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

Generalized Newton’s rings with vortex beams

Open Access Open Access

Abstract

The Newton’s rings are interference patterns with concentric rings, and Newton’s rings experiment is one of the most famous classic optics experiments. Here, we show that if we use a vortex beam, we can obtain generalized Newton’s rings. Unlike traditional Newton’s rings, the generalized ones are no longer concentric rings but spiral arms, and fork-shaped dislocations appear in spiral arms. More interesting, we reveal that both the number of spiral arms and the number of fork-shaped dislocations are equal to the value of topological charge of incident vortex beams. Our theoretical results are demonstrated experimentally. This novel interference pattern can be used for measuring the topological charge of vortex beams.

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

1. Introduction

Optical vortices have been widely studied since Allen et al. shown the connection between the spiral phase singularity of the Laguerre-Gaussian (LG) beam and the quantized orbital angular momentum (OAM) [1]. Light beam carrying OAM is associated with an azimuthal phase structure $\exp (il\varphi )$, with $\varphi$ and $l(l\in \mathbb {Z})$ the angular coordinate and azimuthal index, respectively. The azimuthal index is also called as the topological charge (TC) of the OAM beam, which presents an OAM of $l\hbar$ per photon ($\hbar =h/(2\pi )$ is reduced Planck’s constant) [1]. Due to the phase singularity at the center, the intensity profile of a vortex beam is always doughnut-shaped. OAM beams have been widely used in a variety of interesting applications [24], such as generation of Pancharatnam–Berry phase [5], optical microscopy [68], optical micromanipulation [912], quantum information [13,14], free-space and fiber optical communication [1519], and etc.

To distinguish different OAM states is one of the most fundamental steps towards the utilization of OAM beams. There are some existing methods to achieve this goal, such as interference with a plane wave or a spherical wave [20,21], bifurcation fringes of its self-interference [20], Cartesian to log-polar coordinate transformation [22,23], and diffraction patterns of various apertures or slits [2429]. Meanwhile, several classic interference experiments adopting OAM beams have been reported [30,31]. It is well known that, interference is a key evidence of the light’s wave nature and the interference pattern presents rich information of intensity and phase which is important in high precision measurement. Sztul and Alfano revealed the interference characteristics of OAM beams in Young’s double-slit interference experiment and pointed out the twist fringes are closely related to the topological charges of incident OAM beams [30].

Besides the double-slit interference, the Newton’s rings experiment is also one of the classic examples in optics. The Newton’s rings pattern shows equal thickness interference fringes and is usually observed in the interferometry with a spherical surface [32]. Due to the rotationally symmetric spherical surface and rotationally symmetric illumination such as the planar illumination and the Gaussian beams, the rotationally symmetric optical path length difference is induced and therefore the Newton’s rings pattern is consisted of a set of concentric circular fringes. The position and the curvature radius of a spherical surface can be measured by analyzing these Newton’s rings patterns. Although the concentric Newton’s rings pattern is similar to the doughnut-shaped intensity pattern of a vortex beam, the underlying physical mechanisms are totally different. What will happen when a vortex beam encounters the spherical surface is not unveiled till now and arises our curiosity.

In this paper we put forward the generalized Newton’s rings interference experiments with optical LG beams as illumination. We find that the generalized Newton’s rings pattern is consisted of the spiral pattern and the fork-shaped pattern which are significantly related to the topological charge of the incident optical vortex beam. This new phenomenon may lead to a better understanding and interpretation of the nature of the OAM of vortex beams. Due to its simplicity, the generalized Newton’s rings can be used for detecting and measuring the topological charges of vortex beams. While, it is noted that the generalized Newton’s rings patterns will become more complicated and difficult to recognized for vortex beams with high topological charges.

2. Methods and discussions

2.1 Theoretical discussions

Without loss of generality, LG beam was chosen to theoretical and experimental demonstrate the generalized Newton’s rings patterns in this work, for it is one of the most typical vortex beams and it has a complex field amplitude shown as followed [12]:

$$\begin{aligned}LG_p^l(\rho,\varphi,z)=&\frac{\omega_0}{\omega(z)}\sqrt{\frac{2p!}{\pi(|l|+p)!}}\bigg(\frac{\sqrt{2}\rho}{\omega(z)}\bigg)^{|l|}L_p^{|l|}\bigg[2\bigg(\frac{\rho}{\omega(z)}\bigg)^2\bigg] \\ &\times\exp[i(2p+|l|+1)\xi(z)] \\ &\times\exp\bigg[-\bigg(\frac{\rho}{\omega(z)}\bigg)^2\bigg]\exp\bigg[-\frac{ik\rho^2}{2R(z)}\bigg]\exp(il\varphi) \end{aligned}$$
where $\omega _0$ is the waist size of the beam, $L_p^{|l|}(2\rho ^2/\omega (z)^2)$ is the associated Laguerre polynomial, $l$ is the azimuthal index, while the $p$ is the radial index of the LG beam, a set of parameters $\omega (z)$, $\xi (z)$, $R(z)$ are defined as: $\omega (z)=\omega _0\sqrt {1+(z/z_R)^2}$, $\xi (z)=\arctan (z/z_R)$, $R(z)=z(1+(z_R/z)^2$, while the $z_R$ here is the Rayleigh range [12,33]. In most cases, the field amplitude of LG beam can always be separated into radial part and azimuthal part as followed:
$$LG_p^l(\rho,\varphi,z)=A_p^l(\rho,z)\exp(il\varphi)$$
for the sake of simplicity.

In a standard Newton’s rings interference experiment, the spherical surface is usually induced by a plano-convex lens (PCL). With plane wave illumination, the interference pattern on the screen is introduced by the reflected beams from each surface of the PCL and the optical path difference between these two beams is resulted from the geometry of the spherical surface. When a plane wave illuminates vertically, the two reflected beams propagate along the same axis with a spherical phase difference caused by the PCL and there is no more additional phase difference. The interference pattern will therefore be a series of concentric circular fringes formed around the axis of the two reflected beams and the PCL. The derivations of the intensity distribution of the standard Newton’s rings pattern are given as:

$$I(\rho,\varphi)=\bigg|E_0\exp\bigg(ik\frac{\lambda}{2}\bigg)+E_0\exp\bigg(ik\frac{\rho^2}{R}\bigg)\bigg|^2\propto\cos^2\bigg(\frac{k\rho^2}{2R}-\frac{\pi}{2}\bigg)$$

Particularly, Eq. (3) shows the intensity distribution of standard Newton’s rings pattern is azimuthally isotropic. And furthermore, it indicates that the classic Newton’s rings interference experiment with vertical plane wave illumination contains rotational symmetry. While the rotational symmetry is one of the intrinsic properties of the optical vortex beams. Therefore, it is predictable that the rotational symmetry would be conserved if the vertical illumination carried OAM, and the interference pattern would be the similar concentric circular fringes as the case with plane wave illumination.

In contrast, the same experiment setup but using oblique illumination will break this rotational symmetry and will exhibit generalized interference patterns with extra details due to this symmetry broken. Two reflected beams are separated due to the refraction caused by the PCL when incident beam obliquely illuminates, especially the two reflected beams propagated in parallel but separated spatially when the incident beam obliquely illuminates at the center of the spherical side of PCL. Therefore, in Newton’s rings experiment with oblique illumination, the interference patterns of these two reflected beams are not only resulted from the spherical phase induced by the PCL but also additionally resulted from the spatial dislocation caused by the refraction in the PCL due to the oblique incidence.

Based on the analysis by geometrical optics, we can find out the spatial dislocation of the two reflected beams resulted from the PCL with an oblique illumination. As shown in Fig. 1(a), when a LG beam passes through the PCL and reflects at the center of the spherical side of the lens, the reflected light and the incident light inside the PCL are symmetric about the axis of the lens due to the law of reflection and the geometry of sphere. Therefore, reflected beam $E_1$ comes from the spherical side of the lens will exit along the propagation direction which is mirrored to the incident beam about the axis of the PCL. While the reflected beam $E_2$ comes from the flatten side of the lens will also emit mirroredly to the incident beam about the normal, which is parallel to the axis of the PCL, these two reflected beams will propagate in parallel as shown in Fig. 1(a). Moreover, the spatial dislocation between these two beams can be derived from the geometric schematic shown in Fig. 1(a) as followed:

$$\Delta=2d\tan\bigg(\arcsin\bigg(\frac{\sin\theta}{n}\bigg)\bigg)\cos\theta$$
where $d$ is the thickness of the lens, $\theta$ is the incident angle and $n$ is the relative refractive index. If the incident angle $\theta$ is small enough, the spatial dislocation can be approximated to:
$$\Delta\approx \frac{2\theta d}{n}$$

 figure: Fig. 1.

Fig. 1. (a) Spatial dislocation $\Delta$ between the two reflect beams $E_1$, $E_2$ induced by the refraction of the illumination through the PCL; (b) cross-section schematic of the two reflect beams and the coordinate setup at the observing screen plane.

Download Full Size | PDF

Therefore, if we set a coordinate at the observing screen and set the origin point at the center of beam $E_2$ the one reflected by the flatten side of the PCL as shown in Fig. 1(b), the intensity distribution of the generalized Newton’s rings introduced by LG beam illumination carrying topological charge $l$ is given as:

$$\begin{aligned}I(x,y)\propto&\bigg|LG_{p,-l}(x,y)\exp\bigg(ik\frac{\lambda}{2}\bigg)+LG_{p,-l}(x,y+\Delta)\exp\bigg(ik\frac{x^2+(y+\Delta)^2}{R^2}\bigg)\bigg|^2 \\ \propto&\bigg|A_{p,-l}(x,y)\exp({-}il\varphi)\exp\bigg(ik\frac{\lambda}{2}\bigg) \\ &+A_{p,-l}(x,y+\Delta)\exp({-}il\varphi')\exp\bigg(ik\frac{x^2+(y+\Delta)^2}{R^2}\bigg)\bigg|^2 \end{aligned}$$
where $\Delta$ is the spatial dislocation between the two reflect beams as Eq. (4)(5), and the change of the topological charge from $l$ of illumination to $-l$ of reflection is the fact that reflection of vortex beams will invert their topological charges as discussed by previous researches [3436].

From Eq. (46), it is obvious that the intensity distribution of generalized Newton’s rings generated by LG beam is connected with the topological charge of the illumination when its incident angle $\theta$ is non-zero. The interference patterns calculated from Eq. (6) can be graphically expressed as Figs. 2(a), (c) and (e), where the topological charges of LG beams are modeling from $l=+1$ to $l=+3$. In this study, we ignore the radial index $p$ of LG beams, since for the cases $p\neq 0$, the interference patterns will be very complicated and it is not easy to find the regularity any longer. The numerical results shown in Fig. 2 indicate that the generalized Newton’s rings are the combinations of the spiral patterns and the fork-shaped patterns instead of the concentric circular fringes in classic Newton’s rings experiment. The fork-shaped dislocation clings to the arm of the spiral pattern, and the numbers of the spiral arms and the order of fork-shaped patterns are equal to the topological charge of the corresponding incident LG beam. While the topological charges of the illuminations invert into $l=-1, \dots, -3$, as shown in Figs. 2(b), (d) and (f), the chirality of the spiral arms and the direction of the fork patterns invert simultaneously.

 figure: Fig. 2.

Fig. 2. Numerical calculations of the generalized Newton’s rings obtained from vortex beams with different topological charges. (a) $l=+1$, (b) $l=-1$, (c) $l=+2$, (d) $l=-2$, (e) $l=+3$, (f) $l=-3$. The number of spiral arms in the interference pattern indicates the topological charge $l$, and the fork-shaped dislocation (marked by the dashed green circles) indicates the topological charge $l$ as well, namely, one interference fringe splits into $|l|+1$ fringes.

Download Full Size | PDF

2.2 Experimental method

With aforementioned theoretical analyses, the experimental setup to investigate the generalized Newton’s rings patterns generated by LG illuminations is designed as shown in Fig. 3. In our experiment we used a PCL, where the difference between the curvature on each side of the lens can introduce Newton’s rings in its reflected beams. The thickness $d$ of the PCL is $3.5 mm$ and its relative refractive index $n$ is $1.517$. A He-Ne laser ($\lambda = 632.8 nm$) beam properly expanded and collimated by a couple of lenses, i.e. Lens 1 and Lens 2, illuminates a reflective spatial light modulator (SLM) parallel with the PCL. Spiral phases and proper blazed phase are displayed on SLM to generate the illuminating LG beams with topological charges $l=1,2,\dots,4$ and incident angle $\theta$ about $36^{\circ }$. Since the SLM only modulates specifically polarized light, a polarizer is placed before the SLM to match its polarization. The illuminating LG beam is then selected by an aperture slot (AS 2) placing in front of the PCL, allowing only one mode of LG beam to pass through. Due to the LG beams carrying spherical phases naturally, an additional lens (Lens 3) is chosen to reshape the wavefronts and relocate the waist positions of the illuminating LG beams in order to optimize the interference results. The two reflected beams and their interference patterns from the PCL are recorded and measured by a CCD camera.

 figure: Fig. 3.

Fig. 3. Schematic of the experimental setup. AS, aperture slot; PCL, plano-convex lens; SLM, reflective spatial light modulator; CCD, charge-coupled device.

Download Full Size | PDF

3. Experimental results

Figures 4(a)-(d) show the generalized Newton’s rings interference fringes experimentally for $l =+1,+2,\dots,+4$. As we predicted in theory, the interference pattern in Fig. 4(a) is consisted of spiral pattern with one spiral arm at right and fold-1 fork-shaped pattern at left of the figure, which is corresponding to illumination LG beam with topological charge $l=+1$. Figures 4(b)-(d) show the fringe patterns generated by LG beams with high order topological charges, i.e., $l=+2,\dots,+4$. These interference patterns show the spiral patterns, which carry the same orders to the corresponding topological charges, located at right of the generalized Newton’s rings patterns.

 figure: Fig. 4.

Fig. 4. Experimental generalized Newton’s rings interference patterns obtained from (a) the $l=+1$ beam, (b) $l=+2$ beam, (c) $l=+3$ beam, (d) $l=+4$ beam.

Download Full Size | PDF

However, instead of the fork-shaped patterns with the same orders to incident LG beams shown in numerical results, the experimental results show that the fork-shaped patterns of generalized Newton’s rings generated by high-order LG beams, i.e., $l>1$, are split and separated into the corresponding numbers of fold-1 fork-shaped patterns. For example, with the LG beam carried $l=+3$ illuminating, the generalized Newton’s rings shown in Fig. 4(c) is consisted of a triple of fold-1 fork-shaped patterns instead of an order-3 fork-shaped pattern. The separation of the high-order fork-shaped dislocation in the generalized Newton’s rings in our experiment may be explained by the separating of the interference patterns due to the instability of high-order fork-shaped interference patterns discussed by Ricci et al. [37], or maybe it is because the high-order incident vortex beam doesn’t actually exist of perfect mode with high order topological charge.

4. Conclusions

As the example of interference fringes of equal thickness, the Newton’s rings pattern is one of the most classic interference patterns due to its concentric rings pattern induced by a spherical surface as the plane wave illuminating. Our demonstration shows that the generalized Newton’s rings pattern can be generated from the classic Newton’s rings interference experiment setup with oblique Laguerre-Gaussian beam illumination, and the generalized Newton’s rings are interference patterns consisted of spiral patterns and fork-shaped patterns instead of the simple concentric rings. Moreover, our further theoretical analyses and experiments show that both the number of spiral arms and the order of fork-shaped dislocations are equal to the topological charge of the incident LG beam. Considering these novel interference patterns are induced by the helical phase structures of the incident LG beams. Therefore, the generalized Newton’s rings experiment we revealed may be a new approach to determine the topological charges of optical vortex beams. Moreover, the generalized Newton’s rings of the fractional vortex beams [3841] is also an interesting topic and should be studied in the future. Although this is not the first candidate to measure a given optical vortex beam, this generalized Newton’s rings interference experiment provides an interesting and novel attempt to make a connection between the classic Newton’s rings interference experiment and the novel optical vortex beams.

Funding

National Natural Science Foundation of China (11874102, 12174047); Sichuan Province Science and Technology Support Program (2020JDRC0006); Fundamental Research Funds for the Central Universities (ZYGX2019J102).

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

2. Y. Shen, X. Wang, Z. Xie, C. Min, X. Fu, Q. Liu, M. Gong, and X. Yuan, “Optical vortices 30 years on: OAM manipulation from topological charge to multiple singularities,” Light: Sci. Appl. 8(1), 90 (2019). [CrossRef]  

3. J. Chen, C. Wan, and Q. Zhan, “Engineering photonic angular momentum with structured light: a review,” Adv. Photonics 3(06), 064001 (2021). [CrossRef]  

4. A. Forbes, “New twist to twisted light,” Adv. Photonics 4(03), 2 (2022). [CrossRef]  

5. M. Tymchenko, J. S. Gomez-Diaz, J. Lee, N. Nookala, M. A. Belkin, and A. Alí, “Gradient Nonlinear Pancharatnam-Berry Metasurfaces,” Phys. Rev. Lett. 115(20), 207403 (2015). [CrossRef]  

6. K. I. Willig, S. O. Rizzoli, V. Westphal, R. Jahn, and S. W. Hell, “STED microscopy reveals that synaptotagmin remains clustered after synaptic vesicle exocytosis,” Nature 440(7086), 935–939 (2006). [CrossRef]  

7. F. Tamburini, G. Anzolin, G. Umbriaco, A. Bianchini, and C. Barbieri, “Overcoming the Rayleigh Criterion Limit with Optical Vortices,” Phys. Rev. Lett. 97(16), 163903 (2006). [CrossRef]  

8. X. Xie, Y. Chen, K. Yang, and J. Zhou, “Harnessing the Point-Spread Function for High-Resolution Far-Field Optical Microscopy,” Phys. Rev. Lett. 113(26), 263901 (2014). [CrossRef]  

9. H. He, M. E. J. Friese, N. R. Heckenberg, and H. Rubinsztein-Dunlop, “Direct Observation of Transfer of Angular Momentum to Absorptive Particles from a Laser Beam with a Phase Singularity,” Phys. Rev. Lett. 75(5), 826–829 (1995). [CrossRef]  

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

11. S. C. Chapin, V. Germain, and E. R. Dufresne, “Automated trapping, assembly, and sorting with holographic optical tweezers,” Opt. Express 14(26), 13095 (2006). [CrossRef]  

12. Y. Yang, Y.-X. Ren, M. Chen, Y. Arita, and C. Rosales-Guzmín, “Optical trapping with structured light: a review,” Adv. Photonics 3(03), 034001 (2021). [CrossRef]  

13. M. Erhard, R. Fickler, M. Krenn, and A. Zeilinger, “Twisted photons: new quantum perspectives in high dimensions,” Light: Sci. Appl. 7(3), 17146 (2018). [CrossRef]  

14. X.-L. Wang, Y.-H. Luo, H.-L. Huang, M.-C. Chen, Z.-E. Su, C. Liu, C. Chen, W. Li, Y.-Q. Fang, X. Jiang, J. Zhang, L. Li, N.-L. Liu, C.-Y. Lu, and J.-W. Pan, “18-Qubit Entanglement with Six Photons’ Three Degrees of Freedom,” Phys. Rev. Lett. 120(26), 260502 (2018). [CrossRef]  

15. M. P. J. Lavery, C. Peuntinger, K. Günthner, P. Banzer, D. Elser, R. W. Boyd, M. J. Padgett, C. Marquardt, and G. Leuchs, “Free-space propagation of high-dimensional structured optical fields in an urban environment,” Sci. Adv. 3(10), e1700552 (2017). [CrossRef]  

16. J. A. Anguita and J. E. Cisternas, “Algorithmic decoding of dense OAM signal constellations for optical communications in turbulence,” Opt. Express 30(8), 13540 (2022). [CrossRef]  

17. L. Zhu, A. Wang, S. Chen, J. Liu, Q. Mo, C. Du, and J. Wang, “Orbital angular momentum mode groups multiplexing transmission over 26-km conventional multi-mode fiber,” Opt. Express 25(21), 25637 (2017). [CrossRef]  

18. B. de Lima Bernardo and F. Moraes, “Data transmission by hypergeometric modes through a hyperbolic-index medium,” Opt. Express 19(12), 11264 (2011). [CrossRef]  

19. Y. Li, K. Morgan, W. Li, J. K. Miller, R. Watkins, and E. G. Johnson, “Multi-dimensional QAM equivalent constellation using coherently coupled orbital angular momentum (OAM) modes in optical communication,” Opt. Express 26(23), 30969 (2018). [CrossRef]  

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

21. T. Omatsu, K. Miyamoto, and A. J. Lee, “Wavelength-versatile optical vortex lasers,” J. Opt. 19(12), 123002 (2017). [CrossRef]  

22. G. C. G. Berkhout, M. P. J. Lavery, J. Courtial, M. W. Beijersbergen, and M. J. Padgett, “Efficient Sorting of Orbital Angular Momentum States of Light,” Phys. Rev. Lett. 105(15), 153601 (2010). [CrossRef]  

23. G. C. G. Berkhout, M. P. J. Lavery, M. J. Padgett, and M. W. Beijersbergen, “Measuring orbital angular momentum superpositions of light by mode transformation,” Opt. Lett. 36(10), 1863 (2011). [CrossRef]  

24. Q. Zhao, M. Dong, Y. Bai, and Y. Yang, “Measuring high orbital angular momentum of vortex beams with an improved multipoint interferometer,” Photonics Res. 8(5), 745 (2020). [CrossRef]  

25. D. P. Ghai, P. Senthilkumaran, and R. Sirohi, “Single-slit diffraction of an optical beam with phase singularity,” Opt. Lasers Eng. 47(1), 123–126 (2009). [CrossRef]  

26. J. M. Hickmann, E. J. S. Fonseca, W. C. Soares, and S. Chívez-Cerda, “Unveiling a Truncated Optical Lattice Associated with a Triangular Aperture Using Light’s Orbital Angular Momentum,” Phys. Rev. Lett. 105(5), 053904 (2010). [CrossRef]  

27. A. Ambuj, R. Vyas, and S. Singh, “Diffraction of orbital angular momentum carrying optical beams by a circular aperture,” Opt. Lett. 39(19), 5475 (2014). [CrossRef]  

28. W. Z. Wuhong Zhang, Z. W. Ziwen Wu, J. W. Jikang Wang, and A. L. C. Lixiang Chen, “Experimental demonstration of twisted light’s diffraction theory based on digital spiral imaging,” Chin. Opt. Lett. 14(11), 110501 (2016). [CrossRef]  

29. Y. Bai, H. Lv, X. Fu, and Y. Yang, “Vortex beam: generation and detection of orbital angular momentum [Invited],” Chin. Opt. Lett. 20(1), 012601 (2022). [CrossRef]  

30. H. I. Sztul and R. R. Alfano, “Double-slit interference with Laguerre-Gaussian beams,” Opt. Lett. 31(7), 999 (2006). [CrossRef]  

31. H. Zhou, S. Yan, J. Dong, and X. Zhang, “Double metal subwavelength slit arrays interference to measure the orbital angular momentum and the polarization of light,” Opt. Lett. 39(11), 3173 (2014). [CrossRef]  

32. E. Hecht, Optics (Pearson Education, Inc., 2017), 5th ed.

33. K. A. Forbes and D. L. Andrews, “Orbital angular momentum of twisted light: chirality and optical activity,” JPhys Photonics 3(2), 022007 (2021). [CrossRef]  

34. L. Zhang, B. Shen, X. Zhang, S. Huang, Y. Shi, C. Liu, W. Wang, J. Xu, Z. Pei, and Z. Xu, “Deflection of a Reflected Intense Vortex Laser Beam,” Phys. Rev. Lett. 117(11), 113904 (2016). [CrossRef]  

35. V. G. Fedoseyev, “Spin-independent transverse shift of the centre of gravity of a reflected and of a refracted light beam,” Opt. Commun. 193(1-6), 9–18 (2001). [CrossRef]  

36. E. Wolf, Progress in Optics (Elsevier, 2001).

37. F. Ricci, W. Löffler, and M. van Exter, “Instability of higher-order optical vortices analyzed with a multi-pinhole interferometer,” Opt. Express 20(20), 22961 (2012). [CrossRef]  

38. K. Dai, J. K. Miller, W. Li, R. J. Watkins, and E. G. Johnson, “Fractional orbital angular momentum conversion in second-harmonic generation with an asymmetric perfect vortex beam,” Opt. Lett. 46(14), 3332 (2021). [CrossRef]  

39. V. Kotlyar, A. Kovalev, A. Porfirev, and E. Kozlova, “Orbital angular momentum of a laser beam behind an off-axis spiral phase plate,” Opt. Lett. 44(15), 3673 (2019). [CrossRef]  

40. J. B. Götte, S. Franke-Arnold, R. Zambrini, and S. M. Barnett, “Quantum formulation of fractional orbital angular momentum,” J. Mod. Opt. 54(12), 1723–1738 (2007). [CrossRef]  

41. J. Leach, E. Yao, and M. J. Padgett, “Observation of the vortex structure of a non-integer vortex beam,” New J. Phys. 6, 71 (2004). [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 (4)

Fig. 1.
Fig. 1. (a) Spatial dislocation $\Delta$ between the two reflect beams $E_1$, $E_2$ induced by the refraction of the illumination through the PCL; (b) cross-section schematic of the two reflect beams and the coordinate setup at the observing screen plane.
Fig. 2.
Fig. 2. Numerical calculations of the generalized Newton’s rings obtained from vortex beams with different topological charges. (a) $l=+1$, (b) $l=-1$, (c) $l=+2$, (d) $l=-2$, (e) $l=+3$, (f) $l=-3$. The number of spiral arms in the interference pattern indicates the topological charge $l$, and the fork-shaped dislocation (marked by the dashed green circles) indicates the topological charge $l$ as well, namely, one interference fringe splits into $|l|+1$ fringes.
Fig. 3.
Fig. 3. Schematic of the experimental setup. AS, aperture slot; PCL, plano-convex lens; SLM, reflective spatial light modulator; CCD, charge-coupled device.
Fig. 4.
Fig. 4. Experimental generalized Newton’s rings interference patterns obtained from (a) the $l=+1$ beam, (b) $l=+2$ beam, (c) $l=+3$ beam, (d) $l=+4$ beam.

Equations (6)

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

L G p l ( ρ , φ , z ) = ω 0 ω ( z ) 2 p ! π ( | l | + p ) ! ( 2 ρ ω ( z ) ) | l | L p | l | [ 2 ( ρ ω ( z ) ) 2 ] × exp [ i ( 2 p + | l | + 1 ) ξ ( z ) ] × exp [ ( ρ ω ( z ) ) 2 ] exp [ i k ρ 2 2 R ( z ) ] exp ( i l φ )
L G p l ( ρ , φ , z ) = A p l ( ρ , z ) exp ( i l φ )
I ( ρ , φ ) = | E 0 exp ( i k λ 2 ) + E 0 exp ( i k ρ 2 R ) | 2 cos 2 ( k ρ 2 2 R π 2 )
Δ = 2 d tan ( arcsin ( sin θ n ) ) cos θ
Δ 2 θ d n
I ( x , y ) | L G p , l ( x , y ) exp ( i k λ 2 ) + L G p , l ( x , y + Δ ) exp ( i k x 2 + ( y + Δ ) 2 R 2 ) | 2 | A p , l ( x , y ) exp ( i l φ ) exp ( i k λ 2 ) + A p , l ( x , y + Δ ) exp ( i l φ ) exp ( i k x 2 + ( y + Δ ) 2 R 2 ) | 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.