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

A non-singular, field-only surface integral method for interactions between electric and magnetic dipoles and nano-structures

Abstract

With the development of condensed-matter physics and nanotechnology, attention has turned to the fields near and on surfaces that result from interactions between electric dipole radiation and mesoscale structures. It is hoped that studying these fields will further our understanding of optical phenomena in nano-optics, quantum mechanics, electromagnetics and sensing using solid-state photon emitters. Here, we describe a method for implementing dynamic electric and magnetic dipoles in the frequency domain into a non-singular field-only surface integral method. We show that the effect of dipoles can conveniently be described as a relatively simple term in the integral equations, which fully represents how they drive the fields and interactions. Also, due to the non-singularity, our method can calculate the electric and magnetic fields on the surfaces of objects in both near and far fields with the same accuracy, which makes it an ideal tool to investigate nano-optical phenomena. The derivation of the framework is given and tested against a Mie theory alike formula. Some interesting examples are shown involving the interaction of dipoles with different types of mesoscale structures including parabolic nano-antennas and gold probes.

keywords:
Near-surface fields, Circularly polarized dipole

Qiang Sun* Evert Klaseboer

\dedication
{affiliations}

Dr. Qiang Sun
Australian Research Council Centre of Excellence for Nanoscale BioPhotonics, School of Science, RMIT University, Melbourne, VIC 3001, Australia
Email Address: [email protected]

Dr. Evert Klaseboer
Institute of High Performance Computing, 1 Fusionopolis Way, Singapore, 138632, Singapore

1 Introduction

The interaction of the radiation originating from dynamic electric dipoles with matter has a long history which can be traced back to the celebrated work of Lord Rayleigh [1] (see also the classic book of van de Hulst [2]) and can lead to interesting phenomena. Some communication technologies were developed based on such phenomena [3], mainly using the far field pattern of dipole radiation since there was essentially little need to investigate near field phenomena. Recently however, with the development of condensed-matter physics and nanotechnology, the near fields / surface fields of interactions between dynamic electromagnetic dipole radiation and mesoscale structures have become the center of interest [4]. A better understanding from a fundamental physics point of view can help to design better applications, such as in biophysics, novel optical phenomena in nano-optics [5, 6, 7, 8, 9, 10], quantum mechanics electromagnetics [11], imaging [12] and sensing [13] using solid-state photon emitters [14] (for instance the defects in diamonds). As such, it is desirable to have at our disposal a robust method to effectively and efficiently calculate the electromagnetic fields driven by the dipole radiation and its interaction with the structure’s surface, both in the near and in the far field.

Due to the limited amount of available analytical solutions, even for simple geometric shapes, the interactions between dynamic electric dipoles and matter need to be calculated using numerical methods. From a computational point of view, numerical methods can be divided into roughly two classes; volume based methods and surface based methods. Almost all the differential equation-based methods, such as the finite element method (FEM) [15], the finite-difference time-domain method (FDTD) [16] and the discrete-dipole approximation (DDA) [17] belong to the volume based method class. Some integral equation-based methods such as the volumetric Method of Moments algorithm (V-MoM) [18] and the (discontinuous) Galerkin time-domain [19] can also be regarded as belonging to the volume based method class. In order to use a volume based method to study the interaction between dipoles and objects, the complete simulation domain which includes both the scatterers and the surrounding medium needs to be discretized, which thus introduces significant computational overhead. Also, special attention is needed to handle the mesh in the domain around a dipole and how the mesh evolves to the surrounding space [20]. Surface based methods, such as the method of moments (MoM) [21] and the ϕ𝑨\phi-\boldsymbol{A} potential boundary element method [22], usually only require the discretization of the scattered surfaces. The most popular surface based method is MoM which was traditionally used to calculate the electromagnetic field of an electric dipole antenna for applications in Through-the-Earth communication. MoM can calculate the far field, but in spite of its popularity, performs much less accurate when accessing the near field or the surface field due to the congenital hypersingularies of this method. The ϕ𝑨\phi-\boldsymbol{A} potentials boundary element method solves for the scalar and vector potentials ϕ\phi and 𝑨\boldsymbol{A}. In order to get back the electric field, one must take a further step to calculate the electric field by differentiation, which might reduce the accuracy of the calculation. So we are left with the undesirable situation that both the popular volume-based and surface-based methods are not ideally suited to study the interactions between the radiation of dynamic electric dipoles and objects in mesoscale for nano-optics and an alternative method which can deliver more accurate results would be welcomed.

Recently, a new type of surface integral method, the fully non-singular field-only surface method has been developed by us to study the interactions between matter and an incoming electromagnetic wave [23, 24, 25, 26, 27]. In this paper, we develop a method to include electric dipoles into this framework. Before we step into the details, we briefly summarise the advantages of our method:

  • The electric dipole is implemented explicitly on the right hand side of the numerical system, which is derived analytically. From a physics point of view, such a system fully represents how dynamic dipoles drive the fields and interactions.

  • Our method only needs a mesh on the boundary of the scatterer. For volume methods, particular attention on meshing near the dipole is needed. It is even more difficult to handle multiple dipoles, while in our implementation this is straightforward.

  • The surface integration method automatically takes care of radiation conditions at infinity.

  • Compared to other surface integral methods, our method has two obvious advantages. Firstly, our method solves the fields of physical interest directly, namely the electric and magnetic fields. However, other surface integral methods solve for some intermediate quantities, such as surface currents or potentials [28, 22], while the electric and magnetic fields are obtained by post-processing in which the accuracy might be compromised. Secondly, there are no singularities in the surface integrals over the scatterers due to the Green’s function and its derivative. Consequently, high order surface elements, such as quadratic triangular elements, can be easily employed to obtain highly accurate results with fewer degrees of unknowns.

  • As our method is fully free of singularities, we can get the electric and magnetic fields on the surfaces of objects, in both near and far fields with the same accuracy, which makes our method an ideal tool to investigate nano-optical phenomena.

The paper is outlined as follows. In the next section, the description of how to include the electric dipole in the numerical framework is given, followed by the validation using the Mie theory. Then some examples for nano-optics are shown, and after a discussion section on magnetic dipoles, the paper is terminated with the conclusions.

2 Non-singular field-only surface integral method

To clearly illustrate our non-singular field-only surface method for the interactions between dynamic electromagnetic dipoles and mesoscale objects, we start with a simple case of an electric dipole near a perfect electric conductor. Then, we extend the framework for the interaction between an electric dipole and a dielectric object. The same idea can be further extended to multiple electric and magnetic dipoles and multiple scatterers, as discussed in Sec. 4.

2.1 Electric dipole near a perfect electric conductor

In the frequency domain, the electric and magnetic field due to an electric dipole are, respectively,

𝑬Ed\displaystyle\boldsymbol{E}^{\mathrm{Ed}} =14πϵ0ϵr×{[exp(ik|𝒙𝒙d|)|𝒙𝒙d|]×𝒑},\displaystyle=\frac{1}{4\pi\epsilon_{0}\epsilon_{r}}\nabla\times\left\{\nabla\left[\frac{\exp(\mathrm{i}k|\boldsymbol{x}-\boldsymbol{x}_{d}|)}{|\boldsymbol{x}-\boldsymbol{x}_{d}|}\right]\times\boldsymbol{p}\right\}, (1a)
𝑯Ed\displaystyle\boldsymbol{H}^{\mathrm{Ed}} =ω4πi{[exp(ik|𝒙𝒙d|)|𝒙𝒙d|]×𝒑}\displaystyle=\frac{\omega}{4\pi\mathrm{i}}\left\{\nabla\left[\frac{\exp(\mathrm{i}k|\boldsymbol{x}-\boldsymbol{x}_{d}|)}{|\boldsymbol{x}-\boldsymbol{x}_{d}|}\right]\times\boldsymbol{p}\right\} (1b)

where 𝒑\boldsymbol{p} is the electric dipole momentum vector (with dimension Coulomb meter), 𝒙\boldsymbol{x} is the observation location and 𝒙d\boldsymbol{x}_{d} the location of the dipole (presumed to be fixed in space, to avoid Doppler effects), k=ω/ck=\omega/c is the wavenumber with ω\omega the angular frequency and cc the speed of light, ϵ0\epsilon_{0} and ϵr\epsilon_{r} are the permittivity of free space and relative permittivity of the medium respectively.

In the domain that excludes the dipole, thus the domain outside the scatterer SS and a tiny spherical surface SdS_{d} that encloses the dipole as shown in Fig. 1, the Maxwell equations in the frequency domain reduce to the following two equations for the electric field 𝑬\boldsymbol{E}

2𝑬+k2𝑬=𝟎,\displaystyle\nabla^{2}\boldsymbol{E}+k^{2}\boldsymbol{E}=\boldsymbol{0}, (2a)
𝑬=0.\displaystyle\nabla\cdot\boldsymbol{E}=0. (2b)

Thus the electric field satisfies a vector Helmholtz equation and is simultaneously divergence free. To obtain the interactions between one or more dipoles and scatterers in a homogeneous medium, it is necessary to efficiently and accurately solve Eq. (2).

Refer to caption
Figure 1: Object with surface SS embedded in an infinite outer domain with parameters koutk_{\mathrm{out}}, ϵout\epsilon_{\mathrm{out}}, μout\mu_{\mathrm{out}} (the wavenumber, permittivity and permeability respectively). An electric dipole is situated at the position 𝒙d\boldsymbol{x}_{d}. The normal vector 𝒏\boldsymbol{n} is pointing into the object. For a dielectric object we also take into account the inner parameters kink_{\mathrm{in}}, ϵin\epsilon_{\mathrm{in}}, μin\mu_{\mathrm{in}}. In the theory a vanishing small volume around the dipole with surface SdS_{d} and radius rdr_{d} is excluded.

It has been demonstrated that Eq. (2a) can be solved effectively and accurately by the recently developed non-singular field only surface integral method [25, 26]. To obtain the electric field on the boundary, each Cartesian component of the 𝑬\boldsymbol{E} field in Eq. (2a), which obeys the scalar Helmholtz equation (an elliptic partial differential equation), is represented by the non-singular boundary integral equation [29] as (see Appendix A for derivation)

4π𝑬(𝒙0)+S[𝑬(𝒙)Hk(𝒙,𝒙0)𝑭(𝒙)H0(𝒙,𝒙0)]dS(𝒙)\displaystyle 4\pi\boldsymbol{E}(\boldsymbol{x}_{0})+\int_{S}\left[\boldsymbol{E}(\boldsymbol{x})H_{k}(\boldsymbol{x},\boldsymbol{x}_{0})-\boldsymbol{F}(\boldsymbol{x})H_{0}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x}) (3)
=S[𝑬(𝒙)nGk(𝒙,𝒙0)𝑭(𝒙)nG0(𝒙,𝒙0)]dS(𝒙)𝑰Ed.\displaystyle=\int_{S}\left[\frac{\partial\boldsymbol{E}(\boldsymbol{x})}{\partial n}G_{k}(\boldsymbol{x},\boldsymbol{x}_{0})-\frac{\partial\boldsymbol{F}(\boldsymbol{x})}{\partial n}G_{0}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x})-\boldsymbol{I}^{\mathrm{Ed}}.

where 𝒙0\boldsymbol{x}_{0} is an observation point and 𝒙\boldsymbol{x} is an integration point, as shown in Fig. 1. Gk(𝒙,𝒙0)=exp(ik|𝒙𝒙0|)/|𝒙𝒙0|G_{k}(\boldsymbol{x},\boldsymbol{x}_{0})=\exp(\mathrm{i}k|\boldsymbol{x}-\boldsymbol{x}_{0}|)/|\boldsymbol{x}-\boldsymbol{x}_{0}| is the Green’s function for the Helmholtz equation and G0=1/|𝒙𝒙0|G_{0}=1/|\boldsymbol{x}-\boldsymbol{x}_{0}| the Laplace equivalent. Hk=Gk/nH_{k}=\partial G_{k}/\partial n is the normal derivative of the Green’s function as in the standard boundary element framework with /n=𝒏\partial/\partial n=\boldsymbol{n}\cdot\nabla where 𝒏\boldsymbol{n} is the unit normal vector pointing into the object. The classical boundary integral equation is made non-singular analytically by subtracting a function 𝑭(𝒙)\boldsymbol{F}(\boldsymbol{x}):

𝑭(𝒙)=𝑬(𝒙0)+𝒏(𝒙0)(𝒙𝒙0)𝑬(𝒙0)n.\displaystyle\boldsymbol{F}(\boldsymbol{x})=\boldsymbol{E}(\boldsymbol{x}_{0})+\boldsymbol{n}(\boldsymbol{x}_{0})\cdot(\boldsymbol{x}-\boldsymbol{x}_{0})\frac{\partial\boldsymbol{E}(\boldsymbol{x}_{0})}{\partial n}. (4)

which satisfies the Laplace equation as 2𝑭(𝒙)=𝟎\nabla^{2}\boldsymbol{F}(\boldsymbol{x})=\boldsymbol{0}. Note that both terms in each bracket of Eq. (3) cancel each other out when 𝒙\boldsymbol{x} approaches 𝒙0\boldsymbol{x}_{0} and effectively eliminate the singularities analytically which arise due to the Green’s function.

The term 𝑰Ed\boldsymbol{I}^{\mathrm{Ed}} represents the integral on SdS_{d} which is a tiny spherical surface with vanishing radius rdr_{d} that encloses the dipole:

𝑰Ed=limrd0Sd[Hk(𝒙,𝒙0)𝑬(𝒙)Gk(𝒙,𝒙0)𝑬(𝒙)n]dS(𝒙).\displaystyle\boldsymbol{I}^{\mathrm{Ed}}=\lim_{r_{d}\rightarrow 0}\int_{S_{d}}\left[H_{k}(\boldsymbol{x},\boldsymbol{x}_{0})\boldsymbol{E}(\boldsymbol{x})-G_{k}(\boldsymbol{x},\boldsymbol{x}_{0})\frac{\partial\boldsymbol{E}(\boldsymbol{x})}{\partial n}\right]\mathrm{d}S(\boldsymbol{x}). (5)

In Appendix A, it is shown that the term 𝑰Ed\boldsymbol{I}^{\mathrm{Ed}} can be written as:

𝑰Ed\displaystyle\boldsymbol{I}^{\mathrm{Ed}} =1ϵ0ϵrd×{dGk(𝒙d,𝒙0)×𝒑}=1ϵ0ϵr×{Gk(𝒙0,𝒙d)×𝒑}\displaystyle=-\frac{1}{\epsilon_{0}\epsilon_{r}}\nabla_{d}\times\left\{\nabla_{d}G_{k}(\boldsymbol{x}_{d},\boldsymbol{x}_{0})\times\boldsymbol{p}\right\}=-\frac{1}{\epsilon_{0}\epsilon_{r}}\nabla\times\left\{\nabla G_{k}(\boldsymbol{x}_{0},\boldsymbol{x}_{d})\times\boldsymbol{p}\right\} (6)
=exp(ikr0d)ϵ0ϵrr0d3{(k2r0d23ikr0d+3)(𝒙0d𝒑)r0d2𝒙0d+(k2r0d2+ikr0d1)𝒑}\displaystyle=-\frac{\exp(\mathrm{i}kr_{0d})}{\epsilon_{0}\epsilon_{r}r_{0d}^{3}}\left\{(-k^{2}r_{0d}^{2}-3\mathrm{i}kr_{0d}+3)\frac{(\boldsymbol{x}_{0d}\cdot\boldsymbol{p})}{r_{0d}^{2}}\boldsymbol{x}_{0d}+(k^{2}r_{0d}^{2}+\mathrm{i}kr_{0d}-1)\boldsymbol{p}\right\}

in which 𝒙0d=𝒙0𝒙d\boldsymbol{x}_{0d}=\boldsymbol{x}_{0}-\boldsymbol{x}_{d}, r0d=|𝒙0d|r_{0d}=|\boldsymbol{x}_{0d}| and Gk(𝒙d,𝒙0)=Gk(𝒙0,𝒙d)G_{k}(\boldsymbol{x}_{d},\boldsymbol{x}_{0})=G_{k}(\boldsymbol{x}_{0},\boldsymbol{x}_{d}) are used.

The above boundary integral equation (3) gives a relationship between the Cartesian components of the electric field (ExE_{x}, EyE_{y}, EzE_{z}) and their normal derivatives (Ex/n\partial E_{x}/\partial n, Ey/n\partial E_{y}/\partial n, Ez/n\partial E_{z}/\partial n) on the surface SS. However, the divergence free condition in the domain, 𝑬=0\nabla\cdot\boldsymbol{E}=0, also needs to be satisfied simultaneously. Sun et al. [25] demonstrated that when the divergence free condition is satisfied everywhere on the surface of the object, it is automatically satisfied everywhere in the domain. It has also been shown in Sun et al. [25] that the divergence free condition 𝑬=0\nabla\cdot\boldsymbol{E}=0 is equal to

𝒏𝑬nκEn+Et1t1+Et2t2=0.\boldsymbol{n}\cdot\frac{\partial\boldsymbol{E}}{\partial n}-\kappa E_{n}+\frac{\partial E_{t_{1}}}{\partial t_{1}}+\frac{\partial E_{t2}}{\partial t_{2}}=0. (7)

Here κ\kappa is the curvature on the surface, En𝑬𝒏E_{n}\equiv\boldsymbol{E}\cdot\boldsymbol{n} is the normal component of the electric field on the surface, and Et1𝑬𝒕1E_{t1}\equiv\boldsymbol{E}\cdot\boldsymbol{t}_{1}, Et2𝑬𝒕2E_{t2}\equiv\boldsymbol{E}\cdot\boldsymbol{t}_{2} are the tangential components with 𝒕1\boldsymbol{t}_{1} and 𝒕2\boldsymbol{t}_{2} the two orthogonal surface unit tangential vectors and /t1=𝒕1\partial/\partial t_{1}=\boldsymbol{t}_{1}\cdot\nabla (and similar for /t2\partial/\partial t_{2}).

Since the boundary conditions are given in terms of tangential components of the electric field, it turns out to be easier to use the normal and tangential components as unknowns in our framework. As such, we rewrite the electric field and its normal derivative as

𝑬\displaystyle\boldsymbol{E} =Ex𝒆x+Ey𝒆y+Ez𝒆z=En𝒏+Et1𝒕1+Et2𝒕2,\displaystyle=E_{x}\boldsymbol{e}_{x}+E_{y}\boldsymbol{e}_{y}+E_{z}\boldsymbol{e}_{z}=E_{n}\boldsymbol{n}+E_{t1}\boldsymbol{t}_{1}+E_{t2}\boldsymbol{t}_{2}, (8)
𝑬n\displaystyle\frac{\partial\boldsymbol{E}}{\partial n} =Exn𝒆x+Eyn𝒆y+Ezn𝒆z=(𝒏𝑬n)𝒏+(𝒕1𝑬n)𝒕1+(𝒕2𝑬n)𝒕2.\displaystyle=\frac{\partial E_{x}}{\partial n}\boldsymbol{e}_{x}+\frac{\partial E_{y}}{\partial n}\boldsymbol{e}_{y}+\frac{\partial E_{z}}{\partial n}\boldsymbol{e}_{z}=\left(\boldsymbol{n}\cdot\frac{\partial\boldsymbol{E}}{\partial n}\right)\boldsymbol{n}+\left(\boldsymbol{t}_{1}\cdot\frac{\partial\boldsymbol{E}}{\partial n}\right)\boldsymbol{t}_{1}+\left(\boldsymbol{t}_{2}\cdot\frac{\partial\boldsymbol{E}}{\partial n}\right)\boldsymbol{t}_{2}.

Thus for example ExE_{x} can be expressed as Ex=En(𝒏𝒆x)+Et1(𝒕1𝒆x)+Et2(𝒕2𝒆x)=Ennx+Et1t1x+Et2t2xE_{x}=E_{n}(\boldsymbol{n}\cdot\boldsymbol{e}_{x})+E_{t1}(\boldsymbol{t}_{1}\cdot\boldsymbol{e}_{x})+E_{t2}(\boldsymbol{t}_{2}\cdot\boldsymbol{e}_{x})=E_{n}n_{x}+E_{t1}t_{1x}+E_{t2}t_{2x}.

To numerically solve Eq. (3) for the interaction between a dynamic electric dipole and a PEC object, the object surface is discretised with NN nodes. Then, the left hand side of Eq. (3) results in a matrix equation relating all xx-components of the electric field as: Ex{\cal{H}}E_{x} and its normal derivatives on the right hand side of Eq. (3) as: 𝒢Ex/n{\cal{G}}\partial E_{x}/\partial n. Here, 𝒢{\cal{G}} and {\cal{H}} are N×NN\times N matrices and ExE_{x} and Ex/n\partial E_{x}/\partial n are now column vectors of size NN. A similar procedure applies for the yy and zz-components. To introduce boundary condition in Eq. (7) into the linear system, the Cartesian components of the electric field and its normal derivatives can be reconstructed from the normal and tangential components with Eq. 8. Also, for a PEC object, the surface tangential components of the electric field are zeros. As such, we get the following 3N×3N3N\times 3N matrix system as

[nx(κ𝒢)t1x𝒢t2x𝒢ny(κ𝒢)t1y𝒢t2y𝒢nz(κ𝒢)t1z𝒢t2z𝒢][Ensc𝒕1𝑬scn𝒕2𝑬scn]=[IxEdIyEdIzEd],\begin{bmatrix}n_{x}({\cal{H}}-\kappa{\cal{G}})&-t_{1x}{\cal{G}}&-t_{2x}{\cal{G}}\\ n_{y}({\cal{H}}-\kappa{\cal{G}})&-t_{1y}{\cal{G}}&-t_{2y}{\cal{G}}\\ n_{z}({\cal{H}}-\kappa{\cal{G}})&-t_{1z}{\cal{G}}&-t_{2z}{\cal{G}}\end{bmatrix}\begin{bmatrix}E_{n}^{\mathrm{sc}}\\ \boldsymbol{t}_{1}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{sc}}}{\partial n}\\ \boldsymbol{t}_{2}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{sc}}}{\partial n}\end{bmatrix}=\begin{bmatrix}-I_{x}^{\mathrm{Ed}}\\ -I_{y}^{\mathrm{Ed}}\\ -I_{z}^{\mathrm{Ed}}\end{bmatrix}, (9)

from which the electric field on the PEC surface can be obtained. The interested reader can find more details on the implementation in Sun et al. [25].

2.2 Electric dipole in and/or near a dielectric object

For a dielectric object, we need to calculate the field inside the object 𝑬in\boldsymbol{E}^{\mathrm{in}} as well as the external electric field on both sides of the surface 𝑬out\boldsymbol{E}^{\mathrm{out}}. The internal domain is characterised by kink_{in}, ϵin\epsilon_{\mathrm{in}}, μin\mu_{\mathrm{in}} while the outer domain by koutk_{\mathrm{out}}, ϵout\epsilon_{\mathrm{out}} and μout\mu_{\mathrm{out}}. The field only surface integral equation for the external domain s:

4π𝑬out(𝒙0)+S[𝑬out(𝒙)Hkout(𝒙,𝒙0)𝑭out(𝒙)H0(𝒙,𝒙0)]dS(𝒙)\displaystyle 4\pi\boldsymbol{E}^{\mathrm{out}}(\boldsymbol{x}_{0})+\int_{S}\left[\boldsymbol{E}^{\mathrm{out}}(\boldsymbol{x})H_{\mathrm{kout}}(\boldsymbol{x},\boldsymbol{x}_{0})-\boldsymbol{F}^{\mathrm{out}}(\boldsymbol{x})H_{0}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x}) (10)
=S[𝑬out(𝒙)nGkout(𝒙,𝒙0)𝑭out(𝒙)nG0(𝒙,𝒙0)]dS(𝒙)𝑰outEd.\displaystyle=\int_{S}\left[\frac{\partial\boldsymbol{E}^{\mathrm{out}}(\boldsymbol{x})}{\partial n}G_{\mathrm{kout}}(\boldsymbol{x},\boldsymbol{x}_{0})-\frac{\partial\boldsymbol{F}^{\mathrm{out}}(\boldsymbol{x})}{\partial n}G_{0}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x})-\boldsymbol{I}^{\mathrm{Ed}}_{\mathrm{out}}.

For the internal domain we find:

S[𝑬in(𝒙)Hkin(𝒙,𝒙0)𝑭in(𝒙)H0(𝒙,𝒙0)]dS(𝒙)\displaystyle\int_{S}\left[\boldsymbol{E}^{\mathrm{in}}(\boldsymbol{x})H_{\mathrm{kin}}(\boldsymbol{x},\boldsymbol{x}_{0})-\boldsymbol{F}^{\mathrm{in}}(\boldsymbol{x})H_{0}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x}) (11)
=\displaystyle= S[𝑬in(𝒙)nGkin(𝒙,𝒙0)𝑭in(𝒙)nG0(𝒙,𝒙0)]dS(𝒙)𝑰inEd.\displaystyle\int_{S}\left[\frac{\partial\boldsymbol{E}^{\mathrm{in}}(\boldsymbol{x})}{\partial n}G_{\mathrm{kin}}(\boldsymbol{x},\boldsymbol{x}_{0})-\frac{\partial\boldsymbol{F}^{\mathrm{in}}(\boldsymbol{x})}{\partial n}G_{0}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x})-\boldsymbol{I}^{\mathrm{Ed}}_{\mathrm{in}}.

In Eqs. (10) and (11), 𝑭out(𝒙)=𝑬out(𝒙0)+𝒏(𝒙0)(𝒙𝒙0)𝑬out(𝒙0)n\boldsymbol{F}^{\mathrm{out}}(\boldsymbol{x})=\boldsymbol{E}^{\mathrm{out}}(\boldsymbol{x}_{0})+\boldsymbol{n}(\boldsymbol{x}_{0})\cdot(\boldsymbol{x}-\boldsymbol{x}_{0})\frac{\partial\boldsymbol{E}^{\mathrm{out}}(\boldsymbol{x}_{0})}{\partial n} and similar for 𝑭in\boldsymbol{F}^{in} as 𝑭in(𝒙)=𝑬in(𝒙0)+𝒏(𝒙0)(𝒙𝒙0)𝑬in(𝒙0)n\boldsymbol{F}^{\mathrm{in}}(\boldsymbol{x})=\boldsymbol{E}^{\mathrm{in}}(\boldsymbol{x}_{0})+\boldsymbol{n}(\boldsymbol{x}_{0})\cdot(\boldsymbol{x}-\boldsymbol{x}_{0})\frac{\partial\boldsymbol{E}^{\mathrm{in}}(\boldsymbol{x}_{0})}{\partial n}. Note the difference of a term with 4π𝑬out4\pi\boldsymbol{E}^{\mathrm{out}} between Eqs. (10) and (11) due to the absence of integrals at infinity for the internal problem in Eq. (11) (see Appendix A).

The boundary conditions for the electric field, expressed in terms of the normal and tangential components, are:

Enin\displaystyle E_{n}^{\mathrm{in}} =ϵoiEnout\displaystyle=\epsilon_{\mathrm{oi}}E_{n}^{\mathrm{out}} (12a)
Et1in\displaystyle E_{t1}^{\mathrm{in}} =Et1out\displaystyle=E_{t1}^{\mathrm{out}} (12b)
Et2in\displaystyle E_{t2}^{\mathrm{in}} =Et2out\displaystyle=E_{t2}^{\mathrm{out}} (12c)

with ϵoi=ϵout/ϵin\epsilon_{\mathrm{oi}}=\epsilon_{\mathrm{out}}/\epsilon_{\mathrm{in}} the ratio of permittivities between the two media. Meanwhile, the divergence condition of Eq. (7) is applied to 𝑬out\boldsymbol{E}^{\mathrm{out}} as 𝒏𝑬outnκEnout+Et1outt1+Et2outt2=0\boldsymbol{n}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{out}}}{\partial n}-\kappa E_{n}^{\mathrm{out}}+\frac{\partial E_{t_{1}}^{\mathrm{out}}}{\partial t_{1}}+\frac{\partial E_{t2}^{\mathrm{out}}}{\partial t_{2}}=0 and to 𝑬in\boldsymbol{E}^{\mathrm{in}} as 𝒏𝑬innκEnin+Et1int1+Et2mint2=0\boldsymbol{n}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{in}}}{\partial n}-\kappa E_{n}^{\mathrm{in}}+\frac{\partial E_{t_{1}}^{\mathrm{in}}}{\partial t_{1}}+\frac{\partial E_{t2}^{\min}}{\partial t_{2}}=0. In order to get rid of the tangential derivative terms these two equations are subtracted while the boundary conditions of Eq. (12b) and (12c) are used. As such, we can express EninE_{n}^{\mathrm{in}} in terms of EnoutE_{n}^{\mathrm{out}} with Eq. (12a):

𝒏𝑬inn=κ(ϵoi1)Enout+𝒏𝑬outn.\boldsymbol{n}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{in}}}{\partial n}=\kappa\left(\epsilon_{\mathrm{oi}}-1\right)E_{n}^{\mathrm{out}}+\boldsymbol{n}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{out}}}{\partial n}. (13)

This equation relates the normal component of the normal derivatives, 𝒏(𝑬/n)\boldsymbol{n}\cdot(\partial\boldsymbol{E}/\partial n), at both sides of the boundary of the object. We still need another equation to relate both tangential components of the normal derivatives: 𝒕(𝑬/n)\boldsymbol{t}\cdot(\partial\boldsymbol{E}/\partial n). This can be done by using the tangential continuity of the magnetic field on the dielectric surface: Ht1in=Ht1outH_{t1}^{\mathrm{in}}=H_{t1}^{\mathrm{out}} and Ht2in=Ht2outH_{t2}^{\mathrm{in}}=H_{t2}^{\mathrm{out}}. Using the convention 𝒏=𝒕1×𝒕2\boldsymbol{n}=\boldsymbol{t}_{1}\times\boldsymbol{t}_{2}, 𝒕1𝒕2=0\boldsymbol{t}_{1}\cdot\boldsymbol{t}_{2}=0, we express Ht1H_{t1} and Ht2H_{t2} in terms of the electric field on the surface as

Ht1=𝒕1𝑯=𝒕2(𝒏×𝑯)=1iωμ𝒕2(𝒏××𝑬)=1iωμ[𝒕1𝑬n𝒏𝑬t1],\displaystyle H_{t1}=\boldsymbol{t}_{1}\cdot\boldsymbol{H}=\boldsymbol{t}_{2}\cdot(\boldsymbol{n}\times\boldsymbol{H})=\frac{1}{\mathrm{i}\omega\mu}\boldsymbol{t}_{2}\cdot(\boldsymbol{n}\times\nabla\times\boldsymbol{E})=\frac{1}{\mathrm{i}\omega\mu}\left[\boldsymbol{t}_{1}\cdot\frac{\partial\boldsymbol{E}}{\partial n}-\boldsymbol{n}\cdot\frac{\partial\boldsymbol{E}}{\partial t_{1}}\right], (14)
Ht2=𝒕2𝑯=𝒕1(𝒏×𝑯)=1iωμ𝒕1(𝒏××𝑬)=1iωμ[𝒕2𝑬n𝒏𝑬t2].\displaystyle-H_{t2}=-\boldsymbol{t}_{2}\cdot\boldsymbol{H}=\boldsymbol{t}_{1}\cdot(\boldsymbol{n}\times\boldsymbol{H})=\frac{1}{\mathrm{i}\omega\mu}\boldsymbol{t}_{1}\cdot(\boldsymbol{n}\times\nabla\times\boldsymbol{E})=\frac{1}{\mathrm{i}\omega\mu}\left[\boldsymbol{t}_{2}\cdot\frac{\partial\boldsymbol{E}}{\partial n}-\boldsymbol{n}\cdot\frac{\partial\boldsymbol{E}}{\partial t_{2}}\right].

It is worth noting the inversion of t1t_{1} and t2t_{2} and the appearance of a minus sign in the Ht2H_{t2} term in the above equation.

For simplicity, we assume that μin=μout\mu_{\mathrm{in}}=\mu_{\mathrm{out}} (if this is not the case, see Sun et al. [26]), and then the tangential continuity of 𝑯\boldsymbol{H} across the dielectric surface is identical to

𝒕1𝑬inn𝒏𝑬int1=𝒕1𝑬outn𝒏𝑬outt1,\displaystyle\boldsymbol{t}_{1}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{in}}}{\partial n}-\boldsymbol{n}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{in}}}{\partial t_{1}}=\boldsymbol{t}_{1}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{out}}}{\partial n}-\boldsymbol{n}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{out}}}{\partial t_{1}}, (15)
𝒕2𝑬inn𝒏𝑬int2=𝒕2𝑬outn𝒏𝑬outt2.\displaystyle\boldsymbol{t}_{2}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{in}}}{\partial n}-\boldsymbol{n}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{in}}}{\partial t_{2}}=\boldsymbol{t}_{2}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{out}}}{\partial n}-\boldsymbol{n}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{out}}}{\partial t_{2}}.

The terms with 𝒏𝑬/t1\boldsymbol{n}\cdot\partial\boldsymbol{E}/\partial t_{1} and 𝒏𝑬/t2\boldsymbol{n}\cdot\partial\boldsymbol{E}/\partial t_{2} can be rewritten as111This can easiest be seen by writing the vector 𝑬\boldsymbol{E} into normal and tangential components as: 𝑬/t1=t1[En𝒏+Et1𝒕1+Et2𝒕2]=𝒏Ent1+En𝒏t1+𝒕1Et1t1+Et1𝒕1t1+𝒕2Et2t1+Et2𝒕2t1\partial\boldsymbol{E}/\partial t_{1}=\frac{\partial}{\partial t_{1}}[E_{n}\boldsymbol{n}+E_{t1}\boldsymbol{t}_{1}+E_{t2}\boldsymbol{t}_{2}]=\boldsymbol{n}\frac{\partial E_{n}}{\partial t_{1}}+E_{n}\frac{\partial\boldsymbol{n}}{\partial t_{1}}+\boldsymbol{t}_{1}\frac{\partial E_{t1}}{\partial t_{1}}+E_{t1}\frac{\partial\boldsymbol{t}_{1}}{\partial t_{1}}+\boldsymbol{t}_{2}\frac{\partial E_{t2}}{\partial t_{1}}+E_{t2}\frac{\partial\boldsymbol{t}_{2}}{\partial t_{1}}. Using 𝒏/t1=κ1𝒕1\partial\boldsymbol{n}/\partial t_{1}=-\kappa_{1}\boldsymbol{t}_{1}, 𝒕1/t1=κ1𝒏\partial\boldsymbol{t}_{1}/\partial t_{1}=\kappa_{1}\boldsymbol{n} and 𝒕2/t1=𝟎\partial\boldsymbol{t}_{2}/\partial t_{1}=\boldsymbol{0}, we get 𝒏𝑬t1=Ent1+κ1Et1\boldsymbol{n}\cdot\frac{\partial\boldsymbol{E}}{\partial t_{1}}=\frac{\partial E_{n}}{\partial t_{1}}+\kappa_{1}E_{t1} and similar for 𝒏𝑬t2=Ent2+κ2Et2\boldsymbol{n}\cdot\frac{\partial\boldsymbol{E}}{\partial t_{2}}=\frac{\partial E_{n}}{\partial t_{2}}+\kappa_{2}E_{t2}. 𝒏𝑬/t1=En/t1+κ1Et1\boldsymbol{n}\cdot\partial\boldsymbol{E}/\partial t_{1}=\partial E_{n}/\partial t_{1}+\kappa_{1}E_{t1} and 𝒏𝑬/t2=En/t2+κ2Et2\boldsymbol{n}\cdot\partial\boldsymbol{E}/\partial t_{2}=\partial E_{n}/\partial t_{2}+\kappa_{2}E_{t2}, with κ1\kappa_{1} the curvature in the 𝒕1\boldsymbol{t}_{1} direction and κ2\kappa_{2} the curvature in the 𝒕2\boldsymbol{t}_{2} direction. Putting this into Eq. (15) where the terms with κ1Et1\kappa_{1}E_{t1} and κ2Et2\kappa_{2}E_{t2} cancel out due to Eqs. (12b) and (12c), and further using Eq. (12a) to express EninE_{n}^{\mathrm{in}} in terms of EnoutE_{n}^{\mathrm{out}}, we have Enoutt1+κ1Et1outEnint1κ1Et1in=Enoutt1(1ϵoi)\frac{\partial E_{n}^{\mathrm{out}}}{\partial t_{1}}+\kappa_{1}E_{t1}^{\mathrm{out}}-\frac{\partial E_{n}^{\mathrm{in}}}{\partial t_{1}}-\kappa_{1}E_{t1}^{\mathrm{in}}=\frac{\partial E_{n}^{\mathrm{out}}}{\partial t_{1}}(1-\epsilon_{\mathrm{oi}}). As such,

𝒕1𝑬inn=𝒕1𝑬outn+Enoutt1(ϵoi1),\displaystyle\boldsymbol{t}_{1}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{in}}}{\partial n}=\boldsymbol{t}_{1}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{out}}}{\partial n}+\frac{\partial E_{n}^{\mathrm{out}}}{\partial t_{1}}(\epsilon_{\mathrm{oi}}-1), (16)
𝒕2𝑬inn=𝒕2𝑬outn+Enoutt2(ϵoi1).\displaystyle\boldsymbol{t}_{2}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{in}}}{\partial n}=\boldsymbol{t}_{2}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{out}}}{\partial n}+\frac{\partial E_{n}^{\mathrm{out}}}{\partial t_{2}}(\epsilon_{\mathrm{oi}}-1).

With Eq. (12), we now can express all the components of 𝑬in\boldsymbol{E}^{\mathrm{in}} in terms of 𝑬out\boldsymbol{E}^{\mathrm{out}}. All components of the normal derivatives of the internal field can also be expressed in terms of outer field variables with Eq. (13) and Eq. (16). The Cartesian xx, yy and zz components can then be obtained with Eq. (8).

To numerically solve Eqs. (10) and (11) for the interaction between a dynamic electric dipole and a dielectric object, the object surface is discretised with NN nodes. Meanwhile, the xx, yy, zz components of 𝑬in\boldsymbol{E}^{\mathrm{in}} and 𝑬out\boldsymbol{E}^{\mathrm{out}} and their normal derivatives are expressed in terms of EnoutE_{n}^{\mathrm{out}}, Et1outE_{t1}^{\mathrm{out}}, Et2outE_{t2}^{\mathrm{out}}, 𝒏(𝑬out/n)\boldsymbol{n}\cdot(\partial\boldsymbol{E}^{\mathrm{out}}/\partial n), 𝒕1(𝑬out/n)\boldsymbol{t}_{1}\cdot(\partial\boldsymbol{E}^{\mathrm{out}}/\partial n) and 𝒕2(𝑬out/n)\boldsymbol{t}_{2}\cdot(\partial\boldsymbol{E}^{\mathrm{out}}/\partial n), which results in a 6N×6N6N\times 6N matrix system as

[nxoutt1xoutt2xoutnx𝒢outt1x𝒢outt2x𝒢outnyoutt1youtt2youtny𝒢outt1y𝒢outt2y𝒢outnzoutt1zoutt2zoutnz𝒢outt1z𝒢outt2z𝒢out¯inxt1xint2xinnx𝒢int1x𝒢int2x𝒢in¯inyt1yint2yinny𝒢int1y𝒢int2y𝒢in¯inzt1zint2zinnz𝒢int1z𝒢int2z𝒢in][EnoutEt1outEt2out𝒏𝑬outn𝒕1𝑬outn𝒕2𝑬outn]=[Iout,xEdIout,yEdIout,zEdIin,xEdIin,yEdIin,zEd],\begin{bmatrix}n_{x}{\cal{H}}_{\mathrm{out}}&t_{1x}{\cal{H}}_{\mathrm{out}}&t_{2x}{\cal{H}}_{\mathrm{out}}&-{n_{x}}{\cal{G}}_{\mathrm{out}}&-{t_{1x}}{\cal{G}}_{\mathrm{out}}&-{t_{2x}}{\cal{G}}_{\mathrm{out}}\\ n_{y}{\cal{H}}_{\mathrm{out}}&t_{1y}{\cal{H}}_{\mathrm{out}}&t_{2y}{\cal{H}}_{\mathrm{out}}&-{n_{y}}{\cal{G}}_{\mathrm{out}}&-{t_{1y}}{\cal{G}}_{\mathrm{out}}&-{t_{2y}}{\cal{G}}_{\mathrm{out}}\\ n_{z}{\cal{H}}_{\mathrm{out}}&t_{1z}{\cal{H}}_{\mathrm{out}}&t_{2z}{\cal{H}}_{\mathrm{out}}&-{n_{z}}{\cal{G}}_{\mathrm{out}}&-{t_{1z}}{\cal{G}}_{\mathrm{out}}&-{t_{2z}}{\cal{G}_{\mathrm{out}}}\\ \bar{{\cal{H}}}^{x}_{\mathrm{in}}&t_{1x}{\cal{H}}_{\mathrm{in}}&t_{2x}{\cal{H}}_{\mathrm{in}}&-n_{x}{\cal{G}}_{\mathrm{in}}&-t_{1x}{\cal{G}}_{\mathrm{in}}&-t_{2x}{\cal{G}}_{\mathrm{in}}\\ \bar{{\cal{H}}}^{y}_{\mathrm{in}}&t_{1y}{\cal{H}}_{\mathrm{in}}&t_{2y}{\cal{H}}_{\mathrm{in}}&-n_{y}{\cal{G}}_{\mathrm{in}}&-t_{1y}{\cal{G}}_{\mathrm{in}}&-t_{2y}{\cal{G}}_{\mathrm{in}}\\ \bar{{\cal{H}}}^{z}_{\mathrm{in}}&t_{1z}{\cal{H}}_{\mathrm{in}}&t_{2z}{\cal{H}}_{\mathrm{in}}&-n_{z}{\cal{G}}_{\mathrm{in}}&-t_{1z}{\cal{G}}_{\mathrm{in}}&-t_{2z}{\cal{G}}_{\mathrm{in}}\end{bmatrix}\cdot\begin{bmatrix}E_{n}^{\mathrm{out}}\\ E_{t_{1}}^{\mathrm{out}}\\ E_{t_{2}}^{\mathrm{out}}\\ \boldsymbol{n}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{out}}}{\partial n}\\ \boldsymbol{t}_{1}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{out}}}{\partial n}\\ \boldsymbol{t}_{2}\cdot\frac{\partial\boldsymbol{E}^{\mathrm{out}}}{\partial n}\end{bmatrix}=\begin{bmatrix}I^{\mathrm{Ed}}_{\mathrm{out},x}\\ I^{\mathrm{Ed}}_{\mathrm{out},y}\\ I^{\mathrm{Ed}}_{\mathrm{out},z}\\ I^{\mathrm{Ed}}_{\mathrm{in},x}\\ I^{\mathrm{Ed}}_{\mathrm{in},y}\\ I^{\mathrm{Ed}}_{\mathrm{in},z}\end{bmatrix}, (17a)
where (with α=x,y,z.\alpha=x,y,z.)
¯inα\displaystyle\bar{{\cal{H}}}^{\alpha}_{\mathrm{in}} =ϵoinαin+(ϵoi1)𝒢in[κnαt1αt1t2αt2],\displaystyle=\epsilon_{\mathrm{oi}}n_{\alpha}{\cal{H}}_{\mathrm{in}}+\left(\epsilon_{\mathrm{oi}}-1\right){\cal{G}}_{\mathrm{in}}\left[-\kappa n_{\alpha}-t_{1\alpha}\frac{\partial}{\partial t_{1}}-t_{2\alpha}\frac{\partial}{\partial t_{2}}\right], (17b)

In the above matrix system, the first line corresponds to outExout=𝒢outExout/n{\cal{H}}_{\mathrm{out}}E_{x}^{\mathrm{out}}={\cal{G}}_{\mathrm{out}}\partial E_{x}^{\mathrm{out}}/\partial n, the matrix equivalent of Eq. (10) for the xx-component. Line 2 and 3 correspond to the yy and zz-component, respectively. Lines 4, 5 and 6 correspond to the matrix equivalent of Eq. (11) for the xx, yy and zz-component of 𝑬in\boldsymbol{E}^{\mathrm{in}}, respectively. The implementation is straightforward, expect perhaps for the tangential derivative matrices /t1\partial/\partial t_{1} and /t2\partial/\partial t_{2} (for more details see Sun et al. [26]). Once the matrix system is solved we can get back the Cartesian components with Eq. (8).

As we now have the electric field and its normal derivative on the surface for both the internal and external domain, we can get the electric field anywhere in the domain by postprocessing (see Sun et al. [30]).

3 Results

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 2: Comparison between the analytical results for a dielectric sphere with a dipole at its center placed in an outer infinite domain and the numerical simulation. The dipole is placed at the center of the sphere, see (a). The parameters used are kouta=1k_{\mathrm{out}}a=1, kina=2k_{\mathrm{in}}a=2, ϵin/ϵout=4\epsilon_{\mathrm{in}}/\epsilon_{\mathrm{out}}=4 and μin=μout=1\mu_{\mathrm{in}}=\mu_{\mathrm{out}}=1. Excellent agreement is obtained with the analytical result, see (b). In (a) the electric field vectors of the outer domain are shown as well on the surface. (c) Convergence study and computational time. Even with 642 nodes the relative error is only 0.1 percent.

3.1 Verification with Mie alike theory

The numerical framework was tested with an electric dipole placed at the centre of a dielectric sphere for which an analytical solution exists. The radius is indicated with aa, and the physical parameters are chosen to be kouta=1k_{\mathrm{out}}a=1, kina=2k_{\mathrm{in}}a=2, ϵin/ϵout=4\epsilon_{\mathrm{in}}/\epsilon_{\mathrm{out}}=4 and μin=μout=1\mu_{\mathrm{in}}=\mu_{\mathrm{out}}=1. The number of nodes is 642 and 320 quadratic triangular elements were used to represent the sphere surface in the simulation with the dipole moment being 𝒑=(0,0,1)\boldsymbol{p}=(0,0,1). The analytical solution is described in Appendix B. We checked the numerical solutions of 𝑬\boldsymbol{E} on both sides of the surface and they show excellent agreement with the theory (not shown here). We also chose a circle with radius 1.2a1.2a at the plane y=0y=0 to confirm that the electric field is correct in the domain close to the sphere. 72 sample locations were seeded along that circle with equal angular interval of polar angle θ\theta where sample 1 corresponds to θ=0\theta=0. The results are shown in Fig. 2b, where the numerical calculations of the real parts of ExoutE_{x}^{\mathrm{out}} (black symbols) and EzoutE_{z}^{\mathrm{out}} (red symbols) are plotted as a function of the polar angle θ\theta (as indicated in Fig. 2a) and compared to their theoretical counterparts (lines). The theoretical and numerical solutions are virtually identical.In Fig. 2c, we investigated the convergence performance of our method with respect to the example demonstrated in Fig. 2b. As we can see, as the number of surface nodes increases, the average relative error between the analytical solution decreases and the computational time222All the cases in Fig. 2c were performed on a MacBook Pro with 2.6 GHz 6-Core Intel Core i7 processor and 32 GB 2400 MHz DDR4 Memory without parallel computation. increases. The relative error in that figure is defined as

Relative error=172|𝑬anaout𝑬numout|172|𝑬anaout|\displaystyle\text{Relative error}=\frac{\sum_{1}^{72}|\boldsymbol{E}^{\mathrm{out}}_{\text{ana}}-\boldsymbol{E}^{\mathrm{out}}_{\text{num}}|}{\sum_{1}^{72}|\boldsymbol{E}^{\mathrm{out}}_{\text{ana}}|} (18)

where the subscript ‘ana’ stands for the analytical solution and ‘num’ for the numerical solution.

Some other tests to examine the correctness and accuracy of the constructed non-singular field only boundary element framework were also performed: if the surface SS in Eq. (3) is very small or if this surface is situated far away from both the dipole and the observation point, then the electric field goes back to the field of the undisturbed dipole as it should, because both integrals in Eq. (3) will disappear for these cases.

These tests give us the confidence that the constructed framework is correct and robust. In the next sections we will investigate some more physically interesting examples.

3.2 Nano optical-antenna

Refer to caption
(a) 𝒑=(22,0,22i)\boldsymbol{p}=\left(\frac{\sqrt{2}}{2},0,-\frac{\sqrt{2}}{2}\mathrm{i}\right)
Refer to caption
(b) 𝒑=(1,0,0)\boldsymbol{p}=(1,0,0)
Refer to caption
(c) 𝒑=(22,0,22i)\boldsymbol{p}=\left(\frac{\sqrt{2}}{2},0,\frac{\sqrt{2}}{2}\mathrm{i}\right)
Refer to caption
(d) 𝒑=(0,0,i)\boldsymbol{p}=(0,0,\mathrm{i})
Figure 3: A dynamic electromagnetic dipole placed at the focal point of a parabolic mirror (PEC) with ka=20ka=20. In (a) and (c) the dipole is circular polarized with 𝒑=(1,0,i)2/2\boldsymbol{p}=(1,0,-\mathrm{i})\sqrt{2}/2 in (a) and oppositely circular polarized with 𝒑=(1,0,i)2/2\boldsymbol{p}=(1,0,\mathrm{i})\sqrt{2}/2 in (c). Note that the direction of the ‘beam’ is slightly changed to the left in (a) and to the right in (c). In (b) and (d) the dipole is linearly polarized in the xx-direction as 𝒑=(1,0,0)\boldsymbol{p}=(1,0,0) and along the zz-direction with 𝒑=(0,0,i)\boldsymbol{p}=(0,0,\mathrm{i}), respectively. When the dipole is polarized along the axis of the mirror (zz-direction in d), virtually no focused energy transfer is observed.

An optical antenna can be designed, synthesised and used to efficiently convert the freely propagating optical radiation to localised energy to enhance the interaction between light and mesoscopic structures [31]. In analogy to the radio antenna design, we investigate the interaction between a dynamic electric dipole and a parabolic nano-antenna transmitter device, as shown in Fig. 3. The shape and location of the nano parabolic transmitter are given by (x,y,z)=(asinθcosφ,asinθsinφ, 0.1a(cosθ1)+0.3sin2θ)(x,y,z)=(a\sin{\theta}\cos{\varphi},\,a\sin{\theta}\sin{\varphi},\,0.1a(\cos{\theta}-1)+0.3\sin^{2}{\theta}) with θ\theta the polar angle measured from the zz-axis, φ\varphi the azimuthal angle and aa the characteristic size of the nano-antenna. The dynamic dipole is positioned at the focal point of this antenna at (0, 0, 1.35a)(0,\,0,\,1.35a) which is along the symmetry axis of the parabolic nano-antenna (zz-axis). In this example, we chose the characteristic non-dimensional value ka=20ka=20. For such a kaka, when the light wavelength is λ=550nm\lambda=550\,\text{nm}, the corresponding characteristic size of the parabolic nano-antenna would be a=1.75μma=1.75\,\mu\text{m}. Also, we assume that this antenna is well coated to be a perfect mirror (PEC) when the surrounding medium is air (nair=1n_{\text{air}}=1), and use the theory and numerical method demonstrated in Sec. 2.1 to perform the calculations. For this test, 4002 nodes and 2000 quadratic triangular surface elements were used on the antenna surface.

In Fig. 3(b) and (d), the distribution of the electric field magnitude in the area surrounding the nano-antenna when a linearly polarised dipole interacts with it are shown. When the direction of dipole moment, 𝒑\boldsymbol{p}, is along the symmetry axis of the antenna, the conversion of the energy of the free radiation of the dipole is very low as shown in Fig. 3(d). On the other hand, when 𝒑\boldsymbol{p} is perpendicular to the symmetry axis of the parabolic nano-antenna, a strong localised energy flow is formed as shown in Fig. 3(b). For the circularly polarised dipole examples of Fig. 3(a) and (c), other than the obvious conversion of energy, note that an interesting phenomenon is the changing direction of the energy flow depending on the direction of the circular polarisation. Such a change of the direction can illustrate information of the spin angular momentum of the dipole.

3.3 Near field gold (Au) probes

Refer to caption
(a) 𝒑=(1,0,0)\boldsymbol{p}=(1,0,0)
Refer to caption
(b) 𝒑=(1,0,0)\boldsymbol{p}=(1,0,0)
Refer to caption
(c) 𝒑=(1,0,0)\boldsymbol{p}=(1,0,0)
Refer to caption
(d) 𝒑=(0,0,1)\boldsymbol{p}=(0,0,1)
Refer to caption
(e) 𝒑=(0,0,1)\boldsymbol{p}=(0,0,1)
Refer to caption
(f) 𝒑=(0,0,1)\boldsymbol{p}=(0,0,1)
Figure 4: Two Au probes in air, with an electric dipole located exactly in the center between them ((a) and (d)). The parameters are nAu=0.19+i4.71n_{\text{Au}}=0.19+\mathrm{i}4.71, nair=1.00n_{\text{air}}=1.00 and λ=800nm\lambda=800\,\text{nm}. For this case, 5762 nodes connected by 2880 quadratic triangular surface elements have been used on the surface of each probe. In the zoomed-in areas of (b) and (e) the Poynting vector 𝑺\boldsymbol{S} is plotted on the plane given by z=0z=0 and in (b) and (d) on the plane y=0y=0.

Duan et al.[32] showed that classical electromagnetic calculations are capable of capturing the essential physics of a nanoplasmonic system even down to nanometer scale sized gaps and bridges. Near field optical probes, such as metal tips, are one of the key components for near field microscopes in nano-optics [4]. The sensitivity and resolution of such a device is strongly dependent on the field distribution around the tip of the probe. We simulate the interaction between a dynamic electric dipole with two Au probes. The dipole is located at the origin of the reference coordinate system which radiates light with wavelength λ=800nm\lambda=800\,\text{nm}. The shape of the Au probe is described by: x=asinθcosφ[1+2cos2(θ/2)]/10x=a\sin{\theta}\cos{\varphi}[1+2\cos^{2}{(\theta/2)}]/10, y=asinθsinφ[1+2cos2(θ/2)]/10y=a\sin{\theta}\sin{\varphi}[1+2\cos^{2}{(\theta/2)}]/10 and z=acosθz=a\cos{\theta} where θ\theta is the polar angle measured from the zz-axis, φ\varphi is the azimuthal angle and aa is the characteristic size of the probe. As a result, the tip of the Au probe points up along the zz-axis. In this particular example, we use two such Au probes with length of 2 μ\mum (a=1μma=1\,\mu\text{m}). We rotate one probe 90o90^{\text{o}} around the yy-axis clockwise and the other counter-clockwise to make their symmetry axis along the xx-axis with their tips pointing to each other. We then position the probes along the xx-axis with the separation between the tips as 100nm100\,\text{nm} which implies the distance between the dipole to the tip of one probe to be 50nm50\,\text{nm}. With such a gap size, classical electromagnetic phenomena are dominant, whereas quantum effects [33] are deemed not significant.

We consider two scenarios: when the electric dipole moment is along the axis of the tips, 𝒑=(1,0,0)\boldsymbol{p}=(1,0,0) and when it is perpendicular to the axis of the tips, 𝒑=(0,0,1)\boldsymbol{p}=(0,0,1). In both scenarios, the surrounding medium is set to be air with nair=1.00n_{\text{air}}=1.00. When λ=800nm\lambda=800\,\text{nm}, the complex refractive index of Au is nAu=0.19+i4.71n_{\text{Au}}=0.19+\mathrm{i}4.71. As such, the coupling effects between the internal and external electromagnetic fields of the Au probes were solved using the framework described in Sec. 2.2.

In Figs. 4a and Fig. 4d, the distribution of the instantaneous induced surface charge (which is equivalent to the normal component of the surface electric field: Re(𝑬𝒏)Re(\boldsymbol{E}\cdot\boldsymbol{n})) is presented. When the electric dipole moment is along the axis of the tips, the electric polarisation of the radiation light from that dipole is parallel to the tip axis which induces axis-symmetric surface charge density with the highest amplitude at the ends of the tips, as demonstrated in Fig. 4a. However, when the electric dipole moment is perpendicular to the axis of the tips, the electric polarisation of the radiation light from that dipole is also perpendicular to the tip axis. This leads to opposite surface charges on the diametrically opposed surface points on the probes with neutral foremost tips, as demonstrated in Fig. 4d. The difference is also reflected by the energy propagation within the probes. We compared the amplitude of the Poynting vector 𝑺=1/2Re[𝑬×(𝑯)]\boldsymbol{S}=1/2\text{Re}[\boldsymbol{E}\times(\boldsymbol{H})^{*}] (with indicating the conjugate) for those two scenarios in Figs. 4b,c and Figs. 4e,f, respectively, where 𝑯\boldsymbol{H} was solved by the framework illustrated in Sec. 4. We can see that the electromagnetic energy penetrates deeper into the Au probes when the electric dipole moment is along the tip axis than when it is perpendicular to that axis.

4 Discussion

The focus of the previous sections was on the electric field. Given the symmetry between the 𝑬\boldsymbol{E} and 𝑯\boldsymbol{H} field in the Maxwell equations, in the domain that excluding the dipole the magnetic field 𝑯\boldsymbol{H} also satisfies the vector Helmholtz equation and the divergence free condition:

2𝑯+k2𝑯=𝟎,\displaystyle\nabla^{2}\boldsymbol{H}+k^{2}\boldsymbol{H}=\boldsymbol{0}, (19a)
𝑯=0.\displaystyle\nabla\cdot\boldsymbol{H}=0. (19b)

As such, we can equally well use the same framework from Sec. 2.2 for the magnetic field for the interaction between an electric dipole and a dielectric object when Eq. (1b) is used.

Though our method is demonstrated with the interaction between a single electric dipole in the previous section, it is obvious that such a method can be extended to multiple dipoles straightforwardly. Moreover, the idea to calculate 𝑰Ed\boldsymbol{I}^{\mathrm{Ed}} in Appendix A is a general one that can be used for other point sources, such as magnetic dynamic dipoles and dynamic quadrupoles. Since higher order multipoles can alternatively be represented by a few dipoles very close to each other, we concentrate on dipoles in this work. Besides electric dipoles, there also exist magnetic dipoles. For a magnetic dipole, the electric and magnetic fields are, respectively [34]

𝑬Hd=iωμ0μr4π{[exp(ik|𝒙𝒙d|)|𝒙𝒙d|]×𝒎},\displaystyle\boldsymbol{E}^{\mathrm{Hd}}=\frac{\mathrm{i}\omega\mu_{0}\mu_{r}}{4\pi}\left\{\nabla\left[\frac{\exp(\mathrm{i}k|\boldsymbol{x}-\boldsymbol{x}_{d}|)}{|\boldsymbol{x}-\boldsymbol{x}_{d}|}\right]\times\boldsymbol{m}\right\}, (20a)
𝑯Hd=14π×{[exp(ik|𝒙𝒙d|)|𝒙𝒙d|]×𝒎}\displaystyle\boldsymbol{H}^{\mathrm{Hd}}=\frac{1}{4\pi}\nabla\times\left\{\nabla\left[\frac{\exp(\mathrm{i}k|\boldsymbol{x}-\boldsymbol{x}_{d}|)}{|\boldsymbol{x}-\boldsymbol{x}_{d}|}\right]\times\boldsymbol{m}\right\} (20b)

where 𝒎\boldsymbol{m} is the magnetic dipole moment.

In summary, the electric and magnetic fields on dielectric scatterer surfaces driven by the multiple dynamic electric and magnetic dipoles in the unbounded external domain are

4π𝑬out(𝒙0)+S[𝑬out(𝒙)Hkout(𝒙,𝒙0)𝑭Eout(𝒙)H0(𝒙,𝒙0)]dS(𝒙)\displaystyle 4\pi\boldsymbol{E}^{\mathrm{out}}(\boldsymbol{x}_{0})+\int_{S}\left[\boldsymbol{E}^{\mathrm{out}}(\boldsymbol{x})H^{\mathrm{out}}_{k}(\boldsymbol{x},\boldsymbol{x}_{0})-\boldsymbol{F}^{\mathrm{out}}_{E}(\boldsymbol{x})H_{0}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x})
=\displaystyle= S[𝑬out(𝒙)nGkout(𝒙,𝒙0)𝑭Eout(𝒙)nG0(𝒙,𝒙0)]dS(𝒙)\displaystyle\int_{S}\left[\frac{\partial\boldsymbol{E}^{\mathrm{out}}(\boldsymbol{x})}{\partial n}G^{\mathrm{out}}_{k}(\boldsymbol{x},\boldsymbol{x}_{0})-\frac{\partial\boldsymbol{F}^{\mathrm{out}}_{E}(\boldsymbol{x})}{\partial n}G_{0}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x})
+i=1MEout1ϵ0ϵout×{Gkout(𝒙0,𝒙d,i)×𝒑i}+j=1MHout(iωμ0μout){Gkout(𝒙0,𝒙d,j)×𝒎j},\displaystyle+\sum_{i=1}^{M^{\mathrm{out}}_{E}}\frac{1}{\epsilon_{0}\epsilon_{\mathrm{out}}}\nabla\times\left\{\nabla G^{\mathrm{out}}_{k}(\boldsymbol{x}_{0},\boldsymbol{x}_{d,i})\times\boldsymbol{p}_{i}\right\}+\sum_{j=1}^{M^{\mathrm{out}}_{H}}(\mathrm{i}\omega\mu_{0}\mu_{\mathrm{out}})\left\{\nabla G^{\mathrm{out}}_{k}(\boldsymbol{x}_{0},\boldsymbol{x}_{d,j})\times\boldsymbol{m}_{j}\right\}, (21a)
𝑬out=0on scatterer surfaces;\displaystyle\nabla\cdot\boldsymbol{E}^{\mathrm{out}}=0\qquad\text{on scatterer surfaces}; (21b)
4π𝑯out(𝒙0)+S[𝑯out(𝒙)Hkout(𝒙,𝒙0)𝑭Hout(𝒙)H0(𝒙,𝒙0)]dS(𝒙)\displaystyle 4\pi\boldsymbol{H}^{\mathrm{out}}(\boldsymbol{x}_{0})+\int_{S}\left[\boldsymbol{H}^{\mathrm{out}}(\boldsymbol{x})H^{\mathrm{out}}_{k}(\boldsymbol{x},\boldsymbol{x}_{0})-\boldsymbol{F}^{\mathrm{out}}_{H}(\boldsymbol{x})H_{0}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x})
=\displaystyle= S[𝑯out(𝒙)nGkout(𝒙,𝒙0)𝑭Hout(𝒙)nG0(𝒙,𝒙0)]dS(𝒙)\displaystyle\int_{S}\left[\frac{\partial\boldsymbol{H}^{\mathrm{out}}(\boldsymbol{x})}{\partial n}G^{\mathrm{out}}_{k}(\boldsymbol{x},\boldsymbol{x}_{0})-\frac{\partial\boldsymbol{F}^{\mathrm{out}}_{H}(\boldsymbol{x})}{\partial n}G_{0}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x})
+i=1MEoutωi{Gkout(𝒙0,𝒙d,i)×𝒑i}+j=1MHout×{Gkout(𝒙0,𝒙d,j)×𝒎j},\displaystyle+\sum_{i=1}^{M^{\mathrm{out}}_{E}}\frac{\omega}{\mathrm{i}}\left\{\nabla G^{\mathrm{out}}_{k}(\boldsymbol{x}_{0},\boldsymbol{x}_{d,i})\times\boldsymbol{p}_{i}\right\}+\sum_{j=1}^{M^{\mathrm{out}}_{H}}\nabla\times\left\{\nabla G^{\mathrm{out}}_{k}(\boldsymbol{x}_{0},\boldsymbol{x}_{d,j})\times\boldsymbol{m}_{j}\right\}, (22a)
𝑯out=0on scatterer surfaces.\displaystyle\nabla\cdot\boldsymbol{H}^{\mathrm{out}}=0\qquad\text{on scatterer surfaces}. (22b)

where MEinM^{\mathrm{in}}_{E} and MHinM^{\mathrm{in}}_{H} are, respectively, the number of dynamic electric and magnetic dipoles in the external domain, and 𝑭Eout(𝒙)\boldsymbol{F}^{\mathrm{out}}_{E}(\boldsymbol{x}) 𝑭Hout(𝒙)\boldsymbol{F}^{\mathrm{out}}_{H}(\boldsymbol{x}) are, respectively,

𝑭Eout(𝒙)\displaystyle\boldsymbol{F}^{\mathrm{out}}_{E}(\boldsymbol{x}) =𝑬out(𝒙0)+[𝒏(𝒙0)(𝒙𝒙0)]𝑬out(𝒙0)n\displaystyle=\boldsymbol{E}^{\mathrm{out}}(\boldsymbol{x}_{0})+[\boldsymbol{n}(\boldsymbol{x}_{0})\cdot(\boldsymbol{x}-\boldsymbol{x}_{0})]\frac{\partial\boldsymbol{E}^{\mathrm{out}}(\boldsymbol{x}_{0})}{\partial n} (23a)
𝑭Hout(𝒙)\displaystyle\boldsymbol{F}^{\mathrm{out}}_{H}(\boldsymbol{x}) =𝑯out(𝒙0)+[𝒏(𝒙0)(𝒙𝒙0)]𝑯out(𝒙0)n.\displaystyle=\boldsymbol{H}^{\mathrm{out}}(\boldsymbol{x}_{0})+[\boldsymbol{n}(\boldsymbol{x}_{0})\cdot(\boldsymbol{x}-\boldsymbol{x}_{0})]\frac{\partial\boldsymbol{H}^{\mathrm{out}}(\boldsymbol{x}_{0})}{\partial n}. (23b)

For the internal field on the dielectric scatterer surface, we have

S[𝑬in(𝒙)Hkin(𝒙,𝒙0)𝑭Ein(𝒙)H0(𝒙,𝒙0)]dS(𝒙)\displaystyle\int_{S}\left[\boldsymbol{E}^{\mathrm{in}}(\boldsymbol{x})H^{\mathrm{in}}_{k}(\boldsymbol{x},\boldsymbol{x}_{0})-\boldsymbol{F}^{\mathrm{in}}_{E}(\boldsymbol{x})H_{0}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x})
=\displaystyle= S[𝑬in(𝒙)nGkin(𝒙,𝒙0)𝑭Ein(𝒙)nG0(𝒙,𝒙0)]dS(𝒙)\displaystyle\int_{S}\left[\frac{\partial\boldsymbol{E}^{\mathrm{in}}(\boldsymbol{x})}{\partial n}G^{\mathrm{in}}_{k}(\boldsymbol{x},\boldsymbol{x}_{0})-\frac{\partial\boldsymbol{F}^{\mathrm{in}}_{E}(\boldsymbol{x})}{\partial n}G_{0}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x})
i=1MEin1ϵ0ϵin×{Gkin(𝒙0,𝒙d,i)×𝒑i}j=1MHin(iωμ0μin){Gkin(𝒙0,𝒙d,j)×𝒎j},\displaystyle-\sum_{i=1}^{M^{\mathrm{in}}_{E}}\frac{1}{\epsilon_{0}\epsilon_{\mathrm{in}}}\nabla\times\left\{\nabla G^{\mathrm{in}}_{k}(\boldsymbol{x}_{0},\boldsymbol{x}_{d,i})\times\boldsymbol{p}_{i}\right\}-\sum_{j=1}^{M^{\mathrm{in}}_{H}}(\mathrm{i}\omega\mu_{0}\mu_{\mathrm{in}})\left\{\nabla G^{\mathrm{in}}_{k}(\boldsymbol{x}_{0},\boldsymbol{x}_{d,j})\times\boldsymbol{m}_{j}\right\}, (24a)
𝑬in=0on scatterer surfaces;\displaystyle\nabla\cdot\boldsymbol{E}^{\mathrm{in}}=0\qquad\text{on scatterer surfaces}; (24b)
S[𝑯in(𝒙)Hkin(𝒙,𝒙0)𝑭Hin(𝒙)H0(𝒙,𝒙0)]dS(𝒙)\displaystyle\int_{S}\left[\boldsymbol{H}^{\mathrm{in}}(\boldsymbol{x})H^{\mathrm{in}}_{k}(\boldsymbol{x},\boldsymbol{x}_{0})-\boldsymbol{F}^{\mathrm{in}}_{H}(\boldsymbol{x})H_{0}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x})
=\displaystyle= S[𝑯in(𝒙)nGkin(𝒙,𝒙0)𝑭Hin(𝒙)nG0(𝒙,𝒙0)]dS(𝒙)\displaystyle\int_{S}\left[\frac{\partial\boldsymbol{H}^{\mathrm{in}}(\boldsymbol{x})}{\partial n}G^{\mathrm{in}}_{k}(\boldsymbol{x},\boldsymbol{x}_{0})-\frac{\partial\boldsymbol{F}^{\mathrm{in}}_{H}(\boldsymbol{x})}{\partial n}G_{0}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x})
i=1MEinωi{Gkin(𝒙0,𝒙d,i)×𝒑i}j=1MHin×{Gkin(𝒙0,𝒙d,j)×𝒎j},\displaystyle-\sum_{i=1}^{M^{\mathrm{in}}_{E}}\frac{\omega}{\mathrm{i}}\left\{\nabla G^{\mathrm{in}}_{k}(\boldsymbol{x}_{0},\boldsymbol{x}_{d,i})\times\boldsymbol{p}_{i}\right\}-\sum_{j=1}^{M^{\mathrm{in}}_{H}}\nabla\times\left\{\nabla G^{\mathrm{in}}_{k}(\boldsymbol{x}_{0},\boldsymbol{x}_{d,j})\times\boldsymbol{m}_{j}\right\}, (25a)
𝑯in=0on scatterer surfaces.\displaystyle\nabla\cdot\boldsymbol{H}^{\mathrm{in}}=0\qquad\text{on scatterer surfaces}. (25b)

where MEinM^{\mathrm{in}}_{E} and MHinM^{\mathrm{in}}_{H} are, respectively, the number of dynamic electric and magnetic dipoles inside the scatterer, and 𝑭Ein(𝒙)\boldsymbol{F}^{\mathrm{in}}_{E}(\boldsymbol{x}) 𝑭Hin(𝒙)\boldsymbol{F}^{\mathrm{in}}_{H}(\boldsymbol{x}) are, respectively,

𝑭Ein(𝒙)\displaystyle\boldsymbol{F}^{\mathrm{in}}_{E}(\boldsymbol{x}) =𝑬in(𝒙0)+[𝒏(𝒙0)(𝒙𝒙0)]𝑬in(𝒙0)n\displaystyle=\boldsymbol{E}^{\mathrm{in}}(\boldsymbol{x}_{0})+[\boldsymbol{n}(\boldsymbol{x}_{0})\cdot(\boldsymbol{x}-\boldsymbol{x}_{0})]\frac{\partial\boldsymbol{E}^{\mathrm{in}}(\boldsymbol{x}_{0})}{\partial n} (26a)
𝑭Hin(𝒙)\displaystyle\boldsymbol{F}^{\mathrm{in}}_{H}(\boldsymbol{x}) =𝑯in(𝒙0)+[𝒏(𝒙0)(𝒙𝒙0)]𝑯in(𝒙0)n.\displaystyle=\boldsymbol{H}^{\mathrm{in}}(\boldsymbol{x}_{0})+[\boldsymbol{n}(\boldsymbol{x}_{0})\cdot(\boldsymbol{x}-\boldsymbol{x}_{0})]\frac{\partial\boldsymbol{H}^{\mathrm{in}}(\boldsymbol{x}_{0})}{\partial n}. (26b)

It is worth emphasizing that the sums with MEoutM^{\mathrm{out}}_{E}, MHoutM^{\mathrm{out}}_{H}, MEinM^{\mathrm{in}}_{E} and MHinM^{\mathrm{in}}_{H} will end up on the right hand side of the resulting numerical matrix system, and will thus require barely any noticeable additional computational resources, regardless how many dipoles are situated in the domain.

It is also worth pointing out that, for the examples shown in Secs. 3.2 and 3.3, we used the surface meshes that were transformed from the mesh of a sphere, which is just for our own convenience. Of course, for practical applications, our method can accept any surface mesh of a closed object.

5 Conclusions

In this work, we illustrate a non-singular field-only surface integral method for the interactions between electric and magnetic dipoles with mesoscale structures. We have shown that it is relatively easy to implement such volume sources, electric or magnetic dipoles, in our framework to simulate electromagnetic phenomena. The results were verified with a Mie alike theoretical solution. Two examples were given; one with a parabolic nano-antenna with a dipole at its focal point and one with two gold (Au) probes situated around a dipole. Our framework can also deal with multiple dipoles straightforwardly without any substantial increase in the computational requirements. We believe that our method can make a good contribution to addressing one of the key problems in nano-optics to determine the electromagnetic field distributions near and on mesoscale structures. Also, considering that the Maxwell equations describe electromagnetic phenomena for length scales ranging over more than 10 orders of magnitude, our framework can also be used beyond nano-optical topics.

Acknowledgements

Q.S. acknowledges the support from the Australian Research Council grants DE150100169, FT160100357 and CE140100003. This research was partially undertaken with the assistance of resources from the National Computational Infrastructure (NCI Australia), an NCRIS enabled capability supported by the Australian Government (Grant No. LE160100051).

Appendix A Derivation of the surface integral equations

In this appendix the derivations of the integral equations of Eq. (3) are given, including the desingularization and the treatment of the singular dipole terms. The standard boundary integral equation for each component of the electric field ExE_{x}, EyE_{y} and EzE_{z} is

C(𝒙0)𝑬(𝒙0)+S[𝑬(𝒙)Hk(𝒙,𝒙0)𝑬(𝒙)nGk(𝒙,𝒙0)]dS(𝒙)+𝑰Ed=𝟎,\displaystyle C(\boldsymbol{x}_{0})\boldsymbol{E}(\boldsymbol{x}_{0})+\int_{S}\left[\boldsymbol{E}(\boldsymbol{x})H_{k}(\boldsymbol{x},\boldsymbol{x}_{0})-\frac{\partial\boldsymbol{E}(\boldsymbol{x})}{\partial n}G_{k}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x})+\boldsymbol{I}^{\mathrm{Ed}}=\boldsymbol{0}, (27)

where we have excluded the integrals over a small sphere containing the dipole with surface SdS_{d} (see Fig. 1) as

𝑰Ed=Sd[𝑬(𝒙)Hk(𝒙,𝒙0)𝑬(𝒙)nGk(𝒙,𝒙0)]dS(𝒙).\displaystyle\boldsymbol{I}^{\mathrm{Ed}}=\int_{S_{d}}\left[\boldsymbol{E}(\boldsymbol{x})H_{k}(\boldsymbol{x},\boldsymbol{x}_{0})-\frac{\partial\boldsymbol{E}(\boldsymbol{x})}{\partial n}G_{k}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x}). (28)

In Eq. (27), C(𝒙0)C(\boldsymbol{x}_{0}) is the solid angle at 𝒙0\boldsymbol{x}_{0}. Numerically, due care needs to be taken in calculating the integrals since the integrals become singular at 𝒙=𝒙0\boldsymbol{x}=\boldsymbol{x}_{0} because of the presence of HkH_{k} and GkG_{k} in Eq. (27). However, these singularities can be analytically removed by subtracting other integrals with the same singular behavior. One possible choice would be:

C(𝒙0)𝑭(𝒙0)+S[𝑭(𝒙)H0(𝒙,𝒙0)𝑭(𝒙)nG0(𝒙,𝒙0)]dS(𝒙)=𝟎,\displaystyle C(\boldsymbol{x}_{0})\boldsymbol{F}(\boldsymbol{x}_{0})+\int_{S}\left[\boldsymbol{F}(\boldsymbol{x})H_{0}(\boldsymbol{x},\boldsymbol{x}_{0})-\frac{\partial\boldsymbol{F}(\boldsymbol{x})}{\partial n}G_{0}(\boldsymbol{x},\boldsymbol{x}_{0})\right]\mathrm{d}S(\boldsymbol{x})=\boldsymbol{0}, (29)

where 𝑭(𝒙)\boldsymbol{F}(\boldsymbol{x}) satisfies the Laplace equation as 2𝑭(𝒙)=𝟎\nabla^{2}\boldsymbol{F}(\boldsymbol{x})=\boldsymbol{0}, and G0G_{0} and H0H_{0} are the Green’s function for the Laplace equation and its normal derivative. G0G_{0} has the same 1/r1/r singular behavior as GkG_{k}. If we choose 𝑭(𝒙)=𝑬(𝒙0)+𝒏(𝒙0)(𝒙𝒙0)𝑬(𝒙0)/n\boldsymbol{F}(\boldsymbol{x})=\boldsymbol{E}(\boldsymbol{x}_{0})+\boldsymbol{n}(\boldsymbol{x}_{0})\cdot(\boldsymbol{x}-\boldsymbol{x}_{0})\partial\boldsymbol{E}(\boldsymbol{x}_{0})/\partial n, such that 𝑭\boldsymbol{F} approaches 𝑬(𝒙0)\boldsymbol{E}(\boldsymbol{x}_{0}) when 𝒙\boldsymbol{x} approaches 𝒙0\boldsymbol{x}_{0} and the normal derivative lim𝒙𝒙0F(𝒙0)/n=lim𝒙𝒙0𝒏(𝒙)𝒏(𝒙0)E(𝒙0)/n=E(𝒙0)/n\lim_{\boldsymbol{x}\rightarrow\boldsymbol{x}_{0}}\partial F(\boldsymbol{x}_{0})/\partial n=\lim_{\boldsymbol{x}\rightarrow\boldsymbol{x}_{0}}\boldsymbol{n}(\boldsymbol{x})\cdot\boldsymbol{n}(\boldsymbol{x}_{0})\partial E(\boldsymbol{x}_{0})/\partial n=\partial E(\boldsymbol{x}_{0})/\partial n. By subtracting Eq. (29) from Eq. (27), we end up with Eq. (3) of the main text. It is worth noting that the term with the solid angle C(𝒙0)C(\boldsymbol{x}_{0}) disappears. For an unbounded external domain, however, a term 4π𝑬(𝒙0)4\pi\boldsymbol{E}(\boldsymbol{x}_{0}) will appear due to the first term of the 𝑭(𝒙)\boldsymbol{F}(\boldsymbol{x}) function (the integral over a domain at infinity will no longer be zero but will generate this 4π𝑬(𝒙0)4\pi\boldsymbol{E}(\boldsymbol{x}_{0}) term. On the other hand, for internal domains, this term will not appear, see Eq. (11) for example.

The last term on the right hand side of Eq. (3), 𝑰Ed\boldsymbol{I}^{\mathrm{Ed}}, which is defined in Eq. (5), can be calculated analytically as follows. Since the size of SdS_{d} is small, the electric field on SdS_{d} is dominated by the electric dipole. As such, when Eq. (1a) is used, 𝑰Ed\boldsymbol{I}^{\mathrm{Ed}} can be calculated as

𝑰Ed=\displaystyle\boldsymbol{I}^{\mathrm{Ed}}= 14πϵ0ϵrlimrd0SdHk(𝒙,𝒙0)×{[exp(ik|𝒙𝒙d|)|𝒙𝒙d|]×𝒑}dS(𝒙)\displaystyle\quad\;\frac{1}{4\pi\epsilon_{0}\epsilon_{r}}\lim_{r_{d}\rightarrow 0}\int_{S_{d}}H_{k}(\boldsymbol{x},\boldsymbol{x}_{0})\,\nabla\times\left\{\nabla\left[\frac{\exp(\mathrm{i}k|\boldsymbol{x}-\boldsymbol{x}_{d}|)}{|\boldsymbol{x}-\boldsymbol{x}_{d}|}\right]\times\boldsymbol{p}\right\}\mathrm{d}S(\boldsymbol{x}) (30)
14πϵ0ϵrlimrd0SdGk(𝒙,𝒙0)n(×{[exp(ik|𝒙𝒙d|)|𝒙𝒙d|]×𝒑})dS(𝒙).\displaystyle-\frac{1}{4\pi\epsilon_{0}\epsilon_{r}}\lim_{r_{d}\rightarrow 0}\int_{S_{d}}G_{k}(\boldsymbol{x},\boldsymbol{x}_{0})\,\frac{\partial}{\partial n}\Bigg{(}\nabla\times\left\{\nabla\left[\frac{\exp(\mathrm{i}k|\boldsymbol{x}-\boldsymbol{x}_{d}|)}{|\boldsymbol{x}-\boldsymbol{x}_{d}|}\right]\times\boldsymbol{p}\right\}\Bigg{)}\mathrm{d}S(\boldsymbol{x}).

Since F(𝒙𝒙d)=dF(𝒙𝒙d)\nabla F(\boldsymbol{x}-\boldsymbol{x}_{d})=-\nabla_{d}F(\boldsymbol{x}-\boldsymbol{x}_{d}) and ×F(𝒙𝒙d)=d×F(𝒙𝒙d)\nabla\times F(\boldsymbol{x}-\boldsymbol{x}_{d})=-\nabla_{d}\times F(\boldsymbol{x}-\boldsymbol{x}_{d}) where d\nabla_{d} is the gradient with respect to 𝒙𝒅\boldsymbol{x_{d}} and FF any regular function, we can take d\nabla_{d} out of the integral and write :

𝑰Ed=\displaystyle\boldsymbol{I}^{\mathrm{Ed}}= 14πϵ0ϵrd×{dlimrd0SdHk(𝒙,𝒙0)[exp(ik|𝒙𝒙d|)|𝒙𝒙d|]dS(𝒙)×𝒑}\displaystyle\quad\;\frac{1}{4\pi\epsilon_{0}\epsilon_{r}}\,\nabla_{d}\times\left\{\nabla_{d}\lim_{r_{d}\rightarrow 0}\int_{S_{d}}H_{k}(\boldsymbol{x},\boldsymbol{x}_{0})\left[\frac{\exp(\mathrm{i}k|\boldsymbol{x}-\boldsymbol{x}_{d}|)}{|\boldsymbol{x}-\boldsymbol{x}_{d}|}\right]\mathrm{d}S(\boldsymbol{x})\times\boldsymbol{p}\right\} (31)
14πϵ0ϵrd×{dlimrd0SdGk(𝒙,𝒙0)n[exp(ik|𝒙𝒙d|)|𝒙𝒙d|]dS(𝒙)×𝒑}.\displaystyle-\frac{1}{4\pi\epsilon_{0}\epsilon_{r}}\nabla_{d}\times\left\{\nabla_{d}\lim_{r_{d}\rightarrow 0}\int_{S_{d}}G_{k}(\boldsymbol{x},\boldsymbol{x}_{0})\,\frac{\partial}{\partial n}\left[\frac{\exp(\mathrm{i}k|\boldsymbol{x}-\boldsymbol{x}_{d}|)}{|\boldsymbol{x}-\boldsymbol{x}_{d}|}\right]\mathrm{d}S(\boldsymbol{x})\times\boldsymbol{p}\right\}.

The integrals are now easier to deal with. For instance,

limrd0SdHk(𝒙,𝒙0)[exp(ik|𝒙𝒙d|)|𝒙𝒙d|]dS(𝒙)\displaystyle\lim_{r_{d}\rightarrow 0}\int_{S_{d}}H_{k}(\boldsymbol{x},\boldsymbol{x}_{0})\left[\frac{\exp(\mathrm{i}k|\boldsymbol{x}-\boldsymbol{x}_{d}|)}{|\boldsymbol{x}-\boldsymbol{x}_{d}|}\right]\mathrm{d}S(\boldsymbol{x}) (32)
\displaystyle\approxeq Hk(𝒙d,𝒙0)limrd0Sdexp(ik|𝒙𝒙d|)|𝒙𝒙d|dS(𝒙)=0\displaystyle H_{k}(\boldsymbol{x}_{d},\boldsymbol{x}_{0})\lim_{r_{d}\rightarrow 0}\int_{S_{d}}\frac{\exp(\mathrm{i}k|\boldsymbol{x}-\boldsymbol{x}_{d}|)}{|\boldsymbol{x}-\boldsymbol{x}_{d}|}\mathrm{d}S(\boldsymbol{x})=0

since |𝒙𝒙d|1/rd|\boldsymbol{x}-\boldsymbol{x}_{d}|\sim 1/r_{d}, but dS(𝒙)4πrd2\mathrm{d}S(\boldsymbol{x})\sim 4\pi r_{d}^{2}. The second integral can be calculated using /n=d/drd\partial/\partial n=-\mathrm{d}/\mathrm{d}r_{d} and [exp(ik|𝒙𝒙d|)/|𝒙𝒙d|]/n=(ikrd1)/rd2\partial\left[\exp(\mathrm{i}k|\boldsymbol{x}-\boldsymbol{x}_{d}|)/|\boldsymbol{x}-\boldsymbol{x}_{d}|\right]/\partial n=-(\mathrm{i}kr_{d}-1)/r_{d}^{2}, which leads to:

limrd0SdGk(𝒙,𝒙0)n[exp(ik|𝒙𝒙d|)|𝒙𝒙d|]dS(𝒙)\displaystyle\lim_{r_{d}\rightarrow 0}\int_{S_{d}}G_{k}(\boldsymbol{x},\boldsymbol{x}_{0})\,\frac{\partial}{\partial n}\left[\frac{\exp(\mathrm{i}k|\boldsymbol{x}-\boldsymbol{x}_{d}|)}{|\boldsymbol{x}-\boldsymbol{x}_{d}|}\right]\mathrm{d}S(\boldsymbol{x}) (33)
\displaystyle\approxeq Gk(𝒙d,𝒙0)limrd0Sdn[exp(ik|𝒙𝒙d|)|𝒙𝒙d|]dS(𝒙)\displaystyle G_{k}(\boldsymbol{x}_{d},\boldsymbol{x}_{0})\lim_{r_{d}\rightarrow 0}\int_{S_{d}}\,\frac{\partial}{\partial n}\left[\frac{\exp(\mathrm{i}k|\boldsymbol{x}-\boldsymbol{x}_{d}|)}{|\boldsymbol{x}-\boldsymbol{x}_{d}|}\right]\mathrm{d}S(\boldsymbol{x})
=\displaystyle= Gk(𝒙d,𝒙0)limrd0ikrd1rd24πrd2=4πGk(𝒙d,𝒙0).\displaystyle-G_{k}(\boldsymbol{x}_{d},\boldsymbol{x}_{0})\lim_{r_{d}\rightarrow 0}\frac{\mathrm{i}kr_{d}-1}{r_{d}^{2}}4\pi r_{d}^{2}=4\pi G_{k}(\boldsymbol{x}_{d},\boldsymbol{x}_{0}).

Eventually,

𝑰Ed\displaystyle\boldsymbol{I}^{\mathrm{Ed}} =1ϵ0ϵrd×{dGk(𝒙d,𝒙0)×𝒑}=1ϵ0ϵr×{Gk(𝒙0,𝒙d)×𝒑}\displaystyle=-\frac{1}{\epsilon_{0}\epsilon_{r}}\nabla_{d}\times\left\{\nabla_{d}G_{k}(\boldsymbol{x}_{d},\boldsymbol{x}_{0})\times\boldsymbol{p}\right\}=-\frac{1}{\epsilon_{0}\epsilon_{r}}\nabla\times\left\{\nabla G_{k}(\boldsymbol{x}_{0},\boldsymbol{x}_{d})\times\boldsymbol{p}\right\} (34)
=exp(ikr0d)ϵ0ϵrr0d3{(k2r0d23ikr0d+3)(𝒙0d𝒑)r0d2𝒙0d+(k2r0d2+ikr0d1)𝒑}\displaystyle=-\frac{\exp(\mathrm{i}kr_{0d})}{\epsilon_{0}\epsilon_{r}r_{0d}^{3}}\left\{(-k^{2}r_{0d}^{2}-3\mathrm{i}kr_{0d}+3)\frac{(\boldsymbol{x}_{0d}\cdot\boldsymbol{p})}{r_{0d}^{2}}\boldsymbol{x}_{0d}+(k^{2}r_{0d}^{2}+\mathrm{i}kr_{0d}-1)\boldsymbol{p}\right\}

which is the same as Eq. (6) in the main text.

Appendix B Analytical solution

In this appendix, the analytical solution of the field driven by an electric dipole with dipole moment 𝒑\boldsymbol{p} located at the center of a spherical dielectric scatter is derived. As shown in Fig. 2a, the centre of a spherical object is set at the origin of a Cartesian coordinate system. The axis of symmetry of this spherical scatterer is chosen along the zz-axis which aligns with the direction of the electric dipole moment 𝒑=(0, 0,p)\boldsymbol{p}=(0,\,0,\,p). To solve the electromagnetic field of such a system, it is most convenient to employ a spherical coordinate system, (r,θ,φ)(r,\,\theta,\,\varphi), which origin is at the center of the spherical scatterer and the polar angle is measured from the zz-axis. In this spherical system, 𝒑=pcosθ𝒆rpsinθ𝒆θ\boldsymbol{p}=p\cos{\theta}\,\boldsymbol{e}_{r}-p\sin{\theta}\,\boldsymbol{e}_{\theta} where 𝒆r\boldsymbol{e}_{r} is the unit vector along the rr-axis and 𝒆θ\boldsymbol{e}_{\theta} the unit vector along the θ\theta-axis.

From Eq. (1), the components of the electric and magnetic field in the spherical coordinate system driven by such an electric dipole inside the sphere are

ErEd\displaystyle E^{\mathrm{Ed}}_{r} =14πϵ0ϵ2p(2r32ik2r2)exp(ik2r)cosθ\displaystyle=\frac{1}{4\pi\epsilon_{0}\epsilon_{2}}p\left(\frac{2}{r^{3}}-\frac{2\mathrm{i}k_{2}}{r^{2}}\right)\exp{(\mathrm{i}k_{2}r)}\cos{\theta} (35a)
EθEd\displaystyle E^{\mathrm{Ed}}_{\theta} =14πϵ0ϵ2p(1r3ik2r2k22r)exp(ik2r)sinθ\displaystyle=\frac{1}{4\pi\epsilon_{0}\epsilon_{2}}p\left(\frac{1}{r^{3}}-\frac{\mathrm{i}k_{2}}{r^{2}}-\frac{k_{2}^{2}}{r}\right)\exp{(\mathrm{i}k_{2}r)}\sin{\theta} (35b)
EφEd\displaystyle E^{\mathrm{Ed}}_{\varphi} =0\displaystyle=0 (35c)
HrEd\displaystyle H^{\mathrm{Ed}}_{r} =0\displaystyle=0 (35d)
HθEd\displaystyle H^{\mathrm{Ed}}_{\theta} =0\displaystyle=0 (35e)
HφEd\displaystyle H^{\mathrm{Ed}}_{\varphi} =k234πϵ0ϵ2μ0μ2ωp(1ik2r21r)exp(ik2r)sinθ\displaystyle=\frac{k_{2}^{3}}{4\pi\epsilon_{0}\epsilon_{2}\mu_{0}\mu_{2}\omega}p\left(\frac{1}{\mathrm{i}k_{2}r^{2}}-\frac{1}{r}\right)\exp{(\mathrm{i}k_{2}r)}\sin{\theta} (35f)

where k2k_{2} is the wavenumber, c2c_{2} is the speed of light, and ϵ2\epsilon_{2} is the relative permittivity of the sphere.

Based on the Mie theory [35, 2], when the Debye potentials uu and vv are used, we can write the electric and magnetic fields as

𝑬\displaystyle\boldsymbol{E} =𝑴vi𝑵u,\displaystyle=\boldsymbol{M}_{v}-\mathrm{i}\boldsymbol{N}_{u}, (36a)
𝑯\displaystyle\boldsymbol{H} =ϵ0ϵrμ0μr(i𝑵v𝑴u)\displaystyle=\sqrt{\frac{\epsilon_{0}\epsilon_{r}}{\mu_{0}\mu_{r}}}(-\mathrm{i}\boldsymbol{N}_{v}-\boldsymbol{M}_{u}) (36b)

with

𝑴u=×(𝒙u),𝑴v=×(𝒙v);\displaystyle\boldsymbol{M}_{u}=\nabla\times(\boldsymbol{x}u),\qquad\boldsymbol{M}_{v}=\nabla\times(\boldsymbol{x}v); (37)

where we have used

×𝑴u\displaystyle\nabla\times\boldsymbol{M}_{u} =ω(ϵ0ϵrμ0μr)12𝑵u,×𝑴v=ω(ϵ0ϵrμ0μr)12𝑵v,\displaystyle=\omega(\epsilon_{0}\epsilon_{r}\mu_{0}\mu_{r})^{\frac{1}{2}}\boldsymbol{N}_{u},\qquad\nabla\times\boldsymbol{M}_{v}=\omega(\epsilon_{0}\epsilon_{r}\mu_{0}\mu_{r})^{\frac{1}{2}}\boldsymbol{N}_{v}, (38a)
×𝑵u\displaystyle\nabla\times\boldsymbol{N}_{u} =ω(ϵ0ϵrμ0μr)12𝑴u,×𝑵v=ω(ϵ0ϵrμ0μr)12𝑴v;\displaystyle=\omega(\epsilon_{0}\epsilon_{r}\mu_{0}\mu_{r})^{\frac{1}{2}}\boldsymbol{M}_{u},\qquad\nabla\times\boldsymbol{N}_{v}=\omega(\epsilon_{0}\epsilon_{r}\mu_{0}\mu_{r})^{\frac{1}{2}}\boldsymbol{M}_{v}; (38b)

and

×𝑬=iωμ0μr𝑯,×𝑯=iωϵ0ϵr𝑬.\displaystyle\nabla\times\boldsymbol{E}=\mathrm{i}\omega\mu_{0}\mu_{r}\boldsymbol{H},\qquad\nabla\times\boldsymbol{H}=-\mathrm{i}\omega\epsilon_{0}\epsilon_{r}\boldsymbol{E}. (39)

The components of 𝑴u\boldsymbol{M}_{u} and 𝑵u\boldsymbol{N}_{u} in spherical coordinates are:

Mur=0,\displaystyle M_{u_{r}}=0, Muθ=1rsinθ(ru)φ,Muφ=1r(ru)θ;\displaystyle\qquad M_{u_{\theta}}=\frac{1}{r\sin{\theta}}\frac{\partial(ru)}{\partial\varphi},\qquad M_{u_{\varphi}}=-\frac{1}{r}\frac{\partial(ru)}{\partial\theta}; (40a)
Nur=1k2(ru)r2+kru,\displaystyle N_{u_{r}}=\frac{1}{k}\frac{\partial^{2}(ru)}{\partial r^{2}}+kru, Nuθ=1kr2(ru)rθ,Nuφ=1krsinθ2(ru)rφ.\displaystyle\qquad N_{u_{\theta}}=\frac{1}{kr}\frac{\partial^{2}(ru)}{\partial r\partial\theta},\qquad N_{u_{\varphi}}=\frac{1}{kr\sin\theta}\frac{\partial^{2}(ru)}{\partial r\partial\varphi}. (40b)

The above formulation can also be used to get the components of 𝑴v\boldsymbol{M}_{v} and 𝑵v\boldsymbol{N}_{v} when the potential uu is replaced by the potential vv.

Both Debye potentials, uu and vv, satisfy the scalar Helmholtz equation 2ϕ+k2ϕ=0\nabla^{2}\phi+k^{2}\phi=0 which elementary solutions in the spherical coordinate system are of the form:

ϕ(l,n)=\displaystyle\phi_{(l,n)}= cos(lφ)Pnl(cosθ)zn(kr),\displaystyle\cos{(l\varphi)}P_{n}^{l}(\cos{\theta})z_{n}(kr), (41a)
ϕ(l,n)=\displaystyle\phi_{(l,n)}= sin(lφ)Pnl(cosθ)zn(kr).\displaystyle\sin{(l\varphi)}P_{n}^{l}(\cos{\theta})z_{n}(kr). (41b)

In Eq. (41), ll and nn are integers with nl0n\geq l\geq 0, Pnl(cosθ)P_{n}^{l}(\cos{\theta}) is the associated Legendre polynomial, and zn(kr)z_{n}(kr) is the spherical Bessel function. The following rules are applied to determine the choice of the function zn(kr)z_{n}(kr). In the bounded domain comprising the origin, zn(kr)jn(kr)z_{n}(kr)\equiv j_{n}(kr), the spherical Bessel function of the first kind, is used since jn(kr)j_{n}(kr) is finite at the origin. In the unbounded external domain, the spherical Hankel function zn(kr)hn(kr)=jn(kr)+iyn(kr)z_{n}(kr)\equiv h_{n}(kr)=j_{n}(kr)+\mathrm{i}y_{n}(kr) is used since ikhn(kr)inexp(ikr)/r\mathrm{i}kh_{n}(kr)\sim\mathrm{i}^{n}\exp{(\mathrm{i}kr)}/r, represents an outgoing wave, where yn(kr)y_{n}(kr) is the spherical Bessel function of the second kind.

It is worth noting that the potentials uu and vv correspond to cos(lφ)\cos(l\varphi) and sin(lφ)\sin(l\varphi) formulations in Eq. (41), respectively. Nevertheless, according to Eq. (35), the fields driven by a concentric dipole in a sphere do not depend on φ\varphi. As such, only terms with l=0l=0 in Eq. (41) exist, which means just one potential is needed. Let us use the potential uu here. Moreover, when comparing the components given in Eq. (40) and Eq. (35), we can see that only the terms with n=1n=1 in Eq. (41) are needed. As a result, to obtain the induced field inside the sphere and the radiation field outside the sphere, we let

uout\displaystyle u^{\mathrm{out}} =C(0,1)(1)h1(1)(k1r)cosθ,\displaystyle=C^{(1)}_{(0,1)}\,h_{1}^{(1)}(k_{1}r)\cos{\theta}, (42a)
uin\displaystyle u^{\mathrm{in}} =C(0,1)(2)j1(k2r)cosθ,\displaystyle=C^{(2)}_{(0,1)}\,j_{1}(k_{2}r)\cos{\theta}, (42b)

where

h1(1)(k1r)\displaystyle h_{1}^{(1)}(k_{1}r) =sin(k1r)(k1r)2cos(k1r)k1r+i[cos(k1r)(k1r)2sin(k1r)k1r],\displaystyle=\frac{\sin(k_{1}r)}{(k_{1}r)^{2}}-\frac{\cos(k_{1}r)}{k_{1}r}+\mathrm{i}\left[-\frac{\cos(k_{1}r)}{(k_{1}r)^{2}}-\frac{\sin(k_{1}r)}{k_{1}r}\right], (43a)
j1(k2r)\displaystyle j_{1}(k_{2}r) =sin(k2r)(k2r)2cos(k2r)k2r,\displaystyle=\frac{\sin(k_{2}r)}{(k_{2}r)^{2}}-\frac{\cos(k_{2}r)}{k_{2}r}, (43b)

and C(0,1)(1)C^{(1)}_{(0,1)} and C(0,1)(2)C^{(2)}_{(0,1)} are the coefficients to be determined by the boundary conditions.

Introducing Eq. (42) into Eq. (36) and using Eq. (40), we obtain

Erout\displaystyle E_{r}^{\mathrm{out}} =i[1k12(ruout)r2+k1ruout]=C(0,1)(1)2ik1rh1(1)(k1r)cosθ,\displaystyle=-\mathrm{i}\left[\frac{1}{k_{1}}\frac{\partial^{2}(ru^{\mathrm{out}})}{\partial r^{2}}+k_{1}ru^{\mathrm{out}}\right]=-C^{(1)}_{(0,1)}\frac{2\mathrm{i}}{k_{1}r}h_{1}^{(1)}(k_{1}r)\cos{\theta}, (44a)
Eθout\displaystyle E_{\theta}^{\mathrm{out}} =ik1r2(ruout)rθ=C(0,1)(1)ik1r[2h1(1)(k1r)k1rh2(2)(k1r)]sinθ,\displaystyle=-\frac{\mathrm{i}}{k_{1}r}\frac{\partial^{2}(ru^{\mathrm{out}})}{\partial r\partial\theta}=C^{(1)}_{(0,1)}\frac{\mathrm{i}}{k_{1}r}\left[2h_{1}^{(1)}(k_{1}r)-k_{1}rh_{2}^{(2)}(k_{1}r)\right]\sin{\theta}, (44b)
Eφout\displaystyle E_{\varphi}^{\mathrm{out}} =0,\displaystyle=0, (44c)
Erin\displaystyle E_{r}^{\mathrm{in}} =i[1k22(ruin)r2+k2ruin]=C(0,1)(2)2ik2rj1(k2r)cosθ,\displaystyle=-\mathrm{i}\left[\frac{1}{k_{2}}\frac{\partial^{2}(ru^{\mathrm{in}})}{\partial r^{2}}+k_{2}ru^{\mathrm{in}}\right]=-C^{(2)}_{(0,1)}\frac{2\mathrm{i}}{k_{2}r}j_{1}(k_{2}r)\cos{\theta}, (44d)
Eθin\displaystyle E_{\theta}^{\mathrm{in}} =ik1r2(ruin)rθ=C(0,1)(2)ik2r[2j1(k2r)k2rj2(k2r)]sinθ,\displaystyle=-\frac{\mathrm{i}}{k_{1}r}\frac{\partial^{2}(ru^{\mathrm{in}})}{\partial r\partial\theta}=C^{(2)}_{(0,1)}\frac{\mathrm{i}}{k_{2}r}\left[2j_{1}(k_{2}r)-k_{2}rj_{2}(k_{2}r)\right]\sin{\theta}, (44e)
Eφin\displaystyle E_{\varphi}^{\mathrm{in}} =0;\displaystyle=0; (44f)

and

Hrout\displaystyle H_{r}^{\mathrm{out}} =0,\displaystyle=0, (45a)
Hθout\displaystyle H_{\theta}^{\mathrm{out}} =0,\displaystyle=0, (45b)
Hφout\displaystyle H_{\varphi}^{\mathrm{out}} =ϵ0ϵ1μ0μ11r(ruout)θ=ϵ0ϵ1μ0μ1C(0,1)(1)h1(1)(k1r)sinθ,\displaystyle=\sqrt{\frac{\epsilon_{0}\epsilon_{1}}{\mu_{0}\mu_{1}}}\frac{1}{r}\frac{\partial(ru^{\mathrm{out}})}{\partial\theta}=-\sqrt{\frac{\epsilon_{0}\epsilon_{1}}{\mu_{0}\mu_{1}}}C^{(1)}_{(0,1)}h_{1}^{(1)}(k_{1}r)\sin{\theta}, (45c)
Hrin\displaystyle H_{r}^{\mathrm{in}} =0,\displaystyle=0, (45d)
Hθin\displaystyle H_{\theta}^{\mathrm{in}} =0,\displaystyle=0, (45e)
Hφin\displaystyle H_{\varphi}^{\mathrm{in}} =ϵ0ϵ2μ0μ21r(ruin)θ=ϵ0ϵ2μ0μ2C(0,1)(2)j1(k2r)sinθ.\displaystyle=\sqrt{\frac{\epsilon_{0}\epsilon_{2}}{\mu_{0}\mu_{2}}}\frac{1}{r}\frac{\partial(ru^{\mathrm{in}})}{\partial\theta}=-\sqrt{\frac{\epsilon_{0}\epsilon_{2}}{\mu_{0}\mu_{2}}}C^{(2)}_{(0,1)}j_{1}(k_{2}r)\sin{\theta}. (45f)

The boundary conditions across the sphere surface at r=ar=a are:

𝑬out\displaystyle\boldsymbol{E}^{\mathrm{out}}_{\parallel} =𝑬in+𝑬Ed,\displaystyle=\boldsymbol{E}^{\mathrm{in}}_{\parallel}+\boldsymbol{E}^{\mathrm{Ed}}_{\parallel}, (46a)
𝑯out\displaystyle\boldsymbol{H}^{\mathrm{out}}_{\parallel} =𝑯in+𝑯Ed,\displaystyle=\boldsymbol{H}^{\mathrm{in}}_{\parallel}+\boldsymbol{H}^{\mathrm{Ed}}_{\parallel}, (46b)

which means Eθout=Eθin+EθEdE_{\theta}^{\mathrm{out}}=E_{\theta}^{\mathrm{in}}+E_{\theta}^{\mathrm{Ed}} and Hφout=Hφin+HφEdH_{\varphi}^{\mathrm{out}}=H_{\varphi}^{\mathrm{in}}+H_{\varphi}^{\mathrm{Ed}}. As such, using the relative formulations in Eqs. (35), (44) and (45), we have

C(0,1)(1)ik1a[2h1(1)(k1a)k1ah2(2)(k1a)]C(0,1)(2)ik2a[2j1(k2a)k2aj2(k2a)]\displaystyle C^{(1)}_{(0,1)}\frac{\mathrm{i}}{k_{1}a}\left[2h_{1}^{(1)}(k_{1}a)-k_{1}ah_{2}^{(2)}(k_{1}a)\right]-C^{(2)}_{(0,1)}\frac{\mathrm{i}}{k_{2}a}\left[2j_{1}(k_{2}a)-k_{2}aj_{2}(k_{2}a)\right]
=\displaystyle= 14πϵ0ϵ2p(1a3ik2a2k22a)exp(ik2a),\displaystyle\frac{1}{4\pi\epsilon_{0}\epsilon_{2}}p\left(\frac{1}{a^{3}}-\frac{\mathrm{i}k_{2}}{a^{2}}-\frac{k_{2}^{2}}{a}\right)\exp{(\mathrm{i}k_{2}a)}, (47a)
ϵ0ϵ1μ0μ1C(0,1)(1)h1(1)(k1a)+ϵ0ϵ2μ0μ2C(0,1)(2)j1(k2a)\displaystyle-\sqrt{\frac{\epsilon_{0}\epsilon_{1}}{\mu_{0}\mu_{1}}}C^{(1)}_{(0,1)}h_{1}^{(1)}(k_{1}a)+\sqrt{\frac{\epsilon_{0}\epsilon_{2}}{\mu_{0}\mu_{2}}}C^{(2)}_{(0,1)}j_{1}(k_{2}a)
=\displaystyle= k234πϵ0ϵ2μ0μ2ωp(1ik2a21a)exp(ik2a).\displaystyle\frac{k_{2}^{3}}{4\pi\epsilon_{0}\epsilon_{2}\mu_{0}\mu_{2}\omega}p\left(\frac{1}{\mathrm{i}k_{2}a^{2}}-\frac{1}{a}\right)\exp{(\mathrm{i}k_{2}a)}. (47b)

Once C(0,1)(1)C^{(1)}_{(0,1)} and C(0,1)(2)C^{(2)}_{(0,1)} are obtained by solving Eq. (47), the electric and magnetic fields inside, outside and on the sphere surface can be obtained.

References

  • [1] J. Strutt, The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science 1871, 41, 271 107.
  • [2] H. C. van de Hulst, Light scattering by small particles, Dover Publications, New York, 1981.
  • [3] R. W. P. King, M. Owens, T. T. Wu, Lateral Electromagnetic Waves, Springer New York, 1992.
  • [4] L. Novotny, B. Hecht, Principles of Nano-Optics, Cambridge University Press, 2006.
  • [5] F. J. Rodriguez-Fortuno, G. Marino, P. Ginzburg, D. O'Connor, A. Martinez, G. A. Wurtz, A. V. Zayats, Science 2013, 340, 6130 328.
  • [6] B. D. Bartolo, J. Collins, L. Silvestri, editors, Nano-Optics: Principles Enabling Basic Research and Applications, Springer Netherlands, 2017.
  • [7] M. F. Picardi, A. Manjavacas, A. V. Zayats, F. J. Rodríguez-Fortuño, Physical Review B 2017, 95, 24.
  • [8] M. F. Picardi, A. V. Zayats, F. J. Rodríguez-Fortuño, Physical Review Letters 2018, 120, 11.
  • [9] M. F. Picardi, M. Neugebauer, J. S. Eismann, G. Leuchs, P. Banzer, F. J. Rodríguez-Fortuño, A. V. Zayats, Light: Science & Applications 2019, 8, 1.
  • [10] J. E. Vázquez-Lozano, A. Martínez, F. J. Rodríguez-Fortuño, Physical Review Applied 2019, 12, 2.
  • [11] O. Keller, Quantum Theory of Near-Field Electrodynamics, Springer Berlin Heidelberg, 2011.
  • [12] J. Enderlein, E. Toprak, P. R. Selvin, Optics Express 2006, 14, 18 8111.
  • [13] T. J. Antosiewicz, M. Käll, The Journal of Physical Chemistry C 2016, 120, 37 20692.
  • [14] I. Aharonovich, D. Englund, M. Toth, Nature Photonics 2016, 10, 10 631.
  • [15] J.-M. Jin, The finite element method in electromagnetics, John Wiley & Sons Inc, Hoboken. New Jersey, 3rd edition, 2014.
  • [16] S. Karimpour, M. Noori, Annalen der Physik 2020, 532, 5 2000026.
  • [17] K.-H. Kim, J.-K. An, W.-S. Rim, Annalen der Physik 2020, 532, 1 1900383.
  • [18] Z. He, J. H. Gu, W. E. I. Sha, R. S. Chen, Optics Express 2018, 26, 19 25037.
  • [19] B. Cockburn, G. E. Karniadakis, C.-W. Shu, editors, Discontinuous Galerkin Methods, Springer Berlin Heidelberg, 2000.
  • [20] K. M. Kalayeh, J. S. Graf, M. K. Gobbert, In COMSOL Conference 2015. Boston, MA, The United States, 2015 1–6.
  • [21] B. Gallinet, J. Butet, O. J. F. Martin, Laser & Photonics Reviews 2015, 9, 6 577.
  • [22] V. Myroshnychenko, E. Carbó-Argibay, I. Pastoriza-Santos, J. Pérez-Juste, L. M. Liz-Marzán, F. J. García de Abajo, Advanced Materials 2008, 20 4288.
  • [23] E. Klaseboer, Q. Sun, D. Y. C. Chan, IEEE Transactions on Antennas and Propagation 2017, 65, 2 972.
  • [24] Q. Sun, E. Klaseboer, D. Y. C. Chan, Phys. Rev. B 2017, 95, 4 045137.
  • [25] Q. Sun, E. Klaseboer, A. J. Yuffa, D. Y. C. Chan, Journal of the Optical Society of America A 2020, 37, 2 276.
  • [26] Q. Sun, E. Klaseboer, A. J. Yuffa, D. Y. C. Chan, Journal of the Optical Society of America A 2020, 37, 2 284.
  • [27] Q. Sun, K. Dholakia, A. D. Greentree, ACS Photonics 2021, 8, 4 1103.
  • [28] U. Hohenester, A. Trügler, Computer Physics Communications 2012, 183, 2 370.
  • [29] Q. Sun, E. Klaseboer, Including monopoles to a fully desingularized boundary element method for acoustics (submitted), 2021.
  • [30] Q. Sun, E. Klaseboer, B.-C. Khoo, D. Y. C. Chan, Royal Society Open Science 2015, 2, 1 140520.
  • [31] P. Bharadwaj, B. Deutsch, L. Novotny, Advances in Optics and Photonics 2009, 1, 3 438.
  • [32] H. Duan, A. I. Fernández-Domínguez, M. Bosman, S. A. Maier, J. K. W. Yang, Nano Letters 2012, 12 1683.
  • [33] C. Tserkezis, N. A. Mortensen, M. Wubs, Physical Review B 2017, 96, 8.
  • [34] J. Jackson, Classical electrodynamics, Wiley, New York, 1999.
  • [35] G. Mie, Annalen der Physik 1908, 330, 3 377.