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

Synthesis of multiple longitudinal polarization vortex structures and its application in sorting chiral nanoparticles

Open Access Open Access

Abstract

As the essential properties of organisms, detection and characterization of chirality are of supreme importance in physiology and pharmacology. In this work, we propose an optical technique to sort chiral materials by use of longitudinal polarization vortex (LPV) structures, which is generated with tightly focusing Pancharatnam-Berry tailored Laguerre-Gaussian beam. The nonparaxial propagation of the focusing field leads to the creation of multiple pairs of dual LPV structures with arbitrary topological charge and location, which can be independently controlled by the spatial phase modulation applied on the illumination. More importantly, the opposite spin angular momentums carried by each pair of dual foci lead to different energy flow directions, making it suitable to sort nanoparticles by their handedness. In addition, the LPV structures would also bring different dynamic behaviors to the enantiomers, providing a feasible route toward all-optical enantiopure chemical syntheses and enantiomer separations in pharmaceuticals.

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

1. Introduction

As reflected in the properties of the electric field, a light beam may possess not only linear momentum but also angular momentum (AM), the latter of which can be further classified as spin angular momentum (SAM) and orbital angular momentum (OAM). The intrinsic SAM is associated with the polarization state of light beam, where σ± = ±1 correspond to the right- and left-hand (RH and LH) circular polarizations respectively [1], while OAM arises from the spatial structure of the optical field which can take arbitrary integer values. It can be characterized by the intrinsic OAM associated with inhomogeneous transversal phase distributions, commonly referring to the spiral phase gradient [2,3], and the extrinsic one relating to the beam trajectory [46]. Usually, it is considered that OAM and SAM are independent in a uniform and isotropic transparent medium. However, if an incident spin photon encounters a structure with anisotropic inhomogeneous boundaries, there would be spin-orbit interaction and leads to interesting phenomena such as optical spin-Hall effect [7,8]. In addition, the mutual conversion between SAM and OAM of photons can be expected in a high numerical aperture (NA) focusing system [911]. For example, when a circularly polarized vortex beam is tightly focused by a high NA lens, the spin-to-orbit conversion (SOC) would result in a strong longitudinal component of electric field carrying spiral phase at the focal region [12], which has plenty of attractive applications in fluorescent imaging, Raman spectroscopy and optical tweezers [13,14].

The phenomenon of SOC can be observed through the momentum transfer between optical field and nanoparticles [15]. When interacting with a light beam, the nanoparticle would inherit the SAM or OAM of the light, giving rise to the generation of optical forces that drag the nanoparticle to rotate around its own axis [16] or the optical axis [13,17] (namely spin or orbital motion). Since Ashkin and colleagues reported the first stable three-dimensional (3D) optical trapping by using the radiation pressure from a single focused laser beam [18], the continuous development of optical tweezers has become an important tool for research in biology, physical chemistry and soft matter physics [1921]. Usually, a tightly focused light beam is necessary in an optical tweezers, since the large gradient force would stably confine the nanoparticles in the location that having the largest intensity. In order to manipulate various nanoparticles in separated regions simultaneously, focal fields with multiple foci are necessary, which can be realized by tailoring the spatial distribution of an incident vector light. Most of the previous works focus on the generation of identical multiple focal spots with specific spatial intensity distribution (such as spherical solid/hollow spot [2225], optical bubble [26], optical chain [2730], optical needle [31], etc.). Recently, dual coaxial focal spots with different AMs have been reported to rotate nanoparticles located on the optical axis [32]. However, the full control of the focal field in terms of number of focal spots, SAM, OAM, and location in 3D space is still a challenging task. Due to the noncontact and holding nature of optical tweezers, the optical sorting of chiral materials has attracted increasing attentions in recent years. Chirality refers to the geometric property that a substance cannot overlap with its mirror image through translation and rotation operations [33]. The mirror images of a chiral structure are enantiomers, which are widely existed in various macro- and micro- structures. The chiral detection of substances has important applications in pharmacology, toxicology and pharmacokinetics [3436], since the change of the original chirality of a biomolecule may lead many serious diseases [37,38]. Different from conventional materials, the optical response of chiral nanoparticle not only depends on the electric and magnetic polarizability, but also has close connection to the electromagnetic/magnetoelectric polarizability (also known as chiral polarizability). The interaction between light field and chiral nanoparticles leads to many novel optical force effects, such as attractive/repulsive force [39,40], lateral optical force [4143], azimuthal/longitudinal optical torque [44,45], which are capable of sorting chiral nanoparticles. In this work, we firstly demonstrate an effective method to generate dual longitudinal polarization vortex (LPV) structure in the transverse plane of the focal region. By sculpting the spatial distribution of a light field at the pupil plane of a high NA lens, the focal field would be split into two focal spots with arbitrary OAM and opposite SAMs. Then this design principle is extended to the creation and manipulation of multiple pairs of dual LPV structures with prescribed locations in 3D space. Furthermore, the feasibility of sorting enantiomers with the use of LPV structure has been validated. The numerical calculation results show that the chiral nanoparticles would be separated by their handedness, while the dynamic behavior of each type of enantiomers can be adjusted individually, providing a feasible route toward all-optical enantiopure chemical syntheses and enantiomers separation in pharmaceuticals.

2. Generation of multiple pairs of dual foci with arbitrary LPV structure in 3D space

Considering a structured optical field with inhomogeneous distribution of states of polarization (SoP), its local electric field can be expressed in Cartesian coordinate system as:

$${{\boldsymbol E}_0}\textrm{(}x\textrm{,}y\textrm{)} = {A_0}\textrm{(}x\textrm{,}y\textrm{)}[\cos {\xi _0}\textrm{(}x\textrm{,}y\textrm{)}{{\boldsymbol e}_x} + {e^{ - i\Delta \theta \textrm{(}x\textrm{,}y\textrm{)}}}\sin {\xi _0}\textrm{(}x\textrm{,}y\textrm{)}{{\boldsymbol e}_y}] = {A_0}[{e^{i{\xi _0}\textrm{(}x\textrm{,}y\textrm{)}}}{{\boldsymbol e}_1} + {e^{ - i{\xi _0}\textrm{(}x\textrm{,}y\textrm{)}}}{{\boldsymbol e}_2}]/2,$$
where A0 represents the amplitude and ξ0 denotes the direction of the polarization orientation, while Δθ indicates the phase difference between x and y components of the electric field. Furthermore, a pair of orthogonal base vector (e1, e2) can be applied to analyze the polarization structure of the optical field with e1 = ex − ei(π/2−Δθ) ey and e2 = ex + ei(π/2−Δθ) ey. From Eq. (1), one can easily find that a complex optical field can be decomposed into two orthogonal polarizations with opposite phase terms, which are also related to the spatial coordinate. For example, a locally linear polarization (Δθ = 0) is a combination of both the RH and LH circular polarizations. In addition, it is known that the spatial phase difference of ξ0(x, y) between e1 and e2 would lead to the splitting of the optical field under the tight focusing condition, and two axial separated focal spots would be obtained in the focal volume with orthogonal polarization states. Here, we will study the effects of spatial phase distribution of the illumination on the focal field splitting phenomenon and demonstrate the feasibility of creating multiple pairs of dual foci with controllable AM in 3D space.

Figure 1 illustrates the conceptual scheme of generation of multiple focal spots with arbitrary LPV structure. The incident light is firstly modulated by a phase mask then focused by an objective lens with NA of 0.95. Without loss of generality, we consider a generalized linearly polarized Laguerre-Gaussian (LG) vortex beam with λ = 532 nm and field function l0(θ) = (21/2sinθ/sinα)|l|exp(−sin2θ/sin2α)expilϕ, where l denotes the topological charge (TC), α is the maximum focusing angle determined by the NA of lens, and ϕ is the azimuthal angle. According to the Richards-Wolf vector diffraction theory, the electric field in the vicinity of the focus of the lens can be calculated by [46,47]:

$${\boldsymbol E}(x,y,z) ={-} \frac{{ikf}}{{2\pi }}\int\limits_0^\alpha {\int\limits_0^{2\pi } {{{\boldsymbol E}_1}} (\theta ,\phi ){e^{ik({\boldsymbol s}\cdot {\boldsymbol r})}}} \sin \theta d\phi d\theta .$$

Here k = n1k0 with k0 = 2π being the wave vector in free space and n1 being the refractive index of the surrounding medium, f is the focal distance, s is the unit vector along the propagation direction of the ray in image space and r is the position vector from the focal spot to the observation point. Usually, the incident light would be focused by the lens with focal point O(0, 0, 0). However, if we assume that the focal field is split into two separated foci centered at O'x, 0, 0) and O''(−Δx, 0, 0), it has:

$${\boldsymbol s} \cdot ({\boldsymbol r} - {{\boldsymbol r}_{o{o^{\prime}}}}) = {\boldsymbol s} \cdot {\boldsymbol r} - {\boldsymbol s} \cdot {{\boldsymbol r}_{o{o^{\prime}}}} ={-} \rho \sin \theta \cos (\phi - \varphi ) + z\cos \theta \mp \Delta x\sin \theta \cos \phi .$$

 figure: Fig. 1.

Fig. 1. Schematic diagram of setup for creating dual LPV structure. The spatial phase distribution of an incident LG mode light is modulated by a phase mask in the pupil plane. Under the tight focusing condition, components of the RH and LH vibrations of the illumination are focused into separated focal spots along x axis, enabling the arbitrary TC for the longitudinal field of each focal spot.

Download Full Size | PDF

Consequently, the required spatial phase modulation applied in the pupil plane can be expressed as:

$$\xi (\theta ,\phi ) ={-} \Delta x\sin \theta \cos \phi + m\phi ,$$
where θ is the focusing angle, is the introduced Pancharatnam-Berry (PB) phase [48,49]. In this case, apodized field E1 after refraction can be expressed as:
$${{\boldsymbol E}_1}(\theta ,\phi ) = {l_0}(\theta )[{e^{i{\kern 1pt} \xi (\theta ,{\kern 1pt} \phi )}}{{\boldsymbol e}_r} + {e^{ - i{\kern 1pt} \xi (\theta ,{\kern 1pt} \phi )}}{{\boldsymbol e}_l}].$$
where er and el are the RH and LH circular polarization basis vector, respectively. It can be seen that the refracted light field can be decomposed into the RH and LH circularly polarized components with opposite phase modulation functions, leading to the creation of dual foci with different spins and locations that are symmetrical with respect to the origin. In this case, the focal field given in Eq. (2) can be rewritten as:
$$\begin{aligned}{{\boldsymbol E}_x} &={-} \frac{{ikf}}{{2\pi }}\int\limits_0^\alpha {\int\limits_0^{2\pi } {{l_0}} (\theta )\sin \theta \sqrt {\cos \theta } {e^{ik[ - \rho \sin \theta \cos (\phi - \varphi ) + z\cos \theta ]}}} \\ & \quad \times [\sin (\phi - \varPsi )\sin \phi + \cos (\phi - \varPsi )\cos \theta \cos \phi ]d\phi d\theta ,\end{aligned}$$
$$\begin{aligned}{{\boldsymbol E}_y} &={-} \frac{{ikf}}{{2\pi }}\int\limits_0^\alpha {\int\limits_0^{2\pi } {{l_0}} (\theta )\sin \theta \sqrt {\cos \theta } {e^{ik[ - \rho \sin \theta \cos (\phi - \varphi ) + z\cos \theta ]}}} \\ & \quad \times [ - \sin (\phi - \varPsi )\cos \phi + \cos (\phi - \varPsi )\cos \theta \sin \phi ]d\phi d\theta ,\end{aligned}$$
$${{\boldsymbol E}_z} ={-} \frac{{ikf}}{{2\pi }}\int\limits_0^\alpha {\int\limits_0^{2\pi } {{l_0}} (\theta ){{\sin }^2}\theta \sqrt {\cos \theta } {e^{ik[ - \rho \sin \theta \cos (\phi - \varphi ) + z\cos \theta ]}}} \cos (\phi - \varPsi )d\phi d\theta ,$$
where Ψ = −kΔxsinθ cosϕ+mϕ. By using identity involving Bessel function, Eq. (8) can be simplified to:
$$\begin{array}{l} {{\boldsymbol E}_z} ={-} \frac{{ikf}}{{2\pi }}\int\limits_0^\alpha {\sqrt {\cos \theta } {{\left( {\frac{{\sqrt 2 \beta \sin \theta }}{{\sin \alpha }}} \right)}^{|l|}}{e^{\left( { - \frac{{{\beta^2}{{\sin }^2}\theta }}{{{{\sin }^2}\alpha }}} \right)}}} [{i^{l + m - 1}}{e^{i(l + m - 1)\varphi }}{J_{l + m - 1}}( - k(\rho \cos \varphi - \Delta x)\sin \theta )\\ \;\;\;\;\;\;\; + {i^{l - m + 1}}{e^{i(l - m + 1)\varphi }}{J_{l - m + 1}}( - k(\rho \cos \varphi - \Delta x)\sin \theta )]{e^{ikz\cos \theta }}{\sin ^2}\theta d\theta , \end{array}$$
where Jn(x) represents the nth order Bessel function of the first kind. It is clearly shown that the longitudinal electric field Ez has two different phase vortices located at ±Dx with TC of l+m−1 and l − m+1, respectively. By changing the TC of illumination (l) and PB phase applied at the input pupil (m), OAM of the longitudinal component of the two foci can be tuned. In addition, the spacing distance 2Δx between them can be adjusted by tailoring the phase modulation in Eq. (4).

To verify the theoretical prediction, we employ LG beam with l = 1 and set phase modulation according to Eq. (4) with (m, Δx) = (−2, 3λ). Based on the above analysis, it is expecting that two foci with TC l+m–1 = −2 and l − m+1 = 4 in the Ez component at x = −3λ and 3λ, respectively. The required spatial phase distribution modulated in the pupil plane is shown in Fig. 2(a), and the corresponding total, transverse and longitudinal intensity distributions of the focal field are presented in Fig. 2(b)–2(d), showing donut shape with zero amplitude on the optical axis. As shown in Fig. 2(b), the left and right foci are located at r1 = (−3λ, 0, 0) and r2 = (3λ, 0, 0) respectively in the focal region. In addition, polarization map superimposed on a zoom-in intensity distribution is also presented as the inset of Fig. 2(b) to demonstrate that the transverse electric field possesses circular polarization with opposite handedness within the main lobe, where blue and orange colors represent the RH and LH, respectively. Consequently, it is demonstrated that left and right focal fields are contributed by the input component of the RH and LH vibration respectively, leading to the different energy flow directions. Specially, the phase distributions of the longitudinal component of the focal field is calculated and presented in the inset of Fig. 2(d). The TC of Ez can be visualized from the phase profile, which are found to be −2 and 4 for left and right focal spots respectively, showing good agreement between theoretical predications and numerical calculations.

 figure: Fig. 2.

Fig. 2. Lateral dual foci created by a tightly focused LG illumination with (l, m, Δx) = (1, −1, 3λ). (a) Spatial phase modulation of the phase mask in the pupil plane. (b) Total, (c) transverse and (d) longitudinal components of the intensity distributions of the focal field. The insets in (b) and (d) illustrate SoP and phase map within the main lobe, respectively.

Download Full Size | PDF

Furthermore, this strategy can be easily adapted to the separation of multiple pairs of LPV structure in 3D space. Assuming the dual foci are located at O'x, Δy, Δz) and O''(−Δx, −Δy, −Δz), Eq. (3) would be rewritten as:

$${\boldsymbol s} \cdot ({\boldsymbol r} - {{\boldsymbol r}_{o{o^{\prime}}}}){\kern 1pt} ={-} \rho \sin \theta \cos (\phi - \varphi ) + z\cos \theta - ({\pm} \Delta x\sin \theta \cos \phi \pm \Delta y\sin \theta \sin \phi \pm \Delta z\cos \theta ),$$
which allows us to move the foci freely in the image space. In addition, if n pairs of dual foci are required to be generated simultaneously, the corresponding phase modulation function P(θ, ϕ) is expressed as [22]
$$P(\theta ,\phi ) = \sum\limits_{j = 1}^n {\textrm{exp} [ - ik(\Delta {x_j}\sin \theta \cos \phi + \Delta {y_j}\sin \theta \sin \phi + \Delta {z_j}\cos \theta )]} \textrm{,}$$

Then, the electric field can be obtained from Eq. (68) as:

$$\begin{aligned}\left[ {\begin{array}{{c}} {{{\boldsymbol E}_x}}\\ {{{\boldsymbol E}_y}}\\ {{{\boldsymbol E}_z}} \end{array}} \right] &={-} \frac{{ikf}}{{2\pi }}\int\limits_0^\alpha {\int\limits_0^{2\pi } {\sin \theta \sqrt {\cos \theta } {e^{ik[ - \rho \sin \theta \cos (\phi - \varphi ) + z\cos \theta ]}} \sum\limits_{j = 1}^n {{{\left( {\frac{{\sqrt 2 \beta \sin \theta }}{{\sin \alpha }}} \right)}^{|{l_j}|}}{e^{\left( { - \frac{{{\beta^2}{{\sin }^2}\theta }}{{{{\sin }^2}\alpha }}} \right)}}} } } {e^{i{l_j}\phi }}\\ & \quad \times\left[ {\begin{array}{{c}} {{{\boldsymbol V}_{{E_x}_{_j}}}}\\ {{{\boldsymbol V}_{{E_y}_{_j}}}}\\ {{{\boldsymbol V}_{{E_z}_{_j}}}} \end{array}} \right]d\phi d\theta \end{aligned}$$

The electric polarization vectors VE with three components can be expressed as

$$\left\{ {\begin{array}{{c}} {{{\boldsymbol V}_{{E_x}_{_j}}} = \sin (\phi - {\Psi_j})\sin \phi + \cos (\phi - {\Psi_j})\cos \theta \cos \phi \textrm{,}}\\ {{{\boldsymbol V}_{{E_y}_{_j}}} ={-} \sin (\phi - {\Psi_j})\cos \phi + \cos (\phi - {\Psi_j})\cos \theta \sin \phi ,}\\ {{{\boldsymbol V}_{{E_z}_{_j}}} = \cos (\phi - {\Psi_j})\sin \theta ,} \end{array}} \right.$$
where Ψj = −kxj sinθ cosϕyj sinθ sinϕzj cosθ)+mjϕ.

Consequently, the theoretical derivations indicate that when addition degree of freedom (Δx and Δy) are introduced, arbitrary number of paired LPV structure with independent TC can be achieved at arbitrary location in the focal volume. To prove it, we consider a complex input beam with parameters given in Tab. 1, and the corresponding spatial phase modulation is presented in Fig. 3(a). The simulated electric field intensities in the focal plane show six focal spots for the total (Fig. 3(b)) and longitudinal (Fig. 3(d)) fields, in which the locations of each pair of dual foci is symmetrical about the origin. It can be seen that the longitudinal component is quite different from the total field, which exhibit solid or donut-shaped profiles but different sizes. In this case, the six foci should have different TCs limi +1 = (3, 2, 4) and li + mi −1 = (−1, −2, 0), which can be found out by the phase map in the inset of Fig. 3(d). As the Stokes image shown in Fig. 3(c), each pair of foci still hold opposite handedness. Moreover, the tunable separation between foci in 3D space is considered. With the spatial phase modulation given in Tab. 2 and Fig. 4(a), the focal field consists of two pairs of foci with TC of (3, −1) and (2, −2) which are located at (±3λ, ±3λ, ±3λ) and (±3λ, ${\mp}$ 3λ, ±3λ), respectively. Consequently, we demonstrate that this simple method is capable to generate and separate multiple LPV structures.

 figure: Fig. 3.

Fig. 3. Three pairs of dual foci created by a tightly focused LG illumination in 2D space. (a) Spatial phase modulation of the phase mask in the pupil plane. (b) Total, (c) stokes parameter S3 and (d) longitudinal components of the intensity distributions of the focal field. The insets in (d) illustrates the phase distribution of the main lobe. The intensities of several foci have been multiplied by a factor of 3 in order to obtain a clear image.

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. Two pairs of dual foci created by a tightly focused LG illumination in 3D space. (a) Spatial phase modulation of the phase mask in the pupil plane. (b) Total intensity distributions of the focal field.

Download Full Size | PDF

Tables Icon

Table 1. Parameters of multiple pairs of dual foci in transverse plane

Tables Icon

Table 2. Parameters of multiple pairs of dual foci in 3D space

3. Optical force and torque under dipole approximation

Next, we would like to discuss the strategy to sort chiral nanoparticles by using these LPV structures. Assuming the size of chiral nanoparticles are much smaller than the wavelength of the illumination, dipole approximation can be applied to study the optical force exerted on the nanoparticle. By introducing chiral nanoparticles into the focal region shown in Fig. 1, its polarizability elements can be expressed as [44,45,50]:

$${\alpha _{ee}} = \frac{{i6\pi {\varepsilon _0}{\varepsilon _m}}}{{{k^3}}}{a_1},{\alpha _{mm}} = \frac{{i6\pi {\mu _m}}}{{{\mu _0}{k^3}}}{b_1},{\alpha _{em}} ={-} \frac{{6\pi {n_1}}}{{{Z_0}{k^3}}}{c_1},$$
where ɛ0 and μ0 are the permittivity and permeability in vacuum, ɛm and μm are the relative permittivity and permeability of the surrounding medium. Z0 = (m0/e0)1/2 is the wave impedance in free space. The corresponding Mie scattering coefficients (a1, b1, c1) can be expressed in terms of the vector spherical wave functions:
$$\begin{array}{l} {a_n} = [{A_n^{(2 )}V_n^{(1 )} + A_n^{(1 )}V_n^{(2 )}} ]{Q_n},\\ {b_n} = [{B_n^{(1 )}W_n^{(2 )} + B_n^{(2 )}W_n^{(1 )}} ]{Q_n}, \\ {c_n} = [{A_n^{(1 )}W_n^{(2 )} - A_n^{(2 )}W_n^{(1 )}} ]{Q_n}, \end{array}$$
with
$$\begin{array}{l} A_n^{(j )} = {Z_S}D_n^{(1 )}({{x_j}} )- D_n^{(1 )}({{x_0}} ),B_n^{(j )} = D_n^{(1 )}({{x_j}} )- {Z_S}D_n^{(1 )}({{x_0}} ),\\ W_n^{(j )} = {Z_S}D_n^{(1 )}({{x_j}} )- D_n^{(3 )}({{x_0}} ),V_n^{(j )} = D_n^{(1 )}({{x_j}} )- {Z_S}D_n^{(3 )}({{x_0}} ),\\ {Q_n} = \frac{{{{{\psi _n}({{x_0}} )} / {{\xi _n}({{x_0}} )}}}}{{V_n^{(1 )}W_n^{(2 )} + V_n^{(2 )}W_n^{(1 )}}}. \end{array}$$

Here x0 = k0r, x1 = k1r, x2 = k2r with r being the nanoparticle radius, ${k_1} = {k_0}(\sqrt {{\varepsilon _r}{\mu _r}} + \kappa )/\sqrt {{\varepsilon _m}{\mu _m}}$ and ${k_2} = {k_0}(\sqrt {{\varepsilon _r}{\mu _r}} - \kappa )/\sqrt {{\varepsilon _m}{\mu _m}} $ being the wave numbers in chiral medium and κ denotes the chirality parameter. Zs = (mrem/er)1/2 is the wave impedance of the nanoparticle. ξn(x) and ψn(x) are the Riccati-Bessel functions of the first and third kinds, respectively. Dn(1)(x) = ψ'n(x)/ψn(x) and Dn(3)(x) = ξ'n(x)/ξn(x) are the corresponding logarithmic derivatives.

Based on the dipole approximation, the time-averaged optical force < F > exerted on a chiral nanoparticle is given by [45,50]:

$$\left\langle {\boldsymbol F} \right\rangle = {F_{chiral}} + {F_{achiral}} = (F_{grad}^c + {F_{rad}} + {F_{vor}} + {F_{spin}}) + (F_{grad}^a + {F_{curl}} + {F_{flow}}),$$
where the expressions of the force components can be found in Tab. 3. Note that c is the speed of light in vacuum and ω = ck0 is the angular frequency of light, 〈S〉 = 1/2Re[E × H*] denotes the time-averaged Poynting vector, 〈Le〉 = Re[ɛ0ɛm/(4)E × E*] and 〈Lm〉 = Re[μ0μm/(4)H × H*] represent the electric and magnetic parts of the time-averaged SAM densities. Cext = Ce+Cm is a sum of contribution from the electric and magnetic dipole channel with Ce = kIm[αee]/ɛ0ɛm and Cm = 0Im[αee]/μm, respectively. Crecoil = −k4μ0/(6πɛ0n12) (Re[αeeα*mm]+|αem|2) describes the recoil force related to the asymmetry parameter. Fgrad and Frad are the optical force correspond to the gradient force and the radiation pressure; Fcurl is referred to the curl-spin force associated with the curl of the SAM densities [51]; Fvor represents the vortex force determined by the energy flow vortex and nanoparticle chirality [52]; Fspin describe the spin density force, which designate a direct coupling of the nanoparticle to the SAM density of the incident wave; Fflow is due to an alternating flow of the “stored energy” [53].

Tables Icon

Table 3. Concrete expressions of the optical force

Besides optical force, optical torque is another important quantity that directly linked to the movement of the nanoparticle [44]:

$$\begin{array}{l} \left\langle {\boldsymbol T} \right\rangle = \left[ { - 2{\mu_0}Re ({\alpha_{em}}) + \frac{{{\mu_0}{k^3}}}{{3\pi {\varepsilon_0}{\varepsilon_m}}}{\mathop{\rm Im}\nolimits} ({\alpha_{ee}}\alpha_{em}^\ast ) + \frac{{\mu_0^2{k^3}}}{{3\pi }}{\mathop{\rm Im}\nolimits} ({\alpha_{mm}}\alpha_{em}^ \ast )} \right]\left\langle {\boldsymbol S} \right\rangle \\ \;\;\;\;\;\;\;\; + \left[ {\frac{{{\mu_0}{k^3}}}{{6\pi {\varepsilon_0}{\varepsilon_m}}}Re ({\alpha_{ee}}\alpha_{em}^\ast ) - \frac{{\mu_0^2{k^3}}}{{6\pi }}Re ({\alpha_{mm}}\alpha_{em}^ \ast )} \right]{\mathop{\rm Im}\nolimits} ({{\boldsymbol E} \times {{\boldsymbol H}^ \ast }} )\\ \;\;\;\;\;\;\;\; - \left[ {\frac{{2\omega }}{{{\varepsilon_0}{\varepsilon_m}}}{\mathop{\rm Im}\nolimits} ({\alpha_{ee}}) - \frac{{\omega {k^3}}}{{3\pi \varepsilon_0^2\varepsilon_m^2}}{\alpha_{ee}}\alpha_{ee}^\ast{-} \frac{{\omega {\mu_0}{k^3}}}{{3\pi {\varepsilon_0}{\varepsilon_m}}}{\alpha_{em}}\alpha_{em}^\ast } \right]\left\langle {{{\boldsymbol L}_e}} \right\rangle \\ \;\;\;\;\;\;\;\; - \left[ {\frac{{2\omega {\mu_0}}}{{{\mu_m}}}{\mathop{\rm Im}\nolimits} ({\alpha_{mm}}) - \frac{{\omega \mu_0^2{k^3}}}{{3\pi {\mu_m}}}{\alpha_{mm}}\alpha_{mm}^\ast{-} \frac{{\omega {\mu_0}{k^3}}}{{3\pi {\varepsilon_0}{\varepsilon_m}{\mu_m}}}{\alpha_{em}}\alpha_{em}^\ast } \right]\left\langle {{{\boldsymbol L}_m}} \right\rangle . \end{array}$$

With Eq. (17) and Eq. (18), the dynamic behavior of the chiral nanoparticles can be precisely simulated. Considering enantiomers with (r, ɛr, κ) = (50 nm, 2.5 + 0.1i, ±0.5 ± 0.01i) immersed in water (n1 = 1.33) are placed in the focal volume presented in Fig. 2, the distributions of the transverse optical force within the dual longitudinal vortex are indicated by the arrows superimposed in Figs. 5(a) and 6(a). In this case, the input power of the illumination is assumed to be 100 mW. It can be clearly seen that chiral nanoparticles with κ = 0.5 + 0.01i and κ = −0.5 − 0.01i are confined within the main lobe of the right and left focal spots with RH and LH spin, respectively. In order to explore more details about the optical force effect, the line-scans of the components of transverse force through the center of the foci are shown in Figs. 5(b) and 6(b). In this case, the achiral force is dominated by the gradient force term ${\boldsymbol F}_{grad}^a$, which tends to trap nanoparticles at positions that have the largest intensity. As for chiral objects, the contribution from chiral force terms becomes nonnegligible, and the equilibrium position can only be formed when the achiral and chiral forces possess the similar trend. Since the direction of chiral force changes with the handedness of both the enantiomers and the focal field, the chiral nanoparticles would be separated and sorted by their chirality. The trapping stability can be demonstrated by the potential depth Ux as the orange curves in Figs. 5(c) and 6(c), which is estimated as the work done by the optical forces along x axis. It can reach 784.2 kBT and 1732 kBT for the equilibrium positions marked by asterisk, where kB is the Boltzmann constant and T = 300 K is the absolute temperature of the ambient. Traditionally, an optical trap with potential depth larger than kBT can be considered as stable [23]. It is worthy of noting that from the force distribution one may conclude that chiral nanoparticles with κ = 0.5 + 0.01i and κ = −0.5 − 0.01i seem to be confined at the dark center of the left and right focal spots respectively, however, the surrounding high potential barrier would prevent these nanoparticles from entering the central region. Although the lateral sorting is used as an example, it is worthy of noting that the chiral nanoparticles can be captured and rotated in any position of the three-dimensional plane by using this strategy.

 figure: Fig. 5.

Fig. 5. Chiral detection with the use of dual LPV structure illustrated in Fig. 2. For chiral nanoparticles with κ = 0.5 + 0.01i, (a) Intensity pattern superimposed with the distribution of the optical force, (b) line-scan of the total, chiral and achiral force along x-axis, (c) line-scan of the longitudinal optical torque and potential depth along x-axis, (d) trajectories of the nanoparticles in the transverse plane of the focal field. The initial positions of the nanoparticles are indicated by the coordinates.

Download Full Size | PDF

 figure: Fig. 6.

Fig. 6. Chiral detection with the use of dual LPV structure illustrated in Fig. 2. For chiral nanoparticles with κ = −0.5 − 0.01i, (a) Intensity pattern superimposed with the distribution of the optical force, (b) line-scan of the total, chiral and achiral force along x-axis, (c) line-scan of the longitudinal optical torque and potential depth along x-axis, (d) trajectories of the nanoparticles in the transverse plane of the focal field. The initial positions of the nanoparticles are indicated by the coordinates.

Download Full Size | PDF

So far, we demonstrate that the chiral nanoparticles can be well restricted along radial direction of the foci. Due to the axial symmetry of the focal spot, the optical force would be constant along the azimuthal direction, giving rise to the longitudinal optical torque (shown in Figs. 5(c) and 6(c)) that drags the nanoparticle rotating around the beam center. Considering the conservation law of AM in this closed physical system, the nanoparticle would inherit the OAM of the LPV structure, therefore its rotation direction is determined by the sign of TC. In this example, chiral nanoparticles with κ = 0.5 + 0.01i and κ = −0.5 − 0.01i would rotate counterclockwise and clockwise around the right and left focal spots, respectively. In order to elaborately describe the dynamic motion of the enantiomers affected by the longitudinal vortex structure, Langevin dynamics (LD) equation is adopted for each chiral nanoparticle [54,55]:

$$m\frac{{{d^2}{\boldsymbol R}}}{{d{t^2}}} = {\boldsymbol F}({\boldsymbol R},t) - \chi \frac{{d{\boldsymbol R}}}{{dt}} + {\boldsymbol \eta }\textrm{,}$$
where R is center position of the nanoparticle with mass m, F is the optical force, the friction coefficient χ = 6πvr is given by Stokes’ law with v = 0.89 mPa·s being the dynamic viscosity of water. η is the stochastic Brownian force satisfying a Gaussian probability distribution <ηi2> = 2kB, where i is the coordinate index.

To solve the LD equation, BAOAB method is applied due to its excellent computation speed and accuracy [56,57]. Figure 5(d) shows the trajectory of several chiral nanoparticles with κ = 0.5 + 0.01i and initial position randomly distributed in the focal plane. The total duration of the simulation is 50 μs (corresponding to 0.5 μs LD time steps for a LD time step of 100). It is clearly seen that the chiral nanoparticle would be bounced off by the left focal spot but be confined and rotating around the right focal spot with RH polarization. Similarly, for chiral nanoparticles with opposite handedness, the left vortex field with LH polarization is capable to sustain a stable trapping and rotating (shown in Fig. 6(d)). Note that the influence of Brownian motion to the trajectory of the nanoparticle is negligible due to the strong optical force, which can also be demonstrated by the calculated potential depth. Consequently, we demonstrate that the significantly different dynamic behaviors of chiral nanoparticles provide a feasible way to easily distinguish the type of enantiomers.

4. Conclusions

In conclusion, we proposed a method to realize dual foci with arbitrary LPV structures in the transverse plane, which is realized by modulating the PB phase of the incident LG beam at the pupil plane of a high NA lens. We also demonstrated that the strategy can be applied to create multiple pairs of dual foci in 3D space, both the location and the TC of each LPV structure can be independently controlled. Furthermore, the induced optical force and torque from the interaction between chiral nanoparticles and the dual LPV structures are calculated using dipole approximation. Numerical results demonstrated that the nanoparticle would experience a transverse chiral optical force with direction determined by the chirality of the material, consequently the enantiomers with different handedness would be trapped and separated by the dual LPV structures with opposite energy flow directions. Moreover, we demonstrated that there is distinct difference in the dynamic behaviors between the RH and LH chiral nanoparticles. It is worthy of noting that the working wavelength of the proposed sorting strategy can be conveniently adjusted. In addition, the high focusing efficiency of the optical trapping system enables to sustain the force stability with relatively low input power. Consequently, the thermal damage can be avoided by tuning the working wavelength and reducing the laser power. The application of this method may open new avenues for all-optical enantiopure chemical syntheses and enantiomers separation in pharmaceuticals.

Funding

National Natural Science Foundation of China (11504049, 11774055, 12074066, 92050202).

Acknowledgment

G. R. acknowledged the support by the Zhishan Young Scholar Program of Southeast University.

Disclosures

The authors declare that there are no conflicts of interest related to this article.

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. J. H. Poynting, “The wave-motion of a revolving shaft, and a suggestion as to the angular momentum in a beam of circularly polarized light,” Proc. R. Soc. Lond. A 82(557), 560–567 (1909). [CrossRef]  

2. L. Allen, M. W. Beijersbergen, R. J. C. 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]  

3. Y. Yang, G. Thirunavukkarasu, M. Babiker, and J. Yuan, “Orbital-angular-momentum mode selection by rotationally symmetric superposition of chiral states with application to electron vortex beams,” Phys. Rev. Lett. 119(9), 094802 (2017). [CrossRef]  

4. A. Bekshaev, K. Y. Bliokh, and M. Soskin, “Internal flows and energy circulation in light beams,” J. Opt. 13(5), 053001 (2011). [CrossRef]  

5. A. M. Yao and M. J. Padgett, “Orbital angular momentum: origins, behavior and applications,” Adv. Opt. Photonics 3(2), 161–204 (2011). [CrossRef]  

6. L. Allen, M. J. Padgett, and M. Babiker, “The orbital angular momentum of light,” Prog. Opt. 39, 291–372 (1999). [CrossRef]  

7. M. Onoda, S. Murakami, and N. Nagaosa, “Hall effect of light,” Phys. Rev. Lett. 93(8), 083901 (2004). [CrossRef]  

8. X. Ling, X. Zhou, K. Huang, Y. Liu, C. W. Qiu, H. Luo, and S. Wen, “Recent advances in the spin Hall effect of light,” Rep. Prog. Phys. 80(6), 066401 (2017). [CrossRef]  

9. O. G. Rodríguez-Herrera, D. Lara, K. Y. Bliokh, E. A. Ostrovskaya, and C. Dainty, “Optical nanoprobing via spin-orbit interaction of light,” Phys. Rev. Lett. 104(25), 253601 (2010). [CrossRef]  

10. Z. Bomzon, M. Gu, and J. Shamir, “Angular momentum and geometrical phases in tight-focused circularly polarized plane waves,” Appl. Phys. Lett. 89(24), 241104 (2006). [CrossRef]  

11. T. A. Nieminen, A. B. Stilgoe, N. R. Heckenberg, and H. Rubinsztein-Dunlop, “Angular momentum of a strongly focused Gaussian beam,” J. Opt. A: Pure Appl. Opt. 10(11), 115005 (2008). [CrossRef]  

12. Y. Zhao, J. S. Edgar, G. D. M. 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]  

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

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

15. Y. Zhao, D. Shapiro, D. Mcgloin, D. T. Chiu, and S. Marchesini, “Direct observation of the transfer of orbital angular momentum to metal particles from a focused circularly polarized Gaussian beam,” Opt. Express 17(25), 23316–23322 (2009). [CrossRef]  

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

17. L. Paterson, M. MacDonald, J. Arlt, W. Sibbett, P. Bryant, and K. Dholakia, “Controlled rotation of optically trapped microscopic particles,” Science 292(5518), 912–914 (2001). [CrossRef]  

18. A. Ashkin, J. Dziedzic, J. Bjorkholm, and S. Chu, “Observation of a single-beam gradient force optical trap for dielectric particles,” Opt. Lett. 11(5), 288–290 (1986). [CrossRef]  

19. L. Oroszi, P. Galajda, H. Kirei, S. Bottka, and P. Ormos, “Direct measurement of torque in an optical trap and its application to doublestrand DNA,” Phys. Rev. Lett. 97(5), 058301 (2006). [CrossRef]  

20. J. D. Wen, L. Lancaster, C. Hodges, A. C. Zeri, S. H. Yoshimura, H. F. Noller, C. Bustamante, and I. Tinoco, “Following translation by single ribosomes one codon at a time,” Nature 452(7187), 598–603 (2008). [CrossRef]  

21. A. Reiserer, C. Nölleke, S. Ritter, and G. Rempe, “Ground-state cooling of a single atom at the center of an optical cavity,” Phys. Rev. Lett. 110(22), 223003 (2013). [CrossRef]  

22. X. Wang, L. Gong, Z. Zhu, B. Gu, and Q. Zhan, “Creation of identical multiple focal spots with three-dimensional arbitrary shifting,” Opt. Express 25(15), 17737–17745 (2017). [CrossRef]  

23. G. Rui, Y. Wang, X. Wang, B. Gu, and Y. Cui, “Trapping of low-refractive-index nanoparticles in a hollow dark spherical spot,” J. Phys. Commun. 2(6), 065015 (2018). [CrossRef]  

24. W. Yan, S. Lin, H. Lin, Y. Shen, Z. Nie, B. Jia, and X. Deng, “Dynamic control of magnetization spot arrays with three-dimensional orientations,” Opt. Express 29(2), 961–973 (2021). [CrossRef]  

25. T. Mu, Z. Chen, S. Pacheco, R. Wu, C. Zhang, and R. Liang, “Generation of a controllable multifocal array from a modulated azimuthally polarized beam,” Opt. Lett. 41(2), 261–264 (2016). [CrossRef]  

26. Y. Zhao, Q. Zhan, and Y. Li, “Design of DOE for the control of optical “bubble” generated by highly focused radially polarized beam,” Proc. SPIE 5514, 616–625 (2004). [CrossRef]  

27. Y. Zhao, Q. Zhan, Y. Zhang, and Y. Li, “Creation of a three-dimensional optical chain for controllable particle delivery,” Opt. Lett. 30(8), 848–850 (2005). [CrossRef]  

28. J. Wang, W. Chen, and Q. Zhan, “Creation of uniform three-dimensional optical chain through tight focusing of space-variant polarized beams,” J. Opt. 14(5), 055004 (2012). [CrossRef]  

29. Y. Yu and Q. Zhan, “Generation of uniform three-dimensional optical chain with controllable characteristics,” J. Opt. 17(10), 105606 (2015). [CrossRef]  

30. X. Han and P. H. Jones, “Evanescent wave optical binding forces on spherical microparticles,” Opt. Lett. 40(17), 4042–4045 (2015). [CrossRef]  

31. J. Luo, H. Zhang, and S. Wang, “Three-dimensional magnetization needle arrays with controllable orientation,” Opt. Lett. 44(4), 727–730 (2019). [CrossRef]  

32. Z. Man, Z. Xi, X. Yuan, R. E. Burge, and H. P. Urbach, “Dual coaxial longitudinal polarization vortex structures,” Phys. Rev. Lett. 124(10), 103901 (2020). [CrossRef]  

33. A. Guijarro and M. Yus, The Origin of Chirality in the Molecules of Life: A Revision from Awareness to the Current Theories and Perspectives of this Unsolved Problem (Royal Society of Chemistry, 2008).

34. H. Caner, E. Groner, L. Levy, and I. Agranat, “Trends in the development of chiral drugs,” Drug Discovery Today 9(3), 105–110 (2004). [CrossRef]  

35. B. S. Sekhon, “Chiral pesticides,” J. Pestic. Sci. 34(1), 1–12 (2009). [CrossRef]  

36. W. H. Brooks, W. C. Guida, and K. G. Daniel, “The significance of chirality in drug design and development,” Curr. Top. Med. Chem. 11(7), 760–770 (2011). [CrossRef]  

37. L. A. Nguyen, H. He, and C. Pham-Huy, “Chiral drugs: An overview,” Int. J. Biomed. Sci. 2(2), 85–100 (2006).

38. C. M. Dobson, “Protein folding and misfolding,” Nature 426(6968), 884–890 (2003). [CrossRef]  

39. G. Cipparrone, A. Mazzulla, A. Pane, R. J. Hernandez, and R. Bartolino, “Chiral self-assembled solid microspheres: a novel multifunctional microphotonic device,” Adv. Mater. 23(48), 5773–5778 (2011). [CrossRef]  

40. R. J. Hernández, A. Mazzulla, A. Pane, K. Volke-Sepúlveda, and G. Cipparrone, “Attractive-repulsive dynamics on light-responsive chiral microparticles induced by polarized tweezers,” Lab Chip 13(3), 459–467 (2013). [CrossRef]  

41. H. Chen, C. Liang, S. Liu, and Z. Lin, “Chirality sorting using two-wave-interference-induced lateral optical force,” Phys. Rev. A 93(5), 053833 (2016). [CrossRef]  

42. H. Chen, Y. Jiang, N. Wang, W. Lu, S. Liu, and Z. Lin, “Lateral optical force on paired chiral nanoparticles in linearly polarized plane waves,” Opt. Lett. 40(23), 5530–5533 (2015). [CrossRef]  

43. Y. Li, G. Rui, S. Zhou, B. Gu, Y. Yu, Y. Cui, and Q. Zhan, “Enantioselective optical trapping of chiral nanoparticles using a transverse optical needle field with a transverse spin,” Opt. Express 28(19), 27808–27822 (2020). [CrossRef]  

44. H. Chen, W. Lu, X. Yu, C. Xue, S. Liu, and Z. Lin, “Optical torque on small chiral particles in generic optical fields,” Opt. Express 25(26), 32867–32878 (2017). [CrossRef]  

45. H. Chen, N. Wang, W. Lu, S. Liu, and Z. Lin, “Tailoring azimuthal optical force on lossy chiral particles in Bessel beams,” Phys. Rev. A 90(4), 043850 (2014). [CrossRef]  

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

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

48. S. Pancharatnam, “Generalized theory of interference and its applications,” Proc. Indian Acad. Sci. 44(5), 247–262 (1956). [CrossRef]  

49. M. V. Berry, “The adiabatic phase and Pancharatnam's phase for polarized light,” J. Mod. Opt. 34(11), 1401–1407 (1987). [CrossRef]  

50. M. Li, S. Yan, Y. Zhang, Y. Liang, P. Zhang, and B. Yao, “Optical sorting of small chiral particles by tightly focused vector beams,” Phys. Rev. A 99(3), 033825 (2019). [CrossRef]  

51. I. Liberal, I. Ederra, R. Gonzalo, and R. W. Ziolkowski, “Near-field electromagnetic trapping through curl-spin forces,” Phys. Rev. A 87(6), 063807 (2013). [CrossRef]  

52. S. Wang and C. Chan, “Lateral optical force on chiral particles near a surface,” Nat. Commun. 5(1), 3307 (2014). [CrossRef]  

53. J. D. Jackson, Classical Electrodynamics, 3rd ed. (Wiley, 1999).

54. N. Sule, S. A. Rice, S. K. Gray, and N. F. Scherer, “An electrodynamics-Langevin dynamics (ED-LD) approach to simulate metal nanoparticle interactions and motion,” Opt. Express 23(23), 29978–29992 (2015). [CrossRef]  

55. R. Biswas and D. R. Hamann, “Simulated annealing of silicon atom clusters in Langevin molecular dynamics,” Phys. Rev. B 34(2), 895–901 (1986). [CrossRef]  

56. B. Leimkuhler and C. Matthews, “Robust and efficient configurational molecular sampling via Langevin dynamics,” J. Chem. Phys. 138(17), 174102 (2013). [CrossRef]  

57. B. Leimkuhler and C. Matthews, “Rational construction of stochastic numerical methods for molecular sampling,” Appl. Math. Res. Express 2013(1), 34–56 (2012). [CrossRef]  

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 (6)

Fig. 1.
Fig. 1. Schematic diagram of setup for creating dual LPV structure. The spatial phase distribution of an incident LG mode light is modulated by a phase mask in the pupil plane. Under the tight focusing condition, components of the RH and LH vibrations of the illumination are focused into separated focal spots along x axis, enabling the arbitrary TC for the longitudinal field of each focal spot.
Fig. 2.
Fig. 2. Lateral dual foci created by a tightly focused LG illumination with (l, m, Δx) = (1, −1, 3λ). (a) Spatial phase modulation of the phase mask in the pupil plane. (b) Total, (c) transverse and (d) longitudinal components of the intensity distributions of the focal field. The insets in (b) and (d) illustrate SoP and phase map within the main lobe, respectively.
Fig. 3.
Fig. 3. Three pairs of dual foci created by a tightly focused LG illumination in 2D space. (a) Spatial phase modulation of the phase mask in the pupil plane. (b) Total, (c) stokes parameter S3 and (d) longitudinal components of the intensity distributions of the focal field. The insets in (d) illustrates the phase distribution of the main lobe. The intensities of several foci have been multiplied by a factor of 3 in order to obtain a clear image.
Fig. 4.
Fig. 4. Two pairs of dual foci created by a tightly focused LG illumination in 3D space. (a) Spatial phase modulation of the phase mask in the pupil plane. (b) Total intensity distributions of the focal field.
Fig. 5.
Fig. 5. Chiral detection with the use of dual LPV structure illustrated in Fig. 2. For chiral nanoparticles with κ = 0.5 + 0.01i, (a) Intensity pattern superimposed with the distribution of the optical force, (b) line-scan of the total, chiral and achiral force along x-axis, (c) line-scan of the longitudinal optical torque and potential depth along x-axis, (d) trajectories of the nanoparticles in the transverse plane of the focal field. The initial positions of the nanoparticles are indicated by the coordinates.
Fig. 6.
Fig. 6. Chiral detection with the use of dual LPV structure illustrated in Fig. 2. For chiral nanoparticles with κ = −0.5 − 0.01i, (a) Intensity pattern superimposed with the distribution of the optical force, (b) line-scan of the total, chiral and achiral force along x-axis, (c) line-scan of the longitudinal optical torque and potential depth along x-axis, (d) trajectories of the nanoparticles in the transverse plane of the focal field. The initial positions of the nanoparticles are indicated by the coordinates.

Tables (3)

Tables Icon

Table 1. Parameters of multiple pairs of dual foci in transverse plane

Tables Icon

Table 2. Parameters of multiple pairs of dual foci in 3D space

Tables Icon

Table 3. Concrete expressions of the optical force

Equations (19)

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

E 0 ( x , y ) = A 0 ( x , y ) [ cos ξ 0 ( x , y ) e x + e i Δ θ ( x , y ) sin ξ 0 ( x , y ) e y ] = A 0 [ e i ξ 0 ( x , y ) e 1 + e i ξ 0 ( x , y ) e 2 ] / 2 ,
E ( x , y , z ) = i k f 2 π 0 α 0 2 π E 1 ( θ , ϕ ) e i k ( s r ) sin θ d ϕ d θ .
s ( r r o o ) = s r s r o o = ρ sin θ cos ( ϕ φ ) + z cos θ Δ x sin θ cos ϕ .
ξ ( θ , ϕ ) = Δ x sin θ cos ϕ + m ϕ ,
E 1 ( θ , ϕ ) = l 0 ( θ ) [ e i ξ ( θ , ϕ ) e r + e i ξ ( θ , ϕ ) e l ] .
E x = i k f 2 π 0 α 0 2 π l 0 ( θ ) sin θ cos θ e i k [ ρ sin θ cos ( ϕ φ ) + z cos θ ] × [ sin ( ϕ Ψ ) sin ϕ + cos ( ϕ Ψ ) cos θ cos ϕ ] d ϕ d θ ,
E y = i k f 2 π 0 α 0 2 π l 0 ( θ ) sin θ cos θ e i k [ ρ sin θ cos ( ϕ φ ) + z cos θ ] × [ sin ( ϕ Ψ ) cos ϕ + cos ( ϕ Ψ ) cos θ sin ϕ ] d ϕ d θ ,
E z = i k f 2 π 0 α 0 2 π l 0 ( θ ) sin 2 θ cos θ e i k [ ρ sin θ cos ( ϕ φ ) + z cos θ ] cos ( ϕ Ψ ) d ϕ d θ ,
E z = i k f 2 π 0 α cos θ ( 2 β sin θ sin α ) | l | e ( β 2 sin 2 θ sin 2 α ) [ i l + m 1 e i ( l + m 1 ) φ J l + m 1 ( k ( ρ cos φ Δ x ) sin θ ) + i l m + 1 e i ( l m + 1 ) φ J l m + 1 ( k ( ρ cos φ Δ x ) sin θ ) ] e i k z cos θ sin 2 θ d θ ,
s ( r r o o ) = ρ sin θ cos ( ϕ φ ) + z cos θ ( ± Δ x sin θ cos ϕ ± Δ y sin θ sin ϕ ± Δ z cos θ ) ,
P ( θ , ϕ ) = j = 1 n exp [ i k ( Δ x j sin θ cos ϕ + Δ y j sin θ sin ϕ + Δ z j cos θ ) ] ,
[ E x E y E z ] = i k f 2 π 0 α 0 2 π sin θ cos θ e i k [ ρ sin θ cos ( ϕ φ ) + z cos θ ] j = 1 n ( 2 β sin θ sin α ) | l j | e ( β 2 sin 2 θ sin 2 α ) e i l j ϕ × [ V E x j V E y j V E z j ] d ϕ d θ
{ V E x j = sin ( ϕ Ψ j ) sin ϕ + cos ( ϕ Ψ j ) cos θ cos ϕ , V E y j = sin ( ϕ Ψ j ) cos ϕ + cos ( ϕ Ψ j ) cos θ sin ϕ , V E z j = cos ( ϕ Ψ j ) sin θ ,
α e e = i 6 π ε 0 ε m k 3 a 1 , α m m = i 6 π μ m μ 0 k 3 b 1 , α e m = 6 π n 1 Z 0 k 3 c 1 ,
a n = [ A n ( 2 ) V n ( 1 ) + A n ( 1 ) V n ( 2 ) ] Q n , b n = [ B n ( 1 ) W n ( 2 ) + B n ( 2 ) W n ( 1 ) ] Q n , c n = [ A n ( 1 ) W n ( 2 ) A n ( 2 ) W n ( 1 ) ] Q n ,
A n ( j ) = Z S D n ( 1 ) ( x j ) D n ( 1 ) ( x 0 ) , B n ( j ) = D n ( 1 ) ( x j ) Z S D n ( 1 ) ( x 0 ) , W n ( j ) = Z S D n ( 1 ) ( x j ) D n ( 3 ) ( x 0 ) , V n ( j ) = D n ( 1 ) ( x j ) Z S D n ( 3 ) ( x 0 ) , Q n = ψ n ( x 0 ) / ξ n ( x 0 ) V n ( 1 ) W n ( 2 ) + V n ( 2 ) W n ( 1 ) .
F = F c h i r a l + F a c h i r a l = ( F g r a d c + F r a d + F v o r + F s p i n ) + ( F g r a d a + F c u r l + F f l o w ) ,
T = [ 2 μ 0 R e ( α e m ) + μ 0 k 3 3 π ε 0 ε m Im ( α e e α e m ) + μ 0 2 k 3 3 π Im ( α m m α e m ) ] S + [ μ 0 k 3 6 π ε 0 ε m R e ( α e e α e m ) μ 0 2 k 3 6 π R e ( α m m α e m ) ] Im ( E × H ) [ 2 ω ε 0 ε m Im ( α e e ) ω k 3 3 π ε 0 2 ε m 2 α e e α e e ω μ 0 k 3 3 π ε 0 ε m α e m α e m ] L e [ 2 ω μ 0 μ m Im ( α m m ) ω μ 0 2 k 3 3 π μ m α m m α m m ω μ 0 k 3 3 π ε 0 ε m μ m α e m α e m ] L m .
m d 2 R d t 2 = F ( R , t ) χ d R d t + η ,
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.