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

Metastability of photonic spin meron lattices in the presence of perturbed spin-orbit coupling

Open Access Open Access

Abstract

Photonic skyrmions and merons are topological quasiparticles characterized by nontrivial electromagnetic textures, which have received increasing research attention recently, providing novel degree of freedom to manipulate light-matter interactions and exhibiting excellent potential in deep-subwavelength imaging and nanometrology. Here, the topological stability of photonic spin meron lattices, which indicates the invariance of skyrmion number and robustness of spin texture under a continuous deformation of the field configuration, is demonstrated by inducing a perturbation to break the C4 symmetry in the presence spin-orbit coupling in an optical field. We revealed that amplitude perturbation would result in an amplitude-dependent shift of spin center, while phase perturbation leads to the deformation of domain walls, manifesting the metastability of photonic meron. Such spin topology is verified through the interference of plasmonic vortices with a broken rotational symmetry. The results provide new insights on optical topological quasiparticles, which may pave the way towards applications in topological photonics, optical information storage and transfer.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Topological spin textures in condensed matter physics such as skyrmions characterized by a magnetization swirl have attracted numerous research attention for the past decades [17], as they possess ultracompact size and great stability due to the topological nature of structure, promising for data storage and spintronic applications [811]. This topological feature was observed in optical system recently, with photonic counterpart of magnetic skyrmions constructed by either spin angular momentums (SAMs) [1216], electromagnetic fields [1720] or Stokes pseudospin vectors [2124]. Among the numerous realizations, photonic spin skyrmions manifest the topological properties in real space through the spin-momentum locking in evanescent fields [2527], and the deep-subwavelength features in local spin distributions have demonstrated advanced applications in nanoscale metrology [28] and magnetic domain imaging [29].

In addition to skyrmions carrying integer topological charge, exploration of topological quasiparticles with a half-integer topological charge known as merons, which originate from a singular solution of Yang-Mills theories [30], is another active topic in both condensed-matter physics and optics. In magnetic materials, merons are stabilized in the form of pairs or square lattices through easy-plane anisotropy [3136]. Similarly, photonic spin merons were demonstrated to be constructed as square lattices in the presence of spin-orbit coupling in the environment with broken rotational symmetry very recently [37,38], where nontrivial winding in the impinging field gives rise to topological nontrivial spin textures with topology determined by the symmetry of field. Higher order merons and bimerons are also suggested in polarization textures [39,40].

While numerous skyrmionic objects have been proposed in optical community, the topological stability under a continuous deformation of the field configuration has not yet been investigated. In this paper, we demonstrate the metastability of photonic spin meron lattices by inducing a perturbation in the presence spin-orbit coupling. Theoretical model is established by breaking the C4 symmetry of the field, indicating that the gradient of perturbation is the key factor to determine the spin topology. By employing the generalized spin-momentum relation, we reveal that the central of spin texture in each unit cell will be shifted with domains unchanged under amplitude perturbation. While phase perturbation will deform the domain walls into trapezoid, and the skyrmion number is conserved, manifesting the metastability of square meron lattices. Finally, a proof-of-concept experiment is performed utilizing plasmonic vortices to demonstrate the spin topology under perturbation.

2. Results and discussion

2.1 Theoretical model

Photonic spin meron lattices are formed due to the spin–orbit coupling in evanescent optical vortex (eOV) under a C4 symmetry of the field, which can be described by an interfered Hertz vector potential in a source free and homogeneous medium as

$$\Psi = \sum\limits_{n = 1}^4 {{A_n}{e^{il{\varphi _n}}}{e^{i{k_r}{\mathbf r} \cdot {{\mathbf e}_n}}}{e^{ - {k_z}z}}}, $$
where An ≡ 1, φn = /2, en = (cosφn, sinφn), r = (x, y) is in-plane coordinate, kr and ikz are the transverse and longitudinal wave vector components satisfying $k_r^2 - k_z^2 = {k^2}$ with k denoting the wave vector, l is an integer corresponding to the topological charge of the eOV (we consider l = 1 in the following). Taking transverse magnetic (TM) mode as an example, time averaging Poynting vector P = Re(E*×H)/2 representing a directional energy flux can be calculated through the Hertz potential as
$${\mathbf P} = \frac{{\omega \varepsilon k_r^2}}{2}{\mathop{\rm Im}\nolimits} ({{\Psi ^ \ast }\nabla \Psi } ),$$
where ω is the angular frequency of the wave and ε is the absolute permittivity of the medium. The generalized spin-momentum relation between the Poynting vector and SAM yields [25,37]
$${\mathbf S} = \frac{1}{{2{\omega ^2}}}\nabla \times {\mathbf P} = \frac{{\varepsilon k_r^2}}{{4\omega }}{\mathop{\rm Im}\nolimits} (\nabla {\Psi ^ \ast } \times \nabla \Psi ) = \frac{{2\varepsilon k_r^3{e^{ - 2{k_z}z}}}}{\omega }\left( {\begin{array}{*{20}{c}} {{k_z}\sin ({k_r}x)\cos ({k_r}y)}\\ {{k_z}\cos ({k_r}x)\sin ({k_r}y)}\\ {{k_r}\cos ({k_r}x)\cos ({k_r}y)} \end{array}} \right). $$

The skyrmionic vector field is constructed from the unit vectors in the direction of a three-component spin in Eq. (3) as n = S/|S|, which possesses distinct domains where nz = 0. The topological invariant is characterized by the skyrmion number as [6,7]

$$N = \frac{1}{{4\pi }}\int\!\!\!\int_\Sigma {{\mathbf n} \cdot ({{\partial_x}{\mathbf n} \times {\partial_y}{\mathbf n}} )} dxdy. $$

For each unit cell of the spin lattice, the integral region Σ in Eq. (4) is |krx − mπ|<π/2, |kry − nπ|<π/2 with (m, n) denoting integer numbers, and the skyrmion number can be calculated as N = ± ½, manifesting the photonic spin meron topology.

The resultant C4 symmetry in the spin meron lattice is a consequence of C4 symmetry of Hertz potentials $\Psi [\hat{R}({\pi / 2}){\mathbf r}] = \Psi ({\mathbf r})\exp(i{\pi / 2})$, where $\hat{R}(\varphi )$ is the rotation matrix along z-axis. As a perturbation is imposed in the presence of spin-orbit coupling to break the C4 symmetry, additional term should be imported to the Hertz potential as Ψp = Ψ + Φ. The SAM can be obtained from Eq. (3) as

$${\mathbf S} \propto \left\langle {\nabla \Psi } \right|\times i|{\nabla \Psi } \rangle + 2\left\langle {\nabla \Phi } \right|\times i|{\nabla \Psi } \rangle + \left\langle {\nabla \Phi } \right|\times i|{\nabla \Phi } \rangle. $$

The first and third terms in Eq. (5) correspond to the Berry curvatures of original Hertz potential Ψ and perturbation Φ respectively, while the second term denotes the interaction between Ψ and Φ, indicating that the gradient of perturbation is the key factor to determine the variation on SAM. Amplitude and phase perturbations are considered to represent Φ in the following to analyze the topological property of spin vectors under perturbation.

2.2 Spin topology under amplitude and phase perturbations

By introducing an amplitude perturbation δ in Eq. (1) (assume A2 is changed to 1 + δ without loss of generality), the C4 symmetry is broken, and the Hertz potential is expressed as

$${\Psi _a} = 2[i\sin ({k_r}x) - \sin ({k_r}y) - \frac{\delta }{2}{e^{ - i{k_r}x}}]{e^{ - {k_z}z}}, $$
where δ is a real number and Φ = −δexp(−ikrx − kzz) accordingly. The curl relation between P and S implies that the spin vectors point along z axis only at the center of Poynting vector vortex where P = 0, which is the zero points of Hertz potential from Eq. (2). Consequently, the amplitude perturbation takes a limitation that |δ|< 2 to guarantee a topological feature in spin textures. This can also be perceived in view of the Poynting vector and SAM distributions, which can be calculated from Eqs. (23) as
$${{\mathbf P}_a} = 2\omega \varepsilon k_r^3{e^{ - 2{k_z}z}}\left( {\begin{array}{*{20}{c}} { - (1 + {\delta / 2})[\cos ({k_r}x)\sin ({k_r}y) + {\delta / 2}]}\\ {(1 + {\delta / 2})\sin ({k_r}x)\cos ({k_r}y)}\\ 0 \end{array}} \right), $$
$${{\mathbf S}_a} = \frac{{2\varepsilon k_r^3{e^{ - 2{k_z}z}}}}{\omega }\left( {\begin{array}{*{20}{c}} {{k_z}(1 + {\delta / 2})\sin ({k_r}x)\cos ({k_r}y)}\\ {{k_z}(1 + {\delta / 2})[\cos ({k_r}x)\sin ({k_r}y) + {\delta / 2}]}\\ {{k_r}(1 + {\delta / 2})\cos ({k_r}x)\cos ({k_r}y)} \end{array}} \right). $$

The domains where the z component of the local spin orientation nz = 0 does not change under amplitude perturbation, since there is only a coefficient correction on Sz. To form a topological spin texture, the local spin orientation should rotate progressively from the central ‘up’ or ‘down’ state where nz = ± 1 to the edge, corresponding to a mapping onto the unit sphere from the north pole. Hence the in-plane SAM components should have zero points inside each unit cell, yielding |δ|< 2 from Eq. (8). And the central of the spin texture where nz = ± 1 would be shifted by |sin-1(δ/2)|/kr along y axis.

The Poynting vector and local SAM distributions at z = 0 plane for δ = 1 are shown in Figs. 1(a)–(c) as an example to demonstrate the spin topology. The zero points of Hertz potential correspond to the phase singularities (red and blue dots in Fig. 1(a)), which refer to the center of each Poynting vector vortex from Eq. (2). The generalized spin-momentum relationship in Eq. (3) indicates that the SAM originates from the vortices of the electromagnetic energy flow. According to right hand screw rule, only Sz is present in the center of Poynting vector vortex with sign determined by the rotation direction. The amplitude perturbation does not change the domains of each unit cell where nz = 0, while shifting the central of spin texture (Fig. 1(b)). In addition, the junction of Poynting vector vortices where P = 0 with nonzero Hertz potentials will be shifted by |cos-1(δ/2)|/kr along x axis (green dots in Figs. 1(a),(b)), resulting in a dislocation between the vortex junction and domain intersection of spin texture. This dislocation disunifies the transverse spin orientation around the vortex junction inside each unit cell, which is in analogy with an antivortex (see zoomed view of local spin orientation around the vortex junction in Fig. 1(c), that the spin vectors point oppositely along the two sides of vortex junction), hindering the formation of local Neel-type domain wall. Consequently, although the polarity of the spin texture in each unit cell is ± 1/2 as the local spin orientation rotates progressively from the central ‘up’ or ‘down’ state to the edge where nz = 0, the skyrmion number is not conserved, since the vorticity denoting the transverse spin orientations is not 1. The skyrmion number can be calculated through Eq. (4) as ± 0.1547, manifesting a destructive accumulation on vorticity.

 figure: Fig. 1.

Fig. 1. Spin topology under amplitude perturbation. (a) Poynting vector (arrows) and phase distribution of Hertz potential (background) at z = 0 plane for δ = 1. Red and blue dots denote the center of the anticlockwise (red arrows) and clockwise (blue arrows) Poynting vector vortices respectively, where the phase of the Hertz potential is singular. (b) Normalized local spin orientations n (arrows) with background representing nz for δ = 1. Red and blue dots denote the position where nz = ± 1. (c) Zoomed view of the yellow rectangle in (b), where the normalized local spin orientation is an antivortex. Green dots in (a)–(c) represent the junction of Poynting vector vortices where P = 0 with nonzero Hertz potentials. (d),(e) The same as in (a),(b) for δ = 3. No Poynting vector vortices are formed in (d) and the skyrmion number shrinks to zero in the spin textures in (e). (f) Skyrmion number (black) and figure of merit (red) dependences on δ (central unit cell of the spin lattice |x,y|<λr/4). The plot ranges in (a),(b) and (d),(e) are from −0.75 λr to 0.75 λr and the scale bar is λr/2 in (a), while the plot range in (c) is from −0.15 λr to 0.15 λr. The in plane wavevector kr of the evanescent field is set to 1.05 k.

Download Full Size | PDF

As a contrast, the Poynting vector and SAM distributions at z = 0 plane for δ = 3 are shown in Figs. 1(d),(e). From Eq. (7), the x component of Poynting vector is negative around the whole plane, cancelling the circulation of electromagnetic energy flow (Fig. 1(d)). And the spin texture exhibits directional flux with the core shifted out of the domain (Fig. 1(e)), preventing the formation of topological objects. A figure of merit (FOM) is defined as Q = exp(−20|N−0.5|) to quantify the stability under perturbation, where N is the skyrmion number. For tiny amplitude perturbation, N is close to 0.5 and the figure of merit is approaching 1. The skyrmion number (black solid line) in the central unit cell (|x,y|<λr/4) and FOM (red solid line) for different value of δ are calculated in Fig. 1(f), demonstrating the perturbation tolerance as |δ|< 0.1 to limit the fluctuation of skyrmion number within 10% (Q > 1/e). It is worth noting that for |δ|≥ 2, the skyrmion number shrinks to zero due to the absence of Poynting vector vortex.

As the amplitude perturbation δ is a spatial function, only periodic perturbation with period λr would preserve the topological spin texture. If the oscillation is along on the propagating direction of perturbated evanescent wave (x axis in our case as the perturbation is imposed on A2), the domain is kept as rectangle with the central of the spin texture shifted along y axis. While for a spatial perturbation with oscillation perpendicular to the propagating direction of evanescent wave, the domain for each unit cell would be reshaped to a curved surface.

As a phase perturbation φ is imposed in the presence of spin-orbit coupling (assume φ2 is changed to φ2 + φ in Eq. (1)), the Hertz potential, Poynting vector and SAM distributions can be calculated as

$${\Psi _p} = [{e^{i{k_r}x}} - {e^{i(\varphi - {k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi })}} - 2\sin ({k_r}y)]{e^{ - {k_z}z}}, $$
$$\scalebox{0.75}{$\displaystyle{{\mathbf P}_p} = \omega \varepsilon k_r^3{e^{ - 2{k_z}z}}\left( {\begin{array}{@{}c@{}} { - 2\sin ({k_r}y)[\cos({k_r}x) + \cos \varphi \cos({k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi )] - (\cos\varphi - 1)[1 - \cos({k_r}x + {k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi )]}\\ {2\cos ({k_r}y)[\sin ({k_r}x) + \sin({k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi )] + \sin\varphi [\cos({k_r}x + {k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi ) - 2\sin ({k_r}y)\cos({k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi ) - 1]}\\ 0 \end{array}} \right),$}$$
$$\scalebox{0.72}{$\displaystyle{{\mathbf S}_p} = \frac{{\varepsilon k_r^3{e^{ - 2{k_z}z}}}}{\omega }\left( {\begin{array}{@{}c@{}} {2{k_z}\cos ({k_r}y)[\sin ({k_r}x) + \sin({k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi )] + {k_z}\sin\varphi [\cos({k_r}x + {k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi ) - 2\sin ({k_r}y)\cos({k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi ) - 1]}\\ {2{k_z}\sin ({k_r}y)[\cos({k_r}x) + \cos \varphi \cos({k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi )] + {k_z}(\cos\varphi - 1)[1 - \cos({k_r}x + {k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi )]}\\ {2{k_r}\cos ({k_r}y)[\cos ({k_r}x) + \cos \varphi \cos({k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi )] - {k_r}\sin\varphi \sin ({k_r}x + {k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi )} \end{array}} \right),$}$$
yielding a complicated dependence on spatial position, where eφ = (cosφ, sinφ). By expanding the Hertz potential in Eq. (9), the zero points which indicate the cores of each Poynting vector vortex can be found in two sets. The first set is the solution of {sin[(krx + krr·eφφ)/2] = 0, sin(kry) = 0}, giving rise to the shift of vortex center along x axis and the deformation of domain walls for the spin texture (in the absence of perturbation, the zero points of Hertz potential are obtained as sin(krx) = sin(kry) = 0). While the second set is the solution of {cos[(krxkrr·eφ + φ)/2] = 0, sin[(krx + krr·eφφ)/2] ± sin(kry) = 0}, which is sparse to be negligible for small perturbation. Figure 2(a) shows the phase distribution of Hertz potential and Poynting vector at z = 0 plane for φ = π/18, with a shift of each Poynting vector vortex core along x axis (red and blue dots). The corresponding normalized local spin texture is shown in Fig. 2(b), demonstrating trapezoid domain walls for each unit cell. The domain intersections are adjacent to the junction of Poynting vector vortices (green dots), preserving the formation of Neel type domain walls and spin meron lattice (Fig. 2(c)), with the skyrmion number calculated as N = ± 0.491 for each unit cell. The periodic spin lattice with trapezoid domain walls for each unit cell is a correspondence of close packing of two-dimensional quadrangle, manifesting the metastability of square meron under small phase perturbation, that the skyrmion number is conserved while deforming the domains. As φ increases, the second set of vortex cores would stain the adjacent domain walls, which breaks the spin meron textures around. The normalized z component of SAM at z = 0 plane for φ = π/6 is shown in Fig. 2(d), exhibiting a collapse of domain walls periodically. The skyrmion number (black solid line) and figure of merit (red solid line) dependences on φ in central unit cell are calculated in Fig. 2(e), demonstrating a destruction of meron topology for large phase perturbation.

2.3 Numerical simulation

The spin topology of meron lattice under perturbation is numerically verified as follows. The spin–orbit coupling in evanescent optical vortex is realized by tightly focusing a radially polarized beam with helical wavefront (topological charge l = 1, wavelength λ = 633 nm) onto a thin silver film (thickness d = 50 nm, n1 = 0.135 + 3.99i) sandwiched by a silica substrate (n2 = 1.515) and air (n0 = 1), which provides an excitation of surface plasmon polaritons (SPPs) sustained at silver/air interface [41]. By modulating the incident beam with intensity mask comprised of C4 symmetry aperture, photonic spin meron lattice is formed due to interference of the SPPs, as shown in Fig. 3(a). According to the Richards-Wolf vectorial diffraction theory, the electromagnetic field near focus in air can be calculated as [42,43]

$$\scalebox{0.98}{$\displaystyle{\mathbf E}(r,\phi ,z) = A\sum\limits_{n = 1}^4 {\int_{{\theta _{\min }}}^{{\theta _{\max }}} {d\theta } \int_{{\varphi _n} - \frac{{\Delta {\varphi _n}}}{2}}^{{\varphi _n} + \frac{{\Delta {\varphi _n}}}{2}} {{t_p}(\theta )\sqrt {\cos \theta } \left( {\begin{array}{@{}c@{}} {{{i{k_{z0}}\cos \varphi } / {{k_0}}}}\\ {{{i{k_{z0}}\sin \varphi } / {{k_0}}}}\\ {{{ - {k_2}\sin \theta } / {{k_0}}}} \end{array}} \right)} {e^{il\varphi }}{e^{i{k_2}r\sin \theta \cos (\varphi - \phi )}}{e^{ - {k_{z0}}z}}\sin \theta d\varphi },$}$$
$${\mathbf H}(r,\phi ,z) = \frac{A}{{{Z_0}}}\sum\limits_{n = 1}^4 {\int_{{\theta _{\min }}}^{{\theta _{\max }}} {d\theta } \int_{{\varphi _n} - \frac{{\Delta {\varphi _n}}}{2}}^{{\varphi _n} + \frac{{\Delta {\varphi _n}}}{2}} {{t_p}(\theta )\sqrt {\cos \theta } \left( {\begin{array}{c} { - \sin \varphi }\\ {\cos \varphi }\\ 0 \end{array}} \right)} {e^{il\varphi }}{e^{i{k_2}r\sin \theta \cos (\varphi - \phi )}}{e^{ - {k_{z0}}z}}\sin \theta d\varphi }, $$
where A is a constant, φn = /2, Δφn = π/18 denoting the angle spread in each aperture for the transmitted beam, k0 is the wave vector in vacuum, ki = nik0 (i =0,1,2) is the wave vector in each medium, ${k_{zi}} = \sqrt {{{({k_2}\sin \theta )}^2} - k_i^2}$ is the longitudinal wave vector component, ${t_p} = \frac{4}{{(1 + \frac{{{\varepsilon _2}{k_{z1}}}}{{{\varepsilon _1}{k_{z2}}}})(1 + \frac{{{\varepsilon _1}{k_{z0}}}}{{{\varepsilon _0}{k_{z1}}}}){e^{{k_{z1}}d}} + (1 - \frac{{{\varepsilon _2}{k_{z1}}}}{{{\varepsilon _1}{k_{z2}}}})(1 - \frac{{{\varepsilon _1}{k_{z0}}}}{{{\varepsilon _0}{k_{z1}}}}){e^{ - {k_{z1}}d}}}}$ is the transmission coefficient with εi = ni2, θmax = sin-1 (NA/n2) is the maximum incident angle with NA = 1.49 denoting the numerical aperture, incident angles below π/4 are blocked by the mask to stabilize the generation of spin meron lattices that θmin = π/4, Z0 is the wave impendence in vacuum. The SAM of the electromagnetic fields can be calculated via S = Im(εE* × E + μH* × H)/4ω, manifesting the spin meron topology.

 figure: Fig. 2.

Fig. 2. Spin topology under phase perturbation. (a) Poynting vector (arrows) and phase distribution of Hertz potential (background) at z = 0 plane for φ = π/18. Red and blue dots denote the center of the anticlockwise and clockwise Poynting vector vortices respectively. (b) Normalized local spin orientations n (arrows) with background representing nz for φ = π/18. Red and blue dots denote the spin center in each unit cell. Green dots in (a),(b) represent the junction of Poynting vector vortices. (c) Side view of the normalized local spin vectors in (b), exhibiting a spin meron lattice. (d) Normalized z component of SAM for φ = π/6. (e) Skyrmion number (black) and figure of merit (red) dependences on φ (central unit cell of the spin lattice). The plot ranges in (a)–(c) are from −0.75 λr to 0.75 λr and the scale bar is λr/2 in (a). While the plot range in (d) is from −3 λr to 3 λr. The in plane wavevector kr of the evanescent field is set to 1.05 k.

Download Full Size | PDF

 figure: Fig. 3.

Fig. 3. Numerical simulations of spin topology by tightly focused vector beam. (a) Schematic diagram of the proposed setup. A radially polarized beam with helical wavefront (topological charge l = 1) modulated by an intensity mask comprised of C4 symmetry apertures is tightly focused on a thin silver film to excited SPPs. Due to interference of the SPPs in the presence of spin-orbit interaction, the spin meron lattice is formed. (b)–(e) Normalized Sz distributions and local spin vector orientations for (b) Δφ2 = π/9 (c) φ2 = π + π/18 (d) Δφ1 = Dφ2 = π/9 and (e) φ1 = π/2 − π/18, φ2 = π + π/18, demonstrating the modulation on spin textures under perturbation. Insets in (b)–(e) denote the opening angle and orientation of each aperture in the intensity mask.

Download Full Size | PDF

The electromagnetic field in Eq. (12) is a consequence of four interfering plasmonic fields propagating along different directions, which is in correspondence to the Hertz potential in Eq. (1). The opening angle Δφn determines the amplitude of SPPs along each direction, while the orientation of apertures φn represents the phase for each interfering SPP wave. By adjusting the opening angle or orientation of apertures, the C4 symmetry of aperture would be broken, which will induce amplitude and phase perturbation in the presence spin-orbit coupling. The normalized Sz distributions and local spin vector orientations for Δφ2 = π/9 corresponding to δ = 1 in Eqs. (6-8) are shown in Fig. 3(b), demonstrating a shifted spin center for each unit cell. On the other hand, by tuning the orientation of aperture as φ2 = π + π/18 (inset of Fig. 3(c)), the domain walls of spin texture are deformed into trapezoid, as shown in Fig. 3(c). These distributions correspond well to the theoretical predictions in Figs. 1(b) and 2(b), revealing spin topologies under respective perturbation. As multiple perturbations are imposed, the modulation on spin textures can be regarded as a superposition of individual variation, along with the interaction between perturbations. The spin texture for Δφ1 = D φ2 = π/9 is shown in Fig. 3(d), demonstrating a shift of spin center along both x and y axis. While the interaction between perturbations slightly compresses the domain wall. Similarly, under multiple phase perturbation as φ1 = π/2 − π/18 and φ2 = π + π/18, trapezoid domains with larger inner angle are formed in Fig. 3(e).

The aforementioned spin topology can be regarded as a consequence of imperfection of incident structured beam, where the amplitude and phase for the four interfering SPPs would be modified. That metastability of photonics meron lattice can also be verified through the interaction between circularly polarized light and fourfold metallic nanoslits [38]. By tuning the width or tilt angle of the nanoslits corresponding to amplitude and phase perturbation, same spin distribution can be obtained as in Figs. 3(b)–(d), which can be considered as an imperfection of nanostructures.

3. Conclusion

In conclusion, we have demonstrated the topological property of photonic spin meron lattice under both amplitude and phase perturbations, which are imposed by breaking the C4 symmetry in the presence of spin-orbit coupling, in correspondence to the imperfection of either impinging field or nanostructure. By employing the generalized spin-momentum relation, we show that amplitude perturbation would result in an amplitude-dependent shift of spin center in each unit cell, while keeping the location of domain walls. However, the perturbation tolerance is small to conserve the skyrmion number, since the dislocation between the vortex junction and domain intersection hinders the formation of local Neel-type domain walls and decreases the vorticity of spin texture. In contrast, phase perturbation leads to the deformation of domain walls with the skyrmion number merely changes, unveiling the metastability of photonic meron. In addition, the spin topology is verified through the focusing property of plasmonic vortices. These novel topological features of electromagnetic waves provide new insights on skyrmionic textures in the realm of optics, which may pave the way for spin-optics, sensing and metrology.

Funding

National Natural Science Foundation of China (12204309, 12274299, 92050202); Science and Technology Commission of Shanghai Municipality (19060502500); Shanghai Rising-Star Program (22YF1415200).

Disclosures

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

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. U. K. Rossler, A. N. Bogdanov, and C. Pfleiderer, “Spontaneous skyrmion ground states in magnetic metals,” Nature 442(7104), 797–801 (2006). [CrossRef]  

2. X. Z. Yu, Y. Onose, N. Kanazawa, J. H. Park, J. H. Han, Y. Matsui, N. Nagaosa, and Y. Tokura, “Real-space observation of a two-dimensional skyrmion crystal,” Nature 465(7300), 901–904 (2010). [CrossRef]  

3. S. Mühlbauer, B. Binz, F. Jonietz, C. Pfleiderer, A. Rosch, A. Neubauer, R. Georgii, and P. Böni, “Skyrmion lattice in a chiral magnet,” Science 323(5916), 915–919 (2009). [CrossRef]  

4. S. Banerjee, J. Rowland, O. Erten, and M. Randeria, “Enhanced Stability of Skyrmions in Two-Dimensional Chiral Magnets with Rashba Spin-Orbit Coupling,” Phys. Rev. X 4(3), 031045 (2014). [CrossRef]  

5. A. Fert, N. Reyren, and V. Cros, “Magnetic skyrmions: advances in physics and potential applications,” Nat. Rev. Mater. 2(7), 17031 (2017). [CrossRef]  

6. N. Nagaosa and Y. Tokura, “Topological properties and dynamics of magnetic skyrmions,” Nat. Nanotechnol. 8(12), 899–911 (2013). [CrossRef]  

7. B. Göbel, I. Mertig, and O. A. Tretiakov, “Beyond skyrmions: Review and perspectives of alternative magnetic quasiparticles,” Phys. Rep. 895, 1–28 (2021). [CrossRef]  

8. S. S. Parkin, M. Hayashi, and L. Thomas, “Magnetic domain-wall racetrack memory,” Science 320(5873), 190–194 (2008). [CrossRef]  

9. D. Foster, C. Kind, P. J. Ackerman, J.-S. B. Tai, M. R. Dennis, and I. I. Smalyukh, “Two-dimensional skyrmion bags in liquid crystals and ferromagnets,” Nat. Phys. 15(7), 655–659 (2019). [CrossRef]  

10. T. Kurumaji, T. Nakajima, M. Hirschberger, A. Kikkawa, Y. Yamasaki, H. Sagayama, H. Nakao, Y. Taguchi, T. H. Arima, and Y. Tokura, “Skyrmion lattice with a giant topological Hall effect in a frustrated triangular-lattice magnet,” Science 365(6456), 914–918 (2019). [CrossRef]  

11. A. J. Hess, G. Poy, J.-S. B. Tai, S. Žumer, and I. I. Smalyukh, “Control of Light by Topological Solitons in Soft Chiral Birefringent Media,” Phys. Rev. X 10(3), 031042 (2020). [CrossRef]  

12. 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(7), 650–654 (2019). [CrossRef]  

13. P. Shi, L. Du, and X. Yuan, “Strong spin–orbit interaction of photonic skyrmions at the general optical interface,” Nanophotonics 9(15), 4619–4628 (2020). [CrossRef]  

14. Q. Zhang, Z. Xie, L. Du, P. Shi, and X. Yuan, “Bloch-type photonic skyrmions in optical chiral multilayers,” Phys. Rev. Res. 3(2), 023109 (2021). [CrossRef]  

15. F. Meng, A. Yang, K. Du, F. Jia, X. Lei, T. Mei, L. Du, and X. Yuan, “Measuring the magnetic topological spin structure of light using an anapole probe,” Light: Sci. Appl. 11(1), 287 (2022). [CrossRef]  

16. Y. Shen, Q. Zhang, P. Shi, L. Du, A. V. Zayats, and X. Yuan, “Topological quasiparticles of light: Optical skyrmions and beyond,” arXiv, arXiv:2205.10329 (2022). [CrossRef]  

17. S. Tsesses, E. Ostrovsky, K. Cohen, B. Gjonaj, N. H. Lindner, and G. Bartal, “Optical skyrmion lattice in evanescent electromagnetic fields,” Science 361(6406), 993–996 (2018). [CrossRef]  

18. 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 368(6489), eaba6415 (2020). [CrossRef]  

19. 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(7), 10320–10328 (2020). [CrossRef]  

20. Z.-L. Deng, T. Shi, A. Krasnok, X. Li, and A. Alù, “Observation of localized magnetic plasmon skyrmions,” Nat. Commun. 13(1), 8 (2022). [CrossRef]  

21. S. Gao, F. C. Speirits, F. Castellucci, S. Franke-Arnold, S. M. Barnett, and J. B. Götte, “Paraxial skyrmionic beams,” Phys. Rev. A 102(5), 053513 (2020). [CrossRef]  

22. W. Lin, Y. Ota, Y. Arakawa, and S. Iwamoto, “Microcavity-based generation of full Poincaré beams with arbitrary skyrmion numbers,” Phys. Rev. Res. 3(2), 023055 (2021). [CrossRef]  

23. R. Gutiérrez-Cuevas and E. Pisanty, “Optical polarization skyrmionic fields in free space,” J. Opt. 23(2), 024004 (2021). [CrossRef]  

24. Y. Shen, E. C. Martínez, and C. Rosales-Guzmán, “Generation of Optical Skyrmions with Tunable Topological Textures,” ACS Photonics 9(1), 296–303 (2022). [CrossRef]  

25. P. Shi, L. Du, C. Li, A. V. Zayats, and X. Yuan, “Transverse spin dynamics in structured electromagnetic guided waves,” Proc. Natl. Acad. Sci. U.S.A. 118(6), e2018816118 (2021). [CrossRef]  

26. T. Van Mechelen and Z. Jacob, “Photonic Dirac monopoles and skyrmions: spin-1 quantization [Invited],” Opt. Mater. Express 9(1), 95 (2019). [CrossRef]  

27. K. Y. Bliokh, F. J. Rodríguez-Fortuño, F. Nori, and A. V. Zayats, “Spin–orbit interactions of light,” Nat. Photonics 9(12), 796–808 (2015). [CrossRef]  

28. L. Du, A. Yang, and X. Yuan, “Ultrasensitive displacement sensing method and device based on local spin characteristics,” (Google Patents, 2021).

29. X. Lei, L. Du, X. Yuan, and A. V. Zayats, “Optical spin–orbit coupling in the presence of magnetization: photonic skyrmion interaction with magnetic domains,” Nanophotonics 10(14), 3667–3675 (2021). [CrossRef]  

30. A. Actor, “Classical solutions of SU(2)Yang—Mills theories,” Rev. Modern Phys. 51(3), 461–525 (1979). [CrossRef]  

31. K. Moon, H. Mori, K. Yang, S. M. Girvin, A. H. MacDonald, L. Zheng, D. Yoshioka, and S. C. Zhang, “Spontaneous interlayer coherence in double-layer quantum Hall systems: Charged vortices and Kosterlitz-Thouless phase transitions,” Phys. Rev. B 51(8), 5138–5170 (1995). [CrossRef]  

32. S. Wintz, C. Bunce, A. Neudert, M. Korner, T. Strache, M. Buhl, A. Erbe, S. Gemming, J. Raabe, C. Quitmann, and J. Fassbender, “Topology and origin of effective spin meron pairs in ferromagnetic multilayer elements,” Phys. Rev. Lett. 110(17), 177201 (2013). [CrossRef]  

33. S.-Z. Lin, A. Saxena, and C. D. Batista, “Skyrmion fractionalization and merons in chiral magnets with easy-plane anisotropy,” Phys. Rev. B 91(22), 224407 (2015). [CrossRef]  

34. Y. A. Kharkov, O. P. Sushkov, and M. Mostovoy, “Bound States of Skyrmions and Merons near the Lifshitz Point,” Phys. Rev. Lett. 119(20), 207201 (2017). [CrossRef]  

35. X. Z. Yu, W. Koshibae, Y. Tokunaga, K. Shibata, Y. Taguchi, N. Nagaosa, and Y. Tokura, “Transformation between meron and skyrmion topological spin textures in a chiral magnet,” Nature 564(7734), 95–98 (2018). [CrossRef]  

36. N. Gao, S. Je, M. Im, J. W. Choi, M. Yang, Q. Li, T. Y. Wang, S. Lee, H. Han, K. Lee, W. Chao, C. Hwang, J. Li, and Z. Q. Qiu, “Creation and annihilation of topological meron pairs in in-plane magnetized films,” Nat. Commun. 10(1), 5603 (2019). [CrossRef]  

37. 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(23), 237403 (2021). [CrossRef]  

38. 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(4), 041413 (2021). [CrossRef]  

39. M. Król, H. Sigurdsson, K. Rechcińska, P. Oliwa, K. Tyszka, W. Bardyszewski, A. Opala, M. Matuszewski, P. Morawiak, R. Mazur, W. Piecek, P. Kula, P. G. Lagoudakis, B. Piętka, and J. Szczytko, “Observation of second-order meron polarization textures in optical microcavities,” Optica 8(2), 255 (2021). [CrossRef]  

40. Y. Shen, “Topological bimeronic beams,” Opt. Lett. 46(15), 3737–3740 (2021). [CrossRef]  

41. A. V. Zayats, I. I. Smolyaninov, and A. A. Maradudin, “Nano-optics of surface plasmon polaritons,” Phys. Rep. 408(3-4), 131–314 (2005). [CrossRef]  

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

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

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (3)

Fig. 1.
Fig. 1. Spin topology under amplitude perturbation. (a) Poynting vector (arrows) and phase distribution of Hertz potential (background) at z = 0 plane for δ = 1. Red and blue dots denote the center of the anticlockwise (red arrows) and clockwise (blue arrows) Poynting vector vortices respectively, where the phase of the Hertz potential is singular. (b) Normalized local spin orientations n (arrows) with background representing nz for δ = 1. Red and blue dots denote the position where nz = ± 1. (c) Zoomed view of the yellow rectangle in (b), where the normalized local spin orientation is an antivortex. Green dots in (a)–(c) represent the junction of Poynting vector vortices where P = 0 with nonzero Hertz potentials. (d),(e) The same as in (a),(b) for δ = 3. No Poynting vector vortices are formed in (d) and the skyrmion number shrinks to zero in the spin textures in (e). (f) Skyrmion number (black) and figure of merit (red) dependences on δ (central unit cell of the spin lattice |x,y|<λr/4). The plot ranges in (a),(b) and (d),(e) are from −0.75 λr to 0.75 λr and the scale bar is λr/2 in (a), while the plot range in (c) is from −0.15 λr to 0.15 λr. The in plane wavevector kr of the evanescent field is set to 1.05 k.
Fig. 2.
Fig. 2. Spin topology under phase perturbation. (a) Poynting vector (arrows) and phase distribution of Hertz potential (background) at z = 0 plane for φ = π/18. Red and blue dots denote the center of the anticlockwise and clockwise Poynting vector vortices respectively. (b) Normalized local spin orientations n (arrows) with background representing nz for φ = π/18. Red and blue dots denote the spin center in each unit cell. Green dots in (a),(b) represent the junction of Poynting vector vortices. (c) Side view of the normalized local spin vectors in (b), exhibiting a spin meron lattice. (d) Normalized z component of SAM for φ = π/6. (e) Skyrmion number (black) and figure of merit (red) dependences on φ (central unit cell of the spin lattice). The plot ranges in (a)–(c) are from −0.75 λr to 0.75 λr and the scale bar is λr/2 in (a). While the plot range in (d) is from −3 λr to 3 λr. The in plane wavevector kr of the evanescent field is set to 1.05 k.
Fig. 3.
Fig. 3. Numerical simulations of spin topology by tightly focused vector beam. (a) Schematic diagram of the proposed setup. A radially polarized beam with helical wavefront (topological charge l = 1) modulated by an intensity mask comprised of C4 symmetry apertures is tightly focused on a thin silver film to excited SPPs. Due to interference of the SPPs in the presence of spin-orbit interaction, the spin meron lattice is formed. (b)–(e) Normalized Sz distributions and local spin vector orientations for (b) Δφ2 = π/9 (c) φ2 = π + π/18 (d) Δφ1 = Dφ2 = π/9 and (e) φ1 = π/2 − π/18, φ2 = π + π/18, demonstrating the modulation on spin textures under perturbation. Insets in (b)–(e) denote the opening angle and orientation of each aperture in the intensity mask.

Equations (13)

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

$$\Psi = \sum\limits_{n = 1}^4 {{A_n}{e^{il{\varphi _n}}}{e^{i{k_r}{\mathbf r} \cdot {{\mathbf e}_n}}}{e^{ - {k_z}z}}}, $$
$${\mathbf P} = \frac{{\omega \varepsilon k_r^2}}{2}{\mathop{\rm Im}\nolimits} ({{\Psi ^ \ast }\nabla \Psi } ),$$
$${\mathbf S} = \frac{1}{{2{\omega ^2}}}\nabla \times {\mathbf P} = \frac{{\varepsilon k_r^2}}{{4\omega }}{\mathop{\rm Im}\nolimits} (\nabla {\Psi ^ \ast } \times \nabla \Psi ) = \frac{{2\varepsilon k_r^3{e^{ - 2{k_z}z}}}}{\omega }\left( {\begin{array}{*{20}{c}} {{k_z}\sin ({k_r}x)\cos ({k_r}y)}\\ {{k_z}\cos ({k_r}x)\sin ({k_r}y)}\\ {{k_r}\cos ({k_r}x)\cos ({k_r}y)} \end{array}} \right). $$
$$N = \frac{1}{{4\pi }}\int\!\!\!\int_\Sigma {{\mathbf n} \cdot ({{\partial_x}{\mathbf n} \times {\partial_y}{\mathbf n}} )} dxdy. $$
$${\mathbf S} \propto \left\langle {\nabla \Psi } \right|\times i|{\nabla \Psi } \rangle + 2\left\langle {\nabla \Phi } \right|\times i|{\nabla \Psi } \rangle + \left\langle {\nabla \Phi } \right|\times i|{\nabla \Phi } \rangle. $$
$${\Psi _a} = 2[i\sin ({k_r}x) - \sin ({k_r}y) - \frac{\delta }{2}{e^{ - i{k_r}x}}]{e^{ - {k_z}z}}, $$
$${{\mathbf P}_a} = 2\omega \varepsilon k_r^3{e^{ - 2{k_z}z}}\left( {\begin{array}{*{20}{c}} { - (1 + {\delta / 2})[\cos ({k_r}x)\sin ({k_r}y) + {\delta / 2}]}\\ {(1 + {\delta / 2})\sin ({k_r}x)\cos ({k_r}y)}\\ 0 \end{array}} \right), $$
$${{\mathbf S}_a} = \frac{{2\varepsilon k_r^3{e^{ - 2{k_z}z}}}}{\omega }\left( {\begin{array}{*{20}{c}} {{k_z}(1 + {\delta / 2})\sin ({k_r}x)\cos ({k_r}y)}\\ {{k_z}(1 + {\delta / 2})[\cos ({k_r}x)\sin ({k_r}y) + {\delta / 2}]}\\ {{k_r}(1 + {\delta / 2})\cos ({k_r}x)\cos ({k_r}y)} \end{array}} \right). $$
$${\Psi _p} = [{e^{i{k_r}x}} - {e^{i(\varphi - {k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi })}} - 2\sin ({k_r}y)]{e^{ - {k_z}z}}, $$
$$\scalebox{0.75}{$\displaystyle{{\mathbf P}_p} = \omega \varepsilon k_r^3{e^{ - 2{k_z}z}}\left( {\begin{array}{@{}c@{}} { - 2\sin ({k_r}y)[\cos({k_r}x) + \cos \varphi \cos({k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi )] - (\cos\varphi - 1)[1 - \cos({k_r}x + {k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi )]}\\ {2\cos ({k_r}y)[\sin ({k_r}x) + \sin({k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi )] + \sin\varphi [\cos({k_r}x + {k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi ) - 2\sin ({k_r}y)\cos({k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi ) - 1]}\\ 0 \end{array}} \right),$}$$
$$\scalebox{0.72}{$\displaystyle{{\mathbf S}_p} = \frac{{\varepsilon k_r^3{e^{ - 2{k_z}z}}}}{\omega }\left( {\begin{array}{@{}c@{}} {2{k_z}\cos ({k_r}y)[\sin ({k_r}x) + \sin({k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi )] + {k_z}\sin\varphi [\cos({k_r}x + {k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi ) - 2\sin ({k_r}y)\cos({k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi ) - 1]}\\ {2{k_z}\sin ({k_r}y)[\cos({k_r}x) + \cos \varphi \cos({k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi )] + {k_z}(\cos\varphi - 1)[1 - \cos({k_r}x + {k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi )]}\\ {2{k_r}\cos ({k_r}y)[\cos ({k_r}x) + \cos \varphi \cos({k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi )] - {k_r}\sin\varphi \sin ({k_r}x + {k_r}{\mathbf r} \cdot {{\mathbf e}_\varphi } - \varphi )} \end{array}} \right),$}$$
$$\scalebox{0.98}{$\displaystyle{\mathbf E}(r,\phi ,z) = A\sum\limits_{n = 1}^4 {\int_{{\theta _{\min }}}^{{\theta _{\max }}} {d\theta } \int_{{\varphi _n} - \frac{{\Delta {\varphi _n}}}{2}}^{{\varphi _n} + \frac{{\Delta {\varphi _n}}}{2}} {{t_p}(\theta )\sqrt {\cos \theta } \left( {\begin{array}{@{}c@{}} {{{i{k_{z0}}\cos \varphi } / {{k_0}}}}\\ {{{i{k_{z0}}\sin \varphi } / {{k_0}}}}\\ {{{ - {k_2}\sin \theta } / {{k_0}}}} \end{array}} \right)} {e^{il\varphi }}{e^{i{k_2}r\sin \theta \cos (\varphi - \phi )}}{e^{ - {k_{z0}}z}}\sin \theta d\varphi },$}$$
$${\mathbf H}(r,\phi ,z) = \frac{A}{{{Z_0}}}\sum\limits_{n = 1}^4 {\int_{{\theta _{\min }}}^{{\theta _{\max }}} {d\theta } \int_{{\varphi _n} - \frac{{\Delta {\varphi _n}}}{2}}^{{\varphi _n} + \frac{{\Delta {\varphi _n}}}{2}} {{t_p}(\theta )\sqrt {\cos \theta } \left( {\begin{array}{c} { - \sin \varphi }\\ {\cos \varphi }\\ 0 \end{array}} \right)} {e^{il\varphi }}{e^{i{k_2}r\sin \theta \cos (\varphi - \phi )}}{e^{ - {k_{z0}}z}}\sin \theta d\varphi }, $$
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.