Startseite Enhanced inverse Faraday effect and time-dependent thermo-transmission in gold nanodisks
Artikel Open Access

Enhanced inverse Faraday effect and time-dependent thermo-transmission in gold nanodisks

  • Alma K. González-Alcalde ORCID logo , Xinping Shi ORCID logo , Victor H. Ortiz ORCID logo , Ji Feng , Richard B. Wilson ORCID logo EMAIL logo und Luat T. Vuong ORCID logo EMAIL logo
Veröffentlicht/Copyright: 5. Februar 2024
Veröffentlichen auch Sie bei De Gruyter Brill

Abstract

Nonmagnetic media can be magnetized by light via processes referred to as an inverse Faraday effect (IFE). With nonmagnetic metal nanostructures, the IFE is dominated by the presence of light-induced solenoidal surface currents or plasmons with orbital angular momenta, whose properties depend on both the light and nanostructure geometry. Here, through a systematic study of gold nanodisks with different sizes, we demonstrate order-of-magnitude enhancement of the IFE compared to a bare gold film. Large IFE signals occur when light excites the dipolar plasmonic resonance of the gold nanodisk. We observe that the spectral response of the IFE signal mirrors the spectral response of time-dependent thermo-transmission signals. Our careful quantitative experimental measurements and analysis offer insight into the magnitude of IFE in plasmonic structures for compact, low-power, magneto-optic applications.

1 Introduction

Circularly polarized light can induce a nonequilibrium magnetization in a nonmagnetic material via the inverse Faraday effect (IFE). The earliest theoretical descriptions of the inverse Faraday effect were proposed for nonabsorbing media [1], [2], [3], [4]. The IFE magnetization generally refers to the induction of charge orbital motion by incident circularly polarized light [4], [5] or the excitation of electrons into electronic states with nonzero spin [6], [7], [8]. However, several theoretical studies predict that linearly polarized light [9], [10] and optical vortex and vector beams can also induce magnetic IFE moments [11], [12]. Our understanding of the IFE is currently being developed.

Prior theoretical and experimental studies have documented the phenomenology of the IFE in simple metals [3], [13], [14], [15], [16], [17]. The magnetic moment produced by circularly polarized light is proportional to the square of the electric-field amplitude [3], i.e., proportional to the light intensity. The magnetic moment can be monitored through time-resolved measurements of the off-diagonal terms of the optical susceptibility tensor of the metal [14], [17], [18]. The coherent motion of charges and the associated changes in the optical susceptibility persist for as long as the metal is excited with circularly polarized light [14]; once excitation with circularly polarized light stops, electron–electron and electron–phonon scattering randomize the angular momentum of the moving charge on a 10-fs time scale [14].

Knowledge of the IFE, particularly in thin films, is relevant to future computing, integrated photonic chips, and hybrid photonics–electronics hardware for applications such as ultrafast optical switching [19], magneto-optical data writing [20], [21], [22], [23], ultrafast microscopy [24], and spintronics [25], [26]. There is significant interest in enhancing the interaction strength between light and magnetic degrees of freedom in materials: ultrafast magneto-optic phenomena, like all-optical switching, requires electric field strengths on the order of 108 V/m [27], which are nontrivial.

For thin-film applications, plasmonic nanostructures represent one of the most popular platforms for controlling, tuning, and enhancing light–matter interactions; excitation close to the plasmonic resonance may produce orders-of-magnitude enhancements in the local field [28], [29]. Several theoretical studies of IFE have predicted significant plasmonic enhancement of the magnetization induced by circularly polarized light [10], [16], [30], [31], [32], [33], [34]. The presence of the IFE can be inferred and observed indirectly via Lorentz-force mechanical effects, electrical potentials, or changes in optical spectra [5], [35], [36], [37], [38], [39], [40], [41]. However, there are only a few time-resolved measurements of the plasmonic IFE [42], [43]. Recent studies scrutinize the role of Lorentz forces in plasmon damping or IFE reversal [44], [45]. Since there are only a few quantitative experimental measurements of the plasmonic IFE, it remains unclear how and to what extent the IFE is enhanced in plasmonic structures, and what effect different types of plasmonic particle geometries and sizes have on the enhancement.

Here, we experimentally study the IFE and measure the Faraday rotation of gold nanodisks. To quantify the plasmonic enhancement, we compare the IFE signals from nanodisks to the IFE signals from a 20-nm thick gold film [Figure 1]. Additionally, we perform time domain thermo-transmission measurements of the nanodisks. The nanodisks that we study have different aspect ratios and, therefore, different longitudinal plasmonic resonance energies [46]. Most of the samples that we study are 190-nm diameter nanodisks [Figure 2(a)] with varied aspect ratios h/D between 0.13 and 0.21: this increase in aspect ratio leads to a blueshift in the plasmon resonance. Additionally, these nanodisks are considered to be large nanoparticles, where the scattering effects dominate over the absorption. We also characterize the IFE of nanodisks with diameters of 130 and 145 nm. When excited at the peak thermo-transmission wavelengths of the plasmonic resonance, we observe an order-of-magnitude enhancement in the IFE signals from nanodisks in comparison to a gold (Au) film. Finally, we observe that the spectral dependence of the IFE measurements mirrors the spectral dependence of time-domain thermo-transmission measurements.

Figure 1: 
Inverse Faraday rotation. (a) Illustration of the light-induced rotation from the inverse Faraday effect in a nonmagnetic plasmonic monolayer of gold nanodisks and 20-nm thick gold film. (b) Measured rotation Δθ = (θ

RCP
 − θ

LCP
)/2, where RCP and LCP are subscripts for right and left circular polarization. Inset: autocorrelation measurement of the laser pulse duration.
Figure 1:

Inverse Faraday rotation. (a) Illustration of the light-induced rotation from the inverse Faraday effect in a nonmagnetic plasmonic monolayer of gold nanodisks and 20-nm thick gold film. (b) Measured rotation Δθ = (θ RCP θ LCP )/2, where RCP and LCP are subscripts for right and left circular polarization. Inset: autocorrelation measurement of the laser pulse duration.

Figure 2: 
Time-resolved pump/probe experiments. (a) Scanning electron microscopy image of a Au nanodisk sample. The nanodisks in the image have 190-nm diameter and 24-nm height. (b) Experimental setup for the pump/probe experiments. The pump and probe wavelengths are 785 and 778 nm. The pump beam polarization is set to RCP or LCP by a combination of a linear polarizer (LP) and achromatic quarter-waveplate (λ/4) before a nonpolarizing beamsplitter (NBS). A 20X objective focuses both pump and probe on the sample. A half-waveplate (λ/2) before a Wollaston prism (WP) is rotated to balance two linear orthogonal polarization states with equal intensities at the detector. (c) Sample time-domain thermo-transmission (TDTT) measurement. The TDTT signal is shown on a linear scale from pump–probe delay times t = 0–10 ps and on a logarithmic scale from t = 10–300 ps. (d) Sample time-domain IFE (TD-IFE) measurements with RCP (blue line) and LCP (red line).
Figure 2:

Time-resolved pump/probe experiments. (a) Scanning electron microscopy image of a Au nanodisk sample. The nanodisks in the image have 190-nm diameter and 24-nm height. (b) Experimental setup for the pump/probe experiments. The pump and probe wavelengths are 785 and 778 nm. The pump beam polarization is set to RCP or LCP by a combination of a linear polarizer (LP) and achromatic quarter-waveplate (λ/4) before a nonpolarizing beamsplitter (NBS). A 20X objective focuses both pump and probe on the sample. A half-waveplate (λ/2) before a Wollaston prism (WP) is rotated to balance two linear orthogonal polarization states with equal intensities at the detector. (c) Sample time-domain thermo-transmission (TDTT) measurement. The TDTT signal is shown on a linear scale from pump–probe delay times t = 0–10 ps and on a logarithmic scale from t = 10–300 ps. (d) Sample time-domain IFE (TD-IFE) measurements with RCP (blue line) and LCP (red line).

2 Time-resolved pump/probe measurements

The experimental setup for time-domain inverse Faraday effect (TD-IFE) and time-domain thermo-transmission experiments utilize an 80-MHz Ti:Sapphire oscillator with 430-fs pulse durations and tunable wavelengths from 690 nm to 1050 nm (Mai-Tai) [Figure 2(b)]. We estimate the pulse duration from the autocorrelation measurement [inset of Figure 1(b)] by assuming both the pump and probe pulses have an intensity versus time described by the function I = s e c h 2 ln 1 + 2 t / τ , where τ is the duration. We perform two types of pump/probe experiments [Figure 2(c) and (d)]: time-domain thermo-transmission (TDTT) measurements that characterize the change in the sample transmission due to temperature or strain and time-domain IFE measurements that identify nonequilibrium changes of the transmitted probe-beam polarization caused by a right or left-circularly polarized pump-beam (RCP or LCP).

To prevent the pump beam from reaching the detector, the pump and probe beams are spectrally shifted from the Ti:Sapphire oscillator’s center wavelength of 783 nm. The pump passes through a long pass filter, and the probe beam passes through a short pass filter. After passing through the filters, the center wavelength of the beams is measured with a spectrometer (Ocean Optics). The pump and probe beams have wavelengths of 785 nm and 778 nm, respectively. A short-pass optical filter in front of the detector blocks the red-shifted pump beam.

For TDTT measurements, the experimental setup is the same as for TD-IFE measurements, but one of the inputs of the balanced detector is blocked. With this change, the signal becomes dominated by the change in the transmitted intensity, instead of the change in the polarization. The pump and probe average powers before the objective lens are kept constant in all of the experiments and set at 1.5 mW and 1.0 mW, respectively. The 20X objective lens has a transmittance of 0.7. The laser spot-size when focused on the sample is 3.9 μm. More details about the experimental setup can be found in Refs. [47], [48].

In Figure 2, we show a schematic of the experimental setup. In Figure 2(c), we show TDTT data taken from −2 ps to 8000 ps. The measurement allows us to monitor changes in transmission of the probe beam due to temperature and acoustic vibrations. The TDTR data from 10 ps to 300 ps identifies acoustic vibrations on the sample and is also used to determine the nanodisks’ dimensions in the region where the laser is focused. The physics of the nanodisk acoustic vibrations are described in detail in Ref. [49].

In Figure 2(d), we present TD-IFE data taken from −1.5 ps to 1.5 ps. To determine the circularly polarized light-induced rotation (Δθ), we perform measurements for RCP (blue solid line) and LCP (red solid line). The circularly polarized light-induced rotation is Δθ = (θ RCP θ LCP )/2. For the experimental setup shown in Figure 2, the TD-IFE signal is a measure of the real part of the Faraday angle, i.e., polarization rotation. The curves have a hyperbolic secant shape that match the autocorrelation measurement of the pump and probe pulse duration. The temporal resolution of our experiments is not sufficient to resolve the lag between when the pump beam turns on and the Faraday rotation starts, which is expected to take 10 fs [16]. Similarly, we cannot resolve the lag between when the pump beam turns off, and the Faraday rotation ceases, which is expected to occur on a 10 fs time-scale due to momentum relaxation of excited electrons [14]. The sign of the polarization rotation is opposite for LCP versus RCP pump excitation. There is an asymmetry in the LCP versus RCP signals [Figure 2(d)] caused by a background TDTT signal. The background is also the cause of a small peak time-domain IFE signals in Figure 2(d) near 0.3 ps. The background is independent of polarization and so does not affect Δθ. The Supplementary Information provides additional discussion of the background.

3 Sample synthesis and characterization

To control the frequency of the localized surface plasmon resonance (LSPR), we synthesize nanodisks with varied aspect ratio (AR = h/D) [50], [51]. For disks with a low aspect ratio, the LSPR frequency in the quasistatic limit is

(1) ω = π h 4 D ω p 2 γ 2

where ω p is the plasma frequency and γ is the damping term. Smaller-AR nanodisks have resonances at lower frequencies, while larger-AR nanodisks are associated with resonances at higher frequencies.

The Au nanodisks were fabricated on a sapphire substrate by using a hole-mask colloidal lithography technique [52]. Sapphire substrates were cleaned by sonicating in acetone, ethanol, and water for 10 min each and then blow-dried. A sacrificial layer of poly(methyl methacrylate) (PMMA, MicroChem 495 A5) was spin-coated onto the precleaned sapphire substrates at 4000 rpm for 60 s and baked at 170 °C for 10 min to evaporate the solvent. The surface was made hydrophilic by a short period of oxygen plasma treatment (EMS 1050X Plasma Asher) at 100 W for 20 s. The coated substrate was immersed in a water solution of poly(diallyldimethylammonium chloride) (PDDA, Sigma Aldrich, 0.2 wt%) for surface modification with positive charges, rinsed with water thoroughly, and then blow-dried. The modified substrate was immersed in a water suspension of well-dispersed polystyrene (PS) beads (0.05 mg/mL, Bangs laboratories, PC02008) for 4 min to load the nanoparticles, rinsed with water thoroughly, and blow-dried. After that, 5 nm of Ti and 30 nm of Au (e-beam evaporation, Temescal BJD 1800) was deposited on the substrate as an etch mask, followed by sonicating in water for 30 s to remove the PS beads. The exposed PMMA layer was fully removed by reactive ion etching (STS Reactive Ion Etcher, 200 mTorr, 100 W, O2, 5 min). The Au nanodisks (with 5-nm Ti adhesion layer) were deposited on the sapphire substrate by E-beam evaporation, followed by a lift-off step (sonicate in acetone for 2 min). The final products were obtained after depositing a 3-nm thick layer of Al2O3 and annealed at 400 °C for 15 min.

The nanodisks were synthesized out of Au for several reasons. The small intrinsic absorption of Au, weak electron–phonon coupling [53], and lack of chemical reactivity make Au the plasmonic material of choice for many applications [54], [55], [56], [57]. Interband transitions between d-bands and the Fermi-level can have a dramatic effect on both the magneto-optical Kerr effect [48] and nonmagnetic Kerr effect [58]. However, such interband transitions in Au require photon energies above 2 eV [59] and so do not affect our experiments on disks in the near-infrared. (Such interband transitions would matter if we were studying spheres because they have plasmonic resonances at higher energies than disks). We would expect similar behaviors from plasmonic disks made from free-electron-like metals without interband transitions in the near-infrared, e.g., Ag or Cu.

We characterize the samples with scanning electron microscopy (SEM) and atomic force microscopy (AFM). We also use TDTT measurements to estimate the diameters of the nanodisk through the acoustic modes. Estimates of the nanodisk diameter are comparable given both acoustic frequency and microscopy images. We estimate that the nanodisk samples have diameters of D = {190, 190, 200, 190, 200} nm and heights of h = {24, 27, 30, 33, 41} nm. We estimate the nanodisks aspect ratios (AR = h/D) AR = {0.13, 0.14, 0.15, 0.17, 0.21}. The samples have an average of {1.6, 1.2, 1.3, 0.9, 1.5} nanodisks per μm2. The role of sample nanodisk density is discussed in Section 5.

We perform both static transmission and TDTT spectral characterizations of the 200-nm diameter nanodisks. We measure the static transmission spectra with direct illumination and detection. The static transmission spectra are normalized by the spectra of the lamp. The white light emission is unpolarized and the beam spot is 1 mm in diameter. The results are presented in Figure 3(a).

Figure 3: 
Spectral characterization of the samples. (a) Static transmission spectra. (b) Time-domain thermo-transmission spectra collected at a pump–probe delay time of 10 ps. (c) Sample resonance wavelength with static transmission (crosses) and TDTT measurements (asterisks).
Figure 3:

Spectral characterization of the samples. (a) Static transmission spectra. (b) Time-domain thermo-transmission spectra collected at a pump–probe delay time of 10 ps. (c) Sample resonance wavelength with static transmission (crosses) and TDTT measurements (asterisks).

In addition to the ≈200-nm nanodisk samples described above, we prepared an additional set of samples with target diameters of 130 and 145-nm and aspect ratios that vary between 0.08 and 0.21. The samples were synthesized in a similar way as described above, but with the following changes. We used PMMA A4 as the sacrificial layer (Kayalu 495 A4). After spin-coating, the substrate was immersed in a water suspension of well-dispersed polystyrene (PS) beads (0.05 mg/mL, Bangs laboratories, PC02005, PC 02006). Due to the usage of different PMMA, the thickness of the PMMA layer changed, which required a change in the reactive ion etching time. In this new set of samples, the exposed PMMA layer was removed by reactive ion etching for 3 min. We performed static spectral characterizations, AFM, and SEM on all samples. We only performed pump/probe measurements on two of the samples: the 130-nm diameter disks with AR = 0.12 and the 145-nm diameter disks with AR = 0.14. These two samples had a transmission minima at the 785-nm wavelength of our laser. SEM images of the samples identified the density of 130-nm and 145-nm diameter nanodisks as 0.7 and 0.6 disks per μm2, respectively.

To collect the TDTT spectra, TDTT measurements versus wavelength are performed at a fixed time-delay of 10 ps. To perform the measurements, we modify the optical setup shown in Figure 2. The main changes regarding the optical setup are the following: (1) pump and probe beam are not collinear (there is a small angle between the pump and probe), (2) the pump beam is made vertically polarized by removing the λ/4 plate in Figure 2(b), (3) all optical filters that blue-shift the probe beam and red-shift the pump beam are removed, (4) an aperture is added to the detection line to spatially filter the pump beam, and (5) a linear polarizer (LP) is added to the detection line to filter the vertically polarized pump beam. The results of the TDTT spectra are presented in Figure 3(b).

As expected from Eq. (1), there is a blue shift in both static transmission peaks and TDTT spectra peaks as the nanodisk AR increases. For the static transmission measurements, the 785-nm wavelength pump is close to transmission extrema for the nanoparticles with an AR of 0.14 (red line) and 0.15 (green line) [Figure 3(a)]. Alternatively, the TDTT spectra extrema is closest to the pump wavelength for the samples with an AR of 0.14 (red line) and 0.13 (blue line) [Figure 3(b)]. In other words, the TDTT spectra are blue-shifted in comparison with respect to the static transmission spectra [Figure 3(c)]. In our discussion below, we label the TDTT LSPR as the extrema in the TDTT spectra. That the TDTT LSPR is blue-shifted from the static transmission spectra LSPR can be explained in terms of the time-dependent, temperature-induced change of the gold nanodisk dielectric constant by the pump beam [60], [61]. The TDTT spectra are a measure of the difference in the transmission spectra at an elevated versus ambient temperature. The sample transient transmission spectra are connected to the change in the dielectric function by

(2) Δ Q ext ( λ , t ) = Q ext ε r d ε r d t + Q ext ε i d ε i d t ,

where Q ext is the extinction cross section, λ is the wavelength, t is the temperature, and ɛ r and ɛ i are real and imaginary parts of the dielectric function. According to Eq. (2), the spectral position of the dipolar resonance is temperature dependent. The TDTT peak blue shifts upon heating and can gain a positive side-wing [Figure 3(b) and (c)]. The TDTT spectra are narrower than the static transmission spectra. This is due to the fact that the transient transmission spectra carry terms associated with the derivative of the extinction cross section [Eq. (2)].

4 Results

Samples with TDTT LSPR near the laser wavelength have order-of-magnitude enhancements in the TD-IFE and TDTT signals [Figure 4]. The TDTT (ΔT/T) data in Figure 4(a) are collected at a pump/probe time delay of 10 ps. The IFE rotation (Δθ) in Figure 4(b) is the maximum Faraday rotation, which we observe at 0 ps delay time. For both Figure 4(a) and (b), we plot the data as a function of TDTT LSPR position [Figure 3(b) and (c)]. Each marker in Figure 4(a) and (b) represents a measurement at a different location on the sample.

Figure 4: 
Time-domain pump/probe measurements of the samples. (a) Time-domain thermo-transmission (TDTT) measurements at 10 ps. (b) Inverse Faraday rotation measurements at 0 ps. The measurements are for Au nanodisk monolayers with different aspect ratios and a 20-nm bare gold film used as a reference. The wavelength of the pump beam is 785 nm. Each marker is a measurement in a different sample location. (c) Average of the inverse Faraday rotation measurements at 0 ps and TDTT measurements at 10 ps.
Figure 4:

Time-domain pump/probe measurements of the samples. (a) Time-domain thermo-transmission (TDTT) measurements at 10 ps. (b) Inverse Faraday rotation measurements at 0 ps. The measurements are for Au nanodisk monolayers with different aspect ratios and a 20-nm bare gold film used as a reference. The wavelength of the pump beam is 785 nm. Each marker is a measurement in a different sample location. (c) Average of the inverse Faraday rotation measurements at 0 ps and TDTT measurements at 10 ps.

For samples with 0.13 ≤ AR 0.15 , the ΔT/T is much larger compared with a 20-nm thick gold film (purple line). Samples with AR of 0.13 (blue marks) and 0.14 (red marks) have the largest (ΔT/T) values [4(a)]. This is because for both samples, the pump laser is close to the TDTT LSPR. On the other hand, we note a decrease in ΔT/T as the AR increases and indeed a change of sign in ΔT/T for AR 0.17 .

Samples with lower AR produce larger inverse Faraday rotations [Figure 4(b)]. The IFE enhancement is 17 × (AR = 0.13), 16 × (AR = 0.14), 12.5 × (AR = 0.15), 1.8 × (AR = 0.17), and 1.3 × (AR = 0.21) larger than the IFE rotation from the 20-nm thick Au film. As expected, the IFE enhancement is largest for nanodisks whose TDTT LSPR is closest to the laser wavelength of 783 nm.

We observe a correlation between the TDTT and IFE measurements. This is shown in Figure 4(c) where we plot the average of all measurements reported in Figure 4(a) and (b). We observe that higher ΔT/T signals are associated with higher rotations Δθ.

To further explore plasmonic enhancement of the IFE, we perform TD-IFE measurements of 130-nm and 145-nm diameter nanodisks with AR of 0.12 and 0.14. Both of the 130 and 145 nm-diameter nanodisk samples have a minima in the static transmission spectra [Figure 5(a)] near 800 nm and a TDTT LSPR wavelength [Figure 5(b)] close to the pump laser (773-nm and 775-nm peak values for 130- and 145-nm diameter nanodisks, respectively). The 130 and 145-nm diameter samples have areal densities that are one-sixth and one-fifth of the 190-nm disks. The IFE rotation measurements relative to the Au thin-film are summarized in Figure 5(c). We include the average IFE enhancement relative to the Au thin-film measured for the 0.13-AR and 0.14-AR, 190-nm diameter disks. The IFE enhancement with the smaller nanodisks is comparable to that of the 190-nm diameter disks even though these lower-density samples have half the extinction of the 190-nm diameter disk samples. We plot the nanodisk-to-thin-film rotation accounting for the areal density δ 1 2 in Figure 5(d). The calculation of δ 1 is described in detail and discussed in Section 5.

Figure 5: 
Inverse Faraday rotation measurements for nanodisks with different diameters and aspect ratios. (a) Static transmission spectra for samples with different diameters and areal densities. (b) Corresponding TDTT spectra at 10 ps: the TDTT LSPR resonances are close to the pump wavelength. (c) Measured Faraday rotation for these samples relative to Au thin-film. (d) Enhancement of the measured rotation δ
1, which accounts for differences in the optical coverage area, see Section 5.
Figure 5:

Inverse Faraday rotation measurements for nanodisks with different diameters and aspect ratios. (a) Static transmission spectra for samples with different diameters and areal densities. (b) Corresponding TDTT spectra at 10 ps: the TDTT LSPR resonances are close to the pump wavelength. (c) Measured Faraday rotation for these samples relative to Au thin-film. (d) Enhancement of the measured rotation δ 1, which accounts for differences in the optical coverage area, see Section 5.

5 Discussion

5.1 Comparison with prior measurements

In response to ≈0.5 TW/m2 of CP excitation, the 190-nm diameter disks with 24-nm thicknesses display approximately 12-μrad of Faraday rotation or ≈24 μrad rotation per TW/m2. Despite significant differences in experimental geometry, this is in reasonable agreement with the results of Cheng et al. [42], who performed pump/probe IFE measurements on 2-mm thick colloidal suspensions of 100-nm diameter Au nanoparticles and reported a Faraday rotation of ≈30 μrad per TW/m2.

The light-induced nonequilibrium rotation on Au nanodisks in our study is significant; however, it is small when compared to materials that are magnetic at equilibrium. Ferromagnetic metals have Kerr rotations on the order of mrad or 100× greater than we observe from the nanodisks. The μrad-scale polarization rotation in the present experiments is similar in magnitude to the Kerr rotations caused by a few A/m of spin accumulation in a Au thin-film [48].

5.2 Magnetization versus Faraday rotation

To understand the magnitude of the magnetic moment induced in the disks, we now frame the relationship between Faraday rotation and magnetism. Consider K as a material-specific constant where the rotation is θ = KM ind . The range of reported values for K in bulk materials is between 4 and 40 nrad per A/m [48], [62]. In our experiments, we expect plasmonic enhancement of the induced magnetization: M disk = δ 1 M Au , where M disk is the magnetization induced in the Au nanodisks and M Au is the magnetization induced in the Au thin-film, and δ 1 > 1 represents an Au geometric nanostructure plasmonic enhancement factor. The induced magnetization M ind is expected to be proportional to the square of the electric field, and this local field is enhanced by plasmonic nanostructures. We are primarily interested in δ 1, i.e., how much the plasmonic resonance increases the CP light–induced magnetization.

Now, we also expect a plasmonic enhancement of K that occurs when we probe the nanodisk samples: K disk = δ 2 K Au . Based on numerous studies or Kerr and Faraday rotations in plasmonic structures such as [63], we also expect δ 2 > 1. To crudely estimate the order of magnitude of δ 1 from our experiments, we assume δ 1 = δ 2, which is reasonable since both factors describe plasmonic enhancements of the disk’s magneto-optical properties. (As shown below, this assumption yields the conclusion that δ 2 is on the order of 10, which is in good agreement with prior studies of the Faraday effect in similar disks [40]). Then, we expect the enhancement factor in our TD-IFE experiments to be θ disk = δ 1 2 θ A u . This means (1) the rotation measurements from the plasmonically enhanced magnetization are also plasmonically enhanced and (2) induced magnetizations that are smaller or larger than what we infer could be possible.

5.3 Effect of areal coverage

Due to the sparse nanodisk coverage in our samples, most of the probe light incident on our samples is transmitted without interacting with the Au nanodisks. Our 190-nm diameter disks with AR = 0.13–0.14 have a surface coverage of 4–5 percent, respectively. Meanwhile, our 130- and 145-nm diameter disks in Figure 5 have surface coverage of only of ≈1 percent. On the other hand, Au thin films have 100-percent surface coverage.

The lower areal coverage means that the measured Faraday rotations underestimate the actual CP light–induced magnetization. The measured Faraday rotation should be scaled by a numerical factor 1/C related to the coverage area, where C < 1. We can estimate C based on the optical cross section of the disks. The measured attenuation of the nanodisks at the laser’s wavelength (Figure 3(a)) is 4 × larger than the areal coverage of the nanodisks. Therefore, we expect the optical cross section of the nanodisks to be 4 × the physical coverage area. (This is in reasonable agreement with reports in the literature for nanodisks with diameters between 100 and 200 nm [64], [65]). We estimate the optical coverage of the AR = 0.13 and 0.14 disks with 200-nm diameter to be C ≈ 0.16 and 0.2. Similarly, the optical coverage for the 130 and 145-nm diameter disks will be C ≈ 0.04.

Accounting for these differences in optical coverage, the plasmon enhancement in the CP light–induced magnetism is δ 1 2 Δ θ disk / Δ θ A u / C . Therefore, δ 1 2 60 and δ 1 2 50 in the 0.13-AR and 0.14-AR, 200-nm diameter nanodisks, respectively. In the 130 and 145-nm diameter disks, δ 1 2 240 and δ 1 2 210 . [The values of Δθ disk θ Au are taken from Figure 5(c)]. So, a reasonable estimate for the IFE enhancement in our experiments is δ 1 ≈ 7 for 190-nm diameter disks, and δ 1 ≈ 15 and δ 1 ≈ 15 for the 130- and 145-nm disks.

In summary, after accounting for differences in optical coverage (C) and Faraday rotation per moment (K disk ), our best estimate from our experimental data is that the magnetization induced by CP light is approximately one order of magnitude larger in Au nanodisks than a 20-nm thick Au film. Having established this experimental evidence, we now turn our attention to comparison with theoretical predictions for IFE in bulk Au and nanoparticles excited at resonance with CP light.

5.4 Comparison to theory

The light-induced magnetization can be modeled as an interaction between the illuminating light and the light-induced charge-density oscillations [3]. Hertel predicts that for uniform light illumination of a solid-state bulk material where the charge conductivity follows a simple Drude model [3], the induced magnetization is

(3) M = e ϵ 0 ω p 2 4 m ω 3 | E | 2 ,

where E is the electric-field amplitude and ω is the frequency of the circularly polarized light, ω p is the plasmon frequency of Au, e is the charge of an electron, m is the free electron mass, and ϵ 0 is the permittivity of free space. For Au, Hertel’s model predicts a magnetic moment of ≈8 A/m per TW/m2, or 1.5 × 1 0 5 μ B per atom per TW/m2. (Hertel’s model neglects the effect of spin-moments induced by interband transitions. However, Mishra and Coh predict that the spin contribution to the IFE due to interband absorption ranges between only 10−7 and 10−6 μ B per atom per TW/m2 for visible light [8]). To our knowledge, quantitative measurements of the magnetization induced by circularly polarized light are not available for Au. However, the result from Eq. (3) is similar in magnitude to what is observed for simple ferromagnetic metals. Choi et al. report that excitation with 10 TW/m2 of circularly polarized light induces a 50–60 A/m magnetization in 10-nm thick Co, Fe, and Ni films capped with 2-nm layers of Au [22].

The framework described above can also be used to predict the IFE in nanoparticles driven at a plasmonic resonance that enhances local electric fields. The effect of the nanoparticle resonance in Eq. (3) is to replace the factor of 1/ω 3 in Hertel’s expression with a factor of 1/γ 2 ω [16],

(4) M = e ϵ 0 ω p 2 4 m ω γ 2 | E | 2 ,

where γ is an effective plasmon damping rate from a Drude model [16]. Eq. (4) predicts that the IFE is enhanced in a nanoparticle by the square of its quality factor, (ω/γ)2 [16], [32]. For our Au nanodisk samples, the width of the absorption peak in the static transmission spectra ranges between 100 and 200 nm [Figures 3(a) and 5(a)], corresponding to a quality factor <10.

Quantum mechanical based theories, which model dynamics without making a Drude-like free-electron assumption, predict somewhat different magnitude for the IFE than the classical models described above [16], [32]. Hurst et al. predict that a 1-nm radius sphere with 510 TW/m2-intensity circularly polarized light induces a total magnetic moment of 10 μ B or 8 × 1 0 5 μ B per atom per TW/m2 [32]. Sinha-Roy et al. report that a 1-nm radius Au sphere with CP light of 13 TW/m2-intensity induces a total magnetic moment of 2 μ B or 6 × 1 0 4 μ B per atom per TW/m2 [16]. We note that these predictions of quantum mechanical models may not be directly comparable to our values for δ 1, since they model a much smaller volume of atoms.

The large δ 1 that we measure in disks with small diameter is notable. In a parallel work, we theoretically show that the Faraday rotation is largely enhanced at the nanodisk LSPR when the extinction efficiency is dominated by absorption. For a single nanodisk, the effect of light scattering is to dampen the free-electrons oscillations, increase the linewidth, and significantly reduce the rotation efficiency [66]. We expect that similar effects may lead to a larger IFE in disks with smaller diameter and flatter nanodisks than wider, taller nanodisks with the same LSPR wavelength [Figure 5(d)].

6 Conclusions

In summary, we perform a systematic experimental study of the IFE in monolayers composed of gold nanodisks with different aspect ratios. Through pump–probe experiments, we measure that the inverse Faraday rotation is enhanced if the pump wavelength is close to the peak TDTT LSPR, which is blue-shifted from the static transmission measurement of the LSPR. We measure Faraday rotations approximately 10× larger compared to 20-nm thick bare gold film. The Faraday rotation per intensity with monolayers of nanodisks with <50-nm thicknesses are comparable to that reported in Ref. [42] with 2-mm thick colloidal suspensions. When further accounting for the measurement and optical cross section, the Faraday rotation per intensity per area is estimated to be as much as 240× larger than those observed from a gold film.


Corresponding authors: Richard B. Wilson and Luat T. Vuong, University of California at Riverside, Riverside, CA, USA, E-mail: (R. B. Wilson), (L. T. Vuong) (R. B. Wilson) (L. T. Vuong)

Funding source: UC MEXUS

Award Identifier / Grant number: CN-20-149

Funding source: Army Research Office

Award Identifier / Grant number: W911NF-20-1-0274

Award Identifier / Grant number: 1921034

  1. Research funding: This work was primarily supported by the U.S. Army Research Laboratory and the U.S. Army Research Office under Contract/Grant No. W911NF-20-1-0247. LTV also acknowledges support from the National Science Foundation Grant No. 1921034 and UC Mexus Grant CN-20-149.

  2. Author contributions: JF and XS fabricated nanodisk samples. VO annealed and encapsulated the samples and provided AFM characterization. AKGA performed measurements and wrote the first draft of the manuscript. Experiments were performed in the RBW lab. LTV and RBW oversaw all aspects of the research and revised the manuscript with AKGA. All authors have accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Conflict of interest: Authors state no conflicts of interest.

  4. Informed consent: Informed consent was obtained from all individuals included in this study.

  5. Ethical approval: The conducted research is not related to either human or animals use.

  6. Data availability: The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.

References

[1] J. P. van der Ziel, P. S. Pershan, and L. D. Malmstrom, “Optically-induced magnetization resulting from the inverse Faraday effect,” Phys. Rev. Lett., vol. 15, no. 5, p. 190, 1965. https://doi.org/10.1103/physrevlett.15.190.Suche in Google Scholar

[2] P. S. Pershan, J. P. van der Ziel, and L. D. Malmstrom, “Theoretical discussion of the inverse Faraday effect, Raman scattering, and related phenomena,” Phys. Rev., vol. 143, no. 2, p. 574, 1966. https://doi.org/10.1103/physrev.143.574.Suche in Google Scholar

[3] R. Hertel, “Theory of the inverse Faraday effect in metals,” J. Magn. Magn. Mater., vol. 303, no. 1, p. L1, 2006. https://doi.org/10.1016/j.jmmm.2005.10.225.Suche in Google Scholar

[4] R. Hertel and M. Fähnle, “Macroscopic drift current in the inverse Faraday effect,” Phys. Rev. B, vol. 91, no. 2, p. 020411, 2015. https://doi.org/10.1103/physrevb.91.020411.Suche in Google Scholar

[5] N. D. Singh, M. Moocarme, B. Edelstein, N. Punnoose, and L. T. Vuong, “Anomalously-large photo-induced magnetic response of metallic nanocolloids in aqueous solution using a solar simulator,” Opt. Express, vol. 20, no. 17, p. 19214, 2012. https://doi.org/10.1364/oe.20.019214.Suche in Google Scholar

[6] M. Berritta, R. Mondal, K. Carva, and P. M. Oppeneer, “Ab initio theory of coherent laser-induced magnetization in metals,” Phys. Rev. Lett., vol. 117, no. 21, p. 137203, 2016. https://doi.org/10.1103/physrevlett.117.137203.Suche in Google Scholar

[7] P. Scheid, G. Malinowski, S. Mangin, and S. Lebègue, “Ab initio theory of magnetization induced by light absorption in ferromagnets,” Phys. Rev. B, vol. 100, no. 21, p. 214402, 2019. https://doi.org/10.1103/physrevb.100.214402.Suche in Google Scholar

[8] S. B. Mishra and S. Coh, “Spin contribution to the inverse Faraday effect of nonmagnetic metals,” Phys. Rev. B, vol. 107, no. 21, p. 214432, 2023. https://doi.org/10.1103/physrevb.107.214432.Suche in Google Scholar

[9] S. Ali, J. R. Davies, and J. T. Mendonca, “Inverse Faraday effect with linearly polarized laser pulses,” Phys. Rev. Lett., vol. 105, no. 3, p. 035001, 2010. https://doi.org/10.1103/physrevlett.105.035001.Suche in Google Scholar

[10] X. Yang, Y. Mou, R. Zapata, B. Reynier, B. Gallas, and M. Mivelle, “An inverse Faraday effect generated by linearly polarized light through a plasmonic nano-antenna,” Nanophotonics, vol. 12, no. 4, p. 687, 2023. https://doi.org/10.1515/nanoph-2022-0488.Suche in Google Scholar

[11] S. Lin, et al.., “All-optical vectorial control of multistate magnetization through anisotropy-mediated spin-orbit coupling,” Nanophotonics, vol. 8, no. 12, p. 2177, 2019. https://doi.org/10.1515/nanoph-2019-0198.Suche in Google Scholar

[12] V. Karakhanyan, C. Eustache, Y. Lefier, and T. Grosjean, “Inverse Faraday effect from the orbital angular momentum of light,” Phys. Rev. B, vol. 105, no. 4, p. 045406, 2022. https://doi.org/10.1103/physrevb.105.045406.Suche in Google Scholar

[13] S. V. Popov, N. I. Zheludev, and Y. P. Svirko, “Coherent and incoherent specular inverse faraday effect: χ(3) measurements in opaque materials,” Opt. Lett., vol. 19, no. 1, p. 13, 1994. https://doi.org/10.1364/ol.19.000013.Suche in Google Scholar PubMed

[14] V. V. Kruglyak, R. J. Hicken, M. Ali, B. J. Hickey, A. T. G. Pym, and B. K. Tanner, “Ultrafast third-order optical nonlinearity of noble and transition metal thin films,” J. Opt. A: Pure Appl. Opt., vol. 7, no. 2, p. S235, 2005. https://doi.org/10.1088/1464-4258/7/2/031.Suche in Google Scholar

[15] J.-Y. Bigot, M. Vomir, and E. Beaurepaire, “Coherent ultrafast magnetism induced by femtosecond laser pulses,” Nat. Phys., vol. 5, no. 7, p. 515, 2009. https://doi.org/10.1038/nphys1285.Suche in Google Scholar

[16] R. Sinha-Roy, J. Hurst, G. Manfredi, and P.-A. Hervieux, “Driving orbital magnetism in metallic nanoparticles through circularly polarized light: a real-time TDDFT study,” ACS Photonics, vol. 7, no. 9, p. 2429, 2020. https://doi.org/10.1021/acsphotonics.0c00462.Suche in Google Scholar

[17] V. H. Ortiz, S. B. Mishra, L. Vuong, S. Coh, and R. B. Wilson, “Specular inverse Faraday effect in transition metals,” Phys. Rev. Mater., vol. 7, no. 12, p. 125202, 2023. https://doi.org/10.1103/physrevmaterials.7.125202.Suche in Google Scholar

[18] G.-M. Choi, H. G. Park, and B.-C. Min, “The orbital moment and field-like torque driven by the inverse faraday effect in metallic ferromagnets,” J. Magn. Magn. Mater., vol. 474, p. 132, 2019. https://doi.org/10.1016/j.jmmm.2018.10.117.Suche in Google Scholar

[19] A. Kirilyuk, A. V. Kimel, and T. Rasing, “Ultrafast optical manipulation of magnetic order,” Rev. Mod. Phys., vol. 82, no. 3, p. 2731, 2010. https://doi.org/10.1103/revmodphys.82.2731.Suche in Google Scholar

[20] D. Jenkins, et al.., “Advanced optical and magneto-optical recording techniques: a review,” Microsyst. Technol., vol. 10, no. 1, p. 66, 2003. https://doi.org/10.1007/s00542-003-0307-x.Suche in Google Scholar

[21] A. R. Khorsand, et al.., “Role of magnetic circular dichroism in all-optical magnetic recording,” Phys. Rev. Lett., vol. 108, no. 12, p. 127205, 2012. https://doi.org/10.1103/physrevlett.108.127205.Suche in Google Scholar

[22] G.-M. Choi, A. Schleife, and D. G. Cahill, “Optical-helicity-driven magnetization dynamics in metallic ferromagnets,” Nat. Commun., vol. 8, no. 1, p. 15085, 2017. https://doi.org/10.1038/ncomms15085.Suche in Google Scholar PubMed PubMed Central

[23] G.-M. Choi, et al.., “Optical spin-orbit torque in heavy metal-ferromagnet heterostructures,” Nat. Commun., vol. 11, no. 1, p. 1482, 2020. https://doi.org/10.1038/s41467-020-15247-3.Suche in Google Scholar PubMed PubMed Central

[24] Y. Dai, A. Ghosh, S. Yang, Z. Zhou, C.-B. Huang, and H. Petek, “Poincaré engineering of surface plasmon polaritons,” Nat. Rev. Phys., vol. 4, no. 9, p. 562, 2022. https://doi.org/10.1038/s42254-022-00492-w.Suche in Google Scholar

[25] M. Fechner, A. Sukhov, L. Chotorlishvili, C. Kenel, J. Berakdar, and N. A. Spaldin, “Magnetophononics: ultrafast spin control through the lattice,” Phys. Rev. Mater., vol. 2, no. 6, p. 064401, 2018. https://doi.org/10.1103/physrevmaterials.2.064401.Suche in Google Scholar

[26] M. L. M. Lalieu, R. Lavrijsen, and B. Koopmans, “Integrating all-optical switching with spintronics,” Nat. Commun., vol. 10, no. 1, p. 110, 2019. https://doi.org/10.1038/s41467-018-08062-4.Suche in Google Scholar PubMed PubMed Central

[27] J. Gorchon, et al.., “Role of electron and phonon temperatures in the helicity-independent all-optical switching of GdFeCo,” Phys. Rev. B, vol. 94, no. 18, p. 184406, 2016. https://doi.org/10.1103/physrevb.94.184406.Suche in Google Scholar

[28] L. Novotny and N. van Hulst, “Antennas for light,” Nat. Photonics, vol. 5, no. 2, p. 83, 2011. https://doi.org/10.1038/nphoton.2010.237.Suche in Google Scholar

[29] J. A. Schuller, E. S. Barnard, W. Cai, Y. C. Jun, J. S. White, and M. L. Brongersma, “Plasmonics for extreme light concentration and manipulation,” Nat. Mater., vol. 9, no. 3, p. 193, 2010. https://doi.org/10.1038/nmat2630.Suche in Google Scholar PubMed

[30] Y. Gu and K. G. Kornev, “Plasmon enhanced direct and inverse faraday effects in non-magnetic nanocomposites,” J. Opt. Soc. Am. B, vol. 27, no. 11, p. 2165, 2010. https://doi.org/10.1364/josab.27.002165.Suche in Google Scholar

[31] A. Nadarajah and M. T. Sheldon, “Optoelectronic phenomena in gold metal nanostructures due to the inverse Faraday effect,” Opt. Express, vol. 25, no. 11, p. 12753, 2017. https://doi.org/10.1364/oe.25.012753.Suche in Google Scholar PubMed

[32] J. Hurst, P. M. Oppeneer, G. Manfredi, and P.-A. Hervieux, “Magnetic moment generation in small gold nanoparticles via the plasmonic inverse Faraday effect,” Phys. Rev. B, vol. 98, no. 13, p. 134439, 2018. https://doi.org/10.1103/physrevb.98.134439.Suche in Google Scholar

[33] V. Karakhanyan, Y. Lefier, C. Eustache, and T. Grosjean, “Optomagnets in nonmagnetic plasmonic nanostructures,” Opt. Lett., vol. 46, no. 3, p. 613, 2021. https://doi.org/10.1364/ol.411108.Suche in Google Scholar PubMed

[34] X. Yang, Y. Mou, B. Gallas, A. Maitre, L. Coolen, and M. Mivelle, “Tesla-range femtosecond pulses of stationary magnetic field, optically generated at the nanoscale in a plasmonic antenna,” ACS Nano, vol. 16, no. 1, p. 386, 2022. https://doi.org/10.1021/acsnano.1c06922.Suche in Google Scholar PubMed

[35] K. Nakagawa, Y. Ashizawa, S. Ohnuki, A. Itoh, and A. Tsukamoto, “Confined circularly polarized light generated by nano-size aperture for high density all-optical magnetic recording,” J. Appl. Phys., vol. 109, no. 7, p. 07B735, 2011. https://doi.org/10.1063/1.3556924.Suche in Google Scholar

[36] M. Kauranen and A. V. Zayats, “Nonlinear plasmonics,” Nat. Photonics, vol. 6, no. 11, p. 737, 2012. https://doi.org/10.1038/nphoton.2012.244.Suche in Google Scholar

[37] M. Moocarme, J. L. Domínguez-Juárez, and L. T. Vuong, “Ultralow-intensity magneto-optical and mechanical effects in metal nanocolloids,” Nano Lett., vol. 14, no. 3, p. 1178, 2014. https://doi.org/10.1021/nl4039357.Suche in Google Scholar PubMed

[38] N. V. Proscia, M. Moocarme, R. Chang, I. Kretzschmar, V. M. Menon, and L. T. Vuong, “Control of photo-induced voltages in plasmonic crystals via spin-orbit interactions,” Opt. Express, vol. 24, no. 10, p. 10402, 2016. https://doi.org/10.1364/oe.24.010402.Suche in Google Scholar PubMed

[39] H. Ji, et al.., “Long-range self-assembly via the mutual Lorentz force of plasmon radiation,” Nano Lett., vol. 18, no. 4, p. 2564, 2018. https://doi.org/10.1021/acs.nanolett.8b00269.Suche in Google Scholar PubMed

[40] B. Sepúlveda, J. B. González-Díaz, A. García-Martín, L. M. Lechuga, and G. Armelles, “Plasmon-induced magneto-optical activity in nanosized gold disks,” Phys. Rev. Lett., vol. 104, no. 14, p. 147401, 2010. https://doi.org/10.1103/physrevlett.104.147401.Suche in Google Scholar

[41] F. Pineider, et al.., “Circular magnetoplasmonic modes in gold nanoparticles,” Nano Lett., vol. 13, no. 10, p. 4785, 2013. https://doi.org/10.1021/nl402394p.Suche in Google Scholar PubMed

[42] O. Cheng, D. H. Son, and M. Sheldon, “Light-induced magnetism in plasmonic gold nanoparticles,” Nat. Photonics, vol. 14, no. 6, p. 1, 2020. https://doi.org/10.1038/s41566-020-0603-3.Suche in Google Scholar

[43] S. Parchenko, et al.., “Plasmon-enhanced optical control of magnetism at the nanoscale via the inverse Faraday effect,” 2023, arXiv:2307.03326 [physics.optics].10.1002/adpr.202400083Suche in Google Scholar

[44] O. H.-C. Cheng, B. Zhao, Z. Brawley, D. H. Son, and M. T. Sheldon, “Active tuning of plasmon damping via light induced magnetism,” Nano Lett., vol. 22, no. 13, p. 5120, 2022. https://doi.org/10.1021/acs.nanolett.2c00571.Suche in Google Scholar PubMed

[45] Y. Mou, X. Yang, B. Gallas, and M. Mivelle, “A Reversed Inverse Faraday Effect,” Adv. Mater. Technol., vol. 8, no. 21, 2023.10.1002/admt.202300770Suche in Google Scholar

[46] D. Barchiesi, S. Kessentini, N. Guillot, M. L. de la Chapelle, and T. Grosges, “Localized surface plasmon resonance in arrays of nano-gold cylinders: inverse problem and propagation of uncertainties,” Opt. Express, vol. 21, no. 2, p. 2245, 2013. https://doi.org/10.1364/oe.21.002245.Suche in Google Scholar PubMed

[47] M. J. Gomez, K. Liu, J. G. Lee, and R. B. Wilson, “High sensitivity pump–probe measurements of magnetic, thermal, and acoustic phenomena with a spectrally tunable oscillator,” Rev. Sci. Instrum., vol. 91, no. 2, p. 023905, 2020. https://doi.org/10.1063/1.5126121.Suche in Google Scholar PubMed

[48] V. H. Ortiz, S. Coh, and R. B. Wilson, “Magneto-optical Kerr spectra of gold induced by spin accumulation,” Phys. Rev. B, vol. 106, no. 1, p. 014410, 2022. https://doi.org/10.1103/physrevb.106.014410.Suche in Google Scholar

[49] W.-S. Chang, et al.., “Tuning the acoustic frequency of a gold nanodisk through its adhesion layer,” Nat. Commun., vol. 6, no. 1, p. 7022, 2015. https://doi.org/10.1038/ncomms8022.Suche in Google Scholar PubMed

[50] I. Zorić, M. Zäch, B. Kasemo, and C. Langhammer, “Gold, platinum, and aluminum nanodisk plasmons: material independence, subradiance, and damping mechanisms,” ACS Nano, vol. 5, no. 4, p. 2535, 2011. https://doi.org/10.1021/nn102166t.Suche in Google Scholar PubMed

[51] C. A. Downing and G. Weick, “Plasmonic modes in cylindrical nanoparticles and dimers,” Proc. R. Soc. A, vol. 476, no. 2244, p. 20200530, 2020. https://doi.org/10.1098/rspa.2020.0530.Suche in Google Scholar PubMed PubMed Central

[52] X. Xie and D. G. Cahill, “Thermometry of plasmonic nanostructures by anti-Stokes electronic Raman scattering,” Appl. Phys. Lett., vol. 109, no. 18, p. 183104, 2016. https://doi.org/10.1063/1.4966289.Suche in Google Scholar

[53] R. B. Wilson and S. Coh, “Parametric dependence of hot electron relaxation timescales on electron-electron and electron-phonon interaction strengths,” Commun. Phys., vol. 3, no. 1, p. 179, 2020. https://doi.org/10.1038/s42005-020-00442-x.Suche in Google Scholar

[54] N. Liu, M. Mesch, T. Weiss, M. Hentschel, and H. Giessen, “Infrared perfect absorber and its application as plasmonic sensor,” Nano Lett., vol. 10, no. 7, p. 2342, 2010. https://doi.org/10.1021/nl9041033.Suche in Google Scholar PubMed

[55] L. Dykman and N. Khlebtsov, “Gold nanoparticles in biomedical applications: recent advances and perspectives,” Chem. Soc. Rev., vol. 41, no. 6, p. 2256, 2012. https://doi.org/10.1039/c1cs15166e.Suche in Google Scholar PubMed

[56] M. Notarianni, K. Vernon, A. Chou, M. Aljada, J. Liu, and N. Motta, “Plasmonic effect of gold nanoparticles in organic solar cells,” Sol. Energy, vol. 106, p. 23, 2014. https://doi.org/10.1016/j.solener.2013.09.026.Suche in Google Scholar

[57] F. Cheng, J. Gao, T. S. Luk, and X. Yang, “Structural color printing based on plasmonic metasurfaces of perfect light absorption,” Sci. Rep., vol. 5, no. 1, p. 11045, 2015. https://doi.org/10.1038/srep11045.Suche in Google Scholar PubMed PubMed Central

[58] F. Hache, D. Ricard, C. Flytzanis, and U. Kreibig, “The optical kerr effect in small metal particles and metal colloids: the case of gold,” Appl. Phys. A, vol. 47, no. 4, p. 347, 1988. https://doi.org/10.1007/bf00615498.Suche in Google Scholar

[59] T. V. Shahbazyan, I. E. Perakis, and J.-Y. Bigot, “Size-dependent surface plasmon dynamics in metal nanoparticles,” Phys. Rev. Lett., vol. 81, no. 15, p. 3120, 1998. https://doi.org/10.1103/physrevlett.81.3120.Suche in Google Scholar

[60] J. H. Hodak, I. Martini, and G. V. Hartland, “Spectroscopy and dynamics of nanometer-sized noble metal particles,” J. Phys. Chem. B, vol. 102, no. 36, p. 6958, 1998. https://doi.org/10.1021/jp9809787.Suche in Google Scholar

[61] A. Aiboushev, et al.., “Spectral properties of the surface plasmon resonance and electron injection from gold nanoparticles to TiO2 mesoporous film: femtosecond study,” Photochem. Photobiol. Sci., vol. 12, no. 4, p. 631, 2013. https://doi.org/10.1039/c2pp25227a.Suche in Google Scholar PubMed

[62] T. Higo, et al.., “Large magneto-optical Kerr effect and imaging of magnetic octupole domains in an antiferromagnetic metal,” Nat. Photonics, vol. 12, no. 2, p. 73, 2018. https://doi.org/10.1038/s41566-017-0086-z.Suche in Google Scholar PubMed PubMed Central

[63] G. Armelles, A. Cebollada, A. García-Martín, and M. U. González, “Magnetoplasmonics: combining magnetic and plasmonic functionalities,” Adv. Opt. Mater., vol. 1, no. 1, p. 10, 2013. https://doi.org/10.1002/adom.201200011.Suche in Google Scholar

[64] C. Langhammer, B. Kasemo, and I. Zorić, “Absorption and scattering of light by Pt, Pd, Ag, and Au nanodisks: absolute cross sections and branching ratios,” J. Chem. Phys., vol. 126, no. 19, p. 194702, 2007. https://doi.org/10.1063/1.2734550.Suche in Google Scholar PubMed

[65] J. Park, J. Huang, W. Wang, C. J. Murphy, and D. G. Cahill, “Heat transport between au nanorods, surrounding liquids, and solid supports,” J. Phys. Chem. C, vol. 116, no. 50, p. 26335, 2012. https://doi.org/10.1021/jp308130d.Suche in Google Scholar

[66] L. T. Vuong, A. K. González-Alcalde, V. Ortiz, X. Shi, J. Feng, and R. B. Wilson, “Studies of inverse faraday effect with varying thickness nanodisks,” in Enhanced Spectroscopies and Nanoimaging, vol. PC12654, P. Verma and Y. D. Suh, Eds., International Society for Optics and Photonics SPIE, 2023, p. PC126540N.10.1117/12.2678121Suche in Google Scholar


Supplementary Material

This article contains supplementary material (https://doi.org/10.1515/nanoph-2023-0777).


Received: 2023-11-03
Accepted: 2024-01-08
Published Online: 2024-02-05

© 2024 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Heruntergeladen am 17.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/nanoph-2023-0777/html
Button zum nach oben scrollen