Structural Slow Waves: Parallels between Photonic Crystals and

Nov 26, 2018 - Université de Bordeaux-CNRS-CEA , Centre Lasers Intenses et Applications, UMR 5107, 351 Cours de la Libération, 33405 Talence , Franc...
0 downloads 0 Views 2MB Size
Perspective pubs.acs.org/journal/apchd5

Cite This: ACS Photonics 2019, 6, 4−17

Structural Slow Waves: Parallels between Photonic Crystals and Plasmonic Waveguides Philippe Lalanne,*,† Steṕ hane Coudert,‡ Guillaume Duchateau,‡ Stefan Dilhaire,§ and Kevin Vynck† †

LP2N, Institut d’Optique Graduate School, CNRS, Université Bordeaux, 33400 Talence, France Université de Bordeaux-CNRS-CEA, Centre Lasers Intenses et Applications, UMR 5107, 351 Cours de la Libération, 33405 Talence, France § Laboratoire Onde et Matière d’Aquitaine (LOMA), UMR 5798, CNRS-Université de Bordeaux, 33400 Talence, France

ACS Photonics 2019.6:4-17. Downloaded from pubs.acs.org by IOWA STATE UNIV on 01/18/19. For personal use only.



ABSTRACT: The slowdown of propagating electromagnetic waves by engineering waveguide structures is receiving considerable attention in nanophotonics. Two types of structures are independently investigated, namely, diffractionlimited photonic-crystal waveguides and highly confined plasmonic waveguides. Here we propose a unified viewpoint on slow waveguide structures by analyzing both approaches within a single theoretical framework. This allows us to clarify the physical mechanisms underlying slownessthereby highlighting their inherent differencesand to remove some inaccurate ideas on the topic. We further discuss the strengths and weaknesses of plasmonic and photonic slow waveguides, assess the technical and scientific challenges recently solved, and critically identify those that still stand in the way. KEYWORDS: slow waveguide, slow surface plasmon, slow photonic-crystal waveguide, LDOS enhancement, roughness-induced backscattering



INTRODUCTION In general, the velocity at which a pulse of light propagates through a medium (group velocity) is given by vg = cng−1 = c(n + ω dn/dω)−1, where c is the velocity of light in a vacuum, ω is the light angular frequency, n is the refractive index of the medium, and ng denotes the group index. Theoretically, there is no fundamental limit on the value to which light may be slowed down. Very low propagation velocities (17 m/s)1 and even stopped light2 have been reported in ultracold atomic gases. These ultraslow regimes are reached when the velocity of light pulses can be described fully in terms of an extremely frequencydependent refractive index of the material, offering a sharp dip in the absorption or gain spectrum in a narrow spectral region. Simple saturation effects can lead to such behavior, as well as more advanced effects such as electromagnetically induced transparency.3,4 Slowness induced by material dispersion (dn/ dω) will not be discussed further in this Perspective article. It is believed that if slow light is delivered at a micro- or nanoscale on-chip, for instance in a fully integrated configuration compatible with standard technological platforms, science and technology will be strongly impacted. This route is mainly followed by relying on structural dispersion. It refers to situations in which the velocity of light pulses is modified by structuring and/or assembling matter at the micro- or nanoscale. The group velocity is no longer related to the refractive index n of a medium, but to an effective index neff that depends on ω. In marked contrast with material slowness, structural slowness is accompanied by a strong electromagnetic field enhancement. There are essentially two schemes for structural slowness: (i) photonic-crystal waveguides, e.g., single-row-missing photoniccrystal waveguides or coupled optical resonators,5 and (ii) © 2018 American Chemical Society

plasmonic waveguides, e.g., metal/insulator/metal (MIM) or insulator/metal/insulator (IMI) heterostructures, metal nanowires, or hyperbolic materials.6 We will indifferently use the terms “slow photon” or “slow photonic waveguide” for dielectric structures and consistently “slow plasmon” or “slow plasmonic waveguide” for metallic ones and “slow wave” or “slow waveguide” in general terms. Photonic-crystal and plasmonic slow waves share many common features, starting from the fieldenhancement property, and are studied, quite naturally, for the same applications in science and technology, e.g., light harvesting,7 biosensing,8 nonclassical light sources,9,10 lasing,5,11−14 quantum information,15,16 nonlinearities and switching,17−20 and integrated photonic circuits.21,22 Photonic crystals are definitely more advantageous for applications such as delay lines, buffering, or bit memory,23 whereas plasmonic structures are considered to be more promising for nanoscale applications.10,24−29 In photonics, the energy is purely electromagnetic. Slowness originates from waves that bounce back and forth as they propagate through a periodic structure (see Figure 1a for an illustration). It immediately appears that a phase-matching condition should be satisfied and that the characteristic dimension of the structure, i.e., the period a, should be comparable to the wavelength λ, i.e., a ≈ λ/(2n). Quite differently, slowness in plasmonics does not necessarily require periodic back-reflections and can be implemented at deepReceived: Revised: Accepted: Published: 4

September 25, 2018 November 22, 2018 November 25, 2018 November 26, 2018 DOI: 10.1021/acsphotonics.8b01337 ACS Photonics 2019, 6, 4−17

ACS Photonics

Perspective

Figure 1. Structural slow electromagnetic waves. (a) Slowness at the microscale with photonic periodic waveguides. It is achieved for lattice periods that are comparable to the wavelength, and the transversal confinement is limited by the diffraction. The slow-mode intensity pattern is periodic along the direction of propagation. (b) Slowness at the nanoscale with plasmonic translation-invariant waveguides. Slow plasmons are achieved with heterostructures having at least one transverse dimension much smaller than the wavelength. The slow-mode intensity pattern is uniform along the direction of propagation. (Courtesy of S. I. Bozhevolnyi.)

knowledge is very important since many physical effects in the linear and nonlinear regimes depend on how the electric field increases with ng. Section 4 highlights the advantages and drawbacks of plasmonic and photonic slowness. It provides a comprehensive picture of their inherent forces and limitations, including the bandwidth, the attenuation due to field enhancement, and fabrication imperfections (roughness, absorption, etc.), which are known to considerably reduce the propagation distance of slow modes in realistic structures.31,37 We also try to understand why slow plasmons seem to promise much for reaching ultraslow regimes and even possibly for stopping light,6 whereas imperfections are known to ruin any hope of observing ultraslow transport with photonic waveguides.37,38 Section 5 addresses the issue of engineering the local density of states (LDOS) with slow waveguides on the illustrative and important example of light emission. Slowness is indeed attached to field enhancement and may thus be used to boost light emission. After a rapid analysis of light emission in slow photonic waveguides, which is nowadays well understood, we focus on emission in slow plasmonic waveguides. The analysis starts with the textbook case of a single metal/insulator (MI) interface, for which we highlight the role of slow plasmons for quenching the emission. The case of MIMs is considered afterward. We evidence that the large LDOS of the slow gap plasmon mode for tiny insulator gaps counterbalances quenching and further discuss how this effect may be exploited for optimizing the efficiency of photon/plasmon sources. Section 6 finally points out some perspectives for the use of slow waveguides in devices and summarizes the work. In the following, it might appear that the ohmic loss is somewhat neglected. For instance, absorption is disregarded in Sections 2 and 3 and is treated as a perturbation in Section 4. There is a good reason for not rooting complex permittivities directly from the beginning into Maxwell’s equations. This simplifies the mathematical treatment of the next sections and allows us to use exactly the same Hermitian theoretical formalism to analyze photons and plasmons, making the intercomparison more quantitative. We would like to emphasize that (i) the role of ohmic losses will be discussed whenever it impacts the reasoning, (ii) the perturbation treatment of Section 4 is exact for translation-invariant waveguides, and exact numerical values will be provided for the main formula. The importance of ohmic losses, which are critical and sometimes underestimated in the literature, is therefore not overlooked hereafter.

subwavelength scales with translational-invariant waveguides.30−34 Slow surface plasmons result from a coupling of electromagnetic fields with the free electrons of metals. As a result, slow plasmons appear when the characteristic dimensions are much smaller than the wavelength, for instance with a MIM dielectric gap (see Figure 1b for an illustration). Consequently, contrary to photonic slow light that occurs in a quite narrow frequency range with effective mode areas on the order of λ2, slow plasmons are inherently broadband with deep-subwavelength mode areas. The literature on structural slow waves appears to be disjointed between slow plasmons and slow photons. Review articles, wherein applications of slow waves are analyzed and discussed,5,6,15,16,18,19,22,35,36 are available for each individual scheme, but to our knowledge, there is no comprehensive overview of structural slow waves in general. Our primary motivation is to address this shortcoming. By treating absorption loss for plasmons as a perturbation, the same theoretical approach, notably the Lorentz reciprocity theorem in its conjugate form, applies indifferently to each scheme. The emergence of a “unified” vision of both schemes is additionally favored by the systematic reference to a single and common physical parameter to quantify the slowness, the group index ng. Section 2 is concerned with the study of the emblematic example of gap modes of MIM waveguides with classical electrodynamic models. The derivation is found in many textbooks on electrodynamics. Our prime interest is to look more specifically to what is happening for small gaps when the group velocity vanishes and to introduce, with a concrete example, many ingredients that will be discussed afterward more abstractly. Although the results are well known, it is recommended that potential readers do not jump over this section. Section 3 is devoted to the basics of the electrodynamics of slow modes. We review the two physical mechanisms underlying slowness and analyze their inherent differences. First, we clarify how ultraslow transport is achieved by associating two counterpropagative power flows. It is generally admitted that the two power flows exactly balance at vanishing speeds. For instance, for slow plasmons, owing to the negative contribution, the power flow in the metal is expected to balance that in the positive-permittivity dielectric cladding layers.6 We revisit this widely admitted picture for photons and plasmons, showing that the power flows are not perfectly balanced and that the slight difference is important, as for instance it allows perfect adiabatic tapering. Then, we derive the scaling laws with ng of the electric and magnetic fields of slow modes, showing that they are considerably different for plasmons and photons. Their 5

DOI: 10.1021/acsphotonics.8b01337 ACS Photonics 2019, 6, 4−17

ACS Photonics

Perspective

Figure 2. Symmetric gap plasmon mode of MIM waveguides translationally invariant along the y axis. (a) z is the direction of propagation, and x denotes the transverse direction. εm < 0 and εd > 0 are the dielectric permittivities of the metal and insulator. As the gap width g becomes much smaller than the wavelength, the mode has no cutoff and the propagation constant β diverges as g−1. (b) Group index for εd = 2.25, εm = −28.0 + 1.5i, and ∂εm = − 84.23 + 2.63i (silver at λ = 800 nm39) in the limit of small gaps (i.e., large β’s). Black circles are numerical data obtained by computing the ∂λ

(

complex-valued effective index β/c with the aperiodic rigorous coupled-wave analysis,40 and then the group index ng = β /c +



SLOW GAP PLASMONS IN METAL/INSULATOR/METAL WAVEGUIDES In this section, we study the properties of slow gap modes guided in MIM heterostructures with tiny gaps employing a classical electrodynamic model. Our prime intention is to introduce, on a concrete example, the concepts and electromagnetic formalisms, e.g., complex group indices, antiparallel power flows, and the Lorentz reciprocity theorem, which will be used further at a more abstract level in the following sections. A second motivation is to have a complete analysis of an important element that plays a key role in nanophotonics, e.g., for energy transport though metal slits or grooves in metallic thin films, artificial magnetism at optical frequency, and anomalous dispersion in hyperbolic media. MIM heterostructures support both transverse-electric (TE) and transverse-magnetic (TM) gap modes.30,32 TE modes may propagate over distances that may even exceed decay lengths observed for TM modes. Here, we are concerned by the fundamental gap-plasmon TM mode (see Figure 2), which does not exhibit cutoff and offers slow propagation as the gap width becomes much smaller than the wavelength. For ultrasmall gap widths, gap plasmons are sensitive to the electron density profile at the metal surfaces and classical

(

ω2

1/2

)

).

electrodynamic models cease to be valid. With a fully quantum mechanical approach,41 the effects of electron spill-out on gap plasmons have been analyzed, and it has been shown that the mode index of ultranarrow gap plasmons converges to the bulk refractive index in the limit of vanishing gap, thereby rectifying the unphysical divergence found with classical models. These effects that take place at the MI interfaces, 1−2 Å from the walls, and introduce additional ohmic losses, are however negligible for the gap widths (g > 1 nm) considered in the following, as we checked with simplified nonlocal models.42 Solving the time-harmonic source-free Maxwell’s equations for the MIM geometry is classical, as it looks very similar to the textbook case of dielectric slab waveguides for TM polarization. A small additional complexity, related to the definition of the square root function in the complex plane, should be addressed if the metal loss is not neglected, Im εm ≠ 0. We will simply consider the case of εm negative and real and introduce absorption with perturbation theory, thereby making all the following quantities, e.g., the propagation constant β and the group index ng, also real. The magnetic field of the symmetric gap mode for TM polarization, propagating along the positive z direction in a MIM waveguide with gap width g, can be written as

l o i i i g yy g yyy g i i o o Hy(x) = h0jjjjexpjjjγdjjjx − zzzzzz + expjjj−γdjjjx + zzzzzzzzzzexp(iβz), |x| ≤ o o o 2 2 2 o {{ {{{ k k k k k o m o o o g yy g i i o o Hy(x) = h0(1 + exp(−γdg ))expjjj−γmjjj|x|− zzzzzzexp(iβz), |x| > o o o 2 {{ 2 k k n

with γα = β 2 − εα c 2

ω dβ c dω

(1)

ω

the limit of small gaps, c g ≪ 1. By assuming that γdg ≪ 1 and γd ≈ γm ≈ β, we get 2ε βg = − d εm (3)

, with α ≡ m, d. Note that γm and γd

are both positive. The magnetic field amplitude h0 is a constant that will be normalized afterward. By using the tangential-field g continuity equation for Ez at x = ± 2 , we obtain the transcendental dispersion equation γ γ − m (1 + exp(−γdg )) = d (1 − exp(−γdg )) εm εd (2)

Equation 3 is important, as it directly shows that, as the gapplasmon spatial extension is shrunk by the gap size, the gapplasmon wavelength 2π/β can be as small as a few nanometers, which is much smaller than that in any known dielectric medium. This is the physical origin of the subwavelength localization beyond the diffraction limit (that is, the diffraction

Effective Index. Equation 2 can be solved accurately by numerically computing β(ω). It is easily solved analytically in 6

DOI: 10.1021/acsphotonics.8b01337 ACS Photonics 2019, 6, 4−17

ACS Photonics

Perspective

important role and can be disregarded. The net benefits are that the group velocity coincides with the energy velocity, providing a simpler interpretation, and that the same intuitive theoretical formalism can be used for both plasmonic and photonic slow waveguides, enabling a direct comparison of each scheme. General Electromagnetic Theorem for Slow Modes. We start by deriving a relation between the power flow and the group velocity of slow guided modes. Throughout the Perspective, we place no restriction on the geometry, and we denote by ε0ε(x, y) the arbitrary cross-section permittivity distribution of the plasmonic waveguide. The waveguide permeability is equal to the vacuum permeability μ0 everywhere for the sake of simplicity. Our analysis relies on basic electromagnetic laws of z-invariant waveguides.43 We consider a guided mode at frequency ω,

limit in a dielectric medium). Equation 3 also implies that both transverse wave vectors, γd and γm, scale as g−1. Field Profile and Mode Normalization. Always in the limit of small gaps, the gap-plasmon field distribution takes the following asymptotic form: Hy(x) = 2h0 exp(iβz) Ex (x) = 2βh0 /(ωε0εd) exp(iβz) Ez(x)=iβh0 /(ωε0εd)[exp(γd(x − g /2)) − exp(− γd(x + g /2))] × exp(iβz) (4a) g

for |x| ≤ 2 , and Hy(x) = 2h0 exp(−γm(|x| − g /2)) exp(iβz)

[En(r), H n(r)] = [e(x , y), h(x , y)]exp i(βz − ωt )

Ex(x) = 2βh0 /(ωε0εm) exp(−γm(|x| − g /2)) exp(iβz) Ez(x) =

with a field distribution that is exponentially decaying away in the claddings and a propagation constant denoted by β. Note that we drop the index “n” in e(x, y) and h(x, y) for simplicity. For any infinitely large transverse cross-section Σ of the waveguide and for Im(ε) = 0, the Lorentz reciprocity theorem applied to waveguides43 leads to the simple relationship

x 2iγ h0 /(ωε0εm) exp( −γm(|x| − g /2)) exp(iβz) |x | m (4b)

g . 2

for |x| > We straightforwardly infer that the Poynting-vector z-components in the gap and in the metals have opposite directions and that the associated power flows are −4h02/ (ωε0εm) and 2h02/(ωε0εm), respectively, so that the net power flow is −2h02/(ωε0εm). Normalizing this net positive flow to 1, we find that h0 is independent of g and must be chosen so that h0 = ( −ωε0εm /2)1/2

1 2

εd Im(εm) 2

|εm|

(5)

1 2



∬Σ Re(e × h*)·ẑ dx dy =

g −1

λ 2 εd ∂εm g πεm 2 ∂λ

dω dβ

(9)

For photonic waveguides, the modes are Bloch modes. The latter are pseudoperiodic functions of z, with a field distribution still given by eq 8, except that e(x, y) and h(x, y) are now periodic functions of z, e.g., e(x, y, z + a) = e(x, y, z). Accounting for this extraperiodicity, eq 9 becomes44

(6)

showing that the imaginary part of β also scales as g−1. It is interesting to note that this result can be obtained directly from eq 3 by taking a complex-valued εm. In fact, one can show that eq 3 rigorously holds in the presence of absorption.30 Group Index. Using the expression of the group velocity ng = c∂β/∂ω, eq 3 eventually leads to the following formula for the group index: ng = −

ÄÅ ÉÑ ÅÅ ∂(ωε) 2 Ñ 2Ñ ÅÅε ÑÑdx dy ·| | + | | μ e h 0 Å ÑÑ 0 ΣÅ ∂ω ÅÇ ÑÖ

∬Σ Re(e × h*)·ẑ dx dy =

Absorption. Because we neglect absorption, Im εm = 0 and eq 3 predicts a real propagation constant. The effect of absorption can be taken into account by applying the Poynting theorem and perturbation theory to a lossy waveguide (Δε = Im(εm) ≠ 0) with the mode profile of eq 4b. The integration of the exponentially decaying field in the metal leads to a complex propagation constant, whose imaginary part is given by Im β = 2

(8)

1 dω 4a dβ

i

y

) 2 |e| + μ0 |h|2 zzzzdx dy dz ∭ jjjjε0 ∂(∂ωε ω k

{

(10)

where the triple integral runs over any unit cell of the periodic waveguide delimited by two infinitely large cross-sections separated by the period a. The energetic interpretation of the classical eqs 9 and 10 evidences that the group velocity dω/dβ of the wave packet is equal to the velocity at which energy is conveyed by the mode along the waveguide. Physics of Slow Waves. There are different ways to understand the wave deceleration in slow waveguides. For photonic waveguides, deceleration is due to a distributed backreflection, and for plasmonic waveguides an equivalent could be a negative Goos−Hänchen phase shift at the metal/insulator interfaces. Perhaps, a most unified way, that we will use hereafter, consists in considering that deceleration originates from antiparallel power flows. As beautifully evidenced by coupled-wave theory,45 slow light in photonic waveguides results from two counter-propagating waves, each carrying some energy, which together constitute a quasi-stationary pattern.45,46 Similarly, for plasmonic waveguides, because the permittivities of metallic and dielectric layers are opposite, the z-component of the Poynting vector is positive in dielectric layers, whereas it is negative in the metallic layers. In this picture, ultimate slowdowns as the group (or the energy) velocity approaches zero are often seen as resulting from

(7)

Like for β, eq 7 is valid in the presence of absorption. We just need to plug in the complex-valued permittivity of the metal to obtain a complex group velocity c/ng, whose real and imaginary parts represent the velocity and the damping of the light pulse, at least when the pulse attenuation is weak. The predictive force of eq 7, in its complex-valued version, is tested in Figure 2b for a silver MIM waveguide at a wavelength of 800 nm.



ELECTROMAGNETISM OF SLOW GUIDED MODES This section is devoted to the derivation of some fundamentals of the electrodynamics of slow modes. For this, absorption loss due to metal in plasmonic waveguides does not play an 7

DOI: 10.1021/acsphotonics.8b01337 ACS Photonics 2019, 6, 4−17

ACS Photonics

Perspective

flows of the slow mode on the right side of the taper tend to exactly balance as vg → 0 and Φs → 0. Applying now the Poynting theorem to the blue boxes, Φ1 + Φf + Φs = 0, one gets Φ1 = −Φf since Φs → 0: we arrive at the contradictory conclusion that any slowdown is automatically accompanied by radiation on the lateral sides of the taper. Perfectly (or adiabatically) stopping light would then be impossible. This is indeed wrong, as we know from slowdown practice for lossless 1D systems, for which there is no loss by absorption nor by scattering in the clads, by construction. It is important to find out where we get the misconception of the exactly balanced flow and to understand in detail the principle of adiabatic tapering. For that purpose, refer to eq 9 or 10 and consider that the power f low (left-hand-side term) is constant (Φs = −Φf) for any cross-section of the taper, even for instance for cross-section planes close to the tip apex where the group velocity vanishes. The right-hand-side term of the equation should then remain constant as vg → 0, implying that the surface integral of the electromagnetic energy density ÅÄ ÑÉ ) 2 |e| + μ0 |h|2 ÑÑÑÑdx dy should diverge with ng, both for ∬Σ ÅÅÅÅÅε0 ∂(∂ωε ω ÑÖ Ç photonic and plasmonic waveguides. There is no problem with energy conservation, as the product of the vanishing velocity and the diverging electromagnetic energy density remains constant. The fact that the divergence of the electromagnetic density is not observed in practice, because of absorption loss or imperfections, will be discussed in Section 4. Scaling Laws at Small Group Velocities. Now that the physics of slow waves is clarified, let us see how the electromagnetic fields scale as vg vanishes. With the power flow being nonzero even for vanishing group velocities, slow modes can be normalized such that their power flow is 1. This is the classical convention adopted hereafter. We start by slow photons. Since the left-hand side of eq 10 equals 1, the equality of the electric and magnetic energy densities directly implies that both the electric and magnetic fields scale as

a perfect balance, for photons when the quasi-stationary pattern of photonic waveguides becomes perfectly stationary or for plasmons when the energy flows in the metal and in the dielectrics are exactly opposite. Albeit intuitive and widespread, see for instance the introduction of the recent review article,6 this picture however carries some flaw. The power flows indeed never exactly cancel out, even as the energy velocity asymptotically tends toward zero. A convincing counterexample is found by considering the MIM gap mode, for which, as shown above, the power flows in the dielectric gap and in the metal clads are respectively Pfor = −4h02/(ωε0εm) > 0 and Pback = 2h02/(ωε0εm) < 0 in the limit of small gap widths (i.e., Pfor = 2 and Pback = −1 for a normalized gap mode). The energy flows do not balance: no matter the smallness of vg and no matter the convention for the normalization, the power flow in the gap is twice as larger as the power flow in the metal on that example. For a general and intuitive argument clearly evidencing that power flows cannot exactly cancel, imagine an adiabatic taper that progressively transforms a fast mode with a power flow Φf incident from the left into an outgoing slow mode with a power flow Φs; see Figure 3. Further wrongly assume that the opposite

Figure 3. Impossibility of adiabatically slowing down up to vanishing vg (stopping light) if modes with vg = 0 have a null power flow. Indeed, perfect adiabaticity is theoretically possible as shown with studies on 1D lossless thin-film slow light channels, for which Φ1 is null by construction.

|En| , |H n| ∝

ng

(11)

Figure 4. Field enhancement of normalized slow waves with unitary power flows. (a) Photonic case: both the electric and magnetic field intensities scale as the first power of ng. Since the taper is almost perfect (0.997 efficiency for this 2D simulation47), a strong ×20 intensity enhancement is observed between the slow mode (ng = 100) and the fast one (ng ≈ 3). Note that the expected ×30 enhancement is not really implemented because the transverse profiles of the slow and fast modes are slightly different. (b) Plasmonic MIM case: because of the additional transversal geometry change (gap width reduction) as the group velocity is lowered, the electric field intensity (not the magnetic one) increases at a much faster rate, proportionally to ng2. The normalized intensity profiles are computed with the aperiodic Fourier modal method.40 8

DOI: 10.1021/acsphotonics.8b01337 ACS Photonics 2019, 6, 4−17

ACS Photonics

Perspective

Table 1. Scaling of the Main Properties of Structural Slow Light and Slow Plasmons (Representative Measured Values Are Given in Square Brackets)

are indeed differences between the photonic and plasmonic cases. Let us denote again by Pfor and Pback the positive powers flowing forward or backward. The ratio u = Pback/Pfor (0 ≤ u < 1) is an important parameter that quantifies the slowness. For photonic waveguide modes operating outside a photonic gap, u is strictly smaller than 1, and for a normalized Bloch mode with a unitary power flow (Pfor − Pback = 1), we have Pfor ≈ Pback ∝ ng ≫ 1 with (Pback − Pfor)/Pfor ∝ vg/c, since the electromagnetic fields scale proportionally to ng . This implies that, in the slow light regime, the quasi-standing-wave patterns of normalized photonic waveguide modes are composed of two counterpropagating waves, each carrying a huge power flow that diverges linearly with ng. The plasmonic case is markedly different. As the group velocity vanishes, the plasmonic electric field scales with ng, but since the confinement also scales with 1/ng, Pback and Pfor both become constant. For instance, for the MIM waveguide studied in Section 2, Pfor = 2 and Pback = 1 for a normalized gap mode. In sharp contrast with photonic waveguides, each flow becomes independent of ng in the limit of small gaps. By way of summary, Table 1 provides a recapitulation of main properties of structural slow photons and slow plasmons.

To derive eq 11, we have additionally neglected the spatial mode profile dispersion. This assumption, which is only approximately valid for intermediately small values of the group velocity, becomes perfectly valid for lower speeds, as vg only significantly decreases in a narrow spectral range. This wellknown scaling law,38 illustrated in Figure 4a, represents the building block for slow light applications based on a strong boost of light−matter interactions. For translation-invariant plasmonic waveguides, eq 9 holds, and by assuming unitary power flows, we obtain 1 c × ng 4

y |e|2 + μ0 |h|2 zzzdx dy = 1 ∬ ijjjkε0 ∂∂ωε ω {

(12)

In sharp contrast with the photonic case, it is essential to realize that one cannot obtain directly the scaling laws for e or h from eq 12, owing to the fact that one should change at least one transverse characteristic dimension of the z-invariant plasmonic waveguide to tune down the group velocity, thereby radically changing the transverse mode profile. At large ng for deepsubwavelength confinements, self-sustained energy oscillations between magnetic and electric energies no longer hold, and the oscillations are restored by considering the kinetic energy of the free carriers involved in the subwavelength plasmon mode.3 A precise analysis of the scaling of e and h with ng requires knowing the exact geometry, but, as shown in Section 2 on the textbook example of MIM waveguides with asymptotically small dielectric gaps, the magnetic energy becomes negligible and the general trend is imposed by the balance between the electron-kinetic and electric-field energies:48 |En| ∝ ng , |H n| ∝ Cte



LIMITATIONS OF REAL WAVEGUIDES Despite their interesting physical properties, slow waves suffer in practice from important limitations, in particular regarding the operation bandwidth and the propagation distance. The available bandwidth, that is, the information content that can be slowed down or stored, is limited by dispersion, whereas the propagation distance, which defines the storage time and the interaction length, is limited by bulk-material absorption and scattering loss due to fabrication imperfections. The latter are well documented in the context of photonic-crystal waveguides.38,50−52 For plasmonic waveguides, there are only a few studies on roughness-induced attenuation, and one generally assumes that the attenuation is dominantly due to absorption. It is important to have the intrinsic differences and limitations in mind to optimally exploit slowness with the appropriate scheme in applications. The following subsections briefly review key aspects.

(13)

for normalized plasmonic modes. Noticeably, the electric-field increase rate for small group velocities is much larger for plasmons than for photons. The reason comes from the additional transversal geometry change (gap width reduction) that is accompanying plasmon slowdowns, as illustrated in Figure 4b. Based on the previously presented difference in scaling between plasmonic and photonic slow waves, it is interesting to reconsider how the antiparallel power flows scale with ng. There 9

DOI: 10.1021/acsphotonics.8b01337 ACS Photonics 2019, 6, 4−17

ACS Photonics

Perspective

Bandwidth. A major source of limitation of slow light structure is dispersion. For any system, the operating frequency is fixed, but in general, the transported signal possesses a bandwidth, and one should consider how the slowness changes over the bandwidth. To quantify the limiting impact of groupvelocity dispersion, it is convenient to define a dimensionless coefficient ) as

Δng Δω =) ω ng

Figure 5. Attenuation channels for a tiny imperfection characterized by a localized permittivity change Δεδ(r − r0) in a slow waveguide. Energy can be either radiated into the cladding, backscattered, or absorbed if Im(Δε) ≠ 0. Radiation into the cladding is equivalent to absorption in metal clads for MIM waveguides.

(14)

which quantifies the variation Δω of ω due to a small variation Δng at a fixed operating value ng. Large bandwidths are achieved for large ) s, and reversely small bandwidths for small values. ) can be evaluated by considering the dispersion relation ω(β). By derivation, it is possible to infer another relation linking the frequency ω and the group velocity, or equivalently the group index ng. Consider first a photonic waveguide with a quadratic dispersion curve ω(β) = (β − π/a)2/2m + ω0 near a band-edge ω0, where a is the period and m =

∂ 2ω ∂k 2

En|2. The absorption by metal cladding for plasmonic waveguides deserves special care.54 The backscattered power Pback is proportional to |En·J|2 (the excitation modal coefficient of the 1 normalized back-propagating mode is 4 En·J ),44 so that Pback ∝ | En|4. These are only general trends, but they all tell us that the influence of material loss or fabrication imperfection on mode structure is strong at small group velocities, where even a tiny amount of loss produces large changes in the dispersion, such that the group velocity is never zero.55−57 We will investigate this aspect in more detail now. Roughness-Induced Attenuation of Photonic Waveguides. The propagation of light in real photonic waveguides is well documented in the literature; absorption can be neglected with photonic materials, so that roughness of the etched pattern is the main limiting technological factor. Since |En| ∝ ng , we have that Prad ∝ ng and Pback ∝ ng2, so that the corresponding attenuation lengths Lp scale with ng−1 and ng−2, respectively.38 These simple scaling laws have been confirmed by ab initio electromagnetic computations performed with Bloch-modeexpansion methods for realistic waveguide geometries incorporating roughness on the sidewalls of the etched holes.50 Theory predicts typical values for the propagation length that vary between Lp = 50 and 500 μm for ng = 100 and between Lp = 103 and 104 μm for ng = 30, depending on the accuracy of the fabrication processes generally on the order of one nanometer or less. These theoretical predictions are confirmed by experimental transmission and reflection measurements performed with 100−1000λ long waveguides,18,23,37,51,58 showing a weak reflected signal for low group indices and a weak transmitted one at large group indices. Absorption Loss in Plasmonic Waveguides. The extremely tight confinement in plasmonic components is due to the large wavevectors of plasmons. It is inevitably accompanied by substantial absorption loss, essentially given by the inverse of the damping rate of the metal. Thus, it is fundamentally difficult to achieve a long propagation distance at slow speeds. To analyze the attenuation due to metal absorption, one naturally may consider the expression of the absorbed power Pabs = ω/2 ∬ ε0 Im(εm)|En|2 dS , where the integral runs over the transverse cross-section of the plasmonic waveguide. Since En ∝ ng for MIM waveguides (see eq 13), we may infer that Pabs ∝ ng2 at first glance. This however neglects the dispersion of the transverse mode profile and therefore largely overestimates the absorption. Instead, we directly rely on eq 9, which is valid for all reciprocal geometries, and we find that the integral of |e|2 or |En|2 over a cross-section scales as ng. This evidences that Pabs actually scales as ng. More insight into the role of mode-profile dispersion can be gained by considering the textbook case of MIM waveguides.

−1

( )

is the

effective photon mass, which describes the flatness of the dispersion curve. For ng = (mc2/2)1/2(ω − ω0)1/2, we find )=

mc 2 ω0ng 2

(15) −2

which evidences that ) scales with ng . The bandwidth near a Brillouin zone boundary therefore rapidly decreases as the operating group velocity decreases. To estimate ) for slow plasmons, we conveniently use the 2ε dispersion relation of MIM waveguides, βg = − ε d ; see eq 3. m

We obtain ng = (2cεd /g )

(

dεm /εm 2 dω

), and using a Drude model

εm = ε∞(1 − ωp /ω ), ) then reads as 2

) = −1/3

2

(16)

and is independent of ng. The bandwidth of photonic waveguides with quadratic dispersion curves can be enhanced by engineering to locally flatten the dispersion curve,53 but even so, slow photonic waveguides operate in a quite narrow frequency range that remains much smaller than that of slow plasmonic waveguides. Propagation Length. Since light cannot propagate indefinitely at low speed due to fabrication imperfections (surface roughness) or intrinsic material limitations (ohmic losses), the delay is finite. In general, the longer the delay for a given device, the narrower the bandwidth, which is why the socalled “delay−bandwidth product” is such a useful figure of merit. Let us then consider the propagation length Lp (defined for a 1/e field attenuation) of slow waveguides. Qualitatively first, the impact of imperfections can be estimated by considering a local permittivity change Δε, a “Dirac” defect, in a waveguide. In the Born approximation, this defect acts as a local current source, J = −jωΔεEnδ(r − r0), which is proportional to Δε and to the driving field (the slow incident mode En). As illustrated in Figure 5, this local source may radiate into the cladding and induce absorption if the cladding or the defect absorbs or backscatter light into the waveguide by exciting the counter-propagating slow mode. The power Prad radiated into the cladding is proportional to |J|2, and since J is proportional to En, Prad ∝ |En|2. Similarly, since the power Pabs absorbed by the defect is proportional to J·En, Pabs ∝ | 10

DOI: 10.1021/acsphotonics.8b01337 ACS Photonics 2019, 6, 4−17

ACS Photonics

Perspective

group velocities lower than c/1000 were seemingly observed.62 For slow plasmons, experiments seemed to evidence that novel metallic waveguide structures drastically reduce the propagation speed and may even stop broadband light, producing the socalled “trapped rainbows”.63,64 As we will now show, the very notion of group velocity loses all meaning at such small group velocities for both photons and plasmons. There are two regimes for slow photons.52,56,65 Far from the band-edges, where the concept of group velocity applies, tiny fabrication errors introduce random spatial dephasing of the propagating field, and, as the interference between multiply scattered propagating waves may be destructive, propagating light is exponentially damped with an attenuation coefficient proportional to ng2 due to backscattering.38 This well-known scaling is valid only in a double limit, when the disorder level tends toward zero at a fast enough rate compared to vg to guarantee that the impact of random imperfections on transport remains perturbative.52 In practice, the disorder level, albeit very weak, is fixed and given by the employed nanofabrication technique. A sudden breakdown of the perturbative regime, which inevitably ceases to be valid at small speeds, thus eventually occurs. In this “ultraslow” regime, light propagating in the randomly perturbed periodic waveguide not only experiences random phase-shifts, but also strong back-reflection due to vg-enhanced impedance mismatches or stop-band reflection for slow light operated near band-edges.52 The back-reflection results in strong interference between multiply scattered waves, leading to the formation of wavelength-scale localized modes, similar in nature to photonic-crystal microcavity modes. In this regime, the wavenumber k cannot be considered as a good quantum number anymore, and the group velocity evidently loses physical significance.56,57 The same physics hold for slow plasmons, except that the situation is even worse since absorption loss additionally comes into play and the localization regime is even not reached in practice; the plasmon is damped in space or in time before forming disordered resonant cavities. The literature is quite misleading on that matter.

We showed in Section 2 that, as the gap width decreases, the gapplasmon effective index neff = cβ/ω ∝ λ/g increases; see eq 3. Thus, the penetration depth of the gap plasmon mode into the c metal, ω (neff 2 − εm)−1, decreases and becomes shorter than the skin depth. The effective area contributing to the cross-section integral in Pabs = ω/2 ∬ ε0 Im(εm)|En|2 dS diminishes. It is thus the field confinement in the immediate vicinity of the MI interfaces that is responsible for the ng scaling of the absorbed power and not the ng2 scaling of our naive approach. Despite this lowering, the attenuation due to absorption is drastic in plasmonic waveguides. For instance, for a Ag/SiO2/Ag waveguide with εAg = −27 + i1.5 at λ = 800, one finds Lp ≈ 300 nm for ng = 10 and barely 100 nm for ng = 30. Roughness-Induced Attenuation in Plasmonic Waveguides. Since Pback ∝ |En|4 and |En| ∝ ng for plasmonic waveguides (compared to |En| ∝ ng for photonic waveguides), roughness-induced backscattering is plasmonic waveguides is scaling as ∝ ng4 and is expected to significantly alter the propagation of slow plasmons. Moreover, since Pabs only scales as ng comparatively, one may wonder if backscattering may become the dominant attenuation mechanism at small group velocities. This possibility has not been explored experimentally. Theoretically since imperfections at metal−dielectric interfaces result in a large Δε, the local current source J is not simply proportional to the plasmon mode field and local-field corrections should be considered.59 Additionally, the plasmonmode profile disperses as the group velocity is lowered. This may explain why theoretical analyses on roughness-induced attenuation in plasmonic waveguides are rare. Imperfections in MIM waveguides with large gap widths (operating away from the slow regime) have first been documented in ref 60; a strong impact has been predicted but for relatively large roughness (∼4 nm) obtained for films deposited by use of conventional e-beam evaporation. Slow propagation and small roughness have been analyzed49 for either chemically synthesized metal films or flakes, for which the metal surfaces are essentially atomically flat, except for some potential monatomic adlayer defects or high-quality polycrystalline metal films composed of large grains with a small residual roughness and separated by deep valleys that are remnants of grain boundaries.61 The predictive conclusions are that backreflection due to imperfection is completely negligible for chemically synthesized metal films. For polycrystalline films, back-reflection is predicted to be much stronger than for photonic crystal waveguides, in accordance with the scaling laws, and become more impactful than absorption for very small speed regimes (ng > 30). We might think that slow photons are much more sensitive to roughness than slow plasmons. This is actually not true, as evidenced by the different scaling laws in ng4 for plasmons and ng2 for photons. The strong impact of roughness is in fact hidden by absorption for usual slowness (ng ≈ 5−20). Despite having an obvious weaker scaling ∝ng2 than backscattering, absorption is the dominant attenuation channel.49 The reason is simply that absorption (∝ng) occurs over the entire metal volume, whereas backscattering (∝ng 4 ) is due only to surface-localized defects. Impossibility to Reach Ultraslow Regimes. Some of the first experimental investigations of the speed of light transport in monomode photonic-crystal waveguides reported group velocities of c/50 and c/150,37 whereas in other types of experiments



LIGHT EMISSION AND LDOS ENGINEERING IN SLOW WAVEGUIDES Controlling light emission at the single-photon level is one of the major challenges in modern photonics, with great promises for quantum communications, quantum computing, and optoelectronics thanks to devices such as highly efficient nonclassical light sources, low-threshold lasers, and single-photon switches or diodes. Probably the most important problem to address to reach this aim is the realization of efficient coupling between an emitter and a single guided mode. This coupling is described by the β-factor that is the ratio between the power emitted into the mode of interest and the total emitted power. It may thus be made close to unity by boosting the emission into the mode and/or reducing the emission into the other decay channels. Quantitatively, the amplitude coupling coefficient between a dipole-current J0δ(r − r0) and translation-invariant43 or periodic waveguides44 is proportional to the electric field of the normalized mode at the source position r0. Slow modes with large field enhancements (∝ ng for photons and ∝ng for plasmons) are thus expected to offer near-unity couplings. In this section, we review the basic principles underlying light emission in slow waveguides and identify the most promising strategies for near-unity β-factors. 11

DOI: 10.1021/acsphotonics.8b01337 ACS Photonics 2019, 6, 4−17

ACS Photonics

Perspective

Figure 6. Efficient funneling of the spontaneous decay of individual optical emitters into slow modes. (a) Slow photons: near-unity funneling into a single slow mode can be achieved provided that the LDOS of the slow mode is much higher than the LDOS of the clad modes. (b) Slow plasmons: even if tiny gaps lead to gap plasmons with extremely large LDOS, near-unity funneling is prohibited because of near-field nonradiative decay in the metal (quenching).

Figure 7. Slow plasmons on single metal−dielectric interfaces. (a) Radiation of a vertical dipole current above a Ag/polymer interface. The color scale represents log10(|E|2) of the total electric field for a dipole−metal separation distance d = 4 nm. The oscillating dipole decay is coupled into radiation modes detected in the far field, into surface plasmons launched on the surface, or directly transformed by near-field coupling as Joule heating in the metal just beneath the source. (b) Usual dispersion relation, Re k̃ − ω, of surface plasmons at a flat interface between a Drude-metal with a relative permittivity εm = 1 − ωp2/(ω2 − iγω) and a dielectric material with a relative permittivity εd = 1. (c) Quasi-normal mode dispersion relation, Re ω̃ − k. Slow plasmons with large k show up, but their lifetime is vanishingly small. The horizontal asymptote (dashed curve) occurs for a complex frequency ω̃ SP, verifying εm(ω̃ SP) + εd = 0. (b and c) Both curves are obtained by solving eq 17. ωp is the plasma frequency and γ = 0.02ωp. The black line is the light line of the air clad.

in the presence of disorder scales as m ,67 with m the effective photon mass. One should therefore preferably opt for photonic waveguides with a large effective photon mass, i.e., flat dispersion curves, for optimal performance. To date, near-unity β-factors (β ≈ 0.98) have been reported with quantum dots embedded in semiclosed photonic-crystal waveguides operating at ng > 50.9 Such a high efficiency is achieved on a narrow spectral range, yet it may be broadened up to a few tens of nanometers at the cost of a slightly lower β-factor by operating at lower ng values. As an alternative approach to embedded quantum dots, important efforts are currently being made to trap cold atoms near photonic structures.68 A recent study reports β ≈ 0.5 for cold atoms trapped in the so-called corrugated “alligator” waveguide designed to achieve long-range interactions between distant atoms for ng ≈ 11.69 The structure suffers from a much lower effective photon mass and fewer degrees of freedom for design compared to a photonic-crystaltype waveguide,67 suggesting that higher performances may be achieved with new geometries.

Emission in Slow Photonic Waveguides. The possibility of achieving near-unity β-factors is well comprehended for slow photonic waveguides; see Figure 6a. On one hand, the enhancement (Purcell effect) of the LDOS scales with ng in 1D systems, thereby boosting the emission into the slow mode when approaching the band-edge. On the other hand, photoniccrystal waveguides offer great flexibility in loss engineering, such that it is also possible to drastically reduce emission into the other decay channels.44 β-Factors above 90% can hence be achieved even without strong LDOS enhancement, i.e., by operating at moderate ng values. Unsurprisingly, the main limitation in terms of performance comes from fabrication imperfections, which yield the formation of localized modes with unpredictable spatial positions, resonant frequencies, and quality factors.66 This limitation may be mitigated by using waveguides shorter than the localization length. In addition, when using emitters with a sharp spectral response, it is necessary to match the emission frequency with a desired ng. One can show that the highest ng that may be aimed at 12

DOI: 10.1021/acsphotonics.8b01337 ACS Photonics 2019, 6, 4−17

ACS Photonics

Perspective

the inverse of Im ω̃ . The relation dispersion, Re ω̃ − k, shown in Figure 7c, is calculated by solving the transcendental eq 17 at complex frequencies for fixed k. Now, plasmon QNMs with arbitrary large k’s (since quantum effects are not considered71) are obtained for a certain complex frequency ω̃ SP, such that εm(ω̃ SP) + εd = 0 i.e., ω̃SP = ωp (1 + εd) + iγ /2 for γ ≪ ωp. The QNMs form a complete basis, so that the field radiated by the dipole at the (real) excitation frequency ω can be reconstructed from a sum over all QNMs.74 For illumination by a local oscillating dipole-current placed in the near field, QNMs with arbitrarily large k’s are excited. Intuitively, we anticipate that all modes are excited in phase at the dipole position, so that their respective contributions add up constructively especially just beneath the dipole, where strong absorption takes place, as evidenced by the hot spot in Figure 7a. Away from the source on the surface, all the sinusoidal QNMs are not phased and the sum of their field is no longer constructive, so that the contribution of high-k modes to the reconstructed field rapidly decreases away from the dipole in the metal, as shown in Figure 7a. Moreover, very high-k modes have a very fast transversal decay (∝k−1) and are thus excited only when the dipole is very close to the surface. This implies that, as the distance d between the dipole and the surface decreases, the hot spot becomes more intense and localized. Thus, with the QNM interpretation, in contrast with the guided mode interpretation, high-k slow plasmons play an essential (albeit detrimental) role in the light emission process; quenching drastically increases when d decreases (the decay rate scales as (ωd/c)−3 in the static limit71) and when the dipole frequency ω matches Re ω̃ SP, since the excitation coefficients of QNMs with a frequency ω̃ are proportional to (ω − ω̃ )−1.74 The intuitive description of the role of “slow” plasmons has been carefully confirmed with numerical simulations obtained for a silver nanorod in air. By computing several hundreds of QNMs, it has been shown that the hot spot can be accurately reconstructed in the QNM basis up to d as small as 1.5 nm; see Figure 3 in ref 75. The spectrum of the plasmon eigenfrequencies does not form a continuum for the nanorod like for the infinite flat interface. Rather they take discrete values. However, for very small separation distance between the dipole and the antenna, the dipole “sees” an almost flat surface. The nanorod spectrum presents an accumulation point at the plasmon resonance frequency of the flat interface for which εm(ω̃ SP) + εd = 0,75 showing that these nanorod QNMs are indeed very similar to those of the flat surface. Dipolar Emission in MIM Waveguides. Let us now consider the emission of an oscillating dipole in the dielectric gap of a MIM waveguide, in the limit of tiny gaps (see Figure 6b). The conclusions on the role of slow plasmons will be completely different. Indeed, in full analogy with the single-interface or antenna cases, the nonradiative decay in the metal is again present. However, as discussed above, the slow gap plasmons (Re k̃ − ω) that are launched in the gap have an extremely large LDOS. Actually, it can be shown analytically that the nonradiative decay rate (quenching) and the slow-plasmon decay rate both scale as (k0g)−3. None of the two decay channels prevail as the gap width g becomes very small. The branching ratio η (the fraction of the total decay that effectively decay into the gap mode) becomes an intrinsic quantity, as it is independent of g, depending solely on εd and on the metal

Emission in Slow Plasmonic Waveguides. Compared to light emission in photonic structures, the decay rate of emitters into slow plasmons70 is larger and the bandwidth is definitely much broader. The only gray area is the efficiency issue because of the prominence of near-field nonradiative decay in the metal for tiny gaps; see Figure 6b. The whole literature is not always clear on the role played by plasmons, especially “slow” plasmons with large parallel momenta, into the decay process. In the remainder of this section, we analyze this issue, step by step, first revisiting the classical light emission by a dipole-current located just above a single MI interface, and then considering the case of light emission by a dipole-current located in an MIM waveguide, the objective being to completely clarify the role of slow plasmons. Dipolar Emission near a Metal Surface. This problem has been reviewed by many authors, including ref 71. It contains a rich variety of physical phenomena. In particular, we know that the lifetime of an atom placed close to the surface can be typically reduced by several orders of magnitude,72 implying that the LDOS close to an interface is drastic. The reduction is due to several decay channels. A part of the oscillating dipole decay is coupled into radiation modes detected in the far field, another part launches surface plasmons on the surface, and another significant part is directly transformed by near-field coupling into ohmic loss in the metal just beneath the source, resulting in a quenching of the far-field emission especially when the dipole is very close to the surface.72 The plasmon dispersion relation of flat MI interfaces is well known:73 k=

ω εdεm /(εm + εd) c

(17)

with k the parallel in-plane wavevector, ω the frequency, εm and εd the relative permittivities of the metal and dielectric materials. We will use a Drude metal εm = 1 − ωp2/(ω2 − iγω) for illustration hereafter. Since εm is complex, either k or ω is complex and two interpretations, in relation with slow plasmons, are usually considered. The first interpretation relies on the dispersion relation obtained by fixing the frequency to a real value (imposed by the driving dipole) and by directly solving eq 17 for complex k̃’s, denoted with a tilde to emphasize their complex nature; the imaginary part of k̃ is related to the inverse of the attenuation length of the plasmon mode. The dispersion relation Re k̃ − ω is shown in Figure 7b. This plasmon mode is the analogue of the photonic-crystal mode of the previous section or the gapplasmon mode analyzed in Section 2; it is a guided mode with a spatial damping due to absorption. Because realistic values for the loss are taken into account in Figure 7, the relation exhibits a back bending for a relatively low εd ωp/c . In this real frequency value of Re k̃, Re k ̃ ≈ 2(1 + εd)

representation, no slow plasmons with large k are launched,73 even when the emission frequency is close to ωp/(1 + εd)1/2 and the dipole is very close to the surface. This result contrasts with the emission in MIM waveguides with small gaps, presented below, for which we will show that guided gap plasmons with large k̃’s are excited at any frequency. Another representation of the same physics relies on the quasinormal modes (QNMs)74 that are the natural plasmon resonances with a complex frequency ω̃ of the MI interface. The QNMs possess a sinusoidal spatial dependence parallel to the surface (no spatial damping) and a finite lifetime proportional to

(

loss Im εm, η ≈ 1 + 13

Imεm 2εd

−1

)

.54,76 DOI: 10.1021/acsphotonics.8b01337 ACS Photonics 2019, 6, 4−17

ACS Photonics

Perspective

of the plasmon amplitude. Nanofocusing devices have been used in many applications, yet many new ones appear every year for high-harmonic generation and hot-electron plasmon-assisted generation; see ref 63 for a review. The second important slow-plasmon device is nanocavity or nanoantenna. At deep-subwavelength dimensions, these components are essentially “Fabry−Perot cavities” involving slow plasmons that propagate in a slow waveguide and are reflected at the waveguide terminations. For instance, the dipolar resonance of a simple nanorod immersed in a dielectric background can be considered as resulting from the bouncing back and forth of two IMI plasmon modes between the wire facets. Similarly, two MIM plasmon modes are involved in the dipole and quadrupole modes of cut-wire-pair resonators and nanopatch antennas. As the transverse dimensions (the rod diameter or gap width) are scaled down, the physics is the same. The fundamental IMI or MIM plasmon modes experience a monotonic increase in transverse localization and effective index (see eq 3), implying that by reducing the nanoresonator length by the same scaling factor, the resonance wavelength is maintained.90,91 This simple model that promotes slow plasmons as an integral part of the near-field response of deep-subwavelength nanoresonators predicts well the resonance frequency. Pushed further, it also explains how important quantities such as the Q factor, the absorption, or the facet reflectivity scale.90 In particular, it predicts that the Purcell factor scales as g−3,54 making slowplasmon nanoresonators or nanoantennas a serious candidate for an ultimate control of light emission,10,91,92 especially if an accurate control of the conversion of the slow gap plasmons into free-space photons is implemented by engineering the tapered facets to boost the photon generation yield. In summary, we have provided a comparative overview of the physical mechanisms by which wave deceleration is achieved in plasmonic and photonic waveguides. The scaling laws of various important physical quantities (field enhancement, bandwidth, attenuation) as a function of ng have been presented; see Table 1. Slow plasmonic and photonic waveguides share many common features, but they have also distinct properties. We think that the key fundamental property responsible for the difference, notably for the mode intensities (∝ng for slow photons and ∝ng2 for slow plasmons), is that the slow mode of photonic waveguides retains the same profile as light is slowed down, whereas the plasmon-mode profile drastically changes for plasmonic waveguides. Slow-plasmon devices are generally operating at much lower ng than slow-photon devices (ng = 15 corresponds to insulator gap width of g = 2 nm at visible wavelengths). One should nevertheless remember that slow plasmons and photons have different field-enhancement scalings (|En| ∝ ng for plasmons and |En| ∝ ng for photons), so that it is rather illogical to compare plasmonic and photonic devices for the same ng. In this respect, a plasmonic taper operating at ng = 15 would be equivalent to a photonic taper operating at ng ≈ 225 in our analogy. Toward strong field enhancement and ultimate miniaturization, the recent literature has promoted slow plasmons as the pinnacle of slowness in nanophotonics. Indeed, field enhancement and bandwidth are decisive figures of merit that render plasmon wave deceleration incomparably more attractive than photon wave deceleration. However, from other perspectives, the benefit of the deep-subwavelength confinement (or miniaturization) brought by plasmons is not as clear. All-optical switches are perhaps the most important building blocks for on-

The good news is that the emission is no longer quenched for tiny gaps so that, provided that the gap plasmon mode is efficiently tapered into a photonic mode,54 large efficiencies are achievable. The downside of the existence of a branching ratio is that the efficiency cannot reach unity. This imposes a limitation on the β-factors that usually not exceed 80% for MIM54,77−79 or IMI80,81 waveguides and remain significantly smaller than their photonics counterparts.



PERSPECTIVES AND CONCLUSIONS The applications of slow photonic waveguides in both linear and nonlinear optics are so diverse that it is impossible to cover them all even briefly. Slow photons are useful to realize compact alloptical buffer memories, which is one of the most important missing components for the construction of various all-optical processing devices such as photonic routers. They are also useful to funnel light emission into a single channel and to enhance various light−matter interactions in nonlinear optics. We have no choice but to note that slow photon applications rarely feature a speed reduction greater than 10−20. For instance, moderate ng values (∼30) are used in photonic-crystal and coupled-resonator waveguides to implement delay lines over millimeter distances.23,82,83 For short devices such as low-power optical switches,20 typical ng values of ∼50 are met in practice. Interesting perspectives may arise with regimes of smaller group velocities. Remembering that the attenuation length Lp of photonic waveguides at λ = 1500 nm and ng = 100 varies from 50 to 500 μm according to the fabrication technology and that Lp ∝ ng−2, we anticipate that slow light at ng = 300−500 may propagate over 10-period long distances before backscattering takes place. Since remarkably short and efficient tapers have been designed47,84 (see Figure 4a) and successfully implemented in many slow-light experiments,9,85−87 devices that feature a sudden reduction of the group velocity followed by a few-periods-long ultraslow section and a reciprocal acceleration to return to the initial speed appear within reach. Applications may include all-optical processing on chip, biosensors, or new resonator-like devices with anomalous step-like spectral transmission. However, as ultraslow photonic waveguides and microcavities have a lot in common, especially for large ng when it becomes difficult to engineer waveguide dispersion,84 attention should be paid not to repeat earlier works with high-Q microcavities. Comparatively, slow-plasmon studies are more prospective. Slow plasmons can only propagate over very small distances, due to absorption. Their applications must therefore inevitably rely on components with very short lengths, not longer than a few hundreds of nanometers. Two main slow-plasmon components have emerged in the past decade, namely, nanofocusing tapers and slow-plasmon nanocavities or nanoantennas. Nanofocusing devices that taper and concentrate optical energy into nanoscale objects and structures probably constitute the most important device using slow plasmons. Various geometries based on MIM and IMI configurations have been successfully implemented, e.g., conical metal rods with a rounded tip for near-field imaging, metal wedges on substrates such as in Figure 3, or exotic structures such as metal− insulator−insulator (MII) wedges88 or tapered chains of coupled nanoparticles;89 see ref 36 for a review. The whole literature suggests that adiabatic tapering into deep-subwavelength slow plasmons is relatively easy to implement, since for sufficiently short tapers the local field enhancement along the taper could efficiently compensate for the dissipative reduction 14

DOI: 10.1021/acsphotonics.8b01337 ACS Photonics 2019, 6, 4−17

ACS Photonics

Perspective

(13) Ma, R.-M.; Oulton, R. F.; Sorger, V. J.; Zhang, X. Plasmon lasers: Coherent light source at molecular scales. Laser Photon. Rev. 2013, 7, 1−21. (14) Xiong, C.; Monat, C.; Clark, A. S.; Grillet, C.; Marshall, G. D.; Steel, M. J.; Li, J.; O’Faolain, L.; Krauss, T. F.; Rarity, J. G.; Eggleton, B. J. Slow-light enhanced correlated photon pair generation in a silicon photonic crystal waveguide. Opt. Lett. 2011, 36, 3413−3415. (15) Tame, M. S.; McEnery, K. R.; Ö zdemir, S. K.; Lee, J.; Maier, S. A.; Kim, M. S. Quantum plasmonics. Nat. Phys. 2013, 9, 329−340. (16) Krauss, T. F. Photonic Crystals shine on. Phys. World 2006, 19, 32−36. (17) Kauranen, M.; Zayats, A. V. Nonlinear plasmonics. Nat. Photonics 2012, 6, 737. (18) Notomi, M. Manipulating light with strongly modulated photonic crystals. Rep. Prog. Phys. 2010, 73, No. 096501. (19) Soljačić, M.; Joannopoulos, J. D. Enhancement of nonlinear effects using photonic crystals. Nat. Mater. 2004, 3, 211−219. (20) Beggs, D. M.; White, T. P.; O’Faolain, L.; Krauss, T. F. Ultracompact and low-power optical switch based on silicon photonic crystals. Opt. Lett. 2008, 33, 147−149. (21) Vlasov, Y. A.; O’Boyle, M.; Hamann, H. F.; McNab, S. J. Active control of slow light on a chip with photonic crystal waveguides. Nature 2005, 438, 65−69. (22) Stockman, M. I. Nanoplasmonics: The physics behind the applications. Phys. Today 2011, 64, 39−44. (23) Sancho, J.; Bourderionnet, J.; Lloret, J.; Combrié, S.; Gasulla, I.; Xavier, S.; Sales, S.; Colman, P.; Lehoucq, G.; Dolfi, D.; Capmany, J.; De Rossi, A. Integrable microwave filter based on a photonic crystal delay line. Nat. Commun. 2012, 3, 1075−84. (24) Giugni, A.; Torre, B.; Toma, A.; Francardi, M.; Malerba, M.; Alabastri, A.; Proietti Zaccaria, R.; Stockman, M. I.; Di Fabrizio, E. Hotelectron nanoscopy using adiabatic compression of surface plasmons. Nat. Nanotechnol. 2013, 8, 845−852. (25) Harutyunyan, H.; Martinson, A. B. F.; Rosenmann, D.; Khosravi Khorashad, L.; Besteiro, L. V.; Govorov, A. O.; Wiederrecht, G. P. Anomalous ultrafast dynamics of hot plasmonic electrons in nanostructures with hot spots. Nat. Nanotechnol. 2015, 10, 770−775. (26) Sorger, V. J.; Ye, Z.; Oulton, R. F.; Wang, Y.; Bartal, G.; Yin, X.; Zhang, X. Experimental demonstration of low-loss optical waveguiding at deep sub-wavelength scales. Nat. Commun. 2011, 2, 331. (27) Eggleston, M. S.; Messer, K.; Zhang, L.; Yablonovitch, E.; Wu, M. C. Optical antenna enhanced spontaneous emission. Proc. Natl. Acad. Sci. U. S. A. 2015, 112, 1704−1709. (28) Demichel, O.; Petit, M.; Viarbitskaya, S.; Méjard, R.; de Fornel, F.; Hertz, E.; Billard, F.; Bouhelier, A.; Cluzel, B. Dynamics, Efficiency, and Energy Distribution of Nonlinear Plasmon-Assisted Generation of Hot Carriers. ACS Photonics 2016, 3, 791−795. (29) Volkov, V. S.; Bozhevolnyi, S. I.; Devaux, E.; Ebbesen, T. W. Bend loss for channel plasmon polaritons. Extremely Large GroupVelocity Dispersion of Line-Defect Waveguides in Photonic Crystal Slabs. Appl. Phys. Lett. 2006, 89, 143108. (30) Takahara, J.; Yamagishi, S.; Taki, H.; Morimoto, A.; Kobayashi, T. Guiding of a one-dimensional optical beam with nanometer diameter. Opt. Lett. 1997, 22, 475−477. (31) Dionne, J. A.; Sweatlock, L. A.; Atwater, H. A.; Polman, A. Planar metal plasmon waveguides: frequency-dependent dispersion, propagation, localization, and loss beyond the free electron model. Phys. Rev. B: Condens. Matter Mater. Phys. 2005, 72, No. 075405. (32) Dionne, J. A.; Sweatlock, L. A.; Atwater, H. A.; Polman, A. Plasmon slot waveguides: Towards chip-scale propagation with subwavelength-scale localization. Phys. Rev. B: Condens. Matter Mater. Phys. 2006, 73, No. 035407. (33) Kurokawa, Y.; Miyazaki, H. T. Metal-insulator-metal plasmon nanocavities: Analysis of optical properties. Phys. Rev. B: Condens. Matter Mater. Phys. 2007, 75, No. 035411. (34) Oulton, R. F.; Bartal, G.; Pile, D. F. P.; Zhang, X. Confinement and propagation characteristics of subwavelength plasmonic modes. New J. Phys. 2008, 10, 105018.

chip all-optical processing. Switching is generally achieved by modulating the cavity transmission intensity by illuminating the cavity with a control light beam, which shifts the resonance wavelength via the intensity-dependent refractive index of the cavity material. The required switching power scales as V/Q2,18 because the light intensity in the cavity is proportional to Q/V, and the required wavelength shift is proportional to Q−1. In contrast to spontaneous rate (Purcell) enhancements,10 it thus becomes unclear if plasmons, with their drastic reduction of V often counterbalanced by an inevitable Q reduction due to absorption, can really be competitive in devices exploiting, for example, optical nonlinearities. We expect that this short comparison leads to a balanced point of view between slow plasmonic and photonic waveguides.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Philippe Lalanne: 0000-0003-1979-2290 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The work was partly supported by the French State managed by the French National Agency for Research (ANR) in the frame of the “Investments for the Future” Programme IdEx Bordeaux− LAPHIA (Grant No. ANR-10-IDEX-03-02) and by Bordeaux University.



REFERENCES

(1) Hau, L. V.; Harris, S. E.; Dutton, Z.; Behroozi, C. H. Light speed reduction to 17 m/s in a cold atomic gas. Nature 1999, 397, 594. (2) Philips, D. F.; Fleischhauer, A.; Mair, A.; Walsworth, R. L.; Lukin, M. D. Storage of light in atomic vapor. Phys. Rev. Lett. 2001, 86, 783− 786. (3) Khurgin, J. B. Slow light in various media: a tutorial. Adv. Opt. Photonics 2010, 2, 287. (4) Boyd, R. W. Material slow light and structural slow light: similarities and differences for nonlinear optics. J. Opt. Soc. Am. B 2011, 28, A38. (5) Baba, T. Slow light in Photonic crystals. Nat. Photonics 2008, 2, 465−473. (6) Tsakmakidis, K. L.; Hess, O.; Boyd, R. W.; Zhang, X. Ultraslow waves on the nanoscale. Science 2017, 358, eaan5196. (7) Atwater, H. A.; Polman, A. Plasmonics for improved photovoltaic devices. Nat. Mater. 2010, 9, 205−213. (8) Scullion, M. G.; Di Falco, A.; Krauss, T. F. Slotted photonic crystal cavities with integrated microfluidics for biosensing applications. Biosens. Bioelectron. 2011, 27, 101−105. (9) Arcari, M.; Sö llner, I.; Javadi, A.; Lindskov Hansen, S.; Mahmoodian, S.; Liu, J.; Thyrrestrup, H.; Lee, E. H.; Song, J. D.; Stobbe, S.; Lodahl, P. Near-Unity Coupling Efficiency of a Quantum Emitter to a Photonic Crystal Waveguide. Phys. Rev. Lett. 2014, 113, No. 093603. (10) Akselrod, G. M.; Argyropoulos, C.; Hoang, T. B.; Ciracì, C.; Fang, C.; Huang, J.; Smith, D. R.; Mikkelsen, M. H. Probing the mechanisms of large Purcell enhancement in plasmonic nanoantennas. Nat. Photonics 2014, 8, 835. (11) Berini, P.; De Leon, I. Surface plasmon-polariton amplifiers and lasers. Nat. Photonics 2012, 6, 16−24. (12) Yang, J.-K.; Noh, H.; Rooks, M. J.; Solomon, G. S.; Vollmer, F.; Cao, H. Lasing in localized modes of a slow light photonic crystal waveguide. Appl. Phys. Lett. 2011, 98, 241107. 15

DOI: 10.1021/acsphotonics.8b01337 ACS Photonics 2019, 6, 4−17

ACS Photonics

Perspective

(35) Stockman, M. I. Nanoplasmonics: past, present, and glimpse into future. Opt. Express 2011, 19, 22029. (36) Gramotnev, D. K.; Bozhevolnyi, S. I. Nanofocusing of electromagnetic radiation. Nat. Photonics 2014, 8, 13−22. (37) Notomi, M.; Yamada, K.; Shinya, A.; Takahashi, J.; Takahashi, C.; Yokohama, I. Extremely large group-velocity dispersion of linedefect waveguides in photonic crystal slabs. Phys. Rev. Lett. 2001, 87, 253902. (38) Hughes, S.; Ramunno, L.; Young, J. F.; Sipe, J. E. Extrinsic optical scattering loss in photonic crystal waveguides: role of fabrication disorder and photon group velocity. Phys. Rev. Lett. 2005, 94, No. 033903. (39) Palik, E.D. Handbook of Optical Constants of Solids; Academic Press: NY, Part II, 1985. (40) Lalanne, P.; Besbes, M.; Hugonin, J. P.; van Haver, S.; Janssen, O. T. A.; Nugrowati, A. M.; Xu, M.; Pereira, S. F.; Urbach, H. P.; van de Nes, A. S.; Bienstman, P.; Granet, G.; Moreau, A.; Helfert, S.; Sukharev, M.; Seideman, T.; Baida, F. I.; Guizal, B.; Van Labeke, D. Numerical analysis of a slit-groove diffraction problem. J. Eur. Opt. Soc. Rapid Publ. 2007, 2, No. 07022. (41) Tserkezis, C.; Mortensen, N. A.; Wubs, M. How nonlocal damping reduces plasmon-enhanced fluorescence in ultranarrow gaps. Phys. Rev. B: Condens. Matter Mater. Phys. 2017, 96, No. 085413. (42) Luo, Y.; Fernandez-Dominguez, A. I.; Wiener, A.; Maier, S. A.; Pendry, J. B. Surface Plasmons and Nonlocality: A Simple Model. Phys. Rev. Lett. 2013, 111, No. 093901. (43) Snyder, A. W.; Love, J. D. Optical Waveguide Theory; Chapman and Hall: NY, 1983. (44) Lecamp, G.; Hugonin, J. P.; Lalanne, P. Theoretical and computational concepts for periodic optical waveguides. Opt. Express 2007, 15, 11042−60. (45) Yeh, P. Optical Waves in Layered Media; J. Wiley and Sons: New York, 1988. (46) Bendickson, J. M.; Dowling, J. P.; Scalora, M. Analytic expressions for the electromagnetic mode density in finite, onedimensional, photonic band-gap structure. Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 1996, 53, 4107−4121. (47) Hugonin, J. P.; Lalanne, P.; White, T.; Krauss, T. F. Coupling into slow-mode photonic crystal waveguides. Opt. Lett. 2007, 32, 2638− 2640. (48) Khurgin, J. B. How to deal with the loss in plasmonics and metamaterials. Nat. Nanotechnol. 2015, 10, 1−6. (49) Coudert, S.; Duchateau, G.; Dilhaire, S.; Lalanne, P. Attenuation of slow Metal-Insulator-Metal plasmonic waveguides, from Joule absorption to roughness-induced backscattering. arXiv:1810.e07389. (50) Mazoyer, S.; Hugonin, J. P.; Lalanne, P. Disorder-induced multiple-scattering in photonic-crystal waveguides. Phys. Rev. Lett. 2009, 103, No. 063903. (51) Le Thomas, N.; Diao, Z.; Zhang, H.; Houdré, R. Statistical analysis of subnanometer residual disorder in photonic crystal waveguides: Correlation between slow light properties and structural properties. J. Vac. Sci. Technol., B: Nanotechnol. Microelectron.: Mater., Process., Meas., Phenom. 2011, 29, No. 051601. (52) Faggiani, R.; Baron, A.; Zang, X.; Lalouat, L.; Schulz, S. A.; O’Regan, B.; Vynck, K.; Cluzel, B.; de Fornel, F.; Krauss, T. F.; Lalanne, P. Lower bounds for the spatial extent of optical localized modes in 1D disordered periodic media. Sci. Rep. 2016, 6, 27037. (53) Frandsen, L. H.; Lavrinenko, A. V.; Fage-Pedersen, J.; Borel, P. I. Photonic crystal waveguides with semi-slow light and tailored dispersion properties. Opt. Express 2006, 14, 9444−9450. (54) Reza, A.; Dignam, M. M.; Hughes, S. Can light be stopped in realistic metamaterials? Nature 2008, 455, E10−E11. (55) Le Thomas, N.; Zhang, H.; Houdré, R. Light transport regimes in slow light photonic crystal waveguides. Phys. Rev. B: Condens. Matter Mater. Phys. 2009, 80, 125332. (56) Huisman, S. R.; Ctistis, G.; Stobbe, S.; Mosk, A. P.; Herek, J. L.; Lagendijk, A.; Lodahl, P.; Vos, W. L.; Pinkse, P. W. H. Measurement of a band-edge tail in the density of states of a photonic-crystal waveguide. Phys. Rev. B: Condens. Matter Mater. Phys. 2012, 86, 155154.

(57) Topolancik, J.; Ilic, B.; Vollmer, F. Experimental Observation of Strong Photon Localization in Disordered Photonic Crystal Waveguides. Phys. Rev. Lett. 2007, 99, 253901. (58) Johnson, S. G.; Povinelli, M. L.; Soljačić, M.; Karalis, A.; Jacobs, S.; Joannopoulos, J. D. Roughness losses and volume-current methods in photonic-crystal waveguides. Appl. Phys. B: Lasers Opt. 2005, 81, 283−293. (59) Min, C.; Veronis, G. Theoretical investigation of fabricationrelated disorders on the properties of subwavelength metal-dielectricmetal plasmonic waveguides. Opt. Express 2010, 18, 20939−20948. (60) Munoz, R. C.; Concha, A.; Mora, F.; Espejo, R.; Vidal, G.; Mulsow, M.; Arenas, C.; Kremer, G.; Moraga, L.; Esparza, R.; Haberle, P. Surface roughness and size effects of thin gold films on mica: Application of quantitative scanning tunneling microscopy. Phys. Rev. B: Condens. Matter Mater. Phys. 2000, 61, 4514−17. (61) Gersen, H.; Karle, T. J.; Engelen, R. J. P.; Bogaerts, W.; Korterik, J. P.; van Hulst, N. F.; Krauss, T. F.; Kuipers, L. Real-Space Observation of Ultraslow Light in Photonic Crystal Waveguides. Phys. Rev. Lett. 2005, 94, No. 073903. (62) Tsakmakidis, K. L.; Boardman, A. D.; Hess, O. ’Trapped rainbow’ storage of light in metamaterials. Nature 2007, 450, 397−401. (63) Gana, Q.; Gao, Y.; Wagner, K.; Vezenov, D.; Ding, Y. J.; Bartoli, F. J. Experimental verification of the rainbow trapping effect in adiabatic plasmonic gratings. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 5169−5173. (64) Engelen, R. J. P.; Mori, D.; Baba, T.; Kuipers, L. Two regimes of slow-light losses revealed by adiabatic reduction of group velocity. Phys. Rev. Lett. 2008, 101, 103901. (65) Smolka, S.; Thyrrestrup, H.; Sapienza, L.; Lehmann, T. B.; Rix, K. R.; Froufe-Pérez, L. S.; García, P. D.; Lodahl, P. Probing the statistical properties of Anderson localization with quantum emitters. New J. Phys. 2011, 13, No. 063044. (66) Zang, X.; Yang, J.; Faggiani, R.; Gill, C.; Petrov, P. G.; Hugonin, J.-P.; Vynck, K.; Bernon, S.; Bouyer, P.; Boyer, V.; Lalanne, P. Interaction between atoms and slow light: a study in waveguide design. Phys. Rev. Appl. 2016, 5, No. 024003. (67) Goban, A.; Hung, C.-L.; Yu, S. P.; Hood, J. D.; Muniz, J. A.; Lee, J. H.; Martin, M. J.; McClung, A. C.; Choi, K. S.; Chang, D. E.; Painter, O.; Kimble, H. J. Atom−light interactions in photonic crystals. Nat. Commun. 2014, 5, 3808. (68) Goban, A.; Hung, C.-L.; Hood, J. D.; Yu, S.-P.; Muniz, J. A.; Painter, O.; Kimble, H. J. Superradiance for atoms trapped along a photonic crystal waveguide. Phys. Rev. Lett. 2015, 115, No. 063601. (69) Bozhevolnyi, S. I.; Khurgin, J. B. Fundamental limitations in spontaneous emission rate of single-photon sources. Optica 2016, 3, 1418−1421. (70) Ford, G. W.; Weber, W. H. Electromagnetic interactions of molecules with metal surfaces. Phys. Rep. 1984, 113, 195−287. (71) Drexhage, K. H. Interaction of light with monomolecular dye layers. Prog. Opt. 1974, 12, 165. (72) Raether, H. Surface Plasmons on Smooth and Rough Surfaces and on Gratings; Springer-Verlag: Berlin, 1988. (73) Lalanne, P.; Yan, W.; Vynck, K.; Sauvan, C.; Hugonin, J.-P. Light interaction with photonic and plasmonic resonances. Laser Photonics Rev. 2018, 12, 1700113. (74) Yan, W.; Faggiani, R.; Lalanne, P. Rigorous modal analysis of plasmonic resonances. Phys. Rev. B: Condens. Matter Mater. Phys. 2018, 97, 205422. (75) Yang, J.; Faggiani, R.; Lalanne, P. Light emission in nanogaps: overcoming quenching. Nanoscale Horizons 2016, 1, 11−13. (76) Martín-Cano, D.; Martín-Moreno, L.; García-Vidal, F. J.; Moreno, E. Resonance Energy Transfer and Superradiance Mediated by Plasmonic Nanowaveguides. Nano Lett. 2010, 10, 3129−3134. (77) Sorger, V. J.; Pholchai, N.; Cubukcu, E.; Oulton, R. F.; Kolchin, P.; Borschel, C.; Gnauck, M.; Ronning, C.; Zhang, X. Strongly Enhanced Molecular Fluorescence inside a Nanoscale Waveguide Gap. Nano Lett. 2011, 11, 49077−4911. (78) Kongsuwan, N.; Demetriadou, A.; Chikkaraddy, R.; Benz, F.; Turek, V. A.; Keyser, U. F.; Baumberg, J. J.; Hess, O. Suppressed 16

DOI: 10.1021/acsphotonics.8b01337 ACS Photonics 2019, 6, 4−17

ACS Photonics

Perspective

Quenching and Strong-Coupling of Purcell-Enhanced Single-Molecule Emission in Plasmonic Nanocavities. ACS Photonics 2018, 5, 186−191. (79) Chen, Y.; Nielsen, T. R.; Gregersen, N.; Lodahl, P.; Mørk, J. Finite-element modeling of spontaneous emission of a quantum emitter at nanoscale proximity to plasmonic waveguides. Phys. Rev. B: Condens. Matter Mater. Phys. 2010, 81, 125431. (80) Yang, Y.; Zhen, B.; Hsu, C. W.; Miller, O. D.; Joannopoulos, J. D.; Soljačić, M. Optically Thin Metallic Films for High-Radiative-Efficiency Plasmonics. Nano Lett. 2016, 16, 4110−4117. (81) Notomi, M.; Kuramochi, E.; Tanabe, T. Large-scale arrays of ultrahigh-Q coupled nanocavities. Nat. Photonics 2008, 2, 741−747. (82) Melloni, A.; Canciamilla, A.; Ferrari, C.; Morichetti, F.; O’Faolain, L.; Krauss, T. F.; De La Rue, R.; Samarelli, A.; Sorel, M. Tunable delay lines in silicon photonics: coupled resonators and photonic crystals, a comparison. IEEE Photonics J. 2010, 2, 181−194. (83) Faggiani, R.; Jang, J.; Hostein, R.; Lalanne, P. Implementing structural slow light on short length scales: the photonic speed bump. Optica 2017, 4, 393−399. (84) Corcoran, B.; Monat, C.; Grillet, C.; Moss, D. J.; Eggleton, B. J.; White, T. P.; O’Faolain, L.; Krauss, T. F. Green light emission in silicon through slow-light enhanced third-harmonic generation in photoniccrystal waveguides. Nat. Photonics 2009, 3, 206−210. (85) Spasenovic, M.; et al. Experimental observation of evanescent modes at the interface to slow-light photonic crystal waveguides. Opt. Lett. 2011, 36, 1170. (86) Lodahl, P.; Mahmoodian, S.; Stobbe, S. Interfacing single photons and single quantum dots with photonic nanostructures. Rev. Mod. Phys. 2015, 87, 347. (87) Verhagen, E.; Kuipers, L. K.; Polman, A. Plasmonic nanofocusing in a dielectric wedge. Nano Lett. 2010, 10, 3665−3669. (88) Fernández-Domínguez, A. I.; Maier, S. A.; Pendry, J. B. Collection and concentration of light by touching spheres: a transformation optics approach. Phys. Rev. Lett. 2010, 105, 266807. (89) Yang, J.; Sauvan, C.; Jouanin, A.; Collin, S.; Pelouard, J. L.; Lalanne, P. Ultrasmall metal-insulator-metal nanoresonators: impact of slow-wave effects on the quality factor. Opt. Express 2012, 20, 16880− 16891. (90) Dorfmüller, J.; Vogelgesang, R.; Weitz, R. T.; Rockstuhl, C.; Etrich, C.; Pertsch, T.; Lederer, F.; Kern, K. Fabry-Perot resonances in one-dimensional plasmonic nanostructures. Nano Lett. 2009, 9, 2372− 2377. (91) Aizpurua, J.; Bryant, G. W.; Richter, L. J.; García de Abajo, F. J. Optical properties of coupled metallic nanorods for field-enhanced spectroscopy. Phys. Rev. B: Condens. Matter Mater. Phys. 2005, 71, 235420. (92) Russel, K. J.; Liu, T. L.; Cui, S.; Hu, E. L. Large spontaneous emission enhancement in plasmonic cavities. Nat. Photonics 2012, 6, 459−462.

17

DOI: 10.1021/acsphotonics.8b01337 ACS Photonics 2019, 6, 4−17