Enhanced cutoff energies for direct and rescattered strong-field photoelectron emission of plasmonic nanoparticles
-
Erfan Saydanzad
, Jeffrey Powell
, Matthias F. Kling
Abstract
The efficient generation, accurate detection, and detailed physical tracking of energetic electrons are of applied interest for high harmonics generation, electron-impact spectroscopy, and femtosecond time-resolved scanning tunneling microscopy. We here investigate the generation of photoelectrons (PEs) by exposing plasmonic nanostructures to intense laser pulses in the infrared (IR) spectral regime and analyze the sensitivity of PE spectra to competing elementary interactions for direct and rescattered photoemission pathways. Specifically, we measured and numerically simulated emitted PE momentum distributions from prototypical spherical gold nanoparticles (NPs) with diameters between 5 and 70 nm generated by short laser pulses with peak intensities of 8.0 × 1012 and 1.2 × 1013 W/cm2, demonstrating the shaping of PE spectra by the Coulomb repulsion between PEs, accumulating residual charges on the NP, and induced plasmonic electric fields. Compared to well-understood rescattering PE cutoff energies for strong-field photoemission from gaseous atomic targets (10× the ponderomotive energy), our measured and simulated PE spectra reveal a dramatic cutoff-energy increase of two orders of magnitude with a significantly higher contribution from direct photoemission. Our findings indicate that direct PEs reach up to 93 % of the rescattered electron cutoff energy, in contrast to 20 % for gaseous atoms, suggesting a novel scheme for the development of compact tunable tabletop electron sources.
1 Introduction
The characterization of photoexcitation and -emission of plasmonic nanostructures is of basic research and applied interest for efficient harmonic up-conversion [1, 2], femtosecond time-resolved scanning tunneling microscopy and spectroscopy [3, 4], electron-impact spectroscopy [5, 6], and the development of compact electron sources [7]. We here show that prototypical plasmonic NPs exposed to intense IR-laser pulses emit PEs over a large kinetic energy range, owing to an intricate dynamical interplay of distinct electronic and photonic interactions. Extensively investigated during the past two decades [8, 9], metal NPs have remarkable optical properties that are primarily related to incident light in the IR to the visible frequency range enforcing the collective motion of conduction electrons. This light-driven excitation of localized surface-charge plasmons (LSP) controls the particles’ light absorption, reflection, and skin depths [10]. It results in a nanoplasmonic field near the NP surface that can greatly amplify the incident-laser electric field [11, 12]. The LSP resonance frequency of metal NPs can be tuned into resonance from IR to visible frequencies by variation of their shape, size, composition, and dielectric environment [8, 9, 13, 14]. This tunable enhanced light absorption and light scattering are key to powerful diagnostic methods, such as surface-enhanced Raman spectroscopy [15], time-resolved nanoplasmonic-field microscopy [12, 16], [17], [18], and biomedical and chemical sensing [19, 20]. The present single-pulse PE imaging investigation is expected to promote future two-pulse pump-probe experimental schemes for the spatiotemporal imaging of induced-plasmonic-field distributions near the surface of metal nanoshells that have recently been proposed in classical [21] and quantum-mechanical [12, 18, 22] numerical simulations.
In this work, we have employed 2-dimensional velocity-map-imaging (VMI) spectroscopy to investigate strong-field electron emission from metal NPs. VMI spectroscopy provides projections of PE momentum distributions onto the plane of a 2-dimensional PE detector. It is established as a powerful technique for studying intense-light interactions with atoms and molecules [23–25]. Through the last decade, this technique was applied to study strong-field photoemission from isolated NPs by intense linearly polarized laser pulses [26–28]. During strong-field emission from atoms and molecules [29], PEs can gain a significant amount of energy while propagating in the oscillating laser electric field. PEs that are “directly” emitted from gaseous atomic targets by linearly polarized laser pulses (without being driven by the external light field to return to the residual ion) gain up to 2 Up(I0) in kinetic energy, while PEs that are accelerated back to the residual ion by the laser electric field to “rescatter” elastically accumulate up to 10 Up(I0) [30–33]. The ponderomotive energy Up(I0) = I0/(4ω2) is the cycle-averaged quiver energy of a free electron in a laser field of frequency ω and peak intensity I0. Unless indicated otherwise, we use atomic units throughout this work. For strong-field PE emission and rescattering from solids [34–40] and nanostructures, such as nanotips [3, 41], [42], [43], [44], [45], [46], isolated clusters [47–51], and dielectric NPs [27, 28, 52], cutoff energies in directly emitted and rescattered photoemission from dielectric NPs were found to be approximately 2 η2Up(I0) and 10 η2Up(I0), respectively [27, 53]. Compared to atomic targets, these limiting PE energies are enhanced by the square of the near-field plasmonic-enhancement factor η. In this work, we (i) measured and numerically modeled VMI spectra resulting from the strong-field PE emission from metal NPs by intense IR-laser pulses and (ii) validated a recent extension [17] of the three-step-model for atomic strong-field ionization [54] to metal NPs. Owing to plasmonic-near-field enhancement of the incident-laser electric field and PE correlation, we found measured and calculated cutoff energies for metal NPs that exceed typical cutoff energies from gaseous atoms and dielectric NPs by two and one order of magnitude, respectively. Interestingly, the cutoff energy for direct electron emission from metal NPs reaches up to 93 % of the corresponding value for rescattered PEs electrons, dramatically exceeding the well-known proportion of 20 %, discussed earlier for gaseous atoms and dielectric NPs [27, 53].
2 Methods
2.1 Experimental setup
The laser system and VMI electron detection apparatus at the James R. Macdonald Laboratory at Kansas State University are described in more detail in [26, 55]. Briefly, the experiments used a Ti:Sapphire-based chirped pulse amplification (CPA) system generating 25 fs pulses FWHIM (10 optical cycles), and central angular frequency ω = 2.415 PHz (corresponding to a central wavelength λ = 780 nm). The laser pulse intersects the stream of isolated single NPs with diameters of 5, 30, or 70 nm that are injected by aerodynamic lens focusing [28, 55], [56], [57], [58]. As shown in the sketch of the experimental setup in Figure 1, PEs are projected onto the detector by the static electric field between the repeller and extractor. This allows the recording of the 2D projection of the PE momentum distributions as VMI maps. PE spectra were captured in a thick-lens, high-energy VMI spectrometer [59] capable of gathering up to 350 eV electron energy. The NPs were purchased from Cytodiagnostics [60]. The NP samples were custom synthesized, characterized for monodispersity (typical polydispersity index < 0.1) and sphericity (>95 %) to ensure sufficient reproducibility between interactions, and extensively purified to remove any source of contamination. We carefully chose the initial NP concentration to avoid the formation of clusters in the NP beam [61].

Schematic of the velocity-map-imaging spectrometer coupled to the nanoparticle source. The dilute beam of isolated gas-phase nanoparticles is injected into vacuum and focused by an aerodynamic lens to intersect 800 nm, 25 fs, 10 kHz-repetition-rate linearly polarized laser pulses. Emitted electrons are focused onto the microchannel plate (MCP)/phosphor assembly. VR and VE are the respective voltages on the PE repeller and extractor plates needed to guide photoelectrons to the MCP phosphorus detector. The MCP is coupled to a phosphor screen, of which a camera records the spatial distribution of photoelectron hits for every laser shot.
2.2 Laser-intensity characterization
The peak laser intensity was determined by analyzing the above-threshold-ionization (ATI) PE energy distribution from gaseous Xe atoms with the VMI spectrometer described above and for the same laser parameters we selected for the strong-field-ionization studies reported in this work. To determine the absolute value of the intensity, the ponderomotive shift of the Xe ATI comb was measured as a function of the input-laser-pulse energy. From this shift, we deduced the ponderomotive energy, UP, for a given pulse energy. Since Up is proportional to the peak laser intensity I0, the latter could be directly determined from this measurement. We determined the values of the intensities used in this work as I0 = 8.0 × 1012 W/cm2 and 1.5 I0 and estimated the accuracy of the intensity calibration to be better than 15 % (see Ref. [55, 62] for details).
2.3 Theoretical model
We numerically investigated PE emission from metallic NPs by IR-laser pulses with a Gaussian temporal profile. Propagating along the x axis and linearly polarized along the z axis, their electric field is given by
where τ is the pulse length at full-width-half-intensity maximum (FWHIM), ω the pulse’s central frequency, φ the carrier-envelope phase, and c the speed of light in vacuum (Figure 1). During the laser – NP interaction, LSPs are excited and induce an inhomogeneous plasmonic field near the NP surface. At the same time, and most significantly at the LSP resonance frequency [63, 64], electrons are excited to electronic states above the Fermi level. Sufficiently high laser intensities generate multiply ionized NPs [26, 56].
The incident laser pulse induces a transient dipole in the NP. Within the electric-dipole approximation, the corresponding transient induced plasmonic-dipole moment,
where k = 2π/λ = ω/c. We calculate the complex NP polarizability, αMie(ω), within Mie theory [66], following Ref. [67], which restricts the applicability of Eq. (2) to size parameters ka ⪅ 0.6 for nanospheres of radius a [68].
We describe strong-field ionization from metal NPs by extending the semi-classical three-step model (also known as “simple-man model”) for atomic strong-field ionization to metal NPs [17]. Our extended three-step model consists of: (1) electron release based on quantum-mechanical tunneling, (2) PE propagation from the NP surface to the detector by sampling over classical trajectories, and (3) PE rescattering and recombination at the NP surface. In comparison with gaseous atomic targets, each of these steps is significantly more intricate for metal NPs, due to their more complex electronic structure, the added morphological structure, and the emission of a much larger number of electrons, emphasizing the effects of PE – PE correlation, residual charges, and PE – nanoplasmonic-field interactions.
We represent the NPs’ static electronic structure in terms of the surface-potential step V0 = ɛ
F
+ φ with the work function φ = 5.1 eV and Fermi energy ɛ
F
= 8.0 eV for bulk gold [69]. Our dynamical numerical simulation divides the NP surface into small surface elements. During successive small time intervals, the surface elements are modeled as spherical square-well potentials. Bound PEs close to the NP surface tunnel out along the radial component of the total electric field at the NP surface,
In our numerical applications in Section 3, we distinguish and compare specular and diffusive PE rescattering at the NP surface. For diffusive rescattering, we uniformly randomize the polar and azimuthal scattering angles relative to the surface normal at the impact site on the NP surface, modeling rescattering in all accessible directions with equal probability.
3 Experimental and simulation results
3.1 Influence of nanoplasmonic field, rescattering, residual-charge interactions, and photoelectron correlation
VMI spectra are sensitive to all PE interactions included in our simulation. In order to track the effects of different electronic interactions on the propagation and rescattering of released PEs, we leave the modeling of the tunneling release of electrons at the NP surface unchanged when selectively switching off individual PE interactions (for identical laser-pulse parameters), assuming for all calculated VMI maps identical tunneling-ionization rates (Eq. (S1.5) in the Supplementary Information) The comparison of simulations in which we selectively include and exclude specific PE interactions during the PE propagation and rescattering, allows us to quantify their specific effects on VMI maps.
Figure 2 shows simulated VMI spectra compared to experimental results for gold nanospheres with 30 nm diameter for the experimental setup depicted in Figure 1. The VMI spectra are projections of the PE momentum distribution on the x–z plane of the MCP detector and show the projected PE yields as functions of the PE asymptotic velocities, v
x
and v
z
, along the laser-propagation and -polarization directions. The first, second, and third column in Figure 2 include, respectively, VMI spectra of direct PEs, rescattered PEs, and the net PE yield for either specular (first and second row) or diffuse rescattering (third and fourth row, cf., Sec. (S3) in the SI). The first and third row show simulations in which only the incident-laser and plasmon fields (

(a–m) Photoelectron VMI spectra simulated for 30 nm diameter gold nanospheres for direct (first column), rescattered (second column), and all (denoted as “Net”, third column) photoelectrons, including either specular (first and second row) or diffuse rescattering (third and fourth row) for incident 780 nm laser pulses with a pulse length of 25 fs (FWHIM) and 8.0 × 1012 W/cm2 peak intensity. μ designates the integrated yield, normalized to the simulated net PE yield. First and third row: simulations where only the incident-laser and plasmon fields (
To allow for a quantitative comparison of direct and rescattered PE yields, we normalized the yields in each row to the corresponding net PE yield in the third column and display the normalized integrated yield μ in each graph of Figure 2. We calculated μ as the v x - and v z -integrated yields from the simulated VMI maps in each row, divided by the corresponding integrated yield of the VMI maps in the third column. The comparison of the VMI spectra in Figure 2 allows us to assess the influence of the distinct PE interactions on VMI spectra, as we discuss next.
3.1.1 Plasmonic-field interactions
The simulated VMI spectra in the first and third row of Figure 2 are calculated under the assumption that released electrons solely interact with the incident-laser and induced plasmonic fields while propagating to the detector. These PE distributions are aligned with the laser-polarization direction and have a dipole-like appearance, owing to the dipole character transferred from the induced plasmonic field and tunneling ionization.
The comparison of Figure 2(a) and (b) with Figure 2(c) for specular rescattering and Figure 2(g) and (h) with Figure 2(i) for diffuse rescattering reveals that directly emitted PEs dominate the low-energy part of the photoemission spectra. Rescattered PEs, in contrast, can gain additional energy from the laser and induced plasmonic fields and establish the higher-energy part of the PE spectrum. Rescattering boosting PE energies is a well-understood phenomenon in strong-field ionization. For gaseous atomic targets, elastically rescattered PEs reach kinetic energies up to 10 Up(I0) [30–33] and larger energies occur for dielectric NPs (SiO2) [26, 27, 56]. By comparing the yield factors μ in the first and second row, we find that approximately 83 % of the detected PEs is directly emitted, while 17 % have rescattered at the NP surface at least once.
3.1.2 All interactions effect
The second (specular rescattering) and forth (diffuse rescattering) row of Figure 2 show simulated VMI spectra including all PE interactions, i.e.,
As noted above, in the absence of PE–PE interactions and diffuse rescattering, the linearly polarized incident-laser and induced plasmonic electric field imprint their dipole character on the VMI spectra. The inclusion of PE–PE interactions and diffuse rescattering partially removes the dipolar emission character and results in more isotropic VMI spectra [17]. For metal NPs, attractive residual-charge interactions are thus much less influential than PE–PE interactions in shaping PE momentum distributions and determining PE cutoff energies.
Comparing the VMI spectra in rows one and two of Figure 2 for specular and in rows three and four for diffuse rescattering, we notice that the combined effect of
Figure 2(n) shows our experimental VMI spectra. With regard to yield, cutoff energy, and isotropic shape of the PE momentum distribution, Figure 2(m) (including all interactions and with diffuse rescattering) is our most comprehensive simulation result and matches the experiment well. The VMI spectra in Figure 2 clearly show that all PE interactions are relevant for shaping the PE angular distribution in the measured VMI spectrum in Figure 2(n).
3.2 Influence of nanoparticle size and laser intensity
Figure 3 shows simulated and experimental VMI spectra for gold nanospheres with diameters of 5, 30, and 70 nm. The first, second, and third column are simulated VMI spectra for, respectively, the direct, rescattered, and net PEs yield for peak laser intensities I0 (first, second, and third row) and 1.5 I0 (fourth, fifth, and sixth row). Experimental results corresponding to the simulations in the third column are shown in the fourth column. The VMI spectra in Figure 3 are (slightly) elongated along the laser-polarization direction, with PE cutoff energies that increase with NP size. As discussed in Section 3.1, isotropic VMI spectra are promoted by PE–PE interactions and diffuse PE rescattering from the NP surface, while incident-laser and induced plasmonic-field interactions tend to imprint a dipolar shape. The detected number of the PEs per laser short for the experimental data shown in Figure 3 varies from 140 for 5 nm diameter NPs at the lower peak laser intensity (I0) to 600 for 70 nm NPs at 1.5 I0. However, as discussed in detail in Ref. [55], these numbers do not directly reflect the number of PEs that hit the detector due to the PE energy-dependent detector saturation in our experiment. The saturation effect is most prominent in the central detection area, where low-energy electrons (which dominate the total PE yield) hit the MCP.

Comparison of simulated direct (first column), rescattered (second column) and net (i.e., including direct and rescattered yields, third column) photoelectron VMI spectra with experimental (forth column) VMI spectra for gold nanospheres with 5, 30, and 70 nm diameter and laser peak intensities of I0 = 8.0 × 1012 W/cm2 (first – third row) and 1.5 I0 (forth – sixth row). The laser-pulse length and wavelength are 25 fs and 780 nm. Red dashed circles in (a–x) indicate simulated and experimental photoelectron cutoff energies. μ is the integrated photoelectron yield normalized to the integrated net yields in third column.
To allow for a quantitative comparison of direct and rescattered PE yields, we normalized the direct and rescattered PE yields in each row to the corresponding net PE yield in the third column and displayed the normalized integrated yield μ(a, I0) in each graph. μ reveals that the yield of direct PEs decreases as a function of the NP size and intensity, being more sensitive to the size. This observation is compatible with PEs having a higher probability to rescatter off larger NPs. In addition, increasing laser intensity leads to a stronger radial attractive force, due to an increase in the number of residual charges on the NP surface, leading to more PE rescattering events. The direct and rescattered PE yield can be controlled by the intensity of the laser pulse and size of the NP. The measured and simulated VMI maps also reveal a large increase in the direct and rescattered PE cutoff energy with the laser peak intensity and NP size. We quantify this laser-intensity and NP-size-dependent effect in the following subsection.
3.3 Angle-integrated photoelectron yields and cutoff energies
Figure 4 shows simulations corresponding to the VMI spectra in Figure 3. It includes (i) all interactions for the direct PE yield (denoted as “All_Direct”), (ii) all interactions for the rescattered PE yield (“All_Rescat”), and (iii) all interactions for the net PE yield (“All_Net”). In addition, Figure 4 displays (iv) simulations only including incident- and plasmonic-field interactions for the net PE yield (denoted as “Inc + Pl_Net”) and (v) integrated experimental yields as a function of the PE kinetic energy. Due to the detector saturation at the center of the MCP phosphor detector (Figure 1), the experimental yields for kinetic energies below approximately 8 eV (corresponding to PE velocities below 0.8 a.u.) are not accurate. To be able to compare experimental integrated yields to one another and to the simulation results, we have removed the low energy part of the integrated yields from both, experimental and simulated data.

Photoemission yields as functions of the photoelectron kinetic energy for gold nanospheres with (a) and (d) 5, (b) and (e) 30, and (c) and (f) 70 nm diameter and laser peak intensities of (a–c) I0 = 8.0 × 1012 W/cm2 and (d–f) 1.5 I0. The laser-pulse length and wavelength are 25 fs and 780 nm. Simulated photoelectron yields including all interactions are shown for directly emitted (“All_Direct”), rescattered (“All_Rescat”), and net (“All_Net”, i.e., direct and rescattered) photoelectrons. Simulations only including incident- and plasmonic-field interactions are denoted as “Inc + Pl_Net”. Black dots show experimental yields.
The overall agreement between experimental and simulated integrated PE yields in Figure 4 is not perfect for several reasons. With regard to the simulation, an important uncertainty derives from our implementation of approximate modified Fowler–Nordheim tunneling rates. With regard to the experiment, the above-mentioned detector saturation decreases the reliability of the low-energy portion of our spectra. While the low-energy portion of the simulation data was truncated to allow for a better comparison with the experiment, the detection uncertainty due to saturation is not completely removed and tends to affect predominantly our measurements for the largest NP size (70 nm diameter) and the higher laser peak intensity (1.5 I0), due to larger numbers of emitted PEs per laser shot. This is consistent with the agreement between simulation and measurement being better for 30 nm NPs at the lower peak intensity (I0) in Figure 4 than for 70 nm NPs in Figure 4(c) and at the higher laser intensity of 1.5 I0 in Figure 4(d)−(f). However, in view of hardly avoidable inaccuracies in the detailed modeling of this complex interaction scenario and NP-size- and laser-intensity-dependent experimental background noise, we cannot exclude that the exceptionally good match between experimental simulated results shown in 4(b) compared to the other graphs in this figure is serendipitous.
Integration of the VMI-projected PE momentum distributions y(v x , v z ) in Figure 3(a)–(x) over the PE detection angle ϕ in the VMI plane results in PE yields
as functions of the PE energy,
For simulated yields, we define the PE cutoff energy Ecutoff as the energy up to which 99.5 % of the net PE yield has accumulated,
The experimental cutoff energy was extracted from the experimental VMI maps as described in [26, 55], for which the upper energy boundaries of the full 3D momentum sphere and the 2D projection are identical. The radial distribution of these projections along the polarization direction accurately determines the maximum PE energy.
The PE cutoff energies, shown as red dashed circles in Figure 3 in Section 3.2, increase with the NP size and peak laser intensity. Figure 5(a) and (b) display cutoff energies as a function of NP size for peak intensities of I0 = 8.0 × 1012 W/cm2 and 1.5 I0, in units of the incident-laser ponderomotive energies Up(I0) and Up(1.5 I0), respectively. Blue diamonds and red circles show, respectively, simulated cutoff energies (including all interactions) for the direct (denoted as “All_Direct”) and net (“All_Net”) PE yields. Gray squares with error bars are experimental cutoff energies (“Experiment”). Simulation results including all interactions for rescattered PEs are not shown, because they coincide with the “All_Net” yield. For gaseous atomic targets, the cutoff energy is equal to 10 Up [30–33]. Cutoff energies obtained by scaling this well-known expression by the plasmonic intensity enhancement of the incident-laser pulse, η2, are shown as yellow “plus” markers. As expected, they tend to merge with the cutoff energies computed while only including incident-laser-pulse and plasmonic-field interactions (represented as green triangles). We calculated the applied value for η within Mie theory at the poles (relative to the laser-polarization direction) of the NPs [66–68]. In contrast to the 10 η2Up scaling, the comparison of Figure 5(a) and (b), shows that our theoretical cutoff energies predict intensity-dependent changes that become more pronounced for larger NPs. Within the experimental error this theoretical prediction is compatible with our experimental results.

Comparison of simulated and experimental photoelectron cutoff energies scaled by the incident-laser ponderomotive energy Up(I0) for 5, 30, and 70 nm diameter gold nanospheres and laser peak intensities of (a) I0 = 8.0 × 1012 W/cm2 and (b) 1.5 I0. The laser-pulse length and wavelength are the same as in Figure 4. Simulated cutoff energies including all interactions for, respectively, direct (“All_Direct”) and net (“All_Net”, i.e., direct and rescattered) photoemission. Simulations only including incident- and plasmonic-field interactions are denoted as “Inc + Pl_Net”. Yellow “plus” markers show atomic cutoff energies, 10 Up, scaled by the plasmonic intensity enhancement η2.
Based on the discussion in Section 3.1 of different PE interactions and their influence on VMI maps, we investigated two plausible causes for the numerically predicted increase of the PE yield and cutoff energy with the NP size. First, the lowering and narrowing of the surface-potential barrier by the more significant nanoplasmonic-field enhancement near larger NPs [12, 18, 21, 68] promotes strong-field tunneling ionization. However, this not only tends to augment the measured PE yield. Since PEs gain more energy in a more strongly enhanced field, it also entails higher cutoff energies for larger NPs. Second, as the NP size increases, a larger surface area becomes available from where more electrons are emitted, increasing the PE yield. The cutoff energy rises with the PE yield due to the increased repulsive Coulomb energy between PEs upon their release from the NP surface. In principle, a third cause for larger yields and cutoff energies can be laser-pulse-propagation effects inside the NP that result in higher local-field enhancements for larger NPs [56]. However, for the NP sizes investigated here, we did not find this effect to be relevant.
As discussed in Section 3.1, the consequences of residual-charge interactions and PE–PE interactions oppose each other. While attractive residual-charge – PE interactions reduce both, PE yields and cutoff energies, PE Coulomb repulsion increases them. A detailed numerical comparison of these competing interactions is shown in Figure 3 of Ref. [17]. Our numerical results indicate that PE Coulomb repulsion overcompensates residual-charge – PE interactions with regard to the cutoff energy, leading to an overall cutoff-energy increase, especially for larger NPs.
The green triangles (denoted as “Inc + Pl_Net”) in Figure 5 are cutoff energies calculated under the assumption that released electrons solely interact with the incident-laser and induced plasmonic field while propagating to the detector. In Sec. (S4) of the SI, we derive a closed-form analytical heuristic expression for the cutoff energy in direct photoemission, based on a simplified central-field approximation of residual-charge interactions and PE correlation. By comparing Eq. (S4.12) in the Sec. (S4) with the known respective 2 Up and 10 UP limits for direct and rescattered emission for atomic strong-field ionization, we infer the cutoff energy for rescattered PEs,
where
The heuristic Eq. (5) qualitatively explains all experimental results in Figure 5. For very small
An unexpected and interesting result derives from the fact that the simulated cutoff energies for direct PEs (“All_Direct”) and rescattered PEs (“All_Net”) are comparable. The direct cutoff energy for 5, 30, and 70 nm NPs are, respectively, 93, 85, and 89 % of the rescattered PE cutoffs for the intensity I0 and 93, 87, and 84 % for 1.5 I0. This value is 20 % for strong-field ionization of gaseous atoms, molecules, and dielectric NPs.
4 Summary and conclusions
We measured and numerically simulated VMI maps to model strong-field ionization from metal NPs. Our experimental and simulated results scrutinize a complex dynamical interplay of PE emission, propagation, recombination, and rescattering. Augmented by strong plasmonic-field enhancement, a large number of PEs tunnel ionize from metal NPs and result in high PE yields and cutoff energies. We analyzed the size and laser-intensity dependence of PE angular distributions in light of competing contributions from various PE interactions.
We observed that the dipolar shape, imprinted on VMI maps by the incident-laser and induced plasmonic fields, is mostly erased by PE correlation and diffusive rescattering at the NP surface to yield almost isotropic VMI maps. While for gaseous atomic targets, directly emitted PEs acquire no more than about 20 % of the cutoff energy of rescattered PEs [10 Up(I0)], we found direct photoemission from metal NPs to yield cutoff energies up to 303 Up(I0), reaching between 84 and 93 % of the cutoff energy for rescattered PEs. Due to (exponentially) laser-intensity-dependent PE emission, the effects of residual charges and PE–PE interactions are strongly intensity dependent. This leads to a nonlinear intensity-dependence of the PE yield and cutoff energy scaling with Up(I0), contrary to the known linear intensity scaling for gaseous atomic targets.
Our joint experimental and theoretical investigation of a prototypical light-driven nanoplasmonic system supports the use of plasmonic nanostructures towards the development of tunable compact electron and radiation sources for PE and radiation imaging in basic research and for novel photoelectronic detection, catalytic, and light-collecting devices.
Funding source: Air Force Office of Scientific Research
Award Identifier / Grant number: FA9550-21-1-0387
Award Identifier / Grant number: FA9550-17-1-0369
Funding source: U.S. Department of Energy
Award Identifier / Grant number: Unassigned
Funding source: US NSF
Award Identifier / Grant number: NSF/PHY 2110633
Funding source: Chemical Sciences, Geosciences, and Biosciences Division, Office of Basic Energy Sciences, Office of Science, U.S. DOE
Award Identifier / Grant number: DE-FG02-86ER13491
Award Identifier / Grant number: DE-AC02-76SF00515
Award Identifier / Grant number: DE-SC0063
Acknowledgment
We thank Reza Mazloom, and Aram Vajdi for stimulating discussions. J.P, E.S., and C.T. were supported in part by the Air Force Office of Scientific Research award no. FA9550-17-1-0369 and FA9550-21-1-0387 (Recollision physics at the nanoscale). E.S and U.T. were supported in part by NSF grant PHY 2110633 (Transient strong-field dielectric response of nanoparticles). A.R., A.S., S.J.R., and U.T. acknowledge partial support by the Chemical Sciences, Geosciences, and Biosciences Division, Office of Basic Energy Sciences, Office of Science, U.S. DOE under award No. DE-FG02-86ER13491. A.S. and M.F.K.’s work at SLAC is supported by the U.S. Department of Energy, Office of Science, Basic Energy Sciences, Scientific User Facilities Division, under Contract No. DE-AC02-76SF00515 and by the Chemical Sciences, Geosciences, and Biosciences division with award DE-SC0063.
-
Author contribution: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.
-
Research funding: U.S. Air Force Office of Scientific Research (awards FA9550-17-1-0369 and FA9550-21-1-0387); U.S. NSF (grant PHY 2110633); Chemical Sciences, Geosciences, and Biosciences Division, Office of Basic Energy Sciences, Office of Science, U.S. DOE (contracts DE-FG02-86ER13491, DE-AC02-76SF00515, and DE-SC0063).
-
Conflict of interest statement: The authors declare no conflicts of interest regarding this article.
References
[1] C. Winterfeldt, C. Spielmann, and G. Gerber, “Colloquium: optimal control of high-harmonic generation,” Rev. Mod. Phys., vol. 80, no. 1, p. 117, 2008. https://doi.org/10.1103/revmodphys.80.117.Search in Google Scholar
[2] S. Ghimire and D. A. Reis, “High-harmonic generation from solids,” Nat. Phys., vol. 15, no. 1, pp. 10–16, 2019. https://doi.org/10.1038/s41567-018-0315-5.Search in Google Scholar
[3] M. Müller, V. Kravtsov, A. Paarmann, M. B. Raschke, and R. Ernstorfer, “Nanofocused plasmon-driven sub-10 fs electron point source,” ACS Photonics, vol. 3, no. 4, pp. 611–619, 2016. https://doi.org/10.1021/acsphotonics.5b00710.Search in Google Scholar
[4] S. Yoshida, Y. Arashida, H. Hirori, et al.., “Terahertz scanning tunneling microscopy for visualizing ultrafast electron motion in nanoscale potential variations,” ACS Photonics, vol. 8, no. 1, pp. 315–323, 2021. https://doi.org/10.1021/acsphotonics.0c01572.Search in Google Scholar
[5] K. Walzl, C. Koerting, and A. Kuppermann, “Electron-impact spectroscopy of acetaldehyde,” J. Chem. Phys., vol. 87, no. 7, pp. 3796–3803, 1987. https://doi.org/10.1063/1.452935.Search in Google Scholar
[6] K. Venkatraman, B. D. Levin, K. March, P. Rez, and P. A. Crozier, “Vibrational spectroscopy at atomic resolution with electron impact scattering,” Nat. Phys., vol. 15, no. 12, pp. 1237–1241, 2019. https://doi.org/10.1038/s41567-019-0675-5.Search in Google Scholar
[7] J. Schötz, L. Seiffert, A. Maliakkal, et al.., “Onset of charge interaction in strong-field photoemission from nanometric needle tips,” Nanophotonics, vol. 10, no. 14, pp. 3769–3775, 2021. https://doi.org/10.1515/nanoph-2021-0276.Search in Google Scholar
[8] K. L. Kelly, E. Coronado, L. L. Zhao, and G. C. Schatz, “The optical properties of metal nanoparticles: the influence of size, shape, and dielectric environment,” J. Phys. Chem. B, vol. 107, no. 3, pp. 668–677, 2003. https://doi.org/10.1021/jp026731y.Search in Google Scholar
[9] L. Wang, M. Hasanzadeh Kafshgari, and M. Meunier, “Optical properties and applications of plasmonic-metal nanoparticles,” Adv. Funct. Mater., vol. 30, no. 51, p. 2005400, 2020. https://doi.org/10.1002/adfm.202005400.Search in Google Scholar
[10] J. A. Powell, J. Li, A. Summers, et al.., “Strong-field control of plasmonic properties in core–shell nanoparticles,” ACS Photonics, vol. 9, pp. 3515–3521, 2022. https://doi.org/10.1021/acsphotonics.2c00663.Search in Google Scholar
[11] M. I. Stockman, “Nanoplasmonics: the physics behind the applications,” Phys. Today, vol. 64, no. 2, p. 39, 2011. https://doi.org/10.1063/1.3554315.Search in Google Scholar
[12] J. Li, E. Saydanzad, and U. Thumm, “Attosecond time-resolved streaked photoelectron spectroscopy of transition-metal nanospheres,” Phys. Rev. A, vol. 95, p. 043423, 2017. https://doi.org/10.1103/physreva.95.043423.Search in Google Scholar
[13] S. Link and M. A. El-Sayed, “Size and temperature dependence of the plasmon absorption of colloidal gold nanoparticles,” J. Phys. Chem. B, vol. 103, no. 21, pp. 4212–4217, 1999. https://doi.org/10.1021/jp984796o.Search in Google Scholar
[14] P. K. Jain, X. Huang, I. H. El-Sayed, and M. A. El-Sayed, “Review of some interesting surface plasmon resonance-enhanced properties of noble metal nanoparticles and their applications to biosystems,” Plasmonics, vol. 2, no. 3, pp. 107–118, 2007. https://doi.org/10.1007/s11468-007-9031-1.Search in Google Scholar
[15] E. Le Ru and P. Etchegoin, Principles of Surface-Enhanced Raman Spectroscopy: And Related Plasmonic Effects, Oxford, Elsevier, 2008.10.1016/B978-0-444-52779-0.00005-2Search in Google Scholar
[16] M. I. Stockman, M. F. Kling, U. Kleineberg, and F. Krausz, “Attosecond nanoplasmonic-field microscope,” Nat. Photonics, vol. 1, no. 9, pp. 539–544, 2007. https://doi.org/10.1038/nphoton.2007.169.Search in Google Scholar
[17] E. Saydanzad, J. Li, and U. Thumm, “Strong-field ionization of plasmonic nanoparticles,” Phys. Rev. A, vol. 106, p. 033103, 2022. https://doi.org/10.1103/physreva.106.033103.Search in Google Scholar
[18] J. Li, E. Saydanzad, and U. Thumm, “Retrieving plasmonic near-field information: a quantum-mechanical model for streaking photoelectron spectroscopy of gold nanospheres,” Phys. Rev. A, vol. 94, p. 051401, 2016. https://doi.org/10.1103/physreva.94.051401.Search in Google Scholar
[19] A. Kabashin, P. Evans, S. Pastkovsky, et al.., “Plasmonic nanorod metamaterials for biosensing,” Nat. Mater., vol. 8, no. 11, p. 867, 2009. https://doi.org/10.1038/nmat2546.Search in Google Scholar PubMed
[20] K. A. Willets and R. P. Van Duyne, “Localized surface plasmon resonance spectroscopy and sensing,” Annu. Rev. Phys. Chem., vol. 58, pp. 267–297, 2007. https://doi.org/10.1146/annurev.physchem.58.032806.104607.Search in Google Scholar PubMed
[21] E. Saydanzad, J. Li, and U. Thumm, “Spatiotemporal imaging of plasmonic fields near nanoparticles below the diffraction limit,” Phys. Rev. A, vol. 98, p. 063422, 2018. https://doi.org/10.1103/physreva.98.063422.Search in Google Scholar
[22] J. Li, E. Saydanzad, and U. Thumm, “Imaging plasmonic fields with atomic spatiotemporal resolution,” Phys. Rev. Lett., vol. 120, p. 223903, 2018. https://doi.org/10.1103/physrevlett.120.223903.Search in Google Scholar
[23] J. Maurer, D. Dimitrovski, L. Christensen, L. B. Madsen, and H. Stapelfeldt, “Molecular-frame 3d photoelectron momentum distributions by tomographic reconstruction,” Phys. Rev. Lett., vol. 109, p. 123001, 2012. https://doi.org/10.1103/physrevlett.109.123001.Search in Google Scholar PubMed
[24] H. V. S. Lam, T. N. Wangjam, and V. Kumarappan, “Alignment dependence of photoelectron angular distributions in the few-photon ionization of molecules by ultraviolet pulses,” Phys. Rev. A, vol. 105, p. 053109, 2022. https://doi.org/10.1103/physreva.105.053109.Search in Google Scholar
[25] M. Meckel, D. Comtois, D. Zeidler, et al.., “Laser-induced electron tunneling and diffraction,” Science, vol. 320, no. 5882, pp. 1478–1482, 2008. https://doi.org/10.1126/science.1157980.Search in Google Scholar PubMed
[26] J. A. Powell, A. M. Summers, Q. Liu, et al.., “Interplay of pulse duration, peak intensity, and particle size in laser-driven electron emission from silica nanospheres,” Opt. Express, vol. 27, no. 19, pp. 27124–27135, 2019. https://doi.org/10.1364/oe.27.027124.Search in Google Scholar PubMed
[27] S. Zherebtsov, T. Fennel, J. Plenge, et al.., “Controlled near-field enhanced electron acceleration from dielectric nanospheres with intense few-cycle laser fields,” Nat. Phys., vol. 7, no. 8, pp. 656–662, 2011. https://doi.org/10.1038/nphys1983.Search in Google Scholar
[28] L. Seiffert, Q. Liu, S. Zherebtsov, et al.., “Attosecond chronoscopy of electron scattering in dielectric nanoparticles,” Nat. Phys., vol. 13, no. 8, p. 766, 2017. https://doi.org/10.1038/nphys4129.Search in Google Scholar
[29] L. Keldysh, “Ionization in the field of a strong electromagnetic wave,” Sov. Phys. JETP, vol. 20, no. 5, pp. 1307–1314, 1965.Search in Google Scholar
[30] G. G. Paulus, W. Becker, W. Nicklich, and H. Walther, “Rescattering effects in above-threshold ionization: a classical model,” J. Phys. B: At., Mol. Opt. Phys., vol. 27, no. 21, pp. L703–L708, 1994. https://doi.org/10.1088/0953-4075/27/21/003.Search in Google Scholar
[31] B. Walker, B. Sheehy, K. C. Kulander, and L. F. DiMauro, “Elastic rescattering in the strong field tunneling limit,” Phys. Rev. Lett., vol. 77, pp. 5031–5034, 1996. https://doi.org/10.1103/physrevlett.77.5031.Search in Google Scholar PubMed
[32] W. Becker, F. Grasbon, R. Kopold, D. Milošević, G. Paulus, and H. Walther, “Above-threshold ionization: from classical features to quantum effects,” Adv. At., Mol., Opt. Phys., vol. 48, pp. 35–98, 2002.10.1016/S1049-250X(02)80006-4Search in Google Scholar
[33] W. Becker, S. P. Goreslavski, D. B. Milošević, and G. G. Paulus, “The plateau in above-threshold ionization: the keystone of rescattering physics,” J. Phys. B, vol. 51, no. 16, p. 162002, 2018. https://doi.org/10.1088/1361-6455/aad150.Search in Google Scholar
[34] S. Ghimire, A. D. DiChiara, E. Sistrunk, P. Agostini, L. F. DiMauro, and D. A. Reis, “Observation of high-order harmonic generation in a bulk crystal,” Nat. Phys., vol. 7, no. 2, pp. 138–141, 2011. https://doi.org/10.1038/nphys1847.Search in Google Scholar
[35] M. J. Ambrosio and U. Thumm, “Energy-resolved attosecond interferometric photoemission from Ag(111) and Au(111) surfaces,” Phys. Rev. A, vol. 97, p. 043431, 2018. https://doi.org/10.1103/physreva.97.043431.Search in Google Scholar
[36] M. J. Ambrosio and U. Thumm, “Spatiotemporal analysis of a final-state shape resonance in interferometric photoemission from Cu(111) surfaces,” Phys. Rev. A, vol. 100, p. 043412, 2019. https://doi.org/10.1103/physreva.100.043412.Search in Google Scholar
[37] F. Navarrete and U. Thumm, “Two-color-driven enhanced high-order harmonic generation in solids,” Phys. Rev. A, vol. 102, p. 063123, 2020. https://doi.org/10.1103/physreva.102.063123.Search in Google Scholar
[38] V. E. Nefedova, S. Fröhlich, F. Navarrete, et al.., “Enhanced extreme ultraviolet high-harmonic generation from chromium-doped magnesium oxide,” Appl. Phys. Lett., vol. 118, no. 20, p. 201103, 2021. https://doi.org/10.1063/5.0047421.Search in Google Scholar
[39] L. Xue, S. Liu, Y. Hang, et al.., “Unraveling ultrafast photoionization in hexagonal boron nitride,” arXiv preprint arXiv:2101.10429, 2021.Search in Google Scholar
[40] F. Navarrete, M. F. Ciappina, and U. Thumm, “Crystal-momentum-resolved contributions to high-order harmonic generation in solids,” Phys. Rev. A, vol. 100, p. 033405, 2019. https://doi.org/10.1103/physreva.100.033405.Search in Google Scholar
[41] G. Herink, D. R. Solli, M. Gulde, and C. Ropers, “Field-driven photoemission from nanostructures quenches the quiver motion,” Nature, vol. 483, no. 7388, pp. 190–193, 2012. https://doi.org/10.1038/nature10878.Search in Google Scholar PubMed
[42] M. Krüger, M. Schenk, and P. Hommelhoff, “Attosecond control of electrons emitted from a nanoscale metal tip,” Nature, vol. 475, no. 7354, pp. 78–81, 2011. https://doi.org/10.1038/nature10196.Search in Google Scholar PubMed
[43] G. Wachter, “Simulation of condensed matter dynamics in strong femtosecond laser pulses,” Ph.D. thesis, Vienna University of Technology, 2014.Search in Google Scholar
[44] M. Krüger, M. Schenk, P. Hommelhoff, G. Wachter, C. Lemell, and J. Burgdörfer, “Interaction of ultrashort laser pulses with metal nanotips: a model system for strong-field phenomena,” New J. Phys., vol. 14, no. 8, p. 085019, 2012. https://doi.org/10.1088/1367-2630/14/8/085019.Search in Google Scholar
[45] G. Wachter, C. Lemell, J. Burgdörfer, M. Schenk, M. Krüger, and P. Hommelhoff, “Electron rescattering at metal nanotips induced by ultrashort laser pulses,” Phys. Rev. B, vol. 86, p. 035402, 2012. https://doi.org/10.1103/physrevb.86.035402.Search in Google Scholar
[46] H. Kim, M. Garg, S. Mandal, L. Seiffert, T. Fennel, and E. Goulielmakis, “Attosecond field emission,” Nature, vol. 613, pp. 662–666, 2023. https://doi.org/10.1038/s41586-022-05577-1.Search in Google Scholar PubMed PubMed Central
[47] T. Ditmire, J. Zweiback, V. Yanovsky, T. Cowan, G. Hays, and K. Wharton, “Nuclear fusion from explosions of femtosecond laser-heated deuterium clusters,” Nature, vol. 398, no. 6727, pp. 489–492, 1999. https://doi.org/10.1038/19037.Search in Google Scholar
[48] J. Passig, R. Irsig, N. X. Truong, T. Fennel, J. Tiggesbäumker, and K. H. Meiwes-Broer, “Nanoplasmonic electron acceleration in silver clusters studied by angular-resolved electron spectroscopy,” New J. Phys., vol. 14, no. 8, p. 085020, 2012. https://doi.org/10.1088/1367-2630/14/8/085020.Search in Google Scholar
[49] C. Varin, C. Peltz, T. Brabec, and T. Fennel, “Light wave driven electron dynamics in clusters,” Ann. Phys., vol. 526, nos. 3–4, pp. 135–156, 2014. https://doi.org/10.1002/andp.201490001.Search in Google Scholar
[50] T. Lünskens, P. Heister, M. Thämer, C. A. Walenta, A. Kartouzian, and U. Heiz, “Plasmons in supported size-selected silver nanoclusters,” Phys. Chem. Chem. Phys., vol. 17, no. 27, pp. 17541–17544, 2015. https://doi.org/10.1039/c5cp01582k.Search in Google Scholar PubMed PubMed Central
[51] Z. Wang, A. Camacho Garibay, H. Park, et al.., “Universal high-energy photoelectron emission from nanoclusters beyond the atomic limit,” Phys. Rev. Lett., vol. 124, p. 173201, 2020. https://doi.org/10.1103/physrevlett.124.173201.Search in Google Scholar PubMed
[52] F. Süßmann and M. F. Kling, “Attosecond nanoplasmonic streaking of localized fields near metal nanospheres,” Phys. Rev. B, vol. 84, p. 121406, 2011. https://doi.org/10.1103/physrevb.84.121406.Search in Google Scholar
[53] L. Seiffert, S. Zherebtsov, M. F. Kling, and T. Fennel, “Strong-field physics with nanospheres,” Adv. Phys. X, vol. 7, no. 1, p. 2010595, 2022. https://doi.org/10.1080/23746149.2021.2010595.Search in Google Scholar
[54] P. B. Corkum, “Plasma perspective on strong field multiphoton ionization,” Phys. Rev. Lett., vol. 71, pp. 1994–1997, 1993. https://doi.org/10.1103/physrevlett.71.1994.Search in Google Scholar PubMed
[55] J. A. Powell, “Strong-field driven dynamics of metal and dielectric nanoparticles,” Ph.D. thesis, Kansas State University, 2017.Search in Google Scholar
[56] F. Süßmann, L. Seiffert, S. Zherebtsov, et al.., “Field propagation-induced directionality of carrier-envelope phase-controlled photoemission from nanospheres,” Nat. Commun., vol. 6, p. 7944, 2015. https://doi.org/10.1038/ncomms8944.Search in Google Scholar PubMed PubMed Central
[57] J. L. Ellis, D. D. Hickstein, W. Xiong, et al.., “Materials properties and solvated electron dynamics of isolated nanoparticles and nanodroplets probed with ultrafast extreme ultraviolet beams,” J. Phys. Chem. Lett., vol. 7, no. 4, pp. 609–615, 2016. https://doi.org/10.1021/acs.jpclett.5b02772.Search in Google Scholar PubMed
[58] M. Davino, T. Saule, N. Helming, J. A. Powell, and C. Trallero-Herrero, “Characterization of an aerosolized nanoparticle beam beyond the diffraction limit through strong field ionization,” Sci. Rep., vol. 12, p. 9277, 2022. https://doi.org/10.1038/s41598-022-13466-w.Search in Google Scholar PubMed PubMed Central
[59] N. G. Kling, D. Paul, A. Gura, et al.., “Thick-lens velocity-map imaging spectrometer with high resolution for high-energy charged particles,” JINST, vol. 9, no. 05, p. P05005, 2014. https://doi.org/10.1088/1748-0221/9/05/p05005.Search in Google Scholar
[60] We purchased the gold nanoparticles in 2017 from Cytodiagnostics Inc., https://www.cytodiagnostics.com/collections/reactant-free-gold-nanoparticles/Reactant-Free-Gold-Nanoparticles.Search in Google Scholar
[61] P. Rosenberger, P. Rupp, R. Ali, et al.., “Near-field induced reaction yields from nanoparticle clusters,” ACS Photonics, vol. 7, no. 7, pp. 1885–1892, 2020. https://doi.org/10.1021/acsphotonics.0c00823.Search in Google Scholar
[62] A. M. Summers, “Strong-field interactions in atoms and nanosystems: advances in fundamental science and technological capabilities of ultrafast sources,” Ph.D. thesis, Kansas State University, 2019.Search in Google Scholar
[63] C. S. Kumarasinghe, M. Premaratne, Q. Bao, and G. P. Agrawal, “Theoretical analysis of hot electron dynamics in nanorods,” Sci. Rep., vol. 5, no. 12, p. 140, 2015. https://doi.org/10.1038/srep12140.Search in Google Scholar PubMed PubMed Central
[64] P. Narang, R. Sundararaman, and H. A. Atwater, “Plasmonic hot carrier dynamics in solid-state and chemical systems for energy conversion,” Nanophotonics, vol. 5, no. 1, pp. 96–111, 2016. https://doi.org/10.1515/nanoph-2016-0007.Search in Google Scholar
[65] J. D. Jackson, Classical Electrodynamics, 3rd ed., New York, Wiley, 1999.10.1119/1.19136Search in Google Scholar
[66] G. Mie, “Beiträge zur Optik trüber Medien, speziell kolloidaler Metallösungen,” Ann. Phys., vol. 330, no. 3, pp. 377–445, 1908. https://doi.org/10.1002/andp.19083300302.Search in Google Scholar
[67] H. Kuwata, H. Tamaru, K. Esumi, and K. Miyano, “Resonant light scattering from metal nanoparticles: practical analysis beyond the Rayleigh approximation,” Appl. Phys. Lett., vol. 83, no. 22, pp. 4625–4627, 2003. https://doi.org/10.1063/1.1630351.Search in Google Scholar
[68] E. Saydanzad, J. Li, and U. Thumm, “Characterization of induced nanoplasmonic fields in time-resolved photoemission: a classical trajectory approach applied to gold nanospheres,” Phys. Rev. A, vol. 95, p. 053406, 2017. https://doi.org/10.1103/physreva.95.053406.Search in Google Scholar
[69] W. M. Haynes, CRC Handbook of Chemistry and Physics, Boca Raton, CRC Press, 2014.10.1201/b17118Search in Google Scholar
[70] R. H. Fowler and L. Nordheim, “Electron emission in intense electric fields,” Proc. R. Soc. London, A, vol. 119, no. 781, pp. 173–181, 1928.10.1098/rspa.1928.0091Search in Google Scholar
[71] E. L. Murphy and R. GoodJr, “Thermionic emission, field emission, and the transition region,” Phys. Rev., vol. 102, no. 6, p. 1464, 1956. https://doi.org/10.1103/physrev.102.1464.Search in Google Scholar
Supplementary Material
This article contains supplementary material (https://doi.org/10.1515/nanoph-2023-0120).
© 2023 the author(s), published by De Gruyter, Berlin/Boston
This work is licensed under the Creative Commons Attribution 4.0 International License.
Articles in the same Issue
- Frontmatter
- Research Articles
- Reaction nanoscopy of ion emission from sub-wavelength propanediol droplets
- Polariton hybridization phenomena on near-field radiative heat transfer in periodic graphene/α-MoO3 cells
- Experimental observations of communication in blackout, topological waveguiding and Dirac zero-index property in plasma sheath
- Strongly coupled Raman scattering enhancement revealed by scattering-type scanning near-field optical microscopy
- Terahertz nanospectroscopy of plasmon polaritons for the evaluation of doping in quantum devices
- Spin–orbit interactions in plasmonic crystals probed by site-selective cathodoluminescence spectroscopy
- Implementing of infrared camouflage with thermal management based on inverse design and hierarchical metamaterial
- Stable single-mode operation of distributed feedback quantum cascade laser under high current via a grating reflector
- Wide-angle deep ultraviolet antireflective multilayers via discrete-to-continuous optimization
- Active 3D positioning and imaging modulated by single fringe projection with compact metasurface device
- Enhanced cutoff energies for direct and rescattered strong-field photoelectron emission of plasmonic nanoparticles
Articles in the same Issue
- Frontmatter
- Research Articles
- Reaction nanoscopy of ion emission from sub-wavelength propanediol droplets
- Polariton hybridization phenomena on near-field radiative heat transfer in periodic graphene/α-MoO3 cells
- Experimental observations of communication in blackout, topological waveguiding and Dirac zero-index property in plasma sheath
- Strongly coupled Raman scattering enhancement revealed by scattering-type scanning near-field optical microscopy
- Terahertz nanospectroscopy of plasmon polaritons for the evaluation of doping in quantum devices
- Spin–orbit interactions in plasmonic crystals probed by site-selective cathodoluminescence spectroscopy
- Implementing of infrared camouflage with thermal management based on inverse design and hierarchical metamaterial
- Stable single-mode operation of distributed feedback quantum cascade laser under high current via a grating reflector
- Wide-angle deep ultraviolet antireflective multilayers via discrete-to-continuous optimization
- Active 3D positioning and imaging modulated by single fringe projection with compact metasurface device
- Enhanced cutoff energies for direct and rescattered strong-field photoelectron emission of plasmonic nanoparticles