9
Review Fluctuational electrodynamics calculations of near-field heat transfer in non-planar geometries: A brief overview Clayton R. Otey a , Linxiao Zhu a , Sunil Sandhu b , Shanhui Fan b,n a Department of Applied Physics, Stanford University, Stanford, CA 94305, USA b Department of Electrical Engineering, Stanford University, Stanford, CA 94305, USA article info Article history: Received 20 September 2012 Received in revised form 13 March 2013 Accepted 16 April 2013 Available online 3 May 2013 Keywords: Near field Radiative Heat transfer Computational Scattering Boundary element FDTD abstract Near-field electromagnetic heat transfer is of interest for a variety of applications, including energy conversion, and precision heating, cooling and imaging of nano- structures. This past decade has seen considerable progress in the study of near-field electromagnetic heat transfer, but it is only very recently that numerically exact methods have been developed for treating near-field heat transfer in the fluctuational electro- dynamics formalism for non-trivial geometries. In this paper we provide a tutorial review of these exact methods, with an emphasis on the computational aspects of three important methods, which we compare in the context of a canonical example, the coupled dielectric sphere problem. & 2013 Elsevier Ltd. All rights reserved. Contents 1. Introduction ............................................................................................. 3 2. Partial-wave scattering method ............................................................................. 5 2.1. Green function picture .............................................................................. 5 2.2. Field correlator picture .............................................................................. 6 3. Boundary element method ................................................................................. 7 3.1. Green function picture .............................................................................. 7 3.2. Field correlator picture .............................................................................. 7 4. Finite difference time domain method ........................................................................ 8 5. Comparison of computational approaches ..................................................................... 9 6. Conclusion ............................................................................................. 10 Acknowledgment ........................................................................................ 10 Appendix A Electromagnetic operator notation ................................................................ 10 References ............................................................................................. 10 1. Introduction Heat will flow between bodies at different tempera- tures. Among the mechanisms of heat transfer, radiative electromagnetic heat transfer is special in that it may Contents lists available at ScienceDirect journal homepage: www.elsevier.com/locate/jqsrt Journal of Quantitative Spectroscopy & Radiative Transfer 0022-4073/$ - see front matter & 2013 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.jqsrt.2013.04.017 n Corresponding author. E-mail addresses: [email protected], [email protected] (C.R. Otey), [email protected] (L. Zhu), [email protected] (S. Fan). Journal of Quantitative Spectroscopy & Radiative Transfer 132 (2014) 311

Fluctuational electrodynamics calculations of near …web.stanford.edu/group/fan/publication/Otey_JQSRT_132_3...C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Fluctuational electrodynamics calculations of near …web.stanford.edu/group/fan/publication/Otey_JQSRT_132_3...C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative

Contents lists available at ScienceDirect

Journal of Quantitative Spectroscopy &Radiative Transfer

Journal of Quantitative Spectroscopy & Radiative Transfer 132 (2014) 3–11

0022-40http://d

n CorrE-m

clayton.shanhu

journal homepage: www.elsevier.com/locate/jqsrt

Review

Fluctuational electrodynamics calculations of near-field heattransfer in non-planar geometries: A brief overview

Clayton R. Otey a, Linxiao Zhu a, Sunil Sandhu b, Shanhui Fan b,n

a Department of Applied Physics, Stanford University, Stanford, CA 94305, USAb Department of Electrical Engineering, Stanford University, Stanford, CA 94305, USA

a r t i c l e i n f o

Article history:Received 20 September 2012Received in revised form13 March 2013Accepted 16 April 2013Available online 3 May 2013

Keywords:Near fieldRadiativeHeat transferComputationalScatteringBoundary elementFDTD

73/$ - see front matter & 2013 Elsevier Ltd.x.doi.org/10.1016/j.jqsrt.2013.04.017

esponding author.ail addresses: [email protected],[email protected] (C.R. Otey), linxiao@[email protected] (S. Fan).

a b s t r a c t

Near-field electromagnetic heat transfer is of interest for a variety of applications,including energy conversion, and precision heating, cooling and imaging of nano-structures. This past decade has seen considerable progress in the study of near-fieldelectromagnetic heat transfer, but it is only very recently that numerically exact methodshave been developed for treating near-field heat transfer in the fluctuational electro-dynamics formalism for non-trivial geometries. In this paper we provide a tutorial reviewof these exact methods, with an emphasis on the computational aspects of threeimportant methods, which we compare in the context of a canonical example, thecoupled dielectric sphere problem.

& 2013 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32. Partial-wave scattering method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2.1. Green function picture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52.2. Field correlator picture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

3. Boundary element method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73.1. Green function picture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73.2. Field correlator picture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

4. Finite difference time domain method. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85. Comparison of computational approaches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10Appendix A Electromagnetic operator notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

All rights reserved.

d.edu (L. Zhu),

1. Introduction

Heat will flow between bodies at different tempera-tures. Among the mechanisms of heat transfer, radiativeelectromagnetic heat transfer is special in that it may

Page 2: Fluctuational electrodynamics calculations of near …web.stanford.edu/group/fan/publication/Otey_JQSRT_132_3...C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative

C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 132 (2014) 3–114

transmit heat through vacuum. When certain classes ofdielectric bodies are in close proximity, so that theirseparation falls below the characteristic thermal wave-length λT ¼ hc=kBT , we observe the near-field regime,where the heat flow can be enhanced beyond the con-straint of the Planck law that governs far-field heattransfer. Such an enhancement has been demonstrated ina number of recent experiments [1–4], leading to theprospect of applying near-field heat transfer in areasincluding non-contact radiative cooling [5], thermal ima-ging [6,7], thermal circuit elements [8–11], as well asenergy conversion and thermo-photo-voltaics [12–16].

These experimental developments have in turn moti-vated theoretical studies of the mathematical and computa-tional underpinnings of near-field heat transfer. What hasemerged is a clear need for exact methods to calculateelectromagnetic heat transfer in the fluctuational electrody-namics formalism [17]. It is important to state up front thatthe quantum electrodynamics formalism yields the sameresults, provided the thermodynamic interpretation is appro-priate; in particular it must be assumed that the sources arein local equilibrium with a thermal reservoir [18].

Using the fluctuational electrodynamics formalism, we cancalculate heat flux between two bodies separated by a vacuumgap, as shown in Fig. 1a. The problem is completely specifiedby the temperature distribution TðrÞ and relative dielectricfunction of the materials, ϵðr;ωÞ ¼ ϵ′ðr;ωÞ þ iϵ″ðr;ωÞ. We workwith the time-harmonic form of Maxwell's equations, withimplicit ∝e−iωt time-dependent factors, so a positive imagin-ary part of the dielectric functions, ϵ″ðr;ωÞ, corresponds to

Fig. 1. (a) Near-field heat transfer between two non-planar bodies V1 and V2 witemperatures T1 and T2, respectively. (b) Schematic of a fluctuational electrodcorrelations in V1 are specified by the fluctuation–dissipation theorem, and arcorrelations, and thus the ensemble-averaged Poynting vector on the boundary

material loss. We limit our discussion to the case of homo-genous dielectric bodies, so that ϵðr;ωÞ ¼ ϵnðωÞ in the variousregions n∈0;1;2. Free space corresponds to n¼0, and isspatially unbounded, with ϵ0 ¼ limδ-01þ δi. The dielectricconstants are typically assumed to be local, i.e. depends on thefrequency but not the wave vector of the excitation. Oneexpects that such a local dielectric function should be applic-able unless one is in the deep near-field regimewhere the sizeof the vacuum gap is reduced to a few nanometers [7,19]. Wenote that in this regime, other mechanisms of heat transfermay come into play, such as phonon transport [20], but we donot consider these mechanisms in this review. Regarding thetemperature distribution, we assume that the temperature isuniform in each region; we label these temperatures Tn.

The sources of electromagnetic heat transfer are ther-mally fluctuating current density distributions Jiðr;ωÞwithin the bodies; see Fig. 1b. For the purpose of calculat-ing ensemble-averaged heat transfer in a stationary sys-tem, it suffices to know the two-point spectral correlationfunction of the current densities, which for a system inlocal thermal equilibrium is given by the fluctuation–dissipation theorem [21–23]. We use the time-harmonicform of the theorem, and thus we already assume astationary process. Here and throughout, an overbar willdenote an ensemble average. The theorem holds in generalfor dissipative linear systems. In the present case, thedissipation is due to dielectric losses, and we write

Jiðr;ωÞJnkðr′;ωÞ ¼4πωΘðω; TÞδðr−r′ÞIm½ϵðr;ωÞ�ϵ0δik ð1Þ

th dielectric functions ϵ1 and ϵ2, separated by vacuum and maintained atynamics calculation in the Green function picture. The current–currente related via the dyadic Green functions GEE and GHE to the field–fieldof V2.

Page 3: Fluctuational electrodynamics calculations of near …web.stanford.edu/group/fan/publication/Otey_JQSRT_132_3...C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative

C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 132 (2014) 3–11 5

where the mean spectral thermal energy is given byΘðω; TÞ ¼ ℏω=ðexpðℏω=kBTÞ−1Þ, and i and k label differentspatial directions. With the sources specified, the E and Hfields can be expressed in terms of dyadic Green functions(also referred to as Green dyads) for the time-harmonicMaxwell's equations,

Eðr;ωÞ ¼ iωμ0

ZGEEðr; r′;ωÞ � Jðr′;ωÞ ð2Þ

Hðr;ωÞ ¼Z

GHEðr; r′;ωÞ � Jðr′;ωÞ ð3Þ

We will make use of the concise operator notationwhenever possible [24]. A brief review on this notation canbe found in Appendix A. For example, Eqs. (1)–(3) abovebecome

jJ⟩⟨Jj ¼ 4πωΘ Im½ϵ�ϵ0I ð4Þ

jE⟩¼ iωμ0GEEjJ⟩ ð5Þ

jH⟩¼GHEjJ⟩ ð6ÞIn Eq. (4), I is the identity operator over all indices,including a delta

Rdr δ3ðr−r′Þ� term. In general, ϵ is a

tensor, and the operator notation is indifferent to the formof this tensor. In our computational examples, though, wewill assume an isotropic dielectric, which is adequate fordescribing most metal and dielectric systems.

The spectral energy flux into V2 due to sources in V1 canbe calculated by integrating the associated time-harmonicPoynting vector over the surface ∂V2

SðωÞ ¼ 12ReZ∂V2

dn̂ � Eðr;ωÞ �Hnðr;ωÞ ð7Þ

Here, n̂ is the outward pointing normal. The correlation ofthe E and H fields, due to sources inside V1, follows fromEqs. (1)–(3) as

jE⟩⟨Hj ¼ iωμ0GEEjJ⟩⟨JjG†HE

¼ i4πðω=cÞ2Θðω; T1ÞIm½ϵ1�GEEG

†HE ð8Þ

The total heat flux is just the integral of the spectral heatflux in Eq. (7), S¼ R∞

0 dω SðωÞ. Fig. 1b displays a schematicof this formalism.

One can reformulate Eq. (8) as a double volumeintegral, which can be absorbed into a trace in the operatorformalism. Following Krüger et al. [24, Section III.B], wemay write this as

SðωÞ ¼ 2πIm Tr½jE⟩⟨ðωμ0G0

EEÞ−1Ej� ð9Þ

where G0EE is the homogenous, free-space electric field

Green dyad.Of course, heat flows from V2 to V1, as well, and the

physically measurable quantity is the net heat transfer.Fortunately, reciprocity of the Green functions manifests anecessary consequence of detailed balance: at equilibriumthe net heat transfer must vanish, and so when T1¼T2, theheat flux into V2 due to sources in V1 must equal the heatflux into V1 due to sources in V2. From a computationalstandpoint, then, we may therefore limit our attention tosources in V1, effectively treating V2 as a zero temperature

body. To recover the net heat flux when T2≠0, we simplyreplace Θðω; T1Þ by Θðω; T1Þ−Θðω; T2Þ.

A direct implementation of the fluctuational electro-dynamic treatment of near-field electromagnetic heattransfer typically involves integration over a large numberof sources and frequencies, which for arbitrary geometriescan be very computationally demanding. While the basicformalism was developed in the 1950s [17], for severaldecades afterwards exact calculations of near-field elec-tromagnetic heat transfer were restricted to planargeometries [25,26]. The theory for planar surfaces hasbeen subsequently extended to treat, for example,structures with surface roughness [27] and nano-porousmaterials [28].

For non-planar geometries, a number of approximatemethods have been developed [2,29–32]. The dipoleapproximation was developed to treat the heat transferbetween a deep sub-wavelength nano-particle and aplanar substrate [29,31]. The proximity approximation,which was also extensively applied, approximates thetwo bodies in terms of a series of parallel planes [2]. Oneshould note however that the proximity approximation isan uncontrolled approximation in the sense that there isno systematic process through which one can improve theapproximation towards an exact result. Also, the proximityapproximation does not reproduce the correct far-fieldresults when the two bodies are far apart from eachother [32].

One of the most significant recent developments in thepast few years is the developments of exact numericalcalculations of near-field heat transfer between dielectricsin non-planar geometries. In this paper, we provide a brieftutorial review of such developments, with an emphasison the computational aspects involved. We begin with thepartial-wave scattering matrix approach in Section 2,which is well suited for treating highly symmetric geome-tries such as coupled-sphere and sphere-plate. In Section3, we discuss the use of the boundary element method(BEM). In Section 4, we discuss the finite-difference time-domain method, for treating arbitrary geometries in astatistical manner. Finally, in Section 5, we compare allthese methods with an illustrative example, where wesolve a coupled-sphere problem with all three differentmethods.

2. Partial-wave scattering method

2.1. Green function picture

The coupled microsphere geometry was the first non-planar geometry treated exactly with the fluctuationalelectrodynamics formalism. The approach of Narayanas-wamy and Chen in [33] was to calculate the Green dyad'sin Eq. (8) directly, by matching boundary conditions in avector spherical harmonic partial-wave expansion, andperform the source volume and flux surface integrationusing orthonormality properties of this basis. For concre-teness, our discussion below will often make reference tothis original coupled-sphere problem. We will use a moregeneral notation, though, closely following the work ofKrüger et al. [34; 35, Section IV; 24, Section VII]. We aim to

Page 4: Fluctuational electrodynamics calculations of near …web.stanford.edu/group/fan/publication/Otey_JQSRT_132_3...C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative

C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 132 (2014) 3–116

emphasize how this intuitive Green function picture canbe related to an elegant field-correlator picture, whichactually makes no reference to the sources.

To set notation, we work below with a basis of vectorwave function solutions fjΦα⟩g to the Maxwell equation,

∇� ∇� jΦα⟩−ðω=cÞ2jΦα⟩¼ 0 ð10Þ

These basis functions span a Hilbert space, with theusual inner product [36]. Here, and throughout the paper,we sum over repeated indices unless otherwise noted inthe text. A key property of these functions is that they havea well-defined and coordinate-free relationship to thefree-space Green dyads, e.g.

G0EE ¼ iðω=cÞjΦ′out

α ⟩⟨Φ′regα j ð11Þ

Referring now to the coupled-sphere example, we use avector spherical harmonic basis, and α∈fl;m; Pg labels thedegree l, orderm, and polarization P (Mwave or Nwave) ofthe orthonormal basis of spherical harmonics, andjΦ′l;m;P⟩¼ jΦl;−m;P⟩.

G0HE can be decomposed in a similar manner, with the

polarization for the r partial wave swapped (M wavesbecome N waves and vice versa), which we denote with�

the replacement jΦ⟩-jΦ⟩

G0HE ¼ iðω=cÞ2jΦ

� ′out

α ⟩⟨Φ′regα j ð12Þ

The inhomogenous Green functions can be obtained byapplying a scattering operator O to the free-space Greenfunction G0

EE , which acts on the outgoing partial-waveonly, i.e.

GEE ¼OG0EE ð13Þ

GEE ¼ iðω=cÞOαβjΦ′outβ ⟩⟨Φ′reg

α j ð14Þ

O is defined to act on a partial-wave solution of theMaxwell equations in homogenous space (the incidentfield), and return the total field when V1 and V2 are present[37,24].

If we restore the position dependence of the fields anddyads, we may write the field correlations, due to sourcesinside V1, as

⟨rjE⟩⟨Hjr′⟩body ¼4πðω=cÞ2Θðω; T1Þ

�ZV1

dr″ ϵ″1ðω=cÞ3⟨r″jΦ′regα ⟩⟨Φ′reg

β jr″⟩� �

�Oαγ⟨rjΦ′outγ ⟩⟨Φ

� ′out

ρ jr′⟩O†βρ ð15Þ

Note that the volume integral over sources is nowindependent of the scattering, and can be performedanalytically using orthonormality of fjΦα⟩g. The remainderof the computation is thus reduced to determining thescattering operator Oαβ . It suffices to know the T operatorsT1 and T2 for the isolated bodies in a suitable basis, andthe translation operators U12 and U21, which express thelinear relationship between the partial-waves sourced bybody 1 and body 2, in their respective bases. The T

matrices for the individual bodies, which contain thereflection coefficients for the partial waves in the basis

for that body, are defined to satisfy

jΦ′totα ⟩¼ jΦ′reg

α ⟩þ TαβjΦ′outβ ⟩ ð16Þ

where jΦ′totα ⟩ represents the total field resulting from

scattering the homogenous free space solutions jΦ′regα ⟩ in

the presence of the body.In the coupled-sphere problem, the U operators are

matrices with elements given by the vector translationaddition theorem which expresses vector spherical har-monics centered at V1 in terms of a similar basis withorigin at the center of V2. We arrive at a linear system forOby expressing waves sourced from V1 and V2 in a singlebasis using the T and U matrices,

O¼ ðI−T1U12T2U21Þ−1 ð17ÞIn the coupled-sphere problem, we truncate the α basis tof1…lmax;−mmax…mmaxg. Because of azimuthal symmetry inthis case, we find that O splits into blocks, indexed by m.Each m contribution involves a linear solve for 2lmax

unknown partial wave coefficients.Eq. (15) expresses the field correlations in the partial-

wave basis for body 1. In order to obtain the partial-wavecoefficients in the basis for body 2, we multiply byðIþ T2U21Þ. Then we may easily perform the surfaceintegral of the Poynting flux over ∂V2 using orthonormalityof the body 2 basis. A similar procedure can be used in thesphere-plate geometry [34,38,39], and in principle forother systems where orthonormality of the bases can beused to analytically perform the spatial integrals.

2.2. Field correlator picture

Alternatively, one can avoid explicit reference to thecurrent–current correlation function and the Green dyads,in favor of field–field correlators. In Section III.A of [24],Krüger et al. derive an expression for the radiation from abody in terms of the single body T-matrix. This in turnallows them to write the heat transfer as the trace of amatrix, which is constructed entirely from scattering T andtranslation U matrices. Here we briefly outline the idea,closely following their original arguments, but with aslightly different notation.

The starting point is the fluctuation–dissipation theo-rem for the electric field correlator in the absence of thebody [35,24]. In thermal equilibrium the bodies are at thesame temperature and there is no net heat flux, and thefluctuation–dissipation theorem applied to the electric-field fluctuations yields

jE⟩⟨Ejeq ¼ μ0ωð4=πÞΘðω; TÞIm½GEE�¼ aðω; TÞRe½jΦreg

α þ TαβΦoutβ ⟩⟨Φ′reg

α j�¼ aðω; TÞjΦreg

α þ TαβΦregβ ⟩⟨Φ′reg

α j ð18Þ

where aðω; TÞ ¼ 4ω2

πϵ0c3Θðω; TÞ.

One can decompose the field correlator in Eq. (18):

jE⟩⟨Ejeq ¼ jE⟩⟨Ejbody þ jE⟩⟨Ejenv ð19ÞThe environmental correlator is sourced by the free-

space region V0 [22,24,35,30], and can be expressed by

Page 5: Fluctuational electrodynamics calculations of near …web.stanford.edu/group/fan/publication/Otey_JQSRT_132_3...C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative

C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 132 (2014) 3–11 7

applying T-matrices to the partial-wave expansion of G0EE ,

jE⟩⟨Ejjenv ¼ aðω; TÞjΦ′regα þ TαβΦ′out

β ⟩⟨Φ′regα þ TαβΦ′out

β j ð20ÞThe correlator due to sources integrated inside the

body then follows from Eq. (19),

jE⟩⟨Ejbody ¼ aðω; TÞRαβjΦ′outα ⟩⟨Φ′outn

β j ð21Þ

where the R matrix contains all of the information aboutemission, and is defined in terms of the T�matrix, as

R¼−12ðT þ T†Þ−TT† ð22Þ

The presence of the 2nd body can be accounted for byapplying the Oαβ scattering operator from Eq. (17) to eachfield in the tensor product in Eq. (21). This gives theinhomogenous field correlation, and as a result, a traceformula for SðωÞ may be written as

SðωÞ ¼ yðω; TÞ2 2πTrfRn

2WR1W†g ð23Þ

where W¼ ðI−U21T1U12T2Þ−1U21.From a computational perspective, the calculation of

the W�matrix is essentially the same as the calculation ofO, described in the previous section. One needs to deter-mine the reflection coefficients and translation operatorsfor the system and solve a linear system with a truncatedbasis. However, the theoretical framework here is consid-erably more elegant, and can be generalized to anycoordinate system (see [24, Section VII]). Furthermore,for arbitrary geometries, one could calculate elements ofthe T-matrices for each body separately using any numer-ical method at one's disposal, and then use these sametrace formulae [37]. Of course, the coordinate systems foreach body must be appropriate for the geometry. If thereexists a plane that does not intersect either of the bodies,then a plane wave basis will work [40]; if a body can beenclosed by sphere which does not intersect the otherbodies, then a spherical basis is appropriate for that body'sT�matrix.

3. Boundary element method

With the boundary element we consider, instead ofpartial-wave solutions to the homogenous Maxwell'sequations, the scattered fields which are sourced by thetangential E and H fields on the boundaries of the bodies,considered as equivalent current sources. Given any inci-dent field configuration, these tangential surface fields areuniquely determined, and they completely specify thescattered fields.

More specifically, we introduce fjfα⟩g, a set oforthonormal vector-valued functions with support on∂V1;2. We then expand the surface currents

jn̂ �H⟩jE� n̂⟩

!¼ ηαjfα⟩ ð24Þ

and the tangential fields

jEt⟩

jHt⟩

!¼ ϕαjfα⟩ ð25Þ

The coefficients for the surface currents ηα and the

tangential fields ϕα become the objects of interest in thewell known boundary element method (BEM) [41–43]. TheBEM provides an efficient means of relating incident andscattered fields on ∂V1;2.

Using the BEM formalism, we can approach the near-field heat transfer problem from two pictures, very muchthe same as in the previous section. The first is a directapproach based on the Green function picture, i.e. calcu-lating the Green dyads in Eq. (8) explicitly, and integratingover all source configurations. The second approachinvolves constructing the non-equilibrium, inhomogenousfield correlators from the equilibrium, homogenous Greenfunctions and scattering operators.

3.1. Green function picture

In the Green function picture, we consider a set ofvolumetric current density distributions Ji with supportinside V1, as in Fig. 1 and solve for the fields Ei and Hi on∂V2 (See again Fig. 1b). The superscript i indexes this setof current sources. We follow the notation of Reid et al.[42–44], and carry over some content from Chew [45].

For each Ji, the boundary element method gives a linearsystem

Mαβηiβ ¼ ϕi

α ð26Þ

where

M¼Zϵ111 þ Zϵ0

11 Zϵ012

Zϵ021 Zϵ2

22 þ Zϵ022

!ð27Þ

The Zϵ blocks contain interactions between the basisfunctions via the Green dyads Gϵ

EE and GϵEH for a homo-

genous space with dielectric constant ϵ,

ðZϵmnÞαβ ¼ fα

�����iωμ0G

ϵEE Gϵ

EH

−GϵEH iωϵϵ0Gϵ

EE

�����fβ+;

r∈∂Vm

r′∈∂Vn:

*ð28Þ

The solution of Eq. (26) is sufficient for determiningSiðωÞ as

SiðωÞ ¼ 12

Z∂V2

dn̂ � Ei �Hin

¼ 14

Z∂V2

dSEH

� �� n̂ �Hn

En � n̂

!

¼ 14TrfðZ2η

2;iÞðη2;iÞ†g ð29Þ

In the last line, η2;i are those components of η that lie onthe surface of body 2, and we have used the relationϕ2;iα ¼ ðZ2Þαβη2;iβ . The integral is approximated as a trace over

all basis functions with support on ∂V2.With the BEM described above, we can calculate the

matrix M once per frequency, invert it, and use it tocalculate SðωÞ∝∑iS

iðωÞ for a complete set of ϕi associatedwith dipole sources distributed uniformly throughout V1.

3.2. Field correlator picture

There is an interesting connection between the BEMformalism and the field correlator picture discussed inSection 2.2 which gives a direct route to a trace formula

Page 6: Fluctuational electrodynamics calculations of near …web.stanford.edu/group/fan/publication/Otey_JQSRT_132_3...C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative

C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 132 (2014) 3–118

similar to Eq. (23),

SðωÞ ¼ yðω; TÞ2πTr Q 2M

−1Q 1ðM−1Þ† ð30Þ

where

Q 1 ¼12

Zϵ111 þ ðZϵ1

11Þ† 00 0

!ð31Þ

Q 2 ¼12

0 00 Zϵ2

22 þ ðZϵ222Þ†

!ð32Þ

The M and Z matrices are precisely those defined in Eqs.(27) and (28).

This formula was first derived in [44] without referenceto the correlator picture. But the form of Eq. (30) suggestsa connection to the trace formula (23) derived in Section 2.In fact, one can construct Eq. (30) as follows.

First consider a homogenous space with ϵðrÞ ¼ ϵ1. Notethat this is the material that composes the first body, notvacuum. The equilibrium field correlations can be writtenin matrix form, as in the vacuum case,

ϕaϕn

b

� �eq

¼ fajE⟩⟨Ej jE⟩⟨HjjH⟩⟨Ej jH⟩⟨Hj

�����fb+

eq

������*

¼ yðω; TÞ faωμ0 Im½Gϵ1

EE� i Im½Gϵ1HE�

i Im½Gϵ1EH � ωϵ1 Im½Gϵ1

E �

�����fb+�����

*

¼ yðω; TÞ12

Zϵ111 þ ðZϵ1

11ކ� �

aβ ð33Þ

Here we have used the definition of the Z matrices in Eq.(28), and reciprocity of the Green dyads. The equilibriumfield correlations in Eq. (33) contain contributions fromboth the body and from the environment.

We now turn our attention to the correlation of thesurface currents due to sources in body 1. Eq. (26) providesthe necessary connection between the homogenous (inci-dent) fields and inhomogenous (incident plus scattered)fields. The key step is to restrict the incident fields to body1, so that the environmental contribution to the equili-brium correlation is not included when solving for theinhomogenous surface currents. This gives a modified BEMequation,

Mη1

η2

!¼ I 0

0 0

� �ϕ1

ϕ2

!ð34Þ

From Eqs. (33) and (34), we can derive the correlationof the surface current coefficients,

ηηn ¼M−1 I 00 0

� �ϕϕn

eqI 00 0

� �ðM−1Þ†

¼ yðω; TÞM−1Q 1ðM−1Þ† ð35ÞWe finally arrive at Eq. (30) by appalying to the left of

Eq.(35), which isolates the scattered fields on aV2, as in Eq.(29). The BEM trace formula (30) has the same form as theT�matrix formula in Eq. (23). However, the BEM formulais not merely a special case of the T�matrix formula. TheQ matrices are associated with the equilibrium, homo-genous fields in V1, whereas the R operators in Eq. (23) areassociated with the single body radiation fields.

The BEM formalism can also be used to calculateT�matrix elements for bodies which do not have the samesymmetry as the partial-wave coordinate systems. Thismay be accomplished by solving the BEM problem in Eq.(26) Mγβηαβ ¼ ⟨fγ jΦα⟩ for each partial-wave Φα in the trun-cated basis associated with the coordinate system. Then Tn

can be expressed as Tn ¼ Zϵnnnη

α. This type of procedure wasfirst used by McCauley et al. for calculating heat transferbetween a probe and a plane [37], using a cylindricalplane-wave basis.

4. Finite difference time domain method

In this section we describe how the finite-differencetime-domain (FDTD) method [46] can be used to calculateheat transfer by incorporating the Langevin approach toBrownian motion [47–49]. In this approach, we model thepolarization response PðtÞ of a system to a local electricfield EðtÞ and a random force term KðtÞ using the followingequation of motion:

d2P

dt2þ γ

dPdt

þ ω20P¼ sEþ K ð36Þ

where γ is the frictional coefficient of the polarizationsystem, ω0 is its resonance frequency and s is its strength.The random force term KðtÞ is a fluctuating source thatgives rise to thermal radiation.

After discretizing Eq. (36) and specifying KðtÞ, we canperform time-stepping updates of PðtÞ and EðtÞ using theconventional FDTD algorithm [47]. From this numericalsimulation of thermal emission, we can calculate the heatflux spectrum in an arbitrarily shaped geometry.

We first describe how the random force term KðtÞ canbe specified in order to model the thermal fluctuations inthe electromagnetic field. The polarization component ofthe displacement field D¼ ϵ0Eþ P contains a randomcomponent with spectral amplitude [48]

Q ðr;ωÞ ¼ Kðr;ωÞω20−ω2−iγω

ð37Þ

Q ðr;ωÞ is related to the current Jðr;ωÞ ¼−iωQ ðr;ωÞ and

ϵ″ðωÞ ¼ Im ϵ0 þjPjjEj

� �¼ sγω

ðω20−ω2Þ2 þ γ2ω2

:

By combining the fluctuation–dissipation theorem for thecurrents in Eq. (1) with Eq. (37), we find the correlationfunction for Kðr;ωÞ [48],

Kαðr;ωÞKn

βðr′;ωÞ ¼4πsγΘðω; TÞδαβδðr−r′Þ: ð38Þ

This expression for the K correlation function gives usinformation about the distribution of the random forceterm. For FDTD simulations, we generally need Fouriertransform equation (38) in order to get the time-domainvariance jKαðr; tÞj2 of the random force term distribution.

However, if one were to directly implement Eq. (38),the frequency dependency of the correlator needs to beimplemented in terms of a temporal convolution due tothe Θðω; TÞ term, which is computationally expensive.Instead, the key idea is to use an alternative sourceK′ðr;ωÞ ¼Kðr;ωÞ=

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiΘðω; TÞ

pwith a white noise spectrum,

Page 7: Fluctuational electrodynamics calculations of near …web.stanford.edu/group/fan/publication/Otey_JQSRT_132_3...C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative

Table 1A comparison of some computational aspects of three methods forcalculating heat transfer between coupled spheres.

Method Error Totalruntime

Totalmemory

Timecomplexity

Spacecomplexity

Partial-wave

0 3 s o1 Mb Oðζ4Þ Oðζ2Þ

BEM 0.3% 100 h 1.8 Gb Oðζ6Þ Oðζ4ÞFDTD 14% 4000 h 200 Mb Oðζ4Þ Oðζ3Þ

C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 132 (2014) 3–11 9

where the correlation is instantaneous in time [47]. Usinga random force term with correlation functionK′αðr′;ωÞK ′n

α ðr″;ωÞ in the FDTD method will allow us tosimulate a normalized heat flux spectrum SðωÞ=Θðω; TÞ.The actual heat flux spectrum SðωÞ is then calculated bysimply multiplying this normalized heat flux spectrum bythe known function Θðω; TÞ.

The discretized correlation function of K′ðr;ωÞ is

K′αðr;ωÞK ′nβ ðr′;ω′Þ ¼

4sγπΔV

δαβδωω′δrr′: ð39Þ

Eq. (39) can be trivially Fourier transformed into the time-domain:

K′αðr; tÞK ′nβ ðr′; t′Þ ¼

4sγπNΔV

δαβδtt′δrr′ ð40Þ

where N is the number of time steps used in the Fouriertransform. Hence, by performing at each time step arandom drawing of K′ from a distribution with variancejK′αðr; tÞj2 , we can simulate a normalized thermal emissionwith normalized heat flux spectrum SðωÞ=Θðω; TÞ. Since thespecification of the variance jK′αðr; tÞj2 is the only con-straint on the distribution of K′, we can choose a uniformdistribution with a range 7 ð12sγ=πNΔVÞ1=2.

Finally, we note that although in this section we used apermittivity that is modeled by a single-pole Lorentzfunction, the method discussed can be implemented withmaterial permittivity that has multiple poles.

5. Comparison of computational approaches

In order to provide a concrete basis for comparison ofthe computational aspects of the three methods (partialwave scattering, BEM and FDTD), we choose a simple testcase: coupled dielectric spheres. In our example, bothspheres have a radius a¼800 nm, are separated by avacuum gap d¼200 nm, and the dielectrics are n-dopedsilicon with a carrier concentration n¼1020 cm−3, modeledusing [50]. The dielectric functions are shown in Fig. 2a.The small sphere size is chosen in part so that all methodscan converge in a reasonable time.

All these numerical methods require some discretiza-tion or truncation of the problem which introduces error.In the case of BEM, the basis we use in our implementa-tions is the well known RWG (Rao–Wilton–Glisson) func-tions [51,41], which are defined on a triangulated mesh ofthe object boundaries. For the FDTD method we use a non-uniform 3D Cartesian grid. In both methods, the intendedgeometry is only approximated in the spatial discretizationprocedure. At a more fundamental level, the relevantlength scales in the problem must be resolved by incor-porating a sufficient number of unknowns in the compu-tation, so we can characterize the behavior of thesemethods in terms of the scaling behavior of time andmemory costs as we vary a and d. As the bodies grow,more unknowns will be required to account for theadditional degrees of freedom in the fields. Also, as thebodies are brought closer together, more unknowns arerequired to resolve the field variations due to near-fieldinteractions. In the following, we will characterize the

complexity of the system in terms of the single dimen-sionless parameter, ζ¼maxfk0a; a=dg [52].

The numerical results of the three methods, BEM,scattering and FDTD, are shown in Fig. 2b. A summary ofthe computational performance of the three methods isshown in Table 1. All of our calculations were performedon the SDSC Trestles linux cluster [53].

The scattering method acts as a benchmark of thecoupled-sphere problem. There is no discretization error,and the system is small enough that numerical conver-gence to within 1�10−8 can be achieved with fewer than100 unknowns. The complexity of the scattering method isfavorable as well, for highly symmetric geometries. In asystem with azimuthal symmetry, each m¼ 0…mmax givesa separable problem which requires N¼ 2lmax∝ζunknowns. The linear solve is O(N3) time complexity, butin practice, the time spent calculating the N2 matrixelements far exceeds the time for the linear solve, pro-vided N≲104. In general, mmax∝ζ, as well, so for azimuthalsymmetric problems, the total time complexity is Oðζ3Þ inpractice. In the ζ-∞ limit, though, the linear solveeventually dominates so the time complexity is Oðζ4Þ.Space complexity is Oðζ2Þ, and is not at all prohibitive.

The number of unknowns (edges) in the BEM scaleswith the area which needs to be resolved, i.e. as Oðζ2Þ, sooften the limiting computational resource is memory. Forour small test case, we used 7000 unknowns, whichrequires ≈2 Gb to store all the relevant matrices. This isan intentional test to approach maximum problem size wecan easily carry out in a 32-bit memory address space. Wedo not need this many elements to achieve good accuracy.This computation is roughly equally divided betweencomputing BEM matrix elements, and performing matrixoperations. For larger systems, matrix operations willdominate the time complexity. The BEM approach givesnearly identical numerical results, whether using directvolume integration or the closed form boundary traceformula, but the latter is more time efficient for largersystems, and there is no ambiguity in how to choose thesources Ji in the volume integration.

The FDTD method is of a statistical nature; the datashown is an average taken over 40 independent runs, eachof which takes ≈100 h to complete. Even after 40 runs, thenumerical results have not converged, but the results areconsistent with the other methods. Despite these short-comings, the FDTD method is a valuable tool due to thesimplicity of treating arbitrary geometries and inhomo-genous dielectrics.

At least two things are worth reemphasizing. First,FDTD is at volumetric method, it actually requires less

Page 8: Fluctuational electrodynamics calculations of near …web.stanford.edu/group/fan/publication/Otey_JQSRT_132_3...C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative

−50

0

50

Re[

ε],Im

[ε]

Re[ε400] Im[ε400]

Re[ε300] Im[ε300]

1 2 3 4 50

1

2

ω (rad/s) / 1014

S(ω

) [W

/(rad

/s)]

/ 10−2

3

BEMFDTDPartial−wave

Fig. 2. (a) The dielectric function of hot 400 K and cold 300 K 1020 cm−3

n-doped Si spheres of radius 800 nm, separated by a 200 nm vacuum gap.The inset graphic is an artistic rendering of theoretical electric fieldintensity at a hot and cold sphere. (b) Spectral heat transfer SðωÞcalculated using three different computational methods.

C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 132 (2014) 3–1110

memory because the system is sparse, whereas our BEMapproach results in a dense matrix. Second, BEM also haspoor worst case scaling in time complexity, but performsrelatively well in a realistically sized system, and there areoptimizations which can improve this to ζ4 log N using thepre-corrected FFT algorithm [54,55], or better if adaptivemeshes or hierarchical matrices are also used [56].

6. Conclusion

Precision probes of near-field heat transfer have largelyfocused on plate–plate [3] or sphere–plate [1,2,5] geome-tries. However, there are important structures whichcannot be treated efficiently with the methods describedhere. For example, one of the most exciting possibleapplications of near-field heat transfer is near-fieldthermo-photo-voltaics. The design of effective near-fieldemitters and absorbers requires tuning the spectral heattransfer between nano-structured surfaces. There has beensome work towards this end in 1D grating geometries,using the plane-wave scattering approach [57,58] and the

FDTD [49] methods. For more general periodic systems,periodic extensions of the BEM [59] may prove to be anefficient alternative. There is evidence that for nano-scalesystems, non-local dielectric effects may come into play[7]. If these non-local effects can be modeled computa-tionally in arbitrary geometries, then light may be shed onnear-field heat transfer experiments and the properties ofnano-scale dielectrics in general. Certain non-local models[60] can be implemented by modifying the homogenousspace Green's function. Since the BEM matrices are con-structed from homogenous space Green's functions, BEMmethods could perhaps be developed to deal with arbi-trary non-local dielectrics.

Acknowledgment

This work is supported by an AFOSR-MURI program(Grant no. FA9550-08-1-0407) and by the Division ofMaterials Sciences and Engineering, Office of Basic EnergySciences, the U.S. Department of Energy.

Appendix A. Electromagnetic operator notation

We employ the operator algebra in this paper to makethe notation compact. Here we provide a brief discussionof this notation. As an example, in the dyad-vector nota-tion, the electric field EðrÞ relates the current density JðrÞthrough a dyadic Green function,

EiðrÞ ¼Z

dr′ ∑jGijJjðr′Þ ðA:1Þ

This formula describes a linear map between JðrÞ andEðr′Þ. In operator language we write this as

jE⟩¼GEEjJ⟩ ðA:2ÞFrom any two states ϕiðrÞ and ψ jðr′Þ, we can define anoperator

Oijðr; r′Þ≡ϕn

i ðrÞ⊗ψ jðr′Þ ðA:3Þ

We can write this in the operator notation as

O≡jϕ⟩⟨ψj ðA:4ÞFinally, the trace is defined

Tr O≡∑i

Zdr Oiiðr; rÞ ðA:5Þ

where integration occurs over all space.

References

[1] Narayanaswamy A, Shen S, Chen G. Near-field radiative heat transferbetween a sphere and a substrate. Phys Rev B 2008;78:115303-4.

[2] Rousseau E, Siria A, Jourdan G, Volz S, Comin F, Chevrier J, et al.Radiative heat transfer at the nanoscale. Nat Photonics 2009;3:514–7.

[3] Ottens RS, Quetschke V, Wise S, Alemi AA, Lundock R, Mueller G,et al. Near-field radiative heat transfer between macroscopic planarsurfaces. Phys Rev Lett 2011;107:014301-4.

[4] Narayanaswamy A, Gu N. Heat transfer from freely suspendedbimaterial microcantilevers. J Heat Transf 2011;133:042501-6.

[5] Guha B, Otey C, Poitras CB, Fan S, Lipson M. Near-field radiativecooling of nanostructures. Nano Lett 2012;12:4546–50.

Page 9: Fluctuational electrodynamics calculations of near …web.stanford.edu/group/fan/publication/Otey_JQSRT_132_3...C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative

C.R. Otey et al. / Journal of Quantitative Spectroscopy & Radiative Transfer 132 (2014) 3–11 11

[6] De Wilde Y, Formanek F, Carminati R, Gralak B, Lemoine PA, JoulainK, et al. Thermal radiation scanning tunnelling microscopy. Nature2006;444:740–3.

[7] Kittel A, Müller-Hirsch W, Parisi J, Biehs SA, Reddig D, Holthaus M.Near-field heat transfer in a scanning thermal microscope. Phys RevLett 2005;95:224301-4.

[8] Otey CR, Lau WT, Fan S. Thermal rectification through vacuum. PhysRev Lett 2010;104:154301-4.

[9] Basu S, Francoeur M. Near-field radiative transfer based thermalrectification using doped silicon. Appl Phys Lett 2011;98:113106-3.

[10] Zhu L, Otey CR, Fan S. Negative differential thermal conductancethrough vacuum. Appl Phys Lett 2012;100:044104-3.

[11] Iizuka H, Fan S. Rectification of evanescent heat transfer betweendielectric-coated and uncoated silicon carbide plates. J Appl Phys2012;112:024304-7.

[12] Basu S, Zhang ZM, Fu CJ. Review of near-field thermal radiation andits application to energy conversion. Int J Energy Res 2009;33:1203–32.

[13] Whale MD, Cravalho EG. Modeling and performance of microscalethermophotovoltaic energy conversion devices. IEEE Trans EnergyConv 2002;17:130–42.

[14] Narayanaswamy A, Chen G. Surface modes for near field thermo-photovoltaics. Appl Phys Lett 2003;82:3544–6.

[15] Ilic O, Jablan M, Joannopoulos JD, Celanovic I, SoljačićM. Overcomingthe black body limit in plasmonic and graphene near-field thermo-photovoltaic systems. Opt Exp 2012;20:366–84.

[16] Messina R, Ben-Abdallah P. Graphene-based photovoltaic cells fornear-field thermal energy conversion arXiv:1207.1476, 2012.

[17] Rytov SM, Kravtsov YA, Tatarskii VI. Principles of statistical radio-physics. Berlin: Springer-Verlag; 1989.

[18] Janowicz M, Reddig D, Holthaus M. Quantum approach to electro-magnetic energy transfer between two dielectric bodies. Phys Rev A2003;68:043823-17.

[19] Chapuis PO, Volz S, Henkel S, Joulain K, Greffet JJ. Effects of spatialdispersion in near-field radiative heat transfer between two parallelmetallic surfaces. Phys Rev B 2008;77:035431-9.

[20] Sellan DP, Landry ES, Sasihithlu K, Narayanaswamy A, McGaugheyAJH, Amon CH. Phonon transport across a vacuum gap. Phys Rev B2012;85:024118-6.

[21] Haus HA. Electromagnetic noise and quantum optical measure-ments. Berlin: Springer; 2000.

[22] Agarwal GS. Quantum electrodynamics in the presence of dielectricsand conductors. I. Electromagnetic-field response functions andblack-body fluctuations in finite geometries. Phys Rev A 1975;11:230–42.

[23] Calzetta E, Hu BL. Stochastic dynamics of correlations in quantumfield theory: from the Schwinger–Dyson to Boltzmann–Langevinequation. Phys Rev D 1999;61:025012.

[24] Krüger M, Bimonte G, Emig T, Kardar M. Trace formulae for non-equilibrium Casimir interactions, heat radiation and heat transfer forarbitrary objects. Phys Rev B 2012;86:115423.

[25] Polder D, Van Hove M. Theory of radiative heat transfer betweenclosely spaced bodies. Phys Rev B 1971;4:3303–14.

[26] Loomis JJ, Maris HJ. Theory of heat transfer by evanescent electro-magnetic waves. Phys Rev B 1994;50:18517–24.

[27] Biehs SA, Greffet JJ. Influence of roughness on near-field heattransfer between two plates. Phys Rev B 2010;82:245410-8.

[28] Biehs SA, Ben-Abdallah P, Rosa FSS, Joulain K, Greffet JJ. Nanoscaleheat flux between nanoporous materials. Opt Exp 2011;19:1088–103.

[29] Pendry JB. Radiative exchange of heat between nanostructures. JPhys: Condens Matter 1999;11:6621–33.

[30] Volokitin AI, Persson BNJ. Radiative heat transfer between nanos-tructures. Phys Rev B 2001;63:205404-11.

[31] Mulet JP, Joulain K, Carminati R, Greffet JJ. Nanoscale radiative heattransfer between a small particle and a plane surface. Appl Phys Lett2001;78:2931–3.

[32] Sasihithlu K, Narayanaswamy A. Proximity effects in radiative heattransfer. Phys Rev B 2011;83:161406-4.

[33] Narayanaswamy A, Chen G. Thermal near-field radiative transferbetween two spheres. Phys Rev B 2008;77:075125-12.

[34] Krüger M, Emig T, Kardar M. Nonequilibrium electromagneticfluctuations: heat transfer and interactions. Phys Rev Lett 2011;106:210404-4.

[35] Rahi SJ, Emit T, Graham N, Jaffe RL, Kardar M. Scattering theoryapproach to electrodynamic Casimir forces. Phys Rev D 2009;80:085021.

[36] John D, Joannopoulos JD, Johnson SG, Winn JN, Meade RD. Photoniccrystals: molding the flow of light. Princeton: Princeton UniversityPress; 2008.

[37] McCauley AP, Reid MTH, Krüger M, Johnson SG. Modeling near-fieldradiative heat transfer from sharp objects using a general 3dnumerical scattering technique. Phys Rev B 2012;85:165104-4.

[38] Otey C, Fan S. Numerically exact calculation of electromagnetic heattransfer between a dielectric sphere and plate. Phys Rev B 2011;84:245431-7.

[39] Golyk VA, Krüger M, Kardar M. Heat radiation from long cylindricalobjects. Phys Rev E 2012;85:046603-15.

[40] Messina R, Antezza M. Scattering-matrix approach to Casimir–Lifshitz force and heat transfer out of thermal equilibrium betweenarbitrary bodies. Phys Rev A 2011;84:042102-22.

[41] Hänninen I, Taskinen M, Sarvas J. Singularity subtraction integralformulae for surface integral equations with RWG, rooftop andhybrid basis functions. Prog Electromagn Res 2006;63:243–78.

[42] Reid MTH, Rodriguez AW, White J, Johnson SG. Efficient computa-tion of Casimir interactions between arbitrary 3D objects. Phys RevLett 2009;103:040401-4.

[43] Reid MTH, White JK, Johnson SG. Computation of Casimir interac-tions between arbitrary three-dimensional objects with arbitrarymaterial properties. Phys Rev A 2011;84:010503-4.

[44] Rodriguez AW, Reid MTH, Johnson SG. Fluctuating surface-currentformulation of radiative heat transfer for arbitrary geometries. PhysRev B 2012;86:220302-6.

[45] Chew WC. Waves and fields in inhomogenous media. Berlin: VanNostrand Reinhold; 1990.

[46] Taflove A, Hagness SC. Computational electrodynamics: the finite-difference time-domain method. 3rd edBoston: Artech House;2005.

[47] Luo C, Narayanaswamy A, Chen G, Joannopoulos J. Thermal radiationfrom photonic crystals: a direct calculation. Phys Rev Lett 2004;93:19–22.

[48] Chan D, Soljačić M, Joannopoulos J. Thermal emission and design inone-dimensional periodic metallic photonic crystal slabs. Phys Rev E2006;74:1–9.

[49] Rodriguez AW, Ilic O, Bermel P, Celanovic I, Joannopoulos JD, SoljačićM, et al. Frequency-selective near-field radiative heat transferbetween photonic crystal slabs: a computational approach forarbitrary geometries and materials. Phys Rev Lett 2011;107:114302-5.

[50] Basu S, Lee BJ, Zhang ZM. Infrared radiative properties of heavilydoped silicon at room temperature. J Heat Transf 2010;132:023301-8.

[51] Rao SM, Wilton DR, Glisson AW. Electromagnetic scattering bysurfaces of arbitrary shape. IEEE Trans Ant Prop 1982;30(3):1982409.

[52] Sasihithlu K, Narayanaswamy A. Convergence of vector sphericalwave expansion method applied to near-field radiative transfer. OptExp 2011;19(S4):A772–85.

[53] Trestles is an XSEDE San Diego Supercomputer Center clusterconsisting of 324 compute nodes, each with 4�2.4 GHz AMDMagny-Cours processors and 64 GB memory capacity. We use 64-bit floating point calculations throughout our examples ⟨http://www.sdsc.edu/us/resources/trestles/⟩.

[54] Reid MTH, Fluctuating surface currents: a new algorithm forefficient prediction of Casimir interactions among arbitrary materi-als in arbitrary geometries. PhD thesis. MIT; 2011.

[55] Yuan N, Yeo TS, Nie XC, Li LW, Gan YB. Analysis of scattering fromcomposite conducting and dielectric targets using the precorrected-FFT algorithm. J Electromagn Waves Appl 2003;17:499–515.

[56] Bebendorf M, Rjasanow S. Adaptive low-rank approximation ofcollocation matrices. Computing 2003;70:1–24.

[57] Bimonte G. Scattering approach to Casimir forces and radiative heattransfer for nanostructured surfaces out of thermal equilibrium.Phys Rev A 2009;80:042102-7.

[58] Lussange J, Guérout R, Rosa FSS, Greffet JJ, Lambrecht A, Reynaud S.Radiative heat transfer between two dielectric nanogratings in thescattering approach. Phys Rev B 2012;86:085432-5.

[59] Lee SC, Rawat V, Lee JF. A hybrid finite/boundary element methodfor periodic structures on non-periodic meshes using an interiorpenalty formulation for Maxwell's equations. J Comput Phys2010;229:4934–51.

[60] Garcia de Abajo FJ. Nonlocal effects in the plasmons of stronglyinteracting nanoparticles, dimers and waveguides. J Phys Chem C2008;112:17983–7.