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

Optical pulling using evanescent mode in sub-wavelength channels

Open Access Open Access

Abstract

Optical evanescent wave in total internal reflection has been widely used in efficient optical manipulation, where the object is trapped by the intrinsic intensity gradient of the evanescent wave while transported by the scattering force along the orthogonal direction. Here, we propose a distinct optical manipulation scheme using the attenuated modes in subwavelength optical channels, where both the trapping and transportation forces are along the channel direction. We create such a mode in a sub-wavelength photonic crystal waveguide and quantitatively obtain the net pushing and pulling forces, which can overcome the Brownian motion within a critical length. Due to the presence of the physical channel, subwavelength trapping on the transverse direction is natural, and manipulation along bend trajectories is also possible without the assistance of the self-acceleration beams provided a channel is adopted. This optical manipulation method can be extended to any other channels that support attenuation mode, and may provide an alternate way for flexible optical manipulation.

© 2016 Optical Society of America

1. Introduction

Since A. Ashkin’s demonstration of micron-sized particle trapping using tightly focused laser beams [1, 2], micro-manipulation using optical force has witnessed a glorious development [3–5]. Now, optical tweezer has become an indispensable tool in multiple disciplines including atomic physics [6], biological science [7], chemistry [8], and quantum physics [9]. Now, optical force tailoring has been extended to more sophisticated systems including Fano resonance structure [10], chiral material [11], metamaterial [12], and even graded index structures [13, 14]. In optical tweezers, the key point is to generate a large intensity gradient using a high numerical aperture objective lens which focuses the incident beam into a tight spot, where the optical trapping occurs. This mechanism is high efficiency in trapping object around a fixed point while it is lack of flexibility for long distance transportation.

For the purpose of long range transportation, two different mechanisms have been proposed. The first one is utilizing the evanescent wave occurred in total internal reflection on a dielectric interface [15–31]. In this case, the evanescent wave generates a large intensity gradient (exponentially decaying law of evanescent mode) force along the normal direction of the interface, which pulls the object to the interface. At the same time, the scattering force parallel the interface can transport the object over long distance, as shown in Fig. 1(a). In a similar manner, one can also use a waveguide structure (such as a micro-fiber) to generate the evanescent mode. In this case, the guiding mode extends outside of the physical boundary of the waveguide, and can trap and manipulate objects over long distance. Then second more interesting mechanism is the optical tractor beams [32–44], which can transport object backwardly, reverse to the beam propagation direction. Unfortunately, it is a great challenge to achieve an optical tractor beam in practice, where special optical beams [32–34], or special operation objects with exotic optical parameters [36, 45], or a rather complex background configurations are needed [43].

 figure: Fig. 1

Fig. 1 Schematic illustration of electromagnetic force in evanescent modes. (a) Traditional configuration where the gradient force and scattering force are perpendicular to each other. (b) Our new configuration with collinear gradient force and scattering force, which can be regarded as the combination of traditional evanescent mode manipulation and optical tractor beam.

Download Full Size | PDF

In this paper, we propose a distinct optical manipulation method using the attenuation propagation mode, as illustrated in Fig. 1(b), which can be regarded as the combination of evanescent wave trapping and optical tractor beam. In this case, the transportation over relatively long distance is achieved by the attenuation mode, which provides both backward gradient force and forward scattering force. The direction of the total force is determined by their relative amplitudes, which can be tuned by the oscillating frequency and/or the damping rate of the mode. Thus, both the optical pushing force and the tractor-beam-like pulling force can be realized. We also create a solid model in sub-wavelength photonic crystal (PC) waveguides, and obtain the forward and backward forces quantitatively. Both the pushing and pulling forces can overcome the Brownian motion within a critical distance, thus can be used to transport and sort particles. More importantly, due to the physical boundary of the channel, the transverse size can be much smaller than the wavelength, which is advantage for nano-sized object manipulation along straight and bend trajectories. This optical-straw-like manipulation may find potential applications in multidisciplinary fields.

2. Principle, results and discussion

2.1. Optical force in attenuation fields

For an attenuated propagation mode along x direction, the propagation constant β is a complex number of β=αiγ with α and γ being real and positive numbers. Then, the dominant electric fieldEzof the mode can be expressed as

Ez(x,y,z)=Ez0(y,z)eiβx=Ez0(y,z)eiαxγx.
Here Ez0 is the amplitude at the position of x=0. For the case of Rayleigh particle with a radius of r, the time averaged optical gradient force can be obtained by [3]
Fgrad=2πAcn12I(x,y,z)   with   A=n12(m21m2+2)r3.
Here I(x,y,z) is the light intensity, and m=n1/n2 is the relative refractive index of the object. Consider the force projection along the propagation direction of x, one can get straightforwardly

Fgrad,x=2πAcn12xI(x,y,z)=4πγr3c(m21m2+2)I(x,y,z).

Since the attenuation parameter γ is positive, Fgrad,xis a negative force reverse to the propagation direction. However, we still cannot judge the motion direction by this force only since there is also a forward radiation pressure. For the case of Rayleigh particle, we can also derive the scattering force analytically, which is [46]

Fscat,x=8πn1α4r63c(m21m2+2)2I(x,y,z).

From Eq. (4), we can see that Fscat,x is always positive. Finally, the competition between the negative and positive forces gives the total force of

Ftot,x=4πr3c(2n1α4r33m21m2+2γ)(m21m2+2)I(x,y,z).

For the mode with γ=0, one can see that Fgrad,x=0 from Eq. (3). Then the total optical force is determined by the positive scattering force only. For a purely attenuation mode without propagation (i.e., α=0), one can get that Fscat,x=0, and the total force is contributed only by the intrinsic gradient force. For the intermediate cases, i.e., attenuation propagation modes, both the scattering and gradient forces are nonzero, and the total force is determined by their relative amplitude.

In order to get a tractor-beam-like negative force, from Eq. (5) we can see that the propagation constant β=αiγ also plays an important role, apart from the relative refractive index m and radius r of the scatterer. For the usual optical manipulations operated in boundless spaces (such as water or air), β=n2k0 can only be tuned by replacing the background medium (for the same source). Here, we propose to tune the scattering force by introducing optical channel, in which the effective mode refractive n2 can be tuned by selecting different modes. In the extreme case of α=0, the forward scattering force is vanished, and then a tractor-beam-like negative optical force will appear.

2.2. Optical manipulation in sub-wavelength photonic crystal waveguide

Now, we quantitatively consider the damped propagating mode discussed above in a solid structure of two dimensional (2D) photonic crystal (PC) waveguide. As shown in Fig. 2(a), the square lattice PC is formed by silicon rods in water (refractive indexes of na=3.56and nb=1.33, respectively) with the lattice constant of a. A W1 guiding channel is formed by removing a line of silicon rods along the ΓX direction. Then a circular object with a refractive index of no and radius of ro is launched into the channel. For the TM mode (Hx,y and Ez are nonzero), the band structure is calculated using the plane wave expansion method [47], and the results are shown in Fig. 2(b). The shaded region between the guiding mode (upper left curve, blue dashed) and band edge mode (lower right curve, solid red) is the expected attenuation propagation band with complex propagation constants. The upper right inset shows the field pattern of |Ez(x,y)| near the upper band edge of ω1=0.226(2πc/a). It can be seen that the damping rate is small, and |Ez| still keeps significant at a distance of 50a (more than 11 times of the wavelength in vacuum λ0). The lower left inset shows the field pattern of ω2 near the lower edge of the damping band. We can see that the damping rate is relatively large, and |Ez(x,y)| tends to zero within a distance of 10a (about 2λ0). One also can find that the modes are well confined on the transverse direction, of which the dimension is determined by the geometry size of the channel, which is at the sub-wavelength scale. This feature is advantage for the nano-sized object manipulation.

 figure: Fig. 2

Fig. 2 Structure and the evanescent mode. (a) Schematic of the photonic crystal (PC) waveguide channel, which supports attenuation propagating mode. The parameters are na = 3.56 and nb = 1.33. (b) Band structure of TM mode (nonzero component of Hx,y and Ez) of the PC channel. The blue dashed curve (upper left curve) shows the guiding mode of the waveguide, and the red solid (lower right) curve is the band edge of the PC. The shaded region between those two curves is the band of attenuation propagation modes with complex propagation constant. The upper right inset shows the field pattern of |Ez| at the frequency ω1 with a smaller damping rate, and the lower left inset shows the mode of ω2 with a larger damping rate.

Download Full Size | PDF

We calculate the optical force using the integration of Maxwell stress tensor Talong an enclosed surface (it is a closed curve in 2D case), i.e.

F=STn^ds   with   T=12Re[DE*+HB*12I(ED*+HB*)].

Here means the time average operation over an oscillation period, Iis the unit tensor, and is the dyad operator. Re(A) and A* represent the real part and the complex conjunction of A (with A an arbitrary vector). All the involved electric and magnetic components of D, E, B and H are numerically calculated using the finite-difference in time-domain method. Since the object investigated here is embedded in water but not in vacuum, there is some controversy on the proper stress tensor to use. According to our results reported in Ref [48], Minkowski stress tensor can always predict the accurate time averaged optical force even in a non-vacuum background, such as in the inhomogeneous background of interfacial tractor beam case [48]. The Chu and Nelson stress tensors, however, lead to pushing force for the interfacial tractor beam, which is conflict with the experimental observations [48]. Therefore, the Minkowski stress tensor is used here.

For the scatterer with a radius of ro=0.2a and a moderate refractive index of no=1.5, we calculate the optical force Fx on the object when it locates at different positions along the x-axis, and the results are shown in Fig. 3(a). Both positive pushing and negative pulling forces are obtained when different damping rate of the mode is selected by setting different incident frequency, which agrees well with the analysis above. For the frequencies of 0.2259 and 0.2260 (normalized by 2πc/a), the decaying rate γis small, which is overcome by the forward scattering force, and a net pushing force is obtained. On the other hand, for the frequencies of 0.2257 and 0.2258, the decaying rate γ is large enough to generate a strong backward gradient force to overcome the forward scattering force, and then a tractor-beam-like like net pulling force is obtained. This shows that the force can be reversed easily by slight tuning the mode damping rate, or equivalently the incident frequency.

 figure: Fig. 3

Fig. 3 (a) Optical forces change with x for different modes. The radius and refractive index of the object are r=0.2a and no=1.5. (b) Optical force changes with the refractive index and size of the object at the fixed frequency of ω=0.2258. The regions I (III) means Fx is always pulling (pushing) at any positions of x, while the middle region II means Fx maybe pushing or pulling dependent on the position.

Download Full Size | PDF

When the incident frequency is fixed, we calculate the optical forces Fxfor different objects that are characterized by the parameters of no[1.36,1.77] and ro[0.14a,0.36a], and the results are shown in Fig. 3(b). One can see that the entire domain is divided into three regions. For the lower left region I with smaller radius and smaller relative refractive index, Fx is always negative wherever the scatterer is in the channel. For the upper right region III, however, Fx is positive for arbitrary value of x, which means that the pushing force is always larger than the gradient force. The middle region II is the transitional region where the sign of Fx is dependent on the position. Those results show that the optical force can be utilized to sort objects with different size, and/or different refractive index.

Comparing with other tractor beams and optical tweezers investigated in boundless spaces, one advantage of current manipulation is the ability of transporting objects along curved trajectories. In boundless space, self-acceleration beams have to be used for optical manipulations along curved paths, such as the Airy beam that is proposed as an “optical snowblower” [49], or helical profile beam both in phase and amplitude [50, 51]. In current situation, however, the curve manipulation path can be achieved easily by design the channel geometries, rather than by a complex self-acceleration light beam or helical beam, which makes this scheme more flexible in practice.

In order to illustrate this feature, we create different bend channels, and consider the optical forces in them. The bending channels are shown by the gray open circles in Fig. 4, where the angular and radial distances between the units are both a, the lattice constant of the photonic crystal structure, while the whole structure is bend with a radius of R. In Fig. 4(a), the red color shows the field pattern of the mode. One can see clearly that the mode can bend smoothly according to the channel, and the field pattern is almost the same as that in straight. The thick blue curve with arrows shows the motion trajectory of a dipole object pulled by the optical force, which is calculated by Eq. (5). Figure 4(b) is the same as Fig. 4(a) except for the bend radius is smaller of R = 12a. For the case of pushing force, it is much easier to achieve using the mode with smaller damped factor, and not shown here. From those results, one can see that the manipulation along bend trajectories is an intrinsic feature of our mechanism free from complicated self-acceleration beams.

 figure: Fig. 4

Fig. 4 Optical manipulation through a bend channel with bending radius of (a) R = 16a, and (b) R = 12a. The gray open circles show the position of the photonic crystal and red color shows the field pattern of the mode. The blue thick curves with arrows show the motion trajectory of a dipole object driven by the optical force.

Download Full Size | PDF

2.3. Effect of Brownian motion on the manipulation

When Brownian motion is considered, the manipulation property will be modified slightly. Results show that the most significant influence of Brown motion is the decrease of the effective operation range. In order to measure the influence of Brownian motion, we consider the trapping parameter of C(x)=xFx(x')dx'/kBT. Here, kB is the Boltzmann constant, and T is the absolute temperature in Kelvin. Stable trapping and transportation is excepted only when C is larger than certain critical value C0, such as |C(x)|10=C0 [2]. When Fx<0 and C<C0, the object cannot escape from the potential formed by the negative pulling force, and the object will be pulled back to the source direction. Figure 5 shows the profiles of C(x) for different incident frequencies at the incident level of 10 mW, and at the temperature of 300K. We define the effective length Lefffor the optical manipulation parameter by the criteria of |C(Leff)|=C0. We can see that, although the amplitude of force is decaying exponentially with distance, the operation distance can extend as long as several times of wavelength. For the case of tractor-beam-like pulling force, the effective length is about 4λ0 with λ0 the optical wavelength in vacuum. For the object within this distance, the object can be pulled back to the source. For the case of optical pushing force, the effective length is about 5λ0. An object within this distance can be pushed away. What’s more, the effective pushing distance can be increased very easily by tuning the damping rate γ through the incident frequency. It is noted that the effective length (both the pushing length and pulling length) can be tuned by changing the incident power.

 figure: Fig. 5

Fig. 5 The manipulation factor C(x) changes with position at the incident level of 10mW.

Download Full Size | PDF

3. Discussion and conclusion

From those results and discussions presented above, one can find that when the channel mode works as an optical tractor beam like operation, our method possesses several advantages. In the usually investigated optical tractor beam, the incident field firstly exerts a pushing force that is closely related to the momentum per photon of k (with kbeing the wavevector, the reduced Plank’s constant), and then the negative pulling force is possible only when the forward scattering is enhanced enough to generate a larger backward recoil force. Therefore, an intense endeavor is spent to engineer the source and/or object to increase the forward scattering. In our case, the incident pushing force can be greatly reduced due to the present of the waveguide channel, which makes the real part of the propagation constant β much smaller (or even vanish is some extreme cases) than that in homogeneous backgrounds. Therefore, only a moderate damping rate can result in a backward gradient force that overcomes the undesired forward radiation pressure. Also, the waveguide channel provides a physical boundary for the manipulation, which performs like an optical analog of drinking straw for micro particle in the bidirectional transportation process. The physical boundary is also beneficial for the stability on the transverse direction in manipulation, and provides the manipulation ability along curved trajectories (channels). Our scheme releases almost all the other tight restrictions adopted in tractor beams before, such as nonparaxial Bessel beam, gain medium, or negative refractive materials. Those properties are beneficial for practical application.

In summary, we have proposed an optical manipulation method using the attenuation propagation mode in a guiding channel, rather than in a boundless space. The operation can be regarded as a mixing of the optical tweezer and the optical tractor beam with tunable direction of total force. Due to the mode engineering ability and physical boundary of the channel, an object could be bi-directionally transported along straight and bended trajectories. Using the finite-difference in time-domain method, we quantitatively investigated the optical channel manipulation properties in different photonic crystal waveguide channels. This optical pulling force can be extended straightly to other kinds of channels which support attenuation and propagation modes. Therefore, it provides a flexible and universal method for optical manipulations, which can be used in optical sorting and extraction.

Funding

National Natural Science Foundation of China (NSFC) (11474077, and 11404083); Program for Innovation Research of Science in Harbin Institute of Technology (A201411, andB201407); Fundamental Research Funds for the Central Universities of China.

Acknowledgment

The authors thank Prof. Cheng-Wei Qiu of National University of Singapore for useful discussion. T. Zhu and M.R.C. Mahdy contribute equally to this work.

References and links

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

2. 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), 288–290 (1986). [CrossRef]   [PubMed]  

3. K. C. Neuman and S. M. Block, “Optical trapping,” Rev. Sci. Instrum. 75(9), 2787–2809 (2004). [CrossRef]   [PubMed]  

4. K. Dholakia and T. Čižmár, “Shaping the future of manipulation,” Nat. Photonics 5(6), 335–342 (2011). [CrossRef]  

5. D. G. Grier, “A revolution in optical manipulation,” Nature 424(6950), 810–816 (2003). [CrossRef]   [PubMed]  

6. S. Stellmer, B. Pasquiou, R. Grimm, and F. Schreck, “Laser cooling to quantum degeneracy,” Phys. Rev. Lett. 110(26), 263003 (2013). [CrossRef]   [PubMed]  

7. M. Soltani, J. Lin, R. A. Forties, J. T. Inman, S. N. Saraf, R. M. Fulbright, M. Lipson, and M. D. Wang, “Nanophotonic trapping for precise manipulation of biomolecular arrays,” Nat. Nanotechnol. 9(6), 448–452 (2014). [CrossRef]   [PubMed]  

8. K. C. Neuman and A. Nagy, “Single-molecule force spectroscopy: optical tweezers, magnetic tweezers and atomic force microscopy,” Nat. Methods 5(6), 491–505 (2008). [CrossRef]   [PubMed]  

9. J. Chan, T. P. Alegre, A. H. Safavi-Naeini, J. T. Hill, A. Krause, S. Gröblacher, M. Aspelmeyer, and O. Painter, “Laser cooling of a nanomechanical oscillator into its quantum ground state,” Nature 478(7367), 89–92 (2011). [CrossRef]   [PubMed]  

10. T. Cao, L. Mao, D. Gao, W. Ding, and C. W. Qiu, “Fano resonant Ge2Sb2Te5 nanoparticles realize switchable lateral optical force,” Nanoscale 8(10), 5657–5666 (2016). [CrossRef]   [PubMed]  

11. M. Wang, H. Li, D. Gao, L. Gao, J. Xu, and C. W. Qiu, “Radiation pressure of active dispersive chiral slabs,” Opt. Express 23(13), 16546–16553 (2015). [CrossRef]   [PubMed]  

12. T. Cao, L. Zhang, and M. J. Cryan, “Optical forces in metal/dielectric/metal fishnet metamaterials in the visible wavelength regime,” IEEE Photonics J. 4(5), 1861–1869 (2012). [CrossRef]  

13. A. Akbarzadeh, M. Danesh, C.-W. Qiu, and A. J. Danner, “Tracing optical force fields within graded-index media,” New J. Phys. 16(5), 053035 (2014). [CrossRef]  

14. A. Akbarzadeh, J. A. Crosse, M. Danesh, C.-W. Qiu, A. J. Danner, and C. M. Soukoulis, “Interplay of optical force and ray-optic behavior between Luneburg lenses,” ACS Photonics 2(9), 1384–1390 (2015). [CrossRef]  

15. S. Kawata and T. Sugiura, “Movement of micrometer-sized particles in the evanescent field of a laser beam,” Opt. Lett. 17(11), 772–774 (1992). [CrossRef]   [PubMed]  

16. S. Chang, J. H. Jo, and S. S. Lee, “Theoretical calculations of optical force exerted on a dielectric sphere in the evanescent field generated with a totally-reflected focused gaussian beam,” Opt. Commun. 108(1-3), 133–143 (1994). [CrossRef]  

17. M. Lester and M. Nieto-Vesperinas, “Optical forces on microparticles in an evanescent laser field,” Opt. Lett. 24(14), 936–938 (1999). [CrossRef]   [PubMed]  

18. D. Ganic, X. Gan, and M. Gu, “Trapping force and optical lifting under focused evanescent wave illumination,” Opt. Express 12(22), 5533–5538 (2004). [CrossRef]   [PubMed]  

19. M. Gu, J.-B. Haumonte, Y. Micheau, J. W. M. Chon, and X. Gan, “Laser trapping and manipulation under focused evanescent wave illumination,” Appl. Phys. Lett. 84(21), 4236 (2004). [CrossRef]  

20. M. Šiler, T. Čižmár, M. Šerý, and P. Zemánek, “Optical forces generated by evanescent standing waves and their usage for sub-micron particle delivery,” Appl. Phys. B 84(1–2), 157–165 (2006). [CrossRef]  

21. G. Brambilla, G. S. Murugan, J. S. Wilkinson, and D. J. Richardson, “Optical manipulation of microspheres along a subwavelength optical wire,” Opt. Lett. 32(20), 3041–3043 (2007). [CrossRef]   [PubMed]  

22. S. Gaugiran, S. Gétin, J. M. Fedeli, and J. Derouard, “Polarization and particle size dependence of radiative forces on small metallic particles in evanescent optical fields. Evidences for either repulsive or attractive gradient forces,” Opt. Express 15(13), 8146–8156 (2007). [CrossRef]   [PubMed]  

23. V. Ruiz-Cortés and J. P. Vite-Frías, “Lensless optical manipulation with an evanescent field,” Opt. Express 16(9), 6600–6608 (2008). [CrossRef]   [PubMed]  

24. M. Šiler, T. Čižmár, A. Jonáš, and P. Zemánek, “Surface delivery of a single nanoparticle under moving evanescent standing-wave illumination,” New J. Phys. 10(11), 113010 (2008). [CrossRef]  

25. K. Wang, E. Schonbrun, and K. B. Crozier, “Propulsion of gold nanoparticles with surface plasmon polaritons: evidence of enhanced optical force from near-field coupling between gold particle and gold film,” Nano Lett. 9(7), 2623–2629 (2009). [CrossRef]   [PubMed]  

26. S. S. Varghese and M. Gu, “Optical micromanipulations in the non-diffractive regime,” J. Biophotonics 3(4), 207–215 (2010). [CrossRef]   [PubMed]  

27. K. Wang, E. Schonbrun, P. Steinvurzel, and K. B. Crozier, “Scannable plasmonic trapping using a gold stripe,” Nano Lett. 10(9), 3506–3511 (2010). [CrossRef]   [PubMed]  

28. B. H. Liu, L. J. Yang, Y. Wang, and J. L. Yuan, “Nano-manipulation performance with enhanced evanescent field close to near-field optical probes,” Opt. Commun. 284(12), 3039–3046 (2011). [CrossRef]  

29. Y. Yang, J. Li, and Z.-Y. Li, “Fano resonance of the ultrasensitve optical force excited by Gaussian evanescent field,” J. Opt. 17(7), 075004 (2015). [CrossRef]  

30. L. Liu, A. Woolf, A. W. Rodriguez, and F. Capasso, “Absolute position total internal reflection microscopy with an optical tweezer,” Proc. Natl. Acad. Sci. U.S.A. 111(52), E5609–E5615 (2014). [CrossRef]   [PubMed]  

31. N. Irawati, V. John, M. M. Aeinehvand, F. Ibrahim, H. Ahmad, and S. W. Harun, “Evanescent wave optical trapping and transport of polystyrene microspheres on microfibers,” Microw. Opt. Technol. Lett. 56(11), 2630–2634 (2014). [CrossRef]  

32. J. Chen, J. Ng, Z. Lin, and C. T. Chan, “Optical pulling force,” Nat. Photonics 5(9), 531–534 (2011). [CrossRef]  

33. A. Dogariu, S. Sukhov, and J. J. Sáenz, “Optically induced ‘negative forces’,” Nat. Photonics 7(1), 24–27 (2012). [CrossRef]  

34. A. Novitsky, C.-W. Qiu, and H. Wang, “Single gradientless light beam drags particles as tractor beams,” Phys. Rev. Lett. 107(20), 203601 (2011). [CrossRef]   [PubMed]  

35. A. Novitsky, C.-W. Qiu, and A. Lavrinenko, “Material-independent and size-independent tractor beams for dipole objects,” Phys. Rev. Lett. 109(2), 023902 (2012). [CrossRef]   [PubMed]  

36. K. J. Webb and Shivanand, “Negative electromagnetic plane-wave force in gain media,” Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 84(5), 057602 (2011). [CrossRef]   [PubMed]  

37. N. Wang, J. Chen, S. Liu, and Z. Lin, “Dynamical and phase-diagram study on stable optical pulling force in Bessel beams,” Phys. Rev. A 87(6), 063812 (2013). [CrossRef]  

38. S. Sukhov and A. Dogariu, “Negative nonconservative forces: optical “tractor beams” for arbitrary objects,” Phys. Rev. Lett. 107(20), 203602 (2011). [CrossRef]   [PubMed]  

39. D. B. Ruffner and D. G. Grier, “Optical conveyors: a class of active tractor beams,” Phys. Rev. Lett. 109(16), 163903 (2012). [CrossRef]   [PubMed]  

40. N. Wang, W. Lu, J. Ng, and Z. Lin, “Optimized optical “tractor beam” for core-shell nanoparticles,” Opt. Lett. 39(8), 2399–2402 (2014). [CrossRef]   [PubMed]  

41. S. Sukhov and A. Dogariu, “On the concept of “tractor beams”,” Opt. Lett. 35(22), 3847–3849 (2010). [CrossRef]   [PubMed]  

42. O. Brzobohatý, V. Karásek, M. Šiler, L. Chvátal, T. Čižmár, and P. Zemánek, “Experimental demonstration of optical transport, sorting and self-arrangement using a ‘tractor beam’,” Nat. Photonics 7(2), 123–127 (2013). [CrossRef]  

43. V. Kajorndejnukul, W. Ding, S. Sukhov, C.-W. Qiu, and A. Dogariu, “Linear momentum increase and negative optical forces at dielectric interface,” Nat. Photonics 7(10), 787–790 (2013). [CrossRef]  

44. D. Gao, A. Novitsky, T. Zhang, F. C. Cheong, L. Gao, C. T. Lim, B. Luk’yanchuk, and C.-W. Qiu, “Unveiling the correlation between non-diffracting tractor beam and its singularity in Poynting vector,” Laser Photonics Rev. 9(1), 75–82 (2015). [CrossRef]  

45. A. Salandrino and D. N. Christodoulides, “Reverse optical forces in negative index dielectric waveguide arrays,” Opt. Lett. 36(16), 3103–3105 (2011). [CrossRef]   [PubMed]  

46. Y. Harada and T. Asakura, “Radiation forces on a dielectric sphere in the Rayleigh scattering regime,” Opt. Commun. 124(5-6), 529–541 (1996). [CrossRef]  

47. S. Johnson and J. Joannopoulos, “Block-iterative frequency-domain methods for Maxwell’s equations in a planewave basis,” Opt. Express 8(3), 173–190 (2001). [CrossRef]   [PubMed]  

48. C.-W. Qiu, W. Ding, M. R. C. Mahdy, D. Gao, T. Zhang, F. C. Cheong, A. Dogariu, Z. Wang, and C. T. Lim, “Photon momentum transfer in inhomogeneous dielectric mixtures and induced tractor beams,” Light Sci. Appl. 4(4), e278 (2015). [CrossRef]  

49. J. Baumgartl, M. Mazilu, and K. Dholakia, “Optically mediated particle clearing using Airy wavepackets,” Nat. Photonics 2(11), 675–678 (2008). [CrossRef]  

50. V. R. Daria, D. Z. Palima, and J. Glückstad, “Optical twists in phase and amplitude,” Opt. Express 19(2), 476–481 (2011). [CrossRef]   [PubMed]  

51. A. A. Kovalev, V. V. Kotlyar, and A. P. Porfirev, “Optical trapping and moving of microparticles by using asymmetrical Laguerre-Gaussian beams,” Opt. Lett. 41(11), 2426–2429 (2016). [CrossRef]   [PubMed]  

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 Schematic illustration of electromagnetic force in evanescent modes. (a) Traditional configuration where the gradient force and scattering force are perpendicular to each other. (b) Our new configuration with collinear gradient force and scattering force, which can be regarded as the combination of traditional evanescent mode manipulation and optical tractor beam.
Fig. 2
Fig. 2 Structure and the evanescent mode. (a) Schematic of the photonic crystal (PC) waveguide channel, which supports attenuation propagating mode. The parameters are na = 3.56 and nb = 1.33. (b) Band structure of TM mode (nonzero component of Hx,y and Ez) of the PC channel. The blue dashed curve (upper left curve) shows the guiding mode of the waveguide, and the red solid (lower right) curve is the band edge of the PC. The shaded region between those two curves is the band of attenuation propagation modes with complex propagation constant. The upper right inset shows the field pattern of |Ez| at the frequency ω 1 with a smaller damping rate, and the lower left inset shows the mode of ω 2 with a larger damping rate.
Fig. 3
Fig. 3 (a) Optical forces change with x for different modes. The radius and refractive index of the object are r=0.2a and n o =1.5 . (b) Optical force changes with the refractive index and size of the object at the fixed frequency of ω=0.2258 . The regions I (III) means Fx is always pulling (pushing) at any positions of x, while the middle region II means Fx maybe pushing or pulling dependent on the position.
Fig. 4
Fig. 4 Optical manipulation through a bend channel with bending radius of (a) R = 16a, and (b) R = 12a. The gray open circles show the position of the photonic crystal and red color shows the field pattern of the mode. The blue thick curves with arrows show the motion trajectory of a dipole object driven by the optical force.
Fig. 5
Fig. 5 The manipulation factor C(x) changes with position at the incident level of 10mW.

Equations (6)

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

E z ( x,y,z )= E z0 ( y,z ) e iβx = E z0 ( y,z ) e iαxγx .
F grad = 2πA c n 1 2 I( x,y,z )   with   A= n 1 2 ( m 2 1 m 2 +2 ) r 3 .
F grad,x = 2πA c n 1 2 x I( x,y,z )= 4πγ r 3 c ( m 2 1 m 2 +2 )I( x,y,z ).
F scat,x = 8π n 1 α 4 r 6 3c ( m 2 1 m 2 +2 ) 2 I( x,y,z ).
F tot,x = 4π r 3 c ( 2 n 1 α 4 r 3 3 m 2 1 m 2 +2 γ )( m 2 1 m 2 +2 )I( x,y,z ).
F= S T n ^ ds    with    T = 1 2 Re[ D E * +H B * 1 2 I (E D * +H B * ) ].
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.