Skip to content
BY 4.0 license Open Access Published by De Gruyter September 17, 2020

Strong spin–orbit interaction of photonic skyrmions at the general optical interface

  • Peng Shi ORCID logo , Luping Du EMAIL logo and Xiaocong Yuan EMAIL logo
From the journal Nanophotonics

Abstract

Photonic skyrmions have applications in many areas, including the vectorial and chiral optics, optical manipulation, deep-subwavelength imaging and nanometrology. Much effort has been focused on the experimental characterization of photonic skyrmions. Here, we give an insight into the spin and orbital features of photonic skyrmions constructed by the p-polarized and s-polarized surface waves at an interface with various electric and magnetic properties by analyzing the continuity of chirality, energy flow and momentum densities through the electric and magnetic interface. The continuity of chirality density indicates that the photonic skyrmion has a property of the optical transverse spin. Most importantly, the continuity of energy flow and momentum densities results in four spin–orbit interaction quantities, which indicate the gradient of electric polarizability or permeability governs the spin–orbit interaction of photonic skyrmions and leads to the discontinuity and even the reversal of spin orientation through the optical interface. Our investigations on the spin–orbit properties of photonic skyrmions, which can give rise to the spin-dependent force and topological unidirectional transportation, is thorough and can be extended to other classical wave, such as acoustic and fluid waves. The findings help in understanding the spin–orbit feature of photonic topological texture and in constructing further optical manipulation, sensing, quantum and topological techniques.

1 Introduction

Light carries spin angular momentum (SAM) and orbital angular momentum (OAM) [1], [2], [3], [4], [5], [6]. The SAM is associated with the ellipticity of polarization [1], and the OAM is determined by the vortex phase or the trajectory of optical beam [2], [3]. The interplay of these two angular momenta results in many fascinating phenomena [7], [8], [9], including spin–orbit coupling and conversion [10], [11], [12], the spin Hall effect of light [13], [14], [15], [16], [17], [18], the optical Magnus effect [19], [20], [21], the Coriolis effect [22], [23] and the Aharonov–Bohm effect [24]. In particular, the strong spin–orbit interactions between the evanescent waves and interfaces in the optical near-field lead to the discovery of the optical transvers spin [25], [26], [27], [28], [29], [30], [31], [32], [33] and the photonic analog to the quantum spin Hall effect [34], [35], [36], which have potential in the applications of directional scattering and transportation of photons, nanometrology [37], [38], [39], chiral quantum optics [40] and spin-based robust optical surface [41], [42], [43], [44]. Very recently, by considering the spin and complex vectorial properties of surface electromagnetic wave, a photonic skyrmion [45], which has chiral spin texture in line with the topological soliton named from the physicist Tony Skyrme, was discovered and attracts wide interests in the field of spin and topological optics. Most of the relevant research until now has focused on the topological number and the experimental measurement of photonic skyrmion [46], [47], [48], [49], [50]; and its spin and orbital features still need to be explained explicitly.

Here, we give an insight into the spin and orbital features of photonic skyrmions by analyzing the decomposition of energy flow density (EFD) and momentum density (MD) into the spin and orbital parts. Firstly, we investigate the energy flow densities and spin textures on both sides of an optical interface and find the EFD and skyrmion number reversal on the two sides. Secondly, we study the chirality feature of photonic skyrmion and find that their time-averaging chirality density vanishes; hence this spin texture is constituted by the optical transverse spin. In addition, we discuss and demonstrate the spin-dependent force, the topological origins and the resulted spin-momentum locking of photonic skyrmions. Finally and more importantly, we demonstrate that the physically meaningful equations of continuity for the EFD and MD lead to four ‘spin–orbit interaction’ terms, which are in line with electric-magnetic terms of the chirality interactions [51], [52], [53]. For photonic skyrmions constructed by the p-polarized surface waves, the electric polarizability related term governs the spin–orbit interaction of surface waves and results in the EFD reversal and an inverted out-of-plane orientation of spin texture through the optical interface. If a variation of permeability is introduced into the material in lower half-space, the magnetic susceptibility related term would lead to the MD reverse and the accompanied in-plane orientation of spin texture through the optical interface. This theoretical analysis, which is thorough and can be extended to the other classical waves, helps in understanding the spin–orbit features of photonic skyrmions and in constructing further optical manipulation, sensing, quantum and topological techniques.

2 Spin and orbital features of photonic skyrmions

Here, we firstly consider a p-polarized evanescent wave propagating at the z = 0 interface between an isotropic dielectric (in the z > 0 half-space with real permittivity ε+ and permeability μ+) and an isotropic medium (in the z < 0 half-space with real permittivity ε and permeability μ). The corresponding results for the s-polarized evanescent wave at a general interface, including the dispersion relation, field distributions and spin–orbit features, can be found in Supplementary materials II, IV and VI, respectively. The time-averaged Poynting vector Π of this Bessel-type evanescent wave in cylindrical coordinates (r, φ, z) with unit vector (rˆ,φˆ,zˆ) is [54]

(1)Π±=12Re{E±(r)×H±(r)}=|Α|2ωl2ε±rβ2Jl2(βr)Ψφˆ

The EFD propagates in the azimuthal direction and decays exponentially with factor Ψ=1/e±2κ±z in the z-direction, which results in the vanishing of out-of-plane EFD. Here, E(r) and H(r) denote respectively the electric and magnetic field; A is the normalized complex amplitude; Jl is the l-order Bessel function of first kind; the superscript ∗ indicates the complex conjugate operator of a vector; the symbols ‘+’ and ‘−’ indicate physical quantities in the z > 0 and z < 0 half-spaces; ω is the angular frequency; β is the in-plane propagation constant of the evanescent wave; ±± denotes the out-of-pane wave vectors, which satisfy the dispersion relation ω2ε±μ±=β2κ±2 (Supplementary material I). All the quantities, including β, ω, κ±, ε± and μ±, are considered as real numbers throughout the paper. The time-dependent quantity eiωt is ignored because only the time-averaging properties are considered here.

Generally, a Bessel-type surface wave can be excitated by a focused circularly polarized light (carry longitudinal spin) with an additional vortex phase at a dielectric/metal interface. The imaginary characters of the field components arise from the transversality condition ∇·D = 0 [55], which results in the ∓π/2 phase difference between the field in propagating direction Eφ and two transverse direction Er, Ez. Thus, the electric field vectors rotate in the (φ, z) and (r, φ)-planes simultaneously. In other words, a Bessel-type evanescent wave is elliptically polarized in the two propagation planes. The SAM in a complex medium can be expressed as [56], [57], [58]

(2)S±=14ω{ε±ΦE±+μ±ΦH±}=|Α|2β2ε±l2ωε±2r{±κ±Jl2(βr)rˆ,0φˆ,[β+(1η±Z±1)ε±μ±ω2β]Jl(βr)Jl(βr)zˆ}Ψ.

where ε±=[ωε±(r,ω)]/ω and μ±=[ωμ±(r,ω)]/ω are respectively the group permittivity and permeability; the contributions of electric ΦE and magnetic ΦH parts to the SAM, which are essentially related to the polarization ellipticities, are

(3)ΦE±=Im{E±×E±}=|Α|22lr(±β2κ±ε±2Jl2(βr)rˆ+βκ±2ε±2Jl(βr)Jl(βr)zˆ)Ψ

and

(4)ΦH±=Im{(H±×H±)}=|Α|22lrω2βJl(βr)Jl(βr)Ψzˆ,

respectively. Here, the last term of Equation (2) is stemmed from the dispersion and η±=ε±/μ±, Z±=μ±/ε±. The SAM components in the upper and lower half-spaces can be seen in Figure 1(a–f). For a dielectric material in the upper half-space, the out-of-plane and radial SAM components are positive at r = 0 for the topological charge l > 0. Thus, the spin state is ‘up’ in the center, and varies from ‘up’ state to ‘down’ state gradually as shown in Figure 1(g) and (j). Whereas in the lower space with permittivity ε < 0 and permeability μ > 0, the out-of-plane SAM component reverses, which makes the spin texture reversal as shown in Figure 1(h) and (k). In addition, if there is a magnetic material with ε > 0 and μ < 0, the out-of-plane SAM component is unchanged and the radial SAM component reverses due to the evanescent property in the lower half-space. Thus, although the spin state also varies from ‘up’ state to ‘down’ state gradually, the chirality of spin texture is reversal owing to the inversion of radial SAM component as exhibited in Figure 1(i) and (l).

Figure 1: The spin properties of p-polarized Bessel-type surface mode with topological charge l = +2. The radial spin angular momentum (SAM) (a) and out-of-plane SAM (d) components of surface mode in the upper half-space (at z = 10 nm) as the relative permittivity and permeability are 1; the corresponding 1D contour and normalized spin texture can be found in (g) and (j), respectively. The radial SAM (b) and out-of-plane SAM (e) components of surface mode in the lower half-space (at z = −10 nm) as the relative permittivity and permeability are respectively −11.8 and 1; the corresponding 1D contour and normalized spin texture can be found in (h) and (k), respectively. The radial SAM (c) and out-of-plane SAM (f) components of surface mode in the lower half-space (at z = −10 nm) as the relative permittivity and permeability are respectively 11.8 and −1.5; the corresponding 1D contour and normalized spin texture can be found in (i) and (l), respectively. The sign of out-of-plane spin orientation is determined by the permittivity, while the direction of in-plane spin orientation is related to the permeability. The wavelength is 633 nm and all quantities are normalized to the maximal absolute value of total SAM in the upper half-space.
Figure 1:

The spin properties of p-polarized Bessel-type surface mode with topological charge l = +2. The radial spin angular momentum (SAM) (a) and out-of-plane SAM (d) components of surface mode in the upper half-space (at z = 10 nm) as the relative permittivity and permeability are 1; the corresponding 1D contour and normalized spin texture can be found in (g) and (j), respectively. The radial SAM (b) and out-of-plane SAM (e) components of surface mode in the lower half-space (at z = −10 nm) as the relative permittivity and permeability are respectively −11.8 and 1; the corresponding 1D contour and normalized spin texture can be found in (h) and (k), respectively. The radial SAM (c) and out-of-plane SAM (f) components of surface mode in the lower half-space (at z = −10 nm) as the relative permittivity and permeability are respectively 11.8 and −1.5; the corresponding 1D contour and normalized spin texture can be found in (i) and (l), respectively. The sign of out-of-plane spin orientation is determined by the permittivity, while the direction of in-plane spin orientation is related to the permeability. The wavelength is 633 nm and all quantities are normalized to the maximal absolute value of total SAM in the upper half-space.

Recently, it is widely accepted that the electric field vector in the upper half-space has a Néel-type skyrmion-like distribution for the topological charge l = 0 [48], whereas for the topological charge l ≠ 0, the spin texture of this Bessel-type evanescent field is analogous to that of a magnetic skyrmion [45]. Here, we only discuss the skyrmion-like spin texture (l ≠ 0). Since the spin vectors only have radial and out-of-plane components and depend on the radial coordinate r, the directional unit vector of SAM can be converted to sr±=Sr±/|S±|=cosΘr± and sz±=Sz±/|S±|=cosΘr± with |S±|=Sr±2+Sz±2. Simple derivation yields the skyrmion number of the normalized SAM:

(5)n±=14πCs±[(s±/x)×(s±/y)]dxdy=12riri+1cosΘr±Θr±rdr=sgn(ε±l).

Note here that the integral region C, which is defined by the centroid of beam r = ri and the boundary r = ri+1, is circular symmetrical and chosen precisely from the ‘up’ to ‘down’ states of spin vectors. These interesting results indicate the skyrmion-like spin textures are determined by the signs of vortex topological charge and permittivity. As the sign of permittivity reverses in the lower half-space, the spin vectors also reverse, which can be intuitively seen in Figure 1(j–l).

Through these Equations (2)–(4), we can recognize the spin properties of photonic skyrmions in three aspects. Firstly, this great electric-magnetic asymmetry of SAM is originated from the breaking of the inverse symmetry owing to the presence of the optical interface. Naturally, the spin Chern number of photonic skyrmions equals 4 (exactly speaking, +4 in dielectric material, −4 in left-handed material and imaginary in non-Hermitian system; see the derivations in Supplementary material IX), and hence the optical modes are double-degenerate and there would be two pairs of surface modes (s-mode and p-mode). However, the existence of an interface between media with different permittivity and permeability breaks the dual symmetry between the electric and magnetic features and only one polarized state of surface evanescent modes is allowed due to the breaking of the polarization degeneracy [59], [60], [61], [62], [63]. Thus, photonic skyrmions are governed by the topological origin of nontrivial spin Chern number and suffer from spin-momentum locking. Secondly, there are radial and out-of-plane components exist for this Bessel-type evanescent waves, which are corresponding to the rotation of electric field vectors in the (φ, z) and (r, φ)-planes, respectively. Moreover, the directions of SAM are perpendicular to the propagating direction of EFD, which indicates that the SAM of these photonic skyrmions can be regarded as transverse spin.

Thirdly, we now explore an important connection between the spin and chirality of the surface wave to explain this transverse feature of spin. Similar with the energy density and EFD Π, one can derivate that the chirality density and chirality flow Ʃ also satisfy a continuity equation, which can be utilized to characterize the chirality or spin properties of the electromagnetic field [53], [64], [65]. In a homogeneous medium, the time-averaged chirality density and chirality flow can be written as

Here, n±=[ωn±(r,ω)]/ω is the group refractive index. Substituting here the Bessel-type field distributions (Supplementary material III), we immediately reach

Thus, the chirality density vanishes since Im{H*·E}=0 in the whole space independent of the optical interface. A nonzero chirality flow is determined by the ellipticities of the field polarization stemming from the transversality condition. To further understand helical properties of photonic skyrmions, one can compare these results to those of the propagating fields in free space. For the propagating fields, the integral of chirality density in a transverse plane is proportional to the averaged helicity of photons, which is determined by the longitudinal spin: Sˆpˆ/|n|k0 [31]. Here, Sˆ and pˆ are the spin and canonical momentum operators; k0 is wave vector of photons in vacuum; |n| is the absolute value of refractive index; the Dirac notation denotes the inner product. Obviously, the vanishing of the chirality density definitely indicates that the SAM is orthogonal to the momentum and can be regarded as transverse spin universally. On the other hand, the chirality flow is proportional to the SAM [26]:

(8)Σ±v±2=12Im{ε±E±×E±+μ±H±×H±}=2ωS±.

where v±=1/ε±μ±. Thus, although the chirality density vanishes, the appearance of ‘local’ chirality flow can induce helix-dependent torque, which has potential in the applications of optical spinning and manipulation of nanoparticles [66], [67], [68], [69], [70].

To demonstrate this helix-dependent mechanical effect, the optical torque experienced by the particle in the upper half-space can be written as [71], [72]:

(9)Γ=12Re[p×E+m×H]=12Re[αeE×E+αhH×H]=Im(αe)2Im(E×E)+Im(αh)2Im(H×H)=Im(αe)2ΦE+Im(αh)2ΦH

Here, p = αeE (m = αhH) is the electric (magnetic) dipole induced by the electric (magnetic) field of the incident electromagnetic wave with αe and αh the electric and magnetic polarizabilities. The optical torque for photonic skyrmion is obviously helix-dependent. Note here that the SAM and the optical torque are perpendicular to the propagating direction of photonic skyrmion, which indicates the transverse property of SAM and optical torque. In addition, the total optical force experienced by the particle can be written as (Supplementary material VII):

(10)Fi=12Re[pjiEj+mjiHj2k43εijkpjmk],

where εijk is the Levi-Civita tensor, and i, j, or k stands for either x, y or z. For simplicity, we only consider the nonmagnetic spherical Rayleigh particle and ignore the crossover term in Eq. (10). The azimuthal optical force (along the energy propagation direction) is depended by the longitudinal spin (helix) of the incident light. For example, owing to the conservation law of total angular momentum, the photonic skyrmions of ±1 order can be excitated by a circularly polarized light with helical degree σ = ±1. Thus, parts of the longitudinal optical forces are also helix-dependent.

3 Spin and orbital features of photonic skyrmions

Finally, we investigate the spin–orbit coupling in the optical interface. The Poynting vector determines the spin flow density (SFD: Ps) and orbital flow density (OFD: Po) of electromagnetic wave in complex medium (Supplementary material V):

(11)Ps±=18ω×{1μ±ΦE±+1ε±ΦH±}=|Α|2β2l4ωε±2μ±(2κ±2Jl2(βr)r[β+(1η±Z±1)ε±μ±ω2β]r[1rJl(βr)Jl(βr)])Ψφˆ,
(12)Po±=14ωIm{1μ±E±()E±+1ε±H±()H±}=lβ24ωε±2μ±{[β2+l2r2[1+(1η±Z±1)ω2ε±μ±β2]]Jl2(βr)+β2[1+(1η±Z±1)ω2ε±μ±β2]Jl2(βr)}Ψφˆ.

The SFD is determined by the vorticities of the polarization ellipticities ΦE and ΦH, whereas the OFD is essentially related to the canonical momentum of a single wavepacket Po±p±ψ±|pˆ|ψ±/ω with pˆ=i the momentum operator and |ψ±=[εE±,iμH±] the photon wave function and the reduced Planck constant (Supplementary material V). The distributions of EFD, OFD and SFD in the upper and lower half-spaces with vortex topological charge l = +2 can be found in Figure 2. Naturally, the EFD is determined by sgn(l/ε) and the OFD is related to sgn(l/μ′). Thus, the EFD and OFD are parallel in the dielectric material, whereas there are antiparallel in the noble metal or negative magnetic right-handed materials. Moreover, in the left-handed material, there are transformed into parallel. In addition, the SFD is always antiparallel to the OFD, which makes the group velocity of light subluminal.

Figure 2: Energy flow density (EFD), orbital and spin flow densities of p-polarized Bessel-type surface mode with topological charge l = +2. The azimuthal EFD (a), orbital flow density (OFD) (b) and spin flow density (SFD (c) of surface mode in the upper half-space (at z = 10 nm) as the relative permittivity and permeability are 1; the azimuthal EFD (d), OFD (e) and SFD (f) of surface mode in the lower half-space (at z = −10 nm) as the relative permittivity and permeability are respectively −11.8 and 1; the azimuthal EFD (g), OFD (h) and SFD (i) of surface mode in the lower half-space (at z = −10 nm) as the relative permittivity and permeability are respectively 11.8 and −1.5. The EFD is determined by the sgn(l/ε), while the OFD is related to the sgn(l/μ).The wavelength is 633 nm. All quantities are normalized to the maximal absolute value of energy flow in the upper half-space.
Figure 2:

Energy flow density (EFD), orbital and spin flow densities of p-polarized Bessel-type surface mode with topological charge l = +2. The azimuthal EFD (a), orbital flow density (OFD) (b) and spin flow density (SFD (c) of surface mode in the upper half-space (at z = 10 nm) as the relative permittivity and permeability are 1; the azimuthal EFD (d), OFD (e) and SFD (f) of surface mode in the lower half-space (at z = −10 nm) as the relative permittivity and permeability are respectively −11.8 and 1; the azimuthal EFD (g), OFD (h) and SFD (i) of surface mode in the lower half-space (at z = −10 nm) as the relative permittivity and permeability are respectively 11.8 and −1.5. The EFD is determined by the sgn(l/ε), while the OFD is related to the sgn(l/μ).The wavelength is 633 nm. All quantities are normalized to the maximal absolute value of energy flow in the upper half-space.

It is worth noting that, in the complex medium, the relation Π = Po + Ps does not satisfy because there will be an addition dispersive term existing:

(13)PsD±=18ω×{ωμ±(r,ω)/ωμ±μ±ΦE±+ωε±(r,ω)/ωε±ε±ΦH±},

and

(14)PoD±=14ωIm{ωμ±(r,ω)/ωμ±μ±E±()E±+ωε±(r,ω)/ωε±ε±H±()H±}.

where the superscript “D” indicates the dispersion. Thus, there must be

(15)Π=Po+PoD+Ps+PsD

Moreover, the dispersive terms also play critical roles in the spin–orbit interaction. Thus, one can summarize the contributions of spin and orbital parts to the EFD as: Πs=Ps+PsD and Πo=Po+PoD. As in the plane wave cases given in a study by Bliokh et al. [26], the decomposition of EFD into spin and orbit parts works well in a homogeneous medium, but in the presence of inhomogeneity, the spin and orbit parts acquire nonzero divergences, Πs=Πo0, which does not make physical sense since Π=0. One can modify the separation of the spin and orbital parts of EFD to satisfy Πo=Πs=0. In this manner, we obtain Πs=ΠsΞ and Πo=Πo+Ξ with Π=Πo+Πs:

(16)Ξ=ΞHE+ΞEH=18ω1μ×ΦE+18ω1ε×ΦH

These quantities ΞHE and ΞHE arise from the optical potential and can describe the spin–orbit interaction, which vanishes in a homogeneous medium but would play a critical role in optical interfaces. Owing to the abovementioned polarization properties of the Bessel-type surface waves, the electric and magnetic field ellipticities do not vanish. In our case, the inhomogeneity introduced by the optical interface is along the z-direction, which means that ∇(1/ε) and ∇(1/μ) only have out-of-plane components. Thus, the term ΞEH ∝ ∇(1/ε) × ΦH always vanishes for the p-polarized modes, whereas the other term ΞHE ∝ ∇(1/μ) × ΦE is along the azimuthal direction, which results in the strong spin–orbit interaction and even the reversal of EFD. From Eq. (9) with fixed vortex topological charge l, it can be found that the direction of Poynting vector is determined by the permittivity ε. If the ε has an opposite sign with the ε+ (such as surface plasmon polaritons at the air/metal interface), the direction of EFD in the z < 0 half-space is reversed comparing to that in the z > 0 half-space, while the reversal of permeability would not affect the energy propagating direction. However, this term ΞHE is intensely determined by the gradient of permeability. It is nonlogical because there also would be spin–orbit coupling for the nonmagnetic surface mode, for example, the photonic skyrmion suffers from great twisting and the skyrmion number reverses through the optical interface with invariable permeability. Thus, we introduce another two physical quantities to characterize these permittivity or permeability depended spin–orbit interaction.

The Poynting vector determines the density of the MD of the field [73]:

(17)π±(r)=Π±(r)v±2=|Α|2ωμ±l2rβ2Jl2(βr)Ψφˆ.

In our case, for the Bessel-type surface wave, the direction of MD is definitely determined by the permeability μ± from the Equation (17). The sign of MD need not reverse through the nonmagnetic optical interface and hence it is continuous ∇·π = 0. As acted in the homogeneous space, the MD can be decomposed into the contributions of spin and orbital parts. Note here that we do not follow the analysis above in the EFD because it is more concise to employ the spin and orbital parts of MD instead of defining spin and orbital momentum densities or dispersion related spin and orbital momentum densities to analyze the spin–orbit terms. Note here the spin parts and orbital parts are related to the decomposition of EFD into the spin and orbital parts by

(18)πs±=18ω×{ε±ΦE±+μ±ΦH±}=Πs±/v±2,
(19)πo±=14ωIm{ε±E±()E±+μ±H±()H±}=Πs±/v±2.

These two quantities suffer from nonzero divergences: πs=πo0, which do not make physical sense since ∇⋅π = 0. Here, we modify the separation of the SMD and OMD, π=πo+πs, such that πo=πs=0. In this manner, we obtain πs=πsΞ and πo=πo+Ξ with

(20)Ξ=ΞEE+ΞHH=18ωεμε×ΦE+18ωεμμ×ΦH.

It is worth noting that we utilize a factor 1/εμ to unify the dimensions of physical quantities Ξ and Ξ′. For the p-polarized surface waves, due to the discontinuity of ∇ε at the optical interface, strong counter-propagating boundary spin and orbital energy flows arise there. By examining the spin–orbit interaction terms given by Eqs. (16) and (20), we can indicate that ΞEE and ΞHH are the electric polarizability and the magnetic susceptibility related spin–orbit interaction terms, respectively; ΞHE and ΞEH are the mixed electric-magnetic terms. Meanwhile, the boundary EFD is

(21)δΠ=|Α|2ωl2rβ2Jl2(βr)(1ε+1ε)δ(z)φˆ,

where δ(z) is the Dirac function. On the other hand, there is only ΞEE exist for the p-polarized surface wave at the interface of dielectric and nonmagnetic metal, and it can be expressed as

(22)ΞEE=|Α|2lrβ2κ+4ωε+3μ+Jl2(βr)εδ(z)φˆ=κ+ε2ω2ε+2μ+δΠ.

The scaling factor κ+ε/(2ω2ε+2μ+) is positive and hence we can conclude that the strong z gradient of permittivity for p-polarized field leads to the strong spin–orbit coupling and the reversal of EFD, which results in the reversal of out-of-plane spin orientation of photonic skyrmion as shown in Figure 1(b) and (e) at the same time. In addition, if the medium of lower half-space is magnetic with positive permittivity, the boundary MD and the spin–orbit interaction term caused by the z-gradient of permeability are:

(23)δπ=|Α|2ωl2rβ2Jl2(βr)(μ+μ)δ(z)φˆ,

and

(24)ΞHE=|Α|2lrβ2κ+4ωε+2Jl2(βr)1μδ(z)φˆ=κ+2ω2ε+2μ+μδπ.

Obviously, the scaling factor κ+/(2ω2ε+2μ+μ) is positive. Thus, this quantity is also proportional to the boundary MD and can lead to the reversal of MD and the accompanied in-plane spin orientation of photonic skyrmion as shown in Figure 1(c) and (f). In addition, it worth noting that, for the s-polarized surface modes, these spin–orbit interaction processes would be governed by the terms ΞEH and ΞHH(Supplementary material VI). We summarize the physical quantities discussed above, including the spin–orbit interaction terms, EFD, MD, OFD and skyrmion topological number of p-polarized and s-polarized surface waves in various electric, magnetic and left-handed interfaces in Figure 3.

Figure 3: Summary of physical quantities for the p-polarized and s-polarized surface waves in various electric, magnetic and left-handed interfaces. Here, topological charge l > 0; if the l is inverse, the direction of all quantities are inverted. The corresponding available “spin–orbit interaction” terms are given in the end.
Figure 3:

Summary of physical quantities for the p-polarized and s-polarized surface waves in various electric, magnetic and left-handed interfaces. Here, topological charge l > 0; if the l is inverse, the direction of all quantities are inverted. The corresponding available “spin–orbit interaction” terms are given in the end.

Taking into account the “spin–orbit” correction given in Eq. (20), the boundary MDs are

(25)δπs=δπo=|Α|2lβ24ωr[2Jl2(βr)(κ+2ε+κ2ε)+βJl(βr)Jl(βr)r(1ε+1ε)]δ(z)φˆ.

Thus, in the upper half-space, the evanescent waves possess a backward spin part. This backward spin part is subtracted from the forward orbital part to give the total MD and ensures proper local energy transport in evanescent electromagnetic waves. On the other hand, in the lower half-space, the intensity of orbital part (corresponding to the canonical momentum) is much weaker than that of spin part. This is logical because the electromagnetic field in the lower half-space is much more localized and cannot be regarded as propagating field.

4 Conclusion

In conclusion, we investigate the spin and orbital features of photonic skyrmions in both sides of optical interface. For the photonic skyrmion constructed by the p-polarized surface wave, the directions of EFD and spin vector would be reversal due to the inverse of permittivity through the interface, whereas the MD reverse owing to the opposite magnetic permeability in the two sides of interface. By considering the chirality feature of photonic skyrmion, we find that the time-averaging chirality density of photonic skyrmion vanishes, which means that this spin texture is constituted by the optical transverse spin. In addition, by investigating the physically meaningful equations of continuity for the SFD, OFD, spin MD and orbital MD, we obtain four ‘spin–orbit interaction’ terms given by ΞHE, ΞEH, ΞEE and ΞHH, respectively. For the p-polarized surface wave in the dielectric/metal interface, the term ΞEE, which is related to the electric polarizability, governs the spin–orbit interaction and results in the reversal of EFD and the accompanied out-of-plane orientation of skyrmion-like spin texture through the optical interface. If a variation of magnetic permeability is introduced into the material in lower half-space, the term ΞHE, which is determined by the z gradient of the magnetic permeability, would also affect the reverse of the MD and the accompanied in-plane orientation of skyrmion-like spin texture through the optical interface. The presented theoretical analysis have beneficial to understand the spin–orbit features for photonic skyrmions and can be extended to the other classical waves [74], [75], [76], [77], [78], [79], such as: acoustic and fluid waves (Supplementary material VIII). Our findings can construct further applications in the fields of optical manipulation, sensing, quantum and topological techniques.


Corresponding authors: Xiaocong Yuan and Luping Du, Nanophotonics Research Center, Shenzhen Key Laboratory of Micro-Scale Optical Information Technology, Institute of Microscale Optoelectronics, Shenzhen University, Shenzhen, 518060, China, E-mail: (X. Yuan), (L. Du)

Funding source: Leading talents of Guangdong province program

Award Identifier / Grant number: 00201505

Award Identifier / Grant number: U1701661, 61935013, 61490712, 61427819, 61622504, 11504244, 61705135, and 61905163

Award Identifier / Grant number: 2016A030312010

Award Identifier / Grant number: KQTD2015071016560101, KQTD2017033011044403, ZDSYS201703031605029, KQTD2018041218324255, and JCYJ20180507182035270

Award Identifier / Grant number: 2016A030312010

Funding source: Shenzhen University

Award Identifier / Grant number: 2019074

Acknowledgments

The authors acknowledge the support from National Natural Science Foundation of China (NSFC) (grants no. U1701661, 61935013, 61490712, 61427819, 61622504, 11504244, 61705135, and 61905163); Leading talents of Guangdong province program (grant 00201505); the Natural Science Foundation of Guangdong Province (grant 2016A030312010); the Shenzhen Science and Technology Innovation Commission (grants no. KQTD2015071016560101, KQTD2017033011044403, ZDSYS201703031605029, KQTD2018041218324255, and JCYJ20180507182035270); Shenzhen University (2019074). L. D. acknowledges the support given by the Guangdong Special Support Program.

  1. Author contribution: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: This research was supported by the National Natural Science Foundation of China (NSFC) (grants no. U1701661, 61935013, 61490712, 61427819, 61622504, 11504244, 61705135, and 61905163); Leading talents of Guangdong province program (grant 00201505); the Natural Science Foundation of Guangdong Province (grant 2016A030312010); the Shenzhen Science and Technology Innovation Commission (grants no. KQTD2015071016560101, KQTD2017033011044403, ZDSYS201703031605029, KQTD2018041218324255, and JCYJ20180507182035270); Shenzhen University (2019074).

  3. Conflict of interest statement: The authors declare no conflicts of interest.

References

[1] R. A. Beth, “Mechanical detection and measurement of the angular momentum of light,” Phys. Rev., vol. 50, pp. 115–125, 1936, https://doi.org/10.1103/physrev.50.115.Search in Google Scholar

[2] L. Allen, M. W. Beijersbergen, R. J. Spreeuw, and J. P. Woerdman, “Orbital angular momentum of light and the transformation of Laguerre-Gaussian laser modes,” Phys. Rev. A, vol. 45, no. 11, pp. 8185–8189, 1992, https://doi.org/10.1103/physreva.45.8185.Search in Google Scholar

[3] Y. Shen, X. Wang, Z. Xie, et al., “Optical vortices 30 years on: OAM manipulation from topological charge to multiple singularities,” Light: Sci. Appl., vol. 8, pp. 90:1–29, 2019, https://doi.org/10.1038/s41377-019-0194-2.Search in Google Scholar

[4] C.-F. Li, “Spin and orbital angular momentum of a class of nonparaxial light beams having a globally defined polarization,” Phys. Rev. A, vol. 80, no. 6, pp. 063814:1–11, 2009, https://doi.org/10.1103/physreva.80.063814.Search in Google Scholar

[5] D. A. Smirnova, V. M. Travin, K. Y. Bliokh, and F. Nori, “Relativistic spin-orbit interactions of photons and electrons,” Phys. Rev. A, vol. 97, pp. 043840:1–8, 2018, https://doi.org/10.1103/physreva.97.043840.Search in Google Scholar

[6] A. T. O’Neil, I. MacVicar, L. Allen, and M. J. Padgett, “Intrinsic and extrinsic nature of the orbital angular momentum of a light beam,” Phys. Rev. Lett., vol. 88, pp. 053601:1–4, 2002. https://doi.org/10.1103/PhysRevLett.88.053601.Search in Google Scholar

[7] K. Y. Bliokh, F. J. Rodríguez-Fortuño, F. Nori, and A. V. Zayats, “Spin–orbit interactions of light,” Nat. Photonics, vol. 9, no. 12, pp. 796–808, 2015, https://doi.org/10.1038/nphoton.2015.201.Search in Google Scholar

[8] F. Cardano and L. Marrucci, “Spin–orbit photonics,” Nat. Photonics, vol. 9, no. 12, pp. 776–778, 2015, https://doi.org/10.1038/nphoton.2015.232.Search in Google Scholar

[9] L. T. Vuong, A. J. L. Adam, J. M. Brok, P. C. M. Planken, and H. P. Urbach, “Electromagnetic spin-orbit interactions via scattering of subwavelength apertures,” Phys. Rev. Lett., vol. 104, no. 8, p. 083903, 2010, https://doi.org/10.1103/physrevlett.104.083903.Search in Google Scholar

[10] 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., vol. 99, pp. 073901:1–4, 2007, https://doi.org/10.1103/physrevlett.99.073901.Search in Google Scholar

[11] K. Y. Bliokh, E. A. Ostrovskaya, M. A. Alonso, O. G. Rodríguez-Herrera, D. Lara, and C. Dainty, “Spin-toorbital angular momentum conversion in focusing, scattering, and imaging systems,” Opt. Express, vol. 19, no. 27, pp. 26132–26149, 2011, https://doi.org/10.1364/oe.19.026132.Search in Google Scholar

[12] L. Marrucci, C. Manzo, and D. Paparo, “Optical spin-to-orbital angular momentum conversion in inhomogeneous anisotropic media,” Phys. Rev. Lett., vol. 96, no. 16, pp. 163905:1–4, 2006, https://doi.org/10.1103/physrevlett.96.163905.Search in Google Scholar

[13] M. Onoda, S. Murakami, and N. Nagaosa, “Hall effect of light,” Phys. Rev. Lett., vol. 93, pp. 083901:1–4, 2004, https://doi.org/10.1103/physrevlett.93.083901.Search in Google Scholar

[14] K. Y. Bliokh and Y. P. Bliokh, “Conservation of angular momentum, transverse shift, and spin Hall effect in reflection and refraction of an electromagnetic wave packet,” Phys. Rev. Lett., vol. 96, pp. 073903:1–4, 2006, https://doi.org/10.1103/physrevlett.96.073903.Search in Google Scholar

[15] P. Deng, W. Hong, L. Gao, and H. Xu, “Strong spin-orbit interaction of light in plasmonic nanostructures and nanocircuits,” Phys. Rev. Lett., vol. 117, pp. 166803:1–5, 2016, https://doi.org/10.1103/physrevlett.117.166803.Search in Google Scholar

[16] X. Yin, Z. Ye, J. Rho, Y. Wang, and X. Zhang, “Photonic spin Hall effect at metasurfaces,” Science, vol. 339, pp. 1405–1407, 2013, https://doi.org/10.1126/science.1231758.Search in Google Scholar

[17] X. Ling, X. Zhou, K. Huang, et al., “Recent advances in the spin Hall effect of light,” Rep. Prog. Phys., vol. 80, pp. 066401:1–17, 2017, https://doi.org/10.1088/1361-6633/aa5397.Search in Google Scholar

[18] H. Wang and X. Zhang, “Unusual spin Hall effect of a light beam in chiral metamaterials,” Phys. Rev. A, vol. 83, pp. 053820:1–9, 2011, https://doi.org/10.1103/physreva.83.053820.Search in Google Scholar

[19] K. Y. Bliokh, A. Niv, V. Kleiner, and E. Hasman, “Geometrodynamics of spinning light,” Nat. Photoncs, vol. 2, pp. 748–753, 2008, https://doi.org/10.1038/nphoton.2008.229.Search in Google Scholar

[20] K. Y. Bliokh, “Geometrodynamics of polarized light: Berry phase and spin Hall effect in a gradient-index medium,” J. Opt. Pure Appl. Opt., vol. 11, pp. 094009:1–14, 2009, https://doi.org/10.1088/1464-4258/11/9/094009.Search in Google Scholar

[21] H. Luo, S. Wen, W. Shu, and D. Fan, “Role of transverse-momentum currents in the optical Magnus effect in free space,” Phys. Rev. A, vol. 81, pp. 053826:1–10, 2010, https://doi.org/10.1103/physreva.81.053826.Search in Google Scholar

[22] K. Y. Bliokh, Y. Gorodetski, V. Kleiner, and E. Hasman, “Coriolis effect in optics: unified geometric phase and spin-Hall effect,” Phys. Rev. Lett., vol. 101, no. 3, pp. 030404:1–4, 2008, https://doi.org/10.1103/physrevlett.101.030404.Search in Google Scholar

[23] Y. Gorodetski, A. Niv, V. Kleiner, and E. Hasman, “Observation of the spin-based plasmonic effect in nanoscale structures,” Phys. Rev. Lett., vol. 101, no. 4, pp. 043903:1–4, 2008, https://doi.org/10.1103/physrevlett.101.043903.Search in Google Scholar

[24] Y. Gorodetski, S. Nechayev, V. Kleiner, and E. Hasman, “Plasmonic Aharonov-Bohm effect: optical spin as the magnetic flux parameter,” Phys. Rev. B, vol. 82, no. 12, pp. 125433:1–4, 2010, https://doi.org/10.1103/physrevb.82.125433.Search in Google Scholar

[25] A. Aiello, P. Banzer, M. Neugebauer, and G. Leuchs, “From transverse angular momentum to photonic wheels,” Nat. Photonics, vol. 9, no. 12, pp. 789–795, 2015, https://doi.org/10.1038/nphoton.2015.203.Search in Google Scholar

[26] K. Y. Bliokh and F. Nori, “Transverse spin of a surface polariton,” Phys. Rev. A, vol. 85, no. 6, pp. 061801:1–5, 2012, https://doi.org/10.1103/physreva.85.061801.Search in Google Scholar

[27] K. Y. Bliokh, A. Y. Bekshaev, and F. Nori, “Extraordinary momentum and spin in evanescent waves,” Nat. Commun., vol. 5, no. 1, pp. 33001:1–8, 2014, https://doi.org/10.1038/ncomms4300.Search in Google Scholar

[28] M. Neugebauer, T. Bauer, A. Aiello, and P. Banzer, “Measuring the transverse spin density of light,” Phys. Rev. Lett., vol. 114, pp. 063901:1–5, 2015, https://doi.org/10.1103/physrevlett.114.063901.Search in Google Scholar

[29] M. Neugebauer, J. S. Eismann, T. Bauer, and P. Banzer, “Magnetic and electric transverse spin density of spatially confined light,” Phys. Rev. X, vol. 8, pp. 021042:1–8, 2018, https://doi.org/10.1103/physrevx.8.021042.Search in Google Scholar

[30] L. Peng, L. Duan, K. Wang, et al., “Transverse photon spin of bulk electromagnetic waves in bianisotropic media,” Nat. Photonics, vol. 13, pp. 878–882, 2019, https://doi.org/10.1038/s41566-019-0521-4.Search in Google Scholar

[31] K. Y. Bliokh and F. Nori, “Transverse and longitudinal angular momenta of light,” Phys. Rep., vol. 592, pp. 1–38, 2015, https://doi.org/10.1016/j.physrep.2015.06.003.Search in Google Scholar

[32] P. Shi, L.-P. Du, and X.-C. Yuan, “Structured spin angular momentum in highly focused cylindrical vector vortex beams for optical manipulation,” Opt. Express, vol. 26, no. 8, pp. 23449–23459, 2018, https://doi.org/10.1364/oe.26.023449.Search in Google Scholar

[33] P. Shi, L.-P. Du, and X.-C. Yuan, “Optical manipulation with electric and magnetic transverse spin through multilayered focused configuration,” Appl. Phys. Express, vol. 12, pp. 032001:1–5, 2019, https://doi.org/10.7567/1882-0786/aafca1.Search in Google Scholar

[34] K. Y. Bliokh, D. Smirnova, and F. Nori, “Quantum spin Hall effect of light,” Science, vol. 348, pp. 1448–1451, 2015, https://doi.org/10.1126/science.aaa9519.Search in Google Scholar

[35] T. V. Mechelen and Z. Jacob, “Universal spin-momentum locking of evanescent waves,” Optica, vol. 3, no. 2, pp. 118–126, 2016. https://doi.org/10.1364/OPTICA.3.000118.Search in Google Scholar

[36] P. Shi, L.-P. Du, C.-C. Li, A. Zayats, and X.-C. Yuan, “Spin-momentum law for structured guided modes: The generalized quantum spin-Hall effect for light,” arXiv:1910.03904, 2019.Search in Google Scholar

[37] Xi Zheng, Wei Lei, A. J. L. Adam, and H. P. Urbach, “Accurate feeding of nanoantenna by singular optics for nanoscale translational and rotational displacement sensing,” Phys. Rev. Lett., vol. 117, 2016, Art no. 113903, https://doi.org/10.1103/physrevlett.117.113903.Search in Google Scholar

[38] M. Neugebauer, P. Wozniak, A. Bag, G. Leuchs, and Peter Banzer, “Polarization-controlled directional scattering for nanoscopic position sensing,” Nat. Commun., vol. 7, pp. 11286:1–6, 2016, https://doi.org/10.1038/ncomms11286.Search in Google Scholar

[39] L. Wei, A. V. Zayats, and F. J. Rodríguez-Fortuño, “Interferometric evanescent wave excitation of a nanoantenna for ultrasensitive displacement and phase metrology,” Phys. Rev. Lett., vol. 121, pp. 193901:1–6, 2018, https://doi.org/10.1103/physrevlett.121.193901.Search in Google Scholar

[40] P. Lodahl, S. Mahmoodian, S. Stobbe, et al., “Chiral quantum optics,” Nature, vol. 541, pp. 473–480, 2017, https://doi.org/10.1038/nature21037.Search in Google Scholar

[41] J. Petersen, J. Volz, and A. Rauschenbeutel, “Chiral nanophotonic waveguide interface based on spin-orbit interaction of light,” Science, vol. 346, no. 6205, pp. 67–71, 2014, https://doi.org/10.1126/science.1257671.Search in Google Scholar

[42] Z. Shao, J. Zhu, Y. Chen, Y. Zhang, and S. Yu, “Spin-orbit interaction of light induced by transverse spin angular momentum engineering,” Nat. Commun., vol. 9, no. 926, pp. 1–11, 2018, https://doi.org/10.1038/s41467-018-03237-5.Search in Google Scholar

[43] D. O’Connor, P. Ginzburg, F. J. Rodríguez-Fortuño, G. A. Wurtz, and A. V. Zayats, “Spin-orbit coupling in surface plasmon scattering by nanostructures,” Nat. Commun., vol. 5, no. 1, pp. 53271:1–7, 2014. https://doi.org/10.1038/ncomms6327.Search in Google Scholar

[44] F. J. Rodríguez-Fortuño, G. Marino, P. Ginzburg, et al., “Near-field interference for the unidirectional excitation of electromagnetic guided modes,” Science, vol. 340, pp. 328–330, 2013, https://doi.org/10.1126/science.1233739.Search in Google Scholar

[45] L. P. Du, A. P. Yang, A. V. Zayats, and X. C. Yuan, “Deep-subwavelength features of photonic skyrmions in a confined electromagnetic field with orbital angular momentum,” Nat. Phys., vol. 15, pp. 650–654, 2019, https://doi.org/10.1038/s41567-019-0487-7.Search in Google Scholar

[46] T. V. Mechelen and Z. Jacob, “Photonic Dirac monopoles and skyrmions: spin-1 quantization,” Opt. Mater. Express, vol. 9, pp. 95–111, 2019. https://doi.org/10.1364/OME.9.000095.Search in Google Scholar

[47] S. Tsesses, K. Cohen, E. Ostrovsky, B. Gjonaj, and G. Bartal, “Spin-orbit interaction of light in plasmonic lattices,” Nano Lett., vol. 19, pp. 4010–4016, 2019, https://doi.org/10.1021/acs.nanolett.9b01343.Search in Google Scholar

[48] S. Tsesses, E. Ostrovsky, K. Cohen, B. Gjonaj, N. H. Lindner, and G. Bartal, “Optical skyrmion lattice in evanescent electromagnetic fields,” Science, vol. 361, pp. 993–996, 2018, https://doi.org/10.1126/science.aau0227.Search in Google Scholar

[49] T. J. Davis, D. Janoschka, P. Dreher, B. Frank, F. J. Meyer zu Heringdorf, and H. Giessen, “Ultrafast vector imaging of plasmonic Skyrmion dynamics with deep subwavelength resolution,” Science, vol. 368, pp. eaba6415:1–6, 2020. https://doi.org/10.1126/science.aba6415.Search in Google Scholar

[50] Y. Dai, Z. Zhou, A. Ghosh, et al., “Ultrafast microscopy of a plasmonic spin Skyrmion,” arXiv:1912.03826, 2019.Search in Google Scholar

[51] L. D. Barron, Molecular Light Scattering and Optical Activity, Cambridge, England, Cambridge University Press, 2004.10.1017/CBO9780511535468Search in Google Scholar

[52] A. Salam, Molecular Quantum Electrodynamics: Long-range Intermolecular Interactions, New York, Wiley, 2009.10.1002/9780470535462Search in Google Scholar

[53] Y. Tang and A. E. Cohen, “Optical chirality and its interaction with matter,” Phys. Rev. Lett., vol. 104, pp. 163901:1–4, 2010, https://doi.org/10.1103/physrevlett.104.163901.Search in Google Scholar

[54] A. Bekshaev, K. Y Bliokh, and M. Soskin, “Internal flows and energy circulation in light beams,” J. Opt., vol. 13, no. 5, pp. 053001:1–32, 2011, https://doi.org/10.1088/2040-8978/13/5/053001.Search in Google Scholar

[55] K. Y. Bliokh, M. A. Alonso, E. A. Ostrovskaya, and A. Aiello, “Angular momenta and spin-orbit interaction of nonparaxial light in free space,” Phys. Rev. A, vol. 82, no. 6, pp. 063825:1–7, 2010, https://doi.org/10.1103/physreva.82.063825.Search in Google Scholar

[56] F. Alpeggiani, K. Y. Bliokh, F. Nori, and L. Kuipers, “Electromagnetic helicity in complex media,” Phys. Rev. Lett., vol. 120, pp. 243605:1–6, 2018, https://doi.org/10.1103/physrevlett.120.243605.Search in Google Scholar

[57] M. F. Picardi, K. Y. Bliokh, F. J. Rodríguez-Fortuño, F. Alpeggiani, and F. Nori, “Angular momenta, helicity, and other properties of dielectric-fiber and metallic-wire modes,” Optica, vol. 5, pp. 1016–1026, 2018, https://doi.org/10.1364/optica.5.001016.Search in Google Scholar

[58] K. Y. Bliokh, A. Y. Bekshaev, and F. Nori, “Optical momentum, spin, and angular momentum in dispersive media,” Phys. Rev. Lett., vol. 119, pp. 073901:1–6, 2017, https://doi.org/10.1103/physrevlett.119.073901.Search in Google Scholar

[59] A.V. Kats, S. Savel’ev, V. A. Yampol’skii, and F. Nori, “Left-handed interfaces for electromagnetic surface waves,” Phys. Rev. Lett., vol. 98, pp. 073901:1–4, 2007, https://doi.org/10.1103/physrevlett.98.073901.Search in Google Scholar

[60] D. Leykam, K. Y. Bliokh, C. Huang, Y. D. Chong, and F. Nori, “Edge modes, degeneracies, and topological numbers in non-Hermitian systems,” Phys. Rev. Lett., vol. 118, p. 040401, 2017, https://doi.org/10.1103/physrevlett.118.040401.Search in Google Scholar

[61] K.Y. Bliokh, D. Leykam, M. Lein, and F. Nori, “Topological non-Hermitian origin of surface Maxwell waves,” Nat. Commun., vol. 10, no. 580, pp. 1–7, 2019, https://doi.org/10.1038/s41467-019-08397-6.Search in Google Scholar

[62] C. Triolo, A. Cacciola, S. Patanè, R. Saija, S. Savasta, and F. Nori, “Spin-momentum locking in the near field of metal nanoparticles,” ACS Photonics, vol. 4, pp. 2242–2249, 2017, https://doi.org/10.1021/acsphotonics.7b00436.Search in Google Scholar

[63] K. Y. Bliokh, F. J. Rodríguez-Fortuño, A. Y. Bekshaev, Y. S. Kivshar, and F. Nori, “Electric-current-induced unidirectional propagation of surface plasmon-polaritons,” Opt. Lett., vol. 43, pp. 963–966, 2018, https://doi.org/10.1364/ol.43.000963.Search in Google Scholar

[64] D. Lipkin, “Existence of a new conservation law in electromagnetic theory,” J. Math. Phys., vol. 5, pp. 696–700, 1964, https://doi.org/10.1063/1.1704165.Search in Google Scholar

[65] K. Y. Bliokh and F. Nori, “Characterizing optical chirality,” Phys. Rev. A, vol. 83, pp. 021803(R):1–3, 2011, https://doi.org/10.1103/physreva.83.021803.Search in Google Scholar

[66] M. Li, S. Yan, Y. Liang, P. Zhang, and B. Yao, “Transverse spinning of particles in highly focused vector vortex beams,” Phys. Rev. A, vol. 95, no. 5, pp. 053802:1–6, 2017, https://doi.org/10.1103/physreva.95.053802.Search in Google Scholar

[67] C. Min, Z. Shen, J. Shen, et al., “Focused plasmonic trapping of metallic particles,” Nat. Commun., vol. 4, no. 1, pp. 2891–2897, 2013, https://doi.org/10.1038/ncomms3891.Search in Google Scholar

[68] D. G. Grier, “A revolution in optical manipulation,” Nature, vol. 424, no. 6950, pp. 810–816, 2003, https://doi.org/10.1038/nature01935.Search in Google Scholar

[69] E. Narimanov, “Ghost resonance in anisotropic materials: negative refractive index and evanescent field enhancement in lossless media,” Adv. Photonics, vol. 1, no. 4, pp. 046003:1–11, 2019, https://doi.org/10.1117/1.ap.1.5.056003.Search in Google Scholar

[70] D. Cozzolino, E. Polino, M. Valeri, et al., “Air-core fiber distribution of hybrid vector vortex-polarization entangled states,” Adv. Photonics, vol. 1, no. 4, pp. 046005:1–9, 2019, https://doi.org/10.1117/1.ap.1.4.046005.Search in Google Scholar

[71] M. Antognozzi, C. R. Bermingham, R. L. Harniman, et al., “Direct measurements of the extraordinary optical momentum and transverse spin-dependent force using a nano-cantilever,” Nat. Phys., vol. 12, pp. 731–735, 2016, https://doi.org/10.1038/nphys3732.Search in Google Scholar

[72] A.Y. Bekshaev, K. Y. Bliokh, and F. Nori, “Transverse spin and momentum in two-wave interference,” Phys. Rev. X, vol. 5, pp. 011039:1–9, 2015, https://doi.org/10.1103/physrevx.5.011039.Search in Google Scholar

[73] M. V. Berry, “Optical currents,” J. Opt. Pure Appl. Opt., vol. 11, no. 9, pp. 094001:1–12, 2009, https://doi.org/10.1088/1464-4258/11/9/094001.Search in Google Scholar

[74] K. Y. Bliokh and F. Nori, “Spin and orbital angular momenta of acoustic beams,” Phys. Rev. B, vol. 99, pp. 174310:1–9, 2019, https://doi.org/10.1103/physrevb.99.174310.Search in Google Scholar

[75] K. Y. Bliokh and F. Nori, “Transverse spin and surface waves in acoustic metamaterials,” Phys. Rev. B, vol. 99, pp. 020301(R):1–6, 2019, https://doi.org/10.1103/physrevb.99.020301.Search in Google Scholar

[76] D. Leykam, K. Y. Bliokh, and F. Nori, “Edge modes in two-dimensional electromagnetic slab waveguides: analogs of acoustic plasmons,” Phys. Rev. B, vol. 102, pp. 045129:1–6, 2020, https://doi.org/10.1103/physrevb.102.045129.Search in Google Scholar

[77] I. D. Toftul, K. Y. Bliokh, M. I. Petrov, and F. Nori, “Acoustic radiation force and torque on small particles as measures of the canonical momentum and spin densities,” Phys. Rev. Lett., vol. 123, pp. 183901:1–6, 2019, https://doi.org/10.1103/physrevlett.123.183901.Search in Google Scholar

[78] K. Y. Bliokh and F. Nori, “Klein-Gordon representation of acoustic waves and topological origin of surface acoustic modes,” Phys. Rev. Lett., vol. 123, pp. 054301:1–6, 2019, https://doi.org/10.1103/physrevlett.123.054301.Search in Google Scholar

[79] L. Burns, K. Y. Bliokh, F. Nori, and J. Dressel, “Acoustic versus electromagnetic field theory: scalar, vector, spinor representations and the emergence of acoustic spin,” New J. Phys., vol. 22, pp. 053050:1–17, 2020, https://doi.org/10.1088/1367-2630/ab7f91.Search in Google Scholar


Supplementary material

The online version of this article offers supplementary material (https://doi.org/10.1515/nanoph-2020-0430).


Received: 2020-07-28
Accepted: 2020-08-31
Published Online: 2020-09-17

© 2020 Peng Shi et al., published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 20.4.2024 from https://www.degruyter.com/document/doi/10.1515/nanoph-2020-0430/html
Scroll to top button