Photonics Research, Volume. 13, Issue 8, 2033(2025)

Optical backflow for the manipulations of dipolar nanoparticles

Xiangyang Xie1, Peng Shi1、*, Changjun Min1,2, and Xiaocong Yuan1,3
Author Affiliations
  • 1Nanophotonics Research Center, Shenzhen Key Laboratory of Micro-Scale Optical Information Technology & Institute of Microscale Optoelectronics, Shenzhen University, Shenzhen 518060, China
  • 2e-mail: cjmin@szu.edu.cn
  • 3e-mail: xcyuan@szu.edu.cn
  • show less

    Backflow is a counterintuitive phenomenon that is widely predicted in the fields of quantum physics and optics. In contrast to quantum backflow, which is challenging to be observed, optical backflow is prevailing in structured lights. For instance, the azimuthal backflow has been recently observed experimentally in optics via the superposition of two beams carrying different orbital angular momentum topological charges. In this paper, we investigate the spin-momentum characteristics of the superimposed orbital angular momentum beams to confirm the optical azimuthal backflow, which is closely related to off-axis vortex flow and super-oscillations. Furthermore, we extend our study to axial backflow, characterized by a reversed axial energy flow in tightly focused cylindrical vector vortex beams. Then, we explore the application of optical backflow in the manipulation of dipolar nanoparticles. By optimizing material parameters, we achieve on-demand control of optical forces in both azimuthal and axial backflow scenarios. Our findings provide in-depth insights into the optical backflow phenomena with potential applications in optical manipulations.

    1. INTRODUCTION

    Backflow, which is a counterintuitive phenomenon that a particle with positive momentum has a non-zero probability of propagating backwards during specific time intervals, was first discovered in quantum physics [1]. To date, significant attention has been attracted to the study of quantum backflow, and the phenomenon has been predicted in various research fields [24]. Despite extensive theoretical investigations, the experimental observation of quantum backflow has faced a significant challenge. Recently, it has been recognized that backflow is a prevalent effect occurring in the interference of coherent waves [5], and several studies have successfully observed the optical backflow [68].

    The recent experimental observation of optical azimuthal backflow was achieved by the interference of two Laguerre-Gaussian (LG) beams carrying orbital angular momenta (OAMs) [8], which has sparked considerable research interest due to its broad range of applications in optical manipulations [911], communications [12,13], quantum information [14], spin-orbit interactions [15], and light-matter interactions [16,17]. Optical backflow can be classified into two categories: one involving the inverse wavevector (canonical momentum), such as optical azimuthal backflow [8], and the other involving energy backflow (Poynting vector) [1820]. The inverse wavevector is much rarer than the inverse Poynting vector. For instance, in axial energy backflow, the Belinfante’s spin momentum is opposite to the canonical momentum in the backflow region [2123]. Moreover, conventional inverse optical forces can only be achieved under specific conditions [24,25], whereas backflow provides a robust mechanism to overcome this limitation, and offers applications in the particle’s path manipulations, the non-conservative forces [26], and the sub-diffraction limit imaging [19].

    Previously, optical azimuthal backflow was described by scalar wave theory [8]. In this study, we reconsider optical azimuthal backflow from the perspective of vectorial electromagnetic (EM) theory. We first perform an in-depth analysis of the spin-momentum properties of superimposed OAM beams and confirm the existence of optical azimuthal backflow, even if the magnitudes of the two individual OAMs are equal. The spin-momentum properties are crucial for understanding the underlying physics of the optical backflow, such as the relationship between backflow, off-axis vortex flow, and super-oscillations. In addition to optical azimuthal backflow, we also study the optical axial backflow, which occurs in cylindrical vector vortex beams (CVVBs) under a tightly focusing configuration [27].

    In our work, we apply optical backflow to optical manipulations, which have been widely applied to fields such as biomedicine and materials science [28]. Although recent studies on optical magnetism have shown that reverse energy flow can be achieved without OAM [29,30], optical azimuthal backflow generated by OAM beams offers an additional degree of freedom in optical manipulations. Remarkably, by optimizing the material parameters of nanoparticles, we ensure the optical forces parallel to the backflow and mitigate the effects of Brownian motion. Moreover, we analyze the conditions in which optical azimuthal backflow occurs and reveal that the backflow depends on the difference between the topological charges (TCs) of the OAM beams, rather than their absolute values. This implies that the trajectory of trapped particles can be flexibly controlled by adjusting the TCs of the superimposed OAM states, enabling advanced optical manipulation in complex environments. Additionally, by tuning the polarization of CVVBs, we can engineer axial backflow, enabling the control of axial energy flux. Such control is critical for high-precision optical trapping, nanoscale particle manipulation, and advanced imaging techniques.

    2. THEORY

    A. Optical Azimuthal Backflow

    Optical azimuthal backflow occurs in the superposition state of two co-propagating OAM beams. LG beams contain the phase singularity (vortex phase) and OAM characterized by a TC l. In the free space (dielectric constant ε0 and magnetic permeability μ0), the complex amplitude of an LG beam in cylindrical coordinates (r, φ, z) can be expressed as [9] upl=Cplw0w(z)[2rw(z)]|l|Lp|l|[2r2w2(z)]exp[r2w2(z)]exp{i[kzkr22R(z)+φG]}e+ilφeiωt,where Cpl is the normalization coefficient with p the radial index and p=0 in the work; Lp|l| is the generalized Laguerre polynomial; z0=πw02/λ is the Rayleigh range with w0 the beam waist and λ the wavelength; k=2π/λ is the wave number; R(z)=z[1+(z0/z)2] is the curvature radius; w(z)=w0[1+(z0/z)2]1/2 is the beam waist radius, e+ilφ is the vortex phase, and φG=(2p+|l|+1)arctan(z/z0) is the Gouy phase. For numerical illustrations, the value λ=632.8  nm is used throughout the paper, and w0=4λ is taken for analysis of the azimuthal backflow case. In Eq. (1), ω=ck is the light frequency, c being the light velocity; in calculations of time-averaged quantities (energy, energy flow, and momentum densities), the time-dependent term eiωt can be ignored, and it will be omitted in the following expressions.

    Assuming that the two LG beams are x-polarization light, we can thus express the superimposed field as Ex=A1up1l1+A2up2l2, where A1 and A2 are the amplitude factors. Here, we primarily consider the properties of superimposed LG modes at the beam waist plane. By employing the electric Gauss law ·E=0 in the paraxial approximation, the electric field can be calculated as E=[Excosφr^;Exsinφφ^;1ik(1rExφsinφExrcosφ)z^].

    According to Faraday’s law H=(×E)/iωμ0, the magnetic field can be expressed as H=Zμε[Exsinφr^;Excosφφ^;ik(1rExφcosφ+Exrsinφ)z^],where Zμε represents the wave impedance. From Eqs. (2) and (3), the Poynting vector P=Re{E*×H}/2 can be calculated as P=Zμε2Re{ExikExrr^;(sin2φ2ikExExr+ExkrExφ)φ^;ExExz^},and the spin angular momentum (SAM) density S=Imε0E*×E+μ0H*×H/4ω is S=ε4ωkIm{(ExrExφExrExφ)r^;iExExrφ^;0z^},where the superscript * represents the complex conjugate.

    B. Optical Axial Backflow

    Then, we consider the optical axial backflow, which occurs in the tightly focused field of CVVBs. CVVBs expressed by E=[ηxcos(nφ+φ0)x^+ηysin(nφ+φ0)y^]eimφeiωtare a class of structured light with spatially varying polarization structure and helical phase structure. Therefore, they may carry SAM and OAM. Here, ηx and ηy are the amplitude factor of x- and y-polarization components, respectively, n is the polarization TC, φ0 is the initial polarization state, and m is the vortex TC. Like in Section 2.A, the time-dependent multiplier eiωt can be omitted during the energy and momentum calculations; for simplicity, m=0 is assumed in our further analysis. However, the axially mechanical characteristics of tightly focused CVVBs are similar to those of tightly focused cylindrical vector beams (CVBs). In this way, we can use the same technique to achieve the reverse axial optical force for the tightly focused CVVBs. In the focal plane, the vectorial diffraction theory is employed to calculate the focused field [3135] and the electric field vectors are given by E(ρ,ϕ,z)=ifλ0θmax02πP(θ)Γe·eik[zcosθ+ρsinθcos(φϕ)]sinθdθdφ,Γe=[Er(cosθcosφx^cosθsinφy^sinθz^)+Eϕ(sinφx^cosφy^0z^)],where (ρ, ϕ, z) are the cylindrical coordinates in the focal plane, f is the focal length of the tightly focusing systems, θmax=arcsin(NA) is the maximum acceptance angle θ, which is determined by the system’s numerical aperture NA, and P(θ) is the apodization function.

    By substituting Eq. (6) into Eq. (7), the electric field distribution near the focus of the focused CVBs can be calculated as Eρ=iΦ2{2[ηxcos(nϕ+φ0)cosϕ+ηysin(nϕ+φ0)sinϕ](1+cosθ)Jn(α)+(ηx+ηy)(1cosθ)Jn2(α)cos[(n1)ϕ+φ0]+(ηxηy)(1cosθ)Jn+2(α)cos[(n+1)ϕ+φ0]}eikzcosθρ^,Eϕ=iΦ2{2[ηxcos(nϕ+φ0)sinϕηysin(nϕ+φ0)cosϕ](1+cosθ)Jn(α)+(ηx+ηy)(1cosθ)Jn2(α)sin[(n1)ϕ+φ0](ηxηy)(1cosθ)Jn+2(α)sin[(n+1)ϕ+φ0]}eikzcosθϕ^,Ez={(ηx+ηy)sinθJn1(α)cos[(n1)ϕ+φ0]+(ηxηy)sinθJn+1(α)cos[(n+1)ϕ+φ0]}eikzcosθz^,where Φ=AinfP(θ)sinθ/λ, and Jn(α) are the first Bessel functions of order n with argument α=kρsinθ. According to Faraday’s law H=(×E)/iωμ0, the magnetic field near the focus of the focused CVBs can be expressed as Hρ=kΦi2ωμ{2[ηxcos(nϕ+φ0)sinϕηysin(nϕ+φ0)cosϕ](1+cosθ)Jn(α)(ηx+ηy)(1cosθ)Jn2(α)sin[(n1)ϕ+φ0]+(ηxηy)(1cosθ)Jn+2(α)sin[(n+1)ϕ+φ0]}eikzcosθρ^,Hϕ=kΦi2ωμ{2[ηxcos(nϕ+φ0)cosϕ+ηysin(nϕ+φ0)sinϕ](1+cosθ)Jn(α)(ηx+ηy)(1cosθ)Jn2(α)cos[(n1)ϕ+φ0](ηxηy)(1cosθ)Jn+2(α)cos[(n+1)ϕ+φ0]}eikzcosθϕ^,Hz={(ηx+ηy)sinθJn1(α)sin[(n1)ϕ+φ0]+(ηxηy)sinθJn+1αsin[(n+1)ϕ+φ0]}eikzcosθz^.

    From Eq. (8) to Eq. (13), for optical axial backflow, we will focus on the analysis of only the axial component of the Poynting vector Pz=Re{Eρ*HϕEϕ*Hρ}/2, which can be calculated as Pz=k|Φ|28ωμRe{2(ηxηx+ηyηy)(1+cosθ)2Jn2(α)(1cosθ)2(ηx+ηy)(ηx+ηy)Jn22(α)(1cosθ)2(ηxηy)(ηxηy)Jn+22(α)+2(ηxηxηyηy)cos(2nϕ+2φ0)(1+cosθ)2Jn2(α)2(ηxηxηyηy)cos(2nϕ+2φ0)(1cosθ)2Jn2(α)Jn+2(α)},and the SAM density S=Im{ε0E*×E+μ0H*×H}/4ω is Sρ=ε|Φ|24ωRe{(ηxηxηyηy)sinθ(1+cosθ)Jn1(α)Jn(α)(ηxηxηyηy)sinθ(1+cosθ)Jn(α)Jn+1(α)+(ηxηxηyηy)sinθ(1cosθ)Jn2(α)Jn+1(α)+(ηxηxηyηy)sinθ(1cosθ)Jn1(α)Jn+2(α)}sin(2nϕ+2φ0)ρ^,Sϕ=ε|Φ|24ωRe{[(ηx+ηy)(ηx+ηy)sinθ(1+cosθ)Jn1(α)Jn(α)(ηx+ηy)(ηx+ηy)sinθ(1cosθ)Jn2(α)Jn1(α)+(ηxηy)(ηxηy)sinθ(1+cosθ)Jn(α)Jn+1(α)+(ηxηy)(ηxηy)sinθ(1cosθ)Jn+1(α)Jn+2(α)]+[(ηxηxηyηy)sinθ(1cosθ)Jn1(α)Jn+2(α)(ηxηxηyηy)sinθ(1+cosθ)Jn1(α)Jn(α)+(ηxηxηyηy)sinθ(1cosθ)Jn2(α)Jn+1(α)+(ηxηxηyηy)sinθ(1+cosθ)Jn(α)Jn+1(α)]cos(2nϕ+2φ0)}ϕ^,Sz=ε|Φ|24ωIm{(ηyηxηxηy)[(1+cosθ)2Jn(α)Jn(α)+(1cosθ)2Jn2(α)Jn+2(α)]sin(2nϕ+2φ0)}z^.

    The theoretical results presented above provide the foundation for the subsequent analysis and will be applied in the following sections to interpret the results.

    3. RESULTS AND DISCUSSION

    A. Optical Azimuthal Backflow

    We first consider the case of the optical azimuthal backflow. Since the incident light is linearly polarized along the x-axis (i.e., Ey=0), it follows from Faraday’s law that Hx=0. Consequently, the z-component of the SAM vanishes, i.e., Sz=0. The Poynting vector P is proportional to the kinetic Poynting momentum p, i.e., p=P/c2 [22,36,37], which can be decomposed into spin momentum ps and canonical momentum po, i.e., p=ps+po. In the paraxial approximation (A/zkA), the horizontal components (r- and φ-components) of the spin momentum, given by ps=×S/2, are negligible. This implies that the horizontal component of the Poynting momentum p is equal to the canonical momentum po. Additionally, the Minkowski-type canonical momentum can be re-expressed as po=k [3740]. Therefore, the optical azimuthal backflow in superposed LG beams can be quantitatively described by the Poynting momentum p. To identify the conditions for optical azimuthal backflow, we assume that the OAM beam is radially uniform (i.e., upl/r=0) and has the form upl=Aeilφ, where A is a real constant (this assumption is only used in analyzing the occurrence of optical azimuthal backflows). Substituting these conditions into Eq. (4), the azimuthal component of Poynting momentum simplifies to pφA12l1+A22l2+A2A1(l1+l2)cos(l1l2)φ.

    When the amplitudes of the two superimposed beams are equal (i.e., A1=A2), the azimuthal component of the Poynting momentum satisfies pφ[l1+l2+(l1+l2)cos(l1l2)φ]<0 (where l1<0 and l2<0). The azimuthal Poynting momentum of the superimposed state is consistent with that of individual beams, meaning that no backflow phenomenon will occur. When A1A2, for simplicity we set A1=1, and then Eq. (18) can be transformed into pφ[1+A2cos(l1l2)φ]l1+[A22+A2cos(l1l2)φ]l2.

    When cos(l1l2)φ=0, the right-hand side of Eq. (18) simplifies to l1+A22l2. Since the beams have negative TCs, the value is negative. However, when cos(l1l2)φ=1, the right-hand side of Eq. (19) becomes (1A2)l1+(A22A2)l2. The sign of the expression can be changed under the conditions 1A2>0 and (A22A2)>(1A2)l1/l2 (e.g., A2=0.5 and l1/l2<1/2); the right-hand side becomes positive, which indicates the presence of optical azimuthal backflow, as discussed in Ref. [8].

    To intuitively understand the optical azimuthal backflow phenomenon, we present the distributions of the Poynting momentum, canonical momentum, spin momentum, and wavevector at the waist plane (Fig. 1). As depicted in Figs. 1(b) and 1(e), the azimuthal momentum exhibits anomalously positive values in the regions where optical azimuthal backflow occurs. By comparing Figs. 1(c), 1(f), and 1(g), it is evident that the axial component of the ps is small, while the canonical momentum is approximately parallel to the Poynting momentum. When the two beams are superimposed, the Poynting momentum and OAM are not consistently concentrated at the beam center, leading to the formation of off-axis vortex flow [Fig. 1(h)] [19]. This off-axis vortex flow generates a local phase gradient, which increases the local wavevector. The phase gradient further amplifies the local oscillation frequency, resulting in a wavenumber that exceeds the value of a single beam’s wavenumber. As observed in Fig. 1(i), the local normalized wavenumber exceeds k=ω/c in the regions where optical azimuthal backflow occurs, which indicates a necessary condition for super-oscillations [5,20]. In addition, even when A1=A2, the off-axis vortex flow and the wavenumber still exceed k=ω/c, indicating that the optical azimuthal backflow phenomenon still occurs.

    Normalized momentum density and wavevector at the waist plane. The (a) r, (b) φ, (c) z components of Poynting momentum p can be decomposed into the (d) r, (e) φ, (f) z components of po and (g) the z components of ps. Here, the horizontal components of ps are zero. (h) The normalized φ component of k/W indicates off-axis vortexes and (i) the normalized wavenumber |k|/W is larger than one, which is a necessary condition for super-oscillations. Here, l1=−1, l2=−3, A1=A2 throughout the paper.

    Figure 1.Normalized momentum density and wavevector at the waist plane. The (a) r, (b) φ, (c) z components of Poynting momentum p can be decomposed into the (d) r, (e) φ, (f) z components of po and (g) the z components of ps. Here, the horizontal components of ps are zero. (h) The normalized φ component of k/W indicates off-axis vortexes and (i) the normalized wavenumber |k|/W is larger than one, which is a necessary condition for super-oscillations. Here, l1=1, l2=3, A1=A2 throughout the paper.

    Then, we apply the concept of optical azimuthal backflow to optical manipulation. For dipolar particles (with chiral parameter κ, dielectric constant εp, and permeability μp), the time-averaged optical force is expressed as [4143] F=12Re{d·()E+m·()Hck046π(d×m)}.

    Here, the electric and magnetic dipole moments can be expressed as d=αeeE+iαemH, m=αmmHiαemE, where αee, αem, and αmm are the polarizabilities of the particle. Among these, αem is associated with the chiral parameter κ, and for a non-chiral particle, αem=0.

    By substituting the constitutive relations of the particle and the expressions for the EM dipole moments into Eq. (20), the force can be rewritten as [26,44] F=Fgrad+Frad+Fspin+Fcurl+Fflow.

    Here, Fgrad=U represents the optical gradient force, where U=Reαeewe/ε0+Reαmmwm/μ0+ReαemC/ωε0μ0 is associated with the electric energy density we=ε0|E|2/4, the magnetic energy density wm=μ0|H|2/4, and the chiral density C=ωε0μ0ImE*·H. Under the paraxial approximation, the gradient force in the z-component is zero. The radiation force, Frad=σP/c, is caused by the momentum flux of light and is proportional to the Poynting vector or the momentum density. Here, the speed of light in vacuum is c=1/(ε0μ0)1/2, and σ=σe+σm+σ, where σe=kIm{αee}/Re{εp}, σm=kIm{αmm}/Re{μp}, and σ=ck4Re{αemαem*+αeeαmm*}/6π. According to the previous analysis, the canonical momentum plays a dominant role in the total momentum. Therefore, the radiation force should resemble the distribution of canonical momentum. The term Fspin=ωγeSe+ωγmSm is the spin force, which vanishes because both the coefficients γe=2(ωIm{αem}Re{ck4αemαem*}/6πε0) and γm=2(ωIm{αem}Re{ck4αemαem*}/6πμ0) are related to αem, which is equal to zero here. The term Fcurl=2ω2Im{αem}ST2v(σepse+σmpsm) is the curl force, which is related to the optical transverse spin ST=×P/2ω2 [4547], and the spin momentum ps=×S/2 [3740]. Obviously, there is only the axial component in the optical curl force here, since αem=0 and the horizontal components of ps vanish. The last term, Fflow=σImP/c, represents the force associated with the imaginary Poynting vector [48]. Here, the Fflow force is very small, less than 107 of the total force, and can therefore be neglected.

    Motivated by the growing interest in levitated optomechanics [49], Fig. 2 shows the distributions of optical force exerted by the azimuthal optical backflow on achiral gold dipolar particles with a radius of 50 nm in vacuum. As is shown in Fig. 2(b), a sign reversal appears in the azimuthal component of the total optical force Fφ within the backflow region. This means that if the particle moves along the azimuthal direction [i.e., along a circle depicted in Fig. 2(b)], it experiences the azimuthal force of opposite directions [shown by yellow arrows in Fig. 2(b)] at different segments of its trajectory, and it originates from the combined contribution of the gradient force Fgrad and the radiation force Frad. However, the presence of the gradient force Fgrad will cause the misalignment between the Poynting momentum in Fig. 1(b) and optical force in Fig. 2(b), which will make readers confused about the origin of backflow force, particularly in experiment. Therefore, the influence of gradient force and the misalignment should be mitigated by optimizing the material’s properties. On the other hand, the z-component of Fcurl is much smaller than the total force, and thus the Fz aligns parallelly with the axial Poynting momentum.

    The distributions of optical force as the optical azimuthal backflow acts on gold dipolar nanoparticles [50]. The (a) r, (b) φ, (c) z components of total optical force F. The total optical force can be divided into the (d) r, (e) φ, (f) z components of radiation Frad related to Poynting momentum, the (g) r, (h) φ components of gradient force Fgrad related to energy density, and (i) the z component of Fcurl related to the spin momentum. Here, εp=−11.74+1.2611i and μp=1.0; l1=−1 and l2=−3. All the force is normalized by the maximal value of total optical force throughout the paper.

    Figure 2.The distributions of optical force as the optical azimuthal backflow acts on gold dipolar nanoparticles [50]. The (a) r, (b) φ, (c) z components of total optical force F. The total optical force can be divided into the (d) r, (e) φ, (f) z components of radiation Frad related to Poynting momentum, the (g) r, (h) φ components of gradient force Fgrad related to energy density, and (i) the z component of Fcurl related to the spin momentum. Here, εp=11.74+1.2611i and μp=1.0; l1=1 and l2=3. All the force is normalized by the maximal value of total optical force throughout the paper.

    Next, we reconsider the Fgrad, which is related to the optical potential U=Re{αee}we/ε0+Re{αmm}wm/μ0. Here, the electric polarizability αee and the magnetic polarizability αmm coexist, and the electric energy density we and magnetic energy density wm are of the same order of magnitude. However, since ε0/μ01A4s6/(kg2m4), the normalized electric polarizability αee/ε0 is much larger than the magnetic polarizability αmm/μ0, indicating that the electric polarizability dominates the Fgrad. In addition, Brownian motion significantly affects optical trapping stability. For dipolar particles, Brownian motion can be evaluated by the random thermal energy kBT, where kB is the Boltzmann constant and T is the ambient temperature. The optical trapping potential well is given by V=Re{αee}|E|2/2 [51]. Provided that V>kBT, the trapping can be considered stable against the influence of Brownian motion.

    Therefore, to align the optical force with the backflow parallelly and mitigate the influence of Brownian motion, we optimize the nanoparticle permittivity by maximizing ξ1=Imαee/Reαee, thereby ensuring that Frad becomes dominant over Fgrad. Meanwhile, by increasing the trapping parameter ξ2=Re{αee}|E|2/2kBTI to ensure that the optical potential well exceeds the thermal energy, we can achieve the stable trapping. Here, the trapping parameter ξ2 represents the potential V/kBT normalized by the intensity (|E|2), and it is dimensional. To simplify the symbols in the following calculations and figures, we assume that the quantity I only extracts the numerical part of |E|2 to make the ξ2 dimensionless (or setting |E|=1  V/μm uniformly). The corresponding results are shown in Figs. 3(a) and 3(b).

    The optimization of material’s permittivity for eliminating the misalignment between force and Poynting momentum and achieving stable optical trapping. Since the Frad/Fgrad is related to the ξ1=Im{αee}/Re{αee}, we scan the permittivity of nanoparticles for maximal (a) log10|ξ1| to eliminate the misalignment between force and Poynting momentum. The trapping potential well is related to Re{αee} and we scan the permittivity of nanoparticles for maximal (b) log10|ξ2|. Here, μp=1.0.

    Figure 3.The optimization of material’s permittivity for eliminating the misalignment between force and Poynting momentum and achieving stable optical trapping. Since the Frad/Fgrad is related to the ξ1=Im{αee}/Re{αee}, we scan the permittivity of nanoparticles for maximal (a) log10|ξ1| to eliminate the misalignment between force and Poynting momentum. The trapping potential well is related to Re{αee} and we scan the permittivity of nanoparticles for maximal (b) log10|ξ2|. Here, μp=1.0.

    From Fig. 3(a), the maximal value of |ξ1| (ξ1=1.467×107) is attained when εp=1.586+1.653i, corresponding to point A. The total optical force is presented in Figs. 4(a)–4(c), which clearly demonstrate that the misalignment is eliminated. However, the trapping potential well (V/kBT) is much smaller than one [Fig. 4(d)], indicating that the nanoparticle cannot be trapped stably due to Brownian motion.

    The total optical force and the trapping potential well V/kBT. When εp=−1.586+1.653i, the (a) r, (b) φ, (c) z components of F match well with the Poynting momentum given by Figs. 1(a)–1(c). However, the (d) potential well is much smaller than one and the particle cannot be trapped stably. When εp=−2.609−0.339i, the (d) r, (e) φ, (f) z components of F are shown and the (h) potential well is much larger than one. When εp=−2.7458+0.23171i, the (i) r, (j) φ, (k) z components of F are shown and the (l) potential well is much larger than one. Here, l1=−1,l2=−3.

    Figure 4.The total optical force and the trapping potential well V/kBT. When εp=1.586+1.653i, the (a) r, (b) φ, (c) z components of F match well with the Poynting momentum given by Figs. 1(a)–1(c). However, the (d) potential well is much smaller than one and the particle cannot be trapped stably. When εp=2.6090.339i, the (d) r, (e) φ, (f) z components of F are shown and the (h) potential well is much larger than one. When εp=2.7458+0.23171i, the (i) r, (j) φ, (k) z components of F are shown and the (l) potential well is much larger than one. Here, l1=1,l2=3.

    According to Fig. 3(b), the maximal value of |ξ2|(ξ2=1.51×104) is attained when εp=2.6090.339i, corresponding to point B. The trapping potential well is much larger than one, indicating that the stable trapping of nanoparticles can be achieved in this instance. However, the distribution of total optical force shown in Fig. 4(e)–4(g) exhibits a misalignment between the Poynting momentum and optical force in the horizontal components. Although the inverse azimuthal force can also be observed, it is not dominant, and one may argue that this backflow force is not originated from the backflow. It can be observed that there is a tradeoff between the larger radiation force Frad (related to Im{αee}) and the deeper trapping potential well (related to Re{αee}).

    To resolve this problem, we re-check Fig. 3 and find a proper material parameter to achieve the relatively large radiation force and trapping potential well. When εp=2.7458+0.23171i [the permittivity of silver at λ368  nm, and corresponding to point C in Fig. 3(a)], ξ1=4.828, and ξ2=2.514, in this instance [Figs. 4(i)–4(l)], the influence of the gradient force is effectively mitigated, and the effect of Brownian motion is simultaneously suppressed. Here, we ignore the influence of multipoles on the optical force, because we find that the magnitude of quadrupole mode is two orders smaller than that of dipole mode, and the magnitude of hexapole modes is two orders smaller than that of quadrupole mode [52,53].

    In addition, the number of optical azimuthal backflow occurrences is related to the difference in the TC of OAM states, rather than their absolute values. According to Eq. (19), the azimuthal Poynting momentum reverses direction when the sign of cos(l1l2)φ changes. Given the periodic nature of the cosine function, the sign changes |l1l2| times in one period, corresponding to the number of azimuthal backflows equal to |l1l2|. As shown in Fig. 5, we present the variation of Fφ with different TCs, confirming that the number of azimuthal backflow occurrences matches the above analysis. This indicates that the topological structure of beams plays a crucial role in the distribution and manipulation of the optical forces.

    The azimuthal optical force Fφ for different TCs of OAM states. The appearance of optical azimuthal backflow is determined by |l1 −l2|, which can be verified in the Fφ when (a) l1=−1 and l2 =−4, (b) l1=−1 and l2=−5, (c) l1=−1 and l2=−6, (d) l1=−2 and l2 =−5, (e) l1=−2 and l2=−6, (f) l1=−3 and l2=−6. Here, εp=−2.7458+0.23171i and μp=1.0.

    Figure 5.The azimuthal optical force Fφ for different TCs of OAM states. The appearance of optical azimuthal backflow is determined by |l1l2|, which can be verified in the Fφ when (a) l1=1 and l2=4, (b) l1=1 and l2=5, (c) l1=1 and l2=6, (d) l1=2 and l2=5, (e) l1=2 and l2=6, (f) l1=3 and l2=6. Here, εp=2.7458+0.23171i and μp=1.0.

    B. Optical Axial Backflow

    Then, we consider the case of the optical axial backflow. In the axial case, we are interested in identifying the regions where Eq. (14) takes negative values, corresponding to areas of axial backflow. For simplicity, we set ηx=ηy=1, and the azimuthal mode number n=2 and cosθ1, which allows Pz to be simplified as Pz(1+cosθ)2J22(α)(1cosθ)2J02(α).

    Clearly, in the central region where α=0, we have J2(0)=0 and J0(0)=1, resulting in a negative value of Pz. Since the polarization components are set to ηx=ηy=1, we can derive that Sz=0 from Eq. (17). On the other hand, the energy flow is predominant along the z-axis, and this longitudinal energy flow varying in the transverse plane will give rise to the transverse SAM [46,47], indicating that the SAM has only transverse components.

    As previously concluded, the Poynting vector P is proportional to the kinetic Poynting momentum p, which can be decomposed into spin momentum ps and canonical momentum po. The spin momentum is given by ps=×S/2. Since the SAM has only transverse components, ps has only axial components. Moreover, on the focal plane, the Poynting vector has only a z-component, implying that the po also has only an axial component. Consequently, the axial backflow is governed solely by the contributions of the spin and canonical momentum in the z-direction.

    To intuitively understand the optical axial backflow phenomenon, we present the distributions of the Poynting momentum, canonical momentum, and spin momentum at the focal plane (Fig. 6). As shown in Figs. 6(a) and 6(c), both the axial kinetic momentum p and spin momentum ps exhibit anomalous negative values in the central region and in a concentric ring. These features are particularly prominent in Fig. 6(c) and correspond to the regions where optical axial backflow occurs. In contrast, as shown in Fig. 6(b), the canonical momentum po remains unaffected by the backflow.

    Normalized momentum density at the focal plane. (a) The z components of Poynting momentum p can be decomposed into the (b) z components of po and (c) the z components of ps. The horizontal components of p, po, and ps are zero. Here, ηx=ηy=1, numerical aperture NA=0.95, the beam waist w0 =0.26λ, and A=1 throughout the paper.

    Figure 6.Normalized momentum density at the focal plane. (a) The z components of Poynting momentum p can be decomposed into the (b) z components of po and (c) the z components of ps. The horizontal components of p, po, and ps are zero. Here, ηx=ηy=1, numerical aperture NA=0.95, the beam waist w0=0.26λ, and A=1 throughout the paper.

    Similarly, we apply the optical axial backflow to optical manipulation. The optical gradient force Fgrad=U in the z-direction is negligible due to the small gradient of the light field along this axis when the beam is focused. The optical radiation force Frad=σP/c is related to the Poynting vector. According to the previous analysis, the Poynting vector P has only an axial component, which leads to the conclusion that the radiation force also has only an axial component. The optical spin force Fspin=ωγeSe+ωγmSm vanishes because both the coefficients γe and γm are related to αem, which is equal to zero here. The optical curl force Fcurl=2ω2Im{αem}ST2v(σepse+σmpsm) is determined by the transverse spin ST and spin momentum ps. Since αem=0 and the spin angular momentum ps has only an axial component, Fcurl is therefore purely axial. The optical flow force Fflow=σIm{P}/c is related to the imaginary Poynting vector and provides an additional contribution.

    We present the distributions of optical forces acting on achiral gold dipolar nanoparticles with a radius of 70 nm, as shown in Fig. 7. In Figs. 7(a), 7(d), and 7(g), as well as in Figs. 7(b), 7(e), and 7(h), it can be observed that the transverse optical force is primarily contributed by the gradient force Fgrad, whereas the flow force Fflow accounts for only 2% of the total force and can therefore be neglected. According to the previous analysis, the distribution of the optical force F closely resembles the Poynting vector P. However, as shown in Fig. 7(c), the z-component of the total optical force Fz does not exhibit any backward force. Further analysis of Figs. 7(f) and 7(i) reveals that the two components contributing to Fz, namely, the radiation force Frad,z and the curl force Fcurl,z, form an oppositely directed force pair in the central region. This counterbalance prevents Fz from reversing its direction. To investigate the conditions under which optical force reversal may occur, we perform a theoretical reanalysis of Fz, which can be expressed as Fz=Frad,z+Fcurl,z=c(σe+σm+σ)pz2c(σeps,ze+σmps,zm)=cσe(pz2ps,ze)+cσm(pz2ps,zm)+cσpz.

    The distributions of optical force as the optical axial backflow acts on dipolar nanoparticles. The (a) r, (b) φ, (c) z components of total optical force F. The total optical force can be divided into the (d) r, (e) φ components of gradient force Fgrad, (f) the z component of radiation force Frad related to Poynting momentum, the (g) r, (h) φ components of flow force Fflow related to imaginary Poynting vector, and (i) the z component of Fcurl related to the spin momentum. Here ,εp=−11.74+1.2611i and μp=1.0.

    Figure 7.The distributions of optical force as the optical axial backflow acts on dipolar nanoparticles. The (a) r, (b) φ, (c) z components of total optical force F. The total optical force can be divided into the (d) r, (e) φ components of gradient force Fgrad, (f) the z component of radiation force Frad related to Poynting momentum, the (g) r, (h) φ components of flow force Fflow related to imaginary Poynting vector, and (i) the z component of Fcurl related to the spin momentum. Here ,εp=11.74+1.2611i and μp=1.0.

    In the central region, due to the symmetry and mutual coupling between the electric and magnetic fields, their contributions to the spin are equal (i.e., ps,ze=ps,zm0.5ps,z). This indicates that the first two terms in Eq. (23) are approximately equal to c(σe+σm)po in the central region. As shown in Fig. 6(b), the canonical momentum po is always positive, indicating that the sign of the first two terms is determined by σe+σm. Meanwhile, the last term of Eq. (23) is proportional to the Poynting momentum pz and σ. Although Brownian motion may influence stable trapping, the conclusion that its influence can be mitigated by optimizing the nanoparticle properties remains valid in this case. Therefore, to stably generate a reverse optical force and mitigate the influence of Brownian motion, we optimize the permittivity and permeability of the nanoparticle as follows: increase the trapping parameter ξ2, ensuring stable optical trapping by maintaining the trapping potential greater than the thermal energy; ξ3=|(σe+σm)/max(σe,σm)|; decrease the impact of the po on the axial force Fz; and ξ4=σ/max(|σe|,|σm|,|σe+σm|)>0, maximizing ξ4 to ensure the consistency between the axial force Fz and the Poynting momentum pz. The results are shown in Fig. 8.

    The optimization of material’s nanoparticle properties to ensure stable optical trapping and stably generate a reverse optical force. Since the trapping potential well is related to the ξ2 and the Fz is related to ξ3=|(σe+σm)/max(σe,σm)| and ξ4 =σ′/max(|σe|,|σm|,|σe+σm|), we scan the permittivity of nanoparticles for maximal (a), (d) log10|ξ2| to ensure stable optical trapping and scan the permittivity of nanoparticles for minimal (b), (e) log10ξ3 and maximal (c), (f) log10ξ4 to stably generate a reverse optical force. Here, in (a)–(c) μp=1.0+5.0i and in (d)–(f), μp=1.0+10.0i.

    Figure 8.The optimization of material’s nanoparticle properties to ensure stable optical trapping and stably generate a reverse optical force. Since the trapping potential well is related to the ξ2 and the Fz is related to ξ3=|(σe+σm)/max(σe,σm)| and ξ4=σ/max(|σe|,|σm|,|σe+σm|), we scan the permittivity of nanoparticles for maximal (a), (d) log10|ξ2| to ensure stable optical trapping and scan the permittivity of nanoparticles for minimal (b), (e) log10ξ3 and maximal (c), (f) log10ξ4 to stably generate a reverse optical force. Here, in (a)–(c) μp=1.0+5.0i and in (d)–(f), μp=1.0+10.0i.

    Our analysis primarily focuses on the case of μp=1.0+5.0i. As shown in Fig. 8(a), most regions satisfy |ξ2|>1, indicated by the red areas in the figure. To ensure stable axial optical trapping while generating a reverse axial optical force, we exclude parameters that fail to support stable trapping. Consequently, the subsequent discussion is restricted to conditions that guarantee stable trapping.

    According to Fig. 8(b), the minimum value of ξ3  (ξ3=1.297×107) is obtained at point A, which corresponds to εp=7.7952.7825i. As shown in Fig. 9(a), a significant reverse axial optical force appears in the central region. In this case, the axial optical force is influenced by the curl force Fcurl,z, which exhibits a significantly different distribution from the axial Poynting vector pz. According to Fig. 8(c), the value of ξ4=0.572, which confirms the proportional relationship between the axial radiation force and the curl force, as illustrated in Figs. 9(b) and 9(c). Therefore, a trade-off between ξ3 and ξ4 is required to effectively suppress the effect of the curl force. Such an optimization ensures that the generated reverse axial optical force is consistent with the axial backflow. Furthermore, as shown in Fig. 9(d), the trapping potential well is larger than one, indicating that Brownian motion is effectively suppressed. Once the particles are trapped in the potential well, they will move in the z-direction under the influence of Fz.

    The axial optical force and the trapping potential well V/kBT. When εp=−7.795−2.7825i and μp=1.0+5.0i, a significant reverse axial optical force emerges in the (a) z components of F. But Fz have a markedly different distribution from the (b) z components of Frad, due to the influence of the (c) z components of Fcurl and (d) potential well is larger than one. When εp =0.535–3.715i and μp=1.0+5.0i, the distribution of the (e) z components of F aligns with the (f) z components of Frad with tiny (g) z components of Fcurl and (h) potential well is also larger than one. When εp =0.085–3.6525i and μp=1.0+10.0i, the z components of (i) F, (j) Frad, (k) Fcurl and (l) the potential well have similar distributions and results to (e)–(h).

    Figure 9.The axial optical force and the trapping potential well V/kBT. When εp=7.7952.7825i and μp=1.0+5.0i, a significant reverse axial optical force emerges in the (a) z components of F. But Fz have a markedly different distribution from the (b) z components of Frad, due to the influence of the (c) z components of Fcurl and (d) potential well is larger than one. When εp=0.5353.715i and μp=1.0+5.0i, the distribution of the (e) z components of F aligns with the (f) z components of Frad with tiny (g) z components of Fcurl and (h) potential well is also larger than one. When εp=0.0853.6525i and μp=1.0+10.0i, the z components of (i) F, (j) Frad, (k) Fcurl and (l) the potential well have similar distributions and results to (e)–(h).

    Based on the previous analysis, we further explore the parameter space to optimize both the radiation force and the trapping potential well. Specifically, by re-examining Fig. 8, we identify material parameters that enable the achievement of a relatively large radiation force and the suppression of Brownian motion. As shown in Figs. 9(e)–9(h), when εp=0.5353.715i [point B in Fig. 8(c)], we obtain ξ2=5.074, ξ3=1.972, and ξ4=29.866; the influences of curl force and the Brownian motion are eliminated simultaneously. Consequently, the distribution of the axial optical force Fz aligns with the axial radiation force Frad,z (axial Poynting momentum pz). Similarly, as shown in Figs. 9(i)–9(l), when μp=1.0+10.0i and εp=0.0853.6525i [point C in Fig. 8(f)], we obtain ξ2=10.594, ξ3=3.307, and ξ4=364.364. Under these conditions, a similar distribution of the axial optical force Fz is observed.

    The aforementioned material parameters may be achieved by an artificial nanoparticle with its permittivity and permeability described by the Drude-Lorentz model [54,55], εp=1ωpe2ω2+iγeω.

    By substituting the desired real and imaginary parts of εp into this expression, the corresponding plasma frequency ωpe and damping factor γe can be numerically determined. A similar formulation applies to the magnetic permeability μp. Besides, although it is difficult to fabricate such nanoparticles at optical frequencies, the scale law of EM theory allows us to realize the design at microwave frequencies [56]. These artificial particles enable the exploration of back-transport phenomena under controlled conditions, potentially offering new capabilities in long-range particle manipulation and directional transport in tailored photonic environments.

    4. CONCLUSIONS

    In summary, we investigate the spin-momentum properties of both optical azimuthal and axial backflows in the superimposed fields of LG beams and tightly focused CVVBs, respectively. The results demonstrate a close relationship between optical azimuthal backflow, off-axis vortex flow, and super-oscillations. Particularly, the optical azimuthal backflow occurs even when the magnitudes of the two individual OAMs are equal, with the number of azimuthal backflows determined solely by the difference |l1l2|. This indicates that the topological structure of beams plays a crucial role in the generation, control, and application such as optical manipulation with backflow forces. For optical axial backflow, we show that although the axial Poynting momentum pz is inverted, the spin momentum induced by optical spin-orbit couplings will cause the optical force parallel to the canonical momentum and no backflow force appears for conventional non-magnetic nanoparticles.

    Then, we applied both optical azimuthal and axial backflows to optical manipulations. For azimuthal backflow, we optimized the material’s permittivity to eliminate the misalignment between the force and Poynting momentum and mitigate the Brownian motion further. There is a tradeoff between achieving a stronger radiation force and a larger trapping potential well, and thus a proper material parameter must be selected to balance both the radiation force and trapping potential well. In the case of axial backflow, optimization of the material properties was restricted to conditions that guarantee stable trapping. Subsequently, by balancing the electric/magnetic responses and increasing the magnitude of axial optical force, the influence of canonical momentum is suppressed, and the axial optical force is aligned with the Poynting momentum, thereby enabling stable reverse optical force. Our study enriches the in-depth understanding of optical backflow and provides applications in optical trapping, manipulation, and imaging.

    [29] O. V. Minin, I. V. Minin, Y. Cao. Optical magnet for nanoparticles manipulations based on optical vacuum cleaner concept. Proc. SPIE, 11845, 118451G(2021).

    [46] P. Shi, L. Du, C. Li. Transverse spin dynamics in structured electromagnetic guided waves. Proc. Natl. Acad. Sci. USA, 118, e2018816118(2021).

    [56] Y. Cheng, K. A. Oyesina, B. Xue. Directional dipole dice enabled by anisotropic chirality. Proc. Natl. Acad. Sci. USA, 120, e2301620120(2023).

    Tools

    Get Citation

    Copy Citation Text

    Xiangyang Xie, Peng Shi, Changjun Min, Xiaocong Yuan, "Optical backflow for the manipulations of dipolar nanoparticles," Photonics Res. 13, 2033 (2025)

    Download Citation

    EndNote(RIS)BibTexPlain Text
    Save article for my favorites
    Paper Information

    Category: Physical Optics

    Received: Mar. 4, 2025

    Accepted: Apr. 28, 2025

    Published Online: Jul. 18, 2025

    The Author Email: Peng Shi (pittshiustc@gmail.com)

    DOI:10.1364/PRJ.561198

    CSTR:32188.14.PRJ.561198

    Topics