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

Reducing the cross-talk among different orbital angular momentum modes in turbulent atmosphere by using a focusing mirror

Open Access Open Access

Abstract

The cross-talk among different orbital angular momentum (OAM) modes induced by the turbulent atmosphere is a challenging effect commonly occurring in OAM-based free-space optical (FSO) communication. The aim of this study is to propose a simple method to reduce the crosstalk and demonstrate its effect by analytical derivation and numerical simulation. It is found that the crosstalk is largely reduced by using a focusing mirror. Our results will be useful in free-space optical communication.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

The vortex beams carrying OAM keep a subject of intense research because they have important applications in optical manipulation and trapping [1], imaging [2], quantum entanglement [3], free-space optical (FSO) communication [4], fiber optical communication [5] and so on. In particular, significant progress has been made in propagation of vortex beams in turbulent atmosphere. It is well known that the light beam will be disturbed by the random fluctuation of the refraction index of turbulent atmosphere [6–9]. Therefore, there are many OAM states at the received plane when the incident beam with single OAM state propagates in turbulent atmosphere. This is because the initial OAM state will seriously spread to neighbor OAM states. This effect is called as cross-talk of OAM state. How to reduce the cross-talk is still a challenge.

It should be emphasized that there are at least two kinds of manner to be considered to reach the aim. One is the adaptive optics and the other is using special beams. Although the adaptive optics is successfully used to compensate the orbital angular momentum of a vortex beam disturbed by turbulent atmosphere [10,11], its optical system is very complicated. For the special beams, such as the Whittaker-Gaussian beam [12], the autofocusing Airy beam [13], the tailored Airy vortex beam array [14], the Bessel beam and Laguerre–Gaussian beam [15], have been considered to use in FSO communication. They can partially mitigate the effects of turbulence on the detection probability of the signal OAM mode in the weak turbulence regime, but the generation of those beams need to some other complicated instruments. In the present paper, we propose a simple manner but with a good effect to reduce the cross-talk among different OAM modes in turbulent atmosphere.

The paper is organized as follows. Basic derivation for the OAM spectrum of a Gaussian vortex beam in weak turbulence is given in Sec. 2. In Sec. 3, some numerical results and discussion are showed to illustrate the effect of our method. In Sec. 4, the numerical simulation is used to confirm the obtained results. The conclusions of the present paper are given in Sec. 5.

 figure: Fig. 1

Fig. 1 Schematic of the OAM state disturbed by turbulent atmosphere after passing through a focussing mirror.

Download Full Size | PDF

2. Theoretical derivation

The light beam carried OAM propagating in turbulent atmosphere is illustrated in Fig. 1. A light beam with vortex phase passes through a focusing mirror and then propagates in turbulent atmosphere. The source plane is set just behind the mirror. Therefore, the beam profile and phase wavefront in the initial plane (z = 0) can be assumed as [16]

E(r,θ,0)=exp [(1+iC0)2R02r2+il0θ],
where (r,θ) is the polar coordinates and l0 is the quantum number of the OAM mode in the source plane. Here, R0 is initial beam size, C0=kR02/F is the initial beam prefocusing parameter and F is the focal distance or the distance from the source plane to the plane of the receiver, k=2π/λ is the wave number with wavelength λ. The constant C0 occurring in Eq. (1) shows the existence of a mirror. Now, let’s consider this beam propagates in free space. The complex amplitude of this beam in the receiver plane can be obtained with the help of the Huygens-Fresnel integral as
Efree(ρ,φ,z)=iλzexp (ikz)×E(r,θ,0)exp {ik2z[r2+ρ22rρcos (φθ)]}rdrdθ,
where (ρ,ϕ) is the polar coordinates. By substituting Eq. (1) into Eq. (2), and utilizing the following expression [17]
02πexp [inφ1ixcos (φ2φ1)]dφ1=2π(i)nJn(x)exp [inφ2],
one can obtain
Efree(ρ,φ,z)=2π(i)l0(iλz)exp (ikz)exp (il0φ)exp (ik2zρ2)×πkρ8zα3/2exp (k2ρ28αz2)[I0.5l00.5(k2ρ28αz2)I0.5l0+0.5(k2ρ28αz2)],
where α=12R02+iC02R02ik2z, Jn() and In() denote the nth order Bessel function and modified Bessel function, respectively.

Then, we consider this beam propagates in turbulent atmosphere. If we restrict ourselves to considering the weak turbulence regime, the cumulative effect of the turbulence over the propagation path can be thought as a pure phase perturbation on the beam at the output plane z. Using the Rytov approximation, the complex amplitude of the Gaussian vortex beam which propagates in weak turbulence at distance z can be represented as [18]

E(ρ,φ,z)=Efree(ρ,φ,z)exp [ψ(ρ,φ,z)],
where ψ(ρ,φ,z) is the complex phase perturbation of the field due to the atmospheric turbulence along the propagation channel.

In a quantum description, the complex amplitude E(ρ,φ,z) of one light beam can be expressed as the superpositions of states of different OAM [19], i.e.,

E(ρ,φ,z)=12πl=+al(ρ,z)exp (ilφ),
and the expansion coefficient is
al(ρ,z)=12π02πE(ρ,φ,z)exp (ilφ)dφ,
where l denotes the quantum number of the OAM mode in the z plane. Based on Eq. (7) and Eq. (5), in paraxial turbulent channels, the probability density |al(ρ,z)|2 of OAM mode of one vortex beam is given by the integral
|al(ρ,z)|2=12πEfree*(ρ,φ1,z)Efree(ρ,φ2,z)×exp[ψ*(ρ,φ1,z)+ψ(ρ,φ2,z)] exp[il(φ2φ1)]dφ1dφ2,
where the angular brackets represent ensemble average, exp [ψ*(ρ,φ1,z)+ψ(ρ,φ2,z)] is the phase correlation function, and can be given by [20] as
exp [ψ*(ρ,φ1,z)+ψ(ρ,φ2,z)]=exp [2ρ22ρ2cos (φ2φ1)ρ02]

In Eq. (9), ρ0 is the spatial coherence radius, and is given by [20]

ρ0=[π23k2z0κ3ϕn(κ)dκ]1/2.

Here, the Tatarskii spectrum for the spatial power spectrum of the refractive-index fluctuation is adopted as [21]

ϕn(κ)=0.033Cn2κ11/3exp (κ2/κm2),
where Cn2 is the structure constant of the refractive index fluctuation of the turbulence and κm=5.92/li with li being the inner scale of the turbulence. On substituting Eq. (11) into Eq. (10), then we can get
ρ0=(0.5466k2zCn2li1/3)1/2.

By substituting Eq. (4) and Eq. (9) into Eq. (8), after complicated integration, we read

|al(ρ,z)|2=π4k28z2|α1.5|2(1λz)2ρ2exp [k2ρ28α*z2k2ρ28αz2]exp [(2ρ02)ρ2]×Il0l[(2ρ02)ρ2]|[I0.5l00.5(k2ρ28αz2)I0.5l0+0.5(k2ρ28αz2)]|2.

The energy weight for each of the spiral harmonics of the Gaussian vortex beams is defined as the follows

ωl=0R|al(ρ,z)|2ρdρ,
where R is the radius of the receiving aperture. Then, we can get
ωl=π4k28λ2z4|α1.5|20Rρ3exp [qρ2]exp [tρ2]×Il0l[tρ2]|[I0.5l00.5(pρ2)I0.5l0+0.5(pρ2)]|2dρ,
where t=2ρ02, p=k28αz2, q=k28α*z2+k28αz2. The normalized energy weight of the OAM mode of the Gaussian vortex beams in the paraxial region can be expressed as
Pl=ωll=ωl,
where l=ωl is the total energy. In fact, the OAM spectrum is consisted of different Pl. Based on Eqs. (13), (15) and (16), we can numerically calculate the OAM spectrum.

3. Numerical results and discussion

Here some numerical results are used to illustrate a focusing mirror improving the propagating behavior of OAM. In the examples we set λ = 1060 nm, R0=0.02 m, Cn2=1015 m2/3, li=0.001 m, z=1000 m, R=0.05 m, unless specified otherwise.

 figure: Fig. 2

Fig. 2 The probability density of OAM mode for different values of C0 with F= (a)1 km,(b) infinity and (c) 1 km, respectively.

Download Full Size | PDF

Figure 2 shows the variation of the probability density of OAM mode as a function of radial axis for different values of C0 with l=1,2 and l0=1.

 figure: Fig. 3

Fig. 3 The received OAM spectrum for different quantum number. The first row and second row denote there exists a focusing mirror and does not exist a focusing mirror, respectively.

Download Full Size | PDF

It is found that the parameter C0 has obvious influence on the probability density. In particular, one can detect the OAM state (l = 1) with large probability and small space area when C0 is taken a positive value (see Fig. 2(c)). This situation forms strong contrast by comparing with that of C0=0 and C0<0 that correspond to the situation of no focusing mirror and of diverging mirror, respectively. At the same time, one can find that the probability difference of detecting the OAM state of l = 1 and l = 2 becomes large when we change C0 from negative value to zero and to positive value (see the filled area of Fig. 2). It means that the OAM mode crosstalk is reduced by using a focusing mirror. This can be understood as the beam’s interacted area with the turbulence decreases on propagation when the beam is focused. Therefore, adding a focusing mirror can improve the communication quality of the light beam in turbulent atmosphere. To directly see the influence of C0 on the OAM state in turbulent atmosphere, we will show the OAM spectrum of the Gaussian vortex beam in turbulent atmosphere.

 figure: Fig. 4

Fig. 4 The received OAM spectrum varying with different valued of R0 for C0=0 (the first row) and C00 (the second row).

Download Full Size | PDF

Figure 3 plots the normalized energy weight of OAM mode for different quantum number without a focusing mirror C0=0 (the first row) and with a focusing mirror C00 (the second row), respectively. From the first row, one can find the normalized energy weight of central OAM mode is 0.78, 0.7, 0.63 corresponding to l0=1, 2, 3, respectively. However, it rises to 0.9, 0.81, 0.74, respectively (see the second row). This value is larger which means the OAM mode is purer, i.e., the cross talk is weaker. Therefore, we can say that the existing of a focusing mirror can reduce the cross-talk among different OAM modes. We should point out that this is the main result of this paper. In addition, one can find that the rate of the normalized energy weight of central OAM mode growth is 15.4%, 15.7% and 17.5% corresponding to l0=1, 2, 3, respectively. That means the quantum number is larger the action of mirror is stronger.

 figure: Fig. 5

Fig. 5 The varying of the normalized energy weight of the central OAM mode (l=l0=3) as a function of R0 for C0=0 and C00.

Download Full Size | PDF

The influence of the beam size on the OAM spectrum is showed in Fig. 4. One can find that the distribution of the OAM spectrum varies with the changing of beam size (see the second row). For comparing, the OAM spectrum for different values of R0 without the mirror is plotted in the first row. Comparing the first row with the second row, we find that the mirror can reduce the cross-talk and the action of mirror with a large beam size is strong. For stressing the action of the mirror, the varying of the normalized energy weight of the central OAM mode as a function of R0 is plotted in Fig. 5. As reported in [20], for the beam without mirror, there exists an optimal value R0 for the normalized energy weight of the central OAM mode being maximum. However, this situation is different from that of the beam with a mirror that the curve slowly goes up to its asymptotic limit. This is because the increasing of R0 leads to the increasing of C0 and the probability density |al(ρ,z)|2 will increase.

 figure: Fig. 6

Fig. 6 The received intensity and phase (one realization) of the Gaussian vortex beam in turbulent atmosphere for C0=0 and C00.

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 The received OAM spectrum (averaged with 200 realizations) of the Gaussian vortex beam in turbulent atmosphere for C0=0 and C00.

Download Full Size | PDF

4. Numerical simulation

For illustrating the analytical results in above section, we simulate the Gaussian beam with vortex phase propagating in turbulent atmosphere by using the random phase screens [22]. In our simulation, the parameters are set as λ = 1060 nm, R0=0.06 m, l0=3, li=0.001 m, Cn2=1015 m2/3, F = 1 km. The size of the phase screen is set as 0.22 m with 512×512 pixels. The propagation distance is 1 km and with 13 screens. One can find the performance of turbulence resisting for a vortex beam with mirror is better than that of the beam without the mirror in Fig. 6. For further quantitative comparison of the crosstalk performance between the beam with a mirror (C00) and without a mirror (C0=0), the received OAM spectrum are depicted in Fig. 7. It is found that the crosstalk of the vortex beam without the mirror (C0=0) is more serious than that of the beam with a mirror (C00). This result agrees with the conclusion that we discussed in above section.

5. Conclusions

In summary, the propagation of a Gaussian vortex beam with focusing mirror in Kolmogorov turbulence has been derived. The probability density and the normalized energy weight of the received OAM mode were numerically calculated. It was found that the probability of detecting one OAM mode was enhanced and the cross talk of different OAM modes was reduced by adding a focusing mirror. In addition, the source beam size is larger, the action of mirror is stronger. Those results were also confirmed, in principle, by the numerical simulation. Our results will be useful for improving the quality of the FSO communication.

Funding

National Natural Science Fund for Distinguished Young Scholar (11525418); National Natural Science Foundation of China (NSFC) (11604264, 91750201); Natural Science Basic Research Plan in Shaanxi Province of China (Program No. 2017JQ6032); Science Foundation of Northwest University (15NW28).

References

1. H. He, M. Friese, N. 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, 826 (1995). [CrossRef]   [PubMed]  

2. J. A. Davis, D. E. McNamara, D. M. Cottrell, and J. Campos, “Image processing with the radial hilbert transform: theory and experiments,” Opt. Lett. 25, 99–101 (2000). [CrossRef]  

3. A. Mair, A. Vaziri, G. Weihs, and A. Zeilinger, “Entanglement of the orbital angular momentum states of photons,” Nature 412, 313 (2001). [CrossRef]   [PubMed]  

4. G. Gibson, J. Courtial, M. Padgett, M. Vasnetsov, V. Pas’Ko, S. Barnett, and S. Frankearnold, “Free-space information transfer using light beams carrying orbital angular momentum,” Opt. Express 12, 5448–5456 (2004). [CrossRef]   [PubMed]  

5. 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, 1545–1548 (2013). [CrossRef]   [PubMed]  

6. Y. Yuan, T. Lei, Z. Li, Y. Li, S. Gao, Z. Xie, and X. Yuan, “Beam wander relieved orbital angular momentum communication in turbulent atmosphere using Bessel beams,” Sci. Reports 7, 42276 (2017). [CrossRef]  

7. J. Wang, “Advances in communications using optical vortices,” Photonics Res. 4, B14–B28 (2016). [CrossRef]  

8. J. A. Anguita, M. A. Neifeld, and B. V. Vasic, “Turbulence-induced channel crosstalk in an orbital angular momentum-multiplexed free-space optical link,” Appl. Opt. 47, 2414–2429 (2008). [CrossRef]   [PubMed]  

9. M. Krenn, R. Fickler, M. Fink, J. Handsteiner, M. Malik, T. Scheidl, R. Ursin, and A. Zeilinger, “Communication with spatially modulated light through turbulent air across vienna,” New J. Phys. 16, 113028 (2014). [CrossRef]  

10. N. Li, X. Chu, P. Zhang, X. Feng, C. Fan, and C. Qiao, “Compensation for the orbital angular momentum of a vortex beam in turbulent atmosphere by adaptive optics,” Opt. & Laser Technol. 98, 7–11 (2018). [CrossRef]  

11. Y. Ren, G. Xie, H. Huang, C. Bao, Y. Yan, N. Ahmed, M. P. J. Lavery, B. I. Erkmen, S. Dolinar, and M. Tur, “Adaptive optics compensation of multiple orbital angular momentum beams propagating through emulated atmospheric turbulence,” Opt. Lett. 39, 2845–2848 (2014). [CrossRef]   [PubMed]  

12. Y. Zhang, M. Cheng, Y. Zhu, J. Gao, W. Dan, Z. Hu, and F. Zhao, “Influence of atmospheric turbulence on the transmission of orbital angular momentum for Whittaker-Gaussian laser beams,” Opt. Express 22, 22101–22110 (2014). [CrossRef]   [PubMed]  

13. X. Yan, L. Guo, M. Cheng, J. Li, Q. Huang, and R. Sun, “Probability density of orbital angular momentum mode of autofocusing Airy beam carrying power-exponent-phase vortex through weak anisotropic atmosphere turbulence,” Opt. Express 25, 15286–15298 (2017). [CrossRef]   [PubMed]  

14. X. Yan, L. Guo, M. Cheng, and J. Li, “Controlling abruptly autofocusing vortex beams to mitigate crosstalk and vortex splitting in free-space optical communication,” Opt. Express 26, 12605–12619 (2018). [CrossRef]   [PubMed]  

15. S. Fu and C. Gao, “Influences of atmospheric turbulence effects on the orbital angular momentum spectra of vortex beams,” Photonics Res. 4, B1–B4 (2016). [CrossRef]  

16. A. M. Rubenchik, M. P. Fedoruk, and S. K. Turitsyn, “Laser beam self-focusing in the atmosphere,” Phys. Rev. Lett. 102, 233902 (2009). [CrossRef]   [PubMed]  

17. I. S. Gradshteyn and I. M. Ryzhik, Table of integrals, series, and products, 6th ed. (Academic, 2000).

18. C. Paterson, “Atmospheric turbulence and orbital angular momentum of single photons for optical communication,” Phys. Rev. Lett. 94, 153901 (2005). [CrossRef]   [PubMed]  

19. G. Molina-Terriza, J. P. Torres, and L. Torner, “Management of the angular momentum of light: preparation of photons in multidimensional vector states of angular momentum,” Phys. Rev. Lett. 88, 4 (2002).

20. Y. Zhu, M. Chen, Y. Zhang, and Y. Li, “Propagation of the OAM mode carried by partially coherent modified Bessel–Gaussian beams in an anisotropic non-Kolmogorov marine atmosphere,” J. Opt. Soc. Am. A 33, 2277–2283 (2016). [CrossRef]  

21. F. Wang and Y. Cai, “Second-order statistics of a twisted Gaussian Schell-model beam in turbulent atmosphere,” Opt. Express 18, 24661–24672 (2010). [CrossRef]   [PubMed]  

22. J. D. Schmidt, Numerical simulation of optical wave propagation with examples in matlab,(SPIEBellingham, Washington, USA, 2010). [CrossRef]  

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 OAM state disturbed by turbulent atmosphere after passing through a focussing mirror.
Fig. 2
Fig. 2 The probability density of OAM mode for different values of C0 with F = (a) 1 km,(b) infinity and (c) 1 km, respectively.
Fig. 3
Fig. 3 The received OAM spectrum for different quantum number. The first row and second row denote there exists a focusing mirror and does not exist a focusing mirror, respectively.
Fig. 4
Fig. 4 The received OAM spectrum varying with different valued of R0 for C 0 = 0 (the first row) and C 0 0 (the second row).
Fig. 5
Fig. 5 The varying of the normalized energy weight of the central OAM mode ( l = l 0 = 3) as a function of R0 for C 0 = 0 and C 0 0.
Fig. 6
Fig. 6 The received intensity and phase (one realization) of the Gaussian vortex beam in turbulent atmosphere for C 0 = 0 and C 0 0.
Fig. 7
Fig. 7 The received OAM spectrum (averaged with 200 realizations) of the Gaussian vortex beam in turbulent atmosphere for C 0 = 0 and C 0 0.

Equations (16)

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

E ( r , θ , 0 ) = exp  [ ( 1 + i C 0 ) 2 R 0 2 r 2 + i l 0 θ ] ,
E f r e e ( ρ , φ , z ) = i λ z exp   ( i k z ) × E ( r , θ , 0 ) exp   { i k 2 z [ r 2 + ρ 2 2 r ρ cos   ( φ θ ) ] } r d r d θ ,
0 2 π exp   [ i n φ 1 i x cos   ( φ 2 φ 1 ) ] d φ 1 = 2 π ( i ) n J n ( x ) exp   [ i n φ 2 ] ,
E f r e e ( ρ , φ , z ) = 2 π ( i ) l 0 ( i λ z ) exp   ( i k z ) exp   ( i l 0 φ ) exp   ( i k 2 z ρ 2 ) × π k ρ 8 z α 3 / 2 exp   ( k 2 ρ 2 8 α z 2 ) [ I 0.5 l 0 0.5 ( k 2 ρ 2 8 α z 2 ) I 0.5 l 0 + 0.5 ( k 2 ρ 2 8 α z 2 ) ] ,
E ( ρ , φ , z ) = E f r e e ( ρ , φ , z ) exp  [ ψ ( ρ , φ , z ) ] ,
E ( ρ , φ , z ) = 1 2 π l = + a l ( ρ , z ) exp  ( i l φ ) ,
a l ( ρ , z ) = 1 2 π 0 2 π E ( ρ , φ , z ) exp  ( i l φ ) d φ ,
| a l ( ρ , z ) | 2 = 1 2 π E f r e e * ( ρ , φ 1 , z ) E f r e e ( ρ , φ 2 , z ) × exp [ ψ * ( ρ , φ 1 , z ) + ψ ( ρ , φ 2 , z ) ]   exp [ i l ( φ 2 φ 1 ) ] d φ 1 d φ 2 ,
exp   [ ψ * ( ρ , φ 1 , z ) + ψ ( ρ , φ 2 , z ) ] = exp   [ 2 ρ 2 2 ρ 2 cos   ( φ 2 φ 1 ) ρ 0 2 ]
ρ 0 = [ π 2 3 k 2 z 0 κ 3 ϕ n ( κ ) d κ ] 1 / 2 .
ϕ n ( κ ) = 0.033 C n 2 κ 11 / 3 exp  ( κ 2 / κ m 2 ) ,
ρ 0 = ( 0.5466 k 2 z C n 2 l i 1 / 3 ) 1 / 2 .
| a l ( ρ , z ) | 2 = π 4 k 2 8 z 2 | α 1.5 | 2 ( 1 λ z ) 2 ρ 2 exp   [ k 2 ρ 2 8 α * z 2 k 2 ρ 2 8 α z 2 ] exp   [ ( 2 ρ 0 2 ) ρ 2 ] × I l 0 l [ ( 2 ρ 0 2 ) ρ 2 ] | [ I 0.5 l 0 0.5 ( k 2 ρ 2 8 α z 2 ) I 0.5 l 0 + 0.5 ( k 2 ρ 2 8 α z 2 ) ] | 2 .
ω l = 0 R | a l ( ρ , z ) | 2 ρ d ρ ,
ω l = π 4 k 2 8 λ 2 z 4 | α 1.5 | 2 0 R ρ 3 exp   [ q ρ 2 ] exp   [ t ρ 2 ] × I l 0 l [ t ρ 2 ] | [ I 0.5 l 0 0.5 ( p ρ 2 ) I 0.5 l 0 + 0.5 ( p ρ 2 ) ] | 2 d ρ ,
P l = ω l l = ω l ,
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.