Vous êtes sur la page 1sur 8

Turbulence Mitigation by Aperture Averaging in

Wireless Optical Systems


M.A. Khalighi

, N. Aitamer

, N. Schwartz

, S. Bourennane

Ecole Centrale Marseille, Institut Fresnel, Marseille, France

DOTA, ONERA, Ch atillon, France.


Ali.Khalighi@fresnel.fr, Naziha.Aitamer@ec-marseille.fr, Noah.Schwartz@onera.fr, Salah.Bourennane@fresnel.fr
AbstractAtmospheric turbulence can cause a signicant
performance degradation in free space optical (FSO) com-
munication systems. We investigate the impact of aperture
averaging on the performance of FSO systems under differ-
ent atmospheric turbulence regimes. Performance evaluation
is made in terms of the average bit-error-rate. In particular,
we bring clearance on the impact of propagation model on
the benet of aperture averaging. The efciency of channel
coding for aperture averaging receivers is also discussed.
I. INTRODUCTION
Free space optical (FSO) communication has attracted
special attention since a few years as it can provide huge
data transmission rates [1]. In practice, considering clear
sky conditions, atmospheric turbulence causes random
uctuations of the phase and the amplitude of the received
signal, or in other words, channel fading. These intensity
uctuations, called scintillation, can result in a consid-
erable degradation of the system performance. Fading is
more considerable in long-distance transmissions as well
as in the case of communication with a moving platform.
A comprehensive survey of optical-propagation effects
can be found in [2].
It is well known that to mitigate efciently channel fading
one should make use of diversity techniques. We have
considered in a previous work [3], [4] the use of time di-
versity and shown that substantial performance improve-
ment can be obtained by performing channel coding and
interleaving. This solution, however, often imposes long
delay latencies and necessitates the use of large memories
for storing long data frames. Another approach is to
employ spatial diversity for fading reduction. Aperture
averaging can be seen as a simple form of spatial diversity
when the receiver lens aperture is larger than the fading
correlation length [2], [5]. The use of aperture averaging
for reducing the effect of scintillation has been widely
considered in the literature [2], [5], [6], [7], [8], [9], [10],
[11], [12], [13]. It is shown that substantial scintillation
reduction can be obtained, especially in the cases of strong
turbulence.
Here, we consider the case where we work with a
monochromatic laser and a single beam at the transmitter.
Our aim is to study the impact of aperture averaging on
the system performance by considering as criterion the
system average bit-error-rate (BER), which is among the
important parameters in a practical FSO system.
Firstly, we consider the aperture averaging receivers
and evaluate the performance gain for different propa-
gation models by presenting some numerical results. In
particular, we contrast the fading reduction for the cases
of plane and spherical wave propagation and Gaussian-
beam wave models, and also investigate the impact of
the inner scale of turbulence. The efciency of channel
coding in reducing the system BER is also studied.
The remainder of the paper is organized as follows.
In Section II, we present the main assumptions we make
concerning signal transmission and detection. In Section
III, we provide some generalities on turbulence modeling.
The modication of the turbulence model parameters
when using aperture averaging are recalled in Section
IV. The performance improvement by aperture averaging
under different propagation models is studied in Section
V. The impact of channel coding used in combination
with aperture averaging is discussed in Section VI.
II. SYSTEM MODEL AND GENERAL ASSUMPTIONS
For the sake of implementation simplicity, we consider
intensity modulation with direct detection (IM/DD), as it
is used in most current FSO systems. At the transmitter,
binary data are converted to impulses of duration T
s
according to the non return-to-zero (NRZ) on-off keying
(OOK) modulation. At the receiver, the incident beam is
collected on a lens of diameter D, before being focused on
a photo-detector which converts it to an electrical signal.
Let the received signal be:
r = I +n, (1)
where is the optical/electrical conversion efciency, and
n is the sum of thermal, dark, background, and shot
noise. We consider background noise limited receivers
and model n as a Gaussian white stationary random
process. Without loss of generality, we set = 1. After
optical/electrical conversion and sampling, we perform
signal demodulation based on the maximum a posteriori
(MAP) criterion, assuming perfect channel knowledge.
Note that, channel can be estimated with enough accuracy
using only few training symbols [4]. In the results to be
presented, we will not consider any channel coding, unless
otherwise mentioned. We consider the quasi-static channel
model by which the channel fading coefcients remain
constant during the transmission of a frame of symbols.
To compare the receiver performance for different lens
diameters D, we x the noise variance
2
n
and also
consider normalized average received intensity, i.e., we
set E{I} = 1. This, in fact, represents a background noise
10th International Conference on Telecommunications - ConTEL 2009
ISBN: 978-953-184-131-3, June 8-10, 2009, Zagreb, Croatia 59
limited xed eld-of-view (FOV) receiver. In such a case,
by increasing the pupil area by a factor m, the received
signal and noise powers increase by the same factor [14],
and hence, the SNR does not change.
Finally, we assume that the photo-detector area is suf-
ciently large to permit to benet fully from aperture
averaging, irrespective of the turbulence strength.
III. TURBULENCE MODELING
Atmospheric turbulence is characterized by three pa-
rameters: the inner scale l
0
, the outer scale L
0
, and the
index of the refraction structure parameter C
2
n
in units
of m
2/3
, sometimes called the turbulence strength. For
intensity uctuations, we adopt the Gamma-Gamma ()
channel model that can describe any type of turbulence
[2]. Let us denote by I
x
and I
y
the large- and small-
scale irradiance uctuations, respectively. I
x
and I
y
are
assumed to be statistically independent and described by
the Gamma distribution. Based on these assumptions,
the probability density function of the received intensity
I = I
x
I
y
is [2]:
p(I) =
2 ()
(+)/2
() ()
I
(+)/21
K

_
2(I)
1/2

, I > 0
(2)
where and are effective numbers of large- and small-
scale eddies of the scattering environment, respectively,
and are directly related to the atmospheric conditions.
Also, K

(x) is the modied Bessel function of the


second kind and order . Expressions for calculating the
parameters and under different propagation models
can be found in [2]. If we denote by
2
ln x
and
2
ln y
the variances of large- and small-scale log-irradiance,
respectively, the scintillation index is given by:

2
I
= exp(
2
ln x
+
2
ln y
) 1 (3)
Generally, the distinction between the two cases of weak
and strong turbulence are made by considering the Rytov
variance
2
R
:

2
R
= 1.23 C
2
n
k
7/6
L
11/6
, (4)
where k = 2/ is the wave number with the
wavelength, and L is the link distance. In this way,
weak, moderate, and strong turbulence conditions are
characterized by
2
R
< 1,
2
R
1, and
2
R
1,
respectively.
Throughout the paper, we consider the conditions of
homogeneous turbulence, that is, we assume that C
2
n
does
not depend on the distance. Discussion on the case of non-
homogeneous turbulence can be found in [6].
Let us recall some useful denitions that we will use later
in the paper:
Fresnel zone: F =
_
L/k
Coherence radius:
0
= (1.46 C
2
n
k
2
L)
3/5
,
l
0

0
L
0
, (expression valid for plane wave
propagation [2].)
Scattering disk: L/k
0
A. Parameters considered for numerical results
Later in this paper we present numerical results to study
the receiver performance. We specify here the correspond-
ing parameters that we consider in our simulations.
We consider the wavelength = 1550 nm. Three
cases of strong, moderate, and weak turbulence will be
considered, for which we set L = 1500 m and C
2
n
=
4.58 10
13
, L = 500 m and C
2
n
= 4.58 10
13
, and
L = 500 m and C
2
n
= 7 10
16
, respectively. Using
(4), the Rytov variance corresponding to these three cases
is
2
R
= 19.18,
2
R
= 2.56, and
2
R
= 0.004, respec-
tively. Using the results of the experiments conducted by
Andrews et al. [5], the inner scale is l
0
= 4.6 mm for
C
2
n
= 4.58 10
13
and = 1550 nm. The outer scale
L
0
has a negligible impact on the scintillation index in
practice [2]; that is why we assume L
0
in this paper.
For the simulations, we will present the receiver perfor-
mance in terms of average BER as a function of SNR.
We consider the electrical SNR in the form of E
b
/N
0
,
where E
b
is the average received energy per information
bit, and N
0
is the unilateral noise power spectral density.
For an uncoded system, E
b
/N
0
equals the actual SNR.
At the receiver, the electrical signal power is considered
as E{I
2
}. For temporal variation of the turbulence, we
use the quasi-static (frozen) channel model. Each frame
corresponds to 2000 OOK symbols. Note that, since a
quasi-static channel is considered, the frame length has
almost no impact on the results to be presented.
IV. APERTURE AVERAGING
As explained in [2], [9], by aperture averaging, we
average over the relatively fast uctuations caused by the
small-size eddies, and thus, the frequency content of the
irradiance spectrum is shifted to the low spatial frequen-
cies. The remained scintillation is due to the eddies whose
size is larger than the pupil size. The parameter that is
usually used to quantify the fading reduction by aperture
averaging, is the aperture averaging factor, dened as
follows [2].
A =

2
I
(D)

2
I
(0)
(5)
Here,
2
I
(D) and
2
I
(0) denote the scintillation index for
a receiver lens of diameter D and a point receiver
(D 0), respectively. We briey review in the following
the existing theory for different turbulence conditions.
Conventionally, the Kolmogorov power-law spectrum is
used for the refraction index:

n
() = 0.033 C
2
n

11/3
, 1/L
0
1/l
0
, (6)
where is the spatial frequency. This model assumes
that l
0

_
L/k. It is well known that, under moder-
ate to strong turbulence conditions, only eddies of size
smaller than
0
or larger than the scattering disk L/k
0
contribute effectively to the atmospheric turbulence. To
take into account this dependence of the turbulence to the
eddies sizes, it is proposed in [2] to consider an effective
atmosphere spectrum that includes a spatial ltering in
the expression of
n,e
():

n,e
() = 0.033 C
2
n

11/3
[G
x
(, l
0
, L
0
) +G
y
(, l
0
)] .
(7)
Here, G
x
(, l
0
, L
0
) and G
y
(, l
0
) denote respectively the
low pass and high pass spatial lters transfer functions
M.A. Khalighi, N. Aitamer, N. Schwartz, S. Bourennane
ConTEL 2009, ISBN: 978-953-184-131-3 60
that affect large- and small-scale eddies, respectively. (See
[2] for the expressions of these transfer functions.)
The impact of scintillation depends on the propagation
conditions. Here, we just recall the expressions for the
case of plane wave propagation, L
0
, and zero inner-
scale l
0
0 [2]. The expressions of the variances of large-
and small-scale log-irradiance are given by (8) and (9),
respectively:

2
ln x,PL
(D)

=
0.16
2
R

7/6
x
(1 +d
2

x
/4)
7/6
,
2
R
> 1 (8)

2
ln y,PL
(D)

=
1.272
2
R

5/6
y
1 + 0.3 d
2

y
,
2
R
> 1 (9)
where d =
_
kD
2
/4L is the circular aperture radius
scaled by the Fresnel zone, and

x
=
2.61
1 + 1.11
12/5
R
,
y
= 3(1 + 0.69
12/5
R
)
(10)
We used the subscript .
PL
to denote explicitly the assump-
tion of plane wave propagation. From (3), we obtain the
scintillation index as follows.

2
I,PL
(D) = exp
_
0.49
2
R
(1 + 0.653 d
2
+ 1.11
12/5
R
)
7/6
+
0.51
2
R
(1 + 0.69
12/5
R
)
5/6
1 + 0.9 d
2
+ 0.621 d
2

12/5
R
_
1
(11)
Under weak uctuation regime, the aperture averaging
factor A can be approximated as [2]:
A
_
1 + 1.062
_
D
2
4
k
L
__
7/6
(12)
Note that, this expression gives a better interpolation of
the analytical data than the expression given in [6]. Note
also that, (12) assumes that l
0

_
L/k ; however, it can
practically be used for l
0
< 2.73
_
L/k [6].
V. APERTURE AVERAGING EFFECT UNDER DIFFERENT
TURBULENCE MODELS
A. Propagation models
As mentioned above, the effect of aperture averaging
depends on the propagation conditions. In the literature,
usually the plane or spherical wave propagation models
are considered. These models hold with the approximation
of a point optical source. Sometimes, the analyses are
done assuming a collimated Gaussian-beam wave, where
the power prole of the beam is close to a Gaussian.
For optical links through space, the plane wave propa-
gation model is mostly appropriate for space-to-ground
transmissions, whereas the spherical propagation model
is suitable for ground-to-space transmission links. For
horizontal (terrestrial) transmissions, the Gaussian beam
model is a good approximation, however [15]. It should be
noted that, when the Gaussian beam has a relatively large
divergence, its statistical properties are close to the case
of a point source [8]. For such a beam, we can effectively
use the approximations of plane or spherical waves.
10
2
10
1
10
0
10
1
10
2
10
3
10
6
10
5
10
4
10
3
10
2
10
1
10
0
d = (kD
2
/4L)
1/2
A

=

I 2
(
D
)
/

I 2
(
0
)

R
2
=19.18
l
0
=4.6mm

R
2
=19.18
l
0
=0

R
2
=2.56
l
0
=4.6mm

R
2
=2.56
l
0
=0

R
2
=0.004
Figure 1. Aperture averaging factor, plane wave propagation.
The expressions of
2
ln x
(D) and
2
ln y
(D) for a receiver
lens of diameter D are given in [2] for the cases of
spherical and Gaussian-beam waves (respectively in Pages
416 and 420 for l
0
0); we do not recall them here due
to space limitation.
B. Inner and outer scales of turbulence
As stated previously, we assume in this paper that L
0

as it has a negligible impact on the scintillation index
in practice [2]. In the literature, the inner scale is often
neglected too, that is, it is assumed that l
0
0. However,
the effect of the inner scale could be considerable on the
turbulence and namely on
2
I
. Also, the effect of aperture
averaging is more important when considering the inner
scale [2]. The expressions of
2
ln x
(D) and
2
ln y
(D) can
again be found in [2] for l
0
= 0 (Pages 412, 415, and
419, for plane, spherical, and Gaussian beam models,
respectively).
C. Numerical results
We present here some simulation results to study the
effect of aperture averaging under different propagation
conditions.
1) Aperture averaging factor: The aperture averaging
factor A has been already studied in numerous works. We
prefer to recall briey the general points by considering
the three special cases of weak, moderate, and strong
turbulence, specied in Subsection III-A. These results are
useful when discussing other simulation results presented
later in this section.
Let us rst consider the classical case of plane wave
propagation model while neglecting the inner scale, i.e.
assuming l
0
0. As a reminder, we have presented
the curves of the aperture averaging factor A versus the
normalized receiver lens radius d in Fig. 1 for the three
cases of turbulence. Note that l
0
has negligible effect in
the weak turbulence regime. We see from Fig. 1 that for
weak to moderate turbulence regimes, we have almost the
same behavior: A begins to decrease effectively for D >

L, where

L is in fact the size of scintillation spots
Turbulence Mitigation by Aperture Averaging in Wireless Optical Systems
ConTEL 2009, ISBN: 978-953-184-131-3 61
at the receiver in these cases. When the lens diameter D
is larger than

L, it can effectively average over the
turbulence.
For the case of strong turbulence, we have the famous
leveling effect predicted for
0
< D < L/k
0
; A
continues to decrease for D > L/k
0
[9]. From Fig. 1,
the leveling effect can be attributed to about 1 < d < 5,
or equivalently to 38 mm< D <150 mm. Remembering
the parameters from Subsection III-A, for this case, we
have
0
= 3 mm and L/k
0
= 123 mm.
We have also shown in Fig. 1 curves of A versus d
for the case where we do not neglect the inner scale
l
0
. We notice that the leveling effect almost disappears
for l
0
= 0 and A decreases steadily by increasing D.
This can be explained by the fact that larger inner scales
result in larger values of irradiance variance for the strong
turbulence regime [16]. This is also seen in the case
of moderate turbulence but the difference is much less
considerable for l
0
= 0 and l
0
= 0.
To see the effect of aperture averaging for different
propagation models, we have presented the curves of A
versus d for the cases of moderate and strong turbulence,
in Fig. 2. Concerning the Gaussian-beam wave model,
we consider the parameters according to the experiment
presented in [5]: the beam radius is W
0
= 1.59 cm and
the curvature radius of the phase front is F
0
= 69.9 m,
corresponding to a beam divergence of
div
= 0.46 mrad.
For the moderate turbulence case, we have almost the
same trend in A by increasing d for the three models.
For the strong turbulence regime, however, more aperture
averaging gain is predicted for the Gaussian-beam model
for relatively large d. Also, for this case, the leveling
effect is less considerable for the spherical wave model
than for the plane wave model. It does not occur for the
Gaussian-beam model.
2) Average BER performance: Let us now consider the
average BER performance of the receiver. We begin with
the strong turbulence regime. BER curves for the case of
plane wave propagation are shown in Fig. 3 for l
0
= 0
and l
0
= 4.6 mm and different lens diameters D. Pt Rx
denotes the case of a point receiver (D = 0). Note that
due to long Monte Carlo simulations, we have limited the
results to a BER of about 10
6
. (For a given SNR, we
generate as many frames as necessary to obtain at least
1000 frame errors and 5000 bit errors.)
Let us rst consider the case of l
0
= 0. Compared to the
case of a point receiver (absence of aperture averaging),
we notice a considerable performance improvement even
for small receiver lens diameters. For instance, for a BER
of 10
5
, we obtain a gain of 53 dB in SNR by using
a 10 mm lens. Remember that for this case, we have

0
= 3 mm and L/k
0
= 123 mm. So, for D >
0
we
benet considerably from aperture averaging. From Fig.
1, a leveling of A is predicted for 38 mm< D <150 mm.
This is conrmed by the BER curves: We have a small
improvement by increasing D from 50 mm to 100 mm,
but for D = 200 mm, we have again an effective decrease
in BER.
10
2
10
1
10
0
10
1
10
2
10
3
10
4
10
3
10
2
10
1
10
0
d = (kD
2
/4L)
1/2
A

=

I 2
(
D
)
/

I 2
(
0
)

R
2
= 2.56
Plane
Wave
Spherical
Wave
Gaussian
Beam
(a) Moderate turbulence regime
10
2
10
1
10
0
10
1
10
2
10
3
10
5
10
4
10
3
10
2
10
1
10
0
d = (kD
2
/4L)
1/2
A

=

I 2
(
D
)
/

I 2
(
0
)


Gaussian
Beam
Plane
Wave
Spherical
Wave

R
2
= 19.18
(b) Strong turbulence regime
Figure 2. Aperture averaging factor, plane and spherical and Gaussian-
beam propagation models, l
0
= 0.
For the case of l
0
= 4.6 mm (dashed-line curves in
Fig. 3), we have a considerable decrease in BER even
when increasing D from 100 mm to 200 mm. This again
conrms the results of Fig. 1. Note that, although for
relatively large D (D > 100 mm), the performance
for l
0
= 4.6 mm approaches that for l
0
= 0, there
still remains a considerable difference between the corre-
sponding BER curves. In fact, as indicated by Consortini
[16], the turbulence has a more destructive effect for
l
0
= 0 and a poorer performance is obtained, as compared
to the case of l
0
= 0. We see that this is true even with
D as large as 200 mm.
Let us now consider the case of moderate turbulence.
Results are shown in Fig. 4. Remembering the parameters
from Subsection III-A, for this case, we have
0
=
M.A. Khalighi, N. Aitamer, N. Schwartz, S. Bourennane
ConTEL 2009, ISBN: 978-953-184-131-3 62
0 10 20 30 40 50 60 70 80 90 100
10
6
10
5
10
4
10
3
10
2
10
1
10
0
E
b
/N
0
(dB)
B
E
R


l
0
=0, Pt Rx
l
0
=4.6mm, Pt Rx
l
0
=0, D=5mm
l
0
=4.6mm, D=5mm
l
0
=0, D=10mm
l
0
=4.6mm, D=10mm
l
0
=0, D=20mm
l
0
=4.6mm, D=20mm
l
0
=0, D=50mm
l
0
=4.6mm, D=50mm
l
0
=0, D=100mm
l
0
=4.6mm, D=100mm
l
0
=0, D=200mm
l
0
=4.6mm, D=200mm
Figure 3. BER performance in the strong turbulence regime with
2
R
=
19.18, plane wave propagation model is considered. Pt Rx denotes
the point receiver case (D = 0).
5.8 mm and L/k
0
= 21 mm. We notice that for D >
50 mm, the BER curves for l
0
= 0 are very close to the
those for l
0
= 0. The fact that aperture averaging reduces
the effect of l
0
is hence conrmed clearly.
Finally, Fig. 5 shows the BER curves for the weak
turbulence regime. No surprise, the gain obtained by
aperture averaging is far less impressing than for moderate
or strong turbulence regimes. For instance, for a BER of
10
5
, compared to the point receiver, we obtain a gain of
0.33 dB in SNR only, by using a 50 mm lens.
3) Propagation model: Let us now consider the im-
pact of the propagation model. As mentioned previously,
usually the plane or spherical wave propagation models
are considered in the literature. In practice, however,
laser beams used in FSO systems usually have a prole
of near-Gaussian shape. To see the impact of aperture
averaging on the average BER for different propagation
models, we have shown in Fig. 6 the performance curves
for the case of strong turbulence and assuming l
0
= 0.
For the Gaussian-beam model, we set W
0
= 1.59 cm,
F
0
= 69.9 m, and L = 1500 m. As it could be predicted
from Fig. 2, compared to plane and spherical wave
models, more signicant fading reduction is obtained for
the Gaussian-beam model for relatively large apertures
(D 100 mm). For D < 50 mm, the performance
improvement by aperture averaging is far less than that for
the plane wave model. We notice that for D 100 mm
as well as for the point receiver case, the BER curves of
the Gaussian-beam model are close to those of the plane
wave model. In such cases, we may effectively adopt the
approximation of plane wave propagation for a Gaussian
beam.
0 10 20 30 40 50 60 70 80 90 100
10
6
10
5
10
4
10
3
10
2
10
1
10
0
E
b
/N
0
(dB)
B
E
R


l
0
=0, Pt Rx
l
0
=4.6mm, Pt Rx
l
0
=0, D=5mm
l
0
=4.6mm, D=5mm
l
0
=0, D=10mm
l
0
=4.6mm, D=10mm
l
0
=0, D=20mm
l
0
=4.6mm, D=20mm
l
0
=0, D=50mm
l
0
=4.6mm, D=50mm
l
0
=0, D=100mm
l
0
=4.6mm, D=100mm
l
0
=0, D=200mm
l
0
=4.6mm, D=200mm
Figure 4. BER performance in the moderate turbulence regime with

2
R
= 2.56, plane wave propagation model is considered.
5 6 7 8 9 10 11 12 13 14
10
6
10
5
10
4
10
3
10
2
10
1
E
b
/N
0
(dB)
B
E
R
Pt Rx
D = 10 mm
D = 50 mm
D = 100 mm
Figure 5. BER performance in the weak turbulence regime with
2
R
=
0.004, plane wave propagation model
D. Fade statistics
Obviously, when employing aperture averaging at the
receiver, the fade statistics are modied. Since averaging
is specially done over the small-scale uctuations, the
probability density function (PDF) of the channel fades
shifts toward that of the large-scale uctuations [17].
Considering the model, this means that, using a large
diameter lens, we should obtain a Gamma distribution
describing the residual large-scale fading after aperture
averaging. The experimental results, however, show that
the resulting scintillation after the receiver lens is better
Turbulence Mitigation by Aperture Averaging in Wireless Optical Systems
ConTEL 2009, ISBN: 978-953-184-131-3 63
0 10 20 30 40 50 60 70 80 90 100
10
6
10
5
10
4
10
3
10
2
10
1
10
0
E
b
/N
0
(dB)
B
E
R


PL, Pt Rx
SP, Pt Rx
PL, D=5mm
SP, D=5mm
PL, D=10mm
SP, D=10mm
PL, D=20mm
SP, D=20mm
PL, D=50mm
SP, D=50mm
PL, D=100mm
SP, D=100mm
PL, D=200mm
SP, D=200mm
(a) Plane (PL) and spherical (SP) wave models.
0 10 20 30 40 50 60 70 80 90 100
10
6
10
5
10
4
10
3
10
2
10
1
10
0
E
b
/N
0
(dB)
B
E
R
Pt Rx
D=5mm
D=10mm
D=20mm
D=50mm
D=100mm
D=200mm
(b) Gaussian-beam wave model.
Figure 6. Contrasting BER performances for plane (PL), spherical (SP)
and Gaussian-beam models, strong turbulence regime with
2
R
= 19.18
and l
0
= 0.
described by a log-Normal (LN) distribution [5]. This is
in agreement with models like the LNME (Log-Normally
Modulated Exponential) and LNMR (Log-Normally Mod-
ulated Rice), also proposed to describe the moderate to
strong turbulence [5]. The reason for which the model
is usually used, rather than LNME or LNMR, is that, it
has a simpler and more tractable form. Andrews et al.
proposed in [5] to consider the model for relatively
strong turbulence for the case of a point receiver, and the
LN model when using a lens of diameter D
0
. It is
interesting to see the difference of these two models by
considering the average BER performance with different
receiver lens diameters. We have presented in Fig. 7 the
BER curves for the two cases of strong and moderate
turbulence regimes. To obtain these curves, we have used
(11) to calculate
2
I
(D) for the model, that is then
used in the LN PDF model.
We notice from Fig. 7 that, for the strong regime, the
and LN models provide close enough results for about
D 20 mm. For the moderate turbulence regime, this
is true for D 50 mm. Remember that for these cases,
we have
0
= 3 mm and
0
= 5.8 mm, respectively. We
deduce that these models become equivalent for D
0
(roughly D > 6
0
). When taking into account the inner
scale, larger apertures are required before that the two
models become equivalent (results are not shown). We
conclude that, if the inner scale l
0
is negligible, for
the case of strong turbulence regime, we can practically
use the hypothesis of LN distribution for fading as the
receiver effective aperture diameters that are usually used
in practical systems are larger than 10 cm [18].
VI. EFFECT OF CHANNEL CODING
Here, our aim is to study the effectiveness of channel
coding combined with aperture averaging for a quasi-
static channel, that is, assuming that the channel fading
is constant over the duration of a frame of symbols.
Here, we consider a simple rate 1/2 recursive systematic
convolutional (RSC) code of constraint length K = 4,
with the octal representation (1, 15/17). The reason we
chose this code is that, the classical convolutional codes
have been shown to be a suitable choice for use in FSO
systems under any turbulence regime, as they make a good
compromise between complexity and performances [4].
We have shown in Fig. 8, the gain in E
b
/N
0
, required
to obtain BER = 10
5
, achieved by aperture averaging
with and without RSC channel coding. The gain is with
respect to the E
b
/N
0
required for an uncoded system
using a point receiver. (The point D = 1 mm corresponds
to a point receiver; we set this value in order to represent
D in the logarithmic scale.) Note that arguing in terms
of E
b
/N
0
has the advantage of taking into account the
channel coding rate, and thus, to make a fair comparison
with the uncoded case.
From Fig. 8 we notice that we have a relatively consider-
able gain for the case of weak turbulence, irrespective of
D. For instance, for D = 50 mm, we have an SNR gain of
3.6 dB. For the cases of moderate and strong turbulence,
however, channel coding appears to be inefcient except
M.A. Khalighi, N. Aitamer, N. Schwartz, S. Bourennane
ConTEL 2009, ISBN: 978-953-184-131-3 64
0 10 20 30 40 50 60 70 80 90 100
10
6
10
5
10
4
10
3
10
2
10
1
10
0
E
b
/N
0
(dB)
B
E
R


, Pt Rx
LN, Pt Rx
, D = 5 mm
LN, D = 5 mm
, D = 10 mm
LN, D = 10 mm
, D = 20 mm
LN, D = 20 mm
, D = 50 mm
LN, D = 50 mm
, D = 200 mm
LN, D = 200 mm
(a) Strong turbulence regime
0 10 20 30 40 50 60 70 80
10
6
10
5
10
4
10
3
10
2
10
1
10
0
E
b
/N
0
(dB)
B
E
R


, Pt Rx
LN, Pt Rx
, D = 5 mm
LN, D = 5 mm
, D = 10 mm
LN, D = 10 mm
, D = 20 mm
LN, D = 20 mm
, D = 50 mm
LN, D = 50 mm
, D = 100 mm
LN, D = 100 mm
, D = 200 mm
LN, D = 200 mm
(b) Moderate turbulence regime
Figure 7. BER versus SNR for and LN models, plane wave model,
l
0
= 0.
1 5 10 20 50 100 200
0
10
20
30
40
50
60
70
80
D (mm)
S
N
R

g
a
i
n

(
d
B
)
Strong
Weak
Moderate
Figure 8. Gain in E
b
/N
0
(w.r.t. a point receiver without coding) versus
receiver lens diameter D for the three cases of turbulence, plane wave
model, l
0
= 0, BER = 10
5
. Dashed lines: aperture averaging without
coding, Solid lines: aperture averaging + RSC (1, 15/17) coding
for very large D. For instance, for D = 200 mm, we have
SNR gains of 2.2 dB and 1.6 dB, for the moderate and
strong turbulence cases, respectively. There is negligible
improvement in the SNR gain when employing more
powerful channel codes (results are not shown). It is
important to note that by SNR gain we mean a gain in
E
b
/N
0
. We conclude that channel coding is suitable only
for the weak turbulence regime, or when a large aperture
size is used.
VII. CONCLUSION AND DISCUSSIONS
We studied in this paper the effect of aperture averag-
ing on the performance of FSO systems under different
conditions of turbulence and optical wave propagation.
Although a part of the presented results are rather well
known, we brought a different look to the problem by
considering the average bit error rate as the performance
criterion, which is among the important parameters in
a practical system. Our presented results can provide a
clearer perspective for understanding the effective im-
provement achieved by employing aperture averaging in
an FSO system. From the presented results, we can point
out the following main concluding remarks:
The effect of the inner scale l
0
is reduced by aperture
averaging. However, for strong turbulence, its effect
on fading is still considerable even when using very
large apertures.
For moderate and strong turbulence regimes, fade
statistics can be described by a LN distribution when
the lens diameter is about 6 to 10 times larger than
the coherence radius
0
.
Under moderate to strong turbulence conditions,
channel coding does not permit considerable perfor-
mance improvement except for very large aperture
sizes. It is the inverse round for the weak turbulence
regime, where a signicant improvement is achieved
irrespective of the aperture size.
Turbulence Mitigation by Aperture Averaging in Wireless Optical Systems
ConTEL 2009, ISBN: 978-953-184-131-3 65
Note that, in our study, we mostly xed the noise
variance and the average received intensity for any
lens diameter D. We explained in Section II that this
represents the case where background noise dominates
thermal noise and when xed FOV receivers are used.
If diffraction-limited receivers are used, the amount of
received background noise is essentially independent
of the receiver aperture size [14]. In practice, for
reasons of simplifying the tracking task and to alleviate
beam wandering, xed FOV receivers are used in most
FSO systems. On the other hand, if the thermal noise
dominates, in addition to the fading reduction illustrated
in the presented simulation results, we benet from a
gain in the received signal power as well. The main
conclusions of our work remain valid, nevertheless.
Finally, notice that, in general, aperture averaging is
not sufcient for mitigating atmospheric turbulence in a
practical FSO system. It should be used in combination
with channel coding or adaptive optics, for example.
In fact, a complementary efcient way of reducing the
effect of turbulence is to use adaptive optics at the
transmitter and/or at the receiver [19].
ACKNOWLEDGMENT
The authors wish to thank Fr ed eric Chazalet from Shaktiware
Co., Marseille, France, and Prof. Hassan Akhouayri from Institut
Fresnel, Marseille, France, for their fruitful discussions.
REFERENCES
[1] V. W. S. Chan, Free-space optical communications, Journal of
Lightwave Technology, vol. 24, no. 12, pp. 47504762, Dec. 2006.
[2] L. C. Andrews and R. L. Phillips, Laser Beam Propagation
Through Random Media, 2nd ed. Bellingham, Washington: SPIE
Press, 2005.
[3] F. Xu, M. A. Khalighi, P. Causs e, and S. Bourennane, Perfor-
mance of coded time-diversity free-space optical links, Queens
24th Biennial Symposium on Communications (QSBC), pp. 146
149, June 2008, Kingston, Canada.
[4] , Channel coding and time-diversity for optical wireless
links, Optics Express, vol. 17, no. 2, pp. 872887, Jan. 2009.
[5] F. S. Vetelino, C. Young, L. C. Andrews, and J. Recolons,
Aperture averaging effects on the probability density of irradiance
uctuations in moderate-to-strong turbulence, Applied Optics,
vol. 46, no. 11, pp. 20992108, Apr. 2007.
[6] J. H. Churnside, Aperture averaging of optical scintillations in
the turbulent atmosphere, Applied Optics, vol. 30, no. 15, pp.
19821994, May 1991.
[7] L. C. Andrews, Aperture-averaging factor for optical scintillations
of plane and spherical waves in the atmosphere, Journal of
Optical Society of America (JOSA) A, vol. 9, no. 4, pp. 597600,
Apr. 1992.
[8] H. Yuksel, S. Milner, and C. C. Davis, Aperture averaging for
optimizing receiver design and system performance on free-space
optical communication links, Journal of Optical Networking,
vol. 4, no. 8, pp. 462475, Aug. 2005.
[9] L. C. Andrews, R. L. Phillips, and C. Y. Hopen, Aperture
averaging of optical scintillations: power uctuations and the
temporal spectrum, Waves Random Media, vol. 10, pp. 5370,
2000.
[10] F. S. Vetelino, C. Young, and L. C. Andrews, Fade statistics and
aperture averaging for Gaussian beam waves in moderate-to-strong
turbulence, Applied Optics, vol. 46, no. 18, pp. 37803789, June
2007.
[11] G. L. Bastin, L. C. Andrews, R. L. Phillips, R. A. Nelsond,
B. A. Ferrelld, M. R. Borbathe, D. J. Galuse, P. G. Chine, W. G.
Harrisa, J. A. Marna, G. L. B. D. Wayneb, and R. Pescatoreb,
Measurements of aperture averaging on bit-error-rate, Proceed-
ings of SPIE, Atmospheric Optical Modeling, Measurement, and
Simulation, vol. 5891, pp. 021 0212, 2005.
[12] N. Perlot and D. Fritzsche, Aperture-averaging, theory and mea-
surements, Proceedings of SPIE, Free-Space Laser Communica-
tion Technologies XVI, vol. 5338, pp. 233242, 2004.
[13] L. M. Wasiczko and C. C. Davis, Aperture averaging of optical
scintillations in the atmosphere: experimental results, Proceedings
of SPIE, Atmospheric Propagation II, vol. 5793, pp. 197208,
2005.
[14] R. M. Gagliardi and S. Karp, Optical Communications, 2nd ed.
John Wiley & Sons, 1995.
[15] A. K. Majumdar and J. C. Riclkin, Free-Space Laser Communi-
cations: Principles And Advances. Springer-Verlag, 2007.
[16] A. Consortini, E. Cochetti, J. H. Churnside, and R. J. Hill, Inner-
scale effect on irradiance variance measured for weak-to-strong
atmospheric scintillation, Journal of Optical Society of America
(JOSA) A, vol. 10, no. 11, pp. 23542362, Nov. 1993.
[17] J. A. Anguita, M. A. Neifeld, and B. V. Vasic, Spatial correlation
and irradiance statistics in a multiple-beam terrestrial free-space
optical communication link, Applied Optics, vol. 46, no. 26, pp.
65616571, Sept. 2007.
[18] fSONA Co. Optical Wireless Products, http://www.fsona.com.
[19] N. H. Schwartz, N. V edrenne, V. Michau, M.-T. Velluet, and
F. Chazalet, Mitigation of atmospheric effects by adaptive optics
for free-space optical comminucations, Proceedings of the SPIE
Photonics West Conference, 2009.
ConTEL 2009, ISBN: 978-953-184-131-3 66

Vous aimerez peut-être aussi