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

Probability density of orbital angular momentum mode of autofocusing Airy beam carrying power-exponent-phase vortex through weak anisotropic atmosphere turbulence

Open Access Open Access

Abstract

The probability densities of orbital angular momentum (OAM) modes of the autofocusing Airy beam (AAB) carrying power-exponent-phase vortex (PEPV) after passing through the weak anisotropic non-Kolmogorov turbulent atmosphere are theoretically formulated. It is found that the AAB carrying PEPV is the result of the weighted superposition of multiple OAM modes at differing positions within the beam cross-section, and the mutual crosstalk among different OAM modes will compensate the distortion of each OAM mode and be helpful for boosting the anti-jamming performance of the communication link. Based on numerical calculations, the role of the wavelength, waist width, topological charge and power order of PEPV in the probability density distribution variations of OAM modes of the AAB carrying PEPV is explored. Analysis shows that a relatively small beam waist and longer wavelength are good for separating the detection regions between signal OAM mode and crosstalk OAM modes. The probability density distribution of the signal OAM mode does not change obviously with the topological charge variation; but it will be greatly enhanced with the increase of power order. Furthermore, it is found that the detection region center position of crosstalk OAM mode is an emergent property resulting from power order and topological charge. Therefore, the power order can be introduced as an extra steering parameter to modulate the probability density distributions of OAM modes. These results provide guidelines for the design of an optimal detector, which has potential application in optical vortex communication systems.

© 2017 Optical Society of America

1. Introduction

The atmosphere can be divided into the following two parts:1) the free atmosphere (above the atmospheric boundary layer) and 2) the atmospheric boundary layer (up to 1–2 km in altitude) [1]. In the beginning, optical turbulence inside the atmospheric boundary layer is considered isotropic, and the isotropic Kolmogorov spectrum model is generally correct within the inertial subrange. However, a wealth of experimental data and literatures [2–5] have shown that atmospheric turbulence is not flat and horizontal due to the presence of very strong positive temperature gradients in the atmosphere [1]. When a laser beam propagates through the anisotropic atmospheric turbulence, it undergoes extra beam spreading, scintillation, beam wander, and angle-of-arrival fluctuation. These unwanted effects strongly degrade the performance in numerous applications such as remote optical sensing, imaging, laser radar and free-space laser communications [6, 7]. Therefore, the persistent challenge driving most research activities is to mitigate the turbulence-induced degradation.

The key point in investigating these issues is determining how to reduce the diffraction. As of today, various methods have been proposed [8–11]. One method is to use a type of beams with special propagation characteristics (i.e., non-diffracting and self-healing) that leaves their transverse intensity distribution quasi-invariant during propagation despite of the severity of the imposed perturbations. Examples of such non-diffracting and self-healing beams include the Bessel, Laguerre-Gaussian and Airy beams.Among these non-diffracting and self-healing beams [12], has shown that the Airy beam can be controlled with ease, and the peak beam intensity can be delivered to and repositioned on a given target after propagating through disordered or turbulent media. This result opens the possibilities for controlling beam propagation and navigation in turbulent media [12].

It is well known that phase plays an important role in the quality of beam propagation. Optical vortex beams described by a phaseexp(ilφ), which carries an orbital angular momentum (OAM) of l per photon, can resist turbulence-induced distortions [13, 14]. The vortex Airy beam, an optical vortex embedded within an Airy beam, has therefore attracted the attention of many scientists. In recent years, considerable work has been done to investigate the properties of the vortex Airy beam. The beam spreading of a ring Airy beam vortex beam traveling through an anisotropic vertical atmospheric turbulence is analytically explored and numerically calculated [15]. The beam wander of an Airy beam with a spiral phase in a turbulent atmosphere is investigated in detail. Furthermore, the expressions for the second-order moment, the second central moments are derived [16]. The performance of an vortex Airy beam propagating through moderate-to-strong maritime turbulence is simulated and the spatial coherence radius is derived based on the modified Rytov approximation [17]. The propagation properties of a sharply autofocused ring Airy Gaussian vortex beams is numerically investigated by appropriately selecting the distribution factor [18]. The properties of partially coherent Airy beam propagating in turbulent media is studied [19]. The imposed vortices are able to take advantage of the energy of an Airy beam’s intensity peak [20]. In the papers mentioned above, the optical vortices imposed on an Airy beam are always the canonical vortices in which the phase gradient is constant. Since the propagation dynamics of an optical vortex are influenced by the phase function itself, type of noncanonical vortex that carries a spiral phase that varies nonuniformly with the azimuthal angle is investigated to explore the properties and applications of OAM [21, 22]. More interesting than canonical vortex, the anisotropy of the phase profile can be used as a steering parameter of optical vortex [21].

The power-exponent-phase vortex (PEPV) beam is a kind of noncanonical vortex that is characterized by a power-exponent-phase. PEPV was proposed by Li [23], and the autofocusing Airy beam (AAB) carrying PEPV have been experimentally demonstrated. More recently, Lao imposed a PEPV into a Gaussian beam and investigated the properties of the intensity of such a beam on propagation in free space [24]. However, to the best of our knowledge, the AAB carrying PEPV has not yet been used in anisotropic atmospheric turbulence.

In this paper, the Rytov theory is employed to derive an analytical formula for the radial probability density of the OAM mode of the AAB carrying PEPV propagating in the weak anisotropic non-Kolmogorov turbulent atmosphere. Considering the needs for precise tracking and signal reception in certain atmospheric experiments [25], the radial probability density distribution of signal OAM mode and crosstalk OAM modes of the AAB carrying PEPVs are investigated in detail. We also explore the influence of the power order n on the changes in the mode probability density of the OAM modes. We expect that the work will be particularly meaningful in practical applications for free-space optics communication. In Section 2, the model of the probability density for the AAB carrying PEPV in weak non-Kolmogorov turbulence is established. In Section 3, numerical simulations and graphical outputs related to the effects of topological charges, power order, waist width, and wavelength on the probability densities of signal and crosstalk OAM modes of the AAB carrying PEPV along the radial direction are investigated. Conclusions are presented in Section 4.

2. Theoretical model of the AAB carrying PEPV and its propagation in the turbulent atmosphere

Optical vortex beam for communication is to use OAM modes to encode information for communication [26]. However recent works have also shown that [27, 28], the transmitted signal carrying the particular OAM mode launched by the source may actually be susceptible to spatial aberrations after propagation through atmospheric turbulence. The reason is the turbulence induces OAM-mode scattering and leaks the power contained in a given OAM mode from other neighboring OAM modes. So, the received signal actually includes multiple OAM modes, as indicated in Fig. 1. Therefore, to known the precise probability density distributions of OAM modes at the receiver plane is of great significance. In this section, the probability densities of the OAM modes of the AAB carrying PEPV propagating in the weak anisotropic non-Kolmogorov turbulent atmosphere are formulated under a given generalized anisotropic von Karman spectrum based on the Rytov theory.

 figure: Fig. 1

Fig. 1 Schematic diagram of the effects of atmospheric turbulence on an OAM beam. An ideal OAM mode can be break down into multiple OAM modes.

Download Full Size | PDF

The electric fields of the AABs carrying PEPV at the source plane can be expressed as [23]:

E(0)(r,φ,z=0)=Ai(r0rω)exp(ar0rω)exp[i2vπ(φ2π)n],
here, (r,φ) denotes the radial coordinates, and φ ranges from 0 to 2π; Ai denotes the Airy function; r0 and ω refer, respectively, to the radius and width of the main lobe of the AAB; 0a1 is the exponential truncation factor; i=1; v denotes the topological charge of the vortex; n, be either an integer or a fraction, determines the power order of the spiral phase; and v and n collectively encode the angular rotation rate of the wave front. The patterns of the spiral phase for generating the PEPV beam of different topological charges and power orders are displayed in Fig. 2. Figure 2(d) is the same as that given in [24], which verifies the results validity.

 figure: Fig. 2

Fig. 2 Spiral phase patterns of the PEPV beam at the initial plane with different topological charges and power orders. (a) v = 3, n = 1; (b) v = 3, n = 2;(c) v = 3, n = 5;(d) v = 6, n = 2.

Download Full Size | PDF

Within the validity of the paraxial approximation, the electric field distribution of the beam in the receiver plane can be determined, based on the Rytov theory [29], from the knowledge of the initial electric field distribution:

E(ρ,θ,z)=ik2πzexp(ikz)drdφE(0)(r,φ,z=0)×exp{ik[ρ2+r22ρrcos(θφ)]2z+ψ(ρ,θ,z)},
where k=2π/λ is the wavenumber of beam, λ is the wavelength of the AAB in vacuum; ψ(ρ,θ,z), symbolized the random factor in the Rytov method, represents the random part of the complex phase added to the conventional phase of each spherical wave propagating from the input plane to the receiver plane [30].

On substituting from Eq. (1) into Eq. (2), we found the diffraction integral in Eq. (2) cannot be exactly evaluated. P. Zhang [31] develops an accurate description of this integral by introduce a delta ring of amplitude A0, e.g., E(0)(r,φ,z=0)A0δ(r0r). In this situation, atρ=0, the amplitude A0 can be expressed as [31]:

A0ωexp(a33)(1ωa2r0).

By the recall of the formulas of [32]

exp(z2(t1t))=l=Jl(z)tl,Jl(z)=(1)lJl(z),
where Jl() is the Bessel function of the first kind with the integer order l. By considering the Eqs. (1)-(4), and taking the integrations, we finally obtain the AAB carrying PEPV, at the receiver plane in turbulence-free channel and in the paraxial channel, has the analytical expression:
E(ρ,θ,z)=ik2πzexp(ikz)w(r0wa2)exp(ikρ2+ikr022z+a33)×[l=(i)lexp(ilθ)MlJl(kr0ρz)]exp[ψ(ρ,θ,z)],
where [24]

Ml=02πexp[i(2vπ(φ2π)n+lφ)]dφ={j=0h=0ij+hvjlh(2π)j+h+1j!h!(nj+h+1),l0j=0ijvj(2π)j+1j!(nj+1),l=0.

Note that in Eq. (5), the AAB carrying PEPV with a given topological charge v is, in fact, the result of the weighted superposition of multiple propagating OAM beams with phaseexp(ilθ), where l can take a series of integer values, at differing positions within the beam cross-section. As beam wave propagates through the turbulence, the effect of the refractive index fluctuations perturbs the phase of the wave so that it cannot be studied deterministically. Only statistical averages can be considered. In the present cases, we are keen on the average probability density distribution at the receiver plane, which will be discussed below.

The function E(ρ,θ,z) can be written as a superposition of plane waves with phase exp(imθ)as follows [33]:

E(ρ,θ,z)=12πmβm(ρ,z)exp(imθ),
where βm(ρ,z) is given by the integral

βm(ρ,z)=12π02πE(ρ,θ,z)exp(imθ)dθ.

By combining Eqs. (5)-(8), the mode probability density of the AAB carrying PEPV in paraxial channel is given by:

|βm(ρ,z)|2=12π02π02πE(ρ,θ,z)E(ρ,θ,z)exp[im(θθ)]×exp[ψ(ρ,θ,z)+ψ(ρ,θ,z)]dθdθ,
where the angular brackets implies the ensemble average over atmospheric turbulence medium, and asterisk stand for the complex conjugate. The term of the statistical ensemble average over the random complex phase in Eq. (9) is given by the expression [17, 34, 35]:
exp[ψ(ρ,θ,z)+ψ(ρ,θ,z)]=exp(D(ρ,θ,ρ,θ)2)exp(ρ2+ρ22ρρcos(θθ)ρ02),
where D(ρ,θ,ρ,θ) is the wave structure of the random complex phase in Rytov’s method; ρ0is the spatial coherence length of spherical wave propagation in anisotropic non-Kolmogorov turbulence. Here, the higher-order terms are neglected in the Rytov’s phase structure function to obtain the analytical results [30]. In this paper, we use the generalized anisotropic von Karman spectrum to describe the anisotropic non-Kolmogorov turbulence, and takes the form [34]:
ρ0={ξx2+ξy22ξx2ξy2π2k2zA(α)C˜n2/[6(α2)]×[κ2αγexp(κ02κ2)Γ(2α2,κ02κ2)2κ04α]}1/2,
where the non-Kolmogorov power spectrum index α with 3<α<4; ξx and ξy are two anisotropy parameters representing scale-dependent stretching along the x and y directions respectively; Γ() symbols the Gamma function; These parameters C˜n2 (being the effective structure constant with units m3α), A(α), κ0, γ and κ are the same as those given in [34].

On substituting Eq. (10) into Eq. (9), and Making use of the integral formula [36]:

02πexp[im(θθ)]exp[ηcos(θθ)]dθ=2πexp[imθ]Im(η),
where I() is the Bessel function of second kind with m order, thus, we obtain the more general integral formula for the average probability distribution at the receiver plane as follows:

|βm(ρ,z)|2=2π[(kw2πz)2(r0wa2)2exp(2a332ρ2ρ02)]×{l=1[(l=1(Ml+Ml)(Ml+Ml)+M0(Ml+Ml))Jl(kr0ρz)2Ilm(2ρ2ρ02)]+[M0l=1(Ml+Ml)+M0M0]J0(kr0ρz)2Im(2ρ2ρ02)}.

Equation (13) provides a quantitative description to the anisotropic turbulence induced changes in the compositions of OAM modes of the received AAB carrying PEPV.

In our study, v represents the signal OAM mode. For the designated OAM mode m = v, |βm(ρ,z)|2 is defined as the detection probability density of the signal OAM mode at the receiver. A larger value of |βm(ρ,z)|2 means more power remains in the signal OAM mode with index m [37]. For Δm=|vm|, |βΔm(ρ,z)|2 is defined as the crosstalk probability density, which represents the probability that the transmitted OAM state changed from the transmitted channel to the other channels. (This disagrees with [38], which argues that Δm=|lm| denotes the OAM state change from the signal channel to the other channels. It is necessary to be clear that Δm=|lm|, here, has the same meaning as Δl=|ll0| shown in [38]. As described in the literature [9], OAM modes with different l are mutually orthogonal. As is shown in Eq. (5), the AAB carrying PEPV is the result of the weighted superposition of multiple propagating OAM beams at differing positions within the beam cross-section. On this basis, the propagation behavior of the AAB carrying PEPV in turbulence can be modeled by the superposition of the components with appropriate weights. For an isolated signal OAM mode of the basis set, as illustrated in Fig. 1, the signal purity will be reduced by intermodal crosstalk among data channels induced by turbulence. However, for all components, the mutual crosstalk among different channels will compensate for the distortion of each signal OAM mode and will be helpful in allowing the distributive characteristics of these components to get close to the initial transmitted state. The schematic diagram of the propagation behavior of the AAB carrying PEPV is shown in Fig. 3. We just mention a few components of the basis as an examples (l = 1, 2, 3, 11 and 20) and other values can also be adopted. The component weight mappings are colored differently in this figure to indicate the different component channels.

 figure: Fig. 3

Fig. 3 Concept of turbulent atmosphere communication link using PEPV. (a) The turbulence resistance of the beam carrying PEPV. (b) OAM modes weight mapping. The beam carrying PEPV is the result of the weighted superposition of multiple propagating OAM beams at differing positions within the beam cross-section. And the mutual crosstalk among different OAM modes will compensate the distortion of each signal OAM mode and be helpful to let the distributive characteristics of these components get close to the initial state.

Download Full Size | PDF

3. Numerical experiment and discussion

In this section, we will analyze the propagation characteristics of the AAB carrying PEPV in anisotropic atmospheric turbulence using the aforementioned formulae. There is a universal phenomenon that the longer propagating distance and larger structure parameter lead to lower signal OAM mode power and higher crosstalk OAM modes power received. Therefore, we do not discuss the influences of the propagation distance and structure parameter. Instead we mainly devote ourselves to studying the radial probability density distributions of the signal OAM mode and the crosstalk OAM modes of the AAB carrying PEPV under variations in wavelength, waist width, topological charge and power order.

For description convenience in the following, |βm(ρ,z)|2be used to represents the signal OAM mode probability density, and |βΔm(ρ,z)|2 take as the crosstalk OAM modes probability densities, where Δm can take a series positive integer values, (i.e., |βΔm=1(ρ,z)|2 denotes the probability densities of the crosstalk OAM modes adjacent to the signal OAM mode). In addition, The following parameters of numerical calculations are assumed unless otherwise specified:α=3.65, λ=1550nm, ω=0.05m, Cn2=1014m3α, ξx=1 and ξy=3;turbulence outer scale L=10m; inner scale l=0.001m; propagation altitude h=3km; wind speedvws=21m/s; and propagation distance z=1km. The parameters adopted here set as an illustration for theoretical analyses and other values can also be adopted.

The probability density curves in Fig. 4 are plotted as functions of the beam radius. Figure 4(a) shows the probability density curves of the signal OAM modes with n = 1 and 3 and fixed v = 3. There are multiple-peaks in the |βm(ρ,z)|2 curves, but most of the energy is restricted to a small region between ρ = 0 and ρ = 0.05 m. We call it effective detection region of the signal OAM mode, and the effective detection region radius corresponds to the radius of the main lobe of the AAB. Under the condition n = 1, the result reduces to that of the canonical Airy vortex beam. Figure 4(b) shows the |βΔm(ρ,z)|2curves with Δm = 1, 2, 3, 4 and 5, for fixed n = 3 and v = 3. The |βΔm(ρ,z)|2 curves have strong oscillations and the fluctuations fade as Δm increases, which is the same as fractional Bessel Gauss vortex beams given in [38],we attribute this result to the nature of Airy beam as a Bessel beam of fractional order 1/3 [39]. Due to the strong oscillations, in the following section, we focus on the propagation characteristics of the envelope of the|βΔm(ρ,z)|2 curve for a convenient description. Furthermore, the magnitude and position of the envelope peak denote the power and the detected region center of the crosstalk OAM modes. In Fig. 4(b), the peaks of the envelope curves move outward and downward clearly as Δm becomes larger. Therefore, the noise OAM modes that matter the most to the signal OAM mode are the immediately adjacent OAM modes (Δm = 1). Similar to [34, 40], the other cross talk (Δm = 2, 3, 4 and 5) effects on the detection of the signal OAM mode are ignored. In addition, the peak position of the envelope curve is located at ρ = 0.159 m with Δm = 1, which means there is an observable distance between the effective detection region center of |βm(ρ,z)|2 and the detection region center of |βΔm=1(ρ,z)|2. This observable distance is helpful for improving the signal detection accuracy at the receiver plane by choosing a proper position.

 figure: Fig. 4

Fig. 4 The probability density distribution curves of the OAM modes of the AAB carrying PEPV in the anisotropic atmospheric turbulence:(a) signal OAM modes; (b) crosstalk OAM modes.

Download Full Size | PDF

The changes of the normalized probability density curves |βm(ρ,z)|2/|βm(ρ,z)|2max and |βΔm=1(ρ,z)|2/|βΔm=1(ρ,z)|2max are plotted in Figs. 5(a) and 5(b) for different beam waists ω = 0.03, 0.05, 0.08 and 0.10 m. It is shown that the normalized |βm(ρ,z)|2and |βΔm=1(ρ,z)|2 values all increase as the transverse scale ω increases, which is similar to the performance of the Airy Schell beam given in the fractional Bessel Gauss vortex beams given in [40]. It is worth noting that as larger ω increases, the slower the |βm(ρ,z)|2curves decays; meanwhile the effective detection region of the signal OAM mode enlarged. In Fig. 5(b), the peak heights of the envelope curves move upward clearly as ω gets larger, but the peak positions outward slightly. When ω larger than a certain value (i.e. ω>0.05), the detection region of signal OAM mode overlaps with that of crosstalk OAM modes and further influences the detection accuracy of the signal OAM mode.

 figure: Fig. 5

Fig. 5 The normalized probability density curves of the OAM modes against ρforω: (a) signal OAM modes; (b) crosstalk OAM modes with Δm=1.

Download Full Size | PDF

The variation of the normalized |βm(ρ,z)|2 and|βΔm=1(ρ,z)|2 curves against ρ for prescribed λ = 1550, 980, 632.8 and 532 nm are shown in Fig. 6. Figure 6(a) shows that as the λ increases, the decay of the |βm(ρ,z)|2 curves increase, the peak heights of the |βm(ρ,z)|2 curves decrease, the peak positions of the |βm(ρ,z)|2 curves drift toward the outward side of ρ, and the widths of the first rings of the |βm(ρ,z)|2 curves broaden. As shown in Fig. 6(b), the variation tendency of the |βΔm=1(ρ,z)|2 curves is similar to that of |βm(ρ,z)|2, but the difference is that it is not just the first ring that is broader but the entire curve that is broader. Therefore, a relatively small ω and longer λ are good for separate the detection regions between signal OAM mode and crosstalk OAM mode. On the whole, the curves’ variation tendency in Fig. 6 is consistent with that of the Airy Schell beam given in [25].

 figure: Fig. 6

Fig. 6 The normalized probability density curves of the OAM modes against ρ forλ: (a) signal OAM modes; (b) crosstalk OAM modes with Δm=1.

Download Full Size | PDF

The influence of power order n on the |βm(ρ,z)|2 and |βΔm=1(ρ,z)|2 curves of the receiver plane is examined. In this context, for fixed v = 3, it is clear from Fig. 7(a) that as the power order n is increased, |βm(ρ,z)|2 also increases. Figure 7(b) demonstrates that rises in n values will weaken the originally appearing oscillations of the |βΔm=1(ρ,z)|2 curve tails. Furthermore, these rises in n will initially reduce the magnitudes of the envelope curve peaks, while at n values exceeding 3.5, the magnitudes of the peaks will begin to rise again. We attribute this effect to the balance between the two following aspects.1) As the parameter n increases, the light spot size converges [23], which can compensate, to a certain extent, for the beam spreading effect [41] and the smaller probability densities of crosstalk OAM modes are achieved. 2) However when n is increases further (i.e. n>3.5), the light concentrates faster [23], which makes the beam speckle effect induced by atmospheric turbulence increases in strength. Therefore, the effects of turbulence on the probability densities of the OAM modes are worsen. Consequently, taking both of these effects into account, the |βΔm=1(ρ,z)|2 decreases to a minimum value and then increases with the increase of n. It is worth noting that as the power order n increases, the envelope curve peaks of the |βΔm=1(ρ,z)|2 curves drift toward the inward side of ρ.

 figure: Fig. 7

Fig. 7 The normalized probability density curves of the OAM modes plotted against ρ for several values of n = 2, 2.5, 3, 3.5, 4, and 5: (a) signal OAM modes; (b) crosstalk OAM modes with Δm=1.

Download Full Size | PDF

The changes in the |βm(ρ,z)|2 curves profile against topological charges are investigated in Fig. 8, by fixing n to 3. As shown in Fig. 8, |βm(ρ,z)|2 increases lightly with rising values of topological charges, and these rising values of topological charges will cause the envelope peak values of |βΔm=1(ρ,z)|2 curves to be reduced obviously; at the same time, the peak positions will drift toward the inward side of ρ.

 figure: Fig. 8

Fig. 8 The normalized probability density curves of the OAM modes plotted against ρ for several values of v = 1, 2, 3, 4, and 5: (a) signal OAM modes; (b) crosstalk OAM modes withΔm=1.

Download Full Size | PDF

Based on Figs. 7(b) and 8(b), we find that the envelop curve peak position of |βΔm=1(ρ,z)|2 curve is an emergent property that depends on the power order and topological charge. We will now examine the influences of the different v and n values on the |βΔm=1(ρ,z)|2 curves of the AAB carrying PEPV. Figure 9 shows that, for a fixed value of topological charge, the envelop curve peak position of |βΔm=1(ρ,z)|2curve moves to the inward side of ρ with increasing power order until it approaches a nearly stable value at approximately ρ = 0.126 m. And the moving rate of the envelope peak position will become slower as the topological charges increases.

 figure: Fig. 9

Fig. 9 The changes of the detection region center of crosstalk OAM modes with Δm=1 against topological charge v and power order n.

Download Full Size | PDF

4. Conclusions

In this paper, the Rytov theory is employed to derive an analytic formula for the radial probability densities of the OAM modes of the AAB carrying PEPV propagating in the anisotropic non-Kolmogorov turbulent atmosphere. Numerical calculations have shown that there are multiple-peaks in the signal OAM mode and crosstalk OAM mode probability density curves, but most of the energy is restricted in a really small region between ρ = 0 and ρ = 0.05 m, which corresponds to the radius of the main lobe of the AAB. There is an observable distance between the detection region centers of the signal OAM mode and crosstalk OAM modes. This distance is helpful for improving the signal detection accuracy at the receiver by choosing a proper position. But the detection region of signal OAM mode broadens clearly as the ω gets larger and λ becomes shorter, which would lead to the detection regain of signal OAM mode overlaps with that of the crosstalk OAM modes and adversely affect the detection accuracy for the signal OAM mode. Therefore, a relatively small ω and longer λ should be used in real applications. The radial probability density distribution of the signal OAM mode does not change obviously with the topological charge variation; but it will be greatly enhanced with the increase of power order. And rises in the power order values will weaken the originally appearing oscillations of the curve tails. Furthermore, it is found that, for a fixed value of topological charge, the detection region center of crosstalk OAM mode with Δm = 1 moves to the inward side of ρ with increasing values of power order until it approaches a nearly stable value at approximately ρ = 0.126 m, and the moving rate will become slower as the topological charge increases. It has been revealed that power order can be used as a steering parameter to modulate propagation characteristics of the AAB carrying PEPV.

Funding

National Natural Science Foundation of China (NSFC) (61431010, 61621005).

Acknowledgments

The authors would like to thank Xiuxiang Chu in School of Science in Zhejiang Agriculture and Forestry University, Peng Li in School of Science in Northwestern Polytechnical University, and Xinzhong Li in School of Physics and Engineering in Henan University of Science and Technology for the helpful discussions.

References and links

1. I. Toselli, “Introducing the concept of anisotropy at different scales for modeling optical turbulence,” J. Opt. Soc. Am. A 31(8), 1868–1875 (2014). [CrossRef]   [PubMed]  

2. G. Grechko, A. Gurvich, V. Kan, S. Kireev, and S. Savchenko, “Anisotropy of spatial structures in the middle atmosphere,” Adv. Space Res. 12(10), 169–175 (1992). [CrossRef]  

3. C. Robert, J.-M. Conan, V. Michau, J.-B. Renard, C. Robert, and F. Dalaudier, “Retrieving parameters of the anisotropic refractive index fluctuations spectrum in the stratosphere from balloon-borne observations of stellar scintillation,” J. Opt. Soc. Am. A 25(2), 379–393 (2008). [CrossRef]   [PubMed]  

4. M. Yao, I. Toselli, and O. Korotkova, “Propagation of electromagnetic stochastic beams in anisotropic turbulence,” Opt. Express 22(26), 31608–31619 (2014). [CrossRef]   [PubMed]  

5. F. Wang and O. Korotkova, “Random optical beam propagation in anisotropic turbulence along horizontal links,” Opt. Express 24(21), 24422–24434 (2016). [CrossRef]   [PubMed]  

6. M. A. Cox, C. Rosales-Guzmán, M. P. Lavery, D. J. Versfeld, and A. Forbes, “On the resilience of scalar and vector vortex modes in turbulence,” Opt. Express 24(16), 18105–18113 (2016). [CrossRef]   [PubMed]  

7. 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]  

8. J. Wang, S. Zhu, H. Wang, Y. Cai, and Z. Li, “Second-order statistics of a radially polarized cosine-Gaussian correlated Schell-model beam in anisotropic turbulence,” Opt. Express 24(11), 11626–11639 (2016). [CrossRef]   [PubMed]  

9. Y. Ren, Z. Wang, G. Xie, L. Li, Y. Cao, C. Liu, P. Liao, Y. Yan, N. Ahmed, Z. Zhao, A. Willner, N. Ashrafi, S. Ashrafi, R. D. Linquist, R. Bock, M. Tur, A. F. Molisch, and A. E. Willner, “Free-space optical communications using orbital-angular-momentum multiplexing combined with MIMO-based spatial multiplexing,” Opt. Lett. 40(18), 4210–4213 (2015). [CrossRef]   [PubMed]  

10. H. Jian, D. Ke, L. Chao, Z. Peng, J. Dagang, and Y. Zhoushi, “Effectiveness of adaptive optics system in satellite-to-ground coherent optical communication,” Opt. Express 22(13), 16000–16007 (2014). [CrossRef]   [PubMed]  

11. Z. Bouchal, “Nondiffracting optical beams: physical properties, experiments, and applications,” Czech. J. Phys. 53(7), 537–578 (2003). [CrossRef]  

12. Y. Hu, P. Zhang, C. Lou, S. Huang, J. Xu, and Z. Chen, “Optimal control of the ballistic motion of Airy beams,” Opt. Lett. 35(13), 2260–2262 (2010). [CrossRef]   [PubMed]  

13. C. Paterson, “Atmospheric turbulence and orbital angular momentum of single photons for optical communication,” Phys. Rev. Lett. 94(15), 153901 (2005). [CrossRef]   [PubMed]  

14. T. Wang, J. Pu, and Z. Chen, “Beam-spreading and topological charge of vortex beams propagating in a turbulent atmosphere,” Opt. Commun. 282(7), 1255–1259 (2009). [CrossRef]  

15. D. Zhi, R. Tao, P. Zhou, Y. Ma, W. Wu, X. Wang, and L. Si, “Propagation of ring Airy Gaussian beams with optical vortices through anisotropic non-Kolmogorov turbulence,” Opt. Commun. 387, 157–165 (2017). [CrossRef]  

16. W. Wen and X. Chu, “Beam wander of an Airy beam with a spiral phase,” J. Opt. Soc. Am. A 31(4), 685–690 (2014). [CrossRef]   [PubMed]  

17. Y. Zhu, Y. Zhang, and Z. Hu, “Spiral spectrum of Airy beams propagation through moderate-to-strong turbulence of maritime atmosphere,” Opt. Express 24(10), 10847–10857 (2016). [CrossRef]   [PubMed]  

18. B. Chen, C. Chen, X. Peng, Y. Peng, M. Zhou, and D. Deng, “Propagation of sharply autofocused ring Airy Gaussian vortex beams,” Opt. Express 23(15), 19288–19298 (2015). [CrossRef]   [PubMed]  

19. H. T. Eyyuboğlu and E. Sermutlu, “Partially coherent Airy beam and its propagation in turbulent media,” Appl. Phys. B 110(4), 451–457 (2013). [CrossRef]  

20. Y. Zhang, Q. Zhang, X. Ma, Z. Jiang, T. Xu, S. Wu, and X. Wu, “Measurement of Airy-vortex beam topological charges based on a pixelated micropolarizer array,” Appl. Opt. 55(32), 9299–9304 (2016). [CrossRef]   [PubMed]  

21. G.-H. Kim, H. J. Lee, J.-U. Kim, and H. Suk, “Propagation dynamics of optical vortices with anisotropic phase profiles,” J. Opt. Soc. Am. B 20(2), 351–359 (2003). [CrossRef]  

22. G. Molina-Terriza, E. M. Wright, and L. Torner, “Propagation and control of noncanonical optical vortices,” Opt. Lett. 26(3), 163–165 (2001). [CrossRef]   [PubMed]  

23. P. Li, S. Liu, T. Peng, G. Xie, X. Gan, and J. Zhao, “Spiral autofocusing Airy beams carrying power-exponent-phase vortices,” Opt. Express 22(7), 7598–7606 (2014). [CrossRef]   [PubMed]  

24. G. Lao, Z. Zhang, and D. Zhao, “Propagation of the power-exponent-phase vortex beam in paraxial ABCD system,” Opt. Express 24(16), 18082–18094 (2016). [CrossRef]   [PubMed]  

25. Y. Zhu, L. Zhang, and Y. Zhang, “Spiral spectrum of Airy–Schell beams through non-Kolmogorov turbulence,” Chin. Opt. Lett. 14(4), 042101 (2016). [CrossRef]  

26. G. Gibson, J. Courtial, M. Padgett, M. Vasnetsov, V. Pas’ko, S. Barnett, and S. Franke-Arnold, “Free-space information transfer using light beams carrying orbital angular momentum,” Opt. Express 12(22), 5448–5456 (2004). [CrossRef]   [PubMed]  

27. M. Malik, M. O’Sullivan, B. Rodenburg, M. Mirhosseini, J. Leach, M. P. Lavery, M. J. Padgett, and R. W. Boyd, “Influence of atmospheric turbulence on optical communications using orbital angular momentum for encoding,” Opt. Express 20(12), 13195–13200 (2012). [CrossRef]   [PubMed]  

28. Y. Ren, H. Huang, G. Xie, N. Ahmed, Y. Yan, B. I. Erkmen, N. Chandrasekaran, M. P. Lavery, N. K. Steinhoff, M. Tur, S. Dolinar, M. Neifeld, M. J. Padgett, R. W. Boyd, J. H. Shapiro, and A. E. Willner, “Atmospheric turbulence effects on the performance of a free space optical link employing orbital angular momentum multiplexing,” Opt. Lett. 38(20), 4062–4065 (2013). [CrossRef]   [PubMed]  

29. L. C. Andrews and R. L. Phillips, Laser Beam Propagation Through Random Media (SPIE, 2005)

30. L. G. Wang, W. W. Zheng, and L. Q. Wang, “The effect of atmospheric turbulence on the propagation properties of optical vortices formed by using coherent laser beam arrays,” J. Opt. A, Pure Appl. Opt. 11(6), 065703 (2009). [CrossRef]  

31. P. Zhang, J. Prakash, Z. Zhang, M. S. Mills, N. K. Efremidis, D. N. Christodoulides, and Z. Chen, “Trapping and guiding microparticles with morphing autofocusing Airy beams,” Opt. Lett. 36(15), 2883–2885 (2011). [CrossRef]   [PubMed]  

32. M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions: with Formulas, Graphs, and Mathematical Tables (Courier Corporation, 1964)

33. 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(1), 013601 (2001). [CrossRef]   [PubMed]  

34. M. Cheng, L. Guo, J. Li, and Q. Huang, “Propagation properties of an optical vortex carried by a Bessel-Gaussian beam in anisotropic turbulence,” J. Opt. Soc. Am. A 33(8), 1442–1450 (2016). [CrossRef]   [PubMed]  

35. Y. Zhu, L. Zhang, Z. Hu, and Y. Zhang, “Effects of non-Kolmogorov turbulence on the spiral spectrum of Hypergeometric-Gaussian laser beams,” Opt. Express 23(7), 9137–9146 (2015). [CrossRef]   [PubMed]  

36. I. S. Gradshteyn and I. M. Ryzhik, Table of integrals, series, and products (Academic, 2015).

37. C. Chen, H. Yang, S. Tong, and Y. Lou, “Changes in orbital-angular-momentum modes of a propagated vortex Gaussian beam through weak-to-strong atmospheric turbulence,” Opt. Express 24(7), 6959–6975 (2016). [CrossRef]   [PubMed]  

38. J. Gao, Y. Zhang, W. Dan, and Z. Hu, “Turbulent effects of strong irradiance fluctuations on the orbital angular momentum mode of fractional Bessel Gauss beams,” Opt. Express 23(13), 17024–17034 (2015). [CrossRef]   [PubMed]  

39. J. Rogel-Salazar, H. A. Jiménez-Romero, and S. Chávez-Cerda, “Full characterization of Airy beams under physical principles,” Phys. Rev. A 89(2), 204–208 (2014). [CrossRef]  

40. J. Gao, Y. Zhu, D. Wang, Y. Zhang, Z. Hu, and M. Cheng, “Bessel–Gauss photon beams with fractional order vortex propagation in weak non-Kolmogorov turbulence,” Photonics Res. 4(2), 30 (2016). [CrossRef]  

41. Y. Zhu, X. Liu, J. Gao, Y. Zhang, and F. Zhao, “Probability density of the orbital angular momentum mode of Hankel-Bessel beams in an atmospheric turbulence,” Opt. Express 22(7), 7765–7772 (2014). [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 (9)

Fig. 1
Fig. 1 Schematic diagram of the effects of atmospheric turbulence on an OAM beam. An ideal OAM mode can be break down into multiple OAM modes.
Fig. 2
Fig. 2 Spiral phase patterns of the PEPV beam at the initial plane with different topological charges and power orders. (a) v = 3, n = 1; (b) v = 3, n = 2;(c) v = 3, n = 5;(d) v = 6, n = 2.
Fig. 3
Fig. 3 Concept of turbulent atmosphere communication link using PEPV. (a) The turbulence resistance of the beam carrying PEPV. (b) OAM modes weight mapping. The beam carrying PEPV is the result of the weighted superposition of multiple propagating OAM beams at differing positions within the beam cross-section. And the mutual crosstalk among different OAM modes will compensate the distortion of each signal OAM mode and be helpful to let the distributive characteristics of these components get close to the initial state.
Fig. 4
Fig. 4 The probability density distribution curves of the OAM modes of the AAB carrying PEPV in the anisotropic atmospheric turbulence:(a) signal OAM modes; (b) crosstalk OAM modes.
Fig. 5
Fig. 5 The normalized probability density curves of the OAM modes against ρforω: (a) signal OAM modes; (b) crosstalk OAM modes with Δm=1 .
Fig. 6
Fig. 6 The normalized probability density curves of the OAM modes against ρ forλ: (a) signal OAM modes; (b) crosstalk OAM modes with Δm=1 .
Fig. 7
Fig. 7 The normalized probability density curves of the OAM modes plotted against ρ for several values of n = 2, 2.5, 3, 3.5, 4, and 5: (a) signal OAM modes; (b) crosstalk OAM modes with Δm=1 .
Fig. 8
Fig. 8 The normalized probability density curves of the OAM modes plotted against ρ for several values of v = 1, 2, 3, 4, and 5: (a) signal OAM modes; (b) crosstalk OAM modes with Δm=1 .
Fig. 9
Fig. 9 The changes of the detection region center of crosstalk OAM modes with Δm=1 against topological charge v and power order n.

Equations (13)

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

E (0) (r,φ,z=0)=Ai( r 0 r ω )exp( a r 0 r ω )exp[ i2vπ ( φ 2π ) n ],
E( ρ,θ,z )= ik 2πz exp(ikz) drdφ E (0) (r,φ,z=0) ×exp{ ik[ ρ 2 + r 2 2ρrcos(θφ) ] 2z +ψ(ρ,θ,z) },
A 0 ωexp( a 3 3 )(1 ω a 2 r 0 ).
exp( z 2 (t 1 t ))= l= J l (z) t l , J l (z)= (1) l J l (z) ,
E( ρ,θ,z )= ik 2πz exp(ikz)w( r 0 w a 2 )exp( ik ρ 2 +ik r 0 2 2z + a 3 3 ) ×[ l= (i) l exp(ilθ) M l J l ( k r 0 ρ z ) ]exp[ ψ(ρ,θ,z) ],
M l = 0 2π exp[ i(2vπ ( φ 2π ) n +lφ) ] dφ={ j=0 h=0 i j+h v j l h ( 2π ) j+h+1 j!h!(nj+h+1) ,l0 j=0 i j v j ( 2π ) j+1 j!(nj+1) ,l=0.
E( ρ,θ,z )= 1 2π m β m (ρ,z)exp(imθ) ,
β m (ρ,z)= 1 2π 0 2π E( ρ,θ,z )exp(imθ) dθ.
| β m (ρ,z) | 2 = 1 2π 0 2π 0 2π E( ρ,θ,z ) E ( ρ, θ ,z )exp[ im(θ θ ) ] × exp[ ψ(ρ,θ,z)+ψ( ρ , θ ,z ) ] dθd θ ,
exp[ ψ( ρ,θ,z )+ψ( ρ , θ ,z ) ] =exp( D(ρ,θ, ρ , θ ) 2 ) exp( ρ 2 + ρ 2 2ρ ρ cos(θ θ ) ρ 0 2 ),
ρ 0 = { ξ x 2 + ξ y 2 2 ξ x 2 ξ y 2 π 2 k 2 zA(α) C ˜ n 2 /[6(α2)]×[ κ 2α γexp( κ 0 2 κ 2 )Γ(2 α 2 , κ 0 2 κ 2 )2 κ 0 4α ] } 1 /2 ,
0 2π exp[ im(θ θ ) ] exp[ ηcos(θ θ ) ]dθ=2πexp[ im θ ] I m (η),
| β m (ρ,z) | 2 =2π[ ( kw 2πz ) 2 ( r 0 w a 2 ) 2 exp( 2 a 3 3 2 ρ 2 ρ 0 2 ) ] ×{ l=1 [ ( l =1 ( M l + M l )( M l + M l ) + M 0 ( M l + M l ) ) J l ( k r 0 ρ z ) 2 I lm ( 2 ρ 2 ρ 0 2 ) ] +[ M 0 l =1 ( M l + M l )+ M 0 M 0 ] J 0 ( k r 0 ρ z ) 2 I m ( 2 ρ 2 ρ 0 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.