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

Performance analysis of FSO coherent BPSK systems over Rician turbulence channel with pointing errors

Open Access Open Access

Abstract

We investigate the performance of coherent free-space optical (FSO) system in terms of bit error rate (BER) evaluation by adopting the modified Rician distribution based coherent channel model which allows taking into consideration the composite effects of both Rician turbulence, including amplitude fluctuation and optical phase distortion, and pointing errors (PEs). By expanding the Rician distribution, a mathematically traceable expression of the probability density function (PDF) for the composite channel is derived in the form of the Meijer-G function. Based on the composite channel PDF, the exact BER expression is obtained, allowing the analysis of BER performance for single-input single-output (SISO) links. This analysis is extended to single-input multi-output (SIMO) links with maximal ratio combining (MRC). With the help of the moment generating function (MGF), the exact BER expression can be simplified into a single integral, facilitating the analysis with high accuracy and reducing calculation complexity. Engineering insights including high-SNR approximated channel PDF, asymptotic BER expression, coding and diversity gains, are investigated and cross-validated for both SISO and SIMO links. Through both analytical and numerical verifications, the impairment due to PEs as well as the effect of modal compensation on the BER performance are discussed in detail, unveiling the fact that their inner relations should be taken into account for optimization. These verify the effectiveness of our models for both SISO and SIMO links with a wide range of different conditions and can be feasibly applied for different types of coherent FSO links.

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

1. Introduction

Free-space optical (FSO) communication has attracted considerable attention due to those merits of easy deployment, license-free spectrum, high security, and high-speed transmission rates [13]. Recently, with the advent in coherent detection technology that allows for background noise rejection, high sensitivity, and improved spectral efficiency, there has been a regain of interest in coherent FSO communications with coherent homodyne/heterodyne receivers [48]. For instance, a homodyne binary phase shift keying (BPSK) transmission at 5.6 Gbps has been established between two islands [5,6]. However, the atmospheric turbulence in FSO channel induces not only amplitude fluctuation but also phase distortion. Therefore, the efficiency of coherent detection and the link performance are severely degraded [5,7,9,10].

To model the FSO channel, various statistical models have been proposed, such as Log-Normal distribution, $K$ distribution, Gamma-Gamma (G-G) distribution, and the Exponentiated Weibull (EW) distribution [2,1114]. Most of them focus on intensity modulation and direct detection (IM/DD) systems or ideal coherent systems with the assumption of perfect coherent detection where only amplitude fluctuation is considered. Nevertheless, in coherent FSO systems, the optical signal is required to be coherently mixed with a local oscillator (LO) laser to provide shot-noise-limited (SNL) performance. The wavefront matching between the signal and LO is largely deteriorated by both the turbulence-induced amplitude fluctuation and phase distortion [8,15,16]. Even those models could provide sufficient insights into the amplitude fluctuation, the effect of the turbulence-induced phase distortion could hardly be considered or could only be accounted for under certain restrictions. In addition, to cope with the phase distortions, phase compensations of different Zernike modes are usually required in coherent FSO systems. In order to model the composite effects of turbulence-induced amplitude fluctuation, and phase distortion, together with the compensation for phase distortions under varied turbulence cases, the modified Rician distribution has thus been introduced. Its capability and advantage has not only been verified theoretically by Monte Carlo simulation [16] and but also demonstrated experimentally in both maritime and terrestrial coherent FSO links [17]. Furthermore, it has been further employed for the analysis of the coherent FSO relay links [18], etc.

In addition, FSO systems also suffer from pointing errors (PEs) due to such as building sway and vibrations of transceivers [19,20]. For the G-G fading channel, the combined effect of turbulence and PEs has been modeled towards SISO links [20]. However, it is still quite challenging for multi-input multi-output (MIMO) links with PEs due to the mathematical complexity of the composite channel probability density function (PDF) [21,22]. On the other hand, for the Rician channel, the impact of PEs is seldom analyzed. The difficulty arises from the fact that the Rician PDF contains a modified Bessel function of the first kind and precludes from modeling the composite channel distribution of closed form. Conventionally, the Rician PDF could be expanded in Taylor series, which, however, holds only for the condition of negligible-to-small PEs for SISO links [23]. In general, the divergence of the transmitter is rather small (1-2.5 mrad in typical) and the receiver diameter is usually around 5-20 cm level [19,21]. This way, the alignment could be largely deteriorated by wind loads, thermal expansions, and platform vibrations, leading to severe pointing errors (1-5 mrad or larger) and performance degeneration. So far, coherent FSO system with large range of PEs (negligible-to-moderate, and strong-to-severe) has gone largely overlooked in SISO links. Furthermore, analysis of coherent FSO link with PEs in MIMO systems is even more difficult due to the complex high-order integration for BER analysis [22].

To this end, we have been largely motivated to analysis the performance of homodyne BPSK in both SISO and SIMO links over Rician turbulence channel with PEs. To model the composite effects, an exact analytical channel PDF applicable for all range of PEs and different values of SNR is formulated in terms of Meijer-G function via expansion of the Rician model in series of exponential terms. This leads to the derivation of an exact analytical BER expression for SISO links. Thanks to the moment generating function (MGF), the BER expression for (maximal ratio combining) MRC based SIMO links is established in a single integral which can be directly calculated with high accuracy. As it enables both analytical and numerical verifications for almost all different link conditions, the BER performance in both SISO and SIMO links are investigated. On top of that, from an engineering point of view, high-SNR approximated channel model, asymptotic BER, coding and diversity gains are also derived and assessed. The relation between the PEs and the modal compensations are carefully discussed, revealing their impacts on BER performances. Our modeling, in conjunction with the analysis approaches, is thus verified and can be adapted for a wide range of coherent FSO system.

2. System description and channel model

2.1 Coherent communication system

To simplify the analysis, a homodyne coherent BPSK FSO link is considered [8,15]. After being mixed with the LO, the received optical signal is down-converted. The corresponding photo-current can be expressed as

$$i(t)=i_{a c}(t)+i_{n}(t)$$
where ${i_n}\left ( t \right ) = \sqrt {2eR{P_{LO}}\Delta f}$ is the noise term dominated by LO shot noise, and the resulting signal part is given by ${i_{ac}}\left ( t \right ) = 2R\sqrt {{P_s}{P_{LO}}}$. Here, e is the electronic charge, R and $\Delta$f are the responsivity and the noise equivalent bandwidth of the photodetector, $P_s$ and $P_{LO}$ represent the received signal power and the LO power, respectively. Since the $P_{LO}$ is sufficiently large, the coherent detection is assumed to be shot noise limited (SNL). The resulting signal-to-noise ratio (SNR) without turbulence is expressed as ${\gamma _0} = \frac {{2R{P_s}}}{{e\Delta f}}$. From a practical point of view, the SNR suffers from intensity fluctuations due to atmospheric turbulence, path loss, and pointing errors (PEs) [1921]. In detail, such a composite fading channel can be considered as the product of the three different factors, $h = {h_a}{h_l}{h_p}$ [20], where $h_l$ is the path loss, $h_p$ is the attenuation due to geometric spread and pointing errors, and ${h_a}$ is responsible for turbulence-induced intensity fading due to both amplitude fluctuation and phase distortion. Then the resulting SNR is expressed as $\gamma = {\gamma _0}h$ [7,2022]. In the following, the exact PDF of the composited channel would be derived in an analytical form.

2.2 Channel models

For the coherent detection, the efficiency of coherent mixing and wavefront matching are deteriorated by the turbulence-induced amplitude fluctuation and phase distortion, degrading the SNR and system performance. The PDF of the intensity fading $\left ( {{h_a}} \right )$ is known to follow a modified Rician distribution as

$$P_{h_{a}}\left(h_{a}\right)=\frac{1+K}{\left\langle h_{a}\right\rangle} e^{{-}K} \exp \left(-\frac{1+K}{\left\langle h_{a}\right\rangle} h_{a}\right) I_{0}\left[2 \sqrt{\frac{(1+K) K}{\left\langle h_{a}\right\rangle} h_{a}}\right]$$
where ${I_0}\left ( . \right )$ is zero-order modified Bessel function of the first kind and $K$ is the Rician factor depending on turbulence fading characteristics [15,16,23,24].

To determine ${<h_a>}$ and $K$, both the turbulence amplitude fluctuation and the turbulenced-indeced phase distortion should be considered. The log-amplitude variance $\sigma _I^2$ can be described by a scintillation index (SI), i.e., $SI = \exp \left ( {4\sigma _I^2} \right ) - 1$. To compensate the turbulence-induced phase distortions, either zonal or modal compensation should be deployed. The residual phase variance after compensation of $J$ Zernike terms is given by $\sigma _{\phi }^{2}=C_{J}\left (D / r_{o}\right )^{5/3}$ [15,16]. The coefficient $C_J$ is determined by the number of compensated modes, e.g., the $J = 1, 3, 6$ describes the compensation for piston, tip-tilt, and astigmatism, respectively [8,15,23]. And the receiver aperture diameter $D$ is normalized by the wavefront coherence diameter $r_0$, which describes the spatial correlation of phase fluctuations in the receiver plane [15,16,23,24].

The atmospheric attenuation, ${h_l}$, is often deterministic and can be expressed in terms of the visibility and weather condition [21]. As it acts as a scaling factor, with proper approximation, it is always assumed to be equal to 1 [20,21].

Regarding the PE, a general model of misalignment-induced fading given in [19] is considered. Assuming a Gaussian spatial intensity profile of beam waist radius ${\omega _z}$, on the receiver plane at distance $z$ from the transmitter and a circular receive aperture of radius $D$, the PDF of ${h_p}$ is characterized by

$$f_{h_{p}}\left(h_{p}\right)=\frac{\xi^{2}}{A_{0}^{\xi^{2}}} h_{p}^{\xi^{2}-1} \quad 0 \leq h_{p} \leq A_{0}$$
where $\xi = {\omega _{zeq}}/2{\sigma _s}$ is the ratio between the equivalent beam radius and the pointing error displacement standard deviation at the receiver, $\omega _{zeq}=\omega _{z} \sqrt {\sqrt {\pi } erf(v) /\left (2 \operatorname {vexp}\left (-v^{2}\right )\right )}$, $v = \sqrt {\pi /2} \left ( {D/{\omega _z}} \right )$, ${A_0} = {\left [ {erf\left ( v \right )} \right ]^2}$, and $erf\left ( . \right )$ is the error function.

In a typical FSO system with PEs, the normalized pointing error deviation $(\sigma _{\mathrm {s}}/D)$ is assumed to be between 1 and 10, and the normalized beamwith $(\omega _{\mathrm {Z}}/D)$ could be 5 and 10 [1921]. When $\omega _{\mathrm {Z}}/D =10$, the value of $\xi$ is 5 and 0.5 for negligible PEs and severe PEs, respectively, which is in accordance with previous reports [25]. The smaller $\xi$ is, the stronger impact of the PE it implies. Therefore, we believe that it is appropriate to select $\xi = 6.7$ for negligible PEs, $\xi = 0.7$ for strong PEs, and $\xi = 0.4$ for severe PEs in our analysis.

2.3 Composited channel statistical model

Using the individual PDFs discussed above, the unconditional PDF of the composite fading channel is given by [20]

$$f_{h}(h)=\int f_{h h_{a}}\left(h | h_{a}\right) f_{h_{a}}\left(h_{a}\right) d h_{a}$$
where ${f_{h|{h_a}}}\left ( {h|{h_a}} \right )$ is the conditional probability read
$$f_{h|h_{a}}\left(h|h_{a}\right)=\frac{1}{h_{a} h_{l}} \frac{\xi^{2}}{A_{0}^{\xi^{2}}}\left(\frac{h}{h_{a} h_{l}}\right)^{\xi^{2}-1}, \quad \quad 0 \leq h \leq A_{0} h_{a} h_{l}$$
It could be found the integral expression in Eq. (4) is rather complex and could hardly lead to a convenient analytical description for the performance analysis. It is then important to derive analytically tractable formulas for the composite channel model.

2.3.1 Series representation of the Rician PDF

As aforementioned, an analytical expression for the composite channel could bring convenience in the performance analysis. To this end, the ${f_{{h_a}}}\left ( {{h_a}} \right )$ can be tightly approximated by applying Taylor series onto the ${I_0}$ [26, eq. (8.445)], i.e., $I_{0}\left [2 \sqrt {\frac {(1+K) K}{\left \langle h_{a}\right \rangle } h_{a}}\right ]=\sum _{i=1}^{\infty } \frac {1}{(i-1) ! \Gamma (i)}\left (\frac {(1+K) K}{\left \langle h_{a}\right \rangle } h_{a}\right )^{i-1}$. It has been shown that this expansion converges when $i$ is large. Moreover, the convergence of this series is dependent on the channel state and becomes slow until the truncated number $N = \sqrt {\frac {{\left ( {1 + K} \right )K}}{{{\left \langle h_{a}\right \rangle }}}{h_a}}$ [2628]. To achieve sufficient convergence for almost all parameter values, a large value of $N$, e.g., $N = 20$, is selected in our analysis. In this case, the Rician PDF could be tightly approximated and number of terms for the PDF estimation is largely reduced. Being analytically tractable, it converges as follows

$$P_{h_{a}}\left(h_{a}\right)=\sum_{i=1}^{N} \alpha_{i} h_{a}^{i-1} \exp \left(-\frac{1+K}{\left\langle h_{a}\right\rangle} h_{a}\right)$$
where ${\alpha _i} = \frac {{{r^{i - 1}}{e^{ - K}}}}{{\left ( {i - 1} \right )!\Gamma \left ( i \right )}}{\left ( {\frac {{\left ( {1 + K} \right )}}{{\left \langle h_{a}\right \rangle }}} \right )^i}$. Therefore, the Rician PDF now is simplified into the sum of exponential forms and facilitates the derivation of analytical PDF than conventional Taylor series expansion [23,29,30]. In this way, the new representation of the Rician PDF can be applied to derive the composite channel PDF for negligible-to-severe PEs.

2.3.2 Analytical expression of the composite channel PDF

Based on the above Rican PDF, now we can substitute Eqs. (5) and (6) into the integral in Eq. (4), the resulting ${f_h}\left ( h \right )$ reads:

$$f_{h}(h)=\frac{\xi^{2} h^{\xi^{2}-1}}{\left(A_{0} h_{l}\right)^{\xi^{2}}} \sum_{i=1}^{N} \alpha_{i} \int_{\frac{h}{A_{0} h_{i}}} h_{a}^{-\xi^{2}+i-1} e^{-\frac{1+K}{\left\langle h_{a}\right\rangle} h_{a}} d h_{a}$$
Thus a concise analytical result is derived using [26, eq. (3.381.3)] as
$$f_{h}(h)=\frac{\xi^{2}}{\left(A_{0} h_{l}\right)^{\xi^{2}}} h^{\xi^{2}-1} \sum_{i=1}^{N} \alpha_{i}\left(\frac{1+K}{\left\langle h_{a}\right\rangle}\right)^{\xi^{2}-i} \Gamma\left(-\xi^{2}+i, \frac{1+K}{A_{0} h_{l} \left\langle h_{a}\right\rangle} h\right)$$
where $\Gamma \left ( {.,.} \right )$ is the upper incomplete Gamma function defined in [27, eq. (3.381.3)].

Then, an analytical expression of the composite PDF can be derived in terms of the Meijer-G function $G_{p,q}^{m,n}\left [ . \right ]$, defined in [26, eq. (8.350.2)]. By utilizing the $\Gamma (a, b)=G_{1,2}^{2,0}\left [\left .b\right |_{0, a} ^{1}\right ]$ [31, eq. (06.06.26.0005.01)] and the relationship [26, eq. (9.31.5)], it allows us to further simplify the composite PDF as

$${f_h}\left( h \right) = \frac{{{\xi ^2}}}{{{A_0}{h_l}}}\mathop \sum _{i = 1}^N {\alpha _i}{\left( {\frac{{1 + K}}{{{\left\langle h_{a}\right\rangle}}}} \right)^{1 - i}}G_{1,2}^{2,0}\left[ {\frac{{1 + K}}{{{A_0}{h_l}{\left\langle h_{a}\right\rangle}}}h\left| {\begin{array}{*{20}{c}} {{\xi ^2}}\\ {{\xi ^2} - 1,i - 1} \end{array}} \right.} \right]$$
With the above exact composite channel PDF, it is allowed to derive the exact analytical expression for BER and to analyze performance under not only moderate but also severe PEs and this is firstly evaluated in both SISO and SIMO links.

3. Performance analysis for SISO channel

3.1 Exact analytical BER

The exact BER can be obtained by averaging the conditional BER ${P_e}\left ( {\gamma _{o} {|}h} \right )$ over the composite channel PDF as

$$P_{e}=\int_{0}^{+\infty} P_{e}\left(\gamma_{0} | h\right) f_{h}(h) d h$$
where ${P_e}\left ( {{\gamma _0}|h} \right )$ is the conditional BER given by $P_{e}\left (\gamma _{0} | h\right )=Q\left (\sqrt {2 \gamma _{0} h}\right )$ for the BPSK signal, $Q\left ( . \right )$ is the Gaussian Q-function and ${\gamma _0}$ denotes the SNR without turbulence and PEs.

Then we alternatively apply the Q-function in the Meijer-G form ${erfc}\left (\sqrt {\gamma _{0}h}\right )=\frac {1}{\sqrt {\pi }} \boldsymbol {G}_{1,2}^{2,0}\left [\gamma _{0} h\right |_{0,0.5} ^{1}]$ [32, eq. (12)]. Along with the composite channel PDF the integral of the product of two Meijer-G functions [26, eq. (7.811.1)], the exact BER can be analytically expressed as

$$B E R=\frac{\xi^{2}}{2 \sqrt{\pi} A_{0} h_{l}} \frac{1}{\gamma_{0}} \sum_{i=1}^{N} \alpha_{i}\left(\frac{1+K}{\left\langle h_{a}\right\rangle}\right)^{1-i} G_{3,3}^{2,2}\left[\frac{1+K}{A_{0} h_{l} \left\langle h_{a}\right\rangle \gamma_{0}} | \begin{array}{c}{0,-0.5, \xi^{2}} \\ {\xi^{2}-1, i-1,-1}\end{array}\right]$$
If the composite channel PDF in Eq. (8) is applied, the exact BER expression for SISO link could also be obtained with the help of the [26, eq. (2.10.8.2)], as
$$\begin{aligned} B E R=\frac{1}{2} \sum_{i=1}^{N} \alpha_{i}\{& -\frac{\zeta^{2} \Gamma(i+0.5)}{\left(i-\zeta^{2}\right) i \sqrt{\pi}\left(A_{0} \gamma_{0}\right)^{i}}\left[3 F_{2}\left(i-\zeta^{2}, i, i+0.5 ; i-\zeta^{2}+1, i+1 ;-\frac{1}{A_{0} \gamma_{0}} \frac{1+K}{\langle h_{a}\rangle}\right)\right] \\ & +\frac{\Gamma\left(i-\zeta^{2}\right) \Gamma\left(\zeta^{2}+0.5\right)}{\sqrt{\pi}\left(A_{0} \gamma_{0}\right)^{\zeta^{2}}}\left(\frac{1+K}{\langle h_{a}\rangle}\right)^{\zeta^{2}-i} \} \end{aligned}$$
where the $p F_{q}\left (a_{1}, \dots , a_{p} ; b_{1}, \dots , b_{q} ; x\right )$ is the Gauss hypergeometric function. It could found that the Meijer-G form in Eq. (11) is more concise and mathematically traceable. Though only BPSK is considered here, it is also applicable to other modulation formats, such as M-ary PSK and M-ary QAM, by using different Q-functions or conditional BERs. For instance, if the QPSK modulation is applied, the conditional BER could be slightly altered as $P_{e,QPSK}\left (\gamma _{0} | h\right )= 2Q\left (\sqrt {2 \gamma _{0} h}\right )$, and the unconditional BER could be easily calculated.

3.2 High-SNR approximated PDF and asymptotic BER

To get more insights into the coherent FSO system, BER performance in high SNR regime is further evaluated. In typical, high-SNR performance is mainly determined by the behaviors of channel PDF near the origin [29]. As such, the composite PDF in Eq. (9) can be approximated by a single polynomial term as $P_{h \rightarrow 0}(h)=a h^{b}+o\left (h^{b}\right )$ at $h \to 0$, where $o({h^b})$ could be neglected as ${\lim _{h \to 0}}o\left ( {h^{\textrm{b}}} \right )/h = 0$. So as to obtain the expression of ${P_{h \to 0}}\left ( h \right )$, the Meijer-G function in the composite PDF is expanded around $h \to 0$ [31, eq. (07.34.06.0006.01)], as follows

$$G_{1,2}^{2,0}\left[\left.z\right|_{b_{1}, b_{2}} ^{a_{1}}\right] \approx \frac{\Gamma\left(b_{2}-b_{1}\right)}{\Gamma\left(a_{1}-b_{1}\right)} z^{b_{1}}+\frac{\Gamma\left(b_{1}-b_{2}\right)}{\Gamma\left(a_{1}-b_{2}\right)} z^{b_{2}}+o(z)$$
where $\Gamma \left ( \cdot \right )$ is the Gamma function [26], the parameters $a_1, b_1, b_2$ and the argument $z$ are real or complex parameters. The expansion is valid under the assumption of ${a_1} - {b_j} \notin \textrm {Z}$ (integer), $j = 1, 2$. Therefore, different expressions for $a$ and $b$, depending on $\xi$, can be derived as
$$a=\left\{\begin{array}{c}{\frac{\xi^{2} \mathrm{e}^{{-}K}}{\left(\xi^{2}-1\right)} \frac{(1+K)}{A_{0} h_{l} \left\langle h_{a}\right\rangle}, \qquad \qquad \qquad \quad \xi>1} \\ {\xi^{2} \mathrm{e}^{{-}K} \Gamma\left(1-\xi^{2}\right)\left(\frac{1+K}{A_{0} h_{l} \left\langle h_{a}\right\rangle}\right)^{\xi^{2}}, \quad \xi<1}\end{array}\right.$$
$$b=\left\{\begin{array}{c}{0, \quad \qquad \, \xi>1} \\ {\xi^{2}-1, \quad \,\xi<1}\end{array}\right.$$
Therefore, the composite PDF is further approximated as
$$f_{h \rightarrow 0}(h)=\left\{\begin{array}{c}{\frac{\xi^{2} e^{{-}K}}{\left(\xi^{2}-1\right)} \frac{(1+K)}{A_{0} h_{1} \left\langle h_{a}\right\rangle}, \qquad \qquad \qquad \qquad \quad \ \xi>1} \\ {\xi^{2} e^{{-}r} \Gamma\left(1-\xi^{2}\right)\left(\frac{1+K}{A_{0} h_{l} \left\langle h_{a}\right\rangle}\right)^{\xi^{2}} h^{\xi^{2}-1}, \quad \xi<1}\end{array}\right.$$
In order to evaluate the asymptotic BER in high SNR regime, we substitute the approximated PDF into $BER = \int _0^\infty {Q\left ( {\sqrt {2\gamma {}_0h} } \right )a{h^b}dh}$. With the help of the identities [26, eq. (8.359.3) and eq. (6.455.1)], different asymptotic BERs relying on the PEs can be written as
$$BER_{SISO,asym}=\left\{\begin{array}{c}{\frac{\Gamma(1.5) e^{{-}K} \xi^{2}}{2 \sqrt{\pi}\left(\xi^{2}-1\right)} \frac{(1+K)}{\left(A_{0} h_{l} \left\langle h_{a}\right\rangle \right)} \gamma_{0}^{{-}1}, \qquad \qquad \qquad \ \xi>1} \\ {\frac{e^{{-}K} \Gamma\left(1-\xi^{2}\right) \Gamma\left(\xi^{2}+0.5\right)}{2 \sqrt{\pi}}\left(\frac{1+K}{A_{0} h_{l} \left\langle h_{a}\right\rangle}\right)^{\xi^{2}} \gamma_{0}^{-\xi^{2}}, \quad \xi<1}\end{array}\right.$$
For the sake of simplicity, we rewrite the above equation as $BE{R_{\textrm{SISO,asym}}} = {\left ( {{G_{c,SISO}}{\gamma _0}} \right )^{-{G_{d,SISO}}}}$, with ${G_{c,SISO}}$ and ${G_{d,SISO}}$ denoting the coding gain and the diversity gain, respectively [29]. It is shown that the former indicates the shift of the BER curve and the latter determines the slope of the BER curve versus the SNR in a log-log scale. Depending on the PEs, the two gains could be obtained, respectively as
$$G_{\mathrm{c,SISO}}=\left\{\begin{array}{l}{\left(\frac{\mathrm{e}^{{-}K} \xi^{2} \Gamma(1.5)}{2 \sqrt{\pi}\left(\xi^{2}-1\right)} \frac{(1+K)}{\left(A_{0} h_{l} \left\langle h_{a}\right\rangle\right)}\right)^{{-}1}, \qquad \quad \ \;\xi>1} \\ {\left(\frac{\mathrm{e}^{{-}K} \Gamma\left(1-\xi^{2}\right)}{2 \sqrt{\pi}}\left(\frac{1+K}{A_{0} h_{l} \left\langle h_{a}\right\rangle}\right)^{\xi^{2}}\right)^{{-}1 / \xi^{2}}, \quad \xi<1}\end{array}\right.$$
$$G_{\mathrm{d,SISO}}=\left\{\begin{array}{ll}{1,} & {\xi>1} \\ {\xi^{2},} & {\xi<1}\end{array}\right.$$

3.3 Pointing error induced SNR loss factor

When $\xi > 1$ is satisfied, it is found that the diversity gain is independent of negligible-to-moderate PEs, as typically for some FSO sceneries [21,33] . In other words, this requires the equivalent beam radius at the receiver is at least twice the pointing error displacement standard deviation. It is thus convenient to compare the BER performance with varied $\xi$. The impact of PEs can be translated into a sensitivity loss, relative to the case without PEs, as follows

$$\operatorname{Loss}(\textrm{dB})=10 \log _{10}\left(\frac{\zeta^{2}}{A_{0}\left(\zeta^{2}-1\right)}\right), \quad \xi>1$$
Thus it is expected to minimize the $Loss(\textrm {dB})$ by adjusting the equivalent beam width, which is related to $\xi$ as well as $A_{0}$.

4. Performance analysis for SIMO links

In this section, the maximal-ratio combining (MRC) scheme is employed when our model is extended to the analysis for SIMO links.

4.1 Exact BER expression

A typical SIMO link is assumed to include L identical but independent apertures. Different from the SISO link, traditional PDF-based BER analysis is almost mathematically intractable for the the SIMO link. Because it involves a L-fold integral defined by the joint multivariate PDF of the SNR in each branch [22]. When the composite channel is considered in the SIMO link, MGF-based approach is applied to further simplify the BER analysis. In this way, the L-fold integral can be reduced into a single-fold integral [34]. The total SNR with MRC is the sum of the individual SNR at each branch, i.e., ${{{\gamma }}_{MRC}} = \mathop \sum \limits_{l = 1}^L {\gamma _l}$ with ${\gamma _l}$ being the SNR in the $l^{th}$-branch [24,35]. Then, the MGF of ${{{\gamma }}_{MRC}}$ is the product of the MGF of each ${\gamma _l}$ and the BER can be simplified to a single finite integral as

$$\mathrm{BER}=\frac{1}{\pi} \int_{0}^{\frac{\pi}{2}} \operatorname{MGF}_{M R C}\left(s=\frac{1}{\sin ^{2} \theta}\right) d \theta=\frac{1}{\pi} \int_{0}^{\frac{\pi}{2}} \prod_{l=1}^{L} \operatorname{MGF}_{\gamma_{l}}\left(s=\frac{1}{\sin ^{2} \theta}\right) d \theta$$
where $MG{F_{{\gamma _l}}}\left ( s \right ) = \int _0^\infty {{e^{ - s{\gamma _l}}}{f_{{\gamma _l}}}\left ( {{\gamma _l}} \right )d\gamma _l}$ is the MGF of the ${\gamma _l}$ in the $l^{th}$-branch [34]. To calculate $MG{F_{{\gamma _l}}}\left ( s \right )$, the PDF of the SNR for each branch is first obtained by applying the Jacobin transformation ${f_\gamma }\left ( \gamma \right ) = \frac {1}{{{\gamma _o}}}{f_h}\left ( {h = \frac {\gamma }{{{\gamma _o}}}} \right )$ as follows
$$f_{\gamma_{i}}\left(\gamma_{l}\right)=\frac{\xi^{2}}{A_{0} h_{l} \gamma_{o}} \sum_{i=1}^{N} \alpha_{i}\left(\frac{(1+K)}{\left\langle h_{a}\right\rangle}\right)^{1-i} G_{1,2}^{2,0}\left[\frac{(1+K) \gamma_{l}}{A_{0} h_{l}\left\langle h_{a}\right\rangle \gamma_{o}} | \begin{array}{c}{\xi^{2}} \\ {\xi^{2}-1, i-1}\end{array}\right]$$
With $\mathrm {e}^{-x}=G_{0,1}^{1,0}\left [\left .x\right |_{0} ^{-}\right ]$ [32, eq. (11)] and the integral of the product of two Meijer-G functions [26, eq. (7.811.1)], the MGF in the $l^{th}$-branch can be derived as
$$MGF_{\gamma_{i}}(s)=\frac{\xi^{2}}{s A_{0} h_{l} \gamma_{o}} \sum_{i=1}^{N} \alpha_{i}\left(\frac{(1+K)}{\left\langle h_{a}\right\rangle }\right)^{1-i} G_{2,2}^{2,1}\left[\frac{(1+K)}{s A_{0} h_{l} \left\langle h_{a}\right\rangle \gamma_{o}} | \begin{array}{c}{0, \xi^{2}} \\ {\xi^{2}-1, i-1}\end{array}\right]$$
Consequently, the BER can be written as
$$B E R=\frac{1}{\pi} \int_{0}^{\frac{\pi}{2}} \prod_{l=1}^{L}\left\{\frac{\xi^{2}}{A_{0} h_{l} \gamma_{o}} \sum_{i=1}^{N} \alpha_{i}\left(\frac{(1+K)}{\left\langle h_{a}\right\rangle}\right)^{1-i} G_{2,2}^{2,1}\left[\frac{(1+K) \sin ^{2} \theta}{A_{0} h_{l}\left\langle h_{a}\right\rangle \gamma_{o}} | \begin{array}{c}{0, \xi^{2}} \\ {\xi^{2}-1, i-1}\end{array}\right]\right\} d \theta$$
In this way, we only need to consider a single integral for the SIMO link. This is quite feasible and convenient using standard numerical approaches such as Gauss-Chebyshev quadrature (GCQ) formula [27,34].

4.2 Asymptotic BER expression

In the high SNR regime, the MGF can be asymptotically expanded using the dominant term of the $G_{2,2}^{2,1}\left [z | \begin {array}{l}{a_{1}, a_{2}} \\ {b_{1}, b_{2}}\end {array}\right ]$. Depending on $\xi$, the MGF in high SNR regime $\left ( {\textrm{s} \to 0} \right )$ is simplified as

$$MGF_{\gamma}(s)=\left\{\begin{array}{ll}{\left[\frac{\xi^{2} e^{{-}K}(1+K)}{\left(\xi^{2}-1\right) A_{0} h_{1} \left\langle h_{a}\right\rangle \gamma_{o}}\right]^{L} s^{{-}L}+o(s),} & {\xi>1} \\ {e^{{-}r} \Gamma\left(1-\xi^{2}\right) \Gamma\left(1+\xi^{2}\right)\left(\frac{(1+K)}{A_{0} h_{l} \left\langle h_{a}\right\rangle \gamma_{o}}\right)^{\xi^{2}} ]^{L} s^{{-}L \xi^{2}}+o(s),} & {\xi<1}\end{array}\right.$$
where ${o}\left ( \textrm{s} \right )$ stands for higher-ordered series and can be neglected in the high-SNR regime. Thereby, in this case, the PDF of the total SNR can can be obtained by inverse Laplace transform of the MGF [26, Tab. (17.13)], i.e., ${f_\gamma }\left ( x \right ) = {L^{ - 1}}\left \{ {MG{F_\gamma }\left ( s \right )} \right \}$. By using ${L^{ - 1}}\left \{ {{s^{ - m}}} \right \} = \frac {{{x^{m - 1}}}}{{\left ( {m - 1} \right )!}}$ and ${L^{ - 1}}\left \{ {\frac {{{{\Gamma }}\left ( {v + 1} \right )}}{{{s^{v + 1}}}}} \right \} = {x^v}$, the PDF of the MRC-based SIMO link is approximated as
$$f_{MRC,asym}(h)=\left\{\begin{array}{l}{\frac{1}{(L-1) !}\left[\frac{\xi^{2} e^{{-}K}(1+K)}{\left(\xi^{2}-1\right) A_{0} h_{l} \left\langle h_{a}\right\rangle}\right]^{L} h^{(L-1)}, \qquad \qquad \qquad \qquad \qquad \qquad \; \xi>1} \\ {\frac{1}{\Gamma\left(L \xi^{2}\right)}\left[\xi^{2} e^{{-}r} \Gamma\left(1-\xi^{2}\right) \Gamma\left(\xi^{2}\right)\left(\frac{(1+K)}{A_{0} h_{l} \left\langle h_{a}\right\rangle \gamma_{o}}\right)^{\xi^{2}}\right]^{L} h^{L \xi^{2}-1}, \qquad \, \xi<1}\end{array}\right.$$
Especially for $L = 1$, namely for a SISO link, the obtained PDF is in good accordance with that in Eq. (16), which further proves the validity of the approximation in high SNR regime. Similarly, the asymptotic BER expression using $P_{e}=\int _{0}^{+\infty } Q\left (\sqrt {2 \gamma _{0} h}\right ) f_{MRC, asym}(h) dh$ can also be derived as
$$BER_{MRC,asym}=\left\{\begin{array}{l}{\frac{1}{2 \sqrt{\pi} L !}\left[\frac{\xi^{2} e^{{-}K}(1+K)}{\left(\xi^{2}-1\right) A_{0} h_{l} \left\langle h_{a}\right\rangle}\right]^{L} \Gamma(L+0.5)\left(\gamma_{0}\right)^{{-}L}, \qquad \qquad \qquad \qquad \xi>1} \\ {\frac{\Gamma\left(L \xi^{2}+0.5\right)}{2 \sqrt{\pi} \Gamma\left(L \xi^{2}+1\right)}\left[\mathrm{e}^{{-}r} \Gamma\left(1-\xi^{2}\right) \Gamma\left(\xi^{2}+1\right)\left(\frac{(1+K)}{A_{0} h_{l} \left\langle h_{a}\right\rangle \gamma_{o}}\right)^{\xi^{2}}\right]^{L} \gamma_{0}^{{-}L \xi^{2}},\ \xi<1}\end{array}\right.$$
Accordingly, it also suggests that the asymptotic BER in the SIMO link could also be re-written as ${P_{MRC,asym}} = {\left ({{G_{c,MRC}}{\gamma _o}} \right )^{-{G_{d,MRC}}}}$, where the coding gain and the diversity gain are derived, respectively as
$$G_{c,MRC}=\left\{\begin{array}{l}{\left(\frac{\Gamma(L+0.5)}{2 \sqrt{\pi} L !}\left[\frac{\xi^{2} e^{{-}K}(1+K)}{\left(\xi^{2}-1\right) A_{0} h_{l} \left\langle h_{a}\right\rangle}\right]^{\frac{-1}{L}}, \quad \qquad \qquad \qquad \qquad \qquad \qquad \qquad \xi>1\right.} \\ {\left(\frac{\Gamma\left(L \xi^{2}+0.5\right)}{2 \sqrt{\pi} \Gamma\left(L \xi^{2}+1\right)}\left[e^{{-}r} \Gamma\left(1-\xi^{2}\right) \Gamma\left(\xi^{2}+1\right)\left(\frac{(1+K)}{A_{0} h_{l} \left\langle h_{a}\right\rangle \gamma_{o}}\right)^{\xi^{2}}\right]^{{-}1 /\left(L \xi^{2}\right)}, \quad \xi<1\right.}\end{array}\right.$$
$$G_{d,MRC}=\left\{\begin{array}{ll}{L,} & {\xi>1} \\ {L \xi^{2},} & {\xi<1}\end{array}\right.$$
It is noted that those obtained coding gain and diversity order are in accordance to the SISO link ($L = 1$). It further demonstrates that the asymptotic BER and diversity order are only determined by the receiver branches ($L$) when the PE is negligible-to-moderate, i.e., ${\xi > 1}$, as the same way as SISO link. However, as the pointing error is stronger, i.e., ${\xi < 1}$ , the diversity order is partially reduced by the pointing error.

Meanwhile, it is further manifested that the diversity gain is only determined by the receiver branches ($L$) under the condition of negligible-to-moderate PEs, as the same way as the SISO link. In this case, it could be deduced that the diversity gain could be linearly increased by distributing more branches in SIMO links. However, when the impact of PE becomes stronger, i.e., ${\xi < 1}$ , the diversity gain is partially reduced. For instance, when $\xi = 0.4$ , the diversity gain is largely suppressed to merely $16\ \%$.

5. Numerical results and discussion

In this section, detailed analyses on the performance of the previously derived SISO and SIMO FSO systems are discussed. The optical wavelength is assumed to be ${{\lambda }} = 1550\, \textrm{nm}$. For a typical FSO system, a source-to-destination distance of $1 km$ and receiver diameter of $D = 10\ \textrm {cm}$ is assumed. Without loss of generality, the weak, moderate, and strong turbulence is characterized by $SI = 0.1 ,0.33$ , and 1.6, respectively. The path loss factor ${h_l} = 1$ is considered. As for the impacts of PEs, the $\xi$ is varied within $\xi = 6.7,1.4,0.7,0.4$ with ${A_0} = 1$. For instance, generally speaking, ${{\xi }} = 6.7$ represents negligible PE while $\xi = 0.4$ stands for severe PE.

The combined effects of Rician turbulence and PEs on the BER performance for a SISO-FSO link is verified in Fig. 1. Without loss of generality, we consider two cases: weak $\left ( {\textrm{SI} = 0.1} \right )$ and strong turbulence $\left ( {\textrm{SI} = 1.6} \right )$. When the impact of PEs is as large as $\xi = 0.4$, the BER for all SNRs remain above $1.0 \times {10^{-2}}$ under weak turbulence and also above $1.0 \times {10^{-1}}$ under strong turbulence. In this case, the severe PE leads to a noise floor and the BER could hardly be improved by increasing SNRs. For moderate PE, namely $\xi = 1.4$ , the BER at $SNR = \ 30 \ \textrm {dB}$ becomes $6.0\times {10^{-4}}$ and $2.0\times {10^{-2}}$ under weak and strong turbulence, respectively. Within moderate PE regime, even for a large change from $\xi = 6.7$ to $\xi = 1.4$, there is no significant degradations in the BER performance. This indicates that the system seems to be quite tolerable for negligible-to-moderate PEs. According to the discussion in section 3.3, the sensitivity loss due to the PEs could be easily estimated. For instance, at $BER = {10^{-4}}$, the SNR loss is approximated around 3 dB for $\xi = 1.4$ than that for $\xi = 6.7$. As shown in Fig. 1, the exact BER closely matches the numerical results for all the SNRs and the PEs. Moreover, the asymptotic BER approaches the exact BER in the high SNR regime.

 figure: Fig. 1.

Fig. 1. BER performance of coherent FSO-SISO link in the composite channel with Rician turbulence and PEs, under (a) weak turbulence $SI = 0.1$, (b) strong turbulence $SI = 1.6$. To represent the impacts of PEs from negligible to severe, the $\xi$ is varied as $\xi = 6.7,1.4,0.7,0.4$. Analytical results (solid lines), asymptotic results (dashed lines), and numerical simulation (*) results are both included.

Download Full Size | PDF

As the turbulence-induced phase distortions could be improved by modal compensation, here we further investigate the BER performance under different Zernike modal compensations $(J = 1, 3, 6)$ when various PEs are considered. In Fig. 2, the condition of negligible PE $\left ( {\xi = 6.7} \right )$ is firstly assessed. Compared with the case for $J = 1$, the BER at $SNR = \ 30 \ \textrm {dB}$ is improved from the original $2.0 \times {10^{-3}}$ to $4.0 \times {10^{-7}}$ for $J = 3$, and further to $6.0 \times {10^{-9}}$ for $J = 6$. All these results are in accordance with the previous reports [8,35] that low order corrections such as tip-tilt correction $(J = 3)$ are able to improve the BER performance significantly under the condition of $D/{r_0}=2$. When the impacts of PEs are considered, it is obvious that the modal compensation can effectively suppress the impacts of PEs in low-to-moderate SNR regime. On the other hand, degradation in BER performance can be found in high SNR regime for variations in PEs from $\xi = 6.7$ to $\xi = 0.4$. In comparison with the case for $J = 1$, the deterioration of PEs on BER performance now shifts to high SNR regime when modal compensations are applied. This is mainly due to the fact that the mean value $\left \langle {{h_a}} \right \rangle$ gradually approaches the unity and the Rician factor $K$ also becomes stronger when higher order modal compensation is applied. As for a coherent FSO system, the BER performance is mainly dominated by the phase distortion, therefore, the modal compensation is more effective in the small-to-moderate SNR regime. On the contrary, in the high SNR regime, the BER is gradually deteriorated by the PEs, as the misalignment induces extra intensity fading.

 figure: Fig. 2.

Fig. 2. BER performance under various compensation modes $(J = 1, 3, 6)$ over the composite Rician turbulence and PE fading channel. The $J = 1, 3, 6$ describes compensation for piston, tip-tilt, astigmatism, respectively. To represent general turbulence, moderate turbulence of $SI = 0.33$ is considered.

Download Full Size | PDF

The impact of PEs for the MRC-based SIMO link is further investigated in Fig. 3, where two cases are included, Fig.3 (a) without modal compensation, and Fig.3 (b) with $J = 3$ compensated mode in each aperture. To represent a general FSO scenario, moderate turbulence with $SI = 0.33$ is assumed. For the sake of fair comparison, aperture area of each receiver is assumed to be $1/L$ times the aperture area of a SISO link [24,35], ${D_l} = D/\sqrt L ,0 \ll l \le L$ , where $D$ denotes the aperture diameter in SISO link.

 figure: Fig. 3.

Fig. 3. Comparison of BER performance for MRC-SIMO link in cases of different number of receiver branches $(L = 1, 2, 4)$ under moderate turbulence $(SI = 0.33)$ with varied PEs. Two situations are involved, (a) without and (b) with modal compensation $(J = 3)$. The BER for the SISO link $(L = 1)$ is depicted as a benchmark.

Download Full Size | PDF

In general, it can be inferred by Fig. 3(a) that, in the SIMO system, due to the presence of PEs, the more branches are employed, the greater impairment on the BER improvement can be observed. In particular, for negligible PE, the BER could be improved from $2.0 \times {10^{-3}}$ at $SNR = \ 30 \ \textrm {dB}$ for a single receiver, to $1.0 \times {10^{-5}}$ for $L = 2$, and further to $1.0 \times {10^{-10}}$ for $L = 4$. Similar tendencies are observed in the case of moderate PE [red curves in Fig. 3(a)] which is manifested as a quasi-linear BER improvement. This cross-verifies the relation unveiled between the diversity gain and number of branches as previously indicated by Eq. (29). Nevertheless, with the presence of strong-to-severe PEs, such trendency has been largely retard. For instance, in the case of $\xi = 0.7$ [(pink curves in Fig. 3(a)], the BER at $SNR = \ 30 \ \textrm {dB}$ is now only $2.0 \times {10^{-2}}$ for a single receiver and the improvement is narrowed to $2.0 \times {10^{-3}}$ for $L = 2$, and to merely $1.0 \times {10^{-6}}$ for $L = 4$. This tendency is more significant for $\xi = 0.4$ which is, however, not shown to keep the figure clear. Clearly, for all the cases, the asymptotic BERs (dashed curves) are sufficiently tight with the exact ones in the high SNR regime.

In Fig. 3(b), the effect of modal compensation is verified in the SIMO linkwith varied PEs. As can be inferred by the results from the SISO link, the effect of low-order modal compensation is already sufficient. Thus the modal compensation of $J = 3$ is investigated without losing generality. Not surprisingly, the model compensation further facilitates the BER improvement from small to large SNR cases. In addition, the impairments due to the PEs are largely alleviated. This is confirmed as the beginning of the PE-induced BER deterioration is now observed at a higher SNR compared to the case without modal compensation. It should be also noted that the boundary of the BER is almost limited by the PEs and could be further improved with higher-order modal compensations.

Therefore, to ensure BER performance, PEs existing in any practical SISO and SIMO links should be considered and suppressed. Besides, the impairments due to PEs could also be reduced with other techniques, such as optimizing transmitter beamwidth. As only low-ordered modal compensations can provide significant BER improvements and suppression onto the impacts of PEs, deployments of multiple receivers with modal compensations are suggested for achieving sufficient BER performances.

6. Conclusion

We investigate the BER performances for coherent FSO system in both SISO and MRC-based SIMO links, within which the combined effects of Rican turbulence and PEs are both modeled and considered. Different from other channels models, the modified Rician distribution is manifested by its capability to model turbulence amplitude fluctuation, optical phase distortions and compensation for a practical coherent receiver. For the ease of analysis, a novel mathematically traceable expression for the composite channel PDF is first formulated, which allows accounting for a wide range of PEs, namely, negligible-to-severe PEs. Based on this model, an exact analytical BER expression is derived for SISO links while the extension to MRC-based SIMO links is achieved through a simple one-dimension integral of the MGF. To provide in-depth engineering insights for both types of links, high-SNR approximated channel PDF, asymptotic BER expressions, coding and diversity gains, are explored and discussed in detail. For the SISO link, it is instructive to find that the impact of PEs dominates the BER performance but can be efficiently mitigated by the modal compensation in low-to-moderate SNR regime. With higher-order of compensations, the PEs induced BER deterioration is retard as it becomes obvious only at higher SNRs compared with the non-compensated cases. For SIMO links, quite similar tendencies could also be observed while the impact of PEs becomes more severe with the increase of the number of branches. In high SNR regime, thanks to the asymptotic BER evaluation, the SNR penalty due to the PEs is estimated for both types of links. Both analytical and numerical investigations are carried out, constituting a cross-verification for the effectiveness of our modeling. The proposed modeling provides an accurate and more complete description for the Rician channel with the presence of a wide range of PEs with respect to the coherent FSO system. Therefore, it offers an efficient and useful approach for practical performance analysis, enabling further system optimization from an application point of view.

Funding

National Natural Science Foundation of China (61690193, 61805014, 61827807); China Postdoctoral Science Foundation (2018M630082, 2019T120051).

References

1. M. A. Khalighi and M. Uysal, “Survey on free space optical communication: A communication theory perspective,” IEEE Commun. Surv. Tut. 16(4), 2231–2258 (2014). [CrossRef]  

2. H. Kaushal and G. Kaddoum, “Optical communication in space: challenges and mitigation techniques,” IEEE Commun. Surv. Tut. 19(1), 57–96 (2017). [CrossRef]  

3. A. Mansour, R. Mesleh, and M. Abaza, “New challenges in wireless and free space optical communications,” Opt. Lasers Eng. 89, 95–108 (2017). [CrossRef]  

4. C. Tang, R. Li, Y. Shao, N. Chi, J. Yu, Z. Dong, and G. Chang, “Experimental demonstration for 40 km fiber and 2 m wireless transmission of 4 Gb/s OOK signals at 100 GHz carrier,” Chin. Opt. Lett. 11(2), 20608–20610 (2013). [CrossRef]  

5. R. Lange, B. Smutny, B. Wandernoth, R. Czichy, and D. Giggenbach, “142 km, 5.625 Gbps free-space optical link based on homodyne bpsk modulation,” in Free-Space Laser Communication Technologies XVIII, vol. 6105 (International Society for Optics and Photonics, 2006), p. 61050A.

6. Z. Zhu, H. Zhou, W. Xie, J. Qin, and Y. Dong, “10 Gb/s homodyne receiver based on costas loop with enhanced dynamic performance,” in 16th International Conference on Optical Communications and Networks (ICOCN), (IEEE, 2017), pp. 1–3.

7. M. Niu, X. Song, J. Cheng, and J. F. Holzman, “Performance analysis of coherent wireless optical communications with atmospheric turbulence,” Opt. Express 20(6), 6515–6520 (2012). [CrossRef]  

8. C. Liu, S. Chen, X. Li, and H. Xian, “Performance evaluation of adaptive optics for atmospheric coherent laser communications,” Opt. Express 22(13), 15554–15563 (2014). [CrossRef]  

9. K. Zong and J. Zhu, “Performance evaluation and optimization of multiband phase-modulated radio over isowc link with balanced coherent homodyne detection,” Opt. Commun. 413, 152–156 (2018). [CrossRef]  

10. D. Zheng, Y. Li, H. Zhou, Y. Bian, C. Yang, W. Li, J. Qiu, H. Guo, X. Hong, Y. Zuo, P. G. Ian, T. Weijun, and W. Jian, “Performance enhancement of free-space optical communications under atmospheric turbulence using modes diversity coherent receipt,” Opt. Express 26(22), 28879–28890 (2018). [CrossRef]  

11. L. C. Andrews and R. L. Phillips, Laser beam propagation through random media (SPIE, Bellingham, WA, 2005).

12. H. Yura and T. Rose, “Exponentiated weibull distribution family under aperture averaging gaussian beam waves: comment,” Opt. Express 20(18), 20680–20683 (2012). [CrossRef]  

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

14. Z. Zeng, S. Fu, H. Zhang, Y. Dong, and J. Cheng, “A survey of underwater optical wireless communications,” IEEE Commun. Surv. Tut. 19(1), 204–238 (2017). [CrossRef]  

15. A. Belmonte and J. M. Kahn, “Performance of synchronous optical receivers using atmospheric compensation techniques,” Opt. Express 16(18), 14151–14162 (2008). [CrossRef]  

16. S. Aghajanzadeh and M. Uysal, “Diversity–multiplexing trade-off in coherent free-space optical systems with multiple receivers,” J. Opt. Commun. Netw. 2(12), 1087–1094 (2010). [CrossRef]  

17. M. Li and M. Cvijetic, “Coherent free space optics communications over the maritime atmosphere with use of adaptive optics for beam wavefront correction,” Appl. Opt. 54(6), 1453–1462 (2015). [CrossRef]  

18. J. Park, E. Lee, C.-B. Chae, and G. Yoon, “Outage probability analysis of a coherent fso amplify-and-forward relaying system,” IEEE Photonics Technol. Lett. 27(11), 1204–1207 (2015). [CrossRef]  

19. A. A. Farid and S. Hranilovic, “Outage capacity optimization for free-space optical links with pointing errors,” J. Lightwave Technol. 25(7), 1702–1710 (2007). [CrossRef]  

20. H. G. Sandalidis, T. A. Tsiftsis, and G. K. Karagiannidis, “Optical wireless communications with heterodyne detection over turbulence channels with pointing errors,” J. Lightwave Technol. 27(20), 4440–4445 (2009). [CrossRef]  

21. A. García-Zambrana, C. Castillo-Vázquez, and B. Castillo-Vázquez, “Outage performance of MIMO FSO links over strong turbulence and misalignment fading channels,” Opt. Express 19(14), 13480–13496 (2011). [CrossRef]  

22. M. R. Bhatnagar and Z. Ghassemlooy, “Performance analysis of gamma–gamma fading FSO MIMO links with pointing errors,” J. Lightwave Technol. 34(9), 2158–2169 (2016). [CrossRef]  

23. J. Park, E. Lee, C.-B. Chae, and G. Yoon, “Impact of pointing errors on the performance of coherent free-space optical systems,” IEEE Photonics Technol. Lett. 28(2), 181–184 (2016). [CrossRef]  

24. A. Belmonte and J. M. Kahn, “Capacity of coherent free-space optical links using diversity-combining techniques,” Opt. Express 17(15), 12601–12611 (2009). [CrossRef]  

25. E. Zedini, H. Soury, and M.-S. Alouini, “Dual-hop FSO transmission systems over gamma–gamma turbulence with pointing errors,” IEEE Trans. Wirel. Commun. 16(2), 784–796 (2017). [CrossRef]  

26. I. S. Gradshteyn and I. M. Ryzhik, Table of integrals, series, and products (Elsevier, 2014).

27. M. Abramowitz and I. A. Stegun, Handbook of mathematical functions: with formulas, graphs, and mathematical tables, vol. 55 (Courier Corporation, 1965).

28. K. O. Odeyemi, P. A. Owolawi, and V. M. Srivastava, “Optical spatial modulation over gamma–gamma turbulence and pointing error induced fading channels,” Optik 147, 214–223 (2017). [CrossRef]  

29. Z. Wang and G. B. Giannakis, “A simple and general parameterization quantifying performance in fading channels,” IEEE Trans. Commun. 51(8), 1389–1398 (2003). [CrossRef]  

30. Y. Dhungana and C. Tellambura, “New simple approximations for error probability and outage in fading,” IEEE Commun. Lett. 16(11), 1760–1763 (2012). [CrossRef]  

31. Wolfram Research, Inc., “Mathematica,” URL http://functions.wolfram.com.

32. V. Adamchik and O. Marichev, “The algorithm for calculating integrals of hypergeometric type functions and its realization in REDUCE system,” in Proceedings of the international symposium on Symbolic and algebraic computation, (ACM, 1990), pp.212–224.

33. M. A. Amirabadi and V. T. Vakili, “A new optimization problem in FSO communication system,” IEEE Commun. Lett. 22(7), 1442–1445 (2018). [CrossRef]  

34. M. K. Simon and M.-S. Alouini, Digital communication over fading channels (John Wiley & Sons, 2005).

35. J. Park, E. Lee, C.-B. Chae, and G. Yoon, “Performance analysis of coherent free-space optical systems with multiple receivers,” IEEE Photonics Technol. Lett. 27(9), 1010–1013 (2015). [CrossRef]  

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 (3)

Fig. 1.
Fig. 1. BER performance of coherent FSO-SISO link in the composite channel with Rician turbulence and PEs, under (a) weak turbulence $SI = 0.1$, (b) strong turbulence $SI = 1.6$. To represent the impacts of PEs from negligible to severe, the $\xi$ is varied as $\xi = 6.7,1.4,0.7,0.4$. Analytical results (solid lines), asymptotic results (dashed lines), and numerical simulation (*) results are both included.
Fig. 2.
Fig. 2. BER performance under various compensation modes $(J = 1, 3, 6)$ over the composite Rician turbulence and PE fading channel. The $J = 1, 3, 6$ describes compensation for piston, tip-tilt, astigmatism, respectively. To represent general turbulence, moderate turbulence of $SI = 0.33$ is considered.
Fig. 3.
Fig. 3. Comparison of BER performance for MRC-SIMO link in cases of different number of receiver branches $(L = 1, 2, 4)$ under moderate turbulence $(SI = 0.33)$ with varied PEs. Two situations are involved, (a) without and (b) with modal compensation $(J = 3)$. The BER for the SISO link $(L = 1)$ is depicted as a benchmark.

Equations (29)

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

i ( t ) = i a c ( t ) + i n ( t )
P h a ( h a ) = 1 + K h a e K exp ( 1 + K h a h a ) I 0 [ 2 ( 1 + K ) K h a h a ]
f h p ( h p ) = ξ 2 A 0 ξ 2 h p ξ 2 1 0 h p A 0
f h ( h ) = f h h a ( h | h a ) f h a ( h a ) d h a
f h | h a ( h | h a ) = 1 h a h l ξ 2 A 0 ξ 2 ( h h a h l ) ξ 2 1 , 0 h A 0 h a h l
P h a ( h a ) = i = 1 N α i h a i 1 exp ( 1 + K h a h a )
f h ( h ) = ξ 2 h ξ 2 1 ( A 0 h l ) ξ 2 i = 1 N α i h A 0 h i h a ξ 2 + i 1 e 1 + K h a h a d h a
f h ( h ) = ξ 2 ( A 0 h l ) ξ 2 h ξ 2 1 i = 1 N α i ( 1 + K h a ) ξ 2 i Γ ( ξ 2 + i , 1 + K A 0 h l h a h )
f h ( h ) = ξ 2 A 0 h l i = 1 N α i ( 1 + K h a ) 1 i G 1 , 2 2 , 0 [ 1 + K A 0 h l h a h | ξ 2 ξ 2 1 , i 1 ]
P e = 0 + P e ( γ 0 | h ) f h ( h ) d h
B E R = ξ 2 2 π A 0 h l 1 γ 0 i = 1 N α i ( 1 + K h a ) 1 i G 3 , 3 2 , 2 [ 1 + K A 0 h l h a γ 0 | 0 , 0.5 , ξ 2 ξ 2 1 , i 1 , 1 ]
B E R = 1 2 i = 1 N α i { ζ 2 Γ ( i + 0.5 ) ( i ζ 2 ) i π ( A 0 γ 0 ) i [ 3 F 2 ( i ζ 2 , i , i + 0.5 ; i ζ 2 + 1 , i + 1 ; 1 A 0 γ 0 1 + K h a ) ] + Γ ( i ζ 2 ) Γ ( ζ 2 + 0.5 ) π ( A 0 γ 0 ) ζ 2 ( 1 + K h a ) ζ 2 i }
G 1 , 2 2 , 0 [ z | b 1 , b 2 a 1 ] Γ ( b 2 b 1 ) Γ ( a 1 b 1 ) z b 1 + Γ ( b 1 b 2 ) Γ ( a 1 b 2 ) z b 2 + o ( z )
a = { ξ 2 e K ( ξ 2 1 ) ( 1 + K ) A 0 h l h a , ξ > 1 ξ 2 e K Γ ( 1 ξ 2 ) ( 1 + K A 0 h l h a ) ξ 2 , ξ < 1
b = { 0 , ξ > 1 ξ 2 1 , ξ < 1
f h 0 ( h ) = { ξ 2 e K ( ξ 2 1 ) ( 1 + K ) A 0 h 1 h a ,   ξ > 1 ξ 2 e r Γ ( 1 ξ 2 ) ( 1 + K A 0 h l h a ) ξ 2 h ξ 2 1 , ξ < 1
B E R S I S O , a s y m = { Γ ( 1.5 ) e K ξ 2 2 π ( ξ 2 1 ) ( 1 + K ) ( A 0 h l h a ) γ 0 1 ,   ξ > 1 e K Γ ( 1 ξ 2 ) Γ ( ξ 2 + 0.5 ) 2 π ( 1 + K A 0 h l h a ) ξ 2 γ 0 ξ 2 , ξ < 1
G c , S I S O = { ( e K ξ 2 Γ ( 1.5 ) 2 π ( ξ 2 1 ) ( 1 + K ) ( A 0 h l h a ) ) 1 ,   ξ > 1 ( e K Γ ( 1 ξ 2 ) 2 π ( 1 + K A 0 h l h a ) ξ 2 ) 1 / ξ 2 , ξ < 1
G d , S I S O = { 1 , ξ > 1 ξ 2 , ξ < 1
Loss ( dB ) = 10 log 10 ( ζ 2 A 0 ( ζ 2 1 ) ) , ξ > 1
B E R = 1 π 0 π 2 MGF M R C ( s = 1 sin 2 θ ) d θ = 1 π 0 π 2 l = 1 L MGF γ l ( s = 1 sin 2 θ ) d θ
f γ i ( γ l ) = ξ 2 A 0 h l γ o i = 1 N α i ( ( 1 + K ) h a ) 1 i G 1 , 2 2 , 0 [ ( 1 + K ) γ l A 0 h l h a γ o | ξ 2 ξ 2 1 , i 1 ]
M G F γ i ( s ) = ξ 2 s A 0 h l γ o i = 1 N α i ( ( 1 + K ) h a ) 1 i G 2 , 2 2 , 1 [ ( 1 + K ) s A 0 h l h a γ o | 0 , ξ 2 ξ 2 1 , i 1 ]
B E R = 1 π 0 π 2 l = 1 L { ξ 2 A 0 h l γ o i = 1 N α i ( ( 1 + K ) h a ) 1 i G 2 , 2 2 , 1 [ ( 1 + K ) sin 2 θ A 0 h l h a γ o | 0 , ξ 2 ξ 2 1 , i 1 ] } d θ
M G F γ ( s ) = { [ ξ 2 e K ( 1 + K ) ( ξ 2 1 ) A 0 h 1 h a γ o ] L s L + o ( s ) , ξ > 1 e r Γ ( 1 ξ 2 ) Γ ( 1 + ξ 2 ) ( ( 1 + K ) A 0 h l h a γ o ) ξ 2 ] L s L ξ 2 + o ( s ) , ξ < 1
f M R C , a s y m ( h ) = { 1 ( L 1 ) ! [ ξ 2 e K ( 1 + K ) ( ξ 2 1 ) A 0 h l h a ] L h ( L 1 ) , ξ > 1 1 Γ ( L ξ 2 ) [ ξ 2 e r Γ ( 1 ξ 2 ) Γ ( ξ 2 ) ( ( 1 + K ) A 0 h l h a γ o ) ξ 2 ] L h L ξ 2 1 , ξ < 1
B E R M R C , a s y m = { 1 2 π L ! [ ξ 2 e K ( 1 + K ) ( ξ 2 1 ) A 0 h l h a ] L Γ ( L + 0.5 ) ( γ 0 ) L , ξ > 1 Γ ( L ξ 2 + 0.5 ) 2 π Γ ( L ξ 2 + 1 ) [ e r Γ ( 1 ξ 2 ) Γ ( ξ 2 + 1 ) ( ( 1 + K ) A 0 h l h a γ o ) ξ 2 ] L γ 0 L ξ 2 ,   ξ < 1
G c , M R C = { ( Γ ( L + 0.5 ) 2 π L ! [ ξ 2 e K ( 1 + K ) ( ξ 2 1 ) A 0 h l h a ] 1 L , ξ > 1 ( Γ ( L ξ 2 + 0.5 ) 2 π Γ ( L ξ 2 + 1 ) [ e r Γ ( 1 ξ 2 ) Γ ( ξ 2 + 1 ) ( ( 1 + K ) A 0 h l h a γ o ) ξ 2 ] 1 / ( L ξ 2 ) , ξ < 1
G d , M R C = { L , ξ > 1 L ξ 2 , ξ < 1
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.