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

On the Mean First Arrival Time of Brownian Particles on Riemannian Manifolds

M. Nursultanov School of Mathematics and Statistics, University of Sydney [email protected] L. Tzou School of Mathematics and Statistics, University of Sydney [email protected]  and  J.C. Tzou Department of Mathematics and Statistics, Macquarie University [email protected]
Abstract.

We use geometric microlocal methods to compute an asymptotic expansion of mean first arrival time for Brownian particles on Riemannian manifolds. This approach provides a robust way to treat this problem, which has thus far been limited to very special geometries. This paper can be seen as the Riemannian 3-manifold version of the planar result of [1] and thus enable us to see the full effect of the local extrinsic boundary geometry on the mean arrival time of the Brownian particles. Our approach also connects this question to some of the recent progress on boundary rigidity and integral geometry [23, 20].

2010 Mathematics Subject Classification:
Primary: 58J65 Secondary: 60J65, 58G15, 92C37
M. Nursultanov and L. Tzou are partially supported by ARC DP190103302 and ARC DP190103451 during this work. We thank Ben Goldys for the helpful discussion.

1. Introduction

Let (M,g,M)(M,g,\partial M) be a compact connected orientable Riemannian manifold with non-empty smooth boundary and without loss of generality we may assume that it is an open subset of an orientable Riemannian manifold (M~,g)(\tilde{M},g) without boundary oriented by the Riemannian volume form dvolg{\rm dvol}_{g}. Let also (Xt,x)(X_{t},\mathbb{P}_{x}) be the Brownian motion on MM with initial condition at xx, that is, the stochastic process generated by the Laplace-Beltrami operator Δg\Delta_{g} (this article uses the convention Δg=dd\Delta_{g}=-d^{*}d with negative spectrum, where dd is the exterior derivative). For any ΓM\Gamma\subset\partial M open we denote by τΓ\tau_{\Gamma} the first time the Brownian motion XtX_{t} hits Γ\Gamma, that is

τΓ:=inf{t0:XtΓ}.\tau_{\Gamma}:=\inf\{t\geq 0:X_{t}\in\Gamma\}.

In the case when Γ=Γϵ,a\Gamma=\Gamma_{\epsilon,a} is a small elliptic window of eccentricity 1a2\sqrt{1-a^{2}} and size ϵ0+\epsilon\to 0^{+} (to be made precise later), the narrow escape/mean first arrival time problem wishes to derive an asymptotic expansion as ϵ0\epsilon\to 0 for the expected value 𝔼[τΓϵ,a|X0=x]\mathbb{E}[\tau_{\Gamma_{\epsilon,a}}|X_{0}=x] of the first arrival time τΓϵ,a\tau_{\Gamma_{\epsilon,a}} amongst all Brownian particles starting at xx. Another quantity of interest is the average expected value over MM:

|M|1M𝔼[τΓϵ,a|X0=x]dvolg(x).|M|^{-1}\int_{M}\mathbb{E}[\tau_{\Gamma_{\epsilon,a}}|X_{0}=x]{\rm dvol}_{g}(x).

Here |M||M| denotes the Riemannian volume of MM with respect to the metric gg.

Many problems in cellular biology may be formulated as mean first arrival time problems; a collection of analysis methods, results, applications, and references may be found in [10]. For example, cells have been modelled as simply connected two-dimensional domains with small absorbing windows on the boundary representing ion channels or target binding sites; the quantity sought is then the mean time for a diffusing ion or receptor to exit through an ion channel or reach a binding site [28, 8, 24].

There has been much progress for this problem in the setting of planar domains, and we refer the readers to [8, 24, 31, 1] and references therein for a complete bibliography. An important contribution was made in the planar case by [1] to introduce rigor into the computation of [24]. The use of layered potential in [1] also cast this problem in the mainstream language of elliptic PDE and facilitates some of the approach we use in this article.

Few results exists for three dimensional domains in n\mathbb{R}^{n} or Riemannian manifolds; see [3, 27, 30, 5] and references therein. The additional difficulties introduced by higher dimension are highlighted in the introduction of [1] and the challenges in geometry are outlined in [30]. In the case when MM is a domain in 3\mathbb{R}^{3} with Euclidean metric and Γϵ,a\Gamma_{\epsilon,a} is a single small disk absorbing window, [27, 30] gave an expansion for the average of the expected first arrival time, averaged over MM, up to an unspecified O(1)O(1) term:

(1.1) |M|1M𝔼[τΓϵ,a|X0=x]dvolg(x)|M|4ϵ[1ϵπHlogϵ+𝒪(ϵ)].\displaystyle|M|^{-1}\int_{M}\mathbb{E}[\tau_{\Gamma_{\epsilon,a}}|X_{0}=x]{\rm dvol}_{g}(x)\sim\frac{|M|}{4\epsilon}\left[1-\frac{\epsilon}{\pi}H\log\epsilon+\mathcal{O}(\epsilon)\right].

Here, HH is the mean curvature of the boundary at the center of the absorbing window. The case when Γϵ,a\Gamma_{\epsilon,a} is a small elliptic window was also addressed in [27, 30].

When MM is a three dimensional ball with multiple circular absorbing windows on the boundary, an expansion capturing the explicit form of the 𝒪(1)\mathcal{O}(1) correction in equation (1.1) in terms of the Neumann Green’s function and its regular part was done in [3]. The method of matched asymptotic used there required the explicit computation of the Neumann Green’s function, which is only possible in special geometries with high degrees of symmetry/homogeneity. In these results one does not see the full effects of local geometry. This result was also rigorously proved in [2] but with a better estimate for the error term.

In this paper we outline an approach which allows one to derive all the main terms of 𝔼[τΓϵ,a|X0=x]\mathbb{E}[\tau_{\Gamma_{\epsilon,a}}|X_{0}=x] (up to a remainder vanishing as ϵ0\epsilon\to 0) for Riemannian manifolds of dimension three with a multiple number of small absorbing windows which are boundary geodesic balls or ellipses. We will only demonstrate this approach for one absorbing window so as to not obscure the main idea. In the case when the window is a geodesic ball our approach also adapts naturally to Riemannian manifolds of any dimension as the proof of Proposition 1.1 as well as the analysis for inverting a key integral equation on the ball in Section 4 both carry through to higher dimensions.

We discuss briefly here on how to obtain a comprehensive singularity expansion at the boundary for the Neumann Green’s function on a Riemannian manifold as the Euclidean case was of interest in [30] and [3]. We will define in Section 3 the Neumann Green’s function G(x,z)G(x,z) on (M,g,M)(M,g,\partial M) which satisfies

ΔgG(x,z)=δx,νzG(x,z)zM=1|M|,MG(x,z)dvolM(z)=0,\Delta_{g}G(x,z)=-\delta_{x}\ ,\partial_{\nu_{z}}G(x,z)\mid_{z\in\partial M}=\frac{-1}{|\partial M|},\ \ \int_{\partial M}G(x,z){\rm dvol}_{\partial M}(z)=0,

where zMνzz\in\partial M\mapsto\nu_{z} is a outward pointing normal vector field and |M||\partial M| is the area of the boundary.

Singer-Schuss-Holcman in [30] highlighted the difficulty in obtaining a comprehensive singularity expansion of G(x,z)x,zM,xzG(x,z)\mid_{x,z\in\partial M,x\neq z} in a neighbourhood of the diagonal {x=z}\{x=z\} when MM is a bounded domain in n\mathbb{R}^{n}, but it turns out that even when MM is a general Riemannian manifold this question can be treated by the standard pseudodifferential operators approach. We only carry out this calculation in three dimensions as it pertains to our application. Readers who are interested in the higher dimensional analogue can follow our treatment to carry out the (cumbersome) calculations for themselves:

Proposition 1.1.

For x,yMx,y\in\partial M, set H(x)H(x) to be the mean curvature of M\partial M at xx, dh(x,y)d_{h}(x,y) the geodesic distance on the boundary given by metric h:=ιMgh:=\iota_{\partial M}^{*}g, dg(x,y)d_{g}(x,y) the geodesic distance given by the metric gg, and

IIx(V):=IIx(V,V),VTxM{\rm{II}}_{x}(V):={\rm{II}}_{x}(V,V),\ \ V\in T_{x}\partial M

the scalar second fundamental quadratic form (see pages 235 and 381 of [16] for definitions).
i) The map

fC(M)(MG(,y)f(y)𝑑volh(y))|Mf\in C^{\infty}(\partial M)\mapsto\left.\left(\int_{\partial M}G(\cdot,y)f(y)d{\rm vol}_{h}(y)\right)\right|_{\partial M}

is well defined and extends to a map from Hk(M)Hk+1(M)H^{k}(\partial M)\to H^{k+1}(\partial M) for all kk\in\mathbb{R} whose Schwartz kernel we will denote by GM(x,y)𝒟(M×M)G_{\partial M}(x,y)\in{\mathcal{D}}^{\prime}(\partial M\times\partial M). Here the map uH1(M)uMH1/2(M)u\in H^{1}(M)\mapsto u\mid_{\partial M}\in H^{1/2}(\partial M) is the trace map.
ii) There exists an open neighbourhood of the diagonal

Diag:={(x,y)M×Mx=y}{\rm Diag}:=\{(x,y)\in\partial M\times\partial M\mid x=y\}

such that in this neighbourhood, the singularity structure of GM(x,y)G_{\partial M}(x,y) is given by:

GM(x,y)\displaystyle\quad\quad G_{\partial M}(x,y) =\displaystyle= 12πdg(x,y)114πH(x)logdh(y,x)\displaystyle\frac{1}{2\pi}d_{g}(x,y)^{-1}-\frac{1}{4\pi}{H(x)}\log d_{h}(y,x)
+116π(IIx(expx;h1(y)|expx;h1(y)|h)IIx(expx;h1(y)|expx;h1(y)|h))+R(x,y),\displaystyle+\frac{1}{16\pi}\left({\rm{II}}_{x}\left(\frac{\exp_{x;h}^{-1}(y)}{|\exp_{x;h}^{-1}(y)|_{h}}\right)-{\rm{II}}_{x}\left(\frac{*\exp_{x;h}^{-1}(y)}{|\exp_{x;h}^{-1}(y)|_{h}}\right)\right)+R(x,y),

where R(,)C0,μ(M×M)R(\cdot,\cdot)\in C^{0,\mu}(\partial M\times\partial M), for all μ<1\mu<1, is called the regular part of the Green’s function and * is the Hodge-star operator (i.e. rotation by π/2\pi/2 on the surface M\partial M).

We recall the definition of the exponential map. Let (X,g0)(X,g_{0}) be a geodesically complete manifold. For any xXx\in X and VTxXV\in T_{x}X there exists a unique geodesic γg0(t)=γg0(t;V)\gamma_{g_{0}}(t)=\gamma_{g_{0}}(t;V), defined on [0,1][0,1], such that γg0(0)=x\gamma_{g_{0}}(0)=x, γg(0)=V\gamma_{g}^{\prime}(0)=V. The exponential map based at xx is then a map taking TxXXT_{x}X\to X defined by

expx;g0:Vγg0(1;V).\exp_{x;g_{0}}:V\mapsto\gamma_{g_{0}}(1;V).

Observe that when MM is a Euclidean ball the singular term involving the second fundamental forms vanishes due to homogeneity and therefore does not show up. This is consistent with the explicit formula derived in [3].

An explicit formula for the regular part is only possible in special geometries such as the one considered in [3]. However, our approach in arriving at (1.1) also provides a way to numerically compute R(x,y)R(x,y) via a Fredholm integral equation. See Remark 3.3.

We will use the formula in Proposition 1.1 to derive the mean first arrival time of a Brownian particle on a Riemannian manifold with a single absorbing window which is a small geodesic ellipse. As mentioned earlier, our method extends to multiple windows but we present the single window case to simplify notations. We first state the result when the window is a geodesic disk of the boundary M\partial M around a fixed point since the statement is cleaner:

Theorem 1.2.

Let (M,g,M)(M,g,\partial M) be a smooth Riemannian manifold of dimension three with boundary and let |M||M| be its volume.
i) Fix xMx^{*}\in\partial M and let Γϵ\Gamma_{\epsilon} be a boundary geodesic ball centered at xx^{*} of geodesic radius ϵ>0\epsilon>0. For each xΓϵx\notin\Gamma_{\epsilon},

𝔼[τΓϵ|X0=x]=F(x)+Cϵ|M|G(x,x)+rϵ(x),\mathbb{E}[\tau_{\Gamma_{\epsilon}}|X_{0}=x]=F(x)+C_{\epsilon}-|M|G(x,x^{*})+r_{\epsilon}(x),

with rϵCk(K)Ck,Kϵ\|r_{\epsilon}\|_{C^{k}(K)}\leq C_{k,K}\epsilon for any integer kk and compact set KM¯K\subset\overline{M} which does not contain xx^{*}. The function FF is the unique solution to the boundary value problem

ΔgF=1,νF=|M|/|M|,MF=0.\Delta_{g}F=-1,\ \ \partial_{\nu}F=-|M|/|\partial M|,\ \ \int_{\partial M}F=0.

The constant CϵC_{\epsilon} is, modulo an error of O(ϵlogϵ)O(\epsilon\log\epsilon), given by

Cϵ=|M|4ϵ14πH(x)|M|logϵ+R(x,x)|M|F(x)|M|H(x)4π(2log232),\displaystyle C_{\epsilon}=\frac{|M|}{4\epsilon}-\frac{1}{4\pi}H(x^{*})|M|\log\epsilon+R(x^{*},x^{*})|M|-F(x^{*})-\frac{|M|H(x^{*})}{4\pi}\left(2\log 2-\frac{3}{2}\right),

where R(x,x)R(x^{*},x^{*}) is the evaluation at (x,y)=(x,x)(x,y)=(x^{*},x^{*}) of the kernel R(x,y)R(x,y) in (1.1).

ii) One has that the integral of 𝔼[τΓϵ,a|X0=x]\mathbb{E}[\tau_{\Gamma_{\epsilon,a}}|X_{0}=x] over MM satisfies

M𝔼[τΓϵ,a|X0=x]dvolg(x)=MF(x)dvolg(x)+Cϵ|M|F(x)|M|+O(ϵ).\int_{M}\mathbb{E}[\tau_{\Gamma_{\epsilon,a}}|X_{0}=x]{\rm dvol}_{g}(x)=\int_{M}F(x){\rm dvol}_{g}(x)+C_{\epsilon}|M|-F(x^{*})|M|+O(\epsilon).

Theorem 1.2 does not realize the full power of Proposition 1.1 as it does not see the non-homogeneity of the local geometry at xx^{*} (only the mean curvature H(x)H(x^{*}) shows up). This is due to the fact that we are looking at windows which are geodesic balls. If we replace geodesic balls with geodesic ellipses, we see that the second fundamental form term in (1.1) contributes to a term in 𝔼[τΓϵ,a|X0=x]\mathbb{E}[\tau_{\Gamma_{\epsilon,a}}|X_{0}=x] which is the difference of principal curvatures.

To this end let E1(x),E2(x)TxME_{1}(x^{*}),E_{2}(x^{*})\in T_{x^{*}}\partial M be the unit eigenvectors of the shape operator at xx^{*} corresponding respectively to the principal curvatures λ1(x),λ2(x)\lambda_{1}(x^{*}),\ \lambda_{2}(x^{*}). For 1a>01\geq a>0 fixed, let

Γϵ,a:={expx;h(ϵt1E1(x)+ϵt2E2(x))t12+a2t221}\Gamma_{\epsilon,a}:=\{\exp_{x^{*};h}(\epsilon t_{1}E_{1}(x^{*})+\epsilon t_{2}E_{2}(x^{*}))\mid t_{1}^{2}+a^{-2}t_{2}^{2}\leq 1\}

be a small geodesic ellipse.

Theorem 1.3.

Let (M,g,M)(M,g,\partial M) be a smooth Riemannian manifold of dimension three with boundary.
i) For each xM\Γϵ,ax\in M\backslash\Gamma_{\epsilon,a},

𝔼[τΓϵ,a|X0=x]=F(x)+Cϵ,a|M|G(x,x)+rϵ(x)\mathbb{E}[\tau_{\Gamma_{\epsilon,a}}|X_{0}=x]=F(x)+C_{\epsilon,a}-|M|G(x,x^{*})+r_{\epsilon}(x)

with rϵCk(K)Ck,Kϵ\|r_{\epsilon}\|_{C^{k}(K)}\leq C_{k,K}\epsilon for any integer kk and compact set KM¯K\subset\overline{M} which does not contain xx^{*}. The function FF is the unique solution to the boundary value problem

ΔgF=1,νF=|M|/|M|,MF=0.\Delta_{g}F=-1,\ \ \partial_{\nu}F=-|M|/|\partial M|,\ \ \int_{\partial M}F=0.

The constant Cϵ,aC_{\epsilon,a} is given by

(1.3) Cϵ,a=\displaystyle C_{\epsilon,a}= |M|Ka4aϵπ214πH(x)|M|logϵ+aR(x,x)|M|F(x)\displaystyle\frac{|M|K_{a}}{4a\epsilon\pi^{2}}-\frac{1}{4\pi}H(x^{*})|M|\log\epsilon+aR(x^{*},x^{*})|M|-F(x^{*})
|M|H(x)16π3𝔻1(1|s|2)1/2𝔻log((t1s1)2+a2(t2s2)2)1/2(1|t|2)1/2𝑑t𝑑s\displaystyle-\frac{|M|H(x^{*})}{16\pi^{3}}\int_{\mathbb{D}}\frac{1}{(1-|s^{\prime}|^{2})^{1/2}}\int_{\mathbb{D}}\frac{\log\left((t_{1}-s_{1})^{2}+a^{2}(t_{2}-s_{2})^{2}\right)^{1/2}}{(1-|t^{\prime}|^{2})^{1/2}}dt^{\prime}ds^{\prime}
+|M|(λ1λ2)64π3𝔻1(1|s|2)1/2𝔻(t1s1)2a2(t2s2)2(t1s1)2+a2(t2s2)21(1|t|2)1/2𝑑t𝑑s\displaystyle+\frac{|M|(\lambda_{1}-\lambda_{2})}{64\pi^{3}}\int_{\mathbb{D}}\frac{1}{(1-|s^{\prime}|^{2})^{1/2}}\int_{\mathbb{D}}\frac{(t_{1}-s_{1})^{2}-a^{2}(t_{2}-s_{2})^{2}}{(t_{1}-s_{1})^{2}+a^{2}(t_{2}-s_{2})^{2}}\frac{1}{(1-|t^{\prime}|^{2})^{1/2}}dt^{\prime}ds^{\prime}
+O(ϵlogϵ),\displaystyle+O(\epsilon\log\epsilon),

where Ka=π202π(cos2θ+sin2θa2)1/2𝑑θK_{a}=\frac{\pi}{2}\int_{0}^{2\pi}{\left(\cos^{2}\theta+\frac{\sin^{2}\theta}{a^{2}}\right)^{-1/2}}d\theta and 𝔻\mathbb{D} is the two dimensional unit disk centered at the origin.

ii) One has that the integral of 𝔼[τΓϵ,a|X0=x]\mathbb{E}[\tau_{\Gamma_{\epsilon,a}}|X_{0}=x] over MM satisfies

M𝔼[τΓϵ,a|X0=x]dvolg(x)=MF(x)dvolg+Cϵ,a|M|F(x)|M|+O(ϵ).\int_{M}\mathbb{E}[\tau_{\Gamma_{\epsilon,a}}|X_{0}=x]{\rm dvol}_{g}(x)=\int_{M}F(x){\rm dvol}_{g}+C_{\epsilon,a}|M|-F(x^{*})|M|+O(\epsilon).

Note that while the dependence on the eccentricity of the ellipse is hidden in the integrals, the dependence on the difference of the principal curvatures (λ1λ2)(\lambda_{1}-\lambda_{2}) is easy to see in this formula. The integral which multiplies (λ1λ2)(\lambda_{1}-\lambda_{2}) turns out to vanish when a=1a=1 which makes the above result consistent with Theorem 1.2.

The fact that our result is valid on general Riemannian three manifolds allows for the incorporation of spatial heterogeneity such as anisotropic diffusion. In contrast to [30], the fact that we are able to obtain explicitly an expression for the O(1)O(1) term in (1.1) is due to the fact that in Proposition 1.1 we have the expansion of GM(x,z)G_{\partial M}(x,z) all the way to a remainder R(x,y)R(x,y), which is Hölder continuous at the diagonal. We also appeal to some recent advances in integral geometry [29, 20, 23, 22, 12] to address the comment in [1] on the difficulty of treating this problem in higher dimensions.

The strategy and organization of this paper will be as follows. In Section 2 we will give a brief overview of pseudodifferential operators and their associated Schwartz kernels. The machinery of pseudodifferential operators serve as a bridge between the geometric and analytic objects appearing in Proposition 1.1 and we will compute their coordinate expression. In Section 3 we will use the tools we developed in Section 2 to prove Propostion 1.1. A singularity expansion for the Green’s function such as Proposition 1.1 is the gateway for obtaining the asymptotic expansions of Theorems 1.2 and 1.3. However, there is an additional hurdle of inverting an integral transform as mentioned in [1]. Here we make use of some recent advancements in integral geometry and geometric rigidity [23, 12, 20] to overcome these difficulties. This approach is described in Section 4. Finally, in Section 5 we carry out the asymptotic calculation using the tools we have developed. The appendices characterizes the expected first arrival time 𝔼[τΓϵ,a|X0=x]\mathbb{E}[\tau_{\Gamma_{\epsilon,a}}|X_{0}=x] as the solution of an elliptic mixed boundary value problem. This is classical in the Euclidean case (see [26]) but we could not find a reference for the general case of a Riemannian manifold with boundary.

2. Overview of Pseudodifferential Operators

2.1. Basic Definitions

We give some basic definitions and properties of pseudodifferential operators. For a comprehensive treatment we refer the reader to Chapt 7 of [34] or the book [32]. Readers who are already familiar with microlocal analysis can skip this section.

As usual, CC^{\infty} denotes the space of smooth functions. We use notation CcC_{c}^{\infty} for compactly suported smooth functions and DD^{\prime} for its dual. By CkC^{k}, we denote the space of kk time continuously differentiable functions. The spce of functions from CkC^{k}, whose kkth derivatives are Hölder continuous with exponent μ(0,1]\mu\in(0,1], is denoted by Ck,μC^{k,\mu}

Let a(x,ξ)a(x,\xi) be a smooth function on TnT^{*}\mathbb{R}^{n} and for all ll\in\mathbb{R} we say that aS1l(Tn)a\in S^{l}_{1}(T^{*}\mathbb{R}^{n}) (or simply S1lS^{l}_{1}) if for all multi-indices α,β\alpha,\beta there are constants Cα,βC_{\alpha,\beta} such that

(2.1) |DξαDxβa(x,ξ)|Cα,βξl|α|\displaystyle|D_{\xi}^{\alpha}D_{x}^{\beta}a(x,\xi)|\leq C_{\alpha,\beta}\langle\xi\rangle^{l-|\alpha|}

where Dξα=(i)|α|ξαD_{\xi}^{\alpha}=(-i)^{|\alpha|}\partial_{\xi}^{\alpha}, Dxβ=(i)|β|xβD_{x}^{\beta}=(-i)^{|\beta|}\partial_{x}^{\beta} , and ξ:=(1+|ξ|2)1/2\langle\xi\rangle:=(1+|\xi|^{2})^{1/2}. These are the Kohn-Nirenberg symbols. This class of symbols contain the classical symbols, denoted by Scll(Tn)S_{cl}^{l}(T^{*}\mathbb{R}^{n}), which are defined by those a(x,ξ)S1l(Tn)a(x,\xi)\in S^{l}_{1}(T^{*}\mathbb{R}^{n}) satisfying

(2.2) a(x,ξ)m=0alm(x,ξ),\displaystyle a(x,\xi)\sim\sum_{m=0}^{\infty}a_{l-m}(x,\xi),

where each alma_{l-m} are homogeneous in the sense that alm(x,τξ)=τlma(x,ξ)a_{l-m}(x,\tau\xi)=\tau^{l-m}a(x,\xi) for all xnx\in\mathbb{R}^{n}, τ>1\tau>1 and |ξ|>1|\xi|>1. The expression (2.2) means that for all NN,

a(x,ξ)m=0Nalm(x,ξ)S1lN1(Tn).a(x,\xi)-\sum_{m=0}^{N}a_{l-m}(x,\xi)\in S^{l-N-1}_{1}(T^{*}\mathbb{R}^{n}).

If a(x,ξ)S1la(x,\xi)\in S^{l}_{1} we can define an operator a(x,D):Cc(n)𝒟(n)a(x,D):C^{\infty}_{c}(\mathbb{R}^{n})\to{\mathcal{D}}^{\prime}(\mathbb{R}^{n}) by

(2.3) a(x,D)u:=neiξxa(x,ξ)u^(ξ)𝑑ξ,\displaystyle a(x,D)u:=\int_{\mathbb{R}^{n}}e^{i\xi\cdot x}a(x,\xi)\hat{u}(\xi)d\xi,

where u^(ξ):=u:=(2π)nneixξu(x)𝑑x\hat{u}(\xi):={\mathcal{F}}u:=(2\pi)^{n}\int_{\mathbb{R}^{n}}e^{-ix\cdot\xi}u(x)dx is the Fourier transform. Recall that the absolutely convergent integral representation of the Fourier transform is well defined as an automorphism of the Schwartz class functions S(n)S(\mathbb{R}^{n}) but extends to an automorphism of the tempered distributions S(n)S^{\prime}(\mathbb{R}^{n}). (See [4] for a comprehensive guide to distribution theory and definition of these spaces).

Operators taking Cc(n)𝒟(n)C^{\infty}_{c}(\mathbb{R}^{n})\to{\mathcal{D}}^{\prime}(\mathbb{R}^{n}) which have the above representation are said to be in Ψ1l(n)\Psi^{l}_{1}(\mathbb{R}^{n}) and are called pseudodifferential operators. For the symbol class S1l(Tn)S_{1}^{l}(T^{*}\mathbb{R}^{n}), Lemma 1.1 in Chapter 7.1 of [34] extends a(x,D)a(x,D) to map S(n)S(n)S^{\prime}(\mathbb{R}^{n})\to S^{\prime}(\mathbb{R}^{n}).

The classical pseudodifferential operators Ψcll(n)\Psi^{l}_{cl}(\mathbb{R}^{n}) are defined analogously by requiring that a(x,ξ)a(x,\xi) belongs to ScllS^{l}_{cl}. Note that knowing the operator a(x,D)Ψ1l(n)a(x,D)\in\Psi^{l}_{1}(\mathbb{R}^{n}) we can recover a(x,ξ)S1la(x,\xi)\in S^{l}_{1} by the formula

(2.4) a(x,ξ)=eξ(x)a(x,D)eξ,\displaystyle a(x,\xi)=e_{-\xi}(x)a(x,D)e_{\xi},

where eξ(x):=eiξxe_{\xi}(x):=e^{i\xi\cdot x}. Note that if A(x,y)A(x,y) is the Schwartz kernel of the operator a(x,D)a(x,D) then

(2.5) a(0,ξ)=y1(A(0,y))(ξ)=neiξyA(0,y)𝑑y.\displaystyle a(0,\xi)={\mathcal{F}}_{y}^{-1}(A(0,y))(\xi)=\int_{\mathbb{R}^{n}}e^{i\xi\cdot y}A(0,y)dy.

Let XX be a compact manifold without boundary. An operator 𝒜:C(X)𝒟(X)\mathcal{A}:C^{\infty}(X)\to{\mathcal{D}}^{\prime}(X)111We shall avoid a discussion of distributional sections of density bundles by noting that in our setting of XX is always prescribed with a Riemannian volume form which provides a natural trivialization of density bundles. As such all distributional sections of density bundles are identified with sections of the trivial line bundle in this way. is said to be in Ψ1l(X)\Psi^{l}_{1}(X) if there exists coordinate covers {(Oj,Φj)Φj:OjUjn}\{(O_{j},\Phi_{j})\mid\Phi_{j}:O_{j}\to U_{j}\subset\mathbb{R}^{n}\} and a partition of unity {χj}\{\chi_{j}\} subordinate to {Oj}\{O_{j}\} such that the map

(2.6) u(χk𝒜χjΦju)Φk1\displaystyle u\mapsto\left(\chi_{k}\mathcal{A}\chi_{j}\Phi_{j}^{*}u\right)\circ\Phi_{k}^{-1}

from C(Uj)(Uk)C^{\infty}(U_{j})\to{\mathcal{E}}^{\prime}(U_{k}) belongs to Ψ1l(n)\Psi^{l}_{1}(\mathbb{R}^{n}).

If aC(TX)a\in C^{\infty}(T^{*}X) we say that it belongs to the symbol class S1l(TX)S^{l}_{1}(T^{*}X) if

χjΦj1a(Φj1(),Φj)S1l(Tn)\chi_{j}\circ\Phi_{j}^{-1}a(\Phi_{j}^{-1}(\cdot),\Phi_{j}^{*}\cdot)\in S^{l}_{1}(T^{*}\mathbb{R}^{n})

for all jj. The classical pseudodifferential operators Ψcll(X)\Psi^{l}_{cl}(X) and classical symbols Scll(TX)S^{l}_{cl}(T^{*}X) are defined analogously. These definitions depend a-priori on the choice of coordinate systems but turn out to be invariant (see Chapt 7 [34]).

There exists a linear isomorphism

(2.7) σl:Ψcll(X)/Ψcll1(X)Scll(TX)/Scll1(TX)\displaystyle\sigma_{l}:\Psi^{l}_{cl}(X)/\Psi^{l-1}_{cl}(X)\to S^{l}_{cl}(T^{*}X)/S^{l-1}_{cl}(T^{*}X)

called the principal symbol map. For each 𝒜Ψcll(X)\mathcal{A}\in\Psi^{l}_{cl}(X) it can be defined at each xXx\in X by taking a coordinate neighborhood OO containing xx and a χCc(O)\chi\in C^{\infty}_{c}(O) which is identically 11 near xx then considering the operator given in (2.6) for χj=χk=χ\chi_{j}=\chi_{k}=\chi and Φj=Φk=Φ\Phi_{j}=\Phi_{k}=\Phi. As the resulting operator in (2.6) is in Ψcll(n)\Psi^{l}_{cl}(\mathbb{R}^{n}) with symbol aScll(Tn)a\in S^{l}_{cl}(T^{*}\mathbb{R}^{n}), we may set

σl(𝒜)(x,Φξ):=al(Φ(x),ξ)\sigma_{l}(\mathcal{A})(x,\Phi^{*}\xi):=a_{l}(\Phi(x),\xi)

for all xXx\in X and ξTΦj(x)n\xi\in T^{*}_{\Phi_{j}(x)}\mathbb{R}^{n}. This definition depends a-priori on the choice of coordinate systems but turns out to be invariant (see Chapt 5 of [32]). In practice these computations are often done in normal coordinates centered at the point of interest xXx\in X then computing the inverse Fourier transform as in (2.5).

One important property of σl\sigma_{l} which we will use is that it respects the product structure of Ψcll\Psi^{l}_{cl} and ScllS^{l}_{cl}:

(2.8) σl(𝒜)σm()=σl+m(𝒜)\displaystyle\sigma_{l}(\mathcal{A})\sigma_{m}({\mathcal{B}})=\sigma_{l+m}(\mathcal{A}{\mathcal{B}})

for 𝒜Ψcll(X)\mathcal{A}\in\Psi^{l}_{cl}(X) and Ψclm(X){\mathcal{B}}\in\Psi^{m}_{cl}(X).

2.2. Coordinate Calculations

We make some calculations for some geometric objects which will naturally appear in the singularity expansion for GMG_{\partial M}. These identities will be useful in proving Proposition 1.1.

Let (M,g,M)(M,g,\partial M) be a three dimensional Riemannian manifold with non-empty smooth boundary which inherits the metric h:=ιMgh:=\iota_{\partial M}^{*}g. Denote by II{\rm{II}} the scalar second fundamental form on the surface M\partial M and H(x)H(x) be the mean curvature at xMx\in\partial M. Let SMS\partial M denote the unit-sphere bundle over M\partial M,

SM={vTMvh=1}.S\partial M=\{v\in T\partial M\mid\|v\|_{h}=1\}.

For any xMx\in\partial M, let E1(x),E2(x)SxME_{1}(x),E_{2}(x)\in S_{x}\partial M be principal directions (i.e. unit eigenvectors) of the induced shape operator with eigenvalues λ1(x)\lambda_{1}(x) and λ2(x)\lambda_{2}(x). We will drop the dependence in xx from our notation when there is no ambiguity.

We choose E1E_{1} and E2E_{2} such that E1E2νE_{1}^{\flat}\wedge E_{2}^{\flat}\wedge\nu^{\flat} is a positive multiple of the volume form dvolg{\rm dvol}_{g} (see p.26 of [16] for the “musical isomorphism” notation of and ). Here we use ν\nu to denote the outward pointing normal vector field so that it is consistent with most PDE literature. However, in defining II{\rm{II}} and the shape operator we will follow geometry literature (e.g.[16]) and use the inward pointing normal so that the sphere embedded in 3\mathbb{R}^{3} would have positive mean curvature in our convention.

For two points x,yMx,y\in\partial M there are two distances to consider. The first is the shortest path amongst those that stay on the boundary which we denote by dh(x,y)d_{h}(x,y) and the other is the distance measured by paths allowing to enter MM, which we denote by dg(x,y)d_{g}(x,y). Clearly, dh(x,y)dg(x,y)d_{h}(x,y)\geq d_{g}(x,y).

For a fixed x0Mx_{0}\in\partial M, we will denote by Bh(ρ;x0)MB_{h}(\rho;x_{0})\subset\partial M the geodesic disk of radius ρ>0\rho>0 (with respect to the metric hh) centered at x0x_{0} and 𝔻ρ\mathbb{D}_{\rho} to be the Euclidean disk in 2\mathbb{R}^{2} of radius ρ\rho centered at the origin. In what follows ρ\rho will always be smaller than the injectivity radius of (M,h)(\partial M,h). Letting t=(t1,t2,t3)3t=(t_{1},t_{2},t_{3})\in\mathbb{R}^{3}, we will construct a coordinate system x(t;x0)x(t;x_{0}) by the following procedure:
Write t3t\in\mathbb{R}^{3} near the origin as t=(t,t3)t=(t^{\prime},t_{3}) for t=(t1,t2)𝔻ρt^{\prime}=(t_{1},t_{2})\in\mathbb{D}_{\rho}. Define first 222for example it is obvious below that E1E_{1} and E2E_{2} are elements of the tangent space over x0x_{0} as they are inserted into the argument of expx0;h()\exp_{x_{0};h}(\cdot).

x((t,0);x0):=expx0;h(t1E1+t2E2),x((t^{\prime},0);x_{0}):={\rm{exp}}_{x_{0};h}(t_{1}E_{1}+t_{2}E_{2}),

where expx0;h(V){\rm{exp}}_{x_{0};h}(V) denotes the time 11 map of hh-geodesics with initial point x0x_{0} and initial velocity VTx0MV\in T_{x_{0}}\partial M. The coordinate t𝔻ρx((t,0);x0)t^{\prime}\in\mathbb{D}_{\rho}\mapsto x((t^{\prime},0);x_{0}) is then an hh-geodesic coordinate system for a neighborhood of x0x_{0} on the boundary surface M\partial M. We can extend this to become a coordinate system for points in MM near x0x_{0} so that tx(t;x0)t\mapsto x(t;x_{0}) is a boundary normal coordinate system with t3>0t_{3}>0 in MM as the boundary defining function. Readers wishing to know more about boundary normal coordinates can refer to [17] for a brief recollection of the basic properties we use here and Prop 5.26 of [16] for a detailed construction.

For convenience we will write x(t;x0)x(t^{\prime};x_{0}) in place of x((t,0);x0)x((t^{\prime},0);x_{0}). The boundary coordinate system tx(t;x0)t\mapsto x(t;x_{0}) has the advantage that the metric tensor gg can be expressed as

(2.9) j,k=13gj,k(t)dtjdtk=α,β=12hα,β(t,t3)dtαdtβ+dt32,\displaystyle\sum_{j,k=1}^{3}g_{j,k}(t)dt_{j}dt_{k}=\sum_{\alpha,\beta=1}^{2}h_{\alpha,\beta}(t^{\prime},t_{3})dt_{\alpha}dt_{\beta}+dt_{3}^{2},

where hα,β(t,0)=hα,β(t)h_{\alpha,\beta}(t^{\prime},0)=h_{\alpha,\beta}(t^{\prime}) is the expression of the boundary metric hh in the hh-geodesic coordinate system x(t;x0)x(t^{\prime};x_{0}). Note that (gj,k(t))j,k=13(g_{j,k}(t))_{j,k=1}^{3} and (hα,β(t,t3))α,β=12(h_{\alpha,\beta}(t^{\prime},t_{3}))_{\alpha,\beta=1}^{2} are symmetric positive definite 3×33\times 3 and 2×22\times 2 matrices varying smoothly with respect to the variable t=(t,t3)=(t1,t2,t3)t=(t^{\prime},t_{3})=(t_{1},t_{2},t_{3}).

For ϵ>0\epsilon>0 sufficiently small we define the (rescaled) hh-geodesic coordinate by the following map

xϵ(;x0):t=(t1,t2)𝔻x(ϵt;x0)Bh(ϵ;x0),x^{\epsilon}(\cdot;x_{0}):t^{\prime}=(t_{1},t_{2})\in\mathbb{D}\mapsto x(\epsilon t^{\prime};x_{0})\in B_{h}(\epsilon;x_{0}),

where 𝔻\mathbb{D} is the unit disk in 2\mathbb{R}^{2}. We derive some coordinate expressions for some of the geometric objects we will consider later.

Lemma 2.1.

Let h(s)=α,β=12hα,β(s)dsαdsβh(s^{\prime})=\sum_{\alpha,\beta=1}^{2}h_{\alpha,\beta}(s^{\prime})ds_{\alpha}ds_{\beta} be the pullback metric of hh by sx(s;x0)s^{\prime}\mapsto x(s^{\prime};x_{0}) on 𝔻ρ\mathbb{D}_{\rho}. Denote by

r=|st|h(s):=(α,β=12hα,β(s)(sαtα)(sβtβ))1/2r=|s^{\prime}-t^{\prime}|_{h(s^{\prime})}:=\left(\sum_{\alpha,\beta=1}^{2}h_{\alpha,\beta}(s^{\prime})(s_{\alpha}-t_{\alpha})(s_{\beta}-t_{\beta})\right)^{1/2}

so that t=s+rωt^{\prime}=s^{\prime}+r\omega where ωSs𝔻ρ\omega\in S_{s^{\prime}}\mathbb{D}_{\rho}. We have that

dh(x(s;x0),x(t;x0))2=r2j,k=12Hj,k(s,r,ω)ωjωkd_{h}(x(s^{\prime};x_{0}),x(t^{\prime};x_{0}))^{2}=r^{2}\sum\limits_{j,k=1}^{2}H_{j,k}(s^{\prime},r,\omega)\omega_{j}\omega_{k}

for matrix Hj,k(s,r,ω)H_{j,k}(s^{\prime},r,\omega) jointly smooth in (s,r,ω)(s^{\prime},r,\omega). It also satisfies Hj,k(s,0,ω)=hj,k(s)H_{j,k}(s^{\prime},0,\omega)=h_{j,k}(s^{\prime}) and

j,krHj,k(s,0,ω)ωkωj=O(s).\sum_{j,k}\partial_{r}H_{j,k}(s^{\prime},0,\omega)\omega_{k}\omega_{j}=O(s^{\prime}).
Proof.

Expressing tt^{\prime} using (s,r,ω)(s^{\prime},r,\omega) we have that (see e.g. [35] Lemma 4.8)

dh(x(s;x0),x(t;x0))=r(α,βHα,β(s,r,ω)ωαωβ)1/2d_{h}(x(s^{\prime};x_{0}),x(t^{\prime};x_{0}))=r\left(\sum_{\alpha,\beta}H_{\alpha,\beta}(s^{\prime},r,\omega)\omega_{\alpha}\omega_{\beta}\right)^{1/2}

with Hα,βH_{\alpha,\beta} symmetric, even under the map (r,ω)(r,ω)(r,\omega)\mapsto(-r,-\omega), and Hα,β(s,0,ω)=hα,β(s)H_{\alpha,\beta}(s^{\prime},0,\omega)=h_{\alpha,\beta}(s^{\prime}). Setting s=0s^{\prime}=0 and using the fact that we are using normal coordinates we obtain r2=r2(α,βHα,β(0,r,ω)ωαωβ)r^{2}=r^{2}\left(\sum_{\alpha,\beta}H_{\alpha,\beta}(0,r,\omega)\omega_{\alpha}\omega_{\beta}\right). Now Taylor expanding Hα,β(0,r,ω)H_{\alpha,\beta}(0,r,\omega) in rr, we see that

r2=r2(1+α,βrHα,β(0,0,ω)rωαωβ+O(r2)).r^{2}=r^{2}\left(1+\sum_{\alpha,\beta}\partial_{r}H_{\alpha,\beta}(0,0,\omega)r\omega_{\alpha}\omega_{\beta}+O(r^{2})\right).

The r2r^{2} terms on left-hand and right-hand sides cancel, leaving

0=rHα,β(0,0,ω)r3ωαωβ+O(r4).0=\partial_{r}H_{\alpha,\beta}(0,0,\omega)r^{3}\omega_{\alpha}\omega_{\beta}+O(r^{4}).

Divide through by r3r^{3} and take the limit as r0r\to 0 we see that α,βrHα,β(0,0,ω)ωαωβ=0\sum_{\alpha,\beta}\partial_{r}H_{\alpha,\beta}(0,0,\omega)\omega_{\alpha}\omega_{\beta}=0 for any ωS0𝔻\omega\in S_{0}\mathbb{D}. ∎

Lemma 2.2.

We use the same notation for ω\omega and rr as in Lemma 2.1. One has that

dh(x(s;x0),x(t;x0))1=r1+O(s)+O(r),\displaystyle d_{h}(x(s^{\prime};x_{0}),x(t^{\prime};x_{0}))^{-1}=r^{-1}+O(s^{\prime})+O(r),

where O(r)O(r) (respectively O(s)O(s^{\prime})) denotes smooth functions of (s,r,ω)𝔻ρ××Ss𝔻ρ(s^{\prime},r,\omega)\in\mathbb{D}_{\rho}\times\mathbb{R}\times S_{s^{\prime}}\mathbb{D}_{\rho} which vanish to first order as r0r\to 0 (respectively s0s^{\prime}\to 0).

Proof.

From Lemma 2.1 we have that

dh(x(s;x0),x(t;x0))1=r1(α,βhα,β(s)ωαωβ+rrHα,β(s,0,ω)ωαωβ+O(r2))1/2.d_{h}(x(s^{\prime};x_{0}),x(t^{\prime};x_{0}))^{-1}=r^{-1}\left(\sum_{\alpha,\beta}h_{\alpha,\beta}(s^{\prime})\omega_{\alpha}\omega_{\beta}+r\partial_{r}H_{\alpha,\beta}(s^{\prime},0,\omega)\omega_{\alpha}\omega_{\beta}+O(r^{2})\right)^{-1/2}.

Using the fact that ωSs𝔻\omega\in S_{s^{\prime}}\mathbb{D} with respect to the metric given by hα,βh_{\alpha,\beta} we have that

dh(x(s;x0),x(t;x0))1=r1(α,β1+rrHα,β(s,0,ω)ωαωβ+O(r2))1/2.d_{h}(x(s^{\prime};x_{0}),x(t^{\prime};x_{0}))^{-1}=r^{-1}\left(\sum_{\alpha,\beta}1+r\partial_{r}H_{\alpha,\beta}(s^{\prime},0,\omega)\omega_{\alpha}\omega_{\beta}+O(r^{2})\right)^{-1/2}.

For rr and ss^{\prime} sufficiently small we may use Taylor’s expansion to obtain the desired property. The fact that α,βrHα,β(s,0,ω)ωαωβ=O(s)\sum_{\alpha,\beta}\partial_{r}H_{\alpha,\beta}(s^{\prime},0,\omega)\omega_{\alpha}\omega_{\beta}=O(s^{\prime}) is stated in Lemma 2.1

Corollary 2.3.

For ϵ>0\epsilon>0 sufficiently small we have that

dh(xϵ(s;x0),xϵ(t;x0))1=ϵ1r1+ϵr1A(ϵ,s,r,ω)\displaystyle d_{h}(x^{\epsilon}(s^{\prime};x_{0}),x^{\epsilon}(t^{\prime};x_{0}))^{-1}=\epsilon^{-1}r^{-1}+\epsilon r^{-1}A(\epsilon,s^{\prime},r,\omega)

for some smooth function AA in the variables (ϵ,s,r,ω)[0,ϵ0]×𝔻××S1(\epsilon,s^{\prime},r,\omega)\in[0,\epsilon_{0}]\times\mathbb{D}\times\mathbb{R}\times S^{1}. Here we use r=|st|r=|s^{\prime}-t^{\prime}| and t=s+rωt^{\prime}=s+r\omega.

Lemma 2.4.

Let x(;x0)x(\cdot;x_{0}) be the coordinate system at the beginning of this section centered at x0x_{0}. For s,t2s^{\prime},t^{\prime}\in\mathbb{R}^{2} sufficiently small, we have that

dg(x(s;x0),x(t;x0))2=r2(1+r𝐆~(s,ω)+O(r2)),d_{g}(x(s^{\prime};x_{0}),x(t^{\prime};x_{0}))^{2}=r^{2}\left(1+r{\bf\tilde{G}}(s^{\prime},\omega)+O(r^{2})\right),

where r=|st|h(s)r=|s^{\prime}-t^{\prime}|_{h(s^{\prime})} and t=s+rωt^{\prime}=s^{\prime}+r\omega. Here 𝐆~(s,ω){\bf\tilde{G}}(s^{\prime},\omega) is a smooth function of (s,ω)(s^{\prime},\omega) which vanishes at s=0s^{\prime}=0. The O(r2)O(r^{2}) term is a smooth function in (s,r,ω)𝔻ρ××Ss𝔻ρ(s^{\prime},r,\omega)\in\mathbb{D}_{\rho}\times\mathbb{R}\times S_{s^{\prime}}\mathbb{D}_{\rho} which vanishes to second order as r0r\to 0.

Proof.

We begin with the identity that for any ss and tt,

dg(x(s;x0),x(t;x0))2=j,k=13𝐆j,k(s,t)(sjtj)(sktk),d_{g}(x(s;x_{0}),x(t;x_{0}))^{2}=\sum_{j,k=1}^{3}{\bf G}_{j,k}(s,t)(s_{j}-t_{j})(s_{k}-t_{k}),

with 𝐆j,k(s,s)=gj,k(s){\bf G}_{j,k}(s,s)=g_{j,k}(s) given by (2.9). Now set s=(s,0)s=(s^{\prime},0) and t=(t,0)t=(t^{\prime},0) we have

dg(x(s;x0),x(t;x0))2=r2(1+j,k=12r𝐆j,k0(s,0,ω)rωjωk+O(r2))d_{g}(x(s^{\prime};x_{0}),x(t^{\prime};x_{0}))^{2}=r^{2}\left(1+\sum_{j,k=1}^{2}\partial_{r}{\bf G}^{0}_{j,k}(s^{\prime},0,\omega)r\omega_{j}\omega_{k}+O(r^{2})\right)

where 𝐆j,k0(s,r,ω):=𝐆j,k(s,s+rω),{\bf G}^{0}_{j,k}(s^{\prime},r,\omega):={\bf G}_{j,k}(s^{\prime},s^{\prime}+r\omega), is even in (r,ω)(r,ω)(r,\omega)\mapsto-(r,\omega) and O(r2)O(r^{2}) is a smooth function of (s,r,ω)(s^{\prime},r,\omega) which is even and vanishes to second order as r0r\to 0. Observe that ωr𝐆j,k0(s,0,ω)\omega\mapsto\partial_{r}{\bf G}^{0}_{j,k}(s^{\prime},0,\omega) is odd in ω\omega.

We now need to argue that α,βr𝐆j,k0(0,0,ω)ωαωβ=0\sum_{\alpha,\beta}\partial_{r}{\bf G}^{0}_{j,k}(0,0,\omega)\omega_{\alpha}\omega_{\beta}=0. Setting s=0s^{\prime}=0 in the above identity and using the fact that we are using boundary normal coordinates we have

r2=dh(x(0;x0),x(t;x0))2dg(x(0;x0),x(t;x0))2=r2(1+j,k=12r𝐆j,k0(0,0,ω)rωjωk+O(r2)).r^{2}=d_{h}(x(0;x_{0}),x(t^{\prime};x_{0}))^{2}\geq d_{g}(x(0;x_{0}),x(t^{\prime};x_{0}))^{2}=r^{2}\left(1+\sum_{j,k=1}^{2}\partial_{r}{\bf G}^{0}_{j,k}(0,0,\omega)r\omega_{j}\omega_{k}+O(r^{2})\right).

Subtracting off the r2r^{2} terms and dividing by r3r^{3} we see that as r0r\to 0,

j,k=12r𝐆j,k0(0,0,ω)ωjωk0\sum_{j,k=1}^{2}\partial_{r}{\bf G}^{0}_{j,k}(0,0,\omega)\omega_{j}\omega_{k}\leq 0

for all ωS1\omega\in S^{1}. We now use the fact that r𝐆j,k(0,0,ω)\partial_{r}{\bf G}_{j,k}(0,0,\omega) is odd to see that

j,k=12r𝐆j,k0(0,0,ω)ωjωk=0.\sum_{j,k=1}^{2}\partial_{r}{\bf G}^{0}_{j,k}(0,0,\omega)\omega_{j}\omega_{k}=0.

Just as how Lemma 2.2 and Corollary 2.3 followed from Lemma 2.1, we have the following:

Corollary 2.5.

Using the expression r=|st|h(s)r=|s^{\prime}-t^{\prime}|_{h(s^{\prime})} and t=s+rωt^{\prime}=s^{\prime}+r\omega, we have that

dg(x(s;x0),x(t;x0))1=r1+O(s)+O(r),\displaystyle d_{g}(x(s^{\prime};x_{0}),x(t^{\prime};x_{0}))^{-1}=r^{-1}+O(s^{\prime})+O(r),

where O(r)O(r) (respectively O(s)O(s^{\prime})) denotes smooth function of (s,r,ω)𝔻ρ××Ss𝔻ρ(s^{\prime},r,\omega)\in\mathbb{D}_{\rho}\times\mathbb{R}\times S_{s^{\prime}}\mathbb{D}_{\rho} which vanishes to first order as r0r\to 0 (respectively s0s^{\prime}\to 0).

Corollary 2.6.

For ϵ>0\epsilon>0 sufficiently small we have that

dg(xϵ(s;x0),xϵ(t;x0))1=ϵ1r1+ϵr1A(ϵ,s,r,ω)\displaystyle d_{g}(x^{\epsilon}(s^{\prime};x_{0}),x^{\epsilon}(t^{\prime};x_{0}))^{-1}=\epsilon^{-1}r^{-1}+\epsilon r^{-1}A(\epsilon,s^{\prime},r,\omega)

for some smooth function AA in the variables (ϵ,s,r,ω)[0,ϵ0]×𝔻××S1(\epsilon,s^{\prime},r,\omega)\in[0,\epsilon_{0}]\times\mathbb{D}\times\mathbb{R}\times S^{1}. Here we use r=|st|r=|s^{\prime}-t^{\prime}| and t=s+rωt^{\prime}=s+r\omega.

Lemma 2.7.

In the coordinate given by y=x(s;x0)y=x(s^{\prime};x_{0}),

νydg(x0,y)=λ1(x0)s12+λ2(x0)s222|s|+O(|s|2)\partial_{\nu_{y}}d_{g}(x_{0},y)=\frac{\lambda_{1}(x_{0})s_{1}^{2}+\lambda_{2}(x_{0})s_{2}^{2}}{2|s^{\prime}|}+O(|s^{\prime}|^{2})

where O(|s|2)O(|s^{\prime}|^{2}) denotes a smooth function of the variables (|s|,s|s|)(|s^{\prime}|,\frac{s^{\prime}}{|s^{\prime}|}) which vanishes to order 2 at the origin.

Proof.

By Gauss Lemma, for yy near x0x_{0}, gradydg(x0,y)=γ˙(dg(x0,y)){\rm grad}_{y}d_{g}(x_{0},y)=\dot{\gamma}(d_{g}(x_{0},y)) where γ()\gamma(\cdot) is the unit speed geodesic in (M,g)(M,g) from x0x_{0} to yy. Therefore we have that

(2.10) νydg(x0,y)=νy,expy;g1(x0)|expy;g1(x0)|gg.\displaystyle\partial_{\nu_{y}}d_{g}(x_{0},y)=-\left\langle\nu_{y},\frac{{\rm{exp}}_{y;g}^{-1}(x_{0})}{|{\rm{exp}}_{y;g}^{-1}(x_{0})|_{g}}\right\rangle_{g}.

In the coordinates given by sx(s;x0)s\mapsto x(s;x_{0}) the Christoffel symbols are

(2.11) Γ3,33=Γα,33=Γ3,α3=0,Γα,β3=123hα,β.\displaystyle\Gamma_{3,3}^{3}=\Gamma_{\alpha,3}^{3}=\Gamma_{3,\alpha}^{3}=0,\ \ \Gamma_{\alpha,\beta}^{3}=-\frac{1}{2}\partial_{3}h_{\alpha,\beta}.

Choose (s^,0)3(\hat{s}^{\prime},0)\in\mathbb{R}^{3} so that x(s^,0;x0)=yx(\hat{s}^{\prime},0;x_{0})=y. By Lemma 2.4, dg(y,x0)d_{g}(y,x_{0}) is a smooth function of (|s^|,s^|s^|)(|\hat{s}^{\prime}|,\frac{\hat{s}^{\prime}}{|\hat{s}^{\prime}|}) when we write y=x(s^;x0)y=x(\hat{s}^{\prime};x_{0}) in these coordinates. Let V(y):=expy;g1(x0)|expy;g1(x0)|g=j=13Vj(s^)jV(y):=\frac{{\rm{exp}}_{y;g}^{-1}(x_{0})}{|{\rm{exp}}_{y;g}^{-1}(x_{0})|_{g}}=\sum_{j=1}^{3}V_{j}(\hat{s}^{\prime})\partial_{j} be the unique unit vector over yy so that the (M,g)(M,g) unit velocity geodesic in these coordinates starting at yy with initial direction V(y)V(y) reaches x0x_{0} in time dg(y,x0)d_{g}(y,x_{0}). In the coordinates given by sx(s;x0)s^{\prime}\mapsto x(s^{\prime};x_{0}), we want to argue that Vj(s^)V_{j}(\hat{s}^{\prime}) are smooth functions of (|s^|,s^|s^|)(|\hat{s}^{\prime}|,\frac{\hat{s}^{\prime}}{|\hat{s}^{\prime}|}). To do so, observe that in gg-geodesic coordinates centered around x0x_{0} this is of course the case. The result for any other coordinate system can then be obtained via a change of variable.

Note that the outward pointing normal is given by 3-\partial_{3} since s3>0s_{3}>0 in MM. Using (2.10) and the expression for the metric (2.9) we have that in the chosen boundary normal coordinate system

(2.12) νydg(x0,y)=V3.\displaystyle\partial_{\nu_{y}}d_{g}(x_{0},y)=V_{3}.

After time τ\tau, the geodesic with initial position (s^,0)(\hat{s}^{\prime},0) and initial unit velocity VV can be written in the s=(s,s3)s=(s^{\prime},s_{3}) coordinate as

s(τ)=(s^,0)+τV+r(τ;V)s(\tau)=(\hat{s}^{\prime},0)+\tau V+r(\tau;V)

for some remainder r=(r1,r2,r3)r=(r_{1},r_{2},r_{3}) which has initial condition r(0)=r˙(0)=0r(0)=\dot{r}(0)=0. Taylor expanding r(;V)r(\cdot;V) we have that

(2.13) s(τ)=(s^,0)+τV+τ22r¨(0;V)+τ3r(τ;V)\displaystyle s(\tau)=(\hat{s}^{\prime},0)+\tau V+\frac{\tau^{2}}{2}\ddot{r}(0;V)+\tau^{3}r^{\prime}(\tau;V)

for some r(τ;V)r^{\prime}(\tau;V) depending smoothly on τ\tau and VV.

Due to (2.12), we are particularly interested in the evolution of the r3r_{3} component which solves the ODE

r¨3(τ;V)\displaystyle\ddot{r}_{3}(\tau;V) =\displaystyle= j,k=13Γj,k3(s(τ))(Vj+r˙j)(Vk+r˙k)\displaystyle-\sum_{j,k=1}^{3}\Gamma^{3}_{j,k}(s(\tau))(V_{j}+\dot{r}_{j})(V_{k}+\dot{r}_{k})
=\displaystyle= 12α,β=123hα,β(s(τ))(Vα+r˙α)(Vβ+r˙β).\displaystyle\frac{1}{2}\sum_{\alpha,\beta=1}^{2}\partial_{3}h_{\alpha,\beta}(s(\tau))(V_{\alpha}+\dot{r}_{\alpha})(V_{\beta}+\dot{r}_{\beta}).

The last equality comes from (2.11).

Now set τ=dg(y,x0)\tau=d_{g}(y,x_{0}) so that s(τ)=0s(\tau)=0, we have from (2.13) that for α=1,2\alpha=1,2,

Vα=s^αdg(y,x0)+dg(y,x0)fαV_{\alpha}=-\frac{\hat{s}_{\alpha}}{d_{g}(y,x_{0})}+d_{g}(y,x_{0})f_{\alpha}

for functions fαf_{\alpha} which is smooth in VV and dg(y,x0)d_{g}(y,x_{0}) and thus smooth functions of the (s^,s^/|s^|)(\hat{s}^{\prime},\hat{s}^{\prime}/|\hat{s}^{\prime}|) variable. Inserting this into (2.2) yields

dg(x,y)22r¨3(0;V)=14α,β3hα,β(0)s^αs^β+dg(y,x)3f\frac{d_{g}(x,y)^{2}}{2}\ddot{r}_{3}(0;V)=\frac{1}{4}\sum_{\alpha,\beta}\partial_{3}h_{\alpha,\beta}(0)\hat{s}_{\alpha}\hat{s}_{\beta}+d_{g}(y,x)^{3}f

for some function ff which is smooth in the variable (|s^|,s^/|s^|)(|\hat{s}^{\prime}|,\hat{s}^{\prime}/|\hat{s}^{\prime}|). Inserting this expression into (2.13) for τ=dg(x0,y)\tau=d_{g}(x_{0},y) we have that

V3(s^)\displaystyle V_{3}(\hat{s}^{\prime}) =\displaystyle= dg(x,y)2r¨3(0;V)+dg(x0,y)2r3(dg(x0,y),V(s))\displaystyle-\frac{d_{g}(x,y)}{2}\ddot{r}_{3}(0;V)+d_{g}(x_{0},y)^{2}r^{\prime}_{3}(d_{g}(x_{0},y),V(s^{\prime}))
=\displaystyle= 14α,β3hα,β(0)dg(x0,y)s^αs^β+dg(y,x0)2(f+r)\displaystyle-\frac{1}{4}\sum_{\alpha,\beta}\frac{\partial_{3}h_{\alpha,\beta}(0)}{d_{g}(x_{0},y)}\hat{s}_{\alpha}\hat{s}_{\beta}+d_{g}(y,x_{0})^{2}(f+r^{\prime})

Using Lemma 2.4 we have that dg(x0,y)1=|s^|1+F(s^|s^|,s)d_{g}(x_{0},y)^{-1}=|\hat{s}^{\prime}|^{-1}+F(\frac{\hat{s}^{\prime}}{|\hat{s}^{\prime}|},s^{\prime}) for some smooth function F(,)F(\cdot,\cdot). We may thus write

V3(s^)=14α,β3hα,β(0)|s^|s^αs^β+O(|s^|2),V_{3}(\hat{s}^{\prime})=-\frac{1}{4}\sum_{\alpha,\beta}\frac{\partial_{3}h_{\alpha,\beta}(0)}{|\hat{s}^{\prime}|}\hat{s}_{\alpha}\hat{s}_{\beta}+O(|\hat{s}^{\prime}|^{2}),

where O(|s^|2)O(|\hat{s}^{\prime}|^{2}) is a smooth function of (s^|s^|,s^)(\frac{\hat{s}^{\prime}}{|\hat{s}^{\prime}|},\hat{s}^{\prime}) which vanishes to order 22 near the origin. Now use the fact that 123hα,β(0)-\frac{1}{2}\partial_{3}h_{\alpha,\beta}(0) is the coordinate expression for the scalar second fundamental form with respect to the normal given by 3\partial_{3} 333recall that we are using the convention where II{\rm{II}} and shape operator are defined with respect to the inward pointing normal at the point x0x_{0} (see Proposition 8.17 of [16]) and our coordinate system x(s;x0)x(s^{\prime};x_{0}) is chosen so that the shape operator is diagonalized at x0x_{0}. Therefore,

V3(s^)=12α,βλ1s^12+λ2s^22|s^|+O(|s^|2).V_{3}(\hat{s}^{\prime})=\frac{1}{2}\sum_{\alpha,\beta}\frac{\lambda_{1}\hat{s}_{1}^{2}+\lambda_{2}\hat{s}_{2}^{2}}{|\hat{s}^{\prime}|}+O(|\hat{s}^{\prime}|^{2}).

In view of (2.12) we have proven the required identity. ∎

Let x0Mx_{0}\in\partial M and define RII(,),RII(,)L(Bh(ρ;x0)Bh(ρ;x0))R_{{\rm{II}}}(\cdot,\cdot),R_{{\rm{II}}*}(\cdot,\cdot)\in L^{\infty}(B_{h}(\rho;x_{0})B_{h}(\rho;x_{0})) by

(2.15) RII(x,y):=IIx(expx;h1y|expx;h1y|h,expx;h1y|expx,h1y|h),\displaystyle R_{{\rm{II}}}(x,y):={\rm{II}}_{x}\left(\frac{{\rm{exp}}_{x;h}^{-1}y}{|{\rm{exp}}_{x;h}^{-1}y|_{h}},\frac{{\rm{exp}}_{x;h}^{-1}y}{|{\rm{exp}}_{x,h}^{-1}y|_{h}}\right),
RII(x,y):=IIx(expx;h1y|expx;h1y|h,expx;h1y|expx,h1y|h).\displaystyle R_{{\rm{II}}*}(x,y):={\rm{II}}_{x}\left(*\frac{{\rm{exp}}_{x;h}^{-1}y}{|{\rm{exp}}_{x;h}^{-1}y|_{h}},*\frac{{\rm{exp}}_{x;h}^{-1}y}{|{\rm{exp}}_{x,h}^{-1}y|_{h}}\right).

Here * is the Hodge star operator associated to the metric hh.

Lemma 2.8.

Let x(;x0):𝔻ρMx(\cdot;x_{0}):\mathbb{D}_{\rho}\to\partial M be a normal coordinate system for (M,h)(\partial M,h) centred around x0Mx_{0}\in\partial M and let hh denote the pullback metric tensor on on 𝔻ρ\mathbb{D}_{\rho} under this coordinate. Then for all s,t𝔻ρs^{\prime},t^{\prime}\in\mathbb{D}_{\rho} sufficiently close to the origin,

expt;h1(s)|expt;h1(s)|h(t)=j=12ωjj+O(t)+O(r),\frac{\exp^{-1}_{t^{\prime};h}(s^{\prime})}{|\exp^{-1}_{t^{\prime};h}(s^{\prime})|_{h(t^{\prime})}}=\sum_{j=1}^{2}\omega_{j}\partial_{j}+O(t^{\prime})+O(r),

where ω:=strS1\omega:=\frac{s^{\prime}-t^{\prime}}{r}\in S^{1} with r:=|st|r:=|s^{\prime}-t^{\prime}| being the Euclidean distance between ss^{\prime} and tt^{\prime}. The O(t)O(t^{\prime}) (resp O(r)O(r)) term denotes a smooth map of (t,ω,r)𝔻ρ×S2×[0,ρ](t^{\prime},\omega,r)\in\mathbb{D}_{\rho}\times S^{2}\times[0,\rho] which vanishes to order 11 as t0t^{\prime}\to 0 (resp r0r\to 0).

Proof.

This comes from the fact that for some matrix Hj,k(s,t)H_{j,k}(s^{\prime},t^{\prime}) smooth in (s,t)(s^{\prime},t^{\prime})

dh(s,t)2=j=12Hj,k(s,t)(sjtj)(sktk),d_{h}(s^{\prime},t^{\prime})^{2}=\sum_{j=1}^{2}H_{j,k}(s^{\prime},t^{\prime})(s_{j}-t_{j})(s_{k}-t_{k}),

where Hj,k(t,t)=hj,k(t)=δj,k+O(|t|2)H_{j,k}(t^{\prime},t^{\prime})=h_{j,k}(t^{\prime})=\delta_{j,k}+O(|t^{\prime}|^{2}). Therefore

dh(s,t)(s,t)=|st|+|st|F(t,st|st|,|st|)d_{h(s^{\prime},t^{\prime})}(s^{\prime},t^{\prime})=|s^{\prime}-t^{\prime}|+|s^{\prime}-t^{\prime}|F\left(t^{\prime},\frac{s^{\prime}-t^{\prime}}{|s^{\prime}-t^{\prime}|},|s^{\prime}-t^{\prime}|\right)

for some smooth function FC(𝔻ρ×S1×[0,r0])F\in C^{\infty}(\mathbb{D}_{\rho}\times S^{1}\times[0,r_{0}]) which is O(t)+O(r)O(t^{\prime})+O(r).

A coordinate calculation yields that expt,h1(s)|expt,h1(s)|h(t)=gradtdh(s,t)\frac{\exp^{-1}_{t^{\prime},h}(s^{\prime})}{|\exp^{-1}_{t^{\prime},h}(s^{\prime})|_{h(t^{\prime})}}={\rm grad}_{t^{\prime}}d_{h}(s^{\prime},t^{\prime}) so

expt,h1(s)|expt,h1(s)|h(t)=j,k=12hj,k(t)tjdhθ(s,t)(s,t)k=jωjj+O(t)+O(r).\frac{\exp^{-1}_{t^{\prime},h}(s^{\prime})}{|\exp^{-1}_{t^{\prime},h}(s^{\prime})|_{h(t^{\prime})}}=\sum_{j,k=1}^{2}h_{j,k}(t^{\prime})\partial_{t_{j}}d_{h^{\theta}(s^{\prime},t^{\prime})}(s^{\prime},t^{\prime})\partial_{k}=\sum_{j}\omega_{j}\partial_{j}+O(t^{\prime})+O(r).

Corollary 2.9.

Let x0Mx_{0}\in\partial M and λ1(x0)\lambda_{1}(x_{0}) and λ2(x0)\lambda_{2}(x_{0}) be the eigenvalues of the shape operator at x0x_{0}. Then,
i) in the tx(t;x0)t^{\prime}\mapsto x(t^{\prime};x_{0}) coordinate system prescribed at the beginning of this section,

RII(x(t;x0),x(s;x0))(λ1(x0)(s1t1)2|st|2+λ2(x0)(s2t2)2|st|2)=O(r)+O(t)R_{\rm{II}}(x(t^{\prime};x_{0}),x(s^{\prime};x_{0}))-\left(\lambda_{1}(x_{0})\frac{(s_{1}-t_{1})^{2}}{|s^{\prime}-t^{\prime}|^{2}}+\lambda_{2}(x_{0})\frac{(s_{2}-t_{2})^{2}}{|s^{\prime}-t^{\prime}|^{2}}\right)=O(r)+O(t^{\prime})
RII(x(t;x0),x(s;x0))(λ2(x0)(s1t1)2|st|2+λ1(x0)(s2t2)2|st|2)=O(r)+O(t)R_{{\rm{II}}*}(x(t^{\prime};x_{0}),x(s^{\prime};x_{0}))-\left(\lambda_{2}(x_{0})\frac{(s_{1}-t_{1})^{2}}{|s^{\prime}-t^{\prime}|^{2}}+\lambda_{1}(x_{0})\frac{(s_{2}-t_{2})^{2}}{|s^{\prime}-t^{\prime}|^{2}}\right)=O(r)+O(t^{\prime})

The O(t)O(t^{\prime}) (resp O(r)O(r)) term denotes a smooth function of (t,ω,r)𝔻ρ×S2×[0,ρ](t^{\prime},\omega,r)\in\mathbb{D}_{\rho}\times S^{2}\times[0,\rho] which vanishes to order 11 as t0t^{\prime}\to 0 (resp r0r\to 0).

ii)In the txϵ(t;x0)t^{\prime}\mapsto x^{\epsilon}(t^{\prime};x_{0}) coordinate systems prescribed at the beginning of this section,

RII(xϵ(t;x0),xϵ(s;x0))(λ1(x0)(s1t1)2|st|2+λ2(x0)(s2t2)2|st|2)=ϵRϵ(t,ω,r),R_{\rm{II}}(x^{\epsilon}(t^{\prime};x_{0}),x^{\epsilon}(s^{\prime};x_{0}))-\left(\lambda_{1}(x_{0})\frac{(s_{1}-t_{1})^{2}}{|s^{\prime}-t^{\prime}|^{2}}+\lambda_{2}(x_{0})\frac{(s_{2}-t_{2})^{2}}{|s^{\prime}-t^{\prime}|^{2}}\right)=\epsilon R_{\epsilon}(t^{\prime},\omega,r),
RII(xϵ(t;x0),xϵ(s;x0))(λ2(x0)(s1t1)2|st|2+λ1(x0)(s2t2)2|st|2)=ϵRϵ(t,ω,r),R_{{\rm{II}}*}(x^{\epsilon}(t^{\prime};x_{0}),x^{\epsilon}(s^{\prime};x_{0}))-\left(\lambda_{2}(x_{0})\frac{(s_{1}-t_{1})^{2}}{|s^{\prime}-t^{\prime}|^{2}}+\lambda_{1}(x_{0})\frac{(s_{2}-t_{2})^{2}}{|s^{\prime}-t^{\prime}|^{2}}\right)=\epsilon R_{\epsilon}(t^{\prime},\omega,r),

where Rϵ(t,ω,r)R_{\epsilon}(t^{\prime},\omega,r) is smooth with derivatives of all orders uniformly bounded in ϵ\epsilon.

Proof.

Since xϵ(t;x0)=x(ϵt;x0)x^{\epsilon}(t^{\prime};x_{0})=x(\epsilon t^{\prime};x_{0}) we have that ii) is a consequence of i). For i), we will only prove this for RIIR_{\rm{II}} since the statement for RIIR_{{\rm{II}}*} can be obtained via a rotation.

Recall that the normal coordinate system tx(t;x0)t^{\prime}\mapsto x(t^{\prime};x_{0}) is chosen so that at x0x_{0} the coordinate vectors {t1,t2}\{\partial_{t_{1}},\partial_{t_{2}}\} pushes forward under x(;x0x(\cdot;x_{0} to become of eigenvectors of the shape operator. Because of this we have that the pull-back of IIx(,){\rm{II}}_{x}(\cdot,\cdot) under this coordinate system is given by IIx(t;x0)=j,k=12IIj,kdtjdtk{\rm{II}}_{x(t^{\prime};x_{0})}=\sum_{j,k=1}^{2}{\rm{II}}_{j,k}dt_{j}dt_{k} where IIj,k=λj(x0)δj,k+O(t){\rm{II}}_{j,k}=\lambda_{j}(x_{0})\delta_{j,k}+O(t^{\prime}). Here O(t)O(t^{\prime}) denotes a smooth function of tt^{\prime} which vanishes at the origin. Using the expression derived in Lemma 2.8 for the vector expx;h1y|expx;h1y|h\frac{{\rm{exp}}_{x;h}^{-1}y}{|{\rm{exp}}_{x;h}^{-1}y|_{h}} in the coordinate given by x=x(t;x0)x=x(t^{\prime};x_{0}) and y=x(s;x0)y=x(s^{\prime};x_{0}), we have the desired expression for RII(x(t;x0),x(s;x0))R_{{\rm{II}}}(x(t^{\prime};x_{0}),x(s^{\prime};x_{0})). ∎

2.3. Operator Estimates

In this section we derive Sobolev estimates for some integral kernels we will encounter when obtaining the asymptotic expansions of Theorems 1.2 and 1.3. As these depend on a parameter ϵ>0\epsilon>0 and do not immediately fit into the framework of semiclassical calculus, we need to keep track of the bounds by hand.

It is useful to take the Fourier transform with respect to only some variables. Let u(s,t)u(s^{\prime},t^{\prime}) be a family of tempered distributions in t2t^{\prime}\in\mathbb{R}^{2} depending smoothly on the parameter s2s^{\prime}\in\mathbb{R}^{2}. That is, it is the distribution defined by ϕ2u(s,t)ϕ(t)𝑑t\phi\mapsto\int_{\mathbb{R}^{2}}u(s^{\prime},t^{\prime})\phi(t^{\prime})dt^{\prime} for all ϕS(2)\phi\in S(\mathbb{R}^{2}). We denote by t(u(s,t))(ξ)\mathcal{F}_{t^{\prime}}(u(s^{\prime},t^{\prime}))(\xi^{\prime}) to be the Fourier transform with respect to the tt^{\prime} variable only.

Lemma 2.10.

Let A(s,ω)A(s^{\prime},\omega) be a smooth function on (s,ω)2×S1(s^{\prime},\omega)\in\mathbb{R}^{2}\times S^{1}. For j0j\geq 0 and ξ0\xi^{\prime}\neq 0 we have that for any multi-index α\alpha,

Dsαt(A(s,t|t|)|t|j1)(ξ)=t(DsαA(s,t|t|)|t|j1)(ξ).D^{\alpha}_{s^{\prime}}\mathcal{F}_{t^{\prime}}\left(A\left(s^{\prime},\frac{t^{\prime}}{|t^{\prime}|}\right)|t^{\prime}|^{j-1}\right)\left(\xi^{\prime}\right)=\mathcal{F}_{t^{\prime}}\left(D^{\alpha}_{s^{\prime}}A\left(s^{\prime},\frac{t^{\prime}}{|t^{\prime}|}\right)|t^{\prime}|^{j-1}\right)\left(\xi^{\prime}\right).

Furthermore, t(A(s,t|t|)|t|j1)(ξ)\mathcal{F}_{t^{\prime}}\left(A\left(s^{\prime},\frac{t^{\prime}}{|t^{\prime}|}\right)|t^{\prime}|^{j-1}\right)\left(\xi^{\prime}\right) is jointly smooth in (s,ξ)2×2\{0}(s^{\prime},\xi^{\prime})\in\mathbb{R}^{2}\times\mathbb{R}^{2}\backslash\{0\}.

Proof.

Let χCc(2)\chi\in C^{\infty}_{c}(\mathbb{R}^{2}) be identically 11 near the origin. We can write for any positive integer kk and ξ0\xi^{\prime}\neq 0,

t(A(s,t|t|)|t|j1)(ξ)\displaystyle\mathcal{F}_{t^{\prime}}\left(A\left(s^{\prime},\frac{t^{\prime}}{|t^{\prime}|}\right)|t^{\prime}|^{j-1}\right)\left(\xi^{\prime}\right) =\displaystyle= t(A(s,t|t|)|t|j1χ(t))(ξ)\displaystyle\mathcal{F}_{t^{\prime}}\left(A\left(s^{\prime},\frac{t^{\prime}}{|t^{\prime}|}\right)|t^{\prime}|^{j-1}\chi(t^{\prime})\right)\left(\xi^{\prime}\right)
+\displaystyle+ |ξ|2kt(Δtk(A(s,t|t|)|t|j1(1χ(t))))(ξ).\displaystyle|\xi^{\prime}|^{-2k}\mathcal{F}_{t^{\prime}}\left(\Delta_{t^{\prime}}^{k}\left(A\left(s^{\prime},\frac{t^{\prime}}{|t^{\prime}|}\right)|t^{\prime}|^{j-1}(1-\chi(t^{\prime}))\right)\right)\left(\xi^{\prime}\right).

The first integral is absolutely convergent by the χ(t)\chi(t^{\prime}) cut-off. The second Fourier transform is also absolutely convergent owing to the fact that the integrand is smooth and Δtk\Delta_{t^{\prime}}^{k} makes the integrand decay quickly as tt^{\prime}\to\infty provided kk is chosen large enough. ∎

Lemma 2.11.

Let Aϵ(s,r,ω)A_{\epsilon}(s^{\prime},r,\omega) be a family of Cc(2××S1)C^{\infty}_{c}(\mathbb{R}^{2}\times\mathbb{R}\times S^{1}) whose support is uniformly bounded in ϵ[0,ϵ0]\epsilon\in[0,\epsilon_{0}] and whose derivatives are also uniformly bounded in ϵ\epsilon. Then for all l0l\geq 0,

𝒜ϵf:=2Aϵ(s,r,ω)rlf(s+rω)𝑑r𝑑ω\mathcal{A}_{\epsilon}f:=\int_{\mathbb{R}^{2}}A_{\epsilon}(s^{\prime},r,\omega)r^{l}f(s^{\prime}+r\omega)drd\omega

is a map bounded uniformly in ϵ\epsilon from Hm(2)Hm+1+l(2)H^{m}(\mathbb{R}^{2})\to H^{m+1+l}(\mathbb{R}^{2}).

Proof.

We prove the estimates only for Schwartz functions fS(2)f\in S(\mathbb{R}^{2}). We first expand

(2.16) Aϵ=j=1NrjAϵ(s,0,ω)rj+rN+1Rϵ(s,r,ω).\displaystyle A_{\epsilon}=\sum_{j=1}^{N}\partial_{r}^{j}A_{\epsilon}(s^{\prime},0,\omega)r^{j}+r^{N+1}R_{\epsilon}(s^{\prime},r,\omega).

All terms are smooth in its variables with derivatives uniformly bounded in ϵ\epsilon. Estimating the remainder term in (2.16) is easy:

𝒜Rf:=S10rN+1+lRϵ(s,r,ω)f(s+rω)=2|st|N+lRϵ(s,|st|,st|st|)f(t)𝑑t.\mathcal{A}_{R}f:=\int_{S^{1}}\int_{0}^{\infty}r^{N+1+l}R_{\epsilon}(s^{\prime},r,\omega)f(s^{\prime}+r\omega)=\int_{\mathbb{R}^{2}}|s^{\prime}-t^{\prime}|^{N+l}R_{\epsilon}\left(s^{\prime},|s^{\prime}-t^{\prime}|,\frac{s^{\prime}-t^{\prime}}{|s^{\prime}-t^{\prime}|}\right)f(t^{\prime})dt^{\prime}.

For any positive integer mm we write f=D2mD2mff=\langle D\rangle^{2m}\langle D\rangle^{-2m}f. Choose N>>mN>>m so that there is sufficient smoothness in the integral kernel to integrate by parts the formula

𝒜Rf=2|st|N+lRϵ(s,|st|,st|st|)Dt2mDt2mf(t)𝑑t.\mathcal{A}_{R}f=\int_{\mathbb{R}^{2}}|s^{\prime}-t^{\prime}|^{N+l}R_{\epsilon}\left(s^{\prime},|s^{\prime}-t^{\prime}|,\frac{s^{\prime}-t^{\prime}}{|s^{\prime}-t^{\prime}|}\right)\langle D_{t^{\prime}}\rangle^{2m}\langle D_{t^{\prime}}\rangle^{-2m}f(t^{\prime})dt^{\prime}.

We see from this that for a fixed positive integer mm we may choose NN large enough so that 𝒜R:H2m(2)H2m(2)\mathcal{A}_{R}:H^{-2m}(\mathbb{R}^{2})\to H^{2m}(\mathbb{R}^{2}) is uniformly bounded in ϵ\epsilon.

For the integral involving the main term of (2.16), we write

𝒜jf:=S10rjAϵ(s,0,ω)rj+lf(s+rω).\mathcal{A}_{j}f:=\int_{S^{1}}\int_{0}^{\infty}\partial_{r}^{j}A_{\epsilon}(s^{\prime},0,\omega)r^{j+l}f(s^{\prime}+r\omega).

Let χCc(2)\chi\in C^{\infty}_{c}(\mathbb{R}^{2}) be 11 near the origin and write

(2.17) 𝒜jf=𝒜jχ(D)f+𝒜j(1χ(D))f.\displaystyle\mathcal{A}_{j}f=\mathcal{A}_{j}\chi(D)f+\mathcal{A}_{j}(1-\chi(D))f.

To see the mapping property of the first term of (2.17) we write it out in Cartesian coordinates

𝒜jχ(D)f=2rjAϵ(s,0,t|t|)|t|j1+l(χˇf)(st)dt.\mathcal{A}_{j}\chi(D)f=\int_{\mathbb{R}^{2}}\partial_{r}^{j}A_{\epsilon}\left(s^{\prime},0,\frac{t^{\prime}}{|t^{\prime}|}\right)|t^{\prime}|^{j-1+l}(\check{\chi}*f)(s^{\prime}-t^{\prime})dt^{\prime}.

Since rjAϵ(s,0,t|t|)\partial_{r}^{j}A_{\epsilon}\left(s^{\prime},0,\frac{t^{\prime}}{|t^{\prime}|}\right) is smooth in ss^{\prime} with derivatives bounded uniformly in ϵ\epsilon and χ(D)\chi(D) is smoothing, we have that 𝒜jχ(D):Hm(2)Hm(2)\mathcal{A}_{j}\chi(D):H^{-m}(\mathbb{R}^{2})\to H^{m}(\mathbb{R}^{2}) for any positive integer mm with bound uniform in ϵ\epsilon.

The second term of (2.17) is a pseudodifferential operator with full symbol

aj(s,ξ;ϵ):=(1χ(ξ))t(rjAϵ(s,0,t|t|)|t|j1+l)a_{j}(s^{\prime},\xi^{\prime};\epsilon):=(1-\chi(\xi^{\prime}))\mathcal{F}_{t^{\prime}}\left(\partial_{r}^{j}A_{\epsilon}\left(s^{\prime},0,\frac{t^{\prime}}{|t^{\prime}|}\right)|t^{\prime}|^{j-1+l}\right)

and away from ξ=0\xi=0 we can deduce from Lemma 2.10 that

t(rjAϵ(s,0,t|t|)|t|j1+l)=|ξ|jl1a~j(s,ξ/|ξ|;ϵ)\mathcal{F}_{t^{\prime}}\left(\partial_{r}^{j}A_{\epsilon}\left(s^{\prime},0,\frac{t^{\prime}}{|t^{\prime}|}\right)|t^{\prime}|^{j-1+l}\right)=|\xi^{\prime}|^{-j-l-1}\tilde{a}_{j}(s^{\prime},\xi^{\prime}/|\xi^{\prime}|;\epsilon)

for some a~j(s,ω;ϵ)Cc(2×S1)\tilde{a}_{j}(s^{\prime},\omega;\epsilon)\in C_{c}^{\infty}(\mathbb{R}^{2}\times S^{1}) with derivatives uniformly bounded in ϵ\epsilon. The operator Dm+l+1𝒜j(1χ(D))Dm\langle D\rangle^{m+l+1}\mathcal{A}_{j}(1-\chi(D))\langle D\rangle^{-m} then has full symbol in S0S^{0} with symbol seminorms bounded uniformly in ϵ\epsilon. We can now apply Calderón-Vailancourt Theorem to deduce that 𝒜j(1χ(D))\mathcal{A}_{j}(1-\chi(D)) is bounded uniformly in ϵ\epsilon from Hm(2)Hm+l+1(2)H^{m}(\mathbb{R}^{2})\to H^{m+l+1}(\mathbb{R}^{2}). ∎

2.4. Symbol Computations

Compute symbol by taking the Fourier transform and multiply by 2π2\pi. We compute the principal symbols of some of the main operators which we will encounter. The following list of inverse Fourier transforms will be useful for later computations and we will leave its proof to the reader:

Lemma 2.12.

In 2\mathbb{R}^{2} with ξ=(ξ1,ξ2)\xi=(\xi_{1},\xi_{2}) and x=(x1,x2)x=(x_{1},x_{2}) one has that for |ξ|1|\xi|\geq 1,
i) 1(log|x|)(ξ)=2π|ξ|2{\mathcal{F}}^{-1}(\log|x|)(\xi)=-2\pi|\xi|^{-2}1(|x|1)(ξ)=2π|ξ|1{\mathcal{F}}^{-1}(|x|^{-1})(\xi)=2\pi|\xi|^{-1}
ii) 1(xj|x|1)(ξ)=2πiξj|ξ|3{\mathcal{F}}^{-1}(x_{j}|x|^{-1})(\xi)=2\pi i\xi_{j}|\xi|^{-3}
iii) 1(xj2|x|3)(ξ)=2πξk2|ξ|3{\mathcal{F}}^{-1}(x_{j}^{2}|x|^{-3})(\xi)=2\pi\xi_{k}^{2}|\xi|^{-3}, kjk\neq j
iv) 1(xj2|x|2)(ξ)=2π(ξk2ξj2)|ξ|4{\mathcal{F}}^{-1}(x_{j}^{2}|x|^{-2})(\xi)=2\pi(\xi_{k}^{2}-\xi_{j}^{2})|\xi|^{-4}, kjk\neq j.

Remark 2.13.

Note that we ignore the behaviour of 1()\mathcal{F}^{-1}(\cdot) near ξ=0\xi=0 as they are irrelevant to the principal symbol computations we are interested in.

Lemma 2.14.

Let AA be a pseudodifferential operator on M\partial M whose singularity along the diagonal is given by dg(x,y)1d_{g}(x,y)^{-1}. Then σ(A)(x,ξ)2π|ξ|h(x)1\sigma(A)(x,\xi)\equiv 2\pi|\xi|^{-1}_{h(x)}.

Proof.

We compute the symbol at x0Mx_{0}\in\partial M using the normal coordinate tx(t;x0)t^{\prime}\mapsto x(t^{\prime};x_{0}). By Corollary 2.5 d(x(t;x0),x0)1=|t|1+O(|t|)d(x(t^{\prime};x_{0}),x_{0})^{-1}=|t^{\prime}|^{-1}+O(|t^{\prime}|). By Lemma 2.12, the inverse Fourier transform of the leading order singularity is 2π|ξ|12\pi|\xi|^{-1} for ξ\xi large and therefore σ(A)(x,ξ)2π|ξ|h(x)1\sigma(A)(x,\xi)\equiv 2\pi|\xi|^{-1}_{h(x)}. ∎

Lemma 2.15.

Let AA be a pseudodifferential operator on M\partial M whose singularity near the diagonal is given by νydg(x,y)1x,yM×M\partial_{\nu_{y}}d_{g}(x,y)^{-1}\mid_{x,y\in\partial M\times\partial M}. Then σ(A)(x0,ξ)πIIx0(ξ)|ξ|h(x0)3\sigma(A)(x_{0},\xi)\equiv-\pi\frac{{\rm{II}}_{x_{0}}(*\xi^{\sharp})}{|\xi|_{h(x_{0})}^{3}} where ξξ\xi\mapsto\xi^{\sharp} denotes the musical isomorphism from TMT^{*}\partial M to TMT\partial M induced by the boundary metric hh and * is the Hodge star operator in this metric. Here we use IIx(V){\rm{II}}_{x}(V) to denote the quadratic form IIx(V,V){\rm{II}}_{x}(V,V) for VTxMV\in T_{x}\partial M.

Proof.

Using Lemma 2.7 we see that in normal coordinates given by x(s;x0)x(s^{\prime};x_{0}) the leading order singularity of νydg(x0,y=x(s;x0))\partial_{\nu_{y}}d_{g}(x_{0},y=x(s^{\prime};x_{0})) is given by λ1s12+λ2s222|s|3-\frac{\lambda_{1}s_{1}^{2}+\lambda_{2}s_{2}^{2}}{2|s^{\prime}|^{3}}. By Lemma 2.12 we have that

1(λ1s12+λ2s222|s|3)=πλ1ξ22+λ2ξ12|ξ|3.{\mathcal{F}}^{-1}\left(-\frac{\lambda_{1}s_{1}^{2}+\lambda_{2}s_{2}^{2}}{2|s^{\prime}|^{3}}\right)=-\pi\frac{\lambda_{1}\xi_{2}^{2}+\lambda_{2}\xi_{1}^{2}}{|\xi|^{3}}.

This is precisely the normal coordinate expression for πIIx0(ξ)|ξ|h(x0)3-\pi\frac{{\rm{II}}_{x_{0}}(*\xi^{\sharp})}{|\xi|_{h(x_{0})}^{3}}. ∎

Proposition 2.16.

Let H(x)H(x) denote the mean curvature of M\partial M at xx, IIx{\rm{II}}_{x} the second fundamental form of M\partial M at xMx\in\partial M, and IIx(V):=IIx(V,V){\rm{II}}_{x}(V):={\rm{II}}_{x}(V,V) for VTxMV\in T_{x}\partial M. Define

(2.18) K(x,y):=H(x)πlogdh(y,x)π4(IIx(expx1(y)|expx1(y)|h)IIx(expx1(y)|expx1(y)|h)),\displaystyle\quad\quad K(x,y):={H(x)\pi}\log d_{h}(y,x)-\frac{\pi}{4}\left({\rm{II}}_{x}\left(\frac{\exp_{x}^{-1}(y)}{|\exp_{x}^{-1}(y)|_{h}}\right)-{\rm{II}}_{x}\left(\frac{*\exp_{x}^{-1}(y)}{|\exp_{x}^{-1}(y)|_{h}}\right)\right),

where * is the Hodge star operator for the metric hh. Let A:C(M)𝒟(M)A:C^{\infty}(\partial M)\to{\mathcal{D}}^{\prime}(\partial M) be the operator defined by

A:uMK(x,y)u(y)dvolh.A:u\mapsto\int_{\partial M}K(x,y)u(y){\rm dvol}_{h}.

Then AΨcl2(M)A\in\Psi_{cl}^{-2}(\partial M) with principal symbol a(x,ξ)C(TM\{0})a(x,\xi)\in C^{\infty}(T^{*}\partial M\backslash\{0\}) given by

a(x,ξ)2π2|ξ|h4IIx(ξ),a(x,\xi)\equiv-\frac{2\pi^{2}}{|\xi|_{h}^{4}}{\rm{II}}_{x}(*\xi^{\sharp}),

where ξξTM\xi\mapsto\xi^{\sharp}\in T\partial M denotes the raising of index with respect to the metric hh on M\partial M.

Proof.

To see that AA is a classical Ψ\PsiDO, we use Lemma 2.1 and Corollary 2.9 to see that the coordinate expression for the integral kernel K(x,y)K(x,y) satisfies the polyhomogeneous conditions of Prop 2.8 in Chapter 7 of [34]. Therefore AΨcl2(M)A\in\Psi^{-2}_{cl}(\partial M).

The principal symbol computation is done using normal coordinates. Fix x0Mx_{0}\in\partial M and denote by

tx(t;x0):=expx0;h(t)t^{\prime}\mapsto x(t^{\prime};x_{0}):=\exp_{x_{0};h}(t^{\prime})

the normal coordinate system around x0x_{0}. By a rotation we can choose the coordinates so that tjx(0;x0)Tx0M\partial_{t_{j}}x(0;x_{0})\in T_{x_{0}}\partial M is an eigenvector of the shape operator at x0x_{0} with eigenvalue λj\lambda_{j}.

According to Lemma 2.1 and Corollary 2.9, in these coordinates the terms of K(x0,y)K(x_{0},y) can be expressed as

(2.19) K(x0,x(t;x0))=λ1+λ22πlog|t|π4(λ1t12+λ2t22|t|2λ1t22+λ2t12|t|2)\displaystyle K(x_{0},x(t^{\prime};x_{0}))=\frac{\lambda_{1}+\lambda_{2}}{2}\pi\log|t^{\prime}|-\frac{\pi}{4}\left(\frac{\lambda_{1}t_{1}^{2}+\lambda_{2}t_{2}^{2}}{|t^{\prime}|^{2}}-\frac{\lambda_{1}t_{2}^{2}+\lambda_{2}t_{1}^{2}}{|t^{\prime}|^{2}}\right)

for tt^{\prime} close to the origin. Computing the principal symbol a(x0,ξ)a(x_{0},\xi) amounts to taking the inverse Fourier transform of the above expression, and observe the behaviour as |ξ||\xi|\to\infty. Use the formula in Lemma 2.12 , we obtain

a(x0,ξ)2π2|ξ|4(λ1ξ22+λ2ξ12)=2π2|ξ|h4IIx(ξ).a(x_{0},\xi)\equiv-\frac{2\pi^{2}}{|\xi|^{4}}(\lambda_{1}\xi_{2}^{2}+\lambda_{2}\xi_{1}^{2})=-\frac{2\pi^{2}}{|\xi|_{h}^{4}}{\rm{II}}_{x}(*\xi^{\sharp}).

The last equality holds due to the fact that we are using normal coordinates. ∎

3. Proof of Proposition 1.1

In this section we use layer potential methods to pick out the singularity structure of the Neumann Green’s function at the boundary. Assume without loss of generality that MM is an open subset of a compact Riemannian manifold (M~,g)(\tilde{M},g) without boundary. Choose MM~M^{\prime}\subset\tilde{M} a manifold with boundary which compactly contains MM. For all FC0(M)F\in C^{\infty}_{0}(M^{\prime}), standard elliptic theory shows that there exists a unique solution UFC(M)U_{F}\in C^{\infty}(M^{\prime}) to

ΔgUF=F,UFM=0.\Delta_{g}U_{F}=F,\ \ U_{F}\mid_{\partial M^{\prime}}=0.

The map FUFF\mapsto U_{F} is a continuous linear operator from C0(M)𝒟(M)C^{\infty}_{0}(M^{\prime})\to{\mathcal{D}}^{\prime}(M^{\prime}) and is therefore given by a Schwartz kernel E(x,y)𝒟(M×M)E(x,y)\in{\mathcal{D}}^{\prime}(M^{\prime}\times M^{\prime}) which we call the Green’s function. Note that for any uC0(M)u\in C_{0}^{\infty}(M^{\prime}), if we fix xMx\in M^{\prime} then by definition

u(x)=MΔgu(y)E(x,y)𝑑volg(y)=u(),ΔgE(x,).u(x)=\int_{M^{\prime}}\Delta_{g}u(y)E(x,y)d{\rm vol}_{g}(y)=\langle u(\cdot),\Delta_{g}E(x,\cdot)\rangle.

We formally write

(3.1) ΔgE(x,)=δx()onM\displaystyle\Delta_{g}E(x,\cdot)=\delta_{x}(\cdot)\ \ {\rm on}\ \ M^{\prime}

Note that if MM is a bounded domain in 3\mathbb{R}^{3} then M~\tilde{M} can be chosen to be the flat torus and E(x,y)E(x,y) can be chosen to be 14π|xy|\frac{-1}{4\pi|x-y|} for the appropriate constant cc\in\mathbb{R}.

Using standard elliptic parametrix construction in normal coordinates we express E(x,y)E(x,y) in the following way.

Lemma 3.1.

For all x,yMx,y\in M^{\prime}

(3.2) E(x,y)=14πdg(x,y)1+Ψcl4(M~).\displaystyle E(x,y)=-\frac{1}{4\pi}d_{g}(x,y)^{-1}+\Psi_{cl}^{-4}(\tilde{M}).

Here Ψcl4(M~)\Psi_{cl}^{-4}(\tilde{M}) denotes the Schwartz kernel of an operator in Ψcl4(M~)\Psi_{cl}^{-4}(\tilde{M}).

Proof.

Let PΨcl2(M~)P\in\Psi_{cl}^{-2}(\tilde{M}) be a parametrix for the Δg\Delta_{g} on the closed compact manifold M~\tilde{M} without boundary meaning that

ΔgP=I+Ψ(M~).\Delta_{g}P=I+\Psi^{-\infty}(\tilde{M}).

By ellipticity, for any χ0Cc(M)\chi_{0}\in C^{\infty}_{c}(M^{\prime}) we have

χ0(x)χ0(y)(E(x,y)P(x,y))Ψ(M~).\chi_{0}(x)\chi_{0}(y)(E(x,y)-P(x,y))\in\Psi^{-\infty}(\tilde{M}).

Therefore it suffices to show that

(3.3) PP2Ψcl4(M~)\displaystyle P-P_{-2}\in\Psi_{cl}^{-4}(\tilde{M})

where P2Ψ2(M~)P_{-2}\in\Psi^{-2}(\tilde{M}) is defined by

(P2u)(x):=M~χ(x,y)4πdg(x,y)1u(y)𝑑volg(y)(P_{-2}u)(x):=\int_{\tilde{M}}\frac{-\chi(x,y)}{4\pi}d_{g}(x,y)^{-1}u(y)d{\rm vol}_{g}(y)

for some smooth function χ(,)C(M~×M~)\chi(\cdot,\cdot)\in C^{\infty}(\tilde{M}\times\tilde{M}) satisfying χ(x,y)=χ(y,x)\chi(x,y)=\chi(y,x), χ(x,y)=1\chi(x,y)=1 if dg(x,y)<InjM~/4d_{g}(x,y)<{\rm Inj}_{\tilde{M}}/4 and

supp(χ(,)){(x,y)dg(x,y)<InjM~/2}.{\rm supp}(\chi(\cdot,\cdot))\subset\subset\{(x,y)\mid d_{g}(x,y)<{\rm Inj}_{\tilde{M}}/2\}.

Here InjM~{\rm Inj}_{\tilde{M}} is the injectivity radius of the closed compact Riemannian manifold (M~,g)(\tilde{M},g).

By elliptic regularity (3.3) is equivalent to showing that

(3.4) ΔgP2IΨcl2(M~).\displaystyle\Delta_{g}P_{-2}-I\in\Psi_{cl}^{-2}(\tilde{M}).

Taking the adjoint and use the self-adjointness of both Δg\Delta_{g} and P2P_{-2}, this is the same as

(3.5) P2ΔgIΨcl2(M~).\displaystyle P_{-2}\Delta_{g}-I\in\Psi_{cl}^{-2}(\tilde{M}).

Using the principal symbol map defined in (2.7) it amounts to showing that

(3.6) σ1(P2ΔgI)=0\displaystyle\sigma_{-1}(P_{-2}\Delta_{g}-I)=0

as an element of the quotient space Scl1/Scl2S^{-1}_{cl}/S^{-2}_{cl}. In fact, since the symbol is classical, we now choose σ1(P2ΔgI)\sigma_{-1}(P_{-2}\Delta_{g}-I) to be the representative in the equivalence class which is positively homogeneous of degree 1-1.

For each y0M~y_{0}\in\tilde{M} and covector η0Sy0M~\eta_{0}\in S^{*}_{y_{0}}\tilde{M} in the unit cosphere bundle we will show that

(3.7) |σ1(P2ΔgI)(y0,τη0)|Cy0,η0τ2\displaystyle|\sigma_{-1}(P_{-2}\Delta_{g}-I)(y_{0},\tau\eta_{0})|\leq C_{y_{0},\eta_{0}}\tau^{-2}

as τ\tau\to\infty. Homogeneity would then ensure that σ1(P2ΔgI)(y,η)=0\sigma_{-1}(P_{-2}\Delta_{g}-I)(y,\eta)=0 for all (y,η)TM~(y,\eta)\in T^{*}\tilde{M} which would then ensure (3.6).

To this end let V1,V2,V3Sy0M~V_{1},V_{2},V_{3}\in S_{y_{0}}\tilde{M} be three orthonormal vectors and choose normal coordinate given by Φ(t):=expy0(j=13tjVj)\Phi(t):=\exp_{y_{0}}(\sum_{j=1}^{3}t_{j}V_{j}) for |t|3<InjM~/2|t|_{\mathbb{R}^{3}}<{\rm Inj}_{\tilde{M}}/2. Let χ3Cc(3)\chi_{\mathbb{R}^{3}}\in C_{c}^{\infty}(\mathbb{R}^{3}) take the value 11 in an open set containing the origin but supp(χ3){t3|t|3<InjM~/2}{\rm supp}(\chi_{\mathbb{R}^{3}})\subset\subset\{t\in\mathbb{R}^{3}\mid|t|_{\mathbb{R}^{3}}<{\rm Inj}_{\tilde{M}}/2\}. Similarly let χM~C(3)\chi_{\tilde{M}}\in C^{\infty}(\mathbb{R}^{3}) take the value 11 in an open set containing y0y_{0} but supp(χM~){xdg(x,y0)<InjM~/2}{\rm supp}(\chi_{\tilde{M}})\subset\subset\{x\mid d_{g}(x,y_{0})<{\rm Inj}_{\tilde{M}}/2\}.

Define the pullback operators A,B:C(3)C(3)A,B:C^{\infty}(\mathbb{R}^{3})\to C^{\infty}(\mathbb{R}^{3}) by

A:uΦ(χM~P2Φ(χ3u)),B:uΦ(χM~ΔgΦ(χ3u))A:u\mapsto\Phi_{*}\left(\chi_{\tilde{M}}P_{-2}\Phi^{*}\left(\chi_{\mathbb{R}^{3}}u\right)\right),\ B:u\mapsto\Phi_{*}\left(\chi_{\tilde{M}}\Delta_{g}\Phi^{*}\left(\chi_{\mathbb{R}^{3}}u\right)\right)

where Φ\Phi^{*} and Φ\Phi_{*} are pullback by Φ\Phi and Φ1\Phi^{-1} respectively.

Thanks to the invariance of the principal symbol map under symplectomorphism, we have

(3.8) σ1(P2ΔgI)(y0,Φξ)=σ1(ABI)(0,ξ)\displaystyle\sigma_{-1}(P_{-2}\Delta_{g}-I)(y_{0},\Phi^{*}\xi)=\sigma_{-1}(AB-I)(0,\xi)

for all ξT03\xi\in T^{*}_{0}\mathbb{R}^{3}. We see then that (3.6) amounts to showing that ABIAB-I satisfies

(3.9) σ1(ABI)(0,ξ)=0.\displaystyle\sigma_{-1}(AB-I)(0,\xi)=0.

Let a(t,ξ)a(t,\xi) and b(t,ξ)b(t,\xi) be the full symbol of AA and BB respectively. The full symbol of AA can be computed by the formula

a(t,ξ)=eξ(t)(Aeξ)(t)a(t,\xi)=e_{-\xi}(t)\left(Ae_{\xi}\right)(t)

where eξ(t):=eitξe_{\xi}(t):=e^{-it\cdot\xi}. Since we are using the normal coordinate around y0y_{0}, dg(y0,Φ(t))=|t|3d_{g}(y_{0},\Phi(t))=|t|_{\mathbb{R}^{3}} and

ΦdVolg=|g|dt=dt+H2(t)dt\Phi_{*}d{\rm Vol}_{g}=\sqrt{|g|}dt=dt+H_{2}(t)dt

where H2(t)H_{2}(t) is a smooth function vanishing to order 22 at t=0t=0. So

(Aeξ)(0)=34π|t|eitξ𝑑t+3c|t|eitξH~2(t)𝑑t(Ae_{\xi})(0)=\int_{\mathbb{R}^{3}}\frac{-4\pi}{|t|}e^{-it\cdot\xi}dt+\int_{\mathbb{R}^{3}}\frac{c}{|t|}e^{-it\cdot\xi}\tilde{H}_{2}(t)dt

for some smooth and compactly supported function H~2(t)\tilde{H}_{2}(t) vanishing to order 22 at t=0t=0. Computing the first term directly and treat the second term by expanding H2(t)H_{2}(t) using Taylor expansion we see that

(3.10) a(0,ξ)=(Aeξ)(0)=|ξ|2+Scl4(T3).\displaystyle a(0,\xi)=(Ae_{\xi})(0)=|\xi|^{-2}+S_{cl}^{-4}(T^{*}\mathbb{R}^{3}).

Since BB is the Laplace operator in the coordinate given by Φ\Phi we have that

(3.11) b(t,ξ)=j,k=13gj,k(t)ξjξk+1|g|j,k=13tj(|g|gj,k)(t)ξk\displaystyle b(t,\xi)=\sum_{j,k=1}^{3}g^{j,k}(t)\xi_{j}\xi_{k}+\frac{1}{\sqrt{|g|}}\sum_{j,k=1}^{3}\partial_{t_{j}}(\sqrt{|g|}g^{j,k})(t)\xi_{k}

Composition calculus gives that if c(x,ξ)c(x,\xi) is the full symbol of the operator ABAB then

(3.12) c(0,ξ)=a(0,ξ)b(0,ξ)+ij=13ξja(0,ξ)tjb(t,ξ)t=0+Scl2.\displaystyle c(0,\xi)=a(0,\xi)b(0,\xi)+-i\sum_{j=1}^{3}\partial_{\xi_{j}}a(0,\xi)\partial_{t_{j}}b(t,\xi)\mid_{t=0}+S_{cl}^{-2}.

Substituting into (3.12) the expression we have in (3.10), (3.11), and the fact that in normal coordinates gj,k(t)=δj,k+O(|t|2)g^{j,k}(t)=\delta^{j,k}+O(|t|^{2}) for tt in a neighbourhood of the origin we have that the second term in (3.12) drops out. So the full symbol of ABIAB-I at the point (0,ξ)T3(0,\xi)\in T^{*}\mathbb{R}^{3} is

c(0,ξ)1Scl2.c(0,\xi)-1\in S_{cl}^{-2}.

In light of (3.8), for each fixed ξTy0M~\xi\in T^{*}_{y_{0}}\tilde{M},

|σ1(P2ΔgI)(y0,τΦξ)|<Cξτ2|\sigma_{-1}(P_{-2}\Delta_{g}-I)(y_{0},\tau\Phi^{*}\xi)|<C_{\xi}\tau^{-2}

as τ\tau\to\infty. Therefore (3.7) is verified.

For all fC(M)f\in C^{\infty}(\partial M) we define as in [34] the operators S,NΨcl1(M)S,N\in\Psi_{cl}^{-1}(\partial M) by the following

(3.13) Sf(x):=ME(x,y)f(y)volh,Nf(x):=2MνyE(x,y)f(y)volh\displaystyle Sf(x):=\int_{\partial M}E(x,y)f(y){\rm vol}_{h},\ \ Nf(x):=2\int_{\partial M}\partial_{\nu_{y}}E(x,y)f(y){\rm vol}_{h}

for xMx\in\partial M. Note that Nf(x)Nf(x) is different from (see [34] Chapt 7 Sect 11)

limxM,xM2MνyE(x,y)f(y)dy=f(x)+Nf(x).\lim_{x\to\partial M,x\in M}2\int_{\partial M}\partial_{\nu_{y}}E(x,y)f(y)dy=f(x)+Nf(x).

Modulo lower order pseudodifferential operator, SS and NN are given by the integral kernels dg(x,y)1d_{g}(x,y)^{-1} and νydg(x,y)1\partial_{\nu_{y}}d_{g}(x,y)^{-1} respectively. Indeed, using (3.2) and equation (11.14) on page 38 of [34], we see that for (x,y)M×M(x,y)\in\partial M\times\partial M in a neighbourhood of the diagonal,

(3.14) S=14πdg(x,y)1+Ψcl3(M),N=12πνydg(x,y)1+Ψcl2(M).\displaystyle S=-\frac{1}{4\pi}d_{g}(x,y)^{-1}+\Psi_{cl}^{-3}(\partial M),\ \ N=-\frac{1}{2\pi}\partial_{\nu_{y}}d_{g}(x,y)^{-1}+\Psi_{cl}^{-2}(\partial M).

Using (3.1) we can construct the so called Neumann Green’s function on MM via the following procedure. For each fixed xMox\in M^{o} we can solve the following Neumann boundary value problem to obtain the correction term C(x,y)C(x,y) as a function of yMy\in M

ΔgC(x,y)=0,,νyC(x,y)=νyE(x,y)1|M|,MC(x,y)dvolh=ME(x,y)dvolh.\Delta_{g}C(x,y)=0,\ ,\ \partial_{\nu_{y}}C(x,y)=\partial_{\nu_{y}}E(x,y)-\frac{1}{|\partial M|},\ \ \int_{\partial M}C(x,y)d{\rm vol}_{h}=\int_{\partial M}E(x,y)d{\rm vol}_{h}.

Setting G(x,y)=E(x,y)+C(x,y)G(x,y)=-E(x,y)+C(x,y) we get, for each fixed xMx\in M the unique solution (as a distribution in zz) G(x,z)G(x,z) to

(3.15) ΔgG(x,z)=δx(z),νzG(x,z)zM=1|M|,MG(x,z)𝑑volh=0.\displaystyle\Delta_{g}G(x,z)=-\delta_{x}(z)\ ,\partial_{\nu_{z}}G(x,z)\mid_{z\in\partial M}=\frac{-1}{|\partial M|},\ \ \int_{\partial M}G(x,z)d{\rm vol}_{h}=0.

Fix for the time being yzy\neq z in the interior of MM and observe that xG(z,x)x\mapsto G(z,x) is smooth in a neighbourhood of the singularity of the map xG(x,y)x\mapsto G(x,y) and vice versa. Therefore we can integrate by parts the the expression G(z,y)=MG(z,x)ΔxG(y,x)𝑑xG(z,y)=-\int_{M}G(z,x)\Delta_{x}G(y,x)dx to obtain

G(z,y)G(y,z)=MG(y,x)νxG(z,x)G(z,x)νxG(y,x)volh(x).G(z,y)-G(y,z)=\int_{\partial M}G(y,x)\partial_{\nu_{x}}G(z,x)-G(z,x)\partial_{\nu_{x}}G(y,x){\rm vol}_{h}(x).

The boundary and orthogonality conditions in (3.15) ensures that the right side vanishes so we have

(3.16) G(z,y)=G(y,z).\displaystyle G(z,y)=G(y,z).

Let Λ:Hk(M)Hk1(M)\Lambda:H^{k}(\partial M)\to H^{k-1}(\partial M) denote the Dirichlet-to-Neumann map (see [17] for definition) whose range is precisely a codimension one subspace of Hk1H^{k-1} which annihilates the constant function. By the orthogonality condition in (3.15), the behaviour of G(x,y)G(x,y) is uniquely characterized by its action on the range of Λ\Lambda. To this end, for fC(M)f\in C^{\infty}(\partial M), denote its harmonic extension by ufu_{f}. Integrating by parts the expression 0=MΔguf(x)G(x,y)𝑑x0=\int_{M}\Delta_{g}u_{f}(x)G(x,y)dx for zz in the interior of MM we have

u(y)\displaystyle u(y) =\displaystyle= Mνuf(x)G(x,y)dvolh(x)+1|M|Mfdvolh\displaystyle\int_{\partial M}\partial_{\nu}u_{f}(x)G(x,y){\rm dvol}_{h}(x)+\frac{1}{|\partial M|}\int_{\partial M}f{\rm dvol}_{h}
=\displaystyle= M(Λf)(x)G(x,y)dvolh(x)+1|M|Mfdvolh\displaystyle\int_{\partial M}(\Lambda f)(x)G(x,y){\rm dvol}_{h}(x)+\frac{1}{|\partial M|}\int_{\partial M}f{\rm dvol}_{h}

Observe that any f~C(M)\tilde{f}\in C^{\infty}(\partial M) has a unique decomposition f~=c+Λf\tilde{f}=c+\Lambda f for some constant function cc and fC(M)f\in C^{\infty}(\partial M) satisfying Mf=0\int_{\partial M}f=0. Therefore, using the orthogonality condition of (3.15) and taking the trace of (3) we see that the map

f~(Mf~(x)G(x,y)dvolh(x))|yM\tilde{f}\mapsto\left.\left(\int_{\partial M}\tilde{f}(x)G(x,y){\rm dvol}_{h}(x)\right)\right|_{y\in\partial M}

is well defined and takes C(M)C(M)C^{\infty}(\partial M)\to C^{\infty}(\partial M). We denote this operator by GMG_{\partial M} and its Schwartz kernel by GM(x,y)G_{\partial M}(x,y). Going back to (3) we see that

f=GMΛf+Pf,f=G_{\partial M}\Lambda f+Pf,

where

(3.18) Pf:=|M|1Mf\displaystyle Pf:=|\partial M|^{-1}\int_{\partial M}f

is smoothing. In operator form this is

(3.19) I=GMΛ+P.\displaystyle I=G_{\partial M}\Lambda+P.

Since ΛΨcl1(M)\Lambda\in\Psi^{1}_{cl}(\partial M) is elliptic (see [17]) we can conclude that GMΨcl1(M)G_{\partial M}\in\Psi_{cl}^{-1}(\partial M) which maps Hk(M)Hk+1(M)H^{k}(\partial M)\to H^{k+1}(\partial M) for all kk\in\mathbb{R}. This completes the proof of Proposition 1.1 part i).

Remark 3.2.

A quick way to prove part ii) of Proposition 1.1 would be to observe that (3.19) implies GMG_{\partial M} is a parametrix for Λ\Lambda. The symbol expansion for Λ\Lambda has already been computed in [17] so constructing its parametrix follows from standard pseudodifferential calculus. However, we will choose instead to take the layer potential approach since it iwill be more conducive for future numerical implementations. See Remark 3.3 below.

Applying on the right the single-layered potential SS defined in(3.13) and using identity (11.58) of [34] we have

(3.20) GM=2S+GMN+2PS.\displaystyle G_{\partial M}=-2S+G_{\partial M}N^{*}+2PS.

Iterating this equation and using intertwining property (11.59) of [34] we get

(3.21) GM=2S2NS+Ψcl3(M).\displaystyle G_{\partial M}=-2S-2NS+\Psi_{cl}^{-3}(\partial M).

By (2.8) the principal symbol of the operator NSNS is simply the product of the principal symbols of SS with the principal symbol of NN. The leading singularities of the operators SS and NN are given in (3.14) and the principal symbols of these kernels are computed in Lemmas 2.14 and 2.15. Therefore, using Proposition 2.16, we see that modulo Ψcl3(M)\Psi_{cl}^{-3}(\partial M), the integral kernel of NSNS is given by

18πH(x)logdh(y,x)132π(IIx(expx1(y)|expx1(y)|h)IIx(expx1(y)|expx1(y)|h))\frac{1}{8\pi}{H(x)}\log d_{h}(y,x)-\frac{1}{32\pi}\left({\rm{II}}_{x}\left(\frac{\exp_{x}^{-1}(y)}{|\exp_{x}^{-1}(y)|_{h}}\right)-{\rm{II}}_{x}\left(\frac{*\exp_{x}^{-1}(y)}{|\exp_{x}^{-1}(y)|_{h}}\right)\right)

when x,yMx,y\in\partial M are close to each other.

Inserting this into (3.21) we get that when x,yMx,y\in\partial M are close to each other,

GM(x,y)\displaystyle\quad\quad G_{\partial M}(x,y) =\displaystyle= 12πdg(x,y)114πH(x)logdh(y,x)\displaystyle\frac{1}{2\pi}d_{g}(x,y)^{-1}-\frac{1}{4\pi}{H(x)}\log d_{h}(y,x)
+116π(IIx(expx1(y)|expx1(y)|h)IIx(expx1(y)|expx1(y)|h))+R(x,y),\displaystyle+\frac{1}{16\pi}\left({\rm{II}}_{x}\left(\frac{\exp_{x}^{-1}(y)}{|\exp_{x}^{-1}(y)|_{h}}\right)-{\rm{II}}_{x}\left(\frac{*\exp_{x}^{-1}(y)}{|\exp_{x}^{-1}(y)|_{h}}\right)\right)+R(x,y),

where R(x,y)R(x,y) is the Schwartz kernel of an operator in Ψcl3(M)\Psi_{cl}^{-3}(\partial M) which we call the regular part of G(x,y)G(x,y). Observe that since the principal symbol of RR is in Scl3(TM)S_{cl}^{-3}(T^{*}\partial M) and M\partial M is dimension 22, Sobolev embedding yields that

(3.23) R(,)C0,α(M×M)\displaystyle R(\cdot,\cdot)\in C^{0,\alpha}(\partial M\times\partial M)

for all α<1\alpha<1. The proof of Proposition 1.1 is now complete.

Remark 3.3.

Note that (1.1) peels off the "singular part" of the distribution GM(x,y)G_{\partial M}(x,y) and gives us the representation

GM(x,y)=Gsing(x,y)+R(x,y)G_{\partial M}(x,y)=G_{sing}(x,y)+R(x,y)

with the singularity structure of GsingG_{sing} explicitly given by (1.1). Inserting this representation of GMG_{\partial M} into (3.20) gives the following integral equation for the regular part R(x,y)R(x,y):

R(IN)=Gsing2S+GsingN+2PSR(I-N^{*})=-G_{sing}-2S+G_{sing}N^{*}+2PS

where the operators PP, SS, and NN are given by (3.18) and (3.13).

Since NΨcl1(M)N^{*}\in\Psi^{-1}_{cl}(\partial M), it is a compact operator which makes INI-N^{*} Fredholm with index zero. Therefore, numerically computing for R(x,y)R(x,y) amounts to solving a Fredholm boundary integral equation subject to the orthogonality condition

MGM(x,z)dvolh(z)=0.\int_{\partial M}G_{\partial M}(x,z){\rm dvol}_{h}(z)=0.

4. Inverting the Normal Operator

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded convex domain with smooth boundary. We will analyze the mapping properties of the operator

(4.1) L:fΩf(s)|st|n1𝑑s1𝑑sn\displaystyle L:f\mapsto\int_{\Omega}\frac{f(s)}{|s-t|^{n-1}}ds_{1}\dots ds_{n}

and its inverse. Methods do exist [18],[7] for the explicit expression of the inverse of LL when Ω=𝔻\Omega=\mathbb{D} (which is sufficient for our setting). When Ω\Omega is a two dimensional ellipse [27] computed explicitly the inverse of LL acting on the constant function.

The purpose of Section 4.1 is to provide a geometric perspective to the operator LL one of the advantages of which is that it provides an explicit formula for L1(1)L^{-1}(1) when Ω\Omega is a ball of any dimension. Our perspective is based on some of the recent progress on integral geometry (in particular [23], [20], [12]). Since the explicit formulas and estimates will be valid in all dimensions, this will provide the key ingredient in proving Theorem 1.3 in all dimensions. When Ω\Omega is not necessarily 𝔹\mathbb{B}, this geometric point of view may also potentially provide ways to relate some of the quantities of interest to the geometry of Ω\Omega.

Section 4.2 will provide some explicit formulas for the composition of L1L^{-1} with other operators in the case when Ω=𝔻\Omega=\mathbb{D}, the two dimensional disk. Section 4.3 will do the same for when Ω\Omega is the two dimensional ellipse although the formulas will not be as explicit.

4.1. Mapping Properties of L

Denote by

+SΩ:={(x,v)Ω×Sn1vν(x)0}\partial_{+}S\Omega:=\{(x,v)\in\partial\Omega\times S^{n-1}\mid v\cdot\nu(x)\leq 0\}

to be the set of inward pointing unit vectors on Ω\partial\Omega. Note that this is a closed submanifold of the sphere bundle SΩS\Omega and thus inherits its smooth structure. Define the X-Ray transform I:C(Ω¯)C(+SΩ)I:C^{\infty}(\overline{\Omega})\to C^{\infty}(\partial_{+}S\Omega) by If(x,v):=0τ(x,v)f(x+tv)𝑑tIf(x,v):=\int_{0}^{\tau(x,v)}f(x+tv)dt where τ(x,v)\tau(x,v) is the time it takes for a ray of unit velocity vv starting at xΩ¯x\in\overline{\Omega} to reach the boundary Ω\partial\Omega. Note that because Ω\Omega is assumed to be convex, τ(x,v)\tau(x,v) is a smooth function on +SM\partial_{+}SM. Furthermore, II is injective by [23].

By [29] Theorem 4.2.1 this operator extends to an operator I:L2(Ω)Lμ2(+SΩ)I:L^{2}(\Omega)\to L^{2}_{\mu}(\partial_{+}S\Omega) where μ\mu is the measure given by μ=|ν(x)v|dvolΩdvolSn1\mu=|\nu(x)\cdot v|{\rm dvol}_{\partial\Omega}{\rm dvol}_{S^{n-1}}. This L2L^{2} space mapping property allows us to define the adjoint operator II^{*} given by (see [23])

(4.2) Iω(x)=Sn1ω(x+τ(x,v)v)dvolSn1(v)\displaystyle I^{*}\omega(x)=\int_{S^{n-1}}\omega(x+\tau(x,v)v){\rm dvol}_{S^{n-1}}(v)

when acting on smooth functions ω\omega. This allows us to define a self-adjoint normal operator II:L2(Ω)L2(Ω)I^{*}I:L^{2}(\Omega)\to L^{2}(\Omega). It turns out that by [23] the Schwartz kernel of III^{*}I is precisely 2|st|n+12|s-t|^{-n+1} and therefore II=2LI^{*}I=2L. Let dΩ()d_{\Omega}(\cdot) be any smooth positive function on Ω\Omega which is equal to dist(,Ω)dist(\cdot,\partial\Omega) near the boundary. By Theorem 2.2 and Theorem 4.4 of [20] respectively, we have that

(4.3) II:dΩ1/2C(Ω¯)C(Ω¯)\displaystyle I^{*}I:d_{\Omega}^{-1/2}C^{\infty}(\overline{\Omega})\to C^{\infty}(\overline{\Omega})

is a bijection and

(4.4) 2L=II:{uH1/2(n)supp(u)Ω}H1/2(Ω)H1/2(Ω)\displaystyle 2L=I^{*}I:\{u\in H^{-1/2}(\mathbb{R}^{n})\mid{\rm supp}(u)\subset\Omega\}\simeq H^{1/2}(\Omega)^{*}\to H^{1/2}(\Omega)

is a self-adjoint homeomorphism. Thus there exists a unique function u0dΩ1/2C(Ω¯)u_{0}\in d_{\Omega}^{-1/2}C^{\infty}(\overline{\Omega}) such that Lu0=1Lu_{0}=1 which is equivalent to IIu0=2I^{*}Iu_{0}=2. To compute u0u_{0}, observe that if we find u0u_{0} such that

(4.5) Iu0(x,v)=2Vol(Sn1)\displaystyle Iu_{0}(x,v)=\frac{2}{{\rm Vol}(S^{n-1})}

for all (x,v)+SΩ(x,v)\in\partial_{+}S\Omega then by (4.2) we would have IIu0=2I^{*}Iu_{0}=2. The solution of (4.5) is easy to compute explicitly when Ω=𝔹\Omega=\mathbb{B}. Indeed, direct computation shows that choosing

(4.6) u0(x)=2πVol(Sn1)1|x|2\displaystyle u_{0}(x)=\frac{2}{\pi{\rm Vol}(S^{n-1})\sqrt{1-|x|^{2}}}

one satisfies (4.5). In particular if n=2n=2 (which is the case we are interested in) we have that

(4.7) L1(1)=u0(x)=1π21|x|2,indimension 2\displaystyle L^{-1}(1)=u_{0}(x)=\frac{1}{\pi^{2}\sqrt{1-|x|^{2}}},\ \ {\rm in\ dimension\ 2}
Remark 4.1.

This process of computing solution to IIu0=constI^{*}Iu_{0}=const by solving for Iu0=constIu_{0}=const unfortunately only works for Ω=𝔹\Omega=\mathbb{B}. In fact, thanks to the rigidity result of [12], we know that Iu0=1Iu_{0}=1 is solvable iff Ω=𝔹\Omega=\mathbb{B}. However, for more general domains the geometric view presented here could potentially allow one to apply the reconstruction formula for inverting II [22] to solve IIu0=1I^{*}Iu_{0}=1 explicitly. To do so one must first invert II^{*} (which has a large kernel but is surjective) into the range of II and it is not clear how to do this when Ω𝔹\Omega\neq\mathbb{B}.

4.2. Integrals Involving L inverse of 1.

We also define RlogR_{\log} and RR_{\infty} to be operators with kernels log|st|\log|s^{\prime}-t^{\prime}| and (s1t1)2(s2t2)2|st|2\frac{(s_{1}-t_{1})^{2}-(s_{2}-t_{2})^{2}}{|s^{\prime}-t^{\prime}|^{2}} respectively.

Lemma 4.2.

The operators RR_{\infty} and RlogR_{\log} are bounded maps from H1/2(𝔻)H^{1/2}(\mathbb{D})^{*} to H3/2(𝔻)H^{3/2}(\mathbb{D}).

Proof.

Observe that the integral kernels of both RlogR_{\log} and RR_{\infty} extends naturally to kernels representing operators in Ψ(2)\Psi^{\infty}(\mathbb{R}^{2}) which we denote by R~log\tilde{R}_{\log} and R~\tilde{R}_{\infty} respectively. We denote by E:H1/2(𝔻){uH1/2(2)supp(u)Ω}E:H^{1/2}(\mathbb{D})^{*}\to\{u\in H^{-1/2}(\mathbb{R}^{2})\mid{\rm supp}(u)\subset\Omega\} to be the isomorphism obtained by the trivial extension. Let χC0(2)\chi\in C^{\infty}_{0}(\mathbb{R}^{2}) be identically 11 on 𝔻\mathbb{D}. Then we have that

(4.8) (Rlogu)𝔻=(χR~logχEu)𝔻\displaystyle\left(R_{\log}u\right)\mid_{\mathbb{D}}=\left(\chi\tilde{R}_{\log}\chi Eu\right)\mid_{\mathbb{D}}

and the same holds for RR_{\infty}.

By Proposition 7.2.8 of [34] we have that both χR~logχ\chi\tilde{R}_{\log}\chi and χR~χ\chi\tilde{R}_{\infty}\chi are pseudodifferential operators of order 2-2. Therefore by (4.8) both RlogR_{\log} and RR_{\infty} are bounded operators from H1/2(𝔻)H^{1/2}(\mathbb{D})^{*} to H3/2(𝔻)H^{3/2}(\mathbb{D}).

The following lemma was proved in [7, Theorem 4.2].

Lemma 4.3.

For uH12(𝔻)u\in H^{\frac{1}{2}}(\mathbb{D}), it follows

(4.9) L1u,1=1π2𝔻u(t)(1|t|2)12𝑑t1𝑑t2.\displaystyle\left\langle{L}^{-1}u,1\right\rangle=\frac{1}{\pi^{2}}\int_{\mathbb{D}}\frac{u(t^{\prime})}{(1-|t^{\prime}|^{2})^{\frac{1}{2}}}dt_{1}dt_{2}.
Proof.

By (4.4) L:H12(𝔻)H12(𝔻)L:H^{\frac{1}{2}}(\mathbb{D})^{*}\to H^{\frac{1}{2}}(\mathbb{D}) is a self-adjoint homeomorphism. The result of this Lemma is a direct consequence. ∎

Lemma 4.4.

Let

f(s)=RlogL11=𝔻log|ts|[L11](t)𝑑t1𝑑t2,\displaystyle f(s^{\prime})=R_{log}L^{-1}1=\int_{\mathbb{D}}\log|t^{\prime}-s^{\prime}|[L^{-1}1](t^{\prime})dt_{1}dt_{2},

then

f(s)=2πlog|s|+2π(12log|(1|s|2)12+1|12log|(1|s|2)121|)2π(1|s|2)12.\displaystyle f(s^{\prime})=\frac{2}{\pi}\log|s^{\prime}|+\frac{2}{\pi}\left(\frac{1}{2}\log\left|(1-|s^{\prime}|^{2})^{\frac{1}{2}}+1\right|-\frac{1}{2}\log\left|(1-|s^{\prime}|^{2})^{\frac{1}{2}}-1\right|\right)-\frac{2}{\pi}(1-|s^{\prime}|^{2})^{\frac{1}{2}}.
Proof.

Note that 12πlog|ts|\frac{1}{2\pi}\log|t^{\prime}-s^{\prime}| is the fundamental solution for the Laplace operator in 2\mathbb{R}^{2}, therefore,

Δf=2π[L11]=2π1(1|t|2)12\displaystyle\Delta f=2\pi[L^{-1}1]=\frac{2}{\pi}\frac{1}{(1-|t^{\prime}|^{2})^{\frac{1}{2}}}

Since [L11](t)[{L}^{-1}1](t^{\prime}) is radially symmetric, f(t)=f~(r)f(t^{\prime})=\tilde{f}(r) where r=|t|r=|t^{\prime}|. Writing the Laplace operator in polar coordinates, we get

(rf~r)r=2πr(1r2)12.\displaystyle(r\tilde{f}_{r})_{r}=\frac{2}{\pi}\frac{r}{(1-r^{2})^{\frac{1}{2}}}.

Integration gives

f~r(r)=2π(C1r(1r2)12r)\displaystyle\tilde{f}_{r}(r)=\frac{2}{\pi}\left(\frac{C_{1}}{r}-\frac{(1-r^{2})^{\frac{1}{2}}}{r}\right)

and

f~(r)=2π(C1logr12log|(1r2)121|+12log|(1r2)12+1|(1r2)12+C2).\displaystyle\quad\tilde{f}(r)=\frac{2}{\pi}\left(C_{1}\log r-\frac{1}{2}\log\left|(1-r^{2})^{\frac{1}{2}}-1\right|+\frac{1}{2}\log\left|(1-r^{2})^{\frac{1}{2}}+1\right|-(1-r^{2})^{\frac{1}{2}}+C_{2}\right).

Let us find C1C_{1} and C2C_{2}. Note that f~(r)\tilde{f}(r) does not have singularity at r=0r=0, namely,

f~(0)=f(0)=1π2𝔻log|t|(1|t|2)12=2π01rlogr(1r2)12𝑑r=2π(log21).\displaystyle\tilde{f}(0)=f(0)=\frac{1}{\pi^{2}}\int_{\mathbb{D}}\frac{\log|t^{\prime}|}{(1-|t^{\prime}|^{2})^{\frac{1}{2}}}=\frac{2}{\pi}\int_{0}^{1}\frac{r\log r}{(1-r^{2})^{\frac{1}{2}}}dr=\frac{2}{\pi}(\log 2-1).

Therefore, the identities

C1logr12log|(1r2)121|=12log|r2C1(1r2)121|=12log|r2C112r2+O(r4)|,\displaystyle C_{1}\log r-\frac{1}{2}\log\left|(1-r^{2})^{\frac{1}{2}}-1\right|=\frac{1}{2}\log\left|\frac{r^{2C_{1}}}{(1-r^{2})^{\frac{1}{2}}-1}\right|=\frac{1}{2}\log\left|\frac{r^{2C_{1}}}{-\frac{1}{2}r^{2}+O(r^{4})}\right|,

as r0r\rightarrow 0, implies that C1=1C_{1}=1. Hence, putting r=0r=0 into (4.4), gives

2π(log21)=2π(12log2+12log21+C2),\displaystyle\frac{2}{\pi}(\log 2-1)=\frac{2}{\pi}\left(\frac{1}{2}\log 2+\frac{1}{2}\log 2-1+C_{2}\right),

so that C2=0C_{2}=0. ∎

Due to Lemmas 4.4 and 4.3, the following identity is a direct computation:

Lemma 4.5.

The following identity holds

L1RlogL11,1=8π2log26π2.\displaystyle\left\langle L^{-1}R_{log}L^{-1}1,1\right\rangle=\frac{8}{\pi^{2}}\log 2-\frac{6}{\pi^{2}}.
Lemma 4.6.

The following idetity holds

L1RL11,1=0.\displaystyle\left\langle L^{-1}R_{\infty}L^{-1}1,1\right\rangle=0.
Proof.

By Lemmas (4.7) and 4.3, we know that

L1RL11,1=1π2𝔻𝔻(s1t1)2(s2t2)2|st|21(1|s|2)12𝑑s1(1|t|2)12𝑑t.\displaystyle\left\langle L^{-1}R_{\infty}L^{-1}1,1\right\rangle=\frac{1}{\pi^{2}}\int_{\mathbb{D}}\int_{\mathbb{D}}\frac{(s_{1}-t_{1})^{2}-(s_{2}-t_{2})^{2}}{|s^{\prime}-t^{\prime}|^{2}}\frac{1}{(1-|s^{\prime}|^{2})^{\frac{1}{2}}}ds^{\prime}\frac{1}{(1-|t^{\prime}|^{2})^{\frac{1}{2}}}dt^{\prime}.

Consider the following two changes of variables for the right-hand side

(s1,s2,t1,t2)=(rcosϕ,rsinϕ,ρcosθ,ρsinθ),\displaystyle(s_{1},s_{2},t_{1},t_{2})=(r\cos\phi,r\sin\phi,\rho\cos\theta,\rho\sin\theta),
(s1,s2,t1,t2)=(rsinϕ,rcosϕ,ρsinθ,ρcosθ).\displaystyle(s_{1},s_{2},t_{1},t_{2})=(r\sin\phi,r\cos\phi,\rho\sin\theta,\rho\cos\theta).

The results differ by multiplying by 1-1, which means that the right-hand side is 0. ∎

4.3. Explicit Formulas in 2 Dimensional Ellipse

We now compute the inverse of the map faf(s)|st|𝑑sf\mapsto\int_{{\mathcal{E}}_{a}}\frac{f(s^{\prime})}{|s^{\prime}-t^{\prime}|}ds^{\prime} where the domain of integration is the two dimensional ellipse a:={s12+s22a2=1}{\mathcal{E}}_{a}:=\{s_{1}^{2}+\frac{s_{2}^{2}}{a^{2}}=1\} instead of a ball. A change of variable leads us to consider the operator

(4.11) Laf=a𝔻f(s)((t1s1)2+a2(t2s2)2)1/2𝑑sL_{a}f=a\int_{\mathbb{D}}\frac{f(s^{\prime})}{\left((t_{1}-s_{1})^{2}+a^{2}(t_{2}-s_{2})^{2}\right)^{1/2}}ds^{\prime}

acting on functions of the disk 𝔻\mathbb{D}. By [27] we have that

(4.12) La(Ka1(1|t|2)1/2)=1,L_{a}\left({K_{a}}^{-1}{(1-|t^{\prime}|^{2})^{-1/2}}\right)=1,

on 𝔻\mathbb{D} where

Ka=π202π1(cos2θ+sin2θa2)1/2𝑑θ.K_{a}=\frac{\pi}{2}\int_{0}^{2\pi}\frac{1}{\left(\cos^{2}\theta+\frac{\sin^{2}\theta}{a^{2}}\right)^{1/2}}d\theta.

By (4.4) this is the unique solution in H1/2(𝔻)H^{1/2}(\mathbb{D})^{*} to Lau=1L_{a}u=1.

Next we denote

Rlog,af(t):=a𝔻log((t1s1)2+a2(t2s2)2)1/2f(s)ds,R_{\log,a}f(t^{\prime}):=a\int_{\mathbb{D}}\log\left((t_{1}-s_{1})^{2}+a^{2}(t_{2}-s_{2})^{2}\right)^{1/2}f(s^{\prime})ds^{\prime},
R,af(t):=a𝔻(t1s1)2a2(t2s2)2(t1s1)2+a2(t2s2)2f(s)𝑑s.R_{\infty,a}f(t^{\prime}):=a\int_{\mathbb{D}}\frac{(t_{1}-s_{1})^{2}-a^{2}(t_{2}-s_{2})^{2}}{(t_{1}-s_{1})^{2}+a^{2}(t_{2}-s_{2})^{2}}f(s^{\prime})ds^{\prime}.

For general aa, the quantities La1(1),log,aLa1(a)\langle L^{-1}_{a}(1),\mathbb{R}_{\log,a}L_{a}^{-1}(a)\rangle and La1(1),,aLa1(a)\langle L^{-1}_{a}(1),\mathbb{R}_{\infty,a}L_{a}^{-1}(a)\rangle cannot be computed as explicitly as in the case when a=1a=1 in Section 4.2.

5. Asymptotic Expansion of the Singularly Perturbed Problems

5.1. Asymptotic Expansion of Mixed Boundary Value Problems

We are now ready to compute the asymptotic expansion for the mixed boundary value problem uϵH1(M)u_{\epsilon}\in H^{1}(M),

(5.1) Δguϵ=1,uϵΓϵ,a=0,νuϵM\Γϵ,a=0\displaystyle\Delta_{g}u_{\epsilon}=-1,\ \ u_{\epsilon}\mid_{\Gamma_{\epsilon,a}}=0,\ \ \partial_{\nu}u_{\epsilon}\mid_{\partial M\backslash\Gamma_{\epsilon,a}}=0

which gives the compatibility condition

(5.2) Mνuϵdvolh=|M|.\displaystyle\int_{\partial M}\partial_{\nu}u_{\epsilon}{\rm dvol}_{h}=-|M|.

All integrals on open subsets of M\partial M are with respect to the volume form given by the metric hh.

Using Green’s formula as in [9], also in [1], we can deduce that for points xMox\in M^{o}, uϵ(x)u_{\epsilon}(x) satisfies the integral equation

(5.3) uϵ(x)=F(x)+Cϵ,a+MG(x,y)νuϵ(y)dvolh(y),\displaystyle u_{\epsilon}(x)=F(x)+C_{\epsilon,a}+\int_{\partial M}G(x,y)\partial_{\nu}u_{\epsilon}(y){\rm dvol}_{h}(y),

where Cϵ,a=|M|1MuϵC_{\epsilon,a}=|\partial M|^{-1}\int_{\partial M}u_{\epsilon} and F(x)=MG(x,y)F(x)=\int_{M}G(x,y) solves the boundary value problem

(5.4) ΔF=1,νF=|M|/|M|,MFdvolh=0.\displaystyle\Delta F=-1,\ \ \partial_{\nu}F=-|M|/|\partial M|,\ \ \int_{\partial M}F{\rm dvol}_{h}=0.

By Proposition 1.1 we can take the trace of (5.3) to the boundary and restrict to the open subset Γϵ,aoM\Gamma_{\epsilon,a}^{o}\subset\partial M. Using (5.1) we see that

0=F(x)+Cϵ,a+Γϵ,aGM(x,y)νuϵ(y)dvolh(y)0=F(x)+C_{\epsilon,a}+\int_{\Gamma_{\epsilon,a}}G_{\partial M}(x,y)\partial_{\nu}u_{\epsilon}(y){\rm dvol}_{h}(y)

for xΓϵ,aox\in\Gamma_{\epsilon,a}^{o}. We now replace GM(x,y)G_{\partial M}(x,y) for x,yΓϵ,aox,y\in\Gamma_{\epsilon,a}^{o} with the expression in (1.1) to obtain

F(x)Cϵ,a\displaystyle-F(x)-C_{\epsilon,a} =\displaystyle= 12πΓϵ,adg(x,y)1νuϵ(y)dvolh(y)H(x)4πΓϵ,alogdh(x,y)νuϵ(y)dvolh(y)\displaystyle\frac{1}{2\pi}\int_{\Gamma_{\epsilon,a}}d_{g}(x,y)^{-1}\partial_{\nu}u_{\epsilon}(y){\rm dvol}_{h}(y)-\frac{H(x)}{4\pi}\int_{\Gamma_{\epsilon,a}}\log d_{h}(x,y)\partial_{\nu}u_{\epsilon}(y){\rm dvol}_{h}(y)
+116πΓϵ,a(IIx(expx1(y)|expx1(y)|h)IIx(expx1(y)|expx1(y)|h))νuϵ(y)dvolh(y)\displaystyle+\frac{1}{16\pi}\int_{\Gamma_{\epsilon,a}}\left({\rm{II}}_{x}\left(\frac{\exp_{x}^{-1}(y)}{|\exp_{x}^{-1}(y)|_{h}}\right)-{\rm{II}}_{x}\left(\frac{*\exp_{x}^{-1}(y)}{|\exp_{x}^{-1}(y)|_{h}}\right)\right)\partial_{\nu}u_{\epsilon}(y){\rm dvol}_{h}(y)
+Γϵ,aR(x,y)νuϵ(y)dvolh(y).\displaystyle+\int_{\Gamma_{\epsilon,a}}R(x,y)\partial_{\nu}u_{\epsilon}(y){\rm dvol}_{h}(y).

We will write this integral equation in the coordinate system given by

(5.6) 𝔻(s1,s2)xϵ(s1,as2;x)Γϵ,a,\displaystyle\mathbb{D}\ni(s_{1},s_{2})\mapsto x^{\epsilon}(s_{1},as_{2};x^{*})\in\Gamma_{\epsilon,a},

where xϵ(;x):aΓϵ,ax^{\epsilon}(\cdot;x^{*}):{\mathcal{E}}_{a}\to\Gamma_{\epsilon,a} is the coordinate defined in Section 2.2. To simplify notation we will drop the xx^{*} in the notation and denote xϵ(;x)x^{\epsilon}(\cdot;x^{*}) by simply xϵ()x^{\epsilon}(\cdot).

Note that in these coordinates the volume form for M\partial M is given by

(5.7) dvolh=aϵ2(1+ϵ2Qϵ(s))ds1ds2,s𝔻\displaystyle{\rm dvol}_{h}=a\epsilon^{2}(1+\epsilon^{2}Q_{\epsilon}(s^{\prime}))ds_{1}\wedge ds_{2},\ s^{\prime}\in\mathbb{D}

for some smooth function Qϵ(s)Q_{\epsilon}(s^{\prime}) whose derivatives of all orders are bounded uniformly in ϵ\epsilon. We denote

(5.8) ψϵ(s):=νuϵ(xϵ(s1,as2)).\displaystyle\psi_{\epsilon}(s^{\prime}):=\partial_{\nu}u_{\epsilon}(x^{\epsilon}(s_{1},as_{2})).

The compatibility condition (5.2) written using the expression for the volume form (5.7) is then

(5.9) 𝔻ψϵ(s)(1+ϵ2Qϵ(s))𝑑s1𝑑s2=|M|aϵ2.\displaystyle\int_{\mathbb{D}}\psi_{\epsilon}(s^{\prime})(1+\epsilon^{2}Q_{\epsilon}(s^{\prime}))ds_{1}ds_{2}=-\frac{|M|}{a\epsilon^{2}}.

Let us unwrap the right hand side of (5.1) term by term in the coordinate given by xϵ()x^{\epsilon}(\cdot). Write out the integral of the first term using the expression of the volume form (5.7) and the expression for dg(x,y)1d_{g}(x,y)^{-1} in Corollary 2.6 and taking into account that the coordinate system is scaled by a factor aa as in (5.6) gives

(5.10) Γϵ,adg(x,y)1νuϵ(y)dvolh(y)=aϵ𝔻1((t1s1)2+a2(t2s2)2)1/2ψϵ(s)𝑑s+ϵ3𝒜ϵψϵ,\int_{\Gamma_{\epsilon,a}}d_{g}(x,y)^{-1}\partial_{\nu}u_{\epsilon}(y){\rm dvol}_{h}(y)=a\epsilon\int_{\mathbb{D}}\frac{1}{\left((t_{1}-s_{1})^{2}+a^{2}(t_{2}-s_{2})^{2}\right)^{1/2}}\psi_{\epsilon}(s^{\prime})ds+\epsilon^{3}\mathcal{A}_{\epsilon}\psi_{\epsilon},

for some operator 𝒜ϵ\mathcal{A}_{\epsilon} whose Schwartz kernel is given by the second term of the expansion in Corollary 2.6. Here due to Lemma 2.11 we have that

𝒜ϵ:H1/2(𝔻;ds)H1/2(𝔻;ds)\mathcal{A}_{\epsilon}:H^{1/2}(\mathbb{D};ds^{\prime})^{*}\to H^{1/2}(\mathbb{D};ds^{\prime})

with operator norm bounded uniformly in ϵ\epsilon. From here on we will denote by 𝒜ϵ\mathcal{A}_{\epsilon} any operator which takes H1/2(𝔻;ds)H1/2(𝔻;ds)H^{1/2}(\mathbb{D};ds^{\prime})^{*}\to H^{1/2}(\mathbb{D};ds^{\prime}) whose operator norm is bounded uniformly in ϵ\epsilon.

Doing the same thing for the second term of (5.1) while using Lemma 2.2, Lemma 2.11, and (5.9) gives

(5.11) H(x)Γϵ,alogdh(x,y)νuϵ(y)dvolh(y)\displaystyle H(x)\int_{\Gamma_{\epsilon,a}}\log d_{h}(x,y)\partial_{\nu}u_{\epsilon}(y){\rm dvol}_{h}(y) =\displaystyle= H(x)|M|logϵ+ϵ2H(x)Rlog,aψϵ+ϵ3𝒜ϵψϵ\displaystyle-H(x^{*})|M|\log\epsilon+\epsilon^{2}H(x^{*})R_{\log,a}\psi_{\epsilon}+\epsilon^{3}\mathcal{A}_{\epsilon}\psi_{\epsilon}
+OH1/2(𝔻)(ϵlogϵ)\displaystyle+O_{H^{1/2}(\mathbb{D})}(\epsilon\log\epsilon)

where Rlog,aR_{\log,a} is defined at the very beginning of Section 4.2. Here OH1/2(𝔻)(ϵlogϵ)O_{H^{1/2}(\mathbb{D})}(\epsilon\log\epsilon) denotes a function with H1/2(𝔻;ds)H^{1/2}(\mathbb{D};ds^{\prime}) norm vanishing to order ϵlogϵ\epsilon\log\epsilon. Note the volume for we use here is now the Euclidean one rather than dvolh{\rm dvol}_{h} given by (5.7).

Finally, for the third term of (5.1) we get by using the coordinate expression derived in Corollary 2.9 and the estimate of Lemma 2.11:

(5.12) Γϵ,a(IIx(expx1(y)|expx1(y)|h)IIx(expx1(y)|expx1(y)|h))νuϵ(y)\displaystyle\int_{\Gamma_{\epsilon,a}}\left({\rm{II}}_{x}\left(\frac{\exp_{x}^{-1}(y)}{|\exp_{x}^{-1}(y)|_{h}}\right)-{\rm{II}}_{x}\left(\frac{*\exp_{x}^{-1}(y)}{|\exp_{x}^{-1}(y)|_{h}}\right)\right)\partial_{\nu}u_{\epsilon}(y) =\displaystyle= ϵ2(λ1λ2)R,aψϵ\displaystyle\epsilon^{2}(\lambda_{1}-\lambda_{2})R_{\infty,a}\psi_{\epsilon}
+ϵ3𝒜ϵψϵ,\displaystyle+\epsilon^{3}\mathcal{A}_{\epsilon}\psi_{\epsilon},

where R,aR_{\infty,a} is defined in Section 4. Inserting into (5.1) the identities (5.10), (5.11), and (5.12) we have

F(x)Cϵ,aH(x)|M|(4π)1logϵϵ\displaystyle\frac{-F(x^{*})-C_{\epsilon,a}-H(x^{*})|M|(4\pi)^{-1}\log\epsilon}{\epsilon} =\displaystyle= 12πLaψϵϵH(x)4πRlog,aψϵ+ϵλ1λ216πR,aψϵ\displaystyle\frac{1}{2\pi}L_{a}\psi_{\epsilon}-\epsilon\frac{H(x^{*})}{4\pi}R_{\log,a}\psi_{\epsilon}+\epsilon\frac{\lambda_{1}-\lambda_{2}}{16\pi}R_{\infty,a}\psi_{\epsilon}
+aϵ𝔻R(xϵ(t),xϵ(s))ψϵ(s)+ϵ2𝒜ϵψϵ\displaystyle+a\epsilon\int_{\mathbb{D}}R(x^{\epsilon}(t^{\prime}),x^{\epsilon}(s^{\prime}))\psi_{\epsilon}(s^{\prime})+\epsilon^{2}\mathcal{A}_{\epsilon}\psi_{\epsilon}
+OH1/2(𝔻)(logϵ).\displaystyle+O_{H^{1/2}(\mathbb{D})}(\log\epsilon).

We would like to approximate the integral kernel R(xϵ(t),xϵ(s))R(x^{\epsilon}(t^{\prime}),x^{\epsilon}(s^{\prime})) by the constant R(x,x)R(x^{*},x^{*}) and this is the content of

Lemma 5.1.

Let Tϵ:Cc(𝔻)𝒟(𝔻)T_{\epsilon}:C_{c}^{\infty}(\mathbb{D})\to{\mathcal{D}}^{\prime}(\mathbb{D}) be the operator defined by the integral kernel

R(xϵ(t;x),xϵ(s;x))R(x,x).R(x^{\epsilon}(t^{\prime};x^{*}),x^{\epsilon}(s^{\prime};x^{*}))-R(x^{*},x^{*}).

Then

Tϵ(H1/2(𝔻))H1/2(𝔻)Cϵlogϵ.\|T_{\epsilon}\|_{(H^{1/2}(\mathbb{D}))^{*}\to H^{1/2}(\mathbb{D})}\leq C\epsilon\log\epsilon.
Proof.

Set T(t,s):=R(x(t;x),x(s;x))R(x,x)T(t^{\prime},s^{\prime}):=R(x(t^{\prime};x^{*}),x(s^{\prime};x^{*}))-R(x^{*},x^{*}) for tt^{\prime} and ss^{\prime} small and extend it to be a smooth compactly supported kernel otherwise. The kernel for TϵT_{\epsilon} is then T(ϵt,ϵs)T(\epsilon t^{\prime},\epsilon s^{\prime}). Note that

(5.14) T(0,0)=0.\displaystyle T(0,0)=0.

Observe that if χCc(2)\chi\in C_{c}^{\infty}(\mathbb{R}^{2}) which is identically 11 on 𝔻\mathbb{D} then the operator TϵT_{\epsilon} acting on distributions supported in 𝔻\mathbb{D} is given by the Schwartz kernel

χ(s)χ(t)T(ϵt,ϵs)\chi(s^{\prime})\chi(t^{\prime})T(\epsilon t^{\prime},\epsilon s^{\prime})

for t,s𝔻t^{\prime},s^{\prime}\in\mathbb{D} and ϵ>0\epsilon>0 small.

Observe that T(t,s)T(t^{\prime},s^{\prime}) is the integral kernel for an operator in Ψcl3(2)\Psi^{-3}_{cl}(\mathbb{R}^{2}). Applying Prop 2.8 in Chap 7 of [34] we can deduce that for all kk we may choose NN sufficiently large such that

T(t,s)=l=0N(ql(t,ts)+pl(t,ts)log|ts|)+Rk(t,s),T(t^{\prime},s^{\prime})=\sum_{l=0}^{N}\left(q_{l}(t^{\prime},t^{\prime}-s^{\prime})+p_{l}(t^{\prime},t^{\prime}-s^{\prime})\log|t^{\prime}-s^{\prime}|\right)+R_{k}(t^{\prime},s^{\prime}),

where for each integer ll and multi-index γ\gamma, Dtγql(t,)D^{\gamma}_{t^{\prime}}q_{l}(t^{\prime},\cdot) is a bounded continuous function of tt^{\prime} with value in the space of smooth (away from the origin) homogeneous distributions of degree l+1l+1, pl(t,)p_{l}(t^{\prime},\cdot) are homogenous polynomials of degree l+1l+1 with coefficients which are smooth functions of tt^{\prime}, and for all multi-indices γ\gamma, DtγRk(t,)D^{\gamma}_{t^{\prime}}R_{k}(t^{\prime},\cdot) bounded continuous function of tt^{\prime} with value in Ck(2)C^{k}(\mathbb{R}^{2}).

Using (5.14) along with the homogenenity degree of qlq_{l} and plp_{l} we see that Rk(0,0)=0R_{k}(0,0)=0. Therefore, for s,t𝔻s^{\prime},t^{\prime}\in\mathbb{D} the integral kernel of TϵT_{\epsilon} is

(5.15) Tϵ(t,s)\displaystyle T_{\epsilon}(t^{\prime},s^{\prime}) =\displaystyle= l=0N(ϵl+1ql(ϵt,ts)+ϵl+1pl(ϵt,ts)logϵ+ϵl+1pl(ϵt,ts)log|ts|)\displaystyle\sum_{l=0}^{N}\left(\epsilon^{l+1}q_{l}(\epsilon t^{\prime},t^{\prime}-s^{\prime})+\epsilon^{l+1}p_{l}(\epsilon t^{\prime},t^{\prime}-s^{\prime})\log\epsilon+\epsilon^{l+1}p_{l}(\epsilon t^{\prime},t^{\prime}-s^{\prime})\log|t^{\prime}-s^{\prime}|\right)
+\displaystyle+ Rk(ϵt,ϵs).\displaystyle R_{k}(\epsilon t^{\prime},\epsilon s^{\prime}).

The kernel Rk(s,t)R_{k}(s^{\prime},t^{\prime}) is sufficiently smooth. Therefore, by doing a Taylor expansion and using Rk(0,0)=0R_{k}(0,0)=0 we see that the integral kernel χ(s)χ(t)Rk(ϵt,ϵs)\chi(s^{\prime})\chi(t^{\prime})R_{k}(\epsilon t^{\prime},\epsilon s^{\prime}) takes H1/2(𝔻)H1/2(𝔻)H^{1/2}(\mathbb{D})^{*}\to H^{1/2}(\mathbb{D}) with norm ϵ\epsilon. The worst term in the polyhomogeneous expansion part of (5.15) happens when l=0l=0 and this term is given by

ϵq0(ϵt,z)+p0(ϵt,z)log|z|+ϵlogϵp0(ϵt,z),\epsilon q_{0}(\epsilon t^{\prime},z^{\prime})+p_{0}(\epsilon t^{\prime},z^{\prime})\log|z^{\prime}|+\epsilon\log\epsilon p_{0}(\epsilon t^{\prime},z^{\prime}),

where z=tsz^{\prime}=t^{\prime}-s^{\prime}. Recall that both q0q_{0} and p0p_{0} are homogeneous of degree 11 in zz^{\prime} so writing z=rωz^{\prime}=r\omega then applying Lemma 2.11 we have uniform estimates in ϵ\epsilon for the kernels χ(s)χ(t)q0(ϵt,ts)\chi(s^{\prime})\chi(t^{\prime})q_{0}(\epsilon t^{\prime},t^{\prime}-s^{\prime}) and χ(s)χ(t)p0(ϵt,ts)\chi(s^{\prime})\chi(t^{\prime})p_{0}(\epsilon t^{\prime},t^{\prime}-s^{\prime}). For the term involving log|st|\log|s^{\prime}-t^{\prime}|, we use the fact that p0(t,z)p_{0}(t^{\prime},z^{\prime}) is a linear function in zz^{\prime} whose coefficients are smooth functions of tt^{\prime}. Therefore, if uCc(2)u\in C^{\infty}_{c}(\mathbb{R}^{2}),

2χ(t)p0(ϵt,ts)log|ts|u(s)𝑑s\displaystyle\int_{\mathbb{R}^{2}}\chi(t^{\prime})p_{0}(\epsilon t^{\prime},t^{\prime}-s^{\prime})\log|t^{\prime}-s^{\prime}|u(s^{\prime})ds^{\prime} =\displaystyle= χ(t)2jcj(ϵt)(tjsj)log|st|u(s)ds\displaystyle\chi(t^{\prime})\int_{\mathbb{R}^{2}}\sum_{j}c_{j}(\epsilon t^{\prime})(t_{j}-s_{j})\log|s^{\prime}-t^{\prime}|u(s^{\prime})ds^{\prime}
=\displaystyle= χ(t)jcj(ϵt)2aj(ξ)u^(ξ)eitξ𝑑ξ,\displaystyle\chi(t^{\prime})\sum_{j}c_{j}(\epsilon t^{\prime})\int_{\mathbb{R}^{2}}a_{j}(\xi^{\prime})\hat{u}(\xi^{\prime})e^{it^{\prime}\cdot\xi^{\prime}}d\xi^{\prime},

where aj()𝒮(2)a_{j}(\cdot)\in{\mathcal{S}}^{\prime}(\mathbb{R}^{2}) for j=1,2j=1,2 are derivatives of the distribution PF|ξ|2{\rm PF}|\xi|^{-2} with respect to ξj\partial_{\xi_{j}}. We refer the reader to (8.31) in Chapt 3 of [33] for the definition of the the distribution PF|ξ|2{\rm PF}|\xi|^{-2}. From this we see that the integral kernel

χ(t)χ(s)p0(ϵt,ts)log|ts|\chi(t^{\prime})\chi(s^{\prime})p_{0}(\epsilon t^{\prime},t^{\prime}-s^{\prime})\log|t^{\prime}-s^{\prime}|

also maps H1/2(𝔻)H1/2(𝔻)H^{1/2}(\mathbb{D})^{*}\to H^{1/2}(\mathbb{D}) with uniform bound in ϵ\epsilon. ∎

Due to Lemma 5.1 we can write (5.1) as

(5.16) 2πϵ(aR(x,x)|M|F(x)Cϵ,aH(x)|M|logϵ4π)=\displaystyle\frac{2\pi}{\epsilon}\left(aR(x^{*},x^{*})|M|-F(x^{*})-C_{\epsilon,a}-\frac{H(x^{*})|M|\log\epsilon}{4\pi}\right)= (LaϵH(x)2Rlog,a+ϵ(λ1λ2)8R,a)ψϵ\displaystyle\left(L_{a}-\frac{\epsilon H(x^{*})}{2}R_{\log,a}+\frac{\epsilon(\lambda_{1}-\lambda_{2})}{8}R_{\infty,a}\right)\psi_{\epsilon}
+ϵTϵψϵ++ϵ2𝒜ϵψϵ+OH1/2(𝔻)(logϵ).\displaystyle+\epsilon T_{\epsilon}\psi_{\epsilon}++\epsilon^{2}\mathcal{A}_{\epsilon}\psi_{\epsilon}+O_{H^{1/2}(\mathbb{D})}(\log\epsilon).

We hit both sides with La1L_{a}^{-1} and use (4.12) and (4.4) to get the identity

(5.17) 2πϵ(aR(x,x)|M|F(x)Cϵ,aH(x)|M|logϵ4π)1Ka(1|t|2)1/2=\displaystyle\frac{2\pi}{\epsilon}\left(aR(x^{*},x^{*})|M|-F(x^{*})-C_{\epsilon,a}-\frac{H(x^{*})|M|\log\epsilon}{4\pi}\right)\frac{1}{K_{a}(1-|t^{\prime}|^{2})^{1/2}}=
(IϵH(x)2La1Rlog,a+ϵ(λ1λ2)8La1R,a+Tϵ)ψϵ+OH1/2(𝔻)(logϵ).\displaystyle\left(I-\frac{\epsilon H(x^{*})}{2}L_{a}^{-1}R_{\log,a}+\frac{\epsilon(\lambda_{1}-\lambda_{2})}{8}L_{a}^{-1}R_{\infty,a}+T^{\prime}_{\epsilon}\right)\psi_{\epsilon}+O_{H^{1/2}(\mathbb{D})^{*}}(\log\epsilon).

for some Tϵ:H1/2(𝔻)H1/2(𝔻)T^{\prime}_{\epsilon}:H^{1/2}(\mathbb{D})^{*}\to H^{1/2}(\mathbb{D})^{*} with operator norm O(ϵ2logϵ)O(\epsilon^{2}\log\epsilon). Use the mapping properties from Lemma 4.2 we see that the right side can be inverted by Neumann series to deduce

(5.18) ψϵ=2πCϵ,aϵKa(1|t|2)1/2+Cϵ,aOH1/2(𝔻)(1)+OH1/2(𝔻)(ϵ1logϵ).\displaystyle\psi_{\epsilon}=-\frac{2\pi C_{\epsilon,a}}{\epsilon K_{a}(1-|t^{\prime}|^{2})^{1/2}}+C_{\epsilon,a}O_{H^{1/2}(\mathbb{D})^{*}}(1)+O_{H^{1/2}(\mathbb{D})^{*}}(\epsilon^{-1}\log\epsilon).

Insert (5.18) into (5.9) we get that

(5.19) Cϵ,a=|M|Ka4aϵπ2+Cϵ,a\displaystyle C_{\epsilon,a}=\frac{|M|K_{a}}{4a\epsilon\pi^{2}}+C^{\prime}_{\epsilon,a}

with Cϵ,a=O(logϵ)C^{\prime}_{\epsilon,a}=O(\log\epsilon). Insert (5.19) into (5.18)

(5.20) ψϵ=|M|2aπϵ2(1|t|2)1/2+ψϵ=|M|Ka2aπϵ2La1(1)+ψϵ\displaystyle\psi_{\epsilon}=\frac{-|M|}{2a\pi\epsilon^{2}(1-|t^{\prime}|^{2})^{1/2}}+\psi_{\epsilon}^{\prime}=-\frac{|M|K_{a}}{2a\pi\epsilon^{2}}L_{a}^{-1}(1)+\psi^{\prime}_{\epsilon}

with ψϵH1/2(𝔻;ds)Cϵ1logϵ\|\psi_{\epsilon}^{\prime}\|_{H^{1/2}(\mathbb{D};ds^{\prime})^{*}}\leq C\epsilon^{-1}\log\epsilon. Insert (5.19) and (5.20) into (5.17) we get

(5.21) 2πϵ(aR(x,x)|M|F(x)Cϵ,aH(x)|M|logϵ4π)1Ka(1|t|2)1/2=\displaystyle\frac{2\pi}{\epsilon}\left(aR(x^{*},x^{*})|M|-F(x^{*})-C^{\prime}_{\epsilon,a}-\frac{H(x^{*})|M|\log\epsilon}{4\pi}\right)\frac{1}{K_{a}(1-|t^{\prime}|^{2})^{1/2}}=
ψϵ+|M|Ka2πϵ(H(x)2La1Rlog,a(λ1λ2)8La1R,a)La1(1)+OH1/2(𝔻)(logϵ).\displaystyle\psi^{\prime}_{\epsilon}+\frac{|M|K_{a}}{2\pi\epsilon}\left(\frac{H(x^{*})}{2}L_{a}^{-1}R_{\log,a}-\frac{(\lambda_{1}-\lambda_{2})}{8}L_{a}^{-1}R_{\infty,a}\right)L_{a}^{-1}(1)+O_{H^{1/2}(\mathbb{D})^{*}}(\log\epsilon).

Inserting the expression (5.20) into (5.9) we get that

(5.22) 𝔻ψϵ(s)𝑑s1𝑑s2=O(1).\displaystyle\int_{\mathbb{D}}\psi^{\prime}_{\epsilon}(s^{\prime})ds_{1}ds_{2}=O(1).

Multiply (5.21) by ϵ\epsilon then integrate over 𝔻\mathbb{D} . Then (5.22) implies

Cϵ,a\displaystyle C^{\prime}_{\epsilon,a} =\displaystyle= 14πH(x)|M|logϵ+aR(x,x)|M|F(x)\displaystyle-\frac{1}{4\pi}H(x^{*})|M|\log\epsilon+aR(x^{*},x^{*})|M|-F(x^{*})
|M|H(x)Ka216π3𝔻La1Rlog,aLa11(s)𝑑s\displaystyle-\frac{|M|H(x^{*})K_{a}^{2}}{16\pi^{3}}\int_{\mathbb{D}}L_{a}^{-1}R_{\log,a}L_{a}^{-1}1(s^{\prime})ds^{\prime}
+|M|(λ1λ2)Ka264π3𝔻La1R,aLa11(s)𝑑s+O(ϵlogϵ).\displaystyle+\frac{|M|(\lambda_{1}-\lambda_{2})K_{a}^{2}}{64\pi^{3}}\int_{\mathbb{D}}L_{a}^{-1}R_{\infty,a}L_{a}^{-1}1(s^{\prime})ds^{\prime}+O(\epsilon\log\epsilon).

Since La1L_{a}^{-1} is self-adjoint, we can express the last two integrals more explicitly:

𝔻La1Rlog,aLa11(s)𝑑s=Ka2(1|s|2)1/2,Rlog,a(1|s|2)1/2,\displaystyle\int_{\mathbb{D}}L_{a}^{-1}R_{\log,a}L_{a}^{-1}1(s^{\prime})ds^{\prime}=K_{a}^{-2}\langle(1-|s^{\prime}|^{2})^{-1/2},R_{\log,a}(1-|s^{\prime}|^{2})^{-1/2}\rangle,
𝔻La1R,aLa11(s)𝑑s=Ka2(1|s|2)1/2,R,a(1|s|2)1/2.\displaystyle\int_{\mathbb{D}}L_{a}^{-1}R_{\infty,a}L_{a}^{-1}1(s^{\prime})ds^{\prime}=K_{a}^{-2}\langle(1-|s^{\prime}|^{2})^{-1/2},R_{\infty,a}(1-|s^{\prime}|^{2})^{-1/2}\rangle.

We summarize this calculation into the following:

Proposition 5.2.

We have that

ψϵ=|M|2aπϵ2(1|t|2)1/2+ψϵ\psi_{\epsilon}=\frac{-|M|}{2a\pi\epsilon^{2}(1-|t^{\prime}|^{2})^{1/2}}+\psi_{\epsilon}^{\prime}

with ψϵ=OH1/2(𝔻)(ϵ1logϵ)\psi_{\epsilon}^{\prime}=O_{H^{1/2}(\mathbb{D})^{*}}(\epsilon^{-1}\log\epsilon). Furthermore

(5.24) Cϵ,a=\displaystyle C_{\epsilon,a}= |M|Ka4aϵπ214πH(x)|M|logϵ+aR(x,x)|M|F(x)\displaystyle\frac{|M|K_{a}}{4a\epsilon\pi^{2}}-\frac{1}{4\pi}H(x^{*})|M|\log\epsilon+aR(x^{*},x^{*})|M|-F(x^{*})
|M|H(x)16π3(1|s|2)1/2,Rlog,a(1|s|2)1/2\displaystyle-\frac{|M|H(x^{*})}{16\pi^{3}}\langle(1-|s^{\prime}|^{2})^{-1/2},R_{\log,a}(1-|s^{\prime}|^{2})^{-1/2}\rangle
+|M|(λ1λ2)64π3(1|s|2)1/2,R,a(1|s|2)1/2\displaystyle+\frac{|M|(\lambda_{1}-\lambda_{2})}{64\pi^{3}}\langle(1-|s^{\prime}|^{2})^{-1/2},R_{\infty,a}(1-|s^{\prime}|^{2})^{-1/2}\rangle
+O(ϵlogϵ),\displaystyle+O(\epsilon\log\epsilon),

where FF is the unique solution to (5.4) and R(x,x)R(x^{*},x^{*}) is the evaluation at (x,y)=(x,x)(x,y)=(x^{*},x^{*}) of the kernel R(x,y)R(x,y) in (1.1).

Observe that in the case of the disc (i.e. a=1a=1) we have that

(1|s|2)1/2,Rlog,a(1|s|2)1/2=π2(8log26)\langle(1-|s^{\prime}|^{2})^{-1/2},R_{\log,a}(1-|s^{\prime}|^{2})^{-1/2}\rangle=\pi^{2}\left({8}\log 2-{6}\right)

by Lemma 4.5 and

(1|s|2)1/2,R,a(1|s|2)1/2=0\langle(1-|s^{\prime}|^{2})^{-1/2},R_{\infty,a}(1-|s^{\prime}|^{2})^{-1/2}\rangle=0

by Lemma 4.6. Thus the formula (5.24) simplifies to

Cϵ:=Cϵ,1\displaystyle C_{\epsilon}:=C_{\epsilon,1} =\displaystyle= |M|Ka4aϵπ214πH(x)|M|logϵ+aR(x,x)|M|F(x)\displaystyle\frac{|M|K_{a}}{4a\epsilon\pi^{2}}-\frac{1}{4\pi}H(x^{*})|M|\log\epsilon+aR(x^{*},x^{*})|M|-F(x^{*})
\displaystyle- |M|H(x)16π(8log26)+O(ϵlogϵ),\displaystyle\frac{|M|H(x^{*})}{16\pi}\left(8\log 2-6\right)+O(\epsilon\log\epsilon),

when a=1a=1.

5.2. Proof of Theorems 1.2 and 1.3

We now prove Theorem 1.3. Theorem 1.2 follows from the explicit expression for CϵC_{\epsilon} in (5.1).

By the result of Appendix A we have that uϵ=𝔼[τΓϵ,a|X0=x]u_{\epsilon}=\mathbb{E}[\tau_{\Gamma_{\epsilon,a}}|X_{0}=x] solves the mixed boundary value problem (5.1) so using Proposition 5.2, (5.3), and (5.7), the expansion for 𝔼[τΓϵ,a|X0=x]\mathbb{E}[\tau_{\Gamma_{\epsilon,a}}|X_{0}=x] is given by

uϵ(x)=𝔼[τΓϵ,a|X0=x]=F(x)+Cϵ,a|M|G(x,x)+rϵ(x)u_{\epsilon}(x)=\mathbb{E}[\tau_{\Gamma_{\epsilon,a}}|X_{0}=x]=F(x)+C_{\epsilon,a}-|M|G(x,x^{*})+r_{\epsilon}(x)

for each xM\Γϵ,ax\in M\backslash\Gamma_{\epsilon,a}. Here FF is the unique solution to (5.4) and the remainder rϵr_{\epsilon} is given by

(5.26) rϵ(x)=M(G(x,y)G(x,x))νuϵ(y)dvolh(y).\displaystyle r_{\epsilon}(x)=\int_{\partial M}(G(x,y)-G(x,x^{*}))\partial_{\nu}u_{\epsilon}(y){\rm dvol}_{h}(y).

Let KM¯K\subset\overline{M} be a compact subset of M¯\overline{M} which has positive distance from xx^{*} and consider xKx\in K. Writing out this integral in the xϵ(;x)x^{\epsilon}(\cdot;x^{*}) coordinate system and use (5.8), (5.7), and the expression of ψϵ\psi_{\epsilon} derived in Proposition 5.2 we get

rϵ(x)\displaystyle r_{\epsilon}(x) =\displaystyle= ϵ𝔻|M|2π(1|s|2)1/2L(x,ϵs)(1+ϵ2Qϵ(s))𝑑s\displaystyle\epsilon\int_{\mathbb{D}}\frac{-|M|}{2\pi(1-|s^{\prime}|^{2})^{1/2}}L(x,\epsilon s^{\prime})(1+\epsilon^{2}Q_{\epsilon}(s^{\prime}))ds^{\prime}
+\displaystyle+ aϵ3𝔻ψϵ(s)L(x,ϵs)(1+ϵ2Qϵ(s))𝑑s\displaystyle a\epsilon^{3}\int_{\mathbb{D}}\psi^{\prime}_{\epsilon}(s^{\prime})L(x,\epsilon s^{\prime})(1+\epsilon^{2}Q_{\epsilon}(s^{\prime}))ds^{\prime}

for some function L(x,s)L(x,s^{\prime}) jointly smooth in (x,s)K×𝔻(x,s^{\prime})\in K\times\mathbb{D}. The second integral formally denotes the duality between H1/2(𝔻)H^{1/2}(\mathbb{D})^{*} and H1/2(𝔻)H^{1/2}(\mathbb{D}). The estimate for ψϵ\psi^{\prime}_{\epsilon} derived in Proposition 5.2 now gives for any integer kk and any compact set KK not containing xx^{*}, rϵCk(K)Ck,Kϵ\|r_{\epsilon}\|_{C^{k}(K)}\leq C_{k,K}\epsilon.

Our pseudodifferential characterization of GMG_{\partial M} also allows us to compute the asymptotic of the average Muϵ\int_{M}u_{\epsilon}. Indeed, integrating (5.3) over MM we get

(5.28) Muϵdvolg=MF(x)dvolg+Cϵ,a|M|+MΓϵ,aG(x,y)νuϵ(y)dvolh(y)dvolg(x).\int_{M}u_{\epsilon}{\rm dvol}_{g}=\int_{M}F(x){\rm dvol}_{g}+C_{\epsilon,a}|M|+\int_{M}\int_{\Gamma_{\epsilon,a}}G(x,y)\partial_{\nu}u_{\epsilon}(y){\rm dvol}_{h}(y){\rm dvol}_{g}(x).

We compute the last integral by noting that

v(x):=Γϵ,aG(x,y)νuϵ(y)dvolh(y)v(x):=\int_{\Gamma_{\epsilon,a}}G(x,y)\partial_{\nu}u_{\epsilon}(y){\rm dvol}_{h}(y)

is the unique solution to the Dirichlet boundary value problem:

(5.29) Δgv=0,v(x)M=Γϵ,aGM(x,y)uϵ(y)dvolh(y)H1/2(M).\displaystyle\Delta_{g}v=0,\ \ v(x)\mid_{\partial M}=\int_{\Gamma_{\epsilon,a}}G_{\partial M}(x,y)u_{\epsilon}(y){\rm dvol}_{h}(y)\in H^{1/2}(\partial M).

We concluded the boundary value is in H1/2H^{1/2} because GMΨcl1(M)G_{\partial M}\in\Psi^{-1}_{cl}(\partial M) by (1.1).

Let a sequence of smooth functions fjνuϵf_{j}\to\partial_{\nu}u_{\epsilon} in H1/2(M)H^{-1/2}(\partial M) and let vjv_{j} solve

Δgvj=0,vj(x)M=MGM(x,y)fj(y)dvolh(y).\Delta_{g}v_{j}=0,\ \ v_{j}(x)\mid_{\partial M}=\int_{\partial M}G_{\partial M}(x,y)f_{j}(y){\rm dvol}_{h}(y).

Standard elliptic theory shows that vjvv_{j}\to v in H1(M)H^{1}(M). Therefore

MΓϵ,aG(x,y)νuϵ(y)\displaystyle\int_{M}\int_{\Gamma_{\epsilon,a}}G(x,y)\partial_{\nu}u_{\epsilon}(y) =\displaystyle= limjMMG(x,y)fj(y)\displaystyle\lim_{j}\int_{M}\int_{\partial M}G(x,y)f_{j}(y)
=\displaystyle= limjMfj(y)MG(x,y)\displaystyle\lim_{j}\int_{\partial M}f_{j}(y)\int_{M}G(x,y)
=\displaystyle= limjMfj(y)F(y)\displaystyle\lim_{j}\int_{\partial M}f_{j}(y)F(y)
=\displaystyle= νuϵ,F=F(x)|M|+O(ϵ),\displaystyle\langle\partial_{\nu}u_{\epsilon},F\rangle=-F(x^{*})|M|+O(\epsilon),

where FF is the solution to the boundary value problem (5.4) and ,\langle\cdot,\cdot\rangle denotes the pairing between H1/2(M)H^{-1/2}(\partial M) and H1/2(M)H^{1/2}(\partial M). The last equality comes from (5.2), smoothness of FF, and supp(νuϵ)Γϵ,a{\rm supp}(\partial_{\nu}u_{\epsilon})\subset\Gamma_{\epsilon,a}. Inserting this into (5.28) we have

Muϵdvolg=MF(x)dvolg+Cϵ,a|M|F(x)|M|+O(ϵ).\int_{M}u_{\epsilon}{\rm dvol}_{g}=\int_{M}F(x){\rm dvol}_{g}+C_{\epsilon,a}|M|-F(x^{*})|M|+O(\epsilon).

The constant Cϵ,aC_{\epsilon,a} is given by Proposition 5.2.

6. Appendix A -Elliptic Equation for the first passage time

In this appendix we show that u(x):=𝔼[τΓ|X0=x]u(x):=\mathbb{E}[\tau_{\Gamma}|X_{0}=x] satisfies the boundary value problem (5.1). This is standard material but we could not find a suitable reference which precisely addresses our setting. As such we are including this appendix for the convenience of the reader.

Let (M,g,M)(M,g,\partial M) be an orientable compact connected Riemannian manifold with non-empty smooth boundary oriented by dvolg{\rm dvol}_{g}. Let also (Xt,x)(X_{t},\mathbb{P}_{x}) be the Brownian motion on MM starting at xx, that is, the stochastic process generated by the Laplace-Beltrami operator Δg\Delta_{g}. Let Γ\Gamma be a geodesic ball on M\partial M with radius ε>0\varepsilon>0. We denote by τΓ\tau_{\Gamma} the first time the Brownian motion XtX_{t} hits Γ\Gamma, that is

τΓ:=inf{t0:XtΓ}.\tau_{\Gamma}:=\inf\{t\geq 0:X_{t}\in\Gamma\}.

We set

𝒫Γ(t,x):=[τΓt|X0=x].\mathcal{P}_{\Gamma}(t,x):=\mathbb{P}[\tau_{\Gamma}\leq t|X_{0}=x].

Let us note that 𝒫Γ(t,x)\mathcal{P}_{\Gamma}(t,x) is the probability that the Brownian motion hits Γ\Gamma before or at time tt, and therefore, satisfies

(6.1) 𝒫Γ(0,x)=0,xMΓ,\mathcal{P}_{\Gamma}(0,x)=0,\quad x\in M\setminus\Gamma,
(6.2) 𝒫Γ(t,x)=1,(t,x)[0,)×Γ.\mathcal{P}_{\Gamma}(t,x)=1,\quad(t,x)\in[0,\infty)\times\Gamma.

Note that, for any compact subset ΓM\Gamma\subset M, it follows444Note that in [6] and [12], the authors consider the manifold together with its boundary, and Cc(M)C_{c}^{\infty}(M), C0(M)C_{0}^{\infty}(M) denote the set of smooth (up to the boundary) functions with compact support. In case of compact manifold, these sets coincide with C(M¯)C^{\infty}(\overline{M}).

Cap(Γ,M):=infuC(M¯),u|Γ=1Mu2dvolg=0.\text{Cap}(\Gamma,M):=\inf_{u\in C^{\infty}(\overline{M}),\left.u\right|_{\Gamma}=1}\int_{M}\|\nabla u\|^{2}{\rm dvol}_{g}=0.

Then, [13, Theorem 1.5] implies that (M,g,M)(M,g,\partial M) is parabolic, that is, the probability that the Brownian motion ever hits any compact set FF with non-empty interior is 11. Since ΓM\Gamma\subset\partial M is connected with non-empty interior on M\partial M, we can extend MM to a compact connected Riemannian manifold M~\tilde{M} such that M~M¯\overline{\tilde{M}\setminus M} is compact with non-empty interior and M~M¯M=Γ\overline{\tilde{M}\setminus M}\cap M=\Gamma. Note that, the Brownian motion, starting at any point MΓM\setminus\Gamma, hits M~M¯\overline{\tilde{M}\setminus M} if and only if it hits Γ\Gamma. Therefore, the parabolicity condition of (M,g)(M,g) gives

(6.3) limt𝒫Γ(t,x)=1,xM.\lim_{t\rightarrow\infty}\mathcal{P}_{\Gamma}(t,x)=1,\quad x\in M.

Further, let us define the mean first arrival time uu, as

(6.4) u(x):=𝔼[τΓ|X0=x]:=0t𝑑𝒫Γ(t,x),u(x):=\mathbb{E}[\tau_{\Gamma}|X_{0}=x]:=\int_{0}^{\infty}td\mathcal{P}_{\Gamma}(t,x),

where the integral is a Riemann-Stieltjes integral. To investigate uu, let us recall some properties of 𝒫Γ\mathcal{P}_{\Gamma}. By Remmark 2.1 in [6], it follows that

1𝒫Γ(t,x)=(etΔmix1)(x),1-\mathcal{P}_{\Gamma}(t,x)=\left(e^{t\Delta_{mix}}1\right)(x),

where etΔmixe^{t\Delta_{mix}} is the semigroup with infinitesimal generator Δmix\Delta_{mix}, and Δmix\Delta_{mix} is the Laplace operator Δg\Delta_{g} corresponding to the Dirichlet boundary condition on Γ\Gamma and Neumann boundary condition on MΓ\partial M\setminus\Gamma, which is defined as follows

(6.5) D(Δmix):={uH1(M):ΔguL2u|Γ=0,νu|Γc=0}\displaystyle\mathrm{D}(\Delta_{mix}):=\{u\in H^{1}(M):\;\Delta_{g}u\in L^{2}\;\left.u\right|_{\Gamma}=0,\;\left.\partial_{\nu}u\right|_{\Gamma^{c}}=0\}
(6.6) Δmixu=ΔguuD(Δmix).\displaystyle\Delta_{mix}u=\Delta_{g}u\quad u\in\mathrm{D}(\Delta_{mix}).

In (6.5) we define νuH1/2(M)\partial_{\nu}u\in H^{-1/2}(\partial M) using the same method for defining the Dirichlet to Neumann map. That is, for uH1(M)u\in H^{1}(M) such that ΔguL2(M)\Delta_{g}u\in L^{2}(M), the distribution νu|MH1/2(M)\partial_{\nu}u\left.\right|_{\partial M}\in H^{-1/2}(\partial M) acts on fH1/2(M)f\in H^{1/2}(\partial M) via

νu|M,f:=MΔugvf¯𝑑volg+Mg(du,dvf)𝑑volg,\langle\partial_{\nu}u\left.\right|_{\partial M},f\rangle:=\int_{M}\Delta u_{g}\overline{v_{f}}\;dvol_{g}+\int_{M}g(du,dv_{f})\;dvol_{g},

where vfH1(M)v_{f}\in H^{1}(M) is the harmonic extension of ff. We say that νu|ω¯=0\partial_{\nu}u\left.\right|_{\overline{\omega}}=0, for non-empty open set ωM\omega\subset\partial M, if νu|M,f|M=0\langle\partial_{\nu}u\left.\right|_{\partial M},f\left.\right|_{\partial M}\rangle=0 for all fH1/2(M)f\in H^{1/2}(\partial M) such that f|Mω¯=0f|_{\partial M\setminus\overline{\omega}}=0.

Note that if uu sufficiently regular, for instance uH2(M)u\in H^{2}(M), then νu|M,f\langle\partial_{\nu}u\left.\right|_{\partial M},f\rangle is equal to the boundary integral of νu|M\partial_{\nu}u\left.\right|_{\partial M} and ff.

In fact, Δmix\Delta_{mix} can be equivalently defined by quadratic form; see Proposition 7.1 in Appendix B. Moreover, Δmix\Delta_{mix} is the non-positive self-adjoint operator with the discrete spectrum, consisting of negative eigenvalues accumulating at -\infty; see Proposition 7.1 in Appendix. Hence, Δmix\Delta_{mix} satisfies the quadratic estimate

0tΔmix(1+t2Δmix2)1uL22dttCu2,\int_{0}^{\infty}\|t\Delta_{mix}(1+t^{2}\Delta_{mix}^{2})^{-1}u\|^{2}_{L^{2}}\frac{dt}{t}\leq C\|u\|^{2},

for some C>0C>0 and all uL2(M)u\in L^{2}(M); see for instance [19, p. 221]. Therefore, Δmix\Delta_{mix} admits the functional calculus defined in [21].

Remark 6.1.

The functional calculus in [21] is defined for a concrete operator, which is denoted by TT in the notation used in that article. However, Δmix\Delta_{mix} satisfy all necessary conditions to admit this functional calculus.

Therefore, the semigroup etΔmixe^{t\Delta_{mix}}, which is contracting by Hille-Yosida theorem [14, Theorem 8.2.3], can be defined as follows

etΔmixu=12πiγa,αetζ(ζΔmix)1u𝑑ζ,uL2(M),e^{t\Delta_{mix}}u=\frac{1}{2\pi i}\int_{\gamma_{a,\alpha}}e^{t\zeta}(\zeta-\Delta_{mix})^{-1}ud\zeta,\qquad u\in L^{2}(M),

where a(τ,0)a\in(\tau,0), α(0,π2)\alpha\in(0,\frac{\pi}{2}), and γa,α\gamma_{a,\alpha} is the anti-clockwise oriented curve:

γa,α:={ζ:Reζa, |Imζ|=|Reζa|tanα}.\gamma_{a,\alpha}:=\{\zeta\in\mathbb{C}:\textrm{Re}\zeta\leq a,\text{ }|\textrm{Im}\zeta|=|\textrm{Re}\zeta-a|\tan\alpha\}.

Let ε>0\varepsilon>0 such that a+ε<0a+\varepsilon<0. Then Δmix+ε\Delta_{mix}+\varepsilon is also a negative self-adjoint operator, and hence generates contracting semigroup, et(Δmix+ε)e^{t(\Delta_{mix}+\varepsilon)}, as above.

By definition, we obtain, for uL2(M)u\in L^{2}(M),

(6.7) etΔmixu\displaystyle e^{t\Delta_{mix}}u =12πiγa,αetζ(ζΔmix)1u𝑑ζ\displaystyle=\frac{1}{2\pi i}\int_{\gamma_{a,\alpha}}e^{t\zeta}(\zeta-\Delta_{mix})^{-1}ud\zeta
=etε2πiγa,αet(ζ+ε)(ζ+ε(Δmix+ε))1u𝑑ζ\displaystyle=\frac{e^{-t\varepsilon}}{2\pi i}\int_{\gamma_{a,\alpha}}e^{t(\zeta+\varepsilon)}(\zeta+\varepsilon-(\Delta_{mix}+\varepsilon))^{-1}ud\zeta
=etε2πiγa+ε,αetξ(ξ(Δmix+ε))1u𝑑ξ=etεet(Δmix+ε)u,\displaystyle=\frac{e^{-t\varepsilon}}{2\pi i}\int_{\gamma_{a+\varepsilon,\alpha}}e^{t\xi}(\xi-(\Delta_{mix}+\varepsilon))^{-1}ud\xi=e^{-t\varepsilon}e^{t(\Delta_{mix}+\varepsilon)}u,

where γa+ε,α=γa,α+ε{Reξ<0}\gamma_{a+\varepsilon,\alpha}=\gamma_{a,\alpha}+\varepsilon\subset\{\textrm{Re}\xi<0\}. Let f1f_{1} the constant function on MM equals 11. By Theorem 8.2.2 in [14], we know, for λ>0\lambda>0,

(λ(Δmix+ε))1f1=0eλtet(Δmix+ε)f1𝑑t.(\lambda-(\Delta_{mix}+\varepsilon))^{-1}f_{1}=\int_{0}^{\infty}e^{-\lambda t}e^{t(\Delta_{mix}+\varepsilon)}f_{1}dt.

Let us choose λ=ε\lambda=\varepsilon, then, by using (6.7), we obtain

Δmix1f1=0eεtet(Δmix+ε)f1𝑑t=0etΔmixf1𝑑t-\Delta_{mix}^{-1}f_{1}=\int_{0}^{\infty}e^{-\varepsilon t}e^{t(\Delta_{mix}+\varepsilon)}f_{1}dt=\int_{0}^{\infty}e^{t\Delta_{mix}}f_{1}dt

and hence,

(6.8) 01𝒫Γ(t,x)dt=(Δmix1f1)(x)<.\int_{0}^{\infty}1-\mathcal{P}_{\Gamma}(t,x)dt=-(\Delta_{mix}^{-1}f_{1})(x)<\infty.

Therefore, the dominated convergence theorem implies

limb0b(𝒫Γ(b,x)𝒫Γ(t,x))𝑑t=01𝒫Γ(t,x)dt<.\lim_{b\rightarrow\infty}\int_{0}^{b}\left(\mathcal{P}_{\Gamma}(b,x)-\mathcal{P}_{\Gamma}(t,x)\right)dt=\int_{0}^{\infty}1-\mathcal{P}_{\Gamma}(t,x)dt<\infty.

Hence, by using (6.4) and integration by parts, we obtain

u(x)\displaystyle u(x) =limb(𝒫Γ(b,x)b0b𝒫Γ(t,x)𝑑t)=limb0b(𝒫Γ(b,x)𝒫Γ(t,x))𝑑t<\displaystyle=\lim_{b\rightarrow\infty}\left(\mathcal{P}_{\Gamma}(b,x)b-\int_{0}^{b}\mathcal{P}_{\Gamma}(t,x)dt\right)=\lim_{b\rightarrow\infty}\int_{0}^{b}\left(\mathcal{P}_{\Gamma}(b,x)-\mathcal{P}_{\Gamma}(t,x)\right)dt<\infty
=01𝒫Γ(t,x)dt.\displaystyle=\int_{0}^{\infty}1-\mathcal{P}_{\Gamma}(t,x)dt.

Therefore, by (6.8), we obtain

Δmixu=f1=1.\Delta_{mix}u=-f_{1}=-1.

In particular, uD(Δmix)u\in\mathrm{D}(\Delta_{mix}), and hence,

uΓ=0,νuM\Γ=0.u\mid_{\Gamma}=0,\qquad\partial_{\nu}u\mid_{\partial M\backslash\Gamma}=0.

We see that (5.1) is satisfied.

7. Appendix B - Quadratic Form

Let (M,g,M)(M,g,\partial M) be a compact connected Riemannian manifold with non-empty smooth boundary. Let Γ\Gamma be a closed subset of M\partial M such that MΓc¯\partial M\setminus\overline{\Gamma^{c}} is a non-empty open set. Consider the quadratic form

(7.1) a[u,v]:=Mg(du,dv¯)dvolg,u,vD(a):={uH1(M):u|Γ=0}.\displaystyle a[u,v]:=\int_{M}g(du,d\overline{v}){\rm dvol}_{g},\qquad u,v\in\mathrm{D}(a):=\{u\in H^{1}(M):\;\left.u\right|_{\Gamma}=0\}.

Note that D(a)\mathrm{D}(a) is closed subspace of H1(M)H^{1}(M) containing H01(M)H_{0}^{1}(M) and a[,]a[\cdot,\cdot] is a non-negative, closed, densely defined form. Therefore, by Friedrichs Theorem 2.23 in [15], it generates a non-negative self-adjoint operator Δa-\Delta_{a} in L2(M)L^{2}(M) whose domain is contained in D(a)D(a) such that (Δau,u)L2(M)=a[u,u](-\Delta_{a}u,u)_{L^{2}(M)}=a[u,u] for uD(Δa)u\in\mathrm{D}(-\Delta_{a}).

Let us show that the resolvents of Δa-\Delta_{a} are compact. Assume that ss belongs to the resolvent set of Δa-\Delta_{a}. Since Δa(Δas)1:L2(M)L2(M)-\Delta_{a}(-\Delta_{a}-s)^{-1}:L^{2}(M)\rightarrow L^{2}(M) is bounded, it is suffices to show that D(Δa)D(-\Delta_{a}), endowed with the graph norm, compactly embedded into L2(M)L^{2}(M). Since, for uD(Δa)u\in D(-\Delta_{a}),

duL2(M)2=(du,du)L2(M)=a[u,u]=(Δau,u)L2(M)14(ΔauL2(M)+uL2(M))2,\|du\|^{2}_{L^{2}(M)}=(du,du)_{L^{2}(M)}=a[u,u]=(-\Delta_{a}u,u)_{L^{2}(M)}\leq\frac{1}{4}(\|-\Delta_{a}u\|_{L^{2}(M)}+\|u\|_{L^{2}(M)})^{2},

we see that any bounded sequence in D(Δa)D(-\Delta_{a}), endowed with the graph norm, is bounded in H1(M)H^{1}(M), and hence, it contains a Cauchy subsequence in L2(M)L^{2}(M) by Rellich-Kondrachov theorem. This implies that the resolvents of Δa-\Delta_{a} are compact, and hence, the spectrum of Δa-\Delta_{a} is discrete, consisting of non-negative eigenvalues accumulating at ++\infty. Assume that λ=0\lambda=0 is an eigenvalue, and let u0D(a)u_{0}\in\mathrm{D}(a) be a corresponding eigenfunction.

Then the Poincaré-Wirtinger inequality gives, for some C>0C>0,

u1|M|MudvolgL2CduL2=C(Δau,u)L2=0,\left\|u-\frac{1}{|M|}\int_{M}u{\rm dvol}_{g}\right\|_{L^{2}}\leq C\|du\|_{L^{2}}=C(-\Delta_{a}u,u)_{L^{2}}=0,

so that u0u_{0} is a constant in L2(M)L_{2}(M). Since u0H1(M)u_{0}\in H^{1}(M), we conclude that u0=constu_{0}=const in L2(M)L^{2}(\partial M), and hence, u0=0u_{0}=0 by choice of Γ\Gamma. Therefrore λ=0\lambda=0 is not an eigenvalue, and consequently, the spectrum of Δa-\Delta_{a} is positive. For sake of completeness, we prove the following well known result.

Proposition 7.1.

Let Δa-\Delta_{a} be the operator defined above and Δmix-\Delta_{mix} be the operator defined in Section 6, then Δa=Δmix-\Delta_{a}=-\Delta_{mix}. In particular, Δmix-\Delta_{mix} is a self-adjoint operator with the positive discrete spectrum accumulating at infinity.

Proof.

Assume that uu, vH1(M)v\in H^{1}(M) and Δgu\Delta_{g}u, ΔguL2(M)\Delta_{g}u\in L^{2}(M). Let VV be the harmonic extension of v|Mv\left.\right|_{\partial M}, then ω:=VvH01(M)\omega:=V-v\in H^{1}_{0}(M) and ΔgωL2(M)\Delta_{g}\omega\in L^{2}(M). By H2H^{2}-regularity, ωH2(M)\omega\in H^{2}(M). The generalized Green’s identity gives

MuΔgω𝑑volg+MΔguω𝑑volg=νu|M,ω|Mu|M,νω|M,-\int_{M}u\Delta_{g}\omega\;dvol_{g}+\int_{M}\Delta_{g}u\omega\;dvol_{g}=\langle\partial_{\nu}u\left.\right|_{\partial M},\omega\left.\right|_{\partial M}\rangle-\langle u\left.\right|_{\partial M},\partial_{\nu}\omega\left.\right|_{\partial M}\rangle,

where the first term of the right hand side vanishes since ωH01(M)\omega\in H_{0}^{1}(M). Therefore we get

0\displaystyle 0 =MuΔgω𝑑volg+u|M,νω|M+MΔguω𝑑volg\displaystyle=-\int_{M}u\Delta_{g}\omega\;dvol_{g}+\langle u\left.\right|_{\partial M},\partial_{\nu}\omega\left.\right|_{\partial M}\rangle+\int_{M}\Delta_{g}u\omega\;dvol_{g}
=Mg(du,dω¯)𝑑volg+MΔguω𝑑volg.\displaystyle=\int_{M}g(du,d\overline{\omega})\;dvol_{g}+\int_{M}\Delta_{g}u\omega\;dvol_{g}.

Hence, we obtain

νu|M,v|M=MΔguv¯𝑑volg+Mg(du,dv¯)𝑑volg\langle\partial_{\nu}u\left.\right|_{\partial M},v\left.\right|_{\partial M}\rangle=\int_{M}\Delta_{g}u\overline{v}\;dvol_{g}+\int_{M}g(du,d\overline{v})\;dvol_{g}

for uu, vH1(M)v\in H^{1}(M) and Δgu\Delta_{g}u, ΔguL2(M)\Delta_{g}u\in L^{2}(M).

Assume that uD(Δmix)D(a)u\in\mathrm{D}(\Delta_{mix})\subset\mathrm{D}(a) and vD(Δa)v\in\mathrm{D}(\Delta_{a}), then, by above formula,

a[u,v]=Mg(du,dv¯)𝑑volg=MΔguv¯𝑑volg+νu|M,v|M\displaystyle a[u,v]=\int_{M}g(du,d\overline{v})\;dvol_{g}=-\int_{M}\Delta_{g}u\overline{v}\;dvol_{g}+\langle\partial_{\nu}u\left.\right|_{\partial M},v\left.\right|_{\partial M}\rangle

Note that the last term vanishes since uD(Δmix)u\in\mathrm{D}(\Delta_{mix}) and vD(a)v\in\mathrm{D}(a), so that

a[u,v]=MΔguv¯𝑑volg=MΔmixuv¯𝑑volg.a[u,v]=-\int_{M}\Delta_{g}u\overline{v}\;dvol_{g}=-\int_{M}\Delta_{mix}u\overline{v}\;dvol_{g}.

Since this holds for all vD(Δa)v\in\mathrm{D}(\Delta_{a}), it follows from Theorem 2.1., in [15], that uD(Δa)u\in\mathrm{D}(\Delta_{a}) and Δau=Δmixu\Delta_{a}u=\Delta_{mix}u.

Conversely, assume that uD(Δa)u\in\mathrm{D}(\Delta_{a}), then Δgu=ΔauL2(M)-\Delta_{g}u=-\Delta_{a}u\in L^{2}(M). Then, it follows

νu|M,f=MΔugVf¯𝑑volg+Mg(du,dVf)𝑑volg=a[u,Vf]a[u,Vf]=0.\displaystyle\langle\partial_{\nu}u\left.\right|_{\partial M},f\rangle=\int_{M}\Delta u_{g}\overline{V_{f}}\;dvol_{g}+\int_{M}g(du,dV_{f})\;dvol_{g}=a[u,V_{f}]-a[u,V_{f}]=0.

for any fH1/2(M)f\in H^{1/2}(\partial M) such taht f|Γ=0f\left.\right|_{\Gamma}=0. This means that νu|Γc=0\partial_{\nu}\left.u\right|_{\Gamma^{c}}=0, so that uD(Δmix)u\in D(\Delta_{mix}) and Δau=Δmixu\Delta_{a}u=\Delta_{mix}u.

References

  • [1] Habib Ammari, Kostis Kalimeris, Hyeonbae Kang, and Hyundae Lee. Layer potential techniques for the narrow escape problem. J. Math. Pures Appl. (9), 97(1):66–84, 2012.
  • [2] Xinfu Chen and Avner Friedman. Asymptotic analysis for the narrow escape problem. SIAM J. Math. Anal., 43(6):2542–2563, 2011.
  • [3] Alexei F Cheviakov, Michael J Ward, and Ronny Straube. An asymptotic analysis of the mean first passage time for narrow escape problems: Part II: The sphere. Multiscale Modeling & Simulation, 8(3):836–870, 2010.
  • [4] F. G. Friedlander. Introduction to the theory of distributions. Cambridge University Press, Cambridge, 1982.
  • [5] Daniel Gomez and Alexei F. Cheviakov. Asymptotic analysis of narrow escape problems in nonspherical three-dimensional domains. Phys. Rev. E, 91:012137, Jan 2015.
  • [6] Alexander Grigor’yan and Laurent Saloff-Coste. Hitting probabilities for Brownian motion on Riemannian manifolds. J. Math. Pures Appl. (9), 81(2):115–142, 2002.
  • [7] R. Hiptmair, C. Jerez-Hanckes, and C. Urzúa-Torres. Closed-form inverses of the weakly singular and hypersingular operators on disks. Integral Equations Operator Theory, 90(1):Art. 4, 14, 2018.
  • [8] D Holcman and Z Schuss. Escape through a small opening: receptor trafficking in a synaptic membrane. Journal of Statistical Physics, 117(5-6):975–1014, 2004.
  • [9] D. Holcman and Z. Schuss. Escape through a small opening: receptor trafficking in a synaptic membrane. J. Statist. Phys., 117(5-6):975–1014, 2004.
  • [10] David Holcman and Zeev Schuss. Stochastic narrow escape in molecular and cellular biology. Analysis and Applications. Springer, New York, 2015.
  • [11] Lars Hőrmander. Fourier Integral Operators. Acta Math., 127 (1971), 79-183
  • [12] Joonas Ilmavirta and Gabriel P. Paternain. Functions of constant geodesic x-ray transform. Inverse Problems, 35(6):065002, 11, 2019.
  • [13] Debora Impera, Stefano Pigola, and Alberto G. Setti. Potential theory for manifolds with boundary and applications to controlled mean curvature graphs. J. Reine Angew. Math., 733:121–159, 2017.
  • [14] Jürgen Jost. Partial differential equations, volume 214 of Graduate Texts in Mathematics. Springer, New York, third edition, 2013.
  • [15] Tosio Kato. Perturbation theory for linear operators. Classics in Mathematics. Springer-Verlag, Berlin, 1995. Reprint of the 1980 edition.
  • [16] John M. Lee. Introduction to Riemannian manifolds, volume 176 of Graduate Texts in Mathematics. Springer, Cham, 2018. Second edition of [ MR1468735].
  • [17] John M. Lee and Gunther Uhlmann. Determining anisotropic real-analytic conductivities by boundary measurements. Comm. Pure Appl. Math., 42(8):1097–1112, 1989.
  • [18] P. A. Martin. Exact solution of some integral equations over a circular disc. J. Integral Equations Appl., 18(1):39–58, 2006.
  • [19] Alan McIntosh. Operators which have an HH_{\infty} functional calculus. In Miniconference on operator theory and partial differential equations (North Ryde, 1986), volume 14 of Proc. Centre Math. Anal. Austral. Nat. Univ., pages 210–231. Austral. Nat. Univ., Canberra, 1986.
  • [20] François Monard, Richard Nickl, and Gabriel P. Paternain. Efficient nonparametric Bayesian inference for XX-ray transforms. Ann. Statist., 47(2):1113–1147, 2019.
  • [21] Medet Nursultanov and Andreas Rosén. Evolution of time-harmonic electromagnetic and acoustic waves along waveguides. Integral Equations Operator Theory, 90(5):Paper No. 53, 32, 2018.
  • [22] Leonid Pestov and Gunther Uhlmann. On characterization of the range and inversion formulas for the geodesic X-ray transform. Int. Math. Res. Not., (80):4331–4347, 2004.
  • [23] Leonid Pestov and Gunther Uhlmann. Two dimensional compact simple Riemannian manifolds are boundary distance rigid. Ann. of Math. (2), 161(2):1093–1110, 2005.
  • [24] S Pillay, Michael J Ward, A Peirce, and Theodore Kolokolnikov. An asymptotic analysis of the mean first passage time for narrow escape problems: Part I: Two-dimensional domains. Multiscale Modeling & Simulation, 8(3):803–835, 2010.
  • [25] Xavier Saint Raymond Elementary Introduction to the Theory of Pseudodifferential Operators. Studies in Advanced Mathematics, 1991 CRC Press.
  • [26] Zeev Schuss. Theory and applications of stochastic differential equations. John Wiley & Sons, Inc., New York, 1980. Wiley Series in Probability and Statistics.
  • [27] Zeev Schuss, Amit Singer, and David Holcman. Narrow escape, part i. Journal of Statistical Physics, 122(41):437–463, 2006.
  • [28] Zeev Schuss, Amit Singer, and David Holcman. The narrow escape problem for diffusion in cellular microdomains. Proceedings of the National Academy of Sciences, 104(41):16098–16103, 2007.
  • [29] V. A. Sharafutdinov. Integral geometry of tensor fields. Inverse and Ill-posed Problems Series. VSP, Utrecht, 1994.
  • [30] Amit Singer, Z Schuss, and D Holcman. Narrow escape and leakage of Brownian particles. Physical Review E, 78(5):051111, 2008.
  • [31] Amit Singer, Zeev Schuss, and David Holcman. Narrow escape, part II: The circular disk. Journal of statistical physics, 122(3):465–489, 2006.
  • [32] Michael E. Taylor. Pseudodifferential Operators, volume 34 of PRINCETON MATHEMATICAL SERIES. Princeton University Press, first edition, 1981.
  • [33] Michael E. Taylor. Partial differential equations I. Basic theory, volume 115 of Applied Mathematical Sciences. Springer, New York, second edition, 2011.
  • [34] Michael E. Taylor. Partial differential equations II. Qualitative studies of linear equations, volume 116 of Applied Mathematical Sciences. Springer, New York, second edition, 2011.
  • [35] Lefeuvre Thibault. Sur la rigidité des variétés riemanniennes. 2019.