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

Polarization singularities hidden in a deep subwavelength confined electromagnetic field with angular momentum

Open Access Open Access

Abstract

Topologies associated with polarization point and line singularities can provide tools for controlling light propagation. By using the Stokes parameter, we demonstrate the emergence of polarization singularities hidden in deep subwavelength confined electromagnetic fields with angular momentum. We show that when the incoming orbital angular momentum is appropriately chosen, highly confined electromagnetic fields with super-diffraction-limited spatial dimensions can be obtained. At the same time, a conversion of orbital to spin angular momentum occurs, leading to a non-trivial topology. Our method provides a platform for developing topological photonics and studying the behavior of polarization singularities under strong focusing.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Phase singularities—also called vortices or dislocations—are points or lines in space where the phase of an optical field is undefined and the corresponding intensity vanishes [115]. An optical vortex is characterized by a spiral phase and is topological in nature. In 1992, Allen and coworkers demonstrated that a light beam possessing a spiral phase can carry an optical orbital angular momentum (OAM) of per photon, where ћ is the reduced Planck constant, h divided by 2π. [2]. New families of phase singularities in the form of closed, knotted, and linked dislocation lines have also been reported and demonstrated [1621]. Because of their nontrivial spatial features, optical fields possessing phase singularities have formed the basis of a rapidly growing number of unique applications, including super-resolution imaging [2224], optical micromanipulation [25,26], quantum information processing [27,28], and optical communication [2932].

In addition to scalar phase singularities, the free space optical field can also support singularities in its polarization vector that have attracted growing attention [3351]. For the two-dimensional paraxial fields, the polarization singularities in these fields are isolated points, at which at least one of the parameters defining the state of polarization (SoP) is indeterminate. The generic singularities in an ellipse field are called C-points. The C-point is circularly polarized when the orientation is undefined and is surrounded by an elliptical SoP. Similar to C-point singularities, V-point singularities also occur in a vector field and are characterized by integer values of the Poincare–Hopf index. In difference, the V-point is a vector-point singularity at which the direction of the electric vector of a linearly polarized field is undefined. Moreover, in the neighborhood of the V-point, the SoPs are all linearly polarized and undergo changes only in the azimuth. These singularities are structurally stable when perturbed and propagating. In this sense, they are the natural counterpart to phase singularities in scalar optics. Polarization singularities usually appear in random fields, multi-beam interference, and vector beams. Most recently, polarization singularities with torus knot and link topologies as well as topologically more complicated figure-eight knots have been generated and observed [52,53]. In addition, a nonparaxial 3D polarization topology in the form of Möbius strips was found to appear in an appropriately chosen observation plane near the focus of a tightly focused Gaussian beam [54,55].

In this Letter, based on the general integrated analytical formula for tight focusing of arbitrary locally linearly polarized fields, the focusing properties of azimuthally polarized (AP) Laguerre-Gaussian (LG) beams are numerically investigated and discussed in detail. Calculations show that the intrinsic optical degree of freedom of the OAM does provide a simple but powerful solution for tailoring the highly confined field distributions. By manipulating the topological charge, we obtain a superdiffraction-limited electromagnetic field without additional phase or amplitude modulation. Meanwhile, orbital-to-spin angular momentum conversion occurs, resulting in nontrivial polarization topological structures. In particular, by exploiting the two-dimensional Stokes parameters, the topological behavior of polarization hidden in complex polarization distributions is revealed.

2. Deep subwavelength confined electromagnetic field with angular momentum

Polarization, as an intrinsic optical degree of freedom, is one of the salient features of light. From theory, a monochromatic light beam with arbitrary locally linear SoPs may be written in polar coordinates (ρ, ϕ) as [5662]

$$\begin{aligned} {{\boldsymbol E}_0}(\rho ,\;\phi ) &= {A_0}\{{\textrm{exp} [{ - i\sigma (\rho ,\;\phi )} ]{{\hat{{\mathbf e}}}_L} + \textrm{exp} [{i\sigma (\rho ,\;\phi )} ]{{\hat{{\mathbf e}}}_R}} \}\textrm{exp} ({ikz - i\omega t} )\\ &= \sqrt 2 {A_0}\{{\cos [{\sigma (\rho ,\;\phi )} ]{{\hat{{\mathbf e}}}_x} + \sin [{\sigma (\rho ,\;\phi )} ]{{\hat{{\mathbf e}}}_y}} \}\textrm{exp} ({ikz - i\omega t} ),\end{aligned}$$
where A0 denotes the complex amplitude, ${\{ }{\hat{{\mathbf e}}_L}\textrm{,}\;{\hat{{\mathbf e}}_R}\} $ are the left- and right-handed circularly polarized eigenvectors, and $\{ {\hat{{\mathbf e}}_x},\;{\hat{{\mathbf e}}_y}\} $ are the horizontally and vertically polarized eigenvectors. The function σ, which theoretically has an arbitrary spatial distribution that depends on polar radius ρ and azimuthal angle ϕ, determines the relative polarization distribution of the monochromatic light beam. For instance, when σ = c, where c is a constant, the light described by Eq. (1) degenerates into a uniform linearly polarized beam at an angle c with the x axis. And it is, respectively, radially polarized (RP) beam and AP beam when σ = ϕ and ϕ +p/2.

Highly confined electromagnetic fields, which can be generated, for example, in high numerical-aperture (NA) objective lens focusing conditions shown in Fig. 1, are widely used as excellent tools to study detailed nano-optics. Strongly focused light fields have been widely analyzed using the generalized vector Debye integral. Richards and Wolf considered the contributions of input polarization to the focused field and built an analytical model for a uniform linearly polarized beam [63]. Similarly, the electric field E at each point r near the focus can also be derived when the incident polarized light beam is embodied as in Eq. (1).

 figure: Fig. 1.

Fig. 1. Schematic of the high NA objective lens focusing system and coordinate system followed in our calculations. The focal plane is located at z = 0.

Download Full Size | PDF

Here, we employ the LG complex amplitude distribution $\textrm{LG}_p^l$, for which l and p are integers associated with intertwined helices and relate to the topological charge and the number of additional concentric rings. With p = 0, the electromagnetic field in the image space becomes

$$\begin{array}{l} {\boldsymbol E}\;({\boldsymbol r}) ={-} \frac{{ikf}}{{2\mathrm{\pi }}}\int\limits_0^\alpha {\int\limits_0^{2\mathrm{\pi }} {\sqrt {\cos \theta } {e^{\{{ik[{ - r\sin \theta \cos ({\phi - \varphi } )+ z\cos \theta } ]} \}}}} } \\ \;\;\;\;\;\;{\kern 1pt} {\kern 1pt} {\kern 1pt} \;\;\;\;{\left( {\frac{{\sqrt 2 \beta \sin \theta }}{{\sin \alpha }}} \right)^{|l|}}\textrm{exp} \left( { - \frac{{{\beta^2}{{\sin }^2}\theta }}{{{{\sin }^2}\alpha }}} \right)\textrm{exp} (il\phi ){\mathbf V}\sin \theta \textrm{d}\phi \textrm{d}\theta, \end{array}$$
where f denotes the focal distance, α=arcsin(NA/n), where NA = 0.95 is the numerical aperture and n = 1 is the index of refraction of free space, and β is the ratio of the pupil radius to the beam waist that we take as one in the subsequent calculations. The vector V represents the electric field polarization vector in the image space,
$${\mathbf V} = \left[ \begin{array}{l} {\textrm{V}_x}\\ {\textrm{V}_y}\\ {\textrm{V}_z} \end{array} \right] = \left[ \begin{array}{l} \sin ({\phi - \sigma } )\sin \phi + \cos ({\phi - \sigma } )\cos \theta \cos \phi \\ - \sin ({\phi - \sigma } )\cos \phi + \cos ({\phi - \sigma } )\cos \theta \sin \phi \\ \cos ({\phi - \sigma } )\sin \theta \end{array} \right]. $$
Within the above formalism, we now investigate the non-paraxial field distributions in the image space of the tightly focused paraxial LG beams. Usually, an AP Bessel-Gaussian input plane wave can generate a doughnut like spot with sharper focal dark center in a high NA objective lens focusing system [64], which is extremely useful in practical application like STED microscopy [65]. The effect of input OAM on the electric intensity distributions of tightly focused AP beam is depicted in Fig. 2, which shows the total electric intensity distributions in the focal and through-focus planes of three types of AP LG beams with l = 0, 1, and 2 when p = 0, respectively. All these beams show good cylindrically symmetric polarization distributions in the beam cross section. As a result, their focal fields give a perfect circularly symmetric intensity distribution. However, they are different. For l = 0 and 2, they are on-axis energy null and annular intensity distributions, whereas it is a hot spot and strongest on axis for l = 1. So, the electric field distribution firstly shrinks toward the optical axis and then slowly moves away from it with the increase in l, which is very interesting. Furthermore, these electric intensity distributions are totally contributed by the transverse component. For the input AP LG beams, the function σ in Eq. (1) must be chosen in the form σ = ϕ + π/2, where ϕ is the azimuthal angle with respect to x0 axis as shown in Fig. 1. From Eq. (3), the corresponding longitudinal polarization vector in the focal region can be obtained as Vz = 0. Clearly, there is no longitudinal electric field component.

 figure: Fig. 2.

Fig. 2. Calculated total electric field intensity distributions in the focal (left column) and through-focus (right column) planes for the input AP $\textrm{LG}_0^0$, $\textrm{LG}_0^1$ and $\textrm{LG}_0^2$ beams (from the first row to the third row, respectively) All intensities have been normalized by the peak intensity of the total field in the focal volume for each input light.

Download Full Size | PDF

Numerous optical instruments and equipment require the use of sharply focused light fields, ranging from lithography, data storage, as well as microscopy to optical trapping. The vector properties of light have been utilized to obtain the smallest possible spot. It has been demonstrated that a RP input plane wave can lead to a spot size significantly smaller than for linear and circular polarizations [6668]. To be better understanding of the on-axis circularly symmetric hot spot shown in Figs. 2(c) and 2(d), Fig. 3 depicts its corresponding normalized focal field intensity profiles along x and z axes. For the sake of comparison, the results for RP $\textrm{LG}_0^0$ input beam under the same focusing conditions are also given. The RP $\textrm{LG}_0^0$ beam can have a narrow central peak due to the appearance of the strong longitudinal field component that is sharply centered on the optical axis as shown in Fig. 3(a). The beam size contributed by the longitudinal field component has a full width at half-maximum (FWHM) of 0.506λ in the transverse direction. However, the beam size for the total field is as large as 0.778λ, which is larger than the diffraction limit for this focusing lens λ/(2NA) = 0.526λ. Nevertheless, the FWHM value for the total fields of input AP $\textrm{LG}_0^1$ beam is only 0.522λ, which is beyond diffraction limit. From Fig. 3(b), the beam size along optical axis is calculated to be about 1.458λ and 1.416λ for RP $\textrm{LG}_0^0$ and AP $\textrm{LG}_0^1$ beams, respectively. So, an AP $\textrm{LG}_0^1$ input beam can generate a tighter hot spot than RP $\textrm{LG}_0^0$ beam, not only in the transverse direction but also in the longitudinal direction. Most importantly, we believe that this transversely polarized ultra-diffraction-limited spot is an important complement to its longitudinally polarized counterpart, since polarization plays a crucial and even decisive role in light-matter interactions.

 figure: Fig. 3.

Fig. 3. (a) Calculated normalized focal electric intensity profiles in the x axis direction for the longitudinal component and the total fields of input RP $\textrm{LG}_0^0$ beam as well as total fields of input AP $\textrm{LG}_0^1$ beam. (b) The corresponding intensity profiles along z axis for the total fields of the above two beams.

Download Full Size | PDF

For the input paraxial AP $\textrm{LG}_0^1$ beam, it carries optical OAM directed along the propagation axis and no spin angular momentum (SAM). For its strongly focused non-paraxial field, the corresponding angular momentum properties can be studied using SAM and OAM densities, which can be expressed, respectively, as [6971]

$${{\boldsymbol j}_{\textrm{SAM}}} = \frac{{{\varepsilon _0}}}{{4{\omega _0}}}{\mathop{\rm Im}\nolimits} [{{{\boldsymbol E}^\ast } \times {\boldsymbol E}} ], $$
$${{\boldsymbol j}_{\textrm{OAM}}} = \frac{{{\varepsilon _0}}}{{4{\omega _0}}}{\boldsymbol r} \times {\mathop{\rm Im}\nolimits} [{{{\boldsymbol E}^\ast } \cdot (\nabla ){\boldsymbol E}} ], $$
where ɛ0 is the permittivity in vacuum, the superscript asterisk represents complex conjugate, ω0 is the circular frequency.

Figure 4 illustrates the normalized distributions of the SAM and OAM densities in the focal and through-focus planes for the input AP $\textrm{LG}_0^1$ beam. From calculations, only a longitudinal component of the SAM density [Figs. 4(a) and 4(b)] emerges, arising from a missing out-of-plane electric field component [see Figs. 2(c) and 2(d)], and thus results in the absence of a transverse component. Here, the positive and negative values mean that its direction is along the forward and reverse directions of the optical axis, respectively. And the appearance of longitudinal SAM also shows an interesting optical process referred to as orbital-to-spin conversion, in contrast to the well-known spin-to-orbital angular momentum conversion. There are on-axis densities null and annular distributions for both transverse and longitudinal components of optical OAM [see Figs. 4(c)–4(f)]. Additionally, compared with the longitudinal component, the transverse component is much stronger, which is much different from its SAM distribution.

 figure: Fig. 4.

Fig. 4. Optical SAM and OAM density distributions in the focal (left column) and through-focus (right column) planes for the input AP $\textrm{LG}_0^1$ beam. The three rows from top to the bottom are, respectively, the normalized longitudinal SAM, transverse and longitudinal OAM density distributions.

Download Full Size | PDF

3. Polarization singularities hidden in the complex polarization distributions

Polarization plays a vital role in the light-matter interaction. Figures 5(a) and 5(b) illustrates the polarization distributions on the propagation and focal planes for the input AP $\textrm{LG}_0^1$ beam, respectively. The polarization projections in both the x-z and y-z planes are only linearly polarized [see Fig. 5(a)], indicating the spinning axis of the electric field is parallel to the z axis; i.e., the polarization ellipses all lie in the plane parallel to the x-y plane, arising from the absence of longitudinal electric field component. However, the polarization projection on the x-y plane is extremely complicated because the polarization ellipse becomes a solid ellipse in some regions away from the optical axis, indicating that the direction of the long axis of the polarization ellipse rotates as the light propagates. Overall, the polarization distribution on the focal plane exhibits perfect cylindrical symmetry, and linear, circular and elliptical SoPs mixed in the beam cross section [see Fig. 5(b)]. Meanwhile, the local polarization ellipse changes not only in its handedness but also in its ellipticity and orientation in the radial direction, whereas it only changes in the orientation in the azimuthal direction.

 figure: Fig. 5.

Fig. 5. (a) Polarization distributions in the x-z plane when y = 0 and their projections onto the three orthogonal planes. (b) Polarization ellipse fields in the focal plane for the input AP $\textrm{LG}_0^1$ beam.

Download Full Size | PDF

Now, we focus our attention on the polarization topology. The light vibrations lie in the transverse plane for the aforementioned non-paraxial ultra-diffraction-limited light fields, analogous to the traditional 2D paraxial optical fields. Therefore, we can study the topological structures using the same analysis for traditional transverse fields. First, we recall the Stokes parameters, s0, s1, s2, and s3 used in 2D paraxial fields, but adapt them to the current field,

$${s_0} = {|{{E_x}} |^2} + {|{{E_y}} |^2}, $$
$${s_1} = {|{{E_x}} |^2} - {|{{E_y}} |^2}, $$
$${s_2} = 2Re ({E_x^ \ast {E_y}} ), $$
$${s_3} = 2{\mathop{\rm Im}\nolimits} ({E_x^ \ast {E_y}} ). $$

Detailed analysis of the polarization topology can be carried out with the help of the normalized Stokes parameters, namely, Si = si/s0, (i = 0, 1, 2, 3). For a spatially varying polarization distribution, the Stokes parameters are also functions of the position coordinates. The complex Stokes field S12 = S1 + iS2 can be used for probing polarization singularities. The phase distribution corresponding to this Stokes field is expressed as φ12 = arctan(S2/S1). The phase singularities of this complex Stokes field correspond to the polarization singularities in the actual field distribution.

Figure 6 depicts the phase variation of the Stokes field S12 in the focal plane. The phase variation of the Stokes field S12 exhibits a vortex of charge 2 in the center. In the outside of optical axis, four annular boundaries with different widths, which are π/2 phase shifted in the radial direction, separate the phase pattern into several concentric regions. Phase singularities appear on the inner and outer rings of the annulus boundaries. In total, there are about ten positions that exhibit phase singularities as can be found in the inset in Fig. 6.

 figure: Fig. 6.

Fig. 6. Phase variation of the Stokes field S12 on the focal plane for the input AP $\textrm{LG}_0^1$ beam.

Download Full Size | PDF

To further explore the polarization topology hidden in the complex polarization distribution, the normalized stokes parameters S1, S2, and S3 are investigated. Figures 7(a)–7(c) shows, respectively, the contour lines of the normalized stokes parameters S1 = 0, S2 = 0, and S3 = 0. Two straight lines crossing each other on the axis and several circle can be found when S1 = 0 and S2 = 0. However, they are different, the two straight lines in Fig. 7(a) being directed, respectively, along the azimuthal angles of π/4 and 3π/4, whereas they are along the x- and y-axes in Fig. 7(b). For S3 = 0, however, only circles appear. The zero counters of all three normalized Stokes parameters S1, S2, and S3 are also plotted in a single map [see Fig. 7(d)]. It is known that at the C-points, S1 = S2 = 0, and for the L lines, S3 = 0. Interestingly, the straight lines in Fig. 7(a) and 7(b) cross each other on the axis in Fig. 7(d), resulting in a C point. And the circles in Fig. 7(a) and 7(b) are found to be coincident, resulting in the appearance of C lines. These locations should also be the positions where the annular boundaries emerge in the phase variation of the Stokes field S12 [see Fig. 6]. In contrast, the circles in Fig. 7(c) are L lines. For the innermost circle, the light vibrations should be all radially arranged and exhibit perfect circular symmetry [see Fig. 5(b)], and therefore is much different from the traditional L line. Furthermore, the location of this L line should be exactly where the SAM density is zero. On the two sides of this L line, the polarization ellipses have opposite handedness [see Figs. 4(a) and 4(b)].

 figure: Fig. 7.

Fig. 7. Theoretically calculated contour lines of the normalized Stokes parameters with (a) S1 = 0, (b) S2 = 0, (c) S3 = 0, and (d) the zero counters of all three normalized Stokes parameters S1, S2, and S3 in the focal plane for the input AP $\textrm{LG}_0^1$ beam.

Download Full Size | PDF

4. Conclusions

In summary, we reveal the optical polarization topology hidden in a deep-subwavelength confined electromagnetic field with angular momentum. A general integrated analytical formula for calculating tightly focused fields of arbitrary locally linearly polarized beams based on Richards and Wolf’s vector diffraction theory is proposed. Calculations show that the intrinsic optical degree of freedom of the OAM provides a simple but powerful solution for tailoring the focal field distributions of the input AP beam. By manipulating the topological charge, an ultra-diffraction-limit electromagnetic field can be obtained without additional phase or amplitude modulations. Meanwhile, orbital-to-spin angular momentum conversion occurs, resulting in nontrivial polarization topological structures. In particular, by using the Stokes parameter, the topological behaviors of polarization hidden in the complex polarization distributions are revealed. These results may contribute to the research of complex highly confined optical fields.

Funding

National Natural Science Foundation of China (12074224, 61975128); Natural Science Foundation of Shandong Province (ZR2020MA087, ZR2021YQ02).

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 corresponding author upon reasonable request.

References

1. J. F. Nye and M. V. Berry, “Dislocations in Wave Trains,” Proc. R. Soc. Lond. A 336, 165–190 (1974). [CrossRef]  

2. L. Allen, M. W. Beijersbergen, R. Spreeuw, and J. Woerdman, “Orbital angular momentum of light and the transformation of Laguerre-Gaussian laser modes,” Phys. Rev. A 45(11), 8185–8189 (1992). [CrossRef]  

3. M. S. Soskin and M. V. Vasnetsov, “Singular optics,” Prog. Opt. 42(4), 219–276 (2001). [CrossRef]  

4. J. E. Curtis and D. G. Grier, “Structure of Optical Vortices,” Phys. Rev. Lett. 90(13), 133901 (2003). [CrossRef]  

5. J. Leach, E. Yao, and M. J. Padgett, “Observation of the vortex structure of a non-integer vortex beam,” New J. Phys. 6(1), 71 (2004). [CrossRef]  

6. S. Franke-Arnold, L. Allen, and M. J. Padgett, “Advances in optical angular momentum,” Laser Photonics Rev. 2(4), 299–313 (2008). [CrossRef]  

7. M. R. Dennis, K. O’Holleran, and M. J. Padgett, “Singular optics: Optical Vortices and Polarization Singularities,” Prog. Opt. 53, 293–363 (2009). [CrossRef]  

8. N. Yu, P. Genevet, M. A. Kats, F. Aieta, J. Tetienne, F. Capasso, and Z. Gaburro, “Light Propagation with Phase Discontinuities: Generalized Laws of Reflection and Refraction,” Science 334(6054), 333–337 (2011). [CrossRef]  

9. A. M. Yao and M. J. Padgett, “Orbital angular momentum: origins, behavior and applications,” Adv. Opt. Photonics 3(2), 161–204 (2011). [CrossRef]  

10. L. Huang, X. Chen, H. Mühlenbernd, G. Li, B. Bai, Q. Tan, G. Jin, T. Zentgraf, and S. Zhang, “Dispersionless Phase Discontinuities for Controlling Light Propagation,” Nano Lett. 12(11), 5750–5755 (2012). [CrossRef]  

11. M. K. Lilley and R. M. Nyqvist, “Formation of Phase Space Holes and Clumps,” Phys. Rev. Lett. 112(15), 155002 (2014). [CrossRef]  

12. F. G. Mitri, “Vector spherical quasi-Gaussian vortex beams,” Phys. Rev. E 89(2), 023205 (2014). [CrossRef]  

13. M. Soskin, S. V. Boriskina, Y. Chong, M. R. Dennis, and A. Desyatnikov, “Singular optics and topological photonics,” J. Opt. 19(1), 010401 (2017). [CrossRef]  

14. L. Du, Z. Man, Y. Zhang, C. Min, S. Zhu, and X. Yuan, “Manipulating orbital angular momentum of light with tailored in-plane polarization states,” Sci. Rep. 7(1), 41001 (2017). [CrossRef]  

15. S. Nechayev, J. S. Eismann, G. Leuchs, and P. Banzer, “Orbital-to-spin angular momentum conversion employing local helicity,” Phys. Rev. B 99(7), 075155 (2019). [CrossRef]  

16. M. V. Berry and M. R. Dennis, “Knotted and linked phase singularities in monochromatic waves,” Proc. R. Soc. Lond. A 457(2013), 2251–2263 (2001). [CrossRef]  

17. J. Leach, M. R. Dennis, J. Courtial, and M. J. Padgett, “Knotted threads of darkness,” Nature 432(7014), 165 (2004). [CrossRef]  

18. W. T. M. Irvine and D. Bouwmeester, “Linked and knotted beams of light,” Nature Phys. 4(9), 716–720 (2008). [CrossRef]  

19. K. O’Holleran, M. R. Dennis, and M. J. Padgett, “Topology of light’s darkness,” Phys. Rev. Lett. 102(14), 143902 (2009). [CrossRef]  

20. M. R. Dennis, R. P. King, B. Jack, K. O’Holleran, and M. J. Padgett, “Isolated optical vortex knots,” Nature Phys. 6(2), 118–121 (2010). [CrossRef]  

21. X. Guo, P. Li, J. Zhong, S. Liu, B. Wei, W. Zhu, S. Qi, H. Cheng, and J. Zhao, “Tying polarization-switchable optical vortex knots and links via holographic all dielectric metasurfaces,” Laser Photonics Rev. 14(3), 1900366 (2020). [CrossRef]  

22. S. W. Hell, “Far-Field Optical Nanoscopy,” Science 316(5828), 1153–1158 (2007). [CrossRef]  

23. S. W. Hell and J. Wichmann, “Breaking the diffraction resolution limit by stimulated emission: stimulated-emission-depletion fluorescence microscopy,” Opt. Lett. 19(11), 780–782 (1994). [CrossRef]  

24. S. W. Hell, “Toward fluorescence nanoscopy,” Nat. Biotechnol. 21(11), 1347–1355 (2003). [CrossRef]  

25. D. G. Grier, “A revolution in optical manipulation,” Nature 424(6950), 810–816 (2003). [CrossRef]  

26. Y. Zhang, W. Shi, Z. Shen, Z. Man, C. Min, J. Shen, S. Zhu, H. P. Urbach, and X. Yuan, “A plasmonic spanner for metal particle manipulation,” Sci. Rep. 5(1), 15446 (2015). [CrossRef]  

27. A. R. Altman, K. G. Köprülü, E. Corndorf, P. Kumar, and G. A. Barbosa, “Quantum Imaging of Nonlocal Spatial Correlations Induced by Orbital Angular Momentum,” Phys. Rev. Lett. 94(12), 123601 (2005). [CrossRef]  

28. J. Leach, B. Jack, J. Romero, A. K. Jha, A. M. Yao, S. Franke-Arnold, D. G. Ireland, R. W. Boyd, S. M. Barnett, and M. J. Padgett, “Quantum Correlations in Optical Angle-Orbital Angular Momentum,” Science 329(5992), 662–665 (2010). [CrossRef]  

29. C. Paterson, “Atmospheric Turbulence and Orbital Angular Momentum of Single Photons for Optical Communication,” Phys. Rev. Lett. 94(15), 153901 (2005). [CrossRef]  

30. J. Wang, J. Yang, I. M. Fazal, N. Ahmed, Y. Yan, H. Huang, Y. Ren, Y. Yue, S. Dolinar, M. Tur, and A. E. Willner, “Terabit free-space data transmission employing orbital angular momentum multiplexing,” Nat. Photonics 6(7), 488–496 (2012). [CrossRef]  

31. A. E. Willner, H. Huang, Y. Yan, Y. Ren, N. Ahmed, G. Xie, C. Bao, L. Li, Y. Cao, Z. Zhao, J. Wang, M. P. J. Lavery, M. Tur, S. Ramachandran, A. F. Molisch, N. Ashrafi, and S. Ashrafi, “Optical communications using orbital angular momentum beams,” Adv. Opt. Photonics 7(1), 66–106 (2015). [CrossRef]  

32. N. Bozinovic, Y. Yue, Y. Ren, M. Tur, P. Kristensen, H. Huang, A. E. Willner, and S. Ramachandran, “Terabit-Scale Orbital Angular Momentum Mode Division Multiplexing in Fibers,” Science 340(6140), 1545–1548 (2013). [CrossRef]  

33. J. F. Nye, “Polarization effects in the diffraction of electromagnetic waves: the role of disclinations,” Proc. R. Soc. Lond. A 387(1792), 105–132 (1983). [CrossRef]  

34. J. F. Nye, “Lines of circular polarization in electromagnetic wave fields,” Proc. R. Soc. Lond. A 389(1797), 279–290 (1983). [CrossRef]  

35. M. R. Dennis, “Polarization singularities in paraxial vector fields: morphology and statistics,” Opt. Commun. 213(4-6), 201–221 (2002). [CrossRef]  

36. I. Freund, “Polarization flowers,” Opt. Commun. 199(1-4), 47–63 (2001). [CrossRef]  

37. M. V. Berry, M. R. Dennis, and R. L. Lee Jr, “Polarization singularities in the clear sky,” New J. Phys. 6, 162 (2004). [CrossRef]  

38. M. S. Soskin, V. Denisenko, and I. Freund, “Optical polarization singularities and elliptic stationary points,” Opt. Lett. 28(16), 1475–1477 (2003). [CrossRef]  

39. A. M. Beckley, T. G. Brown, and M. A. Alonso, “Full Poincaré beams,” Opt. Express 18(10), 10777–10785 (2010). [CrossRef]  

40. E. J. Galvez, S. Khadka, W. H. Schubert, and S. Nomoto, “Poincaré-beam patterns produced by nonseparable superpositions of Laguerre-Gauss and polarization modes of light,” Appl. Opt. 51(15), 2925–2934 (2012). [CrossRef]  

41. F. Cardano, E. Karimi, L. Marrucci, C. de Lisio, and E. Santamato, “Generation and dynamics of optical beams with polarization singularities,” Opt. Express 21(7), 8815–8820 (2013). [CrossRef]  

42. S. Vyas, Y. Kozawa, and S. Sato, “Polarization singularities in superposition of vector beams,” Opt. Express 21(7), 8972–8986 (2013). [CrossRef]  

43. E. J. Galvez and B. L. Rojec, “Generation of isolated asymmetric umbilics in light’s polarization,” Phys. Rev A 89(3), 031801 (2014). [CrossRef]  

44. B. Khajavi and E. J. Galvez, “High-order disclinations in space-variant polarization,” J. Opt. 18(8), 084003 (2016). [CrossRef]  

45. S. K. Pal and P. Senthilkumaran, “Cultivation of lemon fields,” Opt. Express 24(24), 28008–28013 (2016). [CrossRef]  

46. I. Freund, “Topological events on the lines of circular polarization in nonparaxial vector optical fields,” Opt. Lett. 42(3), 611–614 (2017). [CrossRef]  

47. B. S. B. Ram, P. Ruchi, and Senthilkumaran, “Angular momentum switching and orthogonal field construction of C-points,” Opt. Lett. 43(9), 2157–2160 (2018). [CrossRef]  

48. S. K. Pal and P. Senthilkumaran, “Hexagonal vector field of polarization singularities with a gradient basis structure,” Opt. Lett. 44(8), 2093–2096 (2019). [CrossRef]  

49. L. De Angelis, T. Bauer, F. Alpeggiani, and L. Kuipers, “Index-symmetry breaking of polarization vortices in 2D random vector waves,” Optica 6(9), 1237–1243 (2019). [CrossRef]  

50. G. Zhang, C. Tu, Y. Li, and H. Wang, “Observation of polarization topological singular lines,” Photonics Res. 7(6), 705–710 (2019). [CrossRef]  

51. W. Chen, Y. Chen, and W. Liu, “Singularities and Poincaré Indices of Electromagnetic Multipoles,” Phys. Rev. Lett. 122(15), 153907 (2019). [CrossRef]  

52. H. Larocque, D. Sugic, D. Mortimer, A. J. Taylor, R. Fickler, R. W. Boyd, M. R. Dennis, and E. Karimi, “Reconstructing the topology of optical polarization knots,” Nature Phys. 14(11), 1079–1082 (2018). [CrossRef]  

53. E. Pisanty, G. J. Machado, V. Vicuña-Hernández, A. Picón, A. Celi, J. P. Torres, and M. Lewenstein, “Knotting fractional-order knots with the polarization state of light,” Nat. Photonics 13(8), 569–574 (2019). [CrossRef]  

54. T. Bauer, P. Banzer, E. Karimi, S. Orlov, A. Rubano, L. Marrucci, E. Santamato, R. W. Boyd, and G. Leuchs, “Observation of optical polarization Möbius strips,” Science 347(6225), 964–966 (2015). [CrossRef]  

55. T. Bauer, M. Neugebauer, G. Leuchs, and P. Banzer, “Optical Polarization Möbius strips and Points of Purely Transverse Spin Density,” Phys. Rev. Lett. 117(1), 013601 (2016). [CrossRef]  

56. Q. Zhan, “Cylindrical vector beams: From mathematical concepts to applications,” Adv. Opt. Photonics 1(1), 1–57 (2009). [CrossRef]  

57. H. Rubinsztein-Dunlop, A. Forbes, M. Berry, et al., “Roadmap on structured light,” J. Opt. 19(1), 013001 (2017). [CrossRef]  

58. C. Rosales-Guzmán, B. Ndagano, and A. Forbes, “A review of complex vector light fields and their applications,” J. Opt. 20(12), 123001 (2018). [CrossRef]  

59. Z. Man, C. Min, L. Du, Y. Zhang, S. Zhu, and X. Yuan, “Sub-wavelength sized transversely polarized optical needle with exceptionally suppressed side-lobes,” Opt. Express 24(2), 874–882 (2016). [CrossRef]  

60. Z. Man, X. Dou, and S. Fu, “Pancharatnam-Berry phase shaping for control of the transverse enhancement of focusing,” Opt. Lett. 44(2), 427–430 (2019). [CrossRef]  

61. Z. Man, P. Meng, and S. Fu, “Creation of complex nano-interferometric field structures,” Opt. Lett. 45(1), 37–40 (2020). [CrossRef]  

62. Z. Man, Z. Xi, X. Yuan, R. E. Burge, and H. P. Urbach, “Dual Coaxial Longitudinal Polarization Vortex Structures,” Phys. Rev. Lett. 124(10), 103901 (2020). [CrossRef]  

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

64. K. S. Youngworth and T. G. Brown, “Focusing of high numerical aperture cylindrical vector beams,” Opt. Express 7(2), 77–87 (2000). [CrossRef]  

65. Y. Xue, C. Kuang, S. Li, Z. Gu, and X. Liu, “Sharper fluorescent super-resolution spot generated by azimuthally polarized beam in STED microscopy,” Opt. Express 20(16), 17653–17666 (2012). [CrossRef]  

66. R. Dorn, S. Quabis, and G. Leuchs, “Sharper focus for a radially polarized light beam,” Phys. Rev. Lett. 91(23), 233901 (2003). [CrossRef]  

67. Z. Man, S. Fu, and G. Wei, “Focus engineering based on analytical formulae for tightly focused polarized beams with arbitrary geometric configurations of linear polarization,” J. Opt. Soc. Am. A 34(8), 1384–1391 (2017). [CrossRef]  

68. Z. Man, Z. Bai, J. Li, S. Zhang, X. Li, X. Ge, and S. Fu, “Focus shaping by tailoring arbitrary hybrid polarization states that have a combination of orthogonal linear polarization bases,” Appl. Opt. 57(12), 3047–3055 (2018). [CrossRef]  

69. K. Y. Bliokh, A. Y. Bekshaev, and F. Nori, “Extraordinary momentum and spin in evanescent waves,” Nat. Commun. 5(1), 3300 (2014). [CrossRef]  

70. K. Y. Bliokh and F. Nori, “Transverse and Longitudinal Angular Momenta of Light,” Phys. Rep. 592, 1–38 (2015). [CrossRef]  

71. A. Aiello, P. Banzer, M. Neugebauer, and G. Leuchs, “From Transverse Angular Momentum of Photonic Wheels,” Nat. Photonics 9(12), 789–795 (2015). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the corresponding author 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 (7)

Fig. 1.
Fig. 1. Schematic of the high NA objective lens focusing system and coordinate system followed in our calculations. The focal plane is located at z = 0.
Fig. 2.
Fig. 2. Calculated total electric field intensity distributions in the focal (left column) and through-focus (right column) planes for the input AP $\textrm{LG}_0^0$, $\textrm{LG}_0^1$ and $\textrm{LG}_0^2$ beams (from the first row to the third row, respectively) All intensities have been normalized by the peak intensity of the total field in the focal volume for each input light.
Fig. 3.
Fig. 3. (a) Calculated normalized focal electric intensity profiles in the x axis direction for the longitudinal component and the total fields of input RP $\textrm{LG}_0^0$ beam as well as total fields of input AP $\textrm{LG}_0^1$ beam. (b) The corresponding intensity profiles along z axis for the total fields of the above two beams.
Fig. 4.
Fig. 4. Optical SAM and OAM density distributions in the focal (left column) and through-focus (right column) planes for the input AP $\textrm{LG}_0^1$ beam. The three rows from top to the bottom are, respectively, the normalized longitudinal SAM, transverse and longitudinal OAM density distributions.
Fig. 5.
Fig. 5. (a) Polarization distributions in the x-z plane when y = 0 and their projections onto the three orthogonal planes. (b) Polarization ellipse fields in the focal plane for the input AP $\textrm{LG}_0^1$ beam.
Fig. 6.
Fig. 6. Phase variation of the Stokes field S12 on the focal plane for the input AP $\textrm{LG}_0^1$ beam.
Fig. 7.
Fig. 7. Theoretically calculated contour lines of the normalized Stokes parameters with (a) S1 = 0, (b) S2 = 0, (c) S3 = 0, and (d) the zero counters of all three normalized Stokes parameters S1, S2, and S3 in the focal plane for the input AP $\textrm{LG}_0^1$ beam.

Equations (9)

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

E 0 ( ρ , ϕ ) = A 0 { exp [ i σ ( ρ , ϕ ) ] e ^ L + exp [ i σ ( ρ , ϕ ) ] e ^ R } exp ( i k z i ω t ) = 2 A 0 { cos [ σ ( ρ , ϕ ) ] e ^ x + sin [ σ ( ρ , ϕ ) ] e ^ y } exp ( i k z i ω t ) ,
E ( r ) = i k f 2 π 0 α 0 2 π cos θ e { i k [ r sin θ cos ( ϕ φ ) + z cos θ ] } ( 2 β sin θ sin α ) | l | exp ( β 2 sin 2 θ sin 2 α ) exp ( i l ϕ ) V sin θ d ϕ d θ ,
V = [ V x V y V z ] = [ sin ( ϕ σ ) sin ϕ + cos ( ϕ σ ) cos θ cos ϕ sin ( ϕ σ ) cos ϕ + cos ( ϕ σ ) cos θ sin ϕ cos ( ϕ σ ) sin θ ] .
j SAM = ε 0 4 ω 0 Im [ E × E ] ,
j OAM = ε 0 4 ω 0 r × Im [ E ( ) E ] ,
s 0 = | E x | 2 + | E y | 2 ,
s 1 = | E x | 2 | E y | 2 ,
s 2 = 2 R e ( E x E y ) ,
s 3 = 2 Im ( E x E y ) .
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.