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

Error performance of deep space optical communication with M-ary pulse position modulation over coronal turbulence channels

Open Access Open Access

Abstract

Free space optical (FSO) communication is a promising technology for future deep space exploration, but it encounters coronal turbulence during superior solar conjunction. In this paper, the bit-error rate (BER) performance of deep space FSO communication systems is evaluated for optical waves propagating in the non-Kolmogorov coronal turbulence. By virtue of its high energy efficiency, the pulse position modulation (PPM) technique is adopted to mitigate the influence of turbulence under lognormal distribution channels. The effects of the parameters of the coronal turbulence and of the FSO system, such as the turbulence outer scale, spectral index, symbol number, data bit rate, equivalent load resistor, and average gain, on the BER are investigated and discussed in this paper. In addition, the performance improvement from M-ary PPM and binary phase-shift keying subcarrier intensity modulation is studied. The results of a numerical evaluation illustrate that a deep space FSO communication system with PPM scheme can well be used to mitigate the impact of coronal turbulence.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Due to its advantages of large transmission capacity, narrow beam size, high transmission rate, cost-effectiveness, etc., free space optical (FSO) communication has become a promising scheme and a technology complementary to radio frequency (RF) communication [1, 2]. Recently, FSO communication has exhibited its superiority in many scenarios, such as oceanic communication [3, 4], atmospheric communication [5, 6], and deep space communication [2, 7], and has gained plenty of attention. However, the performance of FSO communication systems is susceptible to degradation by the scintillation effect occurring when the optical waves pass through turbulent channels. The intensity fluctuations of the received signal caused by the adverse turbulence condition further induce a large bit-error rate (BER). Therefore, the FSO communication performance is severely degraded and the communication link can even be interrupted.

Much attention has been devoted in recent years to investigate the BER of FSO communication systems [8–11]. On the one hand, plenty of statistical distribution models have been introduced to understand the turbulence-induced fading process in FSO communication system. The lognormal (LN) distribution [8], Gamma-Gamma (GG) distribution [9], and exponential Weibull (EW) distribution [10] are the most extensively adopted fading models for characterizing the turbulence channels under weak scintillation, weak-to-moderate scintillation, and weak-to-strong scintillation, respectively. On the other hand, various modulation technologies have been used to reduce the influence of the turbulence fading effect on FSO communication systems and to achieve a high transmission rate and excellent BER [12, 13].

An expression of the link BER for FSO systems with on–off keying (OOK) modulation over atmospheric turbulence has been proposed in [8] and was further validated with the experimental data. Considering the effect GG distributed atmospheric turbulence on satellite-to-ground optical communication, the BER with binary phase-shift keying (BPSK) modulation was studied in [9]. In [10, 13–15], the BER of both atmospheric and underwater FSO systems with subcarrier intensity modulation (SIM) under LN, GG, and EW distributed channels were compared and it was demonstrated that the EW distribution has an advantage in characterizing the turbulence intensity for weak-to-strong scintillation.

After that, the BER of optical waves in EW fading channels with OOK and BPSK was further investigated comprehensively in [10]. Due to its simplicity, OOK has been the most studied technique for FSO communication in much of the literature. In addition, the BPSK modulation scheme has also had a wide range of application to FSO systems since it does without an adaptive threshold. However, the lower power efficiency of these modulation schemes restrains their further development. In [16], pulse position modulation (PPM) with high power efficiency was first proposed and it has received much attention. Since PPM offers a high power efficiency at the expense of increased bandwidth, some modified PPM schemes have been proposed. The M-ary PPM techniques are used in many communication system to transmit M different signals in one time slot by non-coherent detection at the receiver [17]. Up to now, several ideas for superimposing different coding and modulation techniques, such as multiple PPM (MPPM)-coded multiple-input multiple-output FSO systems and hybrid M-ary quadrature amplitude modulation, have been proposed to improve both the power efficiency and spectral efficiency simultaneously [18, 19].

As we have mentioned above, many researchers have studied the link performance of optical waves in both atmospheric and oceanic turbulence with different modulation techniques and have achieved abundant meaningful results. However, as far as our knowledge extends, the BER of FSO communication in deep space has not been studied comprehensively. With the development of deep space exploration, FSO communication is an alternative scheme to overcome the long transmission distance [2]. Nevertheless, the link will inevitably pass through the turbulent solar corona during superior solar conjunction when the probe moves on the opposite side of the Sun from to the Earth [20]. Therefore, the BER of optical waves propagating in coronal turbulence is a challenging problem.

In contrast to atmospheric and oceanic turbulence, the solar corona is filled with high density plasma that erupts from the Sun and further expands into interplanetary space (IPS). According to the mounting evidence of astronomical observation, the statistics of these plasma irregularities obey non-Kolmogorov’s spectrum model [21, 22]. Since the refractive index of the solar wind plasma changes, the signal intensity will be distorted when the radio frequency (RF) and optical signal pass through the turbulent coronal regions [20, 24].

Motivated by the above analysis, in this paper, we formulate and study the BER of optical waves propagating through coronal turbulence. The PPM technique is adopted since it is predominantly used for deep space FSO communication, due to its high power efficiency and lack of an adaptive threshold. We investigate the effect on the average BER of different coronal parameters, such as the turbulent outer scale, the non-Kolmogorov spectral index, and FSO system parameters, such as the symbol number, bit rate, equivalent load resistor, and average gain.

In addition, a comparison of the performance of M-ary PPM and BPSK modulation in deep space FSO communication system is made. Numerical evaluation are helpful for understand how coronal turbulence will affect FSO communication, and will also provide a theoretical basis for the design of future deep space FSO links.

The rest of this paper is organized as follows. In Section 2, the formulation of solar scintillation is introduced. After that, the BER with PPM and M-ary PPM techniques under coronal channels with LN distribution is derived. Numerical results are discussed in Section 3, and the conclusions are summarized in Section 4.

2. System model and formulation

2.1. System model and the formulation of solar scintillation

A diagram for a deep space FSO communication system during superior solar conjunction is given in Fig. 1. The relay communications technique is adopted here, where a relay satellite around the Earth first receives the signal from the Earth station via RF and then transmits it by an optical wave. Here, we assume that the photo–detectors are located in both the receiver and the transmitter of the relay satellite and the probe. The distance between the Earth and the Sun is Lse and the distance between the probe and the Sun is Lsp. The distance between the surface of the Sun and the communication link is denoted by r, which is generally in a unit of the solar radius, Rsun. Therefore, both the Sun–Earth–Probe (SEP) angle, α, and the Sun–Probe–Earth (SPE) angle, β, have a direct relation with r, since α=arcsin (r/Lse) and β=arcsin (r/Lsp), as shown in Fig. 1.

 figure: Fig. 1

Fig. 1 Deep space FSO communication between the Earth and probe during superior solar conjunction.

Download Full Size | PDF

As mentioned in Section I, the solar wind is a turbulent plasma erupting from the Sun and ultimately pervading in the entire IPS. The solar wind density and its irregularities are very high in the chromosphere and decrease along its radial direction, as is intuitive. Therefore, the optical waves transmitted between the Earth station and the probe will suffer severe fluctuations due to the irregularities in the solar wind when the FSO communication link passes through those coronal irregularities with strong turbulent channels. These fluctuations are generally referred to as solar scintillation [23]. Consequently, the performance of the link is degraded during this period, as depicted in Fig. 1. Note that the solar wind plasma permeates the whole IPS and its density decreases along its radial direction. Besides, the system model of FSO communication during superior solar conjunction has a short duration but will frequently occurrence. Therefore, we mainly consider the influence of the coronal turbulence on the optical waves that propagate along the transmission link.

According to [7, 24], the dielectric constant of the solar wind plasma, as a direct result, the amplitude fluctuations, is related to the plasma density by

δε(r)=reλ2πδNe(r).
where re is the classical electron radius and λ is the wavelength. Note that the amplitude fluctuations are only related to λ and the fluctuation of the density of the solar wind, δNe. Therefore, research on the influence of ionospheric plasma on RF signals provides a potential way to investigate the impact of coronal turbulence on optical waves [24].

The performance level of deep space FSO system during superior solar conjunction due to intensity fluctuations is usually characterized by the variance of amplitude fluctuations, χ2, and scintillation index. Apart from the impact of coronal turbulence, atmospheric turbulence also has a significant influence on the whole FSO communication system, which has been extensively studied, as we elaborated on in Section I. However, we mainly focus on the impact of coronal turbulence on the deep space FSO system in this paper, for simplicity.

An expression of the amplitude fluctuations under weak solar scintillation has been successfully derived with Rytov’s approximation method in our previous research, and was introduced in [7].

χ2=16π4re 2k20κdκ0Lsin 2(zκ22k)ΦN(κ,z)dz.
where k=2π/λ denotes the wave number.

Here, the amplitude fluctuations are the double integral over the magnitude of the spatial frequency vector, κ, and the optical waves propagation distance, z, since the coronal turbulence extends over the whole link. L represents the length of the FSO communication link, which is the distance between the relay satellite around the Earth and the probe. ΦN(κ,z) is the spatial spectrum of the coronal turbulence. In the following description, the spectrum model is simplified as ΦN(κ) since z can be represented by r according to Fig. 1. Here, a non-Kolmogorov spectrum model is adopted [25].

ΦN(κ)=(2π)3/2(p3)Γ(p/2)Γ[(p1)/2]κop3δNe2κp,

Here, p represents the spectral index fluctuation in the range of 3 to 4 [26], Γ() is the Gamma function, κo=2π/Lo, and Lo is the outer scale of the coronal turbulence. δNe denotes the irregularities of the solar wind density, and means the ensemble average. Note that δNe is normally assumed proportional to the solar wind plasma density, Ne, i.e. δNeηNe, which has been extensively employed in astronomy. Here, η represents the relative solar wind density fluctuations coefficient. Since we mainly consider the effect of coronal turbulence on the optical waves during superior solar conjunction, the following simplified, but widely used, solar wind density model for 1.5Rsunr2Rsun is adopted in this paper [24]. Note that Rsun is the solar radius.

Ne(r)=4×1014(Rsunr)10+3×1014(Rsunr)6

By substituting (3) into (2), a closed form formula for the amplitude fluctuations can be obtained as

χ2=(p3)Γ(p/2)re 2(2π)11/2pπ8Γ[(p1)/2]Γ(1+p/2)η2Ne 2Lo 3pLp2kp21sec (πp/4).

According to [28], the relation between the scintillation index, m2, which is characterized as the standard deviation of the intensity of the waves normalized to the average received intensity, and the variance of the amplitude fluctuations can be written as m2=4χ2.

2.2. BER under PPM and M-ary PPM

The link performance of deep space FSO communication systems during superior solar conjunction is affected by the amplitude fluctuations resulting from turbulent irregularities in the solar wind. Theoretical and experimental research into the propagation of waves in coronal turbulence demonstrate that the communication link will encounter weak turbulence, moderate turbulence, and strong turbulence, successively, with a decrease in the heliocentric distance or SEP angle [7]. Consequently, the intensity of the solar scintillation increases during this period. For the sake of simplicity, only weak solar scintillation is considered in this paper.

In this section, we consider an avalanche photodiode (APD)-based direct-detection receiver. The current induced in the load resistance in the response of the APD is modeled as

y(t)=qn=Gnx(ttn)+n(t)
where q is the charge of the electron (1.602×1019C), and Gn denotes the random gain of the APD for the n-th detected photon. The response of the APD to a photon in the n-th event time, tn, is represented by x(t). n(t) is the thermal noise at the receiver.

According to the derivation in [28], the average number of the received photon in a specific PPM slot duration can be calculated as

Ks=12ρλPrcTw
where ρ means the detector’s quantum efficiency. ℏ and c are the Planck’s constant and the light speed, respectively. Pr is the received average power for one PPM slot. The PPM symbol duration is denoted by Tw=2Ts. Ts is the slot duration for a specific PPM signal with the bit duration Tb=1/Rb, where Rb is the data bit rate.

Without loss of generality, we also assume Ks obeys the LN statistics distribution, which is defined as [27, 28]

f(Ks)=12πσk2Ksexp {[ln (Ks)nk]22σk2},

Here, σk and nk represent the standard deviation and mean value of ln(Ks). Therefore, the unconditional BER, Pb, in an FSO communication system with PPM can be written as [29]

Pb=0Q(Ξ(Ks))f(Ks)dKs,

Note that the Gaussian Q-function is expressed as Q(t)=1/2erfc(t/2) and erfc() is the complementary error function. Note that Ξ(Ks) denotes the signal-to-noise ratio (SNR) and can be further given by [30]

Ξ(Ks)=(Gq)2Ks 2(Gq)2F(Ks+2Kbg)+2σth2
where G denotes the average gain of the APD, which usually falls in the range of 50∼250. Note that σth 2=2σTeTs/RL, where σ is Boltzmann’s constant (1.38×1023J/K), Te and RL represent the effective temperature of the receiver in degree Kelvin and the equivalent load resistor, respectively.

It also should be noted that the noise sources in the FSO communication system can be divided as the shot noise associated to the received signal, the background optical power, the dark current noise, the thermal noise and the total equivalent noise input current associated to the amplifier. For simplicity, only the influence of background radiation noise is considered in (10). Kbg=ρPbgTsλ/(c) denotes the average photon count owing to the background radiation of power Pbg in each PPM slot. Note that Pbg is assumed as a constant value in the following calculation. The excess noise factor of APD in (10) can be expressed as F=2+Gς, where ς denotes the ionization factor.

Substituting (10) and (8) into (9) and using Hermite’s polynomials [30], the BER of the FSO communications system with PPM can be finally recast as

Pb1πi=1NwiQ{exp [2(2σkxi+nk)]Fexp (2σkxi+nk)+Kn},
where Kn=2[FKbg+σth 2/(Gq)2], and σk 2=ln (m2+1). Note that m2 is the scintillation index of the optical waves in coronal turbulence and can be obtained by (5). In addition, [xi]i=1N and [wi]i=1N denote the zero points and weight factors of the N-th order Hermite polynomial, respectively. It is imperative to note that the values of wi, xi, and Q() in (11) are well tabulated and can be easily found in [30]. Moreover, the value of N is chosen depending on the desired value of accuracy, which is taken as 20 in the subsequent calculations. Furthermore, the ensemble average of the photon count, Ks, has a direct relationship with the coronal turbulence as E[Ks]=exp (σk2/2+nk). Note that nk is the mean value of ln (Ks).

In an M-ary PPM system, each optical pulse is divided into M adjacent time slots, which means each PPM symbol slot has log 2M bits. Therefore, the slot duration for the M-ary PPM FSO system can be defined as

Ts=Tblog 2MM

Similarly, for a PPM system working in coronal turbulence conditions, the BER formula of the FSO system with M-ary PPM can be further expressed as

PbMM2πi=1nwiQ{exp [2(2mkxi+nk)]Fexp (2mkxi+nk)+Kn}.

3. Numerical evaluations and discussion

In this section, the BER of a deep space FSO communication system with PPM under coronal turbulence during superior solar conjunction is investigated considering the effect of employing different values for the system parameters. All results were obtained by numerically evaluating (11) and (13) with the Matlab software package. In all the Figs., the receiver temperature is fixed at Te = 100 K and we assume that the APD ionization factor ζ=0.028. The solar wind density fluctuations coefficient is set to be η=10%, which is reasonable in light of the astronomical observation [26]. It should be noted that the communication distance L can be obtained by employing the triangle relation as showing in Fig. 1. In the following calculations, all of the other parameters have been selected from some actual FSO communication systems with deep space missions, as shown in Table 1. The antenna radius for the transmitter and receiver are 0.6 m and 0.15 m.

Tables Icon

Table 1. Parameters used in the real FSO communication systems in deep space missions.

Figure 2 demonstrates the variation of BER as a function of the scintillation index, m2, for different symbol numbers, M. The modulation-related parameters are set to be: Ks = 300, RL=60Ω, G=150. According to deep space missions with optical wave communications [35], the data bit rate is normally lower than 30 Mb/s. However, we set it to be Rb = 50 Mb/s in the subsequent calculations since this is the target for future optical communication systems. Once can see from Fig. 2 that the BER increases with an increase in the scintillation index. This is because the solar scintillation intensity increases from weak to moderate, and even up to strong, when the scintillation index increases. Therefore, the BER decreases along this tendency. In addition, Fig. 2 also shows that a higher modulation order with a large symbol number induces a larger BER. The BER increases from 1×103 to 8.2×102 when M increases from 2 to 128 at m2=0.4. Note that this phenomenon is more obvious when m2 gets larger. These phenomena are reasonable because the optical wave propagation through the coronal turbulence by M-ary PPM with a larger symbol number means the communication system needs a larger transmitted bandwidth, which induces more noise power, and finally gives a larger BER.

 figure: Fig. 2

Fig. 2 BER versus scintillation index for different values of M.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 BER versus scintillation index for different values of G.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 BER versus scintillation index for different values of Ks.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 BER versus the wavelength, λ, for different values of the outer scale.

Download Full Size | PDF

In Fig. 3 and Fig. 4, the influence of the average photodetector gain, G, and the photon count, Ks, are analyzed. The simulation parameters in Fig. 3 were set to be: M = 8, Ks = 300, Rb = 50 Mb/s, RL=60 Ω, and the parameters in Fig. 4 are set as: M = 8, G=150, Rb = 50 Mb/s, RL=60 Ω. It can be seen in Fig. 3 that an increase in G causes a decrease in the BER, especially when G increases from 100 to 150. This can be attributed to the fact that the noise level increases along with an increase of average photodetector gain and finally contributes to the BER. The BER also decreases with an increase of Ks, as shown in Fig. 4. It should be noted that this tendency becomes negligible when the scintillation index increases. For example, the BER decreases from 3×102 to 5×103 when Ks increases from 180 to 260 at m2=0.4. However, it decreases from 7×102 to 2×102 at m2=0.8. This phenomenon can be directly explained from (13) in that a larger photon count dominates the BER degradation. Nevertheless, the scintillation index plays a dominant role in the BER, eclipsing that of the photon count when the optical waves are passing through moderate or even strong coronal turbulence.

As shown in Eq. (5), the coronal turbulence parameters, such as, the outer scale, Lo, spectral index, p, also have a large influence on the BER. Therefore, both the effect of the coronal turbulence parameters and of the FSO system parameters will be investigated, comprehensively, in the following simulations. The simulation parameters have been set as follows unless otherwise specified: M = 8, Ks = 300, G=150, Rb = 50 Mb/s, RL=60 Ω.

In Fig. 5, we illustrate the effect of the optical wavelength, λ, on the BER for different outer scales, Lo. Here, we set p=3.6. According to the long-time astronomical observations and analysis [26], the outer scale is set in the range of 5×107Lo2×108 m. As expected, decreasing the optical wavelength results in decreasing the BER for different Lo. In addition, the BER also decreases with an increase of Lo. The ability of Lo to reduce the BER is more obvious when λ is about 800 nm compared to λ1200 nm. Since we mainly consider the non-Kolmogorov spectrum model in this paper, the BER versus the wavelength for various p values is presented in Fig. 6. An outer scale of Lo=1.5×108 m was taken for this simulation. We can also observe that the BER increases with an increase of λ. In addition, a smaller p induces an higher BER.

The physical interpretation of the results shown in Fig. 5 and Fig. 6 is that, the integrated coronal turbulence strength decreases with an increase of outer scale and spectral index. Thus, it will finally degrade the value of the BER. Note that the decline of BER performance in Fig. 5 and Fig. 6 becomes more gradual with a decrease of outer scale and spectral index, respectively, when the wavelength is large. This phenomenon further indicates that the wavelength plays a vital role in the FSO communication system. Therefore, the communication performance can be significantly enhanced by decreasing the wavelength.

 figure: Fig. 6

Fig. 6 BER versus the wavelength, λ, for different values of the spectral index.

Download Full Size | PDF

Figure 7 and Fig. 8 show the BER versus the heliocentric distance, r, for different data bit rates, Rb, and equivalent load resistors, RL, respectively. The heliocentric distance is set to be 1.75Rsunr1.825Rsun, therefore, the corresponding SEP angle is approximately 0.467α0.487. In addition, we set M = 8, Ks = 300, G=150, Lo=5×107 m, p=3.5. The equivalent load resistor in Fig. 7 is RL=60 Ω and the data bit rate in Fig. 8 is Rb = 50 Mb/s. Note that the data bit rate in Fig. 7 has been chosen in the range of 5∼50 Mb/s according to NASA’s future deep space project [35].

 figure: Fig. 7

Fig. 7 BER versus the heliocentric distance, r, for different data bit rates.

Download Full Size | PDF

 figure: Fig. 8

Fig. 8 BER versus the heliocentric distance, r, for different equivalent load resistors.

Download Full Size | PDF

 figure: Fig. 9

Fig. 9 BER versus the SNR, γ, for different scintillation indices, m2, with 4PPM and BPSK.

Download Full Size | PDF

As shown in Fig. 7, the higher are the Rb, the larger the BER of the FSO communication system. As Rb increases, a larger transmitted bandwidth is needed and finally induces more noise power, according to Eq. (12). In Fig. 8, it is clear that the BER decreases with an increase in RL. This is because a larger load resistor will improve the effective sensitivity of the photodetector, and as a result, bring about a better BER performance in the FSO communication system. It is worth noting that, increasing the load resistor is an alternative way to enhance the BER performance. However, it is not practical, due to the bandwidth constraints.

In Fig. 7 and Fig. 8, the BER decreases with an increase in the heliocentric distance or SEP angle for different Rb and RL. Note that this phenomenon is more obvious for a large heliocentric distance. For example, the BER gradually decreases from 8.9×103 to 7.2×103 when Rb decreases from 50 Mb/s to 35 Mb/s at r=1.78Rsun. However, it decreases rapidly from 4.2×104 to 8.6×105 when r=1.8Rsun. This phenomenon is attributed to the fact that, the intensity of the coronal turbulence decreases with an increase in the heliocentric distance, and the scintillation index decreases at the same time. Therefore, the BER is reduced during this time, according to Eq. (13), along with the decrease in the bit rate and the increase in load resistor. In addition, we can conclude that the effect of solar wind irregularities on the propagation of the optical waves through the coronal turbulence can be negligible when r1.8Rsun with Rb15 Mb/s and RL60 Ω.

To evaluate the BER performance under the M-ary PPM scheme, it is desirable to compare the numerical simulation with theoretical calculations. For an FSO communication system, the BER for M-ary PPM and BPSK over an LN channel model are shown in Fig. 9 as a function of SNR, γ, and scintillation index, m2. We also set: M = 4, Ks = 300, G=150, RL=60 Ω, Lo=5×107 m, and p=3.5. For comparability, the symbol duration of the BPSK scheme is equal to the duration of the time slot of the 4PPM scheme. In addition, both schemes have the same data rate, Rb = 50 Mb/s. Therefore, the bandwidth for BPSK and 4PPM are 100 MHz and 200 MHz, respectively.

As shown in Fig. 9, the results of numerical simulation of the 4PPM scheme are fairly consistent with the theoretical analysis, under different scintillation indices. Therefore, it is reasonable to believe that our theoretical result is correct. It can be seen that the required SNR for 4PPM is 12 dB, whereas for the BPSK it is about 13 dB to obtain a standard BER of 105 at a scintillation index of m2=0.1. Similarly, at the higher scintillation index of m2=0.3, the required SNR for 4PPM is about 15.8 dB, and for BPSK, it is around 17.3 dB to maintain the same BER. Therefore, this demonstrates that the 4PPM avoids the power penalty of 1 dB and 1.5 dB in the lower scintillation index and higher index, respectively, in comparison to the BPSK scheme. Of course, with both modulation schemes, the BER decreases as the scintillation index decreases. This can also be concluded from Fig. 2, Fig. 3, and Fig. 4, and indicates that the scintillation intensity has an intimate relation with the BER performance of FSO communication systems. Moreover, the 4PPM scheme achieves a better BER than the traditional BPSK scheme irrespective of the coronal turbulence conditions.

According to the above discussion, the main contribution of this paper is to understand the influence of various parameters on the performance of a deep space FSO communication link. Inspired by the simulation results, the BER can be significantly improved by increasing the equivalent load resistance, the average gain of the APD and decreasing the data bit rate. In addition, decreasing the wavelength would be an extremely effective way to enhance the BER, but needs investigation due to the system design. Moreover, both M-ary PPM and BPSK modulation schemes are alternative techniques for deep space FSO communication systems but should be further studied due to the variable solar corona environment. It should be noted that our results are strongly dependent on the simplified solar wind density model and fluctuations model, as depicted in Section 2. However, the actual coronal environment is very complicated. Therefore, more attention should be paid to improving the forecast accuracy of the effect of the coronal environment on an FSO communication system.

4. Conclusion

In this paper, we investigated the bit-error rate (BER) of free space optical (FSO) communication systems operating in the non-Kolmogorov coronal turbulence channels with a lognomal distribution. The pulse position modulation (PPM) technique is used to mitigate the BER caused by irregular solar wind turbulence under weak scintillation. In addition, the influence of various coronal parameters and FSO system parameters on the BER has been discussed. The results show that BER decreases with an increase of the turbulence outer scale, of the spectral index, and the heliocentric distance. Furthermore, increasing equivalent load resistance, photodetector gain, and decreasing the symbol number or the data bit rate can remarkably decrease the BER to some extent. Moreover, the M-ary PPM scheme surpasses traditional binary phase-shift keying modulation in improving the BER of an FSO system in coronal turbulence during superior solar conjunction. Therefore, our results have potential value for the design of future deep space FSO communication systems.

Funding

National Natural Science Foundation (NSFC) (61801181, 61831008); Open Research Fund of Shanghai Key Laboratory of Multidimensional Information Processing.

References

1. M. A. Khalighi and M. Uysal, “Survey on Free Space Optical Communication: A Communication Theory Perspective,” IEEE Commun. Surv. Tutor. 16, 2231–2258 (2014). [CrossRef]  

2. H. Kaushal and G. Kaddoum, “Optical Communication in Space: Challenges and Mitigation Techniques,” IEEE Commun. Surv. Tutor. 19, 57–96 (2017). [CrossRef]  

3. Y. Zhao, A. Zhao, L. Zhu, W. Lv, J. Xu, S. Li, and J. Wang, “Performance evaluation of underwater optical communications using spatial modes subjected to bubbles and obstructions,” Opt. Lett. 42, 4699–4702 (2017). [CrossRef]   [PubMed]  

4. M. Sharifzadeh and M. Ahmadirad, “Performance analysis of underwater wireless optical communication systems over a wide range of optical turbulence,” Opt. Commun. 427, 609–616 (2018). [CrossRef]  

5. W. Lim, C. Yun, and K. Kim, “BER performance analysis of radio over free-space optical systems considering laser phase noise under Gamma-Gamma turbulence channels,” Opt. Express 17, 4479–4484 (2009). [CrossRef]   [PubMed]  

6. H. Liu, R. Liao, Z. Wei, Z. Hou, and Y. Qiao, “BER Analysis of a Hybrid Modulation Scheme Based on PPM and MSK Subcarrier Intensity Modulation,” IEEE Photonics J. 7, 7201510 (2015). [CrossRef]  

7. G. Xu and Z. Song, “Amplitude fluctuations for optical waves propagation through non-Kolmogorov coronal solar wind turbulence channels,” Opt. Express 26, 8566–8580 (2018). [CrossRef]   [PubMed]  

8. F. Li, Z. Hou, and Y. Wu, “Experiment and numerical evaluation of bit error rate for free-space communication in turbulent atmosphere,” Opt. Laser Technol. 45, 104–109 (2013). [CrossRef]  

9. J. Ma, K. Li, L. Tan, S. Yu, and Y. Cao, “Performance analysis of satellite-to-ground downlink coherent optical communications with spatial diversity over Gamma–Gamma atmospheric turbulence,” Appl. Opt. 54, 7575–7585 (2015). [CrossRef]   [PubMed]  

10. P. Wang, Z. Lu, L. Guo, H. Feng, T. Shang, R. Wang, and Y. Yang, “Average BER of subcarrier intensity modulated free space optical systems over the exponentiated Weibull fading channels,” Opt. Express 22, 20828–20841 (2014). [CrossRef]   [PubMed]  

11. Y. Zhang, P. Wang, T. Liu, L. Guo, Y. Li, and W. Wang, “Performance analysis of a LDPC coded OAM-based UCA FSO system exploring linear equalization with channel estimation over atmospheric turbulence,” Opt. Express 26, 22182–22196 (2018). [CrossRef]   [PubMed]  

12. D. Liu, Y. Wang, X. Luo, G. Wang, and H. Yin, “Evolution properties of partially coherent four-petal Gaussian beams in oceanic turbulence,” J. Mod. Opt. 64, 1579–1587 (2017). [CrossRef]  

13. M. V. Jamali, A. Mirani, A. Parsay, B. Abolhassani, P. Nabavi, A. Chizari, P. Khorramshahi, S. Abdollahramezani, and J. A. Salehi, “Statistical Studies of Fading in Underwater Wireless Optical Channels in the Presence of Air Bubble, Temperature, and Salinity Random Variations,” IEEE Trans. Commun. 66, 4706–4723 (2018).

14. Y. Zhang, X. Wang, S. H. Zhao, J. Zhao, and B. Y. Deng, “On the performance of 2x2 DF relay mixed RF/FSO airborne system over Exponentiated Weibull fading channel,” Opt. Commun. 425, 190–195 (2018). [CrossRef]  

15. J. Zhao, S. H. Zhao, W. H. Zhao, Y. Liu, and X. Li, “Performance of mixed RF/FSO systems in exponentiated Weibull distributed channels,” Opt. Commun. 405, 244–252 (2017). [CrossRef]  

16. H. Sugiyama and K. Nosu, “MPPM: a method for improving the band-utilization efficiency in optical PPM,” J. Lightwave Technol. 7, 465–472 (1989). [CrossRef]  

17. A. Viswanath, V. K. Jain, and S. Kar, “Analysis of earth-to-satellite free-space optical link performance in the presence of turbulence, beam-wander induced pointing error and weather conditions for different intensity modulation schemes,” IET Commun. 9, 2253–2258 (2015). [CrossRef]  

18. H. S. Khallaf, J. M. Garrido-Balsells, H. M. H. Shalaby, and S. Sampei, “SER Analysis of MPPM-Coded MIMO-FSO System over Uncorrelated and Correlated Gamma-Gamma Atmospheric Turbulence Channels,” Opt. Commun. 356, 530–535 (2015). [CrossRef]  

19. H. S. Khallaf, H. M. H. Shalaby, J. M. Garrido-Balsells, and S. Sampei, “Performance analysis of a hybrid QAM-MPPM technique over turbulence-free and gamma-gamma free-space optical channels,” J. Opt. Commun. Netw. 9, 161–171 (2017). [CrossRef]  

20. D. D. Morabito, S. Shambayati, S. Finley, and D. Fort, “The cassini may 2000 solar conjunction,” IEEE Trans. Antennas Propag. 51, 201–219 (2003). [CrossRef]  

21. A. I. Efimov, L. Samoznaev, M. Bird, I. Chashei, and D. Plettemeier, “Solar wind turbulence during the solar cycle deduced from Galileo coronal radio-sounding experiments,” Adv. Space Res. 42, 117–123 (2008). [CrossRef]  

22. A. Afanasiev and N. Afanasiev, “Diagnostics of near-solar plasma turbulence parameters using the radio sounding technique at small heliocentric distances,” Sol. Phys. 245, 355–367 (2007). [CrossRef]  

23. S. Wing, J. R. Johnson, C. C. Chaston, M. Echim, C. P. Escoubet, B. Lavraud, C. Lemon, K. Nykyri, A. Otto, J. Raeder, and C. P. Wang, “Review of Solar Wind Entry into and Transport Within the Plasma Sheet,” Space Sci. Rev. 184, 33–86 (2014). [CrossRef]  

24. C. M. Ho, D. D. Morabito, and R. Woo, “Solar corona effects on angle of arrival fluctuations for microwave telecommunication links during superior solar conjunction,” Radio Sci. 43, RS2003 (2008). [CrossRef]  

25. G. Xu and Z. Song, “Solar Scintillation Effects on the Deep Space Communication Performance for Radio Wave Propagation Through Non-Kolmogorov Turbulence,” IEEE Antennas Wirel. Propag. Lett. 17, 1505–1509 (2018). [CrossRef]  

26. G. Thejappa and R. J. Macdowall, “Effects of scattering on radio emission from the quiet sun at low frequencies,” Astrophys. J. 676, 1338–1345 (2008). [CrossRef]  

27. J. Singh and V. K. Jain, Performance Analysis of BPPM and M-ary PPM Optical Communication Systems in Atmospheric Turbulence, IETE Tech. Rev. 25, 146–153 (2008). [CrossRef]  

28. K. Kiasaleh, “Performance of APD-based, PPM free-space optical communication systems in atmospheric turbulence,” IEEE Trans. Commun. 53, 1455–1461 (2005). [CrossRef]  

29. Y. Ata, Y. Baykal, and M. C. Gökçe, “M-ary pulse position modulation performance in non-Kolmogorov turbulent atmosphere,” Appl. Opt. 57, 7006–7011 (2018). [CrossRef]   [PubMed]  

30. Z. Ghassemlooy, W. Popoola, and S. Rajbhandari, Optical Wireless Communications: System and Channel Modelling with MATLAB (CRC, 2013).

31. K. E. Wilson and J. R. Lesh, “An overview of the Galileo Optical Experiment (GOPEX),” Telecommunications and Data Acquisition Progress Report 114, 192–204 (1993).

32. K. E. Wilson, “An Overview of the GOLD Experiment Between the ETS-VI Satellite and the Table Mountain Facility,” Telecommunications and Data Acquisition Progress Report , 1996, 8–19 (1996).

33. D. E. Smith, M. T. Zuber, and J. B. Abshire, “Mars Observer laser altimeter investigation,” J. Geophys. Res. Planet. 97, 7781–7797 (1993).

34. A. Albee, S. Battel, R. Brace, G. Burdick, J. Casani, J. Lavell, C. Leising, D. MacPherson, and P. Burr, “Report on the Loss of the Mars Polar Lander and Deep Space 2 Missions,” NASA Sti/recon Technical Report N (NASA, 2000).

35. D. M. Boroson, A. Biswas, and B. L. Edwards, “MLCD: Overview of NASA’s Mars laser communications demonstration system,” Proceedings of SPIE - The International Society for Optical Engineering , 2004, 5338 (2004).

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (9)

Fig. 1
Fig. 1 Deep space FSO communication between the Earth and probe during superior solar conjunction.
Fig. 2
Fig. 2 BER versus scintillation index for different values of M.
Fig. 3
Fig. 3 BER versus scintillation index for different values of G.
Fig. 4
Fig. 4 BER versus scintillation index for different values of Ks.
Fig. 5
Fig. 5 BER versus the wavelength, λ, for different values of the outer scale.
Fig. 6
Fig. 6 BER versus the wavelength, λ, for different values of the spectral index.
Fig. 7
Fig. 7 BER versus the heliocentric distance, r, for different data bit rates.
Fig. 8
Fig. 8 BER versus the heliocentric distance, r, for different equivalent load resistors.
Fig. 9
Fig. 9 BER versus the SNR, γ, for different scintillation indices, m2, with 4PPM and BPSK.

Tables (1)

Tables Icon

Table 1 Parameters used in the real FSO communication systems in deep space missions.

Equations (13)

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

δ ε ( r ) = r e λ 2 π δ N e ( r ) .
χ 2 = 16 π 4 r e   2 k 2 0 κ d κ 0 L sin  2 ( z κ 2 2 k ) Φ N ( κ , z ) d z .
Φ N ( κ ) = ( 2 π ) 3 / 2 ( p 3 ) Γ ( p / 2 ) Γ [ ( p 1 ) / 2 ] κ o p 3 δ N e 2 κ p ,
N e ( r ) = 4 × 10 14 ( R s u n r ) 10 + 3 × 10 14 ( R s u n r ) 6
χ 2 = ( p 3 ) Γ ( p / 2 ) r e   2 ( 2 π ) 11 / 2 p π 8 Γ [ ( p 1 ) / 2 ] Γ ( 1 + p / 2 ) η 2 N e   2 L o   3 p L p 2 k p 2 1 sec   ( π p / 4 ) .
y ( t ) = q n = G n x ( t t n ) + n ( t )
K s = 1 2 ρ λ P r c T w
f ( K s ) = 1 2 π σ k 2 K s exp  { [ ln  ( K s ) n k ] 2 2 σ k 2 } ,
P b = 0 Q ( Ξ ( K s ) ) f ( K s ) d K s ,
Ξ ( K s ) = ( G q ) 2 K s   2 ( G q ) 2 F ( K s + 2 K b g ) + 2 σ t h 2
P b 1 π i = 1 N w i Q { exp  [ 2 ( 2 σ k x i + n k ) ] F exp  ( 2 σ k x i + n k ) + K n } ,
T s = T b log  2 M M
P b M M 2 π i = 1 n w i Q { exp  [ 2 ( 2 m k x i + n k ) ] F exp  ( 2 m k x i + n k ) + K n } .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.