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

Symmetry-breaking enabled topological phase transitions in spin-orbit optics

Open Access Open Access

Abstract

The topological phase transitions (TPT) of light refers to a topological evolution from one type of spin-orbit interaction to another, which has been recently found in beam scattering at optical interfaces and propagation in uniaxial crystals. In this work, the focusing of off-axis and partially masked circular-polarization Gaussian beams are investigated by using of a full-wave theory. Moreover, two different types of spin-orbit interactions (i.e., spin-dependent vortex generation and photonic spin-Hall effect) in the focusing system are unified from the perspective of TPT. It is demonstrated that as the off-axis distance or the masked area increases, a TPT phenomenon in the focused optical field takes place, evolving from the spin-dependent vortex generation to the spin-Hall shift of the beam centroids. The intrinsic mechanism is attributed to the cylindrical symmetry-breaking of the system. This symmetry-breaking induced TPT based on the method of vortex mode decomposition is further examined. The main difference between the TPT phenomenon observed here and that trigged by oblique incidence at optical interfaces or oblique propagation in uniaxial crystals is also uncovered. Our findings provide fruitful insights for understanding the spin-orbit interactions in optics, providing an opportunity for unifying the TPT phenomena in various spin-orbit photonics systems.

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

1. Introduction

Light can carry spin angular momentum (SAM) and orbital angular momentum (OAM). The latter can be further divided into intrinsic OAM (IOAM) and extrinsic OAM (EOAM) [15]. The SAM, IOAM and EOAM are related to the circular polarization, optical vortices within the beam, and beam trajectories with respect to a given reference point or axis, respectively. The interaction and coupling between the SAM and the two types of OAMs are known as the two main spin-orbit interactions (SOI) in optics [36], i.e., spin-dependent vortex generation and photonic spin-Hall effect (PSHE). The SOIs are inherent in many basic optical processes, such as reflection, refraction, propagation, focusing, and imaging [416], and have attracted wide attention in recent years, playing an important role in the fields of nanophotonics [1719], topological photonics [2022], and plasmonics [2325].

A transition between these two types of SOIs can take place in some optical systems, which is called the topological phase transitions (TPT) [26,27]. The TPT provides a new perspective for unifying the two types of SOIs and could be leveraged for the generation and manipulation of structured optical fields. The TPT phenomenon in spin-orbit photonics has been demonstrated in some systems, such as light beam reflection and refraction at optical interfaces [2629] and light beam propagation in uniaxial crystals [30,31]. When a light beam illuminates an optical interface normally or propagates along the optical axis in a uniaxial crystal, an optical vortex is generated. As the incident angle is gradually increased, the vortex is evolved to an off-axis one, and then to non-vortex states with a transverse shift (i.e., the PSHE), manifested by an incident-angle-driven TPT.

In focusing systems, the coupling between the SAM and the two types of OAMs, such as the spin-dependent vortex generation and the PSHE, may also occur. Particularly, when a circular-polarization beam is focused, a fraction of the transverse focused beam is spin-flip and acquires a vortex phase with a topological charge of ±2, whereas the longitudinal mode (z-polarization component of the light field) gets a vortex phase with a topological charge of ±1 [10,14,15,3236]. This indicates a SAM-to-IOAM conversion in the focusing system. When an asymmetrical circular-polarization beam is tightly focused, the beam centroid is shifted against the focal point. This effect is regarded as a manifestation of the PSHE, indicating a SAM-to-EOAM conversion in the focusing process [3744]. Since there are two different types of SOIs in the focusing system, one is natural to wonder whether they can also be unified from the viewpoint of the TPT. Moreover, the intrinsic connection between the SOIs in the focusing system and at the optical interface is still obscure.

In this work, based on a full-wave theory, the focusing of off-axis and partially masked circular-polarization Gaussian beams is systematically examined, respectively. With the increases of the off-axis distance or masked area, the focused field exhibits a transition from the spin-dependent vortex generation to the PSHE, i.e., two different types of SOIs can be unified by the TPT. The underlying mechanism is attributed to the symmetry-breaking of the focusing system with respect to the central axis. Such type of symmetry-breaking enabled TPT is further explored from the perspective of vortex mode decomposition and is compared with the beam scattering at optical interfaces or propagation in uniaxial crystals. From our work, it is suggested that the breaking of the cylindrical symmetry can lead to the TPT between two different types of SOIs, providing a unique opportunity for a unified understanding of the TPT phenomenon in spin-orbit photonics systems.

2. Full-wave theory of beam focusing

A full-wave theory is first employed to describe the SOI when the beam is focused. The E-field distribution of the light beam on the reference plane under circular polarization basis can be written as follows (superscripts a={i,f} represent the incidence or focusing, respectively):

$${\mathbf{E}^a}(\mathbf{r}_ \bot ^a) = \int {{d^2}\mathbf{k}_ \bot ^a\,\exp (i{\mathbf{k}^a} \cdot {\mathbf{r}^a})\sum\limits_{\xi ={+} , - ,z} {U_\xi ^a({\mathbf{k}^a}){{\hat{\mathbf{V}}}_\xi }({\mathbf{k}^a})} } ,$$
where $U_\xi ^a({\mathbf{k}^a})$ denotes the transverse and the longitudinal patterns of the ath beam, ${\hat{\mathbf{V}}_ \pm }({\mathbf{k}^a}) = (\hat{\mathbf{x}} \pm i\hat{\mathbf{y}})/\sqrt 2 $ refers to the unit vector of circular polarizations in the transverse direction of the beam, and ${\hat{\mathbf{V}}_z}({\mathbf{k}^a}) = \hat{\mathbf{z}}$. A 3 × 3 matrix $\hat{M}$ can be used to relate the focused field with the initial field [12,14]:
$$\hat{M} = \frac{1}{2}\left[ {\begin{array}{ccc} {\cos {\theta^i}\cos {\theta^f} + 1}&{\exp ( - 2i\varphi )(\cos {\theta^i}\cos {\theta^f} - 1)}&{ - \sqrt 2 \exp ( - i\varphi )\cos {\theta^f}\sin {\theta^i}}\\ {\exp (2i\varphi )(\cos {\theta^i}\cos {\theta^f} - 1)}&{\cos {\theta^i}\cos {\theta^f}}&{ - \sqrt 2 \exp (i\varphi )\cos {\theta^f}\sin {\theta^i}}\\ { - \sqrt 2 \exp (i\varphi )\cos {\theta^i}\sin {\theta^f}}&{ - \sqrt 2 \exp ( - i\varphi )\cos {\theta^i}\sin {\theta^f}}&{2\sin {\theta^i}\sin {\theta^f}} \end{array}} \right]$$
where ${\theta ^f}$ represents the angle between the wave vector of any focused plane wave and the z-axis, and $\varphi $ signifies the azimuth angle of the plane wave component in the spherical coordinate system.

It is assumed that the incident beam is a left-handed circular-polarization beam, i.e., $U_\textrm{ + }^i({\mathbf{k}^i}) = \exp [ - {(k_ \bot ^iw)^2}/4]$, $U_ - ^i({\mathbf{k}^i}) \equiv 0$, and $U_z^i({\mathbf{k}^i}) \equiv 0$, where w denotes the half-width of the Gaussian beam waist. Under the normal incidence and paraxial approximation (${\theta ^i} \approx 0$), the following expression can be derived:

$$\left[ {\begin{array}{c} {U_ +^f}\\ {U_ -^f}\\ {U_z^f} \end{array}} \right] = \hat{M}\left[ {\begin{array}{c} {U_ +^i}\\ {U_ -^i}\\ {U_z^i} \end{array}} \right] = \frac{1}{2}\left[ {\begin{array}{c} {(\cos {\theta^f} + 1)}\\ {\exp (2i\varphi )(\cos {\theta^f} - 1)}\\ { - \sqrt 2 \exp (i\varphi )\sin {\theta^f}} \end{array}} \right]U_\textrm{ + }^i.$$

As can be observed, a part of the transverse field retains its circular polarization handedness (called the normal mode), another part undergoes spin-flip transformation and carries a vortex phase with a topological charge of 2 (called the abnormal mode), while the longitudinal mode carries a vortex phase with a topological charge of 1. For the incidence of the right-handed circular-polarization beam, the abnormal and longitudinal modes exhibit vortex phases with opposite topological charges of -2 and -1, respectively.

Following Richards and Wolf's method [45], the E-field distribution of the focused beam is obtained by Debye integral:

$$E_\xi ^f(\rho ,\phi ,z) ={-} \frac{{ikL\,\exp ( - ikL)}}{{2\pi }}\int\limits_{\theta _1^f}^{\theta _2^f} {\int\limits_{{\varphi _1}}^{{\varphi _2}} {U_\xi ^f} } \exp (ikz\cos {\theta ^f})\exp [ik\rho \sin {\theta ^f}\cos (\varphi \textrm{ - }\phi )]\sin {\theta ^f}d\varphi d{\theta ^f},$$
where L signifies the focal distance, k denotes the wave number, ρ and φ represent the radial coordinate and azimuthal angle in the focal area, respectively. When the integral range of ${\theta ^f}$ and $\varphi$ is 0 to $\theta _{\max }^f$ (i.e., the maximum exit angle) and 0 to 2π, respectively, the E-field distribution of the incident beam is cylindrically symmetric about the z-axis. If other values for the integral range of ${\theta ^f}$ and $\varphi$ are used (e.g., $\theta _1^f$ to $\theta _2^f$, ${\varphi _1}$ to ${\varphi _2}$), the E-field distribution of the incident light beam is no longer cylindrically symmetric about the z-axis.

Equation (4) provides the Debye integral in cylindrical coordinates. By substituting ${\theta ^f} = {\tan ^{ - 1}}[{({u^2} + {v^2})^{{\raise0.5ex\hbox{$\scriptstyle 1$}\kern-0.1em/\kern-0.15em\lower0.25ex\hbox{$\scriptstyle 2$}}}}/L]$, $\varphi = {\tan ^{ - 1}}(v/u)$, $\rho = {({x^2} + {y^2})^{{\raise0.5ex\hbox{$\scriptstyle 1$}\kern-0.1em/\kern-0.15em\lower0.25ex\hbox{$\scriptstyle 2$}}}}$, and $\phi = {\tan ^{ - 1}}(y/x)$ into Eq. (4), the integral in Cartesian coordinates can be obtained, where (u, v) are the Cartesian coordinates of the focusing system’s output pupil, and (x, y) are Cartesian coordinates in the focal region.

3. Symmetry-breaking induced TPT in focusing

The focusing of three types of circular-polarization beams, i.e., the off-axis, sector-masked, and horizontally masked beams, is considered here. The intensity patterns and beam shifts after focusing are calculated by selecting the following parameters: w = 40λ, L = 100λ, and k = 2π/λ, where λ is the working wavelength.

3.1 Off-axis beam focusing

Figure 1(a) schematically illustrates the focusing effect of an off-axis beam by a thin lens. The incident beam is a left-handed circular-polarization Gaussian beam whose centroid is located at a distance d from the z-axis [1st column in Fig. 1(b)]. The focusing field has three components: a normal mode $E_ + ^f$ with its spin handedness the same as the incident one, a spin-flip abnormal mode $E_ - ^f$, and a longitudinal mode $E_z^f$ (z-polarization component), as depicted in the 2nd to 4th column in Fig. 1(b). When the off-axis distance is d = 0, the abnormal mode exhibits a vortex phase with a topological charge of 2 and the longitudinal mode has a vortex phase with a topological charge being 1 [10]. As d is increased, the intensity patterns of the abnormal and longitudinal modes significantly deform, which gradually evolves from a perfect vortex state to crescent-shaped vortex states, and finally to non-vortex states with spin-Hall shifts. This is manifested as a TPT process from one type of SOI to another. Although the incident beam is off-axis, no beam shift and intensity deformation in the normal mode take place, as the vortex and beam shift only occur in the abnormal and longitudinal modes orthogonal to the incident one.

 figure: Fig. 1.

Fig. 1. The TPT of light in off-axis beam focusing. (a) Schematic of the problem studied. (b) The light intensity patterns of the normal, abnormal, and longitudinal modes after focusing by a thin lens when d = 0, 0.25w, 0.75w, and 1.25w. (c) The dependence of the beam centroid shift Δx of the abnormal and longitudinal modes on the off-axis distance d.

Download Full Size | PDF

The shifts of the beam centroid positions can be visually observed from the third and fourth columns of Fig. 1(b), which are manifested as the PSHE. The direction of the beam shift is always orthogonal to the off-axis direction of the incident light. That is, the off-axis direction of the incident light is along the y-axis, then the directions of the beam shifts is along the x-axis, and vice versa. This is due to the wavefront rotation of the vortex beam upon propagation, which is usually attributed to the Gouy phase [37,46]. To further understand the TPT induced by the off-axis distance, the following equation is used to calculate the shift Δx of beam centroid [26]:

$$\Delta x = \frac{{\int\!\!\!\int {x|E_\xi ^f{|^2}dxdy} }}{{\int\!\!\!\int {|E_\xi ^f{|^2}dxdy} }}.$$

The shift Δx can be controlled by changing the off-axis distance d. The dependencies of Δx of the abnormal and longitudinal modes on the off-axis distance d after focusing are given in Fig. 1(c). With the increases of d, the values of Δx of the abnormal and longitudinal modes increase from 0, then decrease after reaching a maximum value. This TPT is very similar to that of a beam scattering by a sharp interface, where the beam shift has a peak in the transition region [26]. This is a remarkable feature of this TPT, which originates from the competition between intrinsic OAM and extrinsic OAM.

The beam shifts of the abnormal and longitudinal modes exhibit the same evolutionary behavior. However, the shift magnitude of the abnormal mode is obviously larger than that of the longitudinal mode since the two modes possessed vortex phases with different topological charges, which have different beam sizes. It is also worth noting that the beam shift directions of the right-handed and the left-handed circular-polarization incident light are opposite, as can be ascertained from Fig. 2 [37,4649].

 figure: Fig. 2.

Fig. 2. The TPT of light under the incidence of left- and right-handed circular-polarization light. The first and second rows: intensity patterns of the incident, normal, abnormal, and longitudinal modes when the incident beam is left and right-handed circular-polarization, respectively.

Download Full Size | PDF

3.2 Sector-masked beam focusing

Another kind of asymmetric beam is taken into account, i.e., the sector-masked beam. Figure 3(a) schematically shows the focusing effect of a sector-masked beam by a thin lens. The incident beam is a left-handed circular-polarization Gaussian beam, a sector of which is masked by a sharp obstacle at an angle α. The obstacle is symmetric along the y-axis, and its center locates at the z-axis.

 figure: Fig. 3.

Fig. 3. The TPT of light in sector-masked beam focusing. (a) Schematic of the problem studied. (b) The light intensity patterns of the normal, abnormal, and longitudinal modes after focusing by a thin lens when α=0, π/2, π, and 3π/2. (c) The dependence of the beam centroid shift Δx of the abnormal and longitudinal modes on the masked angle α.

Download Full Size | PDF

When the incident beam is not masked (i.e., α=0), the focused field is actually exactly the same as that of the on-axis focusing [1st row of Fig. 3(b)]. When the mask angle is increased, the intensity patterns of the abnormal and longitudinal modes are significantly deformed [50]. Furthermore, their centroids are displaced away from the coordinate origin, manifesting as a topological transition from the vortex generation to the PSHE. Although the light intensity patterns of the normal mode are also deformed due to the partial masking of the incident beam, no beam shift is detected in the normal mode.

During the process of the TPT, the beam centroid is shifted to the x-axis direction, manifested as the PSHE. To further understand the influence of the masked angle α of the incident beam on the TPT, Eq. (5) is also used to calculate the beam shifts Δx. The dependences of the abnormal and longitudinal modes Δx on the masked angle are displayed in Fig. 3(c). Thus, the shift Δx can be controlled by adjusting the masked angle α. By increasing the value of α, the values of Δx of the abnormal and longitudinal modes increase accordingly. It's worth noting that since the field distribution in the focal region is entirely determined by the far-field, the effects of diffraction are not considered.

3.3 Horizontally masked beam focusing

The focusing effect of a left-handed circular-polarization Gaussian beam that is horizontally masked by a sharp obstacle is also further considered [Fig. 4(a)]. The edge of this sharp obstacle is parallel to the x-axis, and the masked height h represents the distance from the edge of the sharp obstacle to -w on the y-axis.

 figure: Fig. 4.

Fig. 4. The TPT of light in horizontally masked beam focusing. (a) Schematic of the problem studied. (b) The light intensity patterns of the normal, abnormal, and longitudinal modes after focusing by a thin lens when h = 0, 0.5w, w, and 1.5w. (c) The dependence of the beam centroid shift Δx of the abnormal and longitudinal modes on the masked height h.

Download Full Size | PDF

The first row of Fig. 4(b) indicates the light intensity patterns when the beam is not masked (i.e., h = 0), which is the same as that when the incident beam is on-axis or not masked by the sector-masked sharp obstacle. The masked height h is gradually increased, and the first to fourth rows of Fig. 4(b) show the light intensity patterns after focusing when h = 0, 0.5w, w, and 1.5w are used, respectively. The light intensity patterns of the horizontally masked beam are similar to those of the sector-masked beam. The TPT phenomenon exists in the abnormal and longitudinal modes when the horizontally masked beam is focused. The spin-Hall shift Δx increases with h, as can be ascertained from Fig. 4(c). On top of that, the dependency of Δx on the masked height h is very similar to that of Δx on the masked angle α in Fig. 3(c).

3.4 Comparison and discussion

From the above analysis, the two types of SOIs in the focusing system, namely the spin-dependent vortex generation and the PSHE, are unified from the viewpoint of the TPT. When the focusing system has cylindrical symmetry, both the abnormal and longitudinal modes acquire vortex phase with topological charges of ±2 and ±1, respectively. However, when the cylindrical symmetry of the system is broken, such as when the incident beam is off-axis or partially obscured, the abnormal and longitudinal modes in the focused beam exhibit a TPT from the spin-dependent vortex generation to the PSHE. This TPT phenomenon is very akin to that produced by the light beam scattering at optical interfaces [2629] or propagating in uniaxial crystals [30,31]. When a beam is scattered by a sharp interface, the scattered light exhibits a transition from the spin-dependent vortex generation at normal incidence to the PSHE at oblique incidence. A beam propagating in a uniaxial crystal has a similar TPT as the angle between the beam propagation direction and the optical axis gradually increases. Thus, a common origin of the TPT phenomenon is found, which is associated with the breaking of the system’s cylindrical symmetry. In other words, the breaking of cylindrical symmetry will drive a spin-orbit optical systems manifesting from one type of SOI to another. From this point of view, we can describe the TPT in spin-orbit photonics in a unified form from the perspective of symmetry breaking.

4. Understanding the TPT based on vortex mode decomposition

Vortex mode decomposition provides an alternative perspective for understanding the two types of SOIs and is considered an effective method for describing the TPT phenomenon. The vortex mode contents of the abnormal mode are further examined here. The longitudinal mode has a similar behavior and will not be discussed. The abnormal mode can be decomposed into a series of Laguerre–Gaussian (LG) modes. More specifically, $C_p^m$ denotes the normalized weight coefficient of the mth-order azimuthal LG mode with radial index p [27,51,52]:

$$C_p^m = \frac{{\int {\int {E_{LG_p^m}^ \ast E_ - ^fdxdy} } }}{{{{(\int {\int {\textrm{|}{E_{LG_p^m}}{\textrm{|}^2}} } dxdy)}^{1/2}}{{(\int {\int {\textrm{|}E_ - ^f{\textrm{|}^2}} } dxdy)}^{1/2}}}},$$
where ${E_{LG_p^m}}$ stands for the E-field distribution of the LG beam.

Figures 5(a)–5(c) show the normalized vortex mode spectra of the focused optical field under the illumination of three types of left-handed circular-polarization beams, i.e., the off-axis, sector-masked, and horizontally masked beams, respectively. When the incident beam is cylindrically symmetric, the abnormal mode has only a single vortex mode with a topological charge of m = 2 [1st column in Fig. 5]. Multiple vortex modes appear as the symmetry of the incident beam is broken (i.e., 2nd column to 4th column in Fig. 5), competing and superposing with m = 2 vortex mode. Then, the cylindrical symmetry breaking causes the transformation of the vortex mode from a single mode to multimode, which significantly deforms the intensity patterns. Unlike the focusing system in this work, the TPT process of the previously reported two systems [26,27,30] can be described by the competition and superposition of three vortex modes with topological charges of only m = 2, 1, and 0 (or -2, -1, and 0) [27].

 figure: Fig. 5.

Fig. 5. Normalized vortex mode spectra of the focusing field of the (a) off-axis beam, (b) sector-masked beam, and (c) horizontally masked beam.

Download Full Size | PDF

It can also be found that the envelopes of the vortex mode spectra in Fig. 5(a) possess a Gaussian distribution when the cylindrically symmetric incident beams are focused [53]. The envelopes of the spectra in Fig. 5(b) and 5(c) can be described by using sinc functions, which is similar to the single slit diffraction of light [54]. In the single slit diffraction, a restriction of position due to a single slit causes a sinc envelope of the light intensity in the transverse linear momentum or the far field position. Here, the restrictions of the angular position and the horizontal position by the mask cause the sinc envelopes of the vortex mode spectra in the focused optical field. Unlike the vortex spectrum of the TPT produced when the beam is scattered at the optical interface or propagates in the uniaxial crystal, which has only three components [27], the vortex spectra here possess more modes and the shape of the spectra is symmetric about m = 2. The vortex mode spectra of the longitudinal modes exhibit similar behaviors to those of the abnormal modes. It is also worth noting that the vortex mode spectra of the right-handed and the left-handed circular-polarization beams are symmetric about m = 0.

5. Conclusions

The focusing of three asymmetric beams with the full-wave theory has been thoroughly investigated in this work, and a unified description of the two different types of SOIs in the focusing system from the perspective of TPT was presented. The intrinsic mechanism of TPT is attributed to the cylindrical symmetry-breaking of the system. The TPT in the focusing system was compared with that in the previous two systems, i.e., beam scattering at optical interfaces and propagation in uniaxial crystals. It was found that the incidence-angle-driven TPT in the previous two systems can also be considered to be induced by the symmetry-breaking of the systems. Furthermore, the TPT phenomena in the focusing system was explained, in terms of vortex mode decomposition, indicating that it can be accurately described by the competition and superposition of multiple vortex modes. Our work paves the way for understanding the two types of SOIs in other systems, providing an opportunity for the development of a unified theoretical framework of the TPT phenomenon in various spin-orbit photonics systems. More importantly, since SOIs play an important role in the fields of nanophotonics [1719], topological photonics [2022], and plasmonics [2325], our findings can be applied to optical vortex metrology and diagnostics [50], structured light communication [55], and particle manipulation [56].

Funding

National Natural Science Foundation of China (12174091); Hunan Provincial Innovation Foundation for Postgraduate (CX20221288).

Disclosures

The authors declare no conflicts of interest.

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. A. T. O’Neil, I. MacVicar, L. Allen, and M. J. Padgett, “Intrinsic and extrinsic nature of the orbital angular momentum of a light beam,” Phys. Rev. Lett. 88(5), 053601 (2002). [CrossRef]  

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

3. A. Aiello, P. Banzer, M. Neugebauer, and G. Leuchs, “From transverse angular momentum to photonic wheels,” Nat. Photonics 9(12), 789–795 (2015). [CrossRef]  

4. K. Y. Bliokh and N. Franco, “Transverse and longitudinal angular momenta of light,” Phys. Rep. 592, 1–38 (2015). [CrossRef]  

5. 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]  

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

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

8. K. Y. Bliokh and Y. P. Bliokh, “Conservation of angular momentum, transverse shift, and spin Hall effect in reflection and refraction of an electromagnetic wave packet,” Phys. Rev. Lett. 96(7), 073903 (2006). [CrossRef]  

9. O. Hosten and P. Kwiat, “Observation of the spin Hall effect of light via weak measurements,” Science 319(5864), 787–790 (2008). [CrossRef]  

10. Y. Zhao, J. S. Edgar, G. D. Jeffries, D. McGloin, and D. T. Chiu, “Spin–to–orbital angular momentum conversion in a strongly focused optical beam,” Phys. Rev. Lett. 99(7), 073901 (2007). [CrossRef]  

11. L. T. Vuong, A. J. L. Adam, J. M. Brok, P. C. M. Planken, and H. P. Urbach, “Electromagnetic spin–orbit interactions via scattering of subwavelength apertures,” Phys. Rev. Lett. 104(8), 083903 (2010). [CrossRef]  

12. K. Y. Bliokh, E. A. Ostrovskaya, M. A. Alonso, O. G. Rodríguez-Herrera, D. Lara, and C. Dainty, “Spin-to-orbital angular momentum conversion in focusing, scattering, and imaging systems,” Opt. Express 19(27), 26132–26149 (2011). [CrossRef]  

13. S. Fu, C. Guo, G. Liu, Y. Li, H. Yin, Z. Li, and Z. Chen, “Spin-orbit optical Hall effect,” Phys. Rev. Lett. 123(24), 243904 (2019). [CrossRef]  

14. Z. Zhang, W. Mei, J. Cheng, Y. Tan, Z. Dai, and X. Ling, “Revisiting vortex generation in the spin–orbit interactions of refraction and focusing of light,” Phys. Rev. A 106(6), 063520 (2022). [CrossRef]  

15. S. Khonina and I. Golub, “Vectorial spin Hall effect of light upon tight focusing,” Opt. Lett. 47(9), 2166–2169 (2022). [CrossRef]  

16. X. Ling, Z. Zhang, Z. Dai, Z. Wang, H. Luo, and L. Zhou, “Photonic spin–Hall effect at generic interfaces,” Laser Photonics Rev. 17(4), 2200783 (2023). [CrossRef]  

17. E. Maguid, M. Yannai, A. Faerman, I. Yulevich, V. Kleiner, and E. Hasman, “Disorder–induced optical transition from spin Hall to random Rashba effect,” Science 358(6369), 1411–1415 (2017). [CrossRef]  

18. L. Du, Z. Xie, G. Si, A. Yang, C. Li, J. Lin, G. Li, H. Wang, and X. Yuan, “On–Chip Photonic Spin Hall Lens,” ACS Photonics 6(8), 1840–1847 (2019). [CrossRef]  

19. J. Wang, L. Shi, and J. Zi, “Spin Hall Effect of Light via Momentum–Space Topological Vortices around Bound States in the Continuum,” Phys. Rev. Lett. 129(23), 236101 (2022). [CrossRef]  

20. Q. Guo, W. Gao, J. Chen, Y. Liu, and S. Zhang, “Line degeneracy and strong spin–orbit coupling of light with bulk bianisotropic metamaterials,” Phys. Rev. Lett. 115(6), 067402 (2015). [CrossRef]  

21. W. J. M. Kort-Kamp, “Topological phase transitions in the photonic spin Hall effect,” Phys. Rev. Lett. 119(14), 147401 (2017). [CrossRef]  

22. S. Ma, B. Yang, and S. Zhang, “Topological photonics in metamaterials,” Photonics Insights 1(1), R02 (2022). [CrossRef]  

23. D. Pan, H. Wei, L. Gao, and H. Xu, “Strong spin-orbit interaction of light in plasmonic nanostructures and nanocircuits,” Phys. Rev. Lett. 117(16), 166803 (2016). [CrossRef]  

24. Q. Xu, S. Ma, C. Hu, Y. Xu, C. Ouyang, X. Zhang, Y. Li, W. Zhang, Z. Tian, J. Gu, X. Zhang, S. Zhang, J. Han, and W. Zhang, “Coupling-mediated selective spin-to-plasmonic-orbital angular momentum conversion,” Adv. Opt. Mater. 7(20), 1900713 (2019). [CrossRef]  

25. M. Thomaschewski, Y. Yang, C. Wolff, A. S. Roberts, and S. I. Bozhevolnyi, “On–chip detection of optical spin–orbit interactions in plasmonic nanocircuits,” Nano Lett. 19(2), 1166–1171 (2019). [CrossRef]  

26. X. Ling, F. Guan, X. Cai, S. Ma, H. Xu, Q. He, S. Xiao, and L. Zhou, “Topology–induced phase transitions in spin–orbit photonics,” Laser Photonics Rev. 15(6), 2000492 (2021). [CrossRef]  

27. X. Ling, F Guan, Z. Zhang, H. Xu, S. Xiao, and H. Luo, “Vortex mode decomposition of the topology-induced phase transitions in spin-orbit optics,” Phys. Rev. A 104(5), 053504 (2021). [CrossRef]  

28. A. Debnath and N. K. Viswanathan, “Generic optical singularities and beam-field phenomena due to general paraxial beam reflection at a plane dielectric interface,” Phys. Rev. A 106(1), 013522 (2022). [CrossRef]  

29. A. Debnath, N. Kumar, U. Baishya, and N. K. Viswanathan, “Optical singularity dynamics and spin-orbit interaction due to a normal-incident optical beam reflected at a plane dielectric interface,” Phys. Rev. A 107(1), 013522 (2023). [CrossRef]  

30. Z. Zhang, J. Cheng, W. Mei, W. Xiao, Z. Wang, Z. Dai, and X. Ling, “Enhancing the efficiency of the topological phase transitions in spin–orbit photonics,” Appl. Phys. Lett. 120(18), 181102 (2022). [CrossRef]  

31. H. Liu and L. Yuan, “Controlling the spin Hall effect of grafted vortex beams propagating in uniaxial crystal,” Opt. Express 31(6), 10434–10448 (2023). [CrossRef]  

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

33. X. Gao, Y. Pan, G. Zhang, M. Zhao, Z. Ren, C. Tu, Y. Li, and H. Wang, “Redistributing the energy flow of tightly focused ellipticity–variant vector optical fields,” Photonics Res. 5(6), 640–648 (2017). [CrossRef]  

34. W. Shu, C. Lin, J. Wu, S. Chen, X. Ling, X. Zhou, H. Luo, and S. Wen, “Three-dimensional spin Hall effect of light in tight focusing,” Phys. Rev. A 101(2), 023819 (2020). [CrossRef]  

35. H. Li, C. Ma, J. Wang, M. Tang, and X. Li, “Spin–orbit Hall effect in the tight focusing of a radially polarized vortex beam,” Opt. Express 29(24), 39419–39427 (2021). [CrossRef]  

36. C. Prajapati, “Study of electric field vector, angular momentum conservation and Poynting vector of nonparaxial beams,” J. Opt. 23(2), 025604 (2021). [CrossRef]  

37. J. Davis and J. Bentley, “Azimuthal prism effect with partially blocked vortex–producing lenses,” Opt. Lett. 30(23), 3204–3206 (2005). [CrossRef]  

38. V. V. Kotlyar, S. S. Stafeev, and A. P. Porfirev, “Tight focusing of an asymmetric Bessel beam,” Opt. Commun. 357, 45–51 (2015). [CrossRef]  

39. P. Li, S. Liu, Y. Zhang, G. Xie, and J. Zhao, “Experimental realization of focal field engineering of the azimuthally polarized beams modulated by multi-azimuthal masks,” J. Opt. Soc. Am. B 32(9), 1867–1872 (2015). [CrossRef]  

40. C. Huang, Y. Zheng, and H. Li, “Orbital angular momentum and paraxial propagation characteristics of non-coaxial Laguerre–Gaussian beams,” J. Opt. Soc. Am. A 33(11), 2137–2143 (2016). [CrossRef]  

41. P. Srinivas, C. Perumangatt, N. Lal, R. Singh, and B. Srinivasan, “Investigation of propagation dynamics of truncated vector vortex beams,” Opt. Lett. 43(11), 2579–2582 (2018). [CrossRef]  

42. C. Cisowski and R. Correia, “Splitting an optical vortex beam to study photonic orbit–orbit interactions,” Opt. Lett. 43(3), 499–502 (2018). [CrossRef]  

43. A. Bekshaev and L. Mikhaylovskaya, “Displacements of optical vortices in Laguerre–Gaussian beams diffracted by a soft-edge screen,” Opt. Commun. 447, 80–88 (2019). [CrossRef]  

44. S. Khonina and I. Golub, “Breaking the symmetry to structure light,” Opt. Lett. 46(11), 2605–2608 (2021). [CrossRef]  

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

46. H. Cui, X. Wang, B. Gu, Y. Li, J. Chen, and H. Wang, “Angular diffraction of an optical vortex induced by the Gouy phase,” J. Opt. 14(5), 055707 (2012). [CrossRef]  

47. N. B. Baranova, A. Y. Savchenko, and B. Y. Zel’Dovich, “Transverse shift of a focal spot due to switching of the sign of circular polarization,” JETP Lett. 59, 232–234 (1994).

48. B. Y. Zel’Dovich, N. D. Kundikova, and L. F. Rogacheva, “Obsevation transverse shift of a focal spot upon a change in the sign of circular polarization,” JETP Lett. 59, 766–769 (1994).

49. J. Arlt, “Handedness and azimuthal energy flow of optical vortex beams,” J. Mod. Opt. 50(10), 1573–1580 (2003). [CrossRef]  

50. A. Y. Bekshaev, “Spin–orbit interaction of light and diffraction of polarized beams,” J. Opt. 19(8), 085602 (2017). [CrossRef]  

51. A. D’Errico, R. D’Amelio, B. Piccirillo, F. Cardano, and L. Marrucci, “Measuring the complex orbital angular momentum spectrum and spatial mode decomposition of structured light beams,” Optica 4(11), 1350–1357 (2017). [CrossRef]  

52. S. Fu, Y. Zhai, J. Zhang, X. Liu, R. Song, H. Zhou, and C. Gao, “Universal orbital angular momentum spectrum analyzer for beams,” PhotoniX 1(1), 19 (2020). [CrossRef]  

53. M. V. Vasnetsov, V. A. Pas’ko, and M. S. Soskin, “Analysis of orbital angular momentum of a misaligned optical beam,” New J. Phys. 7, 46 (2005). [CrossRef]  

54. B. Jack, M. Padgett M, and S. Franke-Arnold, “Angular diffraction,” New J. Phys. 10(10), 103013 (2008). [CrossRef]  

55. Z. Wan, H. Wang, Q. Liu, X. Fu, and Y. Shen, “Ultra-Degree-of-Freedom Structured Light for Ultracapacity Information Carriers,” ACS Photonics (2023).https://doi.org/10.1021/acsphotonics.2c01640

56. K. Dholakia and T. Čižmár, “Shaping the future of manipulation,” Nat. Photonics 5(6), 335–342 (2011). [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 (5)

Fig. 1.
Fig. 1. The TPT of light in off-axis beam focusing. (a) Schematic of the problem studied. (b) The light intensity patterns of the normal, abnormal, and longitudinal modes after focusing by a thin lens when d = 0, 0.25w, 0.75w, and 1.25w. (c) The dependence of the beam centroid shift Δx of the abnormal and longitudinal modes on the off-axis distance d.
Fig. 2.
Fig. 2. The TPT of light under the incidence of left- and right-handed circular-polarization light. The first and second rows: intensity patterns of the incident, normal, abnormal, and longitudinal modes when the incident beam is left and right-handed circular-polarization, respectively.
Fig. 3.
Fig. 3. The TPT of light in sector-masked beam focusing. (a) Schematic of the problem studied. (b) The light intensity patterns of the normal, abnormal, and longitudinal modes after focusing by a thin lens when α=0, π/2, π, and 3π/2. (c) The dependence of the beam centroid shift Δx of the abnormal and longitudinal modes on the masked angle α.
Fig. 4.
Fig. 4. The TPT of light in horizontally masked beam focusing. (a) Schematic of the problem studied. (b) The light intensity patterns of the normal, abnormal, and longitudinal modes after focusing by a thin lens when h = 0, 0.5w, w, and 1.5w. (c) The dependence of the beam centroid shift Δx of the abnormal and longitudinal modes on the masked height h.
Fig. 5.
Fig. 5. Normalized vortex mode spectra of the focusing field of the (a) off-axis beam, (b) sector-masked beam, and (c) horizontally masked beam.

Equations (6)

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

E a ( r a ) = d 2 k a exp ( i k a r a ) ξ = + , , z U ξ a ( k a ) V ^ ξ ( k a ) ,
M ^ = 1 2 [ cos θ i cos θ f + 1 exp ( 2 i φ ) ( cos θ i cos θ f 1 ) 2 exp ( i φ ) cos θ f sin θ i exp ( 2 i φ ) ( cos θ i cos θ f 1 ) cos θ i cos θ f 2 exp ( i φ ) cos θ f sin θ i 2 exp ( i φ ) cos θ i sin θ f 2 exp ( i φ ) cos θ i sin θ f 2 sin θ i sin θ f ]
[ U + f U f U z f ] = M ^ [ U + i U i U z i ] = 1 2 [ ( cos θ f + 1 ) exp ( 2 i φ ) ( cos θ f 1 ) 2 exp ( i φ ) sin θ f ] U  +  i .
E ξ f ( ρ , ϕ , z ) = i k L exp ( i k L ) 2 π θ 1 f θ 2 f φ 1 φ 2 U ξ f exp ( i k z cos θ f ) exp [ i k ρ sin θ f cos ( φ  -  ϕ ) ] sin θ f d φ d θ f ,
Δ x = x | E ξ f | 2 d x d y | E ξ f | 2 d x d y .
C p m = E L G p m E f d x d y ( | E L G p m | 2 d x d y ) 1 / 2 ( | E f | 2 d x d y ) 1 / 2 ,
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.