Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Light-induced micro-vibrator with controllable amplitude and frequency

Open Access Open Access

Abstract

We propose and demonstrate a light-induced micro-vibrator that can perform an adjustable reciprocating vibration based on the Δα-typed photophoretic force. The vibration amplitudes and periods can be precisely controlled and modulated in real-time, and the maximum average restoring speed is as high as 23.26 μm/s. In addition, by using the self-healing properties of the Bessel-like beam, we achieve the simultaneous driving and modulating of three absorbing micro-vibrators. The proposed absorbing micro-vibrator can be used as a novel light-driven micromotor, which is considered to have potential application value in the field of targeted drug delivery, biosensing, and environmental detection.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Optical manipulation is a technology that uses optical force to trap and drive particles in the fluid environment. It has potential application values in many fields, such as particle trapping [13], particle sorting and transporting [49], condensed matter physics [10], biosensing and cell analysis [1115], micropump [16,17], and light-driven micromotors [1820]. In particular, the light-driven micromotors have received extensive attention from researchers, as they have potential application values in the biomedical field [19,20]. However, it is still a challenge to design stable and adjustable light-driven micromotors.

Optical force, which results from the energy and momentum exchange between object and photon, can be used as the driving source of light-driven micromotors. When a laser beam illuminates particles, the particle will represent two types of photodynamic phenomena. The first one is optical radiation pressure, which results from the momentum transfer directly between photons and particles [1]. In the past three decades, radiation pressure plays an important role in optical trapping [13] and has been creatively designed as the driving source of laser-induced rotary micromotors [18,2026]. However, unfortunately, those laser-induced rotary manipulations are not satisfactory.

On the one hand, these micromotors rely on complex micromachining methods [1420] or are driven by a complex optical system [2124]. On the other hand, the light-to-work conversion ratio is too small, being in the order of 10−15-10−12. Finally, it is worth noting that there are few reports about laser-induced particle vibration. Although Zhao reports a harmonic oscillation manipulation based on optical radiation pressure [27], it is difficult to control the particle vibration state.

The second photodynamic phenomenon, also called the photothermal effect [28] (light-induced heat), is a kind most efficient method to improve the light-to-work conversion for optical manipulation [29]. When light illuminates the particles of photothermal materials in the fluid environment, two types of photophoretic forces exert on the particles [30]. The first one is ΔT-typed photophoretic forces, resulting from the temperature gradient of the surface of the particles. The large temperature gradient distributions provide a strong driving force (in the 10−7 N) order for the particles to overcome the viscous resistance and perform the motion. According to the feature of ΔT-typed photophoretic force, researchers have designed linear transmission micromotors [3134] and rotary micromotors [29,3537]. However, considering the relationship between ΔT-typed photophoretic forces and temperature gradient direction, it is still challenging to manipulate particles to perform harmonic-like vibrations only depending on a single beam laser.

The second photophoretic force is the Δα-typed photophoretic force generated by the difference in heat exchange between the particle surface and the fluid molecules. The direction of Δα-typed photophoretic forces only depends on the accommodation coefficient (α), which represents the degree of heat exchange between particles and surrounding liquid molecules. Therefore, it is necessary to construct two opposite forces to perform the particle vibration. For instance, Shi proposes a method to realize the vibration of nanoparticles [8]. It depends on the dynamic adjustment of light radiation pressure and fluid drag force. These methods show that it is tough to realize the vibration of particles only by using the radiation pressure produced by a single beam of laser. A few reports about the transport and trap of particles are induced by the Δα-typed photophoretic forces in the gas environment [38]. Our previous studies also show that the Δα-typed photophoretic force is related to the incident laser power exerted on the particle surface [35,3942], which indicates that we may perform the modulable motion of a particle by modulating the incident laser power. We have realized the laser-induced hammer-hit vibration of absorbing particles in pure glycerol [40]. However, the vibration amplitude of the absorbing micro-vibrator is only 7.24 μm under the irradiation of the Gaussian beam. Therefore, it is challenging to realize long-distance directional transportation with vibration state adjustment.

This paper proposes and demonstrates a method to perform the controllable reciprocating (harmonic-like) vibration of an absorbing micro-vibrator based on the Δα-typed photophoretic force. In this work, we use an arbitrary waveform generator to precisely adjust the incident laser power, which will modulate the motion (moving speed and direction) of the micromotor. We may perform the modulation of the absorbing vibrator moving with the same amplitude and different frequency (see Visualization 1) or different amplitude and same frequency (see Visualization 2). In addition, we may also perform the simultaneous driving of three vibrators based on the self-healing properties of the Bessel-like beam (see Visualization 3). The proposed method does not require complex micromachining technology, provides a new idea for recycling the micromotors and bidirectional targeted drug transportation, and even provides piston-like mechanical kinetic energy.

2. Principles and methods

The absorbing vibrator (AV) we used in the experiment is a black sphere with a diameter of 6 μm, which is made of carbon black and SiO2(the mass ratio of carbon is 15% and the volume ratio is 10%). When a laser with a constant power irradiates on the AV in the pure glycerin, the motion of the AV is determined by the Δα-typed photophoretic force FΔα (the radiation pressure Frp and the ΔT-typed photophoretic forces FΔT can be ignored for the FΔα) [40,41], which is caused by the difference between the accommodation coefficient α on both sides of the particle. For simplicity, we may obtain the expression of FΔα along the X-axis and Z-axis (see the inset illustration in Fig. 1(a)), respectively. The expression of FΔα along the X-axis and Z-axis are [35]:

$$\begin{aligned}{{\boldsymbol F}_{\Delta \alpha - x}}&=\frac{{\pi \eta _g^2\sigma I}}{{32{\rho _g}{\kappa _g}{R_{AV}}{T_\infty }}}\frac{{\Delta {\alpha _x}}}{{{{\bar{\alpha }}_x}}}{{\hat{{\boldsymbol e}}}_x}\\ \\ {{\boldsymbol F}_{\Delta \alpha - z}}&=\frac{{\pi \eta _g^2\sigma I}}{{32{\rho _g}{\kappa _g}{R_{AV}}{T_\infty }}}\frac{{\Delta {\alpha _z}}}{{{{\bar{\alpha }}_z}}}{{\hat{{\boldsymbol e}}}_z} \end{aligned}$$
where σ is the absorbing cross-section of the AV, T is the room temperature (T=298 K), ηg is the dynamic viscosity coefficient of pure glycerol (ηg = 1.5 Pa·s), ρg is the density of pure glycerol (ρg = 1.261×103 kg∕m3), RAV is the radius of the vibrator (RAV = 3 μm), Kg is the thermal conductivity of the glycerol (Kg=0.286 W/(mK)), I {I=2P/π[ω(Z)]2, is expressed as the intensity on the surface of the particle, P is the incident laser power, and ω(Z) is the beam radius along the Z axis}. ᾱx and ᾱz are the average accommodation coefficients along the X-axis and Z-axis, respectively. The accommodation coefficient α can be defined as [40]:
$$\alpha =\frac{{{T_r} - {T_i}}}{{{T_s} - {T_i}}}$$

 figure: Fig. 1.

Fig. 1. (a) The schematic diagram of the AV driven by the Δα-type photophoretic force FΔα; The larger the FΔα the smaller the distance between the light source and the particle. (b) Simulated results of the light field distribution near the fiber probe (Z=7.5 μm, P=60 mW); (c) simulated result of temperature field distribution introduced by the light field in (b). (d) The calculated result of the X-axis FΔα-x exerting on the AV with different incident laser power. (e) The calculated result of the Z-axis FΔα-z exerting on the AV with different incident laser power. (f) Experimental and simulated results of the relationship between the incident laser power and trapping positions of the AV (here we define the coordinate origin as the center of the fiber end-face).

Download Full Size | PDF

where Ti and Tr are the incident and leaving temperature of the fluid molecules on the particle's surface, respectively, Ts is the average temperature of the particle surface. Here we divide the surface of the AV into two hemispheres σ1-z and σ2-z (see Fig. 1(a), where σ1-z refers to the illuminated hemisphere, σ2-z refers to the non-illuminated hemisphere). As a result, we may obtain the accommodation coefficient α1-z and α2-z (see the inset illustration in Fig. 1(a)), α1-z represents the accommodation coefficient of the illuminated surface of AV, α2-z represents the accommodation coefficient of the dark surface of AV, and Δαz1-z−α2-z and ᾱ=(α1-z2-z)/2. Therefore, the direction of the FΔα-z is determined by Δαz/ᾱz.

When the momentum diffusivity between the particle and the glycerol in the illuminated surface is larger than the dark surface (α1-z2-z), the FΔα-z will push the particle moving away from the fiber tip; and when the momentum diffusivity in the illuminated surface is smaller than the dark surface (α1-z2-z), the FΔα-z will pull the particle moving towards the fiber tip. The incident laser power determines the accommodation coefficient (α1-z and α2-z). Similarly, we can analyze the FΔα-x along the X-axis. We employ the finite-element analysis method to simulate the light field distribution near the fiber tip (see Fig. 1(b), the particle position Z=7.5 μm (the distance between the particle and the end face of the fiber) and the incident laser power P=60 mW) and then we obtain the temperature field distribution (see Fig. 1(c)) introduced by the light field. We may extract the temperature Ti, Tr, and Ts from the simulated results in Fig. 1(c) to calculate the accommodation coefficients.

According to Eq. (1) and Eq. (2), we may calculate the photophoretic force FΔα-x along the X-axis (see Fig. 1(d)) and the photophoretic force FΔα-z along the Z-axis (see Fig. 1(e)) with different incident laser powers (the incident laser power Pic changes from 30 mW to 60 mW). The results indicate that the FΔα-x ensures the AV to be trapped on the main axis(X=0) of the fiber probe and performs different stable trapping positions Z(P) of AV in different incident laser power Pi. The larger the Pi, the smaller the Z(P) (see the upper right corner in Fig. 1(a)). The trapping properties allow us to adjust the motion of AV (moving direction and position) by adjusting the incident laser power.

When we employ a laser beam with a constant power Pi0 to irradiate the AV, the accommodation coefficient α1-z is equal to α2-z, which means the FΔα-z=0, and the AV will be stable in a certain equilibrium position Zi0 in the light field. Then we increase the incident power from Pi0 to Pi1 (Pi1 > Pi0), the AV will move from Zi0 to Zi1 and Zi1< Zi0. The FΔα-z pulls the AV moving towards the laser source. Conversely, if we reduce the incident power from Pi1 to Pi0, the AV will move from Zi1 to Zi0. The FΔα-z pushes the AV moving away from the laser source and goes back to the initial position Z0. The simulated parameters include the density and the complex refractive index of AV are ρ=2.27×103 kg/m3, n=1.908 + 0.519i, and the thermal conductivity of the AV κAV = 3.758 W/(mK).

Calculated results indicate that when the incident power Pic changes from 30 mW to 60 mW, the equilibrium position Zic of AV changes from 108 μm to 8 μm (see Fig. 1(e), here we define the coordinate origin as the center of the fiber end-face). We obtain the fitting equation as Zic=(-3.24×Pic+202.82) [μm] (see the blue line in Fig. 1(f)). We also calibrate the relationship between the position and laser with the experimental methods (see the red point in Fig. 1(f)). In the practical experiment, we increase the incident power Pie from 31 mW to 58 mW, and the trapping position Zie of AV changes from 110.4 μm to 8.7 μm. The calculated results are in good agreement with the experimental results.

The above results indicate that the incident power has a linear relationship with the equilibrium position of AV. It means that when the incident power changes periodically, the AV will move back and forth along the beam’s main axis correspondingly.

3. Experiment and result

3.1 Experimental setup

Figure 2(a) shows the schematic diagram of the experimental setup. We use an arbitrary waveform generator to adjust the period and amplitude of the incident power and employ a bayonet nut connector (BNC) to connect an arbitrary waveform generator (DG1000, Rigol Technologies) and the 980-nm laser source (CLD101x, Thorlabs). Next, we fabricate the all-fiber probe (see Fig. 2(b)) by coaxial welding a single-mode fiber (ClearLite98) and a multimode fiber (Nufe MM-S10).

 figure: Fig. 2.

Fig. 2. (a) The schematic diagram of the experimental setup. (b) Image of the all-fiber probe. Here, SMF means the single-mode fiber, and MMF means the multimode fiber. (c) Profile image of the Bessel-like beam generated in the MMF (produced by a 532-nm laser, the length of the multimode fiber is 1.6373 mm).

Download Full Size | PDF

The length of the multimode fiber is 1.6373 mm (see Fig. 2(b)), and the length of the single-mode fiber is approximately 1 m. The core and cladding refractive index of the single-mode fiber is 1.4661 and 1.4573, respectively. The core and cladding refractive index of the multimode fiber is 1.4446 and 1.4277, respectively. The core diameter of the single-mode fiber is 4.4 μm, and the core diameter of the multimode fiber is 105 μm. The Bessel-like beam will be excited in the multimode fiber when the fundamental Gaussian beam enters from the single-mode fiber (see Fig. 2(c)). We fabricate the sample solution by mixing the AVs into pure glycerin solution. The pure glycerin has almost no convection or Laminar flow in the vicinity of heated AV. Compared with the heat transfer, the momentum diffusivity between the AV and the glycerol molecules plays the most important role, and the Δα-typed photophoretic force will dominate the motion of the AV. We use a 3-dimensional micro-manipulation controller to modulate the position of the fiber probe. We employ an objective lens (20×, numerical aperture of 0.4) and a CCD camera (30 frame/s) to monitor the motion of the AV. We employ the image processing method to analyze the trajectory of AV.

3.2 Modulation of vibrating period τv

The vibrating period τv is determined by the driving period τi of the incident laser power. If we want to perform the vibration of AV with the same vibrating amplitude Ω with different vibrating period τv, we should adjust the power P0 and the period τi of the incident laser correspondingly. Here we perform an example of a series of vibrations with the same vibrating amplitudes (Ω=20.68 ± 0.74 μm) and different vibrating periods: τv1=4 s, τv2=6 s, τv3=8 s, τv4=10 s and τv5=12 s, respectively (see Visualization 1 and see Fig. 3(a)). We use the initial incident power P0=50mW to trap the AV at Zi1=36 μm, and then we employ the arbitrary waveform generator to set the period τi of incident laser power as τi1=4 s with the power as P01=40mW, τi2=6 s with the P02=24mW, τi3=8 s with the P03=20mW, τi4=10 s with the P04=16mW and τi5=12 s with the P05=12mW, respectively. We measure the average restoring speed of AV in a period (see Fig. 3(b)), which are v1=21.18 μm/s, v2=13.41 μm/s, v3=11.01 μm/s, v4=8.37 μm/s and v5=7.22 μm/s, respectively (see Fig. 3(c)). We may perform the modulation of vibrating period τv with the vibrating amplitude unchanged.

 figure: Fig. 3.

Fig. 3. (a) Continuous-time video screenshot of the AV with different vibrating periods with the time interval of 1 s; (b) vibrating trajectory of the AV with different vibrating periods; (c) average restoring speed of the AV with different vibrating periods (Visualization 1).

Download Full Size | PDF

3.3 Modulation of vibrating amplitude Ω

The vibrating amplitude Ω is determined by the change of the incident laser power. Therefore, if we want to perform the vibration with the same vibrating period τv and different vibrating amplitude Ω, we should adjust the amplitude and period of the incident laser correspondingly.

Here we perform an example of a series of vibration with the same vibrating period (τv=5 s) with different vibrating amplitudes: Ω1=6.13 μm, Ω2=9.69 μm, Ω3=16.33 μm, Ω4=23.92 μm, and Ω5=29.07 μm, respectively (see Visualization 2 and see Fig. 4(b)). Originally, we use the initial incident power P01=56 mW to trap the AV at Zi1=17.45 μm. Then, within 5 s (one vibrating period), we modulate the incident laser power from (56 + 8) mW to (56-8) mW (here the change of laser power PΔ is 8 mW), correspondingly the AV moves from the Zmin=14.99 μm to Zmax=27.26 μm, which means the vibrating amplitude of AV is 27.26-14.99 = 6.13 μm. Similarly, we set the change of incident power PΔ as PΔ2=16 mW, PΔ3=24 mW, PΔ4=32 mW, and PΔ5=40 mW, respectively, we perform the modulation of the vibrating amplitude as Ω2=9.69 μm, Ω3=16.33 μm, Ω4=23.92 μm, and Ω5=29.07 μm, respectively (see Visualization 2 and Fig. 4(c)). We measure the average restoring speeds of AV, which are v1=4.91 μm/s, v2=7.75 μm/s, v3=13.06 μm/s, v4=19.14 μm/s and v5=23.26 μm/s, respectively (see Fig. 4(c)). We obtain the fitting equation as v=(0.60×PΔ-0.80) μm/s.

 figure: Fig. 4.

Fig. 4. (a) Continuous-time video screenshot of the AV with different vibrating amplitudes with the time interval of 1s; (b) vibrating trajectory of the AV with different vibrating amplitudes; (c) average restoring speed of the AV with different vibrating amplitudes (Visualization 2).

Download Full Size | PDF

3.4 Driving of multiple AVs

The Bessel-like beam is a beam with self-healing and non-diffraction characteristics. Therefore, we employ the Bessel-like beam to achieve simultaneous driving of multiple AVs (see Visualization 3).

We use the initial incident power Pi=50mW to trap the first vibrator AV1 at Zi1=39.26 μm, the second vibrator AV2 at Zi2=195.35 μm, and the third vibrator AV3 at Zi3=467.52 μm (see Fig. 5(a)). We set the period of incident power τv=5 s and the change of laser power PΔ=24mW. The vibration amplitude of AV1 is Ω1=15.54 μm (see Fig. 5(b)), and the average restoring speed of AV1 is v1=12.43 μm/s, the vibration amplitude of AV2 is Ω2=14.21 μm (see Fig. 5(c)), and the average restoring speed is v2=11.37 μm/s, the vibration amplitude of AV3 is Ω3=16.95 μm (see Fig. 5(d)), and the average restoring speed v3=13.56 μm/s. In addition, the results in Fig. 5 indicate that three AVs move with the same period and amplitude, which are 5 s and ∼15 μm, respectively.

 figure: Fig. 5.

Fig. 5. (a)Continuous-time video screenshot of three AVs with PΔ=24 mW and τv=5 s; (b-d)vibrating trajectory of the 1st AV, the 2nd AV, and the 3rd AV, respectively (Visualization 3).

Download Full Size | PDF

4. Conclusions

In conclusion, by using the Δα-typed photophoretic force FΔα, we achieve the bidirectional vibration of an absorbing vibrator in the liquid glycerol. We may control the vibrating period and amplitude of the vibrator precisely by modulating the incident laser power. In addition, based on the self-healing and non-diffraction property of Bessel-like beams, we perform the simultaneous driving of three AVs. The proposed method of bidirectional micromotors is considered to have potential application value in targeted drug delivery, biosensing, and environmental detection.

Funding

National Key Research and Development Program of China (No.2018YFC1503703); National Natural Science Foundation of China (No.61775047, No.61975039); Natural Science Foundation of Heilongjiang Province (YQ2020F011); 111 Project (B13015); Fundamental Research Funds for Harbin Engineering University of China.

Acknowledgments

We thank the College of Physics and Optoelectronic Engineering of Harbin Engineering University for providing technical support under the Key Lab of In-fiber Integrated Optics, Ministry Education of China.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. A. Ashkin and, “Acceleration and trapping of particles by radiation pressure,” Phys. Rev. Lett. 24(4), 156–159 (1970). [CrossRef]  

2. A. Ashkin and J. Dziedzic, “Optical trapping and manipulation of viruses and bacteria,” Science 235(4795), 1517–1520 (1987). [CrossRef]  

3. A. Ashkin, J. M. Dziedzic, J. E. Bjorkholm, and S. Chu, “Observation of a single-beam gradient force optical trap for dielectric particles,” Opt. Lett. 11(5), 288 (1986). [CrossRef]  

4. M. P. Macdonald, G. C. Spalding, and K. Dholakia, “Microfluidic sorting in an optical lattice,” Nature 426(6965), 421–424 (2003). [CrossRef]  

5. M. M. Wang, E. Tu, D. E. Raymond, J. M. Yang, and W. F. Butler, “Microfluidic sorting of mammalian cells by optical force switching,” Nat. Biotechnol. 23(1), 83–87 (2005). [CrossRef]  

6. E. Almaas and I. Brevik, “Possible sorting mechanism for microparticles in an evanescent field,” Phys. Rev. A 87(6), 063826 (2013). [CrossRef]  

7. K. Setoura, T. Tsuji, S. Ito, S. Kawano, and H. Miyasaka, “Opto-thermophoretic separation and trapping of plasmonic nanoparticles,” Nanoscale 11(44), 1517 (2019). [CrossRef]  

8. Y. Shi, S. Xiong, L. K. Chin, J. Zhang, W. Ser, J. Wu, T. Chen, Z. Yang, Y. Hao, B. Liedberg, P. H. Yap, D. P. Tsai, C. Qiu, and A. Liu, “Nanometer-precision linear sorting with synchronized optofluidic dual barriers,” Sci. Adv. 4(1), eaao0773 (2018). [CrossRef]  

9. Y. Shi, T. Zhu, T. Zhang, A. Mazzulla, D. Tsai, W. Ding, A. Liu, G. Cipparrone, J. J. Sáenz, and C. Qiu, “Chirality-assisted lateral momentum transfer for bidirectional enantioselective separation,” Light. Sci. Appl 9(1), 62 (2020). [CrossRef]  

10. S. Chu, J. E. Bjorkholm, A. Ashkin, and A. E. Cable, “Experimental observation of optically trapped atoms,” Phys. Rev. Lett. 57(3), 314–317 (1986). [CrossRef]  

11. K. Uhrig, R. Kurre, C. Schmitz, J. E. Curtis, T. Haraszti, E. M. Clemen, and J. P. Spatz, “Optical force sensor array in a microfluidic device based on holographic optical tweezers,” Lab Chip 9(5), 661–668 (2009). [CrossRef]  

12. S. Mohanty, “Optically-actuated translational and rotational motion at the microscale for microfluidic manipulation and characterization,” Lab Chip 12(19), 3624–3636 (2012). [CrossRef]  

13. P. R. Sevilla, L. Labrador-Páez, D. J. Garcia, and P. H. Gonzalez, “Optical trapping for biosensing: materials and applications,” J. Mater. Chem. B 10, C7TB01921A (2017). [CrossRef]  

14. H. Xin, Y. Li, D. Xu, Y. Zhang, C. Chen, and B. Li, “Single Upconversion Nanoparticle–Bacterium Cotrapping for Single-Bacterium Labeling and Analysis,” Small 13(14), 1603418 (2017). [CrossRef]  

15. H. Xin, Y. Liu, Y. Zhang, Y. Xiao, and B. Li, “Optical Forces: From Fundamental to Biological Applications,” Adv. Matt. 32(37), 2001994 (2020). [CrossRef]  

16. S. Maruo and H. Inoue, “Optically driven micropump produced by three-dimensional two-photon microfabrication,” Appl. Phys. Lett. 89(14), 144101 (2006). [CrossRef]  

17. M. Shoji, T. Akira, and S. Yohei, “Optically driven micropump with a twin spiral microrotor,” Opt. Express 17(21), 18525–32 (2009). [CrossRef]  

18. U. G. Būtaitė, G. M. Gibson, Y. L. D. Ho, M. Taverne, J. M. Taylor, and D. B. Phillips, “Indirect optical trapping using light driven micro-rotors for reconfigurable hydrodynamic manipulation,” Nat. Commun. 10(1), 1215 (2019). [CrossRef]  

19. D. Li, C. Liu, Y. Yang, L. Wang, and Y. Shen, “Micro-rocket robot with all-optic actuating and tracking in blood,” Light. Sci. Appl 9(1), 84 (2020). [CrossRef]  

20. H. Xin, N. Zhao, Y. Wang, X. Zhao, T. Pan, Y. Shi, and B. Li, “Optically Controlled Living Micromotors for the Manipulation and Disruption of Biological Targets,” Nano Lett. 20(10), 7177–7185 (2020). [CrossRef]  

21. Y. Zong, J. Liu, R. Liu, H. Guo, M. Yang, Z. Li, and K. Chen, “An optically driven bistable Janus rotor with patterned metal coatings,” ACS Nano 9(11), 10844–10851 (2015). [CrossRef]  

22. J. Liu, H. L. Guo, and Z. Li, “Self-propelled round-trip motion of Janus particles in static line optical tweezers,” Nanoscale 8(47), 19894–19900 (2016). [CrossRef]  

23. M. Friese, T. A. Nieminen, N. R. Heckenberg, and H. Rubinsztein-Dunlop, “Optical alignment and spinning of laser-trapped microscopic particles,” Nature 394(6691), 348–350 (1998). [CrossRef]  

24. M. Padgett and R. Bowman, “Tweezers with a twist,” Nat. Photonics 5(6), 343–348 (2011). [CrossRef]  

25. K. Y. Bliokh, F. J. Rodriguez-Fortuno, F. Nori, and A. V. Zayats, “Spin-orbit interactions of light,” Nat. Photonics 9(12), 796–808 (2015). [CrossRef]  

26. S. Tsesses, K. Cohen, E. Ostrovsky, B. Gjonaj, and G. Bartal, “Spin-orbit interaction of light in plasmonic lattices,” Nano Lett. 19(6), 4010–4016 (2019). [CrossRef]  

27. H. Zhao, G. Farrell, P. Wang, and L. Yuan, “Investigation of particle harmonic oscillation using four-core fiber integrated twin-tweezers,” IEEE Photonics Technol. Lett. 28, 461 (2016). [CrossRef]  

28. H. Li, Y. Cao, L. Zhou, X. Xu, T. Zhu, Y. Shi, C. Qiu, and W. Ding, “Optical pulling forces and their applications,” Adv. Opt. Photonics 12(2), 288–366 (2020). [CrossRef]  

29. C. Maggi, F. Saglimbeni, M. Dipalo, F. D. Angelis, and R. D. Leonardo, “Micromotors with asymmetric shape that efficiently convert light into work by thermocapillary effects,” Nat. Commun. 6(1), 7855 (2015). [CrossRef]  

30. V. G. Shvedov, A. V. Rode, Y. V. Izdebskaya, A. S. Desyatnikov, W. Krolikowski, and Y. S. Kivshar, “Giant optical manipulation,” Phys. Rev. Lett. 105(11), 118103 (2010). [CrossRef]  

31. 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). [CrossRef]  

32. Z. Wu, T. Si, W. Gao, X. Lin, J. Wang, and Q. He, “Superfast near-infrared light-driven polymer multilayer rockets,” Small 12(5), 577–582 (2016). [CrossRef]  

33. M. Xuan, Z. Wu, J. Shao, L. Dai, T. Si, and Q. He, “Near infrared light-powered janus mesoporous silica nanoparticle motors,” J. Am. Chem. Soc. 138(20), 6492–6497 (2016). [CrossRef]  

34. T. Luan, F. Meng, P. Tao, W. Shang, J. Wu, C. Song, and T. Deng, “Bubble-enabled underwater motion of a light-driven motor,” Small 15(12), 1804959 (2019). [CrossRef]  

35. Y Zhang, S Lin, Z Liu, Y Zhang, J Zhang, J Yang, and L Yuan, “Laser-induced rotary micromotor with high energy conversion efficiency,” Photonics Res. 8(4), 534 (2020). [CrossRef]  

36. X. Peng, Z. Chen, P. S. Kollipara, Y. Liu, and Y. Zheng, “Opto-thermoelectric microswimmers,” Light. Sci. Appl 9(1), 141 (2020). [CrossRef]  

37. A. P. Porfirev, A. B. Dubman, and D. P. Porfiriev, “Demonstration of a simple technique for controllable orbital rotation of light-absorbing particles in air,” Opt. Lett. 45(6), 1475–1478 (2020). [CrossRef]  

38. G. H. Chen, L. He, M. Y. Wu, and Y. Q. Li, “Temporal dependence of photophoretic force optically induced on absorbing airborne particles by a power-modulated laser,” Phys. Rev. Appl. 10(5), 054027 (2018). [CrossRef]  

39. Z. Liu, J. Wu, Y. Zhang, Y. Zhang, and L. Yuan, “Optical trapping and axial shifting for strongly absorbing particle with single focused TEM00 gaussian beam,” Appl. Phys. Lett. 113(9), 091101 (2018). [CrossRef]  

40. Y. Zhang, Y. Zhang, Z. Liu, X. Tang, X. Yang, J. Zhang, J. Yang, and L. Yuan, “Laser-induced microsphere hammer-hit vibration in liquid,” Phys. Rev. Lett. 121(13), 133901 (2018). [CrossRef]  

41. Y. Zhang, X. Tang, Y. Zhang, Z. Liu, X. Yang, J. Zhang, J. Yang, and L. Yuan, “Optical attraction of strongly absorbing particles in liquids,” Opt. Express 27(9), 12414 (2019). [CrossRef]  

42. Z. Liu, L. Wang, Y. Zhang, S. Lin, Y. Zhang, X. Yang, J. Zhang, J. Yang, and L. Yuan, “X-typed curvilinear transport of strongly absorbing particle in a dual-beam fiber optical trap,” Opt. Express 27(23), 33967 (2019). [CrossRef]  

Supplementary Material (3)

NameDescription
Visualization 1       In this work, we use an arbitrary waveform generator to precisely adjust the incident laser power, which will modulate the motion (moving speed and direction) of the micromotor. We may perform the modulation of the absorbing vibrator moving with the
Visualization 2       In this work, we use an arbitrary waveform generator to precisely adjust the incident laser power, which will modulate the motion (moving speed and direction) of the micromotor.We perform the modulation of the absorbing vibrator moving with the diffe
Visualization 3       In this work, we use an arbitrary waveform generator to precisely adjust the incident laser power, which will modulate the motion (moving speed and direction) of the micromotor.We also perform the simultaneous driving of three vibrators based on the

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. (a) The schematic diagram of the AV driven by the Δα-type photophoretic force FΔα; The larger the FΔα the smaller the distance between the light source and the particle. (b) Simulated results of the light field distribution near the fiber probe (Z=7.5 μm, P=60 mW); (c) simulated result of temperature field distribution introduced by the light field in (b). (d) The calculated result of the X-axis FΔα-x exerting on the AV with different incident laser power. (e) The calculated result of the Z-axis FΔα-z exerting on the AV with different incident laser power. (f) Experimental and simulated results of the relationship between the incident laser power and trapping positions of the AV (here we define the coordinate origin as the center of the fiber end-face).
Fig. 2.
Fig. 2. (a) The schematic diagram of the experimental setup. (b) Image of the all-fiber probe. Here, SMF means the single-mode fiber, and MMF means the multimode fiber. (c) Profile image of the Bessel-like beam generated in the MMF (produced by a 532-nm laser, the length of the multimode fiber is 1.6373 mm).
Fig. 3.
Fig. 3. (a) Continuous-time video screenshot of the AV with different vibrating periods with the time interval of 1 s; (b) vibrating trajectory of the AV with different vibrating periods; (c) average restoring speed of the AV with different vibrating periods (Visualization 1).
Fig. 4.
Fig. 4. (a) Continuous-time video screenshot of the AV with different vibrating amplitudes with the time interval of 1s; (b) vibrating trajectory of the AV with different vibrating amplitudes; (c) average restoring speed of the AV with different vibrating amplitudes (Visualization 2).
Fig. 5.
Fig. 5. (a)Continuous-time video screenshot of three AVs with PΔ=24 mW and τv=5 s; (b-d)vibrating trajectory of the 1st AV, the 2nd AV, and the 3rd AV, respectively (Visualization 3).

Equations (2)

Equations on this page are rendered with MathJax. Learn more.

F Δ α x = π η g 2 σ I 32 ρ g κ g R A V T Δ α x α ¯ x e ^ x F Δ α z = π η g 2 σ I 32 ρ g κ g R A V T Δ α z α ¯ z e ^ z
α = T r T i T s T i
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.