Capacity Limits of Optical Fiber Networks
Capacity Limits of Optical Fiber Networks
Capacity Limits of Optical Fiber Networks
AND
I. INTRODUCTION
ETERMINING an ultimate limit to the rate at which one
can reliably transmit information over a physical medium
in a given environment is an endeavor having both fundamental
(1)
Another representation of the real signal
follows because
its spectrum satisfies
where is the complex
conjugate of . In other words, the negative frequency components of
are redundant for real signals. Thus, we may represent
by using
at positive frequencies only, as shown
in Fig. 1(b). This spectrum is referred to as the single sideband
(SSB) version of
.
Suppose next that we are interested in a passband signal
[22][25]
(2)
where
bandlimited to
663
Fig. 1. Spectral amplitudes of baseband and passband signals. (a) Real baseband signal spectrum. (b) Complex signal spectrum (single-sideband spectrum,
real and imaginary parts). (c) Complex baseband signal spectrum (real and
imaginary parts). (d) Real passband signal spectrum (cosine and sine parts).
Fig. 2. Two adjacent sinc pulses. The value of a sinc pulse at regular sampling
instants is zero except at one sampling instant for each pulse. This property
makes sinc pulses free from intersymbol interference (ISI) between symbols at
these sampling instants.
of the
and
functions. We may represent (2) by
defining a complex baseband signal
(3)
and writing
(4)
where
is the real part of and
. The real signals
and
are sometimes referred to as in-phase (I)
and quadrature (Q) components. The signal
is sometimes
.
referred to as the complex envelope of
The complex signal
has spectral support
as shown in Fig. 1(c), while the signal
has spectral
support
as shown in
Fig. 1(d). We again say that
has bandwidth
Hz. We can
sample
and
at their Nyquist rate of
samples per
second and these samples
and
may be represented
as complex samples
of the signal
.
The signal
is reconstructed from the complex samples as
follows [19, see Fig. 2], [13, Ch. 2], [26, Ch. 6], and [27, Ch. 8]
(5)
The reader might wonder why (1) and (5) have the sampling
and , respectively. In fact, the sampling in (5) also
rates
represents a sampling rate of
real samples per second since
the
in (5) are complex numbers. Mathematically, the passband signal of (2) is projected onto the signals
and
Fig. 3. Symbol rate R , spectral support W , and wavelength-division multiplexing (WDM) channel bandwidth B for a transmit pulse with a square-root
raised-cosine spectrum.
now that
is permitted. In both representations (1) and
(5), a sampling rate of
real samples per signal is necessary
and sufficient.
One sometimes encounters an alternative way of viewing
passband signals, related to SSB modulation. For instance,
we can generate a passband signal with spectrum shown in
Fig. 1(d) corresponding to a real-valued baseband signal
with spectrum shown in Fig. 1(a) by using the following steps.
1) Strip off one sideband of
by using a Hilbert filter [26,
p. 200], [27, Ch. 7], Ch. 7) to generate a complex signal
with spectral support shown in Fig. 1(b). Now frequency
up-convert the signal by
Hz and transmit the
real part of the resulting signal.
2) Modulate
with the carrier
and
then eliminate one sideband by using a bandpass filter.
We observe that SSB modulation, being effectively an alternative way of generating passband signals, does not gain capacity
over dual sideband modulation with complex baseband signals,
as described above.
We remark that we distinguish between the signal spectral
support , the frequency bandwidth assigned to the signal
within the optical network, and the symbol rate
at which one
is modulating the transmit pulse. The symbol rate is
,
where is the symbol period. For example, consider Fig. 3 and
suppose we use passband communication, a transmit pulse with
a square-root raised-cosine spectrum [25], [23] and a roll-off
factor
, and that is 20% larger than . In this case,
we have
and
. In general,
we consider
.
664
Fig. 4. Schematic representation of the main functions of a digital communication channel. The lighter boxes represent the source compression and decompression
functions while the darker boxes represent the communication channel and the related coding and modulation functions. The thicker arrows represent analog
waveforms.
B. Transmitter: Digital-to-Analog
The discrete-time signal representations described earlier are
eminently practical: to generate a passband signal
that is
bandlimited to
Hz one may generate
complex symbols
per second, multiply each symbol by a pulse that is bandlimited
to
Hz, and transmit these pulses in sequence. In other words,
signal generation may be separated into two distinct parts: modulation defined by a discrete set of values called the modulation alphabet or constellation, and pulse shaping to create
the pulse waveforms [3]. The size of the constellation determines the maximum information that each symbol can carry,
while pulse shaping affects the spectral width occupied by the
signal. The pulse shaping can be synthesized digitally using a
digital-to-analog converter (DAC) [28], [29]. Constellations are
treated in Section IV, so we consider pulse shaping next.
The pulse shape may be chosen so that there is no ISI between
successive symbols. For instance, the sinc pulse in Fig. 2 is zero
for all sampling instants
except for
, where
is the symbol period. However, the sinc pulse amplitude decays
slowly 1
so that there is significant ISI if the sampling is
imperfect. Another commonly used pulse shape has the raisedcosine spectrum [26, Ch. 6]
(6)
where
(6) is
(7)
Clearly, we have
for
so there is no ISI.
Moreover, the pulse decays much more quickly 1
than
the sinc pulse if
, thereby significantly reducing ISI with
imperfect sampling. The price paid is that the spectrum has a
larger bandwidth
than the sinc pulse. The choice
recovers the sinc pulse.
A third commonly used pulse has the square-root raisedcosine spectrum
with
given by (6). The corresponding time-domain pulse is not zero at the sampling times
and, therefore, exhibits ISI. The reason for using this pulse
shape is that by placing a square-root raised-cosine spectrum
filter at both the transmitter and receiver, the overall pulse shape
has a raised-cosine spectrum [15, Ch. 11]. Furthermore, the receiver filter now acts as a matched filter for the transmit pulse.
We use the matched filter receiver because it maximizes the
signal-to-noise ratio (SNR) (see Section III-C) for channels with
additive noise. However, note that the nonlinear fiber channel
studied in Part 2 of this paper may have a different optimum receiver structure, which is a topic of further investigations. We
refer to [26, Ch. 6], [14, Ch. 5], and [15, Ch. 11] for more discussion on pulse shapes and matched filters.
Finally, we remark that the signal spectrum can be narrowed below the minimum bandwidth associated to the
by using correlative methods such as partial-resymbol rate
sponse signaling [30], [31] or continuous phase modulation
(CPM) [32], [33]. Correlative methods introduce memory. For
instance, the duobinary signaling in [34, Fig. 6] modulates a
pulse twice as fast as usual, perhaps resulting in
,
and thereby introducing memory into successive Nyquist-rate
symbols while doubling the data rate in a given spectral band
[34]. However, as was pointed out in [35], such modulations
are better viewed as an encoding operation followed by a
memoryless modulator. In other words, correlative methods
such as duobinary signaling or CPM are better viewed as coded
versions of the usual signaling.
For example, the duobinary signaling in [23] and [24] with a
pulse shape
is equivalent to a unit-memory, rate 1, digital
encoder followed by a memoryless modulator with a constellation size of 3 and a pulse shape
.1 A closely related CPM
method is known as minimum-shift keying (MSK) [36], and it
can also be represented as a unit-memory, rate 1, digital encoder
followed by a memoryless modulator and a pulse-shaper [37],
[38]. We can, therefore, focus our attention on finding the capacity of signaling with memoryless modulators and the usual
pulse shaping.
C. Receiver: Analog-to-Digital
At a receiver we may capture all the information in a noisy
bandlimited signal
by filtering the signal to reject noise and
interference outside the band of interest, and subsequently sampling the signal at its Nyquist rate. As an additional practical
step, one usually quantizes the amplitude of the signal samples
1See Section III and Fig. 4 for the meaning of unit-memory, rate 1, and
digital encoder. The encoder for duobinary signaling shapes the spectrum and
this type of coding is called line coding.
at detection to a discrete and finite set of values that are represented by sampling bits [23]. This is done by an analog-to-digital converter (ADC) [39], [29].
and suppose that
Consider the transmitted waveform
each symbol
in (5) takes on one of
complex values. The
combination of an ADC and demodulator that puts out more
values per sample is called a soft-decision detector and
than
it leads to two scenarios for the digital demodulator and subsequent decoder shown in Fig. 4 (see [26, Ch. 8], [14, Ch. 8],
and [27, Ch. 29]). In the first scenario, called hard-decision decoding, the demodulator decides which modulation symbol was
transmitted and passes its decision to the decoder (see Fig. 4);
the decoder operates on these hard decisions. In the second scenario, called soft-decision decoding, some or all of the sampling
bits are passed to the decoder and the decoder uses this soft information to decode [40]. In other words, the digital demodulator
in Fig. 4 is effectively removed. Obviously, using soft-decision
decoding with many quantization levels is preferable for performance, while using hard-decision decoding with few quantization levels reduces complexity. A summary of the performance
and challenges of high-speed ADCs can be found in [41][43].
As we shall see shortly, using discrete-time and discrete-alphabet signals makes sense at both the transmitter and the
receiver because noise limits our ability to extract information.
The process of converting a continuous-time and amplitude
signal
to a discrete-time and discrete-amplitude-and-phase
signal
is referred to as digitization. One of the many key
insights provided by Shannons information theory is that it
suffices to consider digitized signals to approach the ultimate
capacity limits of noisy channels [1]. Shannons work is generally recognized as having given birth to digital communications
and laying the foundation of todays computer and information
age.
665
666
(12)
(10)
where
is the probability that
was transmitted
given that we observe
at the receiver. Using Bayes rule
[3], [45], [46], we have
where
is the joint probability that
. The
average conditional entropy is written as
(16)
where
and
.
Suppose that
and have a joint density
. The
entropies and mutual information are now defined as [3, Ch. 9]5
(17)
(18)
(11)
(19)
Using (8) and (11), we define a quantity that measures the
and , in the
information between two random variables
sense that it measures how much knowing reduces
to
X
hX
HX
HX
667
(20)
where the last step follows by inserting (15) into (17) with
replacing .
Suppose we have the input constraint6
(we soon
interpret as a power). The capacity in bits per symbol is now
[1], [2, Ch. 7], and [3, Ch. 9]
bits per
Observe from (24) that
second. In other words, capacity increases linearly with if we
can use all frequency bands. Moreover, the spectral efficiency
SE in bits per second per Hz is [26, Ch. 7]
(26)
Suppose next that we use a pulse with energy but at the
that is less than
which is less than the WDM
symbol rate
channel bandwidth (see Fig. 3). For the same energy , the
signal power is reduced by a factor of
as compared to
the previous case with sinc pulses at symbol rate . Suppose
we filter the received signal by using a unit-energy matched filter
before sampling. The noise energy is again
by using Parsevals theorem [26, p. 115] and the noise power is
. The
capacity of (24) in bits/second, thus, reduces to
(27)
The spectral efficiency of (26) in bits/second/Hz reduces to
(28)
(21)
It is well known from Shannons work that the optimum has a
density that is a bidimensional Gaussian distribution of the form
[1], [2, Ch. 7], and [3, Ch. 9]
(22)
The resulting capacity in bits per modulation symbol is
(23)
and
is referred to as an SNR.
Suppose, we use a sinc pulse (see Fig. 2) with symbol period
, bandwidth
, and energy
Joules, respectively, for signaling and the same sinc pulse but with unit energy
as a receiver (matched) filter. Suppose the noise is a Gaussian
random process with a (two-sided) power spectral density of
Watts/Hz across all (positive and negative) frequencies of
interest [26, Sec. 7.7], [27, Sec. 25.15]. The noise power after the
receiver sinc filter is, therefore,
Watts per sample
and the noise samples are independent. Using (23), the capacity
in bits per second is [26, Ch. 7], [13, Ch. 5]
(24)
nP
(29)
Equation (28) can thus be expressed as
(30)
Second, the SNR in (29) is based on the energy or power per
modulation symbol. For a fair comparison among modulation
formats, it is convenient to consider the SNR per information
bit which we denote by SNR . Recall that the number of information bits per modulation symbol is
so we
define
(31)
We remark that
is often referred to as
since the energy per information bit is
[26, Ch. 7]. Note that both
SNR and SNR are defined here in a single mode (i.e., single
668
and
for a
are
669
. We provide the details of the calculation of the mutual information for ring constellations in Section IV-B and in
Appendix A.
There are several reasons for choosing ring constellations
[48], [49]. First, such constellations approximate the Gaussian
distribution as the number of rings increases if we choose the
appropriately. A simple choice is to have
ring amplitudes
the rings equally spaced in optical field amplitude and with
equal frequencies of occupation (equal probability for choosing
a transmitted symbol from any of the rings). The ring amplitudes and occupation frequencies could both still be optimized
to better approximate (22) but, as we shall see in Figs. 16 and 17,
respectively, this pragmatic constellation choice already gives
very close to the AWGN channel capacity. We treat
our choice of ring constellations in more detail in Section X-C.
(43)
(44)
where the second step follows because the
are independent
and by using the chain rule for expanding entropy [3, Sec. 2.5],
and the third step follows because conditioning cannot increase
entropy [3, p. 29]. We now use the fact that
is convex in
if the distribution of is held fixed [3, p. 33] . Observe
, are identically distributed with,
that, the
say, the distribution of . Applying Jensens inequality [3, Sec.
2.6] to (44), we thus have
(45)
where
is the output of a channel with input
the channel density is
Inserting (45) into (41), we find that
and where
.
(46)
(42)
670
Fig. 8. Examples of constellations using only one quadrature of the field (here
the real part). The number of bits/symbol is given by log ( ) where
is the
total number of symbols. The number log ( ) of symbols is used as the first
digit of the format label.
Fig. 10. BER as a function of SNR for the modulation formats of Figs. 8 and 9.
that utilize both field quadratures. These constellations generally carry a different number of bits per symbol, depending on
the number of symbols . A constellation can carry a maximum of
information bits per symbol. This maximum
is achieved when all the points in a constellation are used at the
same frequency and in the absence of coding, which means that
. One can design transmitters that use some constellation points more often than others, in which case different frequencies of occupation are associated with different constellation points. The conveyed information is then less than
bits per symbol for such transmitters.
The average power associated with a constellation is given
by
, i.e., the average of the square of all symbol amplitudes. This means that constellations with larger
must have
their points more closely spaced together. One, therefore, expects different constellations to have different BER versus SNR
curves for the AWGN channel. BER curves for the constellations of Figs. 8 and 9 are shown in Fig. 10. These BER curves
apply in the absence of coding and for an identical frequency of
occupation for each constellation point. Moreover, for
,
the bits are mapped into the constellation using Gray mapping
(after Frank Gray, who used the term reflected binary code [52])
that minimizes the number of bit errors for a given symbol error
Fig. 11. BER as a function of SNR per bit for the modulation formats of Figs. 8
and 9.
671
Fig. 12. Capacity as a function of SNR for the modulation formats of Figs. 8
and 9.
Fig. 13. Capacity as a function of SNR per bit of information for the modulation formats of Figs. 8 and 9, respectively.
672
Fig. 14. Bidimensional Gaussian constellation (left plot) optimum for the
AWGN channel and a four ring constellation approximation (right plot).
Fig. 15. Ring constellations with various numbers of rings r from 1 to 16.
In ring constellations, only discrete values of amplitude are allowed while the
phase can assume an arbitrary value (continuous phase). The amplitude of the
outer rings is here equal to a multiple of the amplitude of the inner ring for
constellations larger than one ring.
capacity in terms of
(in bits/s/Hz) as opposed to
(in
bits/symbol), these formats would have the exact same limiting
capacity as their complex equivalents, e.g., 2-ASK/2-PSK
would have the exact same limiting capacity of 4 bits/s/Hz as
16-QAM.
We have considered in this section a number of commonly
used constellations and capacity calculations assuming soft-decision decoding. In Appendix C, we consider a more extensive set of constellations, including constellations of larger sizes
and different shapes. Appendix C also includes the impact of
hard-decision decoding on capacity.
B. Ring Constellations
The constellations considered in Figs. 8 and 9, respectively,
are used in practice because of their discrete nature that facilitates their generation. However, their capacity is limited to
bits/symbol. Such capacity limitations are alleviated by
continuous constellations such as the bidimensional Gaussian
constellation that leads to the Shannon capacity formula of (23).
One way to understand this is that a continuous constellation
allows increasing arbitrarily the effective number of constellation points in both quadratures. It is interesting to note that a
format remains unbounded in capacity as SNR increases even
if only one of the two dimensions of the constellation is continuous while the other is discretized. One way to produce a constellation not bounded in capacity is to discretize the bidimensional Gaussian constellation in amplitude to create concentric
rings. Such a discretization allows to take advantage of continuous rotational symmetries of certain channels, i.e., the fact that
constellation points rotated by an arbitrary value of phase are
equivalent for these channels.
A schematic representation of the discretization in concentric
rings is shown in Fig. 14. A constellation with a single ring is
referred to as phase-shift keying (PSK) while constellations with
two or more rings can be referred to as r-ASK/PSK, where r is
the number of rings. Ring constellations having 1, 2, 4, 8, and
16 rings are shown in Fig. 15 along with their names and labels.
The capacities of the ring constellations as a function of SNR
are shown in Fig. 16. We consider equidistant rings here, i.e.,
constellations where the radii of the outer rings are given by
an integer multiple of the radius of the inner ring. The number
of points on each ring is assumed to be large enough so as to
be considered continuously distributed in phase, with identical
frequency of occupation on each ring. The details of the ring
capacity calculations are given in Appendix A.
Fig. 16. Capacity as a function of SNR for the ring constellations of Fig. 15.
The capacities of constellations with more than 16 rings are also presented.
673
Fig. 17. Same as Fig. 16 but with capacity as a function of SNR per information bit. All constellations converge to an SNR per information bit of
10 log (ln(2))
1:59 dB at low capacity.
0
Fig. 18. Example of signal field using sinc pulses and a ring constellation. (a)
Spectrum, (b) Constellation, and (c) Waveform. A small number of symbols
(32) and rings (2) are represented for clarity.
AND
VI. INTRODUCTION
The vast majority of worldwide data and voice traffic is transported using optical fibers, interconnected to form global fiberoptic networks. As the demand for bandwidth continues to increase exponentially at about 60% per year [54], it is of great
interest to study the transmission capacity between two locations in such optical networks. The aim of Part 2 of this paper
is to provide the most accurate capacity estimate possible for a
fiber channel defined in the context of transporting information in optically-routed networks (ORNs).
Since its foundation [1], information theory has been applied
to several communication channels with great success. The
capacity analysis presented here is also based on information
theory, with specific adaptations to the optical fiber channel.
The most important difference between the optical fiber and
other transmission media that have been considered for capacity analyses is the presence of Kerr-nonlinearity, i.e., the
propagation properties of the medium change with increasing
signal power. As we shall see, this property has important
consequences. While linear physical media perturbed by additive noise generally result in channel capacities that increase
monotonically with transmit power owing to an increasing
SNR, we may find that the negative impact of nonlinear signal
distortions grows at a faster rate than the SNR capacity gain
at high signal powers and for a band-limited channel. This
behavior may turn the channel capacity into a nonmonotonic
function of the transmit power, and the channel capacity will
exhibit a pronounced maximum at a given (finite) signal power
level or SNR. Our fiber channel capacity estimate exhibits such
behavior and therefore differs fundamentally from linear channels whose capacities have been extensively studied [5][18].
Applying information theory to the fiber channel faces several major challenges. An important difficulty originates from
the presence of three phenomena in the fiber channel: noise,
filtering, and Kerr nonlinearity, as visualized in Fig. 19. These
three phenomena are distinct in nature, occur simultaneously,
are distributed along the propagation path, and influence each
other. Note that fiber chromatic dispersion is a form of all-pass
filter and can introduce substantial memory into the channel.
674
Fig. 19. List of the physical phenomena present in the optical path classified
in three groups: 1) Fiber nonlinearities, 2) filtering, and 3) noise. All the phenomena mentioned in the figure are discussed in this paper. The most important
phenomena limiting fiber capacity are in bold.
Fig. 20. (Top) Historical evolution of record capacity and (bottom) spectral
efficiency of hero experiments in fiber-optic communication systems.
time when forward-error correction (FEC) was not used in fiberoptic communication). With the introduction of FEC in the mid1990s, system experiments started to include a coding bit-rate
overhead of typically 7%, and declared error-free transmission if the measured BER was at least as good as the value
needed at the input of state-of-the-art (hard-decision) FEC devices such that the FEC output BER would be between
in the late 1990s to as low as
today. With advances in
FEC technology [77][79] the required value for the measured
[80] for first-generation
input BER has shifted from
for second-generation
7% ReedSolomon FEC to
FECs [79]. Codes with higher overheads (on the order of 25%)
have been investigated and used mostly in the context of subat
marine systems so far. They require about a BER of
the FEC input [79]. Note that all these codes guarantee the correction at the prescribed input BER only for uncorrelated errors
such as for the AWGN channel. On the nonlinear fiber channel,
however, different noise statistics as well as burst errors may
be encountered, which can tighten the BER requirements of a
code [72], [78], [81][84]. This fact is neglected in virtually all
hero experiments due to the experimental difficulties associated with testing full end-to-end transmission including FEC.
Fig. 20 plots the capacity and the spectral efficiency of
hero experiments since the mid-1980s [85]. The lower curve
in the top plot shows the transmission bit rate that could be
obtained on a single optical wavelength and on a single polarization using electronically time-division multiplexed (ETDM)
transmitters. The experienced growth is about one order of
magnitude over two decades or about 12% per year, which
would have been insufficient to fuel the bandwidth demand
of modern data services that grows at about 60% per year
[54]. The highest ETDM bit rate reported today is 100 Gb/s
Fig. 21. Spectral dependence of the fiber loss coefficient for a typical low-loss
optical fiber (SSMF) and a fiber without the water absorption peak (Allwave).
The origin of the main sources of loss are indicated along with the names of the
amplification bands and their wavelength ranges. (Courtesy of D. Peckham.)
675
and optical filtering technologies. Around the turn of the millennium, the bandwidth of optical amplifiers approached their maximum values allowed by the material, and the capacity growth
began to slow down. Capacity growth became mainly driven by
an increase in spectral efficiency (see bottom plot of Fig. 20)
brought by advanced modulation formats that have quickly been
replacing the prevailing OOK systems in the long-haul transport
arena.
Fig. 22. Spectral layout of the WDM channel superposed to the noisy field
originating from ASE.
Early fiber-optic transmission systems provided point-topoint transmission [91] with all WDM channels co-propagating
over the same optical path. These WDM systems have evolved
to now use reconfigurable optical add-drop multiplexers
(ROADMs) [92] at network nodes to form optically-routed
networks (ORNs) [93][96] such as represented in Fig. 23.
The granularity of WDM channels has an important effect
on the design of ORNs. Because the routing granularity of
ROADMs cannot be smaller than the granularity of WDM
channels unless expensive optical-electronic-optical (OEO)
conversion is performed, it becomes economical to establish a
hierarchy of nodes in a network. Nodes with insufficient traffic
to fill a WDM channel are aggregated in a single, bigger node
called a core node. These core nodes are linked together to
form a core ORN as represented in Fig. 23. ROADMs can then
add and drop channels at the WDM channel granularity.
In this paper, we consider an ORN with a generic mesh
network topology [94], [95] as schematically represented in
Fig. 23. The figure illustrates that independent WDM channels can share some optical fiber spans on their propagation
path from their respective transmitters (Tx) to their respective
receivers (Rx), distorting each others waveforms through
fiber Kerr nonlinearity when sharing the same fiber. Individual
users generally do not have access to other WDM channels
at either the transmitter or the receiver, since this would (in
the most general case) involve the exchange of the full optical
field information between all transponders physically separated
by large distances in the ORN. We further assume that the
676
677
(50)
Although shot noise is identified as a nonstationary noise source,
we see from this equation that in a coherent receiver shot noise
can be made arbitrarily stationary by increasing the LO power,
term dominate and
which eventually lets the
dwarfs the shot noise contributions of signal power and noise
power. In the limit of
, we are left with
(51)
Furthermore, we note that at high LO powers, the probability
distribution of shot noise rapidly converges from its Poissonian
nature towards a Gaussian.
3) Is Shot Noise or Beat Noise More Important to Fiber Capacity: Having specified beat noise and shot noise variances,
we are now in the position to answer the question whether shot
noise or beat noise is the more fundamental noise source in
an ideal homodyne receiver when it comes to evaluating fiber
capacity. We can establish a fundamental connection between
the shot noise and beat noise variances as
(52)
for the responsivity of a detector
where we used
with perfect unity quantum efficiency; is Plancks constant,
and is the optical frequency ( is used to represent optical frequencies in this paper). Note that (52) is independent of both
and
. As we will see in Section IX-B, the noise power spectral density
at the receiver for perfectly (ideal) distributed
optical amplification is given by
[cf. (56)]. Inserting
this expression into (52), we see that shot noise is essentially
negligible compared to beat noise whenever
where
is the fiber loss coefficient and the system length, a condition
that is very well satisfied for any reasonable fiber-optic transport
system length. We may neglect shot noise in our further studies
on fiber-optic transport capacities.9
4) Thermal and Electronics Noise: Finally, we acknowledge that practical receivers are fundamentally associated with
9In certain other applications, e.g., in optical satellite communication links
[18], the optical noise power spectral density at the receiver can be much lower
than in an amplified fiber-optic system, which can make shot noise the dominant
noise contribution in such systems.
678
is fully characterized by its autocorrelation [130]. For ideal distributed Raman amplification, it is given by [124], [131], [132]
(53)
where
is the expectation operator [3, Ch. 2], is the Dirac
functional, and
is the power spectral density of the ASE
after a transmission distance [125], [133], [134] given by (56).
An amplification scheme widely used in fiber-optic communication consists of amplifying the signal periodically at
discrete locations along the optical path. This is done by
inserting optical amplifiers, generally Erbium-doped fiber amplifiers (EDFAs) [125], to interconnect passive fiber spans. This
discrete EDFA amplification scheme is shown in Fig. 24(a).
EDFAs are typically unidirectional as they include optical
isolators a nonreciprocal component that allows propagation in
one direction while blocking propagation in the opposite direction [117], [135]. Fiber span lengths between EDFAs typically
range between 40 and 120 km, depending on the network type.
This corresponds to between 8 and 24 dB of loss per fiber
span before amplification by an EDFA can take place. EDFAs
today closely approach the 3-dB theoretical noise figure limit
dictated by quantum mechanics.
To improve the OSNR beyond the capabilities of EDFAs,
one can transform the passive (lossy) fiber into an amplifying
medium by injecting optical pump power (see Fig. 24(b) and
Section IX-D2). Such optical pumps provide gain through a
stimulated Raman scattering (SRS) process [136], [137] in the
transmission fiber and prevent the signal power from dropping
along the optical path, which results in improved delivered
OSNR [74].
We next calculate the delivered OSNR for both the discrete
EDFA and the distributed Raman systems depicted in Fig. 24.
For periodically spaced discrete EDFAs, the noise spectral density per state of polarization
, generated at the end of a
transmission line composed of a chain of
amplifiers spaced
by fiber spans of length
is given by
(54)
where
(55)
(56)
where
is replaced by
, the phonon occupancy factor. It
, where
is given by
[137] with
the Boltzmann constant, the
fiber temperature and
the optical frequency of the Raman
is approxipump providing the distributed gain. The factor
mately 1.13 for Raman amplification of fiber-optic communication systems at room temperature. Experimental demonstrations
of nearly ideal distributed gain can be found in [141], [142].
For this capacity limit study, we choose the ideal distributed
Raman amplification scheme with Raman gain exactly compensating the fiber intrinsic loss, as it maximizes OSNR (and
the SNR) at fixed nonlinear phase which will be discussed in
Section IX-D2. The delivered OSNR and SNR can be calculated using (33) and (34).
2) Double Rayleigh Scattering: As described in Section X,
Rayleigh scattering can be an important source of fiber loss, but
it can also be an important source of noise [119], [143][145].
A fraction of the Rayleigh scattering of the forward propagating signal is recaptured into the guided mode of the fiber and
propagates in the opposite direction of the signal. A fraction
of that back-propagating light is then Rayleigh scattered and
recaptured into the guided mode of forward propagation, hence
co-propagating as double Rayleigh backscatter (DRB) along
with the signal. This double-scattering process is distributed
over the entire fiber length and creates a continuum of echoes
that act as multipath interference (MPI) on the signal [146].
Since we deny the receiver knowledge of the amplitudes and
phases of these continuum of echoes, MPI is considered as a
fundamental source of noise in this context [146].
The power of the DRB light for a lossy fiber with a power
loss coefficient per unit length due to Rayleigh scattering of
is given by [146]
(57)
where
is the signal input power to a fiber of length and
is the dimensionless backscatter recapture fraction that defines
how much of the scattered light is recaptured into the guided
fiber mode for a particular optical fiber type [137]. The parameters and are the distributed gain and fiber loss coefficients per
unit of length, respectively, both assumed to be constant along
the fiber.
From (57), for ideal distributed Raman amplification for
which
, we obtain
(58)
679
Fig. 25. Concatenation of optical filters. (a) Optical filter with significant amplitude roll-off: spectral narrowing occurs as a result of repeated optical filtering
(ten times). (b) Idealized rectangular optical filter: absence of amplitude narrowing for an arbitrary number n of filters.
680
is defined as
(66)
is the effective
(67)
with represents the fiber length. When gain compensates loss
, we have
.
exactly, i.e.,
The integrated nonlinear phase for an arbitrary signal power
evolution is defined as [158]
(68)
10For 100 mW of signal power propagating in a fiber of 100 m effective
GW/m .
area A [158], the intensity of the field 100 mW/100 m
=1
where
is the signal power evolution. Another
measure that relates to nonlinear transmission is the integrated
nonlinear phase spectral density
defined as
where is the WDM channel spacing. The integrated nonlinear
phase spectral density is a measure of the nonlinear phase that
takes into account the spectral density of the WDM signal.
2) Noninstantaneous Kerr Nonlinearity: The noninstantaneous part of the Kerr effect in optical fibers leads to
Brillouin [91], [167][170] and Raman scatterings [91], [136],
[168][170]. These processes can be spontaneous [171] or
stimulated by the presence of an input wave [145], [172]. Both
phenomena can be interpreted as mechanical waves, of low
frequencies (acoustic phonons) for the Brillouin scattering and
of high frequencies (optical phonons) for Raman scattering.
The Raman effect is often modeled as a delayed nonlinear
response [173][178].
The most important phenomenon associated with stimulated
Brillouin scattering (SBS) in fibers is the presence of optical
amplification in the backward direction downshifted by about
10 GHz in frequency from the signal and of bandwidth less
than 100 MHz [171]. This amplification is generally detrimental
to transmission but can be efficiently suppressed using various
techniques with minimal impact on capacity, using for instance
slow (a few tens of kHz) frequency dithering [179]. It is worth
pointing out that SBS gain has some polarization dependence
[99], [180]. Stimulated Raman scattering (SRS) also leads to
optical amplification, but in contrast to SBS, SRS gain is very
broad ( 10 THz) in fibers. As for SBS, SRS gain is also polarization-dependent [172], [174]. We consider below the two
main amplification mechanisms resulting from SRS.
a) Interchannel Stimulated Raman Scattering: In WDM
transmission, SRS can occur between the different wave-lengths
of the WDM spectrum. An important resulting effect of this interchannel SRS is the creation of a tilt of the WDM spectrum
[91]. Such a gain tilt results in high-frequency WDM channels to be depleted and the low-frequency WDM channels to
be amplified. The WDM spectrum tilt can be calculated using
relations derived in [181] and [182]. Some capacity limitations
are expected from the WDM spectrum tilt but tilt compensation
through gain tilt in the opposite direction and pre-equalization
[183] can greatly reduce its impact.
The gain tilt represents only the average effect of
inter-channel SRS. The gain provided by interchannel SRS
originates from the WDM channels that are data and polarization modulated and experience propagation effects. As a result,
the waveform has power and polarization variations in time
which produces time-varying gain [170], [184][186]. These
effects are not incorporated in our analysis. Further studies
are needed to assess their importance in the context of fiber
capacity.
b) Ideal Distributed Raman Amplification: Stimulated
Raman scattering can be exploited to create fiber Raman
amplifiers by pumping the optical fiber at a frequency about
13 THz above the desired frequency of gain (see [187] and
[188]). An important application of SRS in systems is to generate distributed gain [189] in passive transmission fibers. This
amplification scheme is referred to as distributed Raman ampli-
681
Fig. 26. Delivered OSNR for EDFA and ideal distributed amplification for a
typical system: (a) delivered OSNR at fixed input power, (b) nonlinear phase,
and (c) delivered OSNR at constant nonlinear phase.
Fig. 27. Decomposition of the instantaneous fiber Kerr nonlinearities into two
categories: intrachannel and interchannel nonlinearities. A list of elementary
nonlinear interactions for each category is provided. NLPN: nonlinear phase
noise, SPM: self-phase modulation, NL: nonlinear, MI: modulation instability,
XPM: cross-phase modulation, FWM: four-wave mixing, IXPM: intrachannel
XPM, IFWM: intrachannel FWM.
E. Fiber Propagation
X. FIBER CHANNEL
The equation that describes the evolution of the optical field
(that contains all WDM channels) in a fiber using ideal
distributed Raman amplification (gain continuously compensates fiber loss) with ASE generation can be represented by the
This section describes the choices we made in terms of modulation, constellation and digital signal processing to mitigate
nonlinear distortions for our fiber capacity estimate.
682
Fig. 28. Schematic of the distribution of physical effects for a COI for the fiber
channel considered in this paper. Signal and noise fields are represented for both
in-band and out-of-band frequencies.
A. Distributed Impairments
The physical effects present during the propagation from the
transmitter (Tx) to the receiver (Rx) are represented schematically in Fig. 28 in the context of where they appear along
the propagation path. The signal experiences distributed noise
(ASE), fiber nonlinearity, chromatic dispersion (CD), and
periodic filtering from optical bandpass filters (OFs). Two
out-of-band fields are also shown. They stand as a representation of any other WDM channels added and dropped at random
locations in ORNs. In this study, we assume that all neighboring
WDM channels co-propagate with the WDM COI during the
entire optical path but are not available at either the Tx or Rx.
The in-band fields, signal and noise, propagate all the way
to the COI receiver while out-of-band fields may be dropped
or added along the way. To fully capture and understand the
impact of the distributed nature of the impairments of Fig. 28,
it requires a full solution of (70).
B. Choice of Modulation
Studies of capacity limits for band-limited channels [1] have
been developed for linear channels that conserve the signal spectral support. A nonlinear channel can, in general, create new frequencies falling outside the originally transmitted signal spectral support and eventually the channel bandwidth. In the case of
the ORN fiber channel, the signal spectrum is repeatedly confined by optical bandpass filters at ROADMs. Even though, it
is possible to reconstruct a signal truncated by filtering in some
scenarios in the absence of noise [191], [192], we surmise that,
from a capacity standpoint, it is generally preferable to avoid
spectral broadening altogether in a band-limited fiber channel.
Our approach to deal with these difficulties is to place ourselves in a nonlinear regime that limits spectral broadening by
, where
is the signal
fulfilling the condition
(power dependent) transmission length over which a non-negligible amount of spectral components are generated beyond
the signal spectral support, and is the transmission length.
To make
large, we operate in the regime
,
often referred to as the pseudo-linear regime of transmission
[97], [98]. The dispersion length is
[158] where
is the symbol duration and
the GVD, related to
the dispersion by (61). The nonlinear length [158] is
where is the nonlinear coefficient defined in (64)
and is the signal power. This regime produces a waveform
that changes very rapidly with propagation, helping reduce nonlinear effects. However, the large spreading of the waveform
creates a large number of symbols to interact nonlinearly, creating channel memory [162]. One operates in this regime when
683
684
TABLE I
FIBER PARAMETERS
TABLE II
AMPLIFICATION PARAMETERS
TABLE III
SIGNAL AND OPTICAL FILTER PARAMETERS
TABLE IV
MODULATION AND CONSTELLATION PARAMETERS
polarization effects were included. Finally, the modulation parameters are given in Table IV.
We modeled transmission with a large number of WDM channels and found that increasing the number of WDM channels beyond five only slightly impacts our capacity calculations for the
parameters considered; we, therefore, use five WDM channels
in the following calculations and study the central channel as our
COI. We use constellation points that are randomly chosen on
the ring constellation structures, using time sequences varying
from 2048 and 8192 symbols per simulation. The large computation time prevented using larger numbers of points but repeated trials with different noise, data realizations and time offsets (including time offsets of a fraction of symbol duration
) led to variations in capacity estimates of only a few tenths
of bits/s/Hz. Back-propagation applied at both the transmitter
and receiver in variable ratios also produced capacity estimates
within a few tenths of bits/s/Hz for all these scenarios presented
here.
A. Conventional Dispersion Map
Optical transmission systems that are limited mainly by
single-channel nonlinear transmission generally greatly ben-
Fig. 32. Spectral efficiency after transmission over 2000 km for uniform ring
constellations and a conventional dispersion map.
685
TABLE VI
CONSTELLATION OPTIMIZATION PARAMETERS (SEE [195])
Fig. 33. Spectral efficiency after transmission over 2000 km for ring constellations optimized as described in this paper.
Fig. 34. Spectral efficiency after transmission over 2000 km for various
residual dispersion per span.
benefits constellations with more than two rings, so the maximum capacity is not increased. Note that the calculations shown
in Fig. 33 were performed with different data realizations (i.e.,
WDM waveforms) than in the uniform ring constellation case
of Fig. 32 resulting in slight statistical variations in capacities
for the one-ring case.
C. Effect of Dispersion Map
A measure of the impact of dispersion mapping is presented
in Fig. 34 for 16-ring constellations, a number of rings sufficient for our capacity estimate. The residual dispersion per span
(RDPS) has been varied from full dispersion compensation per
span
to the total absence of any in-line dispersion
compensation [
]. In each case, a
dispersion precompensation equal to half the accumulated link
dispersion is used and the dispersion is brought back to zero before coherent detection. One observes that increasing the value
of RDPS (reducing in-line dispersion compensation) increases
capacity, with the maximum capacity being reached in the absence of dispersion compensation.
Two reasons explain this behavior. The first is that back-propagation eliminates all signalsignal intrachannel nonlinearities
rendering extraneous the function of the dispersion map to reduce intrachannel nonlinearities in such systems. The second
reason is that periodic in-line broadband dispersion compensation recorrelates WDM channels in time producing a coherent
accumulation of nonlinear distortions rather than a statistical averaging when no realignment occurs in the absence of in-line
dispersion compensation. A coherent addition of impairments is
more damaging to capacity than a random addition of the same
686
Fig. 35. Spectral efficiency after transmission for various distance. All links
are without dispersion compensation.
Fig. 36. Spectral efficiency for four signal and noise scenarios for the 2000 km
transmission of Fig. 35. (ch: channel.)
687
Fig. 38. Spectral efficiency results for recent record experiments. The capacity
limit estimate curve for 500 km transmission of Fig. 35 is shown for comparison. There is about a factor three between the capacity limit estimate and the
record capacities. The experimental data, labeled (1)(5) in the legend, are from
[200][204]. The upper axes apply only to the capacity limit estimation curve.
Fig. 37. Four-ring constellations (back-rotated) in the absence of ASE and nonlinearity, and after WDM transmission for various values of input power per
channel P . The optimum power for our capacity estimate is between
6
and
3 dB m.
0
0
tion, is by constituting
different subchannels
by
periodic extractions of symbols in time starting at a difst. So, for sufficiently
ferent symbol, from the 0th to the
large , subchannel one,
, involves symbols transmitted
in time slots numbered
.
consists of symbols transmitted in time slots numbered
.
consists of symbols
. Any one of these sub-channels, for sufficiently large , can by itself be treated as an independent memoryless channel, even though the received symbols suffer impairment through memory effects from symbols
sent in neighboring time slots. The symbols in
can be
decoded and any of the now-known nonlinear influence from
subtracted from the received sigsymbols exclusively in
. Then
nals for sub-channels
the symbols for
are decoded and the now-known nonlinear influence of the transmitted symbols decoded in channels described by
and
together subtracted from
the received symbols for
, and
so on. Finally,
would have the now-known nonlinear influences from all the decoded subchannel symbols involving
removed. Since more
and more nonlinear impairment is removed, the capacities of the
sub channels increase with increasing subscript. By processing
in this way it is clear we have increased the capacity over just
treating each subchannel without subtracting away impairments
due to already decoded subchannels. What we do in this paper
. Pursuing
is pessimistically use times the capacity of
such advanced, but much more involved forms of processing
is beyond the scope of this paper. Indeed, as nonlinear interferences are progressively removed, it suggests that progressively more refined signal constellations be used for channels
involving
with larger indexes .
688
to
The Jacobian of the
so the integral in (71) becomes
transformation is
(75)
where
is the modified Bessel function of the first kind of
order zero. We thus have
XIII. CONCLUSION
A framework for the study of the capacity limits of the fiber
channel in optically routed networks has been described. Using
a series of advanced technologies, including advanced modulation formats, digital signal processing for fiber back-propagation, flat square bandpass optical filters and optimum coding,
we showed that a spectral efficiency per polarization of about
9 bits/s/Hz is achievable over 500 km of standard single-mode
fibers.
(76)
We compute (76) numerically as follows. We generate a large
where is chosen uniformly over
number of
, and then generate by adding complex Gaussian noise.
Finally, we compute
and the average in (76). For large
, we use the approximation [39, p. 47]
entries of
(77)
so that the term inside the expectation in (76) becomes
APPENDIX A
INFORMATION RATES OF RING CONSTELLATIONS ON
AWGN CHANNELS
Consider one ring with
distributed over the interval
pute the entropy
(78)
A. Information Rates for Several Rings
, where is uniformly
. Using (20), we need to com. We have
(71)
where
and we have
where
power
is given by (16),
(80)
(72)
where
(73)
Note that
is concave in
which is linear in
as seen in (79). Thus, once the
are chosen one could optiby using convex optimization methods. The best
mize
will be a function of the SNR.
is to use uniform
A simple approach for choosing the
spacing in the field, i.e., choose
(74)
(82)
689
(92)
Fig. 39. General setup of a coherent optical receiver. Time variables for the
various signals are omitted for visual clarity.
(93)
for
ample, if
, where
.
for a large number of
A better approach for choosing the
rings is to use a spacing that approximates Gaussian signaling
in the complex plane. For example, one such choice is a squareroot logarithmic spacing with
(83)
for
, where
APPENDIX B
SIGNAL AND NOISE IN A COHERENT OPTICAL RECEIVER
The structure of a (single-polarization) balanced coherent opoptical
tical I/Q receiver is shown in Fig. 39. It consists of a
hybrid, which combines the incident optical field with an LO in
both quadratures using two beam splitters as well as two pairs
of balanced photodetectors. Their difference signal constitutes
the output signal of the receiver.
With reference to Fig. 39, the optical fields at the four detectors can be written in the following form [117]
(84)
(85)
(86)
(87)
where the sign of the signal term originates from energy conservation within the lossless beam splitters with power transmission (ideally,
), and the multiplication by is due
to the
phase shift of the LO within the 90-degree optical hybrid. After square-law photodetection with responsivities
(in [A/W]; ideally
for all ), the four electrical signals
read
(88)
) all
(94)
(95)
Finally, these expressions are convolved with the opto-elec, which can, e.g.,
tronic front-ends impulse response
implement a matched filter.
The first term on the right-hand side of (94) and (95) is the
desired signal term, and the second term is the beat term between the LO and the optical noise field, which is the only beat
noise term of relevance. Note that both the beat term between
signal and optical noise as well as the noisenoise beat term are
fully eliminated by ideal balanced detection. The linear conversion of signal and noise optical fields into the electrical regime
also implies that the statistics of the noise are fully preserved.
In particular, a Gaussian optical noise field (ASE) will remain
Gaussian in the electrical domain, in contrast to direct detection
receivers or imbalanced coherent receivers, which generally exhibit non-Gaussian detection noise statistics [105][108].
The variance of the beat noise term between the LO and the
optical noise field can be directly calculated from (94) and (95),
as outlined, e.g., in [119] and [205]. One starts by taking the
expectation of the squared magnitude of the beat-noise term
(96)
assuming that
is a zero-mean stochastic process. (An
equivalent procedure can be performed for the beat noise in
the quadrature component.) Writing out the convolution in
its integral form, expanding the real part into the sum of two
complex conjugated terms, expressing the squared magnitude
as a multiplication with the complex conjugate, assuming the
optical noise field to obey
, which holds,
e.g., for circularly symmetric ccG noise by the moment theorem
of Gaussian random variables, and taking note of the fact that
is temporally constant, we arrive at
(89)
(90)
(91)
(97)
690
Further simplifications can be obtained through certain assumptions on the noise autocorrelation
and the detectors impulse response. For example, if we
assume the noise to be white with power spectral density
over the opto-electronic detection bandwidth, we have
, and consequently
(98)
,
where we made use of
being the power equivalent bandwidth of the real-valued im. This result agrees with the commonly used
pulse response
beat noise variance approximation derived in [206].
APPENDIX C
CAPACITIES AND ERROR RATIOS OF DISCRETE MODULATION
SCHEMES ON AWGN CHANNELS
The capacity-achieving input distribution for the AWGN
is circhannel under an average power constraint
cularly complex Gaussian (see (22) and Fig. 14). Any other
constellation, discrete or continuous, is sub-optimum for the
AWGN channel. As we shall see in Subsection D below, one
can approach the AWGN channel capacity (23) with discrete constellations by using a sufficiently large number
of well-placed points. Any remaining SNR gap between the
capacity (23) and the information rate (37) of constellations is
called the shaping gain [207], [208]. For example, for -QAM
with uniform input probabilities, the shaping gain grows to 1.53
(or large SNR). The gap in SNR
dB asymptotically for large
per information bit between uncoded and coded transmission
, is known as the effective (or
at a certain error ratio, e.g.,
net) coding gain [208].
Discrete constellations are often compared in terms of their
uncoded symbol or bit error ratio (SER/BER) performance and
their coded information rates (12) or (37) with hard- or soft-decision receivers, respectively. Different constellations may be
optimum at different SNR values; generally, a globally optimum constellation cannot be found. The problem of placing
points in the signal space such that the capacity at a given
SNR is maximized (or the BER is minimized) is nontrivial and
attracted much attention in early days of digital communications [209][211]. An overview of the historical development
of two-dimensional constellations can be found in [212].
For a given constellation, error ratios and capacities depend
on how the noisy samples are processed in the receiver. Upon
observing the channel output , the receiver might make a
decision on the transmitted channel input . The probability
of a symbol error is minimized by the maximum a posteriori
(MAP) criterion, i.e., by a receiver that, for an observed channel
which maximizes
output , decides on that input symbol
. For equally probable input symbols, this is equivalent to a maximum likelihood (ML) decision which maximizes
[24]. A receiver that makes such hard decisions
maps each point in the received signal space onto a discrete
symbol; the set of values that are mapped onto a given symbol
form its decision region. If a received value is outside the decision region of the transmitted symbol, a symbol error occurs.
For the AWGN channel, the MAP (or ML) criterion reduces
to deciding on the symbol that is closest to the received value,
Fig. 40. Capacities of various binary and quaternary formats with soft (solid
lines) and hard decision (dashed lines): OOK (1b), BPSK (1a), 2-ASK/2-PSK
(2a), QPSK or 4-QAM (2b), ring-1-3 (2c).
691
Fig. 41. Symbol and bit error ratios (SERsolid lines, BERdashed lines) of
various binary and quaternary formats: OOK (1b), BPSK (1a), 2-ASK/2-PSK
(2a), QPSK or 4-QAM (2b), ring-1-3 (2c). The BER of QPSK (2b) is equal to
that of BPSK (1a).
given that
was sent is
(101)
and
692
= 16 points.
M = 2 = 4, and = 8 points.
;M
Fig. 45. Symbol and bit error ratios (SER: solid lines, BER: dashed lines)
of 16-ary formats: hexagonal (4g), QAM (4b), PSK (4c), ring-1-3-5-7 (4f),
2-ASK/8-PSK (4d), 4-ASK/4-PSK (4e).
Fig. 43. Capacities of 8-ary formats with soft (solid lines) and hard decision
(dashed lines): 4-ASK/2-PSK (3a), rectangular 8-QAM (3b), 8-PSK (3c),
, ring-1-7 (3f), hexagonal (3g), and star
2-ASK/4-PSK (3d) with r =r
8-QAM (3h).
=2
. This compares to
for 8-PSK (3c)
for 2-ASK/4-PSK (3d) with a ring ampliand
tude ratio of
. 2-ASK/4-PSK (3d) subsumes other
, we obtain square
well-known constellations. For
8-QAM with
. Setting
delivers an optimum minimum distance of
.
Star 8-QAM (3h) (which is 2-ASK/4-PSK with equal phase angles in both rings) has
. Rectangular 8-QAM (3b),
which are eight points on a 2-by-4 grid, achieves the same min. More unusual
imum distance as square 8-QAM,
8-ary formats are ring-1-7 (3f)
and 8-hex (3g)
. The latter is the optimum 8-ary constellation
in terms of minimum distance [210]. Notice that the constellation is dc-free; because of its asymmetry, this implies that the
innermost constellation point is not located at the origin.
Fig. 43 shows soft- and hard-decision capacity values for
8-ary modulation schemes. It is noteworthy that despite its
having the largest minimum distance, 8-hex performs slightly
worse than ring-1-7 for a wide SNR range. The SER of 8-hex
is smaller than that of ring-1-7 only for very large SNRs
(
dB); in this range, the capacities of both schemes have
already saturated at 3 bits/symbol.
693
Fig. 48. Capacities of -QAM constellations ( = 8; 16; 32; 64; 128; 256;
512; 1024) with soft (solid lines) and hard decision (dashed lines).
Fig. 46. Capacities of various 16-ary formats with soft (solid lines) and hard
decision (dashed lines): hexagonal (4g), QAM (4b), PSK (4c), ring-1-3-5-7 (4f),
2-ASK/8-PSK (4d).
Fig. 47. QAM constellations with 32, 64, 128, 256, 512, and 1024 points.
C. 16-ary Constellations
The most prominent 16-ary constellations are (sorted by
, 16-QAM
minimum distance): 16-hex (4g)
, 4-ASK/4-PSK (4e)
,
(4b)
, ring-1-3-5-7 (4f)
2-ASK/8-PSK (4d)
, 16-PSK (4c)
. Fig. 44 contains
illustrations of these constellations.
The performance of selected 16-ary modulation schemes is
depicted in Figs. 45 and 46, respectively. As expected, the modulation schemes perform according to their minimum distance
at large SNRs. The hexagonal constellation achieves the highest
capacity; however, the gain over 16-QAM is marginal. Because
of the simpler implementation in practical systems, 16-QAM is
generally preferred over 16-hex. Across the SNR range shown in
Fig. 46, 2-ASK/8-PSK and 4-ASK/4-PSK achieve virtually the
same capacity values. 16-PSK performs worst in terms of error
ratio or capacity at all values of SNR shown. With increasing
, the gain obtained from soft decoding increases.
Fig. 49. SER (solid) and BER (dashed) curves for M -QAM constellations.
694
TABLE VII
LIST OF ACRONYMS
TABLE VIII
LIST OF SYMBOLS AND NOTATION
695
696
TABLE IX
LIST OF SYMBOLS AND NOTATION
REFERENCES
[1] C. E. Shannon, A mathematical theory of communication, Bell Syst.
Tech. J., vol. 27, p. 379423 and 623656, 1948.
[2] R. G. Gallager, Information Theory and Reliable Communication.
New York: Wiley, 1968, ch. 2 and 4.
[3] T. M. Cover and J. A. Thomas, Elements of Information Theory, 2nd
ed. New York: Wiley, 2006.
[4] D. J. C. MacKay, Information Theory, Inference and Learning Algorithms. Cambridge, U.K.: Cambridge Univ. Press, 2003.
[5] J. M. Wozencraft and B. Reiffen, Sequential Decoding. New York:
MIT Press and Wiley, 1961.
[6] J. Massey, Channel models for random-access systems, in NATO Advances Studies Institutes Series E142, 1988, pp. 391402.
[7] I. Kalet and S. Shamai, On the capacity of a twisted-wire pair:
Gaussian model, IEEE Trans. Commun., vol. 38, no. 3, pp. 379383,
Mar. 1990.
[8] J. J. Werner, The HDSL environment, IEEE J. Sel. Areas Commun.,
vol. 9, no. 6, pp. 785800, Aug. 1991.
[9] M. Gagnaire, An overview of broadband access technologies, Proc.
IEEE, vol. 85, no. 12, pp. 19581972, Dec. 1997.
[10] A. Sendonaris, V. V. Veeravalli, and B. Aazhang, Joint signaling
strategies for approaching the capacity of twisted-pair channels, IEEE
Trans. Commun., vol. 46, no. 5, pp. 673685, May 1998.
[11] The Communications Handbook, J. D. Gibson, Ed. Boca Raton, FL:
CRC Press, 1997.
[12] G. J. Foschini, Layered space-time architecture for wireless communication in a fading environment when using multi-element antennas,
Bell Labs Tech. J., vol. 1, pp. 4159, 1996.
[13] D. Tse and P. Viswanath, Fundamentals of Wireless Communication.
Cambridge, U.K.: Cambridge Univ. Press, 2005.
[14] A. Goldsmith, Wireless Communications. Cambridge, U.K.: Cambridge Univ. Press, 2005.
[15] A. F. Molisch, Wireless Communications. New York: Wiley, 2005.
[16] J. P. Gordon, Quantum effects in communications systems, Proc.
IRE, vol. 50, pp. 18981908, 1962.
[17] J. Pierce, Optical channels: Practical limits with photon counting,
IEEE Trans. Commun., vol. COM-26, no. 12, pp. 18191821, Dec.
1978.
[18] D. O. Caplan, Laser communication transmitter and receiver design,
J. Opt. Fiber Commun. Res., vol. 4, no. 4, pp. 225362, 2007.
[19] E. T. Whittaker, On the functions which are represented by the expansions of the interpolation theory, Proc. R. Soc. Edinburgh, vol. 35, pp.
181194, 1915.
[20] H. Nyquist, Certain topics of telegraph transmission theory, Trans.
Amer. Inst. Electr. Eng., vol. 47, pp. 617644, 1928.
[21] A. Jerri, The Shannon sampling theoremIts various extensions
and applications: A tutorial review, Proc. IEEE, vol. 65, no. 11, pp.
15651596, Nov. 1977.
[22] J. M. Wozencraft and I. M. Jacobs, Principles of Communication Engineering. New York: Wiley, 1965.
[23] S. Haykin, Communication Systems, 4th ed. New York: Wiley, 2001.
[24] J. G. Proakis, Digital Communications, 4th ed. New York: McGrawHill, 2001.
[25] B. Sklar, Digital Communications, 2nd ed. Englewood Cliffs, NJ:
Prentice-Hall, 2001.
[26] R. G. Gallager, Principles of Digital Communication. Cambridge,
U.K.: Cambridge Univ. Press, 2008.
[27] A. Lapidoth, A Foundation in Digital Communication. Cambridge,
U.K.: Cambridge Univ. Press, 2009.
[28] J. G. Proakis and D. G. Manolakis, Digital Signal Processing: Principles, Algorithms, and Applications, 4th ed. New York: Macmillan,
2007.
[29] R. G. Lyons, Understanding Digital Signal Processing. Englewood
Cliffs, NJ: Prentice Hall, 2004.
[30] P. Kabal and S. Pasupathy, Partial-response signaling, IEEE Trans.
Commun., vol. COM-23, no. 9, pp. 921934, Sep. 1975.
[31] T. Aulin, N. Rydbeck, and C. E. Sundberg, Continuous phase modulationPart 2: Partial response signaling, IEEE Trans. Commun., vol.
COM-29, no. 3, pp. 210225, Mar. 1981.
[32] T. Aulin and C. E. Sundberg, Continuous phase modulationPart 1:
Full response signaling, IEEE Trans. Commun., vol. COM-29, no. 3,
pp. 196209, Mar. 1981.
[33] J. B. Anderson, T. Aulin, and C. E. Sundberg, Digital Phase Modulation. New York: Plenum, 1986.
[34] A. Lender, Correlative level coding for binary data transmission,
IEEE Spectrum, vol. 3, no. 2, pp. 104115, Feb. 1966.
697
[35] J. L. Massey, The how and why of channel coding, in Proc. Int. Zurich
Semin., 1984, pp. 6773.
[36] S. Pasupathy, Minimum shift keying: A spectrally efficient modulation, IEEE Commun. Mag., vol. 17, no. 4, pp. 1422, Jul. 1979.
[37] J. Massey, A generalized formulation of minimum shift keying modulation, in Proc. IEEE Int. Conf. Commun., 1980, pp. 26.5.126.5.4.
[38] B. E. Rimoldi, A decomposition approach to CPM, IEEE Trans. Inf.
Theory, vol. 34, no. 2, pp. 260270, Mar. 1988.
[39] J. G. Proakis and M. Salehi, Digital Communications, 5th ed. New
York: McGraw-Hill, 2007.
[40] S. Lin and D. J. Costello, Error Control Coding, 2nd ed. Englewood
Cliffs, NJ: Prentice-Hall, 2004.
[41] R. H. Walden, Performance trends for analog to digital converters,
IEEE Commun. Mag., vol. 37, no. 2, pp. 96101, Feb. 1999.
[42] R. H. Walden, Analog-to-digital converter survey and analysis, IEEE
J. Sel. Areas Commun., vol. 17, no. 4, pp. 539550, Apr. 1999.
[43] R. H. Walden, Analog-to-digital converters and associated IC technologies, in Proc. IEEE Compound Semicond. Integr. Circuits Symp.,
2008, pp. 12.
[44] R. M. Gray, Source Coding Theory. New York: Springer-Verlag,
1989.
[45] M. Bayes and M. Price, An essay towards solving a problem in the
doctrine of chances, Philos. Trans. R. Soc. Lond., vol. 53, pp. 370418,
1763.
[46] A. Papoulis and S. U. Pillai, Probability, Random Variables and Stochastic Processes. New York: McGraw-Hill, 2002.
[47] G. Kramer, A. Ashikhmin, A. J. van Wijngaarden, and X. Wei,
Spectral efficiency of coded phase-shift keying for fiber-optic communication, J. Lightw. Technol., vol. 21, no. 10, pp. 24382445,
Oct. 2003.
[48] J. Hancock and R. Lucky, Performance of combined amplitude
and phase-modulated communication systems, Commun. Syst., IRE
Trans., vol. 8, no. 4, pp. 232237, 1960.
[49] J. M. Geist, Capacity and cutoff rate for dense M-ary PSK constellations, in Proc. IEEE Mil. Commun. Conf. Record, 1990, pp. 768770.
[50] P. J. Winzer and R. J. Essiambre, Advanced optical modulation formats, Proc. IEEE, vol. 94, no. 5, pp. 952985, May 2006.
[51] P. J. Winzer and R.-J. Essiambre, Advanced optical modulation formats, in Optical Fiber Telecommunications VB, I. Kaminow and T. Li,
Eds. New York: Academic, 2008, pp. 232304.
[52] F. Gray, Pulse Code Communication, U.S. Patent 2 632 058, 1953.
[53] S. Verd, Spectral efficiency in the wideband regime, IEEE Trans.
Inf. Theory, vol. 48, no. 6, pp. 13191343, Jun. 2002.
[54] R. Tkach, Scaling optical communications for the next decade and
beyond, Bell Labs Tech. J., vol. 14, no. 4, pp. 39, 2010.
[55] J. B. Stark, Fundamental limits of information capacity for optical
communications channels, in Proc. Eur. Conf. Opt. Commun., 1999,
pp. I28.
[56] E. Desurvire, A quantum model for optically amplified nonlinear
transmission systems, Opt. Fiber Technol., vol. 8, pp. 210230, 2002.
[57] E. Desurvire, A common quantum noise model for optical amplification and nonlinearity in WDM transmission, in Proc. Eur. Conf. Opt.
Commun., 2002, p. Tu3.1.1.
[58] E. Desurvire, Classical and Quantum Information Theory. Cambridge, U.K.: Cambridge Univ. Press, 2009.
[59] J. Tang, The Shannon channel capacity of dispersion-free nonlinear
optical fiber transmission, J. Lightw. Technol., vol. 19, no. 8, pp.
11041109, Aug. 2001.
[60] P. P. Mitra and J. B. Stark, Nonlinear limits to the information capacity
of optical fibre communications, Nature, vol. 411, pp. 10271030,
2001.
[61] E. Narimanov and P. P. Mitra, The channel capacity of a fiber optics
communication system: Perturbation theory, J. Lightw. Technol., vol.
20, no. 3, pp. 530537, Mar. 2002.
[62] J. Tang, The channel capacity of a multispan DWDM system employing dispersive nonlinear optical fibers and an ideal coherent optical
receiver, J. Lightw. Technol., vol. 20, no. 7, pp. 10951101, Jul. 2002.
[63] K. S. Turitsyn, S. A. Derevyanko, I. V. Yurkevich, and S. K. Turitsyn,
Information capacity of optical fiber channels with zero average dispersion, Phys. Rev. Lett., vol. 91, 2003, Paper 203901.
[64] L. G. L. Wegener, M. L. Povinelli, A. G. Green, P. P. Mitra, J. B. Stark,
and P. B. Littlewood, The effect of propagation nonlinearities on the
information capacity of WDM optical fiber systems: Cross-phase modulation and four-wave mixing, Physica D, vol. 189, pp. 8199, 2004.
[65] M. H. Taghavi, G. C. Papen, and P. H. Siegel, On the multiuser capacity of WDM in a nonlinear optical fiber: Coherent communication,
IEEE Trans. Inf. Theory, vol. 52, no. 11, pp. 50085022, Nov. 2006.
698
[115] J. Pierce, Physical sources of noise, Proc. IRE, vol. 44, no. 5, pp.
601608, 1956.
[116] H. A. Haus, Electromagnetic Noise and Quantum Optical Measurements. New York: Springer-Verlag, 2000.
[117] B. E. A. Saleh and M. C. Teich, Fundamentals of Photonics, 2nd ed.
New York: Wiley, 2007.
[118] B. E. A. Saleh, Photoelectron Statistics: With Applications to Spectroscopy and Optical Communication. New York: Springer-Verlag,
1978.
[119] L. Mandel and E. Wolf, Optical Coherence and Quantum Optics.
Cambridge, U.K.: Cambridge Univ. Press, 1995.
[120] P. J. Winzer, Linking equations between photon statistics and photocurrent statistics for time-varying stochastic photon rates, Quantum
Semiclass. Opt.: J. Eur. Opt. Soc. Part B, vol. 10, pp. 643655, 1998.
[121] E. Sckinger, Broadband Circuits for Optical Fiber Communication.
New York: Wiley, 2005.
[122] S. Alexander, Optical Communication Receiver Design. Herts, U.K.:
Institution of Engineering and Technology, 1997.
[123] A. Einstein, On the quantum theory of radiation, Phys. Zeits., vol. 18,
p. 121, 1917.
[124] C. W. Gardiner and P. Zoller, Quantum Noise: A Handbook of
Markovian and Non-Markovian Quantum Stochastic Methods With
Applications to Quantum Optics, 3rd ed. New York: Springer-Verlag,
2004.
[125] E. Desurvire, D. Bayart, B. Desthieux, and S. Bigo, Erbium-Doped
Fiber Amplifiers and Device and System Developments. New York:
Wiley, 2002.
[126] P. Diament and M. C. Teich, Evolution of the statistical properties
of photons passed through a traveling-wave laser amplifier, IEEE J.
Quantum Electron., vol. 28, no. 5, pp. 13251334, May 1992.
[127] T. Li and M. C. Teich, Photon point process for traveling-wave laser
amplifiers, IEEE J. Quantum Electron., vol. 29, no. 9, pp. 25682578,
Sep. 1993.
[128] J. P. Gordon, W. H. Louisell, and L. R. Walker, Quantum fluctuations and noise in parametric processes II, Phys. Rev., vol. 129, pp.
481485, 1963.
[129] J. P. Gordon, L. R. Walker, and W. H. Louisell, Quantum statistics of
masers and attenuators, Phys. Rev., vol. 130, pp. 806812, 1963.
[130] J. R. Barry, D. G. Messerschmitt, and E. A. Lee, Digital Communication, 5th ed. New York: Springer-Verlag, 2003.
[131] A. Mecozzi, Limits to long-haul coherent transmission set by the Kerr
nonlinearity and noise of the in-line amplifiers, J. Lightw. Technol.,
vol. 12, no. 11, pp. 19932000, Nov. 1994.
[132] C. W. Gardiner, Handbook of Stochastic Methods: For Physics, Chemistry and the Natural Science, 3rd ed. New York: Springer-Verlag,
2004.
[133] P. C. Becker, N. A. Olsson, and J. R. Simpson, Erbium-Doped Fiber
Amplifiers, Fundamentals and Technology. New York: Academic,
1999.
[134] A. Bjarklev, Optical Fiber Amplifiers: Design and System Applications. Boston, MA and London: Artech House, 1993.
[135] M. Born and E. Wolf, Principles of Optics: Electromagnetic Theory
of Propagation, Interference and Diffraction of Light. Cambridge,
U.K.: Cambridge Univ. Press, 1999.
[136] R. H. Stolen and E. P. Ippen, Raman gain in glass optical waveguides,
Appl. Phys. Lett., vol. 22, pp. 276278, 1973.
[137] J. Bromage, Raman amplification for fiber communications systems,
J. Lightw. Technol., vol. 22, no. 1, pp. 7993, Jan. 2004.
[138] V. E. Perlin and H. G. Winful, On trade-off between noise and nonlinearity in WDM systems with distributed Raman amplification, in Proc.
Opt. Fiber Commun. Conf. (OFC), 2002, pp. 178180, paper WB1.
[139] V. E. Perlin and H. G. Winful, Optimizing the noise performance
of broadband WDM systems with distributed Raman amplification,
IEEE Photon. Technol. Lett., vol. 14, pp. 11991201, 2002.
[140] L. F. Mollenauer and J. P. Gordon, Solitons in Optical Fibers: Fundamentals and Applications. New York: Academic, 2006.
[141] J.-C. Bouteiller, K. Brar, J. Bromage, S. Radic, and C. Headley, Dualorder Raman pump, IEEE Photon. Technol. Lett., vol. 15, pp. 212214,
2003.
[142] T. J. Ellingham, J. D. Ania-Castann, R. Ibbotson, X. Chen, L. Zhang,
and S. K. Turitsyn, Quasi-lossless optical links for broadband transmission and data processing, IEEE Photon. Technol. Lett., vol. 18, pp.
268270, 2006.
[143] P. B. Hansen, L. Eskildsen, A. J. Stentz, T. A. Strasser, J. Judkins, J. J.
DeMarco, R. Pedrazzani, and D. J. DiGiovanni, Rayleigh scattering
limitations in distributed Raman pre-amplifiers, , pp. 343345, 1997.
699
700
[173] R. Hellwarth, J. Cherlow, and T. T. Yang, Origin and frequency dependence of nonlinear optical susceptibilities of glasses, Phys. Rev. B,
vol. 11, no. 2, pp. 964967, 1975.
[174] R. Stolen, Nonlinearity in fiber transmission, Proc. IEEE, vol. 68, no.
10, pp. 12321236, Oct. 1980.
[175] R. Stolen, J. Gordon, W. Tomlinson, and H. Haus, Raman response
function of silica-core fibers, J. Opt. Soc. Amer. B, vol. 6, no. 6, pp.
11591166, 1989.
[176] K. Blow and D. Wood, Theoretical description of transient stimulated
Raman scattering in optical fibers, IEEE J. Quantum Electron., vol.
25, no. 12, pp. 26652673, Dec. 1989.
[177] P. Mamyshev and S. Chernikov, Ultrashort-pulse propagation in optical fibers, , vol. 15, no. 19, pp. 10761078, 1990.
[178] S. Chernikov and P. Mamyshev, Femtosecond soliton propagation in
fibers with slowly decreasing dispersion, J. Opt. Soc. Am. B, vol. 8,
no. 8, pp. 16331641, 1991.
[179] L. Eskildsen, P. Hansen, P. Koren, B. Miller, M. Young, and K.
Dreyer, Stimulated Brillouin scattering suppression with low residual
AM using a novel temperature wavelength-dithered DFB laser diode,
IEE Electron. Lett., vol. 32, no. 15, pp. 13871389, 1996.
[180] M. Van Deventer and A. Boot, Polarization properties of stimulated
Brillouin scattering insingle-mode fibers, J. Lightw. Technol., vol. 12,
no. 4, pp. 585590, Apr. 1994.
[181] D. Christodoulides and R. Jander, Evolution of stimulated Raman
crosstalk in wavelength division multiplexed systems, IEEE Photon.
Technol. Lett., vol. 8, no. 12, pp. 17221724, 1996.
[182] M. Zirngibl, Analytical model of Raman gain effects in massive wavelength division multiplexed transmission systems, IEE Electron. Lett.,
vol. 34, pp. 789790, 1998.
[183] A. Chraplyvy, J. Nagel, and R. Tkach, Equalization in amplified
WDM lightwave transmission systems, IEEE Photon. Technol. Lett.,
vol. 4, no. 8, pp. 920922, 1992.
[184] A. Tomita, Cross talk caused by stimulated Raman scattering in singlemode wavelength-division multiplexed systems, Opt. Lett., vol. 8,
no. 7, pp. 412414, 1983.
[185] W. Jiang and P. Ye, Crosstalk in fiber Raman amplification for
WDM systems, J. Lightw. Technol., vol. 7, no. 9, pp. 14071411,
Sep. 1989.
[186] F. Forghieri, R. W. Tkach, and A. R. Chraplyvy, Effect of modulation statistics on Raman crosstalk in WDM systems, IEEE Photon.
Technol. Lett., vol. 7, no. 3, pp. 101103, 1995.
[187] Raman Amplifiers for Telecommunications 1: Physical Principles, M.
Islam, Ed. New York: Springer-Verlag, 2004.
[188] C. Headley and G. P. Agrawal, Raman Amplification in Fiber Optical
Communication Systems. New York: Academic, 2005.
[189] R.-J. Essiambre, P. J. Winzer, J. Bromage, and C. H. Kim, Design
of bidirectionally pumped fiber amplifiers generating double Rayleigh
backscattering, IEEE Photon. Technol. Lett., vol. 14, no. 7, pp.
914916, 2002.
[190] H. A. Haus, The noise figure of optical amplifiers, IEEE Photon.
Technol. Lett., vol. 10, no. 11, pp. 16021604, 1998.
[191] H. J. Landau, On the recovery of a band-limited signal, after instantaneous companding and subsequent band limiting, Bell Syst. Tech. J.,
vol. 39, pp. 351364, 1960.
[192] H. J. Landau and W. L. Miranker, The recovery of distorted bandlimited signals, J. Math Anal. Appl., vol. 2, pp. 97104, 1961.
[193] S. Bigo, Modelling of WDM terrestrial and submarine links for the
design of WDM networks, in Proc. Opt. Fiber Commun. Conf., 2006,
Paper OThD1.
[194] A. Mecozzi, C. B. Clausen, M. Shtaif, S.-G. Park, and A. H. Gnauck,
Cancellation of timing and amplitude jitter in symmetric links using
highly dispersed pulses, IEEE Photon. Technol. Lett., vol. 13, pp.
445447, 2001.
[195] T. Freckmann, R.-J. Essiambre, P. J. Winzer, G. J. Foschini, and G.
Kramer, Fiber capacity limits with optimized ring constellations,
IEEE Photon. Technol. Lett., vol. 21, pp. 14961498, 2009.
[196] M. Eiselt, Does spectrally periodic dispersion compensation reduce
non-linear effects?, in Proc. European Conf. Opt. Commun., 1999, vol.
1, pp. 144145.
[197] G. Bellotti and S. Bigo, Cross-phase modulation suppressor for
multispan dispersion-managed WDM transmissions, IEEE Photon.
Technol. Lett., vol. 12, no. 6, pp. 726728, 2000.
[198] G. Bellotti, S. Bigo, S. Gauchard, P. Cortes, and S. LaRochelle,
10 10 gb/s cross-phase modulation suppressor using WDM narrowband fiber Bragg gratings, IEEE Photon. Technol. Lett., vol. 12,
pp. 14031405, 2000.
701
Gerard J. Foschini (S60M62SM83F86) received the B.S.E.E. degree from New Jersey Institute
of Technology, Newark, the M.E.E. degree from New
York University, Manhattan, and the Ph.D. degree in
mathematics from Stevens Institute of Technology,
Hoboken, NJ.
He has been at Bell Laboratories, Alcatel-Lucent,
Holmdel, NJ, for over four decades, where he is currently a Distinguished Inventor in the Wireless Research Laboratory at Crawford Hill, Holmdel. He has
done extensive research on point-to-point communication systems as well as on networks. He was also engaged in teaching at
Princeton University and is currently a Visiting Professor in the graduate faculty
in the Electrical Engineering Department, Rutgers University, Piscataway, NJ.
He has published more than 100 papers and holds 15 patents.
Dr. Foschini is a Fellow of Bell Laboratories. He is the recipient of the 2000
Bell Labs Inventors Award, for one of his wireless inventions on multiple antenna communications, and the 2002 Thomas Alva Edison Patent Award. He has
also won the Bell Labs Presidents Gold Award for the outstanding level of innovation, technical excellence and business impact. He also won the Bell Labs
Team Award, the 2006 IEEE Eric E. Sumner Award, and the 2008 Alexander
Graham Bell Medal for his work on multielement antenna technology for high
spectral-efficiency communications. He has been currently elected to the National Academy of Engineering.
Bernhard Goebel (S04) received the Diplom-Ingenieur degree in electrical engineering and
information technology from Technische Universitt
Mnchen (TUM), Munich, Germany, in 2004. He
is currently working toward the Dr.-Ing. degree
at TUM. His diploma thesis was on the information-theoretic analysis of complex genetic diseases.
He was at the University of Southampton,
Southampton, U.K., and Siemens Corporate Research, Princeton, NJ, during graduate semesters,
where his research was focused on segmentation
algorithms for medical imaging. Since 2004, he has been a Member of the
technical staff at the Institute for Communications Engineering, TUM. During
2009, he was a visiting Researcher for three months at Alcatel-Lucent Bell
Laboratories, Holmdel, NJ. His current research interests include information-theoretic aspects of optical communication systems. He has published
in the fields of medical imaging, bioinformatics, optical communications and
information theory, and is the inventor of one U.S. patent.
Mr. Goebel is a student member of the IET and VDE/ITG.