• Chinese Optics Letters
  • Vol. 17, Issue 12, 122402 (2019)
Shima Fardad1、2, Eric Schweisberger1, and Alessandro Salandrino1、2、*
Author Affiliations
  • 1Department of Electrical Engineering and Computer Science, The University of Kansas, Lawrence, KS 66045, USA
  • 2Information and Telecommunication Technology Center, The University of Kansas, Lawrence, KS 66045, USA
  • show less
    DOI: 10.3788/COL201917.122402 Cite this Article Set citation alerts
    Shima Fardad, Eric Schweisberger, Alessandro Salandrino. Parametric resonances in nonlinear plasmonics [Invited][J]. Chinese Optics Letters, 2019, 17(12): 122402 Copy Citation Text show less

    Abstract

    In the context of nonlinear plasmonics, we review the recently introduced concept of plasmonic parametric resonance (PPR) and discuss potential applications of such phenomena. PPR arises from the temporal modulation of one or more of the parameters governing the dynamics of a plasmonic system and can lead to the amplification of high-order sub-radiant plasmonic modes. The theory of PPR is reviewed, possible schemes of implementation are proposed, and applications in optical limiting are discussed.

    Localized surface plasmon (LSP) resonances are a salient feature of the optical and electronic response of metallic nanoparticles. These modes can be externally excited by photonic or electronic scattering, leading to strongly localized electric fields in proximity of the nanoparticle’s surfaces. An enhanced optical response is obtained when LSPs are resonantly excited by an incident field at the characteristic frequency of the dipolar eigenmode. The dynamics of these non-propagating coherent electronic oscillations show a strong dependence on the geometry of the particles, their composition, and the dielectric environment in which they are located. Yet, some features are common to all plasmonic configurations. In particular, LSP resonances in nanoparticles of any shape form an infinite discrete set of modes. In the simple case of particles of spherical shape with permittivity ε1(ω) surrounded by a medium with permittivity ε2, for a resonance of any order n, there are 2n+1 degenerate states with complex frequency ωn, such that ε1(ωn)=(1+n)ε2/n. As shown in Fig. 1, for n1, the eigenmodes tend to occur for ε1(ωn1)ε2. The increased modal density for ε1ε2 is a general feature of all plasmonic structures. Such increased modal density in plasmonic particles of different shapes for ε1ε2 arises because the spatial oscillation of the fields along the particle’s surface occurs with a negligible local wavelength compared with the local radius of curvature of the metal–dielectric interface[1]. Accessing such spectrally dense sets of tightly bound resonant modes would greatly enhance nonlinear light–matter interactions at the nanoscale and foster new developments in nonlinear plasmonics[2].

    Spectral distribution of the LSP resonances in a plasmonic sphere.

    Figure 1.Spectral distribution of the LSP resonances in a plasmonic sphere.

    The efficiency with which LSP resonances can be excited by an external incident field depends upon the spatial and spectral overlap between the excitation field and the specific plasmonic mode. For deeply subwavelength plasmonic particles, only the lowest-order mode of an electric dipolar nature is efficiently coupled to radiation states. In order for a particular mode to be efficiently excited, it is necessary for the incident field to be able to induce the appropriate polarization charge distribution. Such polarization charge distributions are illustrated in Fig. 2 for the first few modes of a sphere. From Fig. 2, it is apparent that to induce modes of order n>1 the incident field would have to display strong spatial variations over deeply subwavelength regions. That is why the higher-order eigenmodes tend to be sub-radiant, and by reciprocity, they are nearly decoupled from free-space propagating fields. Therefore, exciting and detecting such higher-order modes requires either near-field scattering techniques[3] or the use of active media to promote surface plasmon amplification by stimulated emission of radiation (SPASER)[4].

    Polarization charge density of the first few resonant modes of a plasmonic sphere.

    Figure 2.Polarization charge density of the first few resonant modes of a plasmonic sphere.

    A different mechanism to drive high-order LSP modes with a spatially uniform optical field relies on the recently introduced concept of plasmonic parametric resonance (PPR)[5]. In contrast with conventional localized plasmonic resonances, in which modes are excited directly by an external field of frequency and spatial profile matching those of a given mode of the plasmonic particle, PPR is a form of amplification in which a pump field transfers energy to a mode in an indirect way. In PPR, in fact, the modes of a plasmonic structure are amplified by means of a temporal modulation of the background permittivity caused by an appropriate pump field. Such permittivity variation translates into a modulation of the modal resonant frequency. Under specific pump conditions, amplification can occur. As shown in Ref. [5], among the unique characteristics of PPR is the possibility of accessing modes of arbitrarily high order with a simple spatially uniform pump, provided that such pump exceeds a certain intensity threshold—a characteristic of all parametric resonances.

    In very general terms, a parametric resonance[6] occurs when one or more of the parameters controlling the evolution of a dynamical system undergo a temporal modulation of appropriate amplitude and frequency. When such conditions are met, the amplitude of the parametrically resonant mode increases exponentially with time as long as the parametric modulation continues. In formal terms, the temporal evolution of a representative dynamical variable X(t) of a system with resonant frequency ω0 and damping γ under the action of a stimulus F(t) is given by Eq. (1) in the case of direct excitation, and by Eq. (2) for parametric excitation: d2X(t)dt2+γdX(t)dt+ω02X(t)=F(t),d2X(t)dt2+γdX(t)dt+{ω02+2ω0δω[F(t)]}X(t)=0.

    In the parametric Eq. (2), the external stimulus F(t) acts indirectly on the system by modifying the instantaneous resonant frequency by the amount δω[F(t)].

    A simple system described by Eq. (1) could be a simple harmonic oscillator, like a mass m attached to a spring with elastic constant k=mω02. In this example, X(t) represents the displacement of the mass from the equilibrium position. The restoring force provided by the spring can be described in terms of a potential energy U=kX2/2. As illustrated in Fig. 3(a), the free evolution of the system is a damped harmonic oscillation. Now, let us assume that the elastic constant of the spring is changed with time through some external mechanism. This new situation can be effectively described by Eq. (2). In the presence of such parametric modulation, the potential landscape varies with time, as shown in Figs. 3(b)3(d), for the case of a sinusoidal modulation around the unperturbed potential. Under the appropriate conditions, the energy that is externally provided to change the potential of the system can be transferred to the harmonic oscillator to compensate or even overcome the dissipation mechanisms. If a certain parametric-modulation threshold is exceeded, the oscillation amplitude grows in time with an exponential envelope, as indicated in Fig. 3(d). In PPR, these concepts are extended and applied to heavily multimode optical resonators such as plasmonic particles.

    (a) Time evolution of the position X of a harmonic oscillator in a parabolic potential U. (b)–(d) Time evolution of a parametrically driven oscillator with time-varying potential. (b) Below threshold, the oscillation decays. (c) At the threshold of parametric regeneration, the dissipations are exactly compensated. (d) Above threshold, the parametric gain causes the oscillations to grow exponentially.

    Figure 3.(a) Time evolution of the position X of a harmonic oscillator in a parabolic potential U. (b)–(d) Time evolution of a parametrically driven oscillator with time-varying potential. (b) Below threshold, the oscillation decays. (c) At the threshold of parametric regeneration, the dissipations are exactly compensated. (d) Above threshold, the parametric gain causes the oscillations to grow exponentially.

    In order to illustrate in general terms the principle of the operation of PPR in a plasmonic nanoparticle of arbitrary shape, the first step is to identify a dynamical variable obeying an evolution equation similar to Eq. (2). The first complication that arises is due to the fact that all the relevant physical quantities involved in such an electromagnetic problem are fields of some sort (like electric potential, electric field, magnetic field, polarization density, and current density), rather than simple kinematic variables like the position X(t) considered in the previous harmonic oscillator example. This issue is addressed by performing a modal decomposition of the field of interest in terms of an appropriate complete and orthogonal set of basis functions. By doing so, the field is expressed as a sum of vector-field basis functions, each weighted with a scalar time-dependent modal amplitude determined by the initial conditions and by the subsequent evolution. As shown in the following, such modal amplitudes can be used as dynamical variables to cast the PPR problem in the form of Eq. (2). Each modal amplitude will evolve in time, according to an equation of the same form as Eq. (1), with an appropriate natural frequency ωn depending on the specific mode, the permittivity ε1 of the particle, and the permittivity ε2 of the surrounding medium. This approach immediately suggests that modulating one of these properties, say ε2, will cast the problem in the desired form of Eq. (2). In the following, for the purpose of illustration of PPR, we will consider a system that is amenable to a close-form solution: a subwavelength plasmonic sphere in a homogeneous dielectric background medium.

    We consider a sphere of radius R and relative permittivity ε1 (medium 1) embedded in a uniform dielectric medium ε2 (medium 2). The radius R is assumed to be much smaller than the free-space wavelength associated with any of the plasmonic eigenmodes of interest, so that a quasi-static approach is applicable for determining the spatial distribution of the electromagnetic field. The dispersion of ε2 is neglected. Medium 1 is assumed to follow a Drude-like frequency-domain dispersion ε1(ω)=εωpl2/(ω2+iωγ), with plasma frequency ωpl, collision frequency γ, and a non-dispersive term accounting for high-frequency spectral features ε. The dispersive term in the ε1(ω) expression is associated with the equation of motion for the free-carrier polarization density P1(r,t) within medium 1: 2P1(r,t)t2+γP1(r,t)t=ε0ωpl2E1(r,t),where E1(r,t) is the electric field within medium 1. Medium 2 is assumed to be endowed with second-order nonlinearity with a dominant term χzzz(2). Under such hypotheses, the polarization density P2(r,t) in medium 2 can be expressed in terms of the total local field E2(r,t) as follows: P2(r,t)=ε0(ε21)E2(r,t)+P2NL(r,t).

    In Eq. (4), P2NL(r,t)=ε0χ(2)·E2(r,t):E2(r,t) is the nonlinear polarization density due to the quadratic nonlinearity of medium 2. The total electric field E2(r,t) in medium 2 is the sum of all the fields due to the plasmonic modes of the particle and a spatially uniform incident field EP(t), henceforth referred to as “pump”.

    In the quasi-static approximation, the polarization density in medium 1 can be expanded in terms of spherical harmonics Yn,m(e/o)(θ,ϕ), defined and normalized as in Ref. [5]: P1(t)=n,m{rnRn1[Pn,m(e)(t)Yn,m(e)(θ,ϕ)+Pn,m(o)(t)Yn,m(o)(θ,ϕ)]}.

    Performing similar expansions for all field quantities in terms of spherical harmonics and applying the boundary conditions at the particle’s interface yields the following evolution equation for the polarization density amplitude associated with any of the electromagnetic angular momentum eigenmodes of the sphere: d2Pn,m(e,o)(t)dt2+γdPn,m(e,o)(t)dt+ωn2Pn,m(e,o)(t)=ωn2Sn,m(e,o)(t)n.

    In Eq. (6), the parameter ωn is the resonant frequency of the eigenmodes of order n in the absence of nonlinear interactions and is given by ωn=nωpl2nε+(n+1)ε2.

    The term Sn,m(e,o)(t) on the right-hand side of Eq. (6) is the projection on the spherical harmonic Yn,m(e,o)(θ,ϕ) of the nonlinear polarization density P2NL evaluated over the surface of the sphere: Sn,m(e,o)(t)=r=RYn,m(e,o)(θ,ϕ)P2NL(R,θ,ϕ,t)·r^sin(θ)dθdϕ.

    Through Eq. (8), various eigenmodes are nonlinearly coupled to one another and to the pump field. The symmetry group of medium 2 and the spatial profile of the pump determine which specific three-wave mixing products contribute to the dynamics of a given eigenmode.

    Let us consider the dynamics of the azimuthally uniform (m=0) resonant mode of order n in the presence of a spatially uniform time-harmonic z-polarized pump EP(t). In this case, the evolution of Eq. (6) assumes the following form: d2Pn,0(e)(t)dt2+γdPn,0(e)(t)dt+[ωn2α1EP(t)]Pn,0(e)(t)=α2[Pn,0(e)(t)]2.

    The expressions of the nonlinear interaction coefficients α1 and α2 are given by α1=4πχzzzn3ωn4ωpl2Gn,0(e,e);α2=χzzznε0ωn6ωpl4Fn,0(e,e,e),Fn,0(e,e,e)=02π0π{cosθz[Rn+2rn+1Yn,0(e)(θ,ϕ)]R×z[Rn+2rn+1Yn,0(e)(θ,ϕ)]RYn,0(e)(θ,ϕ)}sinθdθdϕ,Gn,0(e,e)=02π0π{cosθz[rY1,0(e)(θ,ϕ)]R×z[Rn+2rn+1Yn,0(e)(θ,ϕ)]RYn,0(e)(θ,ϕ)}sinθdθdϕ.

    The PPR threshold is minimized[5] if the pump field oscillates at the second-harmonic frequency of the mode of interest. We start, therefore, by considering a spatially uniform monochromatic pump field of the form EP(t)=Apsin(2ωnt). In solving Eq. (9), we notice that so long as the condition |Pn,m(e,o)(t)|α1Ap/α2 holds, which is the case at the initial stages of the parametric interaction, a solution can be easily obtained in terms of Mathieu functions[7,8]. More intuitive though is the following slowly varying envelope approximate solution: Pn,m(e,o)(t)=p(t)cos[ωntθ(t)]eγ2t,p(t)=p0cosh(α1Ap2ωnt);θ(t)=arccot[exp(α1Ap2ωnt)],where p0 is the initial modal amplitude. From Eq. (11), for p(t), it is clear that the system enters the PPR regime provided that the pump electric-field amplitude Ap exceeds the threshold value APPR=2γωn/α1.

    It is worth pointing out a unique property of plasmonic parametric gain that emerges from the analysis above: a plasmonic mode of any order (m,n) can undergo PPR and be amplified by a spatially uniform modulation of the background permittivity, provided that the corresponding threshold is exceeded. This is in contrast with conventional LSP resonances, which, for a mode of order (m,n), requires a driving field with a matching spatial profile—a condition almost impossible to realize in practice for high-order plasmonic modes of deeply subwavelength particles. For these reasons, PPR is uniquely suitable to access plasmonic resonances of arbitrarily high order in deeply subwavelength structures.

    The unique characteristics of PPR lend themselves to interesting applications in optical limiting. Recently, a new class of nonlinear absorbers termed plasmonic parametric absorbers (PPAs) has been proposed[9]. The key insight informing the PPA idea is that in the PPR process the pump field experiences an absorption rate that strongly depends on the intensity of the pump itself, creating two distinct regimes: one of weak absorption under low intensity illumination and one of strong absorption when the threshold of parametric resonance is reached or exceeded. Such a threshold condition separates distinct dynamics, so that Pn,m(e,o)(t) decreases exponentially for Ap<APPR and increases exponentially for Ap>APPR. Such contrasting modal dynamics are reflected in the distinct absorption regimes that the pump is subjected to. As shown in Ref. [9], the power parametrically transferred from the pump to the resonant mode is given by Wabs(t)=nR3α1App0232ε0ωpl2[2ωn+γsinh(α1Apt2ωn)]eγtnR3α1App02γ64ε0ωpl2exp[(ApAPPR)α1t2ωn].

    Equation (12) highlights the fundamental trait of PPA, which is in stark contrast with linear absorption: in PPA, absorption is vanishingly small for incident fields below the PPR threshold and increases exponentially under high-intensity conditions.

    Clearly, a saturation of the exponential behavior is expected, because, if nothing else, the absorbed power of Eq. (12) cannot exceed the finite power carried by the pump. In reality, a different mechanism limits Eq. (12) before pump depletion occurs. Such a mechanism is the resonance detuning due to additional three-wave mixing processes in Eq. (9) that we have neglected so far. As |Pn,m(e,o)(t)|α1Ap/α2, Eq. (9) can only be integrated numerically. Nevertheless, the following asymptotic expressions as t for Pn,m(e,o)(t) hold for Ap>APPR: Pn,m(e,o)(t)α2Q122ωn2+Q1cos(ωnt+θ1)+α2Q126ωn2cos(2ωnt+2θ1),θ1=12arccos(APPRAp);Q1=ωnα26γωn5(ApAPPR)21.

    Within the range of validity of Eq. (13), the exponentially growing oscillations of the polarization density amplitude level off as t after a sequence of relaxation oscillations. In Fig. 4, the numerical solution of Eq. (9) (indicated in blue) is compared with the predictions of the asymptotic model in Eq. (13), shown in orange.

    Polarization density amplitude term P11,0(e) of a silver sphere immersed in an MNA background medium. In order to better highlight the relaxation oscillations occurring in the system, we show a case in which the PPR threshold is exceeded by a large margin (Ap=20APPR). The dashed lines show the oscillation limits predicted by the asymptotic Eq. (12).

    Figure 4.Polarization density amplitude term P11,0(e) of a silver sphere immersed in an MNA background medium. In order to better highlight the relaxation oscillations occurring in the system, we show a case in which the PPR threshold is exceeded by a large margin (Ap=20APPR). The dashed lines show the oscillation limits predicted by the asymptotic Eq. (12).

    Based on Eq. (13), the average power W¯P transferred from the pump to the plasmonic mode asymptotically approaches the value W¯abs(t)=320nR3γε0ωpl2ωn3α22(ApAPPR)21.

    Using the steady-state asymptotic estimate of Eq. (14) of the absorbed power, it is possible to obtain the PPR contribution to the particle absorption cross-section (in addition to the linear portion): σNL=3nR340ε0ε2ωn3ωpl2α12α22IPPRIp(1IPPRIp),Ip>IPPR.

    In Eq. (15), Ip=Ap2/(2η) is the incident intensity, and IPPR=APPR2/(2η) is the PPR intensity threshold, where η is the intrinsic impedance of the background medium. If particles similar to the one described so far are dispersed with density N in the background medium, the nonlinear pump attenuation coefficient of the composite follows from Eq. (15) as αNL=NσNL.

    Figure 5 shows how the normalized absorption cross-section of a silver sphere of radius R=100nm in a 2-methyl-4-nitroanline (MNA)[10,11] host is affected by various modes undergoing PPR. The absorption cross-section is plotted against the normalized pump intensity, and, for each of the PPR modes considered in Fig. 5, the incident pump field is at twice the value of the corresponding resonant frequency given by Eq. (7). For the case at hand, all the possible PPR resonant wavelengths λn fall in the range λ<λnλ1, where λ=563nm and λ1=448nm. As evident from Fig. 5, as soon as a pump field of frequency 2ωn exceeds the intensity threshold IPPR, the particle’s absorption cross-section increases dramatically due to the contribution of the mode of eigenfrequency ωn undergoing PPR. This phenomenon is a form of reverse saturable absorption and could have interesting applications in optical limiting devices[12], especially given the design versatility of metallic nanoparticles for targeting different spectral regions.

    Absorption cross-section of the plasmonic particle normalized to the geometrical cross-section as a function of the incident intensity.

    Figure 5.Absorption cross-section of the plasmonic particle normalized to the geometrical cross-section as a function of the incident intensity.

    For the purpose of illustration, in Fig. 6, we apply the analysis and the models described thus far to the practically relevant case of a pulsed pump. The pulse considered here is a 30 ps pulse of average power 0.5 W, focused to an area of 25μm2, on parametric resonance with the n=11, m=0 mode (λ11=460nm) of a silver particle of radius 100 nm embedded in an MNA background medium. The orange curve in Fig. 6 shows the total instantaneous power of the pump field. The blue curve shows the instantaneous absorbed power caused by the PPR process. The vertical dashed lines show the time interval in which the pump field exceeds the PPR threshold. A similar behavior is observed both in Figs. 4 and 6 at the onset of PPR, where the modal polarization (and the corresponding absorption) builds up exponentially to slightly surpass the steady-state value and then relaxes to such a value through a series of oscillations (only one is discernible in Fig. 6). Figure 6 confirms the dramatic increase in absorption that PPAs exhibit under high-intensity illumination.

    Orange curve shows the instantaneous pump power Winc. The blue curve shows the instantaneous pump power Wabs absorbed by the particle via PPR of the n=11, m=0 mode. The dashed vertical lines indicate the times at which the pump intensity is equal to the PPR threshold.

    Figure 6.Orange curve shows the instantaneous pump power Winc. The blue curve shows the instantaneous pump power Wabs absorbed by the particle via PPR of the n=11, m=0 mode. The dashed vertical lines indicate the times at which the pump intensity is equal to the PPR threshold.

    In conclusion, we have illustrated the principles of the operation of PPR. Unlike conventional LSP resonances, all of the plasmonic modes of a nanostructure, including the strongly sub-radiant ones, can be resonantly excited by spatially uniform optical pumping, provided that the corresponding threshold is exceeded. Accessing such a high density of strongly localized states holds promise for enhancing nonlinear light–matter interaction at the nanoscale for the development of nonlinear optical metamaterials and for optical limiting applications. In the context of PPR, we have discussed the closely related theory of PPAs. PPAs exhibit a reverse saturable absorption behavior whereby an incident field that is parametrically resonant with one or more of the modes of a plasmonic particle experiences a strongly enhanced absorption whenever its intensity exceeds the relevant PPR threshold. Such effect makes PPAs very promising candidates for optical limiting applications, in addition to being of fundamental interest in the emerging field of nonlinear plasmonics.

    References

    Shima Fardad, Eric Schweisberger, Alessandro Salandrino. Parametric resonances in nonlinear plasmonics [Invited][J]. Chinese Optics Letters, 2019, 17(12): 122402
    Download Citation