Open Access
8 December 2023 Optical communications in turbulence: a tutorial
Author Affiliations +
Abstract

We provide a tutorial on how to create link budgets and bit error rate (BER) and probability of fade calculations for optical communications systems designed to operate in the turbulent channel. It reviews the characterization models necessary for either incoherent or coherent free space optical communications uplink, downlink, and horizontal system analyses in the turbulent channel. Beam wander, scintillation, and receiver noise variance as well as pointing and tracking effects were included in this paper. Comparisons among these models, computer simulations and field measurements are provided throughout the paper. Good agreement is shown among all. An example analysis was provided using this information. The conclusion is that no matter whether the system is incoherent or coherent and/or which signaling format is used, the scintillation index will peg the BER to constant value at high signal-to-noise ratio well above the desired value. Other turbulence mitigation techniques must be employed to get the desired system performance.

1.

Introduction

Optical communications (OC) have attracted new interests in the recent decades as high data rate airborne- and satellite-based backhaul capabilities for large capacity communication networks.18 Although radio frequency (RF) systems have dominated the communications backhaul market, decreasing RF spectrum availability and network security make OC systems more attractive. Specifically, OC systems have reemerged after three decades of dormancy because of new technologies recently available from the fiber optic communications (FOC) community, such as fiber lasers and high-speed detectors. The result is that these systems provide some inherent advantages over RF communications, e.g., compact terminal size, low power consumption, substantial increase in bandwidth, unlicensed optical spectrum, and no electromagnetic interference.1,2 However, optical signals are subject to various undesirable phenomena, including turbulence-induced signal fading, beam wander effects, and noise caused by light due to sunlight and sky radiance.3 Fortunately, turbulence mitigation under high turbulence conditions has been successfully achieved using new system and network technologies.7,8 The question is why some folks are successful in mitigating turbulence effects while others are not.

In most of the early communications link analyses, researchers only paid attention to signal degradation and produced ways of potentially reducing or eliminating phase perturbations to reduce scintillation and increase signal power. Techniques such as aperture averaging and emerging technologies such as adaptive optics were employed with some success.911 Aperture averaging worked well for large receiver apertures and downlink/horizontal geometries. Uplinks experienced large beam wander that aperture averaging could not help reducing. Adaptive optics worked well for weak turbulent conditions or short ranges but proved ineffective in moderate to strong turbulent conditions.12 This is because these turbulent conditions are dominated by turbulent phase fluctuations while moderate to strong conditions or longer ranges are dominated by amplitude scintillation that needs more than just phase correction.912 What was missing from these analyses was the turbulent-intensity noise variance that also affected communications performance. This omission was the reason many of the analyses did not match experimental field results. Andrews and Phillips recognized this void in Ref. 9 and produced a signal-to-noise ratio (SNR) correction to account for this additional noise term. However, they did not address how this impacted the probability bit error/bit error rate (BER), or the fact that turbulence did not obey Gaussian statistics questioning whether the addition was valid as derived.

This paper is a tutorial on optical communications in turbulence that resolves these issues and shows how link budget analyses should be done. It will validate Andrews and Phillips’s SNR correction and that its inclusion in the SNR equation dramatically affects communications system performance. Specifically, this paper will begin with a discussion of the origin of the SNR, then move to the validity of the Andrews and Phillips’s SNR correction and establishment of the scintillation index as the key turbulent-intensity noise variance. In addition, this paper will develop the scintillation indices for uplink, downlink, and horizontal communications link geometries, which also include tracking and untracked receiver effects. An example analysis was provided using this information. The key aspect of this tutorial is that no matter whether the system is incoherent or coherent and/or which signaling format is used, the scintillation index will peg the BER to constant value at high SNR well above the desired value. Other turbulence mitigation techniques must be employed to get the desired system performance.

2.

Electrical Signal-to-Noise Ratio and BER in Incoherent Optical Communications

Statistical detection and estimation theory revolutionized the fields of RF communications and remote sensing, beginning with S. O. Rice’s seminal paper entitled “The Mathematical Analysis of Random Noise.”13 However, the benefit of this theory to electro-optical systems was not recognized until the 1960s. Mandel14 and Helstrom15 published arguably the first key papers in the area and Papoulis16 came out with the first key textbook in the area. The effects of turbulence on communications and remote sensing emerged in the literature around the same time, e.g., Ref. 17. Andrews and Phillips were one of the first to summarize the research characterizing the effects of turbulence on laser radar and communications system performance,9,10 and Andrews and Beason recently updated our understanding of system performance effects in these areas.11

This section discusses the origin of the SNR, the validity of the Andrews and Phillips’s SNR correction, and the scintillation index as the key turbulent-intensity noise variance.

2.1.

Origin of the Signal-to-Noise Ratio

Radar research during World War II showed that system noise negatively affected a receiver’s output signal, hence, a target’s detectability. This led to matched filter detection and SNR as the means and measure to quantify the signal level above the noises.18 In 1953, Middleton advocated the detection of signals using statistical decision theory,19 whereas Peterson et al.20 advanced this approach with the likelihood ratio shortly thereafter. This section follows Ref. 21, pp. 102–106.

The SNR is derived from determining the signal measurement x that will be used to decide between two hypotheses, H0 (null hypothesis) and H1 (signal-present hypothesis). Considering that any measurement can create errors, we need to treat the outcome of any decision strategy as a random variable that is governed by the probability densities functions (PDF) p0(x) and p1(x) for hypotheses H0 and H1, respectively. In general, the PDFs for hypotheses H0 and H1 can be written as

Eq. (1)

pj(x)=12πn0e(xμj)2/2n0j=0,1,
respectively. In the above equation, the parameters {μj;j=0,1} are the means assuming μ0<x<μ1, and n0 is the unilateral spectral noise density for Gaussian PDFs of the two possible hypotheses. Figure 1(a) shows a comparison of the two PDFs.

Fig. 1

(a) PDFs for hypotheses H0 and H1, pinpointing mean levels, μ0 and   μ1, respectively, and (b) same PDFs with threshold xT.

OE_63_4_041207_f001.png

Hypothesis H0’s PDF implies that the measured value for x comes from the noise distribution and has an averaged value of μ0 and variance n0. Equation (1) implies that x will vary around the averaged value most of the time (values near the peak), but also will occasionally be extremely different (values at the tails). One will see similar measurement deviations occurring with hypothesis H1’s PDF, but around an average value of μ1, rather than μ0. The question one might ask is how often each PDF occurs? The fraction of time that H0 happens is ζp and is called the prior probability for this hypothesis. Alternately, the fraction of time that H1 happens is consequently (1ζp) and is called the prior probability for that hypothesis.

Given all the above, what is the strategy for deciding between hypotheses H0 and H1, assuming a measurement x? The obvious choice is that a decision maker picks a value of x, say xT, and then chooses hypothesis H0 when xxT and hypothesis H1 for x>xT [see Fig. 1(b)]. That makes senses until you wonder, “How did you pick xT?” Obviously, the presence of noise causes mistakes to be made in the decision process. This suggests that the decision threshold is the value of xT to minimize any negative effects caused by making mistakes. How do we determine the desired value for xT?

The Neyman–Pearson decision strategy is designed for deciding when hypothesis H1 occurs rarely and H0 dominates the decision process, which is normally true for sonar, radar, optical, and electro-optical systems doing real world surveillance. This strategy essentially comes down to maximizing the probability of detection of a target while minimizing the false-alarm probability. The probability of detection (Pd) is given as

Eq. (2)

Pd=1Q1=1xTp1(x)dx=xTp1(x)dx,
and the probability of false alarm (FA) equals

Eq. (3)

PFA=Q0=xTp0(x)dx.

The parameters Q0 and Q1 represent “the error of the first kind” and the “error of the second kind,” respectively. Referring again to Fig. 1(b), Q0 is the shaded area under the p0(x) curve to the right of xT and Q1 is the shaded area under the p1(x) curve to the left of xT.

To implement this strategy based on a single measurement x, we calculate the likelihood ratio Λ(x)=p1(x)/p0(x) and compare the result to some threshold value ΛT. If Λ(x)ΛT, the decision maker picks hypothesis H0; if Λ(x)>ΛT, the decision maker selects hypothesis H1. The computed likelihood ratio is a random variable and has a probability density function (PDF) P0(Λ) under hypothesis H0. It is related to the known PDF as follows:

Eq. (4)

P0(Λ)dΛ=p0(x)dx.

The probability PFA therefore can be written as

Eq. (5)

PFA=ΛTP0(Λ)dΛ.

The value of ΛT is established by pre-assigning a value to Q0 and inverse computing it. The probability of detection Pd is then determined by calculating

Eq. (6)

Pd=ΛTP1(Λ)dΛ,
where P1(Λ) is the PDF under hypothesis H1. The results of these computations generally are portrayed in single figure, plotting the detection probability (power of the test) as a function of the FA probability (size of the test). This curve is often called the “receiver operating characteristic” or “ROC” curve. The choice between the two possible hypotheses can be based not only on the likelihood ratio Λ(x) but also on monotonic function ϒ=ϒ(Λ(x)) of the likelihood ratio, called the “test statistic.” This implies that

Eq. (7)

PFA=ϒTP0(ϒ)dϒ,
and

Eq. (8)

Pd=ϒTP1(ϒ)dϒ.

As noted in the last section, the best SNR comes from matched-filtering the incoming signal to the initial transmitted waveform. If we assume a known (deterministic) signal, say s(t), to be detected in additive white Gaussian noise, then our desired test statistic can be written as

Eq. (9)

ϒ=0Ts(t)x(t)dt,
where ϒ is the decision test statistic, T is the receiver integration time, and v(t) is the input voltage. Using this statistic, we find that

Eq. (10)

E{ϒ|H1}=0Ts(t)x(t)dt|v(t)=s(t)=0T|s(t)|2dt=Es,
and

Eq. (11)

E{ϒ|H0}=0Ts(t)x(t)dt|v(t)=n(t)=0,
where Es is the energy of the received signal and n(t) is the receiver noise in the absence of an incoming signal with a variance proportional to N0/2. (The average noise level n(t)=0.) The variance of ϒ is given as

Eq. (12)

Varϒ=n0=0T0Ts(t1)s(t2)n(t1)n(t2)dt1dt2

Eq. (13)

=N020T0Ts(t1)s(t2)δ(t2t1)dt1dt2

Eq. (14)

=N020T|s(t)|2dt=N0Es/2.

Using the above equations, we can write the PDFs for hypotheses H1 and H0 as

Eq. (15)

p1(ϒ)=1πN0Ese(ϒEs)2/N0E,
and

Eq. (16)

p0(ϒ)=1πN0Eseϒ2/N0E,
respectively.

The FA and detection probabilities then are given as

Eq. (17)

PFA=Q(xT),
and

Eq. (18)

Pd=Q(xTSNR),
respectively, where

Eq. (19)

Q(y)=12πyeq2/2dq
is the Q-function [=0.5erfc(x/2), with erfc(x) being the complementary error function],

Eq. (20)

xT=ϒ02/N0Es,
is the decision threshold, and

Eq. (21)

SNR=[E{ϒ|H1}]2Varϒ=2EsN0.

There is an important linkage to be made between energy and information transfer using Eq. (21). In his seminal paper on communications, Shannon22 assumed a signal space defined by a dimension 2BeT (defined as the number of Nyquist samples of a function band-limited to electrical bandwidth Be=1/2T, and observed over the time T). This means one can rewrite Eq. (21) as

Eq. (22)

SNR=2EsN0=(RbTRbT)2EsN0=(Rb2T)EbN0=RbBeEbN0=rEbN0,
where Rb is the communications data rate in bits per second (bps), Eb=Es/RbT is the energy per bit in joules per bit, r=Rb/Be=log2(Msymbol) is the system’s spectral efficiency in bits per second per Hertz (bps/Hz), and Msymbol is the number of symbols in the communications alphabet.

2.2.

Signal-to-Noise Ratio in a Direct Detection System

In most communications system analyses, Eq. (21) is written as

Eq. (23)

SNRe=2EsN0(2T2T)=Es/TN0Be=PsPn=is2in2=Rλ2P¯rec2in2,
where Ps is the average electrical signal power, Pn is the average electrical noise power, is is the average signal current, in is the total average noise current, Rλ is the responsivity of the receiver’s photodetector in amperes per watt (A/W), and P¯rec is the incoming signal mean power impinging an optical detector, which includes all the diffraction, refraction, and any atmospheric loss. The incoming received power is given as

Eq. (24)

PrxγtxγrxPtx[AtxArx(λR)2]=γtxγrxPtxFSL,
where γtx is the transmitter transmittances, γrx is the receiver transmittances, FSL=AtxArx/(λR)2 is the Fraunhofer spreading loss (FSL), Ptx is the transmitter laser power, Atx=πDtx2/4, Dtx is the transmitter aperture diameter, Arx=πDrx2/4, Drx is the receiver aperture diameter, λ is the laser wavelength, and R is the link range. Let us now talk about the detected average noise power. Equation (24) is called the Friis optical range equation.

In RF communications, one only deals with thermal noise within the RF receiver. Its spectral density is given by N0=kBT0, where kB=1.38×1023  J/K is Boltzmann’s constant, and T0 is the receiver temperature in degrees Kelvin [K]. It obeys Gaussian statistics. One also has thermal noise to contend with in optical receivers, which obeys Gaussian statistics like in the RF case. Specifically, the thermal noise current PDF is given as

Eq. (25)

Pth(P)=e(PP¯th)2/2σth2,
where P is the received power by the detector, P¯th is the average thermal power, and σth2=4kbT0Be is the thermal noise power variance. In an optical receiver, one generally creates a photo-electron current i(t) that translates into electrical power via the equation P(t)=i2(t)RL, where RL is the receiver’s load resistor in ohms.

On the other hand, dark current in an optical receiver is a Poisson process. It can be shown that the dark current Poisson PDF is written as

Eq. (26)

PDC(K)=μDCKeμDCK!,
where μDC=2qiDCRLBe, q=1.602×1019 coulombs (C) is the charge on an electron, and iDC is the receiver’s mean dark current.

For large K, Eq. (26) tends toward a Gaussian PDF given as

Eq. (27)

PDC(KP)12πμDCe(PP¯DC)2/2μDC.

Mathematically, this can be shown using Stirling’s equation (Ref. 3, pp. 329–330). Figure 2 shows Eq. (26) as a function of K for values of μDC=P¯th=30,50, and 100. From this figure, one can see that μDC30 electrons, which is typical for most systems, creates Poisson and Gaussian PDFs that match well.

Fig. 2

Comparison of Poisson and Gaussian PDFs as a function of K for various values of μDC.

OE_63_4_041207_f002.png

Now, both thermal and dark current exist in an optical receiver whether an incoming signal is present or not. They are created by separate, independent physical mechanisms, so they are independent random variables in the photo-electron creation process (chapter 8 in Ref. 3). This means that the combined effect of both is derived from the “sum of random variables” theory.23 The combined PDF comes from the convolution of Eqs. (24) and (26). The result is another Gaussian PDF with a mean of P¯thermal+P¯DC and a variance of 4kbT0Be+2qiDCRLBe.

Now, both thermal and dark current exist in an optical receiver whether an incoming signal is present or not. They are created by separate, independent physical mechanisms, so they are independent random variables in the photo-electron creation process (chapter 8 in Ref. 3). This means that the combined effect of both is derived from the “sum of random variables” theory.23 The combined PDF comes from the convolution of Eqs. (25) and (27). The result is another Gaussian PDF with a mean of P¯thermal+P¯DC and a variance of 4kbT0Be+2qiDCRLBe.

Now, what happens when external light enters an optical receiver? One gets “signal-shot-noise” in addition to the dark current and thermal noise in the receiver. All photodetectors create this photon fluctuation noise, even when the light input is constant in intensity. This is quantum mechanical effect and comes from the random emission of photoelectrons derived from the external light absorption by the detector. Like dark current, it is a Poisson process and tends toward Gaussian statistics as the mean current increases and becomes appropriate as an approximation around 30 photo-electrons. The two primary external light sources are the transmitted signal beam and scattered solar radiation. Again, using the “sum of random variables” theory, the resulting Gaussian PDF has variance

Eq. (28)

σtotal2=2qRλP¯recRLBe+2qRλP¯bkgRLBe+2qiDCRLBe+4kbT0Be,
and mean

Eq. (29)

μtotal=P¯th+P¯DC+P¯rec+P¯bkg,
where P¯rec is the received optical solar background mean noise power before the detector. Let us look at direct detection (DD) communications receiver.

Figure 3 illustrates a basic DD receiver layout with a p-type—intrinsic—n-type (PIN) photodiode. It has an incoming signal Ps(t) impinges an optical detector and is changed into current that get electrically amplified in an RF amplifier. At this stage, one denotes its gain by G, its quantum efficiency η, and its load resistor as RL. This current is then subjected to post-detection filtering with impulse response hrf(t) and effective electrical bandwidth Be. This filtered result then passes through sample-and-hold circuitry before yielding the result, communications data.

Fig. 3

Schematic of a DD optical receiver.

OE_63_4_041207_f003.png

The electrical SNR for this type of system is given as

Eq. (30a)

SNRe=G2Rλ2RLPrec22qBeG2[Rλ(Prec+Pb)+iD]RL+4kBT0FrfBe

Eq. (30b)

=Prec22qBe[(Prec+Pb)/Rλ+iD/Rλ2]+4kBT0FrfBe/G2Rλ2RL,
where Frf is its amplifier’s noise factor. Looking at Eq. (18), we see that that the probability of detection is a function of the square root of the electrical SNR, which is true in general. The optical community likes to use the optical signal to noise ratio (OSNR) in their link budget analyses, which is defined as

Eq. (31)

OSNR=SNRe=Prec2qBe[(Prec+Pb)/Rλ+iD/Rλ2]+4kBT0FrfBe/G2Rλ2RL.

We will use this convention for the rest of the paper. This means that Eq. (18) can be rewritten as

Eq. (32)

Pd=Q(xTOSNR).

On the other hand, if one has an avalanche photodiode (APD) detector instead of the PIN photodiode, the optical SNR is written as

Eq. (33)

OSNR=Prec2qFAPDBe[(Prec+Pb)/Rλ+iD/Rλ2]+4kBT0FrfBe/G2Rλ2RL,
where FAPD is the excess noise factor associated with the variation in gain G. The excess noise factor equals

Eq. (34)

FAPD(G,kion)=G[1(1kion)(G1G)2],
where kion is the APD’s effective ionization coefficient ratio. Table 1 depicts excess noise factors for various detector material types.

Table 1

Excess noise factor for various detector material types.

Detector typeIonization ratioTypical gainExcess noise factor
Silicon “reach-through”0.021504.9
Silicon “SLiK low k”0.0025003.0
Germanium0.9109.2
InGaAs0.45105.5

2.3.

Q-Parameter

Figure 1 showed a typical data input scenario for the detection threshold decision process. An incoming signal is determined via some mechanism during a synchronized time interval T1 to establish whether either a “1” or a “0” is transmitted. The BER is the number of bit errors per unit time. That is, the number of bit errors divided by the total number of transferred bits during a studied time interval. It is a unit-less performance measure, often expressed as a percentage. On the other hand, the bit error probability PBE is the expected value for the BER. The BER can be considered as an approximate estimate of the bit error probability. This estimate is accurate for a long-time interval and a high number of bit errors measured. We will not distinguish between the two in the discussions to come as most people consider them synonymous.

If the optical receiver is a maximum a posteriori (MAP) receiver, then the BER is given as

Eq. (35)

BER=ζpPr{1|0}+(1ζp)Pr{0|1},
where ζp and (1ζp) are the a priori probabilities of receiving a “1” or a “0,” respectively, and Pr{1|0} and Pr{0|1} are conditional probabilities representing the respective probability of falsely declaring the first argument true, when the second argument actually is true. If we assume equal a priori probabilities, then Eq. (35) reduces to

Eq. (36)

BER=0.5Pr{1|0}+0.5Pr{0|1}.

Let us assume the photo-current for the bits “1” and “0” equals i1 and i0, respectively. The total variance associated with each of these currents then is written as

Eq. (37)

σ12=σ1ss2+σ1-thermal2,
and

Eq. (38)

σ02=σ0ss2+σ0-thermal2,
respectively. Since the average photo-current is different for the bits “1” and “0,” the associated average shot noise current and its variance will be different, i.e., i1>i0, which implies that σ12σ02. Let ith represent the decision threshold. The conditional probabilities can be written as

Eq. (39)

Pr{1|0}=12πσ02ithe(ii0)2/2σ02di=0.5erfc((ithi0)/2σ02),
and

Eq. (40)

Pr{0|1}=12πσ12ithe(ii1)2/2σ12di=0.5erfc((i1ith)/2σ12),
under the Gaussian statistics assumption. Substituting Eqs. (39) and (40) into Eq. (36) yields

Eq. (41)

BER=0.25erfc((ithi0)/2σ02)+0.25erfc((i1ith)/2σ12).

The current threshold that minimizes the BER given in Eq. (41) is

Eq. (42)

ithσ0i1+σ1i0(σ1+σ0),
assuming the system’s thermal or pre-amplifier noise is the largest components in both σ12 and σ02.

Substituting Eq. (42) into either of the two arguments contained in Eq. (41), we create what is called the Q-parameter, which is given as

Eq. (43)

Q(i1i0)(σ1+σ0).

The Q-parameter is a measure of the OSNR. Using Eq. (43), Eq. (41) reduces to

Eq. (44)

BER0.5erfc(Q/2)1Q2πeQ2/2,
where the approximation in Eq. (44) is valid for Q>3. Figure 4 shows this BER as a function of the Q-parameter. For a BER=1012, Q=7 or 8.45 dB.

Fig. 4

BER versus Q-parameter.

OE_63_4_041207_f004.png

2.4.

Receiver Sensitivity

Let us assume direct-detection communications using a PIN photodiode. Without loss of generality, we will let i0=0, which translates into the zero-signal power transmitted for a bit “0,” G=1 (normally true for PIN-detectors), Pb=0, and no atmospheric loss.

In a 10 Gigabit per second (Gbps) intensity-modulated/DD PIN receiver, the key noise sources typically are thermal, dark current-induced, and signal-shot noises. Using Eq. (43), the associated Q parameter is given as

Eq. (45)

QIM=2RλPrec2qBe[2RλPrec+iD]+4kT0FrfBe/RL+2qBeiD+4kT0FrfBe/RL.

Figure 5 plots this Q-parameter in decibels as a function of received average optical signal power for a 10 Gbps PIN system with λ=1.55  μm, Rλ=0.85  A/W, RL=50  Ω, iD=5  nA, Frf=1, T0=300  K, and Be=7.5  GHz. If the system requirement is a BER=1012, this figure shows that the associated Q-factor of 8.45 dB implies a minimum average power Pavgmin to achieve that BER is 19  dBm (decibel-milliwatt) for the system parameters given. This minimum average power Pavgmin is called the receiver sensitivity. From this plot, we also see that a dB reduction in minimum average signal power decreases the Q-parameter by a dB. Let us look at an emerging system approach that has some benefits to turbulence mitigation that we see later in the paper.

Fig. 5

Receiver sensitivity versus average optical signal power for a 10 Gbps PIN system.

OE_63_4_041207_f005.png

To improve communications link performance, some optical engineers add an optical pre-detection optical amplifier, such as an erbium-doped fiber amplifier (EDFA), in front of the PIN photodiode. An EDFA increases the incoming received signal but also will create optical input noise at the detector. The induced noise comes from both spontaneous and stimulated emissions in the fiber amplifier but can be reduced by optical filtering before the detector. Unfortunately, it cannot be reduced totally.24 We now find the three optical noise powers (variances) are

  • (1) Signal-shot noise

    Eq. (46)

    σshot2=2qBeGiavgRL2qBeG[2ηPavg/hν].

  • (2) The signal-spontaneous beat noise variance, which is given as

    Eq. (47)

    σsig-sp2=2GiavgispRLBe/B0,

    Eq. (48)

    =2hνNFoptG(G1)(2Rλ2Pavg)RLBe,
    with NFopt=2nsp being the EDFA optical noise figure and nsp being the EDFA inversion parameter, and (3) the spontaneous-spontaneous beat noise variance, which is equal to

    Eq. (49)

    σsp-sp2=isp2RL[Be(2BoBe)]/(2B02)

    Eq. (50)

    =0.5[RλhνNFopt(G1)]2RLBe(2B0Be),
    where B0 is the optical bandwidth of the receiver.24 Here, iavg=qηPavg/hν=RλPavg, and isp is the photocurrent generated by the spontaneous emissions (including both polarizations) at the output of the detector, which equals

    Eq. (51)

    isp=2qηnspB0(G1)=qηNFoptB0(G1)=RλhνNFoptB0(G1).

The resulting Q-parameter is then

Eq. (52)

QERDA=2RλGPrec2qBe[2RλGPrec+iD]+σsig-sp2+σspsp2+4kT0FrfBe/RL+2qBeiD+σsp-sp2+4kT0FrfBe/RL.

The product GPavg is the amplified average optical signal power necessary for “1” bit. Figure 6 plots receiver sensitivity as a function of average optical signal power for a 10 Gbps EDFA-PIN system for noise variance conditions. Here, we have GPavg=0  dBm, QERDA=7, λ=1.55  μm, Rλ=0.85  A/W, RL=50  Ω, iD=5  nA, Frf=1, T0=300  K, Be=7.5  GHz, B0=25  GHz, and nsp=1.58 [NFopt5  dB]. (The dark-current Q-factor is 54  dB and is not shown on the chart.)

Fig. 6

Receiver sensitivity versus average optical signal power for a 10 Gbps EDFA system.

OE_63_4_041207_f006.png

It is clear from this figure that the receiver sensitivity Psmin40  dBm for the system parameters given. This is >20  dB better than PIN diode system shown above. It also shows that thermal and signal-shot noises are within a couple of dBs of each other, but much weaker than either stimulated-spontaneous and spontaneous-spontaneous beat noises. In addition, we see that a dB reduction in average signal power decreases the Q-factor by around 2/3 of a dB. Finally, the figure indicates that the stimulated-spontaneous beat noise becomes much stronger than the other noises sources when the average signal power exceeds 45  dBm.

Looking at the results in Fig. 6, the ERDA-PIN system Q-parameter can be written as

Eq. (53)

QEDFA2RλGPrec2σsig-sp2=RλGPrec2hνNFoptG(G1)(2Rλ2Pavg)BePrec2hνNFoptPavgBe=ηPrec2hνBe[1ηNFopt]=SNRQLηNFopt,
for G15. It is easy to show that

Eq. (54)

SNREDFA=SNRQLηnsp=QEDFA2,
and the receiver sensitivity is

Eq. (55)

Pavgmin=QEDFAσsigsp2RλG.

The asymptotic behavior of the electrical SNR is shown in Fig. 7, which is an illustration of the various noise powers and the electrical SNRs from classical calculations. Here, we have Pavg=30  dBm, QERDA=7, λ=1.55  μm, Rλ=0.85  A/W, RL=50  Ω, iD=5  nA, Frf=1, T0=290  K, Be=7.5  GHz, B0=75  GHz, and nsp=1.58 [NFopt5  dB]. This figure clearly shows the dominate behavior of signal-spontaneous beat noise variance for G15.

Fig. 7

Various noise powers and SNRs as a function of EDFA amplifier gain.

OE_63_4_041207_f007.png

2.5.

Electrical Signal-to-Noise Ratio in Free Space with Link Tracking Errors

This section provides the basic equations for analyzing link performance of an optical communications system experiencing pointing and tracking errors induced by the link’s transmitter and receiver platforms. This section follows Refs. 25 and 26.

2.5.1.

Received power equation

The average received optical signal power for a satellite communications cross-link in the presence of jitter can be written as

Eq. (56)

PrxγtxγrxPtx[AtxArx(λR)2]LjitterLpp=γtxγrxPtxFSLLjitterLpp,
where

Eq. (57)

Ljitter=θd2/(θd2+σ2)Jitter Loss Factor,

Eq. (58)

θd=λ/DtxTransmitter beam spread,
σrms jitter error, and LppPower Penalty (because of jitter).

2.5.2.

Bit error rate in the presence of platform jitter

Pointing errors have a significant effect on link performance and the BER. It can be shown that the PDF for the normalized received optical irradiance p(I) is a beta distribution and is given as

Eq. (59)

p(I)=βIβ1,0I1,
where I=I/I¯ is the normalized intensity, and

Eq. (60)

β=θd2/4σ2.

The average intensity, I¯, is given as

Eq. (61)

I¯=β/(β+1).

The unconditional BER in the presence of jitter should be averaged with respect to the above PDF to yield the average BER; specifically,

Eq. (62)

BER¯=01pj(I)BER(QrII¯)dI

Eq. (63)

=Qr(β+1)201Iβ1erfc(QrII¯2)dI,
for an NRZ OOK link with a Q-parameter specification Qr as an example. Figure 8 shows the degradation of the BER¯ as a function of Qr and the ratio of beam θd divergence to random jitter σ.

Fig. 8

Average BER as a function of the required SNR parameter and the ratio of beam θd divergence to random jitter σ.

OE_63_4_041207_f008.png

The power penalty Lpp referenced previously results from the variation in received signal irradiance as it affects the average BER, which is defined as

Eq. (64)

Lpp=[Q|BER(Q)=aBERQr|BER¯(Qr)=aBER],
where aBER represents the system BER specification. Figure 9 shows the power penalty Lpp as a function of θd/σ for several specific values of aBER.

Fig. 9

The power penalty Lpp as a function of the ratio of beam θd divergence to random jitter σ.

OE_63_4_041207_f009.png

As a final note, one would like to minimize the power penalty in any system design. Table 2 gives the ratio for θd/σ to minimize the power penalty for a collection of BERs.

Table 2

The optimum ratio of beam θd divergence to random jitter σ for a set of desired BER.

BER(θd/σ)opt
1024.12431
1034.85198
1045.47376
1056.02693
1066.53065
1076.9964
1087.43177
1097.84207
10108.231236
10118.60228
10128.95751

2.5.3.

Optical tracking channel

The fade statistic is needed for the optical tracking channel. Here, we use the probability of fade, which describes the percentage of time the irradiance of the received wave is below a prescribed threshold value (Ref. 10, p. 95).

As noted in Eq. (59), the PDF of the received irradiance is a beta distribution. The average received optical power for the tracking sensor is given as

Eq. (65)

Prects=γtsγrsPtsFSLtsLjitter.

The fade level FT from the average level equal

Eq. (66)

FT=(β+1β)PF1/β,
where

Eq. (67)

PF=0FT(β/β+1)pj(I)dI=[FT(β/β+1)]β
is the allowable fade probability at the tracking sensor. The surge level ST similarly is defined as

Eq. (68)

ST=(β+1β)[1PS]1/β,
where

Eq. (69)

PF=FT(β/β+1)1pj(I)dI=1[ST(β/β+1)]β
is the allowable surge probability at the tracking sensor. Figures 10 and 11 show the cumulative fade and surge probabilities are plotted as a function of fade and surge levels, respectively.

Fig. 10

Cumulative fade probability as a function of fade levels for three ratios of beam θd divergence to random jitter.

OE_63_4_041207_f010.png

Fig. 11

Cumulative surge probability as a function of surge levels for three ratios of beam θd divergence to random jitter.

OE_63_4_041207_f011.png

The dynamic range required for the tracking sensor to cope with only random angular jitter is given as

Eq. (70)

DRjitter=STFT.

In practice, the variation of distance (due to orbital motion of the satellites) should be accounted for in the dynamic range, which changes Eq. (69) to

Eq. (71)

DRtotal=STFT[RmaxRmin]2,
where Rmax and Rmin are the maximum and minimum ranges, respectively, between the two satellites as they traverse their orbital paths. The transmitted laser power should be adjusted as the link range changes so the dynamic range of the tracking sensor is less than DRtotal.

2.5.4.

Satellite cross-link system link budget for a BER=1012

As an example of free space pointing and tracking, this section calculates the link budget for a 10 Gbps NRZ OOK satellite cross-link communications system with 5 dB link margin. Assume we have constellation of communications satellites located in 800 km circular orbits with an approximate 100 min revisit time. Let us further assume the number of satellites allow us to have dynamic intersatellite ranges between 2600 km and 4600 km. Table 3 illustrates an example of the optimum link budget for a cross-link in the presence of jitter. This table shows the link can close with a better than 5 dB margin.

Table 3

Example of an optimum link budget estimated for space-to-space optical tracking and communications cross-link.

ParameterUnitsTrackingCommunications
TX powerWatts1.000E+011.000E+01
TX powerdBm4.000E+014.000E+01
WavelengthMicrons1.550E-061.550E-06
TX aperture diameter =Meters6.656E-026.656E-02
TX aperture areaMeters-squared3.479E-033.479E-03
TX transmittancedB−4.560E+00−4.560E+00
Range =Meters4.600E+064.600E+06
RX aperture diameterMeters6.656E-026.656E-02
RX aperture areaMeters-squared3.479E-033.479E-03
RX transmittancedB−2.000E+00−2.000E+00
FSL =dB−6.623E+01−6.623E+01
TX beam divergenceRadians2.329E-052.329E-05
SigmaRadians2.600E-062.600E-06
Beam divergence/sigmaUnitless8.957E+008.957E+00
Pointing lossdB−2.113E-01−2.113E-01
Power penaltydB−2.099E+00−2.099E+00
Received powerdBm−3.510E+01−3.510E+01
Rcvr sensitivitydBm−4.100E+01−4.100E+01
MargindB5.898E+005.898E+00
Fade leveldB−1.027E+00
Surge leveldB2.090E-01
Max/min range ratio-squareddB4.956E+00
Total dynamic rangedB6.192E+00

Assuming we want a high quality 10 Gbps communications link, then we need an average R1012, which implies an optimum θd/σ is 8.5751 from Table 1. (In the above table, both the probabilities PF and PS are specified to be 102.) If the jitter equals 2.65  μrad, the beam divergence θd must be 23.3  μrad to obtain that value. This means that the pointing loss is 0.2  dB and the power penalty Lpp is 2.1  dB. Assuming we can close the link, what are our general pointing, tracking, and acquisition (PAT) requirements.

Residual radial pointing jitter on the order of <10% of the transmitter beam width (twice the beam divergence) delivers a received power versus time profile that supports reliable communications with an adequate average BER. To minimize random link BER degradation, the transmitting satellite must track the opposite platform with sub-microradian residual LOS jitter to keep the link’s radial LOS pointing jitter around 4.6  μrad or less. A good ball-park estimate for the total root-sum-squared (RSS) static and dynamic terminal error contributions for the fine tracking mode would be about 1/3 of the radial LOS pointing jitter, i.e., <1.5  μrad.

Acquisition (beacon) detection and handoff RSS terminal contributions nominally are 10’s of mrad and a few mrad, respectively, which are less stressing. A complete system design is required to accurately specify all these parameters, but the above gives a reasonable estimate for the PAT requirements necessary to facilitate cross-link communications.

2.6.

Electrical Signal-to-Noise Ratio in a Turbulent Channel

The turbulent channel range equation is a modification of the Friis range equation and is given as

Eq. (72)

PrecγtxγrxγfiberPtxLatm[AtxArx(λR)2]SRRPSRDPγrxγfiberI(0,R)ArxSRDP,
where

Eq. (73)

I(0,R)γtxPtxLatm[Atx(λR)2]SRRP,
is the average intensity at the receiver aperture. In the above equations,

Eq. (74)

Latm=0Reα(h0,ζ)r(h0,ζ)dr,

Eq. (75)

SRRP1[1+m(Dtx/r0t)5/3]6/5,m={1.000;  with no AO;σI2(0,R)>10.280;  with Tip/Tilt only;0.1σI2(0,R)10.052;  with Full AO;σI2(0,R)<0.1,

Eq. (76)

SRDP1[1+m(Drx/r0r)5/3]6/5,m={1.000;  with no AO;σI2(0,R)>10.280;  with Tip/Tilt only;0.1σI2(0,R)10.052;  with Full AO;σI2(0,R)<0.1,

Eq. (77)

r0rx=[16.71  sec(ζ)0RCn2(r)(r/R)5/3dr/λ2]3/5,
and

Eq. (78)

r0tx=[16.71  sec(ζ)0RCn2(r)(1r/R)5/3dr/λ2]3/5,
where Latm is the atmospheric loss; γfiber is the fiber coupling efficiency (EFC) (if applicable); ζ is the link zenith angle; r0rx and r0tx are the uplink spherical-wave Fried parameter for the receiver and transmitter, respectively;27 and Cn2 is the refractive index structure parameter.911 The above equations for the Fried parameter are valid for l0r0L0, with l0 being the inner scale and L0 being the outer scale of the turbulence. In Eq. (74), α is the volume extinction coefficient of the atmosphere, h0 is the height above the ground. (Alternately, Latm can be calculated using MODTRAN.28) The parameters SRRP and SRDP represent the Strehl ratio that is measured at the pupil plane of the receiver and the Strehl ratio that is measured at the detector plane of the receiver, respectively.911 These parameters contain the new adaptive optics (AO) performance model that depends on the on-axis scintillation index σI2(0,R) or if the link range is short.911 The scintillation index is the turbulent intensity variance given as

Eq. (79)

σI2(r,R)=I2(r,R)I(r,R)21.

The parameter m in Eqs. (75) and (76) is a recognition that AO systems only work in weak turbulence or short ranges where phase fluctuations dominate but degrade in effectiveness as the turbulent channel migrates to moderate-to-strong turbulence levels where amplitude fluctuations dominate. Figure 12 shows the validity ranges of Eqs. (75) and (76) based on the experimental link budget comparisons in Ref. 12. For downlinks, the definitions in Eqs. (75) and (76) reverse. For horizontal links, the two Fried parameters are the same.

Fig. 12

Observed square-root of the scintillation index versus predicted square-root of the scintillation index plot with numbers from NTTR/PAX experiments showing the validity ranges of Eqs. (75) and (76).12

OE_63_4_041207_f012.png

When the system uses an EDFA as a pre-detection amplifier, the received laser beam must be coupled into a single-mode fiber at the output of the receiver terminal. However, propagation through atmospheric turbulence degrades the spatial coherence of a laser beam and limits the fiber-coupling efficiency (FCE). That is, the turbulence affects the ability to focus the beam completely within the numerical aperture of the fiber. Dikmelik and Davidson29 derived the following equation for the FCE under these conditions:

Eq. (80)

γfiber=8a20101e(a2+Arx/Ac)(x12+x22)I0(ArxAcx1x2)x1x2dx1dx2,
where γfiber is the FCE, a is the ratio of the receiver lens radius to the radius of the back propagated fiber mode, Arx=πDrx2/4 is the area of the receiver aperture, Drx is the diameter of the receiver aperture, and Ac is the spatial coherence area of the incident beam. In turbulence, the coherence area Ac essentially is the beam’s speckle size, so the ratio AR/Ac represents the number of speckles over the receiver aperture area.29 To first order, the coherence area equals Acπr02/4.

In the absence of turbulence, the incident plane wave is fully coherent and the optimum value of a is 1.12.29 Figure 13 depicts the FCE as a function of the number of speckles, Arx/Ac, over the receiver aperture, with a=1.12.

Fig. 13

FCE as a function of the number of speckles, Arx/Ac, over the receiver aperture. The coupling parameter a equals 1.12 in this curve. The circles in this graph represent the coupling efficiency for optimized values of a derived for six specific number of speckles Arx/Ac.

OE_63_4_041207_f013.png

With turbulence, the optimum value of the coupling-geometry parameter a that maximizes the coupling efficiency depends on the value of the number of speckles AR/Ac. Dikmelik and Davidson compared the coupling efficiencies for six optimum values with the efficiency calculated with the non-turbulence value.29 Figure 13 also shows the resulting data points plotted with the a=1.12 curve. This figure suggests that one can employ a=1.12 for all FCE calculations with turbulence.

Following the procedure proposed by Andrews and Phillips (Ref. 9, pp. 459–460), the SNRs written in Eqs. (45) and (52) now are given as

Eq. (81)

Q=Prec2qBe[(Prec+Pb)/Rλ+iD/Rλ2]+4kBT0Be/G2Rλ2RL+σI2(0,R)Prec2,
and

Eq. (82)

QPrec2hνNFoptPavgBe+σI2(0,R)Prec2,
respectively.

3.

Electrical Signal-to-Noise Ratios in Coherent Optical Communications

This section will describe models for coherent optical communications.

3.1.

Coherent Communications Receiver in Free Space

In a coherent beam detection system, the optical receiver beats its weak incoming (return) signal against a strong local oscillator (LO) signal to yield a more robust signal at the detector. Figure 14 shows a simplified layout of coherent laser communications receiver. The transmitted laser and LO beams interaction shift the incoming signal spectrum to an intermediate frequency (IF). This is known as heterodyne detection (homodyne detection shifts the incoming spectrum to baseband). The resulting signal is typically then sent to a transimpedance amplifier, followed by an IF filter and then a linear or square-law envelope detector. The dominant detection noise source for this receiver configuration is the LO shot noise.

Fig. 14

Simplified schematic of a coherent laser communications receiver.

OE_63_4_041207_f014.png

In a heterodyne coherent detection system, the performance measure can be evaluated using either the carrier-to-noise ratio (CNR) at the IF filter output or the SNR at the low-pass filter (baseband) output at a secondary detector. The electrical SNR (SNRe) at a secondary detector equals

Eq. (83)

SNRecoh=Γ2(1+m/2)[F11(m/2;1;CNR)1]2Γ(1+m)F11(m;1;CNR)Γ2(1+m/2)F11(m/2;1;CNR),
where Γ(x) is the Gamma function and F11(a,b,c) is the confluent hypergeometric function.30

The parameter m can take on values of either 1 or 2 depending on whether we have a linear or square-law detector. For a linear rectifier, Eq. (83) does not reduce to a simple analytical equation but is close to the following expressions under large and small CNR limits:

Eq. (84)

SNRecoh(1)0.92CNR21+0.46CNR={0.92CNR2;CNR12CNR;CNR1.

For a square-law detector, we have30

Eq. (85)

SNRecoh(1)CNR21+2CNR={CNR2;CNR1CNR/2;CNR1.

Figure 15 compares linear and square-law SNR as a function of CNR.30 In the weak signal case, both detectors predict the output SNR to be roughly the square of the input CNR. Here, the signal is suppressed relative to the noise and both detectors behave like a quadratic rectifier. When the signal is strong prior to the envelope detector, the SNR at the output of each detector is proportional to the input CNR. That is, the contribution of the noise in this latter case is suppressed relative to the signal and both detectors behave like a linear system. For the remainder of this paper, we will assume a square-law envelope detector, but the analysis also is applicable to a linear detector.

Fig. 15

Electrical SNR as a function of CNR for both linear and square-law envelope detectors.

OE_63_4_041207_f015.png

3.2.

Coherent Communications Receiver in Turbulence

In turbulence, Eq. (83) is still useful, but the CNR is modified for the degraded signal from the channel, and we include the turbulence-induced intensity variance. Specifically, we have

Eq. (86)

CNR=ηcohηPavg/(hpνBe)=2ηcohSNReQL,
and

Eq. (87)

ηcoh=ηhet[(1+Drx2/4WS2)(1+Drx2/4WLO2)(1+Drx2/8WE2)+(1+Drx2/16FE)2]MTFturb,
where ηhet is the heterodyne efficiency under perfect beam alignment. The bracketed term in Eq. (87) is the beam mismatch loss whose parameters are given as

Eq. (88)

1WE2=1WS2+1WLO2,

Eq. (89)

1FE=1Fs+1FLO,
and if an IF is needed

Eq. (90)

φIF=kRφLOtan1(kDrx2/16FE1+Drx2/8WE2),
where WS and FS are the incoming signal intensity beam radius and phase front radius of curvature, respectively, WLO and FLO are the LO intensity beam radius and phase front radius of curvature, respectively, and φIF and φLO are the IF and LO longitudinal phases, respectively. It is often used as a measure of detector performance because it measures the loss in coherent power that arises when the received field and the LO are not perfectly matched. If the LO characteristics match those of the incoming signal, which means WLO2=WS2, FLO=FS, and φLO=kR, FE=, and WE2=WS2/2, then the heterodyne efficiency is ηhet, which is its maximum value. Otherwise, there will be some additional loss besides ηhet. In Eq. (87), the modulation transfer function (MTF) for the turbulent channel equals

Eq. (91)

MTFturb=ηhetturb=01ue3.443(uDrx/r0)5/3du.

Figure 16(a) shows ηhetturb/ηhet as a function of Drx/r0. To get good mixing efficiency and minimize Strehl ratio losses, the receiver aperture diameter must be less than the Fried parameter at the receiver, e.g., for a 75% normalized heterodyne efficiency, one needs Drx0.5r0.

Fig. 16

Graph of (a) the normalized heterodyne efficiency as function of and (b) the turbulent coherent SNR as a function of the turbulent CNR for a three beam-wander intensity scintillation ratio.

OE_63_4_041207_f016.png

In Eq. (86), SNReQL is the quantum-limited SNR given as

Eq. (92)

SNReQLηPavg2hpνBe,
where η is the detector quantum efficiency, Pavg is the averaged received signal power, hp is Planck’s constant, ν=c0/λ is the laser frequency, c0 is the speed of light in vacuum, λ is the laser wavelength, and Be is the baseband bandwidth derived from the lowpass filtering after the envelope detector.

In turbulence, the coherent electrical SNR at the lowpass output of a square-law detector is written as30

Eq. (93)

SNRecoh(2)=CNR21+2CNR+σI2(Drx)CNR2.

Figure 16(b) shows the turbulent coherent SNR as a function of the turbulent CNR for three beam-wander intensity scintillation ratios. Here, CNR0=2ηhetturbSNReQL. This chart shows that the scintillation indices create constant SNRs for large CNR0, which in turn will generate constant BER curves. This suggests that other means for turbulence mitigation must be used to get the BER down to reasonable levels. However, this is the only issue with a coherent system approach in the turbulent channel. Achieving a high turbulent CNR is hard because the phase perturbations in the received signal’s phase front do not mix well with the LO’s phase front.

4.

Channel Models

Although there are many refractive index structure parameter models in the literature, there are two models that are most popular within the optics community. One is the Hufnagle–Valley (HV5/7) model and the other is the Hufnagle–Andrews–Phillips (HAP) model. This section will summarize both models, as well as present the round-earth model for those situations where the curvature of the earth affects the link analysis.

4.1.

Refractive Index Structure Parameter Models

There are two refractive index structure parameter models used most often by the community. One is the Hufnagel-Valley 5/7 Cn2 model, which is called the 1xHV5/7 model because for λ=0.5  μm and ζ=0  deg, it predicts a 5 cm Fried parameter and 7  μrad isoplanatic angle. Mathematically, it is written as

Eq. (94)

Cn2(h)=M{0.00594(w27)2(h105)10exp(h1000)+2.7×1016exp(h1500)+Cn2(0)exp(h100)},
with M being a scaling to represent the strength of the average high altitude background turbulence Cn2(0)=1.7×1014  m2/3.911 This allows the model to be used when Cn2(0) are not available. (As it turns out, M×Cn2(0) usually turns out to be close to the measured ground value.31) It has long been recognized that the advantage of this model over other atmospheric models is its inclusion of the rms wind speed, w, and the ground level refractive index structure parameter. Their inclusion permits variations in high-altitude wind speed and local near-ground turbulence conditions to better model real-world profiles over an extensive range of geographic locations. It also provides a model consistent with measurements of the Fried parameter r0 and the isoplanatic angle θ0. Let us see an example on how it is used.

Takenaka et al. reported some fiber coupling experimental result using the old version of the HV5/7 equation (Ref. 10, p. 100), a slew rate of 0.35  deg/s, and ground refractive index structure parameter of 1×1013  m2/3.32 The source of some of their parameters like the average wind speed of 88  m/s was Ref. 33. Figure 17 exhibits their EFC measurements (blue data) with their estimate of EFC, 17  dB, (red-dashed line) using this HV5/7 model and their evaluation of Eq. (64) using

Eq. (95)

a=πDrxWm2λF0,
where Wm is the fiber-mode field radius at the fiber end face and F0 is the receiver focal length. Now, the cited ground refractive index structure parameter for ground site is 2.68×1014  m2/3, which is the reported value of that site’s mean nighttime Cn2(0).34 The International Telecommunication Union (ITU) recommends the 1xHV5/7 model for the nighttime operations analyses.35 Using Eq. (80) with a=1.12, the 1xHV5/7 model and the receiver Fried parameter, one obtains an EFC estimate of 16.69  dB (red-dash-dot line), which is a slightly better estimate of the EFC measurements than the Takenaka et al. result. If one uses the average wind speed of 76  m/s used by Ref. 34, then one obtains a better EFC estimate of 15.65  dB (red-dash line). On the other hand, if one uses 0.65xHV5/7 and an average wind speed of 88  m/s, one calculates a very good EFC estimate of 14.5  dB (red-solid line). The importance of this example is to show that measuring the refractive index structure parameter vertical profile via some method like radiosonde, or at least measuring the ground refractive index structure parameter, may provide better accuracy in link analyses.

Fig. 17

EFC measurements with various analytical EFC estimates.

OE_63_4_041207_f017.png

Equation (93) has a slowly decreasing exponential term with altitude.13,31 This conflicts with the Cn2(h) behavior of h4/3 noted by Walters and Kunkel36 and supported by several other early measurements. To better represent this trend, Stotts and Andrews modified the HAP model to yield

Eq. (96)

Cn2(h)=M[1.04×103(w27m/s)2(h+hG105m)10exp(h+hG1200m)+2.7×1016exp(h+hG1700m)]+Cn2(hG)(hGh)4/3,  h>hG,
in units of m2/3, where hG=5  m always.37 This model is referred to as the modified HAP model. For the first few hundred meters, there is considerable difference between these two models. However, at 1  km and higher altitudes, the models are essentially the same with M=1 and same Cn2(0). Equation (96) also is the more appropriate model for optical links close to the ground. To account for various times of the day, the 4/3 power law in the last term can be replaced by a variable power p as described on p. 101 of Ref. 10.

Figure 18 shows a comparison between Eq. (96), evaluated at specific M and Cn2(h0) values, and the annual Korean Cn2(h) profiles at the 1%, 15%, 50%, 85%, and 99% cumulative percentiles.38 The latter data come from an Air Force Research Laboratory’s data base of 85 radiosonde balloon flights over South Korea. This graph shows good agreement between the theory and data. Let us see how this model compares with a typical refractive index structure parameter vertical profile.

Fig. 18

Comparison between Eq. (96) evaluated at specific M and Cn2(h0) values and the annual Korean Cn2(h) profiles at the 1%, 15%, 50%, 85%, and 99% cumulative percentiles.38

OE_63_4_041207_f018.png

Figure 19 depicts Eq. (96) with Cn2(h0)=1×1014  m2/3, M=0.7, and h0=1  m against the mean of the Cn2(h) measurements obtained at the Teide Observatory (OT) during 2004 (dotted - shorted line) and at the Roque de los Muchachos Observatory (ORM) during 2004 (solid line) and 2005 (dashed line).39 The horizontal line indicates the observatory altitude (2400 m). These statistical profiles correspond to the weighted average of the individual monthly statistical profiles. The number of individual profiles measured each month was used as weights.39 These data were derived from the Scintillation Detection and Ranging (SCIDAR) procedure.39 This technique is based on the statistical analysis of scintillation patterns produced at a telescope pupil by the light from two stars of a binary system. The profiles of the refractive index structure parameter come from the inversion of the average normalized autocovariance of many patterns. We see reasonable agreement between theory and data. Reference 37 showed other reasonable comparisons.

Fig. 19

Comparison between Eq. (96) and the mean of the measurements obtained at the Teide Observatory (OT) during 2004 (dotted-shorted line) and at the Roque de los Muchachos Observatory (ORM) during 2004 (solid line) and 2005 (dashed line).39 (Data used with permission of Dr. Julio A. Castro-Almazán and Dr. Begoña García-Lorenzo.)

OE_63_4_041207_f019.png

Figure 20 shows a year-long average of the refractive index structure parameter profile versus altitude estimates as a function of the re-interpolation distance dz (estimate’s vertical spacing) generated by Quatresooz et al., using the high-density profiles from the University of Wyoming (UWYO) Atmospheric Science Radiosonde Archive.40 This set of data profiles is for Hilo, Hawaii. Besides the data curves, curves generated from the 1xHV5/7 Cn2(h) model, a curve-fit model they created, denoted as analytical model in the figure, and the above modified HAP model for M=1.2 and Cn2(h0)=5×1015  m2/3. Again, the modified HAP model does a pretty good job in following most of the averaged data well just using a two-parameter fit. It does not follow the nominal 12-km bump very well, but it does follow the upper and lower atmosphere profiles much better than the HV5/7 model does.

Fig. 20

Plots of the refractive index structure parameter versus altitude estimates derived from data taken at Hilo, Hawaii, and curves generated from the 1xHV5/7 model, a specific analytical mode and Eq. (94).40 (Data used with permission of Mr. Florian Quatresooz and Professor Claude Oestges.]

OE_63_4_041207_f020.png

Mahmood et al. suggested that Cn2(h) comparisons with experimental data can be improved by using their associated wind speed profiles with altitude via the turbulent intensity.41 To evaluate this premise, we modify Eq. (96) to be

Eq. (97)

Cn2(h)=M[1.04×103(9m/sw(h+hg))2(h+hG105m)10exp(h+hG1200m)+2.7×1016exp(h+hG1700m)]+Cn2(h0)(h0h)4/3,  h>h0>hG,
to reflect a turbulent intensity with a wind speed profile dependence, where w(h) is the associated wind speed vertical profile. Figure 21 compares Eq. (97) with Hilo (dz=200  m)-Cn2(h) using that wind speed profile supplied by Mr. Quatresooz and Professor Oestges.40 The two profiles are in better agreement in the 12-km and upper region but deviate some in the lower atmospheric region. However, both figures are in reasonable agreement with the data overall.

Fig. 21

Plots of the refractive index structure parameter versus altitude measurements (Hilo, Hawaii), and Eq. (97) using the wind speed profile provided by Mr. Quatresooz and Professor Oestges40 (data used with their permission).

OE_63_4_041207_f021.png

4.2.

Round Earth Model

Many satellite laser communications analyses are performed using the above refractive index structure parameter vertical profile models and a flat-earth model in which the curvature of the Earth is ignored. For small zenith angles, this model works well. However, for larger zenith angles, the curvature of the earth must be considered. Assuming the Earth is a sphere, the altitude h above the ground reference point h0 at a zenith angle ζ is described as

Eq. (98)

h(z)=(RE+h0)2+2z(RE+h0)cos(ζ)+z2(RE+h0)+hG,
for h>h0, where RE=6371  km (Ref. 10, p. 112).

5.

Communications Bit Error Rate Equations and the Probability of Fade

This section will present the BER equations for most popular signaling formats in free space optical communications (FSOC) as well as the probability of fade equation, which is just Pfade=1Pd, where Pd is the probability of detection. We then will discuss now these equations in analyzing FSOC in turbulence.

5.1.

BER in Free Space

Table 4 summarizes the BER equations for the most popular signaling formats for FSOC. In the MPPM line, the parameter M equals 2kb where kb is the number bits per laser pulse.

Table 4

Most popular FSOC signaling formats and their BER equations.

Signaling formatBER
Non-return to zero on-off keying (NRZ-OOK)3,31BERNRZOOK=0.5erfc(SNRe/22)
Return to zero on-off keying (RZ-OOK)3BERRZOOK=0.5erfc(SNRe/2)
Differential phase shift keying (DPSK)3BERDPSK=0.5erfc(SNRe/2)
M-ary pulse position modulation (MPPM) (Ref. 21, p. 197)BERMPPM=(M/2M1)[112πe(xSNRe)2/2(10.5erfc(x/2)M1dx]
Binary pulse position modulation (Ref. 21, p. 197)BERBPPM0.5erfc(SNRe/2)BERMPPM
Binary phase shift keying1BERBPSK=0.5erfc(SNRe)

5.2.

BER and Probability of Fade in Turbulence

Figure 22 shows typical power in the fiber (PIF) (blue curve) and power out of the fiber (POF) (green curve) signals and their respective PDFs measured over a 183-km, NRZ OOK communications link cited taken at 01:52 UTC on May 18, 2009, at the Nevada Test and Training Range (NTTR).7 (This figure also shows the fielded system had 2 dB higher gain at low PIF than the system measured in the lab caused by minor system upgrade prior to field test.)

Fig. 22

Typical display of the received PIF and POF signals and their respective PDFs derived from 183-km NRZ OOK optical communications link under 5xHV5/7 turbulence conditions.7

OE_63_4_041207_f022.png

Link budgets generally use the 50-percentile received signal power, which is either the average received power for symmetric intensity PDF or the median received power for non-symmetric intensity PDFs. In turbulent conditions, the calcuated signal is the latter. Figure 23 shows the cummulative density function for the received PIF power derived from tests cited in.7 The PIF looks like it is log-normal. These figures indicate that the received signals are above the median 50% of the time and below the median for 50% of the time. It does not show that the higher received signals above the median are sufficient to be called a “1.” Turbulence mitigation must be employed to get the received signal to be above the threshold for a “1,” which is shown in Fig. 22 by those used to create the POF.7

Fig. 23

The statistical distribution of the PIF produces a consistent linear relationship between log optical power and log cumulative distribution.7

OE_63_4_041207_f023.png

To calculate the link availability, one generally uses the fade margin equation, which accounts for the additional power needed to overcome link fading. The basic form of the fade margin equation in free space is

Eq. (99)

Fade Margin(dB)=System Margin(dB)+Ptx(dBm)γtx(dB)γrx(dB)FSL(dB)Pavgmin(dBm).

This implies that the probability of detection is given as

Eq. (100)

Pd=0.5erfc[(fade margin-OSNR)/2]=0.5erfc[(system margin)/2],
and the probability of fade, which is the percentage of time the received intensity is below a prescribed threshold value, equals

Eq. (101)

Pfade=1Pd=10.5erfc[(system margin)/2].

For zero margin, we have

Pfade=1Pd=10.5erfc[0]=0.5,
which is expected as OSNR represents the average (50% level) optical SNR as noted earlier.

Let us now turn to the BER. The unconditional BER in the presence of turbulence-induced scintillation should be averaged with respect to the turbulence-induced intensity PDF to yield the average BER for those conditions. Specifically, we write

Eq. (102)

BER¯=0pI(I)BER(I)dI;  I>0
using the appropriate BER equation for the signaling format utilized and a scintillation index appropriate for the link’s turbulence and geometry conditions. In many cases, the log-normal (LN) PDF is used in Eq. (102), which is given as

Eq. (103)

plnI(I)=1I2πσlnI2(Drx)exp{[ln(I)+0.5σlnI2(Drx)]22σlnI2(Drx)},
with σlnI2(Drx)=ln[1+σI2(Drx)], which is valid in the weak-to-moderate intensity fluctuations regimes when the receiver aperture is large (Ref. 10, p. 94). The alternative is to replace plnI(I) in Eq. (102) by the Gamma–Gamma (GG) PDF that is written as

Eq. (104)

pI(I)=2(αβ)(α+β)/2Γ(α)Γ(β)(I)(α+β)/21Kαβ(2αβI),
where Kν(x) is the modified Bessel function of the second kind,

Eq. (105)

α=1σX2(Drx)
and

Eq. (106)

β=1σY2(Drx)
with σX2(Drx) being the large-scale intensity variance and σY2(Drx) being the small-scale intensity variance (Ref. 10, p. 94).

Equation (104) is valid in the weak-to-strong intensity fluctuations regimes when the receiver aperture is small as well as with certain large values of Drx. Experimental and computer simulation results indicate that Eq. (103) may be valid in deep turbulence when Drx<ρ0 or when Drx>R/kρ0 with ρ0=r0/2.1 being the lateral spatial coherence radius. For apertures sizes between these extremes, Eq. (103) may be the better model (Ref. 10, p. 94).

As was suggested by one of the reviewers, we need to note that a specific PDF model valid for all turbulence conditions is still an open question, particularly after going through a receiver aperture. That is, which PDF tails are correct. This was investigated in chapter 5 of Ref. 11.

6.

Scintillation Indices for Various Link Geometries

Fluctuations in the received signal intensity resulting from the propagation through turbulent channel is commonly described as “scintillation.” The term scintillation includes the temporal variation of received intensity (such as star twinkling) and spatial variation within a receiver aperture (such as speckle). Its magnitude is link-geometry dependent.

This section will present the scintillation indices for uplink, downlink, and horizontal communications link geometries.

6.1.

Uplink Communications Scintillation Indices and PDFs

In uplink communications geometry, a large beam will be incident on the satellite receiver aperture from diffraction and turbulent beam spreading. Beam wander will be several microradians and will create additional scintillation at the receiver, dependent on transmitter beam size. Angle of arrival fluctuations is typically less than a microradian. Aperture averaging cannot be used as the speckle/Fried parameter will be much larger than the receiver diameter. (Note that multiple beams through statistically independent atmospheric paths can reduce scintillation as a kind of aperture averaging, if available.) This means one now has scintillation at the receiver caused by beam wander. Specifically, the turbulence near the Earth’s surface causes a lever-arm effect at the satellite, causing signal intensity to occur. This means that tip/tilt (N=3 Zernike modes) tracking is necessary to reduce beam-wander scintillation. Specifically, we now need a beam-wander intensity PDF that considers whether the incoming beam is tracked or not.

Uplink beam wander has a significant effect on laser communications link performance. This requires tracked/tilt-corrected and untracked beam wander models.40 The key to these models is two sets of nondimensional beam parameters—one set at the transmitter and a second set at the receiver at distance R. For collimated transmitter, we have

Eq. (107)

z=0:  Θ0=1RfFL=1,Λ0=2RkW02=8λRπDtx2,
whereas at the receiver, the corresponding beam parameters are

Eq. (108)

z=R:  {Θ=Θ0Θ02+Λ02,Θ¯=1Θ,Λ=Λ0Θ02+Λ02=2RkW2=λRπW2.

In the above equations, Θ0=1R/fFL describes the amplitude change due to refraction of focusing, Λ0 describes the amplitude change due to aperture diffraction, k=2π/λ, W0 is the e2-intensity beam radius, and (Ref. 10, p. 18–19)

Eq. (109)

W=W0Θ02+Λ02=λRπΛ.

When Dtx/r0t1 or Dtx/r0t1 beam tracking is not necessary and the associated PDF for the incoming intensity ( power in the bucket [PIB]) is just a simple Gamma (G) distribution:

Eq. (110)

pI(I)dI=1Γ(m0)I[m0II(0,R)]mexp{m0II(0,R)}dI,  I>0,
or

Eq. (111)

pI(I)dI=1Γ(m0)I[m0I]mexp{m0I}dI,I>0,
where the shape parameter m0=1/σI,l2(0,R)Tracked, σI,l2(0,R) is the on-axis scintillation index, and I=I/I(0,R) (Ref. 10, p. 113).

When 1<Dtx/r0t<10, beam wander-induced scintillation becomes important, and the beam probability needs to be tracked. If the beam is tracked, then Eq. (110) or Eq. (111) is used with tilt-corrected m0. If the beam is not tracked, the resulting PDF distribution can be modeled as a modulated gamma distribution (MG), which is given as

Eq. (112)

pIu(I)dI=ϑΓ(m1)I(m1I(1+1/ϑ)I(0,R))ϑΓ(mζ,m1I(1+1/ϑ)I(0,R))dI,
or

Eq. (113)

pIu(I)dI=ϑΓ(m1)I(m1I(1+1/ϑ))ϑΓ(m1ζ,m1I(1+1/ϑ))dI,
where Γ(x,a) is the incomplete Gamma function, m1=1/σI,l2(0,L)untracked is the distribution’s shape parameter (Ref. 10, p. 113). In Eqs. (112) and (113), we have

Eq. (114)

ϑ=1+1+σI,l2(0,R)untracked34.29(ΛR/kr0T2)5/6(σpe2/W2)1,
where

Eq. (115)

σpe2=rc2[1(π2[Dtx/(22r0t)]21+π2[Dtx/(22r0t)]2)1/6],
with

Eq. (116)

rc2=[7.25(Rz0)2[Dtx/23/2]1/3]z0RCn2(h)[1zz0Rz0]2dz,
with Rz0=(Hh0)/cosξ, H is the satellite altitude, and zz0=(hh0)/cosξ (Ref. 10, pp. 109–113). In the case of a collimated beam, Eq. (116) becomes

Eq. (117)

rc2=0.54[λRDtx/2]2(Dtx/2r0t)5/3.

The untracked on-axis scintillation index equals

Eq. (118)

σI,l2(0,R)untracked=34.29(ΛRkr0  t2)5/6[σpe2Dtx2/8]+exp{0.49σBu2[1+0.56(1+Θ)σBu12/5]7/6+0.51σBu2[1+0.69σBu12/5]5/6}1,
where

Eq. (119)

σBu2=8.70k7/6(Hh0)5/6sec11/6(ζ)Re{h0HCn2(h)[ξ5/6(Λξ+i(1Θ¯ξ))5/6Λ5/6ξ5/3]dh},

Eq. (120)

ξ=1(hh0)/(Hh0),
with Θ¯=1Θ (Ref. 10, p. 110).

The tip/tilt-corrected on-axis scintillation index equals

Eq. (121)

σI,l2(0,R)tracked=34.29(ΛRkr0t2)5/6[σpeTT2Dtx2/8]+exp{0.49σBu2[1+0.56(1+Θ)σBu12/5]7/6+0.51σBu2[1+0.69σBu12/5]5/6},
where

Eq. (122)

σpeTT2=[rc2TzR]2[1(π2[Dtx/(22r0t)]21+π2[Dtx/(22r0t)]2)1/6],
and (Ref. 10, pp. 109–110)

Eq. (123)

Tz=0.57(λDtx/2)(Dtx/2r0t)5/6.

The on-axis scintillation index for both tracked and untracked beams is shown in Fig. 24 as a function of beam radius W0 along with computer simulation data.42 The tracked results are tip/tilt-corrected based on Eq. (121), whereas the untracked results come from Eq. (118). The computer simulation data and theoretical results are based on the 1xHV-5/7 profile model and ζ=0  deg. The agreement between theoretical and simulation results is good.

Fig. 24

On-axis scintillation index for both tracked (tilt-corrected) and untracked collimated beams to LEO as a function of beam radius (simulation results courtesy of G. J. Baker42).

OE_63_4_041207_f024.png

Figure 25 plots the on-axis scintillation index of an untracked 0.84-μm beam as a function of beam size at two zenith angles (0 deg and 60 deg) along with the Rytov theory result at 0 deg zenith angle. Computer simulation data were taken at geostationary orbit (GEO), so here we chose the larger scaling constant 3.86 instead of π in Eq. (115) for a slightly better fit at GEO. Again, the agreement between theoretical and simulation results is good.

Fig. 25

On-axis scintillation index at GEO for an untracked beam with zenith angles of 0 deg and 60 deg as a function of beam radius. The dotted curve is Rytov theory at 0 deg zenith. The 1xHV-5/7 and flat-Earth models were used for data and theory (adapted from Ref. 9).

OE_63_4_041207_f025.png

Let us now look at how these equations agree with PDF computer simulations. The simulations to be discussed assumes an optical beam from ground to space in 1xHV5/7 turbulence with zero zenith angle and λ=1.55  μm. The figures to come contain both computer simulation data and theoretical PDF models as a function of the Strehl parameter S=I/IFS(0,R) featuring several beam sizes covering various regimes of interest (Ref. 11 in chapter 6). In the Strehl parameter, IFS(0,R) is the mean free space transmitted intensity. The scintillation index and mean Strehl ratio in the legends of these figures are based on theoretical expressions, not simulation results. In all cases, the Fried parameter is 19 cm. As with the last two figures, the next few figures show good agreement between the theoretical and simulation data.

Figure 26 compares three uncorrected uplink beam PDFs as a function of Strehl parameter with computer simulation results (chapter 6 in Ref. 11). In this figure, we have (2W0/r0tx=0.0741). In this case, the tracked and untracked scintillation indices are the same as there is no beam wander. The three PDFs in the following figures are the log-normal (LN), gamma–gamma (GG), and gamma (G) PDFs. The three PDFs appear to agree with one another until the Strehl parameter becomes greater than 1.5. The LN PDF appears to be closer to the data than the other two PDFs, but there are some outliers in the simulation results.

Fig. 26

Graph of three uplink beam PDFs as a function of Strehl parameter for 2W0/r0  tx1 with computer simulation results (Ref. 11, chapter 6).

OE_63_4_041207_f026.png

Figures 27 and 28 are comparisons of the untracked and tilt-corrected uplink beam PDFs as a function of the Strehl parameter with computer simulation results for (0.1<2W0/r0T<10) using two different W0s (chapter 6 in Ref. 11). The PDF used in the (a) figures is an MG PDF and the one used in the (b) figures is a G PDF. It is clear from Figs. 27(a) and 28(a) that the increase in the size of W0 causes the PDF shape to radically change. Figures 27(b) and 28(b) have similar shapes but not values.

Fig. 27

Comparison of (a) untracked and (b) tilt-corrected uplink beam PDFs as a function of Strehl parameter for (0.1<2W0/r0T<10) with computer simulation results (Ref. 11, chapter 6).

OE_63_4_041207_f027.png

Fig. 28

Comparison of (a) untracked and (b) tilt-corrected uplink beam PDFs as a function of Strehl parameter for (0.1<2W0/r0T<10) with computer simulation results (Ref. 11, chapter 6].

OE_63_4_041207_f028.png

Figure 29 shows a comparison of the untracked and tilt-corrected uplink beam PDFs as a function of Strehl parameter with computer simulation results for (2W0/r0T=7.4) (chapter 6 in Ref. 11). The PDF used in the (a) figures is an MG PDF and the one used in the (b) figure is a G PDF. The untracked and tilt-corrected PDFs have more similar shapes but different values. The similarity comes from the fact that the large beam effectively means there is essentially no beam wander.

Fig. 29

Comparison of (a) untracked and (b) tilt-corrected uplink beam PDFs as a function of Strehl parameter for (2W0/r0T=7.4) with computer simulation results (chapter 6 in Ref. 11).

OE_63_4_041207_f029.png

6.2.

Downlink Communications Scintillation Indices

In this case, the beam will be essentially a collimated beam from the satellite. The scintillation index follows that of a plane wave propagating through turbulence contained in the near-earth atmosphere (angle of arrival fluctuations may be several microradians).31 In fact, from space-to-ground the scintillation index (also equal to the Rytov variance in this case) is basically the same as that of an infinite plane wave and is given as

Eq. (124)

σI2σR2=2.25k7/6sec11/6(ζ)h0HCn2(h)(hh0)5/6dh.  H20  km.

If the receiver aperture is larger, then aperture averaging reduce scintillation. The aperture-averaged scintillation index is written as

Eq. (125)

σI2(Drx)=8.70k7/6(Hh0)5/6Re{h0HCn2(h)[(kDrx216R+i(hh0)(Hh0))5/6(kDrx216R)5/6]dh},
for 0 deg-zenith angle (Ref. 10, p. 106). For other zenith angles, Eq. (125) was modified for a slant path assuming the round Earth model. Figure 30 depicts the aperture-averaged scintillation index as a function of the receiver diameter for ζ=0  deg,20  deg,40  deg, and 60 deg. As the receiver diameter gets larger, the aperture-averaged scintillation index gets smaller. For higher turbulence levels, the scintillation index for a point receiver is written as (Ref. 10, p. 106)

Eq. (126)

σI2(L)=exp[0.49σR2(1+1.11σR12/5)7/6+0.51σR2(1+0.69σR12/5)5/6]1.

Fig. 30

Plot of the aperture averaged scintillation index σI2(Drx) as a function of the receiver diameter Drx for four zenith angles.

OE_63_4_041207_f030.png

Scintillation measurements of a 1064 nm laser were made by an optical ground station at the European Space Agency observatory in Tenerife, Spain, while tracking a low Earth orbit satellite during the spring and summer of 2010. Yura and Kozlowski reported the first comparison between theoretical predictions and these measurements.43 Specifically, they compared Yura’s apertured-averaged scintillation index (SI) model44 evaluated using the MAUI3 turbulence profile model45 against said data. Good agreement was obtained. The authors saw minor variation in the SI between daytime and nighttime experimental runs. In addition, their results suggested that the nighttime turbulence profile for the atmosphere above the observatory in Tenerife is like that above Haleakala in Maui, Hawaii.

Figure 31 compares the MAUI3 and AMOS refractive index structure parameter vertical profile models with a 0.1xHAP model where Cn2(h0)=5×1015  m2/3. The latter model follows the MAUI3 model well, except close to the ground. Comparing the MAUI3 profile with Figs. 19 and 20 for the Canary Islands and Hilo, respectively, the basic profile is similar among all three, but each is shifted relative to the others in Cn2(h) values. The modified HAP models do a good job in predicting each profile.

Fig. 31

Comparison of 0.6xHAP, MAUI3, and AMOS refractive index structure parameter vertical profile models.

OE_63_4_041207_f031.png

Figure 32 compares SI measurements as a function of elevation angle for (a) daytime, 09:44:04 local time on June 23, 2010, and (b) nighttime, 24:27:00 local time on July 14, 2010, with both the Yura SI model43 and Eq. (124) were evaluated using the 0.1xHAP model. The Eq. (124) predictions using the 0.1xHAP model follows the measurements as well as the Yura models do using the MAUI3 model. Like in the case of fiber coupling, we lower the multiplier in the Cn2(h) HAP model to get good agreement with measurements. Again, more definitive comparisons could be made if radiosonde or ground scintillometer measurements were made during the experimental runs.

Fig. 32

Comparisons of measured SI data as a function of elevation angle for specific (a) daytime and (b) nighttime experiments with both the Yura SI model and Eq. (124) evaluated using the 0.1xHAP model.

OE_63_4_041207_f032.png

Perlot et al.46 reported a set of experimental measurements derived from the Japanese Optical Inter-orbit Communications Engineering Test Satellite (OICETS) communicating down to an optical receiver located near Munich, Germany. OICETS is an LEO satellite with a sun-synchronous orbit at an altitude of 610 km. It is optically compatible with the Artemis satellite and transmits a 50-Mbit/s NRZ OOK data stream at a wavelength of 847 nm. The experiment’s ground receiver used a 40-cm (12-cm obscuration) Cassegrain telescope receiver. Figure 33 shows OICETS uplink scintillation index measurements derived from the received uplink as well as Eq. (121) predictions assuming a 0.4xHAP model with constant slew rate of 0.003  rad/s and Cn2(h0)=5×1015  m2/3. This figure shows reasonable agreement between analytical predictions and the measurements.

Fig. 33

Comparison of measured OICETS uplink scintillation indices and Eq. (121) assuming a 0.4xHAP Models with Cn2(h0)=5×1015  m2/3 and a constant slew rate of 0.003  rad/s.

OE_63_4_041207_f033.png

Figure 34 shows (a) various measured OICETS downlink scintillation indices created from the received downlink and (b) DIMM-based measurements of the Fried parameter r0. Included in these figures are Eq. (125) and Fried parameter predictions using the 1xHAP and 0.4xHAP models assuming a constant slew rate of 0.003  rad/s and Cn2(h0)=5×1015  m2/3. The resulting wind speed was 74  m/s. These figures show good agreement between analytical predictions and the various data sets. However, the OICETS satellite experiences a variable slew rate that is a function of the incurred elevation angle. Figure 35 is a redo of Fig. 34, but where only the 0.4xHAP model plus the Elevation Angle-Max=90  deg slew rate curve given in fig. 10 of Ref. 46 are used to calculate Eq. (124) and Fried parameter predictions. Figure 34(a) also includes the Rytov variance, which is the non-aperture averaged scintillation index. In Fig. 34(a), the 0.4xHAP model scintillation index expands and deviates a little from the KT3 curve for elevation angles greater than 20 deg. Clearly, the aperture averaging drives the Rytov variance down a lot to yield the Eq. (125) curve. In Fig. 34(b), the resulting curve now reaches a peak at an elevation angle round 18 deg and then decreases. Its shape in more in line with all data curves, except the KT3 curve. It also shows that the variable slew rate has a dramatic effect on key turbulence parameters.

Fig. 34

Comparison of Eq. (125) and Fried parameter calculations for a 1xHAP and a 0.2xHAP models with (a) various measured OICETS downlink scintillation indices and (b) DIMM-based measurements for the Fried parameter r0. The two models assumed Cn2(h0)=5×1015  m2/3 and a constant slew rate of 0.003  rad/s.

OE_63_4_041207_f034.png

Fig. 35

Comparison of Eq. (125) and Fried parameter calculations for a 0.4xHAP models with (a) various measured OICETS downlink scintillation indices and (b) DIMM-based measurements for the Fried parameter r0. The two models assumed Cn2(h0)=5×1015  m2/3 and the Elevation Angle-Max=90  deg slew rate curve given in fig. 10 of Ref. 46.

OE_63_4_041207_f035.png

6.3.

Horizontal Communications Scintillation Index

This section deals with the scintillation index for a horizontal communications link, which involves Gaussian beam propagation. Because of the horizontal geometry, this path can exhibit strong beam wander and large scintillation indices. The following equations are from Ref. 9, section 10.3.5; Ref. 10, pp. 52–53.

For a horizontal FSOC link, the spherical-wave Fried parameters equal

Eq. (127)

r0tx=r0rx=[0.16k2Cn2(htx)R]3/5,
with htx=hrx.

The beam wander variance and pointing error variance in the pupil plane are

Eq. (128)

rc2=2.42Cn2(htx)R3W01/3,
and

Eq. (129)

σpe2=2.42Cn2(htx)R3W01/3[1(π2W02/25r021+π2W02/25r02)1/6],
respectively. The beam wander variance and pointing error variance in the detector plane are

Eq. (130)

rc2(DRx)=2.42Cn2R3W01/3(ΩRxΛ1+ΩRx),
and

Eq. (131)

σpe2(DRx)=2.42Cn2(htx)R3W01/3(ΩRxΛ+ΩRx)[1(π2W02/25r021+π2W02/25r02)1/6],
with ΩRx=2R/k0WRx2 and WRx=Drx/22.

The Rytov variance, long-term beam radius and short-term beam radius in the pupil plane are given as

Eq. (132)

σR2=1.23Cn2(htx)k07/6R11/6,

Eq. (133)

WLT=W(1+1.33σR2Λ5/6)3/5,
and

Eq. (134)

WST=WLT2rc2.

The wander-induced scintillation index in the detector plane is equal to

Eq. (135)

σI,Untracked2(DRx)=4.42σR2ΛST5/6σpe2WST2+exp[σlnX2(DRx,l0)σlnX2(DRx,l0,L0)+σlnY2(DRx)]1,
and the tip/tilt-tracked scintillation index in the detector plane equals

Eq. (136)

σI,Tip-Tilt2=exp[σlnX2(DRx,l0)σlnX2(DRx,l0,L0)+σlnY2(DRx)]1,
with ΛST=2R/k0WST2. In Eqs. (135) and (136), we have

Eq. (137)

σlnX2(DRx,l0)0.49σR2=(ΩRxΛΩRx+Λ)2(1312Θ¯+15Θ¯2)(ηXdQlηXd+Ql)7/6[1+1.75(ηXdηXd+Ql)1/20.25(ηXdηXd+Ql)7/12],

Eq. (138)

σlnX2(DRx,l0,L0)0.49σR2=(ΩRxΛΩRx+Λ)2(1312Θ¯+15Θ¯2)(ηXd0QlηXd0+Ql)7/6×[1+1.75(ηXd0ηXd0+Ql)1/20.25(ηXd0ηXd0+Ql)7/12],

Eq. (139)

σlnY2(DRx,l0)=1.27σR2ηY5/61+0.40ηY/(ΩRx+Λ1),

Eq. (140)

ηX=[0.38/(13.21Θ¯+5.92Θ¯2)+0.47σR2Ql1/6((1312Θ¯+15Θ¯2)/(1+2.20Θ¯))6/7]1,

Eq. (141)

σG2σI,l2(0,L)3.86σR2{0.40[(1+2Θ)2+(2Λ+3/Ql)2]11/12[(1+2Θ)2+4Λ2]1/2×[sin(116φ2+φ1)+2.610[(1+2Θ)2Ql2+(3+2ΛQl)2]1/4sin(43φ2+φ1)0.518[(1+2Θ)2Ql2+(3+2ΛQl)2]7/24sin(54φ2+φ1)]13.401ΛQl11/6[(1+2Θ)2+4Λ2]116[(1+0.31ΛQlQl)5/6+1.096(1+0.27ΛQl)1/3Ql5/60.186(1+0.24ΛQl)1/4Ql5/6]},
and

Eq. (142)

ηY=3(σR/σG)12/5(1+0.69σG12/5),
where ηXd0=ηXdQ^0/(ηXd+Q^0), ηXd=ηX/[1+0.40ηX(2Θ¯1)/(ΩRx+Λ1)], Q^0=64π2R/k0L02, Ql=10.89R/k0l02, φ1=tan1(2Λ/[1+2Θ]), and φ2=tan1[(1+2Θ)Ql/(3+2ΛQl)].

Strömqvist Vetelino et al.47 used Eq. (135) to compare resulting GG and LN PDF against both simulation and experimental data. The range is 1500 m. Figures 36Fig. 3738 show example comparisons of GG and LN PDFs as a function of the normalized log irradiance ln(I) against simulation and experimental results for Drx=1,5, and 13 mm.47 The authors cited Cn2(htx)=6.45×1014  m2/3 for both the computer simulation and experiment. These figures show that both sets of data agreed with both PDFs in the middle but deviate from the LN PDF at the tails for small intensities. However, both sets of data appear to follow the GG PDF well in that regime.

Fig. 36

Comparison of GG and LN PDFs as a function of the normalized log irradiance ln(I) against (a) simulation data and (b) experimental data for Drx=1  mm and Cn2(htx)=6.5×1014  m2/3.

OE_63_4_041207_f036.png

Fig. 37

Comparison of GG and LN PDFs as a function of the normalized log irradiance ln(I) against (a) simulation data and (b) experimental data for Drx=5  mm and Cn2(htx)=6.5×1014  m2/3.

OE_63_4_041207_f037.png

Fig. 38

Comparison of GG and LN PDFs as a function of the normalized log irradiance ln(I) against (a) simulation data and (b) experimental data for Drx=13  mm and Cn2(htx)=6.5×1014  m2/3.

OE_63_4_041207_f038.png

Figures 39Fig. 4041 show example comparisons of GG, LN, and G PDFs as a function of the normalized log irradiance ln(I) against simulation and experimental results for receiver diameters, but with Cn2(htx)=5.36×1013  m2/3  .47 These figures show that both sets of data agreed with all three PDFs in the middle. In Fig. 39(a), these data are closer to the GG PDF for small intensities but are closer to the LN PDF for high intensities. In Fig. 39(b), no data are available for a comparison at small intensities, but other runs showed data closer to GG again. The exhibited data are still closer to the LN PDF for high intensities. The next five figures show that both the simulation and experimental data follow the LN PDF.

Fig. 39

Comparison of G, GG, and LN PDFs as a function of the normalized log irradiance ln(I) against (a) simulation data and (b) experimental data for Drx=1  mm and Cn2(htx)=5.36×1013  m2/3.

OE_63_4_041207_f039.png

Fig. 40

Comparison of G, GG, and LN PDFs as a function of the normalized log irradiance ln(I) against (a) simulation data and (b) experimental data for Drx=5  mm and Cn2(htx)=5.36×1013  m2/3.

OE_63_4_041207_f040.png

Fig. 41

Comparison of G, GG, and LN PDFs as a function of the normalized log irradiance ln(I) against (a) simulation data and (b) experimental data for Drx=13  mm and Cn2(htx)=5.36×1013  m2/3.

OE_63_4_041207_f041.png

Their conclusion was that in the moderate-to-strong fluctuation regime, the GG distribution provides a good fit to the irradiance fluctuations collected by finite-sized apertures that are significantly smaller than the lateral spatial coherence radius ρ0. For apertures larger than or equal to this radius, the irradiance fluctuations are LN distributed. We noted this trend when discussing intensity PDFs earlier in this paper.

7.

Example BER Calculation and Turbulence Mitigation

In this section, we provide an example of BER calculations. If they do not meet the system BER requirement, we will provide a couple of techniques that will show these BERs can be improved. These latter techniques are for illustrative purposes only and not intended to imply these are the only options or the best. However, they both do work in turbulent situations where amplitude fluctuations dominate unlike adaptive optics.12

In reviewing the literature, one finds a few ways of specifying the atmospheric turbulence levels for SATCOM link budget calculations. One way is to specify the Fried parameter for the envisioned scenario. For example, let us assume a 19 cm Fried parameter for a 0-deg zenith angle SATCOM link. Table 5 exhibits a comparison of Fried parameters and tracked/tilt-corrected and untracked scintillation indices for the HAP and HV5/7 models yielding that assumed Fried parameter. We have assumed the following parameters for the HAP model are

hG=5  m,  h0=1  m,  M=4,  Cn2=5×1014  m2/3.

Table 5

Comparison of Fried parameters and tracked, tip/tilt-corrected scintillation indices for HAP and 1xHV/5/7 models yielding the same Fried parameter at ζ=0  deg.

Zenith angle0 deg20 deg40 deg60 deg
HAP model rot (cm)19.418.716.512.8
1xHV/5/7 model rot (cm)19.418.716.512.8
HAP model ror (m)9.569.148.186.43
1xHV/5/7 model ror (m)17.5717.4416.8014.69
HAP model tracked σI2(0,R)0.1050.1150.1520.265
1xHV/5/7 model tracked σI2(0,R)0.0270.0290.0390.067

These parameters are chosen so that the Fried parameter would match that of the 1xHV-5/7 model. It is clear from this table that the HAP models predict much larger scintillation indices than the 1xHV5/7 model even though the Fried parameters are the same. (It should be noted that the receiver Fried parameters yield unity Strehl ratios for all zenith angles.) Let us now look at example link budgets and BER plots for the 1xHV5/7 numbers in Table 5.

Let us assume that we have tip/tilt-corrected, 10 Gigabit per second (Gbps) EDFA-PIN uplink SATCOM link with the following system parameters: Pr=0  dBm=GPavg for “1” signal, BER=1012, H=300  km, F=, Dtx=Drx=0.10  m, η=0.681, γtx=γrx=0.25, λ=1.55  μm, Be=7.5  GHz, and nsp=1.58 [Nopt=5  dB]. The atmospheric transmittance is derived from a Rural (VIS = 23 km) aerosol model atmosphere from 2-m above ground level to space and is calculated using PCModWin®.28 These transmittances were around 90% or higher and have negligible effect on the SNR for this analysis.

Figure 42 shows the average quantum-limited electrical SNR and received power as a function of SATCOM zenith angle. These SNRs are in the range for a reasonable gain to yield 0 dBm. (Note: One must be careful in applying gain to the input signal, so the result does not destroy the photodetector.)

Fig. 42

Quantum-limited electrical SNR and received power as a function of SATCOM zenith angle for a 1xHV5/7 SATCOM uplink.

OE_63_4_041207_f042.png

From Table 4, the NRZ OOK BER in the absence of turbulence is given as

Eq. (143)

BER¯=0.5erfc(SNRe/23/2).

For a tracked uplink, the average BER for NRZ-OOK then is

Eq. (144)

BER¯=0.50pI(I)erfc(123/2SNReQL(I)2ηNFoptI+SNReQLσI,l2(0,R)Tracked)dI,
using Eq. (101) but with Eq. (110) replacing the LN PDF. Figure 43 shows BER performance as a function of the SNReQL for selected zenith angles. The required 1012 BER is not achieved even though SNReQL is large at any of the zenith angles shown. In fact, the BER essentially limits out for SNReQL>25  dB at all zenith angles at values orders of magnitude larger. As it turns out, whether the system is incoherent or coherent and/or which signaling format is used, the scintillation index will peg the BER to constant value at high SNR well above the BER requirement. Other turbulence mitigation techniques must be employed to get the desired system performance, e.g., interweaving forward error correction (FEC) coding, optical automatic gain control (OAGC) circuitry.

Fig. 43

1xHV5/7 BER performance as a function of quantum-limited electrical SNR for selected zenith angles.

OE_63_4_041207_f043.png

Using a Reed-Solomon FEC Code RS(255,199), the BER performance can be improved to levels better than the BER requirement as shown in Fig. 44, e.g., at ζ=60  deg, the BER is <1014. The penalty is that the information rate drops by the ratio of 199/255=0.78. This is a small price to pay given the significant BER performance improvements. [A RS(255,207) code allows all the zenith angle but ζ=60  deg to exceed the 1012 BER requirement. At ζ=60  deg, the BER is level at 8×1012.]

Fig. 44

1xHV5/7 BER using the RS(255,199) code as a function of quantum-limited electrical SNR for selected zenith angles.

OE_63_4_041207_f044.png

An alternative for OOK and DPSK systems is to use the optical automatic gain control invented by David Young from John Hopkins University Applied Physics Laboratory.8,48 It basically reduces the variance in the intensity PDF under long range/high turbulence conditions, among other things.8,49 Let us be more specific.

Figure 45 shows the high-level layout of the OAGC. The first function of the OAGC system is protection of the photodiodes and follow-on electronics from catastrophic damage caused by high received optical powers at the output of the FSOC terminal.8,48 To prevent this, the OAGC optically amplifies or attenuates as necessary through a series of multiple gain stages as presented in Fig. 45 and discussed in Ref. 49 to output a constant power at a level of optimal performance for the detector. In essence, the time-variant optical input [I(t)] is translated into a constant amplitude output with a variable optical signal-to-noise ratio [OSNR(t)]. This is shown in Fig. 46, which presents OSNR and OAGC output power as a function of power into the OAGC (PIF). The maximum gain of the first-generation system was between 40 and 45 dB.49

Fig. 45

OAGC block diagram (CPU = central processing unit).49

OE_63_4_041207_f045.png

Fig. 46

OAGC block diagram (CPU = central processing unit).8

OE_63_4_041207_f046.png

The second function of the OAGC is to provide low-noise optical amplification for improved receiver sensitivity. This aspect of performance is characterized by the noise figure (NF) and gain metrics.8 Because the system NF is set by the first amplifier, this stage is designed for low NF and only has a moderate net gain. In addition, it is designed to gain clamp at input levels above a certain level to aid in maintaining the desired system output. The second stage incorporates a variable optical attenuator with a response time of <4  μs to dynamically respond to input power fluctuations by adjusting the system gain. The third stage handles output regulation and is designed to manage higher output power with an adjustable gain. For the overall system gain metric, the effective OAGC gain is dependent on the pre-set target output level. The third function of the OAGC is to reduce bit errors arising from power fluctuations at the receiver that cause timing jitter in the digital eye. By maintaining a constant output, power transients are prevented from coupling through the follow-on electronics and degrading BER performance in the FSOC link. This performance was shown in Fig. 22 for a 1012 BER, which presents OSNR and OAGC output power as a function of power into the OAGC (which is equivalent to PIF) for 183 km link under 5xHV5/7 conditions.8 The signal modulation scheme was NRZ-OOK. In short, the OAGC reduces the turbulent-intensity PDF to “delta function” (upper righthand corner) eliminating the PDF integration in Eq. (101) and leaving one with the mean intensity as a result.

8.

Summary

This paper provided a tutorial on how to create link budgets and BER and probability of fade calculations for optical communications designed to operate in the turbulent channel. It reviewed the characterization models necessary for either incoherent or coherent FSOC uplink, downlink, and horizontal system analyses in the turbulent channel. Beam wander, scintillation, and receiver noise variance as well as pointing and tracking effects were included in this paper. Comparisons among these models, computer simulations, and field measurements are provided throughout the paper. Good agreement is shown among all. An example analysis was provided using this information. The conclusion is that no matter whether the system is incoherent or coherent and/or which signaling format is used, the scintillation index will peg the BER to constant value at high SNR well above the desired value. Other turbulence mitigation techniques must be employed to get the desired system performance.

Disclosures

The authors declare that there are no conflicts of interest related to this article.

Code and Data Availability

Experimental data presented in this paper may be obtained from the cited authors upon reasonable request.

References

1. 

C. Chen et al., “Demonstration of a bidirectional coherent air-to-ground optical link,” Proc. SPIE, 10524 105240G https://doi.org/10.1117/12.2292848. PSISDG 0277-786X (2018). Google Scholar

2. 

H. Kaushal and G. Kaddoum, “Optical communication in space: challenges and mitigation techniques,” IEEE Commun. Surv. Tutorials, 19 (1), 57 –96 https://doi.org/10.1109/COMST.2016.2603518 (2017). Google Scholar

3. 

L. B. Stotts, Free Space Optical Systems Engineering, John Wiley and Sons( (2017). Google Scholar

4. 

A. Mansour, R. Mesleh and M. Abaza, “New challenges in wireless and free space optical communications,” Opt. Lasers Eng., 89 95 –108 https://doi.org/10.1016/j.optlaseng.2016.03.027 OLENDN 0143-8166 (2017). Google Scholar

5. 

S. Seel et al., “Space to ground bidirectional optical communications link at 5.6 Gbps and EDRS connectivity outlook,” in Aerosp. Conf., 1 –7 (2011). Google Scholar

6. 

R. Lange et al., “142 km, 5.625 Gbps free: space optical link based on homodyne BPSK modulation,” Proc. SPIE, 6105 61050A https://doi.org/10.1117/12.673749 PSISDG 0277-786X (2006). Google Scholar

7. 

Z. C. Bagley et al., “Hybrid optical radio frequency airborne communications,” Opt. Eng., 51 055006 https://doi.org/10.1117/1.OE.51.5.055006 (2012). Google Scholar

8. 

J. C. Juarez et al., “Analysis of link performance for the FOENEX laser communications system (Proceedings Paper),” Proc. SPIE, 8380 838007 https://doi.org/10.1117/12.919928 PSISDG 0277-786X (2012). Google Scholar

9. 

L. C. Andrews and R. L. Phillips, Laser Beam Propagation through Random Media, 2nd ed.SPIE Press, Bellingham, Washington (2005). Google Scholar

10. 

L. C. Andrews, Field Guide to Atmospheric Optics, 2nd ed.SPIE Press, Bellingham, Washington (2019). Google Scholar

11. 

L. C. Andrews and M. K. Beason, Laser Beam Propagation through Random Media: New and Advanced Topics, SPIE Press, Bellingham, Washington (2023). Google Scholar

12. 

L. B. Stotts and L. C. Andrews, “An adaptive optics model characterizing turbulence mitigation for free space optical communications link budgets,” Opt. Express, 29 (13), 20307 –20321 https://doi.org/10.1364/OE.430554 OPEXFF 1094-4087 (2021). Google Scholar

13. 

S. O. Rice, “The mathematical analysis of random noise,” Bell Syst. Tech. J., 23 282 –332 https://doi.org/10.1002/j.1538-7305.1944.tb00874.x BSTJAN 0005-8580 (1945). Google Scholar

14. 

L. Mandel, “Fluctuations of photon beams; the distribution of photoelectrons,” Proc. Phys. Soc., 74 233 –243 https://doi.org/10.1088/0370-1328/74/3/301 (1959). Google Scholar

15. 

C. W. Helstrom, “The detection and resolution of optical signals,” IEEE Trans. Inf. Theory, 10 (4), 275 –287 https://doi.org/10.1109/TIT.1964.1053702 (1964). Google Scholar

16. 

A. Papoulis, Systems and Transforms with Applications in Optics, McGraw-Hill( (1968). Google Scholar

17. 

V. I. Tartarski, Wave Propagation in a Turbulent Medium, McGraw-Hill( (1961). Google Scholar

18. 

D. O. North, “An analysis of the factors which determine sign/noise discrimination in pulsed-carrier systems,” Proc. IEEE, 51 1016 –1027 https://doi.org/10.1109/PROC.1963.2383 IEEPAD 0018-9219 (1963). Google Scholar

19. 

D. Middleton, “Statistical criteria for the detection of pulsed carriers in noise,” J. Appl. Phys., 24 371 –378 https://doi.org/10.1063/1.1721290 JAPIAU 0021-8979 (1953). Google Scholar

20. 

W. W. Peterson, T. G. Birdsall and W. C. Fox, “The theory of signal detectability,” Trans. IRE Prof. Group Inf. Theory, 4 171 –212 https://doi.org/10.1109/TIT.1954.1057460 (1954). Google Scholar

21. 

C.W. Helstrom, Statistical Theory of Signal Detection, 2nd ed.Pergamon Press( (1968). Google Scholar

22. 

C. E. Shannon, “A mathematical theory of communication,” Bell Syst. Tech. J., 27 379623 –423656 https://doi.org/10.1002/j.1538-7305.1948.tb01338.x BSTJAN 0005-8580 (1948). Google Scholar

23. 

H. Pishro-Nik, Introduction to Probability, Statistics, and Random Processes, 321 –323 Kappa Research LLC( (2014). Google Scholar

24. 

P. Becker, N. Olsson and J. Simpson, Erbium-Doped Fiber Amplifiers Fundamentals and Technology, Academic Press( (1999). Google Scholar

25. 

R. G. Marshalek, “Pointing, acquisition and tracking,” Near-Earth Laser Communications, 59 –96 CRC Press( (2009). Google Scholar

26. 

M. Toyoshima et al., “Optimum divergence angle of a Gaussian beam wave in the presence of random jitter in free-space laser communication systems,” J. Opt. Soc. Am. A, 19 (3), 567 –571 https://doi.org/10.1364/JOSAA.19.000567 JOAOD6 0740-3232 (2002). Google Scholar

27. 

R. R. Beland, “Propagation through atmospheric optical turbulence,” The Infrared & Electro-Optical Systems Handbook, 2 Environmental Research Institute of Michigan( (1996). Google Scholar

28. 

L. B. Stotts and J. Schroeder, Atmospheric Modeling Using PcModWin/MODTRAN, SPIE Press, Bellingham, Washington (2019). Google Scholar

29. 

Y. Dikmelik and F. M. Davidson, “Fiber-coupling efficiency for free-space optical communication through atmospheric turbulence,” Appl. Opt., 44 (23), 4946 –4951 https://doi.org/10.1364/AO.44.004946 (2005). Google Scholar

30. 

L. C. Andrews, R. L. Phillips and C. Y. Hopen, Laser Beam Scintillation with Applications, SPIE Press, Bellingham, Washington (2001). Google Scholar

31. 

L. B. Stotts and L. C. Andrews, “Bit error rate performance of a laser ground-to-satellite uplink communications systems in the presence of atmospheric turbulence and loss,” in Proc. 2022 IEEE Int. Conf. Space Opt. Syst. and Appl. (ICSOS), a-16 (2022). Google Scholar

32. 

H. Takenaka, M. Toyoshima and Y. Takayama, “Experimental verification of fiber-coupling efficiency for satellite-to-ground atmospheric laser downlinks,” Opt. Express, 20 (14), 15301 –15308 https://doi.org/10.1364/OE.20.015301 OPEXFF 1094-4087 (2012). Google Scholar

33. 

M. Toyoshima, H. Takenaka and Y. Takayama, “Atmospheric turbulence-induced fading channel model for space-to-ground laser communications links,” Opt. Express, 19 (17), 15965 –15975 https://doi.org/10.1364/OE.19.015965 OPEXFF 1094-4087 (2011). Google Scholar

34. 

K. Yano, T. Yoshihisa and K. Gyoda, “Measurements of the refractive index structure constant and studies on beam wander,” in Proc. Int. Conf. Space Opt. Syst. and Appl. ICSOS) 2014, (2014). Google Scholar

35. 

International Telecommunication Union (ITU), “Propagation data required for the design of earth-space systems operating between 20 THz and 375 THz,” Google Scholar

36. 

D. L. Walters and K. E. Kunkel, “Atmospheric modulation transfer function for desert and mountain locations: the atmospheric effects on r0,” J. Opt. Soc. Am., 71 397 –405 https://doi.org/10.1364/JOSA.71.000397 JOSAAH 0030-3941 (1981). Google Scholar

37. 

L. B. Stotts and L. C. Andrews, “Improving the Hufnagel-Andrews-Phillips refractive index structure parameter model using turbulent intensity,” Opt. Express, 31 (9), 14265 –14277 https://doi.org/10.1364/OE.488544 OPEXFF 1094-4087 (2023). Google Scholar

38. 

H. A. Pike, Google Scholar

39. 

B. García-Lorenzo, J. J. Fuensalida and M. A. C. Rodríguez-Hernández, “Statistical turbulence vertical profiles at the Roque de los Muchachos Observatory and Teide Observatory,” Proc. SPIE, 6747 67470C https://doi.org/10.1117/12.737745 (2007). Google Scholar

40. 

F. Quatresooz, D. Vanhoenacker-Janvier and C. Oestges, “Computation of optical refractive index structure parameter from its statistical definition using radiosonde data,” Radio Sci., 58 (1), e2022RS007624 https://doi.org/10.1029/2022RS007624 (2023). Google Scholar

41. 

D. A. Mahmood et al., “Improving Hufnagel-Andrews-Phillips model for prediction using empirical wind speed profiles,” J. Atmos. Solar-Terr. Phys., 240 105952 https://doi.org/10.1016/j.jastp.2022.105952 JASPF3 1364-6826 (2022). Google Scholar

42. 

G. J. Baker and R. S. Benson, “Gaussian-beam weak scintillation on ground-to-space paths: compact descriptions and Rytov-method applicability,” Opt. Eng., 44 106002 https://doi.org/10.1117/1.2076927 (2005). Google Scholar

43. 

H. T. Yura and D. A. Kozlowski, “Low Earth orbit satellite-to-ground optical scintillation: comparison of experimental observations and theoretical predictions,” Opt. Lett., 36 (13), 2507 –2509 https://doi.org/10.1364/OL.36.002507 OPLEDP 0146-9592 (2011). Google Scholar

44. 

H. Yura and W. McKinley, “Aperture averaging of scintillation for space-to-ground optical communication applications,” Appl. Opt., 22 1608 –1609 https://doi.org/10.1364/AO.22.001608 APOPAI 0003-6935 (1983). Google Scholar

45. 

L. W. Bradford, “Maui4: a 24-hour Haleakala turbulence profile,” in Proc. Adv. Maui Opt. and Space Surveill. Technol. Conf. (AMOS 2010), 499 –516 (2010). Google Scholar

46. 

N. Perlot et al., “Results of the optical downlink experiment KIODO from OICETS satellite to Optical Ground Station Oberpfaffenhofen (OGS-OP),” Proc. SPIE, 6457 645704 https://doi.org/10.1117/12.708413 PSISDG 0277-786X (2007). Google Scholar

47. 

F. Strömqvist Vetelino et al., “Aperture averaging effects on the probability density of irradiance fluctuations in moderate-to-strong turbulence,” Appl. Opt., 46 (11), 2099 –2106 https://doi.org/10.1364/AO.46.002099 APOPAI 0003-6935 (2007). Google Scholar

48. 

L. B. Stotts et al., “Hybrid optical RF airborne communications,” Proc. IEEE, 97 1109 –1127 https://doi.org/10.1109/JPROC.2009.2014969 IEEPAD 0018-9219 (2009). Google Scholar

49. 

J. C. Juarez et al., “High-sensitivity DPSK receiver for high-bandwidth free-space optical communication links,” Opt. Express, 19 (11), 10789 –10796 https://doi.org/10.1364/OE.19.010789 OPEXFF 1094-4087 (2011). Google Scholar

Biography

Larry B. Stotts is a consultant. His interest areas are RF and optical communications, and RF, infrared, and visible surveillance, and reconnaissance. He received his BA degree in applied physics and information sciences and his PhD in electrical engineering (communications systems), both from the University of California at San Diego. He has published over 113 journal articles and authored/co-authored four books. He is a fellow of the IEEE, SPIE, and Optica.

Larry C. Andrews is a Professor Emeritus of Mathematics at the University of Central Florida (UCF) and an associate member of the Townes Laser Institute in the College of Optics/CREOL at UCF. He received his PhD in theoretical mechanics from Michigan State University. He has authored/coauthored 12 textbooks and has published numerous papers and reports. He is a fellow of SPIE and authored three SPIE field guides.

© 2023 Society of Photo-Optical Instrumentation Engineers (SPIE)
Larry B. Stotts and Larry C. Andrews "Optical communications in turbulence: a tutorial," Optical Engineering 63(4), 041207 (8 December 2023). https://doi.org/10.1117/1.OE.63.4.041207
Received: 28 April 2023; Accepted: 17 November 2023; Published: 8 December 2023
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Signal to noise ratio

Scintillation

Receivers

Optical communications

Turbulence

Optical engineering

Sensors

Back to Top