W. J. M. Kort-Kamp's research while affiliated with Los Alamos National Laboratory and other places

What is this page?


This page lists the scientific contributions of an author, who either does not have a ResearchGate profile, or has not yet added these contributions to their profile.

It was automatically created by ResearchGate to create a record of this author's body of work. We create such pages to advance our goal of creating and maintaining the most comprehensive scientific repository possible. In doing so, we process publicly available (personal) data relating to the author as a member of the scientific community.

If you're a ResearchGate member, you can follow this page to keep up with this author's work.

If you are this author, and you don't want us to display this page anymore, please let us know.

Publications (43)


Controlling electric and magnetic Purcell effects in phosphorene via strain engineering
  • Article
  • Full-text available

October 2023

·

106 Reads

·

1 Citation

·

W. J. M. Kort-Kamp

·

·

[...]

·

We investigate the spontaneous emission lifetime of a quantum emitter near a substrate coated with phosphorene under the influence of uniaxial strain. We consider both electric dipole and magnetic dipole-mediated spontaneous transitions from the excited to the ground state. The modeling of phosphorene is achieved by employing a tight-binding model that goes beyond the usual low-energy description. We demonstrate that both electric and magnetic decay rates can be widely tuned by the application of uniform strain, ranging from a near-total suppression of the Purcell effect to a remarkable enhancement of more than 1300%, all due to the high flexibility associated with the puckered lattice structure of phosphorene. We also unveil the use of strain as a mechanism to tailor the most probable decay pathways of the emitted quanta. Our results show that uniaxially strained phosphorene is an efficient and versatile material platform for the active control of light-matter interactions thanks to its extraordinary optomechanical properties.

Download
Share

Entangled two-plasmon generation in carbon nanotubes and graphene-coated wires

April 2022

·

40 Reads

·

1 Citation

We investigate the two-plasmon spontaneous decay of a quantum emitter near single-walled carbon nanotubes (SWCNTs) and graphene-coated wires (GCWs). We demonstrate efficient, enhanced generation of two-plasmon entangled states in SWCNTs due to the strong coupling between tunable guided plasmons and the quantum emitter. We predict two-plasmon emission rates more than twelve orders of magnitude higher than in free space, with average lifetimes of a few dozen nanoseconds. Given their low dimensionality, these systems could be more efficient for generating and detecting entangled plasmons in comparison to extended graphene. Indeed, we achieve a tunable spectrum of emission in GCWs, where sharp resonances occur precisely at the plasmons' minimum excitation frequencies. We show that by changing the material properties of the GCW's dielectric core, one could tailor the dominant modes and frequencies of the emitted entangled plasmons while keeping the decay rate ten orders of magnitude higher than in free space. By unveiling the unique properties of two-plasmon spontaneous emission processes in the presence of low-dimensional carbon-based nanomaterials, our findings set the basis for a novel material platform with applications to on-chip quantum information technologies.


FIG. 1. (a) Top and side views of the graphene family honeycomb structure. The red and blue colored atoms belong to inequivalent sublattices with a finite staggering 2 between them. (b) Schematics of the system under study. Two graphene family monolayers separated by a distance d and subjected to externally applied electric (E z ) and magnetic (B z ) fields, and a circularly polarized laser (). (c) Low-energy band structure around a given K or K point. For a nonzero magnetic field, the quantized Landau levels (red circles) are built on top of the Dirac cone.
FIG. 2. Casimir energy in the (E z , ,) plane for (a) μ = λ SO and E B = 1.2λ SO and (b) E B = 0.8λ SO . (c) Normalized Casimir energy in the (E z , μ) plane for = 0 and E B = λ SO . (d) Normalized Casimir energy as a function of the chemical potential for = 0. In all plots the distance between the plates is given by dλ SO / ¯ hc = 10 and the dissipation is set to zero.
FIG. 3. Casimir energy as a function of the chemical potential of a single monolayer for {eE z , μ 2 }/λ SO = {(1, 0.5), (1, −0.5), (1.5, 0.5)} (solid blue, dashed red, and dash-dotted green, respectively) and E B = λ SO . For all curves, the distance between the plates is long enough so that Eq. (10) is valid and , , = 0.
FIG. 4. (a) Long distance approximated expression given by Eq. (10) subtracted from the Casimir energy as a function of distance. The chosen parameters are = 0, = 0, eE z = λ SO , and μ = {0.5, 1.25, 1.5}λ SO (solid blue, dashed red, and dash-dotted green lines, respectively). The inset shows the longitudinal and Hall conductivity at frequency ¯ hξ = 0.01λ SO as a function of the chemical potential. (b) Casimir energy for = 0, eE z = λ SO , μ = 1.25λ SO , and ¯ h = {0, 0.01, 0.02}λ SO (solid, dashed, and dash-dotted lines, respectively) as a function of distance. The gray dotted lines are the approximated solutions in the long distance regime, given by the sum of Eqs. (10) and (13).
FIG. 5. (a) Dissipationless Casimir energy as a function of temperature for a separation distance of dλ SO / ¯ hc = 10. The dotted lines correspond to the approximated value of the n = 0 Matsubara frequency contribution presented in Eq. (14). The left inset is a zoom of the plot, showing that each curve goes to its correspondent zero-temperature limit (dashed gray lines). The right inset shows the Casimir force in the region where it becomes attractive. (b) Casimir energy as a function of distance for k B T = 10 −3 λ SO and ¯ h = 0.01λ SO . The gray line corresponds to the n = 0 Matsubara contribution [Eq. (15)]. In both plots, = 0, eE z = λ SO , and μ = {0.5, 1.25, 1.5}λ SO (solid blue, dashed red, and dash-dotted green lines, respectively).
Casimir forces in the flatland: Interplay between photoinduced phase transitions and quantum Hall physics

April 2021

·

76 Reads

·

5 Citations

Physical Review Research

We investigate how photoinduced topological phase transitions and the magnetic-field-induced quantum Hall effect simultaneously influence the Casimir force between two parallel sheets of staggered two-dimensional (2D) materials of the graphene family. We show that the interplay between these two effects enables on-demand switching of the force between attractive and repulsive regimes while keeping its quantized characteristics. We also show that doping these 2D materials below their first Landau level allows one to probe the photoinduced topology in the Casimir force without the difficulties imposed by a circularly polarized laser. We demonstrate that the magnetic field has a huge impact on the thermal Casimir effect for dissipationless materials, where the quantized aspect of the energy levels leads to a strong repulsion that could be measured even at room temperature.


Figure 4. (a) Spectral TPSE for a D = 40 nm graphene nano-disk (solid) and a graphene monolayer (dashed). The emitter (ωt = 0.66 eV) is located at ze = 10 nm. Graphene's conductivity is modeled using intra-and inter-band contributions, mobility is µ = 2500 cm 2 V −1 s −1 , and temperature is T = 300 K. (b) Ratio of quantum yields between two-and one-quantum processes for the nano-disk. (c) Quantum efficiency versus distance for the TPSE 4s → 3s transition in hydrogen (µ is in units of cm 2 V −1 s −1 ). Inset: TPSE rate versus ze for the nano-disk (solid) and monolayer (dashed). (d) QE as a function of EF and D. The numbers near each QE profile show the photon-photon Purcell factor Γ ph,ph /Γ0, where µ = 10 4 cm 2 V −1 s −1 and Γ0 is the free-space TPSE rate.
Two-Photon Spontaneous Emission in Atomically Thin Plasmonic Nanostructures

June 2020

·

230 Reads

The ability to harness light-matter interactions at the few-photon level plays a pivotal role in quantum technologies. Single photons - the most elementary states of light - can be generated on-demand in atomic and solid state emitters. Two-photon states are also key quantum assets, but achieving them in individual emitters is challenging because their generation rate is much slower than competing one-photon processes. We demonstrate that atomically thin plasmonic nanostructures can harness two-photon spontaneous emission, resulting in giant far-field two-photon production, a wealth of resonant modes enabling tailored photonic and plasmonic entangled states, and plasmon-assisted single-photon creation orders of magnitude more efficient than standard one-photon emission. We unravel the two-photon spontaneous emission channels and show that their spectral line-shapes emerge from an intricate interplay between Fano and Lorentzian resonances. Enhanced two-photon spontaneous emission in two-dimensional nanostructures paves the way to an alternative efficient source of light-matter entanglement for on-chip quantum information processing and free-space quantum communications.


Spatio-temporally modulated metasurfaces
a An incident beam impinging on an STMM is converted into a different frequency harmonic that can be focused at any desired focal point. b Breakdown of Lorentz reciprocity can be shown by probing the time-reversed process. c Photograph of our STMM. d Top view and cross-section of the unit cell. All geometrical parameters are in mm.
Dynamical wave-front shaping
a Reflectance measurements for the unmodulated metasurface. b Measured on-demand dynamical beam steering of the n = +1 harmonic. c Calculated spatial distribution of the electric field for on-axis focusing for the n=n¯=+1\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$n = \bar n = + 1$$\end{document} frequency harmonic (infinite-sized STMM). d Measured power for on-axis focusing. e Calculated (finite-sized STMM, Lx = 19 cm) and measured gain as a function of ℓ\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\ell$$\end{document} for off-axis focusing. Error bars are determined by the standard deviation of gain over a narrow angular range within the focal region. f Measured power for off-axis focusing. In b–f, kin = 0, ωin = 2π × 6.9 GHz and Ω = 2π × 50 kHz. Input power in the experiments is 6.3 mW. Focusing parameters are (xf, zf) = (0,15) cm and (6,15) cm for on-axis and off-axis, respectively.
Nonreciprocity in beam steering
a A normally-incident (kin = 0) beam is up-converted to ωin + Ω and steered to an angle ≈ +18° with a phase gradient βx = 44 m⁻¹. Three dominant reverse pathways are allowed (b–d). b Momentum nonreciprocity: Reflection from +18° of the up-converted beam undergoes a down-conversion to ωin and is steered to ≈ −36°. c Frequency nonreciprocity: Reflection from +18° undergoes an up-conversion to ωin + 2Ω and is normally reflected. d Frequency and momentum nonreciprocity: Beam from +18° is specularly reflected at ωin + Ω, differing both in frequency and direction of propagation from the original input beam. e Forward measured reflection beyond threshold: the input beam is steered to ≈ +43° for βx = 96 m⁻¹, which is larger than the +30° threshold. f Computed reflected Poynting vector distribution for the forward process in e. g Measured reverse scattering to ωin resulting in photon-to-surface wave conversion and complete optical isolation. h Calculated profile of evanescent surface waves corresponding to the reverse process in g. In all panels ωin = 2π × 6.9 GHz and Ω = 2π × 50 kHz.
Surface-wave-assisted nonreciprocity in focusing
a Measured reflected field power at ω+1 in the forward off-axis focusing experiment (kin = 0). b Calculated snapshot of the reflected electric field distribution for the reverse scattering process (−∇φn¯=+1focus,ωin+Ω)→(−2∇φn¯=+1focus,ωin)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$( - \nabla \varphi _{\bar n = + 1}^{{\mathrm{focus}}},\omega _{{\mathrm{in}}} + {\mathrm{\Omega }}) \to ( - 2\nabla \varphi _{\bar n = + 1}^{{\mathrm{focus}}},\omega _{{\mathrm{in}}})$$\end{document} for the infinite-sized STMM. The black line represents the time-averaged Poynting vector for the same scattering process. c Measured power at ωin in the far-field for the reverse process is mainly confined within the wedge-shaped region. Data at given radii of the arc sector correspond to different locations ℓ\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\ell$$\end{document} of the monopole source. Indicated are the calculated main directions of radiative emission from the right and left edges of the STMM and angles subtended between their crossings with the arc and its center, respectively given by αRth=+53.27∘\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\alpha _{\mathrm{R}}^{{\mathrm{th}}} = + 53.27^\circ$$\end{document} and αLth=−20.36∘\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\alpha _{\mathrm{L}}^{{\mathrm{th}}} = - 20.36^\circ$$\end{document}. The measured scanning angles for the main peaks on the right-side and left-side are α+exp=(+54±1)∘\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\alpha _ + ^{{\mathrm{exp}}} = ( + 54 \pm 1)^\circ$$\end{document} and α−exp=(−26±1)∘\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\alpha _ - ^{{\mathrm{exp}}} = ( - 26 \pm 1)^\circ$$\end{document}. d Same as c but as a function of the scanning angle and ℓ\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\ell$$\end{document}. Parameters are xf = 6 cm and zf = 15 cm (ℓf=16cm)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$(\ell _f = 16\,{\mathrm{cm}})$$\end{document}. In all panels ωin = 2π × 6.9 GHz and Ω = 2π × 50 kHz.
Surface-wave-assisted nonreciprocity in spatio-temporally modulated metasurfaces

March 2020

·

630 Reads

·

105 Citations

Nature Communications

Emerging photonic functionalities are mostly governed by the fundamental principle of Lorentz reciprocity. Lifting the constraints imposed by this principle could circumvent deleterious effects that limit the performance of photonic systems. Most efforts to date have been limited to waveguide platforms. Here, we propose and experimentally demonstrate a spatio-temporally modulated metasurface capable of complete violation of Lorentz reciprocity by reflecting an incident beam into far-field radiation in forward scattering, but into near-field surface waves in reverse scattering. These observations are shown both in nonreciprocal beam steering and nonreciprocal focusing. We also demonstrate nonreciprocal behavior of propagative-only waves in the frequency- and momentum-domains, and simultaneously in both. We develop a generalized Bloch-Floquet theory which offers physical insights into Lorentz nonreciprocity for arbitrary spatial phase gradients, and its predictions are in excellent agreement with experiments. Our work opens exciting opportunities in applications where free-space nonreciprocal wave propagation is desired. Overcoming reciprocity is important for novel functionalities. Here, the authors demonstrate a spatio-temporally modulated metasurface capable of complete violation of Lorentz reciprocity by reflecting an incident beam into far-field radiation in forward scattering, but into near-field surface waves in reverse scattering.


Figure 3. (a) Photograph of the fabricated reflectarray metasurface. (b) Spatial phase profile of the metasurface, where the color coding represents the use of the different 32 discrete phase values listed in Fig. 2(b). (c) A schematic of the experimental setup for characterizing the off-axis metasurface.
Figure 4. (a) Measured (blue) and simulated (red) gain of microwave radiation of the reflectarray as a function of the azimuth scan angle. The elevation angle is 0° and d 1 = 57 cm. (b) Gain profile as a function of the elevation and azimuth angles at the focus (d 1 = 57 cm), and (c) as a function of azimuth angle and separation d 1 .
(a) Schematic of an off-axis focusing phase gradient metasurface reflectarray. For a plane wave with incidence angle θi\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\,{\theta }_{i}$$\end{document} and focal point (xf, yf , zf), the required phase distribution ψ(x,y)\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\psi (x,y)\,$$\end{document}is given by Eq. (1). (b) Unit-cell design, consisting of a pair of rectangular patch resonators (copper) with their longer axis along the electric field of the microwave radiation and separated from a copper ground plane by a thin dielectric spacer.
(a) Simulated reflection phase (solid curves) and amplitude (dotted curves) as a function of frequency for three different resonator unit cells. The structural parameters of resonator #1 (blue) are L1 = 8.0 mm, w1 = 3.0 mm, g1 = 1.0 mm, resonator #17 (red) are L17 = 6.25 mm, w17 = 2.875 mm, g17 = 0.5 mm, and resonator #32 (green) are L32 = 2.5 mm, w32 = 2.75 mm, g32 = 0.5 mm. Choosing the phase of resonator #1 as 0° at 12 GHz, the relative phases for resonators #17 and #32 are 180° and 348.75° at the same frequency (L, w and g parameters for all 32 resonators are given in the supplementary material). (b) Phase distribution of the 32 unit cells of the reflectarray panel, which indicates the discretization of the phase into a finite number of elements (red). The blue curve shows the corresponding simulated reflectance.
Metasurface-based ultra-lightweight high-gain off-axis flat parabolic reflectarray for microwave beam collimation/focusing

December 2019

·

330 Reads

·

12 Citations

Scientific Reports

Ultra-lightweight deployable antennas with high-gain are pivotal communication components for small satellites, which are intrinsically constrained in size, weight, and power. In this work, we design and demonstrate metasurface-based ultra-lightweight flat off-axis reflectarrays for microwave beam collimation and focusing, similar to a parabolic dish-antenna. Our ultra-thin reflectarrays employ resonators of variable sizes to cover the full 2π phase range, and are arranged on the metasurface to realize a two-dimensional parabolic focusing phase distribution. We demonstrate a 30° off-axis focusing reflector that exhibits a measured gain of 27.5 dB at the central operating frequency of 11.8 GHz and a 3 dB directionality


Highly Plasmonic Titanium Nitride by Room-Temperature Sputtering

October 2019

·

1,116 Reads

·

76 Citations

Scientific Reports

Titanium nitride (TiN) has recently emerged as an attractive alternative material for plasmonics. However, the typical high-temperature deposition of plasmonic TiN using either sputtering or atomic layer deposition has greatly limited its potential applications and prevented its integration into existing CMOS device architectures. Here, we demonstrate highly plasmonic TiN thin films and nanostructures by a room-temperature, low-power, and bias-free reactive sputtering process. We investigate the optical properties of the TiN films and their dependence on the sputtering conditions and substrate materials. We find that our TiN possesses one of the largest negative values of the real part of the dielectric function as compared to all other plasmonic TiN films reported to date. Two-dimensional periodic arrays of TiN nanodisks are then fabricated, from which we validate that strong plasmonic resonances are supported. Our room-temperature deposition process can allow for fabricating complex plasmonic TiN nanostructures and be integrated into the fabrication of existing CMOS-based photonic devices to enhance their performance and functionalities.


Fig. 1. Spatio-temporal modulated metasurfaces. (A) Conceptual illustration of STMMs: An incident beam impinging on a dynamic metasurface is converted into a different harmonic that can be focused at any desired focal point. (B) Violation of Lorentz reciprocity can be shown by probing the time-reversed process. (C) Photograph of our STMM. (D) Top and side view of the unit cell showing bias lines to multi-output controller. All geometrical parameters are in mm.
Fig. 2. Dynamical wave-front shaping. (A) Reflectance measurements at different bias voltages for the unmodulated metasurface. (B) Measured on-demand dynamical beam steering of the í µí±› = +1 harmonic. Arbitrary dynamical focusing is shown for the í µí±› = í µí±› … = +1 harmonic: (C) Calculated spatial distribution of the amplitude of electric field for on-axis focusing (infinite-sized STMM). (D) Measured power for on-axis focusing. (E) Calculated (solid lines, finite-sized STMM, í µí°¿ • = 19 cm) and measured (open circles) gain as a function of the position ℓ along the focal axis for off-axis focusing. (F) Measured power for off-axis focusing. Focal point (í µí±¥ ‰ , í µí± § ‰ ) for on-and off-axis is (0,15) cm and (6,15) cm, respectively. In (B-F), í µí²Œ <= = 0, í µí¼” í µí±–í µí±› = 2í µí¼‹ ×6.9 GHz and í µí»º = 2í µí¼‹ ×50 kHz.
Fig. 3. Nonreciprocal beam steering. (A) A normally incident beam is steered to the í µí±› = +1 harmonic at an angle í µí¼ƒ coe ≈ +18 o by applying a linear phase gradient í µí»½ • =44 m -1 . (B) Reverse experiment: reflection from +18 o incidence of the up-converted beam undergoes a downconversion to í µí±› = 0 and is steered to -36 o . Calculated snapshots of the reflected Poynting vector distribution (normalized by the respective incoming intensities) for the forward (C) and reverse (D) processes described in (A) and (B). (E) Forward reflection to í µí¼ƒ coe ≈ +43 o for í µí»½ • =96 m -1 . (F) Calculated Poynting vector distributions of the scattered fields corresponding to (E) and (F) are shown in (G, H). In the four left (right) panels, the arrows show the direction of propagation of the incident (scattered) fields.
Fig. 4. Nonreciprocal focusing. (A) Measured reflected field power at í µí¼” coe in the forward offaxis focusing experiment, as a function of the angle í µí»¼ between the surface normal and the detector, and the distance ℓ along the focal axis. Focal point is at (6,15) cm (ℓ ‰ = 16 cm). (B) Calculated snapshot of the scattered Poynting vector distribution at í µí¼” <= for the reverse process (infinite-sized STMM). (C) Measured power at í µí¼” <= in the far-field for the reverse process. The monopole source emits at frequency í µí¼” coe .
Dynamical Wave-Front Shaping and Nonreciprocity with Spatio-Temporal Modulated Metasurfaces

October 2019

·

647 Reads

We experimentally demonstrate a spatio-temporal modulated metasurface reflectarray that is capable of dynamical wave-front shaping and nonreciprocal propagation of free space electromagnetic waves. Arbitrary space-time phase distributions in reflection are achieved by embedding electronically modulated elements into the resonators of our metasurface. Our experimental measurements reveal on-demand wave-front control of frequency conversion processes. We also demonstrate maximum violation of Lorentz reciprocity in both beam steering and focusing due to nonreciprocal excitation of surface waves. We develop an analytical generalized Bloch-Floquet theory valid for arbitrary modulations, providing excellent agreement with the experiments. Our ultrathin and lightweight spatio-temporal modulated metasurface will enable novel functionalities for wireless communications, imaging, sensing, and nonreciprocal electromagnetic isolation.


FIG. 1: Spectral density function γ(ω; z) of an emitter near a perfect mirror in terms of the dimensionless variable ω/ωeg for three different values of z. We also plot the spectral density function in free space (solid line).
FIG. 2: Spectral enhancement γ(ω; z)/γ0(ω) of an emitter near a perfect mirror as a function of the separation distance z for three given frequencies.
FIG. 3: Spectral enhancement of an emitter near a half-space Polystyrene medium as a function of ω/ωeg for three given values of z. The Polystyrene resonance frequencies are given by ωR1 = 5.54 × 10 14 rad/s and ωR2 = 1.35 × 10 16 rad/s and the corresponding widths by Γ = 1 × 10 11 rad/s 43 . The emitter transition frequency was chosen as 3ωR2.
FIG. 4: Spectral enhancement of an emitter equidistant from two perfect mirrors as a function of ω/ωeg for three given values of L.
FIG. 5: Spectral enhancement in terms of the dimensionless variable kegL/π with the emitter equidistant to both mirrors for three given frequencies.
Quantum two-photon emission in a photonic cavity

June 2019

·

110 Reads

We derive a new expression for the two-photon spontaneous emission (TPSE) rate of an excited quantum emitter in the presence of arbitrary bodies in its vicinities. After investigating the influence of a perfectly conducting plate on the TPSE spectral distribution (Purcell effect), we demonstrate the equivalence of our expression with the more usual formula written in terms of the corresponding dyadic Green's function. We establish a general and convenient relation between the TPSE spectral distribution and the corresponding Purcell factors of the system. Next, we consider an emitter close to a dielectric medium and show that, in the near field regime, the TPSE spectral distribution is substantially enhanced and changes abruptly at the resonance frequencies. Finally, motivated by the suppression that may occur in the one-photon spontaneous emission of an excited atom between two parallel conducting plates, we discuss the TPSE for this same situation and show that complete suppression can never occur for $s \rightarrow s$ transitions.


Fig. 1 System schematics. The system under study consists of a chain of N nanoparticles with diameters D i separated by a center-to-center distance d ij . The nanoparticles are assumed to rotate around the axis of the chain
Rotational dynamics. aHij for a chain of N = 5 identical particles with D = 10 nm made of SiC, which are uniformly distributed with a center-to-center distance d = 1.5D. b Diagonal components of hij (blue) and the corresponding hi0\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$h_i^0$$\end{document} (red). Notice the different scale. c Natural modes of the chain obtained by diagonalizing Hij and the corresponding decay rates. d–f Temporal evolution of the angular velocity of each particle in the chain for three different initial conditions
Natural decay rates. Evolution of the natural decay rates for chains with different N (as indicated in the legend) plotted as a function of the effective momentum keff. The inset shows the evolution of the slowest decay rate λ1 as a function of N. In all of the cases, D = 10 nm and d = 1.5D
Exotic rotational dynamics. a Temporal evolution of the angular velocity of the particles of a chain with N = 3, arranged as shown in the inset, where D = 10 nm and d = 1.5D. b Crossing of the second and third decay rates (i.e., λ2 and λ3) of the N = 3 chain shown in the inset of panel c when varying the ratio D′/D. In this case, D = 10 nm and d = 3D. c Temporal evolution of the angular momentum, L(t) = IΩ(t), for each of the three particles of the chain depicted in the inset, assuming D′/D = 1.914. We analyze two different cases with initial angular velocities given by Ω1/2π = 10 GHz + Ωc/2π, Ω2/2π = −0.39 GHz + Ωc/2π, and Ω3 = Ωc. In one case, Ωc = 0 while, in the other, Ωc/2π = 5 GHz. However, since there is no transfer of angular momentum to particle 3, both cases produce exactly the same results
Driven rotational dynamics. a Steady-state angular velocity for the particles of a N = 30 chain, in which particle 1 is externally driven at an angular velocity 10 GHz. All of the particles have the same diameter D and are uniformly distributed with a center-to-center distance d, as shown in the inset. b Same as a, but, in this case, two particles, which are indicated with small arrows (see legend), are externally driven at angular velocities 10 and −10 GHz, respectively
Nanoscale transfer of angular momentum mediated by the Casimir torque

June 2019

·

289 Reads

·

19 Citations

Communications Physics

Casimir interactions play an important role in the dynamics of nanoscale objects. Here, we investigate the noncontact transfer of angular momentum at the nanoscale through the analysis of the Casimir torque acting on a chain of rotating nanoparticles. We show that this interaction, which arises from the vacuum and thermal fluctuations of the electromagnetic field, enables an efficient transfer of angular momentum between the elements of the chain. Working within the framework of fluctuational electrodynamics, we derive analytical expressions for the Casimir torque acting on each nanoparticle in the chain, which we use to study the synchronization of chains with different geometries and to predict unexpected dynamics, including a “rattleback”-like behavior. Our results provide insights into the Casimir torque and how it can be exploited to achieve efficient noncontact transfer of angular momentum at the nanoscale, and therefore have important implications for the control and manipulation of nanomechanical devices.


Citations (33)


... The TPSE makes the vacuum unstable and is responsible for the initial buildup of the intracavity field in two-photon micromasers [19][20][21]. More recently, it was shown that the simultaneously emitted photons can be indistinguishable and entangled in time and frequency [22][23][24], renewing the interest in this phenomenon [25][26][27][28][29]. This section aims to obtain the TPSE rate directly from first-order perturbation theory in Hamiltonian (11). ...

Reference:

Time-Dependent Effective Hamiltonians for Light–Matter Interactions
Entangled two-plasmon generation in carbon nanotubes and graphene-coated wires
  • Citing Article
  • April 2022

... Nonetheless, the implementation of the metasurface could also be inspired by several innovative prototypes in the microwave range [57]. Many of these devices are based on adjustable biased PIN diodes [58][59][60] and varactors [18,19,61], traditionally implemented in reflectarray and trasmittarray systems. Moreover, novel theoretical designs based on the use of electronically-reconfigurable materials such as graphene [62,63] or transparent conductive oxides [64] open new alternatives for the millimeter-wave and terahertz regimes. ...

Surface-wave-assisted nonreciprocity in spatio-temporally modulated metasurfaces

Nature Communications

... Many studies and papers focusing on different MS applications covering various frequency bands and methods have been published. For example, applications in which MS plays an important role include cloaks [7], flat polarization control [8], and focusing and collimation flat reflect arrays [9]. Furthermore, interdisciplinary reconfigurable MS applications were demonstrated, such as reconfigurable MS antenna [10][11][12] and the extraordinary sensitivity enhancement detection of antibiotics [13]. ...

Metasurface-based ultra-lightweight high-gain off-axis flat parabolic reflectarray for microwave beam collimation/focusing

Scientific Reports

... The TiN thin film exhibits a larger plasmon decay length, particularly in TiN/Al 2 O 3 structures. This characteristic makes it a suitable candidate for various photonic applications, including integration with III-V semiconductors such as GaN and InN, TiN-based photonic/plasmonic on-chip devices, broadband light absorbers, emitters for solar-thermophotovoltaics, and hot-carrier-assisted devices with improved efficiencies [57][58][59][60]. ...

Highly Plasmonic Titanium Nitride by Room-Temperature Sputtering

Scientific Reports

... Our approach not only enables the scalability of the Casimir self-assembly and self-alignment but also paves the way for the integration of Casimir torque effects with colloidal science, nanophotonics, polaritonics, and self-assembly. On a broader account, it is important to mention that Casimir and Casimir torque effects studied here could be relevant for the design of microelectromechanical devices (36) and in future devices relying on an efficient and contactless transfer of angular momentum (37)(38)(39)(40). ...

Nanoscale transfer of angular momentum mediated by the Casimir torque

Communications Physics

... We first consider the quantum emitter as a two-level system dominated by an ED transition between the excited |e⟩ and ground |g⟩ states with energy difference E e − E g = ℏω 0 = ℏk 0 c. The electric Purcell factor (PF) is the modification in the SE rate due to the presence of neighboring objects and can be written as [65] Γ (e) (r) ...

Hysteresis in the spontaneous emission induced by VO2 phase change
Journal of the Optical Society of America B

Journal of the Optical Society of America B

... A blackbody is a perfect and efficient absorber owing to its broadband absorption capability of the illuminating electromagnetic energy. Numerous research efforts so far have been successfully made in realization of high performance perfect nanostructured absorber designs for number of attractive applications as nanoscale fluidics [7], nanoscale heat-source [8], photodetectors [9], microbolometers [10], thermal imaging [11], sensing [12], thermal management [13] as well as solar energy harvesting [6,[14][15][16][17]. The absorber should be capable of absorbing the solar energy from UV to near-infrared frequency, while simultaneously lowering the mid-infrared re-emissions [1]. ...

High-Temperature Refractory Metasurfaces for Solar Thermophotovoltaic Energy Harvesting

Nano Letters

... Furthermore, the PSHE has also been proposed as a tool for probing topological phase transitions in monolayer silicene [21,22]. Moreover, it has been employed to investigate moiré superlattices and twist angles in 2D systems [23][24][25]. Similarly, enhancing the PSHE on the surface of monolayer black phosphorus in the terahertz region [26][27][28] has been explored. ...

Photonic spin Hall effect in bilayer graphene Moiré superlattices

... Recent advancements in passive self-adaptive radiative cooling have led to the emergence of thermochromic materials (e.g., vanadium dioxide (VO 2 ), [37][38][39] germaniumantimony-tellurides, [40,41] perovskites, [42] thermochromic microcapsules, [43,44] thermosensitive hydrogels including poly(N-isopropylacrylamide), [45][46][47] and hydroxypropyl cellulose [48,49] ), thermally induced reconfigurable structures, [50] optically induced materials, [51] and humidity-sensitive materials [52] (Table S2, Supporting Information). These materials are intended for the modulation of thermal infrared emissivity or autonomous switching between reflecting solar energy and allowing light transmission/absorption. [53,54] Despite the progress achieved in relation to these materials, there are still bottlenecks that hinder their large-scale application. ...

Passive Radiative “Thermostat” Enabled by Phase-Change Photonic Nanostructures

ACS Photonics