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

Engineering a square truncated lattice with light's orbital angular momentum

Open Access Open Access

Abstract

We engineer an intensity square lattice using the Fraunhofer diffraction of a Laguerre-Gauss beam by a square aperture. We verify numerically and experimentally that a perfect optical intensity lattice takes place only for even values of the topological charge. We explain the origin of this behavior based on the decomposition of the patterns. We also study the evolution of the lattice formation by observing the transition from one order to the next of the orbital angular momentum varying the topological charge in fractional steps.

©2011 Optical Society of America

1. Introduction

Vortices are a unique system which exhibits remarkable behavior in various branches of physics such as Bose condensates, superconductors, superfluids, fluid flow, and optics [1,2]. Vortices are characterized by a topological charge (TC) whose signal is associated with the direction of circulation. In particular, in optics, this circulation or orbital angular momentum (OAM) is related to the optical phase profile of the optical field in a plane orthogonal to the light’s propagation direction. Such fields possess a well-defined net OAM value and can be written in polar coordinates as E(r,φ)=E0(r)exp(imφ), where m is an integer that we will call net TC. Note that such fields usually possess an optical axis centered phase singularity. The wave front of this field is composed of m intertwined helical surfaces that result in an annular intensity cross-section, with a handedness given by the sign ofm.

Optical lattice formation is currently a very active research area. It has been studied in various media such as nonlinear media [35], Bose-Einstein condensates [6], and periodic photonic structures [2]. It has also been studied in free space through the interference of three plane waves [79]. Recently, optical intensity lattice formation associated with the apertures using light possessing OAM has been observed using, for example, a equilateral triangle [10] and a multipoint interferometer [1113].

In this paper, we study the optical intensity lattice formation through the diffraction of light possessing OAM by a square aperture. In contrast to the work in ref [10], in which a well-shaped hexagonal truncated intensity lattice is always generated for any value of OAM, we find here that a perfect truncated square intensity lattice is formed only for even values of the TC. For odd values of the TC the lattice is not very well formed. We study both cases theoretically and experimentally, as well as the influence of fractional topological charges.

2. Theoretical results

We determine the Fraunhofer diffraction pattern in the far field region of a beam carrying OAM scattered using a square aperture. If we are interested only in the relative intensities at a fixed plane placed at the positionz=z0, the diffracted field Ed is given by the integral [14]:

Ed(k)=+τ(r)Ei(r)eikrdr,

In this integral, the far field distribution Ed(k) is obtained from the Fourier transform of the product of the function describing the square aperture τ(r) and the incident field Ei(r). Note that the transverse wavevector k can be associated with the coordinate system of the far field region playing the role of reciprocal space.

The integral shown in Eq. (1) was numerically evaluated using high-order LG beams as the initial condition for the electric field. For LG beams withp=0, the Fraunhofer diffraction patterns for different values of the TC ranging from 1 to 12 are shown in Fig. 1 . We can observer that a square intensity lattice forms as mincreases but only for even values of the TC with m=2n and n. For odd values ofm, the maxima are not well defined. It can be seen that there is a relation between the number of lateral spots N and the TC, namely m=2N2, but only for even values of m.

 figure: Fig. 1

Fig. 1 Diffraction patterns corresponding to the numerical results of Eq. (1).

Download Full Size | PDF

To understand the formation of the intensity lattices, we analyze the phase diagrams in the Fourier plane depicted in Fig. 2(a) and 2(b) for m = 5 and m = 6, respectively. Note that the phase has a uniform distribution, mainly in the central region, only in Fig. 2(b). Phase jumps, which form a square shape, are clearly observed in Fig. 2(b). In contrast, in some parts in the center of Fig. 2(a) (in which m has an odd value), the phase distribution is sufficiently smooth that it is impossible to clearly identify the phase jumps. Similar behavior is observed for all phase diagrams for the various values of m, in accordance with the amount and the parity of OAM.

 figure: Fig. 2

Fig. 2 Phase patterns corresponding to the diffraction patterns in Fig. 1, for m = 5 (a) and m = 6 (b). The red dashed square was used only to highlight the center of the pattern.

Download Full Size | PDF

3. Experimental setup

The setup used to perform the experiment is depicted in Fig. 3(a) . An Nd:YAG laser operating at 532 nm illuminates a pixilated computer hologram written with a Hamamatsu model X10468-01 spatial light modulator (SLM) to produce high-order LG modes.

 figure: Fig. 3

Fig. 3 (a) Experimental setup; (b) A beam inside of the square aperture (top) and the phase diagram (bottom) for m = 3. In the figure F is a density neutral filter; fi are lenses; SLM is the spatial light modulator; and SF is a spatial filter.

Download Full Size | PDF

Because the LG beam radius increases with m, we used square masks with sides varying from 1.8 mm to 3.3 mm for different LG beams. These masks can be superimposed over the LG hologram in the SLM. The effect is similar to an LG beam incident in a square aperture as illustrated in Fig. 3(b). Note that we have carefully aligned the beam at the center of the aperture to avoid asymmetries in the patterns. We have used a hologram type 1 for coding phase and amplitude as first proposed by Kirk et al [15] and first used for LG beams by Leach et al [16]. The Fourier transform was implemented by a 50-cm lens (f3) and the collected diffracted light was imaged by a 20-mm lens (f4) in a Charge Coupled Device (CCD) camera.

4. Experimental results and discussion

Panels presented in Fig. 4 show experimental results for the diffraction pattern of a LG beam by a square aperture for m from 0 to 12, −1, −2, −7, and −8. We observe a well-formed square optical lattice only for even values of m, confirming the numerical results. The sign of m does not change globally the shape of the pattern because the phase is invariant by π rotation for a square aperture. Naturally, for m = 0 we have the usual square aperture diffraction pattern.

 figure: Fig. 4

Fig. 4 Diffraction patterns corresponding to experimental results for integer topological charges.

Download Full Size | PDF

The experimental results shown in Fig. 4 can be obtained from summing the contributions of single slits and double slits, a result that in quantum optics is known as Born’s rule [17]. For simplicity we can associate a slit to each edge of the square. Note that both a square aperture and a square slit have very similar diffraction patterns.

Figure 5 shows the diffraction patterns for all slit combinations with m from 2 to 6. The overall pattern PABCD is obtained through the following pattern summation [17]:

PABCD=PAB+PAC+PAD+PBC+PBD+PCD2PA2PB2PC2PD,
where A, B, C, and D represent each slit of the square and Pi and Pij, with i,j=A,B,C,D, represent the patterns due to each slit and each pair of slits, respectively. We can observe a very good agreement of the right side column with Fig. 1 and 3.

 figure: Fig. 5

Fig. 5 Diffraction patterns produced by combinations of slits. Each slit combination is shown at the top.

Download Full Size | PDF

The patternsPi, associated with the single slits, show that when m increases the number of fringes increases and all patterns are shifted as expected [18]. By contrast, the patterns Pij, from the pairs of slits exhibit very different behaviors depending if they orthogonal or parallel. For the orthogonal slit configuration, it can be seen that the patterns comprise intensity peaks and the number of these peaks increases with the value of m but without any TC parity dependence. However, for the parallel slit configuration, it can be seen that there are two types of patterns, one for even values of m and the other for odd values of m. Such dependence on the TC parity can be associated to the fact that there is an odd or even multiple of πphase difference between the opposite slits for odd or even TC, respectively. This observation indicates that the patterns PAC andPBD have an important role in the optical intensity lattice formation processes that are dependent on the parity of m.

In Fig. 6 , we superimpose the intensity patterns, PAC and PBD, for m = 5 and m = 6. In the first case, we observe that the intensity peaks inPAC and PBDdo not match. However, for m = 6 the peaks in the center of each pattern coincide. For even values of m this effect produces a shaped optical intensity lattice when all contributions of all patterns are taken into account. For odd values of m this effect smears out the optical lattice.

 figure: Fig. 6

Fig. 6 Superimposed patterns, PAC (intensity) and PBD (intensity contours).

Download Full Size | PDF

Note that for the cases in which the intensity patterns PAC and PBD match in the center, the phase corresponding to the square optical intensity lattice will be similar to the one shown in Fig. 2(b).

Finally, we analyzed the evolution of the pattern for the azimuthal profile rmeimϕfor fractional TC, varying m in increments of 0.1. Note that the use of an azimuthal profile instead of a true LG beam is a good approximation because the potential functions and the Laguerre have approximately the same behavior at the edges of the square. In the panels of Fig. 7 , we see the results of such simulation. It is clear that there is no well-formed lattice pattern between two patterns from two consecutive even values of m for any fractional TC.

 figure: Fig. 7

Fig. 7 Diffraction patterns corresponding to the experimental results for the fractional topological charges

Download Full Size | PDF

5. Conclusions

We studied the Fraunhofer diffraction of LG beams by a square aperture. We observed that for even topological charge values, a truncated square optical intensity lattice, which is composed of a set of rotationally-symmetric intensity peaks, appears. By contrast, the resulting pattern is washed-out for odd topological charge values. To understand this diffraction pattern formation, we focused on the edge of the square rather than on the square aperture. Because the interference comes from the pairs, we decomposed the diffraction patterns of the square slit in patterns that originated from each slit and pair of slits. We found that the patterns that result from parallel slit configurations are responsible for two different types of patterns: one for odd and the other for even values of topological charge. In contrast to the case of odd values of the topological charge, for even topological charge values there is a good match between the intensity maxima in the center of the pattern. This is reflected in the formation of a square-truncated optical intensity lattice. Finally, we showed that there is a continuous transition between the two lattices that correspond to two consecutive even values of m, and there is no intermediate lattice between them for fractional TC.

Acknowledgments

The authors are thankful for the financial support from Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES) Pró-equipamentos/Programa Nacional de Cooperação Acadêmica (PROCAD)/PROCAD-Ação Novas Fronteiras, Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq)/ Ministério da Ciência e Tecnologia (MCT), Programa de Apoio a Núcleos de Excelência (Pronex)/Fundação de Amparo à Pesquisa do Estado de Alagoas (FAPEAL), FAPEAL, Institutos Nacionais de Ciência e Tecnologia (INCT)—Fotônica para Telecomunicações, and INCT–Informação Quântica.

References and links

1. L. M. Pismen, Vortices in Nonlinear Fields: From Liquid Crystals to Superfluids, from Non-Equilibrium Patterns to Cosmic Strings (Oxford University Press, Oxford, 1999).

2. Y. S. Kivshar, and G. P. Agrawal, Optical solitons: from fibers to photonic crystals (Academic Press, Amsterdam; Boston, 2003).

3. V. Tikhonenko, J. Christou, and B. Lutherdaves, “Spiraling bright spatial solitons formed by the breakup of an optical vortex in a saturable self-focusing medium,” J. Opt. Soc. Am. B 12(11), 2046–2052 (1995). [CrossRef]  

4. V. Tikhonenko, J. Christou, and B. Luther-Davies, “Three dimensional bright spatial soliton collision and fusion in a saturable nonlinear medium,” Phys. Rev. Lett. 76(15), 2698–2701 (1996). [CrossRef]   [PubMed]  

5. D. V. Petrov, L. Torner, J. Martorell, R. Vilaseca, J. P. Torres, and C. Cojocaru, “Observation of azimuthal modulational instability and formation of patterns of optical solitons in a quadratic nonlinear crystal,” Opt. Lett. 23(18), 1444–1446 (1998). [CrossRef]   [PubMed]  

6. J. R. Abo-Shaeer, C. Raman, J. M. Vogels, and W. Ketterle, “Observation of vortex lattices in Bose-Einstein condensates,” Science 292(5516), 476–479 (2001). [CrossRef]   [PubMed]  

7. J. Masajada, A. Popiolek-Masajada, and M. Leniec, “Creation of vortex lattices by a wavefront division,” Opt. Express 15(8), 5196–5207 (2007). [CrossRef]   [PubMed]  

8. J. Masajada and B. Dubik, “Optical vortex generation by three plane wave interference,” Opt. Commun. 198(1-3), 21–27 (2001). [CrossRef]  

9. J. Masajada, A. Popiolek-Masajada, and D. M. Wieliczka, “The interferometric system using optical vortices as phase markers,” Opt. Commun. 207(1-6), 85–93 (2002). [CrossRef]  

10. J. M. Hickmann, E. J. S. Fonseca, W. C. Soares, and S. Chávez-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]   [PubMed]  

11. R. W. Schoonover and T. D. Visser, “Creating polarization singularities with an N-pinhole interferometer,” Phys. Rev. A 79(4), 043809 (2009). [CrossRef]  

12. G. C. G. Berkhout and M. W. Beijersbergen, “Method for probing the orbital angular momentum of optical vortices in electromagnetic waves from astronomical objects,” Phys. Rev. Lett. 101(10), 100801 (2008). [CrossRef]   [PubMed]  

13. G. C. G. Berkhout and M. W. Beijersbergen, “Using a multipoint interferometer to measure the orbital angular momentum of light in astrophysics,” J. Opt. A, Pure Appl. Opt. 11(9), 094021 (2009). [CrossRef]  

14. J. W. Goodman, Introduction to Fourier Optics (McGraw-Hill, New York, 1996).

15. J. P. Kirk and A. L. Jones, “Phase-only complex-valued spatial filter,”, J. Opt. Soc. Am. B 61, 1023–1028 (1971).

16. J. Leach, M. Dennis, J. Courtial, and M. Padgett, “Vortex knots in light,” New J. Phys. 7, 55 (2005). [CrossRef]  

17. U. Sinha, C. Couteau, T. Jennewein, R. Laflamme, and G. Weihs, “Ruling out multi-order interference in quantum mechanics,” Science 329(5990), 418–421 (2010). [CrossRef]   [PubMed]  

18. Q. S. Ferreira, A. J. Jesus-Silva, E. J. S. Fonseca, and J. M. Hickmann, “Fraunhofer diffraction of light with orbital angular momentum by a slit,” Opt. Lett. 36(16), 3106–3108 (2011). [CrossRef]   [PubMed]  

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 Diffraction patterns corresponding to the numerical results of Eq. (1).
Fig. 2
Fig. 2 Phase patterns corresponding to the diffraction patterns in Fig. 1, for m = 5 (a) and m = 6 (b). The red dashed square was used only to highlight the center of the pattern.
Fig. 3
Fig. 3 (a) Experimental setup; (b) A beam inside of the square aperture (top) and the phase diagram (bottom) for m = 3. In the figure F is a density neutral filter; fi are lenses; SLM is the spatial light modulator; and SF is a spatial filter.
Fig. 4
Fig. 4 Diffraction patterns corresponding to experimental results for integer topological charges.
Fig. 5
Fig. 5 Diffraction patterns produced by combinations of slits. Each slit combination is shown at the top.
Fig. 6
Fig. 6 Superimposed patterns, P AC (intensity) and P BD (intensity contours).
Fig. 7
Fig. 7 Diffraction patterns corresponding to the experimental results for the fractional topological charges

Equations (2)

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

E d ( k )= + τ( r ) E i ( r ) e i k r d r ,
P ABCD = P AB + P AC + P AD + P BC + P BD + P CD 2 P A 2 P B 2 P C 2 P D ,
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.