INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale...

15
INVITED PAPER Fluctuation-Induced Phenomena in Nanoscale Systems: Harnessing the Power of Noise Computations for fluctuation-induced phenomena, including near-field radiative heat transfer and Casimir forces are reviewed in this paper; basic physics, theoretical analysis methods, and future predictions are presented. By M. T. Homer Reid , Alejandro W. Rodriguez, and Steven G. Johnson ABSTRACT | The famous Johnson–Nyquist formula relating noise current to conductance has a microscopic generalization relating noise current density to microscopic conductivity, with corollary relations governing noise in the components of the electromagnetic fields. These relations, known collectively in physics as fluctuation–dissipation relations, form the basis of the modern understanding of fluctuation-induced phenomena, a field of burgeoning importance in experimental physics and nanotechnology. In this review, we survey recent progress in computational techniques for modeling fluctuation-induced phenomena, focusing on two cases of particular interest: near- field radiative heat transfer and Casimir forces. In each case we review the basic physics of the phenomenon, discuss semi- analytical and numerical algorithms for theoretical analysis, and present recent predictions for novel phenomena in com- plex material and geometric configurations. KEYWORDS | Boundary-element; CAD; Casimir effect; finite- difference; fluctuation; heat transfer; Johnson; modeling; noise; Nyquist; radiation; simulation I. INTRODUCTION Every electrical engineer knows the famous Johnson– Nyquist formula for the noise current through a resistor, hI 2 4kTGf (1) where hI 2 i is the mean-square noise current [see Fig. 1(a)], kT is the temperature in energy units, G ¼ 1=R is the conductance of the resistor, and f is the measurement bandwidth. Equation (1)Vwhich allows designers to quan- tify, and thus compensate for, the unavoidable presence of noise in physical circuitsVis a crucial tool in the circuit designer’s kit and a mainstay of the electrical engineering curriculum from its earliest stages [1]. Perhaps less well-known in the EE community is that (1) is only one manifestation of a profound and far- reaching principle of physicsVthe fluctuation-dissipation theoremVthat relates the mean-square values of various fluctuating quantities to certain physical parameters (known as generalized susceptibilities) associated with the underlying system. In (1), the fluctuating quantity is the noise current through the resistor, and the generalized susceptibility is the conductance; more generally, as we will see below, the fluctuation-dissipation concept allows Manuscript received November 3, 2011; revised February 18, 2012; accepted March 3, 2012. Date of publication July 3, 2012; date of current version January 16, 2013. This work was supported in part by the Defense Advanced Research Projects Agency (DARPA) under Grant N66001-09-1-2070-DOD, by the Army Research Office through the Institute for Soldier Nanotechnologies (ISN) under Grant W911NF-07-D-0004, and by the AFOSR Multidisciplinary Research Program of the University Research Initiative (MURI) for Complex and Robust On-chip Nanophotonics under Grant FA9550-09-1-0704. M. T. H. Reid is with the Research Laboratory of Electronics, Massachusetts Institute of Technology, Cambridge, MA 02139 USA (e-mail: [email protected]). A. W. Rodriguez is with the School of Engineering and Applied Sciences, Harvard University, Cambridge, MA 02138 USA, and the Department of Mathematics, Massachusetts Institute of Technology, Cambridge, MA 02139 USA (e-mail: [email protected]). S. G. Johnson is with the Department of Mathematics, Massachusetts Institute of Technology, Cambridge, MA 02139 USA (e-mail: [email protected]). Digital Object Identifier: 10.1109/JPROC.2012.2191749 Vol. 101, No. 2, February 2013 | Proceedings of the IEEE 531 0018-9219/$31.00 Ó2012 IEEE

Transcript of INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale...

Page 1: INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale ...arod/papers/ReidRodriguez13_review.pdf · tricalengineersVincludingtheT-matrix,finite-difference, and boundary-element

INV ITEDP A P E R

Fluctuation-InducedPhenomena in NanoscaleSystems: Harnessing thePower of NoiseComputations for fluctuation-induced phenomena, including near-field radiative heattransfer and Casimir forces are reviewed in this paper; basic physics, theoreticalanalysis methods, and future predictions are presented.

ByM. T. Homer Reid, Alejandro W. Rodriguez, and Steven G. Johnson

ABSTRACT | The famous Johnson–Nyquist formula relating

noise current to conductance has a microscopic generalization

relating noise current density to microscopic conductivity, with

corollary relations governing noise in the components of the

electromagnetic fields. These relations, known collectively in

physics as fluctuation–dissipation relations, form the basis of

the modern understanding of fluctuation-induced phenomena,

a field of burgeoning importance in experimental physics and

nanotechnology. In this review, we survey recent progress in

computational techniques for modeling fluctuation-induced

phenomena, focusing on two cases of particular interest: near-

field radiative heat transfer and Casimir forces. In each case we

review the basic physics of the phenomenon, discuss semi-

analytical and numerical algorithms for theoretical analysis,

and present recent predictions for novel phenomena in com-

plex material and geometric configurations.

KEYWORDS | Boundary-element; CAD; Casimir effect; finite-

difference; fluctuation; heat transfer; Johnson; modeling;

noise; Nyquist; radiation; simulation

I . INTRODUCTION

Every electrical engineer knows the famous Johnson–Nyquist formula for the noise current through a resistor,

hI2i ! 4kTG!f (1)

where hI2i is the mean-square noise current [see Fig. 1(a)],kT is the temperature in energy units, G ! 1=R is theconductance of the resistor, and !f is the measurementbandwidth. Equation (1)Vwhich allows designers to quan-tify, and thus compensate for, the unavoidable presence ofnoise in physical circuitsVis a crucial tool in the circuitdesigner’s kit and a mainstay of the electrical engineeringcurriculum from its earliest stages [1].

Perhaps less well-known in the EE community is that(1) is only one manifestation of a profound and far-reaching principle of physicsVthe fluctuation-dissipationtheoremVthat relates the mean-square values of variousfluctuating quantities to certain physical parameters(known as generalized susceptibilities) associated with theunderlying system. In (1), the fluctuating quantity is thenoise current through the resistor, and the generalizedsusceptibility is the conductance; more generally, as wewill see below, the fluctuation-dissipation concept allows

Manuscript received November 3, 2011; revised February 18, 2012; acceptedMarch 3, 2012. Date of publication July 3, 2012; date of current version January 16,2013. This work was supported in part by the Defense Advanced Research ProjectsAgency (DARPA) under Grant N66001-09-1-2070-DOD, by the Army ResearchOffice through the Institute for Soldier Nanotechnologies (ISN) underGrant W911NF-07-D-0004, and by the AFOSR Multidisciplinary Research Programof the University Research Initiative (MURI) for Complex and Robust On-chipNanophotonics under Grant FA9550-09-1-0704.M. T. H. Reid is with the Research Laboratory of Electronics, Massachusetts Instituteof Technology, Cambridge, MA 02139 USA (e-mail: [email protected]).A. W. Rodriguez is with the School of Engineering and Applied Sciences, HarvardUniversity, Cambridge, MA 02138 USA, and the Department of Mathematics,Massachusetts Institute of Technology, Cambridge, MA 02139 USA(e-mail: [email protected]).S. G. Johnson is with the Department of Mathematics, Massachusetts Instituteof Technology, Cambridge, MA 02139 USA (e-mail: [email protected]).

Digital Object Identifier: 10.1109/JPROC.2012.2191749

Vol. 101, No. 2, February 2013 | Proceedings of the IEEE 5310018-9219/$31.00 !2012 IEEE

Page 2: INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale ...arod/papers/ReidRodriguez13_review.pdf · tricalengineersVincludingtheT-matrix,finite-difference, and boundary-element

us to quantify fluctuations not only in macroscopic devicecurrents but also in microscopic current densities, fromwhich it is a short step to obtain fluctuations in the com-ponents of the electric and magnetic fields inside andoutside material bodies [see Fig. 1(b)]. In this case, we willsee that the key tools turn out to be nothing but the fami-liar dyadic Green’s functions, which describe the electro-magnetic fields of prescribed current sources and arecomputable by any number of standard methods of clas-sical electromagnetism. It is remarkable that many prob-lems in the field of fluctuation-induced phenomena, whichwould at first blush seem to necessitate complex statistical-mechanical and quantum-mechanical reasoning, in factreduce in practice to applications of classical electro-magnetic theory that would be familiar to any electricalengineer.

But why would we want to quantify noise in the in-dividual components of the electromagnetic fields aroundmaterial bodies? The answer is that these microscopic fieldfluctuations can mediate macroscopic transfers of energy ormomentum among the bodies, which become especiallydramatic for bodies at submicron separations. In theformer phenomenonVnear-field radiative heat transferVfluctuating fields in micron-scale gaps between inequal-temperature bodies can lead to a rate of heat transferbetween the bodies that can drastically exceed therate observed at larger separations [2]. In the latterphenomenonVthe Casimir effectVfluctuating fieldsaround bodies give rise to attractive and repulsive forcesbetween the bodies, which generalize the familiar van derWaals interactions between molecules [3]. Both phenom-ena become negligibly small for bodies separated by dis-tances of more than a few microns, which places theirobservation squarely within the domain of nanoscale phy-sics and engineering.

Although the study of electromagnetic field fluctua-tions has been an active area of physics for decades, its

relevance to electrical engineering was limited for most ofthat time to (1) and other relations quantifying noise incircuits. In the past fifteen years, however, this situationhas begun to change; advances in fabrication and measure-ment technology have ushered in a golden age of experi-mental studies of fluctuation-induced phenomena [2],[17], and there is reason to believe that this fledgling fieldof experimental physics will soon become relevant toelectrical engineering in areas such as thermal lithographyand MEMS technology. This experimental progress hascreated a demand for modeling and simulation tools capa-ble of predicting fluctuation phenomena in realistic experi-mental configurations, including the complex, asymmetricgeometries and imperfect materials present in real-worldsystems.

The evolution of theoretical tools for modelingfluctuation-induced phenomena mirrors the historical de-velopment of techniques for solving classical electromag-netic scattering problems. In the latter case, the earliestcalculations were restricted to highly symmetric geome-tries (such as Mie’s 1908 treatment of scattering fromspheres) for which a convenient choice of coordinatesand special-function solutions of Maxwell’s equationsallow the problem to be solved analytically (or at leastsemianalyticallyVthat is, with results obtained as expan-sions in special functions, which in practice are then eval-uated numerically [18]). Later, fully numerical techniquescapable of handling more general geometries graduallybecame available, including the finite-difference, finite-element, and boundary-element methods introduced inthe 1960s, and today the problem of electromagnetic scat-tering is addressed by a wealth of comprehensive off-the-shelf CAD tools capable of handling extremely complexmaterial and geometric configurations.

Advances in the modeling of near-field radiative trans-fer and Casimir phenomena have proceeded in similarorder (Fig. 2). In both cases, the first calculations wererestricted to the simplest parallel-plate geometries [4],[12], [13]; these were later extended to other simple shapessuch as cylinders [8] and spheres [5], [9], [14], [19]–[21],and, more recently, tools for general geometries havebecome available [22]–[24]. All of these developments,however, have lagged their antecedents in the classical-scattering domain by many decades; indeed, even for therelatively simple case of two interacting spheres, theCasimir force was only calculated in 2007 [9] and the near-field radiative transfer only in 2008 [14]. One reason forthis lag is the relative paucity of experimental data,whichVas noted aboveVare significantly more difficult togather for fluctuation-induced phenomena than for clas-sical scattering. But perhaps the main reason that practicalcomputations of fluctuation-induced phenomena havebeen so long in coming is simply that the problems pre-sent extraordinary computational challenges. Indeed, aswe will see below, calculations of near-field radiative-transfer and Casimir phenomena may be reduced in

Fig. 1. Frommacroscopic tomicroscopic noise. (a) The current through

a resistor exhibits thermal noise with mean-square amplitude

proportional to the conductance [the Johnson–Nyquist formula, (1)].

(b) More generally, the microscopic current density inside a slab

of conducting material exhibits fluctuations with mean-square

amplitude proportional to the microscopic conductivity [the

fluctuation-dissipation theorem, (2)]. Knowledge of these microscopic

current fluctuations, together with the dyadic Green’s functions

of the system, allow us to predict the mean-square fluctuations

in the components of the electromagnetic fields in space

[(7) and (12)].

Reid et al. : Fluctuation-Induced Phenomena in Nanoscale Systems: Harnessing the Power of Noise

532 Proceedings of the IEEE | Vol. 101, No. 2, February 2013

Page 3: INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale ...arod/papers/ReidRodriguez13_review.pdf · tricalengineersVincludingtheT-matrix,finite-difference, and boundary-element

practice to the solution of classical scattering problemsVbut a great number, thousands or even millions, of sepa-rate scattering problems must be solved to compute theheat transfer or Casimir force for a single geometricconfiguration.

As a result, algorithms for predicting fluctuationphenomena tend to start with techniques familiar to elec-trical engineersVincluding the T-matrix, finite-difference,and boundary-element methods of computationalelectromagnetismVbut then proceed to combine andmodify these techniques in novel ways to obtain compu-tational procedures that can run in a reasonable length oftime. The goal of this review is to describe these compu-tational techniquesVand some of the results that theyhave predictedVin ways that will make sense to electricalengineers.

II . FLUCTUATIONS INELECTROMAGNETIC SOURCES ANDFIELDS: THE Johnson–NYQUIST LAWAND BEYOND

The microscopic generalization of (1) is [13], [25]

Ji"!;x#J$j "!;x0#

D E

! 1

!"ij""x% x0# "h!

2coth

"h!

2kT

! "# $#"!;x# (2)

where Ji"!;x# is the ith cartesian component of the mi-croscopic current density at position x and frequency !; "his Planck’s constant, and #"!;x# is the position- andfrequency-dependent conductivity. [# is related to theimaginary part of the dielectric permittivity according to

#"!;x# ! ! & Im$"!;x#; here and throughout we assumethat $ is linear and isotropic.] In signal-processing languagefamiliar to electrical engineers, the right-hand side of (2) isthe power spectral density (PSD) of a colored-noise process;the fact that the PSD is frequency-dependent (i.e., the factthat this is Bcolored[ instead of white noise) corresponds,in the time domain, to the nonvanishing of correlationsbetween currents at nearby time points.

The similarity between (1) and (2) is obvious: on theleft-hand side we have a mean product of currents, whileon the right-hand side we have a temperature-dependentfactor and a measure of conductivity. However, the mi-croscopic (2) extends the macroscopic (1) in two impor-tant ways.

First, whereas (1) is a low-frequency, high-temperatureapproximation that neglects quantum-mechanical effects,(2) remains valid at all temperatures and frequencies andexplicitly includes quantum-mechanical effects. Indeed,taking the low-temperature limit of the bracketed factor in(2), we find

limT!0

"h!

2coth

"h!

2kT

! "# $! "h!

2(3a)

and (2) thus predicts nonzero current fluctuations even atzero temperature: the well-known quantum-mechanicalzero-point fluctuations. In the high-temperature limit, onthe other hand, we have

limT!1

"h!

2coth

"h!

2kT

! "# $! kT (3b)

Fig. 2. Selective timeline indicating the most complex geometries for which rigorous calculations of Casimir interactions (upper) or near-field

radiative heat transfer (lower) were possible at various historical epochs. Note that computational techniques such as finite-difference grids and

boundary-element discretization, which have been used in electrical engineering for decades, have only been introduced to the study of

fluctuation-induced phenomena within the past five years.

Reid et al.: Fluctuation-Induced Phenomena in Nanoscale Systems: Harnessing the Power of Noise

Vol. 101, No. 2, February 2013 | Proceedings of the IEEE 533

Page 4: INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale ...arod/papers/ReidRodriguez13_review.pdf · tricalengineersVincludingtheT-matrix,finite-difference, and boundary-element

this is the classical regime, in which all dependence on "h islost and we recover the simple linear temperature de-pendence of (1). The classical regime is defined by thecondition

T ' "h!

2kor T in kelvin ' !

4 & 1012 rad/s(4)

a requirement that in practice is always satisfied in circuit-design problems, but which may be readily violated forinfrared and optical frequencies "! > 1015 rad/s#.

The second way in which (2) extends the reach of theJohnson–Nyquist result is that, whereas (1) describes onlymacroscopic currents, (2) gives information on the micros-copic current density, which in turn can be used to predictfluctuations in the components of the electric and mag-netic fields. The relevant tools for this purpose are thedyadic Green’s functions (DGFs), well-known to electricalengineers from problems ranging from radar and antennadesign to microwave device modeling [18]. To recall thedefinition of these quantities, suppose we have a materialconfiguration characterized by spatially-varying linear per-mittivity and permeability functions f$"!;x#; %"!;x#g.(In most of the problems we consider, $ and % will bepiecewise constant in space.) Then the electric DGF de-scribes the field due to a point source in the presence ofthe material configuration

GEij $; %;!;x;x0" #

! "i% component of electric field at x due to a

j% directed point current source at x0# (5)

while the magnetic DGF GM similarly gives the magneticfield of a point current source. (Here and throughout, allfields and currents are understood to have time depen-dence( e%i!t.) In (5) we have indicated the dependence ofG on the spatially-varying material properties $ and %; theDGFs for a given material configuration can be computedusing standard techniques in computational electromag-netism, after which the fields at arbitrary points in spacedue to a prescribed current distribution may be computedaccording to

Ei"!;x# !Z

GEij "!;x;x

0#Jj"!;x0# dx0 (6a)

Hi"!;x# !Z

GMij "!;x;x

0#Jj"!;x0# dx0: (6b)

Note that the long-range nature of the G dyadics ensuresthat the fields are nonvanishing even at points x in emptyspace, i.e., points at which there are no currents ormaterials.

Armed with (2) and (6), we can now make predictionsfor noise in the components of the electromagnetic fields.For example, the mean Poynting flux at a point x is a sumof terms of the form (with ! arguments to E, G, and Jsuppressed)

Ei"x#H$j "x#

D E

!Z

dx0 dx00GEik"x;x

0#GM$j‘ "x;x00# Jk"x0#J‘"x00#h i:

Inserting (2)

!Z

dx0GEik"x;x

0#GM$jk "x;x0## !; T"x0#) *#"!;x0# (7)

where T"x# is the local temperature and #)!; T* !""h!=2!# coth "h!=2kT is the statistical factor in (2).(Summation over repeated tensor indices is implied hereand throughout.)

The obvious advantage of an equation like (7) is that itreduces a problem in quantum statistical mechanics (de-termination of the electromagnetic field fluctuations at x)to a problem in classical electromagnetic scattering (com-putation of the DGFs GE;M). The difficulty of this approachlies in the great number of scattering problems that must besolved. Indeed, (7) says that, to compute the Poynting fluxat a single point x, we need the DGFs connecting x to allpoints x0 at which the conductivity is nonvanishing; for atypical problem involving two dissipative bodies invacuum, this amounts to a solving a separate scatteringproblem for each point in the volume of each body. More-over, even after completing all of these calculations wehave still only computed the Poynting flux at a single pointx; in general we will want to integrate this flux over asurface to get the total power transfer at a given frequency,and subsequently to integrate over all frequencies to getthe total power transfer.

Thus the fluctuation-dissipation concept, in the formof (2) or (7), performs the great conceptual service ofreducing predictions of noise phenomena to problems ofclassical electromagnetic scattering, but leaves in its wakethe practical problem of how to solve the formidablenumber of scattering problems that result. This difficultyhas been addressed in a variety of ways, some of which wewill review in the following sections.

Reid et al. : Fluctuation-Induced Phenomena in Nanoscale Systems: Harnessing the Power of Noise

534 Proceedings of the IEEE | Vol. 101, No. 2, February 2013

Page 5: INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale ...arod/papers/ReidRodriguez13_review.pdf · tricalengineersVincludingtheT-matrix,finite-difference, and boundary-element

III . NEAR-FIELD HEAT RADIATION:FLUCTUATION-INDUCED ENERGYTRANSFER IN NANOSCOPIC SYSTEMS

Fluctuating currents in finite-temperature bodies giverise to radiated fields which carry away energy. If thereare other bodies (or an embedding environment) presentat the same temperature, then any energy lost by onebody to radiation is replenished by an equal energyabsorbed from the radiation of other bodies. However,between objects at different temperatures there is a nettransfer of power, whose rate we can calculate in terms ofthe temperatures and electromagnetic properties of thebodies.

Historically, the first step in this direction was theStefan–Boltzmann law, a triumph of 19th-century physicswhich held that the power radiated per unit surface area ofa temperature-T body was simply &#SBT4, where #SB is auniversal constant and &, the emissivity, is a dimensionlessnumber between 0 and 1 characterizing the electricalproperties of the body (specifically, its propensity to emitradiation relative to that of a perfect emitter or black body).The Stefan–Boltzmann prediction is based on an approxi-mation that simplifies the electromagnetic analysis: itconsiders only propagating electromagnetic waves, neglect-ing the evanescent portions of the E andH fields that existin the vicinity of object surfaces. This is a good approxi-mation when computing the power transfer between asingle body and its environment, or between two inequal-temperature bodies separated by large distances.

However, when inequal-temperature bodies are sepa-rated by short distances, evanescent fields can contributesignificantly to the Poynting flux and the rate of powertransfer may deviate significantly from the Stefan–Boltzmann prediction. The length scale below which dis-tances are to be considered Bshort[ is the thermalwavelength

'T ! "hc

kT+ 7:6 %m & 300 K

T

! "

and thus, in practice, observing deviations from theStefan–Boltzmann law requires measuring the heat fluxbetween two bodies maintained at inequal temperaturesand at a surfaceVsurface separation of a few microns. Thisformidable experimental challenge has recently been metby several groups [2], [26], and this progress has spurredthe development of new theoretical techniques for pre-dicting the heat flux between closely-spaced bodies withrealistic material properties and various shapes, which wenow describe.

A. Radiative Heat Transfer as a Scattering ProblemConsider two homogeneous bodies B1;2 separated by a

short distance and maintained at separate internal thermal

equilibria at temperatures T1;2. (We will consider thebodies to exist in vacuum; the case of a finite-temperatureembedding environment is a straightforward generaliza-tion.) The rate at which energy is absorbed or lost by body 1is given as a surface integral of the mean Poynting flux

P1"!# !1

2

Z

S1

E"!;x# ,H$"!;x#h i & dS (8)

where S1 is the surface of body 1 (or, equivalently, afictitious bounding surface containing body 1 and no otherbodies) and dS is the inward-pointing surface normal.Applying (7) reduces the quantity in brackets to integralsover the volumes of the bodies (again suppressing !arguments to G)

P1"!#!"ijk2

Z

S1

(

#1"!##)!; T1*Z

B1

GEi‘"x;x

0#GM$j‘ "x;x0#dx0

- #2"!##)!; T2*Z

B2

GEi‘"x;x

0#GM$j‘ "x;x0#dx0

)

dSk (9)

where #1;2 are the conductivities of the bodies. [Here wehave used the Levi-Civita symbol "ijk to write thecomponents of the cross product as "A,B#k ! "ijkAiBj.]Note that (9) includes integrations over the volumesof both bodies, since there are fluctuating sources pre-sent in both bodies. Intuitively one might expect thatreciprocity arguments could be exploited to relate thetwo terms to one another and hence streamline thecalculation to involve integration over just one body; thisintuition is indeed born out in practice, as discussedbelow [15].

Equation (9) reduces the calculation of the net energytransfer to or from a body to the classical electromagneticscattering problem of computing the DGFs for a geometryconsisting of our two material bodies B1;2. The difficulty,as anticipated above, is that we must solve a great numberof scattering problems; in principle, for each surface pointx and each volume point x0 in the combined surfaceVvolume integrals in (9) we must solve a separate scatteringproblem (computing the fields at x due to individual pointsources at x0). This challenge is in fact so formidable thatcomputations for geometries even as simple as two sphereshave only become available in the past few years, usingtechniques which we now review.

B. Semi-Analytical Approaches to Radiative TransferA first strategy for evaluating (9) is to consider certain

highly symmetric geometries for which a convenientchoice of coordinates allows the DGFs to be evaluated

Reid et al.: Fluctuation-Induced Phenomena in Nanoscale Systems: Harnessing the Power of Noise

Vol. 101, No. 2, February 2013 | Proceedings of the IEEE 535

Page 6: INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale ...arod/papers/ReidRodriguez13_review.pdf · tricalengineersVincludingtheT-matrix,finite-difference, and boundary-element

analytically. For example, the earliest near-field heat-transfer calculations [12], [13] took the two objects to besemi-infinite planar slabs, in which case the DGFs areanalytically calculable. More recently, several groups haveextended this approach to other highly symmetric geom-etries in which special-function expansions of the DGFsare available [14], [27]–[32]. A particularly convenient toolhere is the Bmatrix[ approach to electromagnetic scatter-ing, a technique first discussed in these Proceedings ofthe IEEE in 1965 [33]. To solve scattering problems in thisapproach, one begins by writing down two sets offunctions, fEin

n "x#g and fEoutn "x#g; these are solutions

of Maxwell’s equations, in an appropriate coordinatesystem, which respectively describe electromagnetic wavespropagating inward from infinity to our scattering geom-etry and outward from the scatterer into open space. (Forexample, in spherical geometries the fEng will be productsof vector spherical harmonics and spherical Bessel functions[18].) The disturbance in the electromagnetic field due to ascattering object is then entirely encapsulated in theobject’s T-matrix, denoted T, whose m; n element gives theamplitude of the mth outgoing wave for a scattererilluminated by the nth incoming wave. In other words

if the incident field is Einc"x# !Einn "x#

then the scattered field is Escat"x# !X

m

TmnEoutm "x#:

Because the T-matrix for a body encodes all informa-tion needed to understand its scattering properties, it isoften possible to express the solution to a radiative-transferproblem in terms of simple matrix operations on theT-matrices of the objects involved. As an illustration of thesort of concise expression that can result from this pro-cedure, the methods of [27] lead to a simple trace formulafor the spectral density of heat radiation from a singlesphere at temperature T [34]

H"!; T#!%2#0)!; T*X

n

ReTnn"!# - Tnn"!#j j2% &

(10)

where T is the T-matrix of the sphere, the sum runs overits diagonal elements, and #0 is just # minus thecontribution of the zero-point energy term.

The obvious advantage of an equation like (10) is that itis simple enough to be implemented in a few lines ofMathematica or Matlab for objects whose T-matrix isknown analytically. The difficulty is that there are notmany such objects; indeed, the only lossy scatterers forwhich the T-matrix may be obtained in closed form arespheres, infinite-length cylinders, and semi-infiniteslabs. (Idealizing the materials as lossless metals extendsthe list of shapes for which the T-matrix is known

analytically [35], but this is not useful for radiative-transferproblems because lossless materials neither absorb norradiate energy.) To make predictions for shapes outsidethis narrow catalog we must turn instead to numericalmethods.

C. Numerical Approaches to Radiative TransferOne approach to numerical heat-transfer modeling is to

combine matrix-trace formulas in the spirit of (10) with anumerical technique for computing the T-matrices ofirregularly-shaped objects. This technique was pursued in[16], which investigated heat transfer from hot tips of va-rious shapes to a cool planar substrate at micron-scale dis-tances. In this work, a boundary-element scattering codewas used to compute numerical T-matrices for finite cylin-ders and finite-length cones; a surprising conclusion wasthat conical tips, despite tapering to a point, nonethelessexhibit less spatial concentration (i.e., a larger and morediffuse spot size) of heat transferred to the substrate ascompared to cylindrical tips.

An alternative numerical approach to heat-transfercalculations is to bypass the T-matrix approach in favor of amore direct assault on (9) [15], [36]; here a Bbrute-force[approach can deliver great generality with minimal prog-ramming time, at the expense of much computer time.Physically, the situation described by (9) is that we haverandomly fluctuating currents distributed throughout theinterior of our material bodies, and we wish to computethe fields to which these currents give rise. A particularlyconvenient way to do this computation is to run a time-domain simulation, in which we calculate the fields due toa random time-varying current distribution whose corre-lation function in the frequency domain satifies (2); byrepeating this calculation for many randomly-generatedcurrent distributions and averaging the results, we obtainapproximate ensemble averages of the time-domain E andH fields, which we may then Fourier-analyze to obtainfrequency spectra. This approach is rendered computa-tionally feasible by exploiting several properties of (2) andof Maxwell’s equations. First, absence of spatial correlation:the " function in (2) ensures that currents at differentlocations in space (in particular, currents in different bod-ies) are uncorrelated and may thus be chosen to haveindependent random phases. Second, linearity: although(2) calls for stochastic currents with nonflat spectral den-sity shaped by the factor #)!; T*Vwhat engineers mightthink of as Bcolored noise[Vthe linearity of Maxwell’sequations ensures that we can instead compute the fieldsdue to white-noise currents, which are significantly easierto generate in the time domain, and only later multiply theresulting frequency spectrum by the appropriate shapingfactor. Finally, reciprocity: the flux absorbed by body B2

due to radiating sources in B1 is equal to the flux absorbedby B1 due to sources in B2. This observation allows us toplace our stochastic sources only in B1 and compute theresulting flux only into B2.

Reid et al. : Fluctuation-Induced Phenomena in Nanoscale Systems: Harnessing the Power of Noise

536 Proceedings of the IEEE | Vol. 101, No. 2, February 2013

Page 7: INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale ...arod/papers/ReidRodriguez13_review.pdf · tricalengineersVincludingtheT-matrix,finite-difference, and boundary-element

Combining these arguments leads to a simple expres-sion for the spectral density of the net heat flux betweenbodies [15]

H"!; T1; T2# ! $"!# #)!; T1* %#)!; T2*f g

where $ is the flux into one of the objects due to random(white-noise) current sources in the other object. In prac-tice, $ is computed using a finite-difference time-domaintechnique, with random current sources placed at gridpoints throughout the volume of the bodies and the resultsaveraged over many ((60) simulations.

Fig. 3 illustrates the type of result that may be obtainedusing this method [15]. The solid curve in the figure plotsthe spectral density of power flux between two one-dimensional photonic crystals of SiC separated by a shortdistance d (inset). The dashed curve plots the power fluxbetween unpatterned SiC slabs. (In both cases, the powerflux is normalized by the flux between the same structuresat infinite separation d ! 1.) The patterning of the slabsdrastically modifies the flux spectrum as compared to theunpatterned case.

IV. CASIMIR FORCES:FLUCTUATION-INDUCED MOMENTUMTRANSFER IN NANOSCOPIC SYSTEMS

In the previous section, we considered applications offluctuation-dissipation ideas to situations out of thermal

equilibrium, and we noted the fierce computational chal-lenges that arise from the need to solve separate scatteringproblems for each point in the volume integration in (7).At thermal equilibrium, a major simplification occurswhich significantly reduces computational requirements.The situation is most clearly displayed by considering themean product ofE-field components, which reads, in closeanalogy to (7)

Ei"x#E$j "x0#

D E

!Z

dyGEik"x;y#G

E$jk "x

0;y## !; T"y#) *#"!;y#: (11)

The key point is that, at thermal equilibrium, T"y# . T isspatially constant, whereupon the statistical factor may bepulled out of the integral to yield

!#)!; T*Z

dyGEik"x;y#G

E$jk "x

0;y##"!;y#

! 1

!#)!; T*ImGE

ij "x;x0#: (12)

(In going to the last line here we used a standardidentity in electromagnetic theory which follows di-rectly from Maxwell’s equations [37].) Thus, evaluatinga mean product of field components at thermal equi-librium requires the solution of only a single scatteringproblem, in contrast to the formally infinite number ofscattering problems required for out-of-equilibriumsituations.

Of course, the heat-transfer calculations of the previoussection are not very interesting at thermal equilibrium, inwhich by definition there can be no net transfer of energybetween bodies. However, a different sort of fluctuation-induced phenomenonVthe Casimir effectVgives rise tonontrivial interactions among bodies even at the sametemperature (and even at zero temperature), and consti-tutes a second major branch of the study of electromag-netic fluctuations.

A. The Casimir EffectIn 1948 [38], Casimir and Polder generalized the van

der Waals (or BLondon dispersion[) force between fluc-tuating dipoles of molecules and other small particles,which depends on the distance r between the particles like1=r7, to a Bretarded[ force that varies like 1=r8 at largedistances (typically tens of nanometers) where the finitespeed of light must be taken into account. Later that year[4], Casimir considered the region between two parallelmirrors as a type of electromagnetic cavity, characterized

Fig. 3. Near-field radiative heat transfer between patterned and

unpatterned SiC slabs [15]. The solid black curve plots the spectral

density of power flux between SiC photonic crystals (inset)maintained

at inequal temperatures and surfaceVsurface separation d.The dashed red curve plots the power flux between unpatterned

SiC slabs. (In both cases, the power flux is normalized by the

power flux that would obtain between the same structures at

infinite separations d ! 1.)

Reid et al.: Fluctuation-Induced Phenomena in Nanoscale Systems: Harnessing the Power of Noise

Vol. 101, No. 2, February 2013 | Proceedings of the IEEE 537

Page 8: INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale ...arod/papers/ReidRodriguez13_review.pdf · tricalengineersVincludingtheT-matrix,finite-difference, and boundary-element

by a set of cavity-mode frequencies f!"d#g depending onthe mirror separation distance d. By summing the zero-point energies [see (3a)] of all modes and differentiatingwith respect to d, Casimir predicted an attractive pressurebetween the plates of magnitude

F

A! !2"hc

240d4+ 10%8 atm

"d in %m#4(13)

negligible at macroscopic distances but significant forsurfaceVsurface separations below a few hundrednanometers.

The Casimir effect was subsequently reinterpreted[39], [40] as an interaction among fluctuating charges andcurrents in material bodies, a perspective which allows theuse of fluctuation-dissipation formulas like (12) to predictCasimir forces in situations where the cavity-mode picturewould be unwieldy. In fact, the Casimir effect has beeninterpreted in a bewildering variety of ways; in addition tothe zero-point-energy picture of [4] and the source-fluctuation picture of [39], there are path-integral formu-lations [23], world-line methods [41], and ray-opticsapproaches [42], to name but a few. Each of these per-spectives emphasizes different aspects of the underlyingphysics, although of course all physical interpretations leadultimately to mathematically equivalent final results [24].However, despite the plethora of theoretical perspectives,and even with the simplifications afforded by thermalequilibrium, the calculations remained so challenging thatforce predictions for all but the simplest geometries werepractically out of reach, andVwith experimental progresshampered by the difficulty of measuring nanonewtonforces between bodies at submicron distancesVfor manydecades there was little demand for computational Casimirmethods that could handle general geometries andmaterials.

This situation began to change about 15 years ago withthe advent of precision Casimir metrology [43], and sincethat time the Casimir effect has been experimentally ob-served in an increasingly wide variety of geometric andmaterial configurations (for recent reviews of experimen-tal Casimir physics, see [17], [44]). This experimentalprogress has spurred the development of theoretical tech-niques capable of predicting Casimir forces and torques incomplex, asymmetric geometries with realistic materials,which we now review.

B. The Casimir Effect as a Scattering ProblemAs in Section III-A, we consider two bodies B1;2 in

vacuum. In (8) we integrated the average Poynting fluxover a surface surrounding a body to obtain the rate ofenergy transfer to that body. To compute the rate ofmomentum transfer to the bodyVthat is, the force on thebodyVwe proceed analogously, but now instead of the

Poynting flux we integrate the average Maxwell stresstensor

F "!# !Z

S

T"!;x#h i & dS (14)

where the components of T are given in terms of thecomponents of E and H as

Tij ! $0EiEj - %0HiHj %"ij2)$0EkEk - %0HkHk*:

Inserting (12) and its magnetic analogue into (14) nowyields an expression analogous to (9)Vbut simplified bythe absence of volume integralsVwhich at temperatureT ! 0 takes the form, for the i component of the force

F i"!# !"h!

!Im

Z

S

(

$0GEij "!;x;x# - %0GH

ij "!;x;x#

%"ij2

$0GEkk"!;x;x# - %0 GH

kk"!;x;x#' (

)

dSj (15)

Here G"x;x0#, the scattering part of a DGF G, is the con-tribution to G which remains finite as x0 ! x; this is justthe field at x due to currents induced by a point source atx0, but neglecting the direct contribution of that pointsource. [In (15), GH is the scattering part of the DGF thatrelates magnetic fields to magnetic currents.]

Equation (15), like (9), reduces our problem to that ofdetermining the DGFs for our material configuration, andin principle we could now proceed to evaluate the surfaceintegral in (15) with the integrand computed by standardscattering techniques. For Casimir calculations, however,the situation is complicated by an important subtlety,which we now discuss.

C. Transition to the Imaginary Frequency AxisIn contrast to the heat-transfer problems discussed in the

previous section, for Casimir problems we will not typicallybe interested in the contributions of individual frequen-cies but will instead seek only the total Casimir force on abody, obtained by integrating (15) over all frequencies:

Fi !Z1

0

F i"!# d!: (16)

Reid et al. : Fluctuation-Induced Phenomena in Nanoscale Systems: Harnessing the Power of Noise

538 Proceedings of the IEEE | Vol. 101, No. 2, February 2013

Page 9: INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale ...arod/papers/ReidRodriguez13_review.pdf · tricalengineersVincludingtheT-matrix,finite-difference, and boundary-element

But naBve attempts to evaluate (16) numerically aredoomed to failure by the existence of rapid oscillationsin the integrand, as pictured in Fig. 4(a) for the particularcase of the Casimir force between parallel metallic platesin vacuum. The origin of these oscillations is not hard toidentify: they are related to the existence of electromag-netic resonances in our scattering geometry, which corre-spond mathematically to poles of the integrand in thelower half of the complex ! plane. (The oscillatory natureof the force spectrum was emphasized in [45], and theimplications for numerical computations were discussedin [46].)

But this diagnosis of the problem suggests a cure:thinking of (16) as a contour integral in the complex fre-quency plane, we simply rotate the contour of integration90 degrees and integrate over the imaginary frequency axis[see Fig. 4(b)]. This procedure, known in physics as aWickrotation [47], yields

Fi !Z1

0

F i"(# d( (17)

where ! ! i( and F now involves the DGFs evaluated atimaginary frequencies

F i"(# !"h(

!

Z

S

n$0GE

ij "(;x;x# - %0GHij "(;x;x#

%"ij2, $0GE

kk"(;x;x# - %0 GHkk"(;x;x#

' (odSj:

(18)

The Wick rotation is possible here because the DGFs areanalytic functions in the upper half of the complex ! plane.This is a well-known consequence of causality: the fieldsarise after the current fluctuations that generate them[48]. Another consequence of causality is that, for passivematerials, the permittivity and permeability functions onthe imaginary frequency axis f$"i(#; %"i(#g are guaranteedto be real-valued and positive [49].

Physically, the transition to the imaginary frequencyaxis corresponds to replacing the oscillatory time de-pendence (e%i!t of all fields and currents with anexponentially growing time dependence ( e-(t; forfrequency-domain computational electromagnetism, thishas the effect of replacing the spatially oscillatoryHelmholtz kernel "ei!r=c=4!r# with an exponentially decay-ing kernel "e%(r=c=4!r#. As illustrated in Fig. 4(b), theimaginary-frequency Casimir force integrand F "(# is awell-behaved smooth function that succumbs readily tonumerical quadrature.

Equations (17) and (18) are valid at zero temperature.At finite temperatures T > 0, we must include a factor#)(; T* ( coth i"h(=2kT under the integral sign; in thiscase, it is well-known in physics [40] that the integral (17)over the imaginary frequency axis may be evaluated usingthe method of residues to obtain

Fi !2!kT

"h

X10

n!0

F i"(n# (19)

Fig. 4. Transition to the imaginary frequency axis. (a) As a function

of real frequency !, the Casimir force integrand F "!# of equation(15)Vshown here for the case of parallel metallic plates separated by

a distance a (inset)Vexhibits severe oscillations which effectively

prohibit evaluation of the integral (16) by numerical quadrature.

(For the particular case of parallel plates, the expression for the

Casimir force integrand is known as the Lifshitz formula [40].) These

oscillations are associated with cavity resonances, which show up

mathematically as poles in the lower half of the complex frequency

plane (inset); the real part of the pole corresponds to the resonance

frequency, while the imaginary part corresponds to the width (or the

inverse lifetime) of the resonance. (b) Rotating to the imaginary

frequency axis (inset) moves the contour of integration away from

the cavity-resonance poles, resulting in a smooth integrand that

succumbs readily to numerical quadrature.

Reid et al.: Fluctuation-Induced Phenomena in Nanoscale Systems: Harnessing the Power of Noise

Vol. 101, No. 2, February 2013 | Proceedings of the IEEE 539

Page 10: INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale ...arod/papers/ReidRodriguez13_review.pdf · tricalengineersVincludingtheT-matrix,finite-difference, and boundary-element

where (n ! 2n!kT="h, the Matsubara frequencies, are justthe poles of the coth factor on the imaginary frequencyaxis. [The primed sum in (19) indicates that the n ! 0term enters with weight 1/2.] Computationally, the upshotof (19) is that finite-temperature Casimir forces arecomputed with no more conceptual difficulty than zero-temperature forces, with the integral in (17) simply re-placed by the sum in (19), although the need to evaluate(18) in the limit of zero frequency "( ! 0-# poses chal-lenges for some methods of computational electromagne-tism [50], [51]. The temperature dependence of Casimirinteractions is a topic of recent theoretical [52] andexperimental [53] interest.

D. Semi-Analytical Approaches toCasimir Computations

Like the first studies of near-field radiative transfer,the first generation of theoretical Casimir techniquesfocused on highly symmetric geometries for whichanalytical scattering solutions are available [22], [23],[28], [54]–[58]. As an example of the type of conciseexpression that may be obtained via these methods, thezero-temperature Casimir force between two compactbodies with centerVcenter separation vector R may beexpressed in the form [9]

Fi !"h

2!

Z1

0

Tr M%1"(# & @M"(#@Ri

# $d( (20)

where the matrix M has the block structure

M ! T%11 U"R#

Uy"R# T%12

! "

here Tn is the T-matrix for body n and U"R# is a trans-lation matrix, which relates spherical Helmholtz solutionsabout different origins and for which closed-form analyticexpressions are available [59]. [The partial derivative in(20) is taken with respect to a rigid displacement of onebody in the ith cartesian direction.]

Like (10), the formula (20) is simple enough that it canbe implemented in just a few lines of mathematica ormatlab code for geometries in which the T-matrix isknown analytically. Again, however, such geometries arerare, and for more complicated geometric configurationswe must turn to numerical methods.

E. Numerical Approaches to Casimir ComputationsThe most direct way to apply numerical techniques to

Casimir computations is simply to evaluate the surfaceintegral in (18) by numerical cubature, with the G tensors

at each integrand point x evaluated by solving a numericalscattering problem in which we place a point source at xand compute the scattered fields back at the same point x.In principle, this scattering problem may be solved by anyof the myriad available techniques for numerical solutionof scattering problems (although the need for imaginary-frequency calculations poses something of a limitation inpractice). To date, computational Casimir methods basedon numerical evaluation of (18) have been implementedusing a variety of standard techniques in computationalelectromagnetism: the finite-difference frequency-domainmethod [46], [60], the finite-difference time-domainmethod [with some transformations to convert the integralover frequencies in (18) into an integral over the time-domain response of a current pulse] [61]–[63], and theboundary-element method [64], [65].

Compared to the special-function approaches discussedin Section IV-D, any one of these numerical methods offersthe significant practical advantage of handling arbitrarilycomplex geometries with little more difficulty than simplegeometries. Among the various numerical methods, thefinite-difference methods have the advantage of greatergeneralityVin the sense that they can readily handle arbi-trarily complex material configurations, including ani-sotropic and continuously-varying dielectricsVwhile theboundary-element methods have the advantage of greatercomputational efficiency for the piecewise-homogeneousmaterial configurations typically encountered in practice.

As an illustration of the type of problem that is faci-litated by numerical Casimir methods, Fig. 5 plots theforce between elongated square pistons confined between

Fig. 5. Casimir force between elongated pistons confined between

parallel plates [46]. The lower inset depicts the geometry, while the

upper inset shows the finite-difference grid used to model the

cross-section of this z-invariant structure. The force between

the pistons exhibits a surprising nonmonotonic dependence on the

separation distance h between the pistons and the plates. (The

quantity plotted is the actual force divided by the proximity-force

approximation (PFA) to the force, a convenient h-independentnormalization.)

Reid et al. : Fluctuation-Induced Phenomena in Nanoscale Systems: Harnessing the Power of Noise

540 Proceedings of the IEEE | Vol. 101, No. 2, February 2013

Page 11: INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale ...arod/papers/ReidRodriguez13_review.pdf · tricalengineersVincludingtheT-matrix,finite-difference, and boundary-element

parallel plates (all bodies are perfect conductors), as com-puted using a finite-difference technique [46]. The lowerinset in the figure depicts the geometry, while the upperinset shows the finite-difference grid used to model thecross-section of this z-invariant structure. The force be-tween the pistons exhibits a suprising non-monotonic de-pendence on the separation distance h between the pistonsand the plates.

F. Fluctuating-Surface-Current Approach toCasimir Computations

The finite-difference and boundary-element methodsdescribed above have the advantage of great generality, inthat they treat bodies of arbitrarily complex shapes with nomore difficulty than simple symmetric bodies. However,the need for numerical evaluation of the surface integral in(18) adds a layer of conceptual and computational com-plexity that is absent from the concise expression (20).

An alternative is the recently developed fluctuating-surface-current approach [10], [66]–[68]. In the FSC tech-nique, we begin with a boundary-element-method (BEM)approach to evaluating the DGFs in (18). Instead of pro-ceeding numerically, however, we exploit the structure ofthe BEM technique to obtain compact analytical expres-sions for the DGFs in fully-factorized form, involving pro-ducts of factors depending separately on the source andevaluation points. Inserting these expressions into (18)then turns out to allow the surface integral to be evaluatedanalytically, in closed form, leaving behind only straight-forward matrix manipulations [67], [68]. The final FSCformula for the Casimir force

Fi !"h

2!

Z1

0

Tr M%1"(# & @M"(#@Ri

# $d( (21)

bears a remarkable similarity to (20), but now with a dif-ferent matrix M entering into the matrix manipulations;whereas M in (20) describes the interactions betweenincoming and outgoing waves in a multipole expansion ofthe electromagnetic field, M in (21) describes the interac-tions among surface currents flowing on the surfaces of theinteracting objects in a Casimir geometry. [M"(# in (21) isjust the usual impedance matrix that enters into thePMCHW formulation of the boundary-element method[69], but now evaluated at imaginary frequencies.]

As one example of the type of calculation that is facili-tated by FSC Casimir techniques, [11] investigated theCasimir force on an elongated nanoparticle above a circularaperture in a metallic plate and identified a region of theforce curve in which the force on the particle is repulsive(Fig. 6). This geometry is notable as the only known con-figuration exhibiting repulsive Casimir forces betweennoninterleaved metallic objects in vacuum. (On the other

hand, repulsive forces between dielectric objects immersedin a dielectric liquid have long been known to exist andwere observed experimentally in 2009 [70]; in addition,numerical Casimir tools have been used to demonstratetheoretically the possibility of achieving stable suspensionof objects in fluids [71], and further work in this area mayhave applications in microfluidics.)

V. SUMMARY AND OUTLOOK

Despite spending most of its history confined to the realmof pure physics, the theory and experimental character-ization of fluctuation-induced electromagnetic phenomenais at last poised to take on a new role as a growth area inelectrical engineering. The growing ease and ubiquity ofnanotechnology are making near-field radiative transportand Casimir forces increasingly relevant to the technolo-gies of today and tomorrow, with a corresponding immi-nent need for engineers to account for these phenomena intheir designs. In this connection it is convenient that a hostof powerful computational methods, inspired by tech-niques of classical computational electromagnetism butextending these methods in several ways, have been devel-oped over the past several years to model various fluctua-tion phenomena. We hope to have convinced the reader

Fig. 6. Repulsive Casimir force between metallic objects in vacuum.

Plotted is the z-directed force on an elongated nanoparticle above

a circular aperture in a metallic plate (inset), as a function of the

separation distance d between the center of the nanoparticle and the

center of the plate. The dashed red curves are for the case of perfectly

conducting materials (for two different plate thicknesses), while the

solid blue curve is for the case of finite-conductivity gold. The shaded

region of the force curve indicates the repulsive regime, in which the

nanoparticle is repelled from the plate. (The dashed vertical line

denotes the separating plane, i.e., the value of d beyond which the

nanoparticle is entirely above the plate.) For comparison, the dashed

grey curve indicates the force on a perfectly-conducting spherical

nanoparticle; in this case the force is attractive at all separations.

(Figure reproduced from [11].).

Reid et al.: Fluctuation-Induced Phenomena in Nanoscale Systems: Harnessing the Power of Noise

Vol. 101, No. 2, February 2013 | Proceedings of the IEEE 541

Page 12: INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale ...arod/papers/ReidRodriguez13_review.pdf · tricalengineersVincludingtheT-matrix,finite-difference, and boundary-element

that the sudden conjunction of new theoretical techniques,increasing experimental relevance, and the paucity ofknown results have created burgeoning opportunities forcomputational scienceVindeed, in fields where twospheres represent a novel geometry, the untapped frontiersof design are vast and inviting.

What lies in store for the future of this field? The workreviewed in this article has answered many questions, onlyto pose many more to be addressed in the coming years.Here we give a brief flavor of some challenges that lie onthe horizon.

General-basis trace formulas for heat transfer. UnlikeCasimir forces, the theory of near-field radiation does notyet benefit from a compact trace formula that applies to anarbitrary localized basis. Existing approaches either re-quire the intermediary of a spectral incoming/outgoingwave basis (such as cylindrical or spherical waves) that maybe ill-suited for irregular geometries, or large-scale com-putations involving costly integral evaluations. Is a syn-thesis (in the spirit of the FSC approach of Section IV-F)possible or practical, and what form does it take?

Fast solvers. To date, practical applications of integral-equation Casimir techniques have evaluated the matrixoperations in (21) (matrix inverse, matrix multiplication,and matrix trace) using methods of dense-direct linearalgebra. These methods are appropriate for matrices ofmoderate dimension (D ( 104 or less), but for larger prob-lems the O"D2# memory scaling and O"D3# CPU-time scal-ing of dense-direct linear algebra renders calculationsintractable. A similar bottleneck was encountered manyyears ago in the computational electromagnetism com-munity, where it was remedied by the advent of fastsolversVtechniques such as the fast multipole [72] andprecorrected-FFT [73] methods that employ matrix-sparsification techniques to reduce the asymptotic com-plexity scaling of matrix operations to more manageablelevels;O"D3=2 logD# [74] orO"D logD# [75], [76] are typical.Although such methods could, in principle, be applied tostress-tensor Casimir computations [46], [64], can they bemade practical? Can they be applied to the FSC trace-formulaapproach, and with what performance implications?

New experimental geometries. Until recently, theoreticaltechniques in fluctuation-induced phenomena laggedbehind the forefront of experimental progress (indeed, aswe have seen, it is only in the past few years that complete

theoretical solutions for the simple sphereVplate geom-etry commonly seen in experiments have become avail-able). This situation has recently begun to change; with ahost of new computational methods for near-field radiativetransfer and Casimir phenomena becoming available in thepast five years, we are entering an era in which theoreticalpredictions can be used to guide the design of futureexperimentsVand, ultimately, future technologies. Such areversal is not without precedent in the history of elec-trical engineering. Indeed, whereas the first computationalalgorithms for modeling antennas and transistor circuitswere validated by checking that they correctly reproducedthe behavior of existing laboratory systems, today it wouldbe unthinkable to fabricate a patch antenna or an integ-rated operational amplifier without first carefully vettingthe design using CAD tools. Will the development of so-phisticated modeling tools for near-field radiative transferand Casimir phenomena transform those fields as thor-oughly as SPICE and its descendants transformed circuitengineering? In the former case, can we use modelingtools to design efficient tipVsurface geometries for ther-mal lithography, or to invent new solar-cell configurationsthat exploit the interplay of material and geometric pro-perties to optimize power absorption and retention at solarwavelengths? In the latter case, can we use computationaltools to understand parasitic Casimir interactions amongmoving parts in MEMS devicesVor to invent new MEMSdevices that exploit Casimir forces and torques to usefulends?

All of these are questions for the future of fluctuation-induced phenomena. We hope in this review to havepiqued the curiosity of electrical engineers in this rapidlydeveloping fieldVand to have encouraged readers to staytuned for future developments.

In closing, we note that all of the computational resultspresented in this review were obtained using freely-available open-source software packages for computationalelectromagnetism: meep, a finite-difference solver, andscuff-em, a boundary-element solver. (Both packages areavailable for download at http://ab-initio.mit.edu/wiki.) Inaddition to their general applicability to scatteringcalculations and other problems in computational electro-magnetism, these codes offer specialized modules imple-menting algorithms discussed in this article for numericalmodeling of fluctuation-induced phenomena. h

REFERENCES

[1] P. R. Gray, P. J. Hurst, S. H. Lewis, andR. G. Meyer, Analysis and Design of AnalogIntegrated Circuits. Hoboken, NJ: Wiley,2009.

[2] A. I. Volokitin and B. N. J. Persson,BNear-field radiative heat transfer andnoncontact friction,[ Rev. Mod. Phys.,vol. 79, pp. 1291–1329, Oct. 2007.[Online]. Available: http://link.aps.org/doi/10.1103/RevModPhys.79.1291.

[3] V. Parsegian, Van der Waals Forces: AHandbook for Biologists, Chemists, Engineers,and Physicists. Cambridge, U.K.: CambridgeUniv. Press, 2006.

[4] H. B. G. Casimir, BOn the attractionbetween two perfectly conducting plates,[Koninkl. Ned. Adak. Wetenschap. Proc., vol. 51,p. 793, 1948.

[5] T. H. Boyer, BQuantum electromagneticzero-point energy of a conducting sphericalshell and the Casimir model for a chargedparticle,[ Phys. Rev., vol. 174, pp. 1764–1776,

Oct. 1968. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRev.174.1764.

[6] M. S. Tomas, BCasimir force in absorbingmultilayers,[ Phys. Rev. A., vol. 66, p. 052103,Nov. 2002. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRevA.66.052103.

[7] F. Zhou and L. Spruch, Bvan der Waals andretardation (Casimir) interactions of anelectron or an atom with multilayered walls,[Phys. Rev. A., vol. 52, pp. 297–310, Jul. 1995.[Online]. Available: http://link.aps.org/doi/10.1103/PhysRevA.52.297.

Reid et al. : Fluctuation-Induced Phenomena in Nanoscale Systems: Harnessing the Power of Noise

542 Proceedings of the IEEE | Vol. 101, No. 2, February 2013

Page 13: INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale ...arod/papers/ReidRodriguez13_review.pdf · tricalengineersVincludingtheT-matrix,finite-difference, and boundary-element

[8] R. Buscher and T. Emig, BNonperturbativeapproach to Casimir interactions in periodicgeometries,[ Phys. Rev. A., vol. 69, p. 062101,Jun. 2004. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRevA.69.062101.

[9] T. Emig, N. Graham, R. L. Jaffe, andM. Kardar, BCasimir forces between arbitrarycompact objects,[ Phys. Rev. Lett., vol. 99,p. 170403, Oct. 2007. [Online]. Available:http://link.aps.org/doi/10.1103/PhysRevLett.99.170403.

[10] M. T. H. Reid, A. W. Rodriguez, J. White,and S. G. Johnson, BEfficient computation ofCasimir interactions between arbitrary 3dobjects,[ Phys. Rev. Lett., vol. 103, p. 040401,Jul. 2009. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRevLett.103.040401.

[11] M. Levin, A. P. McCauley, A. W. Rodriguez,M. T. H. Reid, and S. G. Johnson, BCasimirrepulsion between metallic objects invacuum,[ Phys. Rev. Lett., vol. 105, p. 090403,Aug. 2010. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRevLett.105.090403.

[12] S. Rytov, Theory of Electric Fluctuations andThermal Radiation, AFCRC-TR. ElectronicsResearch Directorate, Air Force CambridgeResearch Center, Air Research andDevelopment Command, U.S. Air Force,1959.

[13] D. Polder and M. Van Hove, BTheory ofradiative heat transfer between closely spacedbodies,[ Phys. Rev. B., vol. 4, pp. 3303–3314,Nov. 1971. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRevB.4.3303.

[14] A. Narayanaswamy and G. Chen, BThermalnear-field radiative transfer between twospheres,[ Phys. Rev. B., vol. 77, p. 075125,Feb. 2008. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRevB.77.075125.

[15] A. W. Rodriguez, O. Ilic, P. Bermel,I. Celanovic, J. D. Joannopoulos, M. Soljacic,and S. G. Johnson, BFrequency-selectivenear-field radiative heat transfer betweenphotonic crystal slabs: A computationalapproach for arbitrary geometries andmaterials,[ Phys. Rev. Lett., vol. 107,p. 114302, Sep. 2011. [Online]. Available:http://link.aps.org/doi/10.1103/PhysRevLett.107.114302.

[16] A. P. McCauley, M. T. Homer Reid,M. Kruger, and S. G. Johnson, BModelingnear-field radiative heat transfer fromsharp objects using a general 3d numericalscattering technique,[ ArXiv e-prints,Jul. 2011. [Online]. Available: http://arxiv.org/abs/1107.2111

[17] A. W. Rodriguez, F. Capasso, andS. G. Johnson, BThe Casimir effect inmicrostructured geometries,[ Nature Photon.,vol. 5, pp. 211–221, Mar. 2011, invited review.

[18] R. Harrington, Time-Harmonic ElectromagneticFields. Piscataway, NJ: IEEE Press, 1961,ser. Electromagnetic Wave Theory.

[19] G. L. Klimchitskaya, U. Mohideen, andV. M. Mostepanenko, BCasimir and van derWaals forces between two plates or a sphere(lens) above a plate made of real metals,[Phys. Rev. A., vol. 61, p. 062107, May 2000.[Online]. Available: http://link.aps.org/doi/10.1103/PhysRevA.61.062107.

[20] P. A. Maia Neto, A. Lambrecht, andS. Reynaud, BCasimir energy between a planeand a sphere in electromagnetic vacuum,[Phys. Rev. A., vol. 78, p. 012115, Jul. 2008.[Online]. Available: http://link.aps.org/doi/10.1103/PhysRevA.78.012115.

[21] K. A. Milton and J. Wagner, BExactexpressions for the Casimir interactionbetween semitransparent spheres and

cylinders,[ Phys. Rev. D., vol. 77, p. 045005,Feb. 2008. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRevD.77.045005.

[22] A. Lambrecht, P. A. M. Neto, and S. Reynaud,BThe Casimir effect within scattering theory,[New J. Phys., vol. 8, no. 10, p. 243, 2006.[Online]. Available: http://stacks.iop.org/1367-2630/8/i=10/a=243.

[23] S. J. Rahi, T. Emig, N. Graham, R. L. Jaffe,and M. Kardar, BScattering theory approachto electrodynamic Casimir forces,[ Phys. Rev.D., vol. 80, p. 085021, Oct. 2009. [Online].Available: http://link.aps.org/doi/10.1103/PhysRevD.80.085021.

[24] S. G. Johnson, BNumerical methods forcomputing Casimir interactions,[ arXiv.orge-Print archive, Jul. 2010, to appear inupcoming Lecture Notes in Physics bookon Casimir Physics. [Online]. Available: http://arxiv.org/abs/1007.0966

[25] H. B. Callen and T. A. Welton, BIrreversibilityand generalized noise,[ Phys. Rev., vol. 83,pp. 34–40, Jul. 1951. [Online]. Available:http://link.aps.org/doi/10.1103/PhysRev.83.34.

[26] A. Narayanaswamy, S. Shen, and G. Chen,BNear-field radiative heat transfer between asphere and a substrate,[ Phys. Rev. B.,vol. 78, p. 115303, Sep. 2008. [Online].Available: http://link.aps.org/doi/10.1103/PhysRevB.78.115303.

[27] M. Kruger, T. Emig, and M. Kardar,BNonequilibrium electromagneticfluctuations: Heat transfer and interactions,[Phys. Rev. Lett., vol. 106, p. 210404, May 2011.[Online]. Available: http://link.aps.org/doi/10.1103/PhysRevLett.106.210404.

[28] G. Bimonte, BScattering approach to Casimirforces and radiative heat transfer fornanostructured surfaces out of thermalequilibrium,[ Phys. Rev. A., vol. 80, p. 042102,Oct. 2009. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRevA.80.042102.

[29] C. Otey and S. Fan, BExact microscopic theoryof electromagnetic heat transfer between adielectric sphere and plate,[ ArXiv e-prints,Mar. 2011. [Online]. Available: http://arxiv.org/abs/1103.2668

[30] R. Messina and M. Antezza, BCasimir-Lifshitzforce out of thermal equilibrium and heattransfer between arbitrary bodies,[ Europhys.Lett. (EPL)., vol. 95, no. 6, p. 61002, 2011.[Online]. Available: http://stacks.iop.org/0295-5075/95/i=6/a=61002.

[31] R. Messina and M. Antezza,BScattering-matrix approach toCasimir-Lifshitz force and heat transfer outof thermal equilibrium between arbitrarybodies,[ Phys. Rev. A, vol. 84, p. 042102,Oct. 2011. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRevA.84.042102.

[32] R. Guerout, J. Lussange, F. S. S. Rosa,J.-P. Hugonin, D. A. R. Dalvit, J.-J. Greffet,A. Lambrecht, and S. Reynaud, BEnhancedradiative heat transfer betweennanostructured gold plates,[ ArXiv e-prints,Mar. 2012. [Online]. Available: http://arxiv.org/abs/1203.1496.

[33] P. Waterman, BMatrix formulationof electromagnetic scattering,[ Proc.IEEE, vol. 53, no. 8, pp. 805–812,Aug. 1965.

[34] M. Kruger et al., to appear.

[35] M. F. Maghrebi, S. J. Rahi, T. Emig,N. Graham, R. L. Jaffe, and M. Kardar,BAnalytical results on Casimir forces forconductors with edges and tips,[ Proc. Nat.Academy Sci., vol. 108, pp. 6867–6871,Apr. 2011.

[36] C. Luo, A. Narayanaswamy, G. Chen, andJ. D. Joannopoulos, BThermal radiation fromphotonic crystals: A direct calculation,[ Phys.Rev. Lett., vol. 93, pp. 213 905–213 908,Nov. 2004. [Online]. Available: http://link.aps.org/abstract/PRL/v93/e213905.

[37] S. Scheel and S. Yoshi Buhmann,BMacroscopic QEDVConcepts andapplications,[ Acta Physica Slovaca,vol. 58, pp. 675–809, 2008.

[38] H. B. G. Casimir and D. Polder, BTheinfluence of retardation on the London-vander Waals forces,[ Phys. Rev., vol. 73,pp. 360–372, Feb. 1948. [Online]. Available:http://link.aps.org/doi/10.1103/PhysRev.73.360.

[39] E. M. L. I. E. Dzyaloshinkii andL. P. Pitaevskii, Sov. Phys. Usp.vol. 4, p. 153, 1961.

[40] E. M. Lifschitz and L. P. Pitaevskii, StatisticalPhysics: Part 2. Oxford, U.K.: Pergamon,1980.

[41] H. Gies and K. Klingmuller, BWorldlinealgorithms for Casimir configurations,[ Phys.Rev. D, vol. 74, p. 045002, Aug. 2006.[Online]. Available: http://link.aps.org/doi/10.1103/PhysRevD.74.045002.

[42] R. L. Jaffe and A. Scardicchio, BCasimir effectand geometric optics,[ Phys. Rev. Lett., vol. 92,p. 070402, Feb. 2004. [Online]. Available:http://link.aps.org/doi/10.1103/PhysRevLett.92.070402.

[43] S. K. Lamoreaux, BDemonstration of theCasimir force in the 0.6 to 6 %m range,[ Phys.Rev. Lett., vol. 78, pp. 5–8, Jan. 1997.[Online]. Available: http://link.aps.org/doi/10.1103/PhysRevLett.78.5.

[44] F. Capasso, J. Munday, D. Iannuzzi, andH. Chan, BCasimir forces and quantumelectrodynamical torques: Physics andnanomechanics,[ IEEE J. Sel. Topics QuantumElectron., vol. 13, no. 2, pp. 400–414,Mar./Apr. 2007.

[45] L. H. Ford, BSpectrum of the Casimir effectand the Lifshitz theory,[ Phys. Rev. A., vol. 48,pp. 2962–2967, Oct. 1993. [Online].Available: http://link.aps.org/doi/10.1103/PhysRevA.48.2962.

[46] A. Rodriguez, M. Ibanescu, D. Iannuzzi,J. D. Joannopoulos, and S. G. Johnson,BVirtual photons in imaginary time:Computing exact Casimir forces via standardnumerical electromagnetism techniques,[Phys. Rev. A., vol. 76, p. 032106, Sep. 2007.[Online]. Available: http://link.aps.org/doi/10.1103/PhysRevA.76.032106.

[47] S. Weinberg, The Quantum Theory of Fields,Volume 1. Cambridge, U.K.: CambridgeUniv. Press, 1996.

[48] J. D. Jackson, Classical Electrodynamics.Hoboken, NJ: Wiley, 1999.

[49] L. Landau and E. Lifshits, Electrodynamicsof Continuous Media, vol. 8. Oxford, U.K.:Pergamon Press, 1960.

[50] J.-S. Zhao and W. C. Chew, BIntegral equationsolution of Maxwell’s equations from zerofrequency to microwave frequencies,[ IEEETrans. Antennas Propag., vol. 48, no. 10,pp. 1635–1645, Oct. 2000.

[51] C. L. Epstein and L. Greengard, BDebyesources and the numerical solution of thetime harmonic Maxwell equations,[ Commun.Pure Appl. Math., vol. 63, no. 4, pp. 413–463,2010. [Online]. Available: http://dx.doi.org/10.1002/cpa.20313.

[52] A. W. Rodriguez, D. Woolf, A. P. McCauley,F. Capasso, J. D. Joannopoulos, andS. G. Johnson, BAchieving a stronglytemperature-dependent Casimir effect,[ Phys.

Reid et al.: Fluctuation-Induced Phenomena in Nanoscale Systems: Harnessing the Power of Noise

Vol. 101, No. 2, February 2013 | Proceedings of the IEEE 543

Page 14: INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale ...arod/papers/ReidRodriguez13_review.pdf · tricalengineersVincludingtheT-matrix,finite-difference, and boundary-element

Rev. Lett., vol. 105, p. 060401, Aug. 2010.[Online]. Available: http://link.aps.org/doi/10.1103/PhysRevLett.105.060401.

[53] A. O. Shuskov, W. J. Kim, D. A. R. Dalvit, andS. K. Lamoreaux, BObservation of the thermalCasimir force,[ Nature Phys., pp. 230–233,Feb. 2011.

[54] K. A. Milton and J. Wagner, BMultiplescattering methods in Casimir calculations,[J. Phys. A: Mathematical and Theoretical.,vol. 41, no. 15, p. 155402, 2008. [Online].Available: http://stacks.iop.org/1751-8121/41/i=15/a=155402.

[55] C. Genet, A. Lambrecht, and S. Reynaud,BCasimir force and the quantum theory oflossy optical cavities,[ Phys. Rev. A., vol. 67,p. 043811, Apr. 2003. [Online]. Available:http://link.aps.org/doi/10.1103/PhysRevA.67.043811.

[56] P. A. Maia Neto, A. Lambrecht, andS. Reynaud, BCasimir energy between a planeand a sphere in electromagnetic vacuum,[Phys. Rev. A., vol. 78, p. 012115, Jul. 2008.[Online]. Available: http://link.aps.org/doi/10.1103/PhysRevA.78.012115.

[57] P. S. Davids, F. Intravaia, F. S. S. Rosa, andD. A. R. Dalvit, BModal approach to Casimirforces in periodic structures,[ Phys. Rev. A.,vol. 82, p. 062111, Dec. 2010. [Online].Available: http://link.aps.org/doi/10.1103/PhysRevA.82.062111.

[58] O. Kenneth and I. Klich, BCasimir forces in aT-operator approach,[ Phys. Rev. B., vol. 78,p. 014103, Jul. 2008. [Online]. Available:http://link.aps.org/doi/10.1103/PhysRevB.78.014103.

[59] R. Wittmann, BSpherical wave operatorsand the translation formulas,[ IEEETrans. Antennas Propag., vol. 36, no. 8,pp. 1078–1087, Aug. 1988.

[60] S. Pasquali and A. Maggs,BFluctuation-induced interactions betweendielectrics in general geometries,[ J. Chem.Phys., vol. 129, no. 1, 2008.

[61] A. W. Rodriguez, A. P. McCauley,J. D. Joannopoulos, and S. G. Johnson,BCasimir forces in the time domain: Theory,[

Phys. Rev. A., vol. 80, p. 012115, Jul. 2009.[Online]. Available: http://link.aps.org/doi/10.1103/PhysRevA.80.012115.

[62] A. P. McCauley, A. W. Rodriguez,J. D. Joannopoulos, and S. G. Johnson,BCasimir forces in the time domain:Applications,[ Phys. Rev. A., vol. 81, p. 012119,Jan. 2010. [Online]. Available: http://link.aps.org/doi/10.1103/PhysRevA.81.012119.

[63] K. Pan, A. P. McCauley, A. W. Rodriguez,M. T. H. Reid, J. K. White, and S. G. Johnson,BCalculation of nonzero-temperature Casimirforces in the time domain,[ Phys. Rev. A.,vol. 83, p. 040503, Apr. 2011. [Online].Available: http://link.aps.org/doi/10.1103/PhysRevA.83.040503.

[64] J. L. Xiong and W. C. Chew, BEfficientevaluation of Casimir force in z-invariantgeometries by integral equationmethods,[ Appl. Phys. Lett., vol. 95, no. 15,pp. 154 102–154 102-3, Oct. 2009.

[65] J. L. Xiong, M. S. Tong, P. Atkins, andW. C. Chew, BEfficient evaluation ofCasimir force in arbitrary three-dimensionalgeometries by integral equation methods,[Phys. Lett. A, vol. 374, pp. 2517–2520,May 2010.

[66] M. T. H. Reid, J. White, and S. G. Johnson,BComputation of Casimir interactionsbetween arbitrary three-dimensional objectswith arbitrary material properties,[ Phys. Rev.A., vol. 84, p. 010503, Jul. 2011. [Online].Available: http://link.aps.org/doi/10.1103/PhysRevA.84.010503.

[67] M. T. H. Reid, BFluctuating Surface Currents:A New Algorithm for Efficient Predictionof Casimir Interactions Among ArbitraryMaterials in Arbitrary Geometries,[ Ph.D.dissertation, Massachusetts Inst. Technol.,Cambridge, MA, Feb. 2011.

[68] M. T. H. Reid, J. White, and S. G. Johnson,BFluctuating surface currents: A newalgorithm for efficient prediction of Casimirinteractions among arbitrary materials inarbitrary geometries. I. Theory,[ ArXive-prints, Feb. 2012. [Online]. Available:http://arxiv.org/abs/1203.0075v2.

[69] L. N. Medgyesi-Mitschang, J. M. Putnam,and M. B. Gedera, BGeneralized method ofmoments for three-dimensional penetrablescatterers,[ J. Opt. Soc. Amer. A., vol. 11, no. 4,pp. 1383–1398, Apr. 1994. [Online].Available: http://josaa.osa.org/abstract.cfm?URI=josaa-11-4-1383.

[70] J. N. Munday, F. Capasso, and V. Parsegian,BMeasured long-range repulsiveCasimir-Lifshitz forces,[ Nature (London),vol. 457, p. 170, 2009.

[71] A. W. Rodriguez, A. P. McCauley, D. Woolf,F. Capasso, J. D. Joannopoulos, andS. G. Johnson, BNontouching nanoparticlediclusters bound by repulsive and attractiveCasimir forces,[ Phys. Rev. Lett., vol. 104,p. 160402, Apr. 2010. [Online]. Available:http://link.aps.org/doi/10.1103/PhysRevLett.104.160402.

[72] L. Greengard and V. Rokhlin, BA fastalgorithm for particle simulations,[ J. Comput.Phys., vol. 73, no. 2, pp. 325–348, 1987.[Online]. Available: http://www.sciencedirect.com/science/article/pii/0021999187901409.

[73] J. Phillips and J. White, BA precorrected-FFTmethod for electrostatic analysis ofcomplicated 3-d structures,[ IEEE Trans.Comput.-Aided Design Integr. Circuits Systs.,vol. 16, no. 10, pp. 1059–1072, Oct. 1997.

[74] X.-C. Nie, L.-W. Li, and N. Yuan,BPrecorrected-FFT algorithm for solvingcombined field integral equations inelectromagnetic scattering,[ in Proc. Int.Symp. Antennas Propag. Soc., 2002, vol. 3,pp. 574–577.

[75] N. Yuan, T. S. Yeo, X.-C. Nie, and L. W. Li,BA fast analysis of scattering and radiationof large microstrip antenna arrays,[ IEEETrans. Antennas Propag., vol. 51, no. 9,pp. 2218–2226, Sep. 2003.

[76] T. Moselhy, X. Hu, and L. Daniel, BPFFT inFastMaxwell: A fast impedance extractionsolver for 3d conductor structures oversubstrate,[ in Proc. Design, Autom., Test EuropeConf. Exhibition, Apr. 2007, pp. 1–6.

ABOUT THE AUTHORS

M. T. Homer Reid received the B.A. degree in

physics from Princeton University, Princeton, NJ,

in 1998 and the Ph.D. degree in physics from the

Massachusetts Institute of Technology (MIT),

Cambridge, in 2011.

From 1998 to 2003, he was Member of Techni-

cal Staff (analog and RF integrated circuit design)

at Lucent Technologies Microelectronics and

Agere Systems. He is currently a Postdoctoral Re-

search Associate in the Research Laboratory of

Electronics at MIT. His research interests include computational methods

for classical electromagnetism, fluctuation-induced phenomena, quan-

tum field theory, electronic structure, and quantum chemistry. He is de-

veloper and distributor of SCUFF-EM, a free, open-source software package

for boundary-element analysis of problems in electromagnetism, in-

cluding nanophotonics, passive RF component modeling, and Casimir

phenomena.

Alejandro W. Rodriguez received the B.Sc. and

Ph.D. degrees in physics from the Massachusetts

Institute of Technology (MIT), Cambridge, in 2006

and 2010, respectively.

He is currently a joint Postdoctoral Fellow at the

Harvard School of Engineering and Applied Sciences

and at the MIT Department of Mathematics, working

in the areas of computational fluctuation-induced

interactions and nanophotonics. His work comprises

of some of the earliest numerical methods for

Casimir calculations, including the first demonstration of unusual, nonad-

ditive, three-body Casimir effects. He is the coauthor of over 35 publications

and four patents in the areas of nonlinear nanophotonics, Casimir and

optomechanical forces, and nonequilibriumnear-field radiative transport. In

addition to his research interests, he participates actively in educational

initiatives aimed at motivating young students to pursue careers in science

and engineering: he was featured in the Spanish-language network

Univision, and on the APS Physics central website, as part of educational

campaigns to increase the number of graduates and underrepresented

minorities in STEM fields, and is currently the professional advisor at the

Harvard Society of Mexican American Engineers and Scientists.

Dr. Rodriguez was named a Department of Energy Computational

Science Graduate Fellow from 2006-2010, was the recipient of the 2011

Department of Energy Fredrick Howes Award in Computational Science,

and was chosen as a World Economic Forum Global Shaper in 2011.

Reid et al. : Fluctuation-Induced Phenomena in Nanoscale Systems: Harnessing the Power of Noise

544 Proceedings of the IEEE | Vol. 101, No. 2, February 2013

Page 15: INVITED P A P E R Fluctuation-Induced PhenomenainNanoscale ...arod/papers/ReidRodriguez13_review.pdf · tricalengineersVincludingtheT-matrix,finite-difference, and boundary-element

Steven G. Johnson received the B.Sc. degree in

physics, mathematics, and computer science from

the Massachusetts Institute of Technology (MIT),

Cambridge, in 1995 and the Ph.D. degree in physics

from MIT in 2001.

Currently, he is Associate Professor of Ap-

plied Mathematics at MIT, where he joined the

Faculty of the Department of Mathematics in

2004. He is active in the field of nanophotonicsV

electromagnetism in media structured on the

wavelength scale, especially in the infrared and optical regimesVwhere

he works on many aspects of the theory, design, and computational

modeling of nanophotonic devices. He is coauthor of over 150 papers and

over 25 patents, including the second edition of the textbook Photonic

Crystals: Molding the Flow of Light, and was ranked among the top ten

most-cited authors in the field of photonic crystals by ScienceWatch.com

in 2008. Since 2007, his work has extended from primarily classical

nanophotonics into the modeling of interactions induced by quantum

and thermal electromagnetic interactions, and in particular, Casimir

forces and near-field radiative transport. In addition to traditional

publications, he distributes several widely used free-software packages

for scientific computation, including the MPB and Meep electromagnetic

simulation tools (cited in over 1000 papers to date) and the FFTW fast

Fourier transform library.

Dr. Johnson received the J. H. Wilkinson Prize for Numerical Software

in 1999.

Reid et al.: Fluctuation-Induced Phenomena in Nanoscale Systems: Harnessing the Power of Noise

Vol. 101, No. 2, February 2013 | Proceedings of the IEEE 545