Top Banner
Coherent Spontaneous Emission of Light Due to Surface Waves Jean-Jacques Greffet 1 , Remi Carminati 1 , Karl Joulain 1 , Jean-Philippe Mulet 1 , Carsten Henkel 2 , Stephane Mainguy 3 , and Yong Chen 4 1 Laboratoire d’Energ´ etique Mol´ eculaire et Macroscopique, Combustion, Ecole Centrale Paris, Centre National de la Recherche Scientifique, 92295 Chˆ atenay-Malabry Cedex, France [email protected] 2 Institut f¨ ur Physik, Universit¨at Potsdam, 14469 Potsdam, Germany 3 CEA/CESTA, 33114 Le Barp, France 4 Laboratoire de Photonique et Nanostructures, Centre National de la Recherche Scientifique, Route de Nozay, 91460 Marcoussis, France Abstract. Surface-phonon polaritons produce peaks in the local density of elec- tromagnetic states close to the interface. We show that this affects dramatically the spontaneous emission of light by thermal sources in the near field. We discuss several effects such as coherence properties in the near field, radiative heat transfer at short distances and the design of coherent thermal sources in the near field and in the far field. 1 Introduction When describing different possible sources of light, lasers are often presented as the typical coherent source of light and thermal sources such as a light bulb are often presented as the typical incoherent source. Indeed, for conven- tional thermal sources, the spectrum is chiefly given by the Planck function, so that the degree of temporal coherence is close to that of blackbody ra- diation. Concerning spatial coherence, it has been shown that light accross a planar Lambertian source, at a given wavelength λ, is correlated over a dis- tance on the order of λ/2[1]. These two results support the statement that conventional thermal sources have a very low degree of temporal and spatial coherence, and therefore are good approximations of incoherent sources. In deriving the above results, the surface waves are ignored. This is correct insofar as the far field is concerned. However, study of the phenomenon of emission of light into the near field, (i.e. at distances smaller than the peak wavelength from the source) may reveal unexpected behavior. It has been shown recently that the energy density close to an interface increases by several orders of magnitude and acquires a very narrow spectrum [2]. This effect has been shown to be due to the excitation of surface modes. This J. Tominaga and D.P. Tsai (Eds.): Optical Nanotechnologies, Topics Appl. Phys. 88, 163–182 (2003) c Springer-Verlag Berlin Heidelberg 2003
20

Coherent Spontaneous Emission of Light Due to Surface Waves

May 14, 2023

Download

Documents

Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Coherent Spontaneous Emission of Light Due to Surface Waves

Coherent Spontaneous Emission of Light

Due to Surface Waves

Jean-Jacques Greffet1, Remi Carminati1, Karl Joulain1,Jean-Philippe Mulet1, Carsten Henkel2,Stephane Mainguy3, and Yong Chen4

1 Laboratoire d’Energetique Moleculaire et Macroscopique, Combustion,Ecole Centrale Paris, Centre National de la Recherche Scientifique,92295 Chatenay-Malabry Cedex, [email protected]

2 Institut fur Physik, Universitat Potsdam,14469 Potsdam, Germany

3 CEA/CESTA, 33114 Le Barp, France4 Laboratoire de Photonique et Nanostructures, Centre National de la RechercheScientifique, Route de Nozay, 91460 Marcoussis, France

Abstract. Surface-phonon polaritons produce peaks in the local density of elec-tromagnetic states close to the interface. We show that this affects dramaticallythe spontaneous emission of light by thermal sources in the near field. We discussseveral effects such as coherence properties in the near field, radiative heat transferat short distances and the design of coherent thermal sources in the near field andin the far field.

1 Introduction

When describing different possible sources of light, lasers are often presentedas the typical coherent source of light and thermal sources such as a lightbulb are often presented as the typical incoherent source. Indeed, for conven-tional thermal sources, the spectrum is chiefly given by the Planck function,so that the degree of temporal coherence is close to that of blackbody ra-diation. Concerning spatial coherence, it has been shown that light accrossa planar Lambertian source, at a given wavelength λ, is correlated over a dis-tance on the order of λ/2 [1]. These two results support the statement thatconventional thermal sources have a very low degree of temporal and spatialcoherence, and therefore are good approximations of incoherent sources.

In deriving the above results, the surface waves are ignored. This is correctinsofar as the far field is concerned. However, study of the phenomenon ofemission of light into the near field, (i.e. at distances smaller than the peakwavelength from the source) may reveal unexpected behavior. It has beenshown recently that the energy density close to an interface increases byseveral orders of magnitude and acquires a very narrow spectrum [2]. Thiseffect has been shown to be due to the excitation of surface modes. This

J. Tominaga and D.P. Tsai (Eds.): Optical Nanotechnologies,Topics Appl. Phys. 88, 163–182 (2003)c© Springer-Verlag Berlin Heidelberg 2003

Page 2: Coherent Spontaneous Emission of Light Due to Surface Waves

164 Jean-Jacques Greffet et al.

suggests that heat transfer in the near field may become very important forvery small spacings. This is an important feature for most of the devicesthat are currently being developed to operate at nanoscales [3,4], such aslocal probe thermal microscopy [5]. The influence of evanescent waves wasstudied in pioneering works almost forty years ago [6,7], and more recentlyby Loomis and Maris [8]. However, the dramatic enhancement due to theresonant excitation of surface modes was ignored. We will show that theradiative flux may be enhanced by several orders of magnitude due to non-propagating fields, and that it may occur at particular frequencies [9].

A major difference between a laser and a thermal source is that thelaser produces a highly directional beam, whereas thermal sources are quasi-lambertian. It has been shown recently [10,11] that the excitation of surfacewaves modifies dramatically the spatial coherence of surface waves. As a con-sequence, directional thermal sources or, in other words, partially spatiallycoherent sources can be designed [12]. We will show that it is possible tofabricate a thermal source that emits light in narrow lobes as does a radioantenna. The physical mechanism will be discussed.

2 Spectral Properties of Emitted Thermal Near Fields

In this section, we consider the emission of a thermal source in the near field.The thermal source is modeled using the geometry depicted in Fig. 1. An in-terface at z = 0 separates a vacuum (medium z < 0) from a semi-infinite ab-sorbing material (medium z > 0), held at a temperature T . The medium is as-sumed to be isotropic, local and non-magnetic. It is described macroscopicallyby its complex frequency-dependent dielectric constant ε(ω) = ε′(ω)+iε′′(ω).

Fig. 1. Geometry of the system used to study ther-mal emission spectra and near-field spatial coher-ence. The upper half-space is a vacuum, the lowerhalf-space is a dielectric material with dielectricconstant ε at temperature T

2.1 Spectrum of the Thermally Emitted Light

Due to thermal fluctuations inside the material, each volume element behavesas a time-dependent fluctuating dipole. The field radiated by the dipolesthrough the interface is the thermally emitted field. Note that since we areinterested in the properties of the emitted field, we assume that there are noother sources in the problem (there is no incident field from the vacuum side).In particular, this is a non-equilibrium situation. The current density j(r, t)

Page 3: Coherent Spontaneous Emission of Light Due to Surface Waves

Coherent Spontaneous Emission of Light Due to Surface Waves 165

at a point r = (x, y, z) inside the material is a random variable, which isstationary in time [13]. The emitted field E(r, t) is also a random variable.When we take the ensemble average of the field over many realizations ofthe random currents, we find a zero mean value of the currents and thereforea zero mean value of the emitted field. However, we are interested in themean value of the density of energy and of the radiative flux. Since theseare quadratic quantities, their mean value is not zero. From a more generalpoint of view, the basic quantity to compute is the electric-field cross-spectraldensity tensor Wjk defined by [1]

〈Ej(r1, ω)E∗k(r2, ω

′)〉 =Wjk(r1, r2, ω) δ (ω − ω′) , (1)

where the superscript ∗ denotes the complex conjugate and the brackets de-note a statistical ensemble average. The tensorWjk is a measure of the corre-lation of the field at two different points and at a given frequency. To computethis second-order quantity, we need to know the spatial correlation functionof the currents inside the material. It is given by the fluctuation-dissipationtheorem. The details of the calculation are given in [10]. The electric energydensity I(r, ω) is proportional to the ensemble average of the square modulusof the electric field. It is deduced from the cross-spectral density by

I(r, ω) =ε02

∑k=x,y,z

Wkk(r, r, ω) . (2)

2.2 Examples: Near-Field Thermal Emission of SiC and Glass

Computed thermal emission spectra for silicon carbide (SiC) at differentdistances z from the interface are shown in Fig. 2 in the frequency range0 < ω < 400× 1012 s−1 for T = 300K. We have used the dielectric constantvalues given in [15]. In the far field (z = 100µm), the spectrum is given by thePlanck function, except for the frequency 150×1012 s−1 < ω < 180×1012 s−1,where SiC is highly reflective, so that it does not emit light. When movingtowards subwavelength distances from the source plane (the peak emissionwavelength at T = 300K is about 10µm), the spectrum changes dramati-cally. The emission spectrum at z = 2µm has a peak at ω = 178.7×1012 s−1.At a distance of 100 nm (Fig. 2c), the emission is almost monochromatic, ata frequency not represented in the far-field spectrum. Note that the verticalscale is very different in the three figures. The broad spectrum represented inFig. 2a has not disappeared, it is merely overwhelmed by the enhancementof the density of energy close to the peak. To summarize, we have two majormodifications of the spectral density of energy. First, the spectrum becomesquasi-monochromatic. Second, the electric density of energy is increased byseveral orders of magnitude. How can we explain these surprising results?The presence of a sharp peak in the emission near-field spectrum can be ex-plained in terms of the density of electromagnetic modes. Indeed, it is known

Page 4: Coherent Spontaneous Emission of Light Due to Surface Waves

166 Jean-Jacques Greffet et al.

Fig. 2. Thermal emission spectra ofa semi-infinite medium of SiC at T =300K, for three different height zabove the interface. Note the tran-sition from a broad spectrum toa quasi-monochromatic spectrum aswell as the dramatic increase of theelectric energy density as z decreases.The data are normalized by the peakvalue for z = 100µm

from statistical physics that the density of energy is the product of threeterms: the energy of each photon, the density of states and the mean numberof photons occupying each state. The latter is given by the Bose–Einsteindistribution at equilibrium. Thus, to explain an increase of the density ofenergy close to the interface, we have to assume that the density of stateshas been locally increased. The fact that the density of states increases closeto the interface is due to the existence of additional solutions of Maxwell’sequations: the surface waves. Surface waves are solutions of Maxwell’s equa-tions that propagate along the interface and decay exponentially away fromthe interface. The wave vector k|| along the interface is given by [17]

k2||(ω) =

ω2

c2ε(ω)

ε(ω) + 1, (3)

where k|| = |k|||. These solutions exist only when the real part of the di-electric constant is smaller than −1. Thus they exist only for some ma-terials and within certain parts of the spectrum. They correspond to theexcitation of coupled mechanical and electromagnetic vibrations. For met-als, these are electron density waves and are called surface-plasmon polari-tons. For polar dielectrics, they are due to the ion vibrations and are calledsurface-phonon polaritons. Each surface wave is an additional solution thatcan host photons. This explains why the energy density increases when ap-

Page 5: Coherent Spontaneous Emission of Light Due to Surface Waves

Coherent Spontaneous Emission of Light Due to Surface Waves 167

Fig. 3. Dispersion relation of surface-phonon polaritons at a vacuum–SiC planarinterface. Re(ω) is plotted versus k||

proaching the interface. In order to understand why the spectrum becomesquasi-monochromatic, we have to study in more detail the surface wave. Thedispersion relation ω(k||) of the surface wave is shown in Fig. 3. A remarkablefeature is that there is an asymptote for a particular frequency. Each couple(ω(k||),k||) is a mode of the problem. Thus, close to the frequency of theasymptote, the local number of modes increases dramatically. This explainsthe large enhancement of the energy density for a particular frequency andclose to the interface.

It has been shown in [2] that the electric energy density can be writtenin the form:

I(z, ω) = θ(ω, T )N(z, ω) . (4)

In this expression, 2πh is the Planck constant and N(z, ω) is the local den-sity of electromagnetic modes (propagating and evanescent) that are excitedduring the emission process.

The mean energy of the quantum harmonic oscillator in thermal equi-librium at temperature T is Θ(ω, T ) = hω/[exp(hω/kBT ) − 1], where kBis the Boltzmann constant. It is the product of the energy hω of a particleand the mean number of particles per oscillator given by the Bose–Einsteindistribution. At short distance z (compared to the peak wavelength of thePlanck function), an asymptotic expression of I(z, ω) can be derived [13].Upon identification with (4), one obtains

N(z, ω) =ε′′(ω)

|ε(ω) + 1|21

16π2ωz3. (5)

The 1/z3 contribution is a well-known quasi-static property of thermalfields near the source plane [11,13]. It exists for any material and is notrelated to surface waves. At a given distance z in the near field, the prefactorε′′(ω)/|1 + ε(ω)|2 is responsible for the peak in the spectrum observed inFig. 2 at z = 100 nm. At a frequency ε(ω)max such that ε′(ωmax) ≈ −1, thedensity of modes displays a sharp peak due to surface modes. Indeed, thepeak corresponds clearly to the asymptote of the dispersion relation of the

Page 6: Coherent Spontaneous Emission of Light Due to Surface Waves

168 Jean-Jacques Greffet et al.

surface waves, as can be seen by comparing with (3). It is also interestingto note that this resonant denominator also appears in the expression of thevan der Waals force [14]. Finally, we note that by expanding the real part ofthe dielectric constant ε′ around ωmax it is easy to show that the peak hasa Lorentzian shape with a width given by [ε′′/(∂ε′/∂ω)]1/2.

We show in Fig. 4 a similar behavior for a very common material, namelyamorphous glass. In the far field (Fig. 4a), the spectrum is very close tothat of a blackbody source. When approaching the surface (Fig. 4b,c), thespectrum changes dramatically and a strong peak emerges. Note that at thelocation of this peak, the energy density has increased by almost four ordersof magnitude. This increase in energy density has important consequenceswhen one considers the radiative heat transfer between two bodies at closedistance. This point will be discussed in Sect. 4. The location of this peakalso corresponds to a frequency ωmax such that ε′(ωmax) ≈ −1, and thus tothe thermal excitation of resonant surface waves.

Fig. 4. Thermal emission spectrum of a semi-infinite medium of glass at T = 300K,for three different heights z above the interface. The data are normalized by thepeak value for z = 100µm

Page 7: Coherent Spontaneous Emission of Light Due to Surface Waves

Coherent Spontaneous Emission of Light Due to Surface Waves 169

2.3 Potential Applications

We now turn to the discussion of some implications of the above effects. Weshall first discuss the possibility of performing solid-state spectroscopy basedon a measurement of the near-field thermal emission spectrum.

Let us assume that a near-field optical measurement yields a signal pro-portional to the electric energy density1. According to (4) and (5), the signalis then proportional to ε′′/|1+ε|2 = 0.5 Im[(ε−1)/(ε+1)]. The real and imagi-nary parts of (ε−1)/(ε+1) are linked by the Kramers–Kronig relations (it canbe shown that the quantity (ε−1)/(ε+1) is causal) so that Re[(ε−1)/(ε+1)]is also obtained. Once these two quantities are known, one can calculate thereal and imaginary parts of the complex dielectric constant. In Fig. 5, we haveplotted ε(ω) obtained by reflectivity measurements on glass [15] (referencecurve) and ε(ω) deduced from a near-field theoretical spectrum calculatedat z = 100 nm. There is a good agreement between the result and the ref-erence curve. This procedure suggests a new method to perform solid-statespectroscopy on surfaces at a submicrometer scale [2,16]. Note that because

Fig. 5. Real and imaginary parts of the dielectric constant of glass, versus frequency.εrefl is the reference value, obtained by reflectivity measurements [15], and εspec isthe value numerically reconstructed from the near-field spectrum using Kramers–Kronig relations

1 In order to account for the polarization dependence of the signal, the correctexpression of the signal [18] should be used.

Page 8: Coherent Spontaneous Emission of Light Due to Surface Waves

170 Jean-Jacques Greffet et al.

the near-field spectrum is sharply peaked (Fig. 4c), a narrow-band spectrumaround the peak frequency is sufficient to get enough information, making iteasier to use the Kramers–Kronig relations. A second application could bethe use of such thermal sources as near-field infrared quasi-monochromaticsources. Semiconductors display resonance frequencies in the range 2–100µm.With doped semiconductors, the peak frequency is controlled by the impuritydensity.

A third application is the increase of the efficiency of thermophotovoltaicgenerators. These are heaters that burn gas to produce infrared radiation.This radiation is converted into electricity by the photovoltaic effect usinglow-gap materials. A major drawback of all photovoltaic systems is that thesource has a wide frequency spectrum mainly given by the Planck function.Thus, all the photons with energy lower than the energy gap are lost. Forphotons with energy higher than the gap, the difference hν − Egap is alsolost. It is clear that the mismatch between the absorption spectrum and theincident spectrum is the major cause of the reduction of the efficiency ofphotovoltaic devices. Since the spectrum in the near field becomes almostmonochromatic, one may anticipate a tremendous increase in the efficiencyof these systems. In what follows, we address the question of the enhancementof the heat transfer due to the excitation of surface-phonon polaritons.

3 Radiative Heat Transfer at Nanometric Distances

We shall now discuss the issue of heat transfer by radiation between closely-spaced bodies.

3.1 Introduction

Besides the application to photovoltaics, there are a number of other fieldswhere heat transfer and temperature control at the submicronic scale is a ma-jor issue. For instance, heat transfer in a microprocessor is not very wellunderstood [3]. Thermal probe microscopes with nanometric resolution havebeen designed [5], and understanding the images is a challenging probleminvolving nanometer-scale heat transfer. In this section, we study the ra-diative heat transfer between two semi-infinite bodies separated by a smallvacuum gap (Fig. 6). We show that, under certain conditions, the radiativeheat transfer due to evanescent waves can be dominant and larger than theclassical radiative heat transfer by several orders of magnitude. Moreover, theradiative transfer at short distance may occur at particular frequencies. Thisis a direct consequence of the near-field spectral behavior discussed in theprevious section. Our goal is to investigate how the radiative heat transferbetween two plane semi-infinite media depends on the separation distance d(Fig. 6).

Page 9: Coherent Spontaneous Emission of Light Due to Surface Waves

Coherent Spontaneous Emission of Light Due to Surface Waves 171

Fig. 6. Geometry of the system used to study nanometer-scale radiative transfer

This problem has already been adressed in the past. Cravalho et al. [6]were the first to point out the role of tunneling in radiative heat transfer. Ry-tov [13], Polder and Van Hove [7] and Loomis and Maris [8] did a completecalculation based on fluctuational electrodynamics. The radiative transfer iscalculated as follows. With reference to Fig. 6, the thermal (fluctuating) cur-rents in medium 1 (temperature T1) produce an electromagnetic field in allspace. The energy dissipated by these fields in medium 2 (temperature T2)is the first contribution to the radiation heat exchange. An analogous contri-bution comes from the dissipation in medium 1 of the energy carried by thefields produced by the thermal currents in medium 2. The difference betweenthe two contributions gives the net radiative flux exchanged between bothmedia.

The first contribution to the radiative heat exchange is obtained by cal-culating the flux of the ensemble average of the Poynting vector 〈S〉 =0.5 〈Re(E×H∗)〉 through a plane just above the interface separating medium2 and the gap. In this case, E and H are the electric field and the magneticfield in medium 2 generated by the thermal currents in medium 1.

The main points in this calculation are that (1) it accounts for bothpropagating and evanescent waves (near fields) and (2) the electromagneticproperties of the interfaces are completely taken into account in the electro-magnetic model [10]. In particular, the excitation of resonant surface wavesis fully described by the poles of the reflection Fresnel factor appearing inthe Green’s function. In the following, we will show that the contributionof the evanescent waves may enhance the heat transfer by several orders ofmagnitude when resonant surface waves are involved. We will also see thatunder these conditions, the radiative transfer becomes quasi-monochromatic.

3.2 Contribution of Resonant Surface Waves

We now evaluate numerically the radiative heat transfer between two half-spaces of absorbing material, separated by a gap with nanometric width d.First of all, let us remark that for most materials, when a gas at atmosphericpressure is present in the gap, the radiative heat transfer remains lower thanthe conductive heat transfer due to the ballistic flight of molecules betweenthe two bodies [19]. At ambient temperature and pressure and for smalldistances between the two media, this mechanism yields a conductive heat

Page 10: Coherent Spontaneous Emission of Light Due to Surface Waves

172 Jean-Jacques Greffet et al.

transfer coefficient2 hC, which is about 4× 104Wm−2K−1 [19]. For the sakeof comparison, the value of the heat-transfer coefficient between air and a ver-tical surface due to natural convection is about 5Wm−2K−1!

Following the approach of [7], a radiative heat transfer coefficient hR (inWm−2K−1) can be introduced when T1 ≈ T2. In many cases, hR remainsmuch lower than hC. Nevertheless, we will see that hR may be dramaticallyenhanced by the thermal excitation of resonant surface waves.

We now consider materials constituting the two media that can supportsurface waves at a wavelength close to the maximum of the Planck functionat 300K (around λ = 10µm). Many materials such as glass, SiC, and II–VIand III–V semiconductors belong to this category. In Fig. 7 we display hR

(T = 300K, d) versus the gap width d for SiC and glass. When d is muchlarger than the wavelength of the maximum of the Planck spectrum (10µmhere), hR does not depend on d. This is the result that would be obtained inclassical heat transfer theory, where only propagating waves are taken intoaccount. When d is lower than 1µm, the heat transfer increases as d−2, inagreement with previous results [8]. Note that for both SiC and glass, hR

reaches the typical value of hC at ambient pressure when d ≈ 10 nm. In fact,the radiative heat transfer at small distances is dramatically enhanced bythe excitation of resonant surface waves along the interfaces. For example,we have seen that at a distance of 10 nm, hR is four orders of magnitude largerfor glass than for chromium, a material which does not support surface wavesin the infrared.

The physical origin of the heat-transfer enhancement can be explainedfrom the argument developed in Sect. 2. The electromagnetic energy den-sity above an interface separating a lossy medium from vacuum increases as

Fig. 7. Radiative heat-transfer coefficient hR at T = 300K versus the gap width d,for two semi-infinite media of SiC or glass

2 Using this coefficient, the conductive flux per unit surface is written φ =hC(T1 − T2).

Page 11: Coherent Spontaneous Emission of Light Due to Surface Waves

Coherent Spontaneous Emission of Light Due to Surface Waves 173

1/z3 at nanometric distances, as shown in (4) and (5). Moreover, at a givendistance, the energy density becomes very large at a frequency ωmax suchthat ε′(ωmax) ≈ −1. As shown in Sect. 2, this enhancement is related to theexistence of additional electromagnetic modes due to the thermal excitationof resonant surface waves. When two media are brought together at a closedistance, these evanescent electromagnetic modes contribute to the radiativetransfer, through optical tunneling. Near the peak frequency, ωmax, signifi-cant radiative energy transfer will take place. A microscopic point of viewcan be introduced – the heat transfer is due to the coupling between surfacephotons in the lower and upper interfaces. In other words, when the spacingdecreases, there are phonon–phonon collisions.

Because of the existence of this peak frequency, the radiative heat transferat small distances exhibits peculiar spectral effects. In fact, the frequenciesat which the heat transfer occurs strongly depend on the distance d! Theincrease of the heat transfer coefficients, due to the contribution of evanescentelectromagnetic modes, exists at all frequencies but is much larger near thefrequency ωmax. To analyze this effect, we can introduce a monochromaticradiative heat transfer coefficient hR

ω . It is plotted in Fig. 8 in the case of SiCand glass, for d = 10 nm. We see that the heat transfer coefficient exhibitslarge peaks at particular frequencies. For SiC, the heat transfer is quasi-monochromatic! This is a very unusual situation in heat transfer by radiationbetween two thermal sources. An asymptotic expansion of hR

ω can be donein the same way as the expansion had been done for the density of modes inSect. 2. One obtains

hRω∼= 1π2d2

ε′′1ε′′2

|1 + ε1|2|1 + ε2|2 kB(hω

kBT

)2 exp[hω/(kBT )]{exp[hω/(kBT )]}2

, (6)

Fig. 8. Monochromatic heat-transfer coefficient hRω at T = 300K, versus frequency,

for two semi-infinite media of SiC and glass; d = 10 nm. It is seen that most of thetransfer takes place in a very narrow range of frequencies

Page 12: Coherent Spontaneous Emission of Light Due to Surface Waves

174 Jean-Jacques Greffet et al.

where ε′′1 and ε′′2 are the imaginary parts of the dielectric constants of medium

1 and medium 2, respectively. We see in this equation that the heat transfer isenhanced at the resonant frequencies ωmax of each material. The enhancementis particularly strong when the materials are identical because the resonancesamplify each other. This asymptotic formula also explains why the integratedheat transfer coefficient hR behaves as d−2 at small distances in Fig. 7. Themonochromatic heat-transfer coefficient hR

ω behaves as d−2 in (6). Because itdisplays narrow quasi-monochromatic peaks, its integral over ω also behavesas d−2. Also note that the temperature dependence of hR is given by (6),using ω ≈ ωmax in the last term. A behavior very different from the T 4 lawof blackbody radiation.

4 Spatial Coherence of Thermal Sourcesin the Near Field

The study of emission spectra in the near field has shown that thermal emis-sion can be quasi-monochromatic due to the excitation of resonant surfacewaves. From the point of view of coherence theory, this also demonstrates thatsuch thermal sources exhibit a high degree of temporal coherence. From (5)we were able to derive the spectral width of the resonance. The coherencetime in the very near field is roughly given by its inverse. It thus stronglydepends on the losses at the peak frequency. Again, this large time-coherencein the near field is due to the peak of the local density of states due to thepresence of surface waves.

In this section, we will study the issue of the spatial coherence of thermalsources. A well-established result of coherence theory states that light accrossa planar Lambertian source (assumed to be a good model for a conventionalthermal source), at a given wavelength λ, is spatially correlated over a dis-tance on the order of λ/2 [1]. In deriving this result, the near-field part of theemitted light is disregarded, because it plays no role in the far-field propertiesof emission from planar sources. Nevertheless, we have seen in Sect. 2 thatthe non-propagating (evanescent) fields play a substantial role in the spec-tral properties of thermal sources. We will see that they also dramaticallyinfluence the spatial coherence.

4.1 Exact Calculations of the Spatial Correlation of the Field

When dealing with spatial coherence, we investigate the correlations of thefield at different points and equal time. An alternative point of view, is to fo-cus on a particular frequency of the spectrum. It is essential to introduce thisfrequency analysis because the materials have very different behaviors for dif-ferent frequencies. Here, we are particularly concerned with the possible pres-ence of surface waves. Therefore, one must study a spatial correlation functionof the electric field at a well-defined frequency. In the context of coherence

Page 13: Coherent Spontaneous Emission of Light Due to Surface Waves

Coherent Spontaneous Emission of Light Due to Surface Waves 175

theory, it is called the electric-field cross-spectral density and its definitionis given by (1). This quantity can be computed following [10]. We start byshowing some typical results obtained close to the interface. Let us first com-pare the spatial correlation of the field emitted by lossy glass and tungsten,the latter being a metal which does not support surface waves in the visiblepart of the spectrum. We plot in Fig. 9 the diagonal element Wxx(r1, r2, ω)of the cross-spectral density tensor, at a wavelength λ = 2π/k = 500 nm. Atthis wavelength, the dielectric constant [15] of a lossy glass is ε = 2.25+0.001jand that of tungsten is ε = 4.35 + 18.05j. The calculation is performed ina plane z = z0 above the surface of the emitting material. Both r1 and r2

are along the x-axis, and the result is plotted versus ρ = |r1 − r2|, and nor-malised by its value at ρ = 0. In the very near field (z0 = 0.01λ), the curvecorresponding to glass (solid line) drops to negligible values after ρ = λ/2,showing that the correlation length of the x-component of the field is λ/2. Infact, the solid curve in Fig. 9 strongly resembles the sin(kr)/kr shape of thecross-spectral density in the source plane of a Lambertian source, previouslyobtained in the scalar approximation [1].

In comparison, the case of tungsten (dotted curve) is completely differ-ent. The correlation length is much smaller than λ/2 (i.e. much smaller thanfor the blackbody radiation), on the order of 0.06λ. This subwavelength cor-relation length is a pure near-field effect, due to non-radiative evanescentfields. At a distance z0 = 0.1λ, we see that the correlation length for tung-sten (dashed curve in Fig. 9) is much larger (on the order of 0.4λ) than thatobtained with z0 = 0.01λ (dotted curve).

We now turn to the study of spatial coherence in light emission from ma-terials supporting resonant surface waves, such as surface-plasmon or surface-phonon polaritons [17]. The thermal excitation of a surface polariton inducessome spatial correlation in the field close to the surface, and we may expecta large increase of the correlation length. We illustrate in Fig. 10 the effect ofsurface-plasmon (Fig. 10a) and surface-phonon (Fig. 10b) polaritons on thespatial coherence of the thermal near field. We plot in Fig. 10a the element

Fig. 9. Wxx(r1, r2, ω) in the plane z = z0 versus ρ = |r1 − r2|; r1 and r2 are onthe x-axis; λ = 500 nm. Two materials are considered: lossy glass (z0 = 0.01λ) andtungsten (z0 = 0.01λ and 0.1λ). All curves are normalized by their maximum valueat ρ = 0. Reprinted from [10]

Page 14: Coherent Spontaneous Emission of Light Due to Surface Waves

176 Jean-Jacques Greffet et al.

Fig. 10. Same as Fig. 9, with z0 = 0.05λ. (a) Case of three metals (tungsten, goldand silver), λ = 620 nm. (b) Case of SiC with λ = 620µm and λ = 11.36 µm.At 620 nm, the metal dielectric constants are ε = 4.6 + 20.5j for tungsten, ε =−8.26 + 1.12j for gold and ε = −15.04 + 1.02j for silver. Reproduced from [10]

Wxx of the cross-spectral density tensor at the wavelength λ = 620 nm, andin the plane z0 = 0.05λ, for three different metals.

Both gold and silver exhibit surface-plasmon resonances at this wave-length. We clearly see that whereas the spatial correlation length for tungstenis a fraction of the wavelength (as in Fig. 9), the correlation length for goldand silver is much larger. In fact, although Fig. 10 is limited to ρ < 5λ forthe sake of visibility, the correlation extends over a larger distance given bythe attenuation length of the surface-plasmon polariton. For gold and silver,the attenuation lengths are 16λ and 65λ, respectively. The same effect is seenin Fig. 10b for a SiC crystal, which exhibits a surface-phonon polariton reso-nance at the wavelength λ = 11.36µm (ε = −7.56+ 0.41j) and no resonanceat λ = 9.1µm (ε = 1.80+4.07j). The difference of behavior of this material atthe two different wavelengths is striking in Fig. 10b. The correlation length ismuch higher in the presence of the resonant surface-wave (dashed line) thanin the case where no surface wave is excited (solid line). The propagationdistance of the surface-phonon polariton in this case is 36λ (λ = 11.36µm).

4.2 Qualitative Discussion

An asymptotic evaluation of the cross-spectral density tensor can be per-formed in the near field [11]. This analysis allows us to retrieve the previousresults analytically and yields physical insight into the mechanisms responsi-ble for near-field spatial coherence effects. In particular, several contributionsto the cross-spectral density Wjk can be identified: the quasi-static field (ex-treme near field), surface-phonon or surface-plasmon polaritons, skin-layercurrents and far-field contributions. In fact, the asymptotic analysis showsthatWjk is the sum of several terms, corresponding to each contribution. De-

Page 15: Coherent Spontaneous Emission of Light Due to Surface Waves

Coherent Spontaneous Emission of Light Due to Surface Waves 177

pending on the distance z to the interface, one of them may dominate. Con-cerning spatial coherence, surface waves such as surface-plasmon polaritonsor surface-phonon polaritons yield long-range spatial coherence on a scaleof the surface wave propagation length which may be much larger than thewavelength when absorption is small. These are the dominant contributionsat a distance from the surface of the order of a wavelength in vacuum. Onthe contrary, skin-layer currents and quasi-static fields dominate at a distancemuch shorter than the wavelength. They lead to a much shorter spatial co-herence length that only depends on the distance to the interface [11]. Note,however, that this conclusion is based on the assumption of a local medium.The ultimately limiting scale is thus given by the electron screening lengthor the electron Fermi wavelength, whatever is larger.

The interested reader is refered to [11] for further details. In what follows,we will give a heuristic description of the mechanism that transforms a ran-dom source with uncorrelated currents into a spatially coherent source. Letus consider the contribution of free electrons in a metal for the sake of clarity.The currents produced by the electrons have two components: a random com-ponent due to the random thermal motion and an induced component due tothe fields in the medium. The fluctuation-dissipation theorem together withthe assumption of a local medium implies that the fluctuating currents aredelta correlated. The question that arises is how can we obtain fields whichare correlated over tens of wavelengths from such incoherent currents?

It is known that two slits will produce interference in transmission whenilluminated by the light coming from the sun. The reason is that the coher-ence of the field increases upon propagation. This is known as the Zernike–van Cittert theorem. It can be shown that the cohence length in the planeperpendicular to the beam is given by λ/θ, where θ is the angle subtendedby the source. The reason is that a given point in the source illuminates bothslits when they are far apart from the source plane. This creates a correlationfor the fields at both slits. On the contrary, if the slit is close to the source,it is mainly illuminated by points of the sources lying just beneath it. There-fore, the fields at both slits are uncorrelated. This is exactly what happenswhen we consider the electric field very close to the interface. Then, the fieldis mainly due to the source elements which lie just beneath the observationpoint, because the electrostatic components varying like 1/r3 dominate. Thisis why we observe a very short coherence length in Fig. 9 for tungsten. Thetypical coherence length depends only on the distance between the observa-tion point and the source plane.

What happens now if there is a surface wave? Any point source can excitea surface wave. Since a surface wave is a delocalized mode of the system, itbuilds up over the interface with a spatial extension given by its decay length.Therefore, each volume element radiating light can excite a surface wave thatspreads over tens of wavelengths along the interface. There is an analogywith a piano chord. The source is a hammer that strikes the chord at a single

Page 16: Coherent Spontaneous Emission of Light Due to Surface Waves

178 Jean-Jacques Greffet et al.

point. Yet, the mode of the chord is excited and vibration takes place alongthe full length of the chord. This is an example of a delocalized mode. Fora thermal light source, the source current due to the random thermal motionof the charges is random and uncorrelated. However, the induced currentsdue to the excitation of the surface wave are delocalized and produce a largecoherence length.

5 A Spatially Coherent Thermal Source

We have just discussed the spatial coherence of a thermal planar source. Inthis section we describe an experimental result [12] that proves that thermalsources can be partially spatially coherent. The basic consequence of thetransverse spatial coherence for a laser is that it propagates with a narrowangular divergence approximately given by the ratio λ/w, where w is thebeam waist. By contrast, as mentionned in the introduction, a thermal sourcesuch as a light bulb is a quasi-lambertian source. This is related to the lackof coherence of the source. Consider two different points of the source; if theyare uncorrelated, the fields that they emit cannot interfere. Since each pointhas a quasi-lambertian angular emission pattern, the overall emission angularpattern is also quasi-lambertian.

We have seen that the field may be coherent along the interface. However,since this coherence is only due to surface waves, it cannot be detected in thefar field. To transfer the near-field coherence into the far field, we can rulea grating on the interface. The surface waves are then coupled with propagat-ing waves. Since they propagate along distances of the order of l ∼ 10–20λ,the field scattered by each groove of the grating can interfere, producingmaxima of emission in well-defined directions with an angular width approx-imately given by λ/l. The particular properties of such sources were firstreported by Hesketh et al. [20] and later by Kreiter et al. [21], although therole of surface waves and coherence was not fully understood at that time.Emission by surface waves using coupling by a prism has also been widelystudied by Zhizhin et al. [22].

In order to produce such a source, it is essential to properly design thegrating so that the coupling efficiency of the grating is as high as possible. Tothis end, we have used numerical simulations as described in [23,24] in orderto find the optimum parameters. The material chosen was SiC, because thismaterial may support a surface wave in the spectral range close to 10µm,which is roughly the peak wavelength of the Planck spectrum at ambienttemperature. The grating was then fabricated using standard techniques. Animage taken with an atomic force microscope is seen in Fig. 11. The angularemission measurements at 773K are shown in Fig. 12. The most strikingfeature is the fact that the emission pattern looks like an antenna emissionpattern. This is a signature of the partial spatial coherence of the source, asdiscussed in [12]. An indirect measurement of the emissivity ε can be done

Page 17: Coherent Spontaneous Emission of Light Due to Surface Waves

Coherent Spontaneous Emission of Light Due to Surface Waves 179

Fig. 11. Atomic force microscope image of the grating. The period is 6.25µm andthe height is 0.284µm. The parameters were optimized so that the emissivity is lfor a wavelength 11.36 µm. Reproduced from [12]

Fig. 12. Polar plot of the emissivity of the grating at a wavelength 11.36 µm andtemperature 773K. Red : experimental data; green: theoretical calculation. The the-oretical result has been convoluted with an angular window to account for theexperimental angular resolution. The disagreement is due to the fact that the cal-culations use the values of the optical constants at ambient temperature. It is seenthat most of the light is emitted in a narrow angular cone. Reproduced from [12]

by measuring the reflectivity R of the sample. Indeed, Kirchhoff’s [25] lawstates that R = 1− ε.

We were able to measure the dispersion relation of the surface-phononpolariton by studying the spectral reflectivity of the grating. The result isshown in Fig. 13. By looking at the lower branch of the dispersion relation itis seen that a particular frequency is associated with a particular componentof the wave vector parallel to the interface. The wave vector is related to theemission direction in the far field by the simple relation k|| = (ω/c) sin θ. Itis thus seen that the spectrum depends on the angle of observation. This is

Page 18: Coherent Spontaneous Emission of Light Due to Surface Waves

180 Jean-Jacques Greffet et al.

Fig. 13. Reflectivity of p-polarized light by the grating as a function of the wave-length. The dip observed at λ = 11.36µm coincides with the emission peak observedin Fig. 12. The figure shows clearly a quasi-total absorption of the incident lightdue to the resonant excitation of a surface-phonon polariton. A flat surface wouldhave a reflectivity on the order of 0.94. Note the very good agreement between ex-periment and theory. This agreement shows that the disagreement in the emissionmeasurements is due to the change of the index with temperature

0 100 200 300 400 500 600 700 800700

750

800

850

900

950

1000

k (cm−1)

ω (

cm−

1 )

Fig. 14. Dispersion relation of the surface-phonon polariton on the grating. Greencurve: theoretical dispersion relation. Red curve: theoretical dispersion relation forthe flat surface. Data points: experimental measurements. The dispersion relationis constructed from reflectivity measurements. A spectrum is taken of the specularreflectivity of the grating at a fixed angle θ using a Fourier Transformation Infra-Red spectrometer. The frequency points of each minimum of the reflectivity andthe value (2π/λ) sin θ yield the coordinates of a point of the dispersion relation. Itis seen that there is excellent agreement between the experiment and the theory

Page 19: Coherent Spontaneous Emission of Light Due to Surface Waves

Coherent Spontaneous Emission of Light Due to Surface Waves 181

standard behavior for any coherent source such as an antenna. For randomsources, it is far from trivial. It is a manifestation of the interplay betweenthe propagation and correlation of the field known as the Wolf effect [26]. Ithas been observed for secondary partially coherent sources so far. The gratinghas this property. Indeed, by making a measurement of emission in the farfield, one selects a particular angle and therefore a particular wave vector.The dispersion relation of the surface waves that produce the coherence showsthat for a given wave vector, there will be peak emissions at some frequencies.If we now consider higher frequencies, it is seen that the dispersion relationdoes almost not depend on the wave vector. This means that the emission isalmost isotropic, with a large peak at some particular frequencies where thedispersion relation has a flat asymptote. This can be used to design infraredlight emitters.

6 Conclusions

In this paper, we have reviewed recent results concerning the near-field prop-erties of thermal sources of light. The existence of surface waves producesa very intense peak of the local density of states close to the interface forsome particular frequencies. We have highlighted several new properties ofthermal sources in the near field by exploring the implications of this largedensity of states.

We have shown that thermal sources may produce quasi-monochromaticnear fields. In light of this result, the possibilities of performing near-fieldsolid-state spectroscopy and of designing near-field infrared sources have beendiscussed. The implication for photovoltaic applications has also been dis-cussed.

The problem of radiative transfer between two thermal sources held atsubwavelength distance has been studied. We have shown that the radiativeflux may be enhanced by several orders of magnitude due to the excitationof resonant surface waves, and that it may occur at particular frequencies.

Finally, we have studied the spatial coherence of thermal sources and thesubstantial influence of the near field. We have shown that surface wavesmay induce long-range spatial correlations on a scale much larger than thewavelength. Sources that emit light within narrow angular lobes in the farfield have been demonstrated. Conversely, quasi-static contributions, as wellas skin-layer currents, induce correlations on small length scales (as far asa macroscopic and local description of the materials is correct).

With the recent development of local (optical and thermal) probe mi-croscopy and the advent of nanotechnology, it is necessary to revisit theemission of light by plane surfaces. The results presented in this paper showthe crucial role of surface waves in this respect.

Page 20: Coherent Spontaneous Emission of Light Due to Surface Waves

182 Jean-Jacques Greffet et al.

References

1. L. Mandel, E. Wolf: Optical Coherence and Quantum Optics (Cambridge Uni-versity Press, Cambridge 1995) 163, 165, 174, 175

2. A.V. Shchegrov, K. Joulain, R. Carminati, J.-J. Greffet: Phys. Rev. Lett. 85,1548 (2000) 163, 167, 169

3. C. L. Tien, G. Chen: J. Heat Transfer 114, 799 (1994) 164, 1704. A.R. Abramson, C. L. Tien: Microscale Thermophys. Engin. 3, 229 (1999) 1645. C.C. Williams, H.K. Wickramasinghe: Appl. Phys. Lett. 49, 1587 (1986) 164,

1706. E.G. Cravalho, C. L. Tien, R. P. Caren: J. Heat Transfer 89, 351 (1967) 164,

1717. D. Polder, M. Van Hove: Phys. Rev. B 4, 3303 (1971) 164, 171, 1728. J. J. Loomis, H. J. Maris: Phys. Rev. B 50, 18517 (1994) 164, 171, 1729. J.-P. Mulet, K. Joulain, R. Carminati, J.-J. Greffet: Appl. Phys. Lett. 78, 2931

(2001) 16410. R. Carminati, J.-J. Greffet: Phys. Rev. Lett. 82, 1660 (1999) 164, 165, 171,

175, 17611. C. Henkel, K. Joulain, R. Carminati, J.-J. Greffet: Opt. Commun. 186, 57

(2000) 164, 167, 176, 17712. J. J. Greffet, R. Carminati, K. Joulain, J. P. Mulet, S. Mainguy, Y. Chen: Co-

herent emission of light by thermal sources, Nature 416, 61 (2002), 164, 178,179

13. S.M. Rytov, Yu. A. Kravtsov, V. I. Tatarskii: Principles of Statistical Radio-physics 3 (Springer Verlag, Berlin, Heidelberg 1989) Chap. 3 165, 167, 171

14. L.D. Landau, E.M. Lifshitz: Electrodynamics of Continuous Media (Pergamon,Oxford 1960) 168

15. E.W. Palik: Handbook of Optical Constants of Solids (Academic Press, SanDiego 1985) 165, 169, 175

16. K. Joulain, J.-P. Mulet, R. Carminati, J.-J. Greffet, A.V. Shchegrov: Proc. HeatTransfer and Transport Phenomena in Microsystems Conf., Banff, Canada,2000 169

17. V.M. Agranovich, D.L. Mills (Eds.): Surface Polaritons (North-Holland, Am-sterdam 1982) 166, 175

18. J.A. Porto, R. Carminati, J.-J. Greffet: J. Appl. Phys. 88, 4845 (2000) 16919. J. Xu: Heat Transfer Between Two Metallic Surfaces at Small Distance, Dis-

sertation, University of Konstanz, Germany (1993) 171, 17220. P. J. Hesketh, J. N. Zemel, B. Gebhart: Nature (London) 324, 549 (1986) 17821. M. Kreiter, J. Oster, R. Sambles, S. Herminghaus, S. Mittler-Neher, W. Knoll:

Thermally induced emission of light from a metallic diffraction grating, medi-ated by surface plasmons, Opt. Commun. 168, 117–122 (1999) 178

22. G.N. Zhizhin, E. A. Vinogradov, M.A. Moskalova, V.A. Yakovlev: Appl. Spec-trosc. Rev. 18, 171 (1982) 178

23. A. Sentenac, J. J. Greffet: J. Opt. Soc. Am. A 9, 996 (1992) 17824. J. Le Gall, M. Olivier, J.-J. Greffet: Phys. Rev. B 55, 10105 (1997) 17825. J. J. Greffet, M. Nieto-Vesperinas: J. Opt. Soc. Am. A 10, 2735 (1998) 17926. E. Wolf: Phys. Rev. Lett. 56, 1370 (1986) 181