License: CC BY 4.0
arXiv:2309.03435v3 [quant-ph] 09 Apr 2024

Spectral Theory for Non-linear Superconducting Microwave Systems: Extracting Relaxation Rates and Mode Hybridization

Dung N. Pham Department of Electrical and Computer Engineering, Princeton University, Princeton, NJ 08544, USA    Richard D. Li Department of Physics, Yale University, New Haven, CT 06511, USA    Hakan E. Türeci Department of Electrical and Computer Engineering, Princeton University, Princeton, NJ 08544, USA
(April 9, 2024)
Abstract

The accurate modeling of mode hybridization and calculation of radiative relaxation rates have been crucial to the design and optimization of superconducting quantum devices. In this work, we introduce a spectral theory for the electrohydrodynamics of superconductors that enables the extraction of the relaxation rates of excitations in a general three-dimensional distribution of superconducting bodies. Our approach addresses the long-standing problem of formulating a modal description of open systems that is both efficient and allows for second quantization of the radiative hybridized fields. This is achieved through the implementation of finite but transparent boundaries through which radiation can propagate into and out of the computational domain. The resulting spectral problem is defined within a coarse-grained formulation of the electrohydrodynamical equations that is suitable for the analysis of the non-equilibrium dynamics of multiscale superconducting quantum systems.

I Introduction

Engineered superconducting systems employed in analog quantum simulation [1, 2], quantum sensing [3, 4] and quantum computing [5, 6] are described by the electrohydrodynamic theory of superconducting systems (EHDS), first proposed without justification by R.P. Feynman in a special lecture he delivered in the 1960s [7]. Few attempts have been made to either justify [8, 9] or solve [10] this model’s equations of motion exactly, but simplified versions [11, 12, 13, 12] have been considered via various modifications to minimal coupling. This low-energy theory can be credited as the underpinning of Josephson phenomena, flux quantization, and, generally, the physics of vortices. The second-quantized version of it constitutes the basis of the formulation known as circuit quantum electrodynamics (cQED) [14], which has been the workhorse behind the current understanding of the physics governing superconducting quantum computers. As systems grow increasingly sophisticated and their electromagnetic environments become more complex, efficient computational strategies in both classical and quantum regimes are much needed to produce accurate reduced models containing the degrees of freedom of interest. This has given rise to an active research area at the intersection of computational electromagnetism and quantum electrodynamics of superconducting devices [15, 16, 17, 18, 19, 20, 21], to analyze and model the physics of these circuits.

An indispensable component in the electrodynamic modeling of superconducting circuits is the extraction of relaxation rates. Purcell modification of radiative lifetimes of qubits is today an essential mechanism for the protection of qubits and plays an important role in the electromagnetic design of the entire processor [22, 23, 24, 25]. Additionally, in applications that require strong RF or microwave excitations of individual oscillators, new dissipative processes and inter-level transitions can be activated [26, 27, 28, 29, 30, 31, 32, 33, 34]. These processes, in turn, depend strongly on the spectral characteristics of the electromagnetic environment in which the non-linear elements are embedded [30]. Resource-intensive numerical simulations are needed to extract such information for complex quantum-electrodynamic systems. Such simulations are not only computationally demanding but also face conceptual difficulties arising from the infinite degrees of freedom inherent in quantum electrodynamics (QED) [35]. These conceptual issues [22, 36, 18, 37] can be effectively addressed by the introduction of a rigorous spectral theory [38]. This approach allows the accurate quantification of limitations brought about by noise [39, 40, 41], dissipation [42, 43] and undesired interactions [44, 45] within the system and its electromagnetic environment.

In the present work, we introduce a spectral theory of EHDS from which the relaxation rates of general three-dimensional superconducting circuits can be obtained. This spectral theory is adapted to a coarse-grained description of the EHDS equations, known as DEC-QED, previously derived and analyzed in Ref. [10]. DEC-QED provides the following advantages: (1) through its structure-preserving geometric discretization procedure, known as discrete exterior calculus (DEC), this formulation enables stable long-time simulations. (2) By virtue of the fundamental fields being hybridized gauge-invariant fields rather than the standard electromagnetic potentials, different materials can be handled in a uniform fashion. (3) At a superconducting-normal-superconducting (SNS) junction, the hybridized field is identical to the gauge-invariant Josephson phase across the junction. (4) For transmission lines and lumped-element circuits and in the limit of a perfect superconductor (λLsubscript𝜆𝐿\lambda_{L}italic_λ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT = 0), DEC-QED equations reduce to the standard cQED equations.

The difficulty with developing a spectral theory stems from the desire to solve the EHDS equations, a set of non-linear PDEs describing the evolution of the order parameter of a charged fluid coupled to Maxwell’s equations, in a finite spatial domain. Ideally, one would like to keep the computational domain as small as possible, just as large as the superconducting system itself. However, this is not possible, because we also need to keep the volume as large as possible to allow for radiative relaxation of the excitations in the superconductor. The optimal solution to avoid this trade-off is to use a modal description for a finite but open domain subject to transparent boundary conditions for electromagnetic (EM) radiation. This allows us to keep the computational domain small enough to be computationally feasible, while still allowing for the radiative relaxation of the excitations. While boundary conditions, such as the perfectly matched layer (PML) [46, 47] or absorbing boundary conditions [48, 49] are known and very useful in classical EM problems, it is also desirable to implement a transparent boundary so that solving the corresponding spectral problem would provide a modal decomposition which can serve as the basis for second quantizing the EHDS equations. This, however, has been a difficult problem that remains unaddressed since the full formulation of quantum electrodynamics [50] because such boundary conditions result in non-Hermitian modes [51, 52, 53]. Recently, it was shown that by using the Heisenberg equations of motion for quantum field operators, this issue can be circumvented in the context of one-dimensional transmission line systems through the use of singular function expansion and a suitable spectral problem [38]. Here we generalize the statement of this spectral problem to the solution of the non-linear EHDS equations (the original spectral problem has been defined [38] in the context of cQED which assumes fields do not penetrate the superconductors), and to general three dimensional domains. We do not address the issue of quantization but focus on the spectral problem that arises through the linearization of the EHDS equations, the resulting non-Hermitian modes, and the calculation of dissipation rates for all modes. The dissipation rates for “qubit-like” modes are precisely their Purcell radiative lifetimes [54], while for the more spatially extended modes, they represent their losses due to their hybridization with radiative channels in the system [55, 56, 57].

With reference to the prior work on modeling of superconducting devices, our approach is within the class of full-wave methods [58, 59]. With respect to existing full-wave methods for superconducting systems, DEC-QED has two important distinctions: 1. Calculation of Radiative Losses: The boundary conditions implemented in existing approaches to capture radiative losses often implement absorbing boundary conditions. Most frequently, radiative losses at a 3D cavity’s ports are modeled by adapting from the lumped-element models and terminating the ports with 50Ω50Ω50\,\Omega50 roman_Ω matched resistors [16] or by covering the holes on the cavity walls with resistive disks [20]. Although this approach is effective in obtaining the frequencies of the eigenemodes, it is known to fail to accurately capture the spatial structure of the open modes. More importantly, no known method exists to second quantize using such modes, and therefore such EM full wave simulations can not rigorously be used to synthesize a quantum mechanical Hamiltonian or Liouvillian. Often the quantum Hamiltonian is derived from simulations ignoring radiative losses, and the relaxation rates are introduced through a perturbative approach. In contrast, DEC-QED builds on prior work in classical electromagnetic systems for the implementation of correct open (transparent) boundary conditions [60, 61], which recently has been shown to serve as a mathematically consistent basis for second quantization of the modes of a transmission line cavity [38]. 2. Time-dependent Evolution: DEC-QED aims at numerically solving the set of non-linear PDEs describing the electrohydrodynamics of the superconducting condensate coupled to fields described by Maxwell equations. This allows DEC-QED to describe the evolution of gauge-invariant hybridized fields living throughout the free space and the superconductors, forgoing issues that arise in defining boundary conditions at superconductor/vacuum boundaries whose accuracy is not controlled. Compared to existing full-wave methods, this is a unique feature of our approach. The linearized spectral solutions of DEC-QED then serves as a basis for expanding this non-linear dynamics. In the current work, we do not discuss the non-linear aspects of evolution, which is discussed at length in Ref.  [10].

The main results presented in this paper are organized as follows: in Section II, we discuss the electrohydrodynamic model used to describe the dynamics of the superconducting condensate coupled to the EM environment and derive the spectral problem to be solved. Then in Section III we briefly discuss the formulation of DEC-QED used in the rest of the article for coarse-grained calculations. Next, we demonstrate the convergence in the calculations for the spectral profile of a sample superconducting cavity. Both simplicial and cubical meshing strategies are benchmarked in Section IV.1. In Section IV.2 we then show how mode hybridization due to accidental degeneracies can be encountered in the calculations of the modes for a symmetric cavity and that our method can distinguish the degenerate modes through the inclusion of a small perturbation to the cavity shape. Next, in Section IV.3 we demonstrate the coarse-grained calculations of hybridized modes for multiscale systems containing multiple components. Finally, we present two main approaches for implementing open boundary conditions based on (1) Green’s boundary integrals in Sections V.1-V.2 and (2) vector spherical harmonic expansions in Section V.3, along with numerical examples, to demonstrate the accuracy and effectiveness of these methods. We have also made the implementation of the formulation discussed here available as part of an open-source DEC-QED repository [62].

II Electrohydrodynamic formulation of superconducting materials

Consider a multiply connected region composed of different dielectric and superconducting materials. Well below the critical temperature, where most superconducting devices operate, the superconducting material can be modeled as a charged condensate [8, 9] trapped by a background of positive charge (ρsrcsubscript𝜌src\rho_{\text{src}}italic_ρ start_POSTSUBSCRIPT src end_POSTSUBSCRIPT). The dynamics of the condensate are wholly determined through minimal coupling to the dynamical electromagnetic potentials (𝑨𝑨\bm{A}bold_italic_A, V𝑉Vitalic_V) and the Coulomb attraction to the static positive background (U𝑈Uitalic_U) defining the superconductor. We refer to the resulting equations as the ElectroHydroDynamics of Superconductors (EHDS) [7]

iΨ(𝐫,t)t=[12m(iq𝐀)2+qV(𝐫,t)+U(𝐫)]Ψ(𝐫,t)𝑖Planck-constant-over-2-piΨ𝐫𝑡𝑡delimited-[]12𝑚superscript𝑖Planck-constant-over-2-pi𝑞𝐀2𝑞𝑉𝐫𝑡𝑈𝐫Ψ𝐫𝑡i\hbar\frac{\partial\Psi({\bf r},t)}{\partial t}=\bigg{[}\frac{1}{2m}\Big{(}\!% \!-i\hbar{\bf\nabla}-q{\bf A}\Big{)}^{2}+qV({\bf r},t)+U({\bf r})\bigg{]}\Psi% \left({\bf r},t\right)italic_i roman_ℏ divide start_ARG ∂ roman_Ψ ( bold_r , italic_t ) end_ARG start_ARG ∂ italic_t end_ARG = [ divide start_ARG 1 end_ARG start_ARG 2 italic_m end_ARG ( - italic_i roman_ℏ ∇ - italic_q bold_A ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_q italic_V ( bold_r , italic_t ) + italic_U ( bold_r ) ] roman_Ψ ( bold_r , italic_t ) (1)

and Maxwell’s equations

××𝐀+μ0ϵ𝐀¨𝐀subscript𝜇0italic-ϵ¨𝐀\displaystyle{\bf\nabla}\times{\bf\nabla}\times{\bf A}+\mu_{0}\epsilon\ddot{{% \bf A}}∇ × ∇ × bold_A + italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ϵ over¨ start_ARG bold_A end_ARG =μ0(𝐉s+𝐉src)μ0ϵV˙,absentsubscript𝜇0subscript𝐉𝑠subscript𝐉srcsubscript𝜇0italic-ϵ˙𝑉\displaystyle=\mu_{0}({\bf J}_{s}+{\bf J}_{\text{src}})-\mu_{0}\epsilon\nabla% \dot{V},= italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_J start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + bold_J start_POSTSUBSCRIPT src end_POSTSUBSCRIPT ) - italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ϵ ∇ over˙ start_ARG italic_V end_ARG , (2)
2V+t(𝐀)superscript2𝑉𝑡𝐀\displaystyle\nabla^{2}V+\frac{\partial}{\partial t}({\mathbf{\nabla}}\cdot{% \mathbf{A}})∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_V + divide start_ARG ∂ end_ARG start_ARG ∂ italic_t end_ARG ( ∇ ⋅ bold_A ) =qϵ(ρ+ρsrc).absent𝑞italic-ϵ𝜌subscript𝜌src\displaystyle=-\frac{q}{\epsilon}(\rho+\rho_{\text{src}}).= - divide start_ARG italic_q end_ARG start_ARG italic_ϵ end_ARG ( italic_ρ + italic_ρ start_POSTSUBSCRIPT src end_POSTSUBSCRIPT ) . (3)

where ρ𝜌\rhoitalic_ρ and 𝐉ssubscript𝐉𝑠{\bf J}_{s}bold_J start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT are the condensate density and the supercurrent, respectively, and ρsrcsubscript𝜌src\rho_{\text{src}}italic_ρ start_POSTSUBSCRIPT src end_POSTSUBSCRIPT and 𝐉srcsubscript𝐉src{\bf J}_{\text{src}}bold_J start_POSTSUBSCRIPT src end_POSTSUBSCRIPT are the external charge and current sources. Here, q=2e𝑞2𝑒q=2eitalic_q = 2 italic_e and m=2me𝑚2subscript𝑚𝑒m=2m_{e}italic_m = 2 italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT are the charge and mass of a cooper pair, respectively, which are twice the charge e𝑒eitalic_e and mass mesubscript𝑚𝑒m_{e}italic_m start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT of an electron. Generally, we include the positive background defining the superconducting regions in ρsrcsubscript𝜌src\rho_{\text{src}}italic_ρ start_POSTSUBSCRIPT src end_POSTSUBSCRIPT. Dielectric regions are defined by ϵ(𝒓)=ϵ~(𝒓)ϵ0italic-ϵ𝒓~italic-ϵ𝒓subscriptitalic-ϵ0\epsilon(\bm{r})=\tilde{\epsilon}(\bm{r})\epsilon_{0}italic_ϵ ( bold_italic_r ) = over~ start_ARG italic_ϵ end_ARG ( bold_italic_r ) italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, where ϵ~~italic-ϵ\tilde{\epsilon}over~ start_ARG italic_ϵ end_ARG is the step-wise constant relative permittivity function that is unity where dielectric is not present and is greater than one otherwise. Using the Madelung representation for the condensate wavefunction, Ψ(𝐫,t)=ρ(𝐫,t)eiθ(𝐫,t)Ψ𝐫𝑡𝜌𝐫𝑡superscript𝑒𝑖𝜃𝐫𝑡\Psi({\bf r},t)=\sqrt{\rho({\bf r},t)}e^{i\theta({\bf r},t)}roman_Ψ ( bold_r , italic_t ) = square-root start_ARG italic_ρ ( bold_r , italic_t ) end_ARG italic_e start_POSTSUPERSCRIPT italic_i italic_θ ( bold_r , italic_t ) end_POSTSUPERSCRIPT, and introducing the gauge-invariant hybridized field 𝓐=𝐀qθ𝓐𝐀Planck-constant-over-2-pi𝑞𝜃\bm{\mathcal{A}}={\bf A}-\frac{\hbar}{q}\nabla\thetabold_caligraphic_A = bold_A - divide start_ARG roman_ℏ end_ARG start_ARG italic_q end_ARG ∇ italic_θ, Eqs. (1)-(3) can be written in the form [10]

××𝓐+μ0𝓐subscript𝜇0\displaystyle{\bf\nabla}\!\times\!{\bf\nabla}\!\times\!\bm{\mathcal{A}}+\mu_{0}∇ × ∇ × bold_caligraphic_A + italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ϵ2𝓐t2+μ0q2mρ𝓐μ0ϵq2mt|𝓐|2italic-ϵsuperscript2𝓐superscript𝑡2subscript𝜇0superscript𝑞2𝑚𝜌𝓐subscript𝜇0italic-ϵ𝑞2𝑚𝑡superscript𝓐2\displaystyle\epsilon\frac{\partial^{2}\bm{\mathcal{A}}}{\partial t^{2}}+\frac% {\mu_{0}q^{2}}{m}\rho\bm{\mathcal{A}}-\frac{\mu_{0}\epsilon q}{2m}\frac{% \partial}{\partial t}\nabla\big{|}\bm{\mathcal{A}}\big{|}^{2}italic_ϵ divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_caligraphic_A end_ARG start_ARG ∂ italic_t start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_m end_ARG italic_ρ bold_caligraphic_A - divide start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ϵ italic_q end_ARG start_ARG 2 italic_m end_ARG divide start_ARG ∂ end_ARG start_ARG ∂ italic_t end_ARG ∇ | bold_caligraphic_A | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
+μ0ϵ22mqt[2(ρ)ρ]=μ0𝐉src,subscript𝜇0italic-ϵsuperscriptPlanck-constant-over-2-pi22𝑚𝑞𝑡superscript2𝜌𝜌subscript𝜇0subscript𝐉𝑠𝑟𝑐\displaystyle+\frac{\mu_{0}\epsilon\hbar^{2}}{2mq}\frac{\partial}{\partial t}% \nabla\bigg{[}\frac{\nabla^{2}(\sqrt{\rho})}{\sqrt{\rho}}\bigg{]}=\mu_{0}{\bf J% }_{src},+ divide start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ϵ roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m italic_q end_ARG divide start_ARG ∂ end_ARG start_ARG ∂ italic_t end_ARG ∇ [ divide start_ARG ∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( square-root start_ARG italic_ρ end_ARG ) end_ARG start_ARG square-root start_ARG italic_ρ end_ARG end_ARG ] = italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT bold_J start_POSTSUBSCRIPT italic_s italic_r italic_c end_POSTSUBSCRIPT , (4)

and

ρt=[qmρ𝓐𝐉srcq]ρsrct.𝜌𝑡delimited-[]𝑞𝑚𝜌𝓐subscript𝐉𝑠𝑟𝑐𝑞subscript𝜌𝑠𝑟𝑐𝑡\frac{\partial\rho}{\partial t}={\bf\nabla}\cdot\Bigg{[}\frac{q}{m}\rho\bm{% \mathcal{A}}-\frac{{\bf J}_{src}}{q}\Bigg{]}-\frac{\partial\rho_{src}}{% \partial t}.divide start_ARG ∂ italic_ρ end_ARG start_ARG ∂ italic_t end_ARG = ∇ ⋅ [ divide start_ARG italic_q end_ARG start_ARG italic_m end_ARG italic_ρ bold_caligraphic_A - divide start_ARG bold_J start_POSTSUBSCRIPT italic_s italic_r italic_c end_POSTSUBSCRIPT end_ARG start_ARG italic_q end_ARG ] - divide start_ARG ∂ italic_ρ start_POSTSUBSCRIPT italic_s italic_r italic_c end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_t end_ARG . (5)

These equations were derived and the resulting real-time dynamics under specific conditions were analyzed in Ref. [10]. Here, we are interested in the spectral problem associated with these non-linear equations. Linearization can be done by splitting the condensate density ρ𝜌\rhoitalic_ρ into the mean value ρ0subscript𝜌0\rho_{0}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT that exactly balances the positively charged ionic background and the fluctuation δρ𝛿𝜌\delta\rhoitalic_δ italic_ρ arising from interactions with the external EM field. The linear sector of Eq. (II) that corresponds to the transverse excitations then reproduces London theory [63] and is sourced by the fluctuation in the source current. A derivation of the linearization is provided in Appendix A. The resulting inhomogeneous source-field equations can be solved through the spectral problem of a vector Helmholtz equation for the gauge-invariant hybridized field 𝓐𝓐\bm{{\mathcal{A}}}bold_caligraphic_A:

××𝓐+(1λL2(𝐫)n2(𝐫)k2)𝓐=0,𝓐1superscriptsubscript𝜆𝐿2𝐫superscript𝑛2𝐫superscript𝑘2𝓐0\displaystyle{\mathbf{\nabla}}\times{\mathbf{\nabla}}\times\bm{{\mathcal{A}}}+% \bigg{(}\frac{1}{\lambda_{L}^{2}({\mathbf{r}})}-n^{2}({\mathbf{r}})k^{2}\bigg{% )}\bm{{\mathcal{A}}}=0,∇ × ∇ × bold_caligraphic_A + ( divide start_ARG 1 end_ARG start_ARG italic_λ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_r ) end_ARG - italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_r ) italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) bold_caligraphic_A = 0 , (6)

where the distribution of superconducting and dielectric materials is given by the local London penetration depth function λL(𝐫)=mμ0q2ρ0(𝐫)subscript𝜆𝐿𝐫𝑚subscript𝜇0superscript𝑞2subscript𝜌0𝐫\lambda_{L}({\mathbf{r}})=\sqrt{\frac{m}{\mu_{0}q^{2}\rho_{0}({\mathbf{r}})}}italic_λ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( bold_r ) = square-root start_ARG divide start_ARG italic_m end_ARG start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_r ) end_ARG end_ARG and the refractive index function n(𝐫)=ϵ~(𝐫)𝑛𝐫~italic-ϵ𝐫n({\mathbf{r}})=\sqrt{\tilde{\epsilon}({\mathbf{r}})}italic_n ( bold_r ) = square-root start_ARG over~ start_ARG italic_ϵ end_ARG ( bold_r ) end_ARG, respectively, and k𝑘kitalic_k is the wavenumber.

III DEC-QED formulation

DEC-QED provides a spatially coarse-grained description of a physical system governed by nonlinear PDEs. The fundamental field variables are small integrals of the original microscopic continuous fields over finite spatial intervals. This results in a discretized model that is computationally efficient and still accurate within the resolution of a given measurement apparatus. In this section, we provide a minimal introduction to the geometric constructions in DEC that are needed for computing electromagnetic modes in systems that may contain an arbitrary distribution of superconducting and dielectric materials, all within a finite computational domain. For a more comprehensive discussion of the formulation, we refer to Ref. [10].

In DEC, the discretization of PDEs requires a dual-mesh construction, within which the d𝑑ditalic_d-dimensional computational space is discretized by a primal mesh M𝑀Mitalic_M that conforms to the boundaries of the enclosed physical domain and the interfaces between materials. The fundamental building blocks of the primal mesh can be simplices (i.e. triangles in 2D and tetrahedra in 3D) or cubical elements. The vertices of the dual mesh Msuperscript𝑀M^{\dagger}italic_M start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT are then circumcenters of the primal dlimit-from𝑑d-italic_d -cells, and the edges of Msuperscript𝑀M^{\dagger}italic_M start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT are generated by connecting the neighboring circumcenters. Throughout this paper, we will use \dagger to denote a dual quantity. For elements strictly inside the computational space, the mappings from vertices (v𝑣vitalic_v), edges (e𝑒eitalic_e), faces (f𝑓fitalic_f), and cells (c𝑐citalic_c) in M𝑀Mitalic_M to cells (vsuperscript𝑣v^{\dagger}italic_v start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT), faces (esuperscript𝑒e^{\dagger}italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT), edges (fsuperscript𝑓f^{\dagger}italic_f start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT), vertices (csuperscript𝑐c^{\dagger}italic_c start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT) respectively in Msuperscript𝑀M^{\dagger}italic_M start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT are one-to-one. This bijective correspondence is not applicable at the computational boundary, where the dual elements are truncated, which leads to auxiliary dual nodes lying on the boundary of M𝑀Mitalic_M (See Figs. (1a) and (1b)). During numerical calculations, this truncation of the dual mesh is taken care of by the appropriate application of boundary conditions.

Refer to caption
Figure 1: (a) An example primal mesh. The highlighted edges at the lower-left corner form the boundary of an auxiliary dual volume. (b) Close-up view of a node v𝑣vitalic_v that lies on the boundary of the primal mesh. The neighboring dual nodes are labeled in dark green, while the auxiliary boundary dual nodes are labeled in bright green. The shaded volume is the truncated volume vsuperscript𝑣v^{\dagger}italic_v start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT dual to the node v𝑣vitalic_v. (c) 2D example of a vertex that lies at the interface among multiple material regions. The material properties assigned to this node is the weighted average of the values in the surrounding regions. (d) 3D example of an edge lying at a material interface composed of multiple cells cisubscript𝑐𝑖{c_{i}}italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The darkened triangle is the intersection of the dual face esuperscript𝑒e^{\dagger}italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT with the cell cisubscript𝑐𝑖c_{i}italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, and the shaded volume is the portion of the support volume of e𝑒eitalic_e that lies inside cisubscript𝑐𝑖c_{i}italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT.

In this DEC framework, scalar fields live on either primal vertices v𝑣vitalic_v or their dual volume vsuperscript𝑣v^{\dagger}italic_v start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT, while vectorial quantities are projected onto the discrete primal edges e𝑒eitalic_e or assigned to the dual faces esuperscript𝑒e^{\dagger}italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT. For example, given a vector field 𝓐𝓐\bm{{\mathcal{A}}}bold_caligraphic_A we construct the coarse-grained edge field

Φ(e)=e𝐝𝓐,Φ𝑒subscript𝑒differential-d𝓐\Phi(e)=\int_{e}{\mathbf{d}\ell}\cdot\bm{{\mathcal{A}}},roman_Φ ( italic_e ) = ∫ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT bold_d roman_ℓ ⋅ bold_caligraphic_A , (7)

where the integral is done along the primal edge e𝑒eitalic_e, and given a scalar function ρ𝜌\rhoitalic_ρ we define the coarse-grained variable

Q(v)=v𝑑Vρ,𝑄superscript𝑣subscriptsuperscript𝑣differential-d𝑉𝜌Q(v^{\dagger})=\int_{v^{\dagger}}dV\rho,italic_Q ( italic_v start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) = ∫ start_POSTSUBSCRIPT italic_v start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_d italic_V italic_ρ , (8)

where the integral is done over the dual volume vsuperscript𝑣v^{\dagger}italic_v start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT. By solving the equations governing these coarse-grained fields, we can probe the properties of the system.

To properly account for the distribution of different materials within the computational domain, material properties such as dielectric function n2superscript𝑛2n^{2}italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT or 1/λL21superscriptsubscript𝜆𝐿21/\lambda_{L}^{2}1 / italic_λ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (which is proportional to the bulk condensate density) can be assigned to objects living on the dual mesh. For example, for every edge e𝑒eitalic_e a value for 1/λL21superscriptsubscript𝜆𝐿21/\lambda_{L}^{2}1 / italic_λ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is assigned to its dual esuperscript𝑒e^{\dagger}italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT. This procedure is particularly convenient if e𝑒eitalic_e lies at the interface between multiple materials. The dual esuperscript𝑒e^{\dagger}italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT in such cases would intersect with all the different material domains that share this edge, and the effective value 1/λL2¯(e)¯1superscriptsubscript𝜆𝐿2superscript𝑒\overline{1/\lambda_{L}^{2}}(e^{\dagger})over¯ start_ARG 1 / italic_λ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) there will be the weighted average of the values in the surrounding domains (See Fig. 1d). In a three-dimensional setting, this is formally defined as followed: let {ci}subscript𝑐𝑖\{c_{i}\}{ italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT } be the list of all the cells that share an edge e𝑒eitalic_e, then

(1λL2)¯(e)=ci|eci|ΔA(e)1λL2(ci),¯1superscriptsubscript𝜆𝐿2superscript𝑒subscriptsubscript𝑐𝑖superscript𝑒subscript𝑐𝑖Δ𝐴superscript𝑒1superscriptsubscript𝜆𝐿2subscript𝑐𝑖\overline{\bigg{(}\frac{1}{\lambda_{L}^{2}}\bigg{)}}(e^{\dagger})=\sum_{c_{i}}% \frac{|e^{\dagger}\cap c_{i}|}{\Delta A(e^{\dagger})}\frac{1}{\lambda_{L}^{2}(% c_{i})},over¯ start_ARG ( divide start_ARG 1 end_ARG start_ARG italic_λ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) end_ARG ( italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) = ∑ start_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT divide start_ARG | italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ∩ italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | end_ARG start_ARG roman_Δ italic_A ( italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) end_ARG divide start_ARG 1 end_ARG start_ARG italic_λ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) end_ARG , (9)

where |eci|superscript𝑒subscript𝑐𝑖|e^{\dagger}\cap c_{i}|| italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ∩ italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | in the area of the intersection between esuperscript𝑒e^{\dagger}italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT and cisubscript𝑐𝑖c_{i}italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, and ΔA(e)Δ𝐴superscript𝑒\Delta A(e^{\dagger})roman_Δ italic_A ( italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) is the area of esuperscript𝑒e^{\dagger}italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT.

This DEC formulation of the electrohydrodynamics of coarse-grained quantities offers several benefits: (1) it provides a natural framework for describing gauge-invariant fields that are consistent with the standard flux-based description of Josephson dynamics. (2) Since the gauge-invariant fields (𝓐,ρ)𝓐𝜌(\bm{{\mathcal{A}}},\rho)( bold_caligraphic_A , italic_ρ ) permeate all of space, the issues related to boundary conditions at material interfaces are simplified by introducing effective material properties. (3) Because the Josephson phase is, by definition, a gauge-invariant and coarse-grained generalized flux variable, our approach is a rigorous generalization of the coarse-grained formulation beyond the junction and apply to the entire discretized 3D(2D) computational space. It is therefore different from other fullwave approaches because here the fundamental variables of interest are the coarse-grained fields that hybridize light and matter. The quantization of this electromagnetic theory therefore follows a similar procedure as is typically done for 1D transmission line cQED, but now with fluxes living on the edges and charges living on the nodes of the fully discretized 3D(2D) system. This approach, in which all of the classical and quantum analysis is done at the EM field level, is distinct from circuit diagram-based methods such as blackbox quantization [15, 16] or energy-participation ratio [20]. We sacrifice the simplicity brought about by the reduction to the circuit picture to obtain accurate mode structure and the full transparency in the ensuing quantization.

That being said, before the theory can be quantization-ready, a rigorous spectral theory for the coarse-grained quantities that can correctly account for radiative losses and mode hybridization, both in the eigenfrequencies and in the spatial profile of the modes (including and particularly on the open boundaries), is needed and is the focus of this paper. In the following sections, we demonstrate with specific examples how this is achieved in our approach.

IV Modes of closed superconducting systems

IV.1 Performance benchmarks for simplicial- and cubical-DEC

Refer to caption
Figure 2: (a) Convergence in the eigenvalues E=k2𝐸superscript𝑘2E=k^{2}italic_E = italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT of the vector Helmholtz equation applied to a perfect superconducting cavity as a function of the number of edges Nesubscript𝑁𝑒N_{e}italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT used in the discretization. The solid lines correspond to results obtained using simplicial meshing, while the dashed lines are results obtained using cubical meshing. The relative error is calculated with respect to the exact analytical solution. (b) The relative error in the edge fields of the eigenmodes with respect to analytical solutions. The errors are averaged over all edges and plotted as a function of Nesubscript𝑁𝑒N_{e}italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT. The cavity dimensions are Lx=1 cm,Ly=1.5 cmformulae-sequencesubscript𝐿𝑥1 cmsubscript𝐿𝑦1.5 cmL_{x}=1\text{\,cm},L_{y}=1.5\text{\,cm}italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = 1 cm , italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = 1.5 cm, and Lz=2 cmsubscript𝐿𝑧2 cmL_{z}=2\text{\,cm}italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 2 cm.

We are interested in solving Eq. (6) for the electromagnetic modes of systems composed of possibly spatially disjoint superconducting structures. To numerically compute the modes of such systems, we derive the DEC equations corresponding to Eq. (6)

e0(e)subscriptsubscript𝑒0superscript𝑒\displaystyle\sum_{e_{0}\in\partial(e^{\dagger})}∑ start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ ∂ ( italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT e1(e0)Δ(e0)ΔA(e0)Φ(e1)subscriptsubscript𝑒1superscriptsubscript𝑒0Δsubscript𝑒0Δ𝐴superscriptsubscript𝑒0Φsubscript𝑒1\displaystyle\sum_{e_{1}\in\partial(e_{0}^{\dagger})}\frac{\Delta\ell(e_{0})}{% \Delta A(e_{0}^{\dagger})}\Phi(e_{1})∑ start_POSTSUBSCRIPT italic_e start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∈ ∂ ( italic_e start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) end_POSTSUBSCRIPT divide start_ARG roman_Δ roman_ℓ ( italic_e start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_ARG start_ARG roman_Δ italic_A ( italic_e start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) end_ARG roman_Φ ( italic_e start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) (10)
+(1λL2¯(e)n2¯(e)k2)ΔA(e)Δ(e)Φ(e)=0,¯1superscriptsubscript𝜆𝐿2superscript𝑒¯superscript𝑛2superscript𝑒superscript𝑘2Δ𝐴superscript𝑒Δ𝑒Φ𝑒0\displaystyle+\bigg{(}\overline{\frac{1}{\lambda_{L}^{2}}}(e^{\dagger})-% \overline{n^{2}}(e^{\dagger})k^{2}\!\bigg{)}\frac{\Delta A(e^{\dagger})}{% \Delta\ell(e)}\Phi(e)=0,+ ( over¯ start_ARG divide start_ARG 1 end_ARG start_ARG italic_λ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG ( italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) - over¯ start_ARG italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) divide start_ARG roman_Δ italic_A ( italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) end_ARG start_ARG roman_Δ roman_ℓ ( italic_e ) end_ARG roman_Φ ( italic_e ) = 0 ,

where (e)superscript𝑒\partial(e^{\dagger})∂ ( italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ) is the boundary of the dual face esuperscript𝑒e^{\dagger}italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT, and {1/λL2¯,n2¯}¯1subscriptsuperscript𝜆2𝐿¯superscript𝑛2\{\overline{1/\lambda^{2}_{L}},\overline{n^{2}}\}{ over¯ start_ARG 1 / italic_λ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT end_ARG , over¯ start_ARG italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG } are defined as in Eq. (9). This form is universal and is independent of the mesh elements used.

First, we investigate the numerical convergence of DEC equations using two types of elements: simplicial and cubical. Simplicial elements are generally the preferred choice for a meshing scheme that conforms to arbitrary superconducting domains. To analyze the convergence of simplicial meshing, we consider shapes of superconducting cavities for which analytical solutions are available. A good choice is a rectangular cavity, which conforms well to cubical elements. A comparison of the convergence of error of simplicial and cubical meshing would illustrate the efficacy of simplicial meshing.

Consider a three-dimensional rectangular cavity with dimensions Lx=1 cm,Ly=1.5 cmformulae-sequencesubscript𝐿𝑥1 cmsubscript𝐿𝑦1.5 cmL_{x}=1\text{\,cm},L_{y}=1.5\text{\,cm}italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = 1 cm , italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = 1.5 cm, and Lz=2 cmsubscript𝐿𝑧2 cmL_{z}=2\text{\,cm}italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 2 cm, where Lx,Ly,subscript𝐿𝑥subscript𝐿𝑦L_{x},L_{y},italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , and Lzsubscript𝐿𝑧L_{z}italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT are the sizes of the cavity along x,y,𝑥𝑦x,y,italic_x , italic_y , and z𝑧zitalic_z respectively. To allow for direct comparisons with analytical solutions, we first assume the cavity walls are perfect superconductors (λL0)subscript𝜆𝐿0(\lambda_{L}\rightarrow 0)( italic_λ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT → 0 ), so that the tangential component of 𝓐𝓐\bm{{\mathcal{A}}}bold_caligraphic_A vanishes at the boundary. This is equivalent to the boundary condition

Φ(et)=0,Φsubscript𝑒𝑡0\Phi(e_{t})=0,roman_Φ ( italic_e start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) = 0 , (11)

where etsubscript𝑒𝑡e_{t}italic_e start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT is an edge lying tangentially on the cavity boundary. The eigenmodes of the cavity are computed with two choices of meshing: tetrahedral and cubical. The convergence with respect to analytical solutions is shown in Fig. 2. As shown in Fig. 2a, the error in the eigenvalues E=k2𝐸superscript𝑘2E=k^{2}italic_E = italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT of both mesh choices converge at the same rate as the number of edges Nesubscript𝑁𝑒N_{e}italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT in the discretized domain increases. The effectiveness of simplical-DEC is therefore shown to be comparable to cubical-DEC, even when it is applied to a geometry where cubical meshing holds an advantage due to its elements sharing the same symmetry as the cavity. In Fig. 2b, we also study how simplicial-DEC produces accurate solutions to the edge fields of the eigenmodes. The convergence to analytical solutions of the edge fields is achieved.

These results help justify our shift to using simplicial meshing onwards, as its flexibility allows us to apply DEC to complicated, realistic structures where cubical symmetry is rarely present. Moreover, simplicial meshing allows for the implementation of different spatial resolutions in different regions, an important feature needed for the efficient application of DEC to systems comprised of multiple spatial scales.

Refer to caption
Figure 3: Comparison between the convergence of computed eigenvalues for a cavity embedded in a superconducting shell and those for a perfect cavity with the hard-wall boundary condition imposed. Nesubscript𝑁𝑒N_{e}italic_N start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is the number of edges used in discretizing the cavity. The dimensions of the vacuum regions within both cavities are Lx=1 cm,Ly=1.5 cmformulae-sequencesubscript𝐿𝑥1 cmsubscript𝐿𝑦1.5 cmL_{x}=1\text{\,cm},L_{y}=1.5\text{\,cm}italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT = 1 cm , italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = 1.5 cm, and Lz=2 cmsubscript𝐿𝑧2 cmL_{z}=2\text{\,cm}italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 2 cm. A schematic of the embedded cavity is shown in the inset at the center top, while the inset at the top-right corner shows schematically how the effective penetration depth is assigned on the inner boundary of this cavity via coarse-graining.

Next, we consider a rectangular cavity enclosed by a superconducting shell that has a finite thickness. The penetration depth of this shell is set to be short enough so that the field inside the cavity decays immediately at the material interface, i.e. the inner walls of the cavity. Instead of directly imposing the Dirichlet boundary condition at the cavity inner walls as in the previous case, the values of the field are left floating there, and we only impose a “hard-wall” boundary condition (Eq. (11)) at the outer boundary of the shell, which is also the boundary of the computational domain. The purpose of this numerical experiment is to investigate the validity of DEC when there are sharp interfaces, such as the vacuum-superconductor boundary here, where the field undergoes abrupt variations. The procedure for applying material properties to the edges lying on the interface is given in Eq. (9). In a non-uniform tetrahedral mesh, the partitioning of the dual faces esuperscript𝑒e^{\dagger}italic_e start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT that lie on multi-material interfaces to the neighboring material domains is also highly non-uniform. This leads to the effective penetration depth of each edge on the same boundary being different from one another. One can imagine the vacuum-material interface made of many small patches, each with a slightly different material property, as shown in the top-right inset in Fig. 3. We compute the eigenmodes of Eq. (10) for an embedded cavity that has the same dimensional ratios as in the perfect cavity case, and the convergence is shown in Fig. 3. The embedded cavity calculation achieves a similar order of accuracy as the perfect cavity case, and the two converge at the same rate as mesh density increases.

IV.2 Detection and removal of hybridization between degenerate modes

Refer to caption
Refer to caption
Refer to caption
Figure 4: Demonstration of the appearances and removals of hybridization between degenerate modes in a system with hidden symmetries. (a) The degenerate 3rdsuperscript3rd3^{\text{rd}}3 start_POSTSUPERSCRIPT rd end_POSTSUPERSCRIPT and 4thsuperscript4th4^{\text{th}}4 start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT modes of a rectangular cavity with perfect superconducting boundaries and dimension ratios of 1:1.5:2:11.5:21\!:\!1.5\!:\!21 : 1.5 : 2. (b) The 3rdsuperscript3rd3^{\text{rd}}3 start_POSTSUPERSCRIPT rd end_POSTSUPERSCRIPT and 4thsuperscript4th4^{\text{th}}4 start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT modes of a perturbed cavity, when the dimension ratios are now 1:1.5:2.01:11.5:2.011\!:\!1.5\!:\!2.011 : 1.5 : 2.01. The two modes are now well-characterized by their respective set of quantum numbers, which are schematically shown below the field distributions.

In symmetric structures such as the rectangular cavity discussed earlier in this article, there is a possibility of hybridization of degenerate modes in the numerically obtained eigenspectrum. These degeneracies are sometimes called “accidental” because they are not predicted by the symmetry group of the Hamiltonian but originate from a hidden symmetry of the system [64]. In such cases, numerical solvers tend to face difficulties in distinguishing these degenerate modes. For a rectangular superconducting cavity, the eigenvalues corresponding to each field component 𝒜isubscript𝒜𝑖{\mathcal{A}_{i}}caligraphic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT of the field 𝓐𝓐\bm{{\mathcal{A}}}bold_caligraphic_A reads

Ei(nx2Lx2+ny2Ly2+nz2Lz2),similar-tosubscript𝐸𝑖superscriptsubscript𝑛𝑥2superscriptsubscript𝐿𝑥2superscriptsubscript𝑛𝑦2superscriptsubscript𝐿𝑦2superscriptsubscript𝑛𝑧2superscriptsubscript𝐿𝑧2E_{i}\sim\bigg{(}\frac{n_{x}^{2}}{L_{x}^{2}}+\frac{n_{y}^{2}}{L_{y}^{2}}+\frac% {n_{z}^{2}}{L_{z}^{2}}\bigg{)},italic_E start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∼ ( divide start_ARG italic_n start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_n start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_n start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (12)

with the corresponding eigenfunction that is separable in cartesian coordinates

𝒜i=Xi(x)Yi(y)Zi(z).subscript𝒜𝑖subscript𝑋𝑖𝑥subscript𝑌𝑖𝑦subscript𝑍𝑖𝑧{\mathcal{A}_{i}}=X_{i}(x)Y_{i}(y)Z_{i}(z).caligraphic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_X start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x ) italic_Y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_y ) italic_Z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_z ) . (13)

If the ratio between any two of the three lengths Lx,Lysubscript𝐿𝑥subscript𝐿𝑦L_{x},L_{y}italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT or Lzsubscript𝐿𝑧L_{z}italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT are integers, then there can be degeneracies. To numerically lift these degeneracies, the hidden symmetry needs to be eliminated. In this specific case, we can do so by either extending the cavity boundary to have a finite thickness, as was done in our earlier discussion on the embedded cavity, or by introducing a small perturbation to the cavity shape. To demonstrate this, we consider again the perfect cavity with dimensions Lx:Ly:Lz=1:1.5:2:subscript𝐿𝑥subscript𝐿𝑦:subscript𝐿𝑧1:1.5:2L_{x}\!\!:\!\!L_{y}\!\!:\!\!L_{z}=1\!:\!1.5\!:\!2italic_L start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT : italic_L start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT : italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 1 : 1.5 : 2. In Fig. 4a, where the 3rdsuperscript3rd3^{\text{rd}}3 start_POSTSUPERSCRIPT rd end_POSTSUPERSCRIPT and 4thsuperscript4th4^{\text{th}}4 start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT modes of the unperturbed cavity are plotted, we can see that each field component of the two modes is not characterized by any single set of quantum numbers (nx,ny,nz)subscript𝑛𝑥subscript𝑛𝑦subscript𝑛𝑧(n_{x},n_{y},n_{z})( italic_n start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ), but rather a linear combination of the two accidentally degenerate modes. In Fig. 4b, we plot the same modes, but the cavity is now slightly perturbed by an amount ΔLz=0.01Δsubscript𝐿𝑧0.01\Delta L_{z}=0.01roman_Δ italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 0.01 along z𝑧zitalic_z, so that now Lz=Lz+ΔLz=2.01subscriptsuperscript𝐿𝑧subscript𝐿𝑧Δsubscript𝐿𝑧2.01L^{\prime}_{z}=L_{z}+\Delta L_{z}=2.01italic_L start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT + roman_Δ italic_L start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 2.01. The degeneracy is then lifted, and each of the two modes is now well-described by a distinct set of quantum numbers, which helps demonstrate the sensitivity of our numerical scheme to small changes in the geometry of the system simulated.

IV.3 Calculations for closed multiscale systems

Refer to caption
Figure 5: (a) Schematic of a system consisting of a three-dimensional superconducting cavity containing two spatially separated chips. Each chip is a dielectric substrate on top of which a Josephson qubit is mounted. The qubit is made of two superconducting capacitor pads connected together by a bridge containing a Josephson junction. A few exemplary modes of the system are plotted, such as (b) the mode where only qubit 1 participates, (c) only qubit 2 participates, (d) both qubits are activated, and (e) a hybridization of the two qubits and the cavity field. The units for all the axes in plots (b)-(e) are in millimeters.

So far, in this article, we have validated the accuracy of DEC by applying the method to simple systems with a large degree of symmetry and comparing the results to analytical solutions. The strength of DEC, however, lies in its ability to correctly capture the properties of systems containing multiple spatial scales through coarse-graining. The ability to simulate the full multiscale system is especially important for analyzing package modes and how they affect the on-chip operations [45] - a study that requires a computational mesh capable of encompassing multiple spatial scales to cover the volume of the entire package while simultaneously resolve the fine details on the quantum chip. Another example of a multiscale system is a three-dimensional superconducting cavity containing one or a few superconducting chips. The dimensions of a 3D cavity are typically on the order of 1similar-toabsent1\sim\!1∼ 1 cm, while a dielectric substrate holding the qubits is a few mm in size, and a qubit itself can have its smallest components in the range from micrometers down to tens of nanometers. To model such systems, one possible workaround to avoid computational bottlenecks is dividing the problem into individual simulations of separate parts. However, due to hybridization, the spectral characteristics of the entire system composed of these devices being in the vicinity of each other can be vastly different from that of the individual components. Hence, it is imperative to be able to model the entire structure and directly extract its modes.

A schematic of the system we shall consider is shown in Fig. 5a. It consists of a 3D cavity containing two dielectric substrates. A transmon qubit is mounted on each substrate. Each transmon qubit is composed of two superconducting capacitor pads connected by a Josephson junction. It is a well-known property of Josephson junctions that their dynamics are well characterized by the coarse-grained phase across it [65]. In other words, to an observer outside and away from the junction and who can only make measurements with a limited precision, the detailed dynamics inside the junction is unimportant to the dynamics that results from its interaction with the surrounding electromagnetic environment. This property can be utilized in coarse-grained calculations using DEC to reduce the complexity of the mesh needed; this is done by modeling a junction by a single edge instead of finely meshing the junction geometry. In the calculations of electromagnetic modes, this is equivalent to the linearization of the junction across the edge. To obtain modes whose wavelengths are orders of magnitude larger than the longitudinal size of the junction, the fine details of the material distribution within the junction are irrelevant and do not need to be resolved during meshing. In Fig. 5b-5e, a few example modes of this multiscale system are shown. In this calculation, we have set the dimensions of the cavity to be 1160×160×55011601605501160\!\times\!160\!\times\!5501160 × 160 × 550 (mm), while the sizes of the two identical substrates are both 25×50×32550325\!\times\!50\!\times\!325 × 50 × 3 (mm) and are separated by a distance of 600600600600 mm. The size of the capacitor pads of the qubits is 6×3×0.1630.16\!\times\!3\!\times\!0.16 × 3 × 0.1 (mm), and the two pads belonging to the same qubit are connected by a 0.20.20.20.2 mm-long bridge. We show the results for different types of modes that the system supports; in Figs. 5b and 5c, single qubit modes are shown, where the hybridization with the other qubit and the cavity field is suppressed. On the other hand, there can also be modes in which both qubits participate, such as the one shown in Fig. 5d. Finally, in Fig. 5e, we demonstrate a hybridized mode in which both the qubits and the cavity participate.

V Calculations of open modes

So far in this article, we have considered only closed systems, i.e. systems that only support modes that decay exponentially beyond their boundaries (e.g. a system surrounded by a superconducting enclosure much thicker than the relevant penetration depths). Here, we would like to extend DEC-QED to allow for a mathematically rigorous consideration of radiative losses. The correct modeling of open systems, where fields can propagate into and out of a confined domain, has been of great interest since it is directly related to the quantification of qubit lifetimes and radiative losses. A superconducting qubit placed in an open cavity acquires a relaxation rate through hybridization of the qubit with the electromagnetic modes of the cavity. Therefore its relaxation rate is given by the imaginary part of the complex-valued eigenfrequency of the qubit-like mode, while the difference between the real part in the frequency of a hybridized mode with the internal natural frequency of the qubit gives the Lamb shift [55]. In a similar vein, radiative loss of cavity modes is given by the imaginary parts of the cavity-like modes and is modified through hybridization with the qubit. Beyond the realm of superconducting microwave circuits, the problem has also been important to the broader scope of electromagnetic devices. We are particularly interested in the implementation of finite, open boundaries that are transparent so that fields can propagate through without any reflection. Moreover, the formulation needs to be compatible with the eventual second quantization of the electromagnetic field everywhere within the system, including the boundary itself. Here, we provide the derivations of two such approaches and present numerical demonstrations.

V.1 Green’s boundary integral formulation for scalar fields

Refer to caption
Figure 6: Schematic of an open system D𝐷Ditalic_D consisting of multiple material regions with refractive indices nisubscript𝑛𝑖n_{i}italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The Surface of Last Scattering (SoLS), the imaginary (transparent) boundary 𝒟𝒟\partial\mathcal{D}∂ caligraphic_D (the dashed contour) containing all material systems can have an arbitrary shape as long as all regions of interest are enclosed within it. The secondary boundary 𝒟superscript𝒟\partial\mathcal{D}^{\prime}∂ caligraphic_D start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT is created by contracting everywhere the original boundary 𝒟𝒟\partial\mathcal{D}∂ caligraphic_D by an amount ΔrΔ𝑟\Delta rroman_Δ italic_r.

First, we discuss an implementation of open boundaries using Green’s boundary integral formalism. The definition of radiative boundary conditions is analytically stated only at spatial infinity (r)𝑟(r\!\rightarrow\!\infty)( italic_r → ∞ ). However, our goal is to state a mathematically equivalent condition at a finite distance from the cavity to keep the computationally volume as small as possible. We refer to this imaginary boundary that includes all material systems within its volume as the “Surface of Last Scattering” (SoLS). We then employ the analytically known free-space frequency-domain Green’s function to ‘propagate’ back the boundary condition at infinity to the chosen SoLS.

Consider a domain 𝒟𝒟\mathcal{D}caligraphic_D consisting of possibly multiple disjoint regions all enclosed by an imaginary boundary surface. Before considering the spectral problem associated with the vector Helmholtz equation 6, we consider the warm-up problem of the Helmholtz equation for a scalar field ϕ(𝐫)italic-ϕ𝐫\phi({\mathbf{r}})italic_ϕ ( bold_r ):

2ϕ(𝐫)+ni2(𝐫)k2ϕ(𝐫)=0,superscript2italic-ϕ𝐫superscriptsubscript𝑛𝑖2𝐫superscript𝑘2italic-ϕ𝐫0\nabla^{2}\phi({\mathbf{r}})+n_{i}^{2}({\mathbf{r}})k^{2}\phi({\mathbf{r}})=0,∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ϕ ( bold_r ) + italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_r ) italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ϕ ( bold_r ) = 0 , (14)

where n(𝐫)𝑛𝐫n({\mathbf{r}})italic_n ( bold_r ) is the dielectric function of the ithsuperscript𝑖thi^{\text{th}}italic_i start_POSTSUPERSCRIPT th end_POSTSUPERSCRIPT region ΓisubscriptΓ𝑖\Gamma_{i}roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, and k𝑘kitalic_k is again the wavenumber. The Green’s function of the Helmholtz operator reads

G(𝐫,𝐫,k)={i4H0(1)(nik|𝐫𝐫|)in 2D,einik|𝐫𝐫|4π|𝐫𝐫|in 3D,𝐺𝐫superscript𝐫𝑘cases𝑖4superscriptsubscript𝐻01subscript𝑛𝑖𝑘𝐫superscript𝐫in 2D,superscript𝑒𝑖subscript𝑛𝑖𝑘𝐫superscript𝐫4𝜋𝐫superscript𝐫in 3D,\displaystyle G({\bf r,r^{\prime}},k)=\begin{cases}-\frac{i}{4}H_{0}^{(1)}(n_{% i}k|{\mathbf{r}}-{\mathbf{r}^{\prime}}|)&\text{in 2D,}\\ -\frac{e^{in_{i}k|{\mathbf{r}}-{\mathbf{r}^{\prime}}|}}{4\pi|{\mathbf{r}}-{% \mathbf{r}^{\prime}}|}&\text{in 3D,}\end{cases}italic_G ( bold_r , bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_k ) = { start_ROW start_CELL - divide start_ARG italic_i end_ARG start_ARG 4 end_ARG italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_k | bold_r - bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | ) end_CELL start_CELL in 2D, end_CELL end_ROW start_ROW start_CELL - divide start_ARG italic_e start_POSTSUPERSCRIPT italic_i italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_k | bold_r - bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_π | bold_r - bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | end_ARG end_CELL start_CELL in 3D, end_CELL end_ROW (15)

where H0(1)superscriptsubscript𝐻01H_{0}^{(1)}italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT is the Hankel function of the first kind. At first glance, the Green’s functions in Eq. (15) and their derivatives diverging at 𝐫=𝐫𝐫superscript𝐫{\mathbf{r}}\!=\!{\mathbf{r}^{\prime}}bold_r = bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT might seem like a problem. This difficulty can be circumvented by casting the boundary condition in an integral form to regularize the Green’s function singularity [66, 67]. Upon applying Green’s second identity and taking the limit as 𝐫Γisuperscript𝐫subscriptΓ𝑖{\mathbf{r}^{\prime}}\rightarrow\partial\Gamma_{i}bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT → ∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, the field value at a point 𝐫superscript𝐫{\mathbf{r}^{\prime}}bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT that lies on the boundary ΓisubscriptΓ𝑖\partial\Gamma_{i}∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT of a domain is [61]

ϕ(𝐫)=2𝒫d1Γi[ϕ(𝐫)G(𝐫,𝐫,k)G(𝐫,𝐫,k)ϕ(𝐫)]𝐝s,italic-ϕsuperscript𝐫2superscript𝒫𝑑1subscriptsubscriptΓ𝑖delimited-[]italic-ϕ𝐫𝐺𝐫superscript𝐫𝑘𝐺𝐫superscript𝐫𝑘italic-ϕ𝐫differential-d𝑠\phi({\bf r^{\prime}})=2\mathcal{P}^{d-1}\!\!\int_{\partial\Gamma_{i}}\!\!\big% {[}\phi({\bf r})\nabla G({\bf r,r^{\prime}},k)-G({\bf r,r^{\prime}},k)\nabla% \phi({\bf r})\big{]}\cdot{\mathbf{d}s},italic_ϕ ( bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = 2 caligraphic_P start_POSTSUPERSCRIPT italic_d - 1 end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT ∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_ϕ ( bold_r ) ∇ italic_G ( bold_r , bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_k ) - italic_G ( bold_r , bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_k ) ∇ italic_ϕ ( bold_r ) ] ⋅ bold_d italic_s , (16)

where 𝒫d1superscript𝒫𝑑1\mathcal{P}^{d-1}caligraphic_P start_POSTSUPERSCRIPT italic_d - 1 end_POSTSUPERSCRIPT indicates the (d1)𝑑1(d\!-\!1)( italic_d - 1 )-dimensional Cauchy principal value integral, with d𝑑ditalic_d being the dimension of the domain. A detailed derivation of Eq. (16) is given in Appendix B for completeness. In Eq. (16), the field at a point on the boundary is determined by the field and its gradient everywhere else on the same boundary. Eq. (16) is applicable to smooth boundaries of any form and shape. They can also be made up of different disjoint but closed segments.

To use Eq. (16) for applying boundary condition at the domain boundary 𝒟𝒟\partial\mathcal{D}∂ caligraphic_D, we need to numerically evaluate the normal gradient of the boundary field in a way that is self-contained within the field living on the boundary itself. To do so, consider a secondary boundary 𝒟superscript𝒟\partial\mathcal{D}^{\prime}∂ caligraphic_D start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT that is formed by contracting 𝒟𝒟\partial\mathcal{D}∂ caligraphic_D by an amount ΔrΔ𝑟\Delta rroman_Δ italic_r along the normal direction everywhere on the surface, as shown schematically in Fig. 6. There is now a thin layer bounded by 𝒟𝒟𝒟superscript𝒟\partial\mathcal{D}\cup\partial\mathcal{D}^{\prime}∂ caligraphic_D ∪ ∂ caligraphic_D start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT over which Green’s identity can be performed to obtain Eq. (16) that determines ϕ(𝐫)italic-ϕsuperscript𝐫\phi({\bf r^{\prime}})italic_ϕ ( bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) for every point 𝐫𝒟superscript𝐫𝒟{\mathbf{r}^{\prime}}\in\partial\mathcal{D}bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ ∂ caligraphic_D. The surface integral is now done over 𝒟𝒟𝒟superscript𝒟\partial\mathcal{D}\cup\partial\mathcal{D}^{\prime}∂ caligraphic_D ∪ ∂ caligraphic_D start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT. Note that during the evaluation of Eq. (16) the unit normal vector n^^𝑛\hat{n}over^ start_ARG italic_n end_ARG on 𝒟superscript𝒟\partial\mathcal{D}^{\prime}∂ caligraphic_D start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT points into 𝒟𝒟\mathcal{D}caligraphic_D.

The application of open boundaries using Eq. (16) results in a boundary condition that depends parametrically on k𝑘kitalic_k. To solve for the eigenmodes of Eq. (14), we rewrite it into the following matrix representation

[s(k)+k2]Φ=0,delimited-[]subscript𝑠𝑘superscript𝑘2Φ0\big{[}\mathbb{H}_{s}(k)+k^{2}\big{]}\Phi=0,[ blackboard_H start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_k ) + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] roman_Φ = 0 , (17)

where s(k)subscript𝑠𝑘\mathbb{H}_{s}(k)blackboard_H start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_k ) is the scalar Helmholtz operator, and the vector ΦΦ\Phiroman_Φ contains the scalar field evaluated at all vertices in the computational mesh. The task of finding the eigenvalues of s(k)subscript𝑠𝑘\mathbb{H}_{s}(k)blackboard_H start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_k ) can be cast into a singular value decomposition (SVD) problem [68]. The eigenvalues kαsubscript𝑘𝛼k_{\alpha}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT correspond to locations of local minima in the lowest singular value of 𝕄s(k)=s(k)+k2subscript𝕄𝑠𝑘subscript𝑠𝑘superscript𝑘2\mathbb{M}_{s}(k)=\mathbb{H}_{s}(k)+k^{2}blackboard_M start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_k ) = blackboard_H start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_k ) + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

Refer to caption
Figure 7: (a) The lowest singular value of 𝕄s(k)subscript𝕄𝑠𝑘\mathbb{M}_{s}(k)blackboard_M start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_k ) computed for a sample range of k𝑘kitalic_k. The physical system is a dielectric disk with radius Rd=5subscript𝑅𝑑5R_{d}=5italic_R start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 5 mm and n=1.5𝑛1.5n=1.5italic_n = 1.5. A schematic of the system is given in the top-left corner. (b) The eigenvalues knsubscript𝑘𝑛k_{n}italic_k start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT of s(k)subscript𝑠𝑘\mathbb{H}_{s}(k)blackboard_H start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_k ), are indicated by the locations of the local minima of the lowest singular value of 𝕄s(k)subscript𝕄𝑠𝑘\mathbb{M}_{s}(k)blackboard_M start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_k ). The semi-analytical results are plotted in blue, while the numerically computed values are plotted in orange.
Refer to caption
Figure 8: A selection of open modes of the scalar Helmholtz equation on a dielectric disk whose radius is Rd=5subscript𝑅𝑑5R_{d}=5italic_R start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 5 mm and refractive index is n=1.5𝑛1.5n=1.5italic_n = 1.5. In each panel: the surface plot shows the distribution of Re(ϕ)𝑅𝑒italic-ϕRe(\phi)italic_R italic_e ( italic_ϕ ) in the entire domain inside the boundary radius R𝑅Ritalic_R. The blue curve shows the value of Re(ϕ)𝑅𝑒italic-ϕRe(\phi)italic_R italic_e ( italic_ϕ ) at a fixed angle θ=4π/5𝜃4𝜋5\theta=4\pi/5italic_θ = 4 italic_π / 5 as a function of the radial coordinate r𝑟ritalic_r. The units for the horizontal axes are in mm. The orange vertical line indicates the edge of the disk. The red curve shows the field at a fixed radius r=4𝑟4r=4italic_r = 4 as a function of the angular coordinate θ𝜃\thetaitalic_θ.

We demonstrate the formulation presented here through the calculations of the open modes of Eq. (14) applied to a 2D dielectric disk placed in a vacuum. The disk has radius Rd=5subscript𝑅𝑑5R_{d}\!=\!5italic_R start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 5 mm and n=1.5𝑛1.5n=1.5italic_n = 1.5, while the boundary is chosen to be a concentric circle with radius R=8𝑅8R\!=\!8italic_R = 8 mm. The results for the lowest singular value of 𝕄s(k)subscript𝕄𝑠𝑘\mathbb{M}_{s}(k)blackboard_M start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( italic_k ) is shown in Fig. 7a for k𝑘kitalic_k within the range 0Re(kRd)4π0𝑅𝑒𝑘subscript𝑅𝑑4𝜋0\leq Re(kR_{d})\leq 4\pi0 ≤ italic_R italic_e ( italic_k italic_R start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) ≤ 4 italic_π and 3.5πIm(kRd)<03.5𝜋𝐼𝑚𝑘subscript𝑅𝑑0-3.5\pi\leq Im(kR_{d})<0- 3.5 italic_π ≤ italic_I italic_m ( italic_k italic_R start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) < 0. Since the system supports incoming and outgoing modes equally, we only need to focus on the outgoing modes, whose real part of k𝑘kitalic_k is positive. The distribution of the eigenvalues has a mirror symmetry across Re(k)=0𝑅𝑒𝑘0Re(k)=0italic_R italic_e ( italic_k ) = 0, and the incoming modes are only different from the outgoing ones by the sign of Re(k)𝑅𝑒𝑘Re(k)italic_R italic_e ( italic_k ). The local minima in Fig. 7a, whose locations correspond to the eigenvalues kαsubscript𝑘𝛼k_{\alpha}italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT, are collected and plotted in Fig. 7b. Note that since all modes are radiative, their eigenvalues are complex with the imaginary parts giving the corresponding relaxation rates. We also compare these numerically computed eigenvalues with the semi-analytical solutions obtained by solving the transcendental equation arising from matching the continuity condition of the field and its normal derivative at the edge of the disk. As seen in Fig. 7b, the numerically obtained eigenvalues (orange) agree with those obtained through the analytical equation (blue) over a wide range of k𝑘kitalic_k.

The field distributions of a few eigenmodes are presented in Fig. 8, where the real part of the fields, Re(ϕ)𝑅𝑒italic-ϕRe(\phi)italic_R italic_e ( italic_ϕ ), are plotted. The numerical labeling of the modes are done first in ascending order of Im(kαRd)𝐼𝑚subscript𝑘𝛼subscript𝑅𝑑Im(k_{\alpha}R_{d})italic_I italic_m ( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) within the range {3.5π,0}3.5𝜋0\{-3.5\pi,0\}{ - 3.5 italic_π , 0 }, then in ascending order of Re(kαRd)𝑅𝑒subscript𝑘𝛼subscript𝑅𝑑Re(k_{\alpha}R_{d})italic_R italic_e ( italic_k start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) in the range {0,4π}04𝜋\{0,4\pi\}{ 0 , 4 italic_π }. The features exhibited by these distributions can be understood through analytical considerations; in polar coordinates, one can write the fundamental solution to the scalar Helmholtz equation as ϕ(r,θ)=R(r)Θ(θ)italic-ϕ𝑟𝜃𝑅𝑟Θ𝜃\phi(r,\theta)\!=\!R(r)\Theta(\theta)italic_ϕ ( italic_r , italic_θ ) = italic_R ( italic_r ) roman_Θ ( italic_θ ). The angular dependence of the fields is of the form Θ(θ)e±imθsimilar-toΘ𝜃superscript𝑒plus-or-minus𝑖𝑚𝜃\Theta(\theta)\sim e^{\pm im\theta}roman_Θ ( italic_θ ) ∼ italic_e start_POSTSUPERSCRIPT ± italic_i italic_m italic_θ end_POSTSUPERSCRIPT, which is confirmed by the curves in red in Fig. 8, where the angular dependence of the fields are plotted at a fixed radius. The radial component, on the other hand, is given by R(r)Zm(nkr)similar-to𝑅𝑟subscript𝑍𝑚𝑛𝑘𝑟R(r)\sim Z_{m}(nkr)italic_R ( italic_r ) ∼ italic_Z start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ( italic_n italic_k italic_r ), where Zmsubscript𝑍𝑚Z_{m}italic_Z start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT is either the Bessel function of the first kind Jmsubscript𝐽𝑚J_{m}italic_J start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT, Bessel function of the second kind Ymsubscript𝑌𝑚Y_{m}italic_Y start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT, or the Hankel functions Hm(1,2)subscriptsuperscript𝐻12𝑚H^{(1,2)}_{m}italic_H start_POSTSUPERSCRIPT ( 1 , 2 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT. Inside the dielectric disk, the radial component of the field is given by Jmsubscript𝐽𝑚J_{m}italic_J start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT, while the field outside the disk is described by Hm(1)superscriptsubscript𝐻𝑚1H_{m}^{(1)}italic_H start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT (Hm(2))H_{m}^{(2)}\big{)}italic_H start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ) for an outgoing (incoming) wave. This radial dependence is shown in the blue curves in Fig. 8.

V.2 Green’s boundary integral formulation for vector fields

A problem of greater interest is, however, the implementation of open boundaries for vector fields. For this, we have extended Green’s boundary integral method for scalar fields to address the vector Helmholtz problem as well. Consider a divergence-free vector field 𝓐(𝐫)𝓐superscript𝐫\bm{\mathcal{A}}({\mathbf{r}^{\prime}})bold_caligraphic_A ( bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) that satisfies Eq. (6) defined over the domain 𝒟𝒟\mathcal{D}caligraphic_D. The dependence of the field value at a location 𝐫Dsuperscript𝐫𝐷{\mathbf{r}^{\prime}}\in\partial Dbold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ ∂ italic_D on the field everywhere else on the same boundary is given by

𝓐(𝐫)=2𝓐superscript𝐫2\displaystyle\bm{\mathcal{A}}({\mathbf{r}^{\prime}})=-2bold_caligraphic_A ( bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = - 2 𝒫d1D{G(𝐫,𝐫,k)[(×𝓐)×n^]\displaystyle\mathcal{P}^{d-1}\int_{\partial D}\bigg{\{}G({\mathbf{r}},{% \mathbf{r}^{\prime}},k)\big{[}({\mathbf{\nabla}}\times\bm{\mathcal{A}})\times% \hat{n}\big{]}caligraphic_P start_POSTSUPERSCRIPT italic_d - 1 end_POSTSUPERSCRIPT ∫ start_POSTSUBSCRIPT ∂ italic_D end_POSTSUBSCRIPT { italic_G ( bold_r , bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_k ) [ ( ∇ × bold_caligraphic_A ) × over^ start_ARG italic_n end_ARG ] (18)
G(𝐫,𝐫,k)(𝓐n^)G×(𝓐×n^)}ds,\displaystyle-\nabla G({\mathbf{r}},{\mathbf{r}^{\prime}},k)(\bm{\mathcal{A}}% \cdot\hat{n})-\nabla G\times({\bm{\mathcal{A}}\times\hat{n}})\bigg{\}}ds,- ∇ italic_G ( bold_r , bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_k ) ( bold_caligraphic_A ⋅ over^ start_ARG italic_n end_ARG ) - ∇ italic_G × ( bold_caligraphic_A × over^ start_ARG italic_n end_ARG ) } italic_d italic_s ,

where G(𝐫,𝐫,k)𝐺𝐫superscript𝐫𝑘G({\mathbf{r}},{\mathbf{r}^{\prime}},k)italic_G ( bold_r , bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_k ) is again the scalar Green’s function as defined in Eq. (15), and the principal value integral is generalized to the vector case. A detailed derivation of Eq. (18) is given in Appendix B.

Refer to caption
Figure 9: Open-mode calculations of the vector Helmholtz problem for a dielectric disk placed in a vacuum. The disk has radius Rd=5subscript𝑅𝑑5R_{d}=5italic_R start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 5 mm and n=1.5𝑛1.5n=1.5italic_n = 1.5. (a) A schematic of the polar grid used in the calculation. (b) The surface plot of the lowest singular value of the operator 𝕄v(k)subscript𝕄𝑣𝑘\mathbb{M}_{v}(k)blackboard_M start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_k ) as a function of k𝑘kitalic_k. The local minima indicate the eigenvalues of the vector Helmholtz operator v(k)subscript𝑣𝑘\mathbb{H}_{v}(k)blackboard_H start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_k ). (c) The distributions of Re(𝓐)𝑅𝑒𝓐Re(\bm{\mathcal{A}})italic_R italic_e ( bold_caligraphic_A ) in a number of radiative modes (the units for all axes are in mm).

We demonstrate the implementation of the boundary condition in Eq. (18) in DEC by calculating the modes of Eq. (6) applied to the same 2D dielectric disk studied in the scalar case. Consider a discretization of the computational domain into a polar grid that has NRsubscript𝑁𝑅N_{R}italic_N start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT and Nθsubscript𝑁𝜃N_{\theta}italic_N start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT vertices along each radial ray and circle, respectively. In the case of open BC, it is natural to allow the mesh to also be “open” such that there are additional edges on the physical boundary of the system that protrude outwards in the direction normal to the surface, as shown in Fig. 9a. A primal edge e𝑒eitalic_e in this setup is identified by nrsubscript𝑛𝑟n_{r}italic_n start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and nθsubscript𝑛𝜃n_{\theta}italic_n start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT, its radial and angular indices, accompanied by the superscript {r,θ}𝑟𝜃\{r,\theta\}{ italic_r , italic_θ } that indicates the direction of the edge. In DEC, applying BC to a vector field translates to the application of BC for the normal and tangential edge fields Φ(eNR,nθr)Φsubscriptsuperscript𝑒𝑟subscript𝑁𝑅subscript𝑛𝜃\Phi(e^{r}_{N_{R},n_{\theta}})roman_Φ ( italic_e start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) and Φ(eNR,nθθ)Φsubscriptsuperscript𝑒𝜃subscript𝑁𝑅subscript𝑛𝜃\Phi(e^{\theta}_{N_{R},n_{\theta}})roman_Φ ( italic_e start_POSTSUPERSCRIPT italic_θ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) living on the boundary edges. The condition in Eq. (18), when applied to a 2D circular boundary, then becomes

Φ(eNR,nθr)=Φ(eNR1,nθr)+θ2{Gr[ΦNR+1r(θ)+ΦNRr(θ)]RΔθGθ[Φnθ1θ(R)+Φnθθ(R)]ΔRR},Φsubscriptsuperscript𝑒𝑟subscript𝑁𝑅subscript𝑛𝜃Φsubscriptsuperscript𝑒𝑟subscript𝑁𝑅1subscript𝑛𝜃subscript𝜃2𝐺𝑟delimited-[]subscriptsuperscriptΦ𝑟subscript𝑁𝑅1𝜃superscriptsubscriptΦsubscript𝑁𝑅𝑟𝜃𝑅Δ𝜃𝐺𝜃delimited-[]subscriptsuperscriptΦ𝜃subscript𝑛𝜃1𝑅subscriptsuperscriptΦ𝜃subscript𝑛𝜃𝑅Δ𝑅𝑅\displaystyle\Phi(e^{r}_{N_{R},n_{\theta}})=-\Phi(e^{r}_{N_{R}-1,n_{\theta}})+% \sum_{\theta}2\bigg{\{}\frac{\partial G}{\partial r}\Big{[}\Phi^{r}_{N_{R}+1}(% \theta)+\Phi_{N_{R}}^{r}(\theta)\Big{]}R\Delta\theta-\frac{\partial G}{% \partial\theta}\Big{[}\Phi^{\theta}_{n_{\theta}-1}(R)+\Phi^{\theta}_{n_{\theta% }}(R)\Big{]}\frac{\Delta R}{R}\bigg{\}},roman_Φ ( italic_e start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) = - roman_Φ ( italic_e start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT - 1 , italic_n start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) + ∑ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT 2 { divide start_ARG ∂ italic_G end_ARG start_ARG ∂ italic_r end_ARG [ roman_Φ start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT + 1 end_POSTSUBSCRIPT ( italic_θ ) + roman_Φ start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_θ ) ] italic_R roman_Δ italic_θ - divide start_ARG ∂ italic_G end_ARG start_ARG ∂ italic_θ end_ARG [ roman_Φ start_POSTSUPERSCRIPT italic_θ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ( italic_R ) + roman_Φ start_POSTSUPERSCRIPT italic_θ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_R ) ] divide start_ARG roman_Δ italic_R end_ARG start_ARG italic_R end_ARG } , (19)

for the radial boundary edge, and

Φ(eNR,nθθ)=Φ(eNR,nθ1θ)θΦsubscriptsuperscript𝑒𝜃subscript𝑁𝑅subscript𝑛𝜃Φsubscriptsuperscript𝑒𝜃subscript𝑁𝑅subscript𝑛𝜃1subscript𝜃\displaystyle\Phi(e^{\theta}_{N_{R},n_{\theta}})=-\Phi(e^{\theta}_{N_{R},n_{% \theta}-1})-\sum_{\theta}roman_Φ ( italic_e start_POSTSUPERSCRIPT italic_θ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT end_POSTSUBSCRIPT ) = - roman_Φ ( italic_e start_POSTSUPERSCRIPT italic_θ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ) - ∑ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT {2Δθ[G(𝐫,𝐫,k)(1+RΔR)RGr](Φnθθ(R)+Φnθ1θ(R))\displaystyle\bigg{\{}2\Delta\theta\bigg{[}G({\mathbf{r}},{\mathbf{r}^{\prime}% },k)\Big{(}1+\frac{R}{\Delta R}\Big{)}-R\frac{\partial G}{\partial r}\bigg{]}% \Big{(}\Phi_{n_{\theta}}^{\theta}(R)+\Phi_{n_{\theta}-1}^{\theta}(R)\Big{)}{ 2 roman_Δ italic_θ [ italic_G ( bold_r , bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_k ) ( 1 + divide start_ARG italic_R end_ARG start_ARG roman_Δ italic_R end_ARG ) - italic_R divide start_ARG ∂ italic_G end_ARG start_ARG ∂ italic_r end_ARG ] ( roman_Φ start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_θ end_POSTSUPERSCRIPT ( italic_R ) + roman_Φ start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_θ end_POSTSUPERSCRIPT ( italic_R ) ) (20)
2GRΔθΔR(Φnθθ(RΔR)+Φnθ1θ(RΔR))2RΔθ2ΔRGθ[ΦNR+1r(θ)+ΦNRr(θ)]2𝐺𝑅Δ𝜃Δ𝑅superscriptsubscriptΦsubscript𝑛𝜃𝜃𝑅Δ𝑅superscriptsubscriptΦsubscript𝑛𝜃1𝜃𝑅Δ𝑅2𝑅Δsuperscript𝜃2Δ𝑅𝐺𝜃delimited-[]superscriptsubscriptΦsubscript𝑁𝑅1𝑟𝜃superscriptsubscriptΦsubscript𝑁𝑅𝑟𝜃\displaystyle-2G\frac{R\Delta\theta}{\Delta R}\Big{(}\Phi_{n_{\theta}}^{\theta% }(R-\Delta R)+\Phi_{n_{\theta}-1}^{\theta}(R-\Delta R)\Big{)}-2R\frac{\Delta% \theta^{2}}{\Delta R}\frac{\partial G}{\partial\theta}\Big{[}\Phi_{N_{R}+1}^{r% }(\theta)+\Phi_{N_{R}}^{r}(\theta)\Big{]}- 2 italic_G divide start_ARG italic_R roman_Δ italic_θ end_ARG start_ARG roman_Δ italic_R end_ARG ( roman_Φ start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_θ end_POSTSUPERSCRIPT ( italic_R - roman_Δ italic_R ) + roman_Φ start_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_θ end_POSTSUPERSCRIPT ( italic_R - roman_Δ italic_R ) ) - 2 italic_R divide start_ARG roman_Δ italic_θ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG roman_Δ italic_R end_ARG divide start_ARG ∂ italic_G end_ARG start_ARG ∂ italic_θ end_ARG [ roman_Φ start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_θ ) + roman_Φ start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_θ ) ]
GRΔθΔR[ΦNR+1r(θ+Δθ)+ΦNRr(θ+Δθ)ΦNR+1r(θΔθ)ΦNRr(θΔθ)]}\displaystyle-G\frac{R\Delta\theta}{\Delta R}\Big{[}\Phi_{N_{R}+1}^{r}(\theta+% \Delta\theta)+\Phi_{N_{R}}^{r}(\theta+\Delta\theta)-\Phi_{N_{R}+1}^{r}(\theta-% \Delta\theta)-\Phi_{N_{R}}^{r}(\theta-\Delta\theta)\Big{]}\bigg{\}}- italic_G divide start_ARG italic_R roman_Δ italic_θ end_ARG start_ARG roman_Δ italic_R end_ARG [ roman_Φ start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_θ + roman_Δ italic_θ ) + roman_Φ start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_θ + roman_Δ italic_θ ) - roman_Φ start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT + 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_θ - roman_Δ italic_θ ) - roman_Φ start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_θ - roman_Δ italic_θ ) ] }

for the angular boundary edges, where ΔRΔ𝑅\Delta Rroman_Δ italic_R and ΔθΔ𝜃\Delta\thetaroman_Δ italic_θ are the spacing between consecutive grid points on a radial ray and circle, respectively. Similar to the scalar problem, to search for the eigenvalues, we compute the lowest singular values of the operator 𝕄v(k)=v(k)k2subscript𝕄𝑣𝑘subscript𝑣𝑘superscript𝑘2\mathbb{M}_{v}(k)=\mathbb{H}_{v}(k)-k^{2}blackboard_M start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_k ) = blackboard_H start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_k ) - italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, where v(k)subscript𝑣𝑘\mathbb{H}_{v}(k)blackboard_H start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_k ) is the vector Helmholtz operator. The distribution of the local minima of 𝕄v(k)subscript𝕄𝑣𝑘\mathbb{M}_{v}(k)blackboard_M start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_k ) in k𝑘kitalic_k-space is shown in Fig. 9b. The field distributions of a few randomly selected eigenmodes are presented in Fig. 9c, where the plotted vector fields exhibit the correct behavior of how radiations permeate and escape a dielectric disk.

Refer to caption
Figure 10: Calculations of radiative modes of a transmon qubit sandwiched between two cylindrical superconducting capacitors. (a) The schematic of the system considered; the qubit is mounted on a circular dielectric disk that has radius Rd=4subscript𝑅𝑑4R_{d}=4italic_R start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 4 mm, refractive index n=1.5𝑛1.5n=1.5italic_n = 1.5. The qubit is composed of two superconducting islands that are 0.750.750.750.75 mm away from each other and connected by a bridge containing a junction in the middle. The width of both the bridge and the junction is 50μ𝜇\muitalic_μm. (b) An example of a “dipole” mode; the qubit acts like a dipole with the field lines starting from the surface of one island and ending on the surface of another island. (c) An example of a mode where the field lines flow around the individual capacitor islands with no normal component at each island’s surface. The boundary of the qubit is painted in yellow in the enlarged insets, and all the axes of the plots are in units of mm. (d) Spontaneous emission rate of a qubit-like mode as a function of the qubit frequency.

To demonstrate the versatility of DEC in computing radiative modes of arbitrarily shaped systems, we apply the formulation to the modeling of a superconducting qubit chip placed in the region between two cylindrical capacitors, as schematically shown in Fig. 10a. The qubit is made of two superconducting islands that are separated by a distance of 0.750.750.750.75 mm and connected by a bridge that contains a Josephson junction. The width of both the bridge and the junction is 50μ50𝜇50\mu50 italic_μm, and the qubit is mounted on a circular dielectric disk that has a radius Rd=4subscript𝑅𝑑4R_{d}=4italic_R start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 4 mm with refractive index n=1.5𝑛1.5n=1.5italic_n = 1.5. We assume the two ideal superconducting cylindrical capacitors extend indefinitely towards both ends, and the qubit is placed in the middle of the slit separating them. The width hhitalic_h of the vacuum spacing between the two cylinders is taken to be very small compared to their radius Rcsubscript𝑅𝑐R_{c}italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. This allows us to effectively discretize the distance between the surfaces of the two cylinders by two edges. Due to mirror symmetry at z=0𝑧0z=0italic_z = 0, the amplitude of the coarse-grained field on one z𝑧zitalic_z-edge is identical to that of its mirroring edge. This allows us to decouple the in-plane field component from the out-of-plane component, and we can focus on solving for the modes of Eq. (10) applied to the 2D in-plane field on the z=0𝑧0z=0italic_z = 0 slice where the substrate containing the qubit is located. We choose the computational boundary to be a circle whose radius R=8𝑅8R=8italic_R = 8 mm is half that of the cylindrical capacitors Rcsubscript𝑅𝑐R_{c}italic_R start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. There are two notable types of modes, whose examples are shown in Figs. 10b-10c. Due to the geometry of the qubit being two islands connected by a bridge, it can act as a dipole. A mode that behaves this way is shown in Fig. 10b, where the field lines exit in the normal direction from one island and enter the other island. The field also decays in the bulks of the islands at the rate determined by their penetration depths. In another scenario, due to their separation, the two islands can behave as individual superconducting objects around which the field lines flow with no normal component at the surfaces of the islands. A demonstration of such a mode is shown in Fig. 10c. Finally, in Fig. 10d, we show how one can tune the internal resonance frequency of the qubit through a cavity resonance, exposing the Purcell enhancement of the qubit relaxation rate [54]. The relaxation rate due to radiative loss of a qubitlike mode, i.e. the spontaneous emission rate, is plotted as a function of the qubit frequency. We observe a finite peak as the qubit freqency is on-resonant with a cavitylike mode, signifiying an enhanced emission rate due to the hybridization of this qubit mode with the cavity field [18]. In this case, the Purcell enhancement factor with respect to its value away from resonance is not large, because we intentionally chose a very (radiatively) lossy cavity represented by placing two cylinders facing each other, a situation that corresponds to the regime of overlapping resonances (finesse 1much-less-thanabsent1\ll 1≪ 1).

As was mentioned earlier, the formulation for calculating radiative fields using Green’s function method is flexible in terms of the shape and topology of boundary surfaces it can be applied. The physical boundary of the system can be made of multiple, possibly disjoint, but closed segments that bound a non-simply connected structure. However, the vector field being calculated has to be divergent-free as a prerequisite. Although this restriction is not a concern in many useful cases, such as the calculation of the electric field in a source-less region or of the magnetic vector potential in the Coulomb gauge, it is sometimes helpful to have the freedom of not necessarily choosing a divergence-less field. In the following section, we introduce an alternative formulation that achieves this.

V.3 Vector spherical harmonics expansion

In this section, we discuss the implementation of open boundaries using vector spherical harmonics (VSH) expansions. The VSH [69] is an extension for vector fields of the perhaps more well-known scalar spherical harmonics. There are three sub-classes of these vectors, defined in spherical coordinates (r,θ,φ)𝑟𝜃𝜑(r,\theta,\varphi)( italic_r , italic_θ , italic_φ ) as

𝐘lmsubscript𝐘𝑙𝑚\displaystyle{\mathbf{Y}_{lm}}bold_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT =Ylm(θ,φ)r^absentsubscript𝑌𝑙𝑚𝜃𝜑^𝑟\displaystyle=Y_{lm}(\theta,\varphi)\hat{r}= italic_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( italic_θ , italic_φ ) over^ start_ARG italic_r end_ARG (21)
𝚿lmsubscript𝚿𝑙𝑚\displaystyle{\mathbf{\Psi}_{lm}}bold_Ψ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT =rYlm(θ,φ)absent𝑟subscript𝑌𝑙𝑚𝜃𝜑\displaystyle=r\nabla Y_{lm}(\theta,\varphi)= italic_r ∇ italic_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( italic_θ , italic_φ ) (22)
𝚽lmsubscript𝚽𝑙𝑚\displaystyle{\mathbf{\Phi}_{lm}}bold_Φ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT =𝐫×Ylm(θ,φ)absent𝐫subscript𝑌𝑙𝑚𝜃𝜑\displaystyle=\mathbf{r}\times\nabla Y_{lm}(\theta,\varphi)= bold_r × ∇ italic_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( italic_θ , italic_φ ) (23)

that altogether form an orthonormal and complete basis. Any vector field can then be expanded as follows

𝓐(𝐫)=l=0m=ll𝒜lmr𝐘lm+𝒜lm(1)𝚿lm+𝒜lm(2)𝚽lm,𝓐𝐫superscriptsubscript𝑙0superscriptsubscript𝑚𝑙𝑙subscriptsuperscript𝒜𝑟𝑙𝑚subscript𝐘𝑙𝑚superscriptsubscript𝒜𝑙𝑚1subscript𝚿𝑙𝑚superscriptsubscript𝒜𝑙𝑚2subscript𝚽𝑙𝑚\displaystyle\bm{\mathcal{A}}({\mathbf{r}})=\sum_{l=0}^{\infty}\sum_{m=-l}^{l}% \mathcal{A}^{r}_{lm}{\mathbf{Y}_{lm}}+\mathcal{A}_{lm}^{(1)}{\mathbf{\Psi}_{lm% }}+\mathcal{A}_{lm}^{(2)}{\mathbf{\Phi}_{lm}},bold_caligraphic_A ( bold_r ) = ∑ start_POSTSUBSCRIPT italic_l = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_m = - italic_l end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT caligraphic_A start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT bold_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT + caligraphic_A start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT bold_Ψ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT + caligraphic_A start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT bold_Φ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT , (24)

where coefficients 𝒜lmr,𝒜lm(1)subscriptsuperscript𝒜𝑟𝑙𝑚superscriptsubscript𝒜𝑙𝑚1\mathcal{A}^{r}_{lm},\mathcal{A}_{lm}^{(1)}caligraphic_A start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT , caligraphic_A start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT and 𝒜lm(2)superscriptsubscript𝒜𝑙𝑚2\mathcal{A}_{lm}^{(2)}caligraphic_A start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT are functions of the radial coordinate r𝑟ritalic_r. In Eq. (24) above, of the three mutually orthogonal terms, the first term on the right-hand side is the radial component of the field, while the remaining two terms are angular contributions. Note that neither of the two angular terms aligns with azimuthal or polar directions but is a linear combination of both. Using Eq. (24) as an ansatz for the field, Eq. (6) can then be split into three decoupled ODEs, each for one of the three coefficients. Their solutions are

𝒜lmr=superscriptsubscript𝒜𝑙𝑚𝑟absent\displaystyle\mathcal{A}_{lm}^{r}=caligraphic_A start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT = Zl+1/2(kr)r3/2,subscript𝑍𝑙12𝑘𝑟superscript𝑟32\displaystyle\frac{Z_{l+1/2}(kr)}{r^{3/2}},divide start_ARG italic_Z start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT end_ARG , (25)
𝒜lm(1)=superscriptsubscript𝒜𝑙𝑚1absent\displaystyle\mathcal{A}_{lm}^{(1)}=caligraphic_A start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT = 1ł(l+1){Zl+1/2(kr)r3/2+kr[Zl1/2(kr)\displaystyle\frac{1}{\l(l+1)}\bigg{\{}\frac{Z_{l+1/2}(kr)}{r^{3/2}}+\frac{k}{% \sqrt{r}}\Big{[}Z_{l-1/2}(kr)divide start_ARG 1 end_ARG start_ARG italic_ł ( italic_l + 1 ) end_ARG { divide start_ARG italic_Z start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_k end_ARG start_ARG square-root start_ARG italic_r end_ARG end_ARG [ italic_Z start_POSTSUBSCRIPT italic_l - 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_r )
Zl+3/2(kr)]},\displaystyle\hskip 72.26999pt-Z_{l+3/2}(kr)\Big{]}\bigg{\}},- italic_Z start_POSTSUBSCRIPT italic_l + 3 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) ] } , (26)
𝒜lm(2)=superscriptsubscript𝒜𝑙𝑚2absent\displaystyle\mathcal{A}_{lm}^{(2)}=caligraphic_A start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT = π2krZl+1/2(kr)𝜋2𝑘𝑟subscript𝑍𝑙12𝑘𝑟\displaystyle\frac{\pi}{2kr}Z_{l+1/2}(kr)divide start_ARG italic_π end_ARG start_ARG 2 italic_k italic_r end_ARG italic_Z start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) (27)

where Z𝑍Zitalic_Z in general can be the Bessel function of the first kind J𝐽Jitalic_J, the Bessel function of the second kind Y𝑌Yitalic_Y, or the Hankel functions H(1,2)superscript𝐻12H^{(1,2)}italic_H start_POSTSUPERSCRIPT ( 1 , 2 ) end_POSTSUPERSCRIPT. For the outgoing(incoming) waves, the Hankel functions of the first(second) kind provide the appropriate description. After the VSH expansion in Eq. (24), the implementation of the open boundary condition on the field 𝓐𝓐\bm{\mathcal{A}}bold_caligraphic_A now reduces to the BC on each of the three field components 𝓐𝒓,𝓐(𝟏)superscript𝓐𝒓superscript𝓐1\bm{\mathcal{A}^{r}},\bm{\mathcal{A}^{(1)}}bold_caligraphic_A start_POSTSUPERSCRIPT bold_italic_r end_POSTSUPERSCRIPT , bold_caligraphic_A start_POSTSUPERSCRIPT bold_( bold_1 bold_) end_POSTSUPERSCRIPT and 𝓐(𝟐)superscript𝓐2\bm{\mathcal{A}^{(2)}}bold_caligraphic_A start_POSTSUPERSCRIPT bold_( bold_2 bold_) end_POSTSUPERSCRIPT. Here we discuss the BC for 𝒜rsuperscript𝒜𝑟\mathcal{A}^{r}caligraphic_A start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT as an example, while the details about BCs for 𝒜(1)superscript𝒜1\mathcal{A}^{(1)}caligraphic_A start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT and 𝒜(2)superscript𝒜2\mathcal{A}^{(2)}caligraphic_A start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT are given in Appendix C. Consider a spherical boundary of radius R𝑅Ritalic_R. The series expansion for the radial component of the radiating field on the sphere is given by

𝓐r(R,θ,φ)superscript𝓐𝑟𝑅𝜃𝜑\displaystyle\bm{\mathcal{A}}^{r}(R,\theta,\varphi)bold_caligraphic_A start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_R , italic_θ , italic_φ ) =l,ma~lmrAlmr(R)𝐘lm(R,θ,φ)absentsubscript𝑙𝑚superscriptsubscript~𝑎𝑙𝑚𝑟superscriptsubscript𝐴𝑙𝑚𝑟𝑅subscript𝐘𝑙𝑚𝑅𝜃𝜑\displaystyle=\sum_{l,m}\tilde{a}_{lm}^{r}A_{lm}^{r}(R){\mathbf{Y}_{lm}}(R,% \theta,\varphi)= ∑ start_POSTSUBSCRIPT italic_l , italic_m end_POSTSUBSCRIPT over~ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT italic_A start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_R ) bold_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( italic_R , italic_θ , italic_φ )
=l,ma~lmrHl+1/2(kR)R3/2Ylmr^,absentsubscript𝑙𝑚superscriptsubscript~𝑎𝑙𝑚𝑟subscript𝐻𝑙12𝑘𝑅superscript𝑅32subscript𝑌𝑙𝑚^𝑟\displaystyle=\sum_{l,m}\tilde{a}_{lm}^{r}\frac{H_{l+1/2}(kR)}{R^{-3/2}}Y_{lm}% \hat{r},= ∑ start_POSTSUBSCRIPT italic_l , italic_m end_POSTSUBSCRIPT over~ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT divide start_ARG italic_H start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) end_ARG start_ARG italic_R start_POSTSUPERSCRIPT - 3 / 2 end_POSTSUPERSCRIPT end_ARG italic_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT over^ start_ARG italic_r end_ARG , (28)

where a~lmrsuperscriptsubscript~𝑎𝑙𝑚𝑟\tilde{a}_{lm}^{r}over~ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT are the expansion coefficients. Performing an integral over the surface of the sphere and utilizing the orthonormality condition of VSH, we arrive at an expression for the expansion coefficients as follows

almrsubscriptsuperscript𝑎𝑟𝑙𝑚\displaystyle a^{r}_{lm}italic_a start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT a~lmrHl+1/2(kR)absentsuperscriptsubscript~𝑎𝑙𝑚𝑟subscript𝐻𝑙12𝑘𝑅\displaystyle\equiv\tilde{a}_{lm}^{r}H_{l+1/2}(kR)≡ over~ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT italic_H start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) (29)
=R3/2𝑑Ω𝓐r𝐘lm*absentsuperscript𝑅32differential-dΩsuperscript𝓐𝑟subscriptsuperscript𝐘𝑙𝑚\displaystyle=R^{3/2}\int d\Omega\bm{\mathcal{A}}^{r}\cdot{\mathbf{Y}^{*}_{lm}}= italic_R start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT ∫ italic_d roman_Ω bold_caligraphic_A start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ⋅ bold_Y start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT
=R1/2i,j𝒜r(R,θi,φj)Ylm*(θi,φj)ΔA(f),absentsuperscript𝑅12subscript𝑖𝑗superscript𝒜𝑟𝑅subscript𝜃𝑖subscript𝜑𝑗superscriptsubscript𝑌𝑙𝑚subscript𝜃𝑖subscript𝜑𝑗Δ𝐴𝑓\displaystyle=R^{-1/2}\sum_{i,j}\mathcal{A}^{r}(R,\theta_{i},\varphi_{j})Y_{lm% }^{*}(\theta_{i},\varphi_{j})\Delta A(f),= italic_R start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT caligraphic_A start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_R , italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_φ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) italic_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ( italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_φ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) roman_Δ italic_A ( italic_f ) ,

where 𝑑Ωdifferential-dΩ\int d\Omega∫ italic_d roman_Ω is the solid angle integral that covers the entire sphere, and the last line of Eq. (29) above is the discretized version of the integral written as a sum over all the triangular patches f𝑓fitalic_f whose centers are located at (θi,φj)subscript𝜃𝑖subscript𝜑𝑗(\theta_{i},\varphi_{j})( italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_φ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) on the spherical boundary. To apply the BCs, consider the sphere beneath the boundary that has radius RΔr𝑅Δ𝑟R\!-\!\Delta ritalic_R - roman_Δ italic_r (i.e. the second-to-last layer). The radial derivative of 𝒜rsuperscript𝒜𝑟\mathcal{A}^{r}caligraphic_A start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT is given by

𝒜r(R,θ,φ)𝒜r(RΔR,θ,φ)ΔR=superscript𝒜𝑟𝑅𝜃𝜑superscript𝒜𝑟𝑅Δ𝑅𝜃𝜑Δ𝑅absent\displaystyle\frac{\mathcal{A}^{r}(R,\theta,\varphi)-\mathcal{A}^{r}(R-\Delta R% ,\theta,\varphi)}{\Delta R}=divide start_ARG caligraphic_A start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_R , italic_θ , italic_φ ) - caligraphic_A start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_R - roman_Δ italic_R , italic_θ , italic_φ ) end_ARG start_ARG roman_Δ italic_R end_ARG = (30)
l,ma~lmr[32R5/2Hl+1/2(kR)+kR3/2Hl+1/2(kR)]Yl,m,subscript𝑙𝑚subscriptsuperscript~𝑎𝑟𝑙𝑚delimited-[]32superscript𝑅52subscript𝐻𝑙12𝑘𝑅𝑘superscript𝑅32subscriptsuperscript𝐻𝑙12𝑘𝑅subscript𝑌𝑙𝑚\displaystyle\sum_{l,m}\tilde{a}^{r}_{lm}\bigg{[}\frac{-3}{2R^{5/2}}H_{l+1/2}(% kR)+\frac{k}{R^{3/2}}H^{\prime}_{l+1/2}(kR)\bigg{]}Y_{l,m},∑ start_POSTSUBSCRIPT italic_l , italic_m end_POSTSUBSCRIPT over~ start_ARG italic_a end_ARG start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT [ divide start_ARG - 3 end_ARG start_ARG 2 italic_R start_POSTSUPERSCRIPT 5 / 2 end_POSTSUPERSCRIPT end_ARG italic_H start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) + divide start_ARG italic_k end_ARG start_ARG italic_R start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT end_ARG italic_H start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) ] italic_Y start_POSTSUBSCRIPT italic_l , italic_m end_POSTSUBSCRIPT ,

which allows us to write the field at the boundary 𝒜r(R,θ,φ)superscript𝒜𝑟𝑅𝜃𝜑\mathcal{A}^{r}(R,\theta,\varphi)caligraphic_A start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_R , italic_θ , italic_φ ) in terms of the expansion coefficients a~lmrsubscriptsuperscript~𝑎𝑟𝑙𝑚\tilde{a}^{r}_{lm}over~ start_ARG italic_a end_ARG start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT and the field at the layer with radius RΔR𝑅Δ𝑅R\!-\!\Delta Ritalic_R - roman_Δ italic_R. Using Eq. (30) and the expression for a~lmrsubscriptsuperscript~𝑎𝑟𝑙𝑚\tilde{a}^{r}_{lm}over~ start_ARG italic_a end_ARG start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT derived in Eq. (29), the boundary condition for 𝒜rsuperscript𝒜𝑟\mathcal{A}^{r}caligraphic_A start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT is then

𝒜rsuperscript𝒜𝑟\displaystyle\mathcal{A}^{r}caligraphic_A start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT (R,θ,φ)=l.mi,j𝒜r(RΔR,θi,φj)Ylm*(θi,φj)ΔA(f)R2𝑅𝜃𝜑subscriptformulae-sequence𝑙𝑚subscript𝑖𝑗superscript𝒜𝑟𝑅Δ𝑅subscript𝜃𝑖subscript𝜑𝑗subscriptsuperscript𝑌𝑙𝑚subscript𝜃𝑖subscript𝜑𝑗Δ𝐴𝑓superscript𝑅2\displaystyle(R,\theta,\varphi)\!=\!\sum_{l.m}\!\sum_{i,j}\!\mathcal{A}^{r}(R% \!-\!\Delta R,\theta_{i},\varphi_{j})Y^{*}_{lm}(\theta_{i},\varphi_{j})\frac{% \Delta A(f)}{R^{2}}( italic_R , italic_θ , italic_φ ) = ∑ start_POSTSUBSCRIPT italic_l . italic_m end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT caligraphic_A start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_R - roman_Δ italic_R , italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_φ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) italic_Y start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_φ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) divide start_ARG roman_Δ italic_A ( italic_f ) end_ARG start_ARG italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG
×{1+ΔR[32R+kHl1/2(kR)Hl+3/2(kR)2Hl+1/2(kR)]}absent1Δ𝑅delimited-[]32𝑅𝑘subscript𝐻𝑙12𝑘𝑅subscript𝐻𝑙32𝑘𝑅2subscript𝐻𝑙12𝑘𝑅\displaystyle\times\bigg{\{}1+\Delta R\Big{[}-\frac{3}{2R}+k\frac{H_{l-1/2}(kR% )-H_{l+3/2}(kR)}{2H_{l+1/2}(kR)}\Big{]}\bigg{\}}× { 1 + roman_Δ italic_R [ - divide start_ARG 3 end_ARG start_ARG 2 italic_R end_ARG + italic_k divide start_ARG italic_H start_POSTSUBSCRIPT italic_l - 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) - italic_H start_POSTSUBSCRIPT italic_l + 3 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) end_ARG start_ARG 2 italic_H start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) end_ARG ] }
×Ylm(θ,φ).absentsubscript𝑌𝑙𝑚𝜃𝜑\displaystyle\hskip 86.72377pt\times Y_{lm}(\theta,\varphi).× italic_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( italic_θ , italic_φ ) . (31)

The expression in Eq. (V.3) is for the radial component of the field, which in DEC is a scalar living on vertices of the primal mesh. To arrive at the appropriate BC for the edge field Φ(e)Φ𝑒\Phi(e)roman_Φ ( italic_e ) governed by the discrete vector Helmholtz equation given in Eq. (10), we need to rewrite 𝒜r(R,θ,φ)superscript𝒜𝑟𝑅𝜃𝜑\mathcal{A}^{r}(R,\theta,\varphi)caligraphic_A start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_R , italic_θ , italic_φ ) and 𝒜r(RΔR,θi,φj)superscript𝒜𝑟𝑅Δ𝑅subscript𝜃𝑖subscript𝜑𝑗\mathcal{A}^{r}(R\!-\!\Delta R,\theta_{i},\varphi_{j})caligraphic_A start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_R - roman_Δ italic_R , italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_φ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) in terms of the edges belonging to the vertex located at (R,θ,φ)𝑅𝜃𝜑(R,\theta,\varphi)( italic_R , italic_θ , italic_φ ). This means two consecutive spherical layers beneath the boundary are needed to compute 𝒜r(RΔR,θi,φj)superscript𝒜𝑟𝑅Δ𝑅subscript𝜃𝑖subscript𝜑𝑗\mathcal{A}^{r}(R\!-\!\Delta R,\theta_{i},\varphi_{j})caligraphic_A start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ( italic_R - roman_Δ italic_R , italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_φ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ). The procedure as a whole can be substantially simplified by designing the mesh so that the boundary surface and the two layers beneath it have identical triangulation patterns. This ensures that for any vertex vbsubscript𝑣𝑏v_{b}italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT located at (R,θ,φ)𝑅𝜃𝜑(R,\theta,\varphi)( italic_R , italic_θ , italic_φ ) on the boundary surface, there is an edge ersuperscript𝑒𝑟e^{r}italic_e start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT oriented radially that connects vbsubscript𝑣𝑏v_{b}italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT with the vertex vbsuperscriptsubscript𝑣𝑏v_{b}^{\prime}italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT at (RΔR,θ,φ)𝑅Δ𝑅𝜃𝜑(R\!-\!\Delta R,\theta,\varphi)( italic_R - roman_Δ italic_R , italic_θ , italic_φ ) that has the same angular coordinates as vbsubscript𝑣𝑏v_{b}italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT. Similarly, there is always an edge connecting vbsuperscriptsubscript𝑣𝑏v_{b}^{\prime}italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT with a vertex vb′′superscriptsubscript𝑣𝑏′′v_{b}^{\prime\prime}italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT at (R2ΔR,θ,φ)𝑅2Δ𝑅𝜃𝜑(R\!-\!2\Delta R,\theta,\varphi)( italic_R - 2 roman_Δ italic_R , italic_θ , italic_φ ). Note that we only need the three outmost layers of the mesh to be spherical and have identical triangulation patterns, while the rest of the internal domain can be freely and randomly discretized with tetrahedra. Similar to the discussion on Green’s function approach in the previous section, here, for open boundaries, it is convenient to use an open mesh such that there are radial edges that protrude out of the boundary surface. The BC applied to these boundary radial edges ebrsubscriptsuperscript𝑒𝑟𝑏e^{r}_{b}italic_e start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT is then

Φ(ebr)=Φ(ebr)+erfv|verl,mΔA(f)3R2Φ(er){1+ΔR[32R+kHl1/2(kR)Hl+3/2(kR)2Hl+1/2(kR)]Ylm*(f)Ylm(θ,φ)},Φsubscriptsuperscript𝑒𝑟𝑏Φsubscriptsuperscript𝑒superscript𝑟𝑏subscriptsuperscript𝑒𝑟subscriptsuperset-of𝑓conditional𝑣𝑣superscript𝑒𝑟subscript𝑙𝑚Δ𝐴𝑓3superscript𝑅2Φsuperscript𝑒𝑟1Δ𝑅delimited-[]32𝑅𝑘subscript𝐻𝑙12𝑘𝑅subscript𝐻𝑙32𝑘𝑅2subscript𝐻𝑙12𝑘𝑅subscriptsuperscript𝑌𝑙𝑚𝑓subscript𝑌𝑙𝑚𝜃𝜑\displaystyle\Phi(e^{r}_{b})=-\Phi(e^{r^{\prime}}_{b})+\sum_{e^{r}}\sum_{f% \supset v|v\subset e^{r}}\sum_{l,m}\frac{\Delta A(f)}{3R^{2}}\Phi(e^{r})\Bigg{% \{}1+\Delta R\bigg{[}-\frac{3}{2R}+k\frac{H_{l-1/2}(kR)-H_{l+3/2}(kR)}{2H_{l+1% /2}(kR)}\bigg{]}Y^{*}_{lm}(f)Y_{lm}(\theta,\varphi)\Bigg{\}},roman_Φ ( italic_e start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) = - roman_Φ ( italic_e start_POSTSUPERSCRIPT italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ) + ∑ start_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_f ⊃ italic_v | italic_v ⊂ italic_e start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_l , italic_m end_POSTSUBSCRIPT divide start_ARG roman_Δ italic_A ( italic_f ) end_ARG start_ARG 3 italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG roman_Φ ( italic_e start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT ) { 1 + roman_Δ italic_R [ - divide start_ARG 3 end_ARG start_ARG 2 italic_R end_ARG + italic_k divide start_ARG italic_H start_POSTSUBSCRIPT italic_l - 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) - italic_H start_POSTSUBSCRIPT italic_l + 3 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) end_ARG start_ARG 2 italic_H start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) end_ARG ] italic_Y start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( italic_f ) italic_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( italic_θ , italic_φ ) } , (32)

where ebrsubscriptsuperscript𝑒superscript𝑟𝑏e^{r^{\prime}}_{b}italic_e start_POSTSUPERSCRIPT italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT is the other radial edge that shares the boundary vertex with ebrsuperscriptsubscript𝑒𝑏𝑟e_{b}^{r}italic_e start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT. In Eq. (32), the first sum is done over all radial edges ersuperscript𝑒𝑟e^{r}italic_e start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT living in the two layers beneath the boundary surface, while the second sum is done over all faces f𝑓fitalic_f that share a vertex with ersuperscript𝑒𝑟e^{r}italic_e start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT. This concludes our derivation for the open BC applied to a radial edge in the DEC framework. The boundary conditions for edges lying tangential to the surface are discussed in Appendix C.

Refer to caption
Figure 11: Results for the open-boundary modes of a dielectric microsphere. (a) Schematic of the system studied: a sphere of radius Rd=12μsubscript𝑅𝑑12𝜇R_{d}=12\muitalic_R start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 12 italic_μm. (b) The lowest singular value of 𝕄v(k)subscript𝕄𝑣𝑘\mathbb{M}_{v}(k)blackboard_M start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_k ) as a function of k𝑘kitalic_k. (c) Error of the lowest eigenvalue as a function of the largest ordering of Bessel terms. (d) Error of the lowest eigenvalue as a function of the boundary radius R𝑅Ritalic_R.

We apply the formulation developed here to calculate the modes for a three-dimensional dielectric microsphere surrounded by a vacuum. The radius of this dielectric sphere is kept fixed at Rd=12μsubscript𝑅𝑑12𝜇R_{d}=12\muitalic_R start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 12 italic_μm throughout all calculations discussed in this section. The computational boundary is an “imaginary” spherical surface at a finite radius R>Rd𝑅subscript𝑅𝑑R>R_{d}italic_R > italic_R start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT, as schematically shown in Fig. 11a. The results of the eigenvalue search found through singular value calculations are shown in Fig. 11b, where we have chosen R=24μ𝑅24𝜇R=24\muitalic_R = 24 italic_μm. With a primal mesh of 2900absent2900\approx 2900≈ 2900 vertices and 18000absent18000\approx 18000≈ 18000 edges, the local minima of the smallest singular value are of order 107superscript10710^{-7}10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT, proving them to be reliable indicators of the eigenvalues of the system. The dielectric microsphere is chosen to demonstrate our formulation because there exist semi-analytical solutions that can be obtained by solving the transcendental equation arising from enforcing the continuity of the field and its normal derivative at the dielectric-vacuum interface. We utilize these solutions to investigate the efficacy of our method in a number of ways. First, the numerical implementation of the series expansion in Eq. (V.3) requires a truncation in the number of terms considered. We, therefore, investigate the dependence of convergence rate on the number of Bessel functions included while keeping the mesh size fixed. The results can be seen in Fig. 11c, where we compute the error of the numerically obtained lowest eigenvalue with the semi-analytical solution. With a manageable number of Lmax=10subscript𝐿𝑚𝑎𝑥10L_{max}=10italic_L start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT = 10, where Lmaxsubscript𝐿𝑚𝑎𝑥L_{max}italic_L start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT is the largest ordering of Bessel terms in the expansion, the error reduces to below 4%percent44\%4 %. Another important factor is how large the computational boundary needs to be in order to achieve precise solutions since accuracy is expected to improve as the boundary is moved further away from where the devices are concentrated, but doing so also increases mesh complexity, and hence the computational burden is exacerbated. In Fig. 11d, the convergence as a function of boundary radius R𝑅Ritalic_R of the lowest eigenvalue is plotted. We see that even with a tight boundary when R=16μ𝑅16𝜇R=16\muitalic_R = 16 italic_μm (while the dielectric sphere is kept at 12μ12𝜇12\mu12 italic_μm) the error is relatively low at 15%percent1515\%15 %, and for R=2Rd𝑅2subscript𝑅𝑑R=2R_{d}italic_R = 2 italic_R start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT it is reduced to <4%absentpercent4<4\%< 4 %. These results verify that our coarse-grained implementation of open boundaries using VSH is able to produce accurate solutions using a boundary of reasonable size.

The formulation based on VSH expansions introduced here allows for effective calculations of open modes while also removing the requirement of divergence-less fields. However, unlike Green’s boundary integral formulation, which is adaptable to arbitrarily-shaped boundaries, this approach needs to be implemented on a spherical surface. This makes the formulation most ideal when applied to three-dimensional structures that have comparable lengths along three directions, such as 3D cavities. In some other specialized cases, this requirement may limit us from drawing the tightest boundary possible, which in turn might cause the computational domain to be larger than it needs to be. An example could be a coplanar waveguide whose length is much larger than its planar width, which in turn is much larger than the thickness along the cross-section of the waveguide. In such a case, replacing a spherical boundary with an ellipsoidal one is preferable since an ellipsoid has three tuning parameters that allow squeezing and stretching in three directions. Fortunately, the series expansion-based approach discussed here is extendable to ellipsoidal coordinates. This relies on the fact that the Helmholtz equation is separable in ellipsoidal coordinates as well, which allows for the expansion of functions, now in terms of ellipsoidal waves. Although the mathematically intensive discussion on how to generalize our method to ellipsoidal coordinates is beyond the scope of this article, in Appendix D, we provide a preamble by discussing a brief proof of the separability of the Helmholtz equation.

VI Conclusion

In this article, we have introduced a spectral theory for the modeling of mode hybridization and relaxation rates in general systems composed of superconducting and dielectric materials. In doing so, we have demonstrated the following important points:

(1) We show that the coarse-grained formulation of electrodynamics using discrete exterior calculus is adaptable to both simplicial and complex meshing strategies with equal precision. Although based on the calculations of coarse-grained quantities, the numerical scheme can still display the high sensitivity to small changes in the geometry and topology of the system that is warranted for instance in the categorization of degenerate modes arising from hidden symmetries when a small perturbation is introduced.

(2) The method, which maps the problem of modeling vector fields to the calculation of averaged projections living on discrete edges, is particularly effective when applied to systems composed of objects of multiple spatial scales, such as those used in superconducting electronic circuits. Since the method keeps track of the average variables, it allows for more efficient meshing strategies that take into account the sizes of the Josephson junctions as well as the finite resolution imposed by a given measurement apparatus. This is an important advantage that this formulation offers because, here, the mesh does not need to be finer than the size of JJs or the measurement apparatus. Future work will focus on adaptive meshers that adapt to the finite spatio-temporal resolution provided by the measurement apparatus interrogating a particular dynamics, allowing for the efficient and accurate simulation of only what is measurable.

(3) We introduce two implementations of open boundaries for the vector wave equation that are applicable to a wide range of electromagnetic systems. By drawing “imaginary,” transparent boundaries that are reasonably sized, we are able to faithfully produce the open modes that agree well with analytical solutions. We also demonstrate the flexibility of the formulation by applying it to the calculation of radiative modes of a superconducting qubit.

The spectral theory presented here is suitable for the second quantization of the electromagnetic field and the time-dependent numerical simulation of the non-linear dynamics of open superconducting systems with arbitrary complexity. We expect the ability to accurately and efficiently extract relaxation rates, as well as the quantification of mode hybridization demonstrated here to be useful in the modeling and optimization of superconducting circuits.

VII Acknowledgements

We gratefully acknowledge discussions with Nicholas Bronn, Thomas G. McConkey, Anil N. Hirani, Thomas Maldonado, Zoe Zager, and Haley M. Cole. We are grateful to Benjamin Lienhard and Wentao Fan for giving the paper their critical read and for the insightful comments. We acknowledge support from the US Department of Energy, Office of Basic Energy Sciences, Division of Materials Sciences and Engineering, under Award No. DESC0016011. The simulations presented in this article were performed on computational resources managed and supported by Princeton Research Computing, a consortium of groups including the Princeton Institute for Computational Science and Engineering (PICSciE) and the Office of Information Technology’s High-Performance Computing Center and Visualization Laboratory at Princeton University.

Appendix A Linearization of the EHDS equations

In this Section, we provide a derivation for the linearization of the EHDS equations which leads to the spectral problem analyzed in this article. We start with the gauge-invariant form of the EHDS equations (Eq. (II) and (5) in the main text) and imposing that charge conservation is individually respected by both the charged condensate ρ𝜌\rhoitalic_ρ and the source ρsrcsubscript𝜌𝑠𝑟𝑐\rho_{src}italic_ρ start_POSTSUBSCRIPT italic_s italic_r italic_c end_POSTSUBSCRIPT, meaning that the last two terms in Eq. (5) cancel each other. Consider a time-harmonic source current 𝐉src=𝐉0eiωtsubscript𝐉𝑠𝑟𝑐subscript𝐉0superscript𝑒𝑖𝜔𝑡{\mathbf{J}}_{src}={\mathbf{J}_{0}}e^{i\omega t}bold_J start_POSTSUBSCRIPT italic_s italic_r italic_c end_POSTSUBSCRIPT = bold_J start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_ω italic_t end_POSTSUPERSCRIPT that, to first order in frequency, results in fluctuations in the gauge-invariant field 𝓐=𝓐𝟎eiωt𝓐subscript𝓐0superscript𝑒𝑖𝜔𝑡\bm{\mathcal{A}}=\bm{\mathcal{A}_{0}}e^{i\omega t}bold_caligraphic_A = bold_caligraphic_A start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_ω italic_t end_POSTSUPERSCRIPT. According to Eq. (5), this in turns leads to a fluctuating condensate

ρ=ρ0+δρ0eiωt𝜌subscript𝜌0𝛿subscript𝜌0superscript𝑒𝑖𝜔𝑡\rho=\rho_{0}+\delta\rho_{0}e^{i\omega t}italic_ρ = italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_δ italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_i italic_ω italic_t end_POSTSUPERSCRIPT (33)

where δρ0ρ0much-less-than𝛿subscript𝜌0subscript𝜌0\delta\rho_{0}\ll\rho_{0}italic_δ italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≪ italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. While the quadratic term t|𝓐|2subscript𝑡superscript𝓐2\partial_{t}\nabla|\bm{\mathcal{A}}|^{2}∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ∇ | bold_caligraphic_A | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT does not have a first-order contribution and can be neglected, the last term on the lhs of Eq. (II), which is related to the rate of change of the quantum pressure, needs careful consideration

μ0ϵ022mqt[2(ρ)ρ]subscript𝜇0subscriptitalic-ϵ0superscriptPlanck-constant-over-2-pi22𝑚𝑞𝑡superscript2𝜌𝜌\displaystyle\frac{\mu_{0}\epsilon_{0}\hbar^{2}}{2mq}\frac{\partial}{\partial t% }\nabla\bigg{[}\frac{\nabla^{2}(\sqrt{\rho})}{\sqrt{\rho}}\bigg{]}divide start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m italic_q end_ARG divide start_ARG ∂ end_ARG start_ARG ∂ italic_t end_ARG ∇ [ divide start_ARG ∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( square-root start_ARG italic_ρ end_ARG ) end_ARG start_ARG square-root start_ARG italic_ρ end_ARG end_ARG ] =μ0ϵ024mq[(1ρ2+|ρ|2ρ32ρρ2ρρ)ρt]absentsubscript𝜇0subscriptitalic-ϵ0superscriptPlanck-constant-over-2-pi24𝑚𝑞1𝜌superscript2superscript𝜌2superscript𝜌3superscript2𝜌superscript𝜌2𝜌𝜌𝜌𝑡\displaystyle=\frac{\mu_{0}\epsilon_{0}\hbar^{2}}{4mq}\nabla\left[\left(\frac{% 1}{\rho}\nabla^{2}+\frac{|\nabla\rho|^{2}}{\rho^{3}}-\frac{\nabla^{2}\rho}{% \rho^{2}}-\frac{\nabla\rho\cdot\nabla}{\rho}\right)\frac{\partial\rho}{% \partial t}\right]= divide start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_m italic_q end_ARG ∇ [ ( divide start_ARG 1 end_ARG start_ARG italic_ρ end_ARG ∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG | ∇ italic_ρ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ρ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG - divide start_ARG ∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ρ end_ARG start_ARG italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - divide start_ARG ∇ italic_ρ ⋅ ∇ end_ARG start_ARG italic_ρ end_ARG ) divide start_ARG ∂ italic_ρ end_ARG start_ARG ∂ italic_t end_ARG ]
=μ0ϵ024m22𝓐,absentsubscript𝜇0subscriptitalic-ϵ0superscriptPlanck-constant-over-2-pi24superscript𝑚2superscript2𝓐\displaystyle=\frac{\mu_{0}\epsilon_{0}\hbar^{2}}{4m^{2}}\nabla\nabla^{2}% \nabla\!\cdot\!\bm{\mathcal{A}},= divide start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_ℏ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∇ ∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∇ ⋅ bold_caligraphic_A , (34)

where in the last line of Eq. (A) we have used the ansatz (33) for ρ𝜌\rhoitalic_ρ with ρ0subscript𝜌0\rho_{0}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT being a step-wise constant function, used Eq. (5) to replace the time derivative tρsubscript𝑡𝜌\partial_{t}\rho∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_ρ, and only kept the leading-order terms. For transverse excitations, the operation on 𝓐𝓐\bm{\mathcal{A}}bold_caligraphic_A in Eq. (A) vanishes, resulting in the inhomogeneous source-field equation

××𝓐𝟎+(1λL2(𝐫)n2(𝐫)k2)𝓐𝟎=𝐉𝟎,subscript𝓐01superscriptsubscript𝜆𝐿2𝐫superscript𝑛2𝐫superscript𝑘2subscript𝓐0subscript𝐉0\displaystyle{\mathbf{\nabla}}\times{\mathbf{\nabla}}\times\bm{{\mathcal{A}}_{% 0}}+\bigg{(}\frac{1}{\lambda_{L}^{2}({\mathbf{r}})}-n^{2}({\mathbf{r}})k^{2}% \bigg{)}\bm{{\mathcal{A}}_{0}}=\mathbf{J_{0}},∇ × ∇ × bold_caligraphic_A start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT + ( divide start_ARG 1 end_ARG start_ARG italic_λ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_r ) end_ARG - italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_r ) italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) bold_caligraphic_A start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT = bold_J start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT , (35)

where λL(𝐫)=mμ0q2ρ0(𝐫)subscript𝜆𝐿𝐫𝑚subscript𝜇0superscript𝑞2subscript𝜌0𝐫\lambda_{L}({\mathbf{r}})=\sqrt{\frac{m}{\mu_{0}q^{2}\rho_{0}({\mathbf{r}})}}italic_λ start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT ( bold_r ) = square-root start_ARG divide start_ARG italic_m end_ARG start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_q start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_r ) end_ARG end_ARG is the London penetration depth, n(𝐫)=ϵ~(𝐫)𝑛𝐫~italic-ϵ𝐫n({\mathbf{r}})=\sqrt{\tilde{\epsilon}({\mathbf{r}})}italic_n ( bold_r ) = square-root start_ARG over~ start_ARG italic_ϵ end_ARG ( bold_r ) end_ARG, and k=μ0ϵ0ω2𝑘subscript𝜇0subscriptitalic-ϵ0superscript𝜔2k=\mu_{0}\epsilon_{0}\omega^{2}italic_k = italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

Appendix B Derivation of Green’s boundary integrals

In this Section, we show the derivations for the integral forms of the boundary fields that obey the scalar and vector Helmholtz equations as were shown in Eqs. (16) and (18), respectively.

Refer to caption
Figure 12: Sketches of how deformations are done to evaluate diverging integrals containing the Green’s function. (a) A deformation on the boundary to include the boundary point. (b) The insertion of a hollow region internal to ΓisubscriptΓ𝑖\Gamma_{i}roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT surrounding the point of interest. (c) Deformation of the boundary to exclude the boundary point of interest.

Consider a homogeneous domain ΓisubscriptΓ𝑖\Gamma_{i}roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT bounded by the boundary ΓisubscriptΓ𝑖\partial\Gamma_{i}∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, the Green’s function is the solution of the impulse response of scalar Helmholtz operator and is defined by

[2+ni2k2]G(𝐫,𝐫,k)=δ(𝐫𝐫),delimited-[]superscript2superscriptsubscript𝑛𝑖2superscript𝑘2𝐺𝐫superscript𝐫𝑘𝛿𝐫superscript𝐫\big{[}\nabla^{2}+n_{i}^{2}k^{2}\big{]}G({\mathbf{r}},{\mathbf{r}^{\prime}},k)% =\delta({\mathbf{r}}-{\mathbf{r}^{\prime}}),[ ∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] italic_G ( bold_r , bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_k ) = italic_δ ( bold_r - bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) , (36)

where δ(𝐫𝐫)𝛿𝐫superscript𝐫\delta({\mathbf{r}}\!-\!{\mathbf{r}^{\prime}})italic_δ ( bold_r - bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) is a dlimit-from𝑑d-italic_d -dimensional Dirac δlimit-from𝛿\delta-italic_δ -function. Multiplying Eq. (13) by G(𝐫,𝐫,k)𝐺𝐫superscript𝐫𝑘G({\mathbf{r}},{\mathbf{r}^{\prime}},k)italic_G ( bold_r , bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_k ) and Eq. (36) by ϕ(𝐫)italic-ϕ𝐫\phi({\mathbf{r}})italic_ϕ ( bold_r ), then applying Green’s theorem gives

ϕ(𝐫)=Γi[ϕ(𝐫)G(𝐫,𝐫,k)G(𝐫,𝐫,k)ϕ(𝐫)]𝐝𝐬,italic-ϕsuperscript𝐫subscriptsubscriptΓ𝑖delimited-[]italic-ϕ𝐫𝐺𝐫superscript𝐫𝑘𝐺𝐫superscript𝐫𝑘italic-ϕ𝐫𝐝𝐬\phi({\bf r^{\prime}})=\int_{\partial\Gamma_{i}}\!\!\big{[}\phi({\bf r})\nabla G% ({\bf r,r^{\prime}},k)-G({\bf r,r^{\prime}},k)\nabla\phi({\bf r})\big{]}\cdot% \mathbf{ds},italic_ϕ ( bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = ∫ start_POSTSUBSCRIPT ∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_ϕ ( bold_r ) ∇ italic_G ( bold_r , bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_k ) - italic_G ( bold_r , bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_k ) ∇ italic_ϕ ( bold_r ) ] ⋅ bold_ds , (37)

for 𝐫superscript𝐫{\mathbf{r}^{\prime}}bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT strictly inside the domain boundary ΓisubscriptΓ𝑖\partial\Gamma_{i}∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. As we take the limit 𝐫Γisuperscript𝐫subscriptΓ𝑖{\mathbf{r}^{\prime}}\rightarrow\partial\Gamma_{i}bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT → ∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, both G(𝐫,𝐫,k)𝐺𝐫superscript𝐫𝑘G({\bf r,r^{\prime}},k)italic_G ( bold_r , bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_k ) and its normal gradient diverges. However, we can show for the 3D case that Eq. (37) is still integrable. A proof for 2D can be found in Ref. 66. We first consider an infinitesimal half-spherical deformation C𝐶Citalic_C of the boundary at 𝐫superscript𝐫{\mathbf{r}^{\prime}}bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT such that the surface now avoids and encloses the singular point inside the domain ΓisubscriptΓ𝑖\Gamma_{i}roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT as seen Fig. (12a). Let the radius of the deformation be ϵitalic-ϵ\epsilonitalic_ϵ, then at the limit ϵ0italic-ϵ0\epsilon\rightarrow 0italic_ϵ → 0 the integral of second term in Eq. (37) over C𝐶Citalic_C vanishes

limϵ0subscriptitalic-ϵ0\displaystyle\lim_{\epsilon\rightarrow 0}roman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT CG(𝐫,𝐫,k)ϕ(𝐫)𝐝𝐬subscript𝐶𝐺𝐫superscript𝐫𝑘italic-ϕ𝐫𝐝𝐬\displaystyle\int_{C}\!\!G({\bf r,r^{\prime}},k)\nabla\phi({\bf r})\cdot% \mathbf{ds}∫ start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT italic_G ( bold_r , bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_k ) ∇ italic_ϕ ( bold_r ) ⋅ bold_ds
=limϵ0ϕn|𝐫14π(1ϵ+inik+)2πϵ2=0,\displaystyle=\lim_{\epsilon\rightarrow 0}-\frac{\partial\phi}{\partial n}% \bigg{\rvert}_{\mathbf{r}^{\prime}}\frac{1}{4\pi}\bigg{(}\frac{1}{\epsilon}+in% _{i}k+...\bigg{)}2\pi\epsilon^{2}=0,= roman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT - divide start_ARG ∂ italic_ϕ end_ARG start_ARG ∂ italic_n end_ARG | start_POSTSUBSCRIPT bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG 4 italic_π end_ARG ( divide start_ARG 1 end_ARG start_ARG italic_ϵ end_ARG + italic_i italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_k + … ) 2 italic_π italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 0 , (38)

while the integral of the first term over C𝐶Citalic_C gives

limϵ0CϕGn𝑑ssubscriptitalic-ϵ0subscript𝐶italic-ϕ𝐺𝑛differential-d𝑠\displaystyle\lim_{\epsilon\rightarrow 0}\int_{C}\!\!\phi\frac{\partial G}{% \partial n}dsroman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT italic_ϕ divide start_ARG ∂ italic_G end_ARG start_ARG ∂ italic_n end_ARG italic_d italic_s =limϵ0ϕ(𝐫)einikϵ4π(inikϵ1ϵ2)2πϵ2absentsubscriptitalic-ϵ0italic-ϕsuperscript𝐫superscript𝑒𝑖subscript𝑛𝑖𝑘italic-ϵ4𝜋𝑖subscript𝑛𝑖𝑘italic-ϵ1superscriptitalic-ϵ22𝜋superscriptitalic-ϵ2\displaystyle=\lim_{\epsilon\rightarrow 0}-\phi({\mathbf{r}^{\prime}})\frac{e^% {in_{i}k\epsilon}}{4\pi}\bigg{(}\frac{in_{i}k}{\epsilon}-\frac{1}{\epsilon^{2}% }\bigg{)}2\pi\epsilon^{2}= roman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT - italic_ϕ ( bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) divide start_ARG italic_e start_POSTSUPERSCRIPT italic_i italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_k italic_ϵ end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_π end_ARG ( divide start_ARG italic_i italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_k end_ARG start_ARG italic_ϵ end_ARG - divide start_ARG 1 end_ARG start_ARG italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) 2 italic_π italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
=12ϕ(𝐫),absent12italic-ϕsuperscript𝐫\displaystyle=\frac{1}{2}\phi({\mathbf{r}^{\prime}}),= divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ϕ ( bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) , (39)

where in the maneuvers above we have used the series expansion of the Green’s function, which reads

G(𝐫,𝐫,k)=einik|𝐫𝐫|4π|𝐫𝐫|=(1|𝐫𝐫|+inik+).𝐺𝐫superscript𝐫𝑘superscript𝑒𝑖subscript𝑛𝑖𝑘𝐫superscript𝐫4𝜋𝐫superscript𝐫1𝐫superscript𝐫𝑖subscript𝑛𝑖𝑘G({\bf r,r^{\prime}},k)=-\frac{e^{in_{i}k|{\mathbf{r}}-{\mathbf{r}^{\prime}}|}% }{4\pi|{\mathbf{r}}-{\mathbf{r}^{\prime}}|}=-\bigg{(}\frac{1}{|{\mathbf{r}}-{% \mathbf{r}^{\prime}}|}+in_{i}k+...\bigg{)}.italic_G ( bold_r , bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_k ) = - divide start_ARG italic_e start_POSTSUPERSCRIPT italic_i italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_k | bold_r - bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_π | bold_r - bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | end_ARG = - ( divide start_ARG 1 end_ARG start_ARG | bold_r - bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | end_ARG + italic_i italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_k + … ) . (40)

From the results in (B) and (B) it follows that the integral in Eq. (37) when 𝐫Γisuperscript𝐫subscriptΓ𝑖{\mathbf{r}^{\prime}}\in\partial\Gamma_{i}bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ ∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT evaluated over the entire boundary surface is indeed the form shown in Eq. (16).

For the vector case, we start by introducing the vector analog of Green’s identity

Γi(𝐮××𝐯𝐯××𝐮)𝑑V=Γi(𝐯××𝐮𝐮××𝐯)𝐝𝐬,subscriptsubscriptΓ𝑖𝐮𝐯𝐯𝐮differential-d𝑉subscriptsubscriptΓ𝑖𝐯𝐮𝐮𝐯𝐝𝐬\displaystyle\int_{\Gamma_{i}}\big{(}{\mathbf{u}}\cdot{\mathbf{\nabla}}\!% \times\!{\mathbf{\nabla}}\!\times\!{\mathbf{v}}-{\mathbf{v}}\cdot{\mathbf{% \nabla}}\!\times\!{\mathbf{\nabla}}\!\times\!{\mathbf{u}}\big{)}dV=\int_{% \partial\Gamma_{i}}\big{(}{\mathbf{v}}\!\times\!\nabla\!\times\!{\mathbf{u}}-{% \mathbf{u}}\!\times\!\nabla\!\times\!{\mathbf{v}}\big{)}\cdot\mathbf{ds},∫ start_POSTSUBSCRIPT roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( bold_u ⋅ ∇ × ∇ × bold_v - bold_v ⋅ ∇ × ∇ × bold_u ) italic_d italic_V = ∫ start_POSTSUBSCRIPT ∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( bold_v × ∇ × bold_u - bold_u × ∇ × bold_v ) ⋅ bold_ds , (41)

where 𝐮𝐮\mathbf{u}bold_u and 𝐯𝐯\mathbf{v}bold_v are vector fields. We now choose 𝐯=𝓐(𝐫)𝐯𝓐𝐫{\mathbf{v}}=\bm{\mathcal{A}}(\mathbf{r})bold_v = bold_caligraphic_A ( bold_r ), our field of interest, and 𝐮=𝐆G(𝐫,𝐫,k)ν^𝐮𝐆𝐺𝐫superscript𝐫𝑘^𝜈{\mathbf{u}}={\mathbf{G}}\equiv G(\mathbf{r},\mathbf{r^{\prime}},k)\hat{\nu}bold_u = bold_G ≡ italic_G ( bold_r , bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_k ) over^ start_ARG italic_ν end_ARG, where ν^^𝜈\hat{\nu}over^ start_ARG italic_ν end_ARG is a unit vector that has a randomly selected, but constant, orientation. For a point 𝐫superscript𝐫{\mathbf{r}^{\prime}}bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT outside of ΓisubscriptΓ𝑖\Gamma_{i}roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, Eq. (41) reduces to

Γi(ν^G)(𝓐)𝑑VsubscriptsubscriptΓ𝑖^𝜈𝐺𝓐differential-d𝑉\displaystyle\int_{\Gamma_{i}}\!(\hat{\nu}\!\cdot\!\nabla G)(\nabla\!\cdot\!% \bm{\mathcal{A}})dV∫ start_POSTSUBSCRIPT roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( over^ start_ARG italic_ν end_ARG ⋅ ∇ italic_G ) ( ∇ ⋅ bold_caligraphic_A ) italic_d italic_V =Γi{(ν^G)(𝓐n^)+𝓐×(G×ν^)𝐆××𝓐}𝐝𝐬absentsubscriptsubscriptΓ𝑖^𝜈𝐺𝓐^𝑛𝓐𝐺^𝜈𝐆𝓐𝐝𝐬\displaystyle=\int_{\partial\Gamma_{i}}\!\!\Big{\{}(\hat{\nu}\!\cdot\!\nabla G% )(\bm{\mathcal{A}}\!\cdot\!\hat{n})+\bm{\mathcal{A}}\!\times\!(\nabla G\!% \times\!\hat{\nu})-\mathbf{G}\!\times\!{\mathbf{\nabla}}\!\times\!\bm{\mathcal% {A}}\Big{\}}\cdot\mathbf{ds}= ∫ start_POSTSUBSCRIPT ∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT { ( over^ start_ARG italic_ν end_ARG ⋅ ∇ italic_G ) ( bold_caligraphic_A ⋅ over^ start_ARG italic_n end_ARG ) + bold_caligraphic_A × ( ∇ italic_G × over^ start_ARG italic_ν end_ARG ) - bold_G × ∇ × bold_caligraphic_A } ⋅ bold_ds (42)
=Γi{(ν^G)(𝓐n^)+(ν^𝓐)(Gn^)(ν^n^)(G𝓐)𝐆[(×𝓐)×n^]}𝑑s,absentsubscriptsubscriptΓ𝑖^𝜈𝐺𝓐^𝑛^𝜈𝓐𝐺^𝑛^𝜈^𝑛𝐺𝓐𝐆delimited-[]𝓐^𝑛differential-d𝑠\displaystyle=\int_{\partial\Gamma_{i}}\!\!\Big{\{}(\hat{\nu}\!\cdot\!\nabla G% )(\bm{\mathcal{A}}\!\cdot\!\hat{n})+(\hat{\nu}\!\cdot\!\bm{\mathcal{A}})(% \nabla G\!\cdot\!\hat{n})-(\hat{\nu}\!\cdot\!\hat{n})(\nabla G\!\cdot\!\bm{% \mathcal{A}})-\mathbf{G}\!\cdot\!\big{[}({\mathbf{\nabla}}\!\times\!\bm{% \mathcal{A}})\!\times\!\hat{n}\big{]}\Big{\}}ds,= ∫ start_POSTSUBSCRIPT ∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT { ( over^ start_ARG italic_ν end_ARG ⋅ ∇ italic_G ) ( bold_caligraphic_A ⋅ over^ start_ARG italic_n end_ARG ) + ( over^ start_ARG italic_ν end_ARG ⋅ bold_caligraphic_A ) ( ∇ italic_G ⋅ over^ start_ARG italic_n end_ARG ) - ( over^ start_ARG italic_ν end_ARG ⋅ over^ start_ARG italic_n end_ARG ) ( ∇ italic_G ⋅ bold_caligraphic_A ) - bold_G ⋅ [ ( ∇ × bold_caligraphic_A ) × over^ start_ARG italic_n end_ARG ] } italic_d italic_s ,

where n^^𝑛\hat{n}over^ start_ARG italic_n end_ARG is the unit normal vector on ΓisubscriptΓ𝑖\partial\Gamma_{i}∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Since ν^^𝜈\hat{\nu}over^ start_ARG italic_ν end_ARG is a common factor in all the terms, it can be dropped and we arrive at

ΓiG(𝓐)𝑑V=Γi{G(𝓐n^)+G×(𝓐×n^)G[(×𝓐)×n^]}𝑑s.subscriptsubscriptΓ𝑖𝐺𝓐differential-d𝑉subscriptsubscriptΓ𝑖𝐺𝓐^𝑛𝐺𝓐^𝑛𝐺delimited-[]𝓐^𝑛differential-d𝑠\displaystyle\int_{\Gamma_{i}}\!\nabla G(\nabla\!\cdot\!\bm{\mathcal{A}})dV=% \int_{\partial\Gamma_{i}}\!\!\Big{\{}\nabla G(\bm{\mathcal{A}}\cdot\hat{n})+% \nabla G\!\times\!(\bm{\mathcal{A}}\!\times\!\hat{n})-G\big{[}({\mathbf{\nabla% }}\!\times\!\bm{\mathcal{A}})\!\times\!\hat{n}\big{]}\Big{\}}ds.∫ start_POSTSUBSCRIPT roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∇ italic_G ( ∇ ⋅ bold_caligraphic_A ) italic_d italic_V = ∫ start_POSTSUBSCRIPT ∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT { ∇ italic_G ( bold_caligraphic_A ⋅ over^ start_ARG italic_n end_ARG ) + ∇ italic_G × ( bold_caligraphic_A × over^ start_ARG italic_n end_ARG ) - italic_G [ ( ∇ × bold_caligraphic_A ) × over^ start_ARG italic_n end_ARG ] } italic_d italic_s . (43)

We are interested in, however, the evaluation of the field inside ΓisubscriptΓ𝑖\Gamma_{i}roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. To proceed, consider when 𝐫superscript𝐫{\mathbf{r}^{\prime}}bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT is inside the material region enclosed by ΓisubscriptΓ𝑖\partial\Gamma_{i}∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, but there exists an infinitesimally small spherical hole S𝑆Sitalic_S of radius ϵitalic-ϵ\epsilonitalic_ϵ centered around 𝐫superscript𝐫{\mathbf{r}^{\prime}}bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT. The boundary of ΓisubscriptΓ𝑖\Gamma_{i}roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT now consists of the original outer surface ΓisubscriptΓ𝑖\partial\Gamma_{i}∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and the boundary of the sphere, as seen in Fig. (12b). At the limit ϵ0italic-ϵ0\epsilon\rightarrow 0italic_ϵ → 0, each term in Eq. (43) evaluated on the sphere gives

limϵ0SG[(×𝓐)×n^]𝑑ssubscriptitalic-ϵ0subscript𝑆𝐺delimited-[]𝓐^𝑛differential-d𝑠\displaystyle\lim_{\epsilon\rightarrow 0}\int_{S}G\big{[}({\mathbf{\nabla}}\!% \times\!\bm{\mathcal{A}})\!\times\!\hat{n}\big{]}dsroman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT italic_G [ ( ∇ × bold_caligraphic_A ) × over^ start_ARG italic_n end_ARG ] italic_d italic_s =limϵ0einjkϵ4πϵ[(×𝓐)×n^]|𝐫4πϵ2\displaystyle=\lim_{\epsilon\rightarrow 0}-\frac{e^{in_{j}k\epsilon}}{4\pi% \epsilon}\big{[}(\nabla\!\times\!\bm{\mathcal{A}})\!\times\!\hat{n}\big{]}% \bigg{\rvert}_{\mathbf{r}^{\prime}}\!4\pi\epsilon^{2}= roman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT - divide start_ARG italic_e start_POSTSUPERSCRIPT italic_i italic_n start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_k italic_ϵ end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_π italic_ϵ end_ARG [ ( ∇ × bold_caligraphic_A ) × over^ start_ARG italic_n end_ARG ] | start_POSTSUBSCRIPT bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT 4 italic_π italic_ϵ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
=0,absent0\displaystyle=0,= 0 , (44)

also

limϵ0S{G(𝓐n^)+G×(𝓐×n^)}𝑑ssubscriptitalic-ϵ0subscript𝑆𝐺𝓐^𝑛𝐺𝓐^𝑛differential-d𝑠\displaystyle\lim_{\epsilon\rightarrow 0}\int_{S}\!\!\Big{\{}\nabla G(\bm{% \mathcal{A}}\cdot\hat{n})+\nabla G\!\times\!(\bm{\mathcal{A}}\!\times\!\hat{n}% )\Big{\}}dsroman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT { ∇ italic_G ( bold_caligraphic_A ⋅ over^ start_ARG italic_n end_ARG ) + ∇ italic_G × ( bold_caligraphic_A × over^ start_ARG italic_n end_ARG ) } italic_d italic_s =limϵ0S(1ϵinik)einikϵ4πϵ𝓐𝑑sabsentsubscriptitalic-ϵ0subscript𝑆1italic-ϵ𝑖subscript𝑛𝑖𝑘superscript𝑒𝑖subscript𝑛𝑖𝑘italic-ϵ4𝜋italic-ϵ𝓐differential-d𝑠\displaystyle=\lim_{\epsilon\rightarrow 0}\int_{S}\Big{(}\frac{1}{\epsilon}-in% _{i}k\Big{)}\frac{e^{in_{i}k\epsilon}}{4\pi\epsilon}\bm{\mathcal{A}}ds= roman_lim start_POSTSUBSCRIPT italic_ϵ → 0 end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ( divide start_ARG 1 end_ARG start_ARG italic_ϵ end_ARG - italic_i italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_k ) divide start_ARG italic_e start_POSTSUPERSCRIPT italic_i italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_k italic_ϵ end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_π italic_ϵ end_ARG bold_caligraphic_A italic_d italic_s
=𝓐(𝐫).absent𝓐superscript𝐫\displaystyle=\bm{\mathcal{A}}{(\mathbf{r}^{\prime}}).= bold_caligraphic_A ( bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) . (45)

Therefore, the integral form of the field at 𝐫superscript𝐫{\mathbf{r}^{\prime}}bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT inside ΓisubscriptΓ𝑖\Gamma_{i}roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is given by

𝓐(𝐫)=Γi{G[(×𝓐)×n^]G(𝓐n^)G×(𝐀×n^)}+ΓiG(𝓐)𝑑V,𝓐superscript𝐫subscriptsubscriptΓ𝑖𝐺delimited-[]𝓐^𝑛𝐺𝓐^𝑛𝐺𝐀^𝑛subscriptsubscriptΓ𝑖𝐺𝓐differential-d𝑉\displaystyle\bm{\mathcal{A}}({\mathbf{r}^{\prime}})=\int_{\partial\Gamma_{i}}% \!\!\bigg{\{}G\big{[}({\mathbf{\nabla}}\times\bm{\mathcal{A}})\times\hat{n}% \big{]}-\nabla G(\bm{\mathcal{A}}\cdot\hat{n})-\nabla G\times({\mathbf{A}% \times\hat{n}})\bigg{\}}+\int_{\Gamma_{i}}\!\nabla G({\mathbf{\nabla}}\cdot\bm% {\mathcal{A}})dV,bold_caligraphic_A ( bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = ∫ start_POSTSUBSCRIPT ∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT { italic_G [ ( ∇ × bold_caligraphic_A ) × over^ start_ARG italic_n end_ARG ] - ∇ italic_G ( bold_caligraphic_A ⋅ over^ start_ARG italic_n end_ARG ) - ∇ italic_G × ( bold_A × over^ start_ARG italic_n end_ARG ) } + ∫ start_POSTSUBSCRIPT roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∇ italic_G ( ∇ ⋅ bold_caligraphic_A ) italic_d italic_V , (46)

where the volume term on the right hand side is an improper integral. However, this term vanishes if the field is divergence-free, in which case the field at anywhere strictly inside ΓisubscriptΓ𝑖\Gamma_{i}roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is determined by the field on the boundary ΓisubscriptΓ𝑖\partial\Gamma_{i}∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT.

Finally, to find the integral form of the field on the boundary, at 𝐫Γisuperscript𝐫subscriptΓ𝑖{\mathbf{r}^{\prime}}\in\partial\Gamma_{i}bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ∈ ∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, we consider a half-spherical deformation of ΓisubscriptΓ𝑖\partial\Gamma_{i}∂ roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT such that the boundary avoids 𝐫superscript𝐫{\mathbf{r}^{\prime}}bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT and leaves it outside of the enclosed domain ΓisubscriptΓ𝑖\Gamma_{i}roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, as shown in Fig. (12c). Note that this deformation is in the opposite direction as in the scalar case discussed earlier, where 𝐫superscript𝐫{\mathbf{r}^{\prime}}bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT was included into ΓisubscriptΓ𝑖\Gamma_{i}roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. By following a similar limiting procedure as was laid out above for 𝐫superscript𝐫{\mathbf{r}^{\prime}}bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT strictly inside ΓisubscriptΓ𝑖\Gamma_{i}roman_Γ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT but now applied to a half-sphere, we arrive at the self-consistent integral form for the boundary field shown in Eq. (18).

Appendix C Coarse-grained open boundary conditions using VSH

In this Section, we derive the boundary conditions for the angular components of the field 𝓐(𝐫)𝓐𝐫\bm{\mathcal{A}}({\mathbf{r}})bold_caligraphic_A ( bold_r ) satisfying the vector Helmholtz equation. We also discuss the DEC implementation of such conditions on the tangential boundary edges.

The series expansion of the 𝓐(1)superscript𝓐1\bm{\mathcal{A}}^{(1)}bold_caligraphic_A start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT component is

𝓐(1)(r,θ,φ)superscript𝓐1𝑟𝜃𝜑\displaystyle\bm{\mathcal{A}}^{(1)}(r,\theta,\varphi)bold_caligraphic_A start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_r , italic_θ , italic_φ ) =l,ma~lm1𝒜lm(1)𝚿lm(r,θ,φ)absentsubscript𝑙𝑚superscriptsubscript~𝑎𝑙𝑚1superscriptsubscript𝒜𝑙𝑚1subscript𝚿𝑙𝑚𝑟𝜃𝜑\displaystyle=\sum_{l,m}\tilde{a}_{lm}^{1}\mathcal{A}_{lm}^{(1)}{\mathbf{\Psi}% }_{lm}(r,\theta,\varphi)= ∑ start_POSTSUBSCRIPT italic_l , italic_m end_POSTSUBSCRIPT over~ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT caligraphic_A start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT bold_Ψ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( italic_r , italic_θ , italic_φ )
=a~lm12l(l+1)[Hl+1/2(kr)r3/2+kr(Hl1/2(kr)Hl+3/2(kr))]𝚿lm(r,θ,φ).absentsubscript~𝑎𝑙𝑚12𝑙𝑙1delimited-[]subscript𝐻𝑙12𝑘𝑟superscript𝑟32𝑘𝑟subscript𝐻𝑙12𝑘𝑟subscript𝐻𝑙32𝑘𝑟subscript𝚿𝑙𝑚𝑟𝜃𝜑\displaystyle=\tilde{a}_{lm}\frac{1}{2l(l+1)}\bigg{[}\frac{H_{l+1/2}(kr)}{r^{3% /2}}+\frac{k}{\sqrt{r}}\big{(}H_{l-1/2}(kr)-H_{l+3/2}(kr)\big{)}\bigg{]}{% \mathbf{\Psi}}_{lm}(r,\theta,\varphi).= over~ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG 2 italic_l ( italic_l + 1 ) end_ARG [ divide start_ARG italic_H start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_k end_ARG start_ARG square-root start_ARG italic_r end_ARG end_ARG ( italic_H start_POSTSUBSCRIPT italic_l - 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) - italic_H start_POSTSUBSCRIPT italic_l + 3 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) ) ] bold_Ψ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( italic_r , italic_θ , italic_φ ) .

The coefficients alm1(R)subscriptsuperscript𝑎1𝑙𝑚𝑅a^{1}_{lm}(R)italic_a start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( italic_R ) in the expansion of the field on the boundary surface are then given by

alm1(R)subscriptsuperscript𝑎1𝑙𝑚𝑅\displaystyle a^{1}_{lm}(R)italic_a start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( italic_R ) =a~lm1𝒜lm(1)(R)12l(l+1)[Hl+1/2(kr)r3/2+kr(Hl1/2(kr)Hl+3/2(kr))]absentsuperscriptsubscript~𝑎𝑙𝑚1superscriptsubscript𝒜𝑙𝑚1𝑅12𝑙𝑙1delimited-[]subscript𝐻𝑙12𝑘𝑟superscript𝑟32𝑘𝑟subscript𝐻𝑙12𝑘𝑟subscript𝐻𝑙32𝑘𝑟\displaystyle=\tilde{a}_{lm}^{1}\mathcal{A}_{lm}^{(1)}(R)\frac{1}{2l(l+1)}% \bigg{[}\frac{H_{l+1/2}(kr)}{r^{3/2}}+\frac{k}{\sqrt{r}}\big{(}H_{l-1/2}(kr)-H% _{l+3/2}(kr)\big{)}\bigg{]}= over~ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT caligraphic_A start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_R ) divide start_ARG 1 end_ARG start_ARG 2 italic_l ( italic_l + 1 ) end_ARG [ divide start_ARG italic_H start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_k end_ARG start_ARG square-root start_ARG italic_r end_ARG end_ARG ( italic_H start_POSTSUBSCRIPT italic_l - 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) - italic_H start_POSTSUBSCRIPT italic_l + 3 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) ) ]
=1l(l+1)𝑑Ω𝓐(1)𝚿lm*absent1𝑙𝑙1differential-dΩsuperscript𝓐1subscriptsuperscript𝚿𝑙𝑚\displaystyle=\frac{1}{l(l+1)}\int d\Omega\bm{\mathcal{A}}^{(1)}\cdot{\mathbf{% \Psi}}^{*}_{lm}= divide start_ARG 1 end_ARG start_ARG italic_l ( italic_l + 1 ) end_ARG ∫ italic_d roman_Ω bold_caligraphic_A start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ⋅ bold_Ψ start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT
=1l(l+1)fΔA(f)R2𝓐t(f)𝚿lm*(f),absent1𝑙𝑙1subscript𝑓Δ𝐴𝑓superscript𝑅2subscript𝓐𝑡𝑓superscriptsubscript𝚿𝑙𝑚𝑓\displaystyle=\frac{1}{l(l+1)}\sum_{f}\frac{\Delta A(f)}{R^{2}}\bm{\mathcal{A}% }_{t}(f)\cdot{\mathbf{\Psi}}_{lm}^{*}(f),= divide start_ARG 1 end_ARG start_ARG italic_l ( italic_l + 1 ) end_ARG ∑ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT divide start_ARG roman_Δ italic_A ( italic_f ) end_ARG start_ARG italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG bold_caligraphic_A start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_f ) ⋅ bold_Ψ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ( italic_f ) , (47)

where 𝓐tsubscript𝓐𝑡\bm{\mathcal{A}}_{t}bold_caligraphic_A start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT is the component of 𝓐(𝐫)𝓐𝐫\bm{\mathcal{A}}({\mathbf{r}})bold_caligraphic_A ( bold_r ) tangential to the boundary surface, and the sum is perform over all the triangular faces f𝑓fitalic_f on the boundary. In the last line of Eq. (C) we have utilized the orthogonality between 𝓐(1)superscript𝓐1\bm{\mathcal{A}}^{(1)}bold_caligraphic_A start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT and 𝚿lm*superscriptsubscript𝚿𝑙𝑚{\mathbf{\Psi}}_{lm}^{*}bold_Ψ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT to switch to using 𝓐tsubscript𝓐𝑡\bm{\mathcal{A}}_{t}bold_caligraphic_A start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT instead of 𝓐(1)superscript𝓐1\bm{\mathcal{A}}^{(1)}bold_caligraphic_A start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT in the computation of alm1(R)subscriptsuperscript𝑎1𝑙𝑚𝑅a^{1}_{lm}(R)italic_a start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( italic_R ), since the former is more readily accessible during numerical implementation. The field 𝓐(1)superscript𝓐1\bm{\mathcal{A}}^{(1)}bold_caligraphic_A start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT can be decomposed into azimuthal and polar contributions, for each of which we derive the boundary condition. The radial derivative of the polar component of 𝓐(1)superscript𝓐1\bm{\mathcal{A}}^{(1)}bold_caligraphic_A start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT at the boundary is given by

𝒜θ(1)(R,θ,φ)𝒜θ(1)(RΔR,θ,φ)ΔR=l,ma~lm(1)12l(l+1)[Hl+1/2(kr)r1/2+kr(Hl1/2(kr)Hl+3/2(kr))]|R(Ylm)θ\displaystyle\frac{\mathcal{A}_{\theta}^{(1)}(R,\theta,\varphi)-\mathcal{A}_{% \theta}^{(1)}(R\!-\!\Delta R,\theta,\varphi)}{\Delta R}=\sum_{l,m}\tilde{a}_{% lm}^{(1)}\frac{1}{2l(l+1)}\bigg{[}\frac{H_{l+1/2}(kr)}{r^{1/2}}+k\sqrt{r}\big{% (}H_{l-1/2}(kr)-H_{l+3/2}(kr)\big{)}\bigg{]}^{\prime}\bigg{\rvert}_{R}(\nabla Y% _{lm})_{\theta}divide start_ARG caligraphic_A start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_R , italic_θ , italic_φ ) - caligraphic_A start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_R - roman_Δ italic_R , italic_θ , italic_φ ) end_ARG start_ARG roman_Δ italic_R end_ARG = ∑ start_POSTSUBSCRIPT italic_l , italic_m end_POSTSUBSCRIPT over~ start_ARG italic_a end_ARG start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 italic_l ( italic_l + 1 ) end_ARG [ divide start_ARG italic_H start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT end_ARG + italic_k square-root start_ARG italic_r end_ARG ( italic_H start_POSTSUBSCRIPT italic_l - 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) - italic_H start_POSTSUBSCRIPT italic_l + 3 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) ) ] start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT ( ∇ italic_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT (48)

from which the expression for 𝓐θ(1)superscriptsubscript𝓐𝜃1\bm{\mathcal{A}}_{\theta}^{(1)}bold_caligraphic_A start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT on the boundary surface can be extracted

𝒜θ(1)(R,θ,φ)=l,malm(1)(R)(R+ΔR{1+R[Hl+1/2(kr)r3/2+kr(Hl1/2(kr)Hl+3/2(kr))]|R[Hl+1/2(kR)R3/2+kR(Hl1/2(kR)Hl+3/2(kR))]})(Ylm)θ,\displaystyle\mathcal{A}_{\theta}^{(1)}(R,\theta,\varphi)=\sum_{l,m}a^{(1)}_{% lm}(R)\left(R+\Delta R\left\{1+R\frac{\Big{[}\frac{H_{l+1/2}(kr)}{r^{3/2}}+% \frac{k}{\sqrt{r}}\big{(}H_{l-1/2}(kr)-H_{l+3/2}(kr)\big{)}\Big{]}^{\prime}% \Big{\rvert}_{R}}{\Big{[}\frac{H_{l+1/2}(kR)}{R^{3/2}}+\frac{k}{\sqrt{R}}\big{% (}H_{l-1/2}(kR)-H_{l+3/2}(kR)\big{)}\Big{]}}\right\}\right)(\nabla Y_{lm})_{% \theta},caligraphic_A start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_R , italic_θ , italic_φ ) = ∑ start_POSTSUBSCRIPT italic_l , italic_m end_POSTSUBSCRIPT italic_a start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( italic_R ) ( italic_R + roman_Δ italic_R { 1 + italic_R divide start_ARG [ divide start_ARG italic_H start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_k end_ARG start_ARG square-root start_ARG italic_r end_ARG end_ARG ( italic_H start_POSTSUBSCRIPT italic_l - 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) - italic_H start_POSTSUBSCRIPT italic_l + 3 / 2 end_POSTSUBSCRIPT ( italic_k italic_r ) ) ] start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT | start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT end_ARG start_ARG [ divide start_ARG italic_H start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) end_ARG start_ARG italic_R start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_k end_ARG start_ARG square-root start_ARG italic_R end_ARG end_ARG ( italic_H start_POSTSUBSCRIPT italic_l - 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) - italic_H start_POSTSUBSCRIPT italic_l + 3 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) ) ] end_ARG } ) ( ∇ italic_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT , (49)

where (Ylm)θsubscriptsubscript𝑌𝑙𝑚𝜃(\nabla Y_{lm})_{\theta}( ∇ italic_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT is the polar component of Ylmsubscript𝑌𝑙𝑚\nabla Y_{lm}∇ italic_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT, and alm1(R)subscriptsuperscript𝑎1𝑙𝑚𝑅a^{1}_{lm}(R)italic_a start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ( italic_R ) given as in Eq. (C). The expression for the azimuthal component 𝒜φ(1)(R,θ,φ)superscriptsubscript𝒜𝜑1𝑅𝜃𝜑\mathcal{A}_{\varphi}^{(1)}(R,\theta,\varphi)caligraphic_A start_POSTSUBSCRIPT italic_φ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_R , italic_θ , italic_φ ) is exactly similar, with a replacement of θφ𝜃𝜑\theta\rightarrow\varphiitalic_θ → italic_φ in the subscripts in Eq. (49). Similar to the steps for 𝒜(1)superscript𝒜1\mathcal{A}^{(1)}caligraphic_A start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT, the derivation of BC for 𝒜(2)superscript𝒜2\mathcal{A}^{(2)}caligraphic_A start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT starts with the expansion

𝓐(2)=l.ma~lm(2)π2kRHl+1/2(kR)𝚽l.m(r,θ,φ)superscript𝓐2subscriptformulae-sequence𝑙𝑚subscriptsuperscript~𝑎2𝑙𝑚𝜋2𝑘𝑅subscript𝐻𝑙12𝑘𝑅subscript𝚽formulae-sequence𝑙𝑚𝑟𝜃𝜑\displaystyle\bm{\mathcal{A}}^{(2)}=\sum_{l.m}\tilde{a}^{(2)}_{lm}\frac{\pi}{2% kR}H_{l+1/2}(kR){\mathbf{\Phi}}_{l.m}(r,\theta,\varphi)bold_caligraphic_A start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_l . italic_m end_POSTSUBSCRIPT over~ start_ARG italic_a end_ARG start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 2 italic_k italic_R end_ARG italic_H start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) bold_Φ start_POSTSUBSCRIPT italic_l . italic_m end_POSTSUBSCRIPT ( italic_r , italic_θ , italic_φ ) (50)

from which the coefficients alm(2)subscriptsuperscript𝑎2𝑙𝑚a^{(2)}_{lm}italic_a start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT can be computed

alm(2)(R)superscriptsubscript𝑎𝑙𝑚2𝑅\displaystyle a_{lm}^{(2)}(R)italic_a start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( italic_R ) =a~lm(2)π2kRHl+1/2(kR)absentsubscriptsuperscript~𝑎2𝑙𝑚𝜋2𝑘𝑅subscript𝐻𝑙12𝑘𝑅\displaystyle=\tilde{a}^{(2)}_{lm}\frac{\pi}{2kR}H_{l+1/2}(kR)= over~ start_ARG italic_a end_ARG start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT divide start_ARG italic_π end_ARG start_ARG 2 italic_k italic_R end_ARG italic_H start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_R )
=1l(l+1)𝑑Ω𝓐(2)𝚽lm*absent1𝑙𝑙1differential-dΩsuperscript𝓐2superscriptsubscript𝚽𝑙𝑚\displaystyle=\frac{1}{l(l+1)}\int d\Omega\bm{\mathcal{A}}^{(2)}\cdot{\mathbf{% \Phi}}_{lm}^{*}= divide start_ARG 1 end_ARG start_ARG italic_l ( italic_l + 1 ) end_ARG ∫ italic_d roman_Ω bold_caligraphic_A start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ⋅ bold_Φ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT
=1l(l+1)fΔA(f)R2𝓐t(f)𝚽lm*(f).absent1𝑙𝑙1subscript𝑓Δ𝐴𝑓superscript𝑅2subscript𝓐𝑡𝑓superscriptsubscript𝚽𝑙𝑚𝑓\displaystyle=\frac{1}{l(l+1)}\sum_{f}\frac{\Delta A(f)}{R^{2}}\bm{\mathcal{A}% }_{t}(f)\cdot{\mathbf{\Phi}}_{lm}^{*}(f).= divide start_ARG 1 end_ARG start_ARG italic_l ( italic_l + 1 ) end_ARG ∑ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT divide start_ARG roman_Δ italic_A ( italic_f ) end_ARG start_ARG italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG bold_caligraphic_A start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_f ) ⋅ bold_Φ start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ( italic_f ) . (51)

We also write the boundary conditions for 𝒜θ(2)superscriptsubscript𝒜𝜃2\mathcal{A}_{\theta}^{(2)}caligraphic_A start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT and 𝒜φ(2)superscriptsubscript𝒜𝜑2\mathcal{A}_{\varphi}^{(2)}caligraphic_A start_POSTSUBSCRIPT italic_φ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT separately using their discrete form of radial derivative at the boundary. This leads to

𝒜θ(2)(R,θ,φ)=l,malm(2)(R){R+ΔR2[1+kRHl1/2(kR)Hl+3/2(kR)Hl+1/2(kR)]}(r^×Ylm)θ,superscriptsubscript𝒜𝜃2𝑅𝜃𝜑subscript𝑙𝑚superscriptsubscript𝑎𝑙𝑚2𝑅𝑅Δ𝑅2delimited-[]1𝑘𝑅subscript𝐻𝑙12𝑘𝑅subscript𝐻𝑙32𝑘𝑅subscript𝐻𝑙12𝑘𝑅subscript^𝑟subscript𝑌𝑙𝑚𝜃\displaystyle\mathcal{A}_{\theta}^{(2)}(R,\theta,\varphi)=\sum_{l,m}a_{lm}^{(2% )}(R)\left\{R+\frac{\Delta R}{2}\left[1+kR\frac{H_{l-1/2}(kR)-H_{l+3/2}(kR)}{H% _{l+1/2}(kR)}\right]\right\}(\hat{r}\times\nabla Y_{lm})_{\theta},caligraphic_A start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( italic_R , italic_θ , italic_φ ) = ∑ start_POSTSUBSCRIPT italic_l , italic_m end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( italic_R ) { italic_R + divide start_ARG roman_Δ italic_R end_ARG start_ARG 2 end_ARG [ 1 + italic_k italic_R divide start_ARG italic_H start_POSTSUBSCRIPT italic_l - 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) - italic_H start_POSTSUBSCRIPT italic_l + 3 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) end_ARG start_ARG italic_H start_POSTSUBSCRIPT italic_l + 1 / 2 end_POSTSUBSCRIPT ( italic_k italic_R ) end_ARG ] } ( over^ start_ARG italic_r end_ARG × ∇ italic_Y start_POSTSUBSCRIPT italic_l italic_m end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT , (52)

and by replacing θ𝜃\thetaitalic_θ with φ𝜑\varphiitalic_φ in the subscripts, we obtain the expression for 𝒜φ(2)(R,θ,φ)superscriptsubscript𝒜𝜑2𝑅𝜃𝜑\mathcal{A}_{\varphi}^{(2)}(R,\theta,\varphi)caligraphic_A start_POSTSUBSCRIPT italic_φ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( italic_R , italic_θ , italic_φ ).

The boundary conditions given in Eqs. (49) and (52) are for vector individual components, which are scalars living on vertices of the primal computational mesh. We would like to translate these BCs to fields living the edges that lie on the boundary surface. First of all, the practical computation of the expansion coefficients in Eqs. (C) and (C) requires knowledge about the tangential field 𝓐tsubscript𝓐𝑡\bm{\mathcal{A}}_{t}bold_caligraphic_A start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT defined on the centers of the triangular faces f𝑓fitalic_f that discretize the boundary surface. In DEC, however, vectors are replaced by their projections onto the discrete edges, which are quantities that we have access to. Therefore, at each triangular face, given the edge fields that are on the three sides of the triangle, we need to determine the value of the vector 𝓐tsubscript𝓐𝑡\bm{\mathcal{A}}_{t}bold_caligraphic_A start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT at the circumcenter fsuperscript𝑓f^{\dagger}italic_f start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT. This mapping from primal edge field to dual vector field is done through a sharp (#) operator, whose formal definition is given in Ref. 70. Now given that the coefficient expansions are obtained, for every vertex on the boundary, the polar(azimuthal) component of 𝒜tsubscript𝒜𝑡\mathcal{A}_{t}caligraphic_A start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT is the sum of projections from 𝒜(1)superscript𝒜1\mathcal{A}^{(1)}caligraphic_A start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT and 𝒜(2)superscript𝒜2\mathcal{A}^{(2)}caligraphic_A start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT

𝒜θ(φ)(R,θ,φ)=𝒜θ(φ)(1)(R,θ,φ)+𝒜θ(φ)(2)(R,θ,φ).subscript𝒜𝜃𝜑𝑅𝜃𝜑superscriptsubscript𝒜𝜃𝜑1𝑅𝜃𝜑superscriptsubscript𝒜𝜃𝜑2𝑅𝜃𝜑\mathcal{A}_{\theta(\varphi)}(R,\theta,\varphi)=\mathcal{A}_{\theta(\varphi)}^% {(1)}(R,\theta,\varphi)+\mathcal{A}_{\theta(\varphi)}^{(2)}(R,\theta,\varphi).caligraphic_A start_POSTSUBSCRIPT italic_θ ( italic_φ ) end_POSTSUBSCRIPT ( italic_R , italic_θ , italic_φ ) = caligraphic_A start_POSTSUBSCRIPT italic_θ ( italic_φ ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 1 ) end_POSTSUPERSCRIPT ( italic_R , italic_θ , italic_φ ) + caligraphic_A start_POSTSUBSCRIPT italic_θ ( italic_φ ) end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( 2 ) end_POSTSUPERSCRIPT ( italic_R , italic_θ , italic_φ ) . (53)

For an edge e[v1,v2]𝑒subscript𝑣1subscript𝑣2e[v_{1},v_{2}]italic_e [ italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ] lying on the boundary, where v1subscript𝑣1v_{1}italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and v2subscript𝑣2v_{2}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are the starting and ending vertices of the edge, the edge field is given by

Φ(e[v1,v2])=[𝓐t(v1)+𝓐t(v2)]2(𝐯2𝐯1),Φ𝑒subscript𝑣1subscript𝑣2delimited-[]subscript𝓐𝑡subscript𝑣1subscript𝓐𝑡subscript𝑣22subscript𝐯2subscript𝐯1\Phi(e[v_{1},v_{2}])=\frac{\big{[}\bm{\mathcal{A}}_{t}(v_{1})+\bm{\mathcal{A}}% _{t}(v_{2})\big{]}}{2}\cdot({\mathbf{v}}_{2}-{\mathbf{v}}_{1}),roman_Φ ( italic_e [ italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ] ) = divide start_ARG [ bold_caligraphic_A start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) + bold_caligraphic_A start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ( italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ] end_ARG start_ARG 2 end_ARG ⋅ ( bold_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - bold_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) , (54)

where 𝐯1subscript𝐯1{\mathbf{v}}_{1}bold_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and 𝐯2subscript𝐯2{\mathbf{v}}_{2}bold_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are the locations of the vertices v1subscript𝑣1v_{1}italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and v2subscript𝑣2v_{2}italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, respectively, with the components of 𝓐tsubscript𝓐𝑡\bm{\mathcal{A}}_{t}bold_caligraphic_A start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT on the boundary given in Eq. (53). Eq. (54) is the boundary condition to be imposed on the edges lying tangentially to the boundary surface.

Appendix D Separability of Helmholtz equation in ellipsoidal coordinates

The ellipsoidal coordinate system is based on the equations

x2ξi2a2+y2ξ2b2+z2ξ2c2=1,superscript𝑥2subscriptsuperscript𝜉2𝑖superscript𝑎2superscript𝑦2superscript𝜉2superscript𝑏2superscript𝑧2superscript𝜉2superscript𝑐21\displaystyle\frac{x^{2}}{\xi^{2}_{i}-a^{2}}+\frac{y^{2}}{\xi^{2}-b^{2}}+\frac% {z^{2}}{\xi^{2}-c^{2}}=1,divide start_ARG italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = 1 , (55)

where abc𝑎𝑏𝑐a\geq b\geq citalic_a ≥ italic_b ≥ italic_c, with i=1,2,3𝑖123i=1,2,3italic_i = 1 , 2 , 3 such that

ξ1>a>ξ2>b>ξ3>c.subscript𝜉1𝑎subscript𝜉2𝑏subscript𝜉3𝑐\xi_{1}>a>\xi_{2}>b>\xi_{3}>c.italic_ξ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT > italic_a > italic_ξ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > italic_b > italic_ξ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT > italic_c . (56)

Eqs. (55) represent three families of confocal quadric surfaces sharing the same foci. In the discussion here, to simplify the algebraic manipulations we consider the case where c=0𝑐0c=0italic_c = 0. The relationship between the ellipsoidal coordinates (ξ1,ξ2,ξ3)subscript𝜉1subscript𝜉2subscript𝜉3(\xi_{1},\xi_{2},\xi_{3})( italic_ξ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_ξ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_ξ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) and the Cartesian coordinates are given by

x𝑥\displaystyle xitalic_x =(ξ12a2)(ξ22a2)(ξ32a2)a2(a2b2),absentsuperscriptsubscript𝜉12superscript𝑎2superscriptsubscript𝜉22superscript𝑎2superscriptsubscript𝜉32superscript𝑎2superscript𝑎2superscript𝑎2superscript𝑏2\displaystyle=\sqrt{\frac{(\xi_{1}^{2}-a^{2})(\xi_{2}^{2}-a^{2})(\xi_{3}^{2}-a% ^{2})}{a^{2}(a^{2}-b^{2})}},= square-root start_ARG divide start_ARG ( italic_ξ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_ξ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_ξ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG end_ARG ,
y𝑦\displaystyle yitalic_y =(ξ12b2)(ξ22b2)(ξ32b2)b2(b2a2),absentsuperscriptsubscript𝜉12superscript𝑏2superscriptsubscript𝜉22superscript𝑏2superscriptsubscript𝜉32superscript𝑏2superscript𝑏2superscript𝑏2superscript𝑎2\displaystyle=\sqrt{\frac{(\xi_{1}^{2}-b^{2})(\xi_{2}^{2}-b^{2})(\xi_{3}^{2}-b% ^{2})}{b^{2}(b^{2}-a^{2})}},= square-root start_ARG divide start_ARG ( italic_ξ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_ξ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_ξ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG start_ARG italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG end_ARG , (57)
z𝑧\displaystyle zitalic_z =ξ1ξ2ξ3ab,absentsubscript𝜉1subscript𝜉2subscript𝜉3𝑎𝑏\displaystyle=\frac{\xi_{1}\xi_{2}\xi_{3}}{ab},= divide start_ARG italic_ξ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG start_ARG italic_a italic_b end_ARG ,

with the scale factors being

h1subscript1\displaystyle h_{1}italic_h start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =(ξ12ξ22)(ξ12ξ32)(ξ12a2)(ξ12b2)absentsuperscriptsubscript𝜉12superscriptsubscript𝜉22superscriptsubscript𝜉12superscriptsubscript𝜉32superscriptsubscript𝜉12superscript𝑎2superscriptsubscript𝜉12superscript𝑏2\displaystyle=\sqrt{\frac{(\xi_{1}^{2}-\xi_{2}^{2})(\xi_{1}^{2}-\xi_{3}^{2})}{% (\xi_{1}^{2}-a^{2})(\xi_{1}^{2}-b^{2})}}= square-root start_ARG divide start_ARG ( italic_ξ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ξ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_ξ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ξ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG start_ARG ( italic_ξ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_ξ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG end_ARG
h2subscript2\displaystyle h_{2}italic_h start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT =(ξ22ξ12)(ξ22ξ32)(ξ22a2)(ξ22b2).absentsuperscriptsubscript𝜉22superscriptsubscript𝜉12superscriptsubscript𝜉22superscriptsubscript𝜉32superscriptsubscript𝜉22superscript𝑎2superscriptsubscript𝜉22superscript𝑏2\displaystyle=\sqrt{\frac{(\xi_{2}^{2}-\xi_{1}^{2})(\xi_{2}^{2}-\xi_{3}^{2})}{% (\xi_{2}^{2}-a^{2})(\xi_{2}^{2}-b^{2})}}.= square-root start_ARG divide start_ARG ( italic_ξ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ξ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_ξ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ξ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG start_ARG ( italic_ξ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_ξ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG end_ARG . (58)

It is known that the Helmholtz equation (Eq. (13) in the main text) is separable in eleven three-dimensional coordinate systems, with the ellipsoidal coordinates being the most general of them and the remaining ten - including the spherical coordinates used in the main text - are derived from it through limiting processes [71]. The separability in ellipsoidal coordinates can be shown by using the ansatz

ϕ(ξ1,ξ2,ξ3)=ψ1(ξ1)ψ2(ξ2)ψ3(ξ3)italic-ϕsubscript𝜉1subscript𝜉2subscript𝜉3subscript𝜓1subscript𝜉1subscript𝜓2subscript𝜉2subscript𝜓3subscript𝜉3\phi(\xi_{1},\xi_{2},\xi_{3})=\psi_{1}(\xi_{1})\psi_{2}(\xi_{2})\psi_{3}(\xi_{% 3})italic_ϕ ( italic_ξ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_ξ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_ξ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) = italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_ξ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_ξ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) italic_ψ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_ξ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) (59)

for the field in Eq. (13), resulting in the separated ODEs of the form

4f(κi)ddκi(f(κi)dψidκi+(λ1+λ2κi+k2κi2))ψi=0,4𝑓subscript𝜅𝑖𝑑𝑑subscript𝜅𝑖𝑓subscript𝜅𝑖𝑑subscript𝜓𝑖𝑑subscript𝜅𝑖subscript𝜆1subscript𝜆2subscript𝜅𝑖superscript𝑘2superscriptsubscript𝜅𝑖2subscript𝜓𝑖0\displaystyle 4\sqrt{f(\kappa_{i})}\frac{d}{d\kappa_{i}}\!\left(\!\sqrt{f(% \kappa_{i})}\frac{d\psi_{i}}{d\kappa_{i}}+(\lambda_{1}+\lambda_{2}\kappa_{i}+k% ^{2}\kappa_{i}^{2})\!\right)\psi_{i}=0,4 square-root start_ARG italic_f ( italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) end_ARG divide start_ARG italic_d end_ARG start_ARG italic_d italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ( square-root start_ARG italic_f ( italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) end_ARG divide start_ARG italic_d italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG + ( italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ) italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 , (60)

where f(κi)=(κia2)(κi2b2)κi𝑓subscript𝜅𝑖subscript𝜅𝑖superscript𝑎2superscriptsubscript𝜅𝑖2superscript𝑏2subscript𝜅𝑖f(\kappa_{i})=\sqrt{(\kappa_{i}-a^{2})(\kappa_{i}^{2}-b^{2})\kappa_{i}}italic_f ( italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) = square-root start_ARG ( italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ( italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG, with κi=ξ2subscript𝜅𝑖superscript𝜉2\kappa_{i}=\xi^{2}italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Upon performing a change of variables

κi=b2sn2(α,b2a2),subscript𝜅𝑖superscript𝑏2𝑠superscript𝑛2𝛼superscript𝑏2superscript𝑎2\kappa_{i}=b^{2}sn^{2}\left(\alpha,\frac{b^{2}}{a^{2}}\right),italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_s italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_α , divide start_ARG italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (61)

where sn(u,m)𝑠𝑛𝑢𝑚sn(u,m)italic_s italic_n ( italic_u , italic_m ) is a Jacobi elliptic function, we can write

f(κi)=b4a2sn2(α,b2a2)dn2(α,b2a2)cn2(α,b2a2),𝑓subscript𝜅𝑖superscript𝑏4superscript𝑎2𝑠superscript𝑛2𝛼superscript𝑏2superscript𝑎2𝑑superscript𝑛2𝛼superscript𝑏2superscript𝑎2𝑐superscript𝑛2𝛼superscript𝑏2superscript𝑎2\displaystyle f(\kappa_{i})=b^{4}a^{2}sn^{2}\left(\alpha,\frac{b^{2}}{a^{2}}% \right)dn^{2}\left(\alpha,\frac{b^{2}}{a^{2}}\right)cn^{2}\left(\alpha,\frac{b% ^{2}}{a^{2}}\right),italic_f ( italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) = italic_b start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_s italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_α , divide start_ARG italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) italic_d italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_α , divide start_ARG italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) italic_c italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_α , divide start_ARG italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (62)

with dn(u,m)𝑑𝑛𝑢𝑚dn(u,m)italic_d italic_n ( italic_u , italic_m ) and cn(u,m)𝑐𝑛𝑢𝑚cn(u,m)italic_c italic_n ( italic_u , italic_m ) also Jacobi elliptic functions. Eq. (60) can then be written in terms of the new variable as

d2ψidα2+[λ1a2+λ2a2sn2(α,b2a2)+k2a2sn4(α,b2a2)]ψi=0.superscript𝑑2subscript𝜓𝑖𝑑superscript𝛼2delimited-[]subscript𝜆1superscript𝑎2subscript𝜆2superscript𝑎2𝑠superscript𝑛2𝛼superscript𝑏2superscript𝑎2superscript𝑘2superscript𝑎2𝑠superscript𝑛4𝛼superscript𝑏2superscript𝑎2subscript𝜓𝑖0\frac{d^{2}\psi_{i}}{d\alpha^{2}}+\left[\frac{\lambda_{1}}{a^{2}}+\frac{% \lambda_{2}}{a^{2}}sn^{2}\left(\alpha,\frac{b^{2}}{a^{2}}\right)+\frac{k^{2}}{% a^{2}}sn^{4}\left(\alpha,\frac{b^{2}}{a^{2}}\right)\right]\psi_{i}=0.divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + [ divide start_ARG italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_s italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_α , divide start_ARG italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) + divide start_ARG italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_s italic_n start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ( italic_α , divide start_ARG italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) ] italic_ψ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0 . (63)

Eq. (63) is the ellipsoidal wave equation whose solutions are known as ellipsoidal wave functions. With the Helmholtz equation being separable in ellipsoidal coordinates with known solutions, the generalization to this coordinate system of our spherical harmonics method for open systems is therefore straightforward.

References

  • Houck et al. [2012] A. A. Houck, H. E. Türeci, and J. Koch, On-chip quantum simulation with superconducting circuits, Nature Physics 8, 292 (2012).
  • Carusotto et al. [2020] I. Carusotto, A. A. Houck, A. J. Kollár, P. Roushan, D. I. Schuster, and J. Simon, Photonic materials in circuit quantum electrodynamics, Nature Physics 16, 268 (2020).
  • Hatridge et al. [2011] M. Hatridge, R. Vijay, D. Slichter, J. Clarke, and I. Siddiqi, Dispersive magnetometry with a quantum limited squid parametric amplifier, Physical Review B 83, 134501 (2011).
  • Macklin et al. [2015] C. Macklin, K. O’brien, D. Hover, M. Schwartz, V. Bolkhovsky, X. Zhang, W. Oliver, and I. Siddiqi, A near–quantum-limited josephson traveling-wave parametric amplifier, Science 350, 307 (2015).
  • Blais et al. [2004] A. Blais, R.-S. Huang, A. Wallraff, S. M. Girvin, and R. J. Schoelkopf, Cavity quantum electrodynamics for superconducting electrical circuits: An architecture for quantum computation, Physical Review A 69, 062320 (2004).
  • Devoret and Schoelkopf [2013] M. H. Devoret and R. J. Schoelkopf, Superconducting circuits for quantum information: an outlook, Science 339, 1169 (2013).
  • Feynman et al. [1965] R. P. Feynman, R. B. Leighton, and M. Sands, The Schrödinger Equation in a Classical Context: A Seminar on Superconductivity, in The Feynman Lectures on Physics, Vol. III (Addison–Wesley, 1965) Chap. 21.
  • Ao et al. [1995] P. Ao, D. J. Thouless, and X.-M. Zhu, Nonlinear Schrödinger equation for superconductors, Modern Physics Letters B 9, 755 (1995).
  • Aitchison et al. [1995] I. J. Aitchison, P. Ao, D. J. Thouless, and X.-M. Zhu, Effective Lagrangians for BCS superconductors at T=0, Physical Review B 51, 6531 (1995).
  • Pham et al. [2023a] D. N. Pham, W. Fan, M. G. Scheer, and H. E. Tureci, Flux-based three-dimensional electrodynamic modeling approach to superconducting circuits and materials, Phys. Rev. A 107, 053704 (2023a).
  • Greiter et al. [1989] M. Greiter, F. Wilczek, and E. Witten, Hydrodynamic relations in superconductivity, Modern Physics Letters B 3, 903 (1989).
  • Fisher et al. [1990] M. P. A. Fisher, G. Grinstein, and S. M. Girvin, Presence of quantum diffusion in two dimensions: Universal resistance at the superconductor-insulator transition, Phys. Rev. Lett. 64, 587 (1990).
  • Salasnich [2009] L. Salasnich, Hydrodynamics of Bose and Fermi superfluids at zero temperature: the superfluid nonlinear Schrödinger equation, Laser physics 19, 642 (2009).
  • Blais et al. [2021] A. Blais, A. L. Grimsmo, S. M. Girvin, and A. Wallraff, Circuit quantum electrodynamics, Rev. Mod. Phys. 93, 025005 (2021).
  • Nigg et al. [2012] S. E. Nigg, H. Paik, B. Vlastakis, G. Kirchmair, S. Shankar, L. Frunzio, M. H. Devoret, R. J. Schoelkopf, and S. M. Girvin, Black-box superconducting circuit quantization, Phys. Revs. Letts. 108, 240502 (2012).
  • Solgun et al. [2014] F. Solgun, D. W. Abraham, and D. P. DiVincenzo, Blackbox quantization of superconducting circuits using exact impedance synthesis, Physical Review B 90, 134504 (2014).
  • Smith et al. [2016] W. C. Smith, A. Kou, U. Vool, I. M. Pop, L. Frunzio, R. J. Schoelkopf, and M. H. Devoret, Quantization of inductively shunted superconducting circuits, Phys. Rev. B 94, 144507 (2016).
  • Malekakhlagh et al. [2017] M. Malekakhlagh, A. Petrescu, and H. E. Türeci, Cutoff-free circuit quantum electrodynamics, Phys. Rev. Lett. 119, 073601 (2017).
  • Parra-Rodriguez et al. [2019] A. Parra-Rodriguez, I. Egusquiza, D. DiVincenzo, and E. Solano, Canonical circuit quantization with linear nonreciprocal devices, Physical Review B 99, 014514 (2019).
  • Minev et al. [2021] Z. K. Minev, Z. Leghtas, S. O. Mundhada, L. Christakis, I. M. Pop, and M. H. Devoret, Energy-participation quantization of josephson circuits, npj Quantum Information 7, 131 (2021).
  • [21] The time-coarse graining toolbox, https://github.com/leonbello/QuantumGraining.jl.
  • Houck et al. [2008] A. Houck, J. Schreier, B. Johnson, J. Chow, J. Koch, J. Gambetta, D. Schuster, L. Frunzio, M. Devoret, S. Girvin, et al., Controlling the spontaneous emission of a superconducting transmon qubit, Physical review letters 101, 080502 (2008).
  • Reed et al. [2010] M. D. Reed, B. R. Johnson, A. A. Houck, L. DiCarlo, J. M. Chow, D. I. Schuster, L. Frunzio, and R. J. Schoelkopf, Fast reset and suppressing spontaneous emission of a superconducting qubit, Applied Physics Letters 96 (2010).
  • Jeffrey et al. [2014] E. Jeffrey, D. Sank, J. Y. Mutus, T. C. White, J. Kelly, R. Barends, Y. Chen, Z. Chen, B. Chiaro, A. Dunsworth, A. Megrant, P. J. J. O’Malley, C. Neill, P. Roushan, A. Vainsencher, J. Wenner, A. N. Cleland, and J. M. Martinis, Fast accurate state measurement with superconducting qubits, Phys. Rev. Lett. 112, 190504 (2014).
  • Bronn et al. [2015] N. T. Bronn, Y. Liu, J. B. Hertzberg, A. D. Córcoles, A. A. Houck, J. M. Gambetta, and J. M. Chow, Broadband filters for abatement of spontaneous emission in circuit quantum electrodynamics, Applied Physics Letters 107 (2015).
  • Boissonneault et al. [2009] M. Boissonneault, J. M. Gambetta, and A. Blais, Dispersive regime of circuit qed: Photon-dependent qubit dephasing and relaxation rates, Phys. Rev. A 79, 013819 (2009).
  • Slichter et al. [2012] D. H. Slichter, R. Vijay, S. J. Weber, S. Boutin, M. Boissonneault, J. M. Gambetta, A. Blais, and I. Siddiqi, Measurement-induced qubit state mixing in circuit qed from up-converted dephasing noise, Phys. Rev. Lett. 109, 153601 (2012).
  • Mundhada et al. [2016] S. Mundhada, S. Shankar, A. Narla, E. Zalys-Geller, S. Girvin, and M. Devoret, Dependence of transmon qubit relaxation rate on readout drive power, APS March Meeting V48 (2016).
  • Minev et al. [2019] Z. Minev, S. Mundhada, S. Shankar, P. Reinhold, R. Gutiérrez-Jáuregui, R. Schoelkopf, M. Mirrahimi, H. Carmichael, and M. Devoret, To catch and reverse a quantum jump mid-flight, Nature 57010.1038/s41586-019-1287-z (2019).
  • Petrescu et al. [2020] A. Petrescu, M. Malekakhlagh, and H. E. Türeci, Lifetime renormalization of driven weakly anharmonic superconducting qubits. ii. the readout problem, Phys. Rev. B 101, 134510 (2020).
  • Hanai et al. [2021] R. Hanai, A. McDonald, and A. Clerk, Intrinsic mechanisms for drive-dependent purcell decay in superconducting quantum circuits, Phys. Rev. Res. 3, 043228 (2021).
  • Gusenkova et al. [2021] D. Gusenkova, M. Spiecker, R. Gebauer, M. Willsch, D. Willsch, F. Valenti, N. Karcher, L. Grünhaupt, I. Takmakov, P. Winkel, D. Rieger, A. V. Ustinov, N. Roch, W. Wernsdorfer, K. Michielsen, O. Sander, and I. M. Pop, Quantum nondemolition dispersive readout of a superconducting artificial atom using large photon numbers, Phys. Rev. Appl. 15, 064030 (2021).
  • Shillito et al. [2022] R. Shillito, A. Petrescu, J. Cohen, J. Beall, M. Hauru, M. Ganahl, A. G. Lewis, G. Vidal, and A. Blais, Dynamics of transmon ionization, Phys. Rev. Appl. 18, 034031 (2022).
  • Cohen et al. [2023] J. Cohen, A. Petrescu, R. Shillito, and A. Blais, Reminiscence of classical chaos in driven transmons, PRX Quantum 4, 020312 (2023).
  • Caldeira and Leggett [1983] A. Caldeira and A. Leggett, Quantum tunnelling in a dissipative system, Annals of Physics 149, 374 (1983).
  • Filipp et al. [2011] S. Filipp, M. Göppl, J. M. Fink, M. Baur, R. Bianchetti, L. Steffen, and A. Wallraff, Multimode mediated qubit-qubit coupling and dark-state symmetries in circuit quantum electrodynamics, Phys. Rev. A 83, 063827 (2011).
  • Gely et al. [2017] M. F. Gely, A. Parra-Rodriguez, D. Bothner, Y. M. Blanter, S. J. Bosman, E. Solano, and G. A. Steele, Convergence of the multimode quantum rabi model of circuit quantum electrodynamics, Phys. Rev. B 95, 245115 (2017).
  • Sinha et al. [2022] K. Sinha, S. A. Khan, E. Cüce, and H. E. Türeci, Radiative properties of an artificial atom coupled to a josephson-junction array, Phys. Rev. A 106, 033714 (2022).
  • Yoshihara et al. [2014] F. Yoshihara, Y. Nakamura, F. Yan, S. Gustavsson, J. Bylander, W. D. Oliver, and J.-S. Tsai, Flux qubit noise spectroscopy using rabi oscillations under strong driving conditions, Physical Review B 89, 020503 (2014).
  • Kumar et al. [2016] P. Kumar, S. Sendelbach, M. Beck, J. Freeland, Z. Wang, H. Wang, C. Y. Clare, R. Wu, D. Pappas, and R. McDermott, Origin and reduction of 1/f magnetic flux noise in superconducting devices, Physical Review Applied 6, 041001 (2016).
  • Wilen et al. [2021] C. D. Wilen, S. Abdullah, N. Kurinsky, C. Stanford, L. Cardani, G. d’Imperio, C. Tomei, L. Faoro, L. Ioffe, C. Liu, et al., Correlated charge noise and relaxation errors in superconducting qubits, Nature 594, 369 (2021).
  • Pop et al. [2014] I. M. Pop, K. Geerlings, G. Catelani, R. J. Schoelkopf, L. I. Glazman, and M. H. Devoret, Coherent suppression of electromagnetic dissipation due to superconducting quasiparticles, Nature 508, 369 (2014).
  • Frattini et al. [2018] N. Frattini, V. Sivak, A. Lingenfelter, S. Shankar, and M. Devoret, Optimizing the nonlinearity and dissipation of a snail parametric amplifier for dynamic range, Physical Review Applied 10, 054020 (2018).
  • Wenner et al. [2011] J. Wenner, M. Neeley, R. C. Bialczak, M. Lenander, E. Lucero, A. D. O’Connell, D. Sank, H. Wang, M. Weides, A. N. Cleland, and J. M. Martinis, Wirebond crosstalk and cavity modes in large chip mounts for superconducting qubits, Supercond. Sci. Technol. 24, 065001 (2011).
  • Huang et al. [2021] S. Huang, B. Lienhard, G. Calusine, A. Vepsäläinen, J. Braumüller, D. K. Kim, A. J. Melville, B. M. Niedzielski, J. L. Yoder, B. Kannan, et al., Microwave package design for superconducting quantum processors, PRX Quantum 2, 020306 (2021).
  • Berenger [1994] J.-P. Berenger, A perfectly matched layer for the absorption of electromagnetic waves, Journal of computational physics 114, 185 (1994).
  • Berenger [1996] J.-P. Berenger, Three-dimensional perfectly matched layer for the absorption of electromagnetic waves, Journal of computational physics 127, 363 (1996).
  • Clayton and Engquist [1977] R. Clayton and B. Engquist, Absorbing boundary conditions for acoustic and elastic wave equations, Bulletin of the seismological society of America 67, 1529 (1977).
  • Higdon [1987] R. L. Higdon, Numerical absorbing boundary conditions for the wave equation, Mathematics of Computation 49, 65 (1987).
  • Gell-Mann and E. [1954] M. Gell-Mann and L. F. E., Quantum electrodynamics at small distances, Phys. Rev. 95, 1300 (1954).
  • Qi et al. [2018] B. Qi, L. Zhang, and L. Ge, Defect states emerging from a non-hermitian flatband of photonic zero modes, Physical review letters 120, 093901 (2018).
  • El-Ganainy et al. [2018] R. El-Ganainy, K. G. Makris, M. Khajavikhan, Z. H. Musslimani, S. Rotter, and D. N. Christodoulides, Non-hermitian physics and pt symmetry, Nature Physics 14, 11 (2018).
  • Gigli et al. [2020] C. Gigli, T. Wu, G. Marino, A. Borne, G. Leo, and P. Lalanne, Quasinormal-mode non-hermitian modeling and design in nonlinear nano-optics, ACS photonics 7, 1197 (2020).
  • Purcell et al. [1946] E. M. Purcell, H. C. Torrey, and R. V. Pound, Resonance absorption by nuclear magnetic moments in a solid, Physical review 69, 37 (1946).
  • Malekakhlagh et al. [2016] M. Malekakhlagh, A. Petrescu, and H. E. Türeci, Non-markovian dynamics of a superconducting qubit in an open multimode resonator, Physical Review A 94, 063848 (2016).
  • Bosman et al. [2017] S. J. Bosman, M. F. Gely, V. Singh, A. Bruno, D. Bothner, and G. A. Steele, Multi-mode ultra-strong coupling in circuit quantum electrodynamics, npj Quantum Information 3, 46 (2017).
  • Puertas Martínez et al. [2019] J. Puertas Martínez, S. Léger, N. Gheeraert, R. Dassonneville, L. Planat, F. Foroughi, Y. Krupko, O. Buisson, C. Naud, W. Hasch-Guichard, et al., A tunable josephson platform to explore many-body quantum optics in circuit-qed, npj Quantum Information 5, 19 (2019).
  • Roth and Chew [2021] T. E. Roth and W. C. Chew, Macroscopic circuit quantum electrodynamics: A new look toward developing full-wave numerical models, IEEE Journal on Multiscale and Multiphysics Computational Techniques 6, 109 (2021).
  • Chew et al. [2016] W. C. Chew, A. Y. Liu, C. Salazar-Lazaro, and W. E. Sha, Quantum electromagnetics: A new look—part ii, IEEE Journal on Multiscale and Multiphysics Computational Techniques 1, 85 (2016).
  • Türeci et al. [2008] H. E. Türeci, L. Ge, S. Rotter, and A. D. Stone, Strong interactions in multimode random lasers, Science 320, 643 (2008).
  • Wiersig [2002] J. Wiersig, Boundary element method for resonances in dielectric microcavities, Journal of Optics A: Pure and Applied Optics 5, 53 (2002).
  • [62] DEC-QED toolbox, https://github.com/dnpham23/DEC-QED.
  • London and London [1935] F. London and H. London, The electromagnetic equations of the supraconductor, Proceedings of the Royal Society of London. Series A - Mathematical and Physical Sciences 149, 71 (1935).
  • Shaw [1974] G. B. Shaw, Degeneracy in the particle-in-a-box problem, Journal of Physics A: Mathematical, Nuclear and General 7, 1537 (1974).
  • Josephson [1965] B. D. Josephson, Supercurrents through barriers, Advances in Physics 14, 419 (1965).
  • Kosztin and Schulten [1997] I. Kosztin and K. Schulten, Boundary integral method for stationary states of two-dimensional quantum systems, International Journal of Modern Physics C 8, 293 (1997).
  • Pham et al. [2023b] D. N. Pham, S. Bharadwaj, S. Rodriguez, L. Rodriguez, and L. R. Ram-Mohan, High-accuracy calculation of singular electromagnetic fields in regions with re-entrant peripheries, Journal of Applied Physics 134, 153102 (2023b).
  • Bäcker [2003] A. Bäcker, Numerical aspects of eigenvalue and eigenfunction computations for chaotic quantum systems, in The Mathematical Aspects of Quantum Maps, edited by M. D. Esposti and S. Graffi (Springer Berlin Heidelberg, Berlin, Heidelberg, 2003) pp. 91–144.
  • Barrera et al. [1985] R. G. Barrera, G. Estevez, and J. Giraldo, Vector spherical harmonics and their application to magnetostatics, European Journal of Physics 6, 287 (1985).
  • Hirani [2003] A. N. Hirani, Discrete Exterior Calculus, Ph.D. thesis, California Institute of Technology (2003).
  • Morse and Feshbach [1954] P. M. Morse and H. Feshbach, Methods of theoretical physics, American Journal of Physics 22, 410 (1954).