Fact-checked by Grok 2 weeks ago

Plasmon

A plasmon is a that represents the quantum of a , arising from the collective motion of free in a conducting medium such as a metal or . These oscillations occur when the electron density in the material is displaced from equilibrium, creating a restoring force due to the resulting , analogous to but on a quantum scale. The concept was formalized by physicist in 1956, who introduced the term "plasmon" to describe these elementary excitations in the electron gas of solids. Plasmons exist in two primary forms: bulk (or volume) plasmons, which are longitudinal oscillations propagating throughout the interior of the material, and surface plasmons, which are confined to the between the and a medium. Bulk plasmons typically resonate at the plasma frequency, determined by the and effective mass, often in the range for noble metals like silver and (around 9 eV after accounting for interband transitions). Surface plasmons, first theoretically predicted by Rufus Ritchie in 1957 through studies of electron energy losses in thin films, couple strongly with electromagnetic waves to form surface plasmon (SPPs), hybrid modes that propagate along the with enhanced fields evanescently decaying perpendicular to it. These require a negative real part of the function (Re(ε) < 0) and low absorption for efficient excitation, conditions met by noble metals in the visible and near-infrared spectrum. The study of plasmons, known as plasmonics, leverages their ability to confine and enhance electromagnetic fields to subwavelength scales, enabling intense light-matter interactions far beyond classical diffraction limits. Key properties include sharp resonances tunable by material composition, geometry (e.g., nanoparticles or nanostructures), and surrounding environment, leading to applications in nanophotonics, biosensing, and spectroscopy. For instance, localized surface plasmons on metallic nanoparticles produce hot spots for surface-enhanced Raman scattering (SERS), amplifying molecular signals by factors up to 10¹⁴, while propagating SPPs underpin surface plasmon resonance (SPR) sensors for real-time detection of biomolecular binding with refractive index sensitivities on the order of 10⁻⁶ RIU. Since the experimental confirmation of surface plasmons in 1959, the field has grown rapidly, with plasmonic technologies integral to over 25% of biosensor research as of 2023.

Fundamentals

Definition

A plasmon is a quasiparticle representing the quantum of plasma oscillations in the free electron gas of metals, semiconductors, or other conducting media. These oscillations arise from the collective motion of conduction electrons, treated as a plasma-like fluid. In contrast to individual electron excitations, which involve single-particle transitions, plasmons describe coherent density waves where many electrons oscillate in phase against a fixed background of positive ions. This collective nature emerges from long-range Coulomb interactions within the electron gas, leading to quantized modes with energy E_p = \hbar \omega_p, where \omega_p is the plasma frequency characteristic of the material. Plasmons are observed in various materials, including noble metals such as and , where they enable strong optical responses in the visible range; semiconductors like , supporting tunable excitations; and two-dimensional systems like , exhibiting highly confined modes. These quasiparticles form a foundational concept for understanding electromagnetic interactions at the nanoscale, underpinning the field of . Particular manifestations, such as at material interfaces, extend these collective effects to boundary-confined geometries.

History

The vibrant red and purple hues observed in medieval stained glass, such as those in the Gothic windows of in Paris, were produced by embedding gold nanoparticles in the glass matrix, where their localized surface plasmons generated intense colors through light scattering and absorption. In 1902, Robert W. Wood reported the first experimental observation of what became known as , a sharp variation in the intensity of diffracted light from metallic gratings exposed to ultraviolet radiation, later recognized as an early indication of plasmonic effects at metal surfaces. Building on classical electron theory, Paul Drude introduced his electron gas model in 1900, treating conduction electrons in metals as a free gas that provided the foundational framework for understanding collective plasma-like behaviors in solids. The concept of plasma oscillations was formalized in 1928 by Irving Langmuir, who described electronic oscillations in ionized gases during studies of gas discharges, coining the term "plasma" and establishing the basis for plasmon excitations in dense electron systems. A pivotal theoretical advancement occurred in 1952 when David Pines and David Bohm proposed plasmons as quantized collective modes—or quasiparticles—arising from density fluctuations in the electron gas of solids, bridging plasma physics with solid-state theory and enabling the treatment of long-range Coulomb interactions. Experimental progress in the 1960s enabled direct observation of surface plasmons; in 1968, E. Kretschmann and H. Raether demonstrated excitation via the using attenuated total reflection in a prism-metal setup, while A. Otto independently developed the for coupling light to non-radiative surface plasmons on thin metal films. The field of plasmonics emerged prominently after 2000, driven by nanoscale fabrication advances that harnessed plasmons for subwavelength optics and enhanced light-matter interactions, as highlighted in Harry A. Atwater's 2007 review outlining its potential for transformative applications in . By the 2020s, research had advanced to highly tunable plasmons in and other two-dimensional materials, enabling efficient terahertz wave manipulation for detectors, modulators, and emitters, with key demonstrations of strong light confinement and low-loss propagation in hybrid structures.

Theoretical Foundations

Derivation of Plasma Frequency

The derivation of the plasma frequency begins with the classical description of a free electron gas in a uniform positive ion background, modeling the behavior of electrons in metals or plasmas under the framework of electrodynamics. This approach assumes non-interacting electrons and initially neglects thermal effects and magnetic fields, focusing on longitudinal electrostatic oscillations. The key equations are Maxwell's equations, particularly Gauss's law, combined with the continuity equation for charge conservation and the equation of motion for electrons. Consider a small perturbation in the electron density from the equilibrium value n_0: n = n_0 + n_1 e^{i(\mathbf{k} \cdot \mathbf{r} - \omega t)}, where n_1 is the amplitude of the perturbation, \mathbf{k} is the wave vector, \mathbf{r} is the position, \omega is the angular frequency, and t is time. The velocity perturbation \mathbf{v} of electrons satisfies the linearized equation of motion, treating electrons as a fluid: m \frac{d\mathbf{v}}{dt} = -e \mathbf{E}, where m is the electron mass, e is the elementary charge magnitude, and \mathbf{E} is the induced electric field. For plane-wave perturbations, this yields \mathbf{v} = \frac{e \mathbf{E}}{i m \omega}. The current density due to this motion is \mathbf{j} = -e n_0 \mathbf{v}, and the continuity equation \frac{\partial n_1}{\partial t} + \nabla \cdot (n_0 \mathbf{v}) = 0 relates the density and velocity perturbations: -i \omega n_1 + i \mathbf{k} \cdot (n_0 \mathbf{v}) = 0, leading to n_1 = \frac{n_0 e \mathbf{k} \cdot \mathbf{E}}{m \omega^2}. From Poisson's equation in the electrostatic approximation, \nabla \cdot \mathbf{E} = -\frac{e n_1}{\epsilon_0}, where \epsilon_0 is the vacuum permittivity, substituting the expression for n_1 gives the dispersion relation for longitudinal waves in the cold plasma limit: \omega^2 = \omega_p^2. Including thermal effects via a pressure gradient term in the momentum equation yields the Bohm-Gross dispersion relation \omega^2 = \omega_p^2 + 3 v_{th}^2 k^2, where v_{th} = \sqrt{\frac{k_B T}{m}} is the thermal velocity (k_B is Boltzmann's constant and T is temperature). In the long-wavelength limit (k \to 0), thermal effects vanish, yielding \omega = \omega_p. The plasma frequency is given by \omega_p = \sqrt{\frac{n_0 e^2}{\epsilon_0 m}}, where n_0 is the equilibrium electron density, e the electron charge, m the electron mass, and \epsilon_0 the permittivity of free space; this represents the natural oscillation frequency of the electron gas under restoring electrostatic forces. While the classical treatment describes collective electron oscillations as waves, the quantum mechanical extension treats plasmons as quantized excitations. Bohm and Pines introduced a collective variable approach using second quantization to describe the electron gas, transforming individual particle interactions into collective plasmon modes with energy \hbar \omega_p per quantum, enabling a many-body description of the system. This quantization bridges classical plasma oscillations to quantum plasmons in solids, assuming a degenerate of non-interacting electrons.

Bulk Plasmons

Bulk plasmons represent collective volume oscillations of the conduction electrons in a homogeneous, infinite electron gas, such as that found in the interior of metals, where the electrons coherently oscillate relative to the fixed positive ion background. These modes are characterized by a uniform density fluctuation throughout the bulk material and occur at the plasma frequency \omega_p, defined as the natural oscillation frequency of the system. As longitudinal excitations, bulk plasmons propagate with the electric field parallel to the wave vector and do not couple directly to transverse electromagnetic radiation, rendering them non-radiative for frequencies \omega > \omega_p. This longitudinal nature arises from the requirement that the of the is non-zero, distinguishing them from radiative photonic modes. Excitation of bulk plasmons is therefore achieved through non-optical methods, primarily inelastic via (EELS), where high-energy electrons interact with the sample and lose discrete amounts of energy corresponding to plasmon creation in thin metallic foils. In metals, bulk plasmon energies typically fall in the range of 5-25 , depending on the density; for instance, in silver, the energy is approximately 9 . These high-energy modes contribute significantly to the of metals, as described by the dielectric function: \epsilon(\omega) = \epsilon_\infty - \frac{\omega_p^2}{\omega(\omega + i\gamma)}, where \epsilon_\infty is the high-frequency constant, \gamma is the damping rate, and for \omega < \omega_p, the real part of \epsilon(\omega) becomes negative, resulting in high reflectivity and the metallic luster observed below the plasma frequency. From a quantum perspective, bulk plasmons are treated as bosonic quasiparticles within the many-body framework of the electron gas, where the second-quantized Hamiltonian for the collective excitations is expressed using bosonic creation (b^\dagger) and annihilation (b) operators that obey the commutation relations [b, b^\dagger] = 1. This quantization allows plasmons to be described as non-interacting bosons in the harmonic approximation, facilitating the study of their interactions with other quasiparticles. In semiconductors, bulk plasmons manifest at much lower energies due to reduced carrier densities; for example, in n-doped with carrier concentration around $10^{17} cm^{-3}, the plasmon energy is on the order of 16 meV, observable via far-infrared spectroscopy.

Surface and Localized Plasmons

Surface Plasmon Polaritons

Surface plasmon polaritons (SPPs) are electromagnetic surface waves that propagate along the interface between a metal and a dielectric medium, arising from the coupling of electromagnetic fields with collective oscillations of free electrons at the boundary. These waves are evanescently confined to the interface, with the electromagnetic fields decaying exponentially away from the boundary into both the metal and the dielectric, typically on the order of hundreds of nanometers. First theoretically described in the context of electron energy loss in thin metal films, SPPs exhibit a hybrid nature, combining photon-like and plasmon-like characteristics, distinct from bulk plasmons which occur as volume modes. The dispersion relation for SPPs at a flat metal-dielectric interface is given by k_{sp} = \frac{\omega}{c} \sqrt{\frac{\epsilon_m \epsilon_d}{\epsilon_m + \epsilon_d}}, where k_{sp} is the SPP wavevector parallel to the interface, \omega is the angular frequency, c is the speed of light in vacuum, \epsilon_m(\omega) is the frequency-dependent permittivity of the metal (often modeled using the Drude form \epsilon_m = 1 - \frac{\omega_p^2}{\omega^2 + i\gamma\omega}, with \omega_p the bulk plasma frequency and \gamma the damping rate), and \epsilon_d is the dielectric constant of the surrounding medium (assumed real and positive). This relation shows that k_{sp} exceeds the free-space light wavevector \frac{\omega}{c}, making direct excitation by far-field light challenging due to momentum mismatch. As frequency increases, k_{sp} increases and the dispersion curve bends away from the light line, approaching an asymptotic frequency \omega_{sp} = \frac{\omega_p}{\sqrt{1 + \epsilon_d}} at large wavevectors, where SPPs remain non-radiative due to the momentum mismatch with free-space light. SPPs require transverse magnetic (TM) polarization, with the magnetic field parallel to the interface and the electric field components both parallel and perpendicular to the propagation direction, enabling the necessary boundary conditions for field continuity and charge accumulation at the interface. Common methods to excite SPPs overcome the momentum mismatch through prism coupling in the , where a thin metal film is deposited on a high-refractive-index prism to allow evanescent wave tunneling from the prism side, or the , which uses a dielectric gap between the prism and metal surface for similar frustrated total internal reflection. Grating coupling provides an alternative by diffracting incident light to match the SPP wavevector via periodic surface structures on the metal. The propagation length of SPPs, defined as L_{sp} = \frac{1}{2 \operatorname{Im}(k_{sp})}, is limited by intrinsic material losses, including ohmic damping from electron scattering in the metal. At visible wavelengths (e.g., around 785 nm), this length is typically on the order of several to tens of microns for noble metals like silver or gold, representing a key practical constraint for applications requiring long-distance waveguiding.

Localized Surface Plasmons

Localized surface plasmons (LSPs) are non-propagating electromagnetic resonances arising from the coherent oscillation of conduction electrons in metallic nanostructures smaller than the wavelength of light, such as gold nanospheres. These modes confine light to subwavelength volumes, contrasting with propagating surface plasmon polaritons at extended interfaces. In nanoparticles, incident light drives dipole moments that oscillate at the plasmon frequency, leading to strong scattering and absorption peaks in the visible or near-infrared spectrum. For spherical particles, Mie theory provides an exact solution to Maxwell's equations, describing the optical response through multipolar expansions. In the quasi-static limit for small spheres (radius r \ll \lambda), the dipole polarizability simplifies to \alpha = 4\pi r^3 \frac{\epsilon_m - \epsilon_d}{\epsilon_m + 2\epsilon_d}, where \epsilon_m and \epsilon_d are the dielectric functions of the metal and surrounding medium, respectively. The resonance condition occurs when the real part of \epsilon_m = -2 \epsilon_d, maximizing the polarizability and resulting in a sharp extinction peak. The resonance wavelength of LSPs depends strongly on nanoparticle size and shape. For gold nanospheres, increasing the diameter from 20 nm to 100 nm causes a redshift of the plasmon peak from approximately 520 nm to 570 nm due to radiative and dynamic depolarization effects. Non-spherical shapes, such as nanorods, enable further tuning; higher aspect ratios shift the longitudinal plasmon mode to longer wavelengths in the near-infrared, while the transverse mode remains near the spherical resonance. LSP excitation produces intense local electric field enhancements near the nanoparticle surface, with magnitudes up to 1000 times the incident field in optimized configurations, facilitating enhanced light-matter interactions like surface-enhanced Raman scattering. These hotspots arise from the collective electron response and are particularly pronounced at sharp features or junctions. Gold and silver nanoparticles supporting LSPs are commonly fabricated via chemical synthesis methods, such as citrate reduction for uniform spheres (), or electron-beam lithography for precise patterning of complex shapes. These techniques yield monodisperse colloids or substrate-bound arrays with sizes from 10 to 200 nm, essential for reproducible plasmonic responses. In two-dimensional materials, LSPs manifest in patterned graphene nanostructures like disks, where Dirac fermions support highly confined mid-infrared plasmons tunable over a wide range. Applying an electrostatic gate voltage modulates the Fermi level, shifting the resonance frequency proportionally to the square root of the carrier density, enabling dynamic control without structural changes.

Properties and Interactions

Dispersion Relations

In plasmonics, the dispersion relation describes how the frequency \omega of plasmonic modes varies with the in-plane wavevector k, providing insight into their propagation characteristics and coupling with photons. This relation unifies the behaviors of different plasmon types, from bulk to surface and localized modes, and is fundamental to understanding confinement and tunability. For ideal lossless cases, the dispersion reveals key features such as asymptotic limits and branch structures, while real systems incorporate material parameters like permittivity. For bulk plasmons in three-dimensional electron gases, the dispersion is nearly flat at the plasma frequency \omega_p for small wavevectors k \ll \omega_p / v_F, where v_F is the Fermi velocity, reflecting the longitudinal, non-radiative nature of these volume modes. Beyond this random-phase approximation regime, the frequency rises parabolically due to Fermi pressure effects, approximated as \omega(k) \approx \omega_p + \frac{3}{10} \frac{v_F^2 k^2}{\omega_p}. This correction, derived from the Lindhard dielectric function with hydrodynamic models providing a classical analog, becomes significant at larger k, enabling coupling to other excitations but remaining gapped from the light line. Surface plasmon polaritons (SPPs) at a metal-dielectric interface exhibit a dispersion that originates near the light line \omega = c k / \sqrt{\epsilon_d} for small k, where c is the speed of light and \epsilon_d the dielectric permittivity, allowing photonic-like propagation. As k increases, the curve bends away from the light line toward an asymptotic surface plasmon frequency \omega_{sp} = \omega_p / \sqrt{1 + \epsilon_d}, forming a plasmonic branch distinct from the photonic one. This hyperbolic-like relation, \omega(k) = c k \sqrt{\epsilon_m \epsilon_d / (\epsilon_m + \epsilon_d)} with Drude metal permittivity \epsilon_m = 1 - \omega_p^2 / \omega^2, arises from boundary conditions on evanescent fields and prevents direct optical excitation without momentum compensation. Localized surface plasmons (LSPs) in subwavelength nanoparticles lack intrinsic wavevector dependence due to their confinement within the particle volume, treated under the quasistatic approximation where retardation is negligible. Instead, the resonance frequency broadens and shifts with particle size, as smaller dimensions enhance restoring forces and increase \omega_{LSP} toward \omega_p / \sqrt{3} for spheres in vacuum, while larger sizes introduce effective dispersion via multipolar modes or interparticle coupling. This size-dependent broadening, quantified by Lorentzian linewidths, stems from the dipole moment induced by uniform fields, limiting LSPs to zero-momentum excitations observable in far-field scattering. In two-dimensional systems like , plasmons display a distinct square-root dispersion \omega \propto \sqrt{k} at long wavelengths, driven by the linear Dirac spectrum of charge carriers and described via random-phase approximation. This acoustic-like scaling, \omega(q) \approx \sqrt{ \frac{e^2 E_F q}{2 \hbar^2 \epsilon_0 \kappa} } with Fermi energy E_F and effective dielectric \kappa, enables strong subwavelength confinement and tunability via gating, contrasting the gapped 3D case and arising from the density of states in . Retardation effects, arising from finite light speed, distinguish non-retarded electrostatic approximations—valid for structures much smaller than the wavelength, yielding local dipole interactions—from full electromagnetic treatments required for larger scales. In the non-retarded limit, dispersion follows Coulombic interactions without phase delays, but full Maxwell solutions introduce radiative corrections that flatten or modify curves, particularly for transverse modes in nanoparticle chains, enhancing propagation lengths while negligible for longitudinal bulk modes. For complex geometries beyond analytical tractability, numerical methods like finite-difference time-domain (FDTD) and finite-element method (FEM) compute dispersion by solving Maxwell's equations on discretized grids. FDTD excels in time-domain simulations of broadband responses, capturing evanescent fields in plasmonic nanostructures such as dimers, while FEM provides modal analysis for frequency-dependent effective indices, both incorporating dispersive permittivities via auxiliary equations. These approaches reveal geometry-induced band structures, with FDTD showing field enhancements up to 15 in nanogaps, validating against quasistatic models.

Damping Mechanisms

Plasmons, as collective oscillations of electrons in metals, are inherently subject to various damping mechanisms that lead to energy dissipation and limit their propagation length and coherence time. These losses arise from interactions within the electron gas, with the surrounding environment, and through radiative processes, ultimately broadening the plasmon resonance and reducing efficiency in plasmonic devices. Ohmic or intraband losses represent a primary non-radiative damping pathway, stemming from resistive heating due to electron scattering events. Electron-phonon scattering contributes significantly, where the oscillating electrons exchange energy with lattice vibrations, with the damping rate increasing with temperature as phonon populations rise. Additionally, occurs when the plasmon wavevector k exceeds \omega / v_F (where \omega is the plasmon frequency and v_F the Fermi velocity), allowing individual electrons to phase-match and absorb energy from the collective mode, particularly relevant in nanoscale structures where spatial dispersion effects are pronounced. These intraband processes dominate in noble metals at visible frequencies and are described within the framework. Interband transitions provide another key damping channel, especially in metals with complex band structures like gold, where direct electronic excitations across band gaps absorb plasmon energy. In gold, significant absorption begins above the d-band edge at approximately 2.4 eV, overlapping with visible plasmon resonances and causing additional broadening. This mechanism is material-specific and can be minimized by selecting metals like silver, which have interband transitions shifted to higher energies in the ultraviolet range. Radiative damping becomes prominent for plasmonic structures larger than about one-tenth of the excitation wavelength, where the oscillating dipole couples efficiently to far-field photons, leading to energy reradiation rather than confinement. This process scales with particle volume and is more relevant for localized surface plasmons in extended nanoparticles or arrays, competing with non-radiative losses at longer wavelengths. The total plasmon linewidth \gamma_{total} encapsulates these contributions as \gamma_{total} = \gamma_{intrinsic} + \gamma_{radiation} + \gamma_{extrinsic}, where \gamma_{intrinsic} includes ohmic and interband losses, \gamma_{radiation} accounts for photon emission, and \gamma_{extrinsic} arises from environmental effects such as substrate-induced scattering or adsorbate interactions. The quality factor Q = \omega / \gamma_{total}, a measure of resonance sharpness, typically ranges from 10 to 100 for visible-frequency plasmons in metals like gold and silver, reflecting the dominance of these losses. Surface plasmons often display somewhat lower damping rates than bulk plasmons due to reduced volume for intraband scattering, though interface effects can introduce additional extrinsic contributions. To mitigate damping and enhance plasmon coherence, strategies include operating at cryogenic temperatures, which suppress electron-phonon scattering by reducing phonon occupancy, thereby narrowing the linewidth. Furthermore, incorporation of gain media, such as laser dyes in spaser configurations, can compensate losses through stimulated emission into the plasmon mode, enabling net amplification despite intrinsic damping.

Advanced Phenomena

Plasmon-Solitons

Plasmon-solitons represent a class of hybrid electromagnetic waves that integrate the collective electron oscillations of with the self-trapping behavior of , forming self-reinforcing pulses whose shape is preserved during propagation through the nonlinear balance between dispersive spreading and self-phase modulation. These structures arise primarily from as the underlying mode, enhanced by Kerr-type nonlinearities in the host medium. In graphene-based waveguides, terahertz plasmon-solitons emerge as particularly promising due to the material's strong Kerr nonlinearity and tunable carrier density, allowing for the excitation of tightly confined modes in the THz regime. The Kerr effect in graphene induces an intensity-dependent refractive index change, which counters the dispersion inherent to plasmonic propagation, enabling stable soliton formation even in discrete nanodisk arrays or multilayer stacks. The theoretical foundation for plasmon-soliton formation relies on an adapted nonlinear Schrödinger equation (NLSE) that captures the evolution of the pulse envelope \psi along the propagation direction z and time t: i \frac{\partial \psi}{\partial z} + \frac{1}{2k} \frac{\partial^2 \psi}{\partial t^2} + \gamma |\psi|^2 \psi = 0, where k is the wave number and \gamma represents the nonlinear coefficient. A fundamental soliton solution takes the form \psi = \sqrt{\frac{1}{\gamma}} \sech(t) e^{i z / 2k}, illustrating the hyperbolic secant profile that maintains temporal integrity without decay or broadening. This equation, derived under slowly varying envelope approximations, highlights how the cubic nonlinearity \gamma |\psi|^2 \psi term stabilizes against linear dispersion \frac{1}{2k} \frac{\partial^2 \psi}{\partial t^2}. Key advantages of plasmon-solitons include their ability to support ultrashort pulses on the femtosecond scale—down to below 10 fs in self-induced transparency configurations—and sustain high peak intensities without significant spreading or diffraction, surpassing the limitations of linear plasmonic modes. These properties stem from the enhanced light-matter interaction at metal-dielectric interfaces, providing subwavelength confinement while mitigating pulse distortion over propagation distances up to several millimeters at moderate input powers. Experimental realizations of plasmon-solitons were first demonstrated in the 2010s using semiconductor slabs, such as chalcogenide films (e.g., Ge<sub>28.1</sub>Sb<sub>6.3</sub>Se<sub>65.6</sub>) in multilayer planar geometries with interfaces, where Kerr self-focusing reduced beam widths to approximately 19 μm and enhanced nonlinearity by up to eightfold compared to non-plasmonic setups. In the 2020s, progress has shifted toward 2D materials, with simulations and theoretical models confirming viable soliton propagation in nanodisk arrays and hybrid structures, driven by external fields in the THz range, though full experimental verification in these systems remains an active pursuit. A distinguishing feature of plasmon-solitons from conventional plasmons lies in their dual localization: while pure plasmons achieve primarily spatial confinement at interfaces, plasmon-solitons extend this to temporal domain via nonlinear , enabling applications in ultrafast . However, intrinsic from material losses poses a challenge for extending over long distances, necessitating designs with compensation or low-loss 2D hosts like .

Quantum and Hybrid Plasmons

Quantum plasmons emerge when collective oscillations in metallic nanostructures are treated quantum mechanically, particularly in confined systems where plasmon states can be realized. In plasmonic cavities, such as nanoparticle-on-mirror configurations, single plasmon excitations have been observed through strong light-matter interactions with few quantum emitters. For instance, vacuum Rabi splitting has been demonstrated in silver bowtie cavities coupled to quantum dots, revealing hybridized states at the single . This splitting arises from the coherent exchange of energy between the plasmon mode and the quantum emitter, with linewidths indicating strong coupling even for dark plasmon modes. Early quantum treatments of plasmons, as developed by in the , laid the foundation for understanding these collective excitations beyond classical descriptions. Plasmon-exciton form in the regime, where plasmons hybridize with excitons in materials like quantum dots or , leading to anticrossing in their dispersion relations. This regime is characterized by the Rabi splitting energy exceeding the damping rates of both constituents, resulting in upper and lower branches that avoid crossing as a function of or detuning. Seminal observations in thin films on metal substrates confirmed this , with splitting energies up to hundreds of meV, enabling coherent energy transfer over nanoscale distances. In such systems, the exhibit modified dispersion, with group velocities reduced compared to uncoupled modes, facilitating applications in . Spasers, or surface plasmon amplification by stimulated emission of radiation, represent plasmonic nanolasers that amplify s using gain media like quantum dots or dyes, achieving thresholds as low as around 100 plasmons due to subwavelength confinement. These devices operate by into plasmon modes rather than photons, with feedback provided by the plasmonic itself, enabling ultrasmall coherent sources below the limit. Experimental realizations in core-shell nanoparticles have shown room-temperature operation with narrow linewidths, highlighting the role of surface gain in overcoming ohmic losses. Topological plasmons introduce robustness through nontrivial band topology in gyromagnetic plasmas, where edge states propagate unidirectionally and are protected against backscattering from defects or , as theorized in post-2015 frameworks. In magnetized plasma structures, such as cylindrical arrays, these edge modes arise from Chern insulators with nonzero topological invariants, ensuring immunity to imperfections. Bulk-edge correspondence in these systems guarantees the existence of gapless edge plasmons, with dispersion relations exhibiting linear crossings at Dirac points. Plasmon-mediated entanglement enables transfer by coupling plasmons to entangled pairs or states, with experiments in the demonstrating protocols via polaritons. In hybrid waveguides, entanglement between plasmon modes and distant qubits has been achieved through -to-plasmon conversion, preserving quantum correlations over propagation distances. These setups leverage the high speed of plasmons for efficient state transfer, though from damping limits fidelity. Hybrid plasmons with phonons, known as plasmon-phonon polaritons, occur in polar materials like silicon carbide (SiC), where electromagnetic modes couple to optical phonons in the Reststrahlen band, yielding hybrid quasiparticles with mixed dispersion. In SiC nanostructures, these polaritons exhibit strong confinement and low losses in the mid-infrared, with coupling strengths enhanced by graphene overlays. The hybridization leads to avoided crossings between plasmon and phonon branches, enabling tunable hyperbolic responses for nanophotonics.

Applications

Sensing and Biosensing

Plasmon-based sensing and biosensing exploit the high sensitivity of surface plasmons to local changes in the , enabling label-free detection of biomolecular interactions. (SPR) sensors operate by monitoring angular shifts in the resonance condition upon binding of analytes, such as biomolecules, to a functionalized metal surface, typically . systems like Biacore utilize this principle to quantify protein-ligand affinities in , where the binding event alters the effective near the sensor surface. The limit of detection (LOD) of conventional SPR sensors is approximately $10^{-6} units (RIU) (or ~1 resonance unit, RU), corresponding to a surface coverage around 1 pg/mm² for protein, with angular sensitivities around 100 deg/RIU, as defined by the resonance unit (RU) where 1 RU ≈ $10^{-6} RIU and equates to roughly 1 pg/mm² of adsorbed protein. This capability has been demonstrated in applications such as the quantification of β-casein in and cheese using SPR immunosensors, marking an early historical example of plasmonic food analysis in the early . More recently, SPR-based platforms have been adapted for rapid antigen detection, achieving high clinical accuracy in nasopharyngeal and nasal swab samples through field-enhanced . Localized surface plasmon resonance (LSPR) in metallic nanoparticles, such as nanorods, extends these techniques to colorimetric sensing, where refractive index changes induce visible spectral shifts observable without complex instrumentation. For instance, functionalized nanorods exhibit LSPR wavelength shifts in response to variations from 6.41 to 8.88, enabling environmental and biosensing applications. To enhance performance, strategies like resonances introduce asymmetric line shapes with sharper peaks, improving resolution by narrowing the (FWHM) compared to profiles. Integration with further optimizes these sensors for low-volume, high-throughput assays, while damping mechanisms influence the overall resolution by broadening the resonance linewidth. A key (FOM) for evaluating plasmonic sensors is defined as the sensitivity divided by the FWHM, providing a measure of detection quality independent of absolute values.

Nanophotonics and Energy Devices

Plasmonics enables precise control of light at the nanoscale, far beyond the diffraction limit of conventional optics, through structures like metal-insulator-metal (MIM) waveguides that confine electromagnetic fields to subwavelength dimensions. These MIM configurations, consisting of thin dielectric layers sandwiched between metal films, support surface plasmon polariton (SPP) modes that propagate with effective indices exceeding 10, allowing for compact photonic circuits. Bandwidths in such waveguides can reach up to 100 THz, facilitating broadband signal processing in integrated nanophotonic systems. This capability is pivotal for overcoming the size constraints of dielectric waveguides, enabling high-density optical interconnects on chips. In enhanced spectroscopy, plasmonic nanostructures amplify Raman scattering signals via surface-enhanced Raman scattering (SERS), where electromagnetic hot spots—regions of intensely enhanced local fields—enable detection down to the single-molecule level. These hot spots arise from coupled plasmons in aggregates or gaps, boosting signal intensities by factors of 10^10 or more, sufficient for resolving vibrational spectra of individual analytes without labels. SERS has thus become a cornerstone for ultrasensitive chemical in , with applications in material characterization and trace detection. For energy devices, plasmons drive hot carrier generation in solar cells by converting absorbed photons into energetic electrons and holes via non-radiative decay of plasmon excitations in metal nanoparticles. In photovoltaic structures incorporating silver () nanoparticles, this mechanism scatters light to increase absorption while injecting hot carriers into the , yielding relative enhancements exceeding 20% in planar designs. Such plasmonic boosting extends carrier lifetimes and broadens spectral response, particularly in thin-film designs where light trapping is critical. Plasmon-enhanced outcoupling in organic light-emitting diodes (OLEDs) and light-emitting diodes (LEDs) mitigates losses by coupling emission to SPP modes at metal-organic interfaces, redirecting trapped light outward. Periodic nanostructures or gratings excite these SPPs to radiate efficiently, increasing external quantum efficiencies by up to 50% in top-emitting configurations while suppressing non-radiative . This approach reduces and SPP mode confinement, which otherwise trap over 70% of generated photons in standard devices. In data transmission, plasmonic interconnects on chips propagate signals at subwavelength scales, bypassing the limit to achieve densities unattainable with photonic wires. Hybrid plasmonic slots, combining metal and elements, support low-loss propagation over micrometer distances at speeds up to 100 Gbps, enabling monolithic integration with for high-, on-chip communication. These structures maintain across visible to near-infrared wavelengths, addressing bandwidth bottlenecks in future computing architectures. Advancements in the 2020s have leveraged plasmonic metasurfaces for , where arrays of nanoantennas impart precise control to wavefronts, reconstructing complex three-dimensional images with high fidelity. By tuning plasmon resonances in or silver nanostructures, these metasurfaces achieve full 2π coverage and multiplexing, enabling dynamic holograms with efficiencies up to 40% in the . Such platforms support applications in displays and secure optical data storage, building on principles for compact, flat .

References

  1. [1]
    [PDF] Introduction to plasmons and plasmonics
    The aim of this chapter is to give an overview of the physics of plasmons and their applications, so-called plasmonics. The emphasis will be on concepts.
  2. [2]
  3. [3]
  4. [4]
    Plasmonic Sensors: A New Frontier in Nanotechnology - PMC - NIH
    Mar 15, 2023 · Plasmonics is the study of surface plasmons formed by the interaction of incident light with electrons to form a surface-bound ...
  5. [5]
    [PDF] Some formulae for evaluation of the polariton and plasmon ... - arXiv
    A plasmon is a quasiparticle resulting from the quantization of plasma oscillations of the free electron gas density with respect to the fixed positive ions.
  6. [6]
    A Collective Description of Electron Interactions. I. Magnetic ...
    A new approach to the treatment of the interactions in a collection of electrons is developed, which we call the collective description.
  7. [7]
    Plasmon theory of electron‐hole pair production - AIP Publishing
    Feb 1, 1973 · The average energy to create an electron‐hole pair, Ex, is thus given essentially by ℏ ωp/n, where ℏ ωp is the plasmon energy and n is the ...
  8. [8]
    Plasmonics beyond noble metals: Exploiting phase and ...
    Aug 25, 2020 · Most common plasmonic materials, such as the noble metals gold and silver, cannot deliver this type of behavior. Therefore, significant ...
  9. [9]
    Signature of Electron-Plasmon Quantum Kinetics in GaAs
    Oct 16, 2000 · It reflects the oscillatory transfer of excitation back and forth between electrons and a mixed plasmon-phonon mode.
  10. [10]
  11. [11]
    Plasmon–emitter interactions at the nanoscale - Nature
    Jan 17, 2020 · Plasmon–emitter interactions are of central importance in modern nanoplasmonics and are generally maximal at short emitter–surface separations.
  12. [12]
    Plasmon-photon interaction in metal nanoparticles
    Plasmons are the quanta of such excitations, which strongly depend on the particle material, environment, and geometry. 1. The resulting possibility of ...
  13. [13]
    Physical Colors in Cultural Heritage: Surface Plasmons in Glass
    Aug 6, 2025 · It was discovered that this color change is due to the presence of gold and silver nanoparticles within the glass substrate [2], which cause a ...
  14. [14]
    Oscillations in Ionized Gases - PNAS
    Soc., 117, 660 (1928). OSCILLA TI&NS IN IONIZED GASES. By IRVING LANGMUIR. RESEARCHLABORATORY, GsERALu EIXCTRIc Co., SCH NsCTADY, N.
  15. [15]
    II. Collective Individual Particle Aspects of the Interactions | Phys. Rev.
    Collective v s Individual Particle Aspects of the Interactions. David Pines · David Bohm* ... 85, 338 – Published 15 January, 1952. DOI: https://doi.org ...
  16. [16]
    Recent progress in terahertz sensors based on graphene ...
    Feb 10, 2025 · This review examines recent advances in graphene MMs-based THz sensors, covering foundational theories and innovative designs.
  17. [17]
    [PDF] Oscillations in Ionized Gas, Tonks and Langmuir (1929)
    Finally, if a local oscillation be traversed by fast electrons having fairly uniform speeds, the alternate acceleration and retardation or the rythmical.
  18. [18]
    Theory of Plasma Oscillations. A. Origin of Medium-Like Behavior
    A theory of electron oscillations of an unbounded plasma of uniform ion density is developed, taking into account the effects of random thermal motions.Missing: Pines 1951
  19. [19]
    [PDF] plasmons and surface plasmons in bulk metals - TKM (KIT)
    ABSTRACT. This article gives an introduction and survey in the theory and spectroscopy of plasmons in the bulk, as well as on boundaries of metals.
  20. [20]
    [PDF] Chapter 1 - Optical Properties of Plasmonic Materials - SPIE
    Figure 1.8 compares the characteristics of metals using the Drude model and the Lorentz model. At the plasma frequency in the Drude model, the dielectric ...
  21. [21]
    [PDF] Optical dielectric function of silver
    Jun 22, 2015 · From Drude analysis of the 0.1–3 eV range, values of the plasma frequency hωp = 8.9 ± 0.2 eV, dielectric function at infinite frequency ...
  22. [22]
    (PDF) Canonical quantization for quantum plasmonics with finite ...
    Oct 28, 2019 · The diagonalized Hamiltonian is quantized by the correspondence principle, introducing creation-annihilation operators in a bosonic Fock space.
  23. [23]
    tJ!\~1
    The loss spectrum for these heavily doped, cleaved GaAs crystals shows two very strong peaks as shown in Fig. 1. The peak at 36 meV is readily identified as the ...
  24. [24]
    Plasma Losses by Fast Electrons in Thin Films
    Plasma Losses by Fast Electrons in Thin Films. R. H. Ritchie. Phys. Rev. 106, 874 – Published 1 June 1957. Article has an altmetric score of 3.
  25. [25]
    [PDF] Nano-optics of surface plasmon polaritons - Duke Physics
    Abstract. A surface plasmon polariton (SPP) is an electromagnetic excitation existing on the surface of a good metal. It is an.
  26. [26]
    Surface plasmon polaritons on metallic surfaces
    Surface plasmon polaritons (SPPs) are electromagnetic waves on metallic surfaces, with long-range modes being useful for practical applications.
  27. [27]
    Main techniques to excite surface plasmon polaritons (SPPs)
    Main techniques to excite surface plasmon polaritons (SPPs): (a) The Kretschmann Configuration; (b) Otto Configuration, and (c) Grating Coupling. Source ...
  28. [28]
    Propagation length of surface plasmon polaritons excited by a 1D ...
    The experimentally measured value propagation length of the SPP (LSPP) at 785 nm wavelength is 13.33 ± 0.13 μm, which is close to the theoretical value of the L ...
  29. [29]
    Review Localized surface plasmon resonance - ScienceDirect.com
    This review serves to describe the physical theory of LSPR formation at the surface of nanostructures, and the potential for this optical technology to serve as ...Review · Abstract · Introduction<|control11|><|separator|>
  30. [30]
    Error (ACS Publications)
    Insufficient relevant content. The provided URL (https://pubs.acs.org/doi/10.1021/ar700238q) leads to a "Page Not Found" error, indicating the page is inaccessible or does not exist. No information on LSP field enhancement, size and shape dependence for gold and silver nanoparticles can be extracted.
  31. [31]
    Developments in Localized Surface Plasmon Resonance
    Nov 4, 2024 · The LSPR properties of metal nanoparticles are heavily dependent on the particle's size. According to Mie's theory, for spherical nanoparticles ...Missing: seminal | Show results with:seminal
  32. [32]
    Two-dimensional Dirac plasmon-polaritons in graphene, 3D ...
    Oct 27, 2022 · The following section addresses general aspects of 2D Dirac plasmon in terms of dispersion relations calculated from RPA. After presenting the ...
  33. [33]
    Bulk-plasmon dispersion relations in metals | Phys. Rev. B
    Jul 15, 1991 · The formulas can be easily modified to incorporate band-structure effects (through an intraband electron effective mass) and core-polarization ...
  34. [34]
    [PDF] Bulk-plasmon dispersion relations in metals
    The formulas can be easily modified to incorporate band-structure effects (through an intraband electron effective mass) and core- polarization effects (through ...
  35. [35]
  36. [36]
  37. [37]
    Electron-electron interactions and plasmon dispersion in graphene
    Dec 2, 2013 · We use a many-body approach to relate plasmon dispersion at long wavelengths to Landau Fermi-liquid interactions and quasiparticle velocity.
  38. [38]
    Retardation effects on the dispersion and propagation of plasmons ...
    Sep 22, 2017 · We further study the impact of retardation effects on the propagation of plasmonic excitations along the chain. While for the longitudinal modes ...
  39. [39]
    Comparison of FDTD numerical computations and analytical ...
    May 26, 2009 · The FDTD method involves discretization of the Maxwell's equations in both the time and the space domains in order to find the E and H fields at ...
  40. [40]
    Plasmon losses due to electron-phonon scattering: The case of ...
    Oct 8, 2014 · In this article we present a theory of collective electron density oscillations in a graphene sheet encapsulated between two hBN semi-infinite slabs.
  41. [41]
    Landau damping of surface plasmons in metal nanostructures
    Dec 23, 2016 · We develop a quantum-mechanical theory for Landau damping of surface plasmons in metal nanostructures of arbitrary shape.Article Text · INTRODUCTION · DECAY RATE OF SURFACE... · CONCLUSIONS
  42. [42]
    Impact of the Interband Transitions in Gold and Silver on the ...
    We give a detailed analysis of Surface Plasmon Polaritons (SPPs) and Localized Surface Plasmons (LPSs) supported by such realistic metal/dielectric interfaces.<|control11|><|separator|>
  43. [43]
    Damping rates of surface plasmons for particles of size from nano
    Total damping rates of multipolar SPs [27] define absorbing and emitting properties of a plasmonic nanoantennas which can be tuned by particle size. Nanoantenna ...
  44. [44]
    Using the Plasmon Linewidth to Calculate the Time and Efficiency of ...
    The total linewidth, Γ, describes the overall decay of the plasmon oscillation and is due to contributions from intrinsic electron scattering in the metal, γb, ...Missing: γ_intrinsic γ_extrinsic
  45. [45]
    Ten years of spasers and plasmonic nanolasers | Light - Nature
    May 25, 2020 · For a good plasmonic metal of Q ~ 10–100, this spasing condition is ... A single-particle spaser has a single source of feedback, which limits the ...
  46. [46]
    Chemical Interface Damping of Surface Plasmon Resonances
    Mar 31, 2021 · From this body of research, we conclude that chemical interface damping is a major component of the total damping rate of the plasmon resonance ...Key References · Introduction · Spectroscopic Methods of... · Mechanisms of CID
  47. [47]
    Reduced damping of surface plasmons at low temperatures
    Jan 27, 2009 · We investigated the effectiveness of low temperatures in reducing plasmon damping by measuring the homogeneous linewidth of plasmon resonances.
  48. [48]
  49. [49]
    Ultrashort Self-Induced Transparency Plasmon Solitons
    Jun 11, 2013 · The intrinsic nonlinearity of two-level gaining media can then be used to obtain transverse localization of dissipative plasmon solitons [12] .
  50. [50]
    Dissipative plasmon solitons in graphene nanodisk arrays
    Schematic view of a discrete plasmon soliton excited by an external plane wave in a chain of graphene nanodisks. ... THz and infrared frequency ranges ...
  51. [51]
    Dissipative plasmon‐solitons in multilayer graphene - Smirnova - 2014
    Dec 20, 2013 · It is predicted that such a structure may support dissipative plasmon-solitons generated and supported by an external laser radiation. Novel ...Missing: review paper
  52. [52]
    [PDF] Plasmon-Soliton - arXiv
    In this letter we present for the first time both an analysis of TM spatial solitons, using the Non- linear Schrödinger Equation (NLSE), where the full ...
  53. [53]
    [PDF] Experimental demonstration of plasmon-soliton waves - HAL
    Abstract: We report the experimental observation of plasmon-soliton waves. The demonstration is performed in a chalcogenide-based four-layer planar geometry ...
  54. [54]
    Vacuum Rabi splitting in a plasmonic cavity at the single quantum ...
    Jun 13, 2016 · Here we demonstrate vacuum Rabi splitting, a manifestation of strong coupling, using silver bowtie plasmonic cavities loaded with semiconductor quantum dots ( ...
  55. [55]
    Vacuum Rabi splitting of a dark plasmonic cavity mode revealed by ...
    Jan 24, 2020 · A dark mode of an individual plasmonic bowtie can interact with a small number of quantum emitters, as evidenced by Rabi-split spectra.
  56. [56]
    Strong Coupling between Surface Plasmons and Excitons in an ...
    We report on the observation of a strong coupling between a surface plasmon and an exciton. Reflectometry experiments are performed on an organic semiconductor.Missing: seminal | Show results with:seminal
  57. [57]
    Plasmon-exciton strong coupling in an organic material - Nature
    Jul 1, 2025 · We report on strong coupling between surface plasmons and excitons in an optically nonmetallic excitonic dye-polymer material of different thicknesses.
  58. [58]
    Topological phases and bulk-edge correspondence of magnetized ...
    Here we systematically map out all the topological phases and establish the bulk-edge correspondence in cold magnetized plasmas.Missing: gyromagnetic post-
  59. [59]
    Quantum teleportation mediated by surface plasmon polariton - PMC
    Quantum teleportation uses entangled photons, Bell-state measurements, and photon-SPP-photon conversion to transfer a quantum state between photons and SPPs.Missing: 2020s | Show results with:2020s
  60. [60]
    Resonance energy transfer and quantum entanglement mediated by ...
    This work shows long-range entanglement and enhanced energy transfer using ENZ and other plasmonic waveguides, with ENZ outperforming others.Missing: 2020s experiments
  61. [61]
    Near-Field Retrieval of the Surface Phonon Polariton Dispersion in ...
    Oct 20, 2022 · Silicon carbide (SiC) is an excellent material for SPhP applications, having a wide RS band (approximately 800–1000 cm–1 or equivalently 10 to ...Introduction · SPhP Dispersion in Thin Films · Conclusions · Supporting Information
  62. [62]
    Coupled surface plasmon–phonon polariton nanocavity arrays ... - NIH
    Silicon carbide (6H-SiC) as the polar dielectric crystal shows a high reflectivity and negative real permittivity within the Reststrahlen band bound by the ...
  63. [63]
    Surface Plasmon Resonance Biosensors: A Review of Molecular ...
    Feb 2, 2024 · Abstract. Surface plasmon resonance (SPR) is a powerful tool for determining molecular interactions quantitatively.Missing: paper | Show results with:paper
  64. [64]
    Surface plasmon resonance biosensors and their medical applications
    Jun 15, 2025 · This work aims to review progress in the development of SPR biosensors for medical diagnostics, focusing mainly on advances in optical platforms and assays.
  65. [65]
    Small Molecule Immunosensing Using Surface Plasmon Resonance
    Resonance units (RU) are often used to quantify refractive index changes in SPR biosensors, with 1 RU = 1 × 10−6 RIU, and so are used as units describing SPR ...
  66. [66]
    [PDF] SPR Sensitivity and Detection Limit | Biosensing Instrument
    Relative Index of Refraction Unit: Another quantity often used to describe SPR sensitivity is the relative change in the index of refraction of the fluid ...
  67. [67]
    Quantification of beta casein in milk and cheese using an optical ...
    beta-Casein was quantified in milk and cheese, using an optical immunosensor, based on surface plasmon resonance (SPR) measurement.Missing: historical | Show results with:historical
  68. [68]
    Pre-clinical and clinical evaluation of a surface plasmon field ...
    Aug 27, 2024 · The SARS-CoV-2 SPFS antigen test showed excellent clinical diagnostic accuracy for nasopharyngeal and nasal swabs, with a rapid turnaround.
  69. [69]
    (PDF) Induced pH-dependent shift by local surface plasmon ...
    Aug 6, 2025 · In this study, we observed LSPR shifts of 11-mercaptoundecanoic acid modified gold nanorods (GNR-MUA) for the pH range of 6.41 to 8.88. We ...
  70. [70]
    Seeing protein monolayers with naked eye through plasmonic Fano ...
    We introduce an ultrasensitive label-free detection technique based on asymmetric Fano resonances in plasmonic nanoholes with far reaching implications for ...Missing: sharper | Show results with:sharper<|control11|><|separator|>
  71. [71]
    Figure-of-merit enhancement of surface plasmon resonance sensors ...
    Hence the figure of merit (FOM), which is defined as the ratio between the sensitivity S and the FWHM, did not change ( FOM = S / FWHM ). In this Letter we ...
  72. [72]
    Nanoplasmonic waveguides: towards applications in integrated ...
    Jun 5, 2015 · In this article, we review recent development on plasmonic waveguides and plasmonic devices, mainly focus on plasmonic components related to digital circuits.
  73. [73]
    Plasmon slot waveguides: Towards chip-scale propagation with ...
    Jan 5, 2006 · MIM waveguides may thus allow optical mode volumes to be reduced to subwavelength scales—with minimal field decay out of the waveguide ...
  74. [74]
    Electrically driven monolithic subwavelength plasmonic interconnect ...
    Oct 20, 2017 · Tailoring photonics for monolithic integration with electronics beyond the diffraction limit opens a new era of chip-level nanoscale electronic- ...
  75. [75]
    A Review on Surface-Enhanced Raman Scattering - PMC - NIH
    The excitation of surface plasmons induces a strong spatial localization and hence amplification of the laser light in small spatial regions, called hot spots.
  76. [76]
    Single-Molecule Surface-Enhanced Raman Spectroscopy
    May 5, 2012 · A general overview of the field of single-molecule (SM) surface-enhanced Raman spectroscopy (SERS) as it stands today is provided.<|separator|>
  77. [77]
    Hot-Carrier Generation in Plasmonic Nanoparticles - ACS Publications
    Jul 20, 2020 · Metal nanoparticles are attractive for plasmon-enhanced generation of hot carriers, which may be harnessed in photochemical reactions.
  78. [78]
    Ultra-thin Ag/Si heterojunction hot-carrier photovoltaic conversion ...
    Apr 3, 2023 · Conventional single-silicon p–n junction solar cells have been developed for more than 40 years to obtain a conversion efficiency of about 20%.
  79. [79]
    Surface Plasmon-Polariton Mediated Red Emission from Organic ...
    Nov 19, 2014 · We demonstrate an effective approach to realize excitation and outcoupling of the SPP modes associated with both cathode/organic and anode/organic interfaces ...
  80. [80]
    Enhanced light emission from top-emitting organic light-emitting ...
    Dec 11, 2015 · We demonstrate enhanced light extraction for monochrome top-emitting organic light-emitting diodes (OLEDs).
  81. [81]
    Controllable hybrid plasmonic integrated circuit | Scientific Reports
    Jun 20, 2023 · The future of integrated circuits will require wireless chip-to-chip communication to overcome the bottlenecks caused by wired connections. The ...
  82. [82]
    Recent advances in multi-dimensional metasurfaces holographic ...
    Oct 19, 2020 · We classify the metasurface holography into several categories such as phase-only holography, amplitude-only holography, complex amplitude holography and so on.
  83. [83]
    Nonlinear Bicolor Holography Using Plasmonic Metasurfaces
    We demonstrate a single-layer nonlinear geometric-phase metasurface made of plasmonic nanostructures for a simultaneous second- and third-harmonic generation.