ArticlePDF Available

T-PPM: A Novel Modulation Scheme for Optical Communication Systems Impaired by Pulse-Width Inaccuracies

Authors:

Abstract and Figures

M-ary pulse-position modulation (PPM) has received considerable attention for direct-detection photon communications over unguided channels. The analysis generally assumes that the signaling set is orthogonal. However, the orthogonality of the signaling set will be destroyed by the finite area and bandwidth of optical detectors, resulting in severe intersymbol interference. This article presents the analysis of a trellis-based pulse-position modulation (T-PPM) scheme for photon communications with non-rectangular pulses. The novelty of the scheme includes the use of a set-partitioning methodology to increase the minimum distance using a simple convolutional encoder. The Viterbi algorithm is used at the receiver to separate the signaling set as part of the demodulation process. It has been shown that T-PPM will restore performance losses due to reduced peak intensity during the detection process. Furthermore, for a large range of background radiation levels, the average number of photons per information bit for T-PPM is smaller than that of the regular PPM. Numerical examples show that for a symbol-error rate of 10±3 when the received pulses extend over 4 PPM slots, the average laser energy per symbol for 256-ary T-PPM could be reduced by as much as 2 dB.
Content may be subject to copyright.
TMO Progress Report 42-135 November 15, 1998
T-PPM: A Novel Modulation Scheme for Optical
Communication Systems Impaired by
Pulse-Width Inaccuracies
K. Kiasaleh1and T.-Y. Yan2
M-ary pulse-position modulation (PPM) has received considerable attention for
direct-detection photon communications over unguided channels. The analysis gen-
erally assumes that the signaling set is orthogonal. However, the orthogonality of
the signaling set will be destroyed by the finite area and bandwidth of optical detec-
tors, resulting in severe intersymbol interference. This article presents the analysis
of a trellis-based pulse-position modulation (T-PPM) scheme for photon commu-
nications with non-rectangular pulses. The novelty of the scheme includes the use
of a set-partitioning methodology to increase the minimum distance using a simple
convolutional encoder. The Viterbi algorithm is used at the receiver to separate the
signaling set as part of the demodulation process. It has been shown that T-PPM
will restore performance losses due to reduced peak intensity during the detection
process. Furthermore, for a large range of background radiation levels, the average
number of photons per information bit for T-PPM is smaller than that of the regu-
lar PPM. Numerical examples show that for a symbol-error rate of 103when the
received pulses extend over 4 PPM slots, the average laser energy per symbol for
256-ary T-PPM could be reduced by as much as 2 dB.
I. Introduction
Pulse-position modulation (PPM) has received considerable attention as the modulation of choice for
direct-detection optical communications over unguided optical channels. There are several key aspects of
this modulation that are critical to the deployment of this modulation scheme for deep-space communica-
tions. First, the presence of a pulse in the symbol frame regardless of the transmitted symbol benefits the
clock recovery subsystem, whereas an on–off-keying (OOK) system may suffer a synchronization loss if a
sequence of 0’s is encountered. Also, if nonreturn-to-zero (NRZ) pulses are used, a sequence of 1’s also
can disrupt the synchronization subsystem of an OOK system. Unlike the OOK scenario, one does not
require a priori knowledge of the signal or background noise radiation levels to implement an optimum
PPM receiver. The other key requirement of systems considered for space applications is the peak laser
power level that must be large enough to survive huge deep-space losses. For this reason, Q-switched
lasers typically are employed for such applications. The current technology, however, does not support
a scenario where a Q-switched laser can be toggled between the “on” and “off” states at a high rate,
1Erik Jonsson School of Engineering and Computer Science, The University of Texas at Dallas, Richardson, Texas.
2Communications Systems and Research Section.
1
limiting severely the data rate that can be supported using an OOK scheme. Hence, an M-ary PPM
with large Mis more suitable than its OOK counterpart for deep-space applications.3Finally, it can
be shown that M-ary PPM for large Mis more energy efficient than its OOK counterpart, which is of
critical importance in a deep-space optical communication environment where energy consumption is a
key constraint of the channel.
For many deep-space applications, due to large distances and pointing inaccuracies, the signal level
received at the detector is significantly attenuated so that only a small number of photons per PPM slot
interval are typically observed. Since thermal noise is present at the receiver, the detection of signal pulses
in the above condition is significantly hampered. For this reason, avalanche photodetectors (APDs) are
utilized to boost the signal level over the additive noise level present at the receiver. An APD, in principle,
magnifies each incident photon to a large number of post-detection electrons. One major problem with
today’s APDs is the excess noise factor of the APD, which manifests itself as a random gain effect. That
is, the ratio of the number of emitted electrons in response to an incident photon is a random variable
with known statistics. Unfortunately, the complicated statistics of the APD response to incident photons
do not lead to a closed-form solution for the error rate of optical PPM communication systems.
One key approximation for the statistics of the number of emitted electrons when the received number
of photons obeys the Poisson statistics is due to Webb [1]. This approximation is shown to be fairly accu-
rate over a wide range of conditions, and for this reason, this approximation is widely used to assess the
error rates of various optical communication systems with APDs. When the number of incident photons
in a slot interval is fairly large (as compared with the APD noise figure) and the APD possesses a large
average gain, it has been shown that the APD’s statistic could be approximated by a Gaussian proba-
bility density function (pdf) [2,3]. This approximation leads to a closed-form bit-error-rate expression
for the PPM channels and usually is valid when large background radiation levels are present (daytime
operation). However, for a large number of cases when the background radiation level is small, the Gaus-
sian approximation fails to accurately predict the statistics of an APD detector. In fact, for a small
number of incident photons, the probability mass function (PMF) of the number of released electrons
significantly departs from a Gaussian form [2,3]. In that case, one has to resort to the Webb–McIntyre–
Conradi (WMC) pdf, which does not lead to a closed-form expression for the probability of error [1].
More significantly, the resulting expression requires a prohibitively large computing time.
In some recent studies, via extensive simulation, it has been demonstrated that when the thermal
noise level is non-negligible, the number of signal photons required to achieve a symbol-error rate in a
range of 102to 103using the Gaussian-approximated model and that predicted by the WMC model
are somewhat similar for background noise levels in excess of 1 photon per PPM slot. Since, for most
applications of interest, one observes in excess of 1 background photon per PPM slot, the Gaussian
model seems to be a reasonable model to exploit for the above scenario. This by no means implies that
the Gaussian assumption is a valid approximation for all symbol-error-rate levels. It only underscores
the negligible loss/gain in the signal photon count associated with the Gaussian assumption for symbol-
error rates in the range of 102to 103. It also is imperative to note that, for most applications, an
uncoded symbol-error rate of 102to 103is of interest. Obviously, with the inclusion of a forward error-
correcting code (FEC), a symbol-error rate of 106, which is required to support data communication,
can be achieved.
Another generic property of optical sensors for detecting narrow laser pulses is the requirement of
high-speed electronics to convert optical signals to electrical currents. In general, for deep-space direct-
detection applications, larger detector areas potentially could collect more photons, thus increasing the
3In an M-ary PPM scenario with large M, the laser remains in an “on” state for 1/M th of the symbol duration, allowing
for log2Mbits to be transmitted in a symbol interval. In an OOK scheme, given the same symbol duration (note that
there exists an upper limit on how fast the laser can be toggled between the two states), only 1 bit can be transmitted
over the same symbol interval.
2
number of photons per information bit. Unfortunately, larger detector sizes imply a lower bandwidth of
the detector, which tends to smear the observed narrow pulses over several PPM slots. The proposed
T-PPM excels in this area by allowing overlapping of the PPM signaling set. That is, as will be demon-
strated in this article, the proposed T-PPM could circumvent this effect without degrading the system
end-to-end performance. Furthermore, it will be shown in the ensuing sections that T-PPM can achieve a
performance similar to that of its PPM counterpart at a reduced number of photons per information bit.
In addition, T-PPM can accommodate lower peak laser power during slot detection by integrating ener-
gies over several slots. The lower peak power is of critical importance in energy-constrained, deep-space
communication systems.4
In this article, we are interested in assessing the impact of T-PPM on enhancing the end-to-end
performance of optical communication systems with pulse-width inaccuracies, which often are present
when high-power Q-switched lasers are employed. The presence of such inaccuracies results in a significant
degradation in the overall performance of conventional PPM systems.
II. System Model
For an M-ary PPM system, the transmitted laser power, P(t), may be described as
P(t)=Ps
X
q=−∞
h(tCqTslot NqTs) (1)
where Psis the peak laser power in watts; h(t) is the pulse shape, which is confined to a PPM symbol
duration Tss; Tslot is the PPM slot duration in seconds; Cqis the PPM symbol taking on the set
{0,1,···,M 1}with equal probability, with Mdenoting the PPM alphabet size; and N1isa
parameter that will be defined shortly. Note that Tslot =Ts/M s. Ideally, the pulse shape h(t) is confined
to a PPM slot duration. In that case, one obtains an ideal PPM signaling scheme. In the scenario
considered here, however, it is assumed that h(t) extends beyond a slot interval, causing significant
degradation in performance. Also, note that, in the formulation shown above, Nis introduced to allow
for one to generate a “silent” period in between symbol durations in which a PPM pulse is transmitted.
This typically is a requirement for operating a high-power laser when the transmission of a symbol is
interrupted by silent periods of several-symbol duration to allow for the recharging of the laser.
Given an unguided optical channel, one can describe the intensity of the received optical signal, λr(t),
as
λr(t)=λb+λs
X
q=−∞
h(tCqTslot NqTs+²qTslot) (2)
where λs=Prη/hν denotes the peak intensity of the received optical signal in photons/s with η,Pr,
h, and νdenoting the quantum efficiency of the detector, the received power collected by the optical
assembly, Planck’s constant, and the operating frequency of the laser in hertz, respectively. Moreover,
λbin Eq. (2) denotes the intensity of the background noise in photons/s, and ²qdenotes the timing
inaccuracy present due to imperfect time tracking at the receiver, which is assumed to be confined to
[1/2,1/2].
4For energy efficiency, one must resort to a large alphabet size PPM scenario. In that event, and for a relatively high data
rate, one must employ fairly short laser pulses. Given the fixed required energy, this implies that the laser must generate
a large peak power. The large peak power then becomes a major constraint of the system.
3
Since an APD is employed at the receiver, the collected primary electrons cause the generation of
a large number of secondary electrons. In general, one can assume that for the mth detected primary
electrons, the APD generates Gmsecondary electrons. The cumulative generating function (CGF) of Gm
is implicitly given by [4]
s=µGm(s)bln ha+(1a)eµGm(s)i(3)
where µGm(s) denotes the CGF, a=[1+κg1)]/¯g, and b=1/(1κ), with ¯gand κdenoting the average
gain and the ionization factor of the APD, respectively. Let F=[E¡G2
m¢]/¯g2=κ¯g+(2[1/¯g]) (1 κ)
denote the APD noise figure. Then, it can be shown that the moment-generating function of Gmmay be
expressed explicitly as [5]
MGm(s)= F
(F1)2h1p12(F1) ¯gsi¯g
F1s+ 1 (4)
Note that, by definition, MGm(s)=eµGm(s).
The response of the APD, in terms of the current in the load resistor of the APD, may be modeled as
follows:
x(t)=
X
j=−∞
GjR(ttj)+n(t) (5)
where tjdenotes the occurrence time of the jth primary electron, governed by a Poisson point process;
R(t) is the current response of the APD; and n(t) is the thermal noise present at the receiver. The n(t)
commonly is modeled as a zero-mean, additive white Gaussian noise (AWGN) with a two-sided power
spectrum density (PSD) level of N0/2=(2kT0)/RLA2/Hz with k,T0, and RLdenoting Boltzmann’s
constant, the effective temperature of the receiver in kelvins, and the APD load resistance in ohms (Ω).
At the receiver, the PPM detector performs Mdistinct integrations of the observed current over the
Mslots of the PPM signal. The symbol associated with the slot with a maximum integrated current is
declared as the demodulated symbol. This detection strategy is optimum when the integrated current
is governed by Gaussian statistics or when a shot-noise-limited (Poisson-statistics) scenario is considered
[2]. It also has been demonstrated that when the statistics of the APD are approximated by WMC
distribution and the observed signal is further corrupted by AWGN, the optimum detection strategy
is identical to that used for purely Gaussian or Poisson channels [6]. Therefore, we use the detection
mechanism outlined above to render a decision regarding the transmitted symbol.5
Let us define Xl,n as the outcome of integrating x(t) over the interval Il,n =[lTslot +NnTs,(l+1)Tslot
+NnTs]. Note that Il,n denotes the lth slot of the nth transmitted symbol.6We then have
Xl,n =
(l+1)Tslot+NnTs
Z
lTslot+NnTs
x(τ);l=0,1,···,M 1 (6)
5In the presence of timing error, the above method is no longer optimal. Since in the ensuing analysis we consider negligible
timing error as compared with the pulse-width inaccuracies, we resort to the aforementioned detection mechanism.
6In the ensuing analysis, and without loss of generality, we consider N= 4 (each transmitted symbol is followed by a
silent period of 3Tss). The selection of N= 4 is dictated by the requirements that currently are imposed on high-power
Q-switched lasers that are being considered for deep-space communication.
4
Given that the APD’s current response is such that it can be approximated with a delta function, we
arrive at
Xl,n =eK2(l, n)+νN(7)
where νNis a zero-mean Gaussian random variable with variance
σ2
ν=2kT0
RL
Tslot (8)
eis the charge of an electron in C, and K2(l, n) is the number of secondary electrons observed over
Il,n. The PMF of K2(l, n) in its exact form originally was discovered by McIntyre [7] and was verified
experimentally by Conradi [8]. This PMF is given by
Pr {K2(l, n)=k2|K1(l, n)=k1}=
k1Γ³k2
1κ+1
´
k2(k2k1)!Γ µκk2
1κ+1+k1
×·1+κg1)
¯g¸k1+κk2/(1κ)·(1 κ)(¯g1)
¯g¸k2k1
(9)
where K1(l, n) is the number of primary electrons observed over the interval Il,n. Given that the primary
electron statistics obey a Poisson model, Webb has demonstrated that the PMF of the secondary electron
count may be approximated by
Pr ©K2(l, n)=k2|¯
K1(l, n)=¯
k1ª=1
(2πC2
1)1/2
1
1+¡k2¯
k1¯g¢
C1C2
3/2
exp
¡k2¯
k1¯g¢2
2C2
1"1+¡k2¯
k1¯g¢
C1C2#
(10)
where ¯
K1(l, n) is the average number of primary electrons observed over Il,n,C2
1g2¯
k1F, and C2
2=
¯
k1F/(F1)2. Given the Poisson statistics, then
¯
K1(l, n)=
(l+1)Tslot+NnTs
Z
lTslot+NnTs
λr(τ) =Kb+KsYl,n (11)
where
Yl,n =1
Tslot
(l+1)Tslot+NnTs
Z
lTslot+NnTs
X
q=−∞
h(τCqTslot NqTs+²qTslot)
5
and Kb=λbTslot denotes the average number of primary electrons observed over a PPM slot interval
due to background noise. Moreover, Ks=λsTslot. Note that the above formulation implies that the
average number of primary electrons due to signal power observed over the lth slot of the nth symbol,
i.e., KsYl,n, decreases as the pulses smear in time, maintaining the average number of primary electrons
observed over a symbol interval (i.e., Ks) constant. That is, we assume that the energy observed over a
symbol interval remains constant as pulses spread over the adjacent slots, leading to less observed energy
per slot.
In the ensuing analysis, we assume that ²qis negligible (since it is a fraction of a slot interval), whereas
the intersymbol interference caused by the shape of the nonideal pulse shape, h(t), is assumed to be quite
significant. More precisely, we assume that h(t) extends beyond a slot boundary and perhaps extends
over 2 to 4 slot intervals. This causes severe degradation in performance in the absence of a strategy
to circumvent intersymbol interference (ISI). In what follows, we assume that the laser pulse obeys a
Gaussian shape. Namely,
h(t)= Tslot
p2πσ2
h
exp
µtTslot
22
2σ2
h
(12)
where we have assumed that the laser pulse is centered around the midpoint of the slot interval and that
the pulse has a standard deviation of σhs. Moreover, note that R
−∞ h(τ) =Tslot. This assumption
is motivated by the fact that if h(t) is replaced with an ideal nonreturn-to-zero (NRZ) pulse of unit
amplitude, a similar result will be obtained. This, in turn, implies that regardless of the value of σ2
h, the
total laser energy over a symbol interval remains constant. That is, as the pulses are smeared in time, the
peak laser power is reduced, maintaining the constant-energy assumption that is critical to space-borne
optical communication systems. Considering that h(t) extends over a few slot intervals and that we have
a silent period of 3Mslots, and assuming that Cn=j,wehave
Zl,j =Yl,n |Cn=j=1
σh2π
(l+1)Tslot
Z
lTslot
exp
µτ(2j+1)Tslot
22
2σ2
h
(13)
Note that the integral on the right-hand side of the above equation is a function of land jonly.
III. Performance Analysis—PPM With Imperfect Pulse Shape
Given the above model, one can obtain the performance of a PPM system with imperfect pulses. A
key obstacle, however, is the absence of a tractable model for the statistics of the APD signal when
corrupted by an AWGN. In a recent simulation study, it was demonstrated that the Ksrequired for
an APD-detected 256-ary PPM signal to achieve an error rate in the range of 102to 103and that
predicted using a Gaussian model differ only slightly over a wide range for Kb(1 Kb1000), allowing
one to benefit from the Gaussian approximation for the above range of symbol-error rates. Furthermore,
we benefit from a union bound to establish an upper bound on the error rate of the uncoded PPM system
with imperfect pulse shapes. To that end,
P(u)
PPM 1
M
M1
X
j=0
M1
X
l=0;j6=l
Pr (El,j) (14)
6
where we have assumed equally likely PPM symbols. In the above equation, Pr (El,j ) is the pair-wise error
probability of making a decision in favor of the lth symbol when the jth symbol actually is transmitted
and P(u)
PPM is the symbol-error rate of the uncoded PPM system. This error event is possible only when
the integrated current of the lth slot exceeds that of the jth slot. As noted earlier, we resort to Gaussian
approximation for the range of error rates stated previously. If one allows for such an approximation,
then Xl,n is a Gaussian random variable when conditioned on Cq(assume ²q= 0). Hence, the mean and
variance of this random variable will be of interest to establish performance. To that end,
ml,j =E{Xl,n|Cn=j}=e¯g(Kb+KsZl,j)+Idc Tslot (15)
and
σ2
l,j =Var{Xl,n|Cn=j}=(e¯g)2F(Kb+KsZl,j)+σ2
ν+eIdcTslot (16)
where E{.|Cn=j}and Var{.|Cn=j}are the expected value and the variance of the random variable
Xl,n conditioned on {Cn=j}and Idc denotes the surface dark current of the APD in amperes, which is
assumed to be non-negligible here. In that event, and considering that an erroneous decision in favor of
the lth symbol is rendered if the integrated APD current over the lth slot exceeds that of the jth slot
(note that this is a comparison between Xl,n and Xj,n, which are a pair of conditionally [when conditioned
on {Cn=j}] independent Gaussian random variables), we have
Pr (El,j)=1
2erfc
mj,j ml,j
r2³σ2
l,j +σ2
j,j ´
;l6=j(17)
where erfc (x) is the complementary error function. Given that the pulse shape follows a Gaussian pattern,
Zl,j and Zj,j may be obtained in terms of erfc (.) or erfc(.). However, we note that
mj,j ml,j =e¯gKs(Zj,j Zl,j) (18)
and
σ2
l,j +σ2
j,j =(e¯g)2F[2Kb+Ks(Zl,j +Zj,j)]+2¡σ2
ν+eIdcTslot ¢(19)
are functions of Zj,j Zl,j and Zj,j +Zl,j only. Therefore, Zj,j Zl,j and Zj,j +Zl,j are only of interest.
With some effort, it may be shown that
Zj,j Zl,j =1erfc µTslot
22σh1
2erfc
Tslot µlj+1
2
2σh
+1
2erfc
Tslot µlj1
2
2σh
(20)
and
7
Zj,j +Zl,j =1erfc µTslot
22σh+1
2erfc
Tslot µlj+1
2
2σh
1
2erfc
Tslot µlj1
2
2σh
(21)
Now, substituting Eqs. (20) and (21) in Eqs. (18) and (19), respectively, and then substituting the
resulting expressions in Eq. (17), we arrive at a closed-form expression for the pair-wise error rate of a
PPM system with pulse-width inaccuracies. The remaining task is to compute the upper bound using
Eq. (14). We postpone a discussion on the performance of the system using Eq. (14) to Section VI of this
article. Instead, we proceed to describe the proposed T-PPM system in the next section.
IV. T-PPM
It is important to note that the use of trellis-coded modulation (TCM) and overlapping PPM (OPPM)
to enhance the capacity of PPM channels was originally discussed in [9]. There are several aspects of the
present analysis that set it apart from its predecessor. First, the analysis in [9] assumes square shape
pulses, whereas we will concern ourselves with Gaussian-type pulse shapes in this analysis. Second, the
analysis in [9] assumes either quantum-limited or shot-noise-limited scenarios, whereas a more realistic
situation (as is assumed in what follows) calls for the inclusion of additive Gaussian noise and APD excess
noise factors.
In the previous section, we described the performance of a PPM system impaired by imperfect pulse
shapes. Needless to say, the impact of an imperfect pulse shape on the performance of a PPM system
may be quite severe for pulse shapes extending over several slot intervals. Hence, the introduction of
TCM is necessary to overcome the impact of severe ISI caused by imperfect pulse shapes. Due to the
regular shape of the pulse, and from Eq. (17), it can readily be seen that Pr (El,j ) is a decreasing function
of |jl|. That is, the pair-wise error rate is a decreasing function of the “distance” between symbols
jand l, where distance is defined in terms of the number of PPM slots that exit between the positions
(in time) of the jth and lth symbols.7In view of this observation, one can proceed with a set-partition
strategy for the generation of T-PPM signals that is aimed at increasing the minimum distance of the
PPM constellation.
We begin by introducing A={1,2,3,···,M}as a set containing the uncoded PPM symbols. Note
that the distance between a pair of symbols jand lselected from Ais simply |jl|. This set is further
divided into sets B0={1,3,5,7,···} and B1={2,4,6,···}. Note that a pair of symbols selected
from either one of these sets is at a minimum distance of 2, whereas the symbols selected from Aare at a
minimum distance of 1. Next, we proceed to subdivide set B0to generate sets C0={1,5,9,···}and C2=
{3,7,11,···}. Set B1also can be subdivided to form C1={2,6,10,···}and C3={4,8,12,···}. Finally,
we perform one last partitioning of the previous sets to form D0={1,9,17,···},D4={5,13,21,···},
D2={3,11,19,···},D6={7,15,23,···},D1={2,10,18,···},D5={6,14,22,···},D3={4,12,20,···},
and D7={8,16,24}. The entire process is depicted in Fig. 1 for M= 256. Note that symbols selected
from Djfor any jare at a minimum distance of 8. Furthermore, symbols selected from Djand Diare at
a minimum distance of |ij|from each other. Although it appears that one can continue this process
in hope of increasing the minimum distance beyond 8, the minimum distance between a pair of paths
through the trellis (which ultimately dictates the overall performance) cannot be increased indefinitely
with further set partitioning. Also, since we are interested only in eliminating the impact of ISI, a
minimum distance of 4 is sufficient to establish orthogonality among the symbols in the set. That is,
when the imperfect pulse h(t) stretches over 2 to 4 PPM slots, a minimum distance of 4 slots among the
symbols selected from Cj(for all j) ensures that there exists no overlap among the pulses in the set.
7Since the jth symbol is represented by a pulse in the jth slot, the distance between the jth and lth symbols is |jl|.
8
d
min
= 4
d
min
= 1
A
= {1,2,3, ..., 256}
01
01
01
B
0 = {1,3,5,7, ...}
B
1 = {2,4,6, ...}
C
0 = {1,5,9, ...}
C
2 = {3,7,11, ...}
C
1 = {2,6,10, ...}
C
3 = {4,8,12, ...}
d
min
= 2
01
D
0 = {1,9,17,...}
D
4 = {5,13,21,...}
D
2 = {3,11,19,...}
D
6 = {7,15,23,...}
D
1 = {2,10,18,...}
D
5 = {6,14,22,...}
D
3 = {4,12,20,...}
D
7 = {8,16,24,...}
d
min
= 8
010101
Fig. 1. Set partitions for the 256-ary PPM signal set.
Without loss of generality, we concern ourselves with rate 1/2 and 2/3 convolutional encoders (CEs)
for the generation of T-PPM signals. As will be shown later, the rate 1/2 CE with a 4-state trellis is
quite suitable for the problem at hand, although the rate 2/3 CE with an 8-state trellis can offer a larger
minimum distance (dmin). We note also that the alphabet size of the PPM is 256 (8 bits), which allows
one to encode 7 data bits per T-PPM symbol (for both the rate 1/2 and rate 2/3 CEs, see Figs. 2 and 3).
Given that for high-power lasers used in deep-space communications the frame rates usually are kept
constant, this implies a loss in the data rate by a factor of 7/8. This, however, is a relatively small price
to pay to circumvent the substantial ISI caused by imperfect pulses. The rate 1/2 and 2/3 convolutional
encoders along with their respective trellises are depicted in Figs. 2 and 3, respectively. Note that, for
the rate 1/2 CE case, a 4-state trellis is used and that the 2 bits produced by the encoder are used to
select a set from the four possible C-type sets: C0,C1,C2, and C3. The remaining 6 bits are used to
select a signal from the selected set (note that there are 26signals in any of the four C-type sets). For
the rate 2/3 code, which, as shown in Fig. 3, corresponds to an 8-state trellis, the 3 bits produced by
the CE are used to select one of the eight possible D-type sets. Given that the code is a rate 2/3 CE,
then the remaining 5 bits are used to select a signal from the selected set. Once again, note that there
are 25signals in any of the eight D-type sets. The above arrangement, then, leads to the existence of a
number of parallel paths in the trellises depicted in Figs. 2 and 3. That is, for each path in the trellis
shown in Fig. 2, there exist 64 parallel paths (the size of the C-type sets). This is a common feature of
TCM systems that is due to the existence of uncoded bits used to select a signal from the constellation.
The number of parallel paths for the trellis in Fig. 3 is 32, which is identical to the size of any of the
D-type sets.
V. T-PPM Performance Analysis
In Figs. 2 and 3, we have depicted the trellises for the rate 1/2 and 2/3 T-PPM systems with 4 and
8 states, respectively. When the rate 1/2, 4-state trellis is used, one easily can conclude that the minimum
distance between the all-zero path and any path that departs from and re-emerges with the all-zero path
for the first time is 4. The path that leads to the minimum distance of 4 is due to a symbol selected
from C0. For the 8-state trellis, we also have depicted the path (dotted line) that is at a minimum
distance of 5 from the all-zero path. Note that parallel transitions in this trellis are selected from the
D-type sets, and, hence, a minimum distance of 8 exists between any pair of parallel paths in Fig. 3. As
noted earlier, although we have been able to increase the minimum distance between parallel paths in the
trellis via further set partitioning, the minimum distance between any path in the trellis and the all-zero
path has not increased dramatically. Due to the complexity of the trellis in Fig. 3 and the fact that a
9
d
C
0
C
2
C
2
C
0
C
1
C
3
C
3
C
1
d
= 0
d
= 1
(00)
(01)
(10)
(11)
C
1
C
1
C
3
C
0
C
0
C
2
C
2
C
0
C
1
C
3
C
3
C
1
Fig. 2. The rate 1/2 convolutional encoder along with its 4-state trellis that uses the set partitioning of Fig. 1 for the
generation of the rate 7/8 T-PPM signals.
D
0
D
0
D
0
D
0
D
4
D
2
D
6
D
0
D
1
D
2
d
min
= 5
D
0
D
4
D
2
D
6
D
1
D
5
D
3
D
7
D
4
D
0
D
6
D
2
D
5
D
1
D
7
D
3
D
2
D
6
D
0
D
4
D
3
D
7
D
1
D
5
D
6
D
2
D
4
D
0
D
7
D
3
D
5
D
1
d
0
d
1 = 11
01 10
00
d
0
d
1
Fig. 3. The rate 2/3 convolutional encoder along with its 8-state trellis that uses the set partitioning of Fig. 1 for the
generation of the rate 7/8 TCM-PPM signals.
10
minimum distance of 4 is sufficient to establish orthogonality among the symbols in the set, in what
follows, we limit our discussion to the rate 1/2, 4-state trellis depicted in Fig. 2. To that end, let us
consider the parallel paths in the trellis. Given that for any path there exist 63 other parallel paths for
M= 256 in Fig. 2 (see Fig. 1), an upper bound on the error rate may be obtained with the aid of a union
bound. This upper bound is given by
P(TCM)
PPM 63 Pr (El,l+4)
where Pr (El,l+4) denotes the pair-wise error rate for a pair of PPM symbols with imperfect pulse shapes
that are separated by 4 time slots. When h(t) extends over 2 to 4 times the duration of a slot, then
Pr (El,l+4) is not a function of l.
VI. Numerical Results
To underscore the impact of imperfect pulse shapes, a pair of adjacent Gaussian-type pulses are plotted
in Figs. 4 through 6. The laser energy (the area covered by the pulse) within one slot time is computed to
be 98, 78.8, and 59.5 percent, respectively, in Figs. 4 through 6. Note that, for a slot duration of 20 ns, for
instance, the case depicted in Fig. 6 corresponds to an 80-ns pulse scenario. It is immediately obvious that
ISI increases in an exponential fashion with σh. Since we are interested in a scenario where h(t) is limited
to 2 to 4 times the duration of a time slot, we limit our analysis to the cases where 0 σh0.6Tslot s.
2.0
1.8
1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
-4-3-2-101234
t
/
T
slot
GAUSSIAN PULSE
Fig. 4. A pair of adjacent Gaussian-type
pulses with s
h
= 0.2
T
slot.
1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
-4-3-2-101234
t
/
T
slot
GAUSSIAN PULSE
Fig. 5. A pair of adjacent Gaussian-type
pulses with s
h
= 0.4
T
slot.
0.7
0.6
0.5
0.4
0.3
0.2
-4-3-2-101234
t
/
T
slot
GAUSSIAN PULSE
Fig. 6. A pair of adjacent Gaussian-type
pulses with s
h
= 0.6
T
slot.
0.1
0.0
11
We first consider the case where a symbol-error rate of 102is of interest, and subsequently we obtain
the required Ksto achieve the desired error rate for a given Kb. Without loss of generality, we limit our
discussion to the following system parameters: M= 256, Tslot = 20 ns, ¯g= 40, κ=0.007, T0= 100 K
(cooled receiver), RL= 146.65 kΩ, and Idc = 2 nA. In Figs. 7 and 8, the Ksrequired to achieve a
symbol-error rate of 102is plotted versus Kbwhen the above set of parameters is used. We note that
the Gaussian assumption considered here does not lead to performance measures that agree with the
simulation results using WMC statistics for the entire range of Kbwhen an error rate of 102is of
interest. In particular, when Kb<50 (Fig. 7), the Gaussian assumption is less reliable than for the case
of Kb>50 (Fig. 8). Nonetheless, the loss in Ksusing the Gaussian approximation is not significant,
and hence we proceed with our analysis using the results shown above. Before doing so, however, it is
imperative to note that the simulation results for the required Ksusing WMC statistics (for the above set
of parameters) are shown to be in close agreement with those predicted using the Gaussian assumption
for a wide range of Kbwhen a symbol-error rate of 103is considered [10]. Since an error rate of 103
is typically of interest, we consider the numerical results shown below for a 103error rate to be a good
approximation of the results obtained using the more realistic WMC statistics for modeling the APD
output statistics.
The first observation that can be made from Figs. 7 and 8 is that, when a severe ISI is present (see the
curves associated with 80- and 60-ns pulses), the performance degrades substantially as compared with
the perfect pulse-shape case. The second significant observation is that, in the event of having perfect
pulses, the performance improves using T-PPM. This may be attributed to the coding gain. Note that
1 of the 7 information bits almost always will be detected correctly, since the major error is due to parallel
transitions in the set.
We also note that the gain in performance as a result of using T-PPM increases almost exponentially
with an increase in σh. In Figs. 9 and 10, a similar set of results is depicted when an error rate of 103
is of interest. We focus on these results, since an error rate of 103typically is needed to ensure a coded
performance8of 106.
140
UPPER BOUND ON REQUIRED
K
s
120
100
80
60
40
20
160
0 5 10 15 20 25 30 35 40 45 50
K
b
Fig. 7. The
K
s
required to achieve a symbol-error
rate of 10-2 for 256-ary PPM with imperfect pulses
when
K
b
is less than 50.
s
h
= 0
s
h
= 0.2
T
slot
s
h
= 0.4
T
slot
s
h
= 0.6
T
slot
UNCODED PPM
T-PPM
s
h
= 0
s
h
= 0.2
T
slot
s
h
= 0.4
T
slot
s
h
= 0.6
T
slot
UNCODED PPM
T-PPM
240
220
200
180
160
140
120
260
50 100 150 200
K
b
Fig. 8. The
K
s
required to achieve a symbol-error
rate of 10-2 for 256-ary PPM with imperfect pulses
when
K
b
is more than 50.
100
80
60
UPPER BOUND ON REQUIRED
K
s
8It is anticipated that T-PPM will be used in conjunction with other more powerful coding schemes. That is, the information
bits provided to T-PPM may be viewed as coded channel symbols provided by an FEC encoder (such as a Reed–Solomon
encoder).
12
200
UPPER BOUND ON REQUIRED
K
s
180
160
140
120
100
80
220
0 5 10 15 20 25 30 35 40 45 50
K
b
Fig. 9. The
K
s
required to achieve a symbol-error
rate of 10-3 for 256-ary PPM with imperfect pulses
when
K
b
is less than 50.
s
h
= 0.2
T
slot
s
h
= 0.4
T
slot
s
h
= 0.6
T
slot
UNCODED PPM
T-PPM
60
40
s
h
= 0
s
h
= 0
s
h
= 0.2
T
slot
s
h
= 0.4
T
slot
s
h
= 0.6
T
slot
300
UPPER BOUND ON REQUIRED
K
s
250
200
150
100
350
50 100 150 200
K
b
Fig. 10. The
K
s
required to achieve a symbol-error
rate of 10-3 for 256-ary PPM with imperfect pulses
when
K
b
is more than 50.
UNCODED PPM
T-PPM
50
Before discussing such results, it is imperative to note that the proposed T-PPM system transmits
7 information bits, as compared with the 8 bits that are transmitted over each symbol duration for the
uncoded system. Hence, it is appropriate to consider the required number of photons per information
bit in comparing the T-PPM scenario with the uncoded PPM case. We then proceed to use the average
number of required photons per information bit as a measure of efficiency of the modulation scheme.
Hence, in Figs. 11 through 14, we depict the average number of photons per information bit that is
required to achieve the desired error rate for a wide range of background radiation photon counts. We
note, however, that the loss of 1 bit in order to achieve the desired performance, given that there exists a
limit on laser energy, is a reasonable consideration. Hence, we discuss the performance of T-PPM using
both the average number of photons as well as the average number of photons per information bit as a
function of background radiation level in what follows.
From Figs. 9 and 10, it can be concluded that when Kb= 1 (night operation), one requires Ks= 133
to achieve an error rate of 103when 80-ns pulses are used (20-ns PPM slots). This number is reduced to
Ks= 79 with T-PPM, a reduction of about 2.2 dB in the required laser energy. However, as noted earlier,
T-PPM conveys only 7 bits of information. If one uses the number of photons per bit, in that case the
required number of photons per bit reduces from 133/8=16.6to79/7=11.2 using T-PPM (see Fig. 13).
This is a substantial gain in the overall system efficiency. For Kb= 100, the Ksrequired to achieve an
error rate of 103when 80-ns pulses are used reduces from 249 to 187 using T-PPM, a reduction of nearly
1.2 dB in the required laser energy. The number of photons per bit improves from 31.3 to 26.7 using
T-PPM (see Fig. 14). It is important to note that for small pulse spreading (60-ns or smaller pulses), the
gain in performance is not substantial and, hence, the gain achieved using T-PPM is noticeable only when
substantial pulse-width inaccuracies are present. In fact, for all cases considered, the system efficiency in
terms of the number of photons per bit increases or remains the same when the pulse width is equal to
or less than 60 ns. However, when the pulse width is increased to 80 ns, a substantial improvement in
system efficiency is observed (see Figs. 11 through 14).
When an error rate of 102is of interest and Kb= 200, from Fig. 8 one can conclude that Ks=
250 photons is needed9when 80-ns pulses are used. This number is reduced to 209 photons for a T-
PPM system, a savings of 0.77 dB in the required laser power. The number of photons per bit improves
from 31.2 to 29.8 (see Fig. 12), a gain smaller than that observed for the 103error-rate scenario.
9Note that, for this large background radiation level, the Gaussian assumption is fairly accurate.
13
REQUIRED NO. OF PHOTONS PER BIT
15
10
20
0510
15 20 25 30 35 40 45 50
K
b
Fig. 11. The average number of photons per
information bit required to achieve a symbol-error
rate of 10-2 for 256-ary PPM with imperfect pulses
when
K
b
is less than 50.
s
h
= 0.2
T
slot
s
h
= 0.4
T
slot
s
h
= 0.6
T
slot
s
h
= 0
UNCODED PPM
T-PPM
5
REQUIRED NO. OF PHOTONS PER BIT
30
25
35
50 100 150 200
K
b
Fig. 12. The average number of photons per
information bit required to achieve a symbol-error
rate of 10-2 for 256-ary PPM with imperfect pulses
when
K
b
is more than 50.
s
h
= 0.2
T
slot
s
h
= 0.4
T
slot
s
h
= 0.6
T
slot
s
h
= 0
UNCODED PPM
T-PPM
15
20
10
REQUIRED NO. OF PHOTONS PER BIT
24
22
26
051015 20 25 30 35 40 45 50
K
b
Fig. 13. The average number of photons per
information bit required to achieve a symbol-error
rate of 10-3 for 256-ary PPM with imperfect pulses
when
K
b
is less than 50.
s
h
= 0.2
T
slot
s
h
= 0.4
T
slot
s
h
= 0.6
T
slot
s
h
= 0
UNCODED PPM
T-PPM
18
20
16
14
12
10
8
6
REQUIRED NO. OF PHOTONS PER BIT
30
25
35
50 100 150 200
K
b
Fig. 14. The average number of photons per
information bit required to achieve a symbol-error
rate of 10-3 for 256-ary PPM with imperfect pulses
when
K
b
is more than 50.
s
h
= 0.2
T
slot
s
h
= 0.4
T
slot
s
h
= 0.6
T
slot
s
h
= 0
UNCODED PPM
T-PPM
15
20
10
40
Finally, in Figs. 15 and 16, we examine the performance of T-PPM when the slot duration is reduced
with the remaining parameters kept fixed. Given that the contribution of thermal noise and APD dark
current to the output of the APD have variances that are directly proportional to the integration interval
(slot duration), as one decreases the slot duration, a gain in performance (in terms of a reduction in
the number of photons per information bit required to achieve a given error rate, 103) is observed.
More significantly, as one decreases the slot duration, for a fixed-background-radiation intensity level, the
average number of received photons per slot decreases accordingly. As seen from these figures, for the
2-ns slot scenario, a significant improvement in performance is observed. This result is not surprising.
To elaborate, as one decreases the slot duration with the total number of signal photons per slot kept
constant (i.e., when a higher-power laser is utilized), the performance approaches that of a quantum-
limited system, leading to the considerable gain in performance observed in Figs. 15 and 16.
14
REQUIRED NO. OF PHOTONS PER BIT
16
14
18
0.0 0.5 1.0 2.0
Fig. 15. The average number of photons per
information bit required to achieve a symbol-error
rate of 10-3 for T-PPM with imperfect pulses
(background intensity less than 2.5 109
photons/s). For all cases, s
h
= 0.6
T
slot.
T
slot = 2 ns
10
12
8
20
BACKGROUND INTENSITY, photons/s ( 109)
2.5
22
T
slot = 20 ns
T
slot = 16 ns
T
slot = 8 ns
1.5 8
REQUIRED NO. OF PHOTONS PER BIT
25
20
30
234 6
Fig. 16. The average number of photons per
information bit required to achieve a symbol-error
rate of 10-3 for T-PPM with imperfect pulses
(background intensity more than 2.5 109
photons/s). For all cases, s
h
= 0.6
T
slot.
T
slot = 2 ns
10
15
35
BACKGROUND INTENSITY, photons/s ( 109)
7
40
T
slot = 20 ns
T
slot = 16 ns
T
slot = 8 ns
5910
VII. Conclusions
This article introduced a robust trellis-based pulse-position modulation (T-PPM) as a technique for
deep-space optical communication and analyzed its performance using union bounds. The analysis as-
sumes the use of a maximal-likelihood receiver for demodulation while the signaling pulses are allowed to
extend over several PPM slots. It has been shown that, using a simple convolutional encoder at the trans-
mitter and a Viterbi algorithm at the receiver, T-PPM restores the performance losses due to reduced
intensity during the detection process.
Furthermore, using the average number of photons per information bit as a performance measure,
T-PPM requires less energy than its regular PPM counterpart by affording a smaller PPM slot width.
Numerical examples show that, for a symbol error of 103when the received pulses extend over 4 PPM
slots, the average laser energy per symbol for 255-ary T-PPM could be reduced by as much as 2 dB. In
addition, the increase in the transmitter efficiency could be more profound if the pulse-width duration
became narrower.
References
[1] P. P Webb, “Properties of Avalanche Photodiodes,” RCA Rev., vol. 35, pp. 234–
278, June 1974.
[2] R. M. Gagliardi and S. Karp, Optical Communications, New York: John Wiley
and Sons, Inc., 1976.
[3] F. M. Davidson and X. Sun, “Gaussian Approximation Versus Nearly Exact
Performance Analysis of Optical Communication Systems With PPM Signaling
and APD Receivers,” IEEE Transactions on Communication, vol. 36, no. 11,
pp. 1185–1192, November 1988.
15
[4] S. D. Personick, “Statistics of a General Class of Avalanche Detectors With Ap-
plications to Optical Communication Systems,” Bell System Technical Journal,
vol. 50, no. 10, pp. 3075–3095, December 1971.
[5] J. T. K. Tang and K. B. Letaief, “The Use of WMC Distribution for Performance
Evaluation of APD Optical Communication Systems,” IEEE Transactions on
Communications, vol. 46, no. 2, pp. 279–285, February 1998.
[6] V. Vilnrotter, M. Simon, and M. Srinivasan, “Maximum Likelihood Detection of
PPM Signals Governed by an Arbitrary Point Process Plus Additive Gaussian
Noise,” JPL Publication 98-7, Jet Propulsion Laboratory, Pasadena, California,
April 1998.
[7] R. J. McIntyre, “The Distribution of Gains in Uniformly Multiplying avalanche
Photodiodes: Theory,” IEEE Transactions on Electron. Devices, vol. ED-19,
pp. 703–713, June 1972.
[8] J. Conradi, “The Distribution of Gains in Uniformly Multiplying Avalanche Pho-
todiodes: Experimental,” IEEE Transactions on Electron. Devices, vol. ED-19,
pp. 714–718, June 1972.
[9] C. N. Georghiades, “Some Implications of TCM for Optical Direct-Detection
Channels,” IEEE Transactions on Communications, vol. 37, no. 5, pp. 481–487,
May 1989.
[10] M. Srinivasan and V. Vilnrotter, “Symbol-Error Probabilities for Pulse-Position
Modulation Signaling With an Avalanche Photodiode Receiver and Gaussian
Thermal Noise,” The Telecommunications and Mission Operations Progress Re-
port 42-134, April–June 1998, Jet Propulsion Laboratory, Pasadena, California,
pp. 1–11, August 15, 1998.
http://tmo.jpl.nasa.gov/tmo/progress report/42-134/134E.pdf
16
... By partitioning the PPM symbol set into subsets of larger minimum distance, and by using convolutional encoding to select the signal subset from which to transmit the next PPM symbol, the symbol-error probability can be reduced substantially. In this article, the trellis-coded PPM (T-PPM) scheme proposed in [2] is reviewed, the decoding algorithm for T-PPM is described, and simulation results are presented for a rate 7/8 T-PPM system with Viterbi decoding using realistic modeling of the APD-based optical receiver. ...
... Following the treatment in [2], the optical system under consideration is described as follows. The intensity of the received optical signal may be written as ...
... In order to combat the performance degradation caused by nonideal pulse shapes, the method of trelliscoded modulation was proposed in [2] for use with PPM signaling. The PPM symbol set is partitioned into subsets that consist of signals with a certain minimum distance, where distance is defined as the number of time slots between PPM symbol positions. ...
Article
This article describes a receiver structure for trellis-coded pulse-position modu- lation (T-PPM) for optical communication systems. Monte Carlo simulation results of the Viterbi decoding algorithm under realistic Webb modeling of the avalanche photodiode (APD)-based receiver are compared with asymptotic bounds on perfor- mance. These results show that T-PPM can reduce the loss in performance incurred by spreading of the PPM optical pulse. Speciflcally, when a PPM symbol-error rate of 0.001 is required, the average signal photon count per information bit is reduced by 5 photons per information bit, or 0.6 to 2.0 dB depending upon the background photon level, when T-PPM is used with the receiver parameters selected here. It is also shown that preserving only a small subset of slot statistics during the decoding process does not result in any signiflcant loss in overall performance. In the receivers considered in this article, an avalanche photodiode (APD) is used to detect the downlink optical signal. The APD generates a random number of electrons in response to each photon absorbed from the received signal. The number of electrons that the APD outputs over a specifled time interval is governed by the McIntyre{Conradi distribution, which is closely approximated by the Webb distribution. In addition, Gaussian thermal noise induced by receiver electronics contributes to the current response of the detector. The APD output is integrated over each slot duration, resulting inM slot statistics per PPM symbol. The density of these slot statistics is modeled as a mixture density consisting of the convolution of Webb and Gaussian density functions. Assuming that the slot statistics are independent random
... The resulting numbers may be compared with those of an APDbased, direct-detection receiver for PPM. In a recent study ( [5], Figs. 9 and 10), an upper bound on the number of received photons required to achieve a given performance was computed as a function of the background radiation level. In Table 1, the required number of signal photons is given for various levels of background radiation for a germanium-type detector when an error rate of 10 −2 is of interest. ...
Article
Full-text available
This article investigates the performance of the heterodyne detection scheme for optical pulse-position modulation (PPM) communications channels when nonideal laser transmitters are employed. Two aspects of this channel will be considered. First, the limitations of unguided, turbulent channels in accommodating coherent laser communications are addressed. This is followed by an investigation of the performance of a heterodyne optical PPM communication system. It is demonstrated here that, due to the presence of laser phase noise, noncoherent (or asynchronous) RF detection is a preferred mode of detection and that the overall performance is severely hampered in the presence of phase noise. The performance of the proposed receiver is compared with that of the avalanche photodiode detector (APD)-based direct-detection receiver. It is shown here that the APD-based detector outperforms its heterodyne counterpart across a wide range of background radiation levels.
Chapter
IntroductionFundamentals of PPMDetection of PPM SignalsPerformance of PPMSynchronizationCapacity of PPM on an Optical ChannelRelated ModulationsConclusion GlossaryCross ReferencesReferences
Article
The capacity is determined for an optical channel employing Pulse Position Modulation (PPM) and an Avalanche PhotoDiode (APD) detector. This channel is dierent from the usual optical channel in that the detector output is characterized by a Webb-plus-Gaussian distribution, not a Poisson distribution. The capacity is expressed as a function of the PPM order, slot width, laser dead time, average number of incident signal and background photons received, and APD parameters. Based on a system using a laser and detector proposed for X2000 second delivery, numerical results provide upper bounds on the data rate and level of background noise that the channel can support while operating at a given BER. For the particular case studied, the capacity-maximizing PPM order is near 2048 for nighttime reception and 16 for daytime reception. Reed-Solomon codes can handle background levels 2.3 to 7.6 dB below the ultimate level that can be handled by codes operating at the Shannon limit. Keywords: Ca...
Article
The basic properties of avalanche photodiodes are reviewed and the conditions under which they can be used to best advantage are discussed. Expressions are presented for calculating the avalanche gain, signal-to-noise ratio, detector excess noise factor, effective noise factor under conditions of mixed injection, diode linearity and saturation, noise equivalent power, quantum efficiency, impulse time response, and frequency response. Several commercially available avalanche photodiode structures are compared, and their virtues and disadvantages are discussed. The properties of a reach-through avalanche photodiode are discussed in detail, and experimental results are presented for single-element avalanche diodes with sensitive areas up to 20 mm**2, and quadrant avalanche diodes with 5 mm**2 sensitive areas. Finally, an equivalent circuit for the reach-through avalanche diode is derived.
Article
Previous results on the statistics of avalanche detectors are generalized to the case where electrons and holes suffer collision ionizations with unequal probability. It is assumed here that the ratio of collision ionization probabilities per unit length of weaker-to-stronger carrier is a constant k independent of position in the high-field region. The moment-generating function of the random avalanche gain G is obtained as a function of k and the average gain Ḡ, and is used to obtain Chernov bounds on error rates of digital optical receivers employing avalanche detectors. It is shown the required energy per pulse to achieve a given error rate decreases as k decreases for fixed Ḡ. For each k > 0, there is an optimal mean gain Ḡopt resulting in minimum required energy per pulse. At k = 0.1, Ḡopt ≅ 100 and the required energy is within 10 dB of that required with very high gains (a few thousand) at k = 0.
Article
The calculation of symbol-error probabilities for uncoded pulse-position mod- ulation using a Webb-modeled avalanche photodiode receiver in the presence of additive Gaussian thermal noise is discussed. Performance curves for 256-ary pulse- position modulation obtained through Monte Carlo simulation as well as a numerical upper bound are presented. The Webb-plus-Gaussian model for the distribution of the received signal is compared with the much simpler Gaussian-only approxima- tion. Results are presented for both flxed and optimized values of the avalanche photodiode gain parameter.
Article
The maximum likelihood decision statistic for pulse-position modulated (PPM) signals governed by an arbitrary discrete point process in the presence of additive Gaussian noise is derived. Sufficient conditions are given for determining when the optimum PPM symbol detection strategy is to choose the PPM symbol corresponding to the maximum slot statistic. In particular, it is shown that for the important case of Webb distributed avalanche photodiode output electrons in the presence of Gaussian noise, the optimum decision rule is to choose the largest slot observable
Article
In the analysis of the bit-error rate (BER) of an optical communication system with avalanche photodiode (APD) detectors, the Webb, McIntyre, Conradi (1974) distribution has been often used to approximate the APD receiver output statistics. This paper presents new simple and closed form BER expressions as well as an importance sampling technique for optical systems employing APD receivers when the WMC model is used. Comparison of the WMC model with the exact model is also carried out. In particular, it is shown that the WMC distribution is quite accurate and has the advantage of simplicity for both simulation and analysis
Article
Consideration is given to the optical direct-detection channel, and it is shown how simple trellis-coded modulation (TCM) can be used to improve performance or increase throughput (in bits per second) without bandwidth expansion or performance loss. In fact, a modest performance gain can be achieved. Although the approach can be used with other signal constellations, the authors concentrate on signals derived from the pulse-position modulation (PPM) format by allowing overlap. Theoretical motivation for using this signal set, known as overlapping PPM (OPPM), was recently given by I. Bar David and G. Kaplan (1984), who showed a capacity gain when overlap is introduced
Article
A 25 Mbit/s direct-detection optical communication system that used Q =4 PPM signaling was constructed and its performance measured under laboratory conditions. The system used a single mode AlGaAs laser diode (λ=834 nm) and low-noise silicon avalanche photodiode (APD). A procedure is given to numerically compute system performance which uses the nearly exact Webb's approximation of the true Conradi distribution for the APD output that does not require excessive amounts of computer time (a few CPU minutes on VAX 8600 per system operating point). Comparison revealed that modeling the APD output as a Gaussian process under conditions of negligible background radiation and low (less than 10<sup>-12</sup>A) APD bulk leakage currents leads to substantial underestimates of optimal APD gain and overestimates of system bit error probability. Examples are given which illustrate the breakdown of the Gaussian approximation in assessing system performance. The measured performance of the system was found to be in excellent agreement with the performance predicted by the nearly exact computational procedure. This system achieved a bit error probability of 10<sup>-6</sup> at a received signal energy corresponding to an average of 60 absorbed photons/bit and optimal APD gain of 700
Article
Expressions are derived for the probability P_{n,m} that a pulse initiated by n electrons (or holes) in a uniformly multiplying semiconductor diode will result in a total number of electrons (or holes) m , to give a gain m/n , and for the probability Q_{n,m} that the gain will be m/n or greater. It is shown that the distributions are far from Gaussian. The gain distribution P_{1,m} for a single photoelectron, for example, is shown to have a maximum value for m = 1 for any value of the average gain M=m/n . The derivations are valid for any electric field distribution and assume only that the hole ionization coefficient beta(E ) can be approximated by the relation beta(E) =kalpha(E) , where alpha(E) is the electron ionization coefficient and k is a constant. A method of determining an effective value of k , for cases where beta=kalpha is not a good approximation, is presented. The results can be used to calculate the average gain and the mean square deviation from the average, giving results in agreement with previously published relations [1], [2]. The implications of this theory on the use of avalanche diodes for low-level photodetection are discussed. It is shown that in the near infrared, cooled avalanche photodiodes can compare favorably with the best available photomultiplier when used either in a photon-counting mode, or for the reliable detection of low-level laser pulses.
Article
Experimental measurements of the gain distribution and noise spectral density of silicon avalanche photodiodes are presented and compared with McIntyre's theories [7], [8]. Excellent agreement is obtained using k eff , the effective ratio of the hole and electron ionization coefficients, as the only adjustable parameter.
  • R M Gagliardi
  • S Karp
R. M. Gagliardi and S. Karp, Optical Communications, New York: John Wiley and Sons, Inc., 1976.