Flexible micro/nano motors responsive to light sources are highly desirable. Conventional photothermal (PT) techniques have provided significant light-actuation methods; however, the dynamic responses of these devices in high frequency domain are severely restricted by the frequency response limitation. To overcome this limitation, the thermal-, electric-, and acoustic-near field interactions in the PT process and its accompanied photoacoustic (PA) process are investigated, and a plasmon enhanced PA actuation method is proposed. The significant improvement in PA oscillation by strong localized surface plasmon resonance provides a powerful means to realize a broadband response laser motor without the frequency response limitation. This PA laser motor could generate PA oscillation of over 5 µm by pulse laser with repetition frequencies of 1 Hz to 20 kHz and has a maximum value of 18.8 µm at the mechanical resonant frequency. It demonstrates the effectiveness of this PA laser motor in offsetting the shortcomings of the PT actuation method. The device requires no electrical or chemical energy, and it has potential benefits, such as bionic research into insect flapping, vocal-cord vibration, and muscular movement in fields including micro/nano physics, biochemistry, and clinical medicine.

Laser motors are essential components in micro-opto-electro-mechanical systems for a wide range of applications, and they have the advantages of non-contact actuation, easy miniaturization, and strong resistance to electromagnetic interference. Actuation based on optical radiation,1,2 photophoretic force,3,4 and photochemical force5,6 are well-known light-driven methods. However, the light forces from these methods are too small (∼pN)2,7,8 to drive motors that are several-hundred micrometers in size, especially when operating in non-liquid environments.9,10 Photothermal (PT) techniques based on nonradiative conversion of absorbed light energy into heat have shown a promising way for laser actuation by providing greater energy transfer depth and breadth. PT actuation can provide large forces and deformations; for example, Palagi et al. demonstrated a liquid-crystal elastomer robot of 1 mm in length and 200–300 µm in diameter that could move with a crawling deformation of 110 µm under laser irradiation pulsed at 2 Hz.11 Zhang et al. investigated a series of polymeric and metallic materials for a PT actuator with a maximum deformation of 12–20 µm and a dynamic response of 1–20 Hz.12–15 Ma et al. investigated all-inorganic PT actuators using a phase-change material to deliver a very large deformation of over 800 µm and fast response of up to 100 Hz, despite a restricted working temperature.16 Other PT actuators/motors using graphene and carbon nanotubes have also been studied.17–23 However, a problem highlighted by these studies is that the dynamic responses of workable PT actuators/motors are no more than 100 Hz, as presented in Table I; here, we refer to this issue as the frequency response limitation for laser motor actuation (for details, see the supplementary material, Note S1 and Fig. S1).

TABLE I.

Studies of PT actuators/motors and their dynamic responses.

Actuator/motor systemLight source (mW/cm2)Upper limit of response (Hz)References
Liquid-crystal elastomer (3–30) × 104 10 11 and 24  
Polymer material (polypropylene, HDPE) (2–2.5) × 104 20 12–14  
Metal (nickel) 3 × 107 20 15  
VO2/CNT 800–1600 77 16  
Graphene oxide-polymer/polycarbonate bilayer 38–106 0.25 17  
PEI-graphene oxide polymer/NOA-73 bilayer 100 0.1 18  
Graphene oxide-CNT/PDMS 250 0.2 19  
Graphene-elastin composite 0.5–5.7 0.2 20  
Single wall CNT/polycarbonate membrane 30–100 21  
Single wall CNT/poly(N-isopropylacrylamide) 9.5 × 106 0.4 22  
CNT/paraffin wax/polyimide 25–100 0.5 23  
Actuator/motor systemLight source (mW/cm2)Upper limit of response (Hz)References
Liquid-crystal elastomer (3–30) × 104 10 11 and 24  
Polymer material (polypropylene, HDPE) (2–2.5) × 104 20 12–14  
Metal (nickel) 3 × 107 20 15  
VO2/CNT 800–1600 77 16  
Graphene oxide-polymer/polycarbonate bilayer 38–106 0.25 17  
PEI-graphene oxide polymer/NOA-73 bilayer 100 0.1 18  
Graphene oxide-CNT/PDMS 250 0.2 19  
Graphene-elastin composite 0.5–5.7 0.2 20  
Single wall CNT/polycarbonate membrane 30–100 21  
Single wall CNT/poly(N-isopropylacrylamide) 9.5 × 106 0.4 22  
CNT/paraffin wax/polyimide 25–100 0.5 23  

The PT method with sharp pulsed laser is capable accompanied by photoacoustic (PA) wave generation, which has been widely used in non-destructive testing25 and microscopy,26 and it was recently exploited as a new power source for laser motor research.9,10 Laser motors actuated by the PA method react much faster in the high-frequency domain because they utilize acoustic energy rather than heat27 and do not suffer from the frequency response limitation. However, in most cases, the accompanying PA response is extremely weak in a laser PT process, and thus, if a strong PA deformation is desirable such as for laser motor applications, much higher energy laser sources must be used.10,28,29 The high-power laser could damage the laser motor and limit its use in most areas such as biological applications.

To overcome these limitations, we theoretically and experimentally investigated the thermal-, electric-, and acoustic-near field interactions in the PT/PA process and the significant improvement in PA responses, including the PA waves and their accompanied PA oscillations (for details about the PA oscillation, see the supplementary material, Note S2 and Fig. S2), by strong localized surface plasmon resonance (LSPR). A function layer was created by depositing Au nanoparticles (AuNPs) on the laser motor, and the AuNPs layer acts as a light absorption enhancer and could potentially address the issues of energy deficiency for strong PA wave generation. The enhancement may be due to two factors: first the AuNPs exhibit a strong LSPR effect, and this allows them to effectively absorb an incident laser.30 Second, because of the close proximity of these nanoabsorbers when in clusters, their laser-induced electric- and acoustic-near field response may overlap spatially and temporally, causing further enhancement of PA effects. The energy carried by PA waves could induce a fast and large oscillation upon the motor beyond the frequency response limitation, which could reach tens of kilohertz. The broadband response of the AuNP-PA laser motor could benefit bionic research on insect flapping, vocal-cord vibration, and muscular movement. Importantly, no electrical or chemical energy is required because the driving force is derived from acoustic dissipation excited by the light source. This device could be used in a wide range of applications in fields including micro/nano physics, biochemistry, and clinical medicine.

The experimental setup is shown in Figs. 1(a) and 1(b), and the setup (for details, see the supplementary material, Fig. S3) was constructed on the technical platform of a laser-Doppler vibrometer microscope (LV-S01-M, Sunny Optical Co., Ltd., China), onto which the sample and the detectors were incorporated. A solid-state pulsed laser (IDOL-S, Xiton Photonics GmbH, Germany) with a wavelength of 480–600 nm, a pulse duration of 50 ns, an average laser power of 0.2 W, and a repetition frequency of 1 Hz to 20 kHz was used to irradiate the sample, and the sample is fixed by two microscope slides. A 590 nm narrow-band pass filter was used to selectively allow only the LSPR wavelength to pass. Micrographs of the laser motor in the yz plane are shown in Fig. 1(c), and its thickness (the x-axis) was 120 µm. The laser irradiated along the x-direction, as shown in the left-half of the bottom picture of Fig. 1(c), and the laser spot was focused to a diameter of 100 µm. An electromagnetic acoustic transducer (EMAT) cooperated with the laser-Doppler vibrometer microscope was used to detect the PT deformation, PA oscillation, and PA waves. The three signals were identified based on their features in response frequencies, curve shapes, and deformation directions, as summarized in Table II, and see the supplementary material, Note S2 and Fig. S2. The EMAT sensor was made according to Ref. 31, and its magnetic-field direction was adjusted to correct the electromagnetic induction during detection. Unless otherwise stated, both the PT and PA experiments used the same laser conditions, that is, the laser was swept at different repetition frequencies with an identical average laser power, laser wavelength, pulse duration, and illumination position. An objective lens with a charge-coupled device (CCD) camera (MC-D500U, Phoenix Optical Co., Ltd., China) was used to visualize and capture the experimental results.

FIG. 1.

(a) and (b) Experimental apparatus for excitation, detection, and observation of PT deformation, PA oscillation, and PA waves, and (c) micrographs of the laser motor in the yz plane.

FIG. 1.

(a) and (b) Experimental apparatus for excitation, detection, and observation of PT deformation, PA oscillation, and PA waves, and (c) micrographs of the laser motor in the yz plane.

Close modal
TABLE II.

Signal features of PT deformation, PA oscillation, and PA wave concerned in this study.

SignalCurve shapeFrequency responseDeformation
PT   0–100 Hz Direction: Stretches in the z axis;  
deformation   amplitude: Dozens of micrometers, and decays 
   with a laser repetition frequency 
PA  4–20 kHz Direction: Oscillates in the y axis; amplitude: 
oscillation   Dozens of micrometers, a maximum value is acquired 
    when laser repetition frequency equals the mechanical 
   resonant frequency of the laser motor 
PA wave  ∼20 MHz Direction: Propagates along the surface of the 
   laser motor; amplitude: Dozens of 
   nanometers, does not change with a laser 
   repetition frequency 
SignalCurve shapeFrequency responseDeformation
PT   0–100 Hz Direction: Stretches in the z axis;  
deformation   amplitude: Dozens of micrometers, and decays 
   with a laser repetition frequency 
PA  4–20 kHz Direction: Oscillates in the y axis; amplitude: 
oscillation   Dozens of micrometers, a maximum value is acquired 
    when laser repetition frequency equals the mechanical 
   resonant frequency of the laser motor 
PA wave  ∼20 MHz Direction: Propagates along the surface of the 
   laser motor; amplitude: Dozens of 
   nanometers, does not change with a laser 
   repetition frequency 
The frequency response limitation in the PT actuation process is analyzed by evaluating the heating of the laser motor after laser irradiation when its shape is deformed by thermal stress. The heat source formed by the laser irradiation can be established as
Q=0tP0αsin(ωt)dt,
(1)
where P0 is the average laser power and has a value of 0.2 W, α is the light absorbance, which is ∼0.8 for AuNPs decorated metallic nickel at an incident light wavelength ∼590 nm, and ω is the angular frequency ω = 2πfL, where fL is the laser repetition frequency. Analysis of the heat-transfer process is described by the heat-conduction equation,
CρTt+(kT)=Q,
(2)
where T, C, ρ, and k are the temperature, specific heat capacity, density, and thermal conductivity of the laser motor, respectively. Considering the thermal exchange between the motor and the external environment, appropriate boundary conditions should be established. A Neumann boundary was applied based on the thermal load Q. Boundaries constraining thermal exchange at the material–air interface are described by32 
n(kT)=h(TTr)+εσ(T4Tr4),
(3)
where n is an outward-facing vector normal to the motor surface, h is the convective heat transfer coefficient of air, which has a value of 30 W/(m2 K) considering the interference factors, ε is the surface emissivity, which is set at 0.2 for nickel, and σ = 5.67 × 10−8 W/(m2 K4) is the Stefan–Boltzmann constant. The room temperature was Tr = 300 K. Other parameters in Eqs. (1)(3) are set based on the experimental design and material characteristics of pure nickel.
The COMSOL Multiphysics software was used to solve Eqs. (1)(3). Figure 2(a) shows the simulated temperature of the laser motor under different laser repetition frequencies, where the temperature is recorded based on the average value of five evenly distributed measuring points along the z-axis of the motor. It can be observed that the temperature has a sine-wave response, with its frequency in sync with the laser repetition frequency. At low repetition frequencies (e.g., 5 Hz), the temperature of the structure cycles between 508 and 615 K—a temperature difference ΔT of more than 107 K. However, at higher repetition frequencies (e.g., 25 Hz), its sinusoidal amplitude decreases, and ΔT shrinks to only 15 K. ΔT becomes smaller with increasing laser repetition frequency because the laser motor cannot heat or cool fast enough. The equilibrium temperature of the thermal transfer process is TC ≈ 557 K. The dynamic thermal deformation ΔD is proportional to ΔT based on the thermal expansion equation,33 
D(z)=K0zT(z)dz,
(4)
where K is the thermal expansion coefficient of nickel (1.5 × 10−6/K). For temperature TC, the laser motor deflects to a fixed position with a deformation of DC, and ΔD is superimposed on this offset deflection, as shown in Fig. 2(b). When the laser repetition frequency is relatively low, ΔD is large due to a bigger ΔT [Fig. 2(b) top], while when the laser repetition frequency increases, ΔD decreases quickly because of the smaller and smaller ΔT [Fig. 2(b) bottom]. In addition to temperature, the thermal deformation is also restricted by the elastic deformation limit DE for linear actuation, which makes the deformation respond no longer in sine but trapezoidal waves. A conclusion can be drawn from the analysis that, limited by thermal transfer on the whole laser motor, the PT deformation generally cannot be as fast as laser irradiation in the high frequency domain, which causes the frequency response limitation.
FIG. 2.

(a) Average temperature variation for the PT deformation as a function of time for irradiation at different laser repetition frequencies. The temperature difference decreases with increasing laser repetition frequency. (b) Comparison of PT deformation under high/low-rate PT processes, where the dynamic PT deformation decreases quickly with a decreasing temperature difference.

FIG. 2.

(a) Average temperature variation for the PT deformation as a function of time for irradiation at different laser repetition frequencies. The temperature difference decreases with increasing laser repetition frequency. (b) Comparison of PT deformation under high/low-rate PT processes, where the dynamic PT deformation decreases quickly with a decreasing temperature difference.

Close modal

A plasmonic–acoustic-based PA actuation method was introduced to improve the performance of the laser motor in the high-frequency domain, as shown in Figs. 3(a) and 3(b), the PA mechanism converts the energy from optical irradiation into a localized temperature increase, which generates a mechanical acoustic wave through the thermal elastic effect.34,35 Material is a crucial factor in the PA principle, where a higher optical energy absorption capability can excite stronger PA waves and larger PA oscillations. For this research, AuNPs with 100 nm diameter were chosen as the PA-generation material due to their high optical energy absorption capabilities at plasmon-resonant frequencies. An absorption band results when the incident light frequency is resonant with the collective oscillation of the conduction-band electrons of the AuNPs, that is, when LSPR occurs.36–38 To facilitate explaining the mechanism that LSPR enhances PA excitation, we present near-field distributions that include electric displacement vectors filling electric fields [Fig. 3(c-i)] and thermal energy field distribution [Fig. 3(c-ii)] simulated using COMSOL software. Surface charges oscillate in the electric field, which is induced by reverse polarity surface charges on surfaces of the nanosphere and the electric field of light (linearly polarized along the z-direction). A large amount of thermal loss caused by electron oscillation is deposited at the interface between the Au nanosphere and nickel substrate, which forms a hotspot. The PA generation is quite different from the slow thermal expansion and contraction process of the PT mechanism,39,40 and it follows the principle of laser based thermal-elastic mechanism. When the hotspot is formed on the surface of the laser motor, the energy is transferred to acoustic deformation through the thermal elastic process within a very short time (∼ns), Fig. 3(d) shows the variation of the acoustic field after the laser irradiation in 25 ns [Fig. 3(d-i)] and 50 ns [Fig. 3(d-ii)] respectively, and the high energy hotspot rapidly heats a localized volume, in which an impulsive thermal expansion induces lattice oscillations and launches acoustic pulses (i.e., PA waves) into the overlying sample.

FIG. 3.

Plasmonic–acoustic-based PA process. (a) Schematic illustration of pulse laser induced PA wave on the laser motor and (b) laser induced LSPR on the AuNP; the surface-plasmon field-enhanced laser heating results in a high local temperature gradient and generates PA wave on the surface of the laser motor. (c) Displacement vector filling electric-field distribution [top row (i)] and energy field distribution [bottom row (ii)] at the xz plane under an excitation wavelength of 590 nm, and the electric-field is estimated from lg |E/E0|, where E0 is the incident electric field. (d) Acoustic field distribution of laser induced PA waves, 25 ns after laser irradiation [top row (i)] and 50 ns after laser irradiation [bottom row (ii)].

FIG. 3.

Plasmonic–acoustic-based PA process. (a) Schematic illustration of pulse laser induced PA wave on the laser motor and (b) laser induced LSPR on the AuNP; the surface-plasmon field-enhanced laser heating results in a high local temperature gradient and generates PA wave on the surface of the laser motor. (c) Displacement vector filling electric-field distribution [top row (i)] and energy field distribution [bottom row (ii)] at the xz plane under an excitation wavelength of 590 nm, and the electric-field is estimated from lg |E/E0|, where E0 is the incident electric field. (d) Acoustic field distribution of laser induced PA waves, 25 ns after laser irradiation [top row (i)] and 50 ns after laser irradiation [bottom row (ii)].

Close modal

On the other hand, the wavelength of the PA wave was calculated to be ∼249 µm based on the equation b = (a2 + v2τ2)1/2,41 where a is the radius of the laser intensity distribution (50 µm), v is the speed of the acoustic wave (∼4900 m/s, see the supplementary material, Note. S3 and Fig. S4), τ is the laser pulse duration (50 ns), and b is the characteristic wavelength of the PA wave. A relationship is drawn that the wavelength of the PA wave is much larger than the overlying sample thickness (120 µm); for the ultrathin plate structure, the radiation force of such PA wave could cause the metal plate-arm of the laser motor to oscillate (for details, see the supplementary material, Note S4 and Fig. S5). The PA generation does not lose efficiency just by changing the laser repetition frequency; the laser motor reacts much faster under the PA mechanism and can surpass the frequency response limitation in the PT response. By comparison, although the same nanosecond laser is used, the laser motor response under the PT mechanism is invalid with a high repetition frequency laser input (e.g., >100 Hz). The different response for PT and PA signals at high frequencies is mainly because global heat conduction cannot transfer as fast as localized thermal interactions.

Significantly, the enhancement might not only occur in the particle–substrate junctions but also in the particle–particle gap. The distribution of AuNPs aggregates on the laser motor, as the SEM picture shown in the inset of Fig. 4(a) is partly like a heptamer, thus a simulated model of an AuNPs heptamer situated on the laser motor substrate is systematically analyzed including the absorption spectra, the plasmonic modes, and their contributions to the formations of hot spots by considering the far field, near field, and their relevance. The cross section of absorption compared to scattering and extinction is shown in Fig. 4(a); for the AuNPs heptamer-substrate system, a broad resonant peak of the simulated absorption spectrum centered at 590 nm originates from hybridized LSPR modes generated on the AuNPs and the nickel substrate. The electric field enhancements would be more complicated because of the competition and synergistic interactions of LSPR of AuNPs, near-field coupling effect of particle–particle and particle–substrate, and the propagating surface plasmons (PSPs) on the substrate.42 First, the nonradiative PSPs on the substrate make a significant contribution to near field enhancement rather than the scattering toward the far field.43 More importantly, the dipoles and higher order multipoles that existed between AuNP aggregates critically influence the near field enhancement. The electric field enhancement (defined as lg |E/E0|) is over 13 [Fig. 4(b-i)] in the particle–particle gaps at 590 nm exciting wavelength, compared to about 8 at the particle–substrate junction. These results can be concluded as the dipole coupling mode between AuNPs heptamer, which belongs to bounded surface charge density wave originating from the collective oscillation of free electrons induced by the incident light,44 that is, there are more surface charges accumulate at interparticle junctions compared to particle–substrate junctions. These enhanced-field distributions corresponding to hybridized LSPR modes show strong energy hot spots at the particle–particle gaps and particle–substrate junctions [Fig. 4(b-ii)], enjoying over 1 × 105 of maximum energy enhancement compared to bare nickel substrate excitation.

FIG. 4.

Absorption spectra, near-field profiles, and acoustic characteristics of 100 nm heptamer AuNPs system. (a) Calculated absorption, scattering, and extinction cross sections of heptamer AuNPs on the laser motor. The insets show the heptamer AuNPs model, and a SEM picture of the AuNPs aggregates on the laser motor, the scale bar is 100 nm. (b) Calculated electric field (i) and energy (ii) enhancement distributions at the xz plane under an excitation wavelength of 590 nm in heptamer AuNPs. (c) PA wave signals induced on the AuNP-decorated PA laser motor (i) and on the undecorated PA laser motor (ii), and amplitude (vertical axis) and timescale (horizontal axis) are 10 nm/div and 100 ns/div, respectively.

FIG. 4.

Absorption spectra, near-field profiles, and acoustic characteristics of 100 nm heptamer AuNPs system. (a) Calculated absorption, scattering, and extinction cross sections of heptamer AuNPs on the laser motor. The insets show the heptamer AuNPs model, and a SEM picture of the AuNPs aggregates on the laser motor, the scale bar is 100 nm. (b) Calculated electric field (i) and energy (ii) enhancement distributions at the xz plane under an excitation wavelength of 590 nm in heptamer AuNPs. (c) PA wave signals induced on the AuNP-decorated PA laser motor (i) and on the undecorated PA laser motor (ii), and amplitude (vertical axis) and timescale (horizontal axis) are 10 nm/div and 100 ns/div, respectively.

Close modal

The energy enhancement of the AuNPs function layer is confirmed by the PA wave signals detected by the EMAT, and the laser-Doppler vibrometer microscope was also used for displacement amplitude calibration. The 100 nm AuNPs were prepared by a seed-mediated growth method, for details see Sec. VI. Then, the AuNPs were anchored on the yz surface of the laser motor through a continuous layer deposition process, for details see Sec. VI. In order to better analyze the PA wave signals, a one-shot laser pulse (pulse energy = 100 µJ) was employed. It can be seen from Fig. 4(c) that both the amplitude and relaxation time of the wave excited on the AuNP-PA laser motor were much larger—∼60 nm and 1 µs for the AuNP-PA laser motor [Fig. 4(c-i)] compared to 20 nm and 0.2 µs for the undecorated motor [Fig. 4(c-ii)], respectively. This meant that the acoustic field of the AuNP-PA laser motor contained more energy and a greater proportion of the incident light was converted into acoustic energy. The pulse width τs of the PA wave was ∼50.8 ns based on equation τs = b/v,39 which was consistent with the laser pulse width and definitively proved that the PA wave was induced by the laser.

The frequency response and amplitude of PT deformation and PA oscillation of the laser motor were detected and compared in the experiment, as shown in Figure 5(a); in order to eliminate the influence of the system vibration noise, all the detection processes were conducted on an optical platform, and the signals were filtered. The PT deformation decreased exponentially with increasing laser repetition frequency, from 17.4 to 0.7 µm at frequencies of 0–100 Hz, respectively. There was no detectable PT deformation at higher frequencies. By comparison, the laser motor under AuNP-PA excitation had strong oscillations of greater than 5 µm across a broad range of frequencies from 1 Hz to 20 kHz. A mechanical resonance with a maximum output of ∼18.8 µm occurred near 4 kHz when the laser repetition frequency matched the mechanical resonant frequency of the laser motor. The real-time response of the PT deformation and PA oscillation was also compared and analyzed. It can be seen from the xy views of the laser motor under the CCD camera that, when the laser irradiated the laser motor at a repetition frequency of 1 Hz, it swelled [Fig. 5(b-i), multimedia view] because the PT deformation of the laser motor beam along the z axis caused the viewing window to become out-of-focus during optical monitoring. While the repetition rate of the laser increased to a higher frequency of 10 kHz, there was no swelling could be observed any longer. In contrast, a rapid and directional forced PA oscillation was visualized at 10 kHz along the y axis [Fig. 5(b-ii), multimedia view]. The dynamic response of the laser motor under the PA and PT mechanisms was quite different even under identical irradiation conditions. In general cases, the PT deformation and PA oscillation would overlap with each other, but their strengths may vary significantly, and they would rarely be noticed at the same time because the PA oscillation is always too weak, as shown by the curve with blue triangles in Fig. 5(a). The undecorated laser motor cannot generate a detectable PA oscillation in most frequency bands, such as 0–1000 Hz and >10 kHz. With proper NP decoration, the AuNP-PA laser motor could produce an output with an amplitude that is comparable to that of the PT deformation but with a much broader frequency response. The reproducibility was also evaluated in the experiments, a batch of three AuNP-PA laser motors was prepared following the same fabrication route, the SEM images of the assembled AuNPs on the laser motor are shown in Fig. 5(c), and the result confirms that similar distribution of AuNPs with only a few relative position variances is achieved on the laser motor. The reproducibility is characterized by the amplitude deviation of PA oscillations, as the error bars shown in Fig. 5(a); by dividing the standard deviation by the mean value, the variation coefficients are calculated to be less than 8%, confirming a high reproducibility of the AuNP-PA laser motor. The fast reaction and large amplitude in the high frequency domain with good stability and high reproducibility demonstrate the effectiveness of this AuNP-PA laser motor in solving the frequency response limitation and being promising for finding applications in micro/nano physics, biochemistry, clinical medicine, and so on.

FIG. 5.

(a) Amplitude–frequency curve of the PT deformation (blocks), AuNP-decorated PA oscillation (circles), and the PA oscillation without decoration (triangles) is also shown for comparison. The inset figure shows an amplitude–time curve of PT deformation and PA oscillation at a frequency of 100 Hz. (b) Visualization experiments involving real-time observation of the dynamic responses of the PT deformation (i) and (ii) PA oscillation. The PT deformation swells in the viewing window due to thermal expansion along the z axis, causing the CCD camera to become out-of-focus, while PA oscillation undergoes a kind of rapid oscillation along the y axis. The deformation patterns of the laser motor under PT/PA excitation are also displayed. (c) SEM images of assembled AuNPs on three laser motors following the same fabrication route, the scale bar is 500 nm (Multimedia available online).

FIG. 5.

(a) Amplitude–frequency curve of the PT deformation (blocks), AuNP-decorated PA oscillation (circles), and the PA oscillation without decoration (triangles) is also shown for comparison. The inset figure shows an amplitude–time curve of PT deformation and PA oscillation at a frequency of 100 Hz. (b) Visualization experiments involving real-time observation of the dynamic responses of the PT deformation (i) and (ii) PA oscillation. The PT deformation swells in the viewing window due to thermal expansion along the z axis, causing the CCD camera to become out-of-focus, while PA oscillation undergoes a kind of rapid oscillation along the y axis. The deformation patterns of the laser motor under PT/PA excitation are also displayed. (c) SEM images of assembled AuNPs on three laser motors following the same fabrication route, the scale bar is 500 nm (Multimedia available online).

Close modal
In this paper, we proposed a new driving method for micro/nano laser motor based on PA oscillation to overcome the frequency response limitation, and the theoretical and experimental results showed that the AuNP-PA laser motor could respond effectively in a very broadband of 1 Hz to 20 kHz, which confirmed the feasibility of such method. Importantly, even though the proposed AuNP-PA laser motor possesses an efficient combination of large displacement amplitude and wide frequency response spectrum, the optimization of the PA oscillation for laser motor applications needs to be analyzed. First, the optimal laser heating conditions should be considered, and it can be judged by the criteria,45,
ττT,
(5)
where τT is the characteristic thermal relaxation time. For an approximation of the thermal relaxation of many closely located NPs, τT can be estimated as τT = R2ρNPCNP/3κS,46 where ρNP and CNP are considered as the average density and heat capacity of NP, R is the radius of NP, and κS is coefficient of thermal conductivity to the surrounding medium. For the R = 50 nm NPs surrounded by ambient air in this study, the characteristic thermal relaxation time τT is roughly 85 ns, see the supplementary material, Table S1 for the parameters of NP and surrounding medium; thus, the 50 ns pulse duration laser employed in the study is within the optimum condition for PA oscillation excitation. The optimum pulse duration should be a bit shorter if operating the AuNP-PA laser motor in a liquid environment, such as water, the pulse duration should be less than 3.5 ns.46 It is worth mentioning that for shorter laser pulses like the picosecond or femtosecond lasers, the excitation process of the PA wave is quite different and does not obey the thermal-elastic mechanism any longer. The generation of acoustic stress using such ultra short laser pulses can no longer be considered an instantaneous process, and the nonequilibrium dynamics of the electrons and phonons diffusion must be considered;47 in this process, the photons absorbed in NPs excite electrons to higher available states, and the nonequilibrium high energy electrons are coupled with the phonons forming a strain source, which launches stress pulses and transfers the energy from the excited electrons to the lattice in the medium. In this study, the most important factor is not only photoacoustic interactions on NPs but also the acoustic coupling between NPs and the nickel substrate through thermal-elastic stress and strain, and strong PA waves propagating on the substrate (i.e., propagating on the whole structure of laser motor) is essential for the PA oscillation excitation; therefore, it needs deeper research to estimate whether such electron–phonon interaction excited using picosecond/femtosecond laser can achieve the same effect as the thermal-elastic mechanism in the laser motor application.

Second, the optimization for NP should be considered; in general, the PA signal is proportional to laser energy, the absorption cross section, and the coefficient of thermal expansion;27 metal nanoparticles, such as AuNPs and AgNPs, are commonly used for PA excitation due to their strong light absorption ability and a high coefficient of thermal expansion. In addition, the size of NP is also a crucial factor because it can influence light absorption by the LSPR effect. The absorption spectrum with different diameter AuNPs shows that the absorption peak shifts to longer wavelength with bigger size NPs, e.g., the absorption peak is located at a wavelength of about 530 nm for 20 nm AuNPs and a wavelength of about 590 nm for 100 nm AuNPs, see the supplementary material, Fig. S6. The maximum absorptivity is roughly the same for AuNPs smaller than 100 nm; thus, the only considered optimum condition for these size NPs is whether the LSPR wavelength matches the laser wavelength. However, in applications, the smaller the NPs the shorter the laser pulse duration required based on the criteria ττT; it will be challenging for technological as well as safety constraints. The NPs larger than 100 nm are not quite suited to the PA method in laser motor application because the absorptivity at LSPR wavelength is relatively much lower compared to NPs smaller than 100 nm, which results in low efficiency of photoacoustic conversion; nevertheless, it may be an option for high laser energy pumped PA motor because smaller NPs are more likely to be melted by high energy laser.48 

The geometry of the AuNP-PA laser motor is also an important factor that needs to be considered; in this study, the laser motor is specially designed with a short-thin arm and a long-wide arm, of which the short-thin arm acts as a PA wave launcher while the long-wide arm functions as an elastically suspended proof-mass to produce an inertial force. The displacement of the PA oscillation is roughly proportional to the motor length based on the geometric relationship, see the supplementary material, Note S4 and Fig. S5, so the long-wide arm can provide larger PA oscillation. On the other hand, benefiting from the ultra-high frequency of the PA oscillation, a huge acceleration can be obtained by the AuNP-PA laser motor, and the acceleration produced by the AuNP-PA laser motor can be up to 3155–94 652 m/s2 depending on the laser driving frequency and oscillation modes (for details, see the supplementary material, Note S5, Note S6, Figs. S7 and S8). Based on such advanced features of the AuNPs laser motor, we look forward to it being designed as a kind of nano biomimetic robot as shown in Fig. 6. The robot can obtain a big push force by the high frequency flapping of the mechanical wings and move with a fast speed; it can be used to remove blood clots and so on. We believe that there will be good application prospects for the AuNP-PA laser motor in the future.

FIG. 6.

A prospect application of the AuNP-PA laser motor: (a) a design of a dragonfly-like nano robot with its wings actuated using the AuNP-PA laser motor and (b) a possible application of the dragonfly-like nano robot for cleaning blockages in blood vessels.

FIG. 6.

A prospect application of the AuNP-PA laser motor: (a) a design of a dragonfly-like nano robot with its wings actuated using the AuNP-PA laser motor and (b) a possible application of the dragonfly-like nano robot for cleaning blockages in blood vessels.

Close modal

Laser induced PT and accompanying PA signals, including PT deformation, PA oscillation, and PA wave, are detected separately. I, detection method for PT deformation. Align the detecting light of the laser Doppler vibrometer along the z direction, when the PT deformation moves along the z axis based on the experimental system, and it can be detected and identified separately because the PA oscillation (move along the y direction) and PA wave (move along the x direction) do not have any motion relative to the laser Doppler vibrometer; thus, they cannot be detected. In addition, the PT deformation can also be identified by aligning the magnetic field direction of the EMAT along the y direction; in this arrangement, only the PT deformation can cut the magnetic induction line and produce signal currents. II, detection method for the PA oscillation. Align the direction of the detecting light of the laser Doppler vibrometer along y and the magnetic field direction of the EMAT along x, and the PA oscillation was identified. III, detection method for the PA wave. Align the direction of the detecting light of the laser Doppler vibrometer along x and the magnetic field direction of the EMAT along z, and the PA waves were identified. All the PT and PA signals were further evaluated using the Fast Fourier transform method according to their frequency response features.

I, preparation for the nickel laser motor. The laser motor was fabricated using LIGA technology, and the material is pure nickel (99.5%); for the procedures of LIGA technology, see the supplementary material, Fig. S9. II, preparation for the AuNPs. The 100 nm Au nanoparticles (AuNPs) were synthesized using a seed-mediated growth method.49 In the first step, add 6 ml of 1 wt. % sodium citrate (CA) solution into 200 ml of 0.01 wt. % boiling HAuCl4 solution and then 16 nm Au seeds were synthesized. In the second step, add 3 ml of 16 nm Au seeds, 0.6 ml of 1 wt. % ascorbic acid (AA), and 0.2 ml of 1 wt. % CA into 20 ml deionized water and then drop 1.1 ml of 0.01 wt. % HAuCl4 solution into it and then the 50 nm AuNPs were synthesized. The third step is to repeat the above procedures with 2 ml of 50 nm Au seeds, 0.4 ml of AA, 0.1 ml of CA, and 0.654 ml of HAuCl4 solution to synthesize the 100 nm of AuNPs. III, preparation for the AuNPs laser motor. The AuNPs function layer was prepared and anchored onto the laser motor according to the literature,50 and 1.5 ml of cyclohexane was dropped into 6 ml of AuNPs colloids to form an immiscible water/cyclohexane interface to prepare monolayered AuNPs. Then, ethanol was added drop by drop until a specular reflection occurred and a shiny gold color appeared. The nickel laser motor was then immersed in the AuNPs monolayer film at a small angle and slowly pulled out; after the cyclohexane was naturally evaporated, the AuNPs monolayer film was successfully transferred to the surface of the laser motor, and the whole process was demonstrated in the supplementary material, Fig. S10.

Laser motors based on the PT actuation method have the advantage of providing large forces and displacements. However, their dynamic responses are severely restricted by the frequency response limitation. Studies on the frequency response limitation in a PT actuation process show that under high-repetition-frequency laser irradiation, the temperature difference of the laser motor becomes smaller and smaller, and the thermal deformation tends to be zero, which causes a failure of PT actuation. By comparison, the AuNP-PA actuation method performed better in the high-frequency domain because the PA oscillations utilized acoustic energy rather than heat. The results showed that the AuNP-PA laser motor could generate PA oscillations of over 5 µm by pulse laser with repetition frequencies of 1 Hz to 20 kHz, exhibiting a maximum value of 18.8 µm at the mechanical resonant frequency. The variation coefficients of different AuNP-PA laser motors following the same fabrication route were calculated to be less than 8%, confirming the high reproducibility of this kind of PA laser motor. The AuNP-PA laser motor effectively offset the shortcomings of PT actuation in the high-frequency domain, yielding comparable output amplitudes. The fast reaction and large amplitude in the high frequency domain with good stability and high reproducibility demonstrate the effectiveness of this method in solving the frequency response limitation. Such an AuNP-PA laser motor requires no electrical or chemical energy and could benefit bionic research in areas including insect flapping, vocal-cord vibration, and muscular movement. These devices could be used in a wide range of applications in fields including micro/nano physics, biochemistry, and clinical medicine.

See the supplementary material for the complete characteristics and performances of the studied AuNP-PA laser motor and the relative experimental instruments.

Financial support from the National Natural Science Foundation of China (Grant Nos. 52105595, 51972292, and 52375550) and the Natural Science Foundation of Zhejiang Province, China (Grant No. LQ22A040002) are gratefully acknowledged.

The authors have no conflicts to disclose.

Fanghao Li: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Visualization (equal); Writing – original draft (equal). Mengru Zhang: Data curation (equal); Software (equal); Validation (equal). Cuixiang Pei: Data curation (equal); Resources (equal); Validation (equal). Xinyao Yu: Data curation (equal); Software (equal). Li Jiang: Funding acquisition (equal); Resources (equal). Yadong Zhou: Resources (equal); Software (equal); Supervision (equal). Fanli Zhang: Methodology (equal); Validation (equal). Yunfeng Song: Project administration (equal). Jian Chen: Conceptualization (equal); Funding acquisition (equal); Project administration (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

1.
L.
Shao
and
M.
Kall
, “
Light-driven rotation of plasmonic nanomotors
,”
Adv. Funct. Mater.
28
,
1706272
(
2018
).
2.
J.
Lu
,
H.
Yang
,
L.
Zhou
,
Y.
Yang
,
S.
Luo
,
Q.
Li
, and
M.
Qiu
, “
Light-induced pulling and pushing by the synergic effect of optical force and photophoretic force
,”
Phys. Rev. Lett.
118
(
4
),
043601
(
2017
).
3.
J.
Lin
,
A. G.
Hart
, and
Y.-q.
Li
, “
Optical pulling of airborne absorbing particles and smut spores over a meter-scale distance with negative photophoretic force
,”
Appl. Phys. Lett.
106
,
171906
(
2015
).
4.
N.
Eckerskorn
,
R.
Bowman
,
R. A.
Kirian
,
S.
Awel
,
M.
Wiedorn
,
J.
Kupper
,
M. J.
Padgett
,
H. N.
Chapman
, and
A. V.
Rode
, “
Optically induced forces imposed in an optical funnel on a stream of particles in air or vacuum
,”
Phys. Rev. Appl.
4
(
6
),
064001
(
2015
).
5.
R.
Lan
,
J.
Sun
,
C.
Shen
,
R.
Huang
,
Z.
Zhang
,
C.
Ma
,
J.
Bao
,
L.
Zhang
,
L.
Wang
,
D.
Yang
, and
H.
Yang
, “
Light‐driven liquid crystalline networks and soft actuators with degree‐of‐freedom‐controlled molecular motors
,”
Adv. Funct. Mater.
30
(
19
),
2000252
(
2020
).
6.
M.
Pacheco
,
B.
Jurado-Sanchez
, and
A.
Escarpa
, “
Visible‐light‐driven Janus microvehicles in biological media
,”
Angew. Chem., Int. Ed.
58
(
50
),
18017
(
2019
).
7.
D.
Gao
,
W.
Ding
,
M.
Nieto-Vesperinas
,
X.
Ding
,
M.
Rahman
,
T.
Zhang
,
C. T.
Lim
, and
C.-W.
Qiu
, “
Optical manipulation from the microscale to the nanoscale: Fundamentals, advances and prospects
,”
Light: Sci. Appl.
6
,
e17039
(
2017
).
8.
P.
Zemanek
,
G.
Volpe
,
A.
Jonas
, and
O.
Brzobohaty
, “
Perspective on light-induced transport of particles: From optical forces to phoretic motion
,”
Adv. Opt. Photonics
11
,
577
(
2019
).
9.
J.
Lu
,
Q.
Li
,
C.-W.
Qiu
,
Y.
Hong
,
P.
Ghosh
, and
M.
Qiu
, “
Nanoscale lamb wave-driven motors in nonliquid environments
,”
Sci. Adv.
5
(
3
),
eaau8271
(
2019
).
10.
F. H.
Li
,
C. X.
Pei
,
L.
Jiang
, and
S. Z.
Jin
, “
Detaching and moving of adhered particles with a photoacoustic micro-resonator
,”
Appl. Phys. Lett.
114
,
081905
(
2019
).
11.
S.
Palagi
,
A. G.
Mark
,
S. Y.
Reigh
,
K.
Melde
,
T.
Qiu
,
H.
Zeng
,
C.
Parmeggiani
,
D.
Martella
,
A.
Sanchez-Castillo
,
N.
Kapernaum
,
F.
Giesselmann
,
D. S.
Wiersma
,
E.
Lauga
, and
P.
Fischer
, “
Structured light enables biomimetic swimming and versatile locomotion of photoresponsive soft microrobots
,”
Nat. Mater.
15
,
647
(
2016
).
12.
D.
Zhang
,
H.
Zhang
,
C.
Liu
, and
J.
Jiang
, “
Theoretical and experimental study of optothermal expansion and optothermal microactuator
,”
Opt. Express
16
(
17
),
13476
(
2008
).
13.
Z.
Zhang
,
Y.
Wang
,
Q.
You
, and
H.
Zhang
, “
Microscopic research on the properties of optothermal microactuators with different lever ratios
,”
Microsc. Res. Tech.
83
(
5
),
464
(
2020
).
14.
Q.
You
,
Y.
Wang
,
Z.
Zhang
,
H.
Zhang
,
T.
Tsuchiya
, and
O.
Tabata
, “
Laser-driven optothermal microactuator operated in water
,”
Appl. Opt.
59
(
6
),
1627
(
2020
).
15.
B.
Shi
,
H. J.
Zhang
,
B.
Wang
,
F. T.
Yi
,
J. Z.
Jiang
, and
D. X.
Zhang
, “
Dynamic properties of a metal photo-thermal micro-actuator
,”
Appl. Opt.
54
(
6
),
1369
(
2015
).
16.
H.
Ma
,
J.
Hou
,
X.
Wang
,
J.
Zhang
,
Z.
Yuan
,
L.
Xiao
,
Y.
Wei
,
S.
Fan
,
K.
Jiang
, and
K.
Liu
, “
Flexible, all-inorganic actuators based on vanadium dioxide and carbon nanotube bimorphs
,”
Nano Lett.
17
(
1
),
421
(
2017
).
17.
Leeladhar
,
P.
Raturi
, and
J. P.
Singh
, “
Sunlight-driven eco-friendly smart curtain based on infrared responsive graphene oxide-polymer photoactuators
,”
Sci. Rep.
8
,
3687
(
2018
).
18.
Y.
Zhang
and
M. C.
Tan
, “
Programmable light-activated gradient materials based on graphene-polymer composites
,”
Adv. Mater. Interfaces
5
(
7
),
1701374
(
2018
).
19.
Y.
Hu
,
G.
Wu
,
T.
Lan
,
J.
Zhao
,
Y.
Liu
, and
W.
Chen
, “
A graphene-based bimorph structure for design of high performance photoactuators
,”
Adv. Mater.
27
(
47
),
7867
(
2015
).
20.
E.
Wang
,
M. S.
Desai
, and
S.-W.
Lee
, “
Light-controlled graphene-elastin composite hydrogel actuators
,”
Nano Lett.
13
(
6
),
2826
(
2013
).
21.
X.
Zhang
,
Z.
Yu
,
C.
Wang
,
D.
Zarrouk
,
J.-W. T.
Seo
,
J. C.
Cheng
,
A. D.
Buchan
,
K.
Takei
,
Y.
Zhao
,
J. W.
Ager
,
J.
Zhang
,
M.
Hettick
,
M. C.
Hersam
,
A. P.
Pisano
,
R. S.
Fearing
, and
A.
Javey
, “
Photoactuators and motors based on carbon nanotubes with selective chirality distributions
,”
Nat. Commun.
5
,
2983
(
2014
).
22.
X.
Zhang
,
C. L.
Pint
,
M. H.
Lee
,
B. E.
Schubert
,
A.
Jamshidi
,
K.
Takei
,
H.
Ko
,
A.
Gillies
,
R.
Bardhan
,
J.
Urban
,
M.
Wu
,
R.
Fearing
, and
A.
Javey
, “
Optically- and thermally-responsive programmable materials based on carbon nanotube-hydrogel polymer composites
,”
Nano Lett.
11
(
8
),
3239
(
2011
).
23.
J.
Deng
,
J.
Li
,
P.
Chen
,
X.
Fang
,
X.
Sun
,
Y.
Jiang
,
W.
Weng
,
B.
Wang
, and
H.
Peng
, “
Tunable photothermal actuators based on a pre-programmed aligned nanostructure
,”
J. Am. Chem. Soc.
138
(
1
),
225
(
2016
).
24.
H.
Shahsavan
,
A.
Aghakhani
,
H.
Zeng
,
Y.
Guo
,
Z. S.
Davidson
,
A.
Priimagi
, and
M.
Sitti
, “
Bioinspired underwater locomotion of light-driven liquid crystal gels
,”
Proc. Natl. Acad. Sci. U. S. A.
117
(
10
),
5125
(
2020
).
25.
C.
Pei
,
K.
Demachi
,
T.
Fukuchi
,
K.
Koyama
, and
M.
Uesaka
, “
Cracks measurement using fiber-phased array laser ultrasound generation
,”
J. Appl. Phys.
113
(
16
),
163101
(
2013
).
26.
K.
Basak
,
X.
Luís Deán-Ben
,
S.
Gottschalk
,
M.
Reiss
, and
D.
Razansky
, “
Non-invasive determination of murine placental and foetal functional parameters with multispectral optoacoustic tomography
,”
Light: Sci. Appl.
8
,
71
(
2019
).
27.
V. P.
Zharov
, “
Ultrasharp nonlinear photothermal and photoacoustic resonances and holes beyond the spectral limit
,”
Nat. Photonics
5
(
2
),
110
(
2011
).
28.
F. H.
Li
,
C. X.
Pei
,
B.
Shi
,
L. B.
Sun
,
H. J.
Zhang
,
J. Z.
Jiang
, and
D. X.
Zhang
, “
Design of ultrafast laser-driven microactuator based on photoacoustic mechanism
,”
Opt. Express
23
(
16
),
20563
(
2015
).
29.
F.
Li
,
C.
Pei
,
T.
Lang
, and
L.
Jiang
, “
Light-controlled detachment of van der Waals adhered particles in solution with photoacoustic resonator
,”
Opt. Commun.
463
,
125480
(
2020
).
30.
Y.
Wang
,
Q.
Zhang
,
Z.
Zhu
,
F.
Lin
,
J.
Deng
,
G.
Ku
,
S.
Dong
,
S.
Song
,
M. K.
Alam
,
D.
Liu
,
Z.
Wang
, and
J.
Bao
, “
Laser streaming: Turning a laser beam into a flow of liquid
,”
Sci. Adv.
3
(
9
),
e1700555
(
2017
).
31.
B.
Dutton
,
S.
Boonsang
, and
R. J.
Dewhurst
, “
A new magnetic configuration for a small in-plane electromagnetic acoustic transducer applied to laser-ultrasound measurements: Modelling and validation
,”
Sens. Actuators, A
125
,
249
(
2006
).
32.
X.
Chen
,
Y.
Chen
,
M.
Yan
, and
M.
Qiu
, “
Nanosecond photothermal effects in plasmonic nanostructures
,”
ACS Nano
6
(
3
),
2550
(
2012
).
33.
R.
Hickey
,
D.
Sameoto
,
T.
Hubbard
, and
M.
Kujath
, “
Time and frequency response of two-arm micromachined thermal actuators
,”
J. Manuf. Syst.
13
,
40
(
2003
).
34.
R. M.
White
, “
Generation of elastic waves by transient surface heating
,”
J. Appl. Phys.
34
(
12
),
3559
(
1963
).
35.
C. B.
Scruby
,
R. J.
Dewhurst
,
D. A.
Hutchins
, and
S. B.
Palmer
, “
Quantitative studies of thermally generated elastic waves in laser-irradiated metals
,”
J. Appl. Phys.
51
(
12
),
6210
(
1980
).
36.
N.
Wu
,
Y.
Tian
,
X.
Zou
,
V.
Silva
,
A.
Chery
, and
X.
Wang
, “
High-efficiency optical ultrasound generation using one-pot synthesized polydimethylsiloxane-gold nanoparticle nanocomposite
,”
J. Opt. Soc. Am. B
29
(
8
),
2016
(
2012
).
37.
S. K.
Ghosh
and
T.
Pal
, “
Interparticle coupling effect on the surface plasmon resonance of gold nanoparticles: From theory to applications
,”
Chem. Rev.
107
,
4797
(
2007
).
38.
X.
Zou
,
N.
Wu
,
Y.
Tian
, and
X.
Wang
, “
Broadband miniature fiber optic ultrasound generator
,”
Opt. Express
22
(
15
),
18119
(
2014
).
39.
Y.
Xu
,
X.
Zhao
,
A.
Li
,
Y.
Yue
,
J.
Jiang
, and
X.
Zhang
, “
Plasmonic heating induced by Au nanoparticles for quasi-ballistic thermal transport in multi-walled carbon nanotubes
,”
Nanoscale
11
,
7572
(
2019
).
40.
L.
Landau
,
L.
Pitaevskii
,
A.
Kosevich
, and
E.
Lifshitz
,
Theory of Elasticity
(
Butterworth-Heinemann
,
Oxford, England
,
1984
).
41.
A. A.
Kolomenskii
,
H. A.
Schuessler
,
V. G.
Mikhalevich
, and
A. A.
Maznev
, “
Interaction of laser-generated surface acoustic pulses with fine particles: Surface cleaning and adhesion studies
,”
J. Appl. Phys.
84
(
5
),
2404
(
1998
).
42.
S.
Chen
,
L.-Y.
Meng
,
H.-Y.
Shan
,
J.-F.
Li
,
L.
Qian
,
C. T.
Williams
,
Z.-L.
Yang
, and
Z.-Q.
Tian
, “
How to light special hot spots in multiparticle-film configurations
,”
ACS Nano
10
(
1
),
581
(
2016
).
43.
J. M.
Pitarke
,
V. M.
Silkin
,
E. V.
Chulkov
, and
P. M.
Echenique
, “
Theory of surface plasmons and surface-plasmon polaritons
,”
Rep. Prog. Phys.
70
,
1
(
2007
).
44.
S.
Chen
,
Y.
Zhang
,
T.-M.
Shih
,
W.
Yang
,
S.
Hu
,
X.
Hu
,
J.
Li
,
B.
Ren
,
B.
Mao
,
Z.
Yang
, and
Z.
Tian
, “
Plasmon-induced magnetic resonance enhanced Raman spectroscopy
,”
Nano Lett.
18
,
2209
(
2018
).
45.
A.
Zerda
,
J.-W.
Kim
,
E. I.
Galanzha
,
S. S.
Gambhir
, and
V. P.
Zharov
, “
Advanced contrast nanoagents for photoacoustic molecular imaging, cytometry, blood test and photothermal theranostics
,”
Contrast Media Mol. Imaging
6
,
346
(
2011
).
46.
G.
Baffou
and
H.
Rigneault
, “
Femtosecond-pulsed optical heating of gold nanoparticles
,”
Phys. Rev. B
84
,
035415
(
2011
).
47.
O. B.
Wright
and
V. E.
Gusev
, “
Ultrafast generation of acoustic waves in copper
,”
IEEE Trans. Ultrason., Ferroelectr. Freq. Control
42
(
3
),
331
(
1995
).
48.
K. K.
Nanda
,
S. N.
Sahu
, and
S. N.
Behera
, “
Liquid-drop model for the size-dependent melting of low-dimensional systems
,”
Phys. Rev. A
66
,
013208
(
2002
).
49.
J. F.
Li
,
X. D.
Tian
,
S. B.
Li
,
J. R.
Anema
,
Z. L.
Yang
,
Y. D.
Ding
,
Y. F.
Wu
,
Y. M.
Zeng
,
Q. Z.
Chen
,
B. R.
Ren
,
Z. L.
Wang
, and
Z. Q.
Tian
, “
Surface analysis using shell-isolated nanoparticle-enhanced Raman spectroscopy
,”
Nat. Protoc.
8
,
52
(
2013
).
50.
T.
Xie
,
Z.
Cao
,
Y.
Li
,
Z.
Li
,
F.-L.
Zhang
,
Y.
Gu
,
C.
Han
,
G.
Yang
, and
L.
Qu
, “
Highly sensitive SERS substrates with multi-hot spots for on-site detection of pesticide residues
,”
Food Chem.
381
,
132208
(
2022
).

Supplementary Material