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

Laplace Green’s Functions for Infinite Ground Planes with Local Roughness

Nail A. Gumerov and Ramani Duraiswami
University of Maryland Institute for Advanced Computer Studies, College Park, MD 20742
Abstract

The Green’s functions for the Laplace equation respectively satisfying the Dirichlet and Neumann boundary conditions on the upper side of an infinite plane with a circular hole are introduced and constructed. These functions enables solution of the boundary value problems in domains where the hole is closed by any surface. This approach enables accounting for arbitrary positive and negative ground elevations inside the domain of interest, which, generally, is not possible to achieve using the regular method of images. Such problems appear in electrostatics, however, the methods developed apply to other domains where the Laplace or Poisson equations govern. Integral and series representations of the Green’s functions are provided. An efficient computational technique based on the boundary element method with fast multipole acceleration is developed. A numerical study of some benchmark problems is presented.

keywords:
Green’s Function, Laplace’s Equation , Electrostatics , infinite ground
MSC:
[2010] 00-01, 99-00

1 Introduction

Many problems of physical interest involve fields satisfying Laplace’s equation with appropriate boundary conditions. Often, the boundaries of the domain might include surfaces of infinite extent, which for analytical convenience, are often considered to be infinite planes, so that they can be treated via the method of images. Examples include electrostatics, magnetostatics, conduction (or diffusion) from a surface at a fixed temperature (or concentration), potential flow of an ideal fluid over a body. Motivating the work reported here are the problems of electric and magnetic fields near an infinite ground, where the surface elevation departs from the plane in both the positive and negative directions, e.g., a ground with bumps and troughs.

Computation of electrostatic fields in scales where the effects of ground should be taken into account can be an important task for a number of applications, for example, for modeling fields in urban environments [1], fields generated by the transmission power lines [2], [3], and fields generated by lightning over rough ocean surface [4]. Scattering of higher frequency electromagnetic waves from rough surfaces is also of interest [5]. Note that if the wavelength is much larger than the scale of the roughness, electrostatic approximation for characterization of scattering properties of the surface is also valid.

Most of numerical methods used for such computations either neglect the presence of objects outside the computational domain or impose some boundary conditions on the domain surface (e.g., see monograph [6] and review [7]). The boundary elements can deliver solutions which provide decay of the electric potential at the infinity, while solving a boundary integral equation for the charge distribution only for the objects located inside the computational domain. It is relatively easy to account the effects of the infinite flat ground in such computations using the method of images [8]. Formally, this can be done by replacing the free space Green’s function in the boundary integral equation by the Green’s function for the semispace. This function takes zero at any point on the ground plane and so any convolution with such a Green’s function produces a solution satisfying the boundary condition on the ground. Some techniques for construction of Green’s functions for the Laplace equation in different 2D domains are developed (e.g., [9]), but they formulated in terms of complex variables and conformal mappings, which makes difficult to extend them for 3D.

In practice, the situation frequently appears to be more complicated as the ground has elevations and dips. The term “ground” can also include buildings or other construction, and can be modeled using surface boundary element meshes. In any case, such modeling is limited to the area where the solution is needed and detailed representation of the environment can be provided only inside the computational domain. The question then is how to take into account the effect of the environment outside the computational domain? One possibility is to assume periodic boundary conditions. In this case some “periodization” techniques for solution of the Laplace equation can be used (e.g., see [10]; quasi-periodic Green’s functions can be found in [11]). Another approach can be to model the environment outside some ball 𝐁\mathbf{B} of radius RR as a flat ground, while provide a detailed meshing of all objects and the ground inside 𝐁\mathbf{B} (sometimes called “locally rough surface”). In the present paper we take the latter approach.

A simple solution of the problem can be found if all points on the ground inside 𝐁\mathbf{B} are located above or on the level of the flat ground at infinity (which we can denote as zero level). In this case the method of images can be applied to any objects in the domain. However, if there exist some dips, or points below the zero level the method of images is not applicable, since it requires placing the image source on the side of the ground plane opposite to the side where the actual source is located, which creates a physically unacceptable (singular) solution. In a number of papers treatment of such situation using boundary integrals for the Helmholtz equation was considered (e.g., [12], [13]). In this paper, we take a different approach and derive an analogue of the Green’s function, which accounts for the effect of infinite flat ground plane outside the computational domain, and which has no singularities besides the source location inside 𝐁\mathbf{B}. This function is suitable for solution of problems with arbitrary location of boundaries inside 𝐁\mathbf{B}. We implemented the method and provide some numerical examples. Note that this method can also be applied to the Helmholtz equation.

The problem described above formally reduces to solution of the Laplace equation with the Dirichlet boundary conditions on the ground. In practice, there appear also situations when the zero Neumann boundary conditions should be imposed on an infinite surface. For example, if we wish to determine underwater electric fields generated by some sources, this can be formulated as the problem with zero Neumann boundary conditions on the sea level (since the electric conductivity of the sea water is much larger than that of the air). The ocean surface can be rough, so that can be taken into account for points close to the surface, while the ocean surface can be considered as flat in the far field (also, see [14], [15], [16]). Another example of the Neumann problem appears in aerodynamics, where the effects of ground should be taken into account to determine correctly the lift force for objects flying close to the ground [17]. For flat surfaces the Neumann problem can be also solved using the method of images, which causes the same problems as for the Dirichlet problem described above. The Green’s functions for the Dirichlet and Neumann boundary conditions are closely related. In this paper we address both problems.

2 Statement of the problem

Refer to caption
Figure 1: The problem geometry and notation. A solution needs to be obtained in domain Ω0\Omega_{0} bounded by the object surface SS, ground surface Sg0S_{g0} and a hemisphere of radius R0R_{0}. The ground surface is flat outside Ω0\Omega_{0}. An extended computational domain Ωe\Omega_{e}, which is a part of Ω\Omega located inside a sphere of radius ReR_{e} can be introduced for efficient computations.

First, we consider the Dirichlet problem. Consider an infinite ground surface SgS_{g} on which the electric potential ϕ\phi is zero and let there be some objects on it. Let the surface be SS the value VV of the potential is given (here and below we use dimensionless equations). In the electrostatics approximation, we have the following Dirichlet problem for the Laplace equation,

2ϕ(𝐲)\displaystyle\nabla^{2}\phi\left(\mathbf{y}\right) =\displaystyle= 0,𝐲Ω,\displaystyle 0,\quad\mathbf{y}\in\Omega, (1)
ϕ|𝐲Sg\displaystyle\left.\phi\right|_{\mathbf{y\in}S_{g}} =\displaystyle= 0,ϕ|𝐲S=V,ϕ||𝐲|==0,\displaystyle 0,\quad\left.\phi\right|_{\mathbf{y\in}S}=V,\quad\left.\phi\right|_{\left|\mathbf{y}\right|\mathbf{=}\infty}=0,

where Ω\Omega is an infinite domain in 3\mathbb{R}^{3} bounded by SS and SgS_{g}. The solution of this problem can be obtained using indirect boundary integral method, which represents the potential in the form of the single layer potential

ϕ(𝐲)=SSgσ(𝐱)G(𝐲,𝐱)𝑑S(𝐱),G(𝐲,𝐱)=14π|𝐲𝐱|,\phi\left(\mathbf{y}\right)=\int_{S\cup S_{g}}\sigma\left(\mathbf{x}\right)G\left(\mathbf{y},\mathbf{x}\right)dS\left(\mathbf{x}\right),\quad G\left(\mathbf{y},\mathbf{x}\right)=\frac{1}{4\pi\left|\mathbf{y}-\mathbf{x}\right|}, (2)

where G(𝐲,𝐱)G\left(\mathbf{y},\mathbf{x}\right) is the free-space Green’s function for the Laplace equation, and σ(𝐱)\sigma\left(\mathbf{x}\right) is the surface charge density, which can be obtained from the solution of the boundary integral equation,

SSgσ(𝐱)G(𝐲,𝐱)𝑑S(𝐱)={V(𝐲),𝐲S,0,𝐲Sg.\int_{S\cup S_{g}}\sigma\left(\mathbf{x}\right)G\left(\mathbf{y},\mathbf{x}\right)dS\left(\mathbf{x}\right)=\left\{\begin{array}[]{c}V\left(\mathbf{y}\right),\quad\mathbf{y}\in S,\\ 0,\quad\mathbf{y}\in S_{g}.\end{array}\right. (3)

Assume further that SgS_{g} can be modelled by a flat horizontal surface (zero elevation, z=0z=0), except for a finite part near the source, and in the region in which we wish to compute the solution in details. The ground plane in this region may have arbitrary positive and negative elevations. We assume that this region is contained inside a ball 𝐁0\mathbf{B}_{0} of radius R0R_{0} centered at the origin of the reference frame. It is also assumed that surface SS is located inside the ball. The ball partitions Ω\Omega into domains Ω0\Omega_{0} and Ω\Omega_{\infty} and SgS_{g} into Sg0S_{g0} and SgS_{g\infty} (see Fig. 1),

Ω0\displaystyle\Omega_{0} =\displaystyle= {𝐲Ω,|𝐲|<R0},Ω={𝐲Ω,|𝐲|R0}={𝐲,|𝐲|R0, z>0},\displaystyle\left\{\mathbf{y}\in\Omega,\quad\left|\mathbf{y}\right|<R_{0}\right\},\quad\Omega_{\infty}=\left\{\mathbf{y}\in\Omega,\quad\left|\mathbf{y}\right|\geqslant R_{0}\right\}=\left\{\mathbf{y},\quad\left|\mathbf{y}\right|\geqslant R_{0},\text{ }z>0\right\}, (4)
Sg0\displaystyle S_{g0} =\displaystyle= {𝐲Sg,|𝐲|<R0},Sg={𝐲Sg,|𝐲|R0}={𝐲,|𝐲|R0, z=0}.\displaystyle\left\{\mathbf{y}\in S_{g},\quad\left|\mathbf{y}\right|<R_{0}\right\},\quad S_{g\infty}=\left\{\mathbf{y}\in S_{g},\quad\left|\mathbf{y}\right|\geqslant R_{0}\right\}=\left\{\mathbf{y},\quad\left|\mathbf{y}\right|\geqslant R_{0},\text{ }z=0\right\}.\mathbf{\quad}

Note now that SgS_{g\infty} is an open surface, for which we have an “upper” side, Sg+S_{g\infty}^{+}, faced towards z>0z>0, and a “lower” side, SgS_{g\infty}^{-} faced towards z<0z<0. The boundary conditions need be satisfied on the upper side only. The Green’s function G(D)(𝐲,𝐱)G^{(D)}\left(\mathbf{y},\mathbf{x}\right) for the zero Dirichlet boundary conditions on Sg+S_{g\infty}^{+} can be introduced as the solution of the problem

𝐲2G(D)(𝐲,𝐱)\displaystyle\nabla_{\mathbf{y}}^{2}G^{(D)}\left(\mathbf{y},\mathbf{x}\right) =\displaystyle= δ(𝐲𝐱),𝐲,𝐱3,\displaystyle-\delta\left(\mathbf{y-x}\right),\quad\quad\mathbf{y,x}\in\mathbb{R}^{3}\mathbf{,} (5)
G(D)(𝐲,𝐱)|𝐲Sg+,𝐲𝐱\displaystyle\left.G^{(D)}\left(\mathbf{y},\mathbf{x}\right)\right|_{\mathbf{y\in}S_{g\infty}^{+},\mathbf{y}\neq\mathbf{x}} =\displaystyle= 0, G(D)(𝐲,𝐱)||𝐲|=,𝐲𝐱=0.\displaystyle 0,\text{\quad}\left.G^{(D)}\left(\mathbf{y},\mathbf{x}\right)\right|_{\left|\mathbf{y}\right|\mathbf{=}\infty,\mathbf{y}\neq\mathbf{x}}=0.

The solution of the Dirichlet problem (1) then can be sought in the form

ϕ(𝐲)=SSg0σ(D)(𝐱)G(D)(𝐲,𝐱)𝑑S(𝐱),\phi\left(\mathbf{y}\right)=\int_{S\cup S_{g0}}\sigma^{(D)}\left(\mathbf{x}\right)G^{(D)}\left(\mathbf{y},\mathbf{x}\right)dS\left(\mathbf{x}\right), (6)

where σ(D)(𝐱)\sigma^{(D)}\left(\mathbf{x}\right) satisfies the boundary integral equation

SSg0σ(D)(𝐱)G(D)(𝐲,𝐱)𝑑S(𝐱)={V(𝐲),𝐲S,0,𝐲Sg0.\int_{S\cup S_{g0}}\sigma^{(D)}\left(\mathbf{x}\right)G^{(D)}\left(\mathbf{y},\mathbf{x}\right)dS\left(\mathbf{x}\right)=\left\{\begin{array}[]{c}V\left(\mathbf{y}\right),\quad\mathbf{y}\in S,\\ 0,\quad\mathbf{y}\in S_{g0}.\end{array}\right. (7)

We seek to derive analytical expressions for, and methods for computation of, the function G(D)(𝐲,𝐱)G^{(D)}\left(\mathbf{y},\mathbf{x}\right).

Similarly, we can also get the Green’s function G(N)(𝐲,𝐱)G^{(N)}\left(\mathbf{y},\mathbf{x}\right) for the zero Neumann boundary conditions on SgS_{g\infty}, which can be found by solving

𝐲2G(N)(𝐲,𝐱)\displaystyle\nabla_{\mathbf{y}}^{2}G^{(N)}\left(\mathbf{y},\mathbf{x}\right) =\displaystyle= δ(𝐲𝐱),𝐲,𝐱3,\displaystyle-\delta\left(\mathbf{y-x}\right),\quad\quad\mathbf{y,x}\in\mathbb{R}^{3}\mathbf{,}\quad (8)
n(𝐲)G(N)(𝐲,𝐱)|𝐲Sg+,𝐲𝐱\displaystyle\left.\frac{\partial}{\partial n\left(\mathbf{y}\right)}G^{(N)}\left(\mathbf{y},\mathbf{x}\right)\right|_{\mathbf{y\in}S_{g\infty}^{+},\mathbf{y}\neq\mathbf{x}} =\displaystyle= 0, G(N)(𝐲,𝐱)||𝐲|=,𝐲𝐱=0.\displaystyle 0,\text{\quad}\left.G^{(N)}\left(\mathbf{y},\mathbf{x}\right)\right|_{\left|\mathbf{y}\right|\mathbf{=}\infty,\mathbf{y}\neq\mathbf{x}}=0.

Construction of this function allows one to solve the problem with mixed boundary conditions

2ϕ(𝐲)\displaystyle\nabla^{2}\phi\left(\mathbf{y}\right) =\displaystyle= 0,𝐲Ω,\displaystyle 0,\quad\mathbf{y}\in\Omega, (9)
n(𝐲)ϕ|𝐲Sg\displaystyle\left.\frac{\partial}{\partial n\left(\mathbf{y}\right)}\phi\right|_{\mathbf{y\in}S_{g}} =\displaystyle= 0,ϕ|𝐲S=V,ϕ||𝐲|==0,\displaystyle 0,\quad\left.\phi\right|_{\mathbf{y\in}S}=V,\quad\left.\phi\right|_{\left|\mathbf{y}\right|\mathbf{=}\infty}=0,

in form (6), where superscript (D)\left(D\right) should be changed to (N)\left(N\right) and the charge density can be determined from the boundary integral equation

SSg0σ(N)(𝐱)G(N)(𝐲,𝐱)𝑑S(𝐱)\displaystyle\int_{S\cup S_{g0}}\sigma^{(N)}\left(\mathbf{x}\right)G^{(N)}\left(\mathbf{y},\mathbf{x}\right)dS\left(\mathbf{x}\right) =\displaystyle= V(𝐲),𝐲S,\displaystyle V\left(\mathbf{y}\right),\quad\mathbf{y}\in S, (10)
n(𝐲)SSg0σ(N)(𝐱)G(N)(𝐲,𝐱)𝑑S(𝐱)\displaystyle\frac{\partial}{\partial n\left(\mathbf{y}\right)}\int_{S\cup S_{g0}}\sigma^{(N)}\left(\mathbf{x}\right)G^{(N)}\left(\mathbf{y},\mathbf{x}\right)dS\left(\mathbf{x}\right) =\displaystyle= 0,𝐲Sg0.\displaystyle 0,\quad\mathbf{y}\in S_{g0}.

Note that functions G(D)(𝐲,𝐱)G^{(D)}\left(\mathbf{y},\mathbf{x}\right) and G(N)(𝐲,𝐱)G^{(N)}\left(\mathbf{y},\mathbf{x}\right) defined above are not unique, since the boundary condition is specified only on Sg+S_{g\infty}^{+}, while an arbitrary boundary condition can be imposed on SgS_{g\infty}^{-}. Depending on the type of condition on SgS_{g\infty}^{-} the functions may or may not satisfy the symmetry typical for Green’s functions for Dirichlet problems for closed or two-sided open surfaces with homogeneous boundary conditions [18]. Despite this non-uniqueness any solution of problems (5) and (8) delivers the Green’s function suitable for representation of solutions in form (6 ), where the charge density is determined from Eqs (7) and (10). Indeed, these relations guarantee that the solution satisfies the Laplace equation and the boundary conditions, and, so, as G(D)G^{(D)} or G(N)G^{(N)} are selected the solution is unique.

We also remark that, despite G(D)(𝐲,𝐱)G^{(D)}\left(\mathbf{y},\mathbf{x}\right) and G(N)(𝐲,𝐱)G^{(N)}\left(\mathbf{y},\mathbf{x}\right) can be computed for any location of the source (𝐱\mathbf{x}) and evaluation (𝐲\mathbf{y}) points, for boundary integral representation of the solution we only consider the case when the source 𝐱\mathbf{x} is located inside the ball 𝐁\mathbf{B}. Moreover, 𝐲\mathbf{y} should be also located in 𝐁\mathbf{B} to solve the boundary integral equations (7) and (10). As soon as the charge density is determined from these equations the values of Green’s functions at arbitrary 𝐲\mathbf{y} can be used.

3 Solution

3.1 Single and double layer potentials

Solutions of the Dirichlet and Neumann problems (5) and (8) can be found using single and double layer potentials,

G(D)(𝐲,𝐱)\displaystyle G^{(D)}\left(\mathbf{y},\mathbf{x}\right) =\displaystyle= G(𝐲,𝐱)+K(D)(𝐲,𝐱),K(D)(𝐲,𝐱)=SgμD(𝐱,𝐱)G(𝐲,𝐱)n(𝐱)𝑑S(𝐱),\displaystyle G\left(\mathbf{y},\mathbf{x}\right)+K^{(D)}\left(\mathbf{y},\mathbf{x}\right),\quad K^{(D)}\left(\mathbf{y,x}\right)=\int_{S_{g\infty}}\mu_{D}\left(\mathbf{x}^{\prime}\mathbf{,x}\right)\frac{\partial G\left(\mathbf{y},\mathbf{x}^{\prime}\right)}{\partial n\left(\mathbf{x}^{\prime}\right)}dS\left(\mathbf{x}^{\prime}\right), (11)
G(N)(𝐲,𝐱)\displaystyle G^{(N)}\left(\mathbf{y},\mathbf{x}\right) =\displaystyle= G(𝐲,𝐱)+K(N)(𝐲,𝐱),K(N)(𝐲,𝐱)=SgσN(𝐱,𝐱)G(𝐲,𝐱)𝑑S(𝐱).\displaystyle G\left(\mathbf{y},\mathbf{x}\right)+K^{(N)}\left(\mathbf{y},\mathbf{x}\right),\quad K^{(N)}\left(\mathbf{y,x}\right)=\int_{S_{g\infty}}\sigma_{N}\left(\mathbf{x}^{\prime}\mathbf{,x}\right)G\left(\mathbf{y},\mathbf{x}^{\prime}\right)dS\left(\mathbf{x}^{\prime}\right).\quad

Here σN(𝐱,𝐱)\sigma_{N}\left(\mathbf{x}^{\prime}\mathbf{,x}\right) and μD(𝐱,𝐱)\mu_{D}\left(\mathbf{x}^{\prime}\mathbf{,x}\right) are the single and double layer densities, which should be found from the boundary conditions. These conditions show that

K(D)(𝐲,𝐱)|𝐲Sg+\displaystyle\left.K^{(D)}\left(\mathbf{y},\mathbf{x}\right)\right|_{\mathbf{y\in}S_{g\infty}^{+}} =\displaystyle= G(𝐲,𝐱)|𝐲Sg,\displaystyle-\left.G\left(\mathbf{y},\mathbf{x}\right)\right|_{\mathbf{y\in}S_{g\infty}}, (12)
K(N)(𝐲,𝐱)n(𝐲)|𝐲Sg+\displaystyle\left.\frac{\partial K^{(N)}\left(\mathbf{y},\mathbf{x}\right)}{\partial n\left(\mathbf{y}\right)}\right|_{\mathbf{y\in}S_{g\infty}^{+}} =\displaystyle= G(𝐲,𝐱)n(𝐲)|𝐲Sg.\displaystyle-\left.\frac{\partial G\left(\mathbf{y},\mathbf{x}\right)}{\partial n\left(\mathbf{y}\right)}\right|_{\mathbf{y\in}S_{g\infty}}.

The single and double layer potentials (11) obey the following symmetry property

K(D)(𝐲,𝐱)|𝐲Sg=K(D)(𝐲,𝐱)|𝐲Sg+,K(N)(𝐲,𝐱)|𝐲Sg=K(N)(𝐲,𝐱)|𝐲Sg+.\left.K^{(D)}\left(\mathbf{y,x}\right)\right|_{\mathbf{y\in}S_{g\infty}^{-}}=-\left.K^{(D)}\left(\mathbf{y,x}\right)\right|_{\mathbf{y\in}S_{g\infty}^{+}},\quad\left.K^{(N)}\left(\mathbf{y,x}\right)\right|_{\mathbf{y\in}S_{g\infty}^{-}}=\left.K^{(N)}\left(\mathbf{y,x}\right)\right|_{\mathbf{y\in}S_{g\infty}^{+}}. (13)

Furthermore, we note that the single and double layer densities are related to the jumps of the potentials and their normal derivatives on the surface as

μD(𝐲,𝐱)\displaystyle\mu_{D}\left(\mathbf{y,x}\right) =\displaystyle= K(D)(𝐲,𝐱)|𝐲Sg+K(D)(𝐲,𝐱)|𝐲Sg,\displaystyle\left.K^{(D)}\left(\mathbf{y,x}\right)\right|_{\mathbf{y\in}S_{g\infty}^{+}}-\left.K^{(D)}\left(\mathbf{y,x}\right)\right|_{\mathbf{y\in}S_{g\infty}^{-}}, (14)
σN(𝐲,𝐱)\displaystyle\sigma_{N}\left(\mathbf{y},\mathbf{x}\right) =\displaystyle= K(N)(𝐲,𝐱)n(𝐲)|𝐲SgK(N)(𝐲,𝐱)n(𝐲)|𝐲Sg+.\displaystyle\left.\frac{\partial K^{(N)}\left(\mathbf{y},\mathbf{x}\right)}{\partial n\left(\mathbf{y}\right)}\right|_{\mathbf{y\in}S_{g\infty}^{-}}-\left.\frac{\partial K^{(N)}\left(\mathbf{y},\mathbf{x}\right)}{\partial n\left(\mathbf{y}\right)}\right|_{\mathbf{y\in}S_{g\infty}^{+}}.

Symmetry (13) then provides

μD(𝐲,𝐱)\displaystyle\mu_{D}\left(\mathbf{y,x}\right) =\displaystyle= 2K(D)(𝐲,𝐱)|𝐲Sg+,\displaystyle 2\left.K^{(D)}\left(\mathbf{y},\mathbf{x}\right)\right|_{\mathbf{y\in}S_{g\infty}^{+}}, (15)
σN(𝐲,𝐱)\displaystyle\sigma_{N}\left(\mathbf{y},\mathbf{x}\right) =\displaystyle= 2K(N)(𝐲,𝐱)n(𝐲)|𝐲Sg+.\displaystyle-2\left.\frac{\partial K^{(N)}\left(\mathbf{y},\mathbf{x}\right)}{\partial n\left(\mathbf{y}\right)}\right|_{\mathbf{y\in}S_{g\infty}^{+}}.

Boundary conditions (12) complete the solution, which can be written in the integral form

K(D)(𝐲,𝐱)\displaystyle K^{(D)}\left(\mathbf{y,x}\right) =\displaystyle= 2SgG(𝐱,𝐱)G(𝐲,𝐱)n(𝐱)𝑑S(𝐱),\displaystyle-2\int_{S_{g\infty}}G\left(\mathbf{x}^{\prime}\mathbf{,x}\right)\frac{\partial G\left(\mathbf{y},\mathbf{x}^{\prime}\right)}{\partial n\left(\mathbf{x}^{\prime}\right)}dS\left(\mathbf{x}^{\prime}\right), (16)
K(N)(𝐲,𝐱)\displaystyle K^{(N)}\left(\mathbf{y,x}\right) =\displaystyle= 2SgG(𝐲,𝐱)G(𝐱,𝐱)n(𝐱)𝑑S(𝐱).\displaystyle 2\int_{S_{g\infty}}G\left(\mathbf{y},\mathbf{x}^{\prime}\right)\frac{\partial G\left(\mathbf{x}^{\prime},\mathbf{x}\right)}{\partial n\left(\mathbf{x}^{\prime}\right)}dS\left(\mathbf{x}^{\prime}\right).\quad

These equations show that 𝐱\mathbf{x} and 𝐲\mathbf{y} enter the expression in a non-symmetric way. However,  due to the symmetry of the free-space Green’s function, we have

K(N)(𝐲,𝐱)=K(D)(𝐱,𝐲),K^{(N)}\left(\mathbf{y,x}\right)=-K^{(D)}\left(\mathbf{x,y}\right), (17)

Therefore, we consider only computation of the function K(D)(𝐲,𝐱)K^{(D)}\left(\mathbf{y,x}\right).

3.2 Integral representations

Equations (16) already provide an integral representation of the solution. Since the domain of integration SS_{\infty} is simple, the integrals can be written in a more explicit form. Let us consider the Cartesian, cylindrical, and spherical coordinates of the points,

𝐫=(x,y,z)=(ρcosφ,ρsinφ,z)=r(sinθcosφ,sinθsinφ,cosθ),\mathbf{r}=\left(x,y,z\right)=\left(\rho\cos\varphi,\rho\sin\varphi,z\right)=r\left(\sin\theta\cos\varphi,\sin\theta\sin\varphi,\cos\theta\right), (18)

and mark the coordinates of 𝐱,𝐲,\mathbf{x,y,} and 𝐱\mathbf{x}^{\prime} with subscripts xx,yy, and the prime, respectively. Note then that for points 𝐱\mathbf{x}^{\prime} on the surface z=0z^{\prime}=0, so

z1|𝐱𝐲||z=0\displaystyle\left.\frac{\partial}{\partial z^{\prime}}\frac{1}{\left|\mathbf{x}^{\prime}-\mathbf{y}\right|}\right|_{z^{\prime}=0} =\displaystyle= zy|𝐱𝐲|3|z=0,\displaystyle\left.\frac{z_{y}}{\left|\mathbf{x}^{\prime}-\mathbf{y}\right|^{3}}\right|_{z^{\prime}=0}, (19)
|𝐱𝐱||z=0\displaystyle\left.\left|\mathbf{x}-\mathbf{x}^{\prime}\right|\right|_{z^{\prime}=0} =\displaystyle= (|𝐱|22𝐱𝐱+|𝐱|2)1/2|z=0=(ρ22ρxρcos(φφx)+rx2)1/2.\displaystyle\left.\left(\left|\mathbf{x}^{\prime}\right|^{2}-2\mathbf{x}^{\prime}\cdot\mathbf{x}+\left|\mathbf{x}\right|^{2}\right)^{1/2}\right|_{z^{\prime}=0}=\left(\rho^{\prime 2}-2\rho_{x}\rho^{\prime}\cos\left(\varphi^{\prime}-\varphi_{x}\right)+r_{x}^{2}\right)^{1/2}.

Using the expression for the free-space Green’s function (2), we obtain from Eq. (16)

K(D)(𝐲,𝐱)\displaystyle K^{(D)}\left(\mathbf{y},\mathbf{x}\right) =\displaystyle= K(D)(ρy,φy,zy;ρx,φx,zx)=zy8π202πR0f𝑑ρ𝑑φ,\displaystyle K^{(D)}\left(\rho_{y},\varphi_{y},z_{y};\rho_{x},\varphi_{x},z_{x}\right)=-\frac{z_{y}}{8\pi^{2}}\int_{0}^{2\pi}\int_{R_{0}}^{\infty}fd\rho^{\prime}d\varphi^{\prime}, (20)
f\displaystyle f =\displaystyle= ρ(ρ22ρyρcos(φφy)+ry2)3/2(ρ22ρxρcos(φφx)+rx2)1/2.\displaystyle\frac{\rho^{\prime}}{\left(\rho^{\prime 2}-2\rho_{y}\rho^{\prime}\cos\left(\varphi^{\prime}-\varphi_{y}\right)+r_{y}^{2}\right)^{3/2}\left(\rho^{\prime 2}-2\rho_{x}\rho^{\prime}\cos\left(\varphi^{\prime}-\varphi_{x}\right)+r_{x}^{2}\right)^{1/2}}.

The integrand has the following asymptotic behavior for large ρ\rho^{\prime}

f=1ρ3+1ρ4(ρxcos(φφx)+3ρycos(φφy))+O(rx2+ry2ρ5).f=\frac{1}{\rho^{\prime 3}}+\frac{1}{\rho^{\prime 4}}\left(\rho_{x}\cos\left(\varphi^{\prime}-\varphi_{x}\right)+3\rho_{y}\cos\left(\varphi^{\prime}-\varphi_{y}\right)\right)+O\left(\frac{r_{x}^{2}+r_{y}^{2}}{\rho^{\prime 5}}\right). (21)

This shows that the integral converges. The infinite domain can be truncated and we have

K(D)=Kt(D)zy8πR2(1+O(rx2+ry2R2)),Kt(D)=zy8π202πR0Rf𝑑ρ𝑑φ,K^{(D)}=K_{t}^{(D)}-\frac{z_{y}}{8\pi R_{\infty}^{2}}\left(1+O\left(\frac{r_{x}^{2}+r_{y}^{2}}{R_{\infty}^{2}}\right)\right),\quad K_{t}^{(D)}=-\frac{z_{y}}{8\pi^{2}}\int_{0}^{2\pi}\int_{R_{0}}^{R_{\infty}}fd\rho^{\prime}d\varphi^{\prime}, (22)

where RR_{\infty} is some cutoff radius. The second term of the asymptotic expansion in Eq. (21) does not contribute to the integral, since being integrated over the angle it produces zero. A closed form for numerical integration can be obtain by substituting the integration variable ρ=R0/η\rho^{\prime}=R_{0}/\eta. We have then from Eq. (20)

K(D)(𝐲,𝐱)=R02zy8π202π01ηdηdφQ(η,φφy,ρy,ry,R0)3Q(η,φφx,ρx,rx,R0),\displaystyle K^{(D)}\left(\mathbf{y},\mathbf{x}\right)=-\frac{R_{0}^{2}z_{y}}{8\pi^{2}}\int_{0}^{2\pi}\int_{0}^{1}\frac{\eta d\eta d\varphi^{\prime}}{Q\left(\eta,\varphi^{\prime}-\varphi_{y},\rho_{y},r_{y},R_{0}\right)^{3}Q\left(\eta,\varphi^{\prime}-\varphi_{x},\rho_{x},r_{x},R_{0}\right)}, (23)
Q(η,φ,ρ,r,R)=(r2η22Rρηcosφ+R2)1/2.\displaystyle Q\left(\eta,\varphi,\rho,r,R\right)=\left(r^{2}\eta^{2}-2R\rho\eta\cos\varphi+R^{2}\right)^{1/2}.

3.3 Spherical harmonic expansions

While the integral representations of K(D)(𝐲,𝐱)K^{(D)}\left(\mathbf{y},\mathbf{x}\right) enable computation of this function for arbitrary locations of the source and evaluation points, numerical integration can be relatively expensive and more efficient way to compute the integral can be of practical interest. Equation (7) shows that to solve the boundary integral equations K(D)(𝐲,𝐱)K^{(D)}\left(\mathbf{y},\mathbf{x}\right) should be evaluated only for the case 𝐱,\mathbf{x,} 𝐲Ω0\mathbf{y}\in\Omega_{0}. In this case the integral can be computed using spherical harmonic expansion. For a single source located at 𝐱Ω0\mathbf{x}\in\Omega_{0} and 𝐱Ω\mathbf{x}^{\prime}\in\Omega_{\infty} the following expansion of the free space Green’s function holds [19], [20]:

G(𝐱,𝐱)=n=0m=nn12n+1Rnm(𝐱)Snm(𝐱),|𝐱|>R0>|𝐱|.\!G\left(\mathbf{x,x}^{\prime}\right)\!=\sum_{n=0}^{\infty}\sum_{m=-n}^{n}\frac{1}{2n+1}R_{n}^{-m}\left(\mathbf{x}\right)S_{n}^{m}\left(\mathbf{x}^{\prime}\right),\quad\!\left|\mathbf{x}^{\prime}\right|\!>R_{0}>\left|\mathbf{x}\right|.\! (24)

Here Rnm(𝐫)R_{n}^{m}\left(\mathbf{r}\right) and Snm(𝐫)S_{n}^{m}\left(\mathbf{r}\right) are the regular and singular (at r=0)r=0) spherical basis functions

Rnm(𝐫)=rnYnm(θ,φ),Snm(𝐫)=rn1Ynm(θ,φ),R_{n}^{m}\left(\mathbf{r}\right)=r^{n}Y_{n}^{m}\left(\theta,\varphi\right),\quad S_{n}^{m}\left(\mathbf{r}\right)=r^{-n-1}Y_{n}^{m}\left(\theta,\varphi\right), (25)

where Ynm(θ,φ)Y_{n}^{m}\left(\theta,\varphi\right) are the orthonormal spherical harmonics,

Ynm(θ,φ)\displaystyle Y_{n}^{m}\left(\theta,\varphi\right) =\displaystyle= NnmPn|m|(ξ)eimφ,ξ=cosθ,\displaystyle N_{n}^{m}P_{n}^{\left|m\right|}(\xi)e^{im\varphi},\quad\xi=\cos\theta,\quad (26)
Nnm\displaystyle N_{n}^{m} =\displaystyle= (1)m2n+14π(n|m|)!(n+|m|)!,n=0,1,2,,m=n,,n.\displaystyle(-1)^{m}\sqrt{\frac{2n+1}{4\pi}\frac{(n-\left|m\right|)!}{(n+\left|m\right|)!}},\quad n=0,1,2,...,\quad m=-n,...,n.

where Pnm(ξ)P_{n}^{m}\left(\xi\right) are the associated Legendre functions defined by the Rodrigue’s formula (see [21]),

Pnm(ξ)=(1)m(1ξ2)m/22nn!dn+mdξn+m(1ξ2)n,n0,m0.P_{n}^{m}\left(\xi\right)=\frac{\left(-1\right)^{m}\left(1-\xi^{2}\right)^{m/2}}{2^{n}n!}\frac{d^{n+m}}{d\xi^{n+m}}\left(1-\xi^{2}\right)^{n},\quad n\geqslant 0,\quad m\geqslant 0. (27)

The derivatives of the basis functions can be expressed via the basis functions (e.g., [20])

zSnm(𝐫)=(2n+1)anmSn+1m(𝐫),\frac{\partial}{\partial z}S_{n}^{m}\left(\mathbf{r}\right)=-\left(2n+1\right)a_{n}^{m}S_{n+1}^{m}\left(\mathbf{r}\right),\quad (28)

where

anm\displaystyle a_{n}^{m} =\displaystyle= anm=(n+1+m)(n+1m)(2n+1)(2n+3),for n|m|.\displaystyle a_{n}^{-m}=\sqrt{\frac{(n+1+m)(n+1-m)}{\left(2n+1\right)\left(2n+3\right)}},\quad\text{for }n\geqslant\left|m\right|. (29)
anm\displaystyle a_{n}^{m} =\displaystyle= 0,for n<|m|.\displaystyle 0,\quad\text{for }n<\left|m\right|.

Thus, we have

G(𝐲,𝐱)z=n=0m=nnanmRnm(𝐲)Sn+1m(𝐱),|𝐱|>R0>|𝐲|,z=z(𝐱).\!\frac{\partial G\left(\mathbf{y,x}^{\prime}\right)}{\partial z^{\prime}}=-\sum_{n=0}^{\infty}\sum_{m=-n}^{n}a_{n}^{m}R_{n}^{-m}\left(\mathbf{y}\right)S_{n+1}^{m}\left(\mathbf{x}^{\prime}\right),\quad\!\left|\mathbf{x}^{\prime}\right|\!>R_{0}>\left|\mathbf{y}\right|,\quad z^{\prime}=z(\mathbf{x}^{\prime}). (30)

For the computation of the integrals in (16) only the values for z=0z^{\prime}=0, or θ=π/2\theta^{\prime}=\pi/2 are needed. Using the value for Pn|m|(0)P_{n}^{|m|}(0) [21], we have from (26)

Ynm(π2,φ)\displaystyle Y_{n}^{m}\left(\frac{\pi}{2},\varphi\right) =\displaystyle= Lnmeimφ,\displaystyle L_{n}^{m}e^{im\varphi}, (31)
Lnm\displaystyle L_{n}^{m} =\displaystyle= NnmPn|m|(0)=lnm(1)(n+|m|)/2(n+|m|)!2n(n|m|2)!(n+|m|2)!Nnm,\displaystyle N_{n}^{m}P_{n}^{|m|}(0)=l_{n}^{m}\frac{\left(-1\right)^{\left(n+\left|m\right|\right)/2}\left(n+\left|m\right|\right)!}{2^{n}\left(\frac{n-\left|m\right|}{2}\right)!\left(\frac{n+\left|m\right|}{2}\right)!}N_{n}^{m},
lnm\displaystyle l_{n}^{m} =\displaystyle= {0,n+m=2l+1,l=0,1,1,n+m=2l,l=0,1,.\displaystyle\left\{\begin{array}[]{c}0,\quad n+m=2l+1,\quad l=0,1,...\\ 1,\quad n+m=2l,\quad l=0,1,...\end{array}\right.. (34)

Hence,

Snm(𝐫)|𝐫S=Lnmρn1eimφ,(ρ|𝐫S=rsinθ|𝐫S=r|𝐫S)\left.S_{n}^{m}\left(\mathbf{r}\right)\right|_{\mathbf{r\in}S_{\infty}}=L_{n}^{m}\rho^{-n-1}e^{im\varphi},\quad\left(\left.\rho\right|_{\mathbf{r\in}S_{\infty}}=\left.r\sin\theta\right|_{\mathbf{r\in}S_{\infty}}=\left.r\right|_{\mathbf{r\in}S_{\infty}}\right) (35)

This shows that

SSn+1m(𝐫)Snm(𝐫)𝑑S(𝐫)\displaystyle\int_{S_{\infty}}S_{n^{\prime}+1}^{m^{\prime}}\left(\mathbf{r}\right)S_{n}^{m}\left(\mathbf{r}\right)dS\left(\mathbf{r}\right) =\displaystyle= Ln+1mLnmRρnn3ρ𝑑ρ02πei(m+m)φ𝑑φ\displaystyle L_{n^{\prime}+1}^{m^{\prime}}L_{n}^{m}\int_{R}^{\infty}\rho^{-n-n^{\prime}-3}\rho d\rho\int_{0}^{2\pi}e^{i\left(m+m^{\prime}\right)\varphi}d\varphi
=\displaystyle= 2πδm,mLn+1mLnmn+n+1Rnn1\displaystyle 2\pi\delta_{m^{\prime},-m}\frac{L_{n^{\prime}+1}^{m^{\prime}}L_{n}^{m}}{n+n^{\prime}+1}R^{-n-n^{\prime}-1}

where δmm\delta_{m^{\prime}m} is the Kronecker delta.

Using Eqs (24), (30), (35), and (3.3) we obtain from Eq. (16)

K(D)(𝐲,𝐱)\displaystyle K^{(D)}\left(\mathbf{y,x}\right) =\displaystyle= n=0m=nnn=0m=nn2anm2n+1Rnm(𝐱)Rnm(𝐲)SSn+1m(𝐱)Snm(𝐱)𝑑S(𝐱)\displaystyle\sum_{n=0}^{\infty}\sum_{m=-n}^{n}\sum_{n^{\prime}=0}^{\infty}\sum_{m^{\prime}=-n^{\prime}}^{n^{\prime}}\frac{2a_{n^{\prime}}^{m^{\prime}}}{2n+1}R_{n}^{-m}\left(\mathbf{x}\right)R_{n^{\prime}}^{-m^{\prime}}\left(\mathbf{y}\right)\int_{S_{\infty}}S_{n^{\prime}+1}^{m^{\prime}}\left(\mathbf{x}^{\prime}\right)S_{n}^{m}\left(\mathbf{x}^{\prime}\right)dS\left(\mathbf{x}^{\prime}\right)
=\displaystyle= n=0m=nnn=0InnmRnm(𝐱)Rnm(𝐲)=m=n=|m|+1n=|m|InnmRnm(𝐱)Rnm(𝐲)\displaystyle\sum_{n=0}^{\infty}\sum_{m=-n}^{n}\sum_{n^{\prime}=0}^{\infty}I_{n^{\prime}n}^{m}R_{n}^{-m}\left(\mathbf{x}\right)R_{n^{\prime}}^{m}\left(\mathbf{y}\right)=\sum_{m=-\infty}^{\infty}\sum_{n=\left|m\right|+1}^{\infty}\sum_{n^{\prime}=\left|m\right|}^{\infty}I_{nn^{\prime}}^{m}R_{n^{\prime}}^{-m}\left(\mathbf{x}\right)R_{n}^{m}\left(\mathbf{y}\right)

where

Innm=JnnmRnn1,Jnnm=4πanmLn+1mLnm(2n+1)(n+n+1).I_{nn^{\prime}}^{m}=J_{nn^{\prime}}^{m}R^{-n-n^{\prime}-1},\quad J_{nn^{\prime}}^{m}=\frac{4\pi a_{n}^{m}L_{n+1}^{m}L_{n^{\prime}}^{m}}{\left(2n^{\prime}+1\right)\left(n+n^{\prime}+1\right)}. (38)

Here we used symmetries anm=anma_{n}^{m}=a_{n}^{-m} and Lnm=LnmL_{n}^{m}=L_{n}^{-m}. Summation in the last sum of Eq (3.3) can start at n=|m|n=\left|m\right|, but at this value Ln+1m=0L_{n+1}^{m}=0 (see Eq. (31)). Note that Innm=InnmI_{nn^{\prime}}^{m}=I_{nn^{\prime}}^{-m}, but the coefficients are non-symmetric with respect to the lower indices, InnmInnmI_{nn^{\prime}}^{m}\neq I_{n^{\prime}n}^{m}. Indeed, we have

InnmInnm=(4π)2anmanmR2n2n2(2n+1)(2n+1)(n+n+1)2LnmLn+1mLnmLn+1m=0.I_{n^{\prime}n}^{m}I_{nn^{\prime}}^{m}=\frac{\left(4\pi\right)^{2}a_{n}^{m}a_{n^{\prime}}^{m}R^{-2n-2n^{\prime}-2}}{\left(2n+1\right)\left(2n^{\prime}+1\right)\left(n+n^{\prime}+1\right)^{2}}L_{n}^{m}L_{n+1}^{m}L_{n^{\prime}}^{m}L_{n^{\prime}+1}^{m}=0. (39)

This is due to the fact LnmLn+1m=0L_{n}^{m}L_{n+1}^{m}=0 (see definition (31)). Hence, if at some nn^{\prime} and nn we have Innm0I_{n^{\prime}n}^{m}\neq 0 (and this is true as not all InnmI_{n^{\prime}n}^{m} are zeros), then due to Eq. (39) Innm=0I_{nn^{\prime}}^{m}=0, means InnmI_{n^{\prime}n}^{m}\neq InnmI_{nn^{\prime}}^{m}. Non-symmetry of the coefficients also shows that K(D)(𝐲,𝐱)K^{(D)}\left(\mathbf{y,x}\right) is non-symmetric, K(D)(𝐲,𝐱)K(D)(𝐱,𝐲)K^{(D)}\left(\mathbf{y,x}\right)\neq K^{(D)}\left(\mathbf{x,y}\right) since the spherical harmonics form a complete orthogonal basis on a sphere (and Rnm(𝐲)R_{n}^{m}\left(\mathbf{y}\right) are proportional to them).

4 Numerical methods

4.1 Dimensionless forms

Since the Laplace equation does not have intrinsic length scales, we have for arbitrary RR (see Eqs (23) and (3.3))

K(D)(𝐲,𝐱;R)\displaystyle K^{(D)}\left(\mathbf{y},\mathbf{x;}R\right) =\displaystyle= 1RK~(D)(𝐲R,𝐱R),\displaystyle\frac{1}{R}\widetilde{K}^{(D)}\left(\frac{\mathbf{y}}{R},\frac{\mathbf{x}}{R}\right), (40)
K~(D)(𝐲,𝐱)\displaystyle\widetilde{K}^{(D)}\left(\mathbf{y},\mathbf{x}\right) =\displaystyle= zy8π202π01ηdηdφQ(η,φφy,ρy,ry,1)3Q(η,φφx,ρx,rx,1),\displaystyle-\frac{z_{y}}{8\pi^{2}}\int_{0}^{2\pi}\int_{0}^{1}\frac{\eta d\eta d\varphi^{\prime}}{Q\left(\eta,\varphi^{\prime}-\varphi_{y},\rho_{y},r_{y},1\right)^{3}Q\left(\eta,\varphi^{\prime}-\varphi_{x},\rho_{x},r_{x},1\right)},
K~(D)(𝐲,𝐱)\displaystyle\widetilde{K}^{(D)}\left(\mathbf{y},\mathbf{x}\right) =\displaystyle= n=0m=nnUnm(𝐱)Rnm(𝐲),Unm(𝐱)=n=|m|JnnmRnm(𝐱),(|𝐲|<1,|𝐱|<1).\displaystyle\sum_{n=0}^{\infty}\sum_{m=-n}^{n}U_{n}^{m}\left(\mathbf{x}\right)R_{n}^{m}\left(\mathbf{y}\right),\quad U_{n}^{m}\left(\mathbf{x}\right)=\sum_{n^{\prime}=\left|m\right|}^{\infty}J_{nn^{\prime}}^{m}R_{n^{\prime}}^{-m}\left(\mathbf{x}\right),\quad\left(\left|\mathbf{y}\right|<1,\left|\mathbf{x}\right|<1\right).

where K~(D)\widetilde{K}^{(D)} is the dimensionless function, which does not depend on parameter RR, which is seen from its integral representation. We put an extra argument RR to the dimensional function K(D)K^{(D)} to show its parametric dependence on RR. The series for K~(D)(𝐲,𝐱)\widetilde{K}^{(D)}\left(\mathbf{y},\mathbf{x}\right) and Unm(𝐱)U_{n}^{m}\left(\mathbf{x}\right)converge for 𝐱\mathbf{x} and 𝐲\mathbf{y} located inside the unit ball.

4.2 Improvement of convergence

Our numerical experiments show that computation of the kernel function via double integral representation (40) provides accurate results, but such computations are somehow slow as they require evaluation of integrals for every pair 𝐱\mathbf{x} and 𝐲\mathbf{y}. On the other hand, recursive computations of the spherical basis functions provide a fast procedure.

For practical use all infinite series should be truncated with some truncation number pp (n<p,n<p, n<pn^{\prime}<p), which should be selected based on the error bounds. The radius of convergence of series (40) is 11 both, with respect to 𝐱\mathbf{x} and 𝐲\mathbf{y}. This means that for points in Ω0\Omega_{0} located closely to the boundary with domain Ω\Omega_{\infty} the truncation number can be large, which may reduce the efficiency of computations.

Several tricks can be proposed to improve the convergence. First, one can simply implement integral and series representations of the kernel and switch between algorithms based on the values |𝐲|\left|\mathbf{y}\right| and |𝐱|\left|\mathbf{x}\right|. Second, we can effectively reduce the size of the domain, where K(D)(𝐲,𝐱)K^{(D)}\left(\mathbf{y},\mathbf{x}\right) should be computed. This is based on the observation that for the evaluation points located on plane z=0z=0 we have from Eqs (25), (26), (31), and (38)

JnnmRnm(𝐲)|zy=0=4πanmLnm(2n+1)(n+n+1)(Ln+1mLnm)ρyneimφy=0,J_{nn^{\prime}}^{m}\left.R_{n}^{m}\left(\mathbf{y}\right)\right|_{z_{y}=0}=\frac{4\pi a_{n}^{m}L_{n^{\prime}}^{m}}{\left(2n^{\prime}+1\right)\left(n+n^{\prime}+1\right)}\left(L_{n+1}^{m}L_{n}^{m}\right)\rho_{y}^{n}e^{-im\varphi_{y}}=0, (41)

due to Ln+1mLnm=0L_{n+1}^{m}L_{n}^{m}=0. Hence,

K(D)(𝐲,𝐱;R)=0,zy=0.K^{(D)}\left(\mathbf{y},\mathbf{x;}R\right)=0,\quad z_{y}=0. (42)

Also, relation (42) follows from the integral representation (20).

Consider now an extended computational domain, which can be denoted as Ωe\Omega_{e}, and which includes Ω0\Omega_{0} and the part of Ω\Omega_{\infty} located between the concentric spheres of radii R0R_{0} and Re>R0.R_{e}>R_{0}. Respectively, SgeS_{ge} denotes the part of ground surface SgS_{g} in domain Ωe\Omega_{e} (see Fig. 1). All formulae derived hold in case if we replace Ω0\Omega_{0} with Ωe\Omega_{e}, R0R_{0} with ReR_{e} and Sg0S_{g0} with SgeS_{ge}. So, instead of Eq. (6) we should have

ϕ(𝐲)=SSgeσe(D)(𝐱)[G(𝐲,𝐱)+K(D)(𝐲,𝐱;Re)]𝑑S(𝐱).\phi\left(\mathbf{y}\right)=\int_{S\cup S_{ge}}\sigma_{e}^{(D)}\left(\mathbf{x}\right)\left[G\left(\mathbf{y},\mathbf{x}\right)+K^{(D)}\left(\mathbf{y},\mathbf{x;}R_{e}\right)\right]dS\left(\mathbf{x}\right). (43)

The charge density σe(D)(𝐱)\sigma_{e}^{(D)}\left(\mathbf{x}\right) can be determined from the boundary conditions

SSgeσe(D)(𝐱)[G(𝐲,𝐱)+K(D)(𝐲,𝐱;Re)]𝑑S(𝐱)\displaystyle\int_{S\cup S_{ge}}\sigma_{e}^{(D)}\left(\mathbf{x}\right)\left[G\left(\mathbf{y},\mathbf{x}\right)+K^{(D)}\left(\mathbf{y},\mathbf{x;}R_{e}\right)\right]dS\left(\mathbf{x}\right) =\displaystyle= {V(𝐲),𝐲S,0,𝐲Sg0,,\displaystyle\left\{\begin{array}[]{c}V\left(\mathbf{y}\right),\quad\mathbf{y}\in S,\\ 0,\quad\mathbf{y}\in S_{g0},\end{array}\right., (46)
SSgeσe(D)(𝐱)G(𝐲,𝐱)𝑑S(𝐱)\displaystyle\int_{S\cup S_{ge}}\sigma_{e}^{(D)}\left(\mathbf{x}\right)G\left(\mathbf{y},\mathbf{x}\right)dS\left(\mathbf{x}\right) =\displaystyle= 0,𝐲Sge\Sg0.\displaystyle 0,\quad\mathbf{y}\in S_{ge}\backslash S_{g0}.

The latter holds since K(D)(𝐲,𝐱;Re)=0,K^{(D)}\left(\mathbf{y},\mathbf{x;}R_{e}\right)=0, for 𝐲Sge\Sg0\mathbf{y}\in S_{ge}\backslash S_{g0}. Hence, both for the field computations (43) and determination of the charge density only values of K(D)(𝐲,𝐱;Re)K^{(D)}\left(\mathbf{y},\mathbf{x;}R_{e}\right) for 𝐲Ω0\mathbf{y}\in\Omega_{0} are needed. On the other hand, K(D)(𝐲,𝐱;Re)K^{(D)}\left(\mathbf{y},\mathbf{x;}R_{e}\right) should be computed for 𝐱Ωe\mathbf{x}\in\Omega_{e} with a notice that for 𝐱Sge\Sg0\mathbf{x}\in S_{ge}\backslash S_{g0} we have zx=0.z_{x}=0. We developed an efficient recursive procedure for such computations described in the next subsection.

4.3 Computations for the ground points on the extended domain

So, the problem for use of the extended domain is to compute function Unm(𝐱)U_{n}^{m}\left(\mathbf{x}\right) represented in Eq. (40) by infinite series at zxz_{x} =0=0, at ξ[R0/Re,1),\xi\in\left[R_{0}/R_{e},1\right),where ξ=|𝐱|\xi=\left|\mathbf{x}\right|. Integral representation of Unm(𝐱)U_{n}^{m}\left(\mathbf{x}\right) can be obtained by substituting of Eq. (30) into integral (16). In the dimensionless form we have

K~(D)(𝐲,𝐱)=n=0m=nnUnm(𝐱)Rnm(𝐲)=n=0m=nn2anmRnm(𝐲)102πG(𝐱,𝐱)Sn+1m(𝐱)𝑑S(𝐱).\widetilde{K}^{(D)}\left(\mathbf{y,x}\right)=\sum_{n=0}^{\infty}\sum_{m=-n}^{n}U_{n}^{m}\left(\mathbf{x}\right)R_{n}^{m}\left(\mathbf{y}\right)=\sum_{n=0}^{\infty}\sum_{m=-n}^{n}2a_{n}^{m}R_{n}^{m}\left(\mathbf{y}\right)\int_{1}^{\infty}\int_{0}^{2\pi}G\left(\mathbf{x},\mathbf{x}^{\prime}\right)S_{n+1}^{-m}\left(\mathbf{x}^{\prime}\right)dS\left(\mathbf{x}^{\prime}\right).\quad (47)

So,

Unm(𝐱)|zx=0=12πanmLn+1meimφxunm(ξ),ξ=ρx,\left.U_{n}^{m}\left(\mathbf{x}\right)\right|_{z_{x}=0}=\frac{1}{2\pi}a_{n}^{m}L_{n+1}^{m}e^{-im\varphi_{x}}u_{n}^{m}\left(\xi\right),\quad\xi=\rho_{x}, (48)
unm(ξ)\displaystyle u_{n}^{m}\left(\xi\right) =\displaystyle= 1dηηn+102πeimφdφ(η22ξηcosφ+ξ2)1/2=1ξn+10ξζnwm(ζ)𝑑ζ,\displaystyle\int_{1}^{\infty}\frac{d\eta}{\eta^{n+1}}\int_{0}^{2\pi}\frac{e^{im\varphi}d\varphi}{\left(\eta^{2}-2\xi\eta\cos\varphi+\xi^{2}\right)^{1/2}}=\frac{1}{\xi^{n+1}}\int_{0}^{\xi}\zeta^{n}w_{m}\left(\zeta\right)d\zeta, (49)
wm(ξ)\displaystyle w_{m}\left(\xi\right) =\displaystyle= 02πeimφdφ(12ξcosφ+ξ2)1/2,unm(ξ)=unm(ξ),wm(ξ)=wm(ξ).\displaystyle\int_{0}^{2\pi}\frac{e^{im\varphi}d\varphi}{\left(1-2\xi\cos\varphi+\xi^{2}\right)^{1/2}},\quad u_{n}^{-m}\left(\xi\right)=u_{n}^{m}\left(\xi\right),\quad w_{-m}\left(\xi\right)=w_{m}\left(\xi\right).

These functions can be expressed via the complete elliptic integrals of the first and second kinds, KK and EE (see[21]), and recursive relations (to shorten notation we drop the argument ξ\xi for functions unm(ξ)u_{n}^{m}\left(\xi\right) and wm(ξ)w_{m}\left(\xi\right))

un0\displaystyle u_{n}^{0} =\displaystyle= 1n2ξ2[4E(ξ2)4n(1ξ2)K(ξ2)+(n1)2un20],n=1,3,5,\displaystyle\frac{1}{n^{2}\xi^{2}}\left[4E\left(\xi^{2}\right)-4n\left(1-\xi^{2}\right)K\left(\xi^{2}\right)+\left(n-1\right)^{2}u_{n-2}^{0}\right],\quad n=1,3,5,... (50)
un+11\displaystyle u_{n+1}^{1} =\displaystyle= 1(2n+3)ξ[(n+1)un0+(n+2)ξ2un+20+4(1ξ2)K(ξ2)8E(ξ2)],n=1,3,5,\displaystyle\frac{1}{\left(2n+3\right)\xi}\left[\left(n+1\right)u_{n}^{0}+\left(n+2\right)\xi^{2}u_{n+2}^{0}+4\left(1-\xi^{2}\right)K\left(\xi^{2}\right)-8E\left(\xi^{2}\right)\right],\quad n=1,3,5,... (51)
un+1m+1\displaystyle u_{n+1}^{m+1} =\displaystyle= 2(2n+3)ξ[(n+1)unm+(n+2)ξ2un+2mvm]un+1m1, m=1,2,,n=m+1,m+3,\displaystyle\frac{2}{\left(2n+3\right)\xi}\left[\left(n+1\right)u_{n}^{m}+\left(n+2\right)\xi^{2}u_{n+2}^{m}-v_{m}\right]-u_{n+1}^{m-1},\text{ }m=1,2,...,\,n=m+1,m+3,... (52)
vm\displaystyle v_{m} =\displaystyle= (1+ξ2)wmξ(wm+1+wm1),m=1,2,\displaystyle\left(1+\xi^{2}\right)w_{m}-\xi\left(w_{m+1}+w_{m-1}\right),\quad m=1,2,... (53)
wm\displaystyle w_{m} =\displaystyle= 1ξ(1+ξ2)2m22m1wm12m32m1wm2,m=2,3,,\displaystyle\frac{1}{\xi}\left(1+\xi^{2}\right)\frac{2m-2}{2m-1}w_{m-1}-\frac{2m-3}{2m-1}w_{m-2},\quad m=2,3,..., (54)
w0\displaystyle w_{0} =\displaystyle= 4K(ξ2),w1=4ξ[K(ξ2)E(ξ2)],\displaystyle 4K\left(\xi^{2}\right),\quad w_{1}=\frac{4}{\xi}\left[K\left(\xi^{2}\right)-E\left(\xi^{2}\right)\right], (55)
K(μ)\displaystyle K\left(\mu\right) =\displaystyle= 0π/2(1μsin2φ)1/2𝑑φ,E(μ)=0π/2(1μsin2φ)1/2𝑑φ.\displaystyle\int_{0}^{\pi/2}\left(1-\mu\sin^{2}\varphi\right)^{-1/2}d\varphi,\quad E\left(\mu\right)=\int_{0}^{\pi/2}\left(1-\mu\sin^{2}\varphi\right)^{1/2}d\varphi. (56)

Appendix A describes how these relations are obtained. We remark that because Ln+1m=0L_{n+1}^{m}=0 when n+mn+m is an even number, to obtain UnmU_{n}^{m} in Eq. (48) we need to compute functions unmu_{n}^{m} only at odd values of n+mn+m, which is reflected in the above recurrences. The recurrence for unmu_{n}^{m} works in layers in subscript nn. So, first, we obtain un0u_{n}^{0} for all nn needed, then un1u_{n}^{1}, and so on until unmu_{n}^{m} at maximum mm needed. For truncation number pp, we have max(n)=p1\left(n\right)=p-1, max(m)=p2\left(m\right)=p-2. As recurrence (52) show that the maximum computed nn at m=p3m=p-3 should be pp, while at m=0m=0 it should be 2p32p-3. It is also remarkable that recurrence (50) can be used first at n=1n=1, which provides the value of u10u_{1}^{0}, and then applied for other un0u_{n}^{0} in the sequence.

4.4 Boundary elements

Equation (46) can be solved using the method of moments (MoM) or collocation boundary element methods (BEM). In the present study we use the BEM. The surface is discretized by triangular elements, the integrals involving Green’s function are computed analytically assuming that the charge density is constant over the elements (constant panel approximation). The integrals involving kernel K(D)(𝐲,𝐱;Rc)K^{(D)}\left(\mathbf{y},\mathbf{x;}R_{c}\right) were computed using center panel quadrature,

Δjσ(𝐱)K(D)(𝐲,𝐱;Rc)𝑑S(𝐱)σjwjK(D)(𝐲,𝐱jc;Rc),\int_{\Delta_{j}}\sigma\left(\mathbf{x}\right)K^{(D)}\left(\mathbf{y},\mathbf{x;}R_{c}\right)dS\left(\mathbf{x}\right)\approx\sigma_{j}w_{j}K^{(D)}\left(\mathbf{y},\mathbf{x}_{j}^{c}\mathbf{;}R_{c}\right), (57)

where σj\sigma_{j} is the mean value of the charge density, while wjw_{j} and 𝐱jc\mathbf{x}_{j}^{c} are the area and the centroid of triangle Δj\Delta_{j}. In the standard BEM realization the integrals over the kernels with Green’s functions were computed analytically.

Δjσ(𝐱)G(𝐲,𝐱)𝑑S(𝐱)σjLj(𝐲),\int_{\Delta_{j}}\sigma\left(\mathbf{x}\right)G\left(\mathbf{y},\mathbf{x}\right)dS\left(\mathbf{x}\right)\approx\sigma_{j}L_{j}\left(\mathbf{y}\right), (58)
Lj(𝐲)\displaystyle L_{j}\left(\mathbf{y}\right) =\displaystyle= ΔjG(𝐲,𝐱)𝑑S(𝐱)=q=13[H(ljqxjq,hj,zjq)H(xjq,hj,zjq)],\displaystyle\int_{\Delta_{j}}G\left(\mathbf{y},\mathbf{x}\right)dS\left(\mathbf{x}\right)=\sum_{q=1}^{3}\left[H\left(l_{jq}-x_{jq},h_{j},z_{jq}\right)-H\left(-x_{jq},h_{j},z_{jq}\right)\right], (59)
xjq\displaystyle x_{jq} =\displaystyle= (𝐲𝐱jq)𝐢jq,hj=|(𝐲𝐱j1)𝐧j|,zjq=(𝐲𝐱jq)𝐧jq,\displaystyle\left(\mathbf{y-x}_{jq}\right)\cdot\mathbf{i}_{jq},\quad h_{j}=\left|\left(\mathbf{y-x}_{j1}\right)\cdot\mathbf{n}_{j}\right|,\quad z_{jq}=\left(\mathbf{y-x}_{jq}\right)\cdot\mathbf{n}_{jq},
𝐢jq\displaystyle\mathbf{i}_{jq} =\displaystyle= 1ljq(𝐱j,q(mod3)+1𝐱jq),ljq=|𝐱j,q(mod3)+1𝐱jq|,𝐧jq=𝐢jq×𝐧j,\displaystyle\frac{1}{l_{jq}}\left(\mathbf{x}_{j,q(\mathop{\mathrm{m}od}3)+1}-\mathbf{x}_{jq}\right),\quad l_{jq}=\left|\mathbf{x}_{j,q(\mathop{\mathrm{m}od}3)+1}-\mathbf{x}_{jq}\right|,\quad\mathbf{n}_{jq}=\mathbf{i}_{jq}\times\mathbf{n}_{j},
H(x,y,z)\displaystyle\quad H\left(x,y,z\right) =\displaystyle= y(arctanxzarctanyxzr)zln|r+x|,r=x2+y2+z2,\displaystyle y\left(\arctan\frac{x}{z}-\arctan\frac{yx}{zr}\right)-z\ln\left|r+x\right|,\quad r=\sqrt{x^{2}+y^{2}+z^{2}},

where 𝐧j\mathbf{n}_{j} is the normal to triangle Δj\Delta_{j} directed into the computational domain.

4.5 Fast multipole method

For large problems the BEM can be accelerated using the fast multipole method (FMM) (e.g., [22]). As shown, the kernel K(D)(𝐲,𝐱;Rc)K^{(D)}\left(\mathbf{y},\mathbf{x;}R_{c}\right) is either zero or can be factored for any pair of points inside the computational domain of radius RcR_{c}. This shows that boundary integral equation (46) can be written as

SSgeσe(D)(𝐱)G(𝐲,𝐱)𝑑S(𝐱)+\displaystyle\int_{S\cup S_{ge}}\!\!\!\!\!\!\!\!\!\!\sigma_{e}^{(D)}\left(\mathbf{x}\right)G\left(\mathbf{y},\mathbf{x}\right)dS\left(\mathbf{x}\right)+ (60)
1Ren=0p1m=nnRnm(𝐲Re)SSgeσe(D)(𝐱)Unm(𝐱Re)𝑑S(𝐱)={V(𝐲),𝐲S,0,𝐲Sge,\displaystyle\frac{1}{R_{e}}\sum_{n=0}^{p-1}\!\sum_{m=-n}^{n}R_{n}^{m}\left(\frac{\mathbf{y}}{R_{e}}\right)\int_{S\cup S_{ge}}\!\!\!\!\!\!\!\!\!\!\sigma_{e}^{(D)}\left(\mathbf{x}\right)U_{n}^{m}\left(\frac{\mathbf{x}}{R_{e}}\right)dS\left(\mathbf{x}\right)=\left\{\begin{array}[]{c}V\left(\mathbf{y}\right),\quad\mathbf{y}\in S,\\ 0,\quad\mathbf{y}\in S_{ge},\end{array}\right. (63)

where the infinite sum over nn is truncated. The first integral in the left hand side can be computed using the FMM with the free space Green’s function. In our implementation, we used some nearfield radius rnfr_{nf}. So, for pairs of points (𝐲,𝐱)\left(\mathbf{y},\mathbf{x}\right) located closer than rnfr_{nf} the integrals over triangles were computed analytically using Eqs (58) and (59), while for larger distances the integrals were computed using center point approximation, similarly to (57).

The most important feature of the global kernel factorization is that the number of operations to compute the matrix-vector product for the kernel is O(p2N)O\left(p^{2}N\right), not O(N2)O\left(N^{2}\right), since all integrals over 𝐱\mathbf{x} can be computed independently from the evaluation points 𝐲\mathbf{y} (NN is the number of boundary elements). So, computation of this sum has the asymptotic complexity consistent with the FMM, and the method can be applied for solution of large problems, where the overall linear system is solved iteratively.

4.6 Real basis functions

Kernel K~(D)\widetilde{K}^{(D)} is real, but is presented in Eq. (40) via sums of complex valued functions RnmR_{n}^{m}. A fast procedure can be developed based on recursive computations of real basis functions (see [23]),

Rnm¯(𝐫)=(1)n+m(n+|m|)!rnPn|m|(cosθ){cosmφ,m0,sinmφ,m<0,\underline{R_{n}^{m}}\left(\mathbf{r}\right)=\frac{\left(-1\right)^{n+m}}{(n+\left|m\right|)!}r^{n}P_{n}^{\left|m\right|}(\cos\theta)\left\{\begin{array}[]{c}\cos m\varphi,\quad m\geqslant 0,\\ \sin m\varphi,\quad m<0,\end{array}\right. (64)

which can be computed recursively as

R00¯\displaystyle\underline{R_{0}^{0}} =\displaystyle= 1,R11¯=12x,R11¯=12y,\displaystyle 1,\quad\underline{R_{1}^{1}}=-\frac{1}{2}x,\quad\underline{R_{1}^{-1}}=\frac{1}{2}y, (65)
Rmm¯\displaystyle\underline{R_{m}^{m}} =\displaystyle= 12m(xRm1m1¯+yRm1(m1)¯),m=2,3,,\displaystyle-\frac{1}{2m}\left(x\underline{R_{m-1}^{m-1}}+y\underline{R_{m-1}^{-\left(m-1\right)}}\right),\quad m=2,3,...,
Rmm¯\displaystyle\quad\underline{R_{m}^{-m}} =\displaystyle= 12m(yRm1m1¯xRm1(m1)¯),m=2,3,,\displaystyle\frac{1}{2m}\left(y\underline{R_{m-1}^{m-1}}-x\underline{R_{m-1}^{-\left(m-1\right)}}\right),\quad m=2,3,...,
Rm+1±m¯\displaystyle\underline{R_{m+1}^{\pm m}} =\displaystyle= zRm±m¯,m=0,1,,\displaystyle-z\underline{R_{m}^{\pm m}},\quad m=0,1,...,
Rn+1±m¯\displaystyle\underline{R_{n+1}^{\pm m}} =\displaystyle= (2n+1)zRn±m¯+r2Rn1±m¯(n+1)2m2,m=0,1,, n=m+1,m+2,.\displaystyle-\frac{\left(2n+1\right)z\underline{R_{n}^{\pm m}}+r^{2}\underline{R_{n-1}^{\pm m}}}{\left(n+1\right)^{2}-m^{2}},\quad m=0,1,...,\text{ }n=m+1,m+2,....\quad

Equations (40), (25), (26), (29), (31),and (38) show then that

K~(D)(𝐲,𝐱)\displaystyle\widetilde{K}^{(D)}\left(\mathbf{y},\mathbf{x}\right) =\displaystyle= n=0m=nnUnm¯(𝐱)Rnm¯(𝐲),Unm¯(𝐱)=2δm,04πνn+1mn=|m|νnmn+n+1Rnm¯(𝐱),\displaystyle\sum_{n=0}^{\infty}\sum_{m=-n}^{n}\underline{U_{n}^{m}}\left(\mathbf{x}\right)\underline{R_{n}^{m}}\left(\mathbf{y}\right),\quad\underline{U_{n}^{m}}\left(\mathbf{x}\right)=-\frac{2-\delta_{m,0}}{4\pi}\nu_{n+1}^{m}\sum_{n^{\prime}=\left|m\right|}^{\infty}\frac{\nu_{n^{\prime}}^{m}}{n^{\prime}+n+1}\underline{R_{n^{\prime}}^{m}}\left(\mathbf{x}\right), (66)
νnm\displaystyle\nu_{n}^{m} =\displaystyle= lnm(1)(n+|m|)/2(n|m|1)!!(n+|m|1)!!.\displaystyle l_{n}^{m}\left(-1\right)^{\left(n+\left|m\right|\right)/2}(n-\left|m\right|-1)!!(n+\left|m\right|-1)!!.

Also, using Eq. (48) we have for points zx=0z_{x}=0

Unm¯(𝐱)=2δm,08π2νn+1munm(ξ){cos(mφx),m0,sin(mφx),m<0.\underline{U_{n}^{m}}\left(\mathbf{x}\right)=-\frac{2-\delta_{m,0}}{8\pi^{2}}\nu_{n+1}^{m}u_{n}^{m}\left(\xi\right)\left\{\begin{array}[]{c}\cos\left(m\varphi_{x}\right),\quad m\geqslant 0,\\ \sin\left(m\varphi_{x}\right),\quad m<0.\end{array}\right. (67)

5 Complexity and optimization

5.1 Kernel computations

Assume there are NN sources and MM receivers in domain Ω0\Omega_{0}. Truncation of the series representation of K(D)(𝐲,𝐱;Re)K^{(D)}\left(\mathbf{y},\mathbf{x;}R_{e}\right) for 𝐲\mathbf{y\in} Ω0\Omega_{0} with truncation number pp produces errors ϵ(R0/Re)p\epsilon\sim(R_{0}/R_{e})^{p}. If there are NeN_{e} sources in the extended domain Ωe\Omega_{e} then the computational cost for factored K(D)(𝐲,𝐱;Re)K^{(D)}\left(\mathbf{y},\mathbf{x;}R_{e}\right) can be estimated as

Cfact=O(Mp2)+O(Np3)+O((NeN)p2),C_{fact}=O\left(Mp^{2}\right)+O\left(Np^{3}\right)+O\left((N_{e}-N)p^{2}\right), (68)

where the first term is the cost to compute p2p^{2} functions RnmR_{n}^{m}(𝐲/Re)\left(\mathbf{y/}R_{e}\right), the second term is the cost to compute p2p^{2} functions UnmU_{n}^{m}(𝐱/Re)\left(\mathbf{x/}R_{e}\right) for 𝐱Ω0\mathbf{x}\in\Omega_{0} using series (66), and the third term is the cost to compute p2p^{2} functions UnmU_{n}^{m}(𝐱/Re)\left(\mathbf{x/}R_{e}\right) for 𝐱Ωe\mathbf{x}\in\Omega_{e} \\backslashΩ0\Omega_{0} (zx=0)\left(z_{x}=0\right) using Eq. (67) and recursions (50)-(56) for unm(ξ)u_{n}^{m}\left(\xi\right). Assume now that the sources are uniformly distributed in domain Ωe\Omega_{e}\\backslashΩ0\Omega_{0}, with the density ρ\rho points per surface area. In this case NeN=πρ(Re2R02)N_{e}-N=\pi\rho\left(R_{e}^{2}-R_{0}^{2}\right). For prescribed accuracy ϵ\epsilon we have Re=R0ϵ1/pR_{e}=R_{0}\epsilon^{-1/p}, which brings the following expression for the computational cost

Cfact=AMp2+BNp3+CπρR02(ϵ2/p1)p2,C_{fact}=AMp^{2}+BNp^{3}+C\pi\rho R_{0}^{2}\left(\epsilon^{-2/p}-1\right)p^{2}, (69)

where A,BA,B, and CC are some asymptotic constants.

Considering CfactC_{fact} as a function of pp, we can see that

dCfactdp=2AMp+3BNp2+2CπρR02p[(ϵ2/p1)ϵ2/p1pln1ϵ].\frac{dC_{fact}}{dp}=2AMp+3BNp^{2}+2C\pi\rho R_{0}^{2}p\left[\left(\epsilon^{-2/p}-1\right)-\epsilon^{-2/p}\frac{1}{p}\ln\frac{1}{\epsilon}\right]. (70)

The term in the square brackets has a single zero at p=pc(ϵ)p=p_{c}\left(\epsilon\right) and is positive at

p>pc(ϵ)=1αln1ϵ,α=0.796812130.8.p>p_{c}\left(\epsilon\right)=\frac{1}{\alpha}\ln\frac{1}{\epsilon},\quad\alpha=0.79681213...\approx 0.8. (71)

So, function Cfact(p)C_{fact}\left(p\right) grows monotonically at p>pc(ϵ)p>p_{c}\left(\epsilon\right) for any values of M,N,M,N, and πρR02.\pi\rho R_{0}^{2}. The function can grow or decay at p<pc(ϵ)p<p_{c}\left(\epsilon\right) depending on M,N,M,N, and πρR02\pi\rho R_{0}^{2}. Note now, that p=pc(ϵ)p=p_{c}\left(\epsilon\right) corresponds to

(ReR0)c=ϵ1/pc=eα=β2.2.\left(\frac{R_{e}}{R_{0}}\right)_{c}=\epsilon^{-1/p_{c}}=e^{\alpha}=\beta\approx 2.2. (72)

This means, that CfactC_{fact} being considered as a function of ReR_{e} decays in the region R0<ReβR0R_{0}<R_{e}\leqslant\beta R_{0} and may continue to decay or start to grow at Re>βR0R_{e}>\beta R_{0}. It is remarkable, that β\beta does not depend on ϵ\epsilon, so for computations using the factored kernel the optimal radius of the extended domain should be always larger than βR0\beta R_{0}. Also, it can be noted that the values of pc(ϵ)p_{c}\left(\epsilon\right) are rather small. For example, at ϵ=104\epsilon=10^{-4} we have pc(ϵ)p_{c}\left(\epsilon\right)\approx 11.611.6, which means that p=12p=12 should be sufficient to achieve the required accuracy at Re=βR0R_{e}=\beta R_{0}. This also shows that at large enough NN and MM computational costs in Eq.(69) are much smaller than that for kernel computations using the integral representation, which formally is O(NM)O\left(NM\right) with the asymptotic constant that can be substantially large as the cost of numerical quadrature also depends on the required accuracy.

Refer to caption
Figure 2: The distribution of the sources (the small dots) and receivers (the bold dots) for the kernel accuracy and computational complexity tests.

To validate these findings we conducted some numerical experiments. First, we did error estimates of the factored solutions using comparisons with the error-controlled integral computations (Matlab integral2 function). For this purpose due to the symmetry of the problem, all receivers can be placed on the intersection of a unit hemisphere (“bump”) with plane y=0y=0. On the other hand, the sources can be distributed over the hemisphere surface and in the extension of the domain (see Fig. 2). Figure 3 shows the relative L2L_{2}-norm errors, ϵ2\epsilon_{2},

ϵ2=ffref2fref2,\epsilon_{2}=\frac{\left\|f-f_{ref}\right\|_{2}}{\left\|f_{ref}\right\|_{2}}, (73)

where ff is the computed and freff_{ref} is the reference solution for all computed source-receiver pairs as the filled contours in plane [Re/R0,p]\left[R_{e}/R_{0},p\right]. It is seen that the boundaries of the regions agree well with the error approximation Re=R0ϵ21/pR_{e}=R_{0}\epsilon_{2}^{-1/p}. Figure 4 is computed by finding for each Re/R0R_{e}/R_{0} the smallest pp at which the computed ϵ2\epsilon_{2} is smaller than the prescribed accuracy and plotting the wall-clock time required for kernel computation via factorization. It is seen that dependences computed at different prescribed ϵ2\epsilon_{2} have global minima in the range 2Re/R03.52\leqslant R_{e}/R_{0}\leqslant 3.5 at 102ϵ210810^{-2}\geqslant\epsilon_{2}\geqslant 10^{-8}. This agrees well with the theoretical estimates provided above. Note that non-monotonous character of the curves shown on this plot is due to discrete change of parameters Re/R0R_{e}/R_{0} and pp and thresholding (the curves show the prescribed error curves, not the actual errors, which can be close to or several times smaller than the prescribed errors). Also, this figure shows that optimization of the extended domain size can speed up computations several times.

Refer to caption
Figure 3: The computed domains (colored) in (Re/R0,p)\left(R_{e}/R_{0},p\right)-plane corresponding to different values of the relative L2L_{2}-norm error ϵ2\epsilon_{2} for the factored kernel representation (Eqs (66) and (67)) using truncation number pp. The reference solution is obtained using the integral representation and the Matlab integral2 procedure with the absolute and relative tolerance 10-12. The boundaries of the domains correlate well with dependences, p=logϵ2/log(R0/Re)p=\log\epsilon_{2}/\log\left(R_{0}/R_{e}\right) shown by the dashed curves. The values of ϵ2\epsilon_{2} are shown near the curves.
Refer to caption
Figure 4: The minimal wall-clock times (in seconds, Matlab implementation) required to compute the factored kernel representation (Eqs (66) and (67)) to achieve the presciribed error ϵ2\epsilon_{2} as a function of Re/R0R_{e}/R_{0}. The values of truncation numbers for optimum domain sizes are shown near the minima of the curves.

5.2 Optimization of the Boundary Element Solution

The extension of the domain can substantially impact the complexity and memory requirements of the BEM and it is preferable not to extend the domain too much. Assume that M=NM=N and Re=R0(1+δ)R_{e}=R_{0}\left(1+\delta\right), δ1\delta\ll 1. In this case, pδ1ln(1/ϵ)p\sim\delta^{-1}\ln\left(1/\epsilon\right), the first term in Eq. (69) can be neglected, and we obtain

CfactBNδ3ln31ϵ(1+πR02A02CBδ2ln1ϵ)BNδ3ln31ϵ.C_{fact}\sim B\frac{N}{\delta^{3}}\ln^{3}\frac{1}{\epsilon}\left(1+\frac{\pi R_{0}^{2}}{A_{0}}\frac{2C}{B}\frac{\delta^{2}}{\ln\frac{1}{\epsilon}}\right)\sim B\frac{N}{\delta^{3}}\ln^{3}\frac{1}{\epsilon}. (74)

Here we assumed that the density of sampling of the domain extension is the same as the density of sampling of surface SSg0S\cup S_{g0} of area A0A_{0}, means ρ=N/A0\rho=N/A_{0}. The last term in the parentheses then can be neglected at δ0\delta\rightarrow 0, due to πR02A0\pi R_{0}^{2}\leqslant A_{0}, ln1ϵ>1,\ln\frac{1}{\epsilon}>1, and assumed CB.C\sim B.

There are different versions of the BEM including those using iterative methods and the FMM, which have different computational costs, e.g., O(Nα)O\left(N^{\alpha}\right), α1\alpha\geqslant 1, for problems of size N.N. So, the cost of the BEM with the extended domain kernel computing at δ1\delta\ll 1 can be estimated as

CBEM\displaystyle C_{BEM} =\displaystyle= Cfact+DNeα=Cfact+DNα(1+NeNN)α\displaystyle C_{fact}+DN_{e}^{\alpha}=C_{fact}+DN^{\alpha}\left(1+\frac{N_{e}-N}{N}\right)^{\alpha}
\displaystyle\sim Cfact+DNα(1+απR02A0(Re2R021))DNα+BNδ3ln31ϵ+2DNααπR02A0δ.\displaystyle C_{fact}+DN^{\alpha}\left(1+\alpha\frac{\pi R_{0}^{2}}{A_{0}}\left(\frac{R_{e}^{2}}{R_{0}^{2}}-1\right)\right)\sim DN^{\alpha}+B\frac{N}{\delta^{3}}\ln^{3}\frac{1}{\epsilon}+2DN^{\alpha}\alpha\frac{\pi R_{0}^{2}}{A_{0}}\delta.

Here DD is some asymptotic constant related to the BEM cost. While it appears that the cost of the last term is much smaller than the cost of the first term, we retained it, as the first term is constant, while the last term reflects the main mechanism of increase of the cost at the increasing δ\delta. Indeed, taking derivative of this expression we can see that CBEMC_{BEM} has a global minimum,

dCBEMdδ\displaystyle\frac{dC_{BEM}}{d\delta} =\displaystyle= 3BNδ4ln31ϵ+2αDπR02A0Nα,dCBEMdδ(δopt)=0,\displaystyle-3B\frac{N}{\delta^{4}}\ln^{3}\frac{1}{\epsilon}+2\alpha D\frac{\pi R_{0}^{2}}{A_{0}}N^{\alpha},\quad\frac{dC_{BEM}}{d\delta}\left(\delta_{opt}\right)=0,\quad (76)
δopt\displaystyle\delta_{opt} =\displaystyle= 3B2αDA0πR02(1Nα1ln31ϵ)1/4=O(1N(α1)/4ln3/41ϵ).\displaystyle\frac{3B}{2\alpha D}\frac{A_{0}}{\pi R_{0}^{2}}\left(\frac{1}{N^{\alpha-1}}\ln^{3}\frac{1}{\epsilon}\right)^{1/4}=O\left(\frac{1}{N^{\left(\alpha-1\right)/4}}\ln^{3/4}\frac{1}{\epsilon}\right).

This shows that, indeed, at α>1\alpha>1 and NN\rightarrow\infty and fixed other parameters the optimal size of the extension, δopt\delta_{opt}, turns to zero. For the regular BEM with exact solvers we have α=3\alpha=3. However, when NN is not very high the cost of the FMM can be limited by the cost of boundary integral computing, in which case α=2\alpha=2. Also, this can be achieved with iterative solvers. In the case of the FMM acceleration α=1\alpha=1 or close to 1. In this case, δopt\delta_{opt} should not depend on NN, and the only reason why δopt1\delta_{opt}\ll 1 can be related to large values of asymptotic constant DD. Usually, this is true for the FMM.

6 BEM computing

Refer to caption
Figure 5: The meshes of “bump” and “dip” and source locations used in computations of Figs 7 and 8. The parts corresponding to Sg0S_{g0} are shaded in the light gray. The darker gray shade shows the parts corresponding to Sge\Sg0S_{ge}\backslash S_{g0}.

For illustrations of the kernel can be used with the boundary elements we consider the following benchmark problems for validation. We assumed that the ground elevation zz is the following function of xx and yy,

Sg:{(x,y,z),z={±1ρ2,ρ<1,0,ρ1.,ρ=x2+y2}.S_{g}:\left\{(x,y,z),\quad z=\left\{\begin{array}[]{c}\pm\sqrt{1-\rho^{2}},\quad\rho<1,\\ 0,\quad\rho\geqslant 1.\end{array}\right.,\quad\rho=\sqrt{x^{2}+y^{2}}\right\}. (77)

Here the sign “+” corresponds to a “bump”, while the sign “-” corresponds to a “dip”. Assume then that the electric field is generated by a monopole source (charge) of unit intensity located at 𝐱s=(0,0,h)\mathbf{x}_{s}=\left(0,0,h\right), and the domain Ω0\Omega_{0} is bounded by a sphere of radius R0=hR_{0}=h (h>1h>1) centered at the origin for the “bump” and , R0=1R_{0}=1 for the “dip” (|h|<1\left|h\right|<1). Figure 5 shows the meshes used in computations.

For validation purposes, the solution of the problem for the bump can be found analytically using the method of images. In this case, the image of the bump is a hemisphere, which together with the “image” bump forms a sphere of radius 1 on which the potential is zero. The image source of negative intensity (sink) is located at 𝐱s=(0,0,h)\mathbf{x}_{s}^{\ast}=\left(0,0,-h\right), and we have the following expression for the potential

ϕ(𝐲)=G(𝐲,𝐱s)G(𝐲,𝐱s)1hG(𝐲,𝐱r+)+1hG(𝐲,𝐱r),𝐱r±=(0,0,±1h),\phi\left(\mathbf{y}\right)=G\left(\mathbf{y},\mathbf{x}_{s}\right)-G\left(\mathbf{y},\mathbf{x}_{s}^{\ast}\right)-\frac{1}{h}G\left(\mathbf{y},\mathbf{x}_{r}^{+}\right)+\frac{1}{h}G\left(\mathbf{y},\mathbf{x}_{r}^{-}\right),\quad\mathbf{x}_{r}^{\pm}=\left(0,0,\pm\frac{1}{h}\right), (78)

where 𝐱r+\mathbf{x}_{r}^{+} and 𝐱r\mathbf{x}_{r}^{-} are the images of 𝐱s\mathbf{x}_{s} and 𝐱s\mathbf{x}_{s}^{\ast} with respect to the sphere. We are not aware about analytical solutions for the case of “dip” for which the method of images cannot be applied.

The boundary integral solution of the problem can be represented in the form

ϕ(𝐲)=G(𝐲,𝐱s)+K(D)(𝐲,𝐱s;Re)+Sgeσ(D)(𝐱)[G(𝐲𝐱)+K(D)(𝐲,𝐱;Re)]𝑑S(𝐱),\phi\left(\mathbf{y}\right)=G\left(\mathbf{y},\mathbf{x}_{s}\right)+K^{(D)}\left(\mathbf{y},\mathbf{x}_{s}\mathbf{;}R_{e}\right)+\int_{S_{ge}}\sigma^{(D)}\left(\mathbf{x}\right)\left[G\left(\mathbf{y}-\mathbf{x}\right)+K^{(D)}\left(\mathbf{y},\mathbf{x;}R_{e}\right)\right]dS\left(\mathbf{x}\right), (79)

where the charge density can be found from

Sgeσ(D)(𝐱)[G(𝐲,𝐱)+K(D)(𝐲,𝐱;Re)]𝑑S(𝐱)=G(𝐲,𝐱s)K(D)(𝐲,𝐱s;Re),𝐲Sge.\int_{S_{ge}}\sigma^{(D)}\left(\mathbf{x}\right)\left[G\left(\mathbf{y},\mathbf{x}\right)+K^{(D)}\left(\mathbf{y},\mathbf{x;}R_{e}\right)\right]dS\left(\mathbf{x}\right)=-G\left(\mathbf{y},\mathbf{x}_{s}\right)-K^{(D)}\left(\mathbf{y},\mathbf{x}_{s}\mathbf{;}R_{e}\right),\quad\mathbf{y}\in S_{ge}. (80)

In the case of “bump” we computed and compared four solutions: 1) analytical; 2) regular BEM using method of images (two symmetric sources and a sphere) (“BEMimage”); 3) regular BEM, where the domain ground plane was simply truncated and the effect of the infinite plane ignored (“BEM”); 4) BEM with the kernel accounting for the infinite plane as described by Eqs (79) and (80) (“BEMinf”). In case of “dip” only solutions 3) and 4) were computed. In all cases we varied ReR_{e}. The truncation number pp for “BEMinf” method was selected based on the theoretical estimate, p=logϵ/log(R0/Re)p=\log\epsilon/\log(R_{0}/R_{e}), where the prescribed error ϵ\epsilon varied in range 10310610^{-3}-10^{-6}.

Refer to caption
Figure 6: The wall-clock times (in seconds) for solution of the “bump” problem shown in Fig. 5 for different Re/R0R_{e}/R_{0}, different prescribed accuracies, and different versions of the BEMinf (with and without the FMM acceleration). The values of pp corresponding to the minima of the dependences are shown near the curves. In the regular BEMinf, the number of faces for surface Sg0S_{g0} is N0=3594N_{0}=3594 and varied in the range Ne=39018106N_{e}=3901-8106 for surface Sge.S_{ge}. In the BEMinf accelerated with the FMM the numbers are N0=13244N_{0}=13244 and Ne=1656830717N_{e}=16568-30717, respectively.
Refer to caption
Figure 7: The induced electric potential, ϕind(𝐲)=ϕ(𝐲)G(𝐲,𝐱s)\phi_{ind}\left(\mathbf{y}\right)=\phi\left(\mathbf{y}\right)-G\left(\mathbf{y},\mathbf{x}_{s}\right), according to the analytical solution (78) , and the absolute errors in domain Ω0\Omega_{0} obtained using different methods of solution for the “bump” case, h=2h=2, δ=0.0935\delta=0.0935, N0=6401N_{0}=6401, and Ne=7661N_{e}=7661. The mesh is shown in Fig. 5 (on the left).

Figure 6 illustrates the effect of the domain extension on the performance for different prescribed accuracies and the BEM complexities for the “bump” case (h=2h=2). Note that here the FMM accelerated BEM was implemented in the way that the boundary integrals were computed within some cutoff radius of the evaluation point using Matlab, while the FMM procedure was executed in the form of a high performance library, so the cost of the FMM BEM is mainly determined by the Matlab integral computations, which is consistent with the kernel evaluation also implemented in Matlab. Also, in the regular BEM the solution time was much smaller than the integral computation time. So, the methods shown can be characterized as O(N2)O\left(N^{2}\right) for the BEM and O(N)O\left(N\right) for the BEM/FMM. It is seen that the curves form minima (sometimes a few local minima), at relatively low δ=Re/R01\delta=R_{e}/R_{0}-1. According to the theory (see Eq. (76)), increasing the prescribed accuracy results in an increase of the optimal ReR_{e}, which is seen on the figure. For δ\delta substantially larger than δopt\delta_{opt} the cost of kernel computation becomes negligible in comparison to the BEM costs, and the curves computed for different ϵ\epsilon become close to each other.

Figure 7 shows the analytical solution (78) (the induced part, ϕind(𝐲)=ϕ(𝐲)G(𝐲,𝐱s)\phi_{ind}\left(\mathbf{y}\right)=\phi\left(\mathbf{y}\right)-G\left(\mathbf{y},\mathbf{x}_{s}\right)), and the absolute errors in the computational domain obtained using different methods of solution for the “bump” case, h=2h=2, δ=0.0935\delta=0.0935, N0=6401N_{0}=6401, and Ne=7661N_{e}=7661, where N0N_{0} and NeN_{e} are the number of boundary elements in Ω0\Omega_{0} and Ωe\Omega_{e}, respectively. The numerical solution using image sources is the most accurate and its relative L2L_{2}-norm error is ϵ2=2.3103\epsilon_{2}=2.3\cdot 10^{-3}. The BEMinf shows error ϵ2=4.5103\epsilon_{2}=4.5\cdot 10^{-3} (at the prescribed accuracy ϵ=104\epsilon=10^{-4}) while the BEM shows ϵ2=3.7102\epsilon_{2}=3.7\cdot 10^{-2}, which is one order of magnitude larger. This figure shows also different error distributions for different methods inside the computational domain.

Refer to caption
Figure 8: The induced electric potential, ϕind(𝐲)=ϕ(𝐲)G(𝐲,𝐱s)\phi_{ind}\left(\mathbf{y}\right)=\phi\left(\mathbf{y}\right)-G\left(\mathbf{y},\mathbf{x}_{s}\right), for the reference solution computed using BEMinf with parameters Re/R0=1.5R_{e}/R_{0}=1.5, N0=6401N_{0}=6401, Ne=14038,N_{e}=14038, and prescribed ϵ=106\epsilon=10^{-6}, and the absolute errors in domain Ω0\Omega_{0} obtained using different methods of solution for the “dip” case, h=0.5h=0.5, δ=0.124\delta=0.124, N0=1592N_{0}=1592, and Ne=2017N_{e}=2017. The mesh is shown in Fig. 5 (on the right).
Refer to caption
Figure 9: The relative L2L_{2}-norm error ϵ2\epsilon_{2} for the “dip” case as a function of parameter Re/R0R_{e}/R_{0} for BEMinf and regular BEM. The solid line shows the dependence ϵ2=C(Re/R0)3\epsilon_{2}=C\left(R_{e}/R_{0}\right)^{-3} with the constant CC fiitting the error of regular BEM at Re/R0=2R_{e}/R_{0}=2.

Figure 8 shows the reference solution (ϕind(𝐲))\left(\phi_{ind}\left(\mathbf{y}\right)\right) for the “dip” case and the absolute errors in the computational domain obtained using BEMinf and the regular BEM, h=0.5h=0.5, δ=0.124\delta=0.124, N0=1592N_{0}=1592, and Ne=2017N_{e}=2017. The reference solution was computed using BEMinf with parameters δ=0.5\delta=0.5, N0=6401N_{0}=6401, and Ne=14038N_{e}=14038 and prescribed ϵ=106\epsilon=10^{-6}. The BEMinf shows error ϵ2=4.7104\epsilon_{2}=4.7\cdot 10^{-4} (at the prescribed accuracy ϵ=104\epsilon=10^{-4}) while the BEM shows ϵ2=5.6102\epsilon_{2}=5.6\cdot 10^{-2}, which is two orders of magnitude larger. Similarly to the case of “bump” one can observe different error distributions for the BEMinf and the BEM. Similarly to the case of “bump” the largest BEMinf errors are observed near the edge of domain Ω0\Omega_{0} coinciding with the ground plane.

Figure 9 plots ϵ2\epsilon_{2} for the “dip” case as a function of parameter Re/R0R_{e}/R_{0}. It is seen that the error is practically constant for BEMinf, while it decays inversely proportionally to the volume of the computational domain for the regular BEM. It is clear then that the results of BEMinf can be obtained by simple extension of the computational domain, but for substantially higher computational costs. For example, to achieve two orders of magnitude in error reduction one should have Re5R0R_{e}\sim 5R_{0}, which may require one order of magnitude increase in the number of boundary elements and increase the computational costs 10-1000 times, based on the solution method used (scaled as O(Nα)O\left(N^{\alpha}\right), 1α31\leqslant\alpha\leqslant 3).

7 Conclusion

We introduced analogues of Green’s function for an infinite plane with a circular hole. We also proposed and tested efficient methods for computing of these functions, which may have broad applications to model infinite ground or ocean surfaces in the problems requiring Laplace equation solvers. While such functions can be expressed in the forms of finite integrals, their computation can be costly. However, a computationally efficient factorization of these functions can be done, which drastically reduces the computational costs, and, in fact, creates rather small overheads to the problems, which simply ignore the existence of ground outside the computational domain. Also, we conducted some optimization study and provided actual boundary element computations for some benchmark problems. The study shows that the accuracy of the solution using the analogues of Green’s function improves substantially compared to a simple domain truncation, which also can become computationally costly for improved accuracy computations.

8 Acknowledgements

This work is supported by Cooperative Research Agreement (W911NF1420118) between the University of Maryland and the Army Research Laboratory, with David Hull, Ross Adelman and Steven Vinci as Technical monitors.

9 References

References

  • [1] R. Adelman, Extremely large, wide-area power-line models, Supercomputing 2017, https://sc17.supercomputing.org/SC17%20Archive/tech_poster/poster_files/post251s2-file3.pdf.
  • [2] M. D’Amore, M. S. Sarto, Simulation models of a dissipative transmission line above a lossy ground for a wide-frequency range. I. Single conductor configuration, IEEE Trans. Electromag. Compat. 38(2) (1996) 127-138. https://doi.org/10.1109/15.494615.
  • [3] M. Trlep, A. Hamler, M. Jesenik, B. Stumberger, Electric field distribution under transmission lines dependent on ground surface, IEEE Trans. Mag. 45(3) (2009) 1748-1751. https://doi.org/10.1109/TMAG.2009.2012806.
  • [4] Q. Zhang, J. Yang, D. Li, Z. Wang, Propagation effects of a fractal rough ocean surface on the vertical electric field generated by lightning return strokes, J. Electrostat. 70(1)(2012) 54-59. https://doi.org/10.1016/j.elstat.2011.10.003.
  • [5] K. S. Chen, T.-D. Wu, L. Tsang, Q. Li, J. Shi, A.K. Fung, Emission of rough surfaces calculated by the integral equation method with comparison to three-dimensional moment method simulations, IEEE Trans. Geo. Rem. Sensing 41(1) (2003) 90-101. https://doi.org/10.1109/TGRS.2002.807587.
  • [6] D. Givoli, Numerical Methods for Problems in Infinite Domains, Elsevier, Amsterdam, 1992.
  • [7] S.V. Tsynkov, Numerical solution of problems on unbounded domains. A review, Appl. Num. Math. 27(4) (1998) 465-532. https://doi.org/10.1016/S0168-9274(98)00025-7.
  • [8] J.D. Jackson, Classical Electrodynamics, third ed., John Wiley and Sons, New York, 1998.
  • [9] D. Crowdy, J. Marshall, Green’s functions for Laplace’s equation in multiply connected domains, IMA J. Appl. Math. 72 (2007) 278-301. https://doi.org/10.1093/imamat/hxm007.
  • [10] N.A. Gumerov and R. Duraiswami, A method to compute periodic sums, J. Comput. Phys. 272 (2014) 307-326. https://doi.org/10.1016/j.jcp.2014.04.039.
  • [11] A. Moroz, Quasi-periodic Green’s functions of the Helmholtz and Laplace equations, J. Phys. A: Math. Gen. 39(36) (2006) 11247. https://doi.org/10.1088%2F0305-4470%2F39%2F36%2F009.
  • [12] G. Bao, J. Lin, Near-field imaging of the surface displacement on an infinite ground plane, Inverse Problems & Imaging, 7(2) (2013), 377-396. https:/doi.org/10.3934/ipi.2013.7.377.
  • [13] S. N. Chandler-Wilde, E. Heinemeyer, R. Potthast, Acoustic scattering by mildly rough unbounded surfaces in three dimensions, SIAM J. Appl. Math. 66 (2006) 1002–1026. https://doi.org/10.1137/050635262.
  • [14] J.J. Holmes, Past, present, and future of underwater sensor arrays to measure the electromagnetic field signatures of naval vessels, Marine Tech. Soc. J. 49(6) (2015) 123-133. https://doi.org/10.4031/MTSJ.49.6.1.
  • [15] R. Yue, P. Hu, J. Zhang, The influence of the seawater and seabed interface on the underwater low frequency electromagnetic field signatures, 2016 IEEE/OES China Ocean Acoustics (COA), Harbin (2016) 1-7. https://doi.org/10.1109/COA.2016.7535694.
  • [16] X. Wang, Q. Xu, J. Zhang, Simulating underwater electric field signal of ship using the boundary element method, Progress In Electromag. Res. M 76 (2018) 43-54. https:/doi.org/10.2528/PIERM18092706.
  • [17] K.V. Rozhdestvensky, Aerodynamics of a Lifting System in Extreme Ground Effect, Springer, Berlin-Heidelberg, 2000.
  • [18] D. G. Duffy, Green’s Functions with Applications, CRC Press, Taylor & Francis Group, FL, 2015.
  • [19] P.M. Morse, H. Feshbach, Methods of Theoretical Physics, McGraw-Hill, NY, 1953.
  • [20] N.A. Gumerov, R. Duraiswamin, Comparison of the efficiency of translation operators used in the fast multipole method for the 3D Laplace equation, UMIACS TR 2005-09, Also issued as Computer Science Technical Report CS-TR-# 4701, University of Maryland, College Park, 2005. https://drum.lib.umd.edu/bitstream/handle/1903/3023/LaplaceTranslation_CSTR_4701.pdf.
  • [21] M. Abramowitz, I.A. Stegun, Handbook of Mathematical Functions, National Bureau of Standards, Washington D.C., 1972.
  • [22] R. Adelman, N.A Gumerov, and R. Duraiswami, FMM/GPU-accelerated boundary element method for computational magnetics and electrostatics, IEEE Trans. Mag. 53(12) (2017), 7002311. https://doi.org/10.1109/TMAG.2017.2725951.
  • [23] N.A. Gumerov and R. Duraiswami, Fast multipole methods on graphics processors, J. Comput. Phys. 227 (2008), 8290-8313. https://doi.org/10.1016/j.jcp.2008.05.023.

Appendix A Derivation of Recurrence Relations

Recurrence (52) can be derived by comparing two representations of integral

Inm\displaystyle I_{n}^{m} =\displaystyle= 1ξn+102πeimφ𝑑φ0ξ(12ζcosφ+ζ2)1/2ζn𝑑ζ\displaystyle\frac{1}{\xi^{n+1}}\int_{0}^{2\pi}e^{im\varphi}d\varphi\int_{0}^{\xi}\left(1-2\zeta\cos\varphi+\zeta^{2}\right)^{1/2}\zeta^{n}d\zeta
=\displaystyle= unm+ξ2un+2mξ(un+1m1+un+1m+1),\displaystyle u_{n}^{m}+\xi^{2}u_{n+2}^{m}-\xi\left(u_{n+1}^{m-1}+u_{n+1}^{m+1}\right),

which can be checked using definition of unm(ξ)u_{n}^{m}\left(\xi\right) (49), and integration by parts:

Inm\displaystyle I_{n}^{m} =\displaystyle= 1(n+1)ξn+102πeimφ𝑑φ0ξ(12ζcosφ+ζ2)1/2𝑑ζn+1\displaystyle\frac{1}{\left(n+1\right)\xi^{n+1}}\int_{0}^{2\pi}e^{im\varphi}d\varphi\int_{0}^{\xi}\left(1-2\zeta\cos\varphi+\zeta^{2}\right)^{1/2}d\zeta^{n+1}
=\displaystyle= 1n+1[vmξ2un+2m+12ξ(un+1m1+un+1m+1)],vm=02πeimφ(12ξcosφ+ξ2)1/2𝑑φ.\displaystyle\frac{1}{n+1}\left[v_{m}-\xi^{2}u_{n+2}^{m}+\frac{1}{2}\xi\left(u_{n+1}^{m-1}+u_{n+1}^{m+1}\right)\right],\quad v_{m}=\int_{0}^{2\pi}e^{im\varphi}\left(1-2\xi\cos\varphi+\xi^{2}\right)^{1/2}d\varphi.

Equation (53) can be checked directly using integral representation of vmv_{m} and definition of wm(ξ).w_{m}\left(\xi\right).

To obtain recurrence (54) we note the following transformation

i02πei(m1)φsinφdφ(12ξcosφ+ξ2)1/2=iξ02πei(m1)φd(12ξcosφ+ξ2)1/2\displaystyle i\int_{0}^{2\pi}\frac{e^{i\left(m-1\right)\varphi}\sin\varphi d\varphi}{\left(1-2\xi\cos\varphi+\xi^{2}\right)^{1/2}}=\frac{i}{\xi}\int_{0}^{2\pi}e^{i\left(m-1\right)\varphi}d\left(1-2\xi\cos\varphi+\xi^{2}\right)^{1/2} (83)
=iξ02π(12ξcosφ+ξ2)1/2𝑑ei(m1)φ=(m1)1ξ02πei(m1)φ(12ξcosφ+ξ2)(12ξcosφ+ξ2)1/2𝑑φ.\displaystyle=-\frac{i}{\xi}\int_{0}^{2\pi}\left(1-2\xi\cos\varphi+\xi^{2}\right)^{1/2}de^{i\left(m-1\right)\varphi}=\left(m-1\right)\frac{1}{\xi}\int_{0}^{2\pi}\frac{e^{i\left(m-1\right)\varphi}\left(1-2\xi\cos\varphi+\xi^{2}\right)}{\left(1-2\xi\cos\varphi+\xi^{2}\right)^{1/2}}d\varphi.

Using this expression, we obtain

wm\displaystyle w_{m} =\displaystyle= 02πei(m1)φeiφdφ(12ξcosφ+ξ2)1/2=02πei(m1)φ(cosφ+isinφ)dφ(12ξcosφ+ξ2)1/2\displaystyle\int_{0}^{2\pi}\frac{e^{i\left(m-1\right)\varphi}e^{i\varphi}d\varphi}{\left(1-2\xi\cos\varphi+\xi^{2}\right)^{1/2}}=\int_{0}^{2\pi}\frac{e^{i\left(m-1\right)\varphi}\left(\cos\varphi+i\sin\varphi\right)d\varphi}{\left(1-2\xi\cos\varphi+\xi^{2}\right)^{1/2}}
=\displaystyle= (m1)1+ξ2ξ02πei(m1)φdφ(12ξcosφ+ξ2)1/22m3202π(eiφ+eiφ)ei(m1)φ(12ξcosφ+ξ2)1/2𝑑φ\displaystyle\left(m-1\right)\frac{1+\xi^{2}}{\xi}\int_{0}^{2\pi}\frac{e^{i\left(m-1\right)\varphi}d\varphi}{\left(1-2\xi\cos\varphi+\xi^{2}\right)^{1/2}}-\frac{2m-3}{2}\int_{0}^{2\pi}\frac{\left(e^{i\varphi}+e^{-i\varphi}\right)e^{i\left(m-1\right)\varphi}}{\left(1-2\xi\cos\varphi+\xi^{2}\right)^{1/2}}d\varphi
=\displaystyle= (m1)1+ξ2ξwm12m32(wm+wm2).\displaystyle\left(m-1\right)\frac{1+\xi^{2}}{\xi}w_{m-1}-\frac{2m-3}{2}\left(w_{m}+w_{m-2}\right).

It is not difficult to see that Eq. (54) follows from this expression.

Functions w0w_{0} and w1w_{1} can be expressed immediately via complete elliptic integrals. Indeed, we have

wm(ξ)\displaystyle w_{m}\left(\xi\right) =\displaystyle= π3πeimφdφ(12ξcosφ+ξ2)1/2=(1)m02πeimφdφ(1+2ξcosφ+ξ2)1/2\displaystyle\int_{\pi}^{3\pi}\frac{e^{im\varphi}d\varphi}{\left(1-2\xi\cos\varphi+\xi^{2}\right)^{1/2}}=\left(-1\right)^{m}\int_{0}^{2\pi}\frac{e^{im\varphi}d\varphi}{\left(1+2\xi\cos\varphi+\xi^{2}\right)^{1/2}}
=\displaystyle= 2(1)m0πcosmφdφ(1+2ξcosφ+ξ2)1/2=4(1)m1+ξ0π/2cos(2mψ)dψ(1μsin2ψ)1/2,μ=4ξ(1+ξ)2.\displaystyle 2\left(-1\right)^{m}\int_{0}^{\pi}\frac{\cos m\varphi d\varphi}{\left(1+2\xi\cos\varphi+\xi^{2}\right)^{1/2}}=\frac{4\left(-1\right)^{m}}{1+\xi}\int_{0}^{\pi/2}\frac{\cos\left(2m\psi\right)d\psi}{\left(1-\mu\sin^{2}\psi\right)^{1/2}},\quad\mu=\frac{4\xi}{\left(1+\xi\right)^{2}}.

So,

w0(ξ)=41+ξK(μ),w1(ξ)=41+ξ[(2μ1)K(μ)2μE(μ)],w_{0}\left(\xi\right)=\frac{4}{1+\xi}K\left(\mu\right),\quad w_{1}\left(\xi\right)=\frac{4}{1+\xi}\left[\left(\frac{2}{\mu}-1\right)K\left(\mu\right)-\frac{2}{\mu}E\left(\mu\right)\right], (86)

where KK and EE are defined by Eq. (56). The Landen transformation of elliptic functions [21] shows that

K(μ)\displaystyle K\left(\mu\right) =\displaystyle= 21+μ11/2K(μ2),μ1=1μ,μ2=(1μ11/21+μ11/2)2\displaystyle\frac{2}{1+\mu_{1}^{1/2}}K\left(\mu_{2}\right),\quad\mu_{1}=1-\mu,\quad\mu_{2}=\left(\frac{1-\mu_{1}^{1/2}}{1+\mu_{1}^{1/2}}\right)^{2} (87)
E(μ)\displaystyle\quad E\left(\mu\right) =\displaystyle= (1+μ11/2)E(μ2)2μ11/21+μ11/2K(μ2).\displaystyle\left(1+\mu_{1}^{1/2}\right)E\left(\mu_{2}\right)-\frac{2\mu_{1}^{1/2}}{1+\mu_{1}^{1/2}}K\left(\mu_{2}\right).

Substituting this into Eq. (86) and using the value of μ\mu from Eq. (A) we obtain Eq. (55).

Note now that relation (52) is also valid in case m=0m=0. In this case we can use symmetry un1=un1u_{n}^{-1}=u_{n}^{1}. Since Eq. (53) also holds in this case, we obtain using symmetry w1=w1w_{-1}=w_{1} and Eq. (55),

v0=(1+ξ2)w02ξw1=8E(ξ2)4(1ξ2)K(ξ2),v_{0}=\left(1+\xi^{2}\right)w_{0}-2\xi w_{1}=8E\left(\xi^{2}\right)-4\left(1-\xi^{2}\right)K\left(\xi^{2}\right), (88)

and confirms recurrence (51).

So, the last relation to prove is Eq. (50). This relation can be obtained using differentiation relations for the complete elliptic integral,

K(μ)=ddμ[μ(1μ)dK(μ)dμ],dK(μ)dμ=E(μ)2μ(1μ)K(μ)2μ,K\left(\mu\right)=\frac{d}{d\mu}\left[\mu\left(1-\mu\right)\frac{dK\left(\mu\right)}{d\mu}\right],\quad\frac{dK\left(\mu\right)}{d\mu}=\frac{E\left(\mu\right)}{2\mu\left(1-\mu\right)}-\frac{K\left(\mu\right)}{2\mu}, (89)

and integration by parts,

un0\displaystyle u_{n}^{0} =\displaystyle= 4ξn+10ξζnK(ζ2)𝑑ζ=8ξn+10ξ2μ(n1)/2ddμ(μ(1μ)dK(μ)dμ)𝑑μ\displaystyle\frac{4}{\xi^{n+1}}\int_{0}^{\xi}\zeta^{n}K\left(\zeta^{2}\right)d\zeta=\frac{8}{\xi^{n+1}}\int_{0}^{\xi^{2}}\mu^{\left(n-1\right)/2}\frac{d}{d\mu}\left(\mu\left(1-\mu\right)\frac{dK\left(\mu\right)}{d\mu}\right)d\mu
=\displaystyle= 8ξn+1[ξn+1(1ξ2)dK(ξ2)d(ξ2)n120ξ2μ(n1)/2(1μ)𝑑K(μ)]\displaystyle\frac{8}{\xi^{n+1}}\left[\xi^{n+1}\left(1-\xi^{2}\right)\frac{dK\left(\xi^{2}\right)}{d\left(\xi^{2}\right)}-\frac{n-1}{2}\int_{0}^{\xi^{2}}\mu^{\left(n-1\right)/2}\left(1-\mu\right)dK\left(\mu\right)\right]
=\displaystyle= 4ξ2[E(ξ2)n(1ξ2)K(ξ2)+(n1)24un20](n21)un0.\displaystyle\frac{4}{\xi^{2}}\left[E\left(\xi^{2}\right)-n\left(1-\xi^{2}\right)K\left(\xi^{2}\right)+\frac{\left(n-1\right)^{2}}{4}u_{n-2}^{0}\right]-\left(n^{2}-1\right)u_{n}^{0}.

Equation (50) is just a rearrangement of terms in this relation.