Open Access
9 December 2023 Giant photoinduced reflectivity modulation of nonlocal resonances in silicon metasurfaces
Andrea Tognazzi, Paolo Franceschini, Olga Sergaeva, Luca Carletti, Ivano Alessandri, Giovanni Finco, Osamu Takayama, Radu Malureanu, Andrei V. Lavrinenko, Alfonso C. Cino, Domenico de Ceglia, Costantino De Angelis
Author Affiliations +
Abstract

Metasurfaces offer a unique playground to tailor the electromagnetic field at subwavelength scale to control polarization, wavefront, and nonlinear processes. Tunability of the optical response of these structures is challenging due to the nanoscale size of their constitutive elements. A long-sought solution to achieve tunability at the nanoscale is all-optical modulation by exploiting the ultrafast nonlinear response of materials. However, the nonlinear response of materials is inherently very weak, and, therefore, requires optical excitations with large values of fluence. We show that by properly tuning the equilibrium optical response of a nonlocal metasurface, it is possible to achieve sizable variation of the photoinduced out-of-equilibrium optical response on the picosecond timescale employing fluences smaller than 250 μJ / cm2, which is 1 order of magnitude lower than previous studies with comparable reflectivity variations in silicon platforms. Our results pave the way to fast devices with large modulation amplitude.

1.

Introduction

In the last few decades, the advent of nanofabrication technologies has boosted the interest in optical devices working in the visible and near-infrared (NIR) region. Metasurfaces are two-dimensional arrangements of nanostructures (or subwavelength elements), called meta-atoms, which allow one to tailor the electromagnetic field at the nanoscale, thus surpassing the capabilities of bulk materials.1,2 The possibility of introducing abrupt changes in phase and amplitude is of paramount importance to focusing,3,4 polarization control,5 nonlinear frequency conversion,6 and wavefront manipulation.7

Providing methods to dynamically control a nanoscale device is mandatory to produce flexible and effective tools to manipulate light. The compact nature of such devices allows one to reach faster modulation speed employing approaches that are traditionally slower tuning mechanisms, such as heating and mechanical strain.811 Liquid crystals have also been employed to rapidly change the optical response of metasurfaces to achieve multifunctional behavior12 and phase tuning.13 Phase-change materials have been proposed as suitable platforms to obtain large all-optical modulation featuring vast versatility of the activation mechanisms.1418 So far, optical excitation provides the fastest modulation source allowing ultrafast dynamic control of nanophotonic devices.1921 Originally, the strong optical nonlinearities in plasmonic meta-atoms were first investigated as a route toward ultrafast nanophotonics.22 Subsequently, their ohmic losses at optical frequencies and the poor compatibility with large-scale CMOS integration opened the way for the advent of all-dielectric structures, which have already been demonstrated to achieve ultrafast control of polarization, amplitude, and wavefront.21,2326 In all-optically modulated devices, the ultrafast two-photon absorption (TPA) and free carrier (FC) generation dynamics come along with thermal effects related to lattice heating, which take place at the nanosecond timescale.27 Recently, it was demonstrated that by properly engineering the metasurface, it is possible to achieve a full recovery of the optical response on the picosecond timescale.23,25 However, the entity of such modulation is often small, usually below a few percent, and requires large fluences, i.e., energy densities.23,25,28 Therefore, the detection of such tiny variations is laborious and requires bulky and expensive experimental setups, annihilating the miniaturization advantages of optical devices.

Here we design a high-contrast nonlocal metasurface (NLM) based on silicon (Si) meta-atoms lying on silica (SiO2) pedestals that exhibits a guided mode resonance (GMR) in the NIR spectral region, and we show giant all-optical relative reflectance variation (up to 60%) on the picosecond timescale. We use time- and wavelength-resolved spectroscopy to investigate the out-of-equilibrium temporal dynamics of the nonlocal NIR resonance following FC injection. In particular, we perform pump–probe experiments to address the carrier and lattice dynamical processes and their role on the time evolution of the spectral signature of the nonlocal resonance. We show a transient photoinduced modulation of the optical properties of NLM (i.e., resonant gratings), which is 1 order of magnitude larger than the unpatterned platform (under the same excitation conditions). The analysis of the experimental data, combining a coupled rate equations model (CREM) and rigorous coupled-wave analysis (RCWA), allows one to disentangle the role of the photonic structure and the modulation of the optical properties (due to FCs and lattice heating effects) and reveals a decrease of the carrier recombination time with respect to bulk material. Our results have direct impact in the design of next-generation ultrafast all-optical modulators and switches with reduced energy consumption, since we demonstrate an increase of 1 order of magnitude in the relative reflectance variation, with fluences below 250  μJ/cm2, with respect to unstructured silicon.

2.

Results and Discussion

We employ an NLM, which we previously demonstrated29 to feature sharp spectral resonance with large amplitude in the third operating window of optical communications, to unveil the pivotal role of the photonic structure to achieve large amplitude modulation with smaller fluence compared to bulk materials. The high-index-contrast NLM under analysis, shown in Fig. 1(a), is based on a Si grating (width w=393  nm, height d=500  nm, and periodicity P=820  nm), supported by silica (SiO2) pedestals (height hp=160  nm and width wp=73  nm) to achieve high-index contrast with the environment, i.e., between air and the lower-lying SiO2 layer. We engrave the Si grating within a silicon-on-insulator substrate combining deep ultraviolet (DUV) lithography and reactive ion etching. The total thickness of the SiO2 layer is h=1160  nm (see Appendix A).

Fig. 1

(a) Sketch of the experiment working principle: a pump pulse (blue) excites the sample and changes the optical properties of silicon (brown). A delayed low-intensity probe pulse (yellow) monitors the reflectivity changes. The image is a false color scanning electron micrograph of the NLM. Dimensions: w=393  nm, d=500  nm, P=820  nm, wp=73  nm, hp=160  nm, and h=1160  nm. (b) Experimental (circles) and calculated (RCWA, red solid line) reflectance spectra (Req) at equilibrium. The inset shows the TE mode spatial distribution in the yz plane (orthogonal to the grating) at 1375-nm wavelength (black arrow).

AP_5_6_066006_f001.png

The equilibrium optical properties of the NLM are shown in Fig. 1(b). Here we display the experimental (markers) reflectance spectrum at equilibrium (Req) as a function of the wavelength (λ) measured by a broadband NIR probe light linearly polarized along the direction parallel to the grating bars, x axis in the present work [see Fig. 1(a)], which corresponds to a transverse electric (TE) polarization (see Appendix B).

The sharp spectral feature at around 1380 nm, arising on top of a background due to Fabry–Perot modes, is ascribed to a GMR (quality factor QF170; see Fig. S3 in the Supplementary Material), which is mediated by a leaky mode propagating along the grating periodicity direction (y axis).29 The leaky nature of the GMR can be deduced by looking at the spatial distribution of the mode in Fig. 1(b) (see inset). To model the NLM equilibrium properties, we calculate the reflectance through RCWA as the normalized intensity of the reflected radiation along the zeroth order (see Appendix C for more details). The resulting calculated reflectance spectrum RRCWAeq(λ) [red solid line in Fig. 1(b)] shows a fairly good agreement with the experimental data, reproducing the line shape of the GMR, manifesting as a Fano resonance. This originates from the coupling of the discrete spectrum of the leaky GMR to the continuum of modes of the incident light.30 Discrepancy in the spectra may originate from the dispersion of the geometrical parameters (due to fabrication tolerances) or minimum differences of the pedestal profile within the sample area detected by the probe beam (see Ref. [29] for more details). Nevertheless, the model reveals the collective (i.e., nonlocal) nature of the resonance and the eigenmode analysis allows one to correctly locate its spectral position at λΓeq1375  nm.

We perform a time- and wavelength-resolved pump–probe spectroscopy experiment to monitor the time evolution of the GMR following photoexcitation. In general, in bulk Si, the carrier relaxation dynamics evolves according to the following well-established picture.31,32 Upon photoexcitation by a pump pulse with photon energy (Ep=2πc/λp, where λp is the pump wavelength), which can be larger (or smaller) than the bandgap Eg, an FC density N is photogenerated within the bands (electrons in the conduction band and holes in the valence band) by linear (or nonlinear) absorption processes. The initial coherently excited carrier states undergo dephasing processes via carrier–carrier and carrier–phonon scattering mechanisms. This coherent regime (30  fs for Si32) is followed by a nonthermal regime (on a timescale of 100 to 300 fs33,34) during which a carrier temperature is established via collisions. This allows one to describe the carrier population in terms of Fermi–Dirac distribution, with the carrier and lattice temperature depending on Ep. Finally, as time evolves, hot carriers cool down via energy exchange with the lattice (accompanied by optical phonon scattering).

In our experiment, as depicted in Fig. 1(a), a short-pump laser pulse centered at λp=410  nm (i.e., in the absorption region of Si) first induces a variation of the optical properties of the NLM via FC generation. Then a broadband probe laser pulse, tuned in the GMR spectral region (1340 to 1400 nm range), impinges at delay time Δt after the pump, thus monitoring the out-of-equilibrium state of the system (see Appendix D for more details). The probe beam spot size [beam waist radius w0(probe)=(75±5)  μm], smaller than the pump one [w0(pump)=(120±10)  μm], has been chosen to ensure the illumination of a minimum number of periods (NP) to avoid the beam size effects (based on Ref. [35], for the NLM under analysis NP=QF/π54).

We measure the relative reflectance variation (ΔR/R) defined as the difference (ΔR=RoutReq) between the reflected light with and without the pump (Rout and Req, respectively) normalized to the reflected light spectrum without the pump.

Measured ΔR/R spectra at various delay times (Δt), in the timescale below 100 ps, are reported in Fig. 2(a) (see also Fig. S2 in the Supplementary Material for additional details on the the corresponding Rout spectra). Here we use an incident pump fluence Finc=(180±30)  μJ/cm2 to ensure that we photoexcite the NLM in the linear regime. We underline that this value is much lower than in other studies with comparable, or even smaller, variations of the reflectivity in Si-based platforms.23,25,28,36 The ΔR/R(λ,Δt) signal is mainly characterized by two long-living components: a negative dip at λ1370  nm and a positive peak at λ1375  nm; these correspond to the spectral regions below and above the minimum of the Fano resonance [see Fig. 1(b)]. The additional change of the sign in the out-of-equilibrium spectra at λ>1380  nm corresponds to the spectral region above the Fano resonance peak. As far as the amplitude of the out-of-equilibrium signal is concerned, we observe a giant relative reflectance variation of 50% (Req=0.18 and Rout=0.27) at Δt=3  ps for λ1373  nm.

Fig. 2

(a) Experimental ΔR/R spectra at different time delays (denoted by the labels on the left side). For clarity, the curves are vertically shifted (labels on the right side). (b) Simulated ΔR/R spectra at the same time delays as in (a) employing CREM + RCWA model. For clarity, the curves are vertically shifted. (c) Real (Re, dashed line) and imaginary (Imag, solid line) parts of the individual components contributing to Δε calculated from the model analysis: TPA (ΔεTPA), FC generation (ΔεFC), and lattice heating (ΔεL). The curves are horizontally shifted by 0.64 ps to visualize negative delay times on a logarithmic scale. (d) Temporal evolution of the shift: ΔλΓ from modal analysis (blue solid line) and Δλr from differential fit (yelllow markers). The data (both solid line and markers) are horizontally shifted by 0.64 ps to visualize negative delay times on a logarithmic scale.

AP_5_6_066006_f002.png

In order to gain further insights into the interplay between the nonequilibrium photoinjected FCs and the geometry-controlled GMR, we simulate the out-of-equilibrium response of NLM with a theoretical approach combining the CREM, accounting for FC concentration and lattice temperature dynamics, and RCWA, describing light–matter interaction in the NLM. The differential spectra shown in Fig. 2(b) are calculated as ΔR/R(λ,Δt)=RRCWAout(λ,Δt)/RRCWAeq(λ)1. The time-dependent out-of-equilibrium reflectivity spectrum RRCWAout(λ,Δt) is obtained from a perturbed Si dielectric function (εout=εSi+Δε, with εSi being the equilibrium Si dielectric function and Δε the photoinduced variation) to account for the effect of photogenerated carriers. In particular, we assume that the total variation Δε includes both linear and nonlinear contributions, where the former is ascribed to mechanisms involving the FCs and the lattice (L), and the latter is related to the TPA; therefore, Δε=ΔεFC+ΔεL+ΔεTPA. For the sake of completeness, we mention that, in addition to the previous effects, also state filling (SF) and bandgap renormalization (BGR) mechanisms37,38 may represent a linear-response contribution to Δε. However, taking into account that the considered probing region is located in a spectrum far from the bandgap, i.e., the spectral region where SF and BGR mainly dominate, we neglect their contribution in the calculation of the ΔR/R spectra.

Regarding FCs, their contribution is mainly described in terms of the Drude model; therefore, we assume the change of the dielectric constant of the material due to FC generation in the form37,39

Eq. (1)

ΔεFC(λ,Δt)=N(Δt)e2·(1iλ2πcτd)·(4π2c2meffε0λ2+1τd2)1,
where e is the electron charge, c is the velocity of light in vacuum, τd is the Drude damping time, meff is the effective mass of electrons and holes, and ε0 is the vacuum permittivity. As mentioned, the relaxation of FCs within the bands, occurring via a first-order trap-assisted process (with characteristic time τrFC) or via second-order bimolecular recombination (at rate γ), is accompanied by lattice heating; its contribution to the optical modulation occurs via the thermo-optic effect, which has been modeled as follows:32,40

Eq. (2)

ΔεL(λ,Δt)=2·εSi(λ)·η1(λ)·T(Δt),
where T is the lattice temperature (recovering the equilibrium value with a characteristic timescale τrL) and η1 is the Si thermo-optic coefficient. Third, the nonlinear absorption (due to TPA) has been implemented as follows:25

Eq. (3)

ΔεTPA(Δt)=R[εSi(λp)]·βTPAI(Δt)λp2π,
where βTPA is the Si nonlinear absorption coefficient and I the impinging pump intensity (see Appendix C). The temporal dependence in Eqs. (1)–(3) is a consequence of the temporal dynamics of the three quantities N(Δt), T(Δt), and I(Δt), described by the two-coupled rate equations in Eq. (4) (see Appendix C for more details).

At this stage, it is useful to underline that the considered model, treating the FC density N, is not suitable for describing any coherence effect of the light–matter interaction under analysis. The chosen design of the theoretical model is based on the assumption that the temporal resolution of our setup (64  fs) is not capable of revealing the signature of coherence, since it is destroyed on a timescale of a few tens of femtoseconds in the case of solid-state platforms.31,32,4144

Figure 2(b) shows the theoretical out-of-equilibrium spectra calculated via CREM + RCWA approach by assuming the parameters reported in Table 1. The theoretical model is capable of quantitatively reproducing the experimental out-of-equilibrium spectra in the whole temporal window under analysis by employing one fixed fitting parameter: the FC relaxation time τrFC (band edge carrier lifetime through single-carrier decay channels). Our analysis provides τrFC150  ps, a value much shorter than that in bulk Si (>50  ns),50 but comparable to that measured in other Si-based photonic structures.51

Table 1

Summary of the material parameter values adopted in this work for CREM + RCWA approach.

ParameterValueUnitsReference
Cp1.66×106J  K1m325
βTPA1.5×106mW125
εSi(5.3410+i0.24127)2 at 410 nm45
εSi, εsub12.2 at 1370 nm45
εSiO22.09 at 1370 nm46
η11.888×104 at 1370 nmK140
meff0.153me47
τd10fs48
γ3×1015m3s149

In order to better understand the role of FCs and lattice to the total Δε, in Fig. 2(c), we show the temporal dynamics of ΔεFC, ΔεL, and ΔεTPA calculated by the CREM. As expected, the contribution due to TPA (black solid curve) is present only at very short time delays (Δt<1  ps). The largest contribution to Δε in the probed time window is due to FC generation (red dashed and solid curves for real and imaginary parts, respectively), since the typical timescale is on the order of tens of picoseconds and lattice heating effects (blue solid and dashed curves for real and imaginary parts, respectively) manifest on longer timescales (Δt>100  ps). ΔεL contribution is negligible in the probed time window due to the low (real-valued) thermo-optic coefficient.

The numerical results reported in Fig. 2(c) offer important insights into the physics of the NLM regarding the photoinduced dynamics of the GMR. Indeed, starting from the calculated Δε(λ,Δt), we can better describe this performing time-dependent modal analysis to retrieve the eigenvalue λΓout(Δt) of the GMR at each delay time. The retrieved time-dependent eigenvalue variation ΔλΓ(Δt)=λΓout(Δt)λΓeq, reported in Fig. 2(d) (blue solid line), suggests that the GMR experiences a blueshift after photoexcitation, reaching a maximum variation of the eigenwavelength |ΔλΓmax|2  nm and then recovering to equilibrium on a timescale of 100 ps. Further, we adopt a differential fit procedure26 based on a Fano line shape52 to model the out-of-equilibrium spectra in Fig. 2(a). The experimental data can be reproduced assuming a resonance blueshift, which represents the FC-induced variation of the dielectric constant31 (see Supplementary Material for more details). The retrieved blueshift temporal dynamics Δλr(Δt), displayed in Fig. 2(d) (yellow markers), overlaps well with the ΔλΓ one.

Summarizing, the results in Fig. 2 highlight a nonnegligible difference regarding the order of magnitude between the optical response of the structure (ΔR/R50% at maximum) and the modulation of the material properties (Δε/ε0.4% and Δλ/λ0.1%) following the moderate excitation of 180  μJ/cm2. Therefore, in order to better investigate the role of the photonic structure in the interplay between the optical response and material properties variation, we report excitation-dependent pump–probe measurements on the NLM, and we compare the results to those obtained from a multilayer thin-film-like platform, consisting in a uniform Si (500 nm)/SiO2 (1160 nm) bilayer29 deposited on the Si substrate. [We underline that the thickness of the Si (SiO2) layer in the thin-film-like sample is the same as d (h) in the NLM (see Appendix D).] In particular, we focus on the nonequilibrium FC-induced effect; therefore, we measure ΔR/R spectra of NLM at Δt=3  ps, where the FC contribution dominates Δε [see Fig. 2(c)], for several excitation fluence (Finc) levels in the 20 to 250-μJ/cm2 range. We note that the maximum amplitude of the detected transient spectra, shown in Fig. 3(a), is 1 order of magnitude larger than that measured on the bilayer structure (see Table 2 and Fig. S5 in the Supplementary Material for more details), under similar excitation conditions and reflectance value in the at-equilibrium state. Therefore, the results clearly demonstrate the key role of the photonic structure (giving rise to spectral resonances exhibiting sharp edges), which allows one to enhance the transient modulation of the optical response [up to ΔR/R=57% at λ=1373nm; thus ΔR=0.103, in Fig. 3(a)] with respect to bulk-form material.

Fig. 3

(a) Experimental ΔR/R spectra measured for increasing fluence values. The inset reports the resonance shift Δλr (retrieved from the measured ΔR/R spectra) as a function of the incident fluence, and the solid black line denotes a linear dependence. The red triangle denotes the probe wavelength value in which the data in (b) are taken. (b) Experimental minimum ΔR/R at λ=1367.5  nm as a function of the fluence (black markers). The black dashed line denotes a linear dependence of the ΔR/R signal upon the fluence.

AP_5_6_066006_f003.png

Table 2

Comparison between different metasurfacesa.

MaterialPlatform|ΔR/R| or |ΔT/T| (%)FluenceReference
SiBulk single crystal (001)716  mJ/cm253
SiHuygens340  mJ/cm254
SiBulk (001)0.042.4  mJ/cm232
Poly-SiNonlocal57250  μJ/cm2This work
Poly-SiNonlocal27130  μJ/cm2This work
Si, SiO2Bilayer2.6120  μJ/cm2This work
a-Si:HHuygens0.630  μJ/cm223
a-Si:HHuygens8800  μJ/cm225
GaAsHuygens60310  μJ/cm224
GaAs-AlAs-InAsQD embedded + Bragg mirror202 to 3  mJ/cm221
Au-ITOPlanar plasmonic crystal80NA22
Au-SiGrating602  mJ/cm255
V2O3Thin film148  mJ/cm256

aQD, quantum dots; a-Si:H, hydrogenated amorphous silicon.

Moreover, the results in Fig. 3(a) reveal the key role of the GMR to achieve giant photoinduced reflectivity variations. From a general point of view, the amplitude of the ΔR/R spectra increases with increasing fluence (i.e., injected FC density); this is also accompanied by a blueshift of the spectra (i.e., the peak and valley of the transient spectra move toward lower wavelengths values as Finc increases). The Δλr(Finc) trace, retrieved from the data by the differential-fit procedure, is shown in the inset of Fig. 3(a), and it is proportional to Finc (black solid line). This behavior is due to the fact that, at the considered delay time (3 ps), ΔεΔεFCN. However, by carefully looking at the ΔR/R signal at 1367.5 nm [see Fig. 3(b)], we note that an initial linear regime (black dashed line) is followed by a saturation for fluences above 200  μJ/cm2. The different behavior of the two quantities (Δλr and ΔR/R) can be interpreted on the basis of the Fano line shape of the GMR at equilibrium [Fig. 1(b)]. Indeed, the maximum achievable transient value is limited by the reflectance contrast (i.e., the difference between the minimum and maximum reflectance values) of the Fano resonance at equilibrium. Therefore, although the blueshift increases linearly with fluence (below the sample damage threshold), the transient signal saturates. Overall, these results testify to the possibility of achieving large amplitude modulations also in the case of structure with moderate quality factor Q (in addition to those with high Q51). Thanks to the spectral properties of the GMR exhibited by the proposed Si-based NLM, the observed sizable modulation is achieved with intensity values much lower (up to 1 order of magnitude) than in other works, as summarized in Table 2. Performances in similar experimental conditions were reported24 in the case of a GaAs-based metasurface. Despite being vastly employed for all-optical modulation due to efficient generation and recombination of FCs, GaAs does not exhibit such a widespread use in microelectronics and on-chip photonics as Si.24

3.

Conclusions

In this work, we proposed an NLM to achieve giant modulation of the reflectivity on an ultrafast timescale. We provided evidence that such variation cannot be induced in bulk materials or thin films employing fluences below 250  μm/cm2 and that the NLM plays a key role, i.e., it provided a sharp spectral variation, which ultimately leads to a giant reflectivity modulation mainly induced by FC generation. We described the experimental results employing a simple, yet effective, numerical approach based on CREM and RCWA simulations. We also unveiled the nonlinear dependence of the modulation depth as a function of the impinging fluence when the FCs are only responsible for the reflectivity variation. In addition to unveiling the physics of the ultrafast dynamics of photoexcited nonlocal Si metasurfaces, our results proved that it is possible to induce large modulation of the optical response inducing modest changes of the dielectric constant, which is of paramount importance in the design of ultrafast modulators and switches of the next generation, with reduced energy consumption and an ultrasmall footprint. Ultimately, the key ingredients for large optical modulation are the presence of a sharp variation in the linear response at equilibrium and its depth.

4.

Appendix A: Fabrication

We oxidize using a wet process (H2O-based) at 1100°C in a furnace (Tempress) a 500-μm-thick single-side polished 100 Si wafer, yielding 1.1  μm thermal Si oxide. Then we coat the wafer with an amorphous silicon (a-Si) layer by means of low-pressure chemical vapor deposition. Silane (SiH4) is used as a precursor. An additional dry oxidation and wet etching (vaporized HF) are performed to accurately determine the final thickness of the a-Si layer. The desired pattern is obtained using DUV lithography. The grating is suspended by performing a controlled and gentle etching of SiO2. Further details about the fabrication process are provided elsewhere.29,57

5.

Appendix B: Equilibrium Linear Spectroscopy

A home-built linear spectroscopy setup is employed to measure the (at equilibrium) reflectance of the NLM at normal incidence. We use a stabilized (fiber-coupled) tungsten–halogen lamp as a broadband light source (Thorlabs SLS201L). The white-light output is collimated by a parabolic mirror (Thorlabs RC12FC-P01). Before impinging on the sample, control over light polarization is accomplished by a polarizer (Thorlabs LPNIR100MP2) and a half-wave plate (Thorlabs AHWP10M-1600). The light passes through a beam splitter (Thorlabs BSN12R) and is focused on the sample by a lens (Thorlabs LA1251, NA = 0.02). The reflected light is collected by a parabolic mirror (Thorlabs RC12FC-P01), which is coupled to an optical fiber attached to a spectrometer (NIR-Quest 512-17).

6.

Appendix C: Numerical Simulations

The temporal dynamics of carrier density [N(Δt)], lattice temperature [T(Δt)], and intensity [I(Δt)]are described by the following CREM consisting of a system of two coupled-ordinary differential equations,

Eq. (4)

{N˙(Δt)=[γN(Δt)+τrFC1]N(Δt)+Ψ(Δt),Cp·T˙(Δt)=Eeh[γN(Δt)+τrFC1]N(Δt),
where γ is the bimolecular recombination rate, τrFC is the FC relaxation time, Cp is the Si volumetric specific heat, Eeh is the electron–hole pair energy, and Ψ(Δt) is the time-dependent FC generation rate per unit volume. Given the model accounting for linear and nonlinear carrier absorption, Eeh is equal to 2πc/λp for single-photon absorption or 4πc/λp for TPA. Moreover, Ψ(Δt)=Φ(Δt)+ΦTPA(Δt), where the linear and nonlinear contributions are calculated as Φ(Δt)=AλpI(Δt)/(2πcd) and ΦTPA(Δt)=ATPA(I)λpI(Δt)/(4πcd), respectively, with A [ATPA=1eβTPA·I(Δt)·d] being the linear (nonlinear) absorption and d being the top Si layer thickness. We numerically solve the system in Eq. (4) assuming a Gaussian temporal profile for the intensity of the pump laser pulse I(Δt)=2/π·(Finc/τp)·exp(2Δt2/τp2), with τp=64  fs.

The equilibrium reflectance (RRCWAeq) of the NLM is calculated29 via RCWA58,59 as the zeroth order back-diffraction efficiency (DEr,0), by assuming geometrical parameters obtained from SEM analysis of the fabricated NLM and dispersionless permittivities of the Si grating (εSi), SiO2 buffer layer (εSiO2), and Si substrate (εsub). In particular, starting from Refs. 45 and 46, the optical properties values are optimized (deviations <1%) to improve the matching between the calculated RRCWAeq and the experimental data [Fig. 1(b)]. The out-of-equilibrium reflectance spectrum is calculated via RCWA as RRCWAout=DEr,0(εSi+Δε,εSiO2,εsub), where the Si permittivity εSi is modified by the variation Δε according to Eqs. (1)–(3), thus assuming only a photoinduced variation of the grating optical properties.

7.

Appendix D: Time-Resolved Spectroscopy

Time-resolved spectroscopy measurements are performed in a home-built pump–probe setup. The system is based on a femtosecond laser (Monaco, Coherent) delivering 300 fs pulses centered at 1035 nm wavelength with 40  μJ energy at a 1 MHz repetition rate. The laser output feeds a hybrid optical parametric amplifier (Opera-F, Coherent) generating idler (IDL) and signal (SIG) beams. The IDL central wavelength is tuned at 1400  nm (bandwidth 100 nm, full width at half-maximum) and represents the probe beam of the experiment. The SIG output radiation, centered at 820 nm, is frequency-doubled in a 6-mm thick barium borate crystal to generate second-harmonic radiation at λp=410  nm (with a maximum pulse energy of 60 nJ); the latter is employed as the pump beam in the experiments. The residual component of the 820 nm signal is removed from the pump radiation by a bandpass filter (Thorlabs FESH0700). The intensity of the pump radiation impinging onto the sample is controlled by a variable neutral density filter (Thorlabs NDC-100C-4). In order to control the delay between the pump and probe arrival time (at sample position), the optical path length of the former is varied by a motorized linear delay stage (Newport M-ILS150CCL). The pump and probe beams are linearly orthogonally polarized, with the former being horizontal and the latter vertical. In order to ensure a proper collection of the reflected radiation, the probe beam impinges onto the sample at slightly off-normal incidence (φ4  deg). After the interaction with the sample, the reflected probe beam is collected and then dispersed on a spectrometer to achieve moderate spectral resolution (1.5 nm in the 1300 to 1500 nm spectral region). In particular, the reflected probe spectrum is measured when the pump does (not) reach the sample, Ron (Roff), and this provides the out-of-equilibrium (at-equilibrium) reflectance Rout (Req). The relative reflectance variation spectrum is then calculated as ΔR/R=(RoutReq)/Req=(RonRoff)/Roff (see Supplementary Material for more details).

Disclosures

The authors declare no conflicts of interests.

Code and Data Availability

All data in support of the findings of this paper are available within the article or as Supplementary Material.

Acknowledgments

C. De Angelis and O. Sergaeva would like to acknowledge the support of the Russian Science Foundation (Grant No. 22-12-00204). C. De Angelis also would like to acknowledge the financial support from the European Community through the “METAFAST” Project (H2020-FETOPEN-2018-2020, Grant No. 899673) and the Ministero Italiano dell’Istruzione (MIUR) through the “METEOR” Project (No. PRIN-2020, 2020EYLJT_002). A. Tognazzi thanks the European Union for the financial support through “FESR o FSE, PON Ricerca e Innovazione 2014-2020-DM 1062/2021” and the University of Palermo through “Fondo Finalizzato alla Ricerca di Ateneo 2023.” A. C. Cino acknowledges financial support from the European Union, NextgenerationEU, MUR D.M 737/2021, through the Project “Eurostart22” (No. PRJ-0988).

References

1. 

D. R. Smith, J. B. Pendry and M. C. K. Wiltshire, “Metamaterials and negative refractive index,” Science, 305 788 –792 https://doi.org/10.1126/science.1096796 SCIEAS 0036-8075 (2004). Google Scholar

2. 

K. Shastri and F. Monticone, “Nonlocal flat optics,” Nat. Photonics, 17 36 –47 https://doi.org/10.1038/s41566-022-01098-5 NPAHBY 1749-4885 (2022). Google Scholar

3. 

M. Khorasaninejad et al., “Metalenses at visible wavelengths: diffraction-limited focusing and subwavelength resolution imaging,” Science, 352 1190 –1194 https://doi.org/10.1126/science.aaf6644 SCIEAS 0036-8075 (2016). Google Scholar

4. 

G.-H. Go et al., “Scannable dual-focus metalens with hybrid phase,” Nano Lett., 23 3152 –3158 https://doi.org/10.1021/acs.nanolett.2c04696 NALEFD 1530-6984 (2023). Google Scholar

5. 

Y. Hu et al., “All-dielectric metasurfaces for polarization manipulation: principles and emerging applications,” Nanophotonics, 9 3755 –3780 https://doi.org/10.1515/nanoph-2020-0220 (2020). Google Scholar

6. 

K. I. Okhlopkov et al., “Tailoring third-harmonic diffraction efficiency by hybrid modes in high-Q metasurfaces,” Nano Lett., 21 10438 –10445 https://doi.org/10.1021/acs.nanolett.1c03790 NALEFD 1530-6984 (2021). Google Scholar

7. 

S. M. Kamali et al., “A review of dielectric optical metasurfaces for wavefront control,” Nanophotonics, 7 1041 –1068 https://doi.org/10.1515/nanoph-2017-0129 (2018). Google Scholar

8. 

M. Rahmani et al., “Reversible thermal tuning of all-dielectric metasurfaces,” Adv. Funct. Mater., 27 1700580 https://doi.org/10.1002/adfm.201700580 AFMDC6 1616-301X (2017). Google Scholar

9. 

Z. Guanxing et al., “Reconfigurable metasurfaces with mechanical actuations: towards flexible and tunable photonic devices,” J. Opt., 23 013001 https://doi.org/10.1088/2040-8986/abcc52 JOOPDB 0150-536X (2020). Google Scholar

10. 

K. Nishida et al., “All-optical scattering control in an all-dielectric quasi-perfect absorbing Huygens’ metasurface,” Nanophotonics, 12 139 –146 https://doi.org/10.1515/nanoph-2022-0597 (2022). Google Scholar

11. 

A. Afridi et al., “Ultrathin tunable optomechanical metalens,” Nano Lett., 23 2496 –2501 https://doi.org/10.1021/acs.nanolett.2c04105 NALEFD 1530-6984 (2023). Google Scholar

12. 

Y. Hu et al., “Electrically tunable multifunctional polarization-dependent metasurfaces integrated with liquid crystals in the visible region,” Nano Lett., 21 4554 –4562 https://doi.org/10.1021/acs.nanolett.1c00104 NALEFD 1530-6984 (2021). Google Scholar

13. 

Z. Yang et al., “Phase-only tuning of extreme Huygens metasurfaces enabled by optical anisotropy,” Adv. Opt. Mater., 10 2101893 https://doi.org/10.1002/adom.202101893 2195-1071 (2021). Google Scholar

14. 

J. Shabanpour, S. Beyraghi and A. Cheldavi, “Ultrafast reprogrammable multifunctional vanadium-dioxide-assisted metasurface for dynamic THz wavefront engineering,” Sci. Rep., 10 8950 https://doi.org/10.1038/s41598-020-65533-9 (2020). Google Scholar

15. 

A. Tripathi et al., “Tunable mie-resonant dielectric metasurfaces based on VO2 phase-transition materials,” ACS Photonics, 8 1206 –1213 https://doi.org/10.1021/acsphotonics.1c00124 (2021). Google Scholar

16. 

A. Tognazzi et al., “Opto-thermal dynamics of thin-film optical limiters based on the VO2 phase transition,” Opt. Mater. Express, 13 41 –52 https://doi.org/10.1364/OME.472347 (2022). Google Scholar

17. 

B. Li et al., “Fundamental limits for transmission modulation in VO2 metasurfaces,” Photonics Res., 11 B40 –B49 https://doi.org/10.1364/PRJ.474328 (2022). Google Scholar

18. 

D. de Ceglia et al., “Transient guided-mode resonance metasurfaces with phase-transition materials,” Opt. Lett., 48 2961 –2964 https://doi.org/10.1364/OL.486733 OPLEDP 0146-9592 (2023). Google Scholar

19. 

Y. Che et al., “Tunable optical metasurfaces enabled by multiple modulation mechanisms,” Nanophotonics, 9 4407 –4431 https://doi.org/10.1515/nanoph-2020-0311 (2020). Google Scholar

20. 

L. Carletti et al., “Reconfigurable nonlinear response of dielectric and semiconductor metasurfaces,” Nanophotonics, 10 4209 –4221 https://doi.org/10.1515/nanoph-2021-0367 (2021). Google Scholar

21. 

P. P. Iyer et al., “Sub-picosecond steering of ultrafast incoherent emission from semiconductor metasurfaces,” Nat. Photonics, 17 588 –593 https://doi.org/10.1038/s41566-023-01172-6 NPAHBY 1749-4885 (2023). Google Scholar

22. 

M. Taghinejad et al., “Ultrafast control of phase and polarization of light expedited by hot-electron transfer,” Nano Lett., 18 5544 –5551 https://doi.org/10.1021/acs.nanolett.8b01946 NALEFD 1530-6984 (2018). Google Scholar

23. 

M. R. Shcherbakov et al., “Ultrafast all-optical switching with magnetic resonances in nonlinear dielectric nanostructures,” Nano Lett., 15 6985 –6990 https://doi.org/10.1021/acs.nanolett.5b02989 NALEFD 1530-6984 (2015). Google Scholar

24. 

M. R. Shcherbakov et al., “Ultrafast all-optical tuning of direct-gap semiconductor metasurfaces,” Nat. Commun., 8 17 https://doi.org/10.1038/s41467-017-00019-3 NCAOBW 2041-1723 (2017). Google Scholar

25. 

G. Della Valle et al., “Nonlinear anisotropic dielectric metasurfaces for ultrafast nanophotonics,” ACS Photonics, 4 2129 –2136 https://doi.org/10.1021/acsphotonics.7b00544 (2017). Google Scholar

26. 

P. Franceschini et al., “Tuning the ultrafast response of Fano resonances in halide perovskite nanoparticles,” ACS Nano, 14 13602 –13610 https://doi.org/10.1021/acsnano.0c05710 ANCAC3 1936-0851 (2020). Google Scholar

27. 

C. Giannetti et al., “Thermomechanical behavior of surface acoustic waves in ordered arrays of nanodisks studied by near-infrared pump–probe diffraction experiments,” Phys. Rev. B, 76 125413 https://doi.org/10.1103/PhysRevB.76.125413 PRBMDO 1098-0121 (2007). Google Scholar

28. 

Z. Chai et al., “Ultrafast all-optical switching,” Adv. Opt. Mater., 5 1600665 https://doi.org/10.1002/adom.201600665 2195-1071 (2016). Google Scholar

29. 

P. Franceschini et al., “Nonlocal resonances in pedestal high-index-contrast metasurfaces based on a silicon-on-insulator platform,” Appl. Phys. Lett., 123 071701 https://doi.org/10.1063/5.0159275 APPLAB 0003-6951 (2023). Google Scholar

30. 

M. F. Limonov et al., “Fano resonances in photonics,” Nat. Photonics, 11 543 –554 https://doi.org/10.1038/nphoton.2017.142 NPAHBY 1749-4885 (2017). Google Scholar

31. 

A. Othonos, “Probing ultrafast carrier and phonon dynamics in semiconductors,” J. Appl. Phys., 83 1789 –1830 https://doi.org/10.1063/1.367411 JAPIAU 0021-8979 (1998). Google Scholar

32. 

A. J. Sabbah and D. M. Riffe, “Femtosecond pump–probe reflectivity study of silicon carrier dynamics,” Phys. Rev. B, 66 165217 https://doi.org/10.1103/PhysRevB.66.165217 (2002). Google Scholar

33. 

J. R. Goldman and J. A. Prybyla, “Ultrafast dynamics of laser-excited electron distributions in silicon,” Phys. Rev. Lett., 72 1364 –1367 https://doi.org/10.1103/PhysRevLett.72.1364 PRLTAO 0031-9007 (1994). Google Scholar

34. 

S. Jeong and J. Bokor, “Ultrafast carrier dynamics near the Si(100)2 × 1 surface,” Phys. Rev. B, 59 4943 –4951 https://doi.org/10.1103/PhysRevB.59.4943 (1999). Google Scholar

35. 

F. Gambino et al., “A review on dielectric resonant gratings: mitigation of finite size and Gaussian beam size effects,” Results Opt., 6 100210 https://doi.org/10.1016/j.rio.2021.100210 (2022). Google Scholar

36. 

A. Di Cicco et al., “Broadband optical ultrafast reflectivity of Si, Ge and GaAs,” Sci. Rep., 10 17363 https://doi.org/10.1038/s41598-020-74068-y (2020). Google Scholar

37. 

B. Bennett, R. Soref and J. Del Alamo, “Carrier-induced change in refractive index of InP, GaAs and InGaAsP,” IEEE J. Quantum Electron., 26 113 –122 https://doi.org/10.1109/3.44924 IEJQA7 0018-9197 (1990). Google Scholar

38. 

K. Sokolowski-Tinten and D. von der Linde, “Generation of dense electron-hole plasmas in silicon,” Phys. Rev. B, 61 2643 –2650 https://doi.org/10.1103/PhysRevB.61.2643 (2000). Google Scholar

39. 

A. Esser et al., “Ultrafast recombination and trapping in amorphous silicon,” Phys. Rev. B, 41 2879 –2884 https://doi.org/10.1103/PhysRevB.41.2879 (1990). Google Scholar

40. 

G. Ghosh, Handbook of Optical Constants of Solids: Handbook of Thermo-Optic Coefficients of Optical Materials with Applications, Academic Press( (1998). Google Scholar

41. 

R. Tomasiunas et al., “Femtosecond dephasing in porous silicon,” Appl. Phys. Lett., 68 3296 –3298 https://doi.org/10.1063/1.116579 APPLAB 0003-6951 (1996). Google Scholar

42. 

H. Petek and S. Ogawa, “Femtosecond time-resolved two-photon photoemission studies of electron dynamics in metals,” Prog. Surf. Sci., 56 239 –310 https://doi.org/10.1016/S0079-6816(98)00002-1 PSSFBP 0079-6816 (1997). Google Scholar

43. 

W. Nessler et al., “Femtosecond time-resolved study of the energy and temperature dependence of hot-electron lifetimes in Bi2CaCu2O8,” Phys. Rev. Lett., 81 4480 –4483 https://doi.org/10.1103/PhysRevLett.81.4480 PRLTAO 0031-9007 (1998). Google Scholar

44. 

P. Franceschini et al., “Coherent control of the orbital occupation driving the insulator-to-metal Mott transition in V2O3,” Phys. Rev. B, 107 L161110 https://doi.org/10.1103/PhysRevB.107.L161110 (2023). Google Scholar

45. 

C. Schinke et al., “Uncertainty analysis for the coefficient of band-to-band absorption of crystalline silicon,” AIP Adv., 5 067168 https://doi.org/10.1063/1.4923379 AAIDBI 2158-3226 (2015). Google Scholar

46. 

I. H. Malitson, “Interspecimen comparison of the refractive index of fused silica,” J. Opt. Soc. Am., 55 1205 –1209 https://doi.org/10.1364/JOSA.55.001205 JOSAAH 0030-3941 (1965). Google Scholar

47. 

H. M. van Driel, “Kinetics of high-density plasmas generated in Si by 1.06- and 0.53-μm picosecond laser pulses,” Phys. Rev. B, 35 8166 –8176 https://doi.org/10.1103/PhysRevB.35.8166 (1987). Google Scholar

48. 

B. E. Sernelius, “Intraband relaxation time in highly excited semiconductors,” Phys. Rev. B, 43 7136 –7144 https://doi.org/10.1103/PhysRevB.43.7136 (1991). Google Scholar

49. 

K. E. Myers, Q. Wang and S. L. Dexheimer, “Ultrafast carrier dynamics in nanocrystalline silicon,” Phys. Rev. B, 64 161309 https://doi.org/10.1103/PhysRevB.64.161309 (2001). Google Scholar

50. 

C.-M. Li, T. Sjodin and H.-L. Dai, “Photoexcited carrier diffusion near a Si(111) surface: non-negligible consequence of carrier-carrier scattering,” Phys. Rev. B, 56 15252 –15255 https://doi.org/10.1103/PhysRevB.56.15252 (1997). Google Scholar

51. 

V. R. Almeida et al., “All-optical control of light on a silicon chip,” Nature, 431 1081 –1084 https://doi.org/10.1038/nature02921 (2004). Google Scholar

52. 

U. Fano, “Effects of configuration interaction on intensities and phase shifts,” Phys. Rev., 124 1866 –1878 https://doi.org/10.1103/PhysRev.124.1866 PHRVAO 0031-899X (1961). Google Scholar

53. 

A. Di Cicco et al., “Broadband optical ultrafast reflectivity of Si, Ge and GaAs,” Sci. Rep., 10 17363 https://doi.org/10.1038/s41598-020-74068-y (2020). Google Scholar

54. 

D. G. Baranov et al., “Nonlinear transient dynamics of photoexcited resonant silicon nanostructures,” ACS Photonics, 3 1546 –1551 https://doi.org/10.1021/acsphotonics.6b00358 (2016). Google Scholar

55. 

J. N. Caspers, N. Rotenberg and H. M. van Driel, “Ultrafast silicon-based active plasmonics at telecom wavelengths,” Opt. Express, 18 19761 –19769 https://doi.org/10.1364/OE.18.019761 OPEXFF 1094-4087 (2010). Google Scholar

56. 

A. Ronchi et al., “Nanoscale self-organization and metastable non-thermal metallicity in mott insulators,” Nat. Commun., 13 3730 https://doi.org/10.1038/s41467-022-31298-0 NCAOBW 2041-1723 (2022). Google Scholar

57. 

L. Y. Beliaev et al., “Pedestal high-contrast gratings for biosensing,” Nanomaterials, 12 1748 https://doi.org/10.3390/nano12101748 (2022). Google Scholar

58. 

M. G. Moharam et al., “Formulation for stable and efficient implementation of the rigorous coupled-wave analysis of binary gratings,” J. Opt. Soc. Am. A, 12 1068 –1076 https://doi.org/10.1364/JOSAA.12.001068 JOAOD6 0740-3232 (1995). Google Scholar

59. 

L. Li, “Formulation and comparison of two recursive matrix algorithms for modeling layered diffraction gratings,” J. Opt. Soc. Am. A, 13 1024 –1035 https://doi.org/10.1364/JOSAA.13.001024 JOAOD6 0740-3232 (1996). Google Scholar

Biography

Andrea Tognazzi is a researcher at the University of Palermo, Italy. He received his bachelor's and MSc degrees from the Catholic University of Sacred Heart in Brescia, Italy, and his PhD from the University of Brescia. His research interests are in the fields of nonlinear optics, reconfigurable metasurfaces, and nanophotonics.

Paolo Franceschini is a postdoctoral researcher at the Department of Information Engineering of University of Brescia (Italy). He received his PhD (dual degree) in 2022 from Catholic University of the Sacred Heart (Italy) and Katholieke Universiteit Leuven (Belgium) with a joint research project within the International Doctoral Program in Science. Research interests focus on nonlinear optics, time-resolved spectroscopy, and photonics.

Olga Sergaeva is a researcher at the Department of Information Engineering of the University of Brescia (Italy). She received her BS, MS, and PhD (2013) at ITMO University (St. Petersburg, Russia), and had a postdoctoral fellowship (2016-2018) at the University of Missouri (Columbia, Missouri, USA). Her research interests include ultrafast light–matter interaction, nonlinear optics, nanophotonics, and reconfigurable metasurfaces.

Giovanni Finco received two MSc degrees, in telecommunication engineering from the University of Padova (Italy) and in photonics engineering from the Technical University of Denmark under the Top International Managers for Engineering Programme. He is currently employed as a doctoral researcher in the Physics Department of the Federal Institute of Technology in Zurich, Switzerland, where he develops lithium-niobate-on-insulator photonic integrated circuits technology for classical and quantum applications.

Andrei V. Lavrinenko received the PhD and DSci degrees from the Belarusian State University, Minsk, Belarus, in 1989 and 2004, respectively. Since 2004, he has been an associate professor with the Department of Electrical and Photonic Engineering at Technical University of Denmark. Since 2008, he has been leading the Metamaterials Group of this department. He is the author of more than 220 journal papers. His research interests include metamaterials, plasmonics, photonic crystals, and numerical methods.

Alfonso C. Cino received his PhD in electronic engineering from the University of Palermo, Italy, in 1998. He was the head of the Optical Technologies Lab at CRES (Center for Electronic Research in Sicily) until 2010. Currently he is an associate professor of electromagnetic fields at the University of Palermo. His research interests have been thin films, integrated optics, nonlinear frequency conversion, optical sensors, terahertz sources, photovoltaics, quantum optics, and optical metasurfaces.

Domenico de Ceglia is associate professor of electromagnetic fields and optics at University of Brescia, Italy. He received his PhD in electronic engineering from Politecnico di Bari, Italy, in 2007. He has been a senior research associate at the US Army C. Bowden Research Center. His research is focused on electromagnetics, optics, and photonics, with emphasis on light–matter interactions in nanostructures. He is a member of IEEE and a senior member of Optica.

Biographies for the other authors are not available.

CC BY: © The Authors. Published by SPIE and CLP under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Andrea Tognazzi, Paolo Franceschini, Olga Sergaeva, Luca Carletti, Ivano Alessandri, Giovanni Finco, Osamu Takayama, Radu Malureanu, Andrei V. Lavrinenko, Alfonso C. Cino, Domenico de Ceglia, and Costantino De Angelis "Giant photoinduced reflectivity modulation of nonlocal resonances in silicon metasurfaces," Advanced Photonics 5(6), 066006 (9 December 2023). https://doi.org/10.1117/1.AP.5.6.066006
Received: 6 September 2023; Accepted: 1 November 2023; Published: 9 December 2023
Advertisement
Advertisement
KEYWORDS
Silicon

Reflectivity

Modulation

Picosecond phenomena

Solids

Ultrafast phenomena

Photonics

Back to Top