ACS Publications. Most Trusted. Most Cited. Most Read
Low-Power Threshold Optical Bistability Enabled by Hydrodynamic Kerr Nonlinearity of Free Carriers in Heavily Doped Semiconductors
My Activity
  • Open Access
Article

Low-Power Threshold Optical Bistability Enabled by Hydrodynamic Kerr Nonlinearity of Free Carriers in Heavily Doped Semiconductors
Click to copy article linkArticle link copied!

Open PDFSupporting Information (2)

ACS Photonics

Cite this: ACS Photonics 2024, 11, 11, 4812–4817
Click to copy citationCitation copied!
https://doi.org/10.1021/acsphotonics.4c01308
Published September 19, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

We develop an efficient numerical model based on the semiclassical hydrodynamic theory for studying Kerr nonlinearity in degenerate electron systems such as heavily doped semiconductors. This model provides direct access to the electromagnetic responses of the quantum nature of the plasmons in heavily doped semiconductors with complex geometries, which is nontrivial for conventional frameworks. Using this model, we demonstrate nanoscale optical bistability at an exceptionally low-power threshold of 1 mW by leveraging Kerr-type hydrodynamic nonlinearities supported by the heavily doped semiconductor’s free carriers. This high nonlinearity is enabled by a strong coupling between metallic gap plasmons and longitudinal bulk plasmons in the semiconductor due to quantum pressure. These findings offer a viable approach to studying Kerr-type nonlinearity and lay the groundwork for developing efficient and ultrafast all-optical nonlinear devices.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
Copyright © 2024 The Authors. Published by American Chemical Society

Introduction

Click to copy section linkSection link copied!

The optical Kerr effect is a third-order nonlinear phenomenon that allows modulating the refractive index of the materials according to the applied electric field intensity. (1) Exploiting such an effect to attain efficient and ultrafast photon–photon interactions has long been sought after for the potential applications in all-optical information processing and storage. (2−10) More recently, the advent of artificial intelligence has led to an exponential increase in the demand for low-power computation solutions, bringing renewed interest in integrated all-optical nonlinear systems.
Semiconductors represent in this context an interesting class of materials, as they are compatible with large-scale fabrication methods for integrated devices while also supporting some of the largest optical nonlinearities. (11,12) When heavily doped, semiconductors can undergo a transition from the quantum regime, where their response is governed by single-particle excitations, to the classical regime of collective plasmon oscillations. (12) Due to the hydrodynamic nature of nonequilibrium carriers, collective plasmons can exhibit an intrinsic and strong free carrier optical nonlinearity. (2,13−17) Although these properties occur, in principle, in all degenerate electron systems, they are significantly larger for materials with low equilibrium density and small effective mass, such as doped InP and InGaAs. (18) In fact, very recently, it has been experimentally shown by measuring third-harmonic generation from heavily doped InGaAs nanoantennas that free electron nonlinearities can easily overcome intrinsic χ(3) nonlinearities. (19) The drawback, however, is that free electrons inside a material mostly respond to external optical fields that oscillate below the plasma frequency as surface plasmons. This confines hydrodynamic nonlinear interactions at the material surface, (13,14,17) severely limiting the Kerr effect.
In this context, longitudinal bulk plasmons (LBPs) – charge density waves occurring in the bulk of spatially dispersive materials (20−22) – become very relevant. The relative distribution of the electrons over a positively charged background generates their characteristic longitudinal electric fields. (23) LBPs have been observed in thin metal films where they naturally occur in the ultraviolet energy range. (24,25) More recently, LBP resonances have been observed in transparent conducting oxide materials in the infrared, such as doped cadmium oxide. (26) Existing above the plasma frequency, the LBP sacrifices the typical field confinement associated with surface plasmons for a much larger active volume for nonlinear interactions, unlocking a new mechanism for low-power Kerr nonlinearity at the nanoscale.
From a theoretical perspective, LBP modes can be excited in finite size systems only if the response of the material is nonlocal. (21,27) One way to account for nonlocality in degenerate electron systems is to model the classical free electron dynamics incorporated with density-dependent energy potentials that encapsulate their quantum properties. This approach is usually referred to as hydrodynamic theory (HT). (28,29) For a proper choice of the kinetic energy potential, it can be shown that the HT is equivalent to the Schrödinger equation for single-particle systems (or more precisely, single orbital systems). (30) Interestingly, the HT maintains its quantum structure for systems characterized by many electrons, as long as they can be considered as a collective quasiparticle, such as plasmons. With this premise, quantum size resonances, especially the intersubband transition plasmons, in quantum-confined systems (i.e., quantum wells) in traditional semiconductors can be seen from a theoretical point of view as equivalent to quantized LBP modes. Quantum well states have been indeed experimentally observed in metals. (31−33) The coupling between semiconductor quantum wells intersubband transitions and photonic systems has been experimentally exploited for linear (34,35) and nonlinear (11,32,33,36) optical modulations. However, modeling such collective intersubband plasmon oscillations in the conduction band of the quantum wells coupled with complex photonic structures typically relies on the combination of separate steps: calculate the band information with effective parameters derived and predict optical responses with separate semiclassical or phenomenological models such as coupled mode theory. (37,38)
In contrast, in this paper, we develop an ad hoc time-domain electromagnetic model integrated with semiclassical HT that can directly access the linear and nonlinear optical responses arising from the quantum nature of electrons, which is typically nontrivial for conventional semiconductor frameworks based on Schrödinger-Poisson equations. (31,32,39) With this model, we consider for the first time the plasmonic hydrodynamic contributions to Kerr nonlinearity in heavily doped semiconductors and demonstrate that LBP strongly coupled with gap surface plasmons can indeed drive a very large Kerr effect with a wide hysteresis in subwavelength volumes. As a result, an ultralow threshold (∼1.55 mW/μm2) and compact optically bistable device is proposed.

Results and Discussion

Click to copy section linkSection link copied!

The many-body nature of a free electron gas driven by electromagnetic fields E(r, t) and H(r, t) can be described, within the HT, through two macroscopic fields, the electron density n(r, t) and velocity v(r, t):
me(t+v·+γ)v=e(E+v×μ0H)δG[n]δn
(1)
with me and e being the effective electron mass and the electron charge, respectively; γ is the phenomenological damping rate; and μ0 is the vacuum permeability. The terms from left to right in eq 1 account for time evolution, convection, dissipation, Coulomb, Lorentz, and internal energy potential, i.e., quantum pressure, respectively. Since we are not interested in electron spill-out effects, here, we consider the energy functional within the Thomas–Fermi (TF) approximation G[n] ≃ TTF[n]. (28) Introducing the polarization field defined as ∂P/∂t = J = −env, expanding the nonlinear terms up to the third order and neglecting the higher-order nonlinear terms, (16) we obtain the following equation (see Supporting Information for detailed derivation):
2Pt2+γPt=n0e2meE+β2(·P)+SNL
(2)
where the first-order quantum pressure term has been written as (δTTFδn)1=meen0β2(·P), where the parameter β is related to the speed of sound in the Fermi-degenerate plasma vF, β2=13vF2=109cTFmen02/3, cTF=me2310(3π2)2/3. All of the higher-order terms containing the nonlinear responses to the external field and electron–electron interactions are encapsulated into the total nonlinear source SNL, which will be elaborated on later in eq 4a and 4b. Its origin should not be confused with the bulk nonlinear susceptibility χ(3) that comes from the anharmonic potential experienced by the electrons in the crystals. The former is mostly a nonlocal response depending on the gradient of the fields, and the latter is a local response that will be considered separately in the wave equation as eq 5.
Let us start with the simplest system supporting quantized LBP modes (hereafter, we directly use LBP to refer to this quantized LBP resonance for simplicity): a thin slab excited by p-polarized light at oblique incidence of θ = 60°, as depicted in Figure 1a. Note that s-polarized light cannot excite the LBP. An electric field with a nonzero normal component at the material interface is required to align with the charge oscillations necessary for driving LBP. The linear properties of LBP can be obtained in the frequency domain by solving the linear part of eq 2 (i.e., SNL = 0) coupled to the wave equation:
××Eεrω2c2Eμ0ω2P=0
(3)
where εr is the relative permittivity accounting for the dielectric local response and c is the speed of light in vacuum.

Figure 1

Figure 1. Optical absorption of a heavily doped InGaAs thin slab. (a) Schematic. (b) Linear absorption spectra of the thin slab with a local response approximation model (LRA, cyan) and nonlocal hydrodynamic theory (HT, purple). (c) Zoom-in plot of the higher-order LBPs. (d) Induced charge ∇·P and electric field of LBPs with an order of n = 1, 3, 5.

We consider a thin layer (g = 10 nm) of heavily doped InGaAs with equilibrium carrier density n0 = 6 × 1018 cm–3. The doped InGaAs can be described as a Drude-like material with εinf = 12, γ = 8.9 ps–1, and me = 0.041m0 (m0 being the electron mass). The parameters were taken by fitting the experimental data in ref (19) such that the screened bulk plasma wavelength lays at λp=2πc(meε0εinf/(n0e2))=9.56μm.
The linear absorption spectra with and without quantum pressure are shown in Figure 1b,c. With local response approximation (LRA), a prominent absorption peak arises due to the epsilon-near-zero (ENZ) enhancement at λp. However, as electron–electron interactions are considered with the HT, the ENZ resonance exhibits a noticeable blueshift, accompanied by the emergence of additional higher-order resonances fulfilling the LBP condition of ε(ω, k) = 0. Only LBP modes with odd orders (n = 1, 3, 5, ···) have a net dipole moment that can be excited by a plane wave. In Figure 1d, the charge density and electric field distributions clearly show the nodes and antinodes of the standing waves formed by the charge density oscillations. The energy and the intensity of these quantized modes greatly depend on the thickness of the slab due to the confinement in that dimension, similar to quantum wells.
Although appreciable, LBP modes in a thin slab display an absorption, in the best case, of only ∼1.5% (similar values are observed for the reflectance). Higher-order resonances, shown in Figure 1c, have even smaller absorption peaks (less than 0.2%). This is not surprising since the LBPs can couple only weakly to the transverse plane wave in free space. This characteristic hinders both the experimental observation and the application in linear and nonlinear devices.
To overcome this issue, we consider a hybrid system constituted by a thin layer of heavily doped InGaAs sandwiched between a silver film and an array of silver stripes (Figure 2a), which is experimentally feasible. (36,38) Such systems are usually referred to as nanopatch antennas, in analogy to their radio frequency homologue. Nanopatch antennas have an absorption cross-section that is much larger than their physical dimensions and can concentrate far-field radiation into deep subwavelength nanogaps. (40) Additionally, the strong nanogap normal fields (Ez) are perfectly aligned with the longitudinal fields of the LBP supported by the semiconductor film (lower panel in Figure 2a).

Figure 2

Figure 2. Hybrid bulk-plasmon slab-nanopatch system. (a) Schematic of the hybrid structure; height h and periodicity Px of the antenna are 50 and 200 nm, respectively. Length, L, can vary. Bottom panels: the induced charges and electric field of the LBP in the thin spacer. (b) Strong coupling between LSP and LBP with a clear Rabi splitting in reflectance spectra, where three particular cases of different L marked by dashed lines are shown in (c) for details. Gray shades in (c) represent the reflectance of the bare nanopatch without nonlocality. Reflectance spectrum of the on-resonance case (L = 124 nm) is fitted with a Rabi splitting of Ω = 33.5 meV. Green shades mark the center wavelengths of the pump for probing Kerr nonlinearities.

The hybrid structure is designed as follows: Since the LBP is mostly determined by the thickness and doping of the InGaAs, we fix g and n0 as the same as those described in Figure 1 such that the third-order LBP resonance is within the range of the nanopatch fundamental mode, i.e., ∼2.15 μm, as shown in Figure 2b. By adjusting the length L of the nanopatch antenna, we can tune the resonances of the antenna localized surface plasmon (LSP) and its spectral overlap with the LBP. Because of the relatively large size and high equilibrium electron density of the silver nanopatches, the nonlocality and free electron nonlinearity can be negligible. In Supporting Information S3, we also apply hydrodynamic theory to silver for validation. Thus, we treat silver as a local Drude metal with εinf = 4.039, ωp = 9.172 eV, and γ = 0.0207 eV. (41)
The detuning between the LBP and the LSP is an essential factor for the coupling. In Figure 2b, we map the reflectance of the system as a function of the incident wavelength and nanopatch L, which highlights the anticrossing dispersion between the LBP and LSP. The map shows antipeaks down to virtually zero reflectance, unveiling a nearly perfect absorption behavior. (42) Even when the two modes are quite detuned (e.g., L = 140 nm), the absorption (81.7%) of the LBP is still enormously enhanced compared to the bare InGaAs slabs (0.16% in Figure 1c). This is due to the broad line width of the LSP resonances of the nanopatch, which can still provide decent field enhancements at certain detunings. This is substantial as the hybrid mode provides a practical and efficient way to access the elusive LBP in the previous experiment. (26)
When the two uncoupled states perfectly overlap, i.e., for L = 124 nm, they hybridize into two modes with identical line width. By fitting the Rabi splitting with a temporal coupled mode theory, (43) it is possible to confirm that the system is working in a strong coupling regime. We have found that the relation between the coupling strength gc (contributing to Rabi splitting Ω = 2gc = 33.5 meV) and the sum of half-width at half-maximum of each bare mode Σii/2) satisfies the well-known condition 4gcii) = 1.26 > 1, which indicates by a large margin that the system is indeed working in the strong coupling regime. As a consequence, the two resulting dressed states (i.e., the two peaks in the spectra) are hybrid bulk-surface plasmons. These hybrid modes hold the potential to inherit both the strong field enhancement from the LSP and the strong free electron nonlinearity from the LBP, providing a promising direction for the development of efficient and low-threshold nonlinear devices.
To prove this, let us investigate the hydrodynamic Kerr nonlinearity supported by the dressed LBP state in the hybrid system. The free carrier Kerr-type nonlinearity in the doped semiconductor is assumed to mainly come from the third-order hydrodynamic contributions in the constitutive relation eq 2, which arise from the convection and TF pressure, i.e., SNL = Scv + STF, where (see Supporting Information for derivations):
Scv=1e2n02[(·P)P˙(·P˙*)+(·P)P˙*(·P˙)+(·P*)P˙(·P˙)+(·P)P˙·P˙*+(·P)P˙*·P˙+(·P*)P˙·P˙+P˙·P˙(·P*)+2P˙·P˙*(·P)]
(4a)
STF=19β2e2n02[(·P)2(·P*)]
(4b)
The “·” over the variables denotes the time derivative, while the “∗” denotes the complex conjugated. Note that because SNL ∝ 1/n02, these nonlinear terms, usually neglected in noble metals, become significant in semiconductors. (18) In our calculation, the equilibrium density n0 in heavily doped InGaAs is nearly 4 orders of magnitude less than that of noble metals, enhancing nonlinear sources by 8 orders of magnitude.
Equations 2 and 4a are solved with the wave equation for the magnetic vector potential A(r, t):
××A+εrc22At2+μ0t(P+PdNL)=0
(5)
The vector potential A works in the Coulomb gauge with a vanishing gradient of the scalar potential due to the absence of a static field such that the electric field E = −∂A/∂t and magnetic field μ0 H = ∇ × A. The linear and nonlinear hydrodynamic contributions are coupled into the wave equation through a polarization vector P. The lattice Kerr nonlinearity from the dielectric χ(3) = 1.6 × 10–18 m2/V2 is considered through another polarization term PdNL = 3ε0χ(3)|E|2E. (1) The contributions from the free carriers (i.e., SNL) and dielectric χ(3) (i.e., PdNL) will be separated and compared. Other nonlinear contributions including higher-order and cascaded processes are beyond the scope of this letter and have been excluded from consideration.
The system of eqs 2, 4a, 4b, and 5 is solved with a finite element method in the time domain through a customized implementation carried out using COMSOL Multiphysics. The driving field is continuous, and the incident intensity is gradually increased and decreased (will be discussed as Figure 3c) to adiabatically sweep out the nonlinear reflectance and hysteresis loop. Also, see video in Supporting Information for the probing scheme. To improve the computational efficiency and reduce the cost, we have implemented a time domain pulse envelope method in which the fast time-varying components from the full time-evolving field are separated, i.e., all the fields are expressed as F(r, t) = (r, t)eik·r+iωt, with F = E, H, P, A. Here, k and ω are the carrier wave vector and frequency, respectively, of the initial field. The tilde indicates the corresponding slowly varying envelope field. Solving for the envelopes allows for a much larger time step (e.g., ∼102 times the wave period). To take into account the transformation above, all of the equations must be modified assuming the following transformations: ∇ → ∇ – ik and ∂/∂t → ∂ /∂t + iω. In our calculations, we neglect the hydrodynamic nonlinear contribution from silver due to its high equilibrium charge density, which leads to a negligible nonlinear contribution. (16) To verify this, we also consider the hydrodynamic contributions from silver in Supporting Information S3; however, this does not result in any significant improvement. This is due to the hydrodynamic contributions in the nonlinear source from InGaAs being much stronger than those from silver (Supporting Information Figure S2c).

Figure 3

Figure 3. Hysteresis loop of two-dimensional nonlinear nanopatch systems. Prominent optical bistabilities with two particular antenna sizes: (a) L = 110 nm and (b) L = 135 nm. (c) Nonlinear reflection (pulse distortion) due to the bistability with L = 135 nm.

To understand the role played by the LBP-LSP coupling in the nonlinear process, we now study in detail three specific detuning situations, as summarized in Figure 2c. As the LBP is red-detuned from the LSP (L = 110 nm, bottom panel of Figure 2c), we use the sharp resonance at 2.142 μm for achieving optical bistability. We set the input signal carrier wavelength slightly above the resonance, at 2.165 μm, to probe the hysteresis loop. As a result, with the increase of pumping power, the Kerr nonlinearity will shift the resonance toward lower energies so that the resonance will gradually overlap with the pump frequency. As shown by the guide-to-eyes gray arrow in Figure 2b for L = 110 nm (orange line), the reflectance would see a decrease until it crosses the dip and then experiences an increase again.
By modulating the input intensity, we can generate the full hysteresis loop shown in Figure 3a, whose reflectance has bistable states with “on” and “off” states between ∼60% and 30% with a threshold starting around 2.45 mW/μm2. This situation could have a higher reflectance at the starting point, but the red-detuned LBP tends to move away from LSP which degrades the low reflectance dip. Likewise, we could switch off the hydrodynamic nonlinear sources (SNL = 0) and study the responses solely from enhanced dielectric nonlinearity PdNL. The results show no bistability, confirming that the hydrodynamic Kerr nonlinearity is dominating (Supporting Information Figure S1a). In addition, we keep increasing the incident intensity to allow dielectric PdNL to trigger the nonlinearity. As shown in Supporting Information Figure S1b, the threshold intensity is 5 orders of magnitude higher than that with hydrodynamic Kerr effect. In addition, the hysteresis window becomes extremely narrow compared with Figure 3a. It means that the combination of the narrow line width and the strong hydrodynamic Kerr effect is the key to the low-threshold wide hysteresis loop.
While the situation in which the LBP and LSP resonances are overlapped (Figure 2c, with L = 124 nm) is interesting from a linear response point of view, it does not represent a viable starting point for carrying out bistability since the LBP sharp resonance is hybridized and hence broadened. This results in a higher intensity threshold and a narrower hysteresis loop.
Let us now consider the opposite situation where the LBP is blue-detuned from the LSP, taking L = 135 nm for instance, and tuning the pumping wavelength at 2.135 μm. Instead of getting away from the LSP, the LBP is now pushed by the nonlinearity toward the LSP, as shown by the guide-to-eyes gray arrow in Figure 2b for L = 135 nm (purple line). The reflectance will decrease and cross the dip, instead of increasing as in the previous situation, due to the broadening of the branch in Rabi splitting. This causes a “flat” off-state in Figure 3b with a 20% reflectance. We can interpret this result as a self-modulated dynamic transition from detuned coupling to resonant strong coupling driven by the nonlocal Kerr nonlinearity. Moreover, the LBP is moving toward the LSP to gradually get a larger field enhancement and greater coupling due to the spectra overlap. This further lowers the intensity threshold for achieving optical bistability: 1.55 mW/μm2 for achieving 70% to 20% branches with 5.4 dB modulation depth, inferring the effectiveness. As shown in Supporting Information Figure S3, this low-intensity threshold can remain robust even when accounting for increased damping γ, which may come from fabrication imperfections or other factors in the experiment. The threshold is maintained at a level of 10 mW/μm2 as the damping grows to 30 ps–1. Even as damping becomes very large (γ = 40 ps–1) such that the LBP modes become weak and flat (Figure S3a), a sigmoid-function-like nonlinear response can still be expected at a level of 10 mW/μm2. Besides, the dielectric χ(3) contribution is again negligible compared with the hydrodynamic ones (Supporting Information Figure S1), revealing the advantage of the free carriers as a nonlinear source for the Kerr effect.
In Figure 3c, we depict the nonlinear reflection against the input light (red dash line). We can observe that the reflected signal is dramatically distorted due to the transitions at a threshold intensity. Other thermal and nonthermal nonlinearity might be considered separately especially when the system is excited by an ultrashort pulse with strong peak intensity. (5,44) These are not considered in the current work because the scope of the present work focuses on the hydrodynamic contributions in Kerr nonlinearity with a much lower excitation intensity. Although this threshold is comparable to previously considered systems, (3,4,7,45,46,47) the required intensity could be further reduced to ∼μW/μm2 if one can overcome the high computational cost and consider a full three-dimensional nanopatch antenna system, which would provide an extra enhancement due to the additional transverse confinement. More interestingly, applying a bias voltage to the semiconductors through the metal patches on top can be expected to allow for the active control of the Kerr nonlinearity due to the charge depletion. Potentially, 1–2 orders of the magnitude drop of the threshold can be expected. (17) Furthermore, the switching time, which determines the limit of the operation frequency, is another important factor. In Supporting Information Figure S4, we simulate the optical switching response of the structure. The switching-on and -off times can be approximately 1 ps. (see detailed discussion in Supporting Information S5). Overall, this low-power threshold enables an efficient switch between two states, while the decent modulation depth ensures the effectiveness and fidelity of the modulation process. Our finding of high hydrodynamic Kerr nonlinearity could enable crucial applications in optical data processing, such as memory, binary switching, and logic functions.

Conclusions

Click to copy section linkSection link copied!

In conclusion, we proposed an ad hoc numerical approach based on the finite element method in the time domain to self-consistently solve intensity-dependent equations at the nanoscale. By selecting appropriate kinetic energies and quantum potentials in the semiclassical HT, this approach is able to describe the behaviors of the quantized resonances in degenerate electron systems, including the intriguing nonlinearities from the intersubband transitions. More importantly, by leveraging the strong coupling between LBP and LSP, we explored the impact of the hydrodynamic Kerr nonlinearity and demonstrated large contrast bistability with a power threshold of ∼1 mW (considering a full patch and a focused beam area to the diffraction limit). Our findings open a viable avenue for studying and designing ultrafast all-optical integrated circuits.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsphotonics.4c01308.

  • Derivation and implementation method of hydrodynamic nonlinear sources and contributions from dielectric χ(3) in the nonlinear reflectance, hydrodynamic contributions from silver and heavily doped InGaAs, robustness of the system with different dampings, and switching time (PDF)

  • Video of temporal evolution of the nonlinear system (MP4)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
  • Authors
  • Funding

    This work was supported by funding from the European Innovation Council through its Horizon Europe program with Grant Agreement No. 101046329.

  • Notes
    The authors declare no competing financial interest.

References

Click to copy section linkSection link copied!

This article references 47 other publications.

  1. 1
    Boyd, R. W. Nonlinear Optics., 3rd ed.; Academic Press, Inc.: USA, 2008.
  2. 2
    Krasavin, A. V.; Ginzburg, P.; Zayats, A. V. Free-electron optical nonlinearities in plasmonic nanostructures: a review of the hydrodynamic description. Laser Photonics Rev. 2018, 12, 1700082  DOI: 10.1002/lpor.201700082
  3. 3
    Almeida, V. R.; Lipson, M. Optical bistability on a silicon chip. Opt. Lett. 2004, 29, 23872389,  DOI: 10.1364/OL.29.002387
  4. 4
    Gibbs, H.; Tarng, S.; Jewell, J.; Weinberger, D.; Tai, K.; Gossard, A.; McCall, S.; Passner, A.; Wiegmann, W. Room-temperature excitonic optical bistability in a GaAs-GaAlAs superlattice étalon. Appl. Phys. Lett. 1982, 41, 221222,  DOI: 10.1063/1.93490
  5. 5
    Un, I.-W.; Sarkar, S.; Sivan, Y. Electronic-based model of the optical nonlinearity of low-electron-density Drude materials. Phys. Rev. Appl. 2023, 19, 044043  DOI: 10.1103/PhysRevApplied.19.044043
  6. 6
    Cox, J. D.; García, J.; de Abajo, F. Electrically tunable nonlinear plasmonics in graphene nanoislands. Nat. Commun. 2014, 5, 5725,  DOI: 10.1038/ncomms6725
  7. 7
    Christensen, T.; Yan, W.; Jauho, A.-P.; Wubs, M.; Mortensen, N. A. Kerr nonlinearity and plasmonic bistability in graphene nanoribbons. Phys. Rev. B 2015, 92, 121407  DOI: 10.1103/PhysRevB.92.121407
  8. 8
    Cox, J. D.; Silveiro, I.; García de Abajo, F. J. Quantum effects in the nonlinear response of graphene plasmons. ACS Nano 2016, 10, 19952003,  DOI: 10.1021/acsnano.5b06110
  9. 9
    Ryabov, D.; Pashina, O.; Zograf, G.; Makarov, S.; Petrov, M. Nonlinear optical heating of all-dielectric super-cavity: efficient light-to-heat conversion through giant thermorefractive bistability. Nanophotonics 2022, 11, 39813991,  DOI: 10.1515/nanoph-2022-0074
  10. 10
    Niu, J.; Shao, H.; Feng, Y.; Gao, B.; Zhang, Y.; Li, Y.; Chen, H.; Qian, H. All-Optical Nonlinear Neuron Based on Metallic Quantum Wells. Adv. Opt. Mater. 2023, 11, 2300223  DOI: 10.1002/adom.202300223
  11. 11
    Lee, J.; Tymchenko, M.; Argyropoulos, C.; Chen, P.-Y.; Lu, F.; Demmerle, F.; Boehm, G.; Amann, M.-C.; Alu, A.; Belkin, M. A. Giant nonlinear response from plasmonic metasurfaces coupled to intersubband transitions. Nature 2014, 511, 6569,  DOI: 10.1038/nature13455
  12. 12
    Taliercio, T.; Biagioni, P. Semiconductor infrared plasmonics. Nanophotonics 2019, 8, 949990,  DOI: 10.1515/nanoph-2019-0077
  13. 13
    Scalora, M.; Vincenti, M. A.; de Ceglia, D.; Roppo, V.; Centini, M.; Akozbek, N.; Bloemer, M. Second-and third-harmonic generation in metal-based structures. Phys. Rev. A 2010, 82, 043828  DOI: 10.1103/PhysRevA.82.043828
  14. 14
    Ciracì, C.; Poutrina, E.; Scalora, M.; Smith, D. R. Second-harmonic generation in metallic nanoparticles: Clarification of the role of the surface. Phys. Rev. B 2012, 86, 115451  DOI: 10.1103/PhysRevB.86.115451
  15. 15
    Khalid, M.; Ciracì, C. Enhancing second-harmonic generation with electron spill-out at metallic surfaces. Commun. Phys. 2020, 3, 214,  DOI: 10.1038/s42005-020-00477-0
  16. 16
    De Luca, F.; Ortolani, M.; Ciracì, C. Free electron nonlinearities in heavily doped semiconductors plasmonics. Phys. Rev. B 2021, 103, 115305  DOI: 10.1103/PhysRevB.103.115305
  17. 17
    De Luca, F.; Ciracì, C. Impact of surface charge depletion on the free electron nonlinear response of heavily doped semiconductors. Phys. Rev. Lett. 2022, 129, 123902  DOI: 10.1103/PhysRevLett.129.123902
  18. 18
    De Luca, F.; Ortolani, M.; Ciracì, C. Free electron harmonic generation in heavily doped semiconductors: the role of the materials properties. EPJ. Appl. Metamat. 2022, 9, 13,  DOI: 10.1051/epjam/2022011
  19. 19
    Rossetti, A. Origin of optical nonlinearity in plasmonic semiconductor nanostructures. arXiv preprint 2024, arXiv:2402.15443, https://arxiv.org/abs/2402.15443 (accessed Feb 23, 2024).
  20. 20
    Ruppin, R. Optical properties of a plasma sphere. Phys. Rev. Lett. 1973, 31, 1434,  DOI: 10.1103/PhysRevLett.31.1434
  21. 21
    Ruppin, R. Extinction properties of thin metallic nanowires. Opt. Commun. 2001, 190, 205209,  DOI: 10.1016/S0030-4018(01)01063-X
  22. 22
    Raza, S.; Toscano, G.; Jauho, A.-P.; Wubs, M.; Mortensen, N. A. Unusual resonances in nanoplasmonic structures due to nonlocal response. Phys. Rev. B 2011, 84, 121412  DOI: 10.1103/PhysRevB.84.121412
  23. 23
    Benisty, H.; Greffet, J.-J.; Lalanne, P. Introduction to nanophotonics; Oxford university press, 2022.
  24. 24
    Anderegg, M.; Feuerbacher, B.; Fitton, B. Optically excited longitudinal plasmons in potassium. Phys. Rev. Lett. 1971, 27, 1565,  DOI: 10.1103/PhysRevLett.27.1565
  25. 25
    Lindau, I.; Nilsson, P. Experimental verification of optically excited longitudinal plasmons. Phys. Scr. 1971, 3, 87,  DOI: 10.1088/0031-8949/3/2/007
  26. 26
    De Ceglia, D.; Scalora, M.; Vincenti, M. A.; Campione, S.; Kelley, K.; Runnerstrom, E. L.; Maria, J.-P.; Keeler, G. A.; Luk, T. S. Viscoelastic optical nonlocality of low-loss epsilon-near-zero nanofilms. Sci. Rep. 2018, 8, 9335,  DOI: 10.1038/s41598-018-27655-z
  27. 27
    Christensen, T.; Yan, W.; Raza, S.; Jauho, A.-P.; Mortensen, N. A.; Wubs, M. Nonlocal response of metallic nanospheres probed by light, electrons, and atoms. ACS Nano 2014, 8, 17451758,  DOI: 10.1021/nn406153k
  28. 28
    Ciracì, C.; Della Sala, F. Quantum hydrodynamic theory for plasmonics: Impact of the electron density tail. Phys. Rev. B 2016, 93, 205405  DOI: 10.1103/PhysRevB.93.205405
  29. 29
    Toscano, G.; Straubel, J.; Kwiatkowski, A.; Rockstuhl, C.; Evers, F.; Xu, H.; Asger Mortensen, N.; Wubs, M. Resonance shifts and spill-out effects in self-consistent hydrodynamic nanoplasmonics. Nat. Commun. 2015, 6, 7132,  DOI: 10.1038/ncomms8132
  30. 30
    Ciracì, C. Current-dependent potential for nonlocal absorption in quantum hydrodynamic theory. Phys. Rev. B 2017, 95, 245434  DOI: 10.1103/PhysRevB.95.245434
  31. 31
    Silberberg, Y.; Sands, T. Optical properties of metallic quantum wells. IEEE J. Quantum Electron. 1992, 28, 16631669,  DOI: 10.1109/3.142553
  32. 32
    Qian, H.; Xiao, Y.; Liu, Z. Giant Kerr response of ultrathin gold films from quantum size effect. Nat. Commun. 2016, 7, 13153  DOI: 10.1038/ncomms13153
  33. 33
    Qian, H.; Li, S.; Chen, C.-F.; Hsu, S.-W.; Bopp, S. E.; Ma, Q.; Tao, A. R.; Liu, Z. Large optical nonlinearity enabled by coupled metallic quantum wells. Light Sci. Appl. 2019, 8, 13,  DOI: 10.1038/s41377-019-0123-4
  34. 34
    Pirotta, S.; Tran, N.-L.; Jollivet, A.; Biasiol, G.; Crozat, P.; Manceau, J.-M.; Bousseksou, A.; Colombelli, R. Fast amplitude modulation up to 1.5 GHz of mid-IR free-space beams at room-temperature. Nat. Commun. 2021, 12, 799,  DOI: 10.1038/s41467-020-20710-2
  35. 35
    Malerba, M.; Pirotta, S.; Aubin, G.; Lucia, L.; Jeannin, M.; Manceau, J.-M.; Bousseksou, A.; Lin, Q.; Lampin, J.-F.; Peytavit, E.; Barbieri, S.; Li, L.; Davies, G.; Linfield, E. H.; Colombelli, R. Ultrafast (≈ 10 GHz) mid-IR modulator based on ultrafast electrical switching of the light–matter coupling. Appl. Phys. Lett. 2024, 125, 041101  DOI: 10.1063/5.0213965
  36. 36
    Jeannin, M.; Cosentino, E.; Pirotta, S.; Malerba, M.; Biasiol, G.; Manceau, J.-M.; Colombelli, R. Low intensity saturation of an ISB transition by a mid-IR quantum cascade laser. Appl. Phys. Lett. 2023, 122, 241107,  DOI: 10.1063/5.0153891
  37. 37
    Zanotto, S.; Mezzapesa, F. P.; Bianco, F.; Biasiol, G.; Baldacci, L.; Vitiello, M. S.; Sorba, L.; Colombelli, R.; Tredicucci, A. Perfect energy-feeding into strongly coupled systems and interferometric control of polariton absorption. Nat. Phys. 2014, 10, 830834,  DOI: 10.1038/nphys3106
  38. 38
    Jeannin, M.; Manceau, J.-M.; Colombelli, R. Unified description of saturation and bistability of intersubband transitions in the weak and strong light-matter coupling regimes. Phys. Rev. Lett. 2021, 127, 187401  DOI: 10.1103/PhysRevLett.127.187401
  39. 39
    Cominotti, R.; Leymann, H.; Nespolo, J.; Manceau, J.-M.; Jeannin, M.; Colombelli, R.; Carusotto, I. Theory of coherent optical nonlinearities of intersubband transitions in semiconductor quantum wells. Phys. Rev. B 2023, 107, 115431  DOI: 10.1103/PhysRevB.107.115431
  40. 40
    Baumberg, J. J.; Aizpurua, J.; Mikkelsen, M. H.; Smith, D. R. Extreme nanophotonics from ultrathin metallic gaps. Nat. Mater. 2019, 18, 668678,  DOI: 10.1038/s41563-019-0290-y
  41. 41
    Johnson, P. B.; Christy, R.-W. Optical constants of the noble metals. Phys. Rev. B 1972, 6, 4370,  DOI: 10.1103/PhysRevB.6.4370
  42. 42
    Moreau, A.; Ciracì, C.; Mock, J. J.; Hill, R. T.; Wang, Q.; Wiley, B. J.; Chilkoti, A.; Smith, D. R. Controlled-reflectance surfaces with film-coupled colloidal nanoantennas. Nature 2012, 492, 8689,  DOI: 10.1038/nature11615
  43. 43
    Hu, H.; Shi, Z.; Zhang, S.; Xu, H. Unified treatment of scattering, absorption, and luminescence spectra from a plasmon–exciton hybrid by temporal coupled-mode theory. J. Chem. Phys. 2021, 155, 074104  DOI: 10.1063/5.0059816
  44. 44
    Silvestri, M.; Sahoo, A.; Assogna, L.; Benassi, P.; Ferrante, C.; Ciattoni, A.; Marini, A. Resonant third-harmonic generation driven by out-of-equilibrium electron dynamics in sodium-based near-zero index thin films. Nanophotonics 2024, 13, 20032013,  DOI: 10.1515/nanoph-2023-0743
  45. 45
    Huang, Z.; Baron, A.; Larouche, S.; Argyropoulos, C.; Smith, D. R. Optical bistability with film-coupled metasurfaces. Opt. Lett. 2015, 40, 56385641,  DOI: 10.1364/OL.40.005638
  46. 46
    Argyropoulos, C.; Chen, P.-Y.; D’Aguanno, G.; Engheta, N.; Alu, A. Boosting optical nonlinearities in ε-near-zero plasmonic channels. Phys. Rev. B 2012, 85, 045129  DOI: 10.1103/PhysRevB.85.045129
  47. 47
    Hu, H.; Álvarez-Pérez, G.; Otomalo, T. O.; Ciracì, C. Low-power threshold optical bistability enabled by hydrodynamic Kerr nonlinearity of free-carriers in heavily doped semiconductors. arXiv preprint 2024, arXiv:2406.07750, https://arxiv.org/abs/2406.07750 (accessed Jun 11, 2024).

Cited By

Click to copy section linkSection link copied!

This article has not yet been cited by other publications.

ACS Photonics

Cite this: ACS Photonics 2024, 11, 11, 4812–4817
Click to copy citationCitation copied!
https://doi.org/10.1021/acsphotonics.4c01308
Published September 19, 2024

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

740

Altmetric

-

Citations

-
Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Optical absorption of a heavily doped InGaAs thin slab. (a) Schematic. (b) Linear absorption spectra of the thin slab with a local response approximation model (LRA, cyan) and nonlocal hydrodynamic theory (HT, purple). (c) Zoom-in plot of the higher-order LBPs. (d) Induced charge ∇·P and electric field of LBPs with an order of n = 1, 3, 5.

    Figure 2

    Figure 2. Hybrid bulk-plasmon slab-nanopatch system. (a) Schematic of the hybrid structure; height h and periodicity Px of the antenna are 50 and 200 nm, respectively. Length, L, can vary. Bottom panels: the induced charges and electric field of the LBP in the thin spacer. (b) Strong coupling between LSP and LBP with a clear Rabi splitting in reflectance spectra, where three particular cases of different L marked by dashed lines are shown in (c) for details. Gray shades in (c) represent the reflectance of the bare nanopatch without nonlocality. Reflectance spectrum of the on-resonance case (L = 124 nm) is fitted with a Rabi splitting of Ω = 33.5 meV. Green shades mark the center wavelengths of the pump for probing Kerr nonlinearities.

    Figure 3

    Figure 3. Hysteresis loop of two-dimensional nonlinear nanopatch systems. Prominent optical bistabilities with two particular antenna sizes: (a) L = 110 nm and (b) L = 135 nm. (c) Nonlinear reflection (pulse distortion) due to the bistability with L = 135 nm.

  • References


    This article references 47 other publications.

    1. 1
      Boyd, R. W. Nonlinear Optics., 3rd ed.; Academic Press, Inc.: USA, 2008.
    2. 2
      Krasavin, A. V.; Ginzburg, P.; Zayats, A. V. Free-electron optical nonlinearities in plasmonic nanostructures: a review of the hydrodynamic description. Laser Photonics Rev. 2018, 12, 1700082  DOI: 10.1002/lpor.201700082
    3. 3
      Almeida, V. R.; Lipson, M. Optical bistability on a silicon chip. Opt. Lett. 2004, 29, 23872389,  DOI: 10.1364/OL.29.002387
    4. 4
      Gibbs, H.; Tarng, S.; Jewell, J.; Weinberger, D.; Tai, K.; Gossard, A.; McCall, S.; Passner, A.; Wiegmann, W. Room-temperature excitonic optical bistability in a GaAs-GaAlAs superlattice étalon. Appl. Phys. Lett. 1982, 41, 221222,  DOI: 10.1063/1.93490
    5. 5
      Un, I.-W.; Sarkar, S.; Sivan, Y. Electronic-based model of the optical nonlinearity of low-electron-density Drude materials. Phys. Rev. Appl. 2023, 19, 044043  DOI: 10.1103/PhysRevApplied.19.044043
    6. 6
      Cox, J. D.; García, J.; de Abajo, F. Electrically tunable nonlinear plasmonics in graphene nanoislands. Nat. Commun. 2014, 5, 5725,  DOI: 10.1038/ncomms6725
    7. 7
      Christensen, T.; Yan, W.; Jauho, A.-P.; Wubs, M.; Mortensen, N. A. Kerr nonlinearity and plasmonic bistability in graphene nanoribbons. Phys. Rev. B 2015, 92, 121407  DOI: 10.1103/PhysRevB.92.121407
    8. 8
      Cox, J. D.; Silveiro, I.; García de Abajo, F. J. Quantum effects in the nonlinear response of graphene plasmons. ACS Nano 2016, 10, 19952003,  DOI: 10.1021/acsnano.5b06110
    9. 9
      Ryabov, D.; Pashina, O.; Zograf, G.; Makarov, S.; Petrov, M. Nonlinear optical heating of all-dielectric super-cavity: efficient light-to-heat conversion through giant thermorefractive bistability. Nanophotonics 2022, 11, 39813991,  DOI: 10.1515/nanoph-2022-0074
    10. 10
      Niu, J.; Shao, H.; Feng, Y.; Gao, B.; Zhang, Y.; Li, Y.; Chen, H.; Qian, H. All-Optical Nonlinear Neuron Based on Metallic Quantum Wells. Adv. Opt. Mater. 2023, 11, 2300223  DOI: 10.1002/adom.202300223
    11. 11
      Lee, J.; Tymchenko, M.; Argyropoulos, C.; Chen, P.-Y.; Lu, F.; Demmerle, F.; Boehm, G.; Amann, M.-C.; Alu, A.; Belkin, M. A. Giant nonlinear response from plasmonic metasurfaces coupled to intersubband transitions. Nature 2014, 511, 6569,  DOI: 10.1038/nature13455
    12. 12
      Taliercio, T.; Biagioni, P. Semiconductor infrared plasmonics. Nanophotonics 2019, 8, 949990,  DOI: 10.1515/nanoph-2019-0077
    13. 13
      Scalora, M.; Vincenti, M. A.; de Ceglia, D.; Roppo, V.; Centini, M.; Akozbek, N.; Bloemer, M. Second-and third-harmonic generation in metal-based structures. Phys. Rev. A 2010, 82, 043828  DOI: 10.1103/PhysRevA.82.043828
    14. 14
      Ciracì, C.; Poutrina, E.; Scalora, M.; Smith, D. R. Second-harmonic generation in metallic nanoparticles: Clarification of the role of the surface. Phys. Rev. B 2012, 86, 115451  DOI: 10.1103/PhysRevB.86.115451
    15. 15
      Khalid, M.; Ciracì, C. Enhancing second-harmonic generation with electron spill-out at metallic surfaces. Commun. Phys. 2020, 3, 214,  DOI: 10.1038/s42005-020-00477-0
    16. 16
      De Luca, F.; Ortolani, M.; Ciracì, C. Free electron nonlinearities in heavily doped semiconductors plasmonics. Phys. Rev. B 2021, 103, 115305  DOI: 10.1103/PhysRevB.103.115305
    17. 17
      De Luca, F.; Ciracì, C. Impact of surface charge depletion on the free electron nonlinear response of heavily doped semiconductors. Phys. Rev. Lett. 2022, 129, 123902  DOI: 10.1103/PhysRevLett.129.123902
    18. 18
      De Luca, F.; Ortolani, M.; Ciracì, C. Free electron harmonic generation in heavily doped semiconductors: the role of the materials properties. EPJ. Appl. Metamat. 2022, 9, 13,  DOI: 10.1051/epjam/2022011
    19. 19
      Rossetti, A. Origin of optical nonlinearity in plasmonic semiconductor nanostructures. arXiv preprint 2024, arXiv:2402.15443, https://arxiv.org/abs/2402.15443 (accessed Feb 23, 2024).
    20. 20
      Ruppin, R. Optical properties of a plasma sphere. Phys. Rev. Lett. 1973, 31, 1434,  DOI: 10.1103/PhysRevLett.31.1434
    21. 21
      Ruppin, R. Extinction properties of thin metallic nanowires. Opt. Commun. 2001, 190, 205209,  DOI: 10.1016/S0030-4018(01)01063-X
    22. 22
      Raza, S.; Toscano, G.; Jauho, A.-P.; Wubs, M.; Mortensen, N. A. Unusual resonances in nanoplasmonic structures due to nonlocal response. Phys. Rev. B 2011, 84, 121412  DOI: 10.1103/PhysRevB.84.121412
    23. 23
      Benisty, H.; Greffet, J.-J.; Lalanne, P. Introduction to nanophotonics; Oxford university press, 2022.
    24. 24
      Anderegg, M.; Feuerbacher, B.; Fitton, B. Optically excited longitudinal plasmons in potassium. Phys. Rev. Lett. 1971, 27, 1565,  DOI: 10.1103/PhysRevLett.27.1565
    25. 25
      Lindau, I.; Nilsson, P. Experimental verification of optically excited longitudinal plasmons. Phys. Scr. 1971, 3, 87,  DOI: 10.1088/0031-8949/3/2/007
    26. 26
      De Ceglia, D.; Scalora, M.; Vincenti, M. A.; Campione, S.; Kelley, K.; Runnerstrom, E. L.; Maria, J.-P.; Keeler, G. A.; Luk, T. S. Viscoelastic optical nonlocality of low-loss epsilon-near-zero nanofilms. Sci. Rep. 2018, 8, 9335,  DOI: 10.1038/s41598-018-27655-z
    27. 27
      Christensen, T.; Yan, W.; Raza, S.; Jauho, A.-P.; Mortensen, N. A.; Wubs, M. Nonlocal response of metallic nanospheres probed by light, electrons, and atoms. ACS Nano 2014, 8, 17451758,  DOI: 10.1021/nn406153k
    28. 28
      Ciracì, C.; Della Sala, F. Quantum hydrodynamic theory for plasmonics: Impact of the electron density tail. Phys. Rev. B 2016, 93, 205405  DOI: 10.1103/PhysRevB.93.205405
    29. 29
      Toscano, G.; Straubel, J.; Kwiatkowski, A.; Rockstuhl, C.; Evers, F.; Xu, H.; Asger Mortensen, N.; Wubs, M. Resonance shifts and spill-out effects in self-consistent hydrodynamic nanoplasmonics. Nat. Commun. 2015, 6, 7132,  DOI: 10.1038/ncomms8132
    30. 30
      Ciracì, C. Current-dependent potential for nonlocal absorption in quantum hydrodynamic theory. Phys. Rev. B 2017, 95, 245434  DOI: 10.1103/PhysRevB.95.245434
    31. 31
      Silberberg, Y.; Sands, T. Optical properties of metallic quantum wells. IEEE J. Quantum Electron. 1992, 28, 16631669,  DOI: 10.1109/3.142553
    32. 32
      Qian, H.; Xiao, Y.; Liu, Z. Giant Kerr response of ultrathin gold films from quantum size effect. Nat. Commun. 2016, 7, 13153  DOI: 10.1038/ncomms13153
    33. 33
      Qian, H.; Li, S.; Chen, C.-F.; Hsu, S.-W.; Bopp, S. E.; Ma, Q.; Tao, A. R.; Liu, Z. Large optical nonlinearity enabled by coupled metallic quantum wells. Light Sci. Appl. 2019, 8, 13,  DOI: 10.1038/s41377-019-0123-4
    34. 34
      Pirotta, S.; Tran, N.-L.; Jollivet, A.; Biasiol, G.; Crozat, P.; Manceau, J.-M.; Bousseksou, A.; Colombelli, R. Fast amplitude modulation up to 1.5 GHz of mid-IR free-space beams at room-temperature. Nat. Commun. 2021, 12, 799,  DOI: 10.1038/s41467-020-20710-2
    35. 35
      Malerba, M.; Pirotta, S.; Aubin, G.; Lucia, L.; Jeannin, M.; Manceau, J.-M.; Bousseksou, A.; Lin, Q.; Lampin, J.-F.; Peytavit, E.; Barbieri, S.; Li, L.; Davies, G.; Linfield, E. H.; Colombelli, R. Ultrafast (≈ 10 GHz) mid-IR modulator based on ultrafast electrical switching of the light–matter coupling. Appl. Phys. Lett. 2024, 125, 041101  DOI: 10.1063/5.0213965
    36. 36
      Jeannin, M.; Cosentino, E.; Pirotta, S.; Malerba, M.; Biasiol, G.; Manceau, J.-M.; Colombelli, R. Low intensity saturation of an ISB transition by a mid-IR quantum cascade laser. Appl. Phys. Lett. 2023, 122, 241107,  DOI: 10.1063/5.0153891
    37. 37
      Zanotto, S.; Mezzapesa, F. P.; Bianco, F.; Biasiol, G.; Baldacci, L.; Vitiello, M. S.; Sorba, L.; Colombelli, R.; Tredicucci, A. Perfect energy-feeding into strongly coupled systems and interferometric control of polariton absorption. Nat. Phys. 2014, 10, 830834,  DOI: 10.1038/nphys3106
    38. 38
      Jeannin, M.; Manceau, J.-M.; Colombelli, R. Unified description of saturation and bistability of intersubband transitions in the weak and strong light-matter coupling regimes. Phys. Rev. Lett. 2021, 127, 187401  DOI: 10.1103/PhysRevLett.127.187401
    39. 39
      Cominotti, R.; Leymann, H.; Nespolo, J.; Manceau, J.-M.; Jeannin, M.; Colombelli, R.; Carusotto, I. Theory of coherent optical nonlinearities of intersubband transitions in semiconductor quantum wells. Phys. Rev. B 2023, 107, 115431  DOI: 10.1103/PhysRevB.107.115431
    40. 40
      Baumberg, J. J.; Aizpurua, J.; Mikkelsen, M. H.; Smith, D. R. Extreme nanophotonics from ultrathin metallic gaps. Nat. Mater. 2019, 18, 668678,  DOI: 10.1038/s41563-019-0290-y
    41. 41
      Johnson, P. B.; Christy, R.-W. Optical constants of the noble metals. Phys. Rev. B 1972, 6, 4370,  DOI: 10.1103/PhysRevB.6.4370
    42. 42
      Moreau, A.; Ciracì, C.; Mock, J. J.; Hill, R. T.; Wang, Q.; Wiley, B. J.; Chilkoti, A.; Smith, D. R. Controlled-reflectance surfaces with film-coupled colloidal nanoantennas. Nature 2012, 492, 8689,  DOI: 10.1038/nature11615
    43. 43
      Hu, H.; Shi, Z.; Zhang, S.; Xu, H. Unified treatment of scattering, absorption, and luminescence spectra from a plasmon–exciton hybrid by temporal coupled-mode theory. J. Chem. Phys. 2021, 155, 074104  DOI: 10.1063/5.0059816
    44. 44
      Silvestri, M.; Sahoo, A.; Assogna, L.; Benassi, P.; Ferrante, C.; Ciattoni, A.; Marini, A. Resonant third-harmonic generation driven by out-of-equilibrium electron dynamics in sodium-based near-zero index thin films. Nanophotonics 2024, 13, 20032013,  DOI: 10.1515/nanoph-2023-0743
    45. 45
      Huang, Z.; Baron, A.; Larouche, S.; Argyropoulos, C.; Smith, D. R. Optical bistability with film-coupled metasurfaces. Opt. Lett. 2015, 40, 56385641,  DOI: 10.1364/OL.40.005638
    46. 46
      Argyropoulos, C.; Chen, P.-Y.; D’Aguanno, G.; Engheta, N.; Alu, A. Boosting optical nonlinearities in ε-near-zero plasmonic channels. Phys. Rev. B 2012, 85, 045129  DOI: 10.1103/PhysRevB.85.045129
    47. 47
      Hu, H.; Álvarez-Pérez, G.; Otomalo, T. O.; Ciracì, C. Low-power threshold optical bistability enabled by hydrodynamic Kerr nonlinearity of free-carriers in heavily doped semiconductors. arXiv preprint 2024, arXiv:2406.07750, https://arxiv.org/abs/2406.07750 (accessed Jun 11, 2024).
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsphotonics.4c01308.

    • Derivation and implementation method of hydrodynamic nonlinear sources and contributions from dielectric χ(3) in the nonlinear reflectance, hydrodynamic contributions from silver and heavily doped InGaAs, robustness of the system with different dampings, and switching time (PDF)

    • Video of temporal evolution of the nonlinear system (MP4)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.