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

Propagation of Gaussian Schell-model beams through a jet engine exhaust

Open Access Open Access

Abstract

Theoretical predictions of light beam interactions with jet engine exhaust are of importance for optimization of various optical systems, including LIDARs, imagers and communication links operating in the vicinity of aircrafts and marine vessels. Here we extend the analysis previously carried out for coherent laser beams propagating in jet engine exhaust, to the broad class of Gaussian Schell-Model (GSM) beams, being capable of treating any degree of coherence in addition to size and radius of curvature. The analytical formulas for the spectral density (SD) and the spectral degree of coherence (DOC) of the GSM beam are obtained and analyzed on passage through a typical jet engine exhaust region. It is shown that for sources with high coherence, the transverse profiles of the SD and the DOC of the GSM beams gradually transition from initially circular to elliptical shape upon propagation at very short ranges. However, such transition is suppressed for sources with lower coherence and disappears in the incoherent source limit, implying that the GSM source with low source coherence is an excellent tool for mitigation of the jet engine exhaust-induced anisotropy of turbulence. The physical interpretation and the illustration are included.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction:

With the rapid development of the optical systems operating through the turbulent boundary layer of Earth, including imaging systems, LIDARs, and free-space optical communications, the statistical changes in light beams propagating through free space and linear random media have been extensively investigated, both theoretically and experimentally [1]. In these studies, the source partial coherence became one of the key mitigation tools of optical turbulence (e.g. [2,3]). After that, the propagation of various classes of partially coherent beams through atmospheric turbulence has received a great deal of attention because of their superiority in self-reshaping of spectra, spectral density and polarization state [423].

In the classic turbulent regime the statistics of the refractive index are predominantly isotropic and homogeneous and can be described by the Kolmogorov’s power spectrum [1]. However, in a number of situations, such as the presence of stratification or the proximity of boundaries, atmospheric turbulence may exhibit different statistical behavior, including anisotropy [24], the presence of coherent structures [25], inverse cascade [26], etc. Since publication of results in [24] the presence of turbulent anisotropy has been measured in a number of campaigns [2732] and studied theoretically, including [33,34] where a convenient theoretical approach was developed (see also [35]) and various statistics of a laser beam interacting with it were calculated, including the long-term beam spread and the scintillation index [36]. Thereafter, the evolution of laser beams and also various partially coherent beams through atmospheric turbulence with anisotropic non-Kolmogorov power spectrum were investigated in a large number of studies, mostly theoretically and by in-lab simulations involving spatial light modulators and specialized air chambers [3752].

On the other hand, recently airborne laser systems arouse considerable interest in view of a number of potential applications relating to security of aircrafts and communications between them [5358]. Usually, airborne platforms meet some challenges for laser systems because of their substantial external perturbations, mainly caused by turbulence in the neighborhood of the jet engine plume. Moreover, it was shown that the power spectrum of the jet engine plume carries anisotropic statistics and might produce similar effects on optical beams, as compared to those caused by natural anisotropic turbulence, e.g., unequal beam spread along two mutually orthogonal directions, transverse to the optical axis.

Therefore, it is quite important not only to investigate the propagation of laser beams through jet engine exhaust (as was already done [53]) but also suggest the ways to mitigate the turbulence-induced beam degradation. The goal of this paper is to suggest the use of partial coherence for mitigation of jet engine plume effects on optical beams, primarily on the anisotropy acquired by the beam’s spectral density (average intensity) and its degree of coherence. As the model for this study we use the beams radiated by scalar Gaussian Schell-Model (GSM) sources, which provide the access to arbitrary choice of size, radius of curvature and degree of coherence [59].

2. Propagation of GSM beams through a jet engine exhaust

Consider a GSM source located in the z = 0 plane and radiating a beam-like optical field propagating along the z-direction, in a jet engine exhaust (see Fig. 1). The two-dimensional position vector in the source plane is denoted as r = (x, y). Let the jet stream plume be ejected from the positive z-direction, along the optical axis. After propagation at distance z from the source, the position vector in the plane transverse to beam axis is denoted as ρ=(ξ,η). The second-order statistical properties of the GSM source are characterized by the cross-spectral density (CSD) function of the form [59]

$${W^{(0)}}({{\boldsymbol r}_1},{{\boldsymbol r}_2},\omega ) = \exp \left( { -\frac{{{\boldsymbol r}_1^2 + {\boldsymbol r}_2^2}}{{4\sigma_0^2}}}\right)\exp \left[ { - \frac{{{{({{\boldsymbol r}_2} -{{\boldsymbol r}_1})}^2}}}{{2\delta_0^2}}} \right],$$
where r1=(x1, y1) and r2=(x2, y2) are two-dimensional position vectors of two source points; σ0 and δ0 denote the incident root-mean-squared (rms) beam width and rms coherence width, respectively, ω stands for angular frequency. For brevity, we will ignore the dependence of the CSD function on ω throughout the paper.

 figure: Fig. 1.

Fig. 1. Schematic diagram for propagation of a GSM beam through a jet engine exhaust.

Download Full Size | PDF

According to the extended Huygens-Fresnel principle and paraxial approximation, the CSD function of a partially coherent beam propagating through a linear turbulent medium at distance z from its source can be expressed as the following formula [59]

$$\begin{aligned} W({{{\boldsymbol \rho }_1},{{\boldsymbol \rho }_2},z} ) &= \frac{1}{{{\lambda ^2}{z^2}}}\int\limits_{ - \infty }^\infty {\int\limits_{ - \infty }^\infty {{W^{(0)}}({{{\boldsymbol r}_1},{{\boldsymbol r}_2}} )} } \exp \left[ { - \frac{{ik}}{{2z}}{{({{\boldsymbol \rho }_1} - {{\boldsymbol r}_1})}^2} + \frac{{ik}}{{2z}}{{({{\boldsymbol \rho }_2} - {{\boldsymbol r}_2})}^2}} \right]\\ &\quad \times \left\langle {\exp [{\psi \ast ({{\boldsymbol r}_1},{{\boldsymbol \rho }_1},z) + \psi ({{\boldsymbol r}_2},{{\boldsymbol \rho }_2},z)} ]} \right\rangle {d^2}{{\boldsymbol r}_1}{d^2}{{\boldsymbol r}_2}, \end{aligned}$$
where ρ1=(ξ1, η1) and ρ2=(ξ2, η2) denote transverse two points in the plane z > 0. Here, k = 2π/λ is the wave number with λ being the wave length of light, while ψ(r, ρ, z) denotes the complex phase perturbation induced by the refractive-index fluctuations of the random medium from point (r, 0) to point (ρ, z). The star denotes complex conjugate and the angular brackets stand for ensemble average over the medium fluctuations.

According to Ref. [60] (see also [39]), the term in the angular brackets <·> in Eq. (2) can be expressed, for general anisotropic power spectra of turbulence, as

$$\begin{array}{l} \left\langle {\exp [{\psi \ast ({{\boldsymbol r}_1},{{\boldsymbol \rho }_1},z) + \psi ({{\boldsymbol r}_2},{{\boldsymbol \rho }_2},z)} ]} \right\rangle \\ \qquad = \exp \left\{ { - 2\pi {k^2}z\int\limits_0^1 {dt\int {{d^2}{\boldsymbol \kappa }{\Phi _n}({\boldsymbol \kappa })[1 - \exp (t{{\boldsymbol \rho }_d} \cdot {\boldsymbol \kappa } + (1 - t){{\boldsymbol r}_d} \cdot {\boldsymbol \kappa })]} } } \right\}, \end{array}$$
where ${{\boldsymbol \rho }_d} = {{\boldsymbol \rho }_1} - {{\boldsymbol \rho }_2}$ and ${{\boldsymbol r}_d} = {{\boldsymbol r}_1} - {{\boldsymbol r}_2}$, and ${\boldsymbol \kappa } = ({\kappa _x},{\kappa _y})$ is the two-dimensional spatial frequency vector. In this paper, we will consider propagation of optical beams within an exhaust region from a jet engine with anisotropic power spectrum of Ref. [53]
$${\Phi _n}({\boldsymbol \kappa }) = 0.033C_n^2 \cdot \left\{ \begin{array}{l} \frac{{{{({{L_{0x}}{L_{0y}}} )}^{11/6}}}}{{{{[{1 + {{({\kappa_x}{L_{0x}})}^2} + {{({\kappa_y}{L_{0y}})}^2}} ]}^{^{11/6}}}}}\exp \left( { - \frac{{\kappa_x^2}}{{\kappa_{mx}^2}} - \frac{{\kappa_y^2}}{{\kappa_{my}^2}}} \right)\\ \quad + Q{\left[ {{{\left( {\frac{{2\pi }}{{{L_s}}}} \right)}^2} + {\kappa^2}} \right]^{ - \frac{{11}}{6}}}\exp ({ - {{{\kappa^2}} \mathord{\left/ {\vphantom {{{\kappa^2}} {\kappa_m^2}}} \right.} {\kappa_m^2}}} )\end{array} \right\}, $$
where L0x and L0y are outer scales in the x and y directions, respectively, κm=c(α)/l0 with ${l_0} = \sqrt {l_{0x}^2 + l_{0y}^2} $ being the inner scale, κmx=c(α)/l0x, κmy=c(α)/l0y with l0x and l0y being the inner scales in the x and y directions, respectively. Comparing with Eq. (2) of Ref. [53], we implement in Eq. (4) the finite inner scale, for the sake of the integral convergence and for obtaining more realistic results (see also Ref. [39]). Provided l0→0, the exponential functions in Eq. (4) will tend to 1, implying that the power spectrum will reduce to Eq. (2) of Ref. [53]. Further, we set for the general power slope α the coefficient:
$$c(\alpha )= {\left[ {\frac{{2\pi \Gamma (5 - \alpha /2)A(\alpha )}}{3}} \right]^{\frac{1}{{\alpha - 5}}}}, $$
where A(α) is defined by expression
$$A(\alpha )= \frac{1}{{4{\pi ^2}}}\Gamma ({\alpha - 1} )\cos \left( {\frac{{\alpha \pi }}{2}} \right),\textrm{ }3\;<\;\alpha\;<\;4. $$
For our calculations α = 11/6. We rewrite Eq. (4) as a sum of two terms
$${\Phi _n}({\boldsymbol \kappa }) = {\Phi _{n1}}({\boldsymbol \kappa }) + {\Phi _{n2}}({\boldsymbol \kappa }), $$
where
$${\Phi _{n1}}({\boldsymbol \kappa }) = 0.033C_n^2 \cdot \frac{{{{({{L_{0x}}{L_{0y}}} )}^{11/6}}}}{{{{[{1 + {{({\kappa_x}{L_{0x}})}^2} + {{({\kappa_y}{L_{0y}})}^2}} ]}^{^{11/6}}}}}\exp \left( { - \frac{{\kappa_x^2}}{{\kappa_{mx}^2}} - \frac{{\kappa_y^2}}{{\kappa_{my}^2}}} \right), $$
$${\Phi _{n2}}({\boldsymbol \kappa }) = 0.033C_n^2 \cdot Q{\left[ {{{\left( {\frac{{2\pi }}{{{L_s}}}} \right)}^2} + {\kappa^2}} \right]^{ - \frac{{11}}{6}}}\exp \left( { - \frac{{{\kappa^2}}}{{\kappa_m^2}}} \right). $$
According to Eqs. (8) and (9), we also rewrite Eq. (3) as a sum of two terms:
$$\left\langle {\exp [{\psi \ast ({{\boldsymbol r}_1},{{\boldsymbol \rho }_1},z) + \psi ({{\boldsymbol r}_2},{{\boldsymbol \rho }_2},z)} ]} \right\rangle = A + B, $$
where
$$A = \exp \left\{ { - 2\pi {k^2}z\int\limits_0^1 {dt\int {{d^2}{\boldsymbol \kappa }{\Phi _{n1}}({\boldsymbol \kappa })[1 - \exp (t{{\boldsymbol \rho }_d} \cdot {\boldsymbol \kappa } + (1 - t){{\boldsymbol r}_d} \cdot {\boldsymbol \kappa })]} } } \right\}, $$
$$B = \exp \left\{ { - 2\pi {k^2}z\int\limits_0^1 {dt\int {{d^2}{\boldsymbol \kappa }{\Phi _{n2}}({\boldsymbol \kappa })[1 - \exp (t{{\boldsymbol \rho }_d} \cdot {\boldsymbol \kappa } + (1 - t){{\boldsymbol r}_d} \cdot {\boldsymbol \kappa })]} } } \right\}. $$

For term A in Eq. (11), we use coordinates changes: ${\kappa _x} = {\kappa ^{\prime}_x}/{\mu _x},{\kappa _y} = {\kappa ^{\prime}_y}/{\mu _y}$, where µx and µy denote the anisotropic factors in two mutually orthogonal directions. In the inertial subrange, when the turbulent eddies transport the energy from the outer scale to the inner scale, we can assume that the ratio of the outer and the inner scale in the x direction to the y direction are constants. According to this assumption, the outer and the inner scales in the x and y direction can be expressed as L0x = µxL0, L0yyL0, l0xxl0 and l0yyl0.

Under these assumptions, Eq. (11) can be written as

$$A = \exp \left\{ { - \frac{{2\pi {k^2}z}}{{{\mu_x}{\mu_y}}}\int\limits_0^1 {dt\int\limits_{ - \infty }^\infty {\int\limits_{ - \infty }^\infty {d{{\kappa^{\prime}}_x}d{{\kappa^{\prime}}_y}{{\Phi^{\prime}}_{n1}}({\boldsymbol \kappa^{\prime}})[1 - \exp (t{{{\boldsymbol \rho^{\prime}}}_d} \cdot {\boldsymbol \kappa^{\prime}} + (1 - t){{{\boldsymbol r^{\prime}}}_d} \cdot {\boldsymbol \kappa^{\prime}})]} } } } \right\}, $$
where ${{\boldsymbol \rho ^{\prime}}_d} = ({{{\xi _d}} \mathord{\left/ {\vphantom {{{\xi_d}} {{\mu_x},}}} \right.} {{\mu _x},}}{{{\eta _d}} \mathord{\left/ {\vphantom {{{\eta_d}} {{\mu_y}}}} \right.} {{\mu _y}}})$ and ${{\boldsymbol r^{\prime}}_d} = ({{{x_d}} \mathord{\left/ {\vphantom {{{x_d}} {{\mu_x},}}} \right.} {{\mu _x},}}{{{y_d}} \mathord{\left/ {\vphantom {{{y_d}} {{\mu_y}}}} \right.} {{\mu _y}}})$ resulting in the modified power spectrum
$${\Phi ^{\prime}_{n1}}(\kappa ^{\prime}) = 0.033C_n^2 \cdot {\left( {\frac{{{\mu_x}{\mu_y}L_0^2}}{{1 + {{\kappa^{\prime}}^2}L_0^2}}} \right)^{\frac{{11}}{6}}}\exp \left( { - \frac{{l_0^2}}{{c{{(\alpha )}^2}}}{{\kappa^{\prime}}^2}} \right). $$
We now change the Cartesian coordinates to polar, i.e., $d{\kappa ^{\prime}_x}d{\kappa ^{\prime}_y} = \kappa ^{\prime}d\kappa ^{\prime}d\theta $ and integrate over θ. Then, Eq. (13) takes form
$$A = \exp \left\{ { - \frac{{4{\pi^2}{k^2}z}}{{{\mu_x}{\mu_y}}}\int\limits_0^1 {dt\int\limits_0^\infty {\kappa^{\prime}d\kappa^{\prime}{{\Phi^{\prime}}_{n1}}(\kappa^{\prime})[1 - {J_0}({\kappa^{\prime}|{t{{{\boldsymbol \rho^{\prime}}}_d} + (1 - t){{{\boldsymbol r^{\prime}}}_d}} |} )]} } } \right\}. $$
If the coordinate points are sufficiently close to the optical axis, i.e., under the paraxial approximation, the Bessel function in Eq. (15) can be expanded to be J0(x)≈1-x2/4 by Taylor series. After integrating over t, κ’, we get
$$A = \exp \left\{ { - \frac{{{\pi^2}{k^2}z{T_A}({\xi_d^2 + {\xi_d}{x_d} + x_d^2} )}}{{3\mu_x^2}}} \right\}\exp \left\{ { - \frac{{{\pi^2}{k^2}z{T_A}({\eta_d^2 + {\eta_d}{y_d} + y_d^2} )}}{{3\mu_y^2}}} \right\}, $$
where
$$\begin{aligned} {T_A} &= \frac{1}{{{\mu _x}{\mu _y}}}\int_0^\infty {{{\kappa ^{\prime}}^3}{{\Phi ^{\prime}}_{n1}}(\kappa ^{\prime})d\kappa ^{\prime}} \\ & = 0\textrm{.033}C_n^2{({\mu _x}{\mu _y})^{\frac{5}{6}}}\left[ {\frac{{\left( {5 + \frac{6}{{\kappa_m^2L_0^2}}} \right)\exp \left( {\frac{1}{{\kappa_m^2L_0^2}}} \right)\Gamma \left( {\frac{1}{6},\frac{1}{{\kappa_m^2L_0^2}}} \right)}}{{10{{\left( {\frac{1}{{\kappa_m^2}}} \right)}^{\frac{1}{6}}}}} - \frac{3}{5}{{\left( {\frac{1}{{L_0^2}}} \right)}^{\frac{1}{6}}}} \right] \end{aligned}. $$
On substituting from Eqs. (1) and (16) into Eq. (2) and after integrating, we find that the CSD function of the GSM beam from term A in the plane z > 0 factorizes as:
$${W_A}({{{\boldsymbol \rho }_1},{{\boldsymbol \rho }_2},z} )= {W_{xA}}({{\xi_1},{\xi_2}} ){W_{yA}}({{\eta_1},{\eta_2}} ), $$
where
$$\begin{aligned} {W_{xA}}({{\xi_1},{\xi_2}} ) &= \frac{1}{{\sqrt {{\Delta _{xA}}(z)} }}\exp \left( { - \frac{{\xi_1^2 + \xi_2^2}}{{4\sigma_0^2{\Delta _{xA}}(z)}}} \right)\exp \left[ { - \frac{{ik}}{{2{R_{xA}}(z)}}({\xi_1^2 - \xi_2^2} )} \right]\\ &\quad \times \textrm{exp}\left\{ { - \left[ {\frac{1}{{2\delta_0^2{\Delta _{xA}}(z)}} + \frac{{{\pi^2}{k^2}{T_A}z}}{{3\mu_x^2}}\left( {1 + \frac{2}{{{\Delta _{xA}}(z)}}} \right) - \frac{{{\pi^4}{k^2}T_A^2{z^4}}}{{18\mu_x^4{\Delta _{xA}}(z)\sigma_0^2}}} \right]{{({{\xi_1} - {\xi_2}} )}^2}} \right\}, \end{aligned}$$
and parameters ΔxA(z) and RxA(z) are given by expressions
$${\Delta _{xA}}(z) = 1 + \left[ {\frac{1}{{4{k^2}\sigma_0^4}} + \frac{1}{{{k^2}\sigma_0^2}}\left( {\frac{1}{{\delta_0^2}} + \frac{{2{\pi^2}{k^2}{T_A}z}}{{3\mu_x^2}}} \right)} \right]{z^2}, $$
$${R_{xA}}(z) = z + \frac{{\sigma _0^2z - {\pi ^2}{T_A}{z^4}/3\mu _x^2}}{{({\Delta _{xA}}(z) - 1)\sigma _0^2 + {\pi ^2}{T_A}{z^3}/3\mu _x^2}}. $$
Factor ${W_{yA}}({{\eta_1},{\eta_2}} )$ has the same form as ${W_{xA}}({{\xi_1},{\xi_2}} )$ by using y, η1and η2 in place of x, ξ1and ξ2 in Eq. (19), respectively.

For term B in Eq. (12), we change the integral variables d2κ from the Cartesian coordinates to polar, i.e., d2κ=κdκdθ, and integrate over θ. Then Eq. (12) turns out to be

$$B = \exp \left\{ { - 4{\pi^2}{k^2}z\int\limits_0^1 {dt\int {\kappa d\kappa {\Phi _{n2}}(\kappa )[1 - {J_0}(\kappa |{t{{\boldsymbol \rho }_d} + (1 - t){{\boldsymbol r}_d}} |)]} } } \right\}. $$
By applying approximation J0(x)≈1-x2/4 and integrating over t, κ, we deduce that
$$B = \exp \left[ { - \frac{1}{3}{\pi^2}{k^2}z{T_B}({\xi_d^2 + {\xi_d}{x_d} + x_d^2} )} \right]\exp \left[ { - \frac{1}{3}{\pi^2}{k^2}z{T_B}({\eta_d^2 + {\eta_d}{y_d} + y_d^2} )} \right], $$
where
$$\begin{aligned} {T_B} &= \int_0^\infty {{\kappa ^3}{\Phi _{n2}}(\kappa )d\kappa } \\ & = 0\textrm{.033}C_n^2Q\left[ {\frac{{\left( {5 + \frac{{24{\pi^2}}}{{\kappa_m^2L_s^2}}} \right)\exp \left( {\frac{{4{\pi^2}}}{{\kappa_m^2L_s^2}}} \right)\Gamma \left( {\frac{1}{6},\frac{{4{\pi^2}}}{{\kappa_m^2L_s^2}}} \right)}}{{10{{\left( {\frac{1}{{\kappa_m^2}}} \right)}^{\frac{1}{6}}}}} - \frac{3}{5}{{\left( {\frac{{4{\pi^2}}}{{L_s^2}}} \right)}^{\frac{1}{6}}}} \right] \end{aligned}. $$
On substituting from Eqs. (1) and (23) into Eq. (2) and after integrating, the CSD function of the GSM beam from term B can be derived in the plane z > 0 as:
$$\begin{aligned} {W_B}({{{\boldsymbol \rho }_1},{{\boldsymbol \rho }_2},z} ) &= \frac{1}{{{\Delta _B}(z)}}\exp \left( { - \frac{{{\boldsymbol \rho }_1^2 + {\boldsymbol \rho }_2^2}}{{4\sigma_0^2{\Delta _B}(z)}}} \right)\exp \left[ { - \frac{{ik}}{{2{R_B}(z)}}({{\boldsymbol \rho }_1^2 - {\boldsymbol \rho }_2^2} )} \right]\\ &\quad \times \textrm{exp}\left\{ { - \left[ {\frac{1}{{2\delta_0^2{\Delta _B}(z)}} + \frac{{{\pi^2}{k^2}{T_B}z}}{3}\left( {1 + \frac{2}{{{\Delta _B}(z)}}} \right) - \frac{{{\pi^4}{k^2}T_B^2{z^4}}}{{18{\Delta _B}(z)\sigma_0^2}}} \right]{{({{{\boldsymbol \rho }_1} - {{\boldsymbol \rho }_2}} )}^2}} \right\}, \end{aligned}$$
where parameters ΔB(z) and RB(z) are given by expressions
$${\Delta _B}(z) = 1 + \left[ {\frac{1}{{4{k^2}\sigma_0^4}} + \frac{1}{{{k^2}\sigma_0^2}}\left( {\frac{1}{{\delta_0^2}} + \frac{{2{\pi^2}{k^2}{T_B}z}}{3}} \right)} \right]{z^2}$$
$${R_B}(z) = z + \frac{{\sigma _0^2z - {\pi ^2}{T_B}{z^4}/3}}{{({\Delta _B}(z) - 1)\sigma _0^2 + {\pi ^2}{T_B}{z^3}/3}}. $$
For scalar random beams the spectral density (SD) and the spectral degree of coherence (DOC) are defined by expressions [59]
$$S({\boldsymbol \rho },\omega ) = {W_{}}({{\boldsymbol \rho },{\boldsymbol \rho },z} )= {W_A}({{\boldsymbol \rho },{\boldsymbol \rho },z} )+ {W_B}({{\boldsymbol \rho },{\boldsymbol \rho },z} ), $$
$$\begin{aligned} \mu ({{\boldsymbol \rho }_1},{{\boldsymbol \rho }_2},z) &= \frac{{{W_{}}({{{\boldsymbol \rho }_1},{{\boldsymbol \rho }_2},z} )}}{{\sqrt {S({{\boldsymbol \rho }_1},z)S({{\boldsymbol \rho }_2},z)} }}\\ & = \frac{{{W_A}({{{\boldsymbol \rho }_1},{{\boldsymbol \rho }_2},z} )+ {W_B}({{{\boldsymbol \rho }_1},{{\boldsymbol \rho }_2},z} )}}{{\sqrt {[{{W_A}({{{\boldsymbol \rho }_1},{{\boldsymbol \rho }_1},z} )+ {W_B}({{{\boldsymbol \rho }_1},{{\boldsymbol \rho }_1},z} )} ][{{W_A}({{{\boldsymbol \rho }_2},{{\boldsymbol \rho }_2},z} )+ {W_B}({{{\boldsymbol \rho }_2},{{\boldsymbol \rho }_2},z} )} ]} }}. \end{aligned}$$

3. Numerical calculation and analysis

In this section, we will illustrate the evolution of the SD and the DOC of the GSM beam propagating through a jet engine exhaust according to the analytical expressions derived in section 2 by a number of numerical examples. The parameters used in numerical calculations are chosen to be σ0=5mm, δ0=5mm, λ=632.8nm, $C_n^2 = 1.6 \times {10^{ - 9}}{\textrm{m}^{ - 2/3}}$, L0=0.5m, LS=1mm, µx=0.7, µy=1.4 and Q = 6 [53] unless different values are specified. In addition, we let the inner scale length be l0=2/3mm ≈ 0.67mm, which originates from the experimental results of Ref. [61] (see also Ref. [62]). In Fig. 2, we plot the anisotropic power spectrum Φn of the refractive index fluctuations along with x and y directions, according to Eq. (4), for different values of inner scale l0. It is shown that the inner scale length has a noticeable impact on the power spectrum only in the high spatial frequency region (above 103 m−1) if l0=1mm, or larger, and the influence of the inner scale on the power spectrum is almost negligible when the l0=0.01mm or smaller.

 figure: Fig. 2.

Fig. 2. Anisotropic power spectrum of the refractive-index fluctuations, from the jet engine exhaust along the x and y directions, according to Eq. (4).

Download Full Size | PDF

Figure 3 presents the density plots of the SD of the GSM beams at some propagation distances for three different values of source coherence width:δ0=5mm, δ0=0.3mm and δ0=5µm. We note right away that our analysis only pertains to very small propagating distances, on the order of meters, in striking difference to the natural anisotropic turbulence where the beam can be analyzed on propagation at kilometer ranges and farther, depending on the local strength of the refractive-index fluctuations (see Ref. [39]). We plot the results at 10 m and 20 m to illustrate that anisotropy effects may increase further with increasing distances, even though such ranges might be out of the realistic scenario. However, if a beam is affected by a jet stream in the beginning of a path and then travels much further in free space or in a natural atmospheric turbulence, it can be considered as an effective phase screen [1] equivalent to somewhat milder extended turbulence with the same anisotropy ratio in transverse direction.

 figure: Fig. 3.

Fig. 3. The spectral density of a GSM beam at some propagation distances from the source for three different values of the source coherence width (First row:δ0=5mm, second row:δ0=0.3mm, third row: δ0=5µm).

Download Full Size | PDF

It is seen from Figs. 3(a)–3(d) that, for the nearly coherent source, the beam profile transitions from circular to elliptical upon propagation. However, with decreasing coherence width, the transition from circular shape to elliptical becomes less noticeable [see Figs. 3(e)–3(h)]. When the source coherence width become much smaller, e.g. δ0=5µm, the SD profile remains circular up to the propagation distance of about z = 20 m. Hence, the GSM beam with very low coherence may weaken the anisotropic effect of turbulence generated from a jet engine as compared to that with highly coherent source.

Next, we will perform some explanations and illustrations. The SD of the GSM beam is a sum of WA in Eq. (18) and WB in Eq. (25). In Eq. (25), there is no anisotropic term. Therefore, WA plays a more important role on determining the anisotropic features of the SD. Hence, in the following, we will emphasize the effect of WA on the whole SD.

The root-mean-square beam width of the GSM beam from term A along the x or y direction in the plane z > 0 can be found by comparing Eq. (18) with Eq. (1), i.e.,

$${\sigma _i}(z) = {\sigma _0}\sqrt {{\Delta _{iA}}(z)} = {\left[ {\sigma_0^2 + \frac{{{z^2}}}{{4{k^2}\sigma_0^2}} + \frac{{{z^2}}}{{{k^2}\delta_0^2}} + \frac{{2{\pi^2}{T_A}{z^3}}}{{3\mu_i^2}}} \right]^{\frac{1}{2}}},\textrm{ }(i = x,y). $$
Equation (30) implies that the propagating beam width has four contributions: the first term is the initial SD width; the second term and third term indicate the free-space diffraction rate of the GSM beam upon propagation, induced by σ0 and δ0, respectively. Diffraction of the second term and third term is isotropic, depending only on the propagation distance and the wavelength. The fourth term indicates turbulence-induced diffraction, which is anisotropic and dependent on the anisotropy parameter µi. Figure 4 shows the second-fourth terms of Eq. (30) versus propagation distance, for four different values of source coherence width. It is clearly seen that the influence of the fourth term is very small when the propagation distance is short. Hence, the SD profile remains circular at small ranges from the source. For the high source coherence δ0=5 mm, with increasing propagation distance z, turbulence-induced diffraction begins to play a more important role because the value of the fourth term changes with z3, which is much greater than those of the second and third term, i.e., z2. As a result, the SD obtains an elliptical shape when the propagation distance is sufficiently large, e.g. z = 20 m. However, with the decrease in the source coherence, the coherence-induced diffraction gradually plays a major role comparing with the turbulence-induced diffraction [see third term in Fig. 4(d)]. In addition, the third term is isotropic. That is why the SD profile retains circular profile when the source coherence is very low. What’s more, the contribution from the fourth term in the x direction is larger than that of the fourth term in the y direction. Hence, the major axis of the SD’s elliptical distribution is the x axis.

 figure: Fig. 4.

Fig. 4. Second-fourth terms of Eq. (30) versus propagation distance, for four different values of source coherence width.

Download Full Size | PDF

The density plots for the modulus of the DOC of the GSM beams at two points ρ1=(ξ, η) and ρ2=(0, 0) are given in Fig. 5. As can be seen from Figs. 5(a)–5(d) the DOC distribution exhibits elliptical shape upon propagation when the source coherence width is large, e.g., δ0=5 mm. And the major axis of the DOC profile is along y direction (in contrast with that for the SD profile). However, with the decreasing source coherence width, the evolution of the DOC profile from circular to elliptical occurs at smaller ranges [see Figs. 5(f)–5(h) and 5(j)–5(l)]. For example, when δ0=5µm, the DOC profile is almost circular at the propagation distance z = 5 m. For larger propagation distances, the DOC profile starts acquiring elliptical shape. In order to give the physical explanation, we examine the root-mean-square value of the DOC of the GSM beam, obtained by comparison of Eq. (18) with Eq. (1):

$${\delta _i}(z) = {\left[ {\frac{1}{{\delta_0^2{\Delta _{iA}}(z)}} + \frac{{2{\pi^2}{k^2}{T_A}z}}{{3\mu_i^2}} + \frac{{2{\pi^2}{k^2}{T_A}z}}{{3\mu_i^2{\Delta _{iA}}(z)}}\left( {2 - \frac{{{\pi^2}{T_A}{z^3}}}{{6\mu_i^2\sigma_0^2}}} \right)} \right]^{ - \frac{1}{2}}},(i = x,y). $$
Equation (31) shows that all of three terms in the square brackets affect the coherence width of the GSM beam. The first term $\textrm{1/}\delta _0^2{\Delta _{iA}}(z)$ indicates the coherence-induced diffraction, which results in the increase of coherence width upon propagation. The second term is the turbulence-induced decoherence effects, attenuating the beam coherence with increasing propagation distance. In the third term, there are complex interactions between the coherence-induced diffraction and the turbulence-induced decoherence effect, which result in a somewhat complicated evolution of the DOC.

 figure: Fig. 5.

Fig. 5. Density plots for the modulus of the DOC of a GSM beam at some propagation distances for three different values of the source coherence width (First row:δ0=5mm, second row:δ0=0.3mm, third row: δ0=5µm).

Download Full Size | PDF

In Fig. 6, we demonstrate the changes in the first-third terms of Eq. (31) and with propagation distance z, for three different values of the source coherence width. As can be seen from Figs. 6(a) and 6(b) the influence of the first term on the coherence width is practically negligible, while the third term displays the non-monotonic changes near the value of zero, with increasing propagation distance z. However, the contribution from the second term becomes dominant already after a short propagation range. Hence, the DOC distribution exhibits elliptical shape upon propagation at relatively short distances when the source coherence widthδ0=5 mm. Moreover, when δ0→∞, i.e., in the coherent limit case, the effect of the first term can be ignored. With decreasing source coherence width, the effect of the first term is dominant for short propagation distance [see Figs. 6(e) and 6(f)], playing a more important role with the decrease of source coherence [see Figs. 6(g) and 6(h)]. Thus, the overall effect of anisotropy of turbulence on the DOC distribution is very weak for short propagation distance. That is why the DOC profile remains nearly circular for short propagation distance [see Fig. 5(j)]. For large propagation distances, the second term becomes dominant, relating to the turbulence-induced decoherence effect. Hence, at sufficiently large distances from the source plane, the DOC always keeps elliptical shape.

 figure: Fig. 6.

Fig. 6. First-third terms of Eq. (31) versus propagation distance, for four different values of the source coherence width.

Download Full Size | PDF

In Fig. 7 we show the evolution of the coherence width with propagation distance for different values of the source coherence widths. It is shown that the coherence widths in the x/y direction decreases with increasing propagation distance when the source coherence width is large [see Fig. 7(a) and 7(b)]. However, for low source coherence, the coherence width of the propagating beam displays the non-monotonic changes with increasing propagation distance [see Figs. 7(c) and 7(d)].

 figure: Fig. 7.

Fig. 7. Coherence widthin x and y direction versus propagation distance for four different values of source coherence width.

Download Full Size | PDF

4. Conclusion

We have investigated in detail the interaction of the scalar GSM beam with a region in space containing a jet engine exhaust. Such medium can be described by means of an anisotropic power spectrum of refractive index fluctuations with Kolmogorov power law, having very high values of the refractive-index structure parameter and very low values of the inner scale. Hence, as compared with natural turbulence, an optical beam propagating in such medium is mostly affected by small scales, which produce the scattering-like effects and limit the paraxial regime to very small ranges. But even over these small distances the effect of anisotropy can be substantial.

Our main contribution to this research area are the derived analytical expressions for the SD and the DOC of the GSM beams and their analysis upon propagation, as a whole quantity and segregated into several terms, indicating the effects of the source coherence, of the turbulent anisotropy and their interplay. Our numerical examples illustrate that the profile of the SD and the DOC of the GSM beams gradually transfer from a circular shape to an elliptical shape upon propagation for highly coherence sources. However, when the source coherence width is short, the SD’s shape remains circular for long propagation distance, while the DOC’s shape remains circular shape for short propagation distance, implying that the GSM beams with sufficiently low source coherence are insusceptible to the anisotropy of turbulence form jet engine exhaust, and thus, can be used as the mitigation tool. This is the major practical outcome of our study. Another possibility, not explored in this paper, but readily seen to be feasible, is modulation of the source DOC in anisotropic manner [63]. If the major semi-axis of the source DOC is orthogonal to that of the jet stream, then at some propagation distances their effects must cancel. This idea can be exploited when the directionality of the beam is also of importance. The obtained results might have some potential applications in optical imaging, ranging, tracking and communication systems that are to operate through the immediate jet engine exhaust environments.

Funding

National Natural Science Foundation of China (61675094, 61575091, 11874321).

Disclosures

The authors declare that there are no conflicts of interest related to this article.

References

1. L. C. Andrews and R. L. Phillips, Laser Beam Propagation through Random Media, 2nd edition ed. (SPIE Publications Bellingham, Washington, 2005).

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

3. 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]  

4. H. Roychowdhury and E. Wolf, “Invariance of spectrum of light generated by a class of quasi-homogenous sources on propagation through turbulence,” Opt. Commun. 241(1-3), 11–15 (2004). [CrossRef]  

5. O. Korotkova, M. Salem, and E. Wolf, “The far-zone behavior of the degree of polarization of electromagnetic beams propagating through atmospheric turbulence,” Opt. Commun. 233(4-6), 225–230 (2004). [CrossRef]  

6. O. Korotkova, M. Salem, A. Dogariu, and E. Wolf, “Changes in the polarization ellipse of random electromagnetic beams propagating through the turbulent atmosphere,” Wave Random Complex 15(3), 353–364 (2005). [CrossRef]  

7. Y. J. Cai and S. L. He, “Propagation of a partially coherent twisted anisotropic Gaussian Schell-model beam in a turbulent atmosphere,” Appl. Phys. Lett. 89(4), 041117 (2006). [CrossRef]  

8. Y. B. Zhu, D. M. Zhao, and X. Y. Du, “Propagation of stochastic Gaussian-Schell model array beams in turbulent atmosphere,” Opt. Express 16(22), 18437–18442 (2008). [CrossRef]  

9. X. L. Ji and G. M. Ji, “Effect of turbulence on the beam quality of apertured partially coherent beams,” J. Opt. Soc. Am. A 25(6), 1246–1252 (2008). [CrossRef]  

10. X. Y. Du and D and M. Zhao, “Polarization modulation of stochastic electromagnetic beams on propagation through the turbulent atmosphere,” Opt. Express 17(6), 4257–4262 (2009). [CrossRef]  

11. H. Mao and D. Zhao, “Second-order intensity-moment characteristics for broadband partially coherent flat-topped beams in atmospheric turbulence,” Opt. Express 18(2), 1741–1755 (2010). [CrossRef]  

12. Y. Gu and G. Gbur, “Scintillation of pseudo-Bessel correlated beams in atmospheric turbulence,” J. Opt. Soc. Am. A 27(12), 2621–2629 (2010). [CrossRef]  

13. X. L. Ji and X. Q. Li, “M2-factor of truncated partially coherent beams propagating through atmospheric turbulence,” J. Opt. Soc. Am. A 28(6), 970–975 (2011). [CrossRef]  

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

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

16. Z. R. 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]  

17. G. Gbur, “Partially coherent beam propagation in atmospheric turbulence [Invited],” J. Opt. Soc. Am. A 31(9), 2038–2045 (2014). [CrossRef]  

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

19. J. Yu, Y. H. Chen, L. Liu, X. L. Liu, and Y. J. Cai, “Splitting and combining properties of an elegant Hermite-Gaussian correlated Schell-model beam in Kolmogorov and non-Kolmogorov turbulence,” Opt. Express 23(10), 13467–13481 (2015). [CrossRef]  

20. C. L. Ding, M. Koivurova, J. Turunen, and L. Z. Pan, “Self-focusing of a partially coherent beam with circular coherence,” J. Opt. Soc. Am. A 34(8), 1441–1447 (2017). [CrossRef]  

21. F. Wang and O. Korotkova, “Circularly symmetric cusped random beams in free space and atmospheric turbulence,” Opt. Express 25(5), 5057–5067 (2017). [CrossRef]  

22. Z. L. Liu and D. M. Zhao, “Spectral density of vortex beams propagating in atmospheric turbulence with new simulation method,” Opt. Lett. 43(11), 2478–2481 (2018). [CrossRef]  

23. J. Li, X. Chen, S. McDuffie, M. A. M. Najjar, S. M. H. Rafsanjani, and O. Korotkova, “Mitigation of atmospheric turbulence with random light carrying OAM,” Opt. Commun. 446, 178–185 (2019). [CrossRef]  

24. A. Consortini, L. Ronchi, and L. Stefanutti, “Investigation of Atmospheric Turbulence by Narrow Laser Beams,” Appl. Opt. 9(11), 2543–2547 (1970). [CrossRef]  

25. J. Jimenez, “Coherent structures in wall-bounded turbulence,” J. Fluid Mech. 842, P1 (2018). [CrossRef]  

26. L. Biferale, S. Musacchio, and F. Toschi, “Inverse Energy Cascade in Three-Dimensional Isotropic Turbulence,” Phys. Rev. Lett. 108(16), 164501 (2012). [CrossRef]  

27. M. S. Belen’kii, J. D. Barchers, S. J. Karis, C. L. Osmon, J. M. Brown II, and R. Q. Fugate, “Preliminary experimental evidence of anisotropy of turbulence and the effect of non-kolmogorov turbulence on wavefront tilt statistics,” in Proceedings of SPIE, (SPIE, 1999), 396–406.

28. M. S. Belen’kii, S. J. Karis, C. L. Osmon, J. M. Brown II, and R. Q. Fugate, “Experimental evidence of the effects of non-Kolmogorov turbulence and anisotropy of turbulence,” in Proceedings of SPIE, (SPIE, 1999), 50–51.

29. J. L. Otten, M. C. Roggemann, B. Al Jones, J. Lane, and D. G. Black, “High bandwidth atmospheric turbulence data collection platform,” in Proceedings of SPIE (SPIE, 1999), 23–32.

30. F. Wang, I. Toselli, J. Li, and O. Korotkova, “Measuring anisotropy ellipse of atmospheric turbulence by intensity correlations of laser light,” Opt. Lett. 42(6), 1129–1132 (2017). [CrossRef]  

31. M. Beason, C. Smith, J. Coffaro, S. Belichki, J. Spychalsky, F. Titus, R. Crabbs, L. Andrews, and R. Phillips, “Near ground measure and theoretical model of plane wave covariance of intensity in anisotropic turbulence,” Opt. Lett. 43(11), 2607–2610 (2018). [CrossRef]  

32. M. Beason, J. Coffaro, C. Smith, J. Spychalsky, S. Belichki, F. Titus, F. Sanzone, B. Berry, R. Crabbs, L. Andrews, and R. Phillips, “Evolution of near-ground optical turbulence over concrete runway throughout multiple days in summer and winter,” J. Opt. Soc. Am. A 35(8), 1393–1400 (2018). [CrossRef]  

33. I. Toselli, L. C. Andrews, R. L. Phillips, and V. Ferrero, “Free space optical system performance for laser beam propagation through non Kolmogorov turbulence,” in Proceedings of SPIE, (SPIE, 2007), 64570T.

34. I. Toselli, L. C. Andrews, R. L. Phillips, and V. Ferrero, “Free-space optical system performance for laser beam propagation through non-Kolmogorov turbulence,” Opt. Eng. 47(2), 026003 (2008). [CrossRef]  

35. L. C. Andrews, R. L. Phillips, and R. Crabbs, “Propagation of a Gaussian-beam wave in general anisotropic turbulence,” in Proceedings of SPIE, (SPIE, 2014), 922402.

36. I. Toselli, B. Agrawal, and S. Restaino, “Light propagation through anisotropic turbulence,” J. Opt. Soc. Am. A 28(3), 483–488 (2011). [CrossRef]  

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

38. I. Toselli, O. Korotkova, X. F. Xiao, and D. G. Voelz, “SLM-based laboratory simulations of Kolmogorov and non-Kolmogorov anisotropic turbulence,” Appl. Opt. 54(15), 4740–4744 (2015). [CrossRef]  

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

40. Y. Zhu, M. Y. Chen, Y. X. 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(12), 2277–2283 (2016). [CrossRef]  

41. Y. H. Zhao, Y. X. Zhang, Z. D. Hu, Y. Li, and D. L. Wang, “Polarization of quantization Gaussian Schell-beams through anisotropic non-Kolmogorov turbulence of marine-atmosphere,” Opt. Commun. 371, 178–183 (2016). [CrossRef]  

42. J. Wang, S. J. Zhu, H. Y. Wang, Y. J. Cai, and Z. H. 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]  

43. D. Zhi, R. M. Tao, P. Zhou, Y. X. Ma, W. M. Wu, X. L. 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]  

44. Y. Zhu, Y. X. Zhang, and G. F. Yang, “Evolution of orbital angular momentum mode of the autofocusing Hypergeometric-Gaussian beams through moderate-to-strong anisotropic non-Kolmogorov turbulence,” Opt. Commun. 405, 66–72 (2017). [CrossRef]  

45. M. J. Cheng, L. X. Guo, and J. T. Li, “Influence of moderate-to-strong anisotropic non-Kolmogorov turbulence on intensity fluctuations of a Gaussian-Schell model beam in marine atmosphere,” Chin. Phys. B 27(5), 054203 (2018). [CrossRef]  

46. J. Wang, H. K. Huang, Y. Chen, H. Y. Wang, S. J. Zhu, Z. H. Li, and Y. J. Cai, “Twisted partially coherent array sources and their transmission in anisotropic turbulence,” Opt. Express 26(20), 25974–25988 (2018). [CrossRef]  

47. Y. P. Huang and A. P. Zeng, “Effect of anisotropic non-Kolmogorov turbulence on the evolution behavior of Gaussian Schell-model vortex beams,” Opt. Commun. 436, 63–68 (2019). [CrossRef]  

48. M. Beason, L. Andrews, and I. Toselli, “Calculating structure function constant from measured C-n(2) in non-Kolmogorov and anisotropic turbulence including inner scale effects,” Appl. Opt. 58(25), 6813–6819 (2019). [CrossRef]  

49. J. Zeng, X. L. Liu, C. L. Zhao, F. Wang, G. Gbur, and Y. J. Cai, “Spiral spectrum of a Laguerre-Gaussian beam propagating in anisotropic non-Kolmogorov turbulent atmosphere along horizontal path,” Opt. Express 27(18), 25342–25356 (2019). [CrossRef]  

50. Y. Ata, Y. Baykal, and M. C. Gokce, “Average channel capacity in anisotropic atmospheric non-Kolmogorov turbulent medium,” Opt. Commun. 451, 129–135 (2019). [CrossRef]  

51. X. Xiao, D. G. Voelz, I. Toselli, and O. Korotkova, “Gaussian beam propagation in anisotropic turbulence along horizontal links: theory, simulation, and laboratory implementation,” Appl. Opt. 55(15), 4079–4084 (2016). [CrossRef]  

52. G. Funes, F. Olivares, C. G. Weinberger, Y. D. Carrasco, L. Nuñez, and D. G. Pérez, “Synthesis of anisotropic optical turbulence at the laboratory,” Opt. Lett. 41(24), 5696–5699 (2016). [CrossRef]  

53. L. Sjöqvist, “Laser beam propagation in jet engine plume environments: a review,” in Proceedings of SPIE, (SPIE, 2008), 71150C.

54. C. B. Hogge and W. L. Visinsky, “Laser Beam Probing of Jet Exhaust Turbulence,” Appl. Opt. 10(4), 889–892 (1971). [CrossRef]  

55. J. L. Barrett and P. A. Budni, “Laser beam propagation through strong turbulence,” J. Appl. Phys. 71(3), 1124–1127 (1992). [CrossRef]  

56. M. Henriksson, L. Sjöqvist, and O. Gustafsson, “Experimental study of mid-IR laser beam wander close to a jet engine exhaust,” in Proceedings of SPIE, (SPIE, 2006), 639709.

57. V. S. Sirazetdinov, “Experimental study and numerical simulation of laser beams propagation through the turbulent aerojet,” Appl. Opt. 47(7), 975–985 (2008). [CrossRef]  

58. M. Henriksson, L. Sjöqvist, D. Seiffer, N. Wendelstein, and E. Sucher, “Laser beam propagation experiments along and across a jet engine plume,” in Proceedings of SPIE, (SPIE, 2008), 71150E.

59. O. Korotkova, Random Light Beams:Theory and Applications (CRC Press, Boca Raton, 2013).

60. H. T. Yura, “Mutual Coherence Function of a Finite Cross Section Optical Beam Propagating in a Turbulent Medium,” Appl. Opt. 11(6), 1399–1406 (1972). [CrossRef]  

61. V. S. Sirazetdinov, I. V. Ivanova, A. D. Starikov, D. H. Titterton, T. A. Sheremetyeva, G. N. Filippov, and Y. N. Yevchenko, “Experimental study of the structure of laser beams disturbed by turbulent stream of aircraft engine,” in Proceedings of SPIE, (SPIE, 2000), 397–405.

62. L. Sjöqvist, O. Gustafsson, and M. Henriksson, “Laser beam propagation in close vicinity to a down-scaled jet engine exhaust,” in Proceedings of SPIE, (SPIE, 2004), 137–148.

63. Y. Li and E. Wolf, “Radiation from anisotropic Gaussian Schell-model sources,” Opt. Lett. 7(6), 256–258 (1982). [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 diagram for propagation of a GSM beam through a jet engine exhaust.
Fig. 2.
Fig. 2. Anisotropic power spectrum of the refractive-index fluctuations, from the jet engine exhaust along the x and y directions, according to Eq. (4).
Fig. 3.
Fig. 3. The spectral density of a GSM beam at some propagation distances from the source for three different values of the source coherence width (First row:δ0=5mm, second row:δ0=0.3mm, third row: δ0=5µm).
Fig. 4.
Fig. 4. Second-fourth terms of Eq. (30) versus propagation distance, for four different values of source coherence width.
Fig. 5.
Fig. 5. Density plots for the modulus of the DOC of a GSM beam at some propagation distances for three different values of the source coherence width (First row:δ0=5mm, second row:δ0=0.3mm, third row: δ0=5µm).
Fig. 6.
Fig. 6. First-third terms of Eq. (31) versus propagation distance, for four different values of the source coherence width.
Fig. 7.
Fig. 7. Coherence widthin x and y direction versus propagation distance for four different values of source coherence width.

Equations (31)

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

W ( 0 ) ( r 1 , r 2 , ω ) = exp ( r 1 2 + r 2 2 4 σ 0 2 ) exp [ ( r 2 r 1 ) 2 2 δ 0 2 ] ,
W ( ρ 1 , ρ 2 , z ) = 1 λ 2 z 2 W ( 0 ) ( r 1 , r 2 ) exp [ i k 2 z ( ρ 1 r 1 ) 2 + i k 2 z ( ρ 2 r 2 ) 2 ] × exp [ ψ ( r 1 , ρ 1 , z ) + ψ ( r 2 , ρ 2 , z ) ] d 2 r 1 d 2 r 2 ,
exp [ ψ ( r 1 , ρ 1 , z ) + ψ ( r 2 , ρ 2 , z ) ] = exp { 2 π k 2 z 0 1 d t d 2 κ Φ n ( κ ) [ 1 exp ( t ρ d κ + ( 1 t ) r d κ ) ] } ,
Φ n ( κ ) = 0.033 C n 2 { ( L 0 x L 0 y ) 11 / 6 [ 1 + ( κ x L 0 x ) 2 + ( κ y L 0 y ) 2 ] 11 / 6 exp ( κ x 2 κ m x 2 κ y 2 κ m y 2 ) + Q [ ( 2 π L s ) 2 + κ 2 ] 11 6 exp ( κ 2 / κ 2 κ m 2 κ m 2 ) } ,
c ( α ) = [ 2 π Γ ( 5 α / 2 ) A ( α ) 3 ] 1 α 5 ,
A ( α ) = 1 4 π 2 Γ ( α 1 ) cos ( α π 2 ) ,   3 < α < 4.
Φ n ( κ ) = Φ n 1 ( κ ) + Φ n 2 ( κ ) ,
Φ n 1 ( κ ) = 0.033 C n 2 ( L 0 x L 0 y ) 11 / 6 [ 1 + ( κ x L 0 x ) 2 + ( κ y L 0 y ) 2 ] 11 / 6 exp ( κ x 2 κ m x 2 κ y 2 κ m y 2 ) ,
Φ n 2 ( κ ) = 0.033 C n 2 Q [ ( 2 π L s ) 2 + κ 2 ] 11 6 exp ( κ 2 κ m 2 ) .
exp [ ψ ( r 1 , ρ 1 , z ) + ψ ( r 2 , ρ 2 , z ) ] = A + B ,
A = exp { 2 π k 2 z 0 1 d t d 2 κ Φ n 1 ( κ ) [ 1 exp ( t ρ d κ + ( 1 t ) r d κ ) ] } ,
B = exp { 2 π k 2 z 0 1 d t d 2 κ Φ n 2 ( κ ) [ 1 exp ( t ρ d κ + ( 1 t ) r d κ ) ] } .
A = exp { 2 π k 2 z μ x μ y 0 1 d t d κ x d κ y Φ n 1 ( κ ) [ 1 exp ( t ρ d κ + ( 1 t ) r d κ ) ] } ,
Φ n 1 ( κ ) = 0.033 C n 2 ( μ x μ y L 0 2 1 + κ 2 L 0 2 ) 11 6 exp ( l 0 2 c ( α ) 2 κ 2 ) .
A = exp { 4 π 2 k 2 z μ x μ y 0 1 d t 0 κ d κ Φ n 1 ( κ ) [ 1 J 0 ( κ | t ρ d + ( 1 t ) r d | ) ] } .
A = exp { π 2 k 2 z T A ( ξ d 2 + ξ d x d + x d 2 ) 3 μ x 2 } exp { π 2 k 2 z T A ( η d 2 + η d y d + y d 2 ) 3 μ y 2 } ,
T A = 1 μ x μ y 0 κ 3 Φ n 1 ( κ ) d κ = 0 .033 C n 2 ( μ x μ y ) 5 6 [ ( 5 + 6 κ m 2 L 0 2 ) exp ( 1 κ m 2 L 0 2 ) Γ ( 1 6 , 1 κ m 2 L 0 2 ) 10 ( 1 κ m 2 ) 1 6 3 5 ( 1 L 0 2 ) 1 6 ] .
W A ( ρ 1 , ρ 2 , z ) = W x A ( ξ 1 , ξ 2 ) W y A ( η 1 , η 2 ) ,
W x A ( ξ 1 , ξ 2 ) = 1 Δ x A ( z ) exp ( ξ 1 2 + ξ 2 2 4 σ 0 2 Δ x A ( z ) ) exp [ i k 2 R x A ( z ) ( ξ 1 2 ξ 2 2 ) ] × exp { [ 1 2 δ 0 2 Δ x A ( z ) + π 2 k 2 T A z 3 μ x 2 ( 1 + 2 Δ x A ( z ) ) π 4 k 2 T A 2 z 4 18 μ x 4 Δ x A ( z ) σ 0 2 ] ( ξ 1 ξ 2 ) 2 } ,
Δ x A ( z ) = 1 + [ 1 4 k 2 σ 0 4 + 1 k 2 σ 0 2 ( 1 δ 0 2 + 2 π 2 k 2 T A z 3 μ x 2 ) ] z 2 ,
R x A ( z ) = z + σ 0 2 z π 2 T A z 4 / 3 μ x 2 ( Δ x A ( z ) 1 ) σ 0 2 + π 2 T A z 3 / 3 μ x 2 .
B = exp { 4 π 2 k 2 z 0 1 d t κ d κ Φ n 2 ( κ ) [ 1 J 0 ( κ | t ρ d + ( 1 t ) r d | ) ] } .
B = exp [ 1 3 π 2 k 2 z T B ( ξ d 2 + ξ d x d + x d 2 ) ] exp [ 1 3 π 2 k 2 z T B ( η d 2 + η d y d + y d 2 ) ] ,
T B = 0 κ 3 Φ n 2 ( κ ) d κ = 0 .033 C n 2 Q [ ( 5 + 24 π 2 κ m 2 L s 2 ) exp ( 4 π 2 κ m 2 L s 2 ) Γ ( 1 6 , 4 π 2 κ m 2 L s 2 ) 10 ( 1 κ m 2 ) 1 6 3 5 ( 4 π 2 L s 2 ) 1 6 ] .
W B ( ρ 1 , ρ 2 , z ) = 1 Δ B ( z ) exp ( ρ 1 2 + ρ 2 2 4 σ 0 2 Δ B ( z ) ) exp [ i k 2 R B ( z ) ( ρ 1 2 ρ 2 2 ) ] × exp { [ 1 2 δ 0 2 Δ B ( z ) + π 2 k 2 T B z 3 ( 1 + 2 Δ B ( z ) ) π 4 k 2 T B 2 z 4 18 Δ B ( z ) σ 0 2 ] ( ρ 1 ρ 2 ) 2 } ,
Δ B ( z ) = 1 + [ 1 4 k 2 σ 0 4 + 1 k 2 σ 0 2 ( 1 δ 0 2 + 2 π 2 k 2 T B z 3 ) ] z 2
R B ( z ) = z + σ 0 2 z π 2 T B z 4 / 3 ( Δ B ( z ) 1 ) σ 0 2 + π 2 T B z 3 / 3 .
S ( ρ , ω ) = W ( ρ , ρ , z ) = W A ( ρ , ρ , z ) + W B ( ρ , ρ , z ) ,
μ ( ρ 1 , ρ 2 , z ) = W ( ρ 1 , ρ 2 , z ) S ( ρ 1 , z ) S ( ρ 2 , z ) = W A ( ρ 1 , ρ 2 , z ) + W B ( ρ 1 , ρ 2 , z ) [ W A ( ρ 1 , ρ 1 , z ) + W B ( ρ 1 , ρ 1 , z ) ] [ W A ( ρ 2 , ρ 2 , z ) + W B ( ρ 2 , ρ 2 , z ) ] .
σ i ( z ) = σ 0 Δ i A ( z ) = [ σ 0 2 + z 2 4 k 2 σ 0 2 + z 2 k 2 δ 0 2 + 2 π 2 T A z 3 3 μ i 2 ] 1 2 ,   ( i = x , y ) .
δ i ( z ) = [ 1 δ 0 2 Δ i A ( z ) + 2 π 2 k 2 T A z 3 μ i 2 + 2 π 2 k 2 T A z 3 μ i 2 Δ i A ( z ) ( 2 π 2 T A z 3 6 μ i 2 σ 0 2 ) ] 1 2 , ( i = x , y ) .
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.