This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

thanks: [email protected]

Transverse spin angular momentum of space-time surface plasmon polariton wave packet

Naoki Ichiji Graduate School of Pure and Applied Sciences, University of Tsukuba, 1-1-1 Tennodai, Tsukuba-shi, Ibaraki 305-8571, Japan    Daigo Oue Instituto Superior Técnico, University of Lisbon, 1049-001 Lisbon, Portugal Department of Physics, Kyoto University, Kyoto 606-8502, Japan The Blackett Laboratory, Imperial College London, London SW7 2AZ, United Kingdom    Murat Yessenov CREOL, The College of Optics & Photonics, University of Central Florida, Orlando, FL 32816, USA    Kenneth L. Schepler CREOL, The College of Optics & Photonics, University of Central Florida, Orlando, FL 32816, USA    Ayman F. Abouraddy CREOL, The College of Optics & Photonics, University of Central Florida, Orlando, FL 32816, USA    Atsushi Kubo Faculty of Pure and Applied Sciences, University of Tsukuba,1-1-1 Tennodai, Tsukuba-shi, Ibaraki 305-8571, Japan
Abstract

In addition to longitudinal spin angular momentum (SAM) along the axis of propagation of light, spatially structured electromagnetic fields such as evanescent waves and focused beams have recently been found to possess transverse SAM in the direction perpendicular to the axis of propagation. In particular, the SAM of SPPs with spatial structure has been extensively studied in the last decade after it became clear that evanescent fields with spatially structured energy flow generate three-dimensional spin texture. Here we present numerical calculations of the space-time surface plasmon polariton (ST-SPP) wave packet, a plasmonic bullet that propagates at an arbitrary group velocity while maintaining its spatial distribution. ST-SPP wave packets with complex spatial structure and energy flow density distribution determined by the group velocity are found to propagate with accompanying three-dimensional spin texture and finite topological charge density. Furthermore, the spatial distribution of the spin texture and topological charge density determined by the spatial structure of the SPP is controllable, and the deformation associated with propagation is negligible. ST-SPP wave packets, which can stably transport customizable three-dimensional spin textures and topological charge densities, can be excellent subjects of observation in studies of spinphotonics and optical topological materials.

preprint: APS/123-QED

I Introduction

Spin angular momentum (SAM) associated with the rotation of electromagnetic polarization [1, 2, 3] plays an important role in both classical and quantum optics [4, 5, 6, 7, 8, 9, 10, 11]. While the SAMs of light have been studied for nearly a century, most discussions until the 2000s focused on the longitudinal SAM (ll-SAM), which is the component parallel to the propagation axis. However, Aiello and Bliokh et al. reported that optical SAM also has a component perpendicular to the propagation axis, transverse SAM (tt-SAM) [12, 13, 14]. tt-SAM has recently received much attention as it is responsible for a long-standing problem of polarization-dependent beam shift (the Imbert-Fedorov shift) [15, 16, 17], has been given new rotational degrees of freedom in optical manipulation [18, 19, 20] and provides strong light-matter coupling, which cannot be achieved by ordinary polarizations [21]. tt-SAM is proportional to the rotation of the energy flow density of the electromagnetic field and arises when the Poynting vector of the electromagnetic field has a spatial gradient in the plane normal to the propagation direction of the optical field, such as in tightly focused beams [22, 23], guided light [24], interference beats between two plane weaves [25], and evanescent fields [26].

Surface plasmon polaritons (SPPs), which are surface electromagnetic waves confined at the metal-insulator interface, are accompanied by an evanescent field that decays in the perpendicular direction to the interface and thus exhibits tt-SAM in the in-plane direction (tt-SAM[27, 28, 29, 30, 31]. The direction of the tt-SAM of the SPP is fixed according to the energy propagation direction, resulting in a broken time-reversal symmetry. Hence, as a counterpart of a quantum spin Hall effect of light [32, 33], the tt-SAM of SPPs has garnered widespread interest in the field of spinphotonics [34, 35, 36, 37, 38].

Furthermore, when the SPP has a two-dimensional spatial structure at the confined interface, its tt-SAM is not limited to the in-plane component but also includes an out-of-plane component (tt-SAM)  [39, 40, 41, 42, 43, 44, 45, 46, 47]. The existence of both tt-SAM and tt-SAM components of SPP indicates that the SPP can have a three-dimensional arrangement of spin textures and thus a finite surface area on the Poincaré polarization vector sphere [48, 49, 50, 51]. Efforts to generate topologically stable three-dimensional spin vector fields by preparing a vortex-like energy flow have been actively pursued in the context of topological quasiparticles [39, 40]. The introduction of integer topological charges into SPPs using the designed light-SPP coupler or vortex beam has enabled the demonstration of various plasmonic topological quasiparticles, including plasmonic spin skyrmion [39, 42, 52, 44], topological melon structure [40, 53], and topological plasmonic field [54, 55] and spin [41, 43] lattice structures. However, the spin textures of SPPs discussed in previous studies are typically excited at specific coordinates and do not involve spatial movement, such as propagation. While numerous studies have examined and experimentally realized the motion control of magnetic skyrmions [56, 57, 58, 59, 60], only a few have considered the spatial movement of the spin texture of SPPs [61, 62].

Shi et al. recently formulated the tt-SAM of the structured electromagnetic guided waves and experimentally demonstrated the tt-SAM of SPP waves with inhomogeneous energy flow density, such as Cosine SPP and Airy SPP waves [46]. Their formulation of the SAM suggests that all SPP waves with both spatial structure and energy flow should be accompanied by tt-SAM, even if these SPP fields are topologically uncharged. Therefore, temporal variation of the SPP fields with spatial structure, such as SPP wave packets with short temporal widths [63], steering SPP beams excited by achromatic incident waves [64], and plasmonic bullets with confinement in all dimensions [65], are expected to be accompanied by tt-SAM in their propagation. However, although several observations of tt-SAM associated with SPPs have been reported [47], the migration of tt-SAM caused by propagating SPP fields has rarely been investigated.

Here, we present theoretical calculations of the tt-SAM for the space-time SPP (ST-SPP) wave packet, a newly proposed type of SPP wave packet classified as a plasmonic bullet [66]. ST-SPP wave packets are structured SPPs with artificially designed propagation characteristics achieved through precise adjustments to the spatial and temporal frequencies of each frequency component of the SPP wave packet. The concept of designing propagation characteristics by adjusting the time-space-frequency correlation of a wave packet was initially proposed and experimentally demonstrated for light pulses and referred to as space-time (ST) waves [67, 68, 69]. ST-SPP wave packets have been theoretically shown to possess attractive properties such as diffraction-free and dispersion-free propagation invariance and arbitrary group velocities. Efforts to experimentally excite and observe them are ongoing [70, 71].

In this study, we conducted an examination of the three-dimensional spin arrangement of ST-SPP wave packets through calculations of each electromagnetic component and energy flow density. Our calculations revealed that ST-SPP wave packets possess a three-dimensional spin texture and a finite topological charge density. The spin texture and topological charge density of ST-SPPs with propagation invariance were found to be spatially robust and maintain a stable spatial distribution during propagation. Interestingly, the topological charge density of ST-SPP wave packets exhibited spatial distributions with only opposite signs on the left and right, with positive and negative positional relationships reversing depending on the group velocity. The proposed calculation results open up the possibility of constructing plasmonic topological quasiparticles and manipulating excited structured SPP waves, which may have potential applications in optical information transfer and optical trapping.

II ST-SPP wave packet

Refer to caption
Figure 1: (a) The spectral representation of conventional plane SPP wave on the surface of the SPP light-cone kx2+ky2=ksp2k_{x}^{2}+k_{y}^{2}=k_{\mathrm{sp}}^{2}. (b) Spatial distribution of the field intensity of conventional plane SPP wave in physical space, I(x,y;t=0)I(x,y;t\!=\!0). (c,d) Same as (a,b) but for a Cosine SPP wave. (e,f) Same as (a,b) but for an ST-SPP wave. The black lines in the side and bottom panels in (b), (d) and (f) are cross sections through the field distributions in the main panels at x=0x\!=\!0 and y=0y\!=\!0. Each profile in (b), (d), and (f) is normalized by each maximum value.

To describe the dispersion relation of a wave packet with a two-dimensional spatial distribution, such as an ST-SPP wave packet, a three-dimensional spatiotemporal spectral diagram consisting of one frequency axis and two wavenumber axes is useful. While the light cone of the light in a vacuum is a cone kl=ω/ck_{\mathrm{l}}\!=\!\omega/c, the three-dimensional spatiotemporal correlation of SPP, the SPP light cone, has a curved surface that reflects the dispersion of SPP (Fig. 1[66, 72],

ksp=ωcϵm(ω)ϵdϵm(ω)+ϵd,\displaystyle k_{\mathrm{sp}}\!=\!\frac{\omega}{c}\sqrt{\frac{\epsilon_{\mathrm{m}}(\omega)\epsilon_{\mathrm{d}}}{\epsilon_{\mathrm{m}}(\omega)+\epsilon_{\mathrm{d}}}}, (1)

where ϵd(m)\epsilon_{\mathrm{d}(\mathrm{m})} is the relative permittivity of the dielectric (metal).

In the case of the conventional pulsed SPP wave with uniform intensity in the transverse direction, the spectral support domain of the SPP wave packet corresponds to the intersection of the SPP-light cone and kx=0k_{x}\!=\!0 plane (Fig. 1(a)). We choose the yy-direction as the axial propagation direction and define the zz-direction as the surface normal of the interface between metal and dielectric so that the SPP is localized on the xyxy-plane. The out-of-plane component of the electric field distribution of the conventional SPP plane wave packet with finite width along the propagation direction (Fig. 1(b)) is given by dωE~(ω)ei(kyyωt)\int\!\differential{\omega}\;\widetilde{E}(\omega)e^{i(k_{y}y-\omega t)}, where ω\omega is the temporal angular frequency and E~(ω)\widetilde{E}(\omega) is the pulse spectrum, the Fourier transform of E(t)E(t) at y=0y\!=\!0. The free-space pulse used to excite the SPP has the form

E(t)=Eoeiωotexp(t2(ΔT/(2ln(2)))2),\displaystyle E(t)\!=\!E_{\mathrm{o}}e^{-i\omega_{\mathrm{o}}t}\exp\left(-\frac{t^{2}}{(\Delta T/(2\ln{2}))^{2}}\right), (2)

where ΔT=100\Delta T\!=\!100 fs, ωo/(2π)=375\omega_{\mathrm{o}}/(2\pi)\!=\!375 THz (800 nm wavelength), and EoE_{\mathrm{o}} is the amplitude of the electric field. The dielectric is assumed to be vacuum (ϵd=1)(\epsilon_{\mathrm{d}}\!=\!1) and the metal is gold (Au). We employ the Lorentz-Drude model [73] for ϵm(ω)\epsilon_{\mathrm{m}}(\omega). Experimentally, the conventional plane SPP wave packet corresponds to a wave packet excited by a linear light-SPP coupling structure with sufficiently long length, such as a slit or gratings [74, 75]. The propagation of a conventional SPP wave on the dispersive sample surface is inevitably accompanied by deformation and chirping due to the frequency dependence of the group velocity [63].

As a comparison, the spectral support domain of the Cosine SPP wave [76] is shown in Fig. 1(c). The Cosine SPP waves are a spatial pattern generated by the interference of two SPPs, corresponding to the two-wave interference in free space [25]. On the SPP light cone surface, the spectral support domain of the Cosine SPP wave is defined by two points that have identical kyk_{y} components and two kxk_{x} components equal in magnitude but with opposite signs. The spatial field distribution EzcosE_{z}^{\mathrm{cos}} is given by:

Ezcos(𝐱|ω)\displaystyle E_{z}^{\mathrm{cos}}(\mathbf{x}\>|\>\omega) =12E~(ω)(ei(kxx+kyyωt)+ei(kxx+kyyωt))\displaystyle=\frac{1}{2}\widetilde{E}(\omega)(e^{i(k_{x}x+k_{y}y-\omega t)}+e^{i(-k_{x}x+k_{y}y-\omega t)}) (3)
=E~(ω)cos(kxx)e+i(kyyωt)\displaystyle=\widetilde{E}(\omega)\cos(k_{x}x)e^{+i(k_{y}y-\omega t)} (4)

where we have defined 𝐱=(x,y;t)\mathbf{x}=(x,y;t). In this study, EzE_{z} is calculated for the insulator side (z>0z>0). Note that kxk_{x} and kyk_{y} are the real components of the wave vector along the transverse (xx) and axial (yy) coordinates, respectively, which satisfy the relation kx2+ky2=ksp2k_{x}^{2}+k_{y}^{2}\!=\!k_{\mathrm{sp}}^{2}. The interference between the two plane waves generates a periodic cosine-shaped field distribution in the transverse direction, as its intensity profile is shown in Fig. 1(d).

Importantly, the axial wavenumber of the Cosine SPP wave, ky=ksp2kx2k_{y}\!=\!\sqrt{k^{2}_{\mathrm{sp}}-k_{x}^{2}} can be arbitrarily set in the range 0<ky<ksp0\!<\!k_{y}\!<\!k_{\mathrm{sp}} by properly selecting kxk_{x}. Therefore, by making all frequency components of the SPP wave packet into Cosine SPP waves with designed kyk_{y}, structured SPP wave packets with arbitrary dispersion on the propagation axis can be constructed. The concept of designing the dispersion relation on the propagation axis by strictly adjusting kxk_{x} and kyk_{y} has already been experimentally demonstrated for light pulses in free space and is known as a Space-Time (ST) wave packet [67, 68, 69]. ST wave packets with a variety of novel propagation characteristics have been reported, including diffraction-free property [67], arbitrary group velocity [77], acceleration and deceleration in unprecedented ranges [78, 79], the introduction of dispersion properties into a light pulse in free space [80, 81], and non-dispersive propagation in a dispersive media [82].

The spectral support domain of an ST-SPP wave packet, which is the target of this study, is defined as a one-dimensional trajectory at the intersection of the SPP light cone, Eq. (1), with a spectral plane Ω=(kyko)ctanθ\Omega\!=\!(k_{y}-k^{\prime}_{o})c\;\mathrm{tan}\,\theta, that is parallel to the kxk_{x}-axis and makes an angle θ\theta with the kyk_{y}-axis (Fig. 1(e)), where kok^{\prime}_{o} is the SPP wave number evaluated at the carrier frequency [i.e., ko=ksp(ω=ωo)k^{\prime}_{o}=k_{\mathrm{sp}}(\omega=\omega_{o})], and Ω=ωωo\Omega\!=\!\omega-\omega_{o} is the frequency measured with respect to ωo\omega_{o}.

The projection of the spectral support domain onto the (ky,ωc)(k_{y},\frac{\omega}{c})-plane is a straight line, and thus the group velocity vg~=dω/dky=ctanθ\widetilde{v_{g}}\!=\!\differential{\omega}/dk_{y}\!=\!c\>\mathrm{tan}\,\theta is a constant independent of ω\omega. Each component of the spatial field distribution of the ST-SPP wave packet on the z=0+z\!=\!0^{+} plane, ExSTE^{\mathrm{ST}}_{x}, EySTE^{\mathrm{ST}}_{y} and EzSTE^{\mathrm{ST}}_{z} is given by [46](See also appendix for the details):

ExST(𝐱)\displaystyle E_{x}^{\mathrm{ST}}(\mathbf{x}) =idωkzksp2xEzcos(𝐱|ω)\displaystyle=i\!\int\!\differential{\omega}\frac{k_{z}}{k_{\mathrm{sp}}^{2}}\partialderivative{x}E_{z}^{\mathrm{cos}}(\mathbf{x}\>|\>\omega) (5)
EyST(𝐱)\displaystyle E_{y}^{\mathrm{ST}}(\mathbf{x}) =idωkzksp2yEzcos(𝐱|ω)\displaystyle=i\!\int\!\differential{\omega}\frac{k_{z}}{k_{\mathrm{sp}}^{2}}\partialderivative{y}E_{z}^{\mathrm{cos}}(\mathbf{x}\>|\>\omega) (6)
EzST(𝐱)\displaystyle E_{z}^{\mathrm{ST}}(\mathbf{x}) =dωEzcos(𝐱|ω).\displaystyle=\!\int\!\differential{\omega}E_{z}^{\mathrm{cos}}(\mathbf{x}\>|\>\omega). (7)

where kz=kl2ϵmksp2k_{z}\!=\!\sqrt{k_{\mathrm{l}}^{2}\epsilon_{\mathrm{m}}-k_{\mathrm{\mathrm{sp}}}^{2}} is the wavenumber in the surface normal that determines the exponential decay in the zz-direction (𝐄(x,y,z;t)=𝐄(𝐱)eikzz(\mathbf{E}(x,y,z;t)=\mathbf{E}(\mathbf{x})e^{ik_{z}z}). Remind that kzk_{z} and kspk_{\mathrm{\mathrm{sp}}} are functions of ω\omega.

While the kxk_{x} of the conventional SPP wave packet is always 0, the kxk_{x} of the ST-SPP wave packet varies with frequency to keep dωdky=ctanθ\derivative*{\omega}{k_{y}}=c\tan\theta. As a result, ST-SPP wave packets possess a finite transverse bandwidth Δkx\Delta k_{x} and thus exhibit spatial confinement in both the axial and transverse directions. Previous experiments on ST wave packets in free space and simulation results of the ST-SPP wave packets have shown that the spatial distributions of the ST wave packets are X-shaped. [see Fig. 1(f)] [69, 83, 71]. The ST-SPP wave packet with a characteristic X-shaped spatial distribution propagates toward the yy-direction with the designed group velocity vg~=ctanθ\widetilde{v_{g}}\!=\!c\,\mathrm{tan}\,\theta, maintaining the identical spatial distribution unless ohmic losses are considered [66].

III Energy flow density

For an SPP wave with an evanescent field, the SAM can be calculated as follows,

𝑺=14ωIm(ϵ𝑬×𝑬+μ𝑯×𝑯).\bm{S}=\frac{1}{4\omega}\mathrm{Im}(\epsilon\bm{E^{*}}\times\bm{E}+\mu\bm{H^{*}}\times\bm{H}). (8)

The electric field 𝑬\bm{E} is given by Eq. (5-7), and the magnetic field 𝑯\bm{H} is calculated from EzE_{z} by employing the Maxwell equations [46](See appendix). The asterisks signify their complex conjugates. We can confirm the SAM is a divergenceless vector field, 𝑺=0\nabla\cdot\bm{S}=0. This implies that the SAM is the rotation of another vector field. Indeed, taking the rotation of the canonical decomposition of momentum density, ×ϵμ(𝑷𝑷o)=××𝑺/2\nabla\times\epsilon\mu(\bm{P}-\bm{P}_{o})=\nabla\times\nabla\times\bm{S}/2, where 𝑷=Re(𝑬×𝑯)/2\bm{P}=\operatorname{Re}(\bm{E}^{*}\times\bm{H})/2 is the total energy flow density, and 𝑷o=Im[μ1𝑬()𝑬+ϵ1𝑯()𝑯]/(4ω)\bm{P}_{o}=\operatorname{Im}[\mu^{-1}\bm{E}^{*}\cdot(\nabla)\bm{E}+\epsilon^{-1}\bm{H}^{*}\cdot(\nabla)\bm{H}]/(4\omega) is the orbital contribution, we can find the following relationship [84, 85, 86, 87, 46]:

𝑺=2ω2×(𝑷𝑷o)=12ω2×𝑷.\bm{S}=\frac{2}{\omega^{2}}\nabla\times(\bm{P}-\bm{P}_{o})=\frac{1}{2\omega^{2}}\nabla\times\bm{P}. (9)

From Eq. (9), it can be confirmed that the SAM of the SPP is represented by a rotation of the energy flow density. In Fig. 2, we plot the in-plane components of the energy flow density in the ST-SPP wave packets calculated at different group velocities. In this study, the group velocity of the conventional SPP wave packet at the center frequency, vsp(ω/(2π)=375THz)=0.91cv_{\mathrm{sp}}(\omega/(2\pi)\!=\!375~{}\mathrm{THz})\!=\!0.91c, is defined as the boundary velocity for superluminal (vg~>vsp\widetilde{v_{g}}\!>\!v_{\mathrm{sp}}) and subluminal (vg~<vsp\widetilde{v_{g}}\!<\!v_{\mathrm{sp}}) (cc: speed of light in vacuum). The tilt angle of the spectral plane θ\theta is determined by the designed group velocity, θ=tan1(vg~/c)\theta\!=\!\mathrm{tan}^{-1}(\widetilde{v_{g}}/c). In this paper, we set ωo/(2π)=360THz(390THz)\omega_{o}/(2\pi)=360\ \mathrm{THz}\ (390\ \mathrm{THz}) for the superluminal (subliminal) ST-SPP wave packet. The spectral projections onto the (kyk_{y},ω\omega)- and (kxk_{x},ω\omega)-planes for an ST-SPP with vg~=c\widetilde{v_{g}}=c, i.e., the superluminal ST-SPP wave packet, are plotted in Fig. 2(c).

The energy flow density in the axial direction PyP_{y} indicates the direction in which the ST-SPP wave packet propagates. The spatial distribution of PyP_{y} shown in Fig. 2(a) corresponds to the spatial distribution of the field intensity profile shown in Fig. 1(f). In contrast to PyP_{y}, which is positively valued everywhere, the spatial distribution of PxP_{x} (i.e., the energy flow density in the direction transverse to the propagation axis) is composed of two diagonal branches with opposite signs [see Fig. 2(b)]. These two branches with opposite signs are constructed by the +kx+k_{x} and kx-k_{x} components of the ST-SPP, indicated by red and blue lines in Fig. 2(c), respectively. As a result of the destructive interference between the two branches, PxP_{x} vanishes on the xx- and yy-axes (at t=0t\!=\!0.). Figs. 2(d-f) show the Poynting vector and the spectral projections for the subluminal ST-SPP wave packet, vg~=0.8c\widetilde{v_{g}}=0.8c. In the subluminal and superluminal ST-SPP wave packets, the energy flow in the propagation direction PyP_{y} exhibits a similar spatial distribution. In contrast, the energy flow in the transverse direction PxP_{x} is reversed (Fig. 2(e)).

Refer to caption
Figure 2: The spatial distributions of the energy flow density for ST-SPP waves at t=0t=0. (a) PyP_{y} and (b) PxP_{x} components of the superluminal ST-SPP wave packet with vg~==1.0c\widetilde{v_{g}}=\!=\!1.0c. (c) The spectral projections onto the (kyk_{y},ω\omega) and (kxk_{x},ω\omega)-plane. (d-f) Same as (a-c) but for the subluminal ST-SPP wave packets with vg~==0.8c\widetilde{v_{g}}=\!=\!0.8c. (g) spatial distributions of the energy flow density of one of the branches comprising the ST-SPP wave packet calculated from the +kx+k_{x} components only. All of each display region is identical to (b,e). The intensities obtained at the x=y=0x\!=\!y\!=\!0 are plotted in the right panel as a function of group velocity. The maximum value of PyP_{y} at vg~=vsp\widetilde{v_{g}}\!=\!v_{\mathrm{sp}} was normalized to 1.

In Fig. 2(g), the calculation for one branch consisting only of the +kxk_{x} components [red lines in Fig. 2(c,f)] is presented, which is useful to understand the spatial distribution of the ST-SPP wave packet and the energy flow. The typical spatial distributions of PyP_{y} and PxP_{x} of +kx+k_{x} branches are plotted in the left panels. While PxP_{x} and PyP_{y} are positive regardless of the group velocity, the spatial distribution of the branches showed a marked dependence on the group velocity. In the superluminal regime, the angle between the branches and the propagation axis is positive, α>0\alpha\!>\!0, while, in the subluminal regime, α<0\alpha\!<\!0. This implies the counterintuitive fact that the spatial angle of the branch is determined independently of the direction of the energy flow (See appendix). In addition, the intensity distributions also show that PyP_{y} and PxP_{x} have different dependencies on the group velocity. PyP_{y} and PxP_{x} at (x,y)=(0,0)(x,y)\!=\!(0,0) for various group velocities ranging from 0.7 cc to 1.1 cc are plotted in the right panel. PyP_{y} reaches the maximum value when vg~=vsp\widetilde{v_{g}}\!=\!v_{\mathrm{sp}} and decreases as the vg~\widetilde{v_{g}} deviates from vspv_{\mathrm{sp}}, and vice versa for PxP_{x}. Since the ratio of kxk_{x} and kyk_{y} represents the tilt angle of each frequency component, an increase in kxk_{x} tends to result in an increase in PxP_{x} in exchange for a decrease in PyP_{y}. It is also noteworthy that the variation of PyP_{y} is less than 1 %\% while that of PxP_{x} exceeds 50%50\% in this calculation range. This correlation between PyP_{y} and PxP_{x} suggests that the controllable group velocity of the ST-SPP wave packet can be interpreted as the projection of the velocity of the diagonal branch onto the propagation axis, which is propagating in a slightly oblique direction from the propagation axis. Note that the velocity projected on the propagation axis corresponds to the velocity of movement of the intersection point of the two branches for the ST-SPP wave packet consisting of ±kx\pm k_{x} components.

IV Spin angular momentum

Refer to caption
Figure 3: (a-c) Spatial field distribution of the spin texture, (a) SxS_{x}, (b) SyS_{y}, and (c) SzS_{z} component. (d) Cross-sections along the dashed lines in (a,c) and energy flow density in the propagation direction, PyP_{y}, obtained at corresponding positions. (e) The L-line singularity map of the in-plane polarization of SPP fields. (f) The inverse of the magnitude of the electric field distribution. The black lines represent the L-line calculated by taking the contour lines of (d) with a value of 100. (g) spatial distributions of the tt-SAM and L-lines before and after propagation calculated at t=0t\!=\!0 and t=334t\!=\!334 fs. The cross sections along the lateral direction at x=0μx\!=\!0~{}\mathrm{\mu}m and 100μ100~{}\mathrm{\mu}m are plotted in (h).

Figures 3(a-c) show each component of the SAM of the superluminal ST-SPP wave packet, calculated from Eq. (8) at vg~=c\widetilde{v_{g}}=c. As the main component of tt-SAM (SxS_{x} and SyS_{y}) arises from the exponential decay of the in-plane energy flow density along the surface normal, we have Sx[1/(2ω2)]Py/zS_{x}\approx-[1/(2\omega^{2})]\partial P_{y}/\partial z and Sy[1/(2ω2)]Px/zS_{y}\approx[1/(2\omega^{2})]\partial P_{x}/\partial z and can confirm these are consistent with the spatial distribution of PyP_{y} and PxP_{x} [see Figs. 3(a,b)].

On the other hand, the tt-SAM (SzS_{z}) arises from the rotation of the in-plane energy flow of the SPP. As the yy-component of the energy flow is approximately 10 times larger than the xx-component, PyPxP_{y}\gg P_{x}, in this study [Fig. 2(g)], it dominantly contributes to the tt-SAM, Sz[1/(2ω2)]Py/xS_{z}\approx[1/(2\omega^{2})]\partial P_{y}/\partial{x}. Figure 3(d) shows the cross-sections of the xx-and zz-components of the SAM along the xx-axis (y=0y\!=\!0) [dashed lines in Figs. 3(a,c)]. Note that the yy-component of the energy flow (dashed line in Figs. 2(a)) is also plotted in the same figure. The spatial distribution of the phase-matched SxS_{x} and π/2\pi/2-shifted SzS_{z} with respect to the PyP_{y} component is consistent with the SAM of the Cosine SPP wave reported previously [46].

To characterize the three-dimensional spin texture of the SAM, we evaluate where the ellipticity of the local polarization of the SAM is minimal in the xyxy plane. This region corresponds to the equator of the Poincaré sphere and can be interpreted as the L-line region where the field is linearly polarized [88, 89]. In Fig. 3(e), we plotted the L-line singularity map |𝑺|/Sz|\bm{S}|/S_{z}. Note that the L-lines shown in Fig. 3(e) are not a one-dimensional trajectory but have a finite width because the L-line map saturates in the region where SzS_{z} is smaller than 1%1\%.

Because the tt-SAM of the ST-SPP wave packet is generated from the cosine-like spatial structure resulting from the interference between the +kx+k_{x} and kx-k_{x} branches, there are almost no SzS_{z} components in regions where the two branches do not spatially overlap. Therefore, the X-shaped region without fringes of the ST-SPP wave packet exhibits linear polarization, which is the polarization state of the conventional SPP wave packet. In contrast, multiple split ring-shaped L-lines, which are derived from the spatial structure of the ST-SPP wave packet, are aligned along the xx-axis (y=0y\!=\!0) with their split points oriented in the opposite direction around the yy-axis. The split ring structures define perimeters where the tt-SAM of the ST-SPP wave packet is distributed on the plane, and the split points on the xx axis are singularities where the electric field amplitude is zero. Around these discontinuous points, there is a steep gradient in the electric field intensity. In Fig. 3(f), we plotted the L-line obtained from Fig. 3(e) overlaid on the inverse of the electric field distribution, 1/|𝑬|1/\bm{|E|}.

Although the calculations up to this point only considered the case t=0t\!=\!0 and propagation loss was not taken into account, the actual propagation of an ST-SPP wave packet on the dispersive metal surface has a frequency-dependent attenuation determined by the imaginary part of the dielectric function. Attenuating the overall intensity is inevitable, and there will be a slight deformation where the ST-SPP wave packet has an ultrabroadband spectral width [66, 70]. To estimate the spin texture stability of the ST-SPP wave packet, calculations of the ST-SPP wave packet before and after propagation were performed with an attenuation model that introduced propagation loss into Eq. (5-7). The propagation loss at each frequency was calculated from Im[ϵm(ω)]\mathrm{Im}[\epsilon_{\mathrm{m}}(\omega)]. For each frequency component with a finite transverse wavenumber, which propagates in a slightly oblique direction, we define the distance of propagation y/cos(ϕ)y/\mathrm{cos}(\phi), where ϕ\phi is the angle between the propagation direction and the yy-axis, ϕ=tan1(kx/ky)\phi\!=\!\mathrm{tan}^{-1}(k_{x}/k_{y}). In Fig. 3(g), we plot the spatial distributions of the tt-SAM calculated by attenuation model at t=0t\!=\!0 and t=334fs)t\!=\!334\,\mathrm{fs}). The L-lines calculated at each time were also overlaid on the plots.

While the SzS_{z} distribution after propagation over 100 μ\mathrm{\mu}m shows an overall attenuation in intensity, the spatial distribution surrounded by L-line is almost unchanged from that before propagation. Moreover, the cross-sections along the dashed lines in Fig. 3(g) almost overlap [see Fig. 3(h)]. This indicates the deformation of the spin texture with propagation is marginal. From these calculations, we conclude that the ST-SPP wave packet can transport the spin texture in a spatially stable manner on the order of the propagation length of the SPP.

V Toplogical charge density

Refer to caption
Figure 4: (a,b) Spatial distributions of the topological charge density DD at t=0t=0 for (a) v~=c\tilde{v}=c, (b) v~=0.8c\tilde{v}=0.8c. The black lines represent the L-line calculated by |𝑺|/Sz|\bm{S}|/S_{z}. The regions around the fourth split point on the left and right from the center are defined as LL and RR, respectively. (c,d) Spatial distributions of each normalized SAM component, SxS^{\prime}_{x}, SyS^{\prime}_{y}, and SzS^{\prime}_{z} for (c) v~=c\tilde{v}=c and (d) v~=0.8c\tilde{v}=0.8c. The expanded views of the dashed squares are plotted in the top panels. (e,f) The expanded view of the right dashed square in Figs. 4 (c,d) of the spatial distribution of each component of the normalized SAM. The vector plots show the in-plane components, and the color lines show the contour of the out-of-plane components (red: -0.9 to -0.6, blue: 0.6 to 0.9). The color map represents the distribution of DD.

We next discuss the topological properties of ST-SPP based on the SAM textures obtained above. We calculate the topological charge density as in studies of topological spin quasiparticles, where the integration of the topological charge density over an area within the boundary of the quasiparticle provides the topological charge (skyrmion number) [40, 49, 45, 42, 41, 39, 44, 43, 53]. The topological charge density is defined by

D=14π𝑺(𝑺x×𝑺y),D=\frac{1}{4\pi}\bm{S^{\prime}}\cdot\left(\frac{\partial\bm{S^{\prime}}}{\partial x}\times\frac{\partial\bm{S^{\prime}}}{\partial y}\right), (10)

where we have normalised the SAM, 𝑺=𝑺/|𝑺|\bm{S^{\prime}}=\bm{S}/|\bm{S}|.

Figure 4(a) shows the calculated DD map for an ST-SPP wave packet with vg~=c\widetilde{v_{g}}=c. The positions of the discretized spatial distribution of DD correspond to the split points of the L-lines. One of the major characteristics of the DD map is its reversed signs on the left and right sides: the spatial distribution exhibits perfect line symmetry around the xx-axis, with positive values for positive xx and negative values for negative xx. In addition, the DD map for v~=0.8c\tilde{v}=0.8c plotted in Fig. 4(b) shows a similar symmetric spatial distribution as in the case of v~=c\tilde{v}=c, but with opposite signs on the left and right. The left-right reversal and the group velocity-dependent reversal of the sign of DD are explained by the spatial distribution of SyS_{y}. In Figs. 4(c,d), we plotted each component of 𝑺\bm{S^{\prime}} (SxS^{\prime}_{x},SyS^{\prime}_{y}, and SzS^{\prime}_{z}) for the superluminal and subluminal ST-SPP waves shown in Figs. 4(a,b), respectively. As shown in Fig. 3, the tt-SAM arising from the decay of evanescent fields has spatial distributions consistent with the in-plane component of 𝑷\bm{P}, PxP_{x} and PyP_{y}. While SxS^{\prime}_{x} is positive throughout the calculation region regardless of the group velocity, SyS^{\prime}_{y} has a different sign in each quadrant. The sign is switched depending on whether the group velocity is subluminal or superluminal. On the other hand, the spatial distribution of tt-SAM is mainly determined by the partial differentiation of PyP_{y} in the xx-direction; it, therefore, does not exhibit quadrant-dependence of the sign unlike SyS^{\prime}_{y}.

To examine the differences in the spatial distribution of each component in detail, the enlarged view around the split points is plotted in the top panels, one on the left (region LL) and one on the right (region RR). In contrast to the consistent spatial distribution of SxS^{\prime}_{x} and SzS^{\prime}_{z}, only the SyS^{\prime}_{y} component shows that the switched signs for right and left, and for superluminal and subluminal. From Eq. 10, we can confirm that the sign of DD is switched when the two vector components of 𝑺\bm{S}^{\prime} are fixed, and only the sign on one component is switched. This is consistent with the calculation results shown in Figs. 3(a-d). It can also be seen that the in-plane and out-of-plane components of 𝑺\bm{S}^{\prime} have complementary spatial distributions at the vicinity of the split points. Both SxS^{\prime}_{x} and SyS^{\prime}_{y} are zero at y=0y\!=\!0 and have a spatial distribution that extends toward the yy-direction from the split point of the L-line. In contrast, SzS^{\prime}_{z} shows the maximum value at y=0y\!=\!0 and has a distribution that extends toward the xx-direction. The vector plots of the in-plane components of 𝑺\bm{S^{\prime}} and the contour plots of the out-of-plane component superimposed on the DD map plotted in Figs. 4(e,f) show more clearly the positional relationship of each 𝑺\bm{S}^{\prime} component. The plotted areas correspond to the region RR. The field distribution of the SxS_{x} in Figs. 4(c,d) and the direction of the in-plane components of the SAM in Figs. 4(e,f) indicate that SxS_{x} is dominant in most of the calculation region. Only in a limited region around the split points of the L-line, each component of SAM has comparable magnitudes, and DD has a finite value.

VI Discussion and Conclusion

The propagation invariance of an ST-SPP preserves the wave packet shape and the L-line structure, which forms the stable frame of the SAM texture. As discussed above, the propagation of the ST-SPP is accompanied by the propagation of topological charge densities with different signs on its left and right sides. In the case of a plasmonic spin skyrmion or meron, a phase singularity (C-point) exists in the core of the vortex where an in-plane rotating electric field (local circular polarization) is formed [39, 42, 40, 44, 49]. It corresponds to the north or the south pole of the Poincaré sphere; the local polarization is circularly polarized, and the SAM vector points in the surface-normal direction. The field exhibits a 2π2\pi-phase rotation around the center with the angular momentum provided by the excitation light field or the geometrical charge of a coupling structure. The ST-SPP, in contrast, does not have the plasmonic vortex like the plasmonic spin skyrmion; the ST-SPP has an X-shaped electric field distribution, with each branch consisting of the +kx+k_{x} and kx-k_{x} branch components of the spectral support domain (Fig. 1(f),  2(a-f)). In the central part of the X-shape, the electric field forms a textured structure and oscillates at the central frequency of the wave packet (See appendix). In the areas on either side, ”holes” in the electric field strength are lined up due to the interference between the +kx+k_{x} and kx-k_{x} branches (Fig. 3(f)). These holes are not accompanied by a 2π2\pi-phase rotation of the field and are therefore distinct from the central core of the plasmon skyrmions mentioned above. Because the ST-SPP propagates at the group velocity vspv_{\mathrm{sp}}, the energy flow density is inherently dominated by the positive-valued PyP_{y} component. The lack of negative PyP_{y} precludes generations of vortices around phase singularities. However, the periodic modulation of PyP_{y} intensity in the transverse direction is responsible for the surface-normal SAM (Fig. 3(d)). The overall SAM vector distribution of ST-SPP is concentrated in the positive SxS_{x}, while it changes steeply only in the vicinity of the holes of the field. The SAM vector is not distributed over the hemisphere or the entire Poincaré sphere, therefore, the topological charge is neither half-integer nor integer. However, the different branches consisting of the +kx+k_{x} and kx-k_{x} components surrounding the hole generate SyS_{y} components of different signs, resulting in the winding of SAM and the related nonzero topological charge density DD (Fig. 4(e,f)). This nonzero DD is specific to ultrashort ST-SPP wave packets, not to single frequency SPP fields, because it originates from the spatial proximity of the +kx+k_{x} and kx-k_{x} branches, and thus should be associated with the geometry of the spectral support domain that generates the ST-SPP.

Since the SPP also has an evanescent field associated with electric polarization on the metal side, it would be useful to describe the SAM interaction with metal electrons and the relation to topological magnetic quasiparticles. Firstly, the SAM in the metal comprises the circular motion of electrons or localized circular currents. As a result, magnetization due to the inverse Faraday effect is expected to occur within the metal [30, 90, 91, 92, 93]. The magnetic field can interact with the electron spin. Indeed, conduction electron spin currents can be driven by the SPPs at a single surface [94, 95, 96], a metallic film [97, 98], and in graphene [99, 100]. Secondly, the SAM vector in the Poincaré sphere is regarded as an optical correspondence of the magnetization vector in magnetic skyrmion systems. The topological charge density is a physical quantity that reflects the degree of twist in the local spin (magnetization) vector. In the context of magnetic topological spin textures, DD can be related to as an emergent magnetic field bemb_{\mathrm{em}} [56, 101, 102]. In two-dimension systems, the magnitude of the field vector directing the surface normal is proportional to DD. Importantly, bemb_{\mathrm{em}} is not merely a fictitious field but can also interact with the spins of other electrons in solid materials. The interaction of bemb_{\mathrm{em}} with electron spins via the magnetic moments described by the Landau-Lifshitz-Gilbert (LLG) equation has been experimentally confirmed [103, 104, 105, 59, 101]. This interaction, known as the topological Hall effect [106, 107, 108], has realized the manipulation of magnetic skyrmions by external currents. It is noteworthy that the ”propagating topological charge density” was found in the context of SPPs, which belong to the class of optical topological systems. Furthermore, the correspondence between the magnetic topological systems and optical spin topological systems demonstrated in various previous studies suggest that the topological charge density calculated in this study can be considered as an optical emergent magnetic field that can interact with the spins of electrons in materials [109]. In addition, the propagation invariance and arbitrary group velocities of the ST-SPP wave packet enable adaptive transport of topological charge density with a stable spatial distribution across subluminal and superluminal regimes. Through the experimental excitation of ST-SPP wave packets using a spatial light modulator (SLM), the intensity, spatial distribution, and polarity of DD can be adjusted externally at the driving frequency of the SLM. This feature enables the exploration of novel opportunities to realize spin information transmission.

In conclusion, we have conducted numerical simulations to examine the spin texture in an ST-SPP wave packet, which is a type of plasmonic bullet that possesses propagation invariance at arbitrary group velocities. Our calculations reveal three-dimensional spin textures in these wave packets, which are found to have stable spatial distributions during propagation over distances on the order of the SPP propagation length. Furthermore, we have discovered that the ST-SPP wave packet also possesses a finite topological charge density, which exhibits a symmetric spatial distribution where only the sign differs between the left and right sides of the propagation axis. As the spin texture and topological charge density are determined by the gradient of the energy flow density, it is possible to design them by manipulating the ST-SPP wave packet. These findings suggest that three-dimensional localized electromagnetic fields can propagate with accompanying spin textures and may offer the possibility of observing optical topological particle propagation phenomena experimentally.

Acknowledgments

A.K. thanks Y. Dai for the valuable discussions.

Funding Sources

This work was supported by the JSPS KAKENHI (JP16823280, JP18967972, JP20J21825); MEXT Q-LEAP ATTO (JPMXS0118068681); and by the U.S. Office of Naval Research (ONR) under contracts N00014-17-1-2458 and N00014-20-1-2789. D.O. is supported by the President’s PhD Scholarships at Imperial College London, by JSPS Overseas Research Fellowship, by the Institution of Engineering and Technology (IET), and by Fundação para a Ciência e a Tecnologia and Instituto de Telecomunicações under project UIDB/50008/2020.

Appendix A The electric and magnetic field

The electric field brought about by an SPP is given by utilizing the transverse polarisation basis with the SPP dispersion relation. As the SPP wave is a transverse magnetic mode, the electric field will be given, in terms of the Hertz potential [110, 111, 112], in the reciprocal space as

𝑬=𝒌×𝒌×ψ𝒖z=(kz𝒌|𝒌|2𝒖z)ψ,\displaystyle\bm{E}=\bm{k}\times\bm{k}\times\psi\bm{u}_{z}=(k_{z}\bm{k}_{\parallel}-|\bm{k}_{\parallel}|^{2}\bm{u}_{z})\psi, (11)

where ψ𝒖z\psi\bm{u}_{z} corresponds to the Hertz vector potential, and we defined the parallel component of the wave vector, 𝒌=kx𝒖x+ky𝒖y\bm{k}_{\parallel}=k_{x}\bm{u}_{x}+k_{y}\bm{u}_{y}. The ratio of Ex,yE_{x,y} to EzE_{z} can be written as

Ex,yEz=kzkx,y|𝒌|2=ikzx,yksp2.\displaystyle\frac{E_{x,y}}{E_{z}}=\frac{k_{z}k_{x,y}}{-|\bm{k}_{\parallel}|^{2}}=\frac{ik_{z}\partial_{x,y}}{k_{\mathrm{sp}}^{2}}. (12)

Note that we have used the dispersion relation |𝒌|2=ksp2|\bm{k}_{\parallel}|^{2}=k_{\mathrm{sp}}^{2} and replaced kx,yix,yk_{x,y}\rightarrow-i\partial_{x,y}. Multiplying EzE_{z} from the right-hand side, we can get

𝑬=ikzksp2Ez.\displaystyle\bm{E}_{\parallel}=\frac{ik_{z}}{k_{\mathrm{sp}}^{2}}\nabla_{\parallel}E_{z}. (13)

where we have defined the parallel component of the electric field, 𝑬=𝑬Ez𝒖z\bm{E}_{\parallel}=\bm{E}-E_{z}\bm{u}_{z}, and the derivative in the direction parallel to the surface, :=𝒖xx+𝒖yy\nabla_{\parallel}:=\bm{u}_{x}\partial_{x}+\bm{u}_{y}\partial_{y}.

Once we get the electric field, the magnetic field can be calculated from Faraday’s law of induction, ×𝑬=ωμ0𝑯\nabla\times\bm{E}=\omega\mu_{0}\bm{H}, as follows:

𝑯\displaystyle\bm{H} =×𝑬ωμ0=(+kz𝒖z)×(𝑬+Ez𝒖z)ωμ0,\displaystyle=\frac{\nabla\times\bm{E}}{\omega\mu_{0}}=\frac{(\nabla_{\parallel}+k_{z}\bm{u}_{z})\times(\bm{E}_{\parallel}+E_{z}\bm{u}_{z})}{\omega\mu_{0}}, (14)
=i1+kz2/ksp2ωμ0𝒖z×Ez,\displaystyle=i\frac{1+k_{z}^{2}/k_{\mathrm{sp}}^{2}}{\omega\mu_{0}}\bm{u}_{z}\times\nabla_{\parallel}E_{z}, (15)
=iωϵϵ0ksp2𝒖z×Ez.\displaystyle=i\frac{\omega\epsilon\epsilon_{0}}{k_{\mathrm{sp}}^{2}}\bm{u}_{z}\times\nabla_{\parallel}E_{z}. (16)

Note that we have applied ksp2+kz2=(ω/c)2ϵk_{\mathrm{sp}}^{2}+k_{z}^{2}=(\omega/c)^{2}\epsilon. Note also that the permittivity is dependent on whether we are on the dielectric or metal side,

ϵ(z,ω)={ϵd(z>0),ϵm(ω)(z<0).\displaystyle\epsilon(z,\omega)=\begin{cases}\epsilon_{\mathrm{d}}&(z>0),\\ \epsilon_{\mathrm{m}}(\omega)&(z<0).\end{cases} (17)

Appendix B Wavefront direction

Refer to caption
Figure S1: (a) Spatial distribution of the ST-SPP wave packet, Re(Ez(x,y;t=0)E_{z}(x,y;t=0)) for different group velocity ranging from 0.8 cc to 1.0 cc. Top panels show the expanded views indicated by dashed squares. The colors of the frames (red, yellow, and blue) correspond to the colors of the dashed squares in the main panels. (b) One of the two branch structures that make up the ST-SPP wave packet, calculated with only the positive kxk_{x} component. Top panels show the expanded views indicated by black dashed squares. White solid lines in main panels and dashed lines in top panels indicate the peak positions of band structures itself and wavefronts of inside the band structures. White arrows are normal direction of the wavefront correspond to the propagate directions. The angles between each line and yy-axis were defined as α\alpha and β\beta. Note that the α\alpha and β\beta are not comparable quantities since the aspect ratio of the expanded views has been modified.

The electric spatial distribution, including fine fringes within the wave packet, is useful for visualizing the intricate motion of the ST-SPP wave packets. Fig. S1 displays the real part of the electric field EzE_{z} (x,y;t=0)(x,y;t=0) of the ST-SPP wave packets calculated for different group velocities ranging from 0.8c0.8~{}c to 1.0c1.0~{}c. The expanded views of the dashed squares are plotted in the top panels.

The entire spatial distributions show that the subluminal and superluminal ST-SPP wave packets with close vdiffv_{\mathrm{diff}} have similar envelope distributions, where vdiffv_{\mathrm{diff}} is a difference of the group velocity between the ST-SPP wave packet and the conventional SPP wave packet (vdiff=|vspvg~|)(v_{\mathrm{diff}}\!=\!|v_{\mathrm{sp}}-\widetilde{v_{g}}|). The angles ±α\pm\alpha of the ±kx\pm k_{x} branches, which constitute the ST-SPP wave packet, depend on the angle θ\theta of the spectral plane and vary with the group velocity. While the entire distribution of the wave packets corresponds, the fringes within the branches are oriented in contrasting directions in the superluminal and subluminal wave packets. The wavefronts of the subluminal ST-SPP are distributed in a rhombic shape surrounding the central region, whereas the wavefronts of the superluminal ST-SPP are distributed radially from the central region. In Fig. S1(b), we plotted the +kx+k_{x} branch calculated from only the +kx+k_{x} component of the ST-SPP wave packet. The angles between the wavefront inside the branch made with the propagation axis are defined as β\beta. The value of α\alpha varies widely with vg~\widetilde{v_{g}}, being negative during subluminal and positive during superluminal. In contrast, the group velocity dependence of β\beta is relatively small compared to α\alpha, and the wavefront faces the upper right direction regardless of vgv_{g}.

Appendix C Spatial distribution

Refer to caption
Figure S2: (a-c) Spatial field distribution of the typical parameters of ST-SPP wave packets in physical space for various group velocities ranging from 0.7c0.7~{}c to 1.1c1.1~{}c, (a) I(x,y,;t=0)I(x,y,;t\!=\!0), (b) tt-SAM(x,y;t=0){}_{\perp}(x,y;t\!=\!0), and (c) D(x,y;t=0)D(x,y;t\!=\!0). White lines in (a) are cross-sections along y=0y\!=\!0. (d) The spectral projections onto the (kyk_{y},ω\omega) and (kxk_{x},ω\omega)-plane. (e) Transverse wavenumber of the ST-SPP wave packets. The red circles and blue line represent the λx\lambda_{x} evaluated from I(x,y;t=0)I(x,y;t\!=\!0) (shwon in Fig. S2 (a)) and 2π/kc2\pi/k_{c} calculated from spectral domain (shown in Fig. S2 (d)), respectively. (f,g) Variation of the (f) SzS_{z} and (g) DD distribution with group velocity. The horizontal profiles correspond to the integral cross sections along the yy-direction of the Sz(x,y;t=0)S_{z}(x,y;t\!=\!0) and D(x,y;t=0)D(x,y;t\!=\!0) in the calculation region shown in Fig. S2(b,c) (30<y<30)μ(-30\!<\!y\!<\!30)~{}\mathrm{\mu}m.

As discussed in Fig. 3 and Fig. 4 in the main text, the spatial distribution of the SAM and topological charge density are determined by the spatial distribution of the ST-SPP wave packet, which vary with the designed group velocity. In Fig. S2(a-c), we plotted the I(x,y;t=0)I(x,y;t\!=\!0), tt-SAM(x,y;t=0){}_{\perp}(x,y;t\!=\!0), and D(x,y;t=0)D(x,y;t\!=\!0) calculated for different group velocities ranging from 0.7 cc to 1.1 cc. The projections of the spectral support domains onto the (ky,ω)(k_{y},\omega) and (kx,ω)(k_{x},\omega) planes at each group velocity were plotted in Fig. S2(d).

In both the superluminal and subluminal cases, the spectral domain projected onto the (ky,ω)(k_{y},\omega) plane moves away from the SPP dispersion curve as the vdiffv_{\mathrm{diff}} increase. Therefore, the transverse wavenumber kx(ω)=ksp2ky2k_{x}(\omega)\!=\!\sqrt{k_{\mathrm{sp}}^{2}-k_{y}^{2}} increases as vdiffv_{\mathrm{diff}} increases, resulting in the transverse fringe density of the ST-SPP wave packet is at its lowest value at vg~=vsp\widetilde{v_{g}}=v_{\mathrm{sp}}. Under the calculation conditions of this paper, in which the spectrum of the ST-SPP wave packet is assumed to have a Fourier-limited pulse with a peak at ωc=375\omega_{c}\!=\!375 THz , the transverse wavelength λx\lambda_{x} of the ST-SPP wave packet is expressed as λx=2π/kc\lambda_{x}\!=\!2\pi/k_{c} (Fig. S2(e)), where kck_{c} is transverse wavenumber evaluated at ωc\omega_{c} (Fig S2(d)).

Since a region extent of the branches distributed on the y=0y\!=\!0 axis is determined by the angle of the branches, the transverse Gaussian spatial distribution of the envelope of the ST-SPP wave packet, wxw_{x} also vary with vgv_{g}. This change of the spatial width can also be interpreted as a change in wave packet width due to the difference in Δkx\Delta k_{x}. For vg~vsp\widetilde{v_{g}}\!\simeq\!v_{\mathrm{sp}}, the spectral domain in (kx,ω)(k_{x},\omega) plane shows a slight curvature but almost constant |kx||k_{x}|, resulting in very small Δkx\Delta k_{x}. The spatial distribution of the wave packet is similar to that of a striped ST-SPP wave packet whose spectral domain is defined by iso-kxk_{x} plane [70], and wxw_{x} is almost infinite. In contrast, for ST-SPP wave packets with large vdiffv_{\mathrm{diff}}, the difference between kyk_{y} and kspk_{\mathrm{sp}} is larger for larger Ω=ωoω\Omega\!=\!\omega_{o}-\omega, resulting in large Δkx\Delta k_{x}. From Fig. S2, we can confirm the inverse correlation between Δkx\Delta k_{x} and wxw_{x}.

A two-dimensional plot of the cross-sections of the tt-SAM (Fig. S2(b)) integrated along the yy-direction is plotted in Fig. S2(f). The vertical axis shows the group velocity normalized by the speed of light in a vacuum. Since the spatial distribution of the tt-SAM reflects the intensity profile of the ST-SPP wave packet, the tt-SAM becomes densely distributed in the central region as vdiffv_{\mathrm{diff}} increases. In addition, the maximum intensity of the SAM also increases with increasing vdiffv_{\mathrm{diff}} because finer fringes imply a larger gradient of the energy flow density. A two-dimensional plot of the integrated cross section of DD plotted in Fig. S2(g) shows a spatial distribution corresponding to the SAM. The left-right reversal of the sign of DD associated with the change in vg~\widetilde{v_{g}} discussed in the main text can be confirmed.

References

  • Poynting [1909] J. H. Poynting, The wave motion of a revolving shaft, and a suggestion as to the angular momentum in a beam of circularly polarised light, Proc. R. Soc. Lond. A 82, 560 (1909).
  • Beth [1935] R. A. Beth, Direct detection of the angular momentum of light, Phys. Rev. 48, 471 (1935).
  • Beth [1936] R. A. Beth, Mechanical detection and measurement of the angular momentum of light, Phys. Rev. 50, 115 (1936).
  • Sugiura et al. [1990] T. Sugiura, S. Kawata, and S. Minami, Optical rotation of small particles by a circulary-polarized laser beam in an optical microscope, J. Spectrosc. Soc. Jpn 39, 342 (1990).
  • Higurashi et al. [1994] E. Higurashi, H. Ukita, H. Tanaka, and O. Ohguchi, Optically induced rotation of anisotropic micro-objects fabricated by surface micromachining, Appl. Phys. Lett. 64, 2209 (1994).
  • Higurashi et al. [1997] E. Higurashi, O. Ohguchi, T. Tamamura, H. Ukita, and R. Sawada, Optically induced rotation of dissymmetrically shaped fluorinated polyimide micro-objects in optical traps, J. Appl. Phys. 82, 2773 (1997).
  • Friese et al. [1998] M. Friese, T. Nieminen, N. Heckenberg, and H. Rubinsztein-Dunlop, Optical alignment and spinning of laser-trapped microscopic particles, Nature 394, 348 (1998).
  • Ahn et al. [2018] J. Ahn, Z. Xu, J. Bang, Y.-H. Deng, T. M. Hoang, Q. Han, R.-M. Ma, and T. Li, Optically levitated nanodumbbell torsion balance and ghz nanomechanical rotor, Phys. Rev. Lett. 121, 033603 (2018).
  • Monteiro et al. [2018] F. Monteiro, S. Ghosh, E. C. van Assendelft, and D. C. Moore, Optical rotation of levitated spheres in high vacuum, Phys. Rev. A 97, 051802(R) (2018).
  • Reimann et al. [2018] R. Reimann, M. Doderer, E. Hebestreit, R. Diehl, M. Frimmer, D. Windey, F. Tebbenjohanns, and L. Novotny, Ghz rotation of an optically trapped nanoparticle in vacuum, Phys. Rev. Lett. 121, 033602 (2018).
  • Jin et al. [2021] Y. Jin, J. Yan, S. J. Rahman, J. Li, X. Yu, and J. Zhang, 6 ghz hyperfast rotation of an optically levitated nanoparticle in vacuum, Photonics Res. 9, 1344 (2021).
  • Aiello et al. [2009a] A. Aiello, N. Lindlein, C. Marquardt, and G. Leuchs, Transverse angular momentum and geometric spin hall effect of light, Phys. Rev. Lett. 103, 100401 (2009a).
  • Aiello et al. [2010a] A. Aiello, C. Marquardt, and G. Leuchs, Transverse angular momentum of photons, Phys. Rev. A 81, 053838 (2010a).
  • Bliokh and Nori [2015] K. Y. Bliokh and F. Nori, Transverse and longitudinal angular momenta of light, Phys. Rep. 592, 1 (2015).
  • Bekshaev [2009] A. Y. Bekshaev, Oblique section of a paraxial light beam: criteria for azimuthal energy flow and orbital angular momentum, J Opt. A: Pure Appl. Opt. 11, 094003 (2009).
  • Aiello et al. [2009b] A. Aiello, N. Lindlein, C. Marquardt, and G. Leuchs, Transverse angular momentum and geometric spin hall effect of light, Phys. Rev. lett. 103, 100401 (2009b).
  • Aiello et al. [2010b] A. Aiello, C. Marquardt, and G. Leuchs, Transverse angular momentum of photons, Phys. Rev. A 81, 053838 (2010b).
  • Antognozzi et al. [2016] M. Antognozzi, C. Bermingham, R. Harniman, S. Simpson, J. Senior, R. Hayward, H. Hoerber, M. Dennis, A. Bekshaev, K. Bliokh, et al., Direct measurements of the extraordinary optical momentum and transverse spin-dependent force using a nano-cantilever, Nat. Phys. 12, 731 (2016).
  • Canaguier-Durand and Genet [2015] A. Canaguier-Durand and C. Genet, Plasmonic lateral forces on chiral spheres, J Opt. 18, 015007 (2015).
  • Canaguier-Durand and Genet [2014] A. Canaguier-Durand and C. Genet, Transverse spinning of a sphere in a plasmonic field, Phys. Rev. A 89, 033841 (2014).
  • Junge et al. [2013] C. Junge, D. O’Shea, J. Volz, and A. Rauschenbeutel, Strong coupling between single atoms and nontransversal photons, Phys. Rev. Lett. 110, 213604 (2013).
  • Aiello et al. [2015] A. Aiello, P. Banzer, M. Neugebaueru, and G. Leuchs, From transverse angular momentum to photonic wheels, Nat. Photonics 9, 789 (2015).
  • Zhang et al. [2022] X. Zhang, B. Shen, Z. Zhu, G. Rui, J. He, Y. Cui, and B. Gu, Understanding of transverse spin angular momentum in tightly focused linearly polarized vortex beams, Opt. Express 30, 5121 (2022).
  • Abujetas and Sanchez-Gil [2020] D. R. Abujetas and J. A. Sanchez-Gil, Spin angular momentum of guided light induced by transverse confinement and intrinsic helicity, ACS Photon. 7, 534 (2020).
  • Bekshaev et al. [2015] A. Y. Bekshaev, K. Y. Bliokh, and F. Nori, Transverse spin and momentum in two-wave interference, Phys. Rev. X 5, 011039 (2015).
  • Bliokh et al. [2014] K. Y. Bliokh, A. Y. Bekshaev, and F. Nori, Extraordinary momentum and spin in evanescent waves, Nat. Commun. 5, 3300 (2014).
  • Bliokh and Nori [2012] K. Y. Bliokh and F. Nori, Transverse spin of a surface polariton, Phys. Rev. A 85, 061801(R) (2012).
  • Kim et al. [2012] K.-Y. Kim, I.-M. Lee, J. Kim, J. Jung, and B. Lee, Time reversal and the spin angular momentum of transverse-electric and transverse-magnetic surface modes, Phys. Rev. A 86, 063805 (2012).
  • Kim and Wang [2015] K.-Y. Kim and A. X. Wang, Spin angular momentum of surface modes from the perspective of optical power flow, Opt. Lett. 40, 2929 (2015).
  • Bliokh et al. [2017] K. Y. Bliokh, A. Y. Bekshaev, and F. Nori, Optical momentum, spin, and angular momentum in dispersive media, Phys. Rev. Lett. 119, 073901 (2017).
  • Dai et al. [2018] Y. Dai, M. Dabrowski, V. A. Apkarian, and H. Petek, Ultrafast microscopy of spin-momentum-locked surface plasmon polaritons, ACS Nano 12, 6588 (2018).
  • Bliokh et al. [2015] K. Y. Bliokh, D. Smirnova, and F. Nori, Quantum spin hall effect of light, Science 348, 1448 (2015).
  • Mechelen and Jacob [2016] T. V. Mechelen and Z. Jacob, Universal spin-momentum locking of evanescent waves, Optica 3, 118 (2016).
  • O’Connor et al. [2014] D. O’Connor, P. Ginzburg, F. J. Rodriguez-Fortuno, G. A. Wurtz, and A. V. Zayats, Spin-orbit coupling in surface plasmon scattering by nanostructures, Nat. Commun. 5, 5327 (2014).
  • Rodriguez-Fortuno et al. [2013] F. J. Rodriguez-Fortuno, G. Marino, P. Ginzburg, D. O’Connor, A. Martinez, G. A. Wurtz, and A. V. Zayats, Near-field interference for the unidirectional excitation of electromagnetic guided modes, Science 340, 328 (2013).
  • Dai and Petek [2019] Y. Dai and H. Petek, Plasmonic spin-hall effect in surface plasmon polariton focusing, ACS Photon. 6, 2005 (2019).
  • Shi et al. [2021a] P. Shi, L. Du, and X. Yuan, Spin photonics: from transverse spin to photonic skyrmions, Nanophotonics 10, 3927 (2021a).
  • Cardano and Marrucci [2015] F. Cardano and L. Marrucci, Spin-orbit photonics, Nat. Photonics 9, 776 (2015).
  • Du et al. [2019] L. Du, A. Yang, A. V. Zayats, and X. Yuan, Deep-subwavelength features of photonic skyrmions in a confined electromagnetic field with orbital angular momentum, Nat. Phys. 15, 650 (2019).
  • Dai et al. [2020] Y. Dai, Z. Zhou, A. Ghosh, R. S. K. Mong, A. Kubo, C.-B. Huang, and H. Petek, Plasmonic topological quasiparticle on the nanometre and femtosecond scales, Nature 588, 616 (2020).
  • Lei et al. [2021] X. Lei, A. Yang, P. Shi, Z. Xie, L. Du, A. V. Zayats, and X. Yuan, Photonic spin lattices: symmetry constraints for skyrmion and meron topologies, Phys. Rev. Lett. 127, 237403 (2021).
  • Zhang et al. [2021] Q. Zhang, Z. Xie, L. Du, P. Shi, and X. Yuan, Bloch-type photonic skyrmions in optical chiral multilayers, Phys. Rev. Res. 3, 023109 (2021).
  • Ghosh et al. [2021] A. Ghosh, S. Yang, Y. Dai, Z. Zhou, T. Wang, C.-B. Huang, and H. Petek, A topological lattice of plasmonic merons, Appl. Phys. Rev. 8, 041413 (2021).
  • Dai et al. [2022a] Y. Dai, Z. Zhou, A. Ghosh, K. Kapoor, M. Dabrowski, A. Kubo, C.-B. Huang, and H. Petek, Ultrafast microscopy of a twisted plasmonic spin skyrmion, Appl. Phys. Rev. 9, 011420 (2022a).
  • Shi et al. [2021b] P. Shi, L. Du, M. Li, and X. Yuan, Symmetry-protected photonic chiral spin textures by spin-orbit coupling, Laser Photon. Rev. 15, 2000554 (2021b).
  • Shi et al. [2021c] P. Shi, L. Du, C. Li, A. Zayats, V, and X. Yuan, Transverse spin dynamics in structured electromagnetic guided waves, Proc. Natl. Acad. Sci. (USA) 118, e2018816118 (2021c).
  • Taneja et al. [2021] C. Taneja, D. Paul, and G. V. P. Kumar, Experimental observation of transverse spin of plasmon polaritons in a single crystalline silver nanowire, Appl. Phys. Lett. 119, 161108 (2021).
  • Dai et al. [2022b] Y. Dai, A. Ghosh, S. Yang, Z. Zhou, C.-b. Huang, and H. Petek, Poincare engineering of surface plasmon polaritons, Nat. Rev. Phys. 4, 562 (2022b).
  • [49] A. Ghosh, S. Yang, Y. Dai, and H. Petek, The spin texture topology of polygonal plasmon fields, ACS Photon. , doi:10.1021/acsphotonics.2c01491.
  • Shen et al. [2022] Y. Shen, E. C. Martinez, and C. Rosales-Guzman, Generation of optical skyrmions with tunable topological textures, ACS Photon. 9, 296 (2022).
  • Shen [2021] Y. Shen, Topological bimeronic beams, Opt. Lett. 46, 3737 (2021).
  • Deng et al. [2022] Z.-L. Deng, T. Shi, A. Krasnok, X. Li, and A. Alu, Observation of localized magnetic plasmon skyrmions, Nat. Commun. 13, 8 (2022).
  • Xiong et al. [2021] L. Xiong, Y. Li, D. Halbertal, M. Sammon, Z. Sun, S. Liu, J. H. Edgar, T. Low, M. M. Fogler, C. R. Dean, A. J. Millis, and D. N. Basov, Polaritonic vortices with a half-integer charge, Nano Lett. 21, 9256 (2021).
  • Tsesses et al. [2018] S. Tsesses, E. Ostrovsky, K. Cohen, B. Gjonaj, N. H. Lindner, and G. Bartal, Optical skyrmion lattice in evanescent electromagnetic fields, Science 361, 993 (2018).
  • Davis et al. [2020] T. J. Davis, D. Janoschka, P. Dreher, B. Frank, F.-J. M. zu Heringdorf, and H. Giessen, Ultrafast vector imaging of plasmonic skyrmion dynamics with deep subwavelength resolution, Science 368, eaba6415 (2020).
  • Nagaosa and Tokura [2013] N. Nagaosa and Y. Tokura, Topological properties and dynamics of magnetic skyrmions, Nat. Nanotechnol. 8, 899 (2013).
  • Jiang et al. [2015] W. Jiang, P. Upadhyaya, W. Zhang, G. Yu, M. B. Jungfleisch, F. Y. Fradin, J. E. Pearson, Y. Tserkovnyak, K. L. Wang, O. Heinonen, S. G. E. te Velthuis, and A. Hoffmann, Blowing magnetic skyrmion bubbles, Science 349, 283 (2015).
  • Woo et al. [2016] S. Woo, K. Litzius, B. Krueger, M.-Y. Im, L. Caretta, K. Richter, M. Mann, A. Krone, R. M. Reeve, M. Weigand, P. Agrawal, I. Lemesh, M.-A. Mawass, P. Fischer, M. Klaeui, and G. R. S. D. Beach, Observation of room-temperature magnetic skyrmions and their current-driven dynamics in ultrathin metallic ferromagnets, Nat. Mat. 15, 501 (2016).
  • Jiang et al. [2017] W. Jiang, G. Chen, K. Liu, J. Zang, S. G. E. te Velthuis, and A. Hoffmann, Skyrmions in magnetic multilayers, Phys. Rep. 704, 1 (2017).
  • Zhang et al. [2018] S. L. Zhang, W. W. Wang, D. M. Burn, H. Peng, H. Berger, A. Bauer, C. Pfleiderer, G. van der Laan, and T. Hesjedal, Manipulation of skyrmion motion by magnetic field gradients, Nat. Commun. 9, 2115 (2018).
  • Bai et al. [2020] C. Bai, J. Chen, Y. Zhang, D. Zhang, and Q. Zhan, Dynamic tailoring of an optical skyrmion lattice in surface plasmon polaritons, Opt. Express 28, 10320 (2020).
  • Lin et al. [2021] M. Lin, W. Zhang, C. Liu, L. Du, and X. Yuan, Photonic spin skyrmion with dynamic position control, ACS Photon. 8, 2567 (2021).
  • Ichiji et al. [2022a] N. Ichiji, Y. Otake, and A. Kubo, Femtosecond imaging of spatial deformation of surface plasmon polariton wave packet during resonant interaction with nanocavity, Nanophotonics 11, 1321 (2022a).
  • Wang et al. [2020] Y. Wang, C. Ming, Y. Zhang, J. Xu, F. Feng, L. Li, and X. Yuan, Spatiotemporal manipulation on focusing and propagation of surface plasmon polariton pulses, Opt. Express 28, 33516 (2020).
  • Karalis and Joannopoulos [2019] A. Karalis and J. D. Joannopoulos, Plasmonic metasurface “bullets” and other “moving objects”: Spatiotemporal dispersion cancellation for linear passive subwavelength slow light, Phys. Rev. Lett. 123, 067403 (2019).
  • Schepler et al. [2020] K. L. Schepler, M. Yessenov, Y. Zhiyenbayev, and A. F. Abouraddy, Space–time surface plasmon polaritons: A new propagation-invariant surface wave packet, ACS Photon. 7, 2966 (2020).
  • Kondakci and Abouraddy [2017] H. E. Kondakci and A. F. Abouraddy, Diffraction-free space-time beams, Nat. Photonics 11, 733 (2017).
  • Bhaduri et al. [2018] B. Bhaduri, M. Yessenov, and A. F. Abouraddy, Meters-long propagation of diffraction-free space-time light sheets, Opt. Express 26, 20111 (2018).
  • Yessenov et al. [2022] M. Yessenov, L. A. Hall, K. L. Schepler, and A. F. Abouraddy, Space-time wave packets, Adv. Opt. Photon. 14, 455 (2022).
  • [70] N. Ichiji, H. Kikuchi, M. Yessenov, K. L. Schepler, A. F. Abouraddy, and A. Kubo, Observation of ultrabroadband striped space-time surface plasmon polaritons, ACS Photon. , doi:10.1021/acsphotonics.2c00296.
  • Ichiji et al. [2022b] N. Ichiji, M. Yessenov, K. L. Schepler, A. F. Abouraddy, and A. Kubo, Exciting space-time surface plasmon polaritons by irradiating a nanoslit structure, arXiv:2212.05450  (2022b).
  • Lassaline et al. [2020] N. Lassaline, R. Brechbuhler, S. J. W. Vonk, K. Ridderbeek, M. Spieser, S. Bisig, B. Le Feber, F. T. Rabouw, and D. J. Norris, Optical fourier surfaces, Nature 582, 506 (2020).
  • Rakić et al. [1998] A. D. Rakić, A. B. Djurišić, J. M. Elazar, and M. L. Majewski, Optical properties of metallic films for vertical-cavity optoelectronic devices, Appl. Opt. 37, 5271 (1998).
  • Kubo et al. [2007] A. Kubo, N. Pontius, and H. Petek, Femtosecond microscopy of surface plasmon polariton wave packet evolution at the silver/vacuum interface, Nano Lett. 7, 470 (2007).
  • Zhang et al. [84] L. Zhang, A. Kubo, L. Wang, H. Petek, and T. Seideman, Imaging of surface plasmon polariton fields excited at a nanometer-scale slit, Phys. Rev. B 2011, 245442 (84).
  • Lin et al. [2012] J. Lin, J. Dellinger, P. Genevet, B. Cluzel, F. de Fornel, and F. Capasso, Cosine-Gauss plasmon beam: A localized long-range nondiffracting surface wave, Phys. Rev. Lett. 109, 093904 (2012).
  • Kondakci and Abouraddy [2019] H. E. Kondakci and A. F. Abouraddy, Optical space-time wave packets of arbitrary group velocity in free space, Nat. Commun. 10, 929 (2019).
  • Yessenov and Abouraddy [2020] M. Yessenov and A. F. Abouraddy, Accelerating and decelerating space-time wave packets in free space, Phys. Rev. Lett. 125, 233901 (2020).
  • Hall et al. [2022] L. A. Hall, M. Yessenov, and A. F. Abouraddy, Arbitrarily accelerating space-time wave packets, Opt. Lett. 47, 694 (2022).
  • Yessenov et al. [2021] M. Yessenov, L. A. Hall, and A. F. Abouraddy, Engineering the optical vacuum: arbitrary magnitude, sign, and order of dispersion in free space using space-time wave packets, ACS Photon. 8, 2274 (2021).
  • Hall and Abouraddy [2021] L. A. Hall and A. F. Abouraddy, Free-space group-velocity dispersion induced in space-time wave packets by v-shaped spectra, Phys. Rev. A 104, 013505 (2021).
  • He et al. [2022] H. He, C. Guo, and M. Xiao, Nondispersive space-time wave packets propagating in dispersive media, Laser Photon. Rev. 16, 2100634 (2022).
  • Saari and Reivelt [1997] P. Saari and K. Reivelt, Evidence of X-shaped propagation-invariant localized light waves, Phys. Rev. Lett. 79, 4135 (1997).
  • Bekshaev and Soskin [2007] A. Y. Bekshaev and M. Soskin, Transverse energy flows in vectorial fields of paraxial beams with singularities, Opt. Commun. 271, 332 (2007).
  • Berry [2009] M. V. Berry, Optical currents, J Opt. A 11, 094001 (2009).
  • Bekshaev et al. [2011] A. Bekshaev, K. Y. Bliokh, and M. Soskin, Internal flows and energy circulation in light beams, J. Opt. 13, 053001 (2011).
  • Bekshaev [2013] A. Y. Bekshaev, Subwavelength particles in an inhomogeneous light field: optical forces associated with the spin and orbital energy flows, J. Opt. 15, 044004 (2013).
  • Fösel et al. [2017] T. Fösel, V. Peano, and F. Marquardt, L lines, c points and chern numbers: understanding band structure topology using polarization fields, New J. Phys. 19, 115013 (2017).
  • Dennis [2002] M. Dennis, Polarization singularities in paraxial vector fields: morphology and statistics, Opt. Commun. 213, 201 (2002).
  • Deschamps et al. [1970] J. Deschamps, M. Fitaire, and M. Lagoutte, Inverse faraday effect in a plasma, Phys. Rev. Lett. 25, 1330 (1970).
  • Sheng and Meyer-ter-Vehn [1996] Z. M. Sheng and J. Meyer-ter-Vehn, Inverse faraday effect and propagation of circularly polarized intense laser beams in plasmas, Phys. Rev. E 54, 1833 (1996).
  • Hertel [2006] R. Hertel, Theory of the inverse faraday effect in metals, J. Magn. Magn. Mater. 303, L1 (2006).
  • Zhang et al. [2009] H.-L. Zhang, Y.-Z. Wang, and X.-J. Chen, A simple explanation for the inverse faraday effect in metals, J. Magn. Magn. Mater. 321, L73 (2009).
  • Oue and Matsuo [2020a] D. Oue and M. Matsuo, Electron spin transport driven by surface plasmon polaritons, Phys. Rev. B 101, 161404(R) (2020a).
  • Wijaya et al. [2020] T. J. Wijaya, D. Oue, M. Matsuo, Y. Ito, K. Elphick, H. Uchida, M. Inoue, and A. Hirohata, Spin current generation by a surface plasmon polariton, arXiv:2004.12348  (2020).
  • Bekshaev [2022] A. Bekshaev, Spin-accumulation and spin-transport effects induced by surface plasmon-polariton waves, J. Opt. Soc. Am. B 39, 3187 (2022).
  • Oue and Matsuo [2020b] D. Oue and M. Matsuo, Effects of surface plasmons on spin currents in a thin film system, New J. Phys. 22, 033040 (2020b).
  • Oue and Matsuo [2020c] D. Oue and M. Matsuo, Optically induced electron spin currents in the Kretschmann configuration, Phys. Rev. B 102, 125431 (2020c).
  • Ukhtary et al. [2021] M. S. Ukhtary, Y. Tian, and R. Saito, Spin current generation by edge plasmons in graphene ribbons, Phys. Rev. B 103, 245428 (2021).
  • Tian et al. [2022] Y. Tian, M. S. Ukhtary, and R. Saito, Switching performance of optically generated spin current at the graphene edge, Phys. Rev. B 106, 045420 (2022).
  • Göbel et al. [2021] B. Göbel, I. Mertig, and O. A. Tretiakov, Beyond skyrmions: review and perspectives of alternative magnetic quasiparticles, Phys. Rep. 895, 1 (2021).
  • Yang et al. [2021] S.-H. Yang, R. Naaman, Y. Paltiel, and S. S. P. Parkin, Chiral spintronics, Nat. Rev. Phys. 3, 328 (2021).
  • Hamamoto et al. [2016] K. Hamamoto, M. Ezawa, and N. Nagaosa, Purely electrical detection of a skyrmion in constricted geometry, Appl. Phys. Lett. 108, 112401 (2016).
  • Maccariello et al. [2018] D. Maccariello, W. Legrand, N. Reyren, K. Garcia, K. Bouzehouane, S. Collin, V. Cros, and A. Fert, Electrical detection of single magnetic skyrmions in metallic multilayers at room temperature, Nat. Nanotech. 13, 233 (2018).
  • Schulz et al. [2012] T. Schulz, R. Ritz, A. Bauer, M. Halder, M. Wagner, C. Franz, C. Pfleiderer, K. Everschor, M. Garst, and A. Rosch, Emergent electrodynamics of skyrmions in a chiral magnet, Nat. Phys. 8, 301 (2012).
  • Bruno et al. [2004] P. Bruno, V. K. Dugaev, and M. Taillefumier, Topological hall effect and berry phase in magnetic nanostructures, Phys. Rev. Lett. 93, 096806 (2004).
  • Neubauer et al. [2009] A. Neubauer, C. Pfleiderer, B. Binz, A. Rosch, R. Ritz, P. G. Niklowitz, and P. Böni, Topological hall effect in the aa phase of mnsi, Phys. Rev. Lett. 102, 186602 (2009).
  • Lee et al. [2009] M. Lee, W. Kang, Y. Onose, Y. Tokura, and N. P. Ong, Unusual hall effect anomaly in mnsi under pressure, Phys. Rev. Lett. 102, 186601 (2009).
  • Jiang et al. [2020] Y. Jiang, H. Y. Yuan, Z.-X. Li, Z. Wang, H. W. Zhang, Y. Cao, and P. Yan, Twisted magnon as a magnetic tweezer, Phys. Rev. Lett. 124, 217204 (2020).
  • Nisbet [1955] A. Nisbet, Hertzian electromagnetic potentials and associated gauge transformations, Proc. R. Soc. Lond. A Math. Phys. Sci. 231, 250 (1955).
  • Hacyan et al. [1990] S. Hacyan, R. Jáuregui, F. Soto, and C. Villarreal, Spectrum of electromagnetic fluctuations in the casimir effect, J. Phys. A Math. Gen. 23, 2401 (1990).
  • Jackson [1999] J. D. Jackson, Classical Electrodynamics, third edition ed. (John Wiley & Sons, Inc., NewYork, 1999).