You are on page 1of 6

Underwater optical communication performance

for laser beam propagation through


weak oceanic turbulence

Xiang Yi,* Zan Li, and Zengji Liu


State Key Laboratory of Integrated Service Networks, Xidian University, Xian 710071, China
*Corresponding author: yixiang@xidian.edu.cn

Received 11 November 2014; accepted 23 December 2014;


posted 6 January 2015 (Doc. ID 226645); published 12 February 2015

In clean ocean water, the performance of a underwater optical communication system is limited mainly
by oceanic turbulence, which is defined as the fluctuations in the index of refraction resulting from
temperature and salinity fluctuations. In this paper, using the refractive index spectrum of oceanic
turbulence under weak turbulence conditions, we carry out, for a horizontally propagating plane wave
and spherical wave, analysis of the aperture-averaged scintillation index, the associated probability of
fade, mean signal-to-noise ratio, and mean bit error rate. Our theoretical results show that for various
values of the rate of dissipation of mean squared temperature and the temperature–salinity balance
parameter, the large-aperture receiver leads to a remarkable decrease of scintillation and consequently
a significant improvement on the system performance. Such an effect is more noticeable in the plane
wave case than in the spherical wave case. © 2015 Optical Society of America
OCIS codes: (010.0010) Atmospheric and oceanic optics; (010.4455) Oceanic propagation; (290.5930)
Scintillation; (010.7060) Turbulence.
http://dx.doi.org/10.1364/AO.54.001273

1. Introduction broadband wireless communications on land, they


There is an increasing interest in exploring the cannot propagate though seawater. In this context,
oceans for scientific, environmental, commercial, underwater optical communication is moving to
and military purposes, and communication among the forefront of research.
various underwater vehicles and sensors is essential. Most of the previous investigations into the perfor-
The transmission of high-resolution video and mance of underwater optical communications were
images between underwater vehicles requires high- concerned with the absorption and scattering effects
data-rate wireless communication links. The amount on optical propagation in different ocean water types
of data collected and stored by underwater sensors [1–6]. In clean ocean water, the transmission is
can be large and should be expected to grow with optimum in the blue–green spectral region. The cor-
the development of underwater sensor networks. responding absorption and scattering coefficients
Although underwater acoustic communication is can approach 0.1 m−1 [1,3]. The calculations of power
well-established and practical in many cases, it budget performed in these environments show that
becomes increasingly difficult to support high-rate, reliable underwater optical communication at high
large-capacity applications due to serious absorption data rates can be achieved as the link distance
and multipath effects at higher carrier frequency. increases from several tens of meters to even hun-
As for radio waves, which are extensively used for dreds of meters [1,2]. However, turbulence effects
are neglected in these studies.
As in the atmosphere, optical turbulence, i.e.,
1559-128X/15/061273-06$15.00/0 random fluctuations in the index of refraction, can
© 2015 Optical Society of America appear in the ocean [7]. It is well-known that

20 February 2015 / Vol. 54, No. 6 / APPLIED OPTICS 1273


atmospheric turbulence resulting from temperature where
and humidity fluctuations causes a variety of effects
on an optical wave related to its irradiance fluctua- AT  C0 C−2
1 DT v
−1
AS  C0 C−2 −1
1 DS v ; (2a)
tions (scintillation) and phase fluctuations [8,9].
Scintillation has an important consequence on laser ATS  0.5C0 C−2 −1
1 DT  DS v ; (2b)
communication. Specifically, scintillation can lead to
fading of the received signal below a detectable 3
threshold and eventually to a large performance pen- δ  C21 κη4∕3  C31 κη2 ; (2c)
alty [10–13]. Oceanic turbulence which is controlled 2
by the salinity and temperature fluctuations [14,15] where κ is the magnitude of the spatial frequency,
may have similar influence on underwater optical α  2.6 × 10−4 L∕ deg is a constant, C0  0.72 is the
communications. In clean ocean where the adsorp- Obukhov–Corrsin constant, C1 ≈ 2.35 is the nondi-
tion and scattering effects are not significant, it is mensional constant determined by comparison with
necessary to understand how oceanic turbulence experiment data, the kinematic viscosity v in water
might affect the performance of the underwater medium is much higher than the molecular thermal
optical communication system. diffusivity DT (v∕DT  7) and more so than the
In recent years there have been several studies molecular salinity transport coefficient DS (v∕DS 
that have investigated oceanic turbulence-induced 700) [14], the Kolmogorov microscale η  v3 ∕ε1∕4
scintillation and the associated bit error rate with ε being the rate of dissipation of turbulent kinetic
(BER) performance [16–18]. In terms of Rytov theory, energy per unit mass of fluid, and χ T is the rate of dis-
numerically evaluated results were presented in sipation of the mean squared temperature altering
[16,17] for the scintillation index (SI) of a horizon- from 10−2 to 10−10 K 2 ∕s (surface water and deep water,
tally propagating plane wave and spherical wave respectively) [7]. The quantity ω varies between −5
that are valid under weak irradiance fluctuations. and 0, defining the contributions of the temperature
The results for the longitudinal and radial scintilla- and salinity distributions to the distribution of the re-
tion associated with the more general Gaussian- fraction index. As ω → −5, the temperature fluctua-
beam wave were also involved in [16]. Very recently tions dominate the refraction index fluctuations,
in [18], the longitudinal component of the SI was whereas at ω → 0, the refraction index fluctuations
formulated for a focused Gaussian beam in weak are mainly determined by the salinity fluctuations.
oceanic turbulence. By using this formulation and
the lognormal distribution of the received irradiance, 3. Aperture-Averaged Scintillation and the Associated
the mean BER was calculated. It has been shown System Performance
that the SI in weak oceanic turbulence increases rap- By means of the ABCD matrix approach illustrated
idly beyond unity within tens of meters, which is in Chapter 6 of [9], the irradiance flux variances over
much shorter than that in atmosphere. the circular aperture for the case of plane wave and
To extend communications distance in the pres- spherical wave under weak turbulence conditions are
ence of oceanic turbulence, one fading-mitigation developed, respectively, by
technique (for example, aperture averaging) should
be intentionally used in underwater optical commu- Z 1Z  
∞ D2 κ 2
nication systems. In this work, based on the refrac- σ 2I;pl D  8π 2 k2 L κΦn κ exp −
tive index spectrum of oceanic turbulence, we first 0 0 16
  2 
formulate the SI over the circular aperture for the Lκ ξ
case of a horizontally propagating plane wave and × 1 − cos dκdξ; (3)
k
spherical wave that are valid under weak irradiance
fluctuations. Then, we investigate the associated Z 1Z  
∞ D2 κ 2 ξ2
probability of fade, mean signal-to-noise ratio σ 2I;sp D  8π 2 k2 L κΦn κ exp −
(SNR), and mean BER. Finally, we present some 0 0 16
numerical results to illustrate the aperture-   2 

averaging effects on the SI and the associated system × 1 − cos ξ1 − ξ dκdξ; (4)
k
performance criteria.
2. Refractive Index Spectrum of Oceanic Turbulence where k  2π∕λ is the wave number at the wave-
length λ, L is the path length, ξ is the normalized
The spatial power spectrum of refractive index fluc-
path length, and D is the diameter of the circular
tuations in locally homogeneous and isotropic oce-
receive aperture.
anic turbulence at the stable stratification (where
After substituting Eq. (1) into Eq. (3) and expressing
the thermal diffusivity K T and the diffusion of the
the cosine function in Eq. (3) in terms of the exponen-
salt K S are equal to each other) is given by [15]
tial function with imaginary argument, we obtain
Φn κ  4π−1 C0 α2 ε−1∕3 κ−11∕3 1  C1 κη2∕3 
σ 2I;pl D  2πC0 α2 χ T ω−2 ε−1∕3 k2 L
χ
× T2 ω2 e−AT δ  e−AS δ − 2ωe−ATS δ ; (1) × ω2 I 1;pl  I 2;pl − 2ωI 3;pl ; (5)
ω

1274 APPLIED OPTICS / Vol. 54, No. 6 / 20 February 2015


where For the spherical wave case, the substitution of
Z1 Z∞ Eq. (1) into Eq. (4) leads to
I i;pl  Re dξ dκκ −8∕3 exp−ai κ4∕3 
0 0 σ 2I;sp D  2πC0 α2 χ T ω−2 ε−1∕3 k2 L
    
D2 2 × ω2 I 1;sp  I 2;sp − 2ωI 3;sp ;
× exp − bi  κ (9)
16
    
D2 2 j16Lξ where
− exp − bi  κ 1
16 16kbi  kD2 Z1 Z∞
Z1 Z∞
 C1 η2∕3 Re dξ dκκ−2 exp−ai κ4∕3  I i;sp  Re dξ dκκ −8∕3 exp−ai κ4∕3 
0 0
  
0 0
      
D2 2 D2 ξ2 2
× exp − bi  κ × exp − bi  κ
16 16
       
D2 2 j16Lξ D2 ξ2 2
− exp − bi  κ 1 ; (6) − exp − bi  κ
16 16kbi  kD2 16
 
j16L
where ai  1.5Ai C21 η4∕3 ; bi  Ai C31 η2 ; i  1, 2, 3 de- × 1 ξ1 − ξ
notes previously pmentioned subscripts T, S, and TS, kbi  kD2 ξ2
 Z1 Z∞
respectively; j  −1; and Rex denotes the real part
 C1 η2∕3 Re dξ dκκ−2 exp−ai κ4∕3 
of x. By expanding the function exp−ai κ4∕3  in a Ma- 0 0
claurin series and performing the inside integral with     
D2 ξ2 2
respect to κ, we derive that × exp − bi  κ
16
 5−4n   
1X ∞
Γ2n∕3 − 5∕6−ai n D2 6 D2 ξ2 2
I i;pl  bi  − exp − bi  κ
2 n0 n! 16 16
Z 1  5−4n   
j16Lξ j16L
× Re 1− 1
6
dξ × 1 ξ1 − ξ :
16kb  kD 2 kbi  kD2 ξ2
0 i
 3−4n
C η2∕3 X∞
Γ2n∕3 − 1∕2−ai n D2 6 Following a similar procedure as has been used in the
 1 bi 
2 n0 n! 16 plane wave case, we deduce that
Z 1  3−4n 
j16Lξ
1X
6 ∞
× Re 1− 1 2
dξ: (7) Γ2n∕3 − 5∕6−ai n
0 16kb i  kD I i;sp 
2 n0 n!
Then, by performing the integral over ξ, we obtain   
5 2n
− 2n 5 1 3 D2
 5 X   2n × b6i2 3 F 1 − ; ; ;−
L 6 ∞ Γ2n∕3 − 5∕6−ai n L − 3 3 6 2 2 16bi
I i;pl  3 Z 1 5−2n 
k n0 11 − 4nn! k D2 2 jL 6 3

   − Re bi  ξ  ξ1 − ξ dξ
11 − 4n 16kbi  kD2 6
5−4n
0 16 k
×
6 16L C η2∕3 X ∞
Γ2n∕3 − 1∕2−ai n
 1
   11−4n 2 n0 n!
16kbi  kD2 2 12
− 1   
16L 1 2n
− 2n 1 1 3 D2
   × b2i2 3 F 1 − ; ; ;−
11 − 4n 16L 3 2 2 2 16bi
× sin tan−1 Z 1 1−2n 
6 16kbi  kD2 D2 2 jL 2 3
 1 X − Re bi  ξ  ξ1 − ξ dξ ; (10)
L 2 ∞ Γ2n∕3 − 1∕2−ai n 0 16 k
 3C1 η2∕3
k n0 9 − 4nn!
 −2n   3−4n where 2 F 1 is the Gauss hypergeometric function.
L 3 9 − 4n 16kbi  kD2 6 The probability density function (PDF) of the ran-
× domly fading irradiance signal often used under
k 6 16L
  2
2 9−4n weak irradiance fluctuations is the lognormal model
16kbi  kD 12
as per Eq. (11.22) of [8]
− 1
16L  
   1 ln u  0.5σ 2I D2
9 − 4n 16L pI u  p exp − ;u > 0:
× sin tan −1
: (8) uσ I D 2π 2σ 2I D
6 16kbi  kD2
(11)

20 February 2015 / Vol. 54, No. 6 / APPLIED OPTICS 1275


The probability of fade describes the percentage of
time the irradiance of the received wave is below
some prescribed threshold value I T . Hence, the
probability of fade as a function of threshold
level is defined
R by the cumulative probability
PrI ≤ I T   0IT pI udu. For the lognormal PDF
[Eq. (11)], the resulting probability of fade leads to
  
1 0.5σ 2I D − 0.23F T
PrI ≤ I T   1  erf p ; (12)
2 2σ I D

where erf x is the error function; and the fade Fig. 1. Irradiance flux variances over circular apertures versus
parameter F T, given in dB, represents the decibel path length and various aperture diameters. Assumed parameters
(dB) level below the on-axis mean irradiance that are χ T  10−5 K 2 ∕s and ω  −2.5.
the threshold I T is set.
The mean SNR at the output of the detector in the
case of a shot-noise-limited systems assumes the increasing receiver collecting lens can take place
form as per Eq. (11.48) of [8] even for relatively small aperture. Specifically, for
the case of the plane wave, the irradiance flux vari-
SNR0 ance at the point receiver increases beyond unity at
hSNRi  q ; (13) the path length of 11 m, whereas at the same path
1  σ 2I DSNR20 length, the irradiance flux variance associated with
the small aperture of 1 cm approaches zero. From
where SNR0 is the SNR in the absence of turbulence. another point of view, we note that the increase in
In the presence of optical turbulence, the probabil- receive apertures can extend the reliable communi-
ity of error is considered a conditional probability cation distance of underwater optical communica-
that must be averaged over the PDF of the random tions systems. As observed, the irradiance flux
signal to determine the unconditional mean BER. In variances for D  0 reach unity at the path length
terms of a normalized signal with unit mean, this of 11 m for the plane wave case and 20 m for the
leads to the expression as per Eq. (11.58) of [8] spherical wave case. This distance extends, respec-
tively, to 120 and 106 m for D  5 cm, which also in-
Z  
1 ∞ hSNRiu dicates that the aperture-averaging effects are more
hBERi  pI uerfc p du; (14) remarkable for plane waves than for spherical waves.
2 0 2 2 Figure 2 provides the irradiance flux variances
where pI u is taken to be the lognormal distribution versus the aperture diameters for various rates of
[Eq. (11)] with unit mean. Equation (14) can be dissipation of mean squared temperature χ T . For
efficiently computed by use of Gauss–Hermite quad- optical waves in the blue–green spectral region,
rature rules [19]. the absorption and scattering coefficient in oceanic
medium are small enough to support a communica-
4. Results and Discussion tion link up to a distance of 100 m [17]. So we set
Next we first show some numerical results of the L  100 m. From Fig. 2 we see that for different val-
analytical irradiance flux variances over circular ues of χ T , the irradiance flux variances first drop off
apertures for the plane wave and spherical wave sharply with increasing values of aperture diame-
cases. To perform the numerical calculations, we ters, and then gradually decrease, approaching zero
set η  10−3 m, ε  10−5 m2 ∕s3 , and λ  0.417 μm as the aperture diameters increase without bound.
which is in the low-absorption and low-scattering If the irradiance flux variance is close to zero, the
range of the ocean. We restrict the values of the irra-
diance flux variance smaller than unity, thus our
results are valid in the weak turbulence regime.
In Fig. 1 we show the irradiance flux variances of a
plane wave [Eq. (5)] and a spherical wave [Eq. (9)] as
a function of the path length and several values of
aperture diameters. In computing the infinite series
in Eqs. (8) and (11), we choose n to be 10 since the
series converges rapidly for typical values of the
involved parameters. D  0 corresponds to a point
receiver. The aperture-averaged receivers with
finite-size receive apertures (D  1 and 5 cm) are
compared to the point receiver as a benchmark. As
excepted, it is observed that the decrease in oceanic Fig. 2. Irradiance flux variances versus aperture diameters and
turbulence-induced scintillation associated with various χ T . Assumed parameters are ω  −2.5 and L  100 m.

1276 APPLIED OPTICS / Vol. 54, No. 6 / 20 February 2015


increases with the value of ω increases. This is
because salinity-driven turbulence induces more sig-
nificant scintillation than temperature-driven turbu-
lence. Also, large χ T leads to increased probability of
fade over that caused by small χ T . We note that for a
large receive aperture of 10 cm, the probability of
fade corresponding to spherical waves is higher than
that of plane waves. This is a consequence of the
behavior in the associated irradiance flux variances
as a function of aperture diameters. For a pointer
receiver, the scintillation of plane waves is larger
than the scintillation of spherical waves, but the
Fig. 3. Irradiance flux variances versus aperture diameters and situation is opposite when the receiver aperture is
various ω. Assumed parameters are χ T  10−6 K 2 ∕s and large enough. This is consistent with the observation
L  100 m.
presented in atmospheric turbulence [10].
In Fig. 5 we plot the mean SNR [Eq. (14)] in dB as a
influence of oceanic turbulence on underwater opti- function of ω for various χ T . SNR0  10 dB. We note
cal communication systems can be ignored. that there is a penalty on the mean SNR with respect
In Fig. 3 the curves for the irradiance flux varian- to SNR0 . Such a penalty becomes large with the
ces associated with different temperature–salinity increasing values of ω and χ T . Under the same oce-
balance parameter ω are plotted against aperture anic conditions, the spherical wave suffers the more
diameters. L is also set to 100 m. Similarly to the re- serious penalty in SNR as compared to the plane
sults illustrated in Fig. 3, we see that the irradiance wave. The reason for this phenomenon is due to
flux variance decreases drastically first and then the behavior of irradiance flux variances over large
moves toward zero with the increases of aperture apertures reported previously.
diameters. We also observe that the reduction in Numerical calculation of the mean BER [Eq. (14)]
temperature-induced scintillation requires much against the mean SNR for various aperture diame-
smaller receive aperture than does the decrease of ters and path lengths leads to the results shown in
salinity-induced scintillation. For example in the Fig. 6. These results show that to achieve an accept-
spherical wave case, to achieve specific irradiance able level of mean BER (typically around 10−6 ) at a
flux variance of 0.4, the 2 cm aperture is large enough communication distance of hundreds of meters in the
for ω  −5, whereas the 5 cm aperture is required for
ω  −0.5. Besides, we note that at a fixed value of χ T ,
the differences between the curves of the aperture-
averaged irradiance flux variances corresponding
to ω  −5 and ω  −2.5 are not significant, but are
obvious between those of ω  −2.5 and ω  −0.1.
This suggests that at a specific depth of ocean, the
receive aperture should be carefully chosen espe-
cially for salinity-induced turbulence.
Figure 4 illustrates the probability of fade
[Eq. (13)] as the variation of the temperature–
salinity balance parameter ω. The irradiance flux
variances in Eq. (13) are obtained by setting D 
10 cm and L  100 m. The fade threshold parameter Fig. 5. Mean SNR versus ω for different χ T . Assumed parameters
are D  10 cm, L  100 m, and SNR0  10 dB.
F T  3 dB. We can see that the probability of fade

Fig. 4. Probability of fade versus ω for different χ T . Assumed Fig. 6. Mean BER for on–off keying direct detection as a function
parameters are D  10 cm, L  100 m, and F T  3 dB. of mean SNR and various D and L. χ T  10−6 K 2 ∕s and ω  −2.5.

20 February 2015 / Vol. 54, No. 6 / APPLIED OPTICS 1277


presence of oceanic turbulence, it will be necessary to 3. F. Hanson and S. Radic, “High bandwidth underwater optical
communication,” Appl. Opt. 47, 277–283 (2008).
utilize a large-aperture collecting lens. 4. S. Arnon, “Underwater optical wireless communication net-
5. Conclusion work,” Opt. Eng. 49, 015001 (2010).
5. P. V. Kumar, S. S. K. Praneeth, and R. B. Narender, “Analysis
In summary, we have studied the aperture-averaged of optical wireless communication for underwater wireless
scintillations of plane and spherical waves that hori- communication,” Int. J. Sci. Eng. Res. 2, 1–9 (2011).
zontally propagate through weak oceanic turbulence. 6. L. J. Johnson, R. J. Green, and M. S. Leeson, “Underwater op-
tical wireless communications: depth dependent variations in
By use of the ABCD matrix method, the exact ana- attenuation,” Appl. Opt. 52, 7867–7873 (2013).
lytic solutions to the irradiance flux variances over 7. S. A. Thorpe, The Turbulent Ocean (Cambridge University,
a circular aperture are obtained after a rigorous 2005).
mathematical derivation. With the newly developed 8. L. C. Andrews and R. L. Phillips, Laser Beam Propagation
through Random Media, 2nd ed. (SPIE, 2005).
expressions, the reduction in oceanic-induced scintil- 9. L. C. Andrews, R. L. Phillips, and C. Y. Hopen, Laser Beam
lation as the result of aperture averaging and the Scintillation with Applications (SPIE, 2001).
corresponding improvements in the associated 10. I. Toselli, L. C. Andrews, and R. L. Phillips, “Free space
probability of fade, mean SNR, and mean BER are optical system performance for laser beam propagation
through non Kolmogorov turbulence,” Proc. SPIE 6457,
analyzed in depth for different values of the path 64570T (2007).
length, the rate of dissipation of the mean squared 11. I. Toselli, L. C. Andrews, R. L. Phillips, and V. Ferrero, “Free
temperature, and the temperature–salinity balance space optical system performance for laser beam propagation
parameter. It has been shown that aperture- through non Kolmogorov turbulence for uplink and downlink
averaging effects in a turbulent ocean environment paths,” Proc. SPIE 6708, 670803 (2007).
12. I. Toselli, L. C. Andrews, R. L. Phillips, and V. Ferrero,
are significant even for relatively small receive aper- “Free space optical system performance for a Gaussian beam
tures. For large-aperture receivers, a plane wave is propagating through non-Kolmogorov weak turbulence,”
preferred with respect to a spherical wave. The IEEE Trans. Antennas Propag. 57, 1783–1788 (2009).
adoption of the aperture-averaging technique in an 13. X. Yi, Z. J. Liu, and P. Yue, “Optical scintillations and fade sta-
tistics for FSO communications through moderate-to-strong
underwater optical communication system can
non-Kolmogorov turbulence,” Opt. Laser Technol. 47,
observably extend its reliable communication dis- 199–207 (2013).
tance. The observations presented in this paper 14. R. J. Hill, “Optical propagation in turbulent water,” J. Opt.
can be helpful in the design of underwater optical Soc. Am. 68, 1067–1072 (1978).
communication systems. 15. V. V. Nikishov and V. I. Nikishov, “Spectrum of turbulent fluc-
tuations of the sea-water refraction index,” Int. J. Fluid Mech.
This work was supported by the New Teacher Res. 27, 82–98 (2000).
Innovation Fund of Xidian University (Grant 16. O. Korotkova, N. Farwell, and E. Shchepakina, “Light scintil-
lation in oceanic turbulence,” Waves Random Complex Media
No. K5051399067). 22, 260–266 (2012).
17. Y. Ata and Y. Baykal, “Scintillations of optical plane and
References spherical waves in underwater turbulence,” J. Opt. Soc.
1. M. A. Chancey, “Short range underwater optical communication Am. A 31, 1552–1556 (2014).
links,” M.S. dissertation (North Carolina State University, 18. H. Gercekcioglu, “Bit error rate of focused Gaussian beams in
2005). weak oceanic turbulence,” J. Opt. Soc. Am. A 31, 1963–1968
2. J. W. Giles and I. N. Bankman, “Underwater optical commu- (2014).
nications systems. Part 2: basic design considerations,” in 19. M. K. Simon and M. S. Alouini, Digital Communication over
Proceedings of MILCOM (IEEE, 2005), pp. 1700–1705. Fading Channels (Wiley, 2000).

1278 APPLIED OPTICS / Vol. 54, No. 6 / 20 February 2015

You might also like