Manifolds with kinks and the asymptotic behavior of the graph Laplacian operator with Gaussian kernel


Abstract

We introduce manifolds with kinks, a class of manifolds with possibly singular boundary that notably contains manifolds with smooth boundary and corners. We derive the asymptotic behavior of the Graph Laplace operator with Gaussian kernel and its deterministic limit on these spaces as bandwidth goes to zero. We show that this asymptotic behavior is determined by the inward sector of the tangent space and, as special cases, we derive its behavior near interior and singular points. Lastly, we show the validity of our theoretical results using numerical simulation.

1 Introduction↩︎

The connection between the graph Laplacian and the Laplace–Beltrami operator is a central theme in geometric data analysis and has attracted substantial interest from both applied and theoretical communities. On the one hand, from a theoretical perspective, this relationship offers a discretization framework for studying differential operators on manifolds via point clouds or sampled data [1][4]. On the other hand, from an applied viewpoint, this connection underpins a wide array of algorithms in machine learning and signal processing, including manifold learning techniques such as Laplacian eigenmaps [5][7] or Diffusion Maps [8], [9]. Understanding the convergence of the graph Laplacian to the Laplace–Beltrami operator — notably the role of scaling, sampling density, and boundary behavior — remains crucial in ensuring that discrete approximations faithfully capture the geometry and analysis of the underlying continuous space. This paper aims at tackling this question on a class of smooth Riemannian manifolds having possibly singular boundary.
To describe our framework and main results, let us consider independent and identically distributed (i.i.d.) random variables \(\{X_i\}_{i \ge 1} \sim X\) with common law \(\mathbb{P}_X\) on a metric measure space \((M,d,\mu)\). For any \(t>0\) and \(n \in \mathbb{N}\backslash \{0\}\), define \[L_{n,t}f(x):=\frac{1}{nt^{d/2+1}}\sum_{i=1}^{n}e^{-\frac{d(x,X_j)^2}{t}}(f(x)-f(X_j))\] for any \(x \in M\) and \(f\) in the space \(\mathcal{C}(M)\) of continous functions on \(M\). Here \(d\) is a positive integer. We call \(L_{n,t}\) the intrinsic Graph Laplacian with \(d\)-dimensional Gaussian kernel associated with the family \(\{X_1,\ldots,X_n\}\) and the parameter \(t\). In case \((M,d,\mu)\) is a subset of some ambient Euclidean space \(\mathbb{R}^D\), then the intrinsic distance \(d\) might be replaced by the extrinsic Euclidean norm \(\|\cdot\|\), in which case we call \(L_{n,t}\) the extrinsic Graph Laplacian.

We define the (intrinsic/extrinsic) \(d\)-dimensional Gaussian operator at time \(t\) as the expected value of \(L_{n,t}\) with respect to \(\mathbb{P}_X\), that is \[L_tf(x):=\mathbb{E}_X\left[\frac{1}{t^{d/2+1}}e^{-\frac{d(x,X)^2}{t}}(f(x)-f(X))\right]\] for any \(x \in X\) and \(f \in \mathcal{C}(M)\). It follows from the Strong Law of Large Numbers that \[L_{n,t}f(x) \to L_tf(x) \qquad \text{almost surely as n \to \infty}\] for any \(x,f,t\) as above. Note that if \(\mathbb{P}_X\) is absolutely continuous with respect to \(\mu\) with Radon–Nikodym derivative \(p \in L^1(M,\mu)\), then the Gaussian operator rewrites as \[L_t f(x) = \frac{1}{t^{d/2+1}} \int_M e^{-\frac{d(x,y)^2}{t}} (f(x) - f(y)) p(y) \, d\mu(y).\]

After the work of various authors, e.g. [6], [10], it is by now known that when \(M\) is a \(d\)-dimensional smooth submanifold of some Euclidean space \(\mathbb{R}^D\), for any sufficiently regular \(f\) and \(p\), the extrinsic Gaussian operator satisfies \[\label{eq:int} L_tf(x)\to c_d \, \Delta_{M,p}f(x) \qquad \text{ as t\to 0}\tag{1}\] at any interior point \(x \in M\), where \(c_d\) is a dimensional constant and \(\Delta_{M,p}\) is the weighted Laplace operator given by \[\Delta_{M,p}f(x) :=p(x)\Delta_M{f}(x) + \frac{1}{2}\nabla{p}(x) \cdot \nabla{f}(x),\] with \(\Delta_M\) the classical Laplace–Beltrami operator of \(M\). In [11] (see also [12]), this pointwise result for the extrinsic Gaussian operator was extended to three types of singular subsets of \(\mathbb{R}^D\), namely submanifolds with boundaries, intersections of them, and submanifolds with edge-type singularities. For instance, if \(x\) belongs to the non-empty boundary of some smooth submanifold \(M \subset \mathbb{R}^D\), then \[\label{eq:bound} L_tf(x) = \frac{c'_d}{\sqrt{t}}p(x)\partial_\nu f(x)+ o\left(\frac{1}{\sqrt{t}}\right) \qquad \text{ as t \to 0,}\tag{2}\] where \(\partial_\nu f(x)\) is the inner normal derivative of \(f\) at \(x\) and \(c_d'\) is another dimensional constant.

Our first main result extends both 1 and 2 in the setting of smooth Riemannian manifolds with kinks. These spaces which we introduce in Section 3 might be understood as manifolds with possibly non-empty and irregular boundary. To distinguish these possibly non-smooth boundaries with classical smooth ones, we use the terminology border : a smooth manifold with kinks \(M\) thus divides into an interior part \(int M\) and a border part \(\partial M\). Manifolds with kinks include notably manifolds without boundary, manifolds with smooth boundary, and manifolds with corners as considered e.g. in [13][18]. Other relevant spaces like cones, pyramids, even cusps, fall into the framework of manifolds with kinks while they are not manifolds with corners.

If \(M\) is a \(d\)-dimensional manifold with kinks, then any \(x \in M\) admits a tangent space \(T_xM \simeq \mathbb{R}^d\) defined as the usual space of smooth derivations at \(x\). A key concept in our analysis is the notion of inward tangent cone \[I_xM \subset T_xM\] which consists in all the directions emanating from \(x\) and pointing towards the interior of \(M\). In particular, \(I_xM\) coincides with \(T_xM\) if \(x\) is an interior point, with a half-subspace of \(T_xM\) if \(x\) is a classical boundary point, and with a binary fraction of \(T_xM\) if \(x\) is a corner point, see Lemma 9.

We also introduce on manifolds with kinks the notions of \(C^k\) functions for \(k \in \mathbb{N}\cup\{\infty\}\), and tangent, cotangent and tensor bundles, see Section 3.3 and 3.4. We then define vector fields as the sections of the tangent bundle, differential forms as alternating sections of the cotangent bundle, Riemannian metrics as the sections of the symmetric, positive definite covariant 2-tensor bundle. Like for manifolds without/with boundary, we define the differential of order \(k\) of a \(\mathcal{C}^k\) function as the suitable symmetric covariant \(k\)-tensor field which acts on contravariant \(k\)-tensor fields. Note that we shall use \(Z^{(k)}\) to denote the contravariant \(k\)-tensor field \(Z \otimes \ldots \otimes Z\) where \(Z\) is a vector field. Any Riemannian metric \(g\) on a manifold with kinks \(M\) allows to specify the gradient and hessian of a \(\mathcal{C}^2\) function \(f\), through the classical relations \[df(Z) = g(\nabla^g f,Z), \qquad d^{(2)}f(Z \otimes Z') = g([\mathrm{Hess}^g f] Z,Z').\] A Riemannian metric additionally identifies those tangent vectors at a point \(x\) that have norm one, whose set we denote by \(S^g_xM\). We also define \[S^gI_xM :=I_x M \cap S^g_xM\] which we endow with the \((d-1)\)-dimensional Hausdorff measure \(\sigma\) associated with the distance induced on \(T_xM\) by the scalar product \(g(x)\).

For the purpose of our analysis, we need minimal regularity on the boundary of the manifolds with kinks that we consider. The weakest assumption we can work with is Lipschitz and Continuously Directionally Differentiable regularity, LCDD for short, see Section 3 for the precise definition. For any positive integer \(\ell\) we set \(c_\ell :=\Gamma((\ell+1)/2)/2\) where \(\Gamma\) is the classical Gamma function. Then our first main result writes as follows.

Theorem 1. Let \((M,g)\) be a smooth \(d\)-dimensional Riemannian manifold with kinks, and \(x \in M\) be either an interior point or an LCDD border point. Then the intrinsic Gaussian operator associated with a density \(p \in \mathcal{C}^2(M)\) with respect to the Riemannian volume measure \(\mathrm{vol}_g\) satisfies \[\begin{align} L_tf(x) & = -\frac{c_d}{\sqrt{t}} \left( p(x) \, \partial_{v_g(x)}f(x) + o(1)\right) - c_{d+1} \bigg( p(x) A_gf(x) + [p,f]_g(x) \bigg)\\ & \qquad \qquad \qquad \qquad \quad \qquad\qquad \qquad \qquad + O(\sqrt{t}) + O(t^{-d}e^{-t^{2\eta-1}}) \qquad \text{ as t \to 0,} \end{align}\] for any \(f \in \mathcal{C}^3(M) \cap L^1(M,p\,\mathrm{vol}_g)\) and \(\eta \in (0,1/2)\), where \[\partial_{v(x)} f(x) :=\mathop{}\!\mathrm{d}_{x}f\left( v_g(x) \right) \quad \text{with} \quad v_g(x) :=\int_{S^gI_xM} \theta \mathop{}\!\mathrm{d}\sigma(\theta),\] and \[A_gf(x) :=\int_{S^gI_xM} \mathop{}\!\mathrm{d}_{x}^{(2)}f \left( \theta^{(2)} \right) \mathop{}\!\mathrm{d}\sigma(\theta), \qquad [p,f]_g(x) :=\frac{1}{2} \int_{S^gI_xM} \mathop{}\!\mathrm{d}_xf(\theta)\mathop{}\!\mathrm{d}_xp(\theta) \mathop{}\!\mathrm{d}\sigma(\theta).\]

The recent preprint [19] provides a result that goes in the same direction as Theorem 1, but with less details on the expression of the second order differential \(A_p\) depending on the singular behaviour at \(x\) of the space under consideration. Our approach allows for a neat identification of the operator in several cases, including the following ones.

  • If \(x\) is an interior point, then \(I_xM\) coincides with \(T_xM\). Then the symmetry of \(T_xM\) with respect to the antipodal map \(z \mapsto -z\) implies that \(v(x)=0\), while \(A_p(x)\) reduces to the weighted Laplacian \(p(x)\Delta_g{f}(x) + \frac{1}{2}g_x(\nabla^g{p}(x), \nabla^g{f}(x))\) where \(\Delta_g\) is the Laplace–Beltrami operator associated with the metric \(g\). This matches up with 1 .

  • In case \(x\) is a boundary point, we obtain \[I_xM = \{ v \in T_xM : g(\partial_\nu f(x),v)\ge 0\}\] where \(\partial_{\nu}^g f(x)\) is the inner normal derivative of \(f\) at \(x\), hence \(\partial_{v_g(x)} f(x)\) coincides with \(\partial_{\nu}^g f(x)\) and we recover 2 .

  • Lastly, if \(x\) is an essential corner of depth \(k\), then we obtain \[\label{eq:Rdk} I_xM \simeq \bigcap_{1 \le i \le k} \{ v \in T_xM : g(\partial_{i} f(x),v)\ge 0\}\tag{3}\] where each \(\partial_{i} f(x)\) corresponds to differentiation along a half-line direction in the local model \[\mathbb{R}^n_k :=\mathbb{R}^k_+ \times \mathbb{R}^{d-k}.\] Then \(\mathop{}\!\mathrm{d}_x f (v_g(x)) = \sum_i \partial_{i} f (x)\).

Theorem 1 opens the door to a natural Neumann boundary condition on manifolds with kinks, that would be \(\partial_{v(x)} f(x)=0\) for any \(x \in \partial M\). A natural problem one could investigate after that is the spectral convergence of the Graph Laplacian generated from finitely many samples \(\{X_1,\ldots,X_n\}\) to the suitable Neumann Laplacian as sample size goes to infinity. Note that this question is studied in [20] with the so-called symmetrized AMV operator on manifolds with smooth boundary.
Our second main result tells us how to choose bandwidth \(t\) as a function of the sample size \(n\) so that the difference between the graph Laplace operator \(L_{n,t}\) and the operator \[\mathcal{L}_tf(x) :=-\frac{c_d}{\sqrt{t}}\left( p(x) \, \partial_{v(x)}f(x) + o(1)\right) - c_{d+1} \bigg( p(x) A_gf(x) + [p,f]_g(x) \bigg)\] identified in Theorem 1 converges to zero in probability or almost surely. We refer to Definition 29 for the notion of \(\alpha\)-subexponentiality, which encompasses boundedness, subexponentiality and subgaussianity for a real-valued random variable.

Theorem 2. Let \((M,g)\) be a smooth \(d\)-dimensional Riemannian manifold with kinks, \(x \in M\) be either an interior point or an LCDD border point, \(X\) a random variable on \(M\) with law having density \(p \in \mathcal{C}^2(M)\) with respect to the Riemannian volume measure \(\mathrm{vol}_g\). Consider a sequence \(\{t_n\} \subset (0,+\infty)\) such that \(t_n \to 0\) as \(n \to \infty\). For \(f \in \mathcal{C}^3(M)\cap L^1(M, p \, \mathrm{vol}_g)\), assume that \(f(X)\) is \(\alpha\)-subexponential for some \(\alpha \in (0,2]\).

  1. If \(\sqrt{n}\, t_n^{\frac{d}{2} + 1} \to \infty\) as \(n \to \infty\), then \[\left| L_{n, t_n} f(x) - \mathcal{L}_tf(x) \right| \xrightarrow{\mathbb{P}} 0.\]

  2. If \(\left(\sqrt{n}\, t_n^{\frac{d}{2} + 1} \right)^{\alpha}/\ln(n)\to \infty\) as \(n \to \infty\), then \[\left| L_{n, t_n} f(x) - \mathcal{L}_tf(x) \right| \xrightarrow{\text{a.s.}} 0.\]

Note the difference in the sufficient condition between (1) and (2) : the former doesn’t depend on the tail decay rate of \(f(X),\) whereas the latter does.

We obtain Theorem 2 in Section 5 through concentration estimates. Such estimates — studied in e.g. [10][12] — quantify the error in approximating the Gauss operator by the graph Laplace operator. Our results generalize this line of work in two key directions:

  1. We work in the setting of Riemannian manifolds with kinks, whose singularities can be significantly more severe than those of manifolds with boundary considered in [11].

  2. Unlike most previous works, we do not assume that the function \(f : M \to \mathbb{R}\) is bounded. In particular, our results hold on noncompact manifolds provided they have finite total volume.

Regarding point (2), previous concentration bounds rely on Hoeffding’s or Bernstein’s inequalities for bounded random variables. In contrast, we employ more general forms of these inequalities that are applicable to subgaussian or subexponential random variables with appropriate tail decay.

Let us conclude by mentioning that Theorem 2 guided us to choose the bandwidth parameter \(t>0\) correctly in the numerical experiments presented in Section 6.

Acknowledgments.↩︎

The authors are funded by the Research Foundation – Flanders (FWO) via the Odysseus II programme no. G0DBZ23N. They both thank Dominic Joyce for helpful answers concerning manifolds with corners, and Iosif Pinelis who suggested the proof of ?? .

2 Preliminaries↩︎

Throughout the paper, we write \(0_d\) for the origin of \(\mathbb{R}^d\), we let \(\mathbb{B}_r^d(x)\) stand for the Euclidean ball of radius \(r\) centered at \(x \in \mathbb{R}^d\), and we write \(\mathbb{B}_r^d\) and \(\mathbb{B}^d\) instead of \(\mathbb{B}_r^d(0_d)\) and \(\mathbb{B}_1^d(0_d)\) respectively. We let \((e_1,\ldots,e_d)\) denote the canonical basis of \(\mathbb{R}^d\). For \(y \in \mathbb{R}^d\) we will often denote by \(y'\) the \(d-1\) first coordinates of \(y\) in the canonical basis of \(\mathbb{R}^d\), and we let \(y_d\) be the last one. We consider the open upper half space \(\mathbb{H}^d:=\{y \in \mathbb{R}^d : y_d > 0\}\) and the Euclidean unit sphere \(\mathbb{S}^{d-1}:=\{y \in \mathbb{R}^d : y_1^2 + \ldots + y_d^2 = 1\}\). We let \(\sigma\) be the usual surface measure on \(\mathbb{S}^{d-1}\) which coincides with the \((d-1)\)-dimensional Hausdorff measure of the Euclidean distance. We also write \(\mathcal{L}^d\) for the \(d\)-dimensional Lebesgue measure in \(\mathbb{R}^d\).

For a subset \(A\) of the Euclidean space \(\mathbb{R}^d\), we let \(\bar{A}\) be the closure of \(A\) and \(int(A)\) be its interior. We recall that a rigid motion of \(\mathbb{R}^d\) is a linear map \(R : y \mapsto Ay+b\) where \(A\in O(d)\) and \(b\in \mathbb{R}^d\). The epigraph and strict epigraph of \(\gamma : \mathbb{R}^d \to \mathbb{R}\) are respectively defined as \[epi(\gamma) :=\{ y \in \mathbb{R}^d : y_d \ge \gamma(y')\} \quad \text{and} \quad \mathring{epi}(\gamma) :=\{ y \in \mathbb{R}^d : y_d > \gamma(y')\}.\] When \(\gamma\) is continuous, we obviously have \[\label{eq:epi61} epi(\gamma) = \overline{\mathring{epi}(\gamma)}.\tag{4}\] The directional derivative of \(\gamma\) at \(x \in \mathbb{R}^d\) in the direction \(v \in \mathbb{R}^d\) is defined as \[\gamma'(x;v) :=\lim_{t \downarrow 0} \frac{\gamma(x+tv)-\gamma(x)}{t}\] whenever the limit exists.

2.1 Boundary points↩︎

For an open set \(\Omega \subset \mathbb{R}^d\) with a non-empty boundary \(\partial \Omega\), we will call boundary points of \(\Omega\) the elements in \(\partial \Omega\). Inspired by [21], we define the regularity of boundary points as follows.

Definition 1. Let \(\Omega\) be an open subset of \(\mathbb{R}^d\) with a non-empty boundary \(\partial \Omega\), and \(x \in \partial \Omega\). Consider \(k \in \mathbb{N}\cup \{\infty\}\) and \(\alpha \in (0,1]\).

  1. We say that \(x\) is a \(\mathcal{C}^k\) boundary point of \(\Omega\) if there exist \(\delta>0\) and \(\gamma \in \mathcal{C}^k(\mathbb{R}^{d-1})\) such that, up to applying a rigid motion, \[\label{eq:boundary95points} \begin{cases} \bar{\Omega} \cap \mathbb{B}_\delta^d(x) = epi(\gamma) \cap \mathbb{B}_\delta^d(x), \nonumber\\ \Omega \cap \mathbb{B}_\delta^d(x) = \mathring{epi}(\gamma) \cap \mathbb{B}_\delta^d(x). \end{cases}\tag{5}\]

  2. We say that \(x\) is a \(\mathcal{C}^{k,\alpha}\) boundary point of \(\Omega\) if it satisfies 5 for some \(\gamma \in \mathcal{C}^{k,\alpha}(\mathbb{R}^{d-1})\), up to a rigid motion.

  3. We say that \(x\) is a directionally differentiable (d.d. for brevity) boundary point if it is a \(\mathcal{C}^0\) boundary point with a function \(\gamma\) as above whose directional derivative \(\gamma'(x';v')\) exists for any \(v' \in \mathbb{R}^{d-1}\), where we identify \(x\) with its image \((x',x_d)\) through the rigid motion from (1).

  4. We say that \(x\) is a continuously directionally differentiable (CDD for brevity) boundary point if it is a d.d. point with a function \(\gamma\) whose directional derivative \(\gamma'(x';\cdot)\) is continuous on \(\mathbb{R}^{d-1}\).

As customary, we may write smooth instead of \(\mathcal{C}^\infty\), and Lipschitz instead of \(\mathcal{C}^{0,1}\). We will essentially work with boundary points that are \(\mathcal{C}^0\), \(\mathcal{C}^1\), Lipschitz, or CDD. Lipschit and CDD shall be abreviated to LCDD.

Remark 3. It should be noted that a boundary point may not be \(\mathcal{C}^0\). For instance, consider \(\Omega:=\{(x,y) \in (0,+\infty)\times \mathbb{R}: y < \sin(1/x)\}.\) Then the boundary \(\partial \Omega\) is the union of \(\{0\} \times (-\infty,1]\) and \(\{y = \sin(1/x)\}\), which cannot be written as the graph of a continuous function locally around \(0_2\), see Figure 1.

Figure 1: 0_2 is not a \mathcal{C}^0 boundary point of \Omega.

Remark 4. The \(\mathcal{C}^0\) regularity of a boundary point is trivially preserved by local homeomorphisms. More precisely, for any \(i \in \{1,2\}\), let \(x_i\) be a boundary point of an open set \(\Omega_i \subset \mathbb{R}^d\) with a non-empty boundary. Assume that there exist open neighborhoods \(V_1, V_2\) in \(\mathbb{R}^d\) of \(x_1,x_2\) respectively, and a homeomorphism \(\Phi : V_1 \cap \bar{\Omega}_1 \xrightarrow{\sim} V_2 \cap \bar{\Omega}_2\) such that \(x_2 = \Phi(x_1)\). If \(x_1\) is \(\mathcal{C}^0\), then the composition of the continuous map \(\gamma_1\) such that 5 holds around \(x_1\) up to rigid motion with the homeomorphism \(\Phi\) yields a continuous map \(\gamma_2 :=\Phi \circ \gamma_1\) which ensures that \(x_2\) is \(\mathcal{C}^0\).

The next proposition shows that, from a topological point of view, there is no difference between a \(\mathcal{C}^0\) boundary point and a boundary point of the upper half space \(\mathbb{H}^d\).

Proposition 5.

Let \(\Omega\) be an open subset of \(\mathbb{R}^d\) admitting a \(\mathcal{C}^0\) boundary point \(x \in \partial \Omega\). Then there exists \(\delta>0\) such that \(\bar{\Omega} \cap \mathbb{B}^d_\delta(x)\) is homeomorphic to \(\bar{\mathbb{H}}^d\cap \mathbb{B}^d_\delta.\)

Proof. Let \(\delta\) and \(\gamma\) be such that 5 holds up to a rigid motion. Then \(\Phi(y) :=x + (y',\gamma(y')+y_{d})\) defines a homeomorphism \(\bar{\mathbb{H}}^d\cap \mathbb{B}^d_\delta \to epi(\gamma) \cap \mathbb{B}^d_\delta(x)\) with inverse \(\Phi^{-1}(\xi) :=(\xi', \xi_d-\gamma(\xi'))-x\). ◻

2.2 Diffeomorphisms and inward sectors↩︎

For \(k\) a positive integer, we shall use the following notion of \(\mathcal{C}^k\) diffeomorphism between possibly non-open subsets of \(\mathbb{R}^d\), see e.g. [17] for details.

Definition 2. We say that a map \(\varphi:A\to B\) between subsets \(A,B\subset \mathbb{R}^d\) is a \(\mathcal{C}^k\) diffeomorphism if there exist open subsets \(\tilde{A},\tilde{B} \subset \mathbb{R}^d\) and a \(\mathcal{C}^k\) diffeomorphism \(\tilde{\varphi}:\tilde{A}\to \tilde{B}\) such that such that \(A \subset \tilde{A}\), \(B \subset \tilde{B}\) and \(\tilde{\varphi}|_A=\varphi.\)

We will mostly use this definition in the case where \(A\) and \(B\) are the closures of open subsets of \(\mathbb{R}^d\) with a non-empty boundary. In this regard, let us introduce a useful definition.

Definition 3. Let \(\Omega\) be an open subset of \(\mathbb{R}^d\) with a non-empty boundary \(\partial \Omega\). For any \(x \in \bar{\Omega}\), we define the inward sector of \(\Omega\) at \(x\) as \[I_x\Omega :=\{c'(0) \in \mathbb{R}^d : c \in \mathcal{C}^1(I,\mathbb{R}^d) \text{ s.t.~I = (-\epsilon,\epsilon) for some \epsilon>0, c(0)=a and c([0,\epsilon)) \subset A} \}.\]

It is easily seen that the inward tangent sector of a point belonging to an open set is \(\mathbb{R}^d\), and that at a \(\mathcal{C}^1\) boundary point it is a half space. Moreover, the inward tangent sector at the origin of \(\mathbb{R}^d_k\) is \(\mathbb{R}^d_k\) itself.

Let us provide relevant properties of diffeomorphisms between closures of open sets.

Lemma 1. Let \(U, V\subset \mathbb{R}^d\) be open subsets and \(\varphi:\bar{U}\to \bar{V}\) a diffeomorphism. Then \(\varphi(U)=V\) and \(\varphi(\partial U)=\partial V.\) Moreover, for any \(x \in \partial U\), the differential \(d_x\varphi : \mathbb{R}^d \to \mathbb{R}^d\) is a well-defined linear map which maps \(I_x U\) to \(I_{\varphi(x)} V.\)

Proof. The first two equalities follow from the fact that a homeomorphism of closures of two open subsets maps interior to interior and boundary two boundary, and that a diffeomorphism between the two is a homeomorphism. For the well-posedness of \(d_x\varphi\), note that any extension \(\tilde{\varphi}\) of \(\varphi\) is \(\mathcal{C}^1\), so that the map \(d\tilde{\varphi} : y \mapsto d_y\tilde{\varphi}\) is continuous on \(\tilde{A}\) and coincides with \(d\varphi\) on \(A\); as a consequence, \(d\tilde{\varphi}\) continuously extends to \(\partial U\) in a unique way. The last point follows from the chain rule applied to any map of the form \(\tilde{\varphi} \circ c\) where \(c\) defines an element \(c'(0)\) of \(I_x U\) and \(\tilde{\varphi}\) is an extension of \(\varphi\). ◻

Remark 6. The first two equalities in the previous lemma do not hold when \(\varphi\) is a diffeomorphism between \(U\) and \(V\) only. For instance, the open sets \(\mathbb{B}^2\) and \(\mathbb{B}^2 \backslash ([0,1)\times \{0\})\) are biholomophic via the Riemann mapping theorem but their boundaries are not homeomorphic. It should be pointed out that, in this case, the diffeomorphism given by the Riemann map does not extend to the boundary.

Before closing this subsection, let us explain how the inward tangent sector provides us with a convenient way to define cusps.

Definition 4. Let \(\Omega\subset \mathbb{R}^d\) be open with a non-empty boundary, and \(x\in \partial{\Omega}\) be a \(\mathcal{C}^0\) boundary point. We say that \(x\) is a cusp if \[int \, I_x \Omega = \emptyset.\]

Examples 7. According to the previous definition, the origin \(0_3\) is a cusp for both \[A :=\{(x,y,z) \in \mathbb{R}^3 : z \ge \sqrt{|x|}\} \qquad \text{and} \qquad B :=\{(x,y,z) \in \mathbb{R}^3 : z \ge (x^2 + y^2)^{1/4}\}.\] Indeed, \(I_{0_3}A = \{(0,0,z) : x \in \mathbb{R}, z \ge 0\}\) and \(I_{0_3}B = \{(x,0,z) : z \ge 0\}\) both have empty interior.

a

b

Figure 2: Sets \(A\) and \(B\) both have \(0_3\) as a cusp.

2.3 Boundary cones of open Euclidean subsets↩︎

In this section, we provide classical notions from convex analysis, see e.g. [22] for more details. We recall that \(C\subset \mathbb{R}^d\) is called a cone if it is invariant under multiplication by a positive number, i.e. if \(\lambda v \in C\) for any \(v\in C\) and \(\lambda > 0.\)

Definition 5. Consider \(A \subset \mathbb{R}^d\) and \(a \in A\).

  1. The Bouligand tangent cone of \(A\) at \(a\) is defined as \[T^B_aA:=\{0_d\}\cup \left\{v\in \mathbb{R}^d\backslash \{0_d\}: \text{there exists \{a_n\}\subset A such thata_n \to a and } \frac{a_n-a}{\left\lVert a_n-a\right\rVert} \stackrel{}{\to} \frac{v}{\left\lVert v\right\rVert} \right\}.\]

  2. The Bouligand tangent sphere of \(A\) at \(a\) is defined as \(S^B_aA:=T^B_aA\cap \mathbb{S}^{d-1}.\)

  3. The feasible direction cone of \(A\) at \(a\) is defined as \[F_a(A):=\{v\in \mathbb{R}^d: \text{there exists t_v>0 such that a+tv \in A for any 0 \le t < t_v}\}.\]

  4. The open feasible direction cone of \(A\) at \(a\) is defined as \[\tilde{F}_a(A):= \{v\in \mathbb{R}^d: \text{there exists t_v>0 such that a+tv \in int(A) for any 0 < t < t_v}\}.\]

The Bouligand tangent cone is also known as tangent cone or contingent cone in the literature. It is a closed set. One can easily check that our definition is equivalent to the one given in [22]. Note also that \[\tilde{F}_a(A) \subset F_a(A) \subset T^B_aA\] with possibly strict inclusion : for instance, \(\tilde{F}_{0_d}(\bar{\mathbb{H}}^d) = \mathbb{H}^d \varsubsetneq \bar{\mathbb{H}}^d = F_{0_d}(\bar{\mathbb{H}}^d)\). Moreover, in our setting, the following holds.

Lemma 2. Let \(x\) be a \(\mathcal{C}^0\) boundary point of an open set \(\Omega \subset \mathbb{R}^d\) with a non-empty boundary. Then \(\tilde{F}_x({\bar{\Omega}})\ne \emptyset.\)

Proof. Let \(\delta\) and \(\gamma\) be such that 5 holds up to a rigid motion. Since \(x \in \partial \Omega\), we have \(x_d = \gamma(x')\) so that \(x_d + t > \gamma(x')\) for any \(t \in (0,\delta)\). Thus \(x+te_d \in \Omega \cap \mathbb{B}_\delta(x) \subset \Omega\) for any such a \(t\), so that \(e_d\in \tilde{F}_x(\bar{\Omega})\). ◻

For suitably regular boundary points, the Bouligand tangent cone satisfies the following additional properties.

Proposition 8. Let \(x=(x',x_d)\) be a LCDD boundary point of an open set \(\Omega \subset \mathbb{R}^d\) with a non-empty boundary, and let \(\gamma\) be such that 5 holds up to a rigid motion. Then \[\label{eq:epi} T^B_x{\bar{\Omega}}=epi(\gamma'(x';\cdot)).\qquad{(1)}\] Moreover, \[\label{eq:epi2} T^B_x{\bar{\Omega}} = \overline{\tilde{F}_x(\bar{\Omega})}, \qquad int(T^B_x{\bar{\Omega}})=int({ \tilde{F}_x(\bar{\Omega}) }), \qquad \partial(T^B_x{\bar{\Omega}})=\partial({ \tilde{F}_x(\bar{\Omega}) }).\qquad{(2)}\] Lastly, \[\label{eq:epi3} \mathcal{L}^n(\partial(T^B_x{\bar{\Omega}})) = 0.\qquad{(3)}\]

Proof. With no loss of generality, we assume that \(x =0_d\). Set \(g(v'):=\gamma'(0_{d-1},v')\) for any \(v' \in \mathbb{R}^{d-1}\).

Consider \(v \in epi(g)\). Then \(g(v')<v_d\). For any \(\epsilon_n \downarrow 0\), \[\begin{align} \gamma(\epsilon_n v') = g(\epsilon_n v') + o(\epsilon_n \|v'\|) = \epsilon_n g(v') + o(\epsilon_n \|v'\|) & < \epsilon_n v_d + o(\epsilon_n \|v'\|)\\ & =\epsilon_n(v_d + o(\|v'\|)) ) =: \tau_n, \end{align}\] so that each \(x_n :=(\epsilon_n v', \tau_n)\) belongs to \(epi(g)\) and the sequence \((x_n)\) satisfies \(x_n \to 0_d\) with \[\frac{x_n}{\|x_n\|} = \frac{(v',v_d+o(\|v'\|))}{\|(v',v_d+o(\|v'\|))\|} \to \frac{v}{\|v\|} \, \cdot\] Thus \(v \in T_{0_d}^B\bar{\Omega}\). This shows the reverse inclusion in ?? .

Consider \(v\in T^B_{0_d}{\bar{\Omega}}\) and \(\{x_n\} \subset \bar{\Omega}\) such that \(x_n \to 0_d\) and \(x_n/\|x_n\| \to v/\|v\|\). By 5 , we can assume that \(\{x_n\} \subset epi(\gamma)\). Set \(\epsilon_n :=\|x_n\|/\|v\|\) for any \(n\). Then \[\gamma(\epsilon_n v') = g(\epsilon_n v') + o(\epsilon_n)\] by definition of \(g\) and \[|\gamma(x_n') - \gamma(\epsilon_n v')| \le L \|x_n' - \epsilon_n v'\| = L \|x_n\| \left|\frac{x_n'}{\|x_n\|} - \frac{v'}{\|v\|}\right| = o(\epsilon_n)\] for some \(L \ge 1\), because \(\gamma\) is Lipschitz. As a consequence, \[\delta_n :=|g(\epsilon_n v') - \gamma(x_n')| = o(\epsilon_n).\] Thus \(\epsilon_n g(v') = g(\epsilon_n v') \le \gamma(x_n') + \delta_n \le (x_n)_d+ \delta_n\) because \(x_n \in epi(\gamma)\). Then \[g(v') \le \frac{(x_n)_d+ \delta_n}{\epsilon_n} = \frac{\|v\|(x_n)_d}{\|x_n\|} + o(\|v\|) \to v_d.\] This shows the direct inclusion in ?? .

From now on, for clarity of the exposition, we do not assume that \(x=0_d\). Let us prove now the first equality in ?? . Using successively ?? and 4 (the latter being available because \(x\) is c.d.d.), we can write: \[\begin{align} T^B_x(\bar{\Omega}) &= epi(g) =\overline{\mathring{epi}(g)}. \end{align}\] But \[\begin{align} \mathring{epi}(g) & = \left\{v \in \mathbb{R}^d : v_d > \lim_{t \to 0^+} \frac{\gamma(x' + t v') - \gamma(x')}{t} \right\}\\ & \subset \left\{ v \in \mathbb{R}^d : \exists \, t_v> 0\, \text{ s.t. } v_d > \frac{\gamma(x' + t v') - \gamma(x')}{t} \, \forall \, t \in (0,t_v) \right\}\\ &= \left\{v \in \mathbb{R}^d : \exists \, t_v> 0\, \text{ s.t. } x_d + t v_d > \gamma(x' + t v') \;\forall\, 0 < t < t_v \right\} \quad (\text{using that } x_d=\gamma(x')) \\ &= \tilde{F}_x(\bar{\Omega}). \end{align}\]

This shows \(T^B_x(\bar{\Omega}) \subset \overline{\tilde{F}_x(\bar{\Omega})}.\) The other inclusion is obvious since the Bouligand tangent cone is closed and contains \(\tilde{F}_x(\bar{\Omega})\).

The third equality in ?? is an obvious consequence of the first and second ones. So we are left with proving the second one. Since \(\tilde{F}_x(\bar{\Omega}) \subset T^B_x{\bar{\Omega}}\) we clearly have \(int(\tilde{F}_x(\bar{\Omega})) \subset int(T^B_x{\bar{\Omega}}),\) hence it is enough to show the converse inclusion \(int(T^B_x{\bar{\Omega}}) \subset int(\tilde{F}_x(\bar{\Omega}))\). By ?? and the continuity of \(g\), this amounts to showing that \(\mathring{epi}(g) \subset int(\tilde{F}_x(\bar{\Omega}))\). Consider \(v \in \mathring{epi}(g)\). Then there exists \(\eta>0\) such that \(g(w')<w_d\) for any \(w \in \mathbb{B}^d_\eta(v)\). By definition of \(g\), for any such a \(w\), there exists \(t_w>0\) such that for any \(t \in (0,t_w)\), \[\frac{\gamma(x'+tw')-\gamma(x')}{t} < w_d.\] Since \(\gamma(x') = x_d\) the previous rewrites as \[\gamma(x'+tw') < x_d + tw_d.\] Thus \(w \in \tilde{F}_x(\bar{\Omega})\). This means that \(\mathbb{B}^{d}_\eta(v) \subset \tilde{F}_x(\bar{\Omega})\), hence \(v \in int(\tilde{F}_x(\bar{\Omega}))\) as claimed.

To conclude, let us explain why ?? holds. By ?? , the set \(\partial(T^B_x{\bar{\Omega}})\) coincides with the graph of \(g\). Since the latter is continuous, its graph has \(n\)-dimensional Lebesgue measure equal to \(0\). ◻

Remark 10. It should be noted that feasible direction cones are not preserved by diffeomorphisms. Indeed, consider the closed subsets \(A :=\{1 \le x \le \sqrt{y}\}\) and \(B :=\{1\le x/2+1/2\le y\}\) of \((0,+\infty)^2\), and \(\Phi : A \to B\) mapping \((x,y)\) to \((x,y^2)\). Then \(\Phi\) is a diffeomorphism in the sense of Definition 2 since it trivially extends to a diffeomorphism of \((0,+\infty)\) onto itself. Now \(F_{(1,1)}A = B \backslash \{x = y\}\) is not closed while \(F_{(1,1)}B = B\) is, so the linear isomorphism \(\mathop{}\!\mathrm{d}_{(1,1)}\Phi\) cannot map \(F_{(1,1)}A\) to \(F_{(1,1)}B\).

a

b

Figure 3: Sets \(A\) and \(B\) in blue with \(\partial F_{(1,1)}A = \partial F_{(1,1)} B\) in red..

The following proposition relates the inward tangent sector with the Bouligand tangent cone at a \(\mathcal{C}^0\) boundary point of an open subset of \(\mathbb{R}^d\) with a non-empty boundary.

Proposition 9. Let \(\Omega\subset \mathbb{R}^d\) be open with a non-empty boundary, and \(x\in \partial{\Omega}\) be a \(\mathcal{C}^0\) boundary point. Then \[I_x\bar{\Omega}\subset T^B_x(\bar{\Omega}).\] Moreover, if \(x\) is LCDD, then \[I_x\bar{\Omega}=T^B_x(\bar{\Omega}).\]

Proof. Let us first show \(\subset.\) Take \(\gamma'(0) \in I_x \overline{\Omega}\) where \(\gamma:I\to \overline{\Omega}\) is smooth and such that \(I = (-\epsilon, \epsilon)\) for some \(\epsilon>0\) and \(\gamma(0)=x.\) Set \(x_n:=\gamma(1/n)\) for any \(n \in \mathbb{N}\) large enough. By Taylor expansion, \(\gamma(t)=\gamma(0) + t\gamma'(0)+ o(t)\) as \(t\to 0+\), hence \(\frac{x_n-x}{\left\lVert x_n-x\right\rVert}\to \frac{\gamma'(0)}{\left\lVert\gamma'(0)\right\rVert}.\) This yields \(\gamma'(0) \in T^B_x\overline{\Omega}\) as desired.

Let us prove the converse inclusion in the LCDD case. By ?? , if \(v \in int(T^B_s(\overline{\Omega})) = int(\tilde{F}_s(\Omega))\), then there exists \(t_v>0\) such that \(x+tv \in \bar{\Omega}\) for any \(t \in (0,t_v)\), so \(v = \gamma'(0) \in I_x \overline{\Omega}\) with \(\gamma(t) :=x+tv\) for any \(t \in [0,t_v)\). Let us now establish that \(\partial(T^B_x(\overline{\Omega})) \subset I_x \overline{\Omega}\). For this, observe that ?? implies that \(\partial(T^B_x(\overline{\Omega}))= \partial {\{(v',v_d)\in \mathbb{R}^d: v_d \ge \gamma'(x',v')\}}= {\{(v',v_d)\in \mathbb{R}^d: v_d = \gamma'(x',v')\}}\) since \(v'\mapsto \gamma'(x',v')\) is continuous. The latter set is equal to \(\{c'(0): c(t)=(c_1(t)\dots c_{d-1}(t), c_d(t)): c_d(t)=\gamma(c_1(t)\dots c_{d-1}(t))\} =\{c'(0): c:[0,\epsilon)\to \partial\Omega \text{ is } \mathcal{C}^1, c(0)=x\}.\) This finishes the proof. ◻

2.4 Non-fluctuating boundary and local blow-ups↩︎

Consider an open subset \(\Omega\subset \mathbb{R}^d\) with a non-empty boundary, and a boundary point \(x\in \partial \Omega.\) For any \(t>0\), set \[\Omega_{x,t}:=\frac{\Omega-x}{{t}} \, \cdot\] We are interested in finding the limit in the sense of convergence of characteristic functions of the sets \(\Omega_{x,t}\) as \(t\to 0\). When this limit exists, we call it the blow up of \(\Omega\) at \(x.\) With this in mind, we introduce the following.

Definition 6. Consider an open subset \(\Omega\subset \mathbb{R}^d\) with a non-empty boundary, and \(x\in \partial \Omega.\)

  1. We say that \(v \in T_x^B\bar{\Omega} \backslash \tilde{F}_x\Omega\) is a non-fluctuating direction at \(x\) if the following holds : for any \(t_n\to 0\) such that \(x+t_nv\notin \Omega\) for any \(n\) there exists \(t_v>0\) such that \(x+tv\notin \Omega\) for any \(t \in (0,t_v)\). Otherwise we say that \(v\) is a fluctuating direction at \(x\).

  2. We say that \(\Omega\) has (a.e.) non-fluctuating boundary at \(x\) if any (\(\sigma\)-a.e.) \(v \in S_x^B\bar{\Omega}\) is a non-fluctuating direction. Otherwise we say that \(\Omega\) has a fluctuating boundary at \(x\).

Roughly speaking, a domain with non-fluctuating boundary at a boundary point \(x\) is so that whenever there exist points on a line spanned by \(v\) arbitrary close to \(x\) that stay outside of \(\Omega,\) then there must be a segment of that line that also remains outside \(\Omega.\) On the co if \(\Omega\) has a fluctuating boundary in the direction \(v,\) then there is a line \(t\mapsto x+tv\) which intersects \(\partial \Omega\) infinitely many times without a segment \(\{x+tv: 0\le t \le t_0\}\) being properly contained in \(\partial \Omega.\) This explains the terms fluctuating and non-fluctuating.

Examples 11. For any \(k \ge 0\), consider \(\Omega_k:=\{(x,y)\in \mathbb{R}^2: y > x^k \sin (1/x), 0<x<1\}.\) For \(k=0,1\), the set \(\Omega_k\) has a fluctuating boundary at the boundary point \(0_2,\) however for \(k\ge 2\), the sole fluctuating direction of \(\Omega_k\) at \(0_2\) is \((1,0)\), and thus the set has a.e. non-fluctuating boundary, see Figure 2.

We shall use the following lemma.

Lemma 3. Let \(\Omega\subset \mathbb{R}^d\) be open with a non-empty boundary, and \(x \in \partial \Omega\). For any non-fluctuating direction \(z\) at \(x\), \[1_{\Omega_{x,t}}(z) \to 1_{\tilde{F}_x(\Omega)}(z) \qquad \text{as } t \to 0^+.\]

Proof. If \(1_{\tilde{F}_x(\Omega)}(z) = 1\), then there exists \(t_0 > 0\) such that for any \(t \in (0,t_0)\) we have \(x + tz \in \Omega\). The latter is obviously equivalent to \(z \in \Omega_{x,t}\), hence we get \(1_{\Omega_{x,t}}(z) = 1\) for any \(t \in (0,t_0)\), thus \(1_{\Omega_{x,t}}(z) \to 1\) as \(t \to 0\). Now if \(1_{\tilde{F}_x(\Omega)}(z)=0\), then \(z \notin \tilde{F}_x(\Omega)\) which implies that for any positive integer \(n\) there exists \(t_n \in (0, \epsilon)\) such that \(x+t_nz \notin \Omega\) for any \(n\). By definition of a non-fluctuating direction, we obtain that \(x+sz \notin \Omega\) for any \(s \in (0,t_z)\) for some \(t_z>0\). Then \(z \notin \Omega_{x,s}\) for any such a \(s\), hence \(1_{\Omega_{x,s}}(z) \to 1\) as \(s \to 0\). ◻

Now we focus on particular boundary points.

Lemma 4. Let \(\Omega\subset \mathbb{R}^d\) be open with a non-empty boundary, and \(x \in \partial \Omega\) be LCDD. Then the fluctuating directions at \(x\) form a \(\mathcal{L}^d\)-negligible set.

Proof. It follows from the definition of fluctuating direction that the set of such directions at \(x\) belongs to \(T^B_x(\bar{\Omega})\setminus \tilde{F}_x(\Omega)\). But \[T^B_x(\bar{\Omega})\setminus \tilde{F}_x(\Omega) \subset T^B_x(\bar{\Omega})\setminus int(\tilde{F}_x(\Omega)) = T^B_x(\bar{\Omega})\setminus int(T^B_x(\bar{\Omega})) = \partial T^B_x(\bar{\Omega})\] where the first equality follows from the second equality in ?? . Then ?? gives the desired result. ◻

We are now in a position to state and prove the following important property.

Proposition 12. Let \(\Omega\subset \mathbb{R}^d\) be open with a non-empty boundary, and \(x \in \partial \Omega\) be LCDD. Then: \[1_{\Omega_{x,t}}(z) \to 1_{\tilde{F}_x(\Omega)}(z) = 1_{T^B_x(\bar{\Omega})}(z) \quad \text{for a.e. }z \in \mathbb{R}^d, \quad t \to 0^+ .\]

Proof. From Lemma 3 and Lemma 4 we get \(1_{\Omega_{x,t}}(z) \to 1_{\tilde{F}_x(\Omega)}(z)\) as \(t \to 0^+\) for a.e. \(z \in \mathbb{R}^d\). Then ?? yields that \(1_{\tilde{F}_x(\Omega)}(z) = 1_{T^B_x(\bar{\Omega})}(z)\) for a.e. \(z \in \mathbb{R}^d\). ◻

Remark 13. There exist domains with a fluctuating boundary which do not blow up at a boundary point to their open feasible direction cone. Consider \(\Omega:=\{(x_1,x_2): x_2 \ne x_1 \sin(1/x_1), -1< x_1 <1\}.\) Then \(\tilde{F}_{(0,0)}(\Omega)=\{(x_1,x_2): x_2 > |x_1| \text{ or } x_2 < - |x_1|, -1< x_1 <1\},\) but \(1_{\Omega_{(0,0),t}}(z) \to 1\) a.e. because the graph of \(x \mapsto x \sin(1/x)\) has two dimensional Lebesgue measure zero.

Let us conclude with providing a large class of domains with non-fluctuating boundary at a boundary point. We recall that a set \(A \subset \mathbb{R}^d\) is convex if the segment \([x,y]:=\{(1-t)x+ty : t \in [0,1]\}\) belongs to \(A\) for any \(x,y \in A\).

Definition 7. We say that \(S \subset \mathbb{R}^d\) is locally convex at \(x \in S\) if there exists \(\delta>0\) such that \(S\cap \mathbb{B}_\delta(x)\) is convex.

Then the following holds.

Proposition 14. Let \(\Omega\) be an open subset of \(\mathbb{R}^d\) with a non-empty boundary, and \(x \in \partial \Omega\). If either \(\bar{\Omega}\) or \(\mathbb{R}^d \setminus \bar{\Omega}\) is locally convex at \(x\), then \(\Omega\) has a non-fluctuating boundary at \(x\).

Proof. Fix \(v\in \mathbb{R}^d.\) Assume that there exists \(t_n\to 0\) such that \(x+t_nv\notin \Omega\) for any \(n\). We need to show that there exists \(t_v>0\) such that \(x+tv\notin \Omega\) for all \(t \in (0,t_v)\). Since \(x+t_nv \in \mathbb{R}^d\setminus \Omega\), if \(\mathbb{R}^d\setminus \Omega\) is locally convex at \(x\), then the claim obviously follows with \(t_v\) equal to \(t_n\). Assume now that \(\Omega\) is locally convex at \(x.\) If the claim were not true, then there would exist \(s_n\to 0\) such that \(t_{n+1}<s_n<t_n\) and \(x+s_nv\in \Omega\) for any \(n\). But this is not possible because once \(x+s_Nv \in \Omega\) for some \(N,\) all the points \(x+sv\) with \(s \in (0,s_n)\) must belong to \(\Omega\) by convexity, in particular \(x+t_{n+1} v\) must belong to \(\Omega\), which is in contradiction with our initial assumption. ◻

3 Manifolds with kinks↩︎

In this section, we introduce the notion of manifolds with kinks.

3.1 Topological manifolds with kinks↩︎

Recall that a topological space is called paracompact if any open cover has a locally finite open refinement, and that a Hausdorff topological space is paracompact if and and only if it admits partitions of unity subordinate to any open cover. For two topological spaces \(X,Y\) and a subset \(A \subset X\), a map \(\varphi : A \to Y\) is a topological embedding if it is an homeomorphism onto its image.

Definition 8. Let \(M\) be a paracompact Hausdorff topological space.

  1. A \(d\)-dimensional chart around \(x \in M\) is a pair \((U,\phi)\) where \(U\) is an open neighborhood of \(x\) in \(M\) and \(\phi : U \to \mathbb{R}^d\) is a topological embedding. We say that \((U,\phi)\) is centered at \(x\) if \(\phi(x)=0_d.\)

  2. A \(d\)-dimensional interior chart around \(x\) is a \(d\)-dimensional chart \((U,\phi)\) such that \(\phi(U)\) is an open subset of \(\mathbb{R}^d.\) Any \(x\) admitting such a chart is called an interior point of \(M.\) The set of interior points of \(M\) is denoted by \(Int(M).\)

  3. A \(d\)-dimensional border chart around \(x\) is a pair \((U,\phi)\), where \(U\) is an open neighborhood of \(x\) in \(M\) and \(\phi : \bar{U} \to \mathbb{R}^d\) is a topological embedding such that \(int(\phi(U))\neq \emptyset\) and \(\phi(x)\) is a \(\mathcal{C}^0\) boundary point of \(int(\phi(U))\). Any \(x\) admitting a border chart is called a border point. The set of border points of \(M\) is denoted by \(\partial M.\)

Remark 15. We use the word border as a synonym for possibly singular boundary of manifolds.

Remark 16. The definition of a border point \(x\) is independent of the choice of a border chart around \(x\). Indeed, let \((U,\phi)\) and \((V,\psi)\) be two such charts. Since \(\phi, \psi\) are defined on \(\bar{U}, \bar{V}\) respectively, the transition map \(\psi \circ \phi^{-1}\) is a homeomorphism from \(\phi(\bar{U}\cap \bar{V}) = \overline{\phi(U \cap V)}\) onto \(\psi(\bar{U}\cap \bar{V}) = \overline{\psi(U \cap V)}\), in particular it maps \(\partial [\phi(U \cap V)]\) to \(\partial [\psi(U \cap V)].\) Then \(\phi(x)\) is a \(\mathcal{C}^0\) boundary point of \(\phi(U),\) if and only if \(\psi(x)\) is a \(\mathcal{C}^0\) boundary point of \(\psi(U),\) because \(\psi \circ \phi^{-1}\) and its inverse are both \(\mathcal{C}^0\), so their composition with a continuous function \(\gamma\) is still continuous.

Note that a border chart cannot be an interior chart, since for \((U,\phi)\) to be an interior chart, \(\phi(x)\) has to be an interior point of \(\phi(U),\) contradicting the fact that for a border chart, \(\phi(x)\) is a boundary point of \(\phi(U).\) What may not be immediate is that a point cannot be simultaneously an interior point with respect to one chart and a border point with respect to another chart. This is precisely the point of the next lemma.

Lemma 5. Let \(M\) be a paracompact Hausdorff topological space. Then \(int(M)\cap \partial M= \emptyset.\)

Proof. It follows from Proposition 5 that for any border chart \((U,\phi)\) centered at a border point \(x \in \partial M\), there exists \(\delta > 0\) such that the set \(int(\phi(U)) \cap \mathbb{B}_\delta\) is homeomorphic to a small relatively open ball around \(0_d\) in the closed upper half-space. But the latter cannot be homeomorphic to a small relatively open ball in the open upper half-space as can be shown by a classical relative homology argument. ◻

We are now in a position to introduce our definition of topological manifold with kinks.

Definition 9. Let \(M\) be a topological manifold with kinks with maximal atlas A \(d\)-dimensional atlas with kinks on \(M\) is a collection \(\{(U_i, \phi_i)\}_{i \in \mathcal{I}}\) of interior or border charts such that \(X=\cup_{i \in \mathcal{I}}U_i\). Such an atlas is called maximal if it is not a proper subcollection of any other atlas. If \(M\) admits a maximal \(d\)-dimensional atlas with kinks, then we say that \(M\) is a \(d\)-dimensional topological manifold with kinks.

The next proposition shows that from a topological point of view, there is no difference between a manifold with kinks and a manifold with boundary, just like there is no topological difference between a manifold with corners and a manifold with boundary. The proof is an immediate consequence of Proposition 5 which implies that the codomain of border charts are all locally homeomorphic.

Proposition 17. Any \(d\)-dimensional topological manifold with kinks \(M\) is locally homeomorphic to a \(d\)-dimensional topological manifold with boundary.

3.2 Differentiable manifolds with kinks↩︎

Let us now consider differentiable structures on topological manifolds with kinks. We let \(k\) be a positive integer that we keep fixed for the whole section.

Definition 10. Let \(M\) be a topological manifold with kinks. We say that two charts \((U,\phi), (V,\psi)\) around \(x\in M\) are \(\mathcal{C}^k\) compatible if the transition map \(\psi \circ \phi^{-1}:\phi(U\cap V)\to \psi(U\cap V)\) is a \(\mathcal{C}^k\) diffeomorphism in the sense of Definition 2.

The previous definition is classical for interior points and becomes relevant for border points only. In particular, it allows to define the regularity of border points thanks to the next key result.

Lemma 6. Let \(M\) be a topological manifold with kinks, and \(x\) a border point of \(M.\) Let \((U,\phi), (U,\psi)\) be two \(\mathcal{C}^1\) compatible border charts around \(x.\) Then \(\phi(x)\) is a Lipschitz (resp. c.d.d., \(\mathcal{C}^1\)) boundary point of \(\phi(U)\) if and only if it is a Lipschitz (resp. c.d.d., \(\mathcal{C}^1\)) boundary point of \(\psi(U)\).

Proof. Immediate from the fact that the transition map \(\psi \circ \phi^{-1}:\phi(U\cap V)\to \psi(U\cap V)\) extends to a \(\mathcal{C}^1\) diffeomorphism which preserves boundary (Lemma 1), and that the composition of a \(\mathcal{C}^1\) map with a Lipschitz (resp. c.d.d., \(\mathcal{C}^1\)) map is Lipschitz (resp. c.d.d., \(\mathcal{C}^1\)). ◻

Note that the previous claim fails if the transition maps are Hölder continuous. On the other hand, increasing the regularity of the border points independently of the border charts would also need us to increase the regularity of the transition maps.
We can now define differentiable manifolds with kinks.

Definition 11. Let \(M\) be a topological manifold with kinks with maximal atlas A \(d\)-dimensional \(\mathcal{C}^k\) atlas with kinks on \(M\) is a collection \(\{(U_i, \phi_i)\}_{i \in \mathcal{I}}\) of pairwise \(\mathcal{C}^k\) compatible interior or border charts such that \(X=\cup_{i \in \mathcal{I}}U_i\). Such an atlas is called maximal if it is not a proper subcollection of any other atlas. If \(M\) admits a maximal \(d\)-dimensional \(\mathcal{C}^k\) atlas with kinks, then we say that \(M\) is a \(d\)-dimensional \(\mathcal{C}^k\) manifold with kinks.

The word “kink” suggests particular boundary points where the manifold presents some type of singularity. With this in mind, we classify border points of \(\mathcal{C}^1\) manifolds with kinks as follows.

Definition 12. Let \(M\) be a \(\mathcal{C}^1\) manifold with kinks.

  1. We say that \(x \in \partial M\) is a \(\mathcal{C}^1\)-boundary point of \(M\) if there exists a local border chart \((U,\phi)\) around \(x\) such that \(\phi(x)\) is a \(\mathcal{C}^1\) boundary point of \(int(\phi(U))\). Otherwise we say that \(x\) is an essential kink.

  2. We say that an essential kink \(x\) is an essential corner of depth \(k\ge 2\) if there exists a border chart \((U,\phi)\) centered at \(x\) such that up to a rigid motion around \(0_d\) the image \(\phi(U)\) locally writes as the epigraph of the function \(\gamma(x_1\dots x_{d-1}):=\sum_{i=1}^{k-1}|x_i|\),

  3. We say that an essential kink \(x\) is an LCDD border point if there exists a border chart \((U,\phi)\) centered at \(x\) such that, up to a rigid motion, the image \(\phi(U)\) locally writes as the epigraph of a Lipschitz c.d.d. function.

Note that, according to our definition, any point on the boundary of a smooth manifold with boundary is a \(\mathcal{C}^1\) boundary point, and any corner on a manifold with corner is an essential corner. But the notion of essential kink captures wilder singularities, as illustrated below. Therefore, the category of manifolds with kinks is strictly larger than the ones of manifolds with corners, of manifolds with boundary, and of manifolds without boundary, and contains them all.

Examples 18.

  1. The space \(M:=\{(x,y,z)\in \mathbb{R}^3: z^2\le x^2+y^2, z>0\}\) is not a manifold with corner, because the point \(0_3\in \partial M\) is an essential kink which is not a corner.

  2. The space \(M:=\{(x,y)\in \mathbb{R}^2: y < \sqrt{|x|} \}\) is not a manifold with corner, because the point \(0_2\in \partial M\) is a cusp; in particular, \(0_2\) is an essential kink which is not a corner.

a
b

Figure 4: Essential kinks that are not essential corners. a — Pyramid, b — Cusp

We conclude this section with some definitions regarding fluctuating boundaries (recall Definition 6 for the case of open Euclidean subsets).

Definition 13. Let \(M\) be a \(\mathcal{C}^1\) manifold with kinks. Then \(x \in \partial M\) is called a (a.e.) non-fluctuating border point if there exists a chart \((U,\phi)\) centered at \(x\) such that \(int(\phi(U))\) has (a.e.) non-fluctuating boundary at \(0_d.\)

Remark 19. Lemma 4 implies that when a border point \(x\) is LCDD then for any chart \((U,\phi)\) centered at \(x\) the set \(int(\phi(U))\) has an a.e. non-fluctuating boundary at \(0_d\).

3.3 Tangent space and inward sector↩︎

In Subsection 2.3, we recalled the concepts of Bouligand tangent cone, feasible direction cone and open feasible direction cone at a point of a subset of \(\mathbb{R}^d\), and we derived some results on these cones at suitably regular boundary points. The goal of the present subsection is to introduce the corresponding notions for smooth manifolds with kinks and to derive peculiar properties at essential kinks.

3.3.1 Tangent space↩︎

From now on, we work only with smooth manifolds with kinks, i.e. those manifolds with kinks for which the transition maps are all \(\mathcal{C}^{\infty}\). This is to make sure that the tangent space defined below as the space of smooth derivations (i.e. linear forms on the space of smooth functions) is finite dimensional. Indeed, for any \(\mathcal{C}^k\) manifold with \(k < +\infty\), the space of \(\mathcal{C}^k\) derivations is infinite dimensional : see e.g. [23].

In order to introduce smooth derivations on manifolds with kinks, we must first define what a smooth function is. This is what we do in the next definition, largely inspired by the context of manifolds with corners (see e.g. [18]).

Definition 14. Let \(M\) be a \(d\)-dimensional smooth manifold with kinks. For any open set \(O \subset M\), a function \(f:O\to \mathbb{R}\) is called smooth in a neighborhood of \(x\in O\) if it is smooth in every chart \((U,\phi)\) centered at \(x\), namely :

  1. if \(x \in int M\), then there exists \(\rho>0\) such that \(f\circ \phi^{-1} : \phi(U) \cap \mathbb{B}_\rho^d \to \mathbb{R}\) is smooth in the classical sense,

  2. if \(x \in \partial M\), then there exist \(\rho>0\) and an open neighborhood \(V\subset \mathbb{R}^d\) of \(0_d\) containing \(\phi(\bar{U}) \cap \mathbb{B}_\rho^d\) to which \(f\circ \phi^{-1} : \phi(\bar{U}) \cap \mathbb{B}_\rho^d \to \mathbb{R}\) extends to a smooth function.

We denote by \(\mathcal{C}^\infty(O)\) the space of smooth functions on \(O\), that is to say, the functions which are smooth at any \(x \in O\).

Remark 20. If \(x \in \partial M\) and \(f\circ \phi^{-1} : \phi(\bar{U}) \cap \mathbb{B}_\rho^d \to \mathbb{R}\) admits Taylor polynomials of arbitrary high order, then the existence of a smooth extenstion of \(f\circ \phi^{-1}\) to \(\mathbb{R}^d\) is ensured by Whitney’s extension theorem [24], see also [25]. Note that in the setting of manifolds with corners, the latter upgrades into the well-posedness of Seeley’s linear extension operator [26], see [17].

We are now in a position to define smooth derivations and tangent spaces as follows.

Definition 15. Let \(M\) be a smooth manifold with kinks. Then the tangent space of \(M\) at \(x \in M\) is the real vector space \[T_xM :=\{\mathcal{D}: \mathcal{C}^{\infty}(M)\to \mathbb{R}\text{ linear }: \mathcal{D}(fg)= f(x) \mathcal{D}(g) + g(x)\mathcal{D}(f) \text{ for any f,g \in \mathcal{C}^{\infty}(M)}\}.\] Any element \(\mathcal{D} \in T_xM\) is called a smooth derivation at \(x\).

We can now define the differential of a smooth function between two smooth manifolds with kinks. There is no change with the case of classical manifolds, see e.g. [27].

Definition 16. Let \(M\) and \(N\) be smooth manifolds with kinks of dimension \(d\) and \(d'\) respectively, and \(O\subset M\) an open set.

  1. A function \(\Phi:O\to N\) is called smooth in a neighborhood of \(x\in O\) if its local expression is smooth in any couple of charts \((U,\phi)\) and \((U',\psi)\) centered at \(x\) and \(\Phi(x)\) respectively, namely:

    1. if \(x \in int M\), then there exists \(\rho>0\) such that \(\psi \circ f\circ \phi^{-1} : \phi(U) \cap \mathbb{B}_\rho^d \to \mathbb{R}^{d'}\) is smooth in the classical sense,

    2. if \(x \in \partial M\), then there exist \(\rho>0\) and an open neighborhood \(V\subset \mathbb{R}^d\) of \(0_d\) containing \(\phi(\bar{U}) \cap \mathbb{B}_\rho^d\) to which \(\psi \circ f\circ \phi^{-1} : \phi(\bar{U}) \cap \mathbb{B}_\rho^d \to \mathbb{R}\) extends to a smooth function.

    We denote by \(\mathcal{C}^\infty(O,N)\) the space of functions \(O\to N\) which are smooth in a neighborhood of any \(x \in O\).

  2. For any \(\Phi \in \mathcal{C}^\infty(O,N)\), the differential of \(\Phi\) at \(x \in M\) is the linear map \[\mathop{}\!\mathrm{d}_x \Phi : T_x M \to T_{\Phi(x)}N\] sending a derivation \(\mathcal{D}\in T_xM\) to the derivation \(\mathop{}\!\mathrm{d}_x \Phi(\mathcal{D}) \in T_{\Phi(x)}N\) defined by : \[\mathop{}\!\mathrm{d}_x \Phi(\mathcal{D})( h) :=\mathcal{D}(h\circ \Phi) \qquad \forall h \in \mathcal{C}^\infty(N).\]

  3. A function \(\Phi \in \mathcal{C}^\infty(O,N)\) is called a smooth diffomorphism onto its image if it is a smooth bijection with smooth inverse.

Remark 21. One can check with no harm that the usual chain rule holds in this context.

Let us now establish the following natural result.

Lemma 7. Let \(M\) be a smooth manifold with kinks and \(x\in \partial M.\) For any local chart \((U,\phi)\) centered at \(x\), the differential \[\mathop{}\!\mathrm{d}_x\phi:T_x M \to T_{0_d}\phi(\bar{U})\] is a linear isomorphism, and \(dim(T_xM)=dim(M).\)

Proof. When \(x\) is an interior point, a \(\mathcal{C}^1\) boundary point or an essential corner, this is already known, see [18]. Let us then assume that \(x\) is an essential kink which is none of the three previous cases. Consider a local chart \((U,\phi)\) centered at \(x\). Like for manifolds without boundary, we obtain from the chain rule applied to the identities \(\phi \circ \phi^{-1} = \mathrm{id}_{\phi(U)}\) and \(\phi \circ \phi^{-1} = \mathrm{id}_{U}\) that the differential \(\mathop{}\!\mathrm{d}_x\phi:T_x M \to T_{0_d}\phi(\bar{U})\) is a linear isomorphism with inverse \(\mathop{}\!\mathrm{d}_{0_d}\phi^{-1}\). But the space of derivations at the \(\mathcal{C}^0\) boundary point \(0_d\) of \(\phi(U)\) coincides with \(\mathbb{R}^n\), since a basis of this space is given by the classical partial differential operators \(\partial \cdot /\partial x_1,\ldots, \, \partial\cdot /\partial x_n\) defined on the open set \(\phi(U)\) and naturally extended to any open set \(V \subset \mathbb{R}^d\) containing \(\overline{\phi(U)}\). ◻

Remark 22. Like for manifolds without/with boundary or corners, the preceding lemma shows that any local chart \((U,\phi)\) centered at a point \(x\) in a manifold with kinks provides a linear isomorphism \(T_xM \simeq \mathbb{R}^d\).

3.3.2 Inward sector↩︎

Let us now introduce the notion of inward tangent sector for smooth manifolds with kinks. This is analogous to the inward tangent sector of Euclidean domains as discussed in Section 2.2. Our definition builds upon the classical characterization of the tangent space in terms of initial velocities of curves, which holds true for manifolds without/with boundary (see e.g. [27]). For manifolds with boundary, one can identify inward tangent vectors at boundary points by specifying the domain of the curves we choose. This is how we came up with the following natural definition.

Definition 17. Let \(M\) be a smooth manifold with kinks. For any \(x \in M\), we define the equivalence relation \(\sim\) on the set of smooth curves \(\gamma : I \to M\) such that \(I=[0,\epsilon)\) for some \(\epsilon >0\) and \(\gamma(0)=x\) by setting : \[\label{eq:equiv} \gamma_1 \sim \gamma_2 \iff (f \circ \gamma_1)'(0)= (f\circ \gamma_2)'(0) \, \, \forall f \in C^\infty(M).\tag{6}\] Then the inward tangent sector of \(M\) at \(x\) is the space of equivalent classes of such curves under \(\sim\) : \[I_xM := \{\gamma \colon I \to M \text{ smooth such that I=[0,\epsilon) for some \epsilon > 0 and \gamma(0)=x}\}/\sim .\]

It follows from 6 that any element \([\gamma] \in I_xM\) canonically defines a smooth derivation \(f \mapsto (f\circ \gamma)'(0)\) belonging to \(T_xM\). As such, \[\label{eq:inclusion} I_xM \subset T_xM.\tag{7}\] The converse is obvious when \(x \in int M\). We discuss the case \(x \in \partial M\) in Lemma 9 below.

We shall use the natural convention which denotes equivalent classes \([\gamma]\) as \(\gamma'(0)\), and think of these objects as initial velocities pointing towards the interior of \(M\).

Lemma 8. Let \(M\) be a smooth manifold with kinks and \((U,\phi)\) a local chart centered at some \(x \in \partial M\). Then for any \(y \in U \cap \partial M\), \[I_yM=[\mathop{}\!\mathrm{d}_{y}\phi]^{-1}(I_{\phi(y)}\phi(\bar{U})).\]

Proof. Let us first establish \(\subset\). If \(\gamma'(0) \in I_yM\) for some \(\gamma : [0,\epsilon) \to M\) of class \(\mathcal{C}^1\) such that \(\gamma(0)=y\), then extend \(\tilde{\gamma}:=\phi \circ \gamma : [0,\epsilon) \to \phi(\bar{U})\) to a \(\mathcal{C}^1\) curve \(\bar{\gamma} : (-\epsilon,\epsilon) \to \phi(\bar{U})\) in any way, for example by symmetrizing \(\tilde{\gamma}\) with respect to \(\phi(y)\). The chain rule yields that \(\bar{\gamma}'(0) = d_x \phi(\tilde{\gamma}'(0))\) so that \(\tilde{\gamma}'(0) = [d_x \phi]^{-1}(\bar{\gamma}'(0)) = [d_{0_d} \phi^{-1}](\bar{\gamma}'(0))\). Since \(\bar{\gamma}'(0) \in I_{0_d} \phi(U)\) we get that \(\tilde{\gamma}'(0) \in [d_{0_d} \phi^{-1}] (I_{0_d} \phi(U))\) as desired. Then \(\supset\) is proved along similar lines. ◻

Remark 23. The same proof shows that if \((U,\phi)\) is a local chart centered at some interior point \(x \in M\), then \(I_y M = [\mathop{}\!\mathrm{d}_{y}\phi]^{-1}(I_{\phi(y)}\phi(U))\) for any \(y \in U\). Since in this case \(\phi(y)\) belongs to the open set \(\phi(U)\), the inward tangent sector \(I_{\phi(y)}\phi(U)\) clearly coincides with \(\mathbb{R}^n\), hence we get \(I_yM = T_yM\). The same holds when \((U,\phi)\) is a local chart centered at a border point \(x\) and \(y\) belongs to \(U \backslash \partial M\).

We are now in a position to characterize border points in terms of their inward tangent sector.

Lemma 9. Let \(M\) be a smooth manifold with kinks, and \(x \in M\). Then the following holds. \[\begin{array}{lll} x \in \partial M & \iff & I_xM \neq T_xM.\\ \text{x is a \mathcal{C}^1 boundary point of M} & \iff & I_xM \simeq \mathbb{H}^d.\\ \text{x is an essential corner of depth k of M} & \iff & I_xM \simeq \mathbb{R}^d_k. \end{array}\]

Proof. If \(x \in \partial M\), then up to rigid motion \(\phi(\bar{U}) \subset \bar{\mathbb{H}}^d\) and the latter is a cone, thus \(I_{0_d}\phi(\bar{U}) \subset \bar{\mathbb{H}}^d\). This prevents \(I_x M\) from coinciding with \(T_xM\). The previous remark implies that if \(x\) is interior, then \(I_x M=T_xM\). Since \(x\) can be either border or interior, the first equivalence is established. To prove the second and third ones, notice that Lemma 9 implies in both cases that \(I_xM \simeq I_{0_d} \phi(\bar{U})\). The conclusion follows from the fact that for small enough \(\rho>0\), the set \(\phi(\bar{U})\cap \mathbb{B}_\rho^d\) is an open neighborhood of \(0_d\) in \(\mathbb{H}^d\) and \(\mathbb{R}^d_k\) respectively, which yields that \(I_{0_d} \phi(\bar{U})\) is \(\mathbb{H}^d\) in the first case only and \(\mathbb{R}^d_k\) in the second one only. ◻

3.3.3 Strictly inward sector↩︎

This section generalizes the Euclidean open feasible direction cone to the setting of manifolds with kinks. A natural attempt to do so might be to consider \[\{ \gamma'(0) : \gamma \colon I \to int(M) \text{ of class } \mathcal{C}^1 \text{ with } I = [0,\epsilon)\text{ for some \epsilon > 0 and } \gamma(0)=x \}/\sim\] where \(x\) is a border point of a smooth manifold with kinks, and \(\sim\) is like in 6 . A problem with this set is that even if a curve \(\gamma\) might entirely lie within \(int(M)\), the initial velocity vector \(\gamma'(0)\) may still be a boundary vector.

For instance, take \(M:=[0,\infty)\times [0,\infty) \subset \mathbb{R}^2, x:=0_2\), and consider \(\gamma(t):=(t+t^2,t^2)\) for \(t>0\) small. Then \(\gamma'(0)=(1,0)\) belongs to the previous set, but if we consider \(M\) as a subset of \(\mathbb{R}^2,\) its open feasible direction cone is \((0,+\infty)\times (0,+\infty)\) that does not contain the boundary vector \((1,0).\)

For this reason, we use a different definition. Before introducing it, let us provide a definition of cusp in this context.

Definition 18. Let \(M\) be a smooth manifold with kinks. An essential kink \(x \in \partial M\) is called a cusp if \[int I_xM = \emptyset.\] Here \(T_xM\) is endowed with the natural topology coming from any identification with \(\mathbb{R}^n\) through the differential of a local chart.

We are now in a position to define the strictly inward tangent sector at a point of a smooth manifold with kinks.

Definition 19. Let \(M\) be a smooth manifold with kinks. Then the strictly inward sector of \(M\) at \(x \in M\) is the set \[\tilde{I}_xM := \begin{cases} int\, I_xM & \text{if x \in \partial M is not a cusp}, \\ I_xM & \text{otherwise.} \end{cases}\]

3.4 Riemannian manifolds with kinks↩︎

In this section, we develop a suitable notion of Riemannian metric for manifolds with kinks. There is no particular difference compare to the case of manifolds without/with boundary, but we provide details for completeness.

3.4.1 Tangent bundle↩︎

Let us first define the tangent bundle on a smooth manifold with kinks. The definition is basically the same as for manifolds without/with boundary.

Definition 20. Let \(M\) be a smooth manifold with kinks. Then the tangent bundle of \(M\) is the vector bundle \[TM:=\bigsqcup_{p\in M}T_pM = \{(p,v):v\in T_pM\}.\]

Let us check the following natural result.

Lemma 10. Let \(M\) be a \(d\)-dimensional smooth manifold with kinks. Then \(TM\) is a \(2d\)-dimensional smooth manifold with kinks.

Proof. Denote by \(\pi\) the quotient map \(TM \to M\) mapping \((p,v)\) to \(p\). We equip \(TM\) with a topological/smooth structure by defining smooth charts as follows: for a local chart \((U, \phi)\) around \(p\in M\), we define a corresponding \(2d\)-dimensional chart \((\mathcal{U}:=\pi^{-1}(U),\Phi)\) for \(TM\) by setting \[\Phi(x,v) := (\phi(x), \mathop{}\!\mathrm{d}_x\phi(v))\] for all \((x,v) \in \mathcal{U}\). It is clear that for two such charts \((\mathcal{U},\Phi), (\mathcal{V}, \Psi)\) the transition map \(\Psi \circ \Phi^{-1}\) is smooth on \(\Phi(\mathcal{U}\cap \mathcal{V})\), because \(\psi \circ \phi^{-1}\) is smooth on \(\phi(\mathcal{U}\cap \mathcal{V})\), and \(D\psi_x \circ (D\phi_x)^{-1}:\mathbb{R}^d\to \mathbb{R}^d\) is smooth as it is linear. We will show that with respect to these charts, the interior and border of \(M\) match up with the interior and border of the tangent bundle \(TM\); in this way we will have constructed a maximal smooth atlas with kinks on \(TM\). This is obvious for the interior since the local charts are defined as in the case of manifolds without boundary. Let us then consider \(p \in \partial M\) and a border chart \((U,\phi)\) centered at \(p\). Then \(\phi(p)=0_d\) is a \(\mathcal{C}^0\) boundary point of \(\operatorname{int}(\phi(U))\). Let us show that for any \(v \in T_pM\), \[\Phi(p,v) = (\phi(p), \mathop{}\!\mathrm{d}_p \phi(v))\] is a \(\mathcal{C}^0\) boundary point of \(int(\Phi(\mathcal{U}))\), where \(\mathcal{U}:=\pi^{-1}(U)\). Shrinking \(U\) if necessary, we can identify \(\mathcal{U}\) with \(U \times \mathbb{R}^d\), so that \(\Phi(\mathcal{U})\) is \(\phi(U) \times \mathbb{R}^d\). Then \(\partial \Phi(\mathcal{U}) = \partial \phi(U) \times \mathbb{R}^d\). Since \(\phi(p)\) is a \(\mathcal{C}^0\) boundary point of \(\phi(U)\), we get that any \((\phi(p), v)\) is a \(\mathcal{C}^0\) boundary point of \(\phi(U) \times \mathbb{R}^d\). ◻

Remark 24. We could also define the cotangent bundle on a smooth manifold with kinks as \(T^*M :=\bigsqcup_{p\in M}T_p^*M = \{(p,\omega):\omega\in T_p^*M\}\), where each \(T_p^*M\) is the dual of \(T_pM\). We do not delve on this notion since we don’t need it in the rest of the paper.

3.4.2 Covariant \(k\)-tensor bundle↩︎

Let \(k\) be a positive integer. Recall that a covariant \(k\)-tensor on a vector space \(V\) is an element of the \(k\)-fold tensor product \(V^* \otimes \dots \otimes V^*\) or, equivalently, a \(k\)-linear map \(V \times V \times \dots \times V \to \mathbb{R}.\) We denote by \(T^k(V^*)\) the space of all covariant \(k\)-tensors on \(V\). Then we can define the covariant \(k\)-tensor bundle on a smooth manifold with kinks as follows.

Definition 21. Let \(M\) be a smooth manifold with kinks. Then the space of covariant \(k\)-tensors on \(M\) is defined as \[T^k(T^*M):=\bigsqcup_{p\in M}T^k(T_p^*M).\]

Acting like in the previous subsection, one can easily show the following. We omit the proof for brevity.

Lemma 11. Let \(M\) be a smooth manifold with kinks. Then \(T^k(T^*M)\) is a smooth manifold with kinks too.

Remark 25. Likewise, we could define \((k,r)\)-tensors on smooth manifolds with kinks, but we do not need them in the present paper so we skip them.

3.4.3 Riemannian metrics↩︎

Let us define covariant tensor fields on manifold with kinks.

Definition 22. Let \(M\) be a smooth manifold with kinks. A \(k\)-tensor field on \(M\) is a section \(g\) of the covariant \(k\)-tensor bundle \({T}^k(T^*M)\). Such a field is of \(\mathcal{C}^\ell\) regularity if \(g:M\to {T}^k(T^*M)\) is a \(\mathcal{C}^{\ell}\) map w.r.t. the smooth structures introduced in the previous section.

We are now in a position to define smooth Riemannian metrics on smooth manifolds with kinks.

Definition 23. Let \(M\) be a smooth manifold with kinks. A smooth Riemannian metric on \(M\) is a smooth, symmetric, positive definite, section of the covariant \(2\)-tensor bundle \({T}^2(T^*M)M\).

3.4.4 Riemannian distance and volume measure↩︎

The Riemannian distance defined via a length-minimizing problem extends with no change to the context of manifolds with kinks. We recall the definition for completeness and refer to [27], for instance, for more details.

Definition 24. Let \(M\) be a smooth connected manifold with kinks admitting a \(\mathcal{C}^1\) Riemannian metric \(g\). The associated Riemannian distance is defined by \[\mathop{}\!\mathrm{d}(x,y) :=\inf \left\{\int_0^1 g_{\gamma(t)}(\gamma'(t),\gamma'(t)) \mathop{}\!\mathrm{d}t : \gamma \in \mathcal{C}^1([0,1],M) \text{ s.t.~}\gamma(0)=x \text{ and } \gamma(1)=y \right\},\] for any \(x,y \in M\).

In the same way, the definition of the Riemannian volume measure carry over to manifolds with kinks and behaves like in the case of manifolds without boundary.

Definition 25. Let \(M\) be a smooth manifold with kink admitting a \(\mathcal{C}^1\) Riemannian metric \(g\). Then the Riemannian volume measure is defined by \[\mathrm{vol}_g(A) :=\sum_{\alpha} \int_{\phi_\alpha(U_\alpha)} \chi_\alpha \circ \phi_\alpha^{-1}\sqrt{\det g}\] for any Borel set \(A \subset M\), where \(\{(U_\alpha, \phi_\alpha\}\) is an atlas of \(M\) and \(\{\chi_\alpha\}\) a partition of unity subordinate to this atlas.

3.5 Extension of Riemannian manifolds with kinks↩︎

For our purposes, we need to extend beyond the border any smooth Riemannian metric on a smooth manifold with kinks \(M\). To this aim, we shall flow \(M\) into its interior using a suitable semiflow. We adapt an argument for manifolds with corners that goes back to [15] at least, see also [16].

3.5.1 Vector fields and semiflows↩︎

Let us begin by defining vector fields and their associated semiflows on smooth manifolds with kinks.

Definition 26. Let \(M\) be a smooth manifold with kinks. A smooth vector field on \(M\) is a smooth section of the tangent bundle \(TM.\) Such a vector field \(\xi\) is called (resp. strictly) inward-pointing if \(\xi_x\in I_xM\) (resp. \(\tilde{I}_xM\)) for any \(x\in M.\)

Let us ensure that any smooth manifold with kinks admits a stricly inward-pointing vector field.

Lemma 12. Let \(M\) be a smooth manifold with kinks. Then \(M\) admits a strictly inward-pointing vector field.

Proof. Let \((U,\phi)\) be a border chart centered at some \(x \in \partial M\). Up to composing \(\phi\) with an isometry, we may assume that there exist \(\delta>0\) and \(\gamma \in \mathcal{C}^0(\mathbb{R}^{n-1})\) such that \[\begin{cases} \phi(\bar{U}) \cap \mathbb{B}_\delta^d = epi(\gamma) \cap \mathbb{B}_\delta^d, \nonumber\\ int \phi(\bar{U}) \cap \mathbb{B}_\delta^d = \mathring{epi}(\gamma) \cap \mathbb{B}_\delta^d. \end{cases}\] Then we set \(\xi_{(U,\phi)}(y) :=(\mathop{}\!\mathrm{d}_{y}\phi)^{-1}(e_d)\) for any \(y \in \bar{U} \cap \phi^{-1}(\mathbb{B}_\delta^d)\). We get from Proposition 8 that \(\xi_{(U,\phi)}(y) \in \tilde{I}_xM\). Consider now an atlas \(\{(U_\alpha, \phi_\alpha\}\) of \(M\) and a partition of unity \(\{\chi_\alpha\}\) subordinate to this atlas. Define \[\xi :=\sum_{\substack{(U_\alpha,\phi_\alpha)\\ \text{border charts}}} \chi_\alpha \, \xi_{(U_\alpha,\phi_\alpha)} : M \to TM.\] Then \(\xi\) is a global smooth vector field on \(M\), and it is strictly inward-pointing by construction. ◻

Recall the definition of integral curve.

Definition 27. Let \(M\) be a smooth manifold with kinks, and \(\xi\) a smooth vector field on it. An integral curve of \(\xi\) is a smooth curve \(\gamma : I \to M\) such that \(\gamma'(t) = \xi(\gamma(t))\) for any \(t \in I\).

Then the following existence result holds.

Theorem 26. Let \(M\) be a smooth manifold with kinks, and \(\xi\) an inward pointing vector field on it. Then there exists a smooth function \(\delta : \partial M \to (0,+\infty)\) and a smooth embedding \(\Phi : \mathcal{P}_\delta \to M\), with \(\mathcal{P}_\delta = \{ (t,x) : x \in \partial M, t \in [0,\delta(x)) \} \subset \mathbb{R}\times \partial M\), such that for any \(x \in \partial M\) the map \([0,\delta(x)) \ni t \mapsto \Phi(t,x)\) is an integral curve of \(\xi\) starting at \(x\).

Proof. The function \(\delta\), the set \(\mathcal{P}_\delta\) and the embedding \(\Phi\) are first defined locally, and then patched together by means of a partition of unity. To define these objects locally around some point \(x \in \partial M\), consider a border chart \((U,\phi)\) centered at \(x\). By definition of border chart, there exists \(\rho>0\) such that up to a rigid motion, \(\phi(U) \cap \mathbb{B}_\rho^d\) writes as the local epigraph of some continuous function \(h:\mathbb{R}^{d-1}\to \mathbb{R}\) such that \(h(0)=0\). Since \(\phi(U \cap \partial M)\) is mapped to the local graph of this function, and since \(\mathop{}\!\mathrm{d}\phi(\xi)\equiv e_d\), we can apply the Cauchy–Lipschitz theorem in \(\mathbb{R}^d\) to get existence of a smooth function \(\delta : \phi(U \cap \partial M) \to \mathbb{R}_+\) such that for any \(y \in U \cap \partial M\) there exists an integral curve \(\gamma_{\phi(y)} : [0,\delta(y))] \to \mathbb{R}^d\) of \(\mathop{}\!\mathrm{d}\phi(\xi)\) starting at \(\phi(y)\), so that the map \([0,\delta(y)) \ni t \mapsto \Phi(t,y) :=\phi^{-1}(\gamma_{\phi(y)}(t))\) is an integral curve of \(\xi\) starting at \(y\). ◻

3.5.2 Submanifolds with kinks↩︎

Let us now provide a definition of submanifold adapted to the context with kinks, inspired by [16] who introduced submanifolds with corners.

Definition 28. Let \(M\) be a \(d\)-dimensional manifolds with kinks, and \(k\) a positive integer at most equal to \(d\). We say that \(N \subset M\) is a \(k\)-dimensional submanifold with kinks of \(M\) if for every \(p\in N,\) there is a chart \((U,\phi)\) of \(M\) centered at \(p\) so that \(\phi(U \cap N) \subset \mathbb{R}^k \times \{0\}^{d-k} \subset \mathbb{R}^d\) and \(\phi(p)=0_d\) is an interior or a \(\mathcal{C}^0\) boundary point of \(\phi(U\cap N).\) We call \((U,\phi)\) a slice chart centered at \(p\) of \(N\)

Our main result now is that any smooth manifold with kinks can be embedded into a manifold with boundary having same dimension. The proof actually embeds a smooth manifold with kinks into its interior.

Theorem 27. Let \(M\) be a \(d\)-dimensional manifolds with kinks. Then there exists a \(d\)-dimensional manifold with boundary \(\tilde{M}\) of which \(M \subset \tilde{M}\) is a smooth submanifold with kinks.

Proof. Consider a strictly inward pointing vector field \(\xi\) on \(M\) as given by Lemma 12. Then the semiflow of \(\xi\) given by Theorem 26 maps \(\partial M\) into \(int(M),\) and \(int(M)\) into \(int(M).\) ◻

3.5.3 Riemannian extension↩︎

We will next show that under this assumption, any \(\mathcal{C}^2\) Riemannian metric can be extended from \(M\) to the manifold \(\tilde{M}\) without boundary given by Lemma 27.

Proposition 28. Let \(M\) be a smooth submanifold with Lipschitz kinks of an open smooth manifold \(\tilde{M}\). Assume that \(M\) admits a \(\mathcal{C}^{2}\) Riemannian metric \(g.\) Then there exist a neighborhood \(O\) of \(M\) in \(\tilde{M},\) i.e. an open subset \(O\subset \tilde{M}\) containing \(M,\) and a \(\mathcal{C}^{2}\) Riemannian metric \(\tilde{g}\) on \(O\) so that \(\tilde{g}\) restricts to \(g\) on \(M.\)

Proof. As usual in many extension theorems, the idea is to extend the Riemannian metric locally to an open subset of \(\tilde{M},\) and then use a smooth partition of unity. We only need to extend the metric at border points of \(M\), since interior points of \(M\) are interior points of \(\tilde{M}\)

Let \((U, \phi)\) be a border chart around a Lipschitz kink \(x \in M\), with \(U \subset \tilde{U}\) open in \(\tilde{M}\), and let \(\tilde{\phi}\) be an extension of \(\phi\) from \(U\) to \(\tilde{U}\). Let \(g\) be a \(\mathcal{C}^{2}\) Riemannian metric on \(U\). The steps to extend \(g\) from \(U\) to \(\tilde{U}\) are:

  1. Pull back \(g\) to \(\phi(U)\) by \(\phi^{-1}\), i.e., consider \((\phi^{-1})^*g\) on \(\phi(U)\). This is a \(\mathcal{C}^2\) Riemannian metric on \(\phi(U)\).

  2. Using Whitney’s theorem, extend each coordinate of \((\phi^{-1})^*g\) to form a Riemannian metric \(h\) on \(V\), an open subset of \(\mathbb{R}^d\) containing \(\phi(x)=0\), such that \(\phi^{-1}(V) \subset \tilde{U}\). Note that this extension may not be unique and depends on the choice of the border chart \((U,\phi)\).

  3. Pull back \(h\) by \(\tilde{\phi}\) to \(\tilde{U}\), i.e., consider \(\tilde{\phi}^* h\) on \(\tilde{U} \supset U\). Since the extension in the previous step is not unique, this pullback may not be unique either. Nevertheless, regardless of the specific extension in Step 2 or the chart used, the tensor \(g\) extends to a \(\mathcal{C}^{2}\) Riemannian metric \(\tilde{g}\) on \(\tilde{U}\) such that its restriction to \(U\) agrees with \(g\), by the contravariant functoriality of pullbacks applied to the composition \(\phi \circ \phi^{-1}=Id\).

Note that the specific extension \(\tilde{g}\) can depend on both the chart chosen and the method of extension in \(\mathbb{R}^d\) used in Step 2.

Next use a smooth partition of unity argument to obtain the extended Riemannian metric on \(O:=\bigcup_{x \in M} \tilde{U}_x,\) which will be \(\mathcal{C}^2,\) as the product of \(\mathcal{C}^2\) and smooth (\(\mathcal{C}^{\infty}\)) (partition of unity) functions are again \(\mathcal{C}^2.\) ◻

4 Asymptotic behavior of the intrinsic Gaussian Operator↩︎

In this section, we prove Theorem 1.

4.1 Localisation↩︎

Recall that for a metric measure space \((M,\mathop{}\!\mathrm{d},\mu)\), the intrinsic \(d\)-dimensional Gaussian operator at time \(t>0\) is defined by \[L_t f(x) :=\frac{1}{t^{d/2+1}}\int_{M} \exp\left(-\frac{\mathop{}\!\mathrm{d}^2(x,y)}{t}\right) (f(x)-f(y)) \mathop{}\!\mathrm{d}\mu(x)\] for any \(f \in \mathcal{C}(M)\) and \(x \in M\). We are interested in the short-time behavior of this operator. In this regard, the next simple lemma implies that all the information about this behavior is contained in small balls. The proof is immediate so we skip it.

Lemma 13. Let \((M,\mathop{}\!\mathrm{d},\mu)\) be a metric measure space and \(\eta \in (0,1/2)\). If \(\mu(M)<+\infty\), then \[\left| \frac{1}{t^{d/2+1}} \int_{M\backslash B_{t^\eta }(x)} \exp\left(-\frac{\mathop{}\!\mathrm{d}^2(x,y)}{t}\right) (f(x)-f(y)) d\mathrm{\mu}(y) \right| \leq [|f(x)|\mu(M) + \left\lVert f\right\rVert_{1}] \, \frac{1}{t^{d/2 +1}} e^{-t^{2\eta -1}}\] for any \(f \in \mathcal{C}(M)\cap L^1(M,\mu)\) and \(x \in X\). If \(\mu(M) = +\infty\), then the previous holds whenever \(f\) is integrable with bounded support, in which case \(\mu(M)\) must be replaced by \(\mu(\mathrm{supp} f)\).

As a consequence, we get \[L_tf(x) = \frac{1}{t^{d/2+1}} \int_{B_{t^\eta }(x)} \exp\left(-\frac{\mathop{}\!\mathrm{d}^2(x,y)}{t}\right) (f(x)-f(y)) d\mathrm{\mu}(y) + O(t^{-d/2-1}e^{-t^{2 \eta -1}})\] and so Theorem 1 follows from studying the asymptotic expansion of the first term in the previous right-hand side. Let us give a name to this quantity: \[\label{eq:localized} \bar{L}_{t,\eta } f (x) :=\frac{1}{t^{d/2+1}} \int_{B_{t^\eta }(x)} \exp\left(-\frac{\mathop{}\!\mathrm{d}^2(x,y)}{t}\right) (f(x)-f(y)) d\mathrm{\mu}(y)\tag{8}\]

4.2 Change of coordinates↩︎

Let us put ourselves in the context of Theorem 1.

We consider a smooth Riemannian manifold with kinks \((M^d,g)\), a density \(p \in \mathcal{C}_{\ge 0}^2(M)\), a function \(f \in \mathcal{C}^3(M)\), a point \(x \in M\) that is LCDD border (note that the case were \(x\) is interior is covered by 1 ), and we choose \(\eta \in (0,1/2)\). We let \((\tilde{M},\tilde{g})\) be the open Riemannian manifold of which \((M,g)\) is a submanifold, see Lemma 27 and Proposition 28. We denote by \(\tilde{B}_r(x)\) the \(\tilde{g}\)-ball of radius \(r\) centered at \(x\), by \(\tilde{\exp}_x\) the \(\tilde{g}\)-exponential map at \(x\), and by \(\tilde{\gamma}_{x,v}\) the unique maximal \(\tilde{g}\)-geodesic with initial point \(x\) and initial velocity \(v \in T_x \tilde{M}\). Since \(x\) is an interior point for \(\tilde{M}\), there exists \(R>0\) such that \(\tilde{\exp}_x\) is a diffeomorphism from \(\mathbb{B}_R^d \subset T_x \tilde{M}\) onto \(\tilde{B}_R(x)\). Moreover, since \((\tilde{M},\tilde{g})\) extends \((M,g)\), the set of \(\mathcal{C}^1\) curves joining \(x\) to \(y \in \tilde{B}_R(x)\) and lying entirely in \(M\) is a subset of those curves lying in \(\tilde{M}\), so that \[\label{eq:ineq95distance} \tilde{\mathop{}\!\mathrm{d}}(x,y) \le \mathop{}\!\mathrm{d}(x,y).\tag{9}\] As a consequence, for any \(r \in (0,R)\), \[B_r(x) \subset \tilde{B}_r(x).\] This implies, in particular, that \((B_R(x), \tilde{\exp}_x^{-1}|_{B_r(x)})\) is a border chart centered at \(x\) of \(M\).

As usual, we identify the Euclidean space \((T_x \tilde{M},g(x))\) with \((\mathbb{R}^d,\cdot)\) by choosing a \(g(x)\)-orthonormal basis \((w_1,\ldots,w_d)\) of \(T_x \tilde{M}\) and mapping each \(w_i\) to the \(i\)-th element of the canonical basis of \(\mathbb{R}^d\). We will freely use this identification in the sequel. Let us also point out that \((T_xM,g(x))\) and \((T_x\tilde{M},\tilde{g}(x))\) coincide as Euclidean spaces. We define \[\Omega :=\tilde{\exp}_x^{-1}(B_R(x)).\] This is an open subset of \(T_x\tilde{M} \simeq \mathbb{R}^d\) having \(0_d\) as a \(\mathcal{C}^0\) boundary point.

Proposition 29. Set \(\tilde{f}:=f \circ \tilde{\exp}_x\) and \(q :=(p \circ \tilde{\exp}_x) \sqrt{\det g}\). Then \[\bar{L}_{t,\eta} f (x) = \frac{1}{t} \int_{\mathbb{B}^d_{t^{\eta}}\cap \tilde{F}_{0_d}\Omega} e^{-\frac{\|\xi\|^2}{t}} (\tilde{f}(0_d) - \tilde{f}(\xi))q(\xi)\mathop{}\!\mathrm{d}\xi + o\left( \frac{1}{\sqrt{t}}\right) \qquad \text{as t\to 0.}\]

Proof. Set \[\begin{align} W & :=\{v \in T_x \tilde{M} \simeq \mathbb{R}^d : \text{ there exists } t_v >0 \text{ s.t.~}\tilde{\gamma}_{x,v}(t) \in M \text{ for all t \in (0,t_v)}\}. \end{align}\] Then \[\bar{L}_{t,\eta} f (x) = I(t) + II(t)\] where \[\begin{align} I(t) & :=\frac{1}{t^{d/2+1}} \int_{B_{t^\eta}(x)\cap \tilde{\exp}_x(W\cap \mathbb{B}_R^d)} \exp\left(-\frac{\mathop{}\!\mathrm{d}^2(x,y)}{t}\right) (f(x)-f(y)) \, d\mathrm{\mu}(y),\\ II(t) & :=\frac{1}{t^{d/2+1}} \int_{B_{t^\eta}(x)\backslash \tilde{\exp}_x(W\cap \mathbb{B}_R^d)} \exp\left(-\frac{\mathop{}\!\mathrm{d}^2(x,y)}{t}\right) (f(x)-f(y)) d\mathrm{\mu}(y). \end{align}\]

Step 1. We show that for any small enough \(t>0\), \[\begin{align} \label{eq:step133} I(t) = \frac{1}{t} \int_{\mathbb{B}^d_{t^{\eta}}\cap \tilde{F}_{0_d}\Omega} e^{-\frac{\|\xi\|^2}{t}} (\tilde{f}(0_d) - \tilde{f}(\xi))q(\xi)\mathop{}\!\mathrm{d}\xi. \end{align}\tag{10}\]

To this purpose, notice that \(W\) is a cone : indeed, if \(v \in W\) and \(\lambda>0\), then \(\tilde{\gamma}_{x,\lambda v}(t) = \tilde{\gamma}_{x,v}(\lambda t)\) for any \(t \in (0,t_v/\lambda)\), so that \(\lambda v \in W\) with \(t_{\lambda v}=t_v/\lambda\).

For any \(v \in W\), set \(s_v :=\sup \{s >0 : \tilde{\gamma}_{x,v}(s') \in M \text{ for any s' \in [0,s]}\} \in (0,+\infty]\). Let us show that \[\label{eq:technical} B_{t^{\eta}}(x) \cap \tilde{\exp}_x(W) = B_{t^{\eta}}(x) \cap \{\tilde{\exp}_x(sv) : v \in W \text{ with } \|v\|_{g_x} = 1 \text{ and } s \in (0,s_v) \}.\tag{11}\] Since \(W\) is a cone, \[W = \bigsqcup_{\substack{v \in W\\ \|v\|_{g_x} = 1}} \mathbb{R}_+ v,\] thus \[B_{t^{\eta}}(x) \cap \tilde{\exp}_x(W) = \bigsqcup_{\substack{v \in W\\ \|v\|_{g_x} = 1}} B_{t^{\eta}}(x) \cap \tilde{\exp}_x(\mathbb{R}_+ v).\] Now for each \(v \in W\) such that \(\|v\|_{g_x} = 1\), \[B_{t^{\eta}}(x) \cap \tilde{\exp}_x(\mathbb{R}_+ v) = \{\tilde{\exp}_x(sv) : s \in (0,s_v) \} \cap B_{t^{\eta}}(x).\] Then \[B_{t^{\eta}}(x) \cap \tilde{\exp}_x(W) = \bigsqcup_{\substack{v \in W\\ \|v\|_{g_x} = 1}} B_{t^{\eta}}(x) \cap \tilde{\exp}_x(\mathbb{R}_+ v) = \bigsqcup_{\substack{v \in W\\ \|v\|_{g_x} = 1}} \{\tilde{\exp}_x(sv) : s \in (0,s_v) \} \cap B_{t^{\eta}}(x)\] hence we get 11 .

Let us now prove that for any \(y \in B_{t^{\eta}}(x) \cap \tilde{\exp}_x(W)\), \[\label{eq:equal95distance} \tilde{\mathop{}\!\mathrm{d}}(x,y) = \mathop{}\!\mathrm{d}(x,y).\tag{12}\] Thanks to 11 , we know that there exist \(v \in W\) with \(\|v\|_{g_x}=1\) and \(s \in [0,s_v)\) such that \(y = \tilde{\exp}_x(sv)\) and \(\tilde{\exp}_x(s'v)\in M\) for any \(s' \in [0,s]\). Since \(g\) and \(\tilde{g}\) coincide on \(M\), the \(g\)-geodesic joining \(x\) and \(y\) coincides with \(s' \mapsto \tilde{\exp}_x(s'v)\). This yields 12 . Lastly, we point out that \[\label{eq:step1951} W = \tilde{F}_{0_d} \Omega\tag{13}\] and \[\label{eq:step1951bis} \tilde{\exp}_x^{-1}(B_{t^\eta}(x)) \cap W = \mathbb{B}^d_{t^\eta} \cap W.\tag{14}\] These are directly resulting from the fact that \(\tilde{\gamma}_{x,v}(s) = \tilde{\exp}_x (sv)\) and \(t\mapsto tv\) is the Euclidean geodesic in \(\mathbb{R}^d\) starting at \(0_d\) with initial velocity \(v\).

We are now in a position to obtain 10 . First, we use 12 and the change of variable \(y = \exp_x(\xi)\) to obtain \[\begin{align} I(t) & = \frac{1}{t^{d/2+1}} \int_{B_{t^\eta}(x)\cap \tilde{\exp}_x(W\cap \mathbb{B}_R^d)} \exp\left(-\frac{\tilde{\mathop{}\!\mathrm{d}}^2(x,y)}{t}\right) (f(x)-f(y)) \, p(y)d\mathrm{vol}_g(y)\\ & = \frac{1}{t^{d/2+1}} \int_{\tilde{\exp}_x^{-1}(B_{t^\eta}(x))\cap W} \exp\left(-\frac{\|\xi\|^2}{t}\right) (\tilde{f}(0_d)-\tilde{f}(\xi)) q(\xi)\, d \xi. \end{align}\] Then we apply 13 and 14 to get 17 as sought.

Step 2. We show that \[\begin{align} \label{eq:step4} II(t) = o\left(\frac{1}{\sqrt{t}}\right) \qquad \text{as t \to 0.} \end{align}\tag{15}\]

To this purpose, let us first establish that as \(t \downarrow 0\), \[\label{eq:step2952} 1_{\frac{\tilde{\exp_x}^{-1}(B_{t^\eta}(x)) }{\sqrt{t}}\backslash W} \to 0 \qquad \text{a.e.}\tag{16}\] For any \(t >0\) such that \(t^{\eta} \le R\), \[\begin{align} 1_{\frac{\tilde{\exp_x}^{-1}(B_{t^\eta}(x)) \backslash W}{\sqrt{t}}} & \le 1_{\frac{\tilde{\exp_x}^{-1}(B_{R}(x)) \backslash W}{\sqrt{t}}} = 1_{\frac{\Omega \backslash W}{\sqrt{t}}}. \end{align}\] Since \(W\) is a cone, the latter characteristic function is equal to \[1_{\frac{\Omega}{\sqrt{t}} \backslash W} = 1_{\frac{\Omega}{\sqrt{t}}} - 1_{W} = 1_{\frac{\Omega}{\sqrt{t}}} - 1_{\tilde{F}_{0_d}\Omega} \stackrel{t \downarrow 0}{\longrightarrow} 1_{\tilde{F}_{0_d}(\Omega)} - 1_{\tilde{F}_{0_d}(\Omega)} = 0 \qquad \text{a.e.}\] Here we use 13 to get the second equality and Proposition 12 for the convergence a.e. as \(t\downarrow 0\).

Let us now estimate \[\begin{align} |II(t)| & \le \frac{1}{t^{d/2+1}} \int_{B_{t^\eta}(x)\backslash \tilde{\exp}_x(W\cap \mathbb{B}_R^d)} \exp\left(-\frac{\mathop{}\!\mathrm{d}^2(x,y)}{t}\right) |f(x)-f(y)| p(y)d\mathrm{vol}_g(y) \\ & \le \frac{1}{t^{d/2+1}}\int_{B_{t^\eta}(x)\backslash \tilde{\exp}_x(W\cap \mathbb{B}_R^d)} \exp\left(-\frac{\tilde{\mathop{}\!\mathrm{d}}^2(x,y)}{t}\right) |f(x)-f(y)| p(y)d\mathrm{vol}_g(y)\\ & = \frac{1}{t^{d/2+1}} \int_{\tilde{\exp}_x^{-1}(B_{t^\eta}(x))\backslash W} \exp\left(-\frac{\|\xi\|^2}{t}\right) |\tilde{f}(0_d)-\tilde{f}(\xi)| q(\xi)d\xi \\ & = \frac{1}{t} \int_{\frac{\tilde{\exp}_x^{-1}(B_{t^\eta}(x))}{\sqrt{t}}\backslash W} \exp\left(-\|\zeta\|^2 \right) |\tilde{f}(0_d)-\tilde{f}(\sqrt{t}\zeta)| q(\sqrt{t}\zeta)d\zeta \end{align}\] where we use 9 to get the second inequality, the change of variable \(y = \tilde{\exp}_x(\xi)\) to get the penultimate line, and \(\zeta = \xi/\sqrt{t}\) to get the last one. Now we use the Taylor theorem with Laplace remainder : for any \(\zeta \in \frac{\tilde{\exp}_x^{-1}(B_{t^\eta}(x))}{\sqrt{t}}\backslash W\) there exist \(s_1,s_2 \in (0,\sqrt{t})\) such that \[\tilde{f}(0_d)-\tilde{f}(\sqrt{t}\zeta) = - \mathop{}\!\mathrm{d}_{0_d} \tilde{f} (\sqrt{t}\zeta) - \frac{1}{2} \mathop{}\!\mathrm{d}^{(2)}_{\sqrt{s_1}\zeta} \tilde{f} ((\sqrt{t}\zeta)^{(2)}) ,\] \[q(\sqrt{t}\zeta) = q(0_d) + \mathop{}\!\mathrm{d}_{\sqrt{s_2}\zeta} q (\sqrt{t}\zeta),\] so that \[\begin{align} |\tilde{f}(0_d)-\tilde{f}(\sqrt{t}\zeta)|q(\sqrt{t}\zeta) & \le \sqrt{t} \, q(0_d) |\mathop{}\!\mathrm{d}_{0_d} \tilde{f} (\zeta)| + \frac{t \, q(0_d)}{2} |\mathop{}\!\mathrm{d}^{(2)}_{\sqrt{s_1}\zeta} \tilde{f} (\zeta^{(2)})| \\ & + t \mathop{}\!\mathrm{d}_{\sqrt{s_2}\zeta} q (\zeta) |\mathop{}\!\mathrm{d}_{0_d} \tilde{f} (\zeta)| + \frac{t^{3/2}}{2} \mathop{}\!\mathrm{d}_{\sqrt{s_2}\zeta} q (\zeta) |\mathop{}\!\mathrm{d}^{(2)}_{\sqrt{s_1}\zeta} \tilde{f} (\zeta^{(2)})| \\ & \le \sqrt{t} \, q(0_d) \| \mathop{}\!\mathrm{d}_{0_d} \tilde{f}\|_{op} \|\zeta\| + \frac{t \, q(0_d) \|\zeta\|^{2} }{2} \sup_{z \in \tilde{\exp}_x^{-1}(B_{t^\eta}(x))\backslash W }\|\mathop{}\!\mathrm{d}^{(2)}_{z} \tilde{f}\|_{op} \\ & + t \|\zeta\|^{2} \|\mathop{}\!\mathrm{d}_{0_d} \tilde{f} \|_{op} \sup_{z \in \tilde{\exp}_x^{-1}(B_{t^\eta}(x))\backslash W } \| \mathop{}\!\mathrm{d}_{z} q \|_{op}\\ & + \frac{t^{3/2}\| \zeta\|^{3} }{2} \sup_{z \in \tilde{\exp}_x^{-1}(B_{t^\eta}(x))\backslash W } \| \mathop{}\!\mathrm{d}_{z} q\|_{op} \|\mathop{}\!\mathrm{d}^{(2)}_{z} \tilde{f}\|_{op}\\ & \le \sqrt{t} \, q(0_d) \| \mathop{}\!\mathrm{d}_{0_d} \tilde{f}\|_{op} \|\zeta\| + t\|\zeta\|^2 C_{x,f,p}(t) \end{align}\] where we have set \[\begin{align} C_{x,f,p}(t) & = \frac{q(0_d) }{2} \sup_{z \in \tilde{\exp}_x^{-1}(B_{t^\eta}(x))\backslash W }\|\mathop{}\!\mathrm{d}^{(2)}_{z} \tilde{f}\|_{op} + \|\mathop{}\!\mathrm{d}_{0_d} \tilde{f} \|_{op} \sup_{z \in \tilde{\exp}_x^{-1}(B_{t^\eta}(x))\backslash W } \| \mathop{}\!\mathrm{d}_{z} q \|_{op}\\ & + \frac{t^{1+\eta} }{2} \sup_{z \in \tilde{\exp}_x^{-1}(B_{t^\eta}(x))\backslash W } \| \mathop{}\!\mathrm{d}_{z} q\|_{op} \|\mathop{}\!\mathrm{d}^{(2)}_{z} \tilde{f}\|_{op}. \end{align}\] Note that \[C_{x,f,p}(t) \to \frac{q(0_d) }{2} \|\mathop{}\!\mathrm{d}^{(2)}_{0_d} \tilde{f}\|_{op} + \|\mathop{}\!\mathrm{d}_{0_d} \tilde{f} \|_{op} \| \mathop{}\!\mathrm{d}_{0_d} q \|_{op}.\] Then we get \[\begin{align} |II(t)| & \le \frac{q(0_d) \| \mathop{}\!\mathrm{d}_{0_d} \tilde{f}\|_{op} }{\sqrt{t}} \int_{\frac{\tilde{\exp}_x^{-1}(B_{t^\eta}(x))}{\sqrt{t}}\backslash W} \exp\left(-\|\zeta\|^2 \right) \|\zeta\| d\zeta \\ & + C_{x,f,p}(t) \int_{\frac{\tilde{\exp}_x^{-1}(B_{t^\eta}(x))}{\sqrt{t}}\backslash W} \exp\left(-\|\zeta\|^2 \right) \|\zeta\|^2 d\zeta. \end{align}\] It follows from 16 that the two previous integrals both converge to \(0\) as \(t \to 0\). This yields 15 as desired. ◻

4.3 Euclidean calculation↩︎

The previous change of coordinates allows us to make a calculation in the Euclidean space \(\mathbb{R}^d\). This is what we do in this subsection. We provide a general result which might allow for considering union of intersecting manifolds with kinks in a future work.

Let us recall the notations. For any \(y \in \mathbb{R}^d\) we set \(y^{(k)} :=y\otimes \ldots \otimes y \in (\mathbb{R}^d)^{\otimes^k}\). If \(f \in \mathcal{C}^{\infty}(U)\) for some open subset \(U \subset \mathbb{R}^d\), we let \(\mathop{}\!\mathrm{d}^{(k)}_xf\) denote the differential of order \(k\) of \(f\) at \(x\), which we see here as a \(k\)-linear symmetric map from \((\mathbb{R}^d)^{\otimes^k}\) to \(\mathbb{R}\), and we set \[\|\mathop{}\!\mathrm{d}^{(k)}_xf\|_{op} :=\sup_{y \in \mathbb{R}^d\backslash \{0_d\}} \frac{\mathop{}\!\mathrm{d}^{(k)}_xf(y^{(k)})}{|y|^{k}} \, \cdot\] We shall also write \(\mathop{}\!\mathrm{d}_x^{(0)}f(y^{(0)})\) for \(f(x)\), in which case \(\|\mathop{}\!\mathrm{d}^{(0)}_xf\|_{op}= f(x)\). We consider a cone \(\mathcal{C} \subset \mathbb{R}^d\) and a scalar product \(g\) on \(\mathbb{R}^d\), and we set \(S^g \mathcal{C} :=\mathcal{C} \cap \{\xi \in \mathbb{R}^d : g(\xi,\xi)=1\}.\)

Proposition 30. Let \(p \in \mathcal{C}^{2}_{\ge 0}(\mathbb{B}^d)\) and \(f \in \mathcal{C}^{3}(\mathbb{B}^d)\) be such that \[\label{eq:p95f} C_{p,f} :=\left( \max_{0 \le j \le 2 } \sup_{\xi \in \mathbb{B}^d}\|\mathop{}\!\mathrm{d}^{(j)}_{\xi}p \|_{op} + \max_{1 \le i \le 3 } \sup_{\xi \in \mathbb{B}^d}\|\mathop{}\!\mathrm{d}^{(i)}_{\xi}f \|_{op} \right) < +\infty.\qquad{(4)}\] For \(\eta \in (0,1/2)\) and \(t \in (0,1)\), define \[L_t^{\mathcal{C},\eta}f(0_d)= \frac{1}{t} \int_{ \mathbb{B}^d_{t^\eta}\cap \mathcal{C}} e^{-\frac{\|y\|^2}{t}} (f(0_d)-f(y))p(y) \mathop{}\!\mathrm{d}y.\] Then as \(t \downarrow 0\), \[\begin{align} \label{eq:eucl95result} L_t^{\mathcal{C},\eta}f(0_d) & = - \frac{c_d}{\sqrt{t}} p(0_d) \partial_{v_{\mathcal{C}}} f(0_d) - c_{d+1} \bigg( p(0_d) A_\mathcal{C}f(0_d) + [p,f]_\mathcal{C}(0_d)\bigg) + O(\sqrt{t}), \end{align}\qquad{(5)}\] where \[\partial_{v_{\mathcal{C}}}f(0_d) :=d_{0_d} f(v_\mathcal{C}) \quad \text{with} \quad v_{\mathcal{C}} :=\int_{S^g \mathcal{C}} \theta \mathop{}\!\mathrm{d}\sigma (\theta),\] and \[A_{\mathcal{C}}f(0_d) :=\int_{S^g\mathcal{C}} \mathop{}\!\mathrm{d}_{0_d}^{(2)}f \left( \theta^{(2)} \right) \mathop{}\!\mathrm{d}\sigma(\theta), \qquad [p,f]_\mathcal{C}(x) :=\frac{1}{2} \int_{S^g\mathcal{C}} \mathop{}\!\mathrm{d}_{0_d}f(\theta)\mathop{}\!\mathrm{d}_{0_d}p(\theta) \mathop{}\!\mathrm{d}\sigma(\theta).\]

To prove this proposition, we need a simple preliminary lemma. We provide a proof for completeness.

Lemma 14. Let \(g \in \mathcal{C}(\mathbb{R}^d)\) be such that there exists \(m \in \mathbb{N}\) for which \[C_g :=\sup_{z \in \mathbb{R}^d \backslash \{0\}} \frac{|g(z)|}{|z|^m} < +\infty.\] Then for any cone \(\mathcal{C}\) in \(\mathbb{R}^d\) and \(a\in(-1/2,0)\), \[\left| \int_{\mathcal{C}\backslash \mathbb{B}^d_{t^a}} e^{-|z|^2} g(z) \, \mathop{}\!\mathrm{d}z \right| = O ( t^{\frac{1}{2}}) \qquad \text{as t \downarrow 0.}\]

Proof. We have \[\begin{align} \left| \int_{\mathcal{C}\backslash \mathbb{B}^d_{t^a}} e^{-|z|^2} g(z) \, \mathop{}\!\mathrm{d}z \right| & \le \int_{\mathcal{C}\backslash \mathbb{B}^d_{t^a}} e^{-|z|^2} | g(z)| \, \mathop{}\!\mathrm{d}z \le C_g \int_{\mathcal{C}\backslash \mathbb{B}^d_{t^a}} e^{-|z|^2} |z|^m \, \mathop{}\!\mathrm{d}z. \end{align}\] Using polar coordinates first, then the change of variable \(\tau=r^2\), we get \[\begin{align} \int_{\mathcal{C}\backslash \mathbb{B}^d_{t^a}} e^{-|z|^2} |z|^m \, \mathop{}\!\mathrm{d}z & = \mathcal{H}^{d-1}(\mathcal{S}^g\mathcal{C}) \int_{t^{a}}^{+\infty} e^{-r^2} r^{m+d-1} \, \mathop{}\!\mathrm{d}r \\ & = \frac{\mathcal{H}^{d-1}(\mathcal{S}^g\mathcal{C})}{2} \int_{t^{2a}}^{+\infty} e^{-\tau} \tau^{\frac{m+d}{2}-1} \, \mathop{}\!\mathrm{d}\tau \\ & = \frac{\mathcal{H}^{d-1}(\mathcal{S}^g\mathcal{C})}{2} \, \Gamma \left( \frac{m+d}{2} , t^{2a} \right). \end{align}\] By asymptotic property of the incomplete Gamma function (see e.g. [28]) \[\Gamma \left( \frac{m+d}{2} , t^{2a} \right) \underset{t \downarrow 0}{\sim} t^{a(m+d-2)} e^{-t^{2a}}.\] Therefore, for small enough \(t>0\), \[\left| \int_{\mathcal{C}\backslash \mathbb{B}^d_{t^a}} e^{-|z|^2} F(z) \, \mathop{}\!\mathrm{d}z \right| \le C_g \, \mathcal{H}^{d-1}(\mathcal{S}^g\mathcal{C}) \, t^{a(m+d-2)} e^{-t^{2a}}.\] The result follows from the fact that \(t^{a(m+d-2)} e^{-t^{2a}} = O(t^{\frac{1}{2}})\) as \(t \downarrow 0\). ◻

Remark 31. Our proof yields the more precise estimate: \[\left| \int_{\mathcal{C}\backslash \mathbb{B}^d_{t^a}} e^{-|z|^2} g(z) \, \mathop{}\!\mathrm{d}z \right| = O \left( t^{a(m+d-2)} e^{-t^{2a}}\right) \qquad \text{as t \downarrow 0.}\]

We are now in a position to prove Proposition 30.

Proof. Step 1. Let us prove that, as \(t \downarrow 0\), \[\begin{align} \label{eq:step1} L_t^{\mathcal{C},\gamma}f(0_d) & = - t^{-\frac{1}{2}} p(0_d) \mathop{}\!\mathrm{d}_{0_d}f\left( \int_\mathcal{C}e^{-|z|^2} z \mathop{}\!\mathrm{d}z\right) - \int_\mathcal{C}e^{-|z|^2} \mathop{}\!\mathrm{d}_{0_d}f(z)\mathop{}\!\mathrm{d}_{0_d}p(z) \mathop{}\!\mathrm{d}z \nonumber \\ & \phantom{=} - \frac{p(0_d)}{2}\mathop{}\!\mathrm{d}_{0_d}^{(2)} f \left(\int_\mathcal{C}e^{-|z|^2} z^{(2)} \mathop{}\!\mathrm{d}z \right) + O(t^{\frac{1}{2}}). \end{align}\tag{17}\] The Taylor theorem with Laplace remainder implies that for any \(y \in \mathbb{B}^d\) there exist \(\xi_y, \zeta_y \in \mathbb{B}^d\) such that \[f(0_d)-f(y) = - \mathop{}\!\mathrm{d}_{0_d} f (y) - \frac{1}{2} \mathop{}\!\mathrm{d}^{(2)}_{0_d} f (y^{(2)}) - \frac{1}{6} \mathop{}\!\mathrm{d}^{(3)}_{\xi_y} f (y^{(3)}),\] \[p(y) = p(0_d) + \mathop{}\!\mathrm{d}_{0_d} p (y) + \frac{1}{2} \mathop{}\!\mathrm{d}^{(2)}_{\zeta_y} p (y^{(2)}).\] Then \[\begin{align} \label{eq:key1} L_t^{\mathcal{C},\gamma} f(0_d) & = \sum_{i=1}^{3}\sum_{j=0}^{2} I_{t}(i,j) \end{align}\tag{18}\] where for any \(i \in \{1,2,3\}\) and \(j \in \{0,1,2\}\), \[I_{t}(i,j) :=- \frac{1}{i!j!t^{\frac{d}{2}+1}} \int_{\mathcal{C} \cap \mathbb{B}^d_{t^\gamma}} e^{-\frac{|y|^2}{t}} g_{i,j}(y) \mathop{}\!\mathrm{d}y.\] Here we have defined, for any \(y \in \mathbb{B}^d\), \[\begin{array}{ll} g_{i,j}(y) :=\mathop{}\!\mathrm{d}_{0_d}^{(i)} f (y^{(i)}) \mathop{}\!\mathrm{d}_{0_d}^{(j)} p(y^{(j)}) & \text{for (i,j) \in \{1,2\} \times \{0,1\}}, \\ g_{3,j}(y) :=\mathop{}\!\mathrm{d}_{\xi_y}^{(3)} f (y^{(3)}) \mathop{}\!\mathrm{d}_{0_d}^{(j)} p(y^{(j)}) & \text{for j\in \{1,2\}},\\ g_{i,3}(y) :=\mathop{}\!\mathrm{d}_{0_d}^{(i)} f (y^{(i)}) \mathop{}\!\mathrm{d}_{\zeta_y}^{(3)} p(y^{(2)}) & \text{for i \in \{1,2\}},\\ g_{3,2}(y) :=\mathop{}\!\mathrm{d}_{\xi_y}^{(3)} f (y^{(3)}) \mathop{}\!\mathrm{d}_{\zeta_y}^{(2)} p(y^{(2)}). & \end{array}\] On one hand, if \(i=3\) or \(j=2\), we know from ?? that for any \(y \in \mathcal{C}\), \[|g_{i,j}(y)| \le C_{p,f}|y|^{i+j}\] hence \[\begin{align} |I_{t}(i,j)| & \le\frac{1}{t^{\frac{d}{2}+1}} \int_{\mathcal{C} \cap \mathbb{B}^d_{t^\gamma}} e^{-\frac{|y|^2}{t}} |g_{i,j}(y)| \mathop{}\!\mathrm{d}y \le \frac{C_{p,f}}{t^{\frac{d}{2}+1}} \int_{\mathcal{C} \cap \mathbb{B}^d_{t^\gamma}} e^{-\frac{|y|^2}{t}} |y|^{i+j} \mathop{}\!\mathrm{d}y \\ & = C_{p,f} t^{\frac{i+j}{2}-1} \int_{\mathcal{C} \cap \mathbb{B}^d_{t^{\gamma-1/2}}} e^{-|z|^2} |z|^{i+j} \mathop{}\!\mathrm{d}z \le C_{p,f} C_{\mathcal{C},i+j} t^{\frac{i+j}{2}-1} \end{align}\] where we have used the change of variable \(z= y/\sqrt{t}\) and we have defined \[C_{\mathcal{C},k} :=\int_{\mathcal{C}} e^{-|z|^2} |z|^{k} \mathop{}\!\mathrm{d}z\] for any integer \(k\). Since \(i=3\) or \(j=2\) implies \(i+j\ge 3\), we have \((i+j)/2-1 \ge 3/2-1 = 1/2\), so that we eventually get \[\begin{align} \label{eq:key2} \sum_{(i,j) \notin \{1,2\} \times \{0,1\}} I_{t}(i,j) = O(t^{1/2}). \end{align}\tag{19}\]

On the other hand, if \((i,j) \in \{1,2\} \times \{0,1\}\), then for any \(z \in \mathbb{R}^d\) and \(t>0\), \[g_{i,j}(\sqrt{t}z) = t^{\frac{i+j}{2}} g_{i,j}(z) \quad \text{and} \quad g_{i,j}(z) \le C_{p,f} |z|^{i+j}.\] Then the change of variable \(z= y/\sqrt{t}\) yields that \[\begin{align} I_{t}(i,j) & = \frac{1}{t} \int_{\mathcal{C} \cap \mathbb{B}^d_{t^{\gamma-1/2}}} e^{-|z|^2} g_{i,j}(\sqrt{t}z) \mathop{}\!\mathrm{d}z = t^{\frac{i+j}{2}-1} \int_{\mathcal{C} \cap \mathbb{B}^d_{t^{\gamma-1/2}}} e^{-|z|^2} g_{i,j}(z) \mathop{}\!\mathrm{d}z. \end{align}\] By Lebesgue dominated convergence theorem, \[\int_{\mathcal{C} \cap \mathbb{B}^d_{t^{\gamma-1/2}}} e^{-|z|^2} g_{i,j}(z) \mathop{}\!\mathrm{d}z \to \int_{\mathcal{C}} e^{-|z|^2} g_{i,j}(z) \mathop{}\!\mathrm{d}z.\] Moreover, by Lemma 14, \[\begin{align} \left| \int_{\mathcal{C} \backslash \mathbb{B}^d_{t^{\gamma-1/2}}} e^{-|z|^2} g_{i,j}(z) \mathop{}\!\mathrm{d}z \right| = O( t^{\frac{1}{2}} ). \end{align}\] Then \[I_{t}(i,j) = t^{\frac{i+j}{2}-1} \int_{\mathcal{C} } e^{-|z|^2} g_{i,j}(z) \mathop{}\!\mathrm{d}z + O( t^{\frac{1}{2}} ).\] As a consequence, \[\begin{align} \label{eq:key3} I_{t}(1,0) & = t^{-\frac{1}{2}} \int_{\mathcal{C} } e^{-|z|^2} g_{1,0}(z) \mathop{}\!\mathrm{d}z + O( t^{\frac{1}{2}} ), \end{align}\tag{20}\] and \[\begin{align} \label{eq:key4} I_{t}(1,1) + I_{t}(2,0) & = \int_{\mathcal{C} } e^{-|z|^2} (g_{1,1}(z)+g_{2,0}(z)) \mathop{}\!\mathrm{d}z + O( t^{\frac{1}{2}}). \end{align}\tag{21}\] Now \[\begin{align} \int_{\mathcal{C} } e^{-|z|^2} g_{1,0}(z) \mathop{}\!\mathrm{d}z & = p(0_d) \int_\mathcal{C}e^{-|z|^2} \mathop{}\!\mathrm{d}_{0_d}f(z) \mathop{}\!\mathrm{d}z \\ \int_{\mathcal{C} } e^{-|z|^2} g_{1,1}(z) \mathop{}\!\mathrm{d}z & = \int_\mathcal{C}e^{-|z|^2} \mathop{}\!\mathrm{d}_{0_d}f(z)\mathop{}\!\mathrm{d}_{0_d}p(z) \mathop{}\!\mathrm{d}z \\ \int_{\mathcal{C} } e^{-|z|^2} g_{2,0}(z) \mathop{}\!\mathrm{d}z & = p(0_d)\int_\mathcal{C}e^{-|z|^2} \mathop{}\!\mathrm{d}_{0_d}^{(2)}f(z^{(2)}) \mathop{}\!\mathrm{d}z . \end{align}\] This implies the desired result.

Step 2. Using polar coordinates, we can write \[\begin{align} \int_\mathcal{C}e^{-|z|^2} z \mathop{}\!\mathrm{d}z = \int_0^{+\infty} \int_{S^g\mathcal{C}} e^{-\rho^2} \rho \theta \rho^{d-1} \mathop{}\!\mathrm{d}\rho \mathop{}\!\mathrm{d}\sigma(\theta) & = \left(\int_0^{+\infty} e^{-\rho^2} \rho^{d} \mathop{}\!\mathrm{d}\rho \right)\underbrace{\int_{S^g\mathcal{C}} \theta \mathop{}\!\mathrm{d}\sigma(\theta)}_{=v_\mathcal{C}}. \end{align}\] The change of variable \(\tau = \rho^2\) yields that \[\int_0^{+\infty} e^{-\rho^2} \rho^{d} \mathop{}\!\mathrm{d}\rho = \Gamma\left(\frac{d+1}{2}\right) = c_{d}.\] In the end, we get \[\int_\mathcal{C}e^{-|z|^2} z \mathop{}\!\mathrm{d}z = c_{d} v_\mathcal{C}.\] In a similar way, \[\begin{align} \int_\mathcal{C}e^{-|z|^2} z^{(2)} \mathop{}\!\mathrm{d}z & = \int_0^{+\infty} \int_{S^g\mathcal{C}} e^{-\rho^2} (\rho \theta)^{(2)} \rho^{d-1} \mathop{}\!\mathrm{d}\rho \mathop{}\!\mathrm{d}\sigma(\theta) = \int_0^{+\infty} \int_{S^g\mathcal{C}} e^{-\rho^2} \theta^{(2)} \rho^{d+1} \mathop{}\!\mathrm{d}\rho \mathop{}\!\mathrm{d}\sigma(\theta)\\ & = \left( \int_0^{+\infty} e^{-\rho^2} \rho^{d+1} \mathop{}\!\mathrm{d}\rho \right)\underbrace{\int_{S^g\mathcal{C}} \theta^{(2)} \mathop{}\!\mathrm{d}\sigma(\theta)}_{=v_\mathcal{C}^{(2)}} = c_{d+1} v_\mathcal{C}^{(2)} \end{align}\] and \[\begin{align} \int_\mathcal{C}e^{-|z|^2} \mathop{}\!\mathrm{d}_{0_d}f(z)\mathop{}\!\mathrm{d}_{0_d}p(z) \mathop{}\!\mathrm{d}z & = \int_0^{+\infty} \int_{S^g\mathcal{C}} e^{-\rho^2} \mathop{}\!\mathrm{d}_{0_d}f(\rho \theta)\mathop{}\!\mathrm{d}_{0_d}p(\rho \theta) \rho^{d-1} \mathop{}\!\mathrm{d}\rho \mathop{}\!\mathrm{d}\sigma(\theta)\\ & = \int_0^{+\infty} \int_{S^g\mathcal{C}} e^{-\rho^2} \mathop{}\!\mathrm{d}_{0_d}f( \theta)\mathop{}\!\mathrm{d}_{0_d}p( \theta) \rho^{d+1} \mathop{}\!\mathrm{d}\rho \mathop{}\!\mathrm{d}\sigma(\theta)\\ & = \left( \int_0^{+\infty} e^{-\rho^2} \rho^{d+1} \mathop{}\!\mathrm{d}\rho \right) \int_{S^g\mathcal{C}} \mathop{}\!\mathrm{d}_{0_d}f( \theta)\mathop{}\!\mathrm{d}_{0_d}p( \theta) \mathop{}\!\mathrm{d}\sigma(\theta)\\ & = c_{d+1} \int_{S^g\mathcal{C}} \mathop{}\!\mathrm{d}_{0_d}f( \theta)\mathop{}\!\mathrm{d}_{0_d}p( \theta) \mathop{}\!\mathrm{d}\sigma(\theta). \end{align}\] Combined with 17 , these yield ?? as desired. ◻

Remark 32. The previous proof may be easily modified to get the following: for any integer \(N \ge 2\), let \(p \in \mathcal{C}^{N}_{\ge 0}(\mathbb{B}^d)\) and \(f \in \mathcal{C}^{N+1}(\mathbb{B}^d)\) be such that \[\max_{0 \le j \le N} \sup_{\xi \in \mathbb{B}^d}\|\mathop{}\!\mathrm{d}^{(j)}_{\xi}p \|_{op} + \max_{1 \le i \le N+1 } \sup_{\xi \in \mathbb{B}^d}\|\mathop{}\!\mathrm{d}^{(i)}_{\xi}f \|_{op} < +\infty.\] Then for any cone \(\mathcal{C}\) in \(\mathbb{R}^d\) and \(\eta \in (0,1/2)\), \[L_t^{\mathcal{C},\eta}f(0_d) = \sum_{i=1}^{N+1} \sum_{j=0}^N \frac{t^{\frac{i+j}{2}-1}c_{d+i+j-1}}{i!j!} \int_{S^g\mathcal{C}} \mathop{}\!\mathrm{d}_{0_d}^{(i)} f (\theta^{(i)}) \mathop{}\!\mathrm{d}_{0_d}^{(j)} p(\theta^{(j)}) \mathop{}\!\mathrm{d}\sigma(\theta) + O(t^{\frac{N-1}{2}}).\]

Remark 33. If \(p(0_d)=0\), then \(L_t^{\mathcal{C},\eta} f(0_d)\) always converges to \(\displaystyle \int_\mathcal{C}e^{-|z|^2} \mathop{}\!\mathrm{d}_{0_d}f(z)\mathop{}\!\mathrm{d}_{0_d}p(z) \mathop{}\!\mathrm{d}z.\)

4.4 Conclusion↩︎

Let us now explain how to reach the conclusion of Theorem 1. From Lemma 13 and Proposition 29, we get \[L_tf(x) = \frac{1}{t} \int_{\mathbb{B}^d_{t^{\eta}}\cap \tilde{F}_{0_d}\Omega} e^{-\frac{\|\xi\|^2}{t}} (\tilde{f}(0_d) - \tilde{f}(\xi))q(\xi)\mathop{}\!\mathrm{d}\xi + o\left( \frac{1}{\sqrt{t}}\right) + O(t^{-d/2-1}e^{-t^{2 \eta -1}}) \qquad \text{as t\to 0.}\] From Proposition 8 and Proposition 9, we can replace \(\tilde{F}_{0_d}\Omega\) by \(I_{0_d}\bar{\Omega}\) in the previous domain of integration. Then we apply Proposition 30 to obtain that \[\begin{align} \frac{1}{t} \int_{\mathbb{B}^d_{t^{\eta}}\cap I_{0_d}\bar{\Omega}} e^{-\frac{\|\xi\|^2}{t}} (\tilde{f}(0_d) - \tilde{f}(\xi))q(\xi)\mathop{}\!\mathrm{d}\xi \end{align}\] equals \[- \frac{c_d}{\sqrt{t}} q(0_d) \partial_{v_{\mathcal{I}_{0_d}\bar{\Omega}}} \tilde{f}(0_d) - c_{d+1} \bigg(q(0_d) A_{I_{0_d}\bar{\Omega}} \tilde{f}(0_d) + [q,\tilde{f}]_{I_{0_d}\bar{\Omega}}(0_d)\bigg) + O(\sqrt{t})\] as \(t \downarrow 0\). But \[q(0_d) = p(\tilde{\exp}_x(0_x)) \sqrt{\det g(0_d)} = p(x),\] \[\begin{align} \partial_{v_{I_{0_d}\bar{\Omega}}} \tilde{f}(0_d) = d_{0_d}\tilde{f}(v_{I_{0_d}\bar{\Omega}}) & = d_xf \left( d_{0_d}\tilde{\exp}_x\left( \int_{S^g I_{0_d}\bar{\Omega}} \theta \mathop{}\!\mathrm{d}\sigma (\theta)\right)\right)\\ & = d_xf \left( \int_{S^g I_{0_d}\bar{\Omega}} d_{0_d}\tilde{\exp}_x(\theta) \mathop{}\!\mathrm{d}\sigma (\theta)\right)\\ & = d_xf \left( \int_{S^gI_xM} \theta \mathop{}\!\mathrm{d}\sigma(\theta)\right) = d_xf(v_g(x)) = \partial_{v_g(x)}f(x), \end{align}\] where we have used Lemma 8 to perform the change of variable which provides us with the last line. Likewise, we obtain \[A_\mathcal{C}\tilde{f}(0_d) = A_gf(x),\] \[\begin{align} [q,\tilde{f}]_{I_{0_d}\bar{\Omega}}(x) = [p,f]_g(x). \end{align}\] This concludes the proof of Theorem 1.

5 Concentration estimates↩︎

The goal of this section is to establish Theorem 2.

5.1 \(\alpha\)-Subexponential Random Variables↩︎

We begin by defining the class of \(\alpha\)-subexponential random variables introduced by Götze, Sambale and Sinulis in [29].

Definition 29. Let \(Z\) be a real-valued random variable. We say that \(Z\) is \(\alpha\)-subexponential for some \(\alpha > 0\) if there exist constants \(K, C > 0\) such that : \[\mathbb{P}[|Z|\ge\epsilon] \le K\exp(-C{\epsilon}^{\alpha}) \qquad \forall \epsilon \ge 0.\]

Remark 34. When \(\alpha = 1\), the previous class coincides with classical subexponential random variables, while for \(\alpha = 2\), it coincides with classical subgaussian random variables, see e.g. [30] for more details about these cases. Note also that \(\alpha > 2\) implies subgaussian (i.e. \(\alpha = 2\)). For values \(\alpha < 1\), heavier tails are permitted, as in the case for Weibull random variables, for instance.

The following lemma is useful for our purposes. We provide a quick proof for completeness.

Lemma 15. Let \(Z\) be \(\alpha\)-subexponential. Let \(W\) be an a.s. bounded real-valued random variable, and \(z\in \mathbb{R}\). Then both the product random variable \(WZ\) and the translated random variable \(Z-z\) are \(\alpha\)-subexponential.

Proof. Let \(m>0\) be such that \(|W|\le m\) a.s. Then for any \(\epsilon \ge 0\), \[\mathbb{P}(|WZ| \ge \epsilon) = \mathbb{P}(|Z| \ge \epsilon/|W|) \le \mathbb{P}(|Z| \ge \epsilon/m) \le K\exp(-C{(\epsilon/m)}^{\alpha}).\] This shows that \(WZ\) is \(\alpha\)-subexponential. As for \(Z-z\), note that \(|Z|+|z| \ge |Z-z|\) implies the first inequality in \[\mathbb{P}(|Z-z| \ge \epsilon ) \le \mathbb{P}(|Z|+|z| \ge \epsilon ) = \mathbb{P}(|Z| \ge \epsilon - |z|) \le K\exp(-C{(\epsilon-|z|)}^{\alpha}) \le K\exp(-(C/2^\alpha){\epsilon}^{\alpha}).\] ◻

The next statement is taken from [29]. It is a concentration result for sample means of \(\alpha\)-subexponential random variables.

Theorem 35. Let \(Z_1, \dots, Z_n \sim Z\) be iid and mean-zero real valued random variables that are \(\alpha\)-subexponential for some \(\alpha \in (0,2]\). Then there exist constants \(c, C> 0\) such that for every \(\epsilon \ge 0\),

  1. if \(\alpha \in (0,1]\), then \[\mathbb{P}\left(\left|\frac{1}{n}\sum_{i=1}^n Z_i\right| \ge \epsilon\right) \le C \exp\left( -c n^{\alpha} \epsilon^{\alpha} \right),\]

  2. if \(\alpha \in (1,2]\), then \[\mathbb{P}\left(\left|\frac{1}{n}\sum_{i=1}^n Z_i\right| \ge \epsilon\right) \le C \exp\left( -c n^{\alpha/2} \epsilon^{\alpha} \right).\]

Remark 36. Note that (1) matches up with the subexponential tail decay given by Bernstein’s inequality in case \(\alpha=1,\) while (2) coincides with the subgaussian tail decay of Hoeffding’s inequality when \(\alpha=2\). We refer to [30] for a nice account on these two classical inequalities.

5.2 Convergence in probability↩︎

In this section, we fall back onto our original setup, i.e. we consider \(\{X_i\}_{\mathbb{N}} \sim X\), a sequence of iid random variables taking values in a Riemannian manifold with kinks \((M,g)\). Assume the law \(\mathbb{P}_X\) has a \(\mathcal{C}^2(M)\) density function \(p,\) and consider a function \(f \in \mathcal{C}^3(M)\) and a number \(t>0\).

We are firstly interested in finding out the concentration of \(L_{n,t}f(x)\) around its expected value \(L_tf(x)\), for \(x \in M\) being either an interior point or an LCDD border point. This is given by the following result.

Proposition 37. Assume \(f(X)\) is \(\alpha\)-subexponential for some \(\alpha \in (0,2]\). Then there exist constants \(C_1, C_2 > 0\) such that for all \(\epsilon \ge 0\),

  1. if \(\alpha \in (0,1]\), then \[\mathbb{P} \left( \left| L_{n,t} f(x) - L_t f(x) \right| \ge \epsilon \right) \le C_1 \exp \left( - C_2 \left( n\, t^{\frac{d}{2} + 1} \epsilon \right)^{\alpha} \right),\]

  2. if \(\alpha \in (1,2]\), then \[\mathbb{P} \left( \left| L_{n,t} f(x) - L_t f(x) \right| \ge \epsilon \right) \le C_1 \exp \left( - C_2 \left( \sqrt{n}\, t^{\frac{d}{2} + 1} \epsilon \right)^{\alpha} \right).\]

Moreover, the constants \(C_1\) and \(C_2\) depend only on \(X\), \(f\), and \(\alpha\).

Proof. Let us provide details for the case \(\alpha \in (0,1]\) only, the other one being analogous. Consider the zero-mean real-valued random variables \[Z_i(t) = e^{-d^2(x,X_i)/t} (f(x)-f(X_i)) - t^{d/2+1}L_tf(x), \qquad i \ge 1.\] We know from Lemma 15 that these are \(\alpha\)-subexponential, hence we can apply Theorem 35 : for any \(\epsilon \ge 0\), \[\mathbb{P} \left( \left| L_{n,t} f(x) - L_t f(x) \right| \ge \epsilon \right) = \mathbb{P} \left( \frac{1}{ t^{d/2+1}} \left| \frac{1}{n}\sum_{i=1}^n Z_i(t) \right| \ge \epsilon \right) \le C_1 \exp \left( - C_2 \left( n\, t^{\frac{d}{2} + 1} \epsilon \right)^{\alpha} \right).\] ◻

Let us recall the operator \[\mathcal{L}_tf(x) :=-\frac{c_d}{\sqrt{t}}\left( p(x) \, \partial_{v(x)}f(x) + o(1)\right) - c_{d+1} \bigg( p(x) A_gf(x) + [p,f]_g(x) \bigg)\] identified in Theorem 1. Since the latter differ from \(L_tf(x)\) of a \(o(\sqrt{t})\) term, we immediately deduce the following from the previous proposition.

Corollary 1. Assume \(f(X)\) is \(\alpha\)-subexponential for some \(\alpha >0\). Then there exist constants \(C_1, C_2 > 0\) such that for all \(\epsilon \ge 0\), we have :

  1. if \(\alpha \in (0,1]\), then \[\mathbb{P} \left( \left| L_{n,t} f(x) - \mathcal{L}_t f(x) \right| \ge \epsilon \right) \le C_1 \exp \left( - C_2 \left( n\, t^{\frac{d}{2} + 1} \epsilon \right)^{\alpha} \right),\]

  2. if \(\alpha \in (1,2]\), then \[\mathbb{P} \left( \left| L_{n,t} f(x) - \mathcal{L}_t f(x) \right| \ge \epsilon \right) \le C_1 \exp \left( - C_2 \left( \sqrt{n}\, t^{\frac{d}{2} + 1} \epsilon \right)^{\alpha} \right).\]

We are now in a position to prove (1) in Theorem 2. Consider \(t_n \to 0\) such that \(\sqrt{n}\, t_n^{\frac{d}{2} + 1} \to \infty\) as \(n \to \infty\). Then \[C_1 \exp \left( - C_2 \left( n\, t^{\frac{d}{2} + 1} \epsilon \right)^{\alpha} \right)\to 0 \qquad \text{and} \qquad C_1 \exp \left( - C_2 \left( \sqrt{n}\, t^{\frac{d}{2} + 1} \epsilon \right)^{\alpha} \right) \to 0 \qquad \text{as n \to +\infty}.\] This shows that \[\mathbb{P} \left( \left| L_{n,t_n} f(x) - \mathcal{L}_{t_n} f(x) \right| \ge \epsilon \right) \to 0 \qquad \text{as n \to +\infty}.\]

5.3 Almost sure convergence↩︎

Let us now prove (2) in Theorem 2. To this aim, recall that a sequence of real-valued random variables \(\{Z_n\}\) completely converges to another real-valued random variable \(Z\) if for any \(\epsilon>0\), \[\sum_{n} \mathbb{P}(|Z_n - Z| \ge \epsilon) < +\infty.\] It is easily seen from the first Borel–Cantelli lemma that complete convergence implies almost sure convergence. We shall need an elementary lemma.

Lemma 16. Let \(\{b_n\}\) be a sequence of positive real numbers such that \[\frac{b_n}{\ln(n)} \stackrel{n\to +\infty}{\longrightarrow} +\infty \quad \text{and} \quad \sum_{n\ge 1} e^{-b_n} < +\infty.\] Then for any \(\delta \in (0,1)\), \[\sum_{n\ge 1} e^{-\delta b_n} < +\infty\]

Proof. Take \(\delta \in (0,1)\). For any \(c>0\) there exists \(N \in \mathbb{N}\) such that for any \(n \ge N\), \[\frac{b_n}{\ln(n)} \ge c.\] For these integers \(n\), \[e^{-\delta b_n} \le e^{-\delta c \ln(n)} = \frac{1}{n^{\delta c}} \, \cdot\] Therefore, if \(c > 1/\delta\), \[\sum e^{-\delta b_n} \le \sum \frac{1}{n^{\delta c}} < +\infty.\] ◻

We can now prove (2) in Theorem 2. Consider \(t_n \to 0\) such that \(\left(\sqrt{n}\, t_n^{\frac{d}{2} + 1} \right)^{\alpha}/\ln(n)\to \infty\) as \(n \to \infty\). For any \(\epsilon >0\), \[\begin{align} \sum_{n} \mathbb{P} \left( \left| L_{n,t_n} f(x) - \mathcal{L}_{t_n} f(x) \right| \ge \epsilon \right) & \le C_1 \sum_{n} \exp \left( - C_2 \left( n\, t^{\frac{d}{2} + 1} \epsilon \right)^{\alpha} \right)\\ & = C_1 \sum_{n} \exp \left( - b_n\epsilon^\alpha \right) \end{align}\] with \(b_n :=C_2 \left( n\, t^{\frac{d}{2} + 1} \right)^{\alpha}\). The assumption on \(\{t_n\}\) implies that \(b_n/\ln(n) \to \infty\) as \(n\to \infty\), hence Lemma 16 yields that \[\sum_{n} \mathbb{P} \left( \left| L_{n,t_n} f(x) - \mathcal{L}_{t_n} f(x) \right| \ge \epsilon \right) \le \sum_{n} \exp \left( - b_n\epsilon^\alpha \right) < +\infty.\] Thus \(|L_{n,t_n}f(x) - \mathcal{L}_{t_n}f(x)|\) completely converges to \(0\), thus it converges almost surely.

6 Numerical simulations↩︎

6.1 Experiments for a 3D ball↩︎

We test the value of \(L_{n,t}f\) at these points, when \(n:=10^8,f(x,y,z):=x+y+z,\) (whose Laplacian is zero) and the kernel bandwidths \(t:=t_n,\) varies logarithmically from \(0.05\) down to \(0.01\) (20 values). Note that Theorem 1.2, part (1) gives us a condition of the convergence in probability of \(L_{n,t_n}\) for \(n\to \infty, t_n\to 0, \sqrt{n}t_n^{\frac{d}{2}+1}\to \infty.\) In our experiments, the values of \(\sqrt{n}t_n^{\frac{d}{2}+1}\) vary between \(\sqrt{10^8}\times (0.01)^{\frac{3}{2}+1}=0.01\) and \(\sqrt{10^8}\times (0.05)^{\frac{3}{2}+1}=5.56.\) These values are not large, but one can already see that the behavior of \(\sqrt{t_n}L_{n,{t_n}}\) matches up with the asymptotics provided by our theory, and the behavior remains consistent as we decrease \(t:=t_n\) in our range.
The two tables below show us how the graph Laplace operator \(L_{n,t}f\) and its expectation \(L_tf\) behave asymptotically. Below ‘int’ corresponds to an interior point (the origin of the ball), and ‘bd’ corresponds to a boundary point (\((1,0)\)), and scaled Laplacians mean \(\sqrt{t}L_{n,t}\) for discrete and \(\sqrt{t}L_t\) for continuous graph Laplacians.

Figure 5: image.

Figure 6: image.

Below are the plots for the above tables:

a
b

Figure 7: Comparison of graph Laplacians on a 3D ball. a — Unscaled graph Laplacian – 3D ball, b — Scaled graph Laplacian – 3D ball

6.2 Experiments for a 3D cube↩︎

We experiment with the asymptotic behavior of the graph Laplacian for a uniform random sample on the unit 3D hypercube \([0,1]^3\) (so \(d:=3\) here) \(abcdefgh,\) at the following points \(I,F,E,V\) below. We test the value of \(L_{n,t}f\) at these points, when \(n:=10^8,f(x,y,z):=x+y+z,\) (whose Laplacian is zero) and the kernel bandwidths \(t:=t_n,\) varies logarithmically from \(0.05\) down to \(0.01\) (20 values). Note that Theorem 1.2, part (1) gives us a condition of the convergence in probability of \(L_{n,t_n}\) for \(n\to \infty, t_n\to 0, \sqrt{n}t_n^{\frac{d}{2}+1}\to \infty.\) In our experiments, the values of \(\sqrt{n}t_n^{\frac{d}{2}+1}\) vary between \(\sqrt{10^8}\times (0.01)^{\frac{3}{2}+1}=0.01\) and \(\sqrt{10^8}\times (0.05)^{\frac{3}{2}+1}=5.56.\) These values are not large, but one can already see that the behavior of \(\sqrt{t_n}L_{n,{t_n}}\) matches up with the asymptotics provided by our theory, and the behavior remains consistent as we decrease \(t:=t_n\) in our range.

  • Interior point : \[I \;=\;\frac{a+b+c+\dots+h}{8}.\]

  • Face midpoint : \[F \;=\;\frac{a+v+d+e}{4}.\]

  • Edge midpoint : \[E \;=\;\frac{a+b}{2}.\]

  • Vertex : \[V = a.\]

Below are the tables and plots giving us the numerical results and the plots.

0.48

\(t\) \(L_{n,t}\) \(L_t\) \(\sqrt t\,L_{n,t}\) \(\sqrt t\,L_t\)
5.000e-02 -4.708e-16 -4.708e-16 -1.053e-16 -1.053e-16
3.737e-02 -3.290e-15 -3.290e-15 -6.360e-16 -6.360e-16
2.474e-02 -6.425e-15 -6.425e-15 -1.010e-15 -1.010e-15
1.421e-02 -1.760e-14 -1.760e-14 -2.098e-15 -2.098e-15
1.000e-02 -2.116e-14 -2.116e-14 -2.116e-15 -2.116e-15

0.48

\(t\) \(L_{n,t}\) \(L_t\) \(\sqrt t\,L_{n,t}\) \(\sqrt t\,L_t\)
5.000e-02 7.002e+00 7.016e+00 1.566e+00 1.569e+00
3.737e-02 8.120e+00 8.142e+00 1.570e+00 1.574e+00
2.474e-02 9.985e+00 1.002e+01 1.570e+00 1.577e+00
1.421e-02 1.318e+01 1.326e+01 1.571e+00 1.581e+00
1.000e-02 1.573e+01 1.586e+01 1.573e+00 1.586e+00

0.48

\(t\) \(L_{n,t}\) \(L_t\) \(\sqrt t\,L_{n,t}\) \(\sqrt t\,L_t\)
5.000e-02 -7.012e+00 -7.027e+00 -1.568e+00 -1.571e+00
3.737e-02 -8.121e+00 -8.144e+00 -1.570e+00 -1.574e+00
2.474e-02 -9.984e+00 -1.002e+01 -1.570e+00 -1.577e+00
1.421e-02 -1.317e+01 -1.326e+01 -1.570e+00 -1.581e+00
1.000e-02 -1.570e+01 -1.586e+01 -1.570e+00 -1.586e+00

0.48

Table 1: Numerical values of the graph Laplacian operators at various points.
\(t\) \(L_{n,t}\) \(L_t\) \(\sqrt t\,L_{n,t}\) \(\sqrt t\,L_t\)
5.000e-02 -5.264e+00 -5.278e+00 -1.177e+00 -1.180e+00
3.737e-02 -6.088e+00 -6.110e+00 -1.177e+00 -1.181e+00
2.474e-02 -7.450e+00 -7.519e+00 -1.177e+00 -1.183e+00
1.421e-02 -9.873e+00 -9.948e+00 -1.177e+00 -1.186e+00
1.000e-02 -1.177e+01 -1.189e+01 -1.177e+00 -1.189e+00
a
b
c
d

Figure 8: For each evaluation point we show discrete \(L_{n,t}f\) (blue \(\bullet\)), continuum \(L_tf\) (orange \(\blacksquare\)), scaled \(\sqrt t\,L_{n,t}f\) (green \(\blacklozenge\)), and scaled \(\sqrt t\,L_tf\) (red \(\times\)); \(t\) is on a log scale.. a — Interior, b — Face midpoint, c — Edge midpoint, d — Vertex

References↩︎

[1]
Zhen Li, Zuoqiang Shi, and Jian Sun. Point integral method for solving poisson-type equations on manifolds from point clouds with convergence guarantees. Communications in Computational Physics, 22(1):228–258, 2017.
[2]
John Harlim, Daniel Sanz-Alonso, and Ruiyi Yang. Kernel methods for Bayesian elliptic inverse problems on manifolds. SIAM/ASA Journal on Uncertainty Quantification, 8(4):1414–1445, 2020.
[3]
Qile Yan, Shixiao Willing Jiang, and John Harlim. Spectral methods for solving elliptic PDEs on unknown manifolds. Journal of Computational Physics, 486:112132, 2023.
[4]
Anran Jiao, Qile Yan, Jhn Harlim, and Lu Lu. Solving forward and inverse PDE problems on unknown manifolds via physics-informed neural operators. arXiv preprint: 2407.05477, 2024.
[5]
Mikhail Belkin and Partha Niyogi. Laplacian eigenmaps for dimensionality reduction and data representation. Neural Computation, 15(6):1373–1396, 2003.
[6]
Mikhail Belkin and Partha Niyogi. Towards a theoretical foundation for Laplacian-based manifold methods. In Proceedings of the 18th Annual Conference on Learning Theory (COLT), pages 486–500, 2005.
[7]
Nicolas Garcia Trillos and Dejan Slepčev. Error estimates for spectral convergence of the graph Laplacian on random geometric graphs towards the Laplace–Beltrami operator. Foundations of Computational Mathematics, 19:197–230, 2019.
[8]
Ronald R. Coifman and Stephane Lafon. Diffusion maps. Applied and Computational Harmonic Analysis, 21(1):5–30, 2006.
[9]
Amit Singer and Hau-Tieng Wu. Vector diffusion maps and the connection Laplacian. Comm. Pure Appl. Math., 65(8):1067–1144, 2012.
[10]
Matthias Hein, Jean-Yves Audibert, and Ulrike von Luxburg. From graphs to manifolds: Weak and strong pointwise consistency of graph Laplacians. In Proceedings of the 18th Annual Conference on Learning Theory (COLT), 2005.
[11]
Mikhail Belkin, Qichao Que, Yusu Wang, and Xueyuan Zhou. Toward understanding complex spaces: Graph Laplacians on manifolds with singularities and boundaries. In Conference on learning theory, pages 36–1. JMLR Workshop and Conference Proceedings, 2012.
[12]
Martin Andersson and Benny Avelin. Exploring singularities in data with the graph Laplacian: An explicit approach. Journal of Computational Mathematics and Data Science, page 100113, 2025.
[13]
Jean Cerf. Topologie de certains espaces de plongements. Bulletin de la Société Mathématique de France, 89:227–380, 1961.
[14]
Adrien Douady. Variétés à bord anguleux et voisinages tubulaires. Séminaire Henri Cartan, 14:1–11, 1961.
[15]
Adrien Douady and Letizia Hérault. Arrondissement des variétés à coins-appendice à" corners and arithmetic groups". Commentarii Mathematici Helvetici, 48:484–491, 1973.
[16]
Peter Michor. Manifolds of differentiable maps. In Differential Topology, pages 117–127. Springer, 1980.
[17]
Richard B. Melrose. Differential analysis on manifolds with corners, 1996.
[18]
Dominic Joyce. On manifolds with corners. In S. Janeczko, J. Li, and D.H. Phong, editors, Advances in Geometric Analysis, volume 21 of Advanced Lectures in Mathematics, pages 225–258. International Press, Boston, 2012. Also available on the Web as arXiv:0910.3518.
[19]
Bernard Akwei, Luke Rogers, and Alexander Teplyaev. Distributional convergence of the empirical Laplacians with integral kernels on domains with boundaries. arXiv preprint: 2503.05633, 2025.
[20]
Manuel Dias and David Tewodrose. Spectral properties of symmetrized AMV operators. arXiv preprint: 2411.10202, 2024.
[21]
Lawrence C Evans. Partial differential equations, volume 19. American Mathematical Society, 2022.
[22]
Ralph T. Rockafellar and Roger J-B Wets. Variational analysis, volume 317. Springer Science & Business Media, 2009.
[23]
Ralph Abraham, Jerrold E Marsden, and Tudor Ratiu. Manifolds, tensor analysis, and applications, volume 75. Springer Science & Business Media, 2012.
[24]
Hassler Whitney. Differentiable functions defined in closed sets. I. Transactions of the American Mathematical Society, 36(2):369–387, 1934.
[25]
Elias M Stein. Singular integrals and differentiability properties of functions. Number 30. Princeton university press, 1970.
[26]
Robert T. Seeley. Extension of \(c^\infty\) functions defined in a half space. Proceedings of the American Mathematical Society, 15(4):625–626, 1964.
[27]
John Lee. Introduction to Smooth Manifolds, volume 218. Springer Science & Business Media, 2012.
[28]
Iosif Pinelis. Exact lower and upper bounds on the incomplete gamma function. Mathematical Inequalities and Applications, 23(4):1261, 2020.
[29]
Friedrich Götze, Holger Sambale, and Arthur Sinulis. Concentration inequalities for polynomials in alpha-sub-exponential random variables. Electronic Journal of Probability, 26, 2021.
[30]
Roman Vershynin. High-dimensional probability: An introduction with applications in data science, volume 47. Cambridge university press, 2018.