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

Spinning and orbiting motion of particles in vortex beams with circular or radial polarizations

Open Access Open Access

Abstract

Focusing fields of optical vortex (OV) beams with circular or radial polarizations carry both spin angular momentum (SAM) and orbital angular momentum (OAM), and can realize non-axial spinning and orbiting motion of absorptive particles. Using the T-matrix method, we evaluate the optical forces and torques exerted on micro-sized particles induced by the OV beams. Numerical results demonstrate that the particle is trapped on the circle of intensity maxima, and experiences a transverse spin torque along azimuthal direction, a longitudinal spin torque, and an orbital torque, respectively. The direction of spinning motion is not only related to the sign of topological charge of the OV beam, but also to the polarization state. However, the topological charge controls the direction of orbiting motion individually. Optically induced rotations of particles with varying sizes and absorptivity are investigated in OV beams with different topological charges and polarization states. These results may be exploited in practical optical manipulation, especially for optically induced rotations of micro-particles.

© 2016 Optical Society of America

1. Introduction

It is well known that optical momentum and angular momentum (AM) are the main dynamical properties of electromagnetic waves, which manifest themselves and have been widely used in light-matter interactions [1], including optical manipulations [2], laser cooling [3], and opto-mechanical systems [4]. When light is scattered by a particle, the transfers of momentum and angular momentum from light to particle generate a radiation force and torque on the particle [5–9]. The optical AM of an electromagnetic wave includes a spin angular momentum (SAM) and an orbital angular momentum (OAM) [10]. The SAM results in the rotation of particle around its own axis, while the OAM causes rotation of particle around the optical beam axis. These light-induced rotations have attracted increasing interests and attentions [11]. For example, circularly polarized beams and cylindrical vector beams are used to spin absorbing particles [12, 13], and optical vortex (OV) beams are employed to rotate metal particles [14, 15]. Nevertheless, in most previous works, the induced rotation direction is along the beam propagation axis [13–17]. The latest research show that transverse SAM appears in various structured fields: evanescent waves [18], interference fields [19], and focused beams [20]. Transverse spinning of particles in plasmonic fields (i.e., evanescent waves) [21] and interference fields [22] have been implemented. However, the effect of transverse spin in focusing fields still needs more attention and much deeper investigation. This paper will show some theoretical results on this effect.

An optical beam with circular polarization is believed to carry a SAM orientated approximately along the propagation direction, i.e., longitudinal SAM. As a result, it is hard to realize transverse spinning of particle in such a beam. However, under tight focusing, the focusing fields of such beams contain appreciable longitudinal components and may potentially carry transverse SAM. For example, tightly focusing a radially polarized beam leads to a strong longitudinal electric field [23], causing the formation of a very strong transverse SAM. Optical vortex beams carrying OAM [24] can drive particles to orbit around the optical beam axis. Most of the previous studies on OV-induced rotation are focused on the orbital rotation of particles [15–17, 25, 26]. Actually, there also exists induced spinning motion of some particles in highly focused OV beams.

In this paper, we propose to realize spinning and orbiting motion of absorptive particles by highly focused OV beams with circular or radial polarizations. In order to understand the rotating properties of the particles, we firstly study the distributions of the focusing fields. Then, we calculate the optical forces and torques on particles based on T-matrix method, and analyze the physical mechanisms of rotation systematically. Furthermore, the influences of the characteristics of the particles and the OV beams on the rotational dynamics are investigated in detail. Some interesting results of spinning and orbiting motion in tightly focused OV beams will be shown and discussed.

2. Theory

Consider a dielectric spherical particle of radius a and index of refraction n2 illuminated by an arbitrary electromagnetic wave with the fields Einc and Hinc. The incident wave is scattered by the particle, yielding the scattered fields Esca and Hsca outside the particle. Assuming that Esca and Hsca are known, then the time-averaged optical forces 〈F〉 and torques 〈Γs〉 on the particle can be calculated by integrating the Maxwell stress tensor over a surface S enclosing the particle

F=Sn^T¯ds,
Γs=Sn^(T¯×r)ds,
here is the outwardly directed normal unit vector to the surface, r is the position vector, and is the Maxwell stress tensor
T¯=ε1EE+μ1HH12(ε1E2+μ1H2)I¯,
with ε1 and μ1 denoting the permittivity and permeability of the medium surrounding the particle, E ( = Einc + Esca) and H ( = Hinc + Hsca) representing the total fields outside the particle, and is the unit dyadic. So far, there still remains the question of how to determine the scattered fields. Here we employ the T-matrix method by Waterman [27] to treat the scattering problem. For the case of isotropic spherical particles, discussed here, the T-matrix method conforms with the well-known Mie scattering method. In the T-matrix method, the incident and scattered fields are expanded as a series of suitable vector spherical wave functions (VSWFs),
Einc(r)=n=1m=nn[amnMmn1(kr)+bmnNmn1(kr)],
Esca(r)=n=1m=nn[pmnMmn3(kr)+qmnNmn3(kr)],
where r is the coordinate of observation point, k is the wave number in the surrounding medium; (amn, bmn) and (pmn, qmn) are the incident and scattered coefficients, respectively; Mmn13(kr) and Nmn13(kr) are VSWFs of the first and third kinds [28].

By applying the boundary conditions on the surface of the particle, the scattered coefficients (pmn, qmn) are found to be related to the incident coefficients (amn, bmn), in matrix notation, as [28],

[pmnqmn]=[Tmnm'n'1100Tmnm'n'22][am'n'bm'n'].
Here the first matrix of the right side of Eq. (6) is the T-matrix, which is diagonal since we are considering a spherical particle.

Substituting the expressions (4) and (5) for the fields into the integrals (1) and (2), it is found that the forces 〈F〉 and torques 〈Γs〉 can be expressed in a sum of the coefficients (amn, bmn) and (pmn, qmn) in quadratic form [29]. Note that in evaluating the integral (2), the origin of the coordinate system is overlapped with the center of the particle. Since the particle is of spherical shape, isotropic and homogenous, the torques thus obtained may be regarded as the intrinsic torques or spin torques on the particle. This is why the subscript s is attached to the torque notation in the Eq. (2). In actual optical trapping, if the equilibrium trapping position of particle is not on the optical axis, the particle experienced an orbital torque, is extrinsic and dependent on the position of the particle. Denoting ro the position vector of particle, the orbital torque 〈Γo〉 is clearly 〈Γo〉 = ro × 〈F〉. In most situations, we are particularly interested in the z-component of orbital torque, which is expressed as

Γo,z=ρoFϕ,
where ρo is radial distance of the particle from the optical axis, <Fϕ> is the time-averaged azimuthal force.

In optical trapping, the field incident on particles is actually the focused field of some input illumination focused by a high numerical aperture (NA) objective lens. According to Richards and Wolf [30], the focused field near the focus can be expressed as an integral of plane waves of the form of

Einc(r)=ikf2π0θmax02πA(θ,ϕ)exp(ikr)sinθdϕdθ.
Here f is the focal length, θmax is the maximal converging angle given by the NA and k is the wave number in the image space where the particle is located; k and r designate the wave vector of plane waves and the position vector of observation point; A(θ, ϕ) stands for the apodized field. The apodized field is related to the input field A0(θ, ϕ) at the entrance pupil according to the following transform rule [31, 32]
A(θ,ϕ)=(cosθ)1/2[eθ00eϕ](A0ρA0ϕ),
where (eθ, eϕ) are the respective unit vectors in the θ and ϕ directions, and (A0ρ, A0ϕ) are the radial and azimuthal components of the input field A0(θ, ϕ). In principle, the torques on trapped particles are a result of the transfer of optical AM of the incident fields to the particles. It is known that the vortex phase factor exp(imϕ) and the polarization helix of optical fields are manifestation of optical AM. In view of this, we put the input field in a form as A0(θ, ϕ) = ul(θ)exp(imϕ), where u is the polarization vector, l(θ) is the amplitude function, and m is the topological charge of the vortex phase. Further we assume a constant amplitude l(θ) = l0 when 0 ≤ θθmax and l(θ) = 0 otherwise, where l0 is a constant factor dependent on the power of the incident light. For the polarization vector u, we focused on two cases: left-handed circular polarization u =(ex + iey) /2and radial polarization u = eρ.

3. Results and discussion

In what follows, we assume that the input power P = 100 mW and the free space wavelength λ0 = 1.064μm; the refractive index of image space n1 = 1.33. The input illumination is focused by an objective lens with NA = 1.26. As mentioned above, the input illumination is set to be circularly polarized vortex beam (CPV) or radially polarized vortex beam (RPV). The intensity distributions of the focused fields along x-axis of the CPV and RPV inputs with topological charge m = 5 or −5 are shown in Fig. 1. The four types of illuminations all give an annular focusing with cylindrical symmetry [31]. However the CPV5 input produces a larger radius of focusing than the CPV-5, while the two RPV inputs yield the same focusing pattern with the radius locating between those of CPV inputs. This result lies in the two facts: 1) the conversion of SAM carried by the CPV to OAM under tight focusing leads to a phase factor of exp[i(m + 1)ϕs] on the field [33, 34]; 2) the larger (absolute) value of topological charge corresponds to a larger radius of focusing.

 figure: Fig. 1

Fig. 1 Line scans along the x-axis of intensity distribution for circularly polarized vortex beams (CPV) and radially polarized vortex beams (RPV) with topological charges of m =+ 5 or −5.

Download Full Size | PDF

The optical forces and torques experienced by a spherical particle in the focused fields of the above four types of illuminations are calculated. The particle has a radius of a = 0.25μm and a complex refractive index n2 = 1.59 + 0.01i. For simplicity, the axial equilibrium position is assumed to be in focal plane and only the transverse trapping is considered. Figure 2 shows the radial trapping forces as a function of the position of the particle. It is seen that for each illumination the equilibrium position is overlapped with the circle of intensity maxima of the focused fields. Consequentially, as far as radial forces are concerned, the RPV5 and RPV-5 illuminations are equivalent, while the CPV5 and CPV-5 illuminations are somewhat different. Having trapped the particle on a circle, the orbital motion is expected to be driven by the azimuthal optical force. To this end, we examine the transverse force distributions in the focal plane for the four illuminations, which are shown in Fig. 3. On the circle of the intensity maxima as well as the equilibrium position, the azimuthal optical force for each illumination exists uniformly along the circle, implying that the particle will orbit around the optical axis. With the topological charge m = 5, the orbital motion is in a clockwise sense around the positive z-axis as shown in Figs. 3(a) and 3(b), while for m = −5 the orbital motion reverses the direction, as shown in Figs. 3(c) and 3(d). Since the sign of the topological charge determines the orientation of the z-component of optical OAM, the results of Fig. 3 are just a reflection of transfer of optical OAM to the particle.

 figure: Fig. 2

Fig. 2 Radial trapping force curves on the particle under (a) CPV and (b) RPV inputs with topological charges of m =+ 5 or −5.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 Transverse force distributions experienced by the particle in the focal plane illuminated, respectively, by CPV inputs with topological charges of (a) m = 5 and (c) m = −5, and by RPV inputs with (b) m = 5 and (d) m = −5.

Download Full Size | PDF

The above calculations show that the particle is to experience an orbital motion, thus obtaining an orbital torque. Since the particle is absorptive, it also actually experiences a spin torque. Table 1 represents a quantitative calculation of the spin and orbital torques on the particle at the equilibrium position when illuminated respectively by the four types of polarized input fields above. It is seen that the x-component of spin torque Γs, x is equal to the negative of the y-component Γs, y for each illumination, indicating that the transverse spin torques are along the positive or negative azimuthal direction. This can be explained by looking at the spin part of the angular momentum of the focusing fields. For time-harmonic fields, the time-averaged SAM density Ls ∝ Im(E × E*) [35]. Under the four types of illuminations above, their focusing fields all exhibit a π/2 phase difference between the radial component and the axial component on the circle of intensity maxima, i.e., the equilibrium position of the particle, while a π phase difference between the azimuthal and axial components [34], leading to a transverse part of the SAM orientated along the azimuthal direction and of course a longitudinal spin component. This fact suggests that the spin of the trapped particle is, in fact, along some direction off the optical axis. In Tab. 1, we also note that under CPV illumination the transverse spin torque as well as the orbital torque Γo, z reverses the direction as the sign of topological charge is changed. Under RPV illumination, the transverse spin torque, however, remains unchanged against the change of the sign of topological charge, while the longitudinal spin torque Γs, z and orbital torque Γo, z experience an inversion. The inversion of the orbital torque is caused directly by the change of the orientation of optical OAM as shown in Fig. 3. As to the spin torque, the orientation of the transverse component of SAM under CPV illumination, as well as the longitudinal component under RPV illumination, is reversed with the change of the sign of topological charge. This phenomenon can also be seen and explained from the corresponding expressions in [34].

Tables Icon

Table 1. Spin and orbital torques exerted on the particle at the equilibrium position under CPV and RPV inputs with topological charge of m = 5 or −5.

To present an institutive picture of the particle’s motion in the focusing fields, we sketch the three-dimensional orientations of the spin and orbital torque of the particle under each illumination, as shown in Fig. 4, where the black arrows represent the spin torque vectors, the black circles around the sphere with arrows denote the directions of the spinning motion (in the counter-clockwise sense), and the white arrows indicate the directions of the orbiting motion. It is seen that the particle when trapped on the circle of intensity maxima will orbit around the optical axis, but suffers different non-axial spinning motion states according to the polarization state and the topological charge sign of the input OV beams.

 figure: Fig. 4

Fig. 4 Sketch of the spin and orbital motion of the particle illuminated, respectively, by CPV inputs with topological charges of (a) m = 5 and (c) m = −5, and by RPV inputs with (b) m = 5 and (d) m = −5.

Download Full Size | PDF

We now examine the dependences of the optical forces and the torques on the particle’s size, the absorptivity (characterized by the imaginary part nʺ of the complex refractive index), and the topological charge value of the OV beams. Figures 5(a1)-5(c1) show the respective azimuthal force, spin and orbital torques at the equilibrium position as a function of the particle’s radius a for a fixed refractive index n2 = 1.59 + 0.01i under CPV5 and RPV5 illuminations. The force and torques increase with increasing size of the particle. We note that the spin torques under CPV5 and RPV5 illuminations show different variation tendency. Under the CPV5 illumination, the longitudinal component is almost equal to the transverse component as a < 0.3μm, thereafter the longitudinal component keeps increasing while the transverse component begins to decrease. Under the RPV5 illumination, the situation is different. When a < 0.3μm, the transverse component have a greater value than the longitudinal component, while for larger size, a > 0.3μm, the case is contrary. Meanwhile, the longitudinal spin torque under the CPV5 illumination is larger than that under RPV5 illumination in all size, and the gap gets smaller and smaller with increasing the size. In calculation, it is found that the equilibrium position of the particle will gradually depart from the circle of intensity maxima and approach to the optical axis with increasing the particle’s radius. The inset in Fig. 5(c1) shows how the radius (R) of stably trapping position changes with the particle’s size. For the particle of radius a < 0.3μm, the stable trapping position is seen to approximately overlap with the circle of intensity maxima so that the orbiting motion as well as the spinning motion is expected observable clearly. The maximum size of the particle that ensures the orbital motion observable is about 1μm for CPV5 and about 0.9μm for RPV5. Much larger particles will be trapped at the optical axis and undergo only axial spinning motion.

 figure: Fig. 5

Fig. 5 The changes of the azimuthal forces (a1)-(a3), the spin torques (b1)-(b3) and the orbital torques (c1)-(c3) exerted on the particle with varying radius a (a1)-(c1), and absorptivity nʺ (a2)-(c2), as well as with the topological charge value m (a3)-(c3) in CPV and RPV inputs. R in the inset of (c1) represents the radius of the equilibrium position of the particle.

Download Full Size | PDF

The dependences of the optical force and the torques on the absorptivity nʺ of the particle under CPV5 and RPV5 illuminations are shown in Figs. 5(a2)-5(c2), in which the particle’s radius a = 0.25μm, and the real part of the refractive index is 1.59. It exhibits a simple linear relation, since the absorptivity is proportional to these quantities. The variation tendency of the spin torques on the particle for CPV5 is significantly larger than that for RPV5. This is because the CPV beam carries the optical SAM, which transfers to the particle and enhances the spin effect.

The behaviors of the optical force and the torques on the particle as a function of the topological charge value m for CPV and RPV illuminations are shown in Figs. 5(a3)-5(c3), where the particle’s radius a = 0.25μm, and the refractive index n2 = 1.59 + 0.01i. Since the focusing spot size of the RPV illumination with m = 1 is too small to guarantee an orbital motion for the particle, the azimuthal force and orbital torque for RPV illumination with m = 1 are dropped from Figs. 5(a3) and 5(c3). An interesting phenomenon concerning the spin torque is observed that the longitudinal spin torque under RPV illumination with m = 1 is much stronger than that under CPV illumination, as shown in Fig. 5(b3), the ratio is roughly 0.894 / 0.494 = 1.81. Furthermore, this value for RPV1 is larger than any other values for all topological charges and polarizations. As we see, the azimuthal force and spin torques are gradually decreasing with the increase of the value of m. This tendency arises from that the radius of the annular focus increases with the charge m increasing, leading to the attenuation of the intensity of the focused ring. While, the orbital torque increases first and then decreases with an optimal topological charge to attain the maximum orbital torque. The values of m corresponding to the maxima of orbital torques are different for CPV and RPV inputs, which are 4 and 7, respectively. This observation can be interpreted from a simple calculation of Eq. (7).

4. Conclusions

In summary, tightly focused optical vortex beams with circular or radial polarizations are used here to study the spinning and orbiting motion of absorptive micro-sized particles. Numerical results show that the particle is trapped on the circle of intensity maxima and orbits around the optical axis while, but also experiences a non-axial spinning motion. The direction of spinning motion is determined by both the sign of topological charge and the polarization state of the vortex beam. For the CPV illumination, the transverse spin torque reverses the direction as the sign of topological charge is changed, but the longitudinal spin torque does not change. Under the RPV illumination, the transverse spin torque, however, remains unchanged against the change of the sign of topological charge, while the longitudinal spin torque experiences an inversion. The direction of the orbiting motion is only controlled by the sign of topological charge. When the topological charge is positive, the orbital torque vector points to the positive z-axis, otherwise, the orbital torque vector points to the negative z-axis. The azimuthal force, spin torque, and orbital torque increase when the particle’s radius and absorptivity increase. With the increase of the topological charge value, the azimuthal force and spin torque will decrease; while the orbital torque will firstly increases and then decreases with an optimal topological charge to attain the maximum orbital torque. These optical trapping properties might be helpful to investigation of optical manipulation, especially for optically induced rotations of micro-particles.

Funding

National Basic Research Program (973 Program) of China (2012CB921900); National Natural Science Foundation of China (NSFC) (11474352 and 11574389).

Acknowledgments

We would like to thank the anonymous reviewers for their exceptionally pertinent and constructive critical comments.

References and Links

1. C. Cohen-Tannoudji, J. Dupont-Roc, G. Grynberg, and P. Thickstun, Atom-Photon Interactions: Basic Processes and Applications (Wiley Online Library, 1992).

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

3. S. Stenholm, “The semiclassical theory of laser cooling,” Rev. Mod. Phys. 58(3), 699–739 (1986). [CrossRef]  

4. M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, “Cavity optomechanics,” Rev. Mod. Phys. 86(4), 1391–1452 (2014). [CrossRef]  

5. A. Ashkin and J. P. Gordon, “Stability of radiation-pressure particle traps: an optical Earnshaw theorem,” Opt. Lett. 8(10), 511–513 (1983). [CrossRef]   [PubMed]  

6. M. Nieto-Vesperinas, J. J. Sáenz, R. Gómez-Medina, and L. Chantada, “Optical forces on small magnetodielectric particles,” Opt. Express 18(11), 11428–11443 (2010). [CrossRef]   [PubMed]  

7. A. Y. Bekshaev, “Subwavelength particles in an inhomogeneous light field: optical forces associated with the spin and orbital energy flows,” J. Opt. 15(4), 044004 (2013). [CrossRef]  

8. A. Canaguier-Durand, A. Cuche, C. Genet, and T. W. Ebbesen, “Force and torque on an electric dipole by spinning light fields,” Phys. Rev. A 88(3), 033831 (2013). [CrossRef]  

9. K. Y. Bliokh, Y. S. Kivshar, and F. Nori, “Magnetoelectric effects in local light-matter interactions,” Phys. Rev. Lett. 113(3), 033601 (2014). [CrossRef]   [PubMed]  

10. M. Padgett, J. Courtial, and L. Allen, “Light’s orbital angular momentum,” Phys. Today 57(5), 35–40 (2004). [CrossRef]  

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

12. M. Li, S. Yan, B. Yao, M. Lei, Y. Yang, J. Min, and D. Dan, “Intrinsic optical torque of cylindrical vector beams on Rayleigh absorptive spherical particles,” J. Opt. Soc. Am. A 31(8), 1710–1715 (2014). [CrossRef]   [PubMed]  

13. A. Lehmuskero, R. Ogier, T. Gschneidtner, P. Johansson, and M. Käll, “Ultrafast spinning of gold nanoparticles in water using circularly polarized light,” Nano Lett. 13(7), 3129–3134 (2013). [CrossRef]   [PubMed]  

14. Z. Yan and N. F. Scherer, “Optical vortex induced rotation of silver nanowires,” J. Phys. Chem. Lett. 4(17), 2937–2942 (2013). [CrossRef]  

15. A. Lehmuskero, Y. Li, P. Johansson, and M. Käll, “Plasmonic particles set into fast orbital motion by an optical vortex beam,” Opt. Express 22(4), 4349–4356 (2014). [CrossRef]   [PubMed]  

16. W.-Y. Tsai, J.-S. Huang, and C.-B. Huang, “Selective trapping or rotation of isotropic dielectric microparticles by optical near field in a plasmonic archimedes spiral,” Nano Lett. 14(2), 547–552 (2014). [CrossRef]   [PubMed]  

17. Y. Cao, T. Zhu, H. Lv, and W. Ding, “Spin-controlled orbital motion in tightly focused high-order Laguerre-Gaussian beams,” Opt. Express 24(4), 3377–3384 (2016). [CrossRef]   [PubMed]  

18. K. Y. Bliokh, A. Y. Bekshaev, and F. Nori, “Extraordinary momentum and spin in evanescent waves,” Nat. Commun. 5, 3300 (2014). [CrossRef]   [PubMed]  

19. A. Y. Bekshaev, K. Y. Bliokh, and F. Nori, “Transverse spin and momentum in two-wave interference,” Phys. Rev. X 5(1), 011039 (2015). [CrossRef]  

20. M. Neugebauer, T. Bauer, A. Aiello, and P. Banzer, “Measuring the transverse spin density of light,” Phys. Rev. Lett. 114(6), 063901 (2015). [CrossRef]   [PubMed]  

21. A. Canaguier-Durand and C. Genet, “Transverse spinning of a sphere in a plasmonic field,” Phys. Rev. A 89(3), 033841 (2014). [CrossRef]  

22. K.-Y. Kim and S. Kim, “Spinning of a submicron sphere by Airy beams,” Opt. Lett. 41(1), 135–138 (2016). [CrossRef]   [PubMed]  

23. K. Youngworth and T. Brown, “Focusing of high numerical aperture cylindrical-vector beams,” Opt. Express 7(2), 77–87 (2000). [CrossRef]   [PubMed]  

24. L. Allen, M. W. Beijersbergen, R. J. Spreeuw, and J. P. Woerdman, “Orbital angular momentum of light and the transformation of Laguerre-Gaussian laser modes,” Phys. Rev. A 45(11), 8185–8189 (1992). [CrossRef]   [PubMed]  

25. M. Šiler, P. Jákl, O. Brzobohatý, and P. Zemánek, “Optical forces induced behavior of a particle in a non-diffracting vortex beam,” Opt. Express 20(22), 24304–24319 (2012). [CrossRef]   [PubMed]  

26. D. B. Ruffner and D. G. Grier, “Optical forces and torques in nonuniform beams of light,” Phys. Rev. Lett. 108(17), 173602 (2012). [CrossRef]   [PubMed]  

27. P. Waterman, “Matrix formulation of electromagnetic scattering,” Proc. IEEE 53(8), 805–812 (1965). [CrossRef]  

28. M. I. Mishchenko, L. D. Travis, and A. A. Lacis, Scattering, Absorption, and Emission of Light by Small Particles (Cambridge University, 2002).

29. J. Barton, D. Alexander, and S. Schaub, “Theoretical determination of net radiation force and torque for a spherical particle illuminated by a focused laser beam,” J. Appl. Phys. 66(10), 4594–4602 (1989). [CrossRef]  

30. B. Richards and E. Wolf, “Electromagnetic diffraction in optical systems. II. Structure of the image field in an aplanatic system,” Proc. R. Soc. Ser. A 253(1274), 358–379 (1959).

31. M. Li, S. Yan, B. Yao, Y. Liang, M. Lei, and Y. Yang, “Optically induced rotation of Rayleigh particles by vortex beams with different states of polarization,” Phys. Lett. A 380(1), 311–315 (2016). [CrossRef]  

32. Q. Zhan, “Cylindrical vector beams: from mathematical concepts to applications,” Adv. Opt. Photonics 1(1), 1–57 (2009). [CrossRef]  

33. Y. Zhao, J. S. Edgar, G. D. Jeffries, D. McGloin, and D. T. Chiu, “Spin-to-orbital angular momentum conversion in a strongly focused optical beam,” Phys. Rev. Lett. 99(7), 073901 (2007). [CrossRef]   [PubMed]  

34. S. Sato and Y. Kozawa, “Hollow vortex beams,” J. Opt. Soc. Am. A 26(1), 142–146 (2009). [CrossRef]   [PubMed]  

35. M. Berry, “Optical currents,” J. Opt. A, Pure Appl. Opt. 11(9), 094001 (2009). [CrossRef]  

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 Line scans along the x-axis of intensity distribution for circularly polarized vortex beams (CPV) and radially polarized vortex beams (RPV) with topological charges of m =+ 5 or −5.
Fig. 2
Fig. 2 Radial trapping force curves on the particle under (a) CPV and (b) RPV inputs with topological charges of m =+ 5 or −5.
Fig. 3
Fig. 3 Transverse force distributions experienced by the particle in the focal plane illuminated, respectively, by CPV inputs with topological charges of (a) m = 5 and (c) m = −5, and by RPV inputs with (b) m = 5 and (d) m = −5.
Fig. 4
Fig. 4 Sketch of the spin and orbital motion of the particle illuminated, respectively, by CPV inputs with topological charges of (a) m = 5 and (c) m = −5, and by RPV inputs with (b) m = 5 and (d) m = −5.
Fig. 5
Fig. 5 The changes of the azimuthal forces (a1)-(a3), the spin torques (b1)-(b3) and the orbital torques (c1)-(c3) exerted on the particle with varying radius a (a1)-(c1), and absorptivity nʺ (a2)-(c2), as well as with the topological charge value m (a3)-(c3) in CPV and RPV inputs. R in the inset of (c1) represents the radius of the equilibrium position of the particle.

Tables (1)

Tables Icon

Table 1 Spin and orbital torques exerted on the particle at the equilibrium position under CPV and RPV inputs with topological charge of m = 5 or −5.

Equations (9)

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

F = S n ^ T ¯ d s ,
Γ s = S n ^ ( T ¯ × r ) d s ,
T ¯ = ε 1 E E + μ 1 H H 1 2 ( ε 1 E 2 + μ 1 H 2 ) I ¯ ,
E i n c ( r ) = n = 1 m = n n [ a m n M m n 1 ( k r ) + b m n N m n 1 ( k r ) ] ,
E s c a ( r ) = n = 1 m = n n [ p m n M m n 3 ( k r ) + q m n N m n 3 ( k r ) ] ,
[ p m n q m n ] = [ T m n m ' n ' 11 0 0 T m n m ' n ' 22 ] [ a m ' n ' b m ' n ' ] .
Γ o , z = ρ o F ϕ ,
E i n c ( r ) = i k f 2 π 0 θ m a x 0 2 π A ( θ , ϕ ) exp ( i k r ) s i n θ d ϕ d θ .
A ( θ , ϕ ) = ( cos θ ) 1 / 2 [ e θ 0 0 e ϕ ] ( A 0 ρ A 0 ϕ ) ,
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.