified that in a scaled limit of the parameters we retrieve meaningful geometric information about the submanifold. As illustrative examples, in the case of a curve we retrieve the inverse squared radius of the osculating circle at a given point, while in the case of a 2-surface we obtain an expression in terms of the second fundamental form and mean curvature. Such coarse extrinsic curvatures can be combined to yield the square magnitude of the mean curvature as a scaled limit.

Declarations

Competing interests

The authors declare no competing interests.

Footnotes

Xue-Mei Li acknowledges support from Engineering and Physical Sciences Research Council grant EP/V026100/1. Benedikt Petko was supported by the Engineering and Physical Sciences Research Council Centre for Doctoral Training in Mathematics of Random Systems: Analysis, Modelling and Simulation (EP/S023925/1).

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

References

  • 1.Ambrosio, L., Gigli, N., Savaré, G.: Diffusion, optimal transport and Ricci curvature for metric measure spaces. Eur. Math. Soc. Newsl. 103, 19–28 (2017) [Google Scholar]
  • 2.Arnaudon, M., Li, X.M., Petko, B.: Coarse Ricci curvature of weighted Riemannian manifolds (2023). arXiv:2303.04228
  • 3.Boissonnat, J.-D., Guibas, L.J., Oudot, S.Y.: Manifold reconstruction in arbitrary dimensions using witness complexes. Discrete Comput. Geom. 42(1), 37–70 (2009) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Bonciocat, A.-I., Sturm, K.-T.: Mass transportation and rough curvature bounds for discrete spaces. J. Funct. Anal. 256(9), 2944–2966 (2009) [Google Scholar]
  • 5.Chazal, F., Cohen-Steiner, D., Lieutier, A., Mérigot, Q., Thibert, B.: Inference of curvature using tubular neighborhoods. In: Najman, L., Romon, P. (eds.) Modern Approaches to Discrete Curvature. Lecture Notes in Mathematics, vol. 2184, pp. 133–158. Springer, Cham (2017) [Google Scholar]
  • 6.Chazal, F., Cohen-Steiner, D., Mérigot, Q.: Boundary measures for geometric inference. Found. Comput. Math. 10(2), 221–240 (2010) [Google Scholar]
  • 7.Eilat, M., Klartag, B.: Rigidity of Riemannian embeddings of discrete metric spaces. Invent. Math. 226(1), 349–391 (2021) [Google Scholar]
  • 8.Erbar, M., Maas, J.: Ricci curvature of finite Markov chains via convexity of the entropy. Arch. Ration. Mech. Anal. 206(3), 997–1038 (2012) [Google Scholar]
  • 9.Faigenbaum-Golovin, S., Levin, D.: Manifold reconstruction and denoising from scattered data in high dimension. J. Comput. Appl. Math. 421, Art. No. 114818 (2023)
  • 10.Federer, H.: Curvature measures. Trans. Amer. Math. Soc. 93, 418–491 (1959) [Google Scholar]
  • 11.Fefferman, C., Ivanov, S., Kurylev, Y., Lassas, M., Narayanan, H.: Reconstruction and interpolation of manifolds. I: the geometric Whitney problem. Found. Comput. Math. 20(5), 1035–1133 (2020)
  • 12.Fefferman, C., Ivanov, S., Lassas, M., Lu, J., Narayanan, H.: Reconstruction and interpolation of manifolds II: inverse problems for Riemannian manifolds with partial distance data (2021). arXiv:2111.14528
  • 13.Fefferman, C., Ivanov, S., Lassas, M., Narayanan, H.: Fitting a manifold of large reach to noisy data (2022). arXiv:1910.05084
  • 14.Genovese, C.R., Perone-Pacifico, M., Verdinelli, I., Wasserman, L.: Manifold estimation and singular deconvolution under Hausdorff loss. Ann. Stat. 40(2), 941–963 (2012) [Google Scholar]
  • 15.Gold, D., Rosenberg, S.: Discretized gradient flow for manifold learning. Int. J. Math. 35(11), Art. No. 2450040 (2024)
  • 16.Gray, A.: Tubes. 2nd edn. Progress in Mathematics, vol. 221. Birkhäuser, Basel (2004)
  • 17.Gu, X., Yau, S.-T.: Optimal transport for generative models. ICCM Not. 10(1), 1–27 (2022) [Google Scholar]
  • 18.van der Hoorn, P., Lippner, G., Trugenberger, C., Krioukov, D.: Ollivier curvature of random geometric graphs converges to Ricci curvature of their Riemannian manifolds. Discrete Comput. Geom. 70(3), 671–712 (2023) [Google Scholar]
  • 19.Jost, J.: Riemannian Geometry and Geometric Analysis, 7th edn. Universitext. Springer, Cham (2017)
  • 20.Last, G., Penrose, M.: Lectures on the Poisson Process. Institute of Mathematical Statistics Textbooks, vol. 7. Cambridge University Press, Cambridge (2018)
  • 21.Lee, J.M.: Introduction to Smooth Manifolds. Graduate Texts in Mathematics, vol. 218. 2nd edn. Springer, New York (2013)
  • 22.Lott, J.: Optimal transport and Ricci curvature for metric-measure spaces. In: Cheeger, J., Grove, K. (eds.) Surveys in Differential Geometry. Vol. XI. Surv. Differ. Geom., vol. 11, pp. 229–257. International Press, Somerville (2007)
  • 23.Lott, J., Villani, C.: Ricci curvature for metric-measure spaces via optimal transport. Ann. Math. 169(3), 903–991 (2009) [Google Scholar]
  • 24.Meyer, W.: Toponogov’s theorem and applications (2004). https://www2.math.upenn.edu/~wziller/math660/TopogonovTheorem-Myer.pdf
  • 25.Ollivier, Y.: Ricci curvature of Markov chains on metric spaces. J. Funct. Anal. 256(3), 810–864 (2009) [Google Scholar]
  • 26.Ollivier, Y.: A visual introduction to Riemannian curvatures and some discrete generalizations. In: Dafni, G., et al. (eds.) Analysis and Geometry of Metric Measure Spaces. CRM Proceedings and Lecture Notes, vol. 56, pp. 197–220. American Mathematical Society, Providence (2013) [Google Scholar]
  • 27.Puchkin, N., Spokoiny, V., Stepanov, E., Trevisan, D.: Reconstruction of manifold embeddings into Euclidean spaces via intrinsic distances. ESAIM Control Optim. Calc. Var. 30, 3 (2024)
  • 28.von Renesse, M.-K., Sturm, K.-T.: Transport inequalities, gradient estimates, entropy, and Ricci curvature. Commun. Pure Appl. Math. 58(7), 923–940 (2005) [Google Scholar]
  • 29.Rout, L., Korotin, A., Burnaev, E.: Generative modeling with optimal transport maps (2022). arXiv:2110.02999
  • 30.Sturm, K.-T.: On the geometry of metric measure spaces. II. Acta Math. 196(1), 133–177 (2006) [Google Scholar]
  • 31.Sturm, K.-T.: Remarks about synthetic upper Ricci bounds for metric measure spaces. Tohoku Math. J. 73(4), 539–564 (2021) [Google Scholar]
  • 32.Weyl, H.: On the volume of tubes. Amer. J. Math. 61(2), 461–472 (1939) [Google Scholar]

Articles from European Journal of Mathematics are provided here courtesy of Springer

RESOURCES