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

Evolution properties of Bessel-Gaussian Schell-model beams in non-Kolmogorov turbulence

Open Access Open Access

Abstract

The analytical expressions for the spectral degree of coherence, the effective radius of curvature and the propagation factor of the Bessel-Gaussian Schell-model (BGSM) beam in turbulent atmosphere are derived based on the extended Huygens-Fresnel principle and the second-order moments of the Wigner distribution function (WDF). The evolution properties of BGSM beams propagating in non-Kolmogorov turbulence are investigated by a set of numerical examples. It is demonstrated that the spectral degree of coherence of the BGSM beam evolves into Gaussian profile twice with the increasing of the propagation distance. The turbulence-induced degradation can be remarkably reduced by using the BGSM beam with the proper source parameters. The effects that the generalized refractive-index structure constant, outer and inner scales, and the spectral index of spatial power spectrum of atmospheric turbulence have on the evolution properties of BGSM beams are also discussed in detail.

© 2015 Optical Society of America

1. Introduction

In recent decades, various types of laser beams propagating through turbulent atmosphere are extensively studied, due to their widespread applications in free-space optical (FSO) communications, laser radar systems, remote sensing, optical imaging, semiconductor waveguide, etc [17]. It has been found that the propagation of laser beams are inevitably affected by random refractive-index fluctuations of the turbulence and the use of partially coherent beams (PCBs) is an effective way to reduce the deleterious effect of the turbulence under certain circumstances [813]. As a result of the constraint of nonnegative definiteness in the choice of the mathematical form of spatial correlation functions for optical fields, most of the papers on propagation of PCBs in the early days are based on the conventional spatial correlation functions, i.e., Gaussian correlated Schell-model functions.

In order to satisfy such a constraint for devising genuine spatial correlation functions of scalar or electromagnetic PCBs, Gori and his collaborators presented a sufficient condition derived from the theory of reproducing kernel Hilbert spaces [14,15]. Since then the PCBs with non-conventional correlation functions, whose spectral degrees of coherence are not Gaussian profiles, have attracted considerable attention. Up to now, a variety of augmented PCBs with non-conventional correlation functions have been introduced, such as beams with locally varying spatial coherence [16], non-uniformly correlated (NUC) beam [1724], multi-Gaussian Schell-model (MGSM) beam [2532], Laguerre-Gaussian Schell-model (LGSM) and Bessel-Gaussian Schell-model (BGSM) beam [3339], cosine-Gaussian Schell-model (CGSM) beam [4046], sinc Schell-model (SSM) [47,48], and specially correlated radially polarized (SCRP) beam [49]. Previous experiments and theoretical analyses have demonstrated that those partially coherent fields with non-conventional correlation functions have some extraordinary propagation characteristics in free space, such as the peculiar self-focusing effect and a laterally shifted intensity maxima of the NUC beam [17,18,20,22], far fields with flat-topped intensity profiles generated by MGSM beams and the first kind SSM (SSM1) beam [25,29,47], far fields with ring-shaped (i.e., dark hollow) intensity profiles formed by LGSM beams, BGSM beams, CGSM beams and the second kind SSM (SSM2) beam [33,34,43,47], far field with shape-invariant double-layer flat-topped intensity profile produced by the electromagnetic sinc Schell-model (EM SSM) beam [48], etc.

To date, a few papers on propagation of PCBs with non-conventional correlation functions in turbulent atmosphere have been published [19,2228,3842,46,48,49]. In [38], the analytical expressions for the average intensity and the beam width of the BGSM beam propagating through a paraxial ABCD optical system in Kolmogorov turbulence were derived, and the evolution properties of the average intensity and the beam width were investigated. However, to the best of the authors’ knowledge, the evolution properties of the spectral degree of coherence (i.e., correlation function), the effective radius of curvature and the propagation factor of the BGSM beam in turbulence have not been known. In this paper, a generalized power spectrum model valid in non-Kolmogorov turbulence is employed, and the explicit analytical expressions for the spectral degree of coherence, the effective radius of curvature and the propagation factor of the BGSM beam in turbulent atmosphere are derived based on the extended Huygens-Fresnel principle and the second-order moments of the Wigner distribution function (WDF). The evolution properties of BGSM beams propagating in non-Kolmogorov turbulence are numerically explored. Some useful observations and conclusions have been proposed.

2. Theory

2.1. Spectral degree of coherence of the BGSM beam in non-Kolmogorov turbulence

Let us consider that a BGSM source generates a beam-like field propagating into the half-space z = 0 filled with turbulent atmosphere nearly parallel to the positive z direction. For brevity, the dependence of all field statistics on the angular frequency ω is implied but omitted in what follows. The cross-spectral density (CSD) of the BGSM beam at the source plane z = 0 is defined by [33]

W(r1,r2,0)=exp(|r1|2+|r2|24σ2)μ(r1,r2,0)
where r1 ≡ (x1, y1) and r2 ≡ (x2, y2) denote two arbitrary transverse position vectors at the source plane, σ is the transverse beam width (i.e., the rms width) of the BGSM beam, and μ(r1, r2, 0) is the spectral degree of coherence of the BGSM beam at z = 0 and and characterized by the following form
μ(r1,r2,0)=exp[|r1r2|22δ2]J0(β|r1r2|δ)
where δ is the transverse spatial coherence width of the BGSM width, β is a real constant, and J0 (·) is the zeroth-order Bessel function of the first kind.

It is clearly seen that the BGSM model represents a family of sources which can reduce to the conventional Gaussian Schell-model (GSM) source when β = 0 and to the J0-correlated source as β → ∞. The condition for a BGSM source to generate a beam-like field is the same as that for a classical GSM source: 1/4σ2 + 1/δ2 ≪ 2π2/λ2 with λ being the wavelength of the source [33].

Within the validity of the paraxial approximation, the propagation of the CSD of the BGSM beam in turbulence can be studied with the help of the following extended Huygens-Fresnel integral [2,50,51]

W(ρs,ρd,z)=(k2πz)2W(rs,rd,0)×exp[ikz(ρsrs)(ρdrd)]×exp[H(ρd,rd,z)]d2rsd2rd
where k = 2π/λ is the wave number with λ being the wavelength of the source. In Eq. (3), the central abscissa coordinate systems are chosen, that is,
rs=r1+r22,rd=r1r2
ρs=ρ1+ρ22,ρd=ρ1ρ2
where ρ1 ≡ (ρ1x, ρ1y) and ρ2 ≡ (ρ2x, ρ2y) are two arbitrary transverse position vectors at the receiver plane, perpendicular to the direction of propagation of the beam, and
W(rs,rd,0)=W(r1,r2,0)=W(rs+rd2,rsrd2,0)

The term exp[−H (ρd, rd; z)] in Eq. (3), represents the contribution of atmospheric turbulence, and H (ρd, rd; z) can be expressed as [25,28,41,42,46,48]

H(ρd,rd,z)=π2k2zT3[rd2+rdρd+ρd2]
where
T=0κ3Φn(κ)dκ
Φn (κ) is the one-dimensional power spectrum of fluctuations in the refractive index of the isotropic turbulent medium, and κ is the scalar spatial wave number (i.e., spatial frequency).

In order to cover a wide scope of atmospheric conditions, our discussion on propagation of the BGSM beam is based on a generalized power spectrum model valid in non-Kolmogorov turbulence [25,28,41,42,46,48,52]

Φn(κ,α)=A(α)C˜n2exp[(κ2/κm2)](κ2+κ02)α/2,0κ<,3<α<4
where the term C˜n2 is a generalized index-of-refraction structure constant (in units of m3−α), κ0 = 2π/L0, κm = c(α)/l0, L0 and l0 being the outer and inner scales of the turbulence, respectively, and
c(α)=[Γ(5α2)A(α)2π3]1/(α5)
A(α)=Γ(α1)cos(απ/2)4π2
with Γ(·) being the Gamma function. The power spectrum expressed by Eq. (9) reduces to the conventional Kolmogorov spectrum when α = 11/3, A(α) = 0.033, C˜n2=Cn2, L0 → ∞ and l0 → 0. With the power spectrum in Eq. (9) the integral in Eq. (8) becomes [25,28,41,42,46,48,52]
T=0κ3Φn(κ)dκ=A(α)2(α2)C˜n2[κm2αμexp(κ02κm2)Γ(2α2,κ02κm2)2κ04α]
where μ=2κ022κm2+ακm2 and Γ(·,·) denotes the incomplete Gamma function.

Applying Eqs. (3), (6) and (7), and calculating the integral with respect to rs by using the following formula [53]

exp(p2x2±qx)dx=πpexp(q24p2),(p>0)
we obtain the expression for the CSD of the BGSM beam in the receiver plane
W(ρs,ρd,z)=Qexp[ard2bρsrd+cρdrd]J0(β|rd|δ)d2rd
where
Q=σ2k22πz2exp[ikzρsρd(π2k2zT3+σ2k22z2)ρd2]
a=18σ2+12δ2+σ2k22z2+π2k2Tz3,b=ikz,c=σ2k2z2π2k2Tz3

With the application of the coordinate transformation, Eq. (14) can be rewritten as

W(ρs,ρd,z)=Q002πexp[ard2|bρscρd|rdcosφ]J0(βrdδ)rddrddφ
where rd is the scalar magnitude of rd, φ is the angle between bρscρd and rd. By using the following formulas [53]
02πexp(±xcosφ)dφ=2πI0(x)
0xexp(γx2)Iν(px)Jν(qx)dx=12γexp(p2q24γ)Jν(pq2γ),[Reγ>0,Reν>1]
where Jν (·) is the ν-order Bessel function of the first kind and Iν (·) is the ν-order modified Bessel function of the first kind, respectively. Eq. (17) is simplified to
W(ρs,ρd,z)=σ2k22az2exp[|bρscρd|24a+ikzρsρd(πk2zT3+σ2k22z2)ρd2]×exp(β24aδ2)J0(β|bρscρd|2aδ)

Accordingly, the spectral density of the BGSM beam in the receiver plane is obtained as S (ρ, z) = W (ρ, 0, z), where ρ ≡ (ρx, ρy) denotes an arbitrary transverse position vector at the receiver plane. Furthermore, the spectral degree of coherence of the BGSM beam in the receiver plane is acquired by

μ(ρ1,ρ2,z)=W(ρ1+ρ22,ρ1ρ2,z)S(ρ1,z)S(ρ2,z)

2.2. Second-order moments of the BGSM beam in non-Kolmogorov turbulence

On the basis of inverse Fourier transform of the Dirac delta function and its property of even function [54], and after some operations as shown in [50], the CSD of the BGSM beam in the receiver plane can be rewritten as

W(ρs,ρd,z)=1(2π)2W(r,ρd+zkκd,0)×exp[iρsκd+iκdrH(ρd,ρd+zkκd,z)]d2rd2κd
where κd ≡ (κdx, κdy) is the position vector in the spatial-frequency domain. The term W(r,ρd+zkκd,0) is obtained by inserting Eqs. (1) and (2) into Eq. (6), as follows
W(r,ρd+zkκd,0)=exp[12σ2r2(18σ2+12δ2)(ρd+zkκd)2]×J0(βδ|ρd+zkκd|)
and the term H(ρd,ρd+zkκd,z) is acquired from Eq. (7), that is
H(ρd,ρd+zkκd,z)=π2k2zT3[3ρd2+3zkκdρd+z2k2κd2]

As is well known, the WDF being closely related to the second-order statistical properties of PCBs on propagation in turbulent atmosphere, is especially suitable for the treatment of PCBs, and the WDF can be expressed in terms of the CSD by the following formula [26,39,50,51]

h(ρs,θ,z)=(k2π)2W(ρs,ρd,z)exp(ikθρd)d2ρd
where θ ≡ (θx, θy) denotes an angle which the vector of interest makes with the z-direction, x and y are the wave vector components along the x-axis and y-axis, respectively.

On substituting from Eqs. (22)(24) into Eq. (25) and calculating the integral with respect to r′ by using Eq. (13), the WDF of the BGSM beam in non-Kolmogorov turbulence is obtained as

h(ρs,θ,z)=k2σ28π3exp[η1ρd2η2κd2η3ρdκdiρsκdikθρd]×exp[H(ρd,ρd+zkκd,z)]J0(βδ|ρd+zkκd|)d2ρdd2κd
where
η1=18σ2+12δ2,η2=(18σ2+12δ2)z2k2+12σ2,η3=(14σ2+1δ2)zk

The moments of order n1 + n2 + m1 + m2 of the WDF of a laser beam are defined by [39,42,50,51]

ρsxn1ρsyn2θxm1θym2=1Pρsxn1ρsyn2θxm1θym2h(ρs,θ,z)d2ρsd2θ
where ρsx and ρsy are the coordinates of of the transverse position vector ρs at the receiver plane, P=h(ρs,θ,z)d2ρsd2θ is the total irradiance of the beam, and P = 2πσ2 for the case of BGSM beams.

Substituting Eq. (26) into Eq. (28), we obtain (after tedious integration) the following expressions for the second-order moments of the WDF of the BGSM beam in non-Kolmogorov turbulence

ρs2=ρsx2+ρsy2=2σ2+(12σ2+2+β2δ2)z2k2+43π2Tz3
θ2=θx2+θx2=(12σ2+2+β2δ2)1k2+4π2Tz
ρsθ=ρsxθx+ρsyθy=(12σ2+2+β2δ2)zk2+2π2Tz2

In the above derivations, we have used the series representation of the Bessel function of the first kind [53]

Jν(x)=n=0(1)n(x/2)2n+νn!Γ(n+ν+1),|x|<
and the following formulas [54]
δ(px)=1|p|δ(x)
δ(n)(s)=12π(ix)nexp(isx)dx,(n=0,1,2)
f(x)δ(n)(x)dx=(1)nf(n)(0),(n=0,1,2)

In accordance with the definition of the effective radius of curvature of a laser beam [26,55,56], we obtain the explicit expression for the effective radius of curvature of the BGSM beam in non-Kolmogorov turbulence with the help of Eqs. (29) and (31)

R(z)=ρs2ρsθ=2σ2+(12σ2+2+β2δ2)z2k2+43π2Tz3(12σ2+2+β2δ2)zk2+2π2Tz2

Based on the second-order moments of the WDF, the propagation factor of a partially coherent beam in turbulence is defined by [39,42,50,51]

M2(z)=k(ρs2θ2ρsθ2)1/2

Applying Eqs. (29)(31) and (37), we obtain the propagation factor of the BGSM beam in non-Kolmogorov turbulence

M2(z)=[2σ2(12σ2+2+β2δ2)+8π2k2σ2Tz+43π2(12σ2+2+β2δ2)Tz3+43π4k2T2z4]12
Eq. (38) indicates that the propagation factor of the BGSM beam in non-Kolmogorov turbulence increases with the propagation distance z, which means the beam quality is in decline on propagation of the beam as a result of the presence of atmospheric turbulence.

Under the condition of Φn (κ) = 0, Eq. (38) reduces to the expression for the propagation factor of the BGSM beam in free space, that is M2 (z) = [1 + 2σ2 (2 + β2)/δ2]1/2. One can find that the propagation factor of the BGSM beam in free space is independent of the propagation distance z, and only determined by the beam width σ, the spatial coherence width δ and the parameter β of the BGSM source. When Φn (κ) = 0 and β = 0, Eq. (38) reduces to the expression for the propagation factor of a GSM beam in free space, i.e., M2 (z) = [1 + 4σ2/δ2]1/2, which is consistent with the previous result reported in [57].

3. Numerical calculation results and analysis

In this section, we will examine the evolution properties of BGSM beams on propagation in non-Kolmogorov turbulence by a set of numerical examples based on the analytical formulas derived in the previous section. Due to the unit of the generalized structure constant C˜n2 depending on the spectral index α, analysis of the beam propagation in non-Kolmogorov turbulence involving C˜n2 for various α may easily cause misleading interpretations. Charnotskii has discussed this issue and suggested to use a dimensional argument of the same dimension or some dimensionless argument to eliminate the unit choice inconsistency, such as coherence radius, Fresnel number and Rytov variance [58,59]. In this paper, the generalized Rytov variance for a plane wave propagating in weak generalized atmospheric turbulence is chosen, which can be expressed by σR2=2Γ(α1)Γ(1α/2)α1cos(απ/2)cos[(α2)/4]C˜n2k3α/2zα/2 [60]. Without loss of generality, the parameters of the source and turbulent atmosphere are chosen to be C˜n2=1014m3α, α = 3.8, L0 = 1 m, l0 = 0.001 m, σ = 0.02 m, δ = 0.01 m, λ = 632.8 nm unless other parameters are specified in each figure caption.

Figure 1 illustrates the evolution properties of the modulus of the transverse spectral degree of coherence of the BGSM beam at several propagation distances in non-Kolmogorov turbulence for different values of the parameter β. For comparison, the spectral degree of coherence of a GSM beam (β = 0) is given under the same conditions, which maintains Gaussian profile for all different propagation distances. From Fig. 1(a), one can find that the spectral degree of coherence of the BGSM beam with β > 0 is non-Gaussian profile, which has several side lobes around the main lobe. The larger the parameter β, the more the side lobes in the spectral degree of coherence. For relatively short propagation distances, the side lobes gradually disappear and the spectral degree of coherence converts to a single Gaussian profile for all β at a distance of approximately 2.6 km (for the given BGSM source and atmospheric parameters) [Figs. 1(b)–1(d)], Moreover, the spectral degree of coherence goes through another change for z > 2.6 km and the side lobes reconstruct successively depending on the values of the parameter β [Figs. 1(e)–1(g)]. In the end, the spectral degree of coherence of the BGSM beam evolves into Gaussian profile again in the far field [Fig. 1(h)]. This interesting phenomenon can be attributed to the combined effect of the source correlation and atmospheric turbulence. To be precise, the effect of the source correlation plays a dominant role at short propagation distances and the influence of atmospheric turbulence at long propagation distances.

 figure: Fig. 1

Fig. 1 Modulus of the transverse spectral degree of coherence of the BGSM beam as a function of ρ1x at several propagation distances in non-Kolmogorov turbulence for different values of parameter β.

Download Full Size | PDF

Figure 2 shows the behaviors of the modulus of the transverse spectral degree of coherence of the BGSM beam at propagation distance z = 2 km in non-Kolmogorov turbulence for different values of spectral index α and generalized Rytov variance σR2. From Fig. 2(a), one can find that the spectral degree of coherence of the BGSM beam is non-Gaussian profile for different values of the spectral index α. As the fluctuations of atmospheric turbulence increase, the spectral degree of coherence of the BGSM beam eventually converges to Gaussian profile, and more rapidly for smaller spectral index α. It can be concluded that the BGSM beam is less affected by the generalized atmospheric turbulence with larger spectral index α.

 figure: Fig. 2

Fig. 2 Modulus of the transverse spectral degree of coherence of the BGSM beam with β = 5 as a function of ρ1x at propagation distance z = 2 km in non-Kolmogorov turbulence for different values of spectral index α and generalized Rytov variance σR2.

Download Full Size | PDF

Moreover, numerical calculations for the normalized propagation factor of the BGSM beam in non-Kolmogorov turbulence were performed on the basis of Eq. (38). Some typical examples are compiled in Figs. 38.

 figure: Fig. 3

Fig. 3 Normalized propagation factor of the BGSM beam for different values of parameter β versus propagation distance z in non-Kolmogorov turbulence with different values of generalized refractive-index structure constant C˜n2.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Normalized propagation factor of the BGSM beam versus propagation distance z in non-Kolmogorov turbulence for different values of parameter β and beam width σ.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 Normalized propagation factor of the BGSM beam with β = 2.5 as a function of propagation distance z in non-Kolmogorov turbulence for different values of (a) outer scale L0 and (b) inner scale l0.

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 Normalized propagation factor of the BGSM beam versus parameter β at propagation distance z = 10 km in non-Kolmogorov turbulence for different values of (a) spatial coherence width δ and (b) wavelength λ.

Download Full Size | PDF

 figure: Fig. 7

Fig. 7 Normalized propagation factor of the BGSM beam versus spatial coherence width δ at propagation distance z = 10 km in non-Kolmogorov turbulence for different values of parameter β.

Download Full Size | PDF

 figure: Fig. 8

Fig. 8 Normalized propagation factor of the BGSM beam versus spectral index α at propagation distance z = 1 km in non-Kolmogorov turbulence with σR2=0.2 for different values of parameter β.

Download Full Size | PDF

Figure 3 plots the normalized propagation factor of the BGSM beam versus propagation distance z in non-Kolmogorov turbulence for different values of parameter β and generalized refractive-index structure constant C˜n2. Figure 4 shows the normalized propagation factor of the BGSM beam versus propagation distance z in non-Kolmogorov turbulence for different values of parameter β and beam width σ. From Fig. 34, it can be clearly seen that the normalized propagation factor of the BGSM beam in turbulence increases with the propagation distance z, which is totally different from that in free space being independent of the propagation distance z. The normalized propagation factor of the BGSM beam with larger parameter β increases more slowly than that with smaller parameter β or a GSM beam, especially in the cases of larger structure constant C˜n2 and larger beam width σ. That implies the influence of atmospheric turbulence can be significantly reduced by using the BGSM beam with larger parameter β.

Figure 5 depicts the normalized propagation factor of the BGSM beam as a function of propagation distance z in non-Kolmogorov turbulence for different values of outer scale L0 and inner scale l0, respectively. From Fig. 5, it can be readily seen that the normalized propagation factor of the BGSM beam increases with outer scale L0 for a fixed propagation distance z while decreasing with inner scale l0, and more sensitive to the change of inner scale l0. The outer scale L0 forms the upper limit of the inertial range and increases with the strength of atmospheric turbulence. The inner scale l0, which forms the lower limit of the inertial range, has a larger value in weak turbulence and a smaller value in strong turbulence. The increasing of outer scale L0 or the decreasing of inner scale l0 is equivalent to increasing the strength of the turbulence. In these cases, the beam will meet more turbulence eddies along its propagation paths. As a result, the beam experiences more spreading and has larger propagation factor [2,52].

Figure 6 plots the normalized propagation factor of the BGSM beam versus parameter β at propagation distance z = 10 km in non-Kolmogorov turbulence for different values of spatial coherence width δ and wavelength λ, respectively. As indicated by Fig. 6, the normalized propagation factor of the BGSM beam decreases monotonically with the parameter β, more apparently for larger spatial coherence width δ and smaller wavelength λ. However, the benefits gained by increasing the parameter β continue to diminish and can be ignored when the parameter β is larger than 15. For smaller values of spatial coherence width δ, the normalized propagation factor of the BGSM beam is always smaller as shown in Fig. 6(a). The reason is that the turbulence-induced spread of a weaker coherent beam is less than that of a stronger coherent beam. In addition, Fig. 6(b) implies that the BGSM beam with larger wavelength has advantage over the BGSM beam with smaller wavelength in reducing the turbulence-induced degradation.

Figure 7 presents the normalized propagation factor of the BGSM beam versus spatial coherence width δ at propagation distance z = 10 km in non-Kolmogorov turbulence for different values of parameter β. As shown in Fig. 7, the normalized propagation factor of the BGSM beam can be reduced by increasing the value of the parameter β when the value of the spatial coherence width δ is fixed, which is in line with the depiction of Fig. 6(a). The normalized propagation factor of the BGSM beam converges to a constant when the spatial coherence width δ is large enough or small enough, whatever the value of the parameter β is. That means the beam quality cannot get improved by changing the value of the parameter while the source is a fully coherent wave or an incoherent wave.

Figure 8 gives the normalized propagation factor of the BGSM beam versus spectral index α at propagation distance z = 1 km in non-Kolmogorov turbulence with σR2=0.2 for different values of parameter β. One can find from Fig. 8 that the relationship between the normalized propagation factor of the BGSM beam and the spectral index α is monotonic. The normalized propagation factor of the BGSM beam decreases with the spectral index α, and the reason is that the influence of atmospheric turbulence becomes weaker with the increasing of the spectral index α. At the same time, we can deduce that the BGSM beam with larger parameter β is less affected by the turbulence than that with smaller parameter β.

Figure 9 represents the effective radius of curvature of the BGSM beam propagating in non-Kolmogorov turbulence for different values of parameter β and generalized refractive-index structure constant C˜n2. As manifested in Fig. 9, the effective radius of curvature of the BGSM beam displays a downward trend in the near field and then turns to increase after reaching a dip, as a result of the diffraction of a collimated beam in atmospheric turbulence. There exists a minimum Rmin(z) as the propagation distance z increases, and the position of Rmin(z) is further close to the source plane for the BGSM beam with larger parameter β than that with smaller parameter β. The difference between the effective radius of curvature of the BGSM beam in free space and that in non-Kolmogorov turbulence is smaller than that of a GSM beam, and the larger the parameter β the smaller the difference. It is also found that the BGSM with larger parameter β is less affected by the turbulence than that with smaller parameter β from the aspect of the evolution properties of the effective radius of curvature.

 figure: Fig. 9

Fig. 9 Effective radius of curvature of the BGSM beam propagating in non-Kolmogorov turbulence for different values of parameter β and generalized refractive-index structure constant C˜n2.

Download Full Size | PDF

4. Concluding remarks

In this paper, we have derived the analytical expressions for the spectral degree of coherence, the effective radius of curvature and the propagation factor of the BGSM beam in non-Kolmogorov turbulence with the help of the extended Huygens-Fresnel principle and the second-order moments of the WDF, which can be simplified to those of a GSM beam when β =0. We have examined the variation of three main statistical properties of the BGSM beam: the spectral degree of coherence, the effective radius of curvature and the normalized propagation factor, on propagation in non-Kolmogorov turbulence by some numerical examples. It is found that the BGSM beam is less affected by the turbulence than a GSM beam, and this advantage is enhanced for larger parameter β, larger beam width σ or in the case of stronger fluctuations of the turbulence. The beam quality of the BGSM beam can further be improved by decreasing of the spatial coherence width δ or increasing the wavelength λ. Furthermore, the effects of outer scale L0, inner scale l0, and the spectral index α of the turbulence on the propagation factor are also explored at great length. The BGSM beam has greater spreading in the generalized atmospheric turbulence with smaller inner scale l0, larger outer scale L0 or smaller spectral index α. Our theoretical results will be useful in the long-distance free-space optical communications. In the future work, we will try to carry out the experiment and find whether the experiment results agree with the theoretical predictions. The experimental generation of partially coherent beams with different complex degrees of coherence was recently reported by Wang et al. [34]. It is possible to generate the BGSM beam with the similar method and realize the experiment of the BGSM beam propagating in non-Kolmogorov turbulence.

Acknowledgments

This work is supported by the National Natural Science Foundation of China under Grant 61001129 and the Fundamental Research Funds for the Central Universities.

References and links

1. L. Mandel and E. Wolf, Optical Coherence and Quantum Optics (Cambridge University, 1995). [CrossRef]  

2. L. C. Andrews and R. L. Phillips, Laser Beam Propagation through Random Media, 2nd ed. (SPIE, 2005). [CrossRef]  

3. H. T. Eyyuboğlu, Ç. Arpali, and Y. Baykal, “Flat topped beams and their characteristics in turbulent media,” Opt. Express 14(10), 4196–4207 (2006). [CrossRef]  

4. Y. Cai and S. He, “Propagation of various dark hollow beams in a turbulent atmosphere,” Opt. Express 14(4), 1353–1367 (2006). [CrossRef]   [PubMed]  

5. Y. Cai and S. He, “Average intensity and spreading of an elliptical Gaussian beam propagating in a turbulent atmosphere,” Opt. Lett. 31(5), 568–570 (2006). [CrossRef]   [PubMed]  

6. Z. Zang, “High-Power (> 110 mW) Superluminescent Diodes by Using Active Multimode Interferometer,” IEEE Photon. Technol. Lett. 22(10), 721–723 (2010). [CrossRef]  

7. Z. Zang, K. Mukai, P. Navaretti, M. Duelk, C. Velez, and K. Hamamoto, “Thermal resistance reduction in high power superluminescent diodes by using active multi-mode interferometer,” Appl. Phys. Lett. 100(3), 031108 (2012). [CrossRef]  

8. J. C. Ricklin and F. M. Davidson, “Atmospheric turbulence effects on a partially coherent Gaussian beam: implications for free-space laser communication,” J. Opt. Soc. Am. A 19(9), 1794–1802 (2002). [CrossRef]  

9. J. C. Ricklin and F. M. Davidson, “Atmospheric optical communication with a Gaussian Schell beam,” J. Opt. Soc. Am. A 20(5), 856–866 (2003). [CrossRef]  

10. A. Dogariu and S. Amarande, “Propagation of partially coherent beams: turbulence-induced degradation,” Opt. Lett. 28(1), 10–12 (2003). [CrossRef]   [PubMed]  

11. M. Salem, T. Shirai, A. Dogariu, and E. Wolf, “Long-distance propagation of partially coherent beams through atmospheric turbulence,” Opt. Commun. 216(4–6), 261–265 (2003). [CrossRef]  

12. O. Korotkova, L. C. Andrews, and R. L. Phillips, “Model for a partially coherent Gaussian beam in atmospheric turbulence with application in Lasercom,” Opt. Eng. 43(2), 330–341 (2004). [CrossRef]  

13. Y. Baykal, H. T. Eyyuboğlu, and Y. Cai, “Scintillations of partially coherent multiple Gaussian beams in turbulence,” Appl. Opt. 48(10), 1943–1954 (2009). [CrossRef]   [PubMed]  

14. F. Gori and M. Santarsiero, “Devising genuine spatial correlation functions,” Opt. Lett. 32(24), 3531–3533 (2007). [CrossRef]   [PubMed]  

15. F. Gori, V. Ramírez-Sánchez, M. Santarsiero, and T. Shirai, “On genuine cross-spectral density matrices,” J. Opt. A, Pure Appl. Opt. 11(8), 085706 (2009). [CrossRef]  

16. L. Waller, G. Situ, and J. W. Fleischer, “Phase-space measurement and coherence synthesis of optical beams,” Nat. Photonics 6(7), 474–479 (2012). [CrossRef]  

17. H. Lajunen and T. Saastamoinen, “Propagation characteristics of partially coherent beams with spatially varying correlations,” Opt. Lett. 36(20), 4104–4106 (2011). [CrossRef]   [PubMed]  

18. Z. Tong and O. Korotkova, “Electromagnetic nonuniformly correlated beams,” J. Opt. Soc. Am. A 29(10), 2154–2158 (2012). [CrossRef]  

19. Z. Mei, Z. Tong, and O. Korotkova, “Electromagnetic non-uniformly correlated beams in turbulent atmosphere,” Opt. Express 20(24), 26458–26463 (2012). [CrossRef]   [PubMed]  

20. H. Lajunen and T. Saastamoinen, “Non-uniformly correlated partially coherent pulses,” Opt. Express 21(1), 190–195 (2013). [CrossRef]   [PubMed]  

21. S. Cui, Z. Chen, L. Zhang, and J. Pu, “Experimental generation of nonuniformly correlated partially coherent light beams,” Opt. Lett. 38(22), 4821–4824 (2013). [CrossRef]   [PubMed]  

22. Z. Mei, “Light sources generating self-focusing beams of variable focal length,” Opt. Lett. 39(2), 347–350 (2014). [CrossRef]   [PubMed]  

23. Z. Tong and O. Korotkova, “Nonuniformly correlated light beams in uniformly correlated media,” Opt. Lett. 37(15), 3240–3242 (2012). [CrossRef]   [PubMed]  

24. Y. Gu and G. Gbur, “Scintillation of nonuniformly correlated beams in atmospheric turbulence,” Opt. Lett. 38(9), 1395–1397 (2013). [CrossRef]   [PubMed]  

25. O. Korotkova, S. Sahin, and E. Shchepakina, “Multi-Gaussian Schell-model beams,” J. Opt. Soc. Am. A 29(10), 2159–2164 (2012). [CrossRef]  

26. S. Du, Y. Yuan, C. Liang, and Y. Cai, “Second-order moments of a multi-Gaussian Schell-model beam in a turbulent atmosphere,” Opt. Laser Technol. 50, 14–19 (2013). [CrossRef]  

27. Y. Yuan, X. Liu, F. Wang, Y. Chen, Y. Cai, J. Qu, and H. T. Eyyuboğlu, “Scintillation index of a multi-Gaussian Schell-model beam in turbulent atmosphere,” Opt. Commun. 305, 57–65 (2013). [CrossRef]  

28. O. Korotkova and E. Shchepakina, “Rectangular Multi-Gaussian Schell-Model beams in atmospheric turbulence,” J. Opt. 16(4), 045704 (2014). [CrossRef]  

29. S. Sahin and O. Korotkova, “Light sources generating far fields with tunable flat profiles,” Opt. Lett. 37(14), 2970–2972 (2012). [CrossRef]   [PubMed]  

30. Y. Zhang and D. Zhao, “Scattering of multi-Gaussian Schell-model beams on a random medium,” Opt. Express 21(21), 24781–24792 (2013). [CrossRef]   [PubMed]  

31. F. Wang, C. Liang, Y. Yuan, and Y. Cai, “Generalized multi-Gaussian correlated Schell-model beam: from theory to experiment,” Opt. Express 22(19), 23456–23464 (2014). [CrossRef]   [PubMed]  

32. Y. Zhang, L. Liu, C. Zhao, and Y. Cai, “Multi-Gaussian Schell-model vortex beam,” Phys. Lett. A 378(9), 750–754 (2014). [CrossRef]  

33. Z. Mei and O. Korotkova, “Random sources generating ring-shaped beams,” Opt. Lett. 38(2), 91–93 (2013). [CrossRef]   [PubMed]  

34. F. Wang, X. Liu, Y. Yuan, and Y. Cai, “Experimental generation of partially coherent beams with different complex degrees of coherence,” Opt. Lett. 38(11), 1814–1816 (2013). [CrossRef]   [PubMed]  

35. Y. Chen and Y. Cai, “Generation of a controllable optical cage by focusing a Laguerre-Gaussian correlated Schell-model beam,” Opt. Lett. 39(9), 2549–2552 (2014). [CrossRef]   [PubMed]  

36. Y. Chen, F. Wang, C. Zhao, and Y. Cai, “Experimental demonstration of a Laguerre-Gaussian correlated Schell-model vortex beam,” Opt. Express 22(5), 5826–5838 (2014). [CrossRef]   [PubMed]  

37. Y. Chen, L. Liu, F. Wang, C. Zhao, and Y. Cai, “Elliptical Laguerre-Gaussian correlated Schell-model beam,” Opt. Express 22(11), 13975–13987 (2014). [CrossRef]   [PubMed]  

38. J. Cang, P. Xiu, and X. Liu, “Propagation of Laguerre-Gaussian and Bessel-Gaussian Schell-model beams through paraxial optical systems in turbulent atmosphere,” Opt. Laser Technol. 54, 35–41 (2013). [CrossRef]  

39. R. Chen, L. Liu, S. Zhu, G. Wu, F. Wang, and Y. Cai, “Statistical properties of a Laguerre-Gaussian Schell-model beam in turbulent atmosphere,” Opt. Express 22(2), 1871–1883 (2014). [CrossRef]   [PubMed]  

40. Z. Mei and O. Korotkova, “Electromagnetic cosine-Gaussian Schell-model beams in free space and atmospheric turbulence,” Opt. Express 21(22), 27246–27259 (2013). [CrossRef]   [PubMed]  

41. Z. Mei, E. Shchepakina, and O. Korotkova, “Propagation of cosine-Gaussian-correlated Schell-model beams in atmospheric turbulence,” Opt. Express 21(15), 17512–17519 (2013). [CrossRef]   [PubMed]  

42. H. Xu, Z. Zhang, J. Qu, and W. Huang, “Propagation factors of cosine-Gaussian-correlated Schell-model beams in non-Kolmogorov turbulence,” Opt. Express 22(19), 22479–22489 (2014). [CrossRef]   [PubMed]  

43. Z. Mei and O. Korotkova, “Cosine-Gaussian Schell-model sources,” Opt. Lett. 38(14), 2578–2580 (2013). [CrossRef]   [PubMed]  

44. L. Pan, C. Ding, and H. Wang, “Diffraction of cosine-Gaussian-correlated Schell-model beams,” Opt. Express 22(10), 11670–11679 (2014). [CrossRef]   [PubMed]  

45. C. Liang, F. Wang, X. Liu, Y. Cai, and O. Korotkova, “Experimental generation of cosine-Gaussian-correlated Schell-model beams with rectangular symmetry,” Opt. Lett. 39(4), 769–772 (2014). [CrossRef]   [PubMed]  

46. Z. Mei, “Light sources generating self-splitting beams and their propagation in non-Kolmogorov turbulence,” Opt. Express 22(11), 13029–13040 (2014). [CrossRef]   [PubMed]  

47. Z. Mei, “Two types of sinc Schell-model beams and their propagation characteristics,” Opt. Lett. 39(14), 4188–4191 (2014). [CrossRef]   [PubMed]  

48. Z. Mei and Y. Mao, “Electromagnetic sinc Schell-model beams and their statistical properties,” Opt. Express 22(19), 22534–22546 (2014). [CrossRef]   [PubMed]  

49. Y. Chen, F. Wang, L. Liu, C. Zhao, Y. Cai, and O. Korotkova, “Generation and propagation of a partially coherent vector beam with special correlation functions,” Phys. Rev. A 89(1), 013801 (2014). [CrossRef]  

50. Y. Dan and B. Zhang, “Beam propagation factor of partially coherent flat-topped beams in a turbulent atmosphere,” Opt. Express 16(20), 15563–15575 (2008). [CrossRef]   [PubMed]  

51. Y. Yuan, Y. Cai, J. Qu, H. T. Eyyuboğlu, Y. Baykal, and O. Korotkova, “M2-factor of coherent and partially coherent dark hollow beams propagating in turbulent atmosphere,” Opt. Express 17(20), 17344–17356 (2009). [CrossRef]   [PubMed]  

52. G. Wu, H. Guo, S. Yu, and B. Luo, “Spreading and direction of Gaussian-Schell model beam through a non-Kolmogorov turbulence,” Opt. Lett. 35(5), 715–717 (2010). [CrossRef]   [PubMed]  

53. I. S. Gradshteyn and I. M. Ryzhik, Table of Integrals, Series, and Products, 7th ed. (Academic, 2007).

54. R. N. Bracewell, The Fourier Transform and Its Applications, 3rd ed. (McGraw-Hill, 2000).

55. H. T. Eyyuboğlu, Y. Baykal, and X. Ji, “Radius of curvature variations for annular, dark hollow and flat topped beams in turbulence,” Appl. Phys. B 99(4), 801–807 (2010). [CrossRef]  

56. X. Ji, H. T. Eyyuboğlu, and Y. Baykal, “Influence of turbulence on the effective radius of curvature of radial Gaussian array beams,” Opt. Express 18(7), 6922–6928 (2010). [CrossRef]   [PubMed]  

57. M. Santarsiero, F. Gori, R. Borghi, G. Cincotti, and P. Vahimaa, “Spreading properties of beams radiated by partially coherent Schell-model sources,” J. Opt. Soc. Am. A 16(1), 106–112 (1999). [CrossRef]  

58. M. Charnotskii, “Common omissions and misconceptions of wave propagation in turbulence: discussion,” J. Opt. Soc. Am. A 29, 711–721 (2012). [CrossRef]  

59. M. Charnotskii, “Intensity fluctuations of flat-topped beam in non-Kolmogorov weak turbulence: comment,” J. Opt. Soc. Am. A 29, 1838–1840 (2012). [CrossRef]  

60. C. Chen, H. Yang, M. Kavehrad, and Y. Lou, “Time-dependent scintillations of pulsed Gaussian-beam waves propagating in generalized atmospheric turbulence,” Opt. Laser Technol. 61, 8–14 (2014). [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 (9)

Fig. 1
Fig. 1 Modulus of the transverse spectral degree of coherence of the BGSM beam as a function of ρ1x at several propagation distances in non-Kolmogorov turbulence for different values of parameter β.
Fig. 2
Fig. 2 Modulus of the transverse spectral degree of coherence of the BGSM beam with β = 5 as a function of ρ1x at propagation distance z = 2 km in non-Kolmogorov turbulence for different values of spectral index α and generalized Rytov variance σ R 2.
Fig. 3
Fig. 3 Normalized propagation factor of the BGSM beam for different values of parameter β versus propagation distance z in non-Kolmogorov turbulence with different values of generalized refractive-index structure constant C ˜ n 2.
Fig. 4
Fig. 4 Normalized propagation factor of the BGSM beam versus propagation distance z in non-Kolmogorov turbulence for different values of parameter β and beam width σ.
Fig. 5
Fig. 5 Normalized propagation factor of the BGSM beam with β = 2.5 as a function of propagation distance z in non-Kolmogorov turbulence for different values of (a) outer scale L0 and (b) inner scale l0.
Fig. 6
Fig. 6 Normalized propagation factor of the BGSM beam versus parameter β at propagation distance z = 10 km in non-Kolmogorov turbulence for different values of (a) spatial coherence width δ and (b) wavelength λ.
Fig. 7
Fig. 7 Normalized propagation factor of the BGSM beam versus spatial coherence width δ at propagation distance z = 10 km in non-Kolmogorov turbulence for different values of parameter β.
Fig. 8
Fig. 8 Normalized propagation factor of the BGSM beam versus spectral index α at propagation distance z = 1 km in non-Kolmogorov turbulence with σ R 2 = 0.2 for different values of parameter β.
Fig. 9
Fig. 9 Effective radius of curvature of the BGSM beam propagating in non-Kolmogorov turbulence for different values of parameter β and generalized refractive-index structure constant C ˜ n 2.

Equations (38)

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

W ( r 1 , r 2 , 0 ) = exp ( | r 1 | 2 + | r 2 | 2 4 σ 2 ) μ ( r 1 , r 2 , 0 )
μ ( r 1 , r 2 , 0 ) = exp [ | r 1 r 2 | 2 2 δ 2 ] J 0 ( β | r 1 r 2 | δ )
W ( ρ s , ρ d , z ) = ( k 2 π z ) 2 W ( r s , r d , 0 ) × exp [ i k z ( ρ s r s ) ( ρ d r d ) ] × exp [ H ( ρ d , r d , z ) ] d 2 r s d 2 r d
r s = r 1 + r 2 2 , r d = r 1 r 2
ρ s = ρ 1 + ρ 2 2 , ρ d = ρ 1 ρ 2
W ( r s , r d , 0 ) = W ( r 1 , r 2 , 0 ) = W ( r s + r d 2 , r s r d 2 , 0 )
H ( ρ d , r d , z ) = π 2 k 2 z T 3 [ r d 2 + r d ρ d + ρ d 2 ]
T = 0 κ 3 Φ n ( κ ) d κ
Φ n ( κ , α ) = A ( α ) C ˜ n 2 exp [ ( κ 2 / κ m 2 ) ] ( κ 2 + κ 0 2 ) α / 2 , 0 κ < , 3 < α < 4
c ( α ) = [ Γ ( 5 α 2 ) A ( α ) 2 π 3 ] 1 / ( α 5 )
A ( α ) = Γ ( α 1 ) cos ( α π / 2 ) 4 π 2
T = 0 κ 3 Φ n ( κ ) d κ = A ( α ) 2 ( α 2 ) C ˜ n 2 [ κ m 2 α μ exp ( κ 0 2 κ m 2 ) Γ ( 2 α 2 , κ 0 2 κ m 2 ) 2 κ 0 4 α ]
exp ( p 2 x 2 ± q x ) d x = π p exp ( q 2 4 p 2 ) , ( p > 0 )
W ( ρ s , ρ d , z ) = Q exp [ a r d 2 b ρ s r d + c ρ d r d ] J 0 ( β | r d | δ ) d 2 r d
Q = σ 2 k 2 2 π z 2 exp [ i k z ρ s ρ d ( π 2 k 2 z T 3 + σ 2 k 2 2 z 2 ) ρ d 2 ]
a = 1 8 σ 2 + 1 2 δ 2 + σ 2 k 2 2 z 2 + π 2 k 2 T z 3 , b = i k z , c = σ 2 k 2 z 2 π 2 k 2 T z 3
W ( ρ s , ρ d , z ) = Q 0 0 2 π exp [ a r d 2 | b ρ s c ρ d | r d cos φ ] J 0 ( β r d δ ) r d d r d d φ
0 2 π exp ( ± x cos φ ) d φ = 2 π I 0 ( x )
0 x exp ( γ x 2 ) I ν ( p x ) J ν ( q x ) d x = 1 2 γ exp ( p 2 q 2 4 γ ) J ν ( p q 2 γ ) , [ Re γ > 0 , Re ν > 1 ]
W ( ρ s , ρ d , z ) = σ 2 k 2 2 a z 2 exp [ | b ρ s c ρ d | 2 4 a + i k z ρ s ρ d ( π k 2 z T 3 + σ 2 k 2 2 z 2 ) ρ d 2 ] × exp ( β 2 4 a δ 2 ) J 0 ( β | b ρ s c ρ d | 2 a δ )
μ ( ρ 1 , ρ 2 , z ) = W ( ρ 1 + ρ 2 2 , ρ 1 ρ 2 , z ) S ( ρ 1 , z ) S ( ρ 2 , z )
W ( ρ s , ρ d , z ) = 1 ( 2 π ) 2 W ( r , ρ d + z k κ d , 0 ) × exp [ i ρ s κ d + i κ d r H ( ρ d , ρ d + z k κ d , z ) ] d 2 r d 2 κ d
W ( r , ρ d + z k κ d , 0 ) = exp [ 1 2 σ 2 r 2 ( 1 8 σ 2 + 1 2 δ 2 ) ( ρ d + z k κ d ) 2 ] × J 0 ( β δ | ρ d + z k κ d | )
H ( ρ d , ρ d + z k κ d , z ) = π 2 k 2 z T 3 [ 3 ρ d 2 + 3 z k κ d ρ d + z 2 k 2 κ d 2 ]
h ( ρ s , θ , z ) = ( k 2 π ) 2 W ( ρ s , ρ d , z ) exp ( i k θ ρ d ) d 2 ρ d
h ( ρ s , θ , z ) = k 2 σ 2 8 π 3 exp [ η 1 ρ d 2 η 2 κ d 2 η 3 ρ d κ d i ρ s κ d i k θ ρ d ] × exp [ H ( ρ d , ρ d + z k κ d , z ) ] J 0 ( β δ | ρ d + z k κ d | ) d 2 ρ d d 2 κ d
η 1 = 1 8 σ 2 + 1 2 δ 2 , η 2 = ( 1 8 σ 2 + 1 2 δ 2 ) z 2 k 2 + 1 2 σ 2 , η 3 = ( 1 4 σ 2 + 1 δ 2 ) z k
ρ s x n 1 ρ s y n 2 θ x m 1 θ y m 2 = 1 P ρ s x n 1 ρ s y n 2 θ x m 1 θ y m 2 h ( ρ s , θ , z ) d 2 ρ s d 2 θ
ρ s 2 = ρ s x 2 + ρ s y 2 = 2 σ 2 + ( 1 2 σ 2 + 2 + β 2 δ 2 ) z 2 k 2 + 4 3 π 2 T z 3
θ 2 = θ x 2 + θ x 2 = ( 1 2 σ 2 + 2 + β 2 δ 2 ) 1 k 2 + 4 π 2 T z
ρ s θ = ρ s x θ x + ρ s y θ y = ( 1 2 σ 2 + 2 + β 2 δ 2 ) z k 2 + 2 π 2 T z 2
J ν ( x ) = n = 0 ( 1 ) n ( x / 2 ) 2 n + ν n ! Γ ( n + ν + 1 ) , | x | <
δ ( p x ) = 1 | p | δ ( x )
δ ( n ) ( s ) = 1 2 π ( i x ) n exp ( i s x ) d x , ( n = 0 , 1 , 2 )
f ( x ) δ ( n ) ( x ) d x = ( 1 ) n f ( n ) ( 0 ) , ( n = 0 , 1 , 2 )
R ( z ) = ρ s 2 ρ s θ = 2 σ 2 + ( 1 2 σ 2 + 2 + β 2 δ 2 ) z 2 k 2 + 4 3 π 2 T z 3 ( 1 2 σ 2 + 2 + β 2 δ 2 ) z k 2 + 2 π 2 T z 2
M 2 ( z ) = k ( ρ s 2 θ 2 ρ s θ 2 ) 1 / 2
M 2 ( z ) = [ 2 σ 2 ( 1 2 σ 2 + 2 + β 2 δ 2 ) + 8 π 2 k 2 σ 2 T z + 4 3 π 2 ( 1 2 σ 2 + 2 + β 2 δ 2 ) T z 3 + 4 3 π 4 k 2 T 2 z 4 ] 1 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.