On the Mean First Arrival Time of Brownian Particles on Riemannian Manifolds
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, 92C371. Introduction
Let 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 without boundary oriented by the Riemannian volume form . Let also be the Brownian motion on with initial condition at , that is, the stochastic process generated by the Laplace-Beltrami operator (this article uses the convention with negative spectrum, where is the exterior derivative). For any open we denote by the first time the Brownian motion hits , that is
In the case when is a small elliptic window of eccentricity and size (to be made precise later), the narrow escape/mean first arrival time problem wishes to derive an asymptotic expansion as for the expected value of the first arrival time amongst all Brownian particles starting at . Another quantity of interest is the average expected value over :
Here denotes the Riemannian volume of with respect to the metric .
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 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 is a domain in with Euclidean metric and is a single small disk absorbing window, [27, 30] gave an expansion for the average of the expected first arrival time, averaged over , up to an unspecified term:
(1.1) |
Here, is the mean curvature of the boundary at the center of the absorbing window. The case when is a small elliptic window was also addressed in [27, 30].
When is a three dimensional ball with multiple circular absorbing windows on the boundary, an expansion capturing the explicit form of the 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 (up to a remainder vanishing as ) 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 on which satisfies
where is a outward pointing normal vector field and is the area of the boundary.
Singer-Schuss-Holcman in [30] highlighted the difficulty in obtaining a comprehensive singularity expansion of in a neighbourhood of the diagonal when is a bounded domain in , but it turns out that even when 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 , set to be the mean curvature of at , the geodesic distance on the boundary given by metric , the geodesic distance given by the metric , and
the scalar second fundamental quadratic form (see pages 235 and 381 of [16] for definitions).
i) The map
is well defined and extends to a map from for all whose Schwartz kernel we will denote by . Here the map is the trace map.
ii)
There exists an open neighbourhood of the diagonal
such that in this neighbourhood, the singularity structure of is given by:
where , for all , is called the regular part of the Green’s function and is the Hodge-star operator (i.e. rotation by on the surface ).
We recall the definition of the exponential map. Let be a geodesically complete manifold. For any and there exists a unique geodesic , defined on , such that , . The exponential map based at is then a map taking defined by
Observe that when 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 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 around a fixed point since the statement is cleaner:
Theorem 1.2.
Let be a smooth Riemannian manifold of dimension three with boundary and let be its volume.
i) Fix and let be a boundary geodesic ball centered at of geodesic radius . For each ,
with for any integer and compact set which does not contain . The function is the unique solution to the boundary value problem
The constant is, modulo an error of , given by
where is the evaluation at of the kernel in (1.1).
ii) One has that the integral of over satisfies
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 (only the mean curvature 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 which is the difference of principal curvatures.
To this end let be the unit eigenvectors of the shape operator at corresponding respectively to the principal curvatures . For fixed, let
be a small geodesic ellipse.
Theorem 1.3.
Let be a smooth Riemannian manifold of dimension three with boundary.
i) For each ,
with for any integer and compact set which does not contain . The function is the unique solution to the boundary value problem
The constant is given by
(1.3) | ||||
where and is the two dimensional unit disk centered at the origin.
ii) One has that the integral of over satisfies
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 is easy to see in this formula. The integral which multiplies turns out to vanish when 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 term in (1.1) is due to the fact that in Proposition 1.1 we have the expansion of all the way to a remainder , 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 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, denotes the space of smooth functions. We use notation for compactly suported smooth functions and for its dual. By , we denote the space of time continuously differentiable functions. The spce of functions from , whose th derivatives are Hölder continuous with exponent , is denoted by
Let be a smooth function on and for all we say that (or simply ) if for all multi-indices there are constants such that
(2.1) |
where , , and . These are the Kohn-Nirenberg symbols. This class of symbols contain the classical symbols, denoted by , which are defined by those satisfying
(2.2) |
where each are homogeneous in the sense that for all , and . The expression (2.2) means that for all ,
If we can define an operator by
(2.3) |
where 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 but extends to an automorphism of the tempered distributions . (See [4] for a comprehensive guide to distribution theory and definition of these spaces).
Operators taking which have the above representation are said to be in and are called pseudodifferential operators. For the symbol class , Lemma 1.1 in Chapter 7.1 of [34] extends to map .
The classical pseudodifferential operators are defined analogously by requiring that belongs to . Note that knowing the operator we can recover by the formula
(2.4) |
where . Note that if is the Schwartz kernel of the operator then
(2.5) |
Let be a compact manifold without boundary. An operator 111We shall avoid a discussion of distributional sections of density bundles by noting that in our setting of 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 if there exists coordinate covers and a partition of unity subordinate to such that the map
(2.6) |
from belongs to .
If we say that it belongs to the symbol class if
for all . The classical pseudodifferential operators and classical symbols 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) |
called the principal symbol map. For each it can be defined at each by taking a coordinate neighborhood containing and a which is identically near then considering the operator given in (2.6) for and . As the resulting operator in (2.6) is in with symbol , we may set
for all and . 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 then computing the inverse Fourier transform as in (2.5).
One important property of which we will use is that it respects the product structure of and :
(2.8) |
for and .
2.2. Coordinate Calculations
We make some calculations for some geometric objects which will naturally appear in the singularity expansion for . These identities will be useful in proving Proposition 1.1.
Let be a three dimensional Riemannian manifold with non-empty smooth boundary which inherits the metric . Denote by the scalar second fundamental form on the surface and be the mean curvature at . Let denote the unit-sphere bundle over ,
For any , let be principal directions (i.e. unit eigenvectors) of the induced shape operator with eigenvalues and . We will drop the dependence in from our notation when there is no ambiguity.
We choose and such that is a positive multiple of the volume form (see p.26 of [16] for the “musical isomorphism” notation of ♭ and ♯). Here we use to denote the outward pointing normal vector field so that it is consistent with most PDE literature. However, in defining and the shape operator we will follow geometry literature (e.g.[16]) and use the inward pointing normal so that the sphere embedded in would have positive mean curvature in our convention.
For two points there are two distances to consider. The first is the shortest path amongst those that stay on the boundary which we denote by and the other is the distance measured by paths allowing to enter , which we denote by . Clearly, .
For a fixed , we will denote by the geodesic disk of radius (with respect to the metric ) centered at and to be the Euclidean disk in of radius centered at the origin. In what follows will always be smaller than the injectivity radius of . Letting , we will construct a coordinate system by the following procedure:
Write near the origin as for . Define first 222for example it is obvious below that and are elements of the tangent space over as they are inserted into the argument of .
where denotes the time map of -geodesics with initial point and initial velocity . The coordinate is then an -geodesic coordinate system for a neighborhood of on the boundary surface . We can extend this to become a coordinate system for points in near so that is a boundary normal coordinate system with in 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 in place of . The boundary coordinate system has the advantage that the metric tensor can be expressed as
(2.9) |
where is the expression of the boundary metric in the -geodesic coordinate system . Note that and are symmetric positive definite and matrices varying smoothly with respect to the variable .
For sufficiently small we define the (rescaled) -geodesic coordinate by the following map
where is the unit disk in . We derive some coordinate expressions for some of the geometric objects we will consider later.
Lemma 2.1.
Let be the pullback metric of by on . Denote by
so that where . We have that
for matrix jointly smooth in . It also satisfies and
Proof.
Expressing using we have that (see e.g. [35] Lemma 4.8)
with symmetric, even under the map , and . Setting and using the fact that we are using normal coordinates we obtain . Now Taylor expanding in , we see that
The terms on left-hand and right-hand sides cancel, leaving
Divide through by and take the limit as we see that for any . ∎
Lemma 2.2.
We use the same notation for and as in Lemma 2.1. One has that
where (respectively ) denotes smooth functions of which vanish to first order as (respectively ).
Proof.
Corollary 2.3.
For sufficiently small we have that
for some smooth function in the variables . Here we use and .
Lemma 2.4.
Let be the coordinate system at the beginning of this section centered at . For sufficiently small, we have that
where and . Here is a smooth function of which vanishes at . The term is a smooth function in which vanishes to second order as .
Proof.
We begin with the identity that for any and ,
with given by (2.9). Now set and we have
where is even in and is a smooth function of which is even and vanishes to second order as . Observe that is odd in .
We now need to argue that . Setting in the above identity and using the fact that we are using boundary normal coordinates we have
Subtracting off the terms and dividing by we see that as ,
for all . We now use the fact that is odd to see that
∎
Corollary 2.5.
Using the expression and , we have that
where (respectively ) denotes smooth function of which vanishes to first order as (respectively ).
Corollary 2.6.
For sufficiently small we have that
for some smooth function in the variables . Here we use and .
Lemma 2.7.
In the coordinate given by ,
where denotes a smooth function of the variables which vanishes to order 2 at the origin.
Proof.
By Gauss Lemma, for near , where is the unit speed geodesic in from to . Therefore we have that
(2.10) |
In the coordinates given by the Christoffel symbols are
(2.11) |
Choose so that . By Lemma 2.4, is a smooth function of when we write in these coordinates. Let be the unique unit vector over so that the unit velocity geodesic in these coordinates starting at with initial direction reaches in time . In the coordinates given by , we want to argue that are smooth functions of . To do so, observe that in -geodesic coordinates centered around 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 since in . Using (2.10) and the expression for the metric (2.9) we have that in the chosen boundary normal coordinate system
(2.12) |
After time , the geodesic with initial position and initial unit velocity can be written in the coordinate as
for some remainder which has initial condition . Taylor expanding we have that
(2.13) |
for some depending smoothly on and .
Due to (2.12), we are particularly interested in the evolution of the component which solves the ODE
The last equality comes from (2.11).
Now set so that , we have from (2.13) that for ,
for functions which is smooth in and and thus smooth functions of the variable. Inserting this into (2.2) yields
for some function which is smooth in the variable . Inserting this expression into (2.13) for we have that
Using Lemma 2.4 we have that for some smooth function . We may thus write
where is a smooth function of which vanishes to order near the origin. Now use the fact that is the coordinate expression for the scalar second fundamental form with respect to the normal given by 333recall that we are using the convention where and shape operator are defined with respect to the inward pointing normal at the point (see Proposition 8.17 of [16]) and our coordinate system is chosen so that the shape operator is diagonalized at . Therefore,
In view of (2.12) we have proven the required identity. ∎
Let and define by
(2.15) | |||
Here is the Hodge star operator associated to the metric .
Lemma 2.8.
Let be a normal coordinate system for centred around and let denote the pullback metric tensor on on under this coordinate. Then for all sufficiently close to the origin,
where with being the Euclidean distance between and . The (resp ) term denotes a smooth map of which vanishes to order as (resp ).
Proof.
This comes from the fact that for some matrix smooth in
where . Therefore
for some smooth function which is .
A coordinate calculation yields that so
∎
Corollary 2.9.
Let and and be the eigenvalues of the shape operator at . Then,
i) in the coordinate system prescribed at the beginning of this section,
The (resp ) term denotes a smooth function of which vanishes to order as (resp ).
ii)In the coordinate systems prescribed at the beginning of this section,
where is smooth with derivatives of all orders uniformly bounded in .
Proof.
Since we have that ii) is a consequence of i). For i), we will only prove this for since the statement for can be obtained via a rotation.
Recall that the normal coordinate system is chosen so that at the coordinate vectors pushes forward under to become of eigenvectors of the shape operator. Because of this we have that the pull-back of under this coordinate system is given by where . Here denotes a smooth function of which vanishes at the origin. Using the expression derived in Lemma 2.8 for the vector in the coordinate given by and , we have the desired expression for . ∎
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 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 be a family of tempered distributions in depending smoothly on the parameter . That is, it is the distribution defined by for all . We denote by to be the Fourier transform with respect to the variable only.
Lemma 2.10.
Let be a smooth function on . For and we have that for any multi-index ,
Furthermore, is jointly smooth in .
Proof.
Let be identically near the origin. We can write for any positive integer and ,
The first integral is absolutely convergent by the cut-off. The second Fourier transform is also absolutely convergent owing to the fact that the integrand is smooth and makes the integrand decay quickly as provided is chosen large enough. ∎
Lemma 2.11.
Let be a family of whose support is uniformly bounded in and whose derivatives are also uniformly bounded in . Then for all ,
is a map bounded uniformly in from .
Proof.
We prove the estimates only for Schwartz functions . We first expand
(2.16) |
All terms are smooth in its variables with derivatives uniformly bounded in . Estimating the remainder term in (2.16) is easy:
For any positive integer we write . Choose so that there is sufficient smoothness in the integral kernel to integrate by parts the formula
We see from this that for a fixed positive integer we may choose large enough so that is uniformly bounded in .
For the integral involving the main term of (2.16), we write
Let be near the origin and write
(2.17) |
To see the mapping property of the first term of (2.17) we write it out in Cartesian coordinates
Since is smooth in with derivatives bounded uniformly in and is smoothing, we have that for any positive integer with bound uniform in .
The second term of (2.17) is a pseudodifferential operator with full symbol
and away from we can deduce from Lemma 2.10 that
for some with derivatives uniformly bounded in . The operator then has full symbol in with symbol seminorms bounded uniformly in . We can now apply Calderón-Vailancourt Theorem to deduce that is bounded uniformly in from . ∎
2.4. Symbol Computations
Compute symbol by taking the Fourier transform and multiply by . 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 with and one has that for ,
i) ,
ii)
iii) ,
iv) , .
Remark 2.13.
Note that we ignore the behaviour of near as they are irrelevant to the principal symbol computations we are interested in.
Lemma 2.14.
Let be a pseudodifferential operator on whose singularity along the diagonal is given by . Then .
Proof.
Lemma 2.15.
Let be a pseudodifferential operator on whose singularity near the diagonal is given by . Then where denotes the musical isomorphism from to induced by the boundary metric and is the Hodge star operator in this metric. Here we use to denote the quadratic form for .
Proof.
Proposition 2.16.
Let denote the mean curvature of at , the second fundamental form of at , and for . Define
(2.18) |
where is the Hodge star operator for the metric . Let be the operator defined by
Then with principal symbol given by
where denotes the raising of index with respect to the metric on .
Proof.
To see that is a classical DO, we use Lemma 2.1 and Corollary 2.9 to see that the coordinate expression for the integral kernel satisfies the polyhomogeneous conditions of Prop 2.8 in Chapter 7 of [34]. Therefore .
The principal symbol computation is done using normal coordinates. Fix and denote by
the normal coordinate system around . By a rotation we can choose the coordinates so that is an eigenvector of the shape operator at with eigenvalue .
According to Lemma 2.1 and Corollary 2.9, in these coordinates the terms of can be expressed as
(2.19) |
for close to the origin. Computing the principal symbol amounts to taking the inverse Fourier transform of the above expression, and observe the behaviour as . Use the formula in Lemma 2.12 , we obtain
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 is an open subset of a compact Riemannian manifold without boundary. Choose a manifold with boundary which compactly contains . For all , standard elliptic theory shows that there exists a unique solution to
The map is a continuous linear operator from and is therefore given by a Schwartz kernel which we call the Green’s function. Note that for any , if we fix then by definition
We formally write
(3.1) |
Note that if is a bounded domain in then can be chosen to be the flat torus and can be chosen to be for the appropriate constant .
Using standard elliptic parametrix construction in normal coordinates we express in the following way.
Lemma 3.1.
For all
(3.2) |
Here denotes the Schwartz kernel of an operator in .
Proof.
Let be a parametrix for the on the closed compact manifold without boundary meaning that
By ellipticity, for any we have
Therefore it suffices to show that
(3.3) |
where is defined by
for some smooth function satisfying , if and
Here is the injectivity radius of the closed compact Riemannian manifold .
By elliptic regularity (3.3) is equivalent to showing that
(3.4) |
Taking the adjoint and use the self-adjointness of both and , this is the same as
(3.5) |
Using the principal symbol map defined in (2.7) it amounts to showing that
(3.6) |
as an element of the quotient space . In fact, since the symbol is classical, we now choose to be the representative in the equivalence class which is positively homogeneous of degree .
For each and covector in the unit cosphere bundle we will show that
(3.7) |
as . Homogeneity would then ensure that for all which would then ensure (3.6).
To this end let be three orthonormal vectors and choose normal coordinate given by for . Let take the value in an open set containing the origin but . Similarly let take the value in an open set containing but .
Define the pullback operators by
where and are pullback by and respectively.
Thanks to the invariance of the principal symbol map under symplectomorphism, we have
(3.8) |
for all . We see then that (3.6) amounts to showing that satisfies
(3.9) |
Let and be the full symbol of and respectively. The full symbol of can be computed by the formula
where . Since we are using the normal coordinate around , and
where is a smooth function vanishing to order at . So
for some smooth and compactly supported function vanishing to order at . Computing the first term directly and treat the second term by expanding using Taylor expansion we see that
(3.10) |
Since is the Laplace operator in the coordinate given by we have that
(3.11) |
Composition calculus gives that if is the full symbol of the operator then
(3.12) |
Substituting into (3.12) the expression we have in (3.10), (3.11), and the fact that in normal coordinates for in a neighbourhood of the origin we have that the second term in (3.12) drops out. So the full symbol of at the point is
In light of (3.8), for each fixed ,
as . Therefore (3.7) is verified.
∎
For all we define as in [34] the operators by the following
(3.13) |
for . Note that is different from (see [34] Chapt 7 Sect 11)
Modulo lower order pseudodifferential operator, and are given by the integral kernels and respectively. Indeed, using (3.2) and equation (11.14) on page 38 of [34], we see that for in a neighbourhood of the diagonal,
(3.14) |
Using (3.1) we can construct the so called Neumann Green’s function on via the following procedure. For each fixed we can solve the following Neumann boundary value problem to obtain the correction term as a function of
Setting we get, for each fixed the unique solution (as a distribution in ) to
(3.15) |
Fix for the time being in the interior of and observe that is smooth in a neighbourhood of the singularity of the map and vice versa. Therefore we can integrate by parts the the expression to obtain
The boundary and orthogonality conditions in (3.15) ensures that the right side vanishes so we have
(3.16) |
Let denote the Dirichlet-to-Neumann map (see [17] for definition) whose range is precisely a codimension one subspace of which annihilates the constant function. By the orthogonality condition in (3.15), the behaviour of is uniquely characterized by its action on the range of . To this end, for , denote its harmonic extension by . Integrating by parts the expression for in the interior of we have
Observe that any has a unique decomposition for some constant function and satisfying . Therefore, using the orthogonality condition of (3.15) and taking the trace of (3) we see that the map
is well defined and takes . We denote this operator by and its Schwartz kernel by . Going back to (3) we see that
where
(3.18) |
is smoothing. In operator form this is
(3.19) |
Since is elliptic (see [17]) we can conclude that which maps for all . 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 is a parametrix for . The symbol expansion for 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 defined in(3.13) and using identity (11.58) of [34] we have
(3.20) |
Iterating this equation and using intertwining property (11.59) of [34] we get
(3.21) |
By (2.8) the principal symbol of the operator is simply the product of the principal symbols of with the principal symbol of . The leading singularities of the operators and 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 , the integral kernel of is given by
when are close to each other.
Inserting this into (3.21) we get that when are close to each other,
where is the Schwartz kernel of an operator in which we call the regular part of . Observe that since the principal symbol of is in and is dimension , Sobolev embedding yields that
(3.23) |
for all . The proof of Proposition 1.1 is now complete.
Remark 3.3.
Note that (1.1) peels off the "singular part" of the distribution and gives us the representation
with the singularity structure of explicitly given by (1.1). Inserting this representation of into (3.20) gives the following integral equation for the regular part :
Since , it is a compact operator which makes Fredholm with index zero. Therefore, numerically computing for amounts to solving a Fredholm boundary integral equation subject to the orthogonality condition
4. Inverting the Normal Operator
Let be a bounded convex domain with smooth boundary. We will analyze the mapping properties of the operator
(4.1) |
and its inverse. Methods do exist [18],[7] for the explicit expression of the inverse of when (which is sufficient for our setting). When is a two dimensional ellipse [27] computed explicitly the inverse of acting on the constant function.
The purpose of Section 4.1 is to provide a geometric perspective to the operator one of the advantages of which is that it provides an explicit formula for when 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 is not necessarily , this geometric point of view may also potentially provide ways to relate some of the quantities of interest to the geometry of .
Section 4.2 will provide some explicit formulas for the composition of with other operators in the case when , the two dimensional disk. Section 4.3 will do the same for when is the two dimensional ellipse although the formulas will not be as explicit.
4.1. Mapping Properties of L
Denote by
to be the set of inward pointing unit vectors on . Note that this is a closed submanifold of the sphere bundle and thus inherits its smooth structure. Define the X-Ray transform by where is the time it takes for a ray of unit velocity starting at to reach the boundary . Note that because is assumed to be convex, is a smooth function on . Furthermore, is injective by [23].
By [29] Theorem 4.2.1 this operator extends to an operator where is the measure given by . This space mapping property allows us to define the adjoint operator given by (see [23])
(4.2) |
when acting on smooth functions . This allows us to define a self-adjoint normal operator . It turns out that by [23] the Schwartz kernel of is precisely and therefore . Let be any smooth positive function on which is equal to near the boundary. By Theorem 2.2 and Theorem 4.4 of [20] respectively, we have that
(4.3) |
is a bijection and
(4.4) |
is a self-adjoint homeomorphism. Thus there exists a unique function such that which is equivalent to . To compute , observe that if we find such that
(4.5) |
for all then by (4.2) we would have . The solution of (4.5) is easy to compute explicitly when . Indeed, direct computation shows that choosing
(4.6) |
one satisfies (4.5). In particular if (which is the case we are interested in) we have that
(4.7) |
Remark 4.1.
This process of computing solution to by solving for unfortunately only works for . In fact, thanks to the rigidity result of [12], we know that is solvable iff . However, for more general domains the geometric view presented here could potentially allow one to apply the reconstruction formula for inverting [22] to solve explicitly. To do so one must first invert (which has a large kernel but is surjective) into the range of and it is not clear how to do this when .
4.2. Integrals Involving L inverse of 1.
We also define and to be operators with kernels and respectively.
Lemma 4.2.
The operators and are bounded maps from to .
Proof.
Observe that the integral kernels of both and extends naturally to kernels representing operators in which we denote by and respectively. We denote by to be the isomorphism obtained by the trivial extension. Let be identically on . Then we have that
(4.8) |
and the same holds for .
By Proposition 7.2.8 of [34] we have that both and are pseudodifferential operators of order . Therefore by (4.8) both and are bounded operators from to .
∎
The following lemma was proved in [7, Theorem 4.2].
Lemma 4.3.
For , it follows
(4.9) |
Proof.
By (4.4) is a self-adjoint homeomorphism. The result of this Lemma is a direct consequence. ∎
Lemma 4.4.
Let
then
Proof.
Note that is the fundamental solution for the Laplace operator in , therefore,
Since is radially symmetric, where . Writing the Laplace operator in polar coordinates, we get
Integration gives
and
Let us find and . Note that does not have singularity at , namely,
Therefore, the identities
as , implies that . Hence, putting into (4.4), gives
so that . ∎
Lemma 4.5.
The following identity holds
Lemma 4.6.
The following idetity holds
4.3. Explicit Formulas in 2 Dimensional Ellipse
We now compute the inverse of the map where the domain of integration is the two dimensional ellipse instead of a ball. A change of variable leads us to consider the operator
(4.11) |
acting on functions of the disk . By [27] we have that
(4.12) |
on where
By (4.4) this is the unique solution in to .
Next we denote
For general , the quantities and cannot be computed as explicitly as in the case when 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 ,
(5.1) |
which gives the compatibility condition
(5.2) |
All integrals on open subsets of are with respect to the volume form given by the metric .
Using Green’s formula as in [9], also in [1], we can deduce that for points , satisfies the integral equation
(5.3) |
where and solves the boundary value problem
(5.4) |
By Proposition 1.1 we can take the trace of (5.3) to the boundary and restrict to the open subset . Using (5.1) we see that
for . We now replace for with the expression in (1.1) to obtain
We will write this integral equation in the coordinate system given by
(5.6) |
where is the coordinate defined in Section 2.2. To simplify notation we will drop the in the notation and denote by simply .
Note that in these coordinates the volume form for is given by
(5.7) |
for some smooth function whose derivatives of all orders are bounded uniformly in . We denote
(5.8) |
The compatibility condition (5.2) written using the expression for the volume form (5.7) is then
(5.9) |
Let us unwrap the right hand side of (5.1) term by term in the coordinate given by . Write out the integral of the first term using the expression of the volume form (5.7) and the expression for in Corollary 2.6 and taking into account that the coordinate system is scaled by a factor as in (5.6) gives
(5.10) |
for some operator 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
with operator norm bounded uniformly in . From here on we will denote by any operator which takes whose operator norm is bounded uniformly in .
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) | |||||
where is defined at the very beginning of Section 4.2. Here denotes a function with norm vanishing to order . Note the volume for we use here is now the Euclidean one rather than 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) | |||||
where is defined in Section 4. Inserting into (5.1) the identities (5.10), (5.11), and (5.12) we have
We would like to approximate the integral kernel by the constant and this is the content of
Lemma 5.1.
Let be the operator defined by the integral kernel
Then
Proof.
Set for and small and extend it to be a smooth compactly supported kernel otherwise. The kernel for is then . Note that
(5.14) |
Observe that if which is identically on then the operator acting on distributions supported in is given by the Schwartz kernel
for and small.
Observe that is the integral kernel for an operator in . Applying Prop 2.8 in Chap 7 of [34] we can deduce that for all we may choose sufficiently large such that
where for each integer and multi-index , is a bounded continuous function of with value in the space of smooth (away from the origin) homogeneous distributions of degree , are homogenous polynomials of degree with coefficients which are smooth functions of , and for all multi-indices , bounded continuous function of with value in .
Using (5.14) along with the homogenenity degree of and we see that . Therefore, for the integral kernel of is
(5.15) | |||||
The kernel is sufficiently smooth. Therefore, by doing a Taylor expansion and using we see that the integral kernel takes with norm . The worst term in the polyhomogeneous expansion part of (5.15) happens when and this term is given by
where . Recall that both and are homogeneous of degree in so writing then applying Lemma 2.11 we have uniform estimates in for the kernels and . For the term involving , we use the fact that is a linear function in whose coefficients are smooth functions of . Therefore, if ,
where for are derivatives of the distribution with respect to . We refer the reader to (8.31) in Chapt 3 of [33] for the definition of the the distribution . From this we see that the integral kernel
also maps with uniform bound in . ∎
Due to Lemma 5.1 we can write (5.1) as
(5.16) | |||||
We hit both sides with and use (4.12) and (4.4) to get the identity
(5.17) | |||
for some with operator norm . Use the mapping properties from Lemma 4.2 we see that the right side can be inverted by Neumann series to deduce
(5.18) |
(5.21) | |||
Inserting the expression (5.20) into (5.9) we get that
(5.22) |
Multiply (5.21) by then integrate over . Then (5.22) implies
Since is self-adjoint, we can express the last two integrals more explicitly:
We summarize this calculation into the following:
Proposition 5.2.
5.2. Proof of Theorems 1.2 and 1.3
By the result of Appendix A we have that solves the mixed boundary value problem (5.1) so using Proposition 5.2, (5.3), and (5.7), the expansion for is given by
for each . Here is the unique solution to (5.4) and the remainder is given by
(5.26) |
Let be a compact subset of which has positive distance from and consider . Writing out this integral in the coordinate system and use (5.8), (5.7), and the expression of derived in Proposition 5.2 we get
for some function jointly smooth in . The second integral formally denotes the duality between and . The estimate for derived in Proposition 5.2 now gives for any integer and any compact set not containing , .
Our pseudodifferential characterization of also allows us to compute the asymptotic of the average . Indeed, integrating (5.3) over we get
(5.28) |
We compute the last integral by noting that
is the unique solution to the Dirichlet boundary value problem:
(5.29) |
We concluded the boundary value is in because by (1.1).
Let a sequence of smooth functions in and let solve
Standard elliptic theory shows that in . Therefore
where is the solution to the boundary value problem (5.4) and denotes the pairing between and . The last equality comes from (5.2), smoothness of , and . Inserting this into (5.28) we have
The constant is given by Proposition 5.2.
6. Appendix A -Elliptic Equation for the first passage time
In this appendix we show that 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 be an orientable compact connected Riemannian manifold with non-empty smooth boundary oriented by . Let also be the Brownian motion on starting at , that is, the stochastic process generated by the Laplace-Beltrami operator . Let be a geodesic ball on with radius . We denote by the first time the Brownian motion hits , that is
We set
Let us note that is the probability that the Brownian motion hits before or at time , and therefore, satisfies
(6.1) |
(6.2) |
Note that, for any compact subset , it follows444Note that in [6] and [12], the authors consider the manifold together with its boundary, and , denote the set of smooth (up to the boundary) functions with compact support. In case of compact manifold, these sets coincide with .
Then, [13, Theorem 1.5] implies that is parabolic, that is, the probability that the Brownian motion ever hits any compact set with non-empty interior is . Since is connected with non-empty interior on , we can extend to a compact connected Riemannian manifold such that is compact with non-empty interior and . Note that, the Brownian motion, starting at any point , hits if and only if it hits . Therefore, the parabolicity condition of gives
(6.3) |
Further, let us define the mean first arrival time , as
(6.4) |
where the integral is a Riemann-Stieltjes integral. To investigate , let us recall some properties of . By Remmark 2.1 in [6], it follows that
where is the semigroup with infinitesimal generator , and is the Laplace operator corresponding to the Dirichlet boundary condition on and Neumann boundary condition on , which is defined as follows
(6.5) | ||||
(6.6) |
In (6.5) we define using the same method for defining the Dirichlet to Neumann map. That is, for such that , the distribution acts on via
where is the harmonic extension of . We say that , for non-empty open set , if for all such that .
Note that if sufficiently regular, for instance , then is equal to the boundary integral of and .
In fact, can be equivalently defined by quadratic form; see Proposition 7.1 in Appendix B. Moreover, is the non-positive self-adjoint operator with the discrete spectrum, consisting of negative eigenvalues accumulating at ; see Proposition 7.1 in Appendix. Hence, satisfies the quadratic estimate
for some and all ; see for instance [19, p. 221]. Therefore, 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 in the notation used in that article. However, satisfy all necessary conditions to admit this functional calculus.
Therefore, the semigroup , which is contracting by Hille-Yosida theorem [14, Theorem 8.2.3], can be defined as follows
where , , and is the anti-clockwise oriented curve:
Let such that . Then is also a negative self-adjoint operator, and hence generates contracting semigroup, , as above.
By definition, we obtain, for ,
(6.7) | ||||
where . Let the constant function on equals . By Theorem 8.2.2 in [14], we know, for ,
Let us choose , then, by using (6.7), we obtain
and hence,
(6.8) |
Therefore, the dominated convergence theorem implies
Hence, by using (6.4) and integration by parts, we obtain
Therefore, by (6.8), we obtain
In particular, , and hence,
We see that (5.1) is satisfied.
7. Appendix B - Quadratic Form
Let be a compact connected Riemannian manifold with non-empty smooth boundary. Let be a closed subset of such that is a non-empty open set. Consider the quadratic form
(7.1) |
Note that is closed subspace of containing and is a non-negative, closed, densely defined form. Therefore, by Friedrichs Theorem 2.23 in [15], it generates a non-negative self-adjoint operator in whose domain is contained in such that for .
Let us show that the resolvents of are compact. Assume that belongs to the resolvent set of . Since is bounded, it is suffices to show that , endowed with the graph norm, compactly embedded into . Since, for ,
we see that any bounded sequence in , endowed with the graph norm, is bounded in , and hence, it contains a Cauchy subsequence in by Rellich-Kondrachov theorem. This implies that the resolvents of are compact, and hence, the spectrum of is discrete, consisting of non-negative eigenvalues accumulating at . Assume that is an eigenvalue, and let be a corresponding eigenfunction.
Then the Poincaré-Wirtinger inequality gives, for some ,
so that is a constant in . Since , we conclude that in , and hence, by choice of . Therefrore is not an eigenvalue, and consequently, the spectrum of is positive. For sake of completeness, we prove the following well known result.
Proposition 7.1.
Let be the operator defined above and be the operator defined in Section 6, then . In particular, is a self-adjoint operator with the positive discrete spectrum accumulating at infinity.
Proof.
Assume that , and , . Let be the harmonic extension of , then and . By -regularity, . The generalized Green’s identity gives
where the first term of the right hand side vanishes since . Therefore we get
Hence, we obtain
for , and , .
Assume that and , then, by above formula,
Note that the last term vanishes since and , so that
Since this holds for all , it follows from Theorem 2.1., in [15], that and .
Conversely, assume that , then . Then, it follows
for any such taht . This means that , so that and .
∎
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 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 -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.