You are on page 1of 11

> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) < 1

Effect of the zenith angle on the average coherence function in the anisotropic non-
Kolmogorov atmospheric turbulence
E. Dakar1, E. Golbraikh2,*, N. Kopeika1, AND A. Zilberman3
1
Department of Electro-optical Engineering, Ben-Gurion University of the Negev, POB 653, Beer-Sheva
84105, Israel.
2
Physics Department, Ben-Gurion University of the Negev, POB 653, Beer-Sheva 84105, Israel.
3
Electrical and Computer Engineering Department, Ben-Gurion University of the Negev, POB
653, Beer-Sheva 653, Beer-Sheva 84105, Israel.
*
Corresponding author: golbref@bgu.ac.il

Abstract
.

1. Introduction

Atmospheric turbulence has a significant degrading impact on the quality of imaging system
due to the random fluctuations of atmospheric refractive-index. This effect can be taken into
account by calculating the modulation transfer function (MTF) of the optical path in the
turbulent atmosphere. Once finished the calculation, one should insert this MTF into the
product of MTFs which gives the imaging system's total MTF.
At the very beginning, researches were focused on the Kolmogorov turbulence, and derived
analytic expressions for the turbulence MTF by using its statistical properties [1] [2]. With the
development of experimental equipment and theoretical investigations, the atmospheric
turbulence has been proved to deviate from prediction of the Kolmogorov model [3] [4] [5] [6]
[7]. In these cases, the non-Kolmogorov effect is associated with a 3D spectrum having an
exponent different from −11/3 (which is the exponent's value for Kolmogorov model). A
series of non-Kolmogorov atmospheric refractive-index fluctuations spectral models,
including the non-Kolmogorov spectrum [8], the generalized exponential spectrum [9] and the
generalized modified atmospheric spectrum [10], have been adopted to derive the theoretical
expressions for the non-Kolmogorov turbulence MTF [9] [11] [12]. In these investigations, the
> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) < 2

isotropic atmospheric turbulence assumption was adopted, so the sizes of turbulence eddies
were assumed to be the same in both vertical and horizontal directions.
However, laboratory and theoretical results have shown that the atmospheric turbulence can
also be anisotropic [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] in the free
atmospheric above the boundary layer. The horizontal size of eddies at these altitudes is
typically from tens of meters to kilometers across, while their vertical size is usually confined
to a few meters. In this case, the free atmosphere is highly anisotropic. Recently, many
researchers have focused on investigating the anisotropic non-Kolmogorov turbulence [24]
[25] [26] [27]. Toselli [24] used the spectrum introduced by Gurvich [28] and Kon [29] to
theoretically investigate the long term beam spread and scintillation index for Gaussian beam
under weak anisotropic non-Kolmogorov turbulence with horizontal path. Gudimetla [25] [26]
studied the log-amplitude correlation function for plane and spherical waves under weak
anisotropic non-Kolmogorov turbulence. For the case when turbulence strength continues to
increase beyond the weak turbulence regime, Andrews [27] developed mathematical models
for the Gaussian beam propagating through weak-to-strong anisotropic non-Kolmogorov
turbulence.
All of the aforementioned investigations regarding the MTF of the anisotropic turbulence are
based on taking the anisotropic structure function as

( ) (1)
2 2 α /2
~2 x + y 2
Dn ( ⃗r , α , ς ) =Cn +z
ς2

Where:
~2
Cn - is the generalized structure constant; α - is the power of the structure function which can
be different than the structure function exponent of the Kolmogorov turbulence; ς - is the
anisotropic parameter for the asymmetry of turbulence cells in both horizontal and vertical
directions. When ς equals one, the isotropic turbulence is shown. As the horizontal turbulence
outer cell is always bigger than the vertical turbulence cell, the value of ς is always bigger than
one. As ς increases, the anisotropic property exhibits more obviously.
However, in this work the MTF calculation is based on the structure function from article [30]
which has the form
~2 2 α (γ ) (2)
D n ( ⃗r , α , ς ) =Cn ⋅ ( √ ρ + z )
2

Where:
> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) < 3

α - is the exponent of the structure function which changes according to γ - the angle of a
direction in space to the horizon (either from above or below the horizon). Since the model in
article [30] is axisymmetric around the vertical axis, the structure function uses the cylindrical
coordinates ρ and z . Because of the fundamental difference in the basic assumptions of the two
models, this article must start its research at the root of the theory of the turbulent medium
MTF. The next section thus presents a mathematical development that start from an article
which is the basis for most of the ensuing papers in its field.

2. The average coherence function

The starting point for the results of this article is the paper by Hufnagel and Stanley [1]. That
paper defines the average mutual coherence factor in an expression for an optical transfer
function. Said optical transfer function is of a system composed of a region of turbulence (see
Figure 1). The paper uses the assumption of local stationarity (at least regarding time and the
coordinates lateral to the optical axis) of the random processes resulting from the turbulence,
in order to define the average coherence function M ( x , y ; z ) as
M ( x , y ; z )= ⟨ V ( x + x ' , y + y ' , z ' , t ) V ¿ ( x ' , y ' , z ' , t ) ⟩ (3)

Where:
V - is the complex scalar representation of the electromagnetic wave emanating from a point-

object; z ' – is the optical axis coordinate; t - is the time variable; ( x ' , y ' )- are the coordinates
lateral to the optical axis; ( x , y ) - are the correlation distance variables, also lateral to the
optical axis; The angle brackets indicate a time average of the enclosed quantity.
> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) < 4

Figure 1 - System Geometry

The average mutual coherence factor is thus the value of the average coherence function at
z=L.

After an involved development of eq.(3) in section 4, article [1] gives an expression for
M (x , y ; z )

⟨ { }⟩
(4)
z

M ( x , y ; z )= ⟨ V ( x + x , y + y , z ,t ) V ( x , y , z , t ) ⟩ = exp i Κ ∫ [ N ( ⃗p1 , z̀ ) −N ( ⃗p 2 , z̀ ) ] d z̀
' ' ¿ ' '

Where:
> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) < 5

Κ – is the optical wavelength; N ( ⃗r ,t ) ≡ [ n ( ⃗r , t ) −n ( ⃗r ,t ) ] /n ( ⃗r ,t ) , is the normalized fluctuating part of

the air index of refraction, and it is assumed n ( r⃗ , t ) =1; ⃗p1 ≡ ⃗p+ ⃗p' and ⃗p2 ≡ ⃗p' so the final result
depends only on ⃗p1−⃗p2.
By the central limit theorem, the quantity S, defined by
z
(5)
S ( ⃗p1 , ⃗p2 , z ) ≡∫ [ N ( ⃗p 1 , z̀ )−N ( ⃗p2 , z̀ ) ] d z̀
0

has a Gaussian distribution with mean zero. As a result

⟨ exp {i Κ ∙ S ( ⃗p , ⃗p , z ) } ⟩=exp {−1


1 2
2
Κ ∙ ⟨ S ( ⃗p , ⃗p , z ) ⟩ }
2 2
1 2
(6)

Where

⟨[ { } ]⟩
z 2 (7)
⟨ S ( ⃗p , ⃗p , z) ⟩= ∫ [ N ( ⃗p , z̀ )−N ( ⃗p , z̀) ] d z̀
2
1 2 1 2 =¿
0

⟨∫∫ ⟩
z z
d z̀ 1 d z̀ 2 [ N ( ⃗p 1 , z̀ 1 )−N ( ⃗p2 , z̀ 1 ) ] ∙ [ N ( ⃗p1 , z̀ 2) −N ( ⃗p2 , z̀ 2 ) ]
0 0

Taking the integration out of the ensemble averaging, and denoting N ( ⃗pa , z̀ b ) ≡ N ab one gets
z z
(8)
⟨ S ( ⃗p , ⃗p , z) ⟩=∫∫ d z̀ d z̀ ⟨ N
2
1 2 1 2 11 N 12−N 11 N 22−N 21 N 12+ N 21 N 22 ⟩
0 0

Using the Markov approximation (see [31], sect. 64), sets

⟨ N ab N cd ⟩=δ ( z̀ b− z̀ d ) ∙ A ( z̀ +2 z̀ , ⃗p −⃗p )
b d
a c
(9)

Where A ( z̀ b + z̀ d
2 ), ⃗p is the 2 D inverse Fourier transform of Φ n ( k z̀ =0 , k ⃗p ) , which by itself is the 3 D

Fourier transform of the exact covariance function ⟨ N ( ⃗r ) N ( ⃗r ' ) ⟩. Said covariance is assumed to
depend only on r⃗ −⃗r ' , which is the variable of integration in the covariance's Fourier transform.
Installing Markov approximation in eq.(8), gives

⟨ S 2 ⟩ =∫ A ( z̀ 1 , 0 ) − A ( z̀ 1 , ⃗p1−⃗p2 ) −¿ A ( z̀1 , ⃗p 1−⃗p2 ) + A ( z̀ 1 , 0 ) d z̀ 1=2 ∙∫ A ( z̀ 1 ,0 )− A ( z̀ 1 , ⃗p1 −⃗p 2 ) d z̀(10)


z z

1¿
0 0

3. The exponent of the average coherence function


> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) < 6

Since A ( z̀ +2 z̀ , ⃗p) is the 2 D inverse Fourier transform of Φ ( k =0 , k ), knowing Φ ( k , k ) makes


b d
n z̀ ⃗p n z̀ p

it possible to calculate (at least numerically) the value of


−1 2 2
Κ ∙ ⟨ S ⟩ through eq.(10). Said value is the exponent of the average coherence function as
2
eq.(6) shows.
Article [30] gives a model for Φ n ( k z̀ , k ⃗p ) in the atmosphere. The model assumes that the
structural function of atmospheric turbulence is anisotropic but with an axial symmetry around
the vertical axis z . The model takes the spectral index of the structural function α to be a
function of the altitude and the propagation angle γ (the supplementary angle to the local
zenith angle). In order to calculate Φ n ( k z̀ , k ⃗p ) from the structure function, article [30] divides the
range of propagation angles into three subranges π /6 ≥ γ ≥ 0 ; π /3≥ γ > π /6 ; π /2≥ γ > π /3, and uses
several approximations to perform the Fourier transform integration. The calculation in article
[30] thus gives 3 different functions for Φ n ( k z̀ , k ⃗p ) , each for a different subrange. In order to
avoid discontinuities, article [30] combines the tree functions into one function that
approximately equals each one of them in the respective subrange. This function has the form:

[ ] [ ]{ [ (❑
} ][ ]
(❑❑⃗ ❑()❑) (11)
❑❑` ) ❑()❑ ]
❑ ❑ ❑
(❑❑` ❑()❑ ) ()
( ❑❑⃗ ❑❑` )
[
❑ ❑
❑ ❑❑ ❑ ( ) ❑❑ ❑
( )


❑❑ ( ❑❑` ❑⃗❑) (❑❑ ❑ )
❑ () ()
⃗ ❑❑`
❑ ❑ ❑ ❑ ❑⃗❑ ❑ ❑

Where ❑()❑, ❑❑ are all parameters that one needs to set for each case. In this article
❑❑ ❑()❑ ❑❑ ❑()❑ ❑❑
Looking at eq.(11), one can see that for the purpose of integrating on the Fourier plane it is
possible to approximate the spectrum. This approximation reduces the spectrum into a single
term which includes only power functions. This approximation thus removes almost all the
parameters from the spectrum and leaves only the amplitude. The approximation's form is:
() ()
( ❑⃗❑❑❑` ) ❑❑` (12)
❑❑ ( ❑❑` ❑⃗❑)
❑⃗❑
> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) < 7

Henceforth, this article will denote the spectrum from eq.(11) as “continuous”, and the
−1 2 2
spectrum from eq.(12) as “compact”. Using the two spectra, one can calculate 2 Κ ∙ ⟨ S ⟩ for

various optical paths in the atmosphere. The optical paths this article calculates for are slant-
paths of either Uplink/Downlink channels. Since these calculations showed no difference
between the uplink and downlink channels, no distinction is needed. Each path is
characterized by the altitude of the receiver/transmitter and by its angle above/below the
horizon. These characteristics determine the total length of the path L. In order to calculate the
exponent one thus needs to insert z=L into eq.(10).
A critical element in the exponent calculation is to determine the value of the generalized
~
structure constant C2n as a function of the altitude. This article uses the H−V 5 /7 turbulence
model [5], [32] (as is article [30]) to calculate C 2n and then normalizes it with the formula


8
) sin ( )
π (13)

π √ L)
( Κ ( )
2
~ 2 2 3 3 α−
3
C =C
Γ ( α +2 ) sin ( α ∙ )
n n

The last element needed for the calculation of the exponent is the turbulence outer scale, L0
which is crucial in calculating the integrals involving Φ n ( k z̀ , k ⃗p ) . This article takes the formula
for L0 from [33], [34]
4 (14)
L0 ( h ) = [ m]
[ ]
2
h−8500
1+
2500

4. Graphical results

This section presents the results of the numerical calculations of the exponent using
MATLAB. Each figure presents results regarding a specific altitude of the receiver/transmitter
and each figure includes four graphs. Each graph gives the exponent as a function of the
propagation angle γ , and each graph is different from the other three by the values of the
correlation distance variables. These variables exist in the plane which is perpendicular to the
optical path. If one observes the plane which includes both the optical path and radius from the
center of the earth to any of the points along the path, then the variable Δ y represents the
> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) < 8

correlation distance along the intersection of the two planes. Δ x represents the correlation
distance along the direction perpendicular to Δ y .

The figures are thus:

Figure 2 – The exponent of the average coherence function vs. the propagation angle (Altitude of receiver/transmitter= 5 [ km ]).

Black graphs – The continuous spectrum; Gray graphs – The compact spectrum.
> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) < 9

Figure 3 - The exponent of the average coherence function vs. the propagation angle (Altitude of receiver/transmitter= 10 [ km ]).

Black graphs – The continuous spectrum; Gray graphs – The compact spectrum.

Figure 4 - The exponent of the average coherence function vs. the propagation angle (Altitude of receiver/transmitter= 15 [ km ]).

Black graphs – The continuous spectrum; Gray graphs – The compact spectrum.
> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) < 10

References

[1] R. E. HUFNAGEL and N. R. STANLEY, "Modulation Transfer Function Associated with Image Transmission through
Turbulent Media," JOSA, vol. 54, no. 1, pp. 52-61, JANUARY 1964.
[2] D. L. Fried, "Optical Resolution Through a Randomly Inhomogeneous Medium for Very Long and Very Short Exposures,"
JOSA, vol. 56, no. 10, pp. 1372-1379, October 1966.
[3] M. S. Belen'kii, S. J. Karis, J. M. Brown II and R. Q. Fugate, "Experimental study of the effect of non-Kolmogorov
stratospheric turbulence on star image motion," in SPIE 3126, Adaptive Optics and Applications., San Diego, CA, United
States, 1997.
[4] M. S. Belen'kii, E. Cuellar, K. A. Hughes and V. A. Rye, "Experimental study of spatial structure of turbulence at Maui Space
Surveillance Site (MSSS)," in Society of Photo-Optical Instrumentation Engineers (SPIE) 6304, 2006.
[5] A. Zilberman, E. Golbraikh, N. S. Kopeika, A. Virtser, I. Kupershmidt and Y. Shtemler, "Lidar study of aerosol turbulence
characteristics in the troposphere: Kolmogorov and non-Kolmogorov turbulence," Atmospheric Research, vol. 88, no. 1, pp.
66-77, April 2008.
[6] A. S. Gurvich and M. S. Belen'kii, "Influence of stratospheric turbulence on infrared imaging," Journal of the Optical Society
of America A, vol. 12, no. 11, pp. 2517-2522, November 1995.
[7] M. S. Belen'kii, "Effect of the stratosphere on star image motion," Optics Letters, vol. 20, no. 12, pp. 1359-1361, 1995.
[8] B. E. Stribling, B. M. Welsh and M. C. Roggemann, "Optical propagation in non-Kolmogorov atmospheric turbulence," in
SPIE 2471, Atmospheric Propagation and Remote Sensing IV, Orlando, FL, United States, 1995.
[9] L.-y. Cui , B.-d. Xue, X.-g. Cao, J.-k. Dong and J.-n. Wang, "Generalized atmospheric turbulence MTF for wave propagating
through non-Kolmogorov turbulence," Optics Express, vol. 18, no. 20, pp. 21269-21283, 2010.
[10] B. Xue, L. Cui, W. Xue, X. Bai and F. Zhou, "Generalized modified atmospheric spectral model for optical wave propagating
through non-Kolmogorov turbulence," Journal of the Optical Society of America A, vol. 28, no. 5, pp. 912-916, 2011.
[11] N. S. Kopeika, A. Zilberman and E. Golbraikh, "Generalized atmospheric turbulence: implications regarding imaging and
communications," in Atmospheric and Oceanic Propagation of Electromagnetic Waves IV, San Francisco, California, United
States, 2010.
[12] B. Xue, L. Cao, L. Cui, X. Bai, X. Cao and F. Zhou, "Analysis of non-Kolmogorov weak turbulence effects on infrared
imaging by atmospheric turbulence MTF," Optics Communications, vol. 300, pp. 114-118, 2013.
[13] A. Consortini, L. Ronchi and L. Stefanutti, "Investigation of Atmospheric Turbulence by Narrow Laser Beams," Applied
Optics, vol. 9, no. 11, pp. 2543-2547, 1970.
[14] R. Manning, "An anisotropic turbulence model for wave propagation near the surface of the Earth," IEEE transactions on
antennas and propagation, vol. 34, no. 2, pp. 258-261, 1986.
[15] V. P. Lukin, "Investigation of the anisotropy of the atmospheric turbulence spectrum in the low-frequency range," in
Atmospheric Propagation and Remote Sensing IV, Orlando, FL, United States, 1995.
[16] L. V. Antoshkin, N. N. Botygina, O. N. Emaleev, L. N. Lavrinova, V. P. Lukin, A. P. Rostov, B. V. Fortes and A. P. Yankov,
"INVESTIGATION OF TURBULENCE SPECTRUM ANISOTROPY IN THE GROUND ATMOSPHERIC LAYER.
PRELIMINARY RESULTS," ATMOSPHERIC AND OCEANIC OPTICS C/C OF OPTIKA ATMOSFERY I OKEANA, vol. 8,
no. 12, pp. 993-996, December 1995.
[17] Belen’kii M. S., J. D. Barchers, S. J. Karis, C. L. Osmon, J. M. Brown, R. Q. Fugate, "Preliminary experimental evidence of
anisotropy of turbulence and the effect of non-Kolmogorov turbulence on wavefront tilt statistics," Adaptive Optics Systems
and Technology, no. 3762, p. 396–406, 1999.
[18] 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 ICO XVIII 18th Congress of the International Commission for
Optics, San Francisco, CA, United States, 1999.
[19] L. R. Tsvang, "Measurements of the spectrum of temperature fluctuations in the free atmosphere," Izvestiya Akademii Nauk
SSSR, Geofizicheskaya 1, pp. 1117-1120, 1960.
[20] G. D. Nastrom and K. S. Gage, "A Climatology of Atmospheric Wavenumber Spectra of Wind and Temperature Observed by
Commercial Aircraft," Journal of the Atmospheric Sciences, vol. 42, no. 9, p. 950–960, 01 May 1985.
[21] F. D. Eaton and G. D. Nastrom, "Preliminary estimates of the vertical profiles of inner and outer scales from White Sands
Missile Range, New Mexico, VHF radar observations," Radio Science, vol. 33, no. 4, pp. 895-903, July-August 1998.
[22] D. T. Kyrazis, J. B. Wissler, D. D. B. Keating, A. J. Preble and K. P. Bishop, "Measurement of optical turbulence in the upper
troposphere and lower stratosphere," in Laser Beam Propagation and Control, Los Angeles, CA, United States, 1994.
[23] A. D. Wheelon, Electromagnetic Scintillation: Volume 1, Geometrical Optics, Cambridge: Cambridge University Press, 2001.
[24] I. Toselli, B. Agrawal and S. Restaino, "Light propagation through anisotropic turbulence," Journal of the Optical Society of
America A, vol. 28, no. 3, pp. 483-488, 2011.
> REPLACE THIS LINE WITH YOUR PAPER IDENTIFICATION NUMBER (DOUBLE-CLICK HERE TO EDIT) < 11

[25] V. S. R. Gudimetla, R. B. Holmes and J. F. Riker, "Analytical expressions for the log-amplitude correlation function for plane
wave propagation in anisotropic non-Kolmogorov refractive turbulence," JOSA A, vol. 29, no. 12, pp. 2622-2627, 2012.
[26] V. S. R. Gudimetla, R. B. Holmes and J. F. Riker, "Analytical expressions for the log-amplitude correlation function for
spherical wave propagation through anisotropic non-Kolmogorov atmosphere," JOSA A, vol. 31, no. 1, pp. 148-154, 2014.
[27] L. C. Andrews, R. L. Phillips, R. Crabbs and T. Leclerc, "Deep turbulence propagation of a Gaussian-beam wave in
anisotropic non-Kolmogorov turbulence," in Laser Communication and Propagation through the Atmosphere and Oceans II,
San Diego, California, United States, 2013.
[28] A. S. Gurvich and A. I. Kon, "Aspect sensitivity of radar returns from anisotropic turbulent irregularities," Journal of
electromagnetic waves and applications, vol. 7, no. 10, pp. 1343-1353, 1993.
[29] Kon, A. I., "Qualitative theory of amplitude and phase fluctuations in a medium with anisotropic turbulent irregularities,"
Waves in Random, no. 4, p. 297–306, 1994.
[30] E. Dakar, E. Golbraikh, N. S. Kopeika and A. Zilberman, "Effect of the Zenith Angle on Optical Wave Propagation in
Anisotropic Non-Kolmogorov Atmospheric Turbulence: A New Experiment-Based Model," IEEE TRANSACTIONS ON
ANTENNAS AND PROPAGATION, vol. 68, no. 8, pp. 6287 - 6295, August 2020.
[31] V. Tatarski, The Effect of rhe Turbulent Atmosphere on Wave Propagation, Springfield, Va.: U.S. Dep. Commerce, 1971.
[32] A. Zilberman, E. Golbraikh and N. S. Kopeika, "Propagation of electromagnetic waves in Kolmogorov and non-Kolmogorov
atmospheric turbulence: three-layer altitude model," Apply Opt., vol. 47, no. 34, p. 6385, 2008.
[33] O. G. Chkhetiani, A. Eidelman and E. Golbraikh, "Large- and small-scale turbulent spectra in MHD and atmospheric,"
Nonlinear Processes Geophys., vol. 13, no. 6, p. 613–620, 2006.
[34] V. P. Lukin, E. V. Nosov and B. S. Fortes, "The Efficient Outer Scale of Atmospheric Turbulence," in ESO Conference and
Workshop Proceedings, 1999.

You might also like