You are on page 1of 13

Opt Quant Electron (2016) 48:65

DOI 10.1007/s11082-015-0338-4

Performance of free space optical links in presence


of turbulence, pointing errors and adverse weather
conditions

Prabhmandeep Kaur1 • Virander Kumar Jain1 •

Subrat Kar1

Received: 15 June 2015 / Accepted: 20 November 2015


 Springer Science+Business Media New York 2015

Abstract The ergodic capacity, bit error rate (BER) and outage probability of a free
space optical communication link is analyzed in the presence of gamma–gamma dis-
tributed turbulence and pointing errors in presence of various weather conditions viz., very
clear air, drizzle, haze and fog. It is seen that the capacity decreases, BER and outage
probability increase as the pointing error jitter increases. The worst degradation of the
above performance parameters occurs in fog conditions while very clear air conditions give
the least degradation. Further, the effect of varying beam divergence angle/beam waist is
studied. It is observed that the beam divergence angle/beam waist can be optimized to
attain maximum capacity, minimum BER and outage probability. The optimum value
depends on the normalized pointing error jitter and increases with increasing jitter.
However, for a fixed value of normalized jitter, it is independent of the weather conditions.
Also, the system performance is more susceptible to channel impairments for beam
divergence angles lower than the optimum value as compared to larger angles.

Keywords Atmospheric attenuation  Free space optical (FSO) communication 


Gamma–gamma distribution  Pointing errors  Turbulence

1 Introduction

Free space optical (FSO) communication has attracted significant research interest in
recent years to meet the high bandwidth and high data rate demands of next generation
communication systems. However, the quality of information carrying optical wave is
degraded as it propagates through the terrestrial atmosphere or free space. A primary

& Prabhmandeep Kaur


pkaur.ee@gmail.com
1
Department of Electrical Engineering, Indian Institute of Technology Delhi, Hauz Khas,
New Delhi 110016, India

123
65 Page 2 of 13 P. Kaur et al.

challenge faced by FSO links is the presence of turbulence in the atmosphere. The vari-
ations in temperature, pressure, wind velocity, etc., lead to formation of optical turbules,
also known as eddies that vary in refractive index. These result in intensity and phase
fluctuations of the propagating optical wave (Andrews et al. 1999; Zhu and Kahn 2002).
Since FSO communication links have a prerequisite of proper alignment of the transmitter
and receiver, blockages in the laser beam path can lead to outright loss of the signal.
Misalignment can also occur due to building motion, earthquakes, vibrations, etc., that
severely affect performance of the link. These links are also adversely affected by the
presence of particulate matter such as aerosols, haze, fog, etc., in the atmosphere. Such
particles cause exponential extinction of the signal carrying optical wave by the phe-
nomena of scattering and absorption (Majumdar and Ricklin 2007).
Many mathematical models have been proposed to investigate the degrading effects of
numerous challenges presented to FSO links. To account for the effect of weather, Beer–
Lambert’s law is applied to determine the path loss or optical extinction caused by various
weather conditions (Nebuloni 2005). The attenuation coefficient is a function of wave-
length and visibility that varies for different weather conditions. Kruse model (Kruse et al.
1962) was the first model used to evaluate attenuation for a range of weather conditions
varying from fog to very clear air conditions. However, it is less reliable for fog conditions
as it was developed originally using data in dense haze conditions. Further studies sug-
gested wavelength independence of optical attenuation in fog conditions with lower visi-
bilities (\0.5 km). Thus, Kruse model was modified using Mie theory to obtain Kim model
(Kim et al. 2001) wherin the wavelength dependence was removed for dense fog condi-
tions and it is valid over the entire visibility range. Al Naboulsi et al. (2004) derived
attenuation coefficients specific to advection and convection fog for a wavelength range of
690–1550 nm and visibilities ranging from 0.05 to 1 km. Grabner and Kvicera (2011)
proposed a wavelength dependent model using Mie single scattering theory in fog and haze
conditions for visibilities lower than 10 km. Of the available models, Kim model is best
suited to estimate attenuation as it conforms to the experimental data and is valid over a
wide visibility range (Ghassemlooy et al. 2012). For turbulence, log-normal model has
been used to describe weak turbulence and for strong turbulence, negative exponential and
K-distribution can be employed. To model a range of turbulence strengths varying from
weak to strong, gamma–gamma model emerges as the preferred choice given its com-
pliance to the practical data (Churnside and Clifford 1987; Andrews et al. 2001). To study
the performance of a FSO link in presence of turbulence only, ergodic channel capacity
and outage probability (Pout) have been evaluated using various models for different tur-
bulence regimes (Laourine et al. 2009; Nistazakis et al. 2009a, 2011). Capacity has also
been determined under the effect of various weather conditions and turbulence strengths
(Kaur et al. 2014a). Also, the impact of pointing error has been studied extensively for
space-based FSO links (Barry and Mecherle 1985; Arnon and Kopeika 1997). Bit error rate
(BER) was calculated considering building sway and log-normal turbulence (Arnon 2003).
The effect of fog with pointing errors was studied to determine the BER by using Monte-
Carlo simulations to model the passage of light through particulate matter (Kedar and
Arnon 2003). Farid and Hranilovic (2007) characterized pointing errors by modeling the
radial displacement of the laser beam at the receiver aperture as a Rayleigh distribution
considering independent and identical jitter in horizontal as well as vertical directions.
However, when taking variable jitter in the orthogonal directions, Hoyt or Nakagami-q
distributions are used (Gappmair et al. 2011). The capacity, BER and Pout of the FSO link
under the combined effect of pointing errors and turbulence were studied (Gappmair et al.
2010; Sandalidis et al. 2008; Borah and Voelz 2009; Jurado-Navas et al. 2012; Farid and

123
Performance of free space optical links in presence of… Page 3 of 13 65

Hranilovic 2012; Yang et al. 2014) and beam width optimization was done (Ren et al.
2010; Liu et al. 2010; Sandalidis 2008). However, the effect of weather on the optimized
beam width was not considered.
In this paper, we evaluate the capacity, BER and Pout of an intensity modulated/direct
detection (IM/DD) link under the combined effect of turbulence, pointing errors and
weather conditions viz., very clear air, clear air/drizzle, haze, and fog. Turbulence is
modeled by a gamma–gamma distribution given its validity over a wide range of turbu-
lence regimes. Rayleigh distribution is used for pointing errors and Beer–Lambert’s law for
accounting the effect of various weather conditions. The effect of increasing pointing error
jitter on link performance is analyzed. Also optimum beam divergence angles are deter-
mined to maximize capacity and minimize BER and Pout.

2 System and channel models

The transmitted signal is intensity modulated using on-off keying (OOK). The OOK sig-
naling scheme is considered because of its low cost and ease of implementation. At the
receiver end, direct detection is employed. The output signal y is given by
y ¼ hRx þ n ð1Þ
where x is the transmitted signal expressed as x 2 f0; 2Pt g, with Pt being the average
transmitted optical power, n the additive white Gaussian noise (AWGN) with variance
r2n ; R the detector responsivity (in A/W) and h the channel state. Since we consider the
combined effects of weather conditions, atmospheric turbulence and pointing errors, h can
be written as (Farid and Hranilovic 2007)
h ¼ hl ha hp ð2Þ
where hl represents the path loss due to various weather conditions, ha fading due to
turbulence and hp the effect of pointing errors.

2.1 Atmospheric channel loss

The loss caused by adverse weather conditions is exponential in nature and can be
determined using the Beer–Lambert’s law (Al Naboulsi et al. 2004) as follows
hl ¼ erz ð3Þ
where z is the FSO link length and r the atmospheric attenuation coefficient which is a
function of the operating wavelength, k and the visibility, V. Visibility is a meteorological
concept that is used to characterize various weather conditions depending upon the
transparency of the atmosphere. These parameters are related as shown in the equation
below (Ghassemlooy et al. 2012).
 
3:91 k q
r¼ ð4Þ
V 550
where q is a parameter that depends on the size and distribution of scattering particles in
the atmosphere. The values of q are taken from the Kim model proposed by Kim et al.
(2001).

123
65 Page 4 of 13 P. Kaur et al.

8
>
> 1:6; for V  50 km
>
>
< 1:3; for 6 km  V\50 km
q ¼ 0:16V þ 0:34; for 1 km  V \6 km ð5Þ
>
>
> V  0:5;
> for 0:5 km  V \1 km
:
0; for V \0:5 km
The values of visibility for various weather conditions viz., very clear air, clear air/drizzle,
haze, fog, etc., and the corresponding attenuation coefficients at an operating wavelength
of 1550 nm are tabulated in Table 1.

2.2 Turbulence model

Turbulence is modeled as a gamma–gamma distribution given its validity for all turbulence
strengths and tractability in calculations. The probability density function (pdf) fha ðha Þ of
the fading ha caused due to turbulence is thus given by Andrews et al. (2001)
aþb
2ðabÞ 2 aþb  pffiffiffiffiffiffiffiffiffiffi
1
fha ðha Þ ¼ ha2 Kab 2 abha ð6Þ
CðaÞCðbÞ
where Kab ð:Þ is the modified Bessel function of the second kind and a and b are the
effective number of large and small scale turbulent eddies, respectively. Their values are a
function of the receiver aperture diameter and Rytov variance that depends on the
refractive index structure parameter Cn2 , the wavelength of operation and link length
(Andrews et al. 2001).
The Bessel function
 in!Eq. (6) can be represented as a Meijer-G function i.e.,
pffiffiffi  

2;0
Km ð2 xÞ ¼ 12 G0;2 x m ; m (Prudnikov et al. 1986, Sect. 8.4.23), to obtain the pdf,
2 2
fha ðha Þ as
aþb aþb1
 !

ðabÞ 2 ha2 2;0  a  bb  a
fha ðha Þ ¼ G abha  ð7Þ
CðaÞCðbÞ 0;2  ;
2 2

2.3 Pointing errors

FSO being a line of sight (LoS) technology is severely affected by misalignment or


pointing error between the transmitter and receiver. The pdf of the pointing loss fhp ðhp Þ was
derived by Farid and Hranilovic (2007) assuming independent and identical pointing error

Table 1 Visibility and attenua-


Weather condition Visibility (km) Attenuation (dB/km)
tion coefficients for various
weather conditions at 1550 nm
wavelength (Kaur et al. 2014b) Very clear air 50.0 0.0647
Clear air/drizzle 20.0 0.2208
Haze 6.0 0.7360
Thin fog 2.0 4.2850
Moderate fog 0.6 25.5160

123
Performance of free space optical links in presence of… Page 5 of 13 65

jitter in both horizontal and vertical directions. The jitter is taken to be Gaussian distributed
with zero mean and variance r2s . Thus the radial displacement of the received laser beam
(with a beam waist wz (Ricklin and Davidson 2002)) from the center of the circular
receiver aperture (with radius a) is modeled as a Rayleigh distribution (Arnon 2003). The
pdf of hp is given as
g2 2
fhp ðhp Þ ¼ 2 hgp 1 ; 0  hp  Ao ð8Þ
Ago
pffiffiffi pffiffiffi
where Ao ¼ ½erfðmÞ2 with m ¼ ð paÞ=ð 2wz Þ and g ¼ we =2rs with we being the equiv-
pffiffiffi 2
alent beam width given by w2e ¼ w2z perfðmÞ=2mem . It may be mentioned that the beam
divergence angle h depends on wz and is given by h ¼ 2wz =z.

2.4 Combined channel model

The pdf of the channel state h combines the effect of weather conditions hl , turbulence ha
and pointing errors hp and can be written as (Farid and Hranilovic 2007)
Z
fh ðhÞ ¼ fhjha ðhjha Þfha ðha Þdha ð9Þ

where fhjha ðhjha Þ ¼ ½fhp ðh=ha hl Þ=ha hl , given that hl can be quantified using Eq. (3). From
Eqs. (7)–(9), we obtain
Z 1  !

aþb 2
2 g 1 2;0  a  bb  a
fh ðhÞ ¼ Mh ha G0;2 abha  dha ð10Þ
Lh  ;
2 2
aþb 2
where Mh ¼ ½g2 hg 1 ðabÞ 2 =½ðhl Ao Þg CðaÞCðbÞ and Lh ¼ h=Ao hl . After integration using
2

(Prudnikov et al. 1986, Sect. 2.24.2), we get


0  1
 aþb
 1  þ g 2
aþb
3;0 B abh  2 C
fh ðhÞ ¼ Nh h 2 1 G1;3 @  A ð11Þ
Ao hl  a þ b 2 ab ba
 þg ; ;
2 2 2
aþb aþb
where Nh ¼ ½g2 ðabÞ 2 =½ðhl Ao Þ 2 CðaÞCðbÞ. Using the translation formula for Meijer-G
functions (Prudnikov et al. 1986, Sect. 8.2.2), in the above equation, fh ðhÞ can be written as
  
g2 ab 3;0 abh 
 g2
fh ðhÞ ¼ G ð12Þ
Ao hl CðaÞCðbÞ 1;3 Ao hl  g2  1; a  1; b  1

3 Performance evaluation

The FSO link performance is evaluated in terms of capacity, BER and outage probability.
The analytic expressions derived for the same are explained in the following subsections.

123
65 Page 6 of 13 P. Kaur et al.

3.1 Capacity

Channel capacity is a random variable that defines the maximum achievable data rate for
reliable communication between transmitter and receiver. Since capacity varies with the
instantaneous SNR i.e., SNR(h) which is a function of the channel state h (Liu et al. 2010),
the average channel capacity can be written as
Z 1
C¼ Blog2 ð1 þ SNRðhÞÞfh ðhÞdh ð13Þ
0

where B is the channel’s bandwidth and SNR(hÞ ¼ ch2 , with c ¼ R2 P2t =r2n being the
average electrical SNR. Expressing
   the log function in Eq. (13) as a Meijer-G function,
 1; 1
1
i.e., log2 ð1 þ xÞ ¼ ln2 1;2
G2;2 x (Prudnikov et al. 1986, Sect. 8.4.6), substituting for
1; 0
fh ðhÞ from Eq. (12) and integrating using (Prudnikov et al. 1986, Sect. 2.24.1), we get
0  1
 1  g2 2  g2 1  a 2  a 1  b 2  b
 1; 1; ; ; ; ; ;
B  2 2 2 2 2 2 C
C ¼ Nc G1;8
8;4 @Mc  A ð14Þ
g2 1  g2
 1;  ; ;0
2 2
where the parameters Nc and Mc are given by ½2aþb3 g2 B=½plnð2ÞCðaÞCðbÞ and
½16cðAo hl Þ2 =½ðabÞ2 , respectively.

3.2 Bit error rate

Error probability is an important performance metric and is evaluated by integrating the


conditional error probability, Pe ðhÞ over the pdf of the channel state, fh ðhÞ (Kaur et al.
2014b).
Z 1
BER ¼ Pe ðhÞfh ðhÞdh ð15Þ
0

Pe ðhÞ is a function of the instantaneous SNR and is given by Sandalidis et al. (2008)
rffiffiffiffiffiffiffi!
1 ch2
Pe ðhÞ ¼ erfc ð16Þ
2 2
qffiffiffiffiffi
ch2
Rewriting the erfc(.) as a Meijer-G function in Eq. (16), i.e., erfc 2 ¼
 !

2  1
pffiffi G
1 2;0 ch 1 (Prudnikov et al. 1986, Sect. 8.4.14) and substituting in Eq. (15), the
p 1;2 2  0;
2
integral can be solved (Prudnikov et al. 1986, Sect. 2.24.1) to obtain the BER as
0  1
 1  g2 2  g2 1  a 2  a 1  b 2  b
 ; ; ; ; ; ; 1
2;6 B  C
BER ¼ Nb G7;4 @Mb  2 2 2 2 2 2 A ð17Þ
 1 g2 1  g2
 0; ;  ;
2 2 2

123
Performance of free space optical links in presence of… Page 7 of 13 65

where the parameters Nb and Mb are given by ½2aþb4 g2 =½p3=2 CðaÞCðbÞ and
½8cðAo hl Þ2 =½ðabÞ2 , respectively. Lowering the order of the Meijer-G function in Eq. (17)
(Prudnikov et al. 1986, Sect. 8.2.2), we get
0  1
 2  g2 1  a 2  a 1  b 2  b
 ; ; ; ; ; 1
2;5 B  C
BER ¼ Nb G6;3 @Mb  2 2 2 2 2 A ð18Þ
 1 g2
 0; ; 
2 2

3.3 Outage probability

Outage probability is the probability that the instantaneous SNR l, falls below a threshold
value cth , leading to unsatisfactory system performance (Sandalidis and Tsiftsis 2008). It is
defined as
Pout ¼ Prðl\cth Þ ¼ Fl ðcth Þ ð19Þ
where Fl ð:Þ is the cumulative distribution function (cdf) of the instantaneous SNR l. The
pdf of l i.e., fl ðlÞ is given by
 sffiffiffiffiffiffiffiffiffiffi
 rffiffiffi
d l  l

fl ðlÞ ¼   fh ð20Þ
dl c  c

Solving Eq. (20), we obtain


 rffiffiffi 
No 1 l g2
fl ðlÞ ¼ ðlcÞ2 G1;3
3;0
Mo ð21Þ
2 c  g2  1; a  1; b  1

where No ¼ ½g2 ab=½CðaÞCðbÞAo hl  and Mo ¼ ½ab=½Ao hl . Integrating Eq. (21) (Prudnikov


et al. 1986, Sect. 2.24.2), we obtain the cdf of l as
rffiffiffi  rffiffiffi 
l 3;1 l 0; g2
Fl ðlÞ ¼ No G2;4 Mo ð22Þ
c c  g2  1; a  1; b  1; 1
Using the translation formula for Meijer-G in Eq. (22) (Prudnikov et al. 1986, Sect. 8.2.2)
and substituting cth as the threshold value, we get the outage probability as
 rffiffiffiffiffi 
g2 3;1 ab cth  1; g2 þ 1
Pout ¼ G ð23Þ
CðaÞCðbÞ 2;4 Ao hl c  g2 ; a; b; 0
Here, c=cth is the normalized average electrical SNR, also known as normalized outage
threshold.

4 Results and discussion

We plot the capacity, BER and Pout for a terrestrial FSO link of length 3.5 km operating at
a wavelength of 1550 nm under the effect of turbulence and pointing errors in various
weather conditions. The turbulence strength is determined by the value of Cn2 , taken as
1:7  1014 m2=3 in our work, valid for moderate turbulence regime (Nistazakis et al.
2009b). Since a strong inverse correlation exists between the turbulence strength and

123
65 Page 8 of 13 P. Kaur et al.

attenuation caused by weather conditions, i.e., fog is highly unlikely in the presence of
strong turbulence and vice versa, we have obtained results for thin fog conditions that can
be present simultaneously with moderate turbulence (Bushuev and Arnon 2006; Farid and
Hranilovic 2007). The diameter of the receiver aperture is taken as 10 cm and the effect of
various weather conditions is accounted by the value of hl calculated from Eq. (3) using
Table 1.
Average channel capacity, hCi=B (in bits/s/Hz) is plotted against the SNR, c in Fig. 1
for weather conditions of very clear air and fog in the absence and presence of pointing
errors. It is seen that the capacity hCi=B, is maximum in the absence of pointing errors and
decreases as the normalized pointing error jitter (rs =a) increases from 1 to 4. This trend is
valid irrespective of the weather condition. The results for clear air/drizzle and haze are not
presented in the graphical form, but included in Table 2. This table lists the SNR penalties
required to obtain a capacity of 1 bits/s/Hz under various weather conditions such as very
clear air, clear air/drizzle, haze and fog for different values of normalized jitter as com-
pared to very clear weather conditions with no pointing error. As expected, the SNR
penalties increase with increasing jitter. Also, as weather conditions deteriorate from very
clear air to fog, the SNR penalties increase over all the jitter values.
Further, the channel capacity hCi=B, is plotted against the beam divergence angle h for
different weather conditions taking rs =a ¼ 2 in Fig. 2a. The results are obtained for Pt ¼
10 mW; r2n ¼ 1014 A2 and R ¼ 0:9 A/W. Since beams with large divergence angles lead
to increased beam divergence/geometric loss, but help compensate for the pointing error
loss and vice versa (Ghassemlooy et al. 2012), an optimum value of the beam divergence
angle exists. At this optimum value, the geometric and pointing error losses are balanced
out, yielding the best performance. From Fig. 2a it is seen that hCi=B is maximized at a
particular value of h, which is the optimum value of beam divergence angle, hopt . This
optimum value remains nearly constant for all the weather conditions of very clear air,
drizzle, haze and fog at a constant rs =a. However, the value of hCi=B at hopt , decreases as
the weather deteriorates from very clear air to fog conditions. We then study the effect of
varying the pointing error jitter on hopt . Figure 2b plots hCi=B against the beam divergence
angle for varying values of normalized jitter in very clear air conditions. Since hopt is
independent of the weather conditions for a given rs =a, weather conditions of clear
air/drizzle, haze and fog are not considered. It is observed that as the normalized jitter

Fig. 1 Average channel capacity


hCi=B versus SNR in the absence
and presence of pointing errors
(PE)

123
Performance of free space optical links in presence of… Page 9 of 13 65

Table 2 SNR penalties with


Weather SNR penalties (in dB)
respect to very clear air and no
conditions
pointing errors for a capacity of rs =a ¼ 0:5 rs =a ¼ 1 rs =a ¼ 2 rs =a ¼ 4
1 bits/s/Hz
Very clear air 9.73 13.17 20.80 34.58
Clear air/drizzle 10.79 14.24 21.90 35.68
Haze 14.40 17.87 25.53 39.29
Fog 26.20 29.64 37.17 51.08

Fig. 2 Average channel capacity hCi=B versus beam divergence angle h for a various weather conditions at
rs =a ¼ 2 and b varying jitter in very clear air conditions

Fig. 3 BER versus SNR in the


absence and presence of pointing
errors (PE)

123
65 Page 10 of 13 P. Kaur et al.

increases, the value of hopt also increases. For a normalized jitter of 1, 2, 3 and 4, hopt
values (in mrad) are 0.06 with hCimax =B ¼ 26:7 bits/s/Hz; 0:11 with hCimax =B ¼
23:7 bits/s/Hz; 0:17 with hCimax =B ¼ 21:5 bits/s/Hz and 0.23 with hCimax =B ¼ 19:7 bits/s/
Hz, respectively under moderate turbulence and very clear air condition. Also, it can be
clearly seen that for h\hopt , there is a sharp decline in capacity as compared to h [ hopt .
Further, BER is plotted against the SNR, c in Fig. 3 for weather conditions of very clear
air and fog in presence of moderate turbulence. The BER performance of the link in
absence of pointing errors is compared with the link performance at different values of
normalized pointing error jitter. As expected, BER is minimum in absence of pointing
errors and degrades (increases) as rs =a increases from 1 to 4, irrespective of the weather
condition. Also the BER performance of the link degrades as the weather conditions
deteriorate from very clear air to fog.
Next, we study the effect of varying h on BER (taking system parameters as mentioned
earlier) in Fig. 4a for the weather conditions of very clear air, clear air/drizzle, haze and
fog at rs =a ¼ 2. In order to obtain best performance of the FSO link in terms of minimum
BER, an optimum value of hopt exists at which a critical balance between the geometric
and pointing error losses is achieved. It is observed that hopt remains almost constant over
all the weather conditions for a given normalized pointing error jitter. At these optimized
values, BER is lowest for the weather condition of very clear air and increases as the
weather conditions deteriorate further to clear air/drizzle and haze, being the maximum in
fog conditions. However, as the pointing error jitter increases, optimum value of the beam
divergence angle also increases. This can be clearly observed from Fig. 4b wherein BER is
plotted against h for noramlized jitter varying from 1 to 4 in very clear air condition. BER
is minimized for hopt (in mrad) of 0.12 with BERmin ¼ 3:8  1013 , 0.24 with BERmin ¼
5  1011 , 0.36 with BERmin ¼ 8:7  1010 and 0.48 with BERmin ¼ 6:5  109 at a
normalized jitter of 1, 2, 3 and 4, respectively under moderate turbulence and very clear air
condition. Also, it is seen that for h\hopt , there is a steeper increase in BER as compared
to h [ hopt .

Fig. 4 BER versus beam divergence angle h for a various weather conditions at rs =a ¼ 2 and b varying
jitter in very clear air conditions

123
Performance of free space optical links in presence of… Page 11 of 13 65

Fig. 5 Outage Probability, Pout


versus normalized outage
threshold, c=cth in the absence
and presence of pointing errors
(PE)

Outage probability, Pout is plotted against the normalized outage threshold, c=cth in
Fig. 5 for weather conditions of very clear air and fog in presence and absence of pointing
errors under moderate turbulence. Pout follows a similar trend as BER over all weather
conditions, being minimum in the absence of pointing errors and degrading (increasing)
with increasing rs =a. The effect of varying h on Pout at a normalized outage threshold of
80 dB is studied in Fig. 6. Figure 6a plots the Pout against h at rs =a ¼ 2 for various
weather conditions viz., very clear air, clear air/drizzle, haze and fog. Further, the variation
of Pout with h is plotted in Fig. 6b, for normalized jitter varying from 1 to 4 in very clear air
condition. As observed for BER variations with h, the optimum beam divergence angle hopt
for obtaining the minimum Pout remains almost constant over different weather conditions
at a fixed pointing error jitter. Also, with an increase in the normalized jitter, hopt increases.

Fig. 6 Outage Probability Pout versus beam divergence angle h for a various weather conditions at rs =a ¼
2 and b varying jitter in very clear air conditions

123
65 Page 12 of 13 P. Kaur et al.

At a normalized outage threshold of 80 dB, Pout is minimized at hopt (in mrad) of 0.11 with
Pout;min ¼ 1:2  109 , 0.23 with Pout;min ¼ 1:5  107 , 0.35 with Pout;min ¼ 2:4  106 and
0.46 with Pout;min ¼ 1:7  105 at a normalized jitter of 1, 2, 3 and 4, respectively in very
clear air condition under the influence of moderate turbulence. Furthermore, over all the
weather conditions and jitter values, it is seen that for h\hopt ; Pout degrades more rapidly
as compared to h [ hopt . The results have also been verified by Monte-Carlo simulations.
Since noticeable differences are not observed between the derived solutions and simula-
tions, the latter results are not shown due to brevity.
From the hopt values determined for the performance parameters of hCi=B, BER and
Pout, the corresponding values of optimum beam waist wz;opt ð¼zhopt =2Þ can be calculated.

5 Conclusion

The performance of a FSO link is evaluated in presence of atmospheric turbulence,


pointing errors and various weather conditions. It is seen that capacity decreases whereas
BER and Pout increase as the weather conditions degrade from very clear air to fog and also
as the pointing error jitter increases. To maximize capacity and to minimize the BER and
Pout, optimum beam divergence angle (hopt )/optimum beam waist (wz;opt ) should be used.
As the normalized jitter is increased, a larger hopt is required implying a broader wopt . With
varying weather conditions viz., very clear air, drizzle, haze and fog, under a constant
normalized jitter, these values however, remain almost constant. Also, beam divergence
angles less than hopt degrade the performance of the FSO system more severely as com-
pared to larger angles.

References
Al Naboulsi, M., Sizun, H., de Fornel, F.: Fog attenuation prediction for optical and infrared waves. Opt.
Eng. 43, 319–329 (2004)
Andrews, L.C., Phillips, R.L., Hopen, C.Y., Al-Habash, M.A.: Theory of optical scintillation. J. Opt. Soc.
Am. A 16, 1417–1429 (1999)
Andrews, L.C., Phillips, R.L., Hopen, C.Y.: Laser Beam Scintillation with Applications. SPIE Press,
Bellingham (2001)
Arnon, S.: Effects of atmospheric turbulence and building sway on optical wireless communication systems.
Opt. Lett. 28, 129–131 (2003)
Arnon, S., Kopeika, N.S.: Laser satellite communication network vibration effect and possible solutions.
Proc. IEEE 85, 1646–1661 (1997)
Barry, J.D., Mecherle, G.S.: Beam pointing error as a significant design parameter for satellite borne, free-
space optical communication systems. Opt. Eng. 24, 241,049–241,054 (1985)
Borah, D.K., Voelz, D.G.: Pointing error effects on free-space optical communication links in the presence
of atmospheric turbulence. J. Lightwave Technol. 27, 3965–3973 (2009)
Bushuev, D., Arnon, S.: Analysis of the performance of a wireless optical multi-input to multi-output
communication system. J. Opt. Soc. Am. A 23, 1722–1730 (2006)
Churnside, J.H., Clifford, S.F.: Log-normal rician probability-density function of optical scintillations in the
turbulent atmosphere. J. Opt. Soc. Am. A 4, 1923–1930 (1987)
Farid, A.A., Hranilovic, S.: Outage capacity optimization for free-space optical links with pointing errors.
J. Lightwave Technol. 25, 1702–1710 (2007)
Farid, A.A., Hranilovic, S.: Diversity gain and outage probability for MIMO free-space optical links with
misalignment. IEEE Trans. Commun. 60, 479–487 (2012)
Gappmair, W., Hranilovic, S., Leitgeb, E.: Performance of PPM on terrestrial FSO links with turbulence and
pointing errors. IEEE Commun. Lett. 14, 468–470 (2010)

123
Performance of free space optical links in presence of… Page 13 of 13 65

Gappmair, W., Hranilovic, S., Leitgeb, E.: OOK performance for terrestrial FSO links in turbulent atmo-
sphere with pointing errors modeled by Hoyt distributions. IEEE Commun. Lett. 15, 875–877 (2011)
Ghassemlooy, Z., Popoola, W., Rajbhandari, S.: Optical Wireless Communications: System and Channel
Modelling with Matlab. CRC Press, Boca Raton (2012)
Grabner, M., Kvicera, V.: The wavelength dependent model of extinction in fog and haze for free space
optical communication. Opt. Express 19, 3379–3386 (2011)
Jurado-Navas, A., Garrido-Balsells, J.M., Paris, J.F., Castillo-Vázquez, M., Puerta-Notario, A.: Impact of
pointing errors on the performance of generalized atmospheric optical channels. Opt. Express 20,
12550–12562 (2012)
Kaur, P., Jain, V.K., Kar, S.: Effect of atmospheric conditions and aperture averaging on capacity of free
space optical links. Opt. Quantum Electron. 46, 1139–1148 (2014a)
Kaur, P., Jain, V.K., Kar, S.: Performance analysis of FSO array receivers in presence of atmospheric
turbulence. IEEE Photonics Technol. Lett. 26, 1165–1168 (2014b)
Kedar, D., Arnon, S.: Optical wireless communication through fog in the presence of pointing errors. Appl.
Opt. 42, 4946–4954 (2003)
Kim, I.I., McArthur, B., Korevaar, E.J.: Comparison of laser beam propagation at 785 nm and 1550 nm in
fog and haze for optical wireless communications. Proc. SPIE 4214, 26–37 (2001)
Kruse, P.W., McGlauchlin, L.D., McQuistan, R.B.: Elements of Infrared Technology: Generation, Trans-
mission and Detection. Wiley, New York (1962)
Laourine, A., Stephenne, A., Affes, S.: On the capacity of log-normal fading channels. IEEE Trans.
Commun. 57, 1603–1607 (2009)
Liu, C., Yao, Y., Sun, Y.X., Xiao, J.J., Zhao, X.H.: Average capacity optimization in free-space optical
communication system over atmospheric turbulence channels with pointing errors. Opt. Lett. 35,
3171–3173 (2010)
Majumdar, A.K., Ricklin, J.C.: Free-Space Laser Communications: Principles and Advances. Springer, New
York (2007)
Nebuloni, R.: Empirical relationships between extinction coefficient and visibility in fog. Appl. Opt. 44,
3795–3804 (2005)
Nistazakis, H.E., Karagianni, E.A., Tsigopoulos, A.D., Fafalios, M.E., Tombras, G.S.: Average capacity of
optical wireless communication systems over atmospheric turbulence channels. J. Lightwave Technol.
27, 974–979 (2009a)
Nistazakis, H.E., Tsiftsis, T.A., Tombras, G.S.: Performance analysis of free-space optical communication
systems over atmospheric turbulence channels. IET Commun. 3, 1402–1409 (2009b)
Nistazakis, H.E., Tsigopoulos, A.D., Hanias, M.P., Psychogios, C.D., Marinos, D., Aidinis, C., Tombras,
G.S.: Estimation of outage capacity for free space optical links over I-K and K turbulent channels.
Radioengineering 20, 493–498 (2011)
Prudnikov, A.P., Brychkov, Y.A., Marichev, O.I.: Integrals and Series. Gordon and Breach Science Pub-
lishers, Amsterdam (1986)
Ren, Y., Dang, A., Luo, B., Guo, H.: Capacities for long-distance free-space optical links under beam
wander effects. IEEE Photonics Technol. Lett. 22, 1069–1071 (2010)
Ricklin, J.C., Davidson, F.M.: Atmospheric turbulence effects on a partially coherent gaussian beam:
implications for free-space laser communication. J. Opt. Soc. Am. A 19, 1794–1802 (2002)
Sandalidis, H.G.: Optimization models for misalignment fading mitigation in optical wireless links. IEEE
Commun. Lett. 12, 395–397 (2008)
Sandalidis, H.G., Tsiftsis, T.A.: Outage probability and ergodic capacity of free-space optical links over
strong turbulence. Electron. Lett. 44, 46–47 (2008)
Sandalidis, H.G., Tsiftsis, T.A., Karagiannidis, G.K., Uysal, M.: BER performance of FSO links over strong
atmospheric turbulence channels with pointing errors. IEEE Commun. Lett. 12, 44–46 (2008)
Yang, F., Cheng, J., Tsiftsis, T.A.: Free-space optical communication with nonzero boresight pointing
errors. IEEE Trans. Commun. 62, 713–725 (2014)
Zhu, X., Kahn, J.M.: Free-space optical communication through atmospheric turbulence channels. IEEE
Trans. Commun. 50, 1293–1300 (2002)

123

You might also like