Sie sind auf Seite 1von 14

1242

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 21, NO. 5, MAY 2003

Modeling and Simulation of Next-Generation


Multimode Fiber Links
Petar Pepeljugoski, Senior Member, IEEE, Steven E. Golowich, Member, IEEE, A. John Ritger,
Paul Kolesar, Member, IEEE, and Aleksandar Risteski, Member, IEEE

AbstractThis paper describes an advanced multimode-fiber-link model that was used to aid the development
of Telecommunication Industry Association standard specifications for a next-generation 50- m-core laser-optimized multimode
fiber. The multimode-link model takes into account the interactions of the laser, the transmitter optical subassembly, and the
fiber, as well as effects of connections and the receiver preamplifier.
We present models for each of these components. Based on these
models, we also develop an efficient and simple formalism for the
calculation of the fiber transfer function and the signal at the link
output in any link configuration. We demonstrate how the model
may be used to develop specifications on transmitters and fibers
that guarantee any desired level of performance.
Index TermsModeling, mode delays, multimode fiber links, optical communications, 10 Gigabit Ethernet.

I. INTRODUCTION

HIS PAPER is the third and last in a set of three papers documenting recent multimode-fiber standards development
in the Telecommunications Industry Association (TIA). It provides an in-depth description of the theoretical model used for
the multimode-fiber standards development. The first paper [1]
focuses on the development of a 500-m 1-Gb/s network solution
using a 62.5- m fiber and short-wavelength (830860 nm) laser
transceivers. The second paper [2] describes the development of
a more generalized approach and focuses on a 300-m 10-Gb/s
solution using a 50- m fiber.
The IEEE 802.3ae Standard, also known as 10 Gigabit Ethernet (10 GbE), was approved for publication in June 2002.
As one of the Physical Media Dependent Layers, it includes
a short wavelength (850-nm) link over a multimode fiber for
lengths up to 300 m. This is an attractive choice for the physical
layer due to both the low cost of vertical-cavity surface-emitting
laser (VCSEL)-based transceivers, which do not require accurate alignment into multimode fibers, as well as legacy support
issues. Although a standard for a laser-optimized multimode
Manuscript received November 16, 2002; revised February 19, 2003.
P. Pepeljugoski is with the IBM T. J. Watson Research Center, Yorktown
Heights, NY 10598 USA (e-mail: petarp@us.ibm.com).
S. E. Golowich is with Bell Laboratories, Lucent Technologies, Murray Hill,
NJ 07974 USA.
A. J. Ritger is with the OFS Optical Fiber Division, Norcross, GA 30071
USA.
P. Kolesar is with the Connectivity Solutions Division, Avaya Laboratories,
Richardson, TX 75082 USA.
A. Risteski is with the St. Cyril and Methodius University, 1000 Skopje,
Macedonia.
Digital Object Identifier 10.1109/JLT.2003.811320

fiber capable of a 300-m reach did not exist at the beginning


of the 10-GbE standardization process, soon afterwards experiments demonstrated the feasibility of such a system [3], [4],
even under highly stressed conditions, using the first samples
of the laser-optimized next-generation fiber developed by the
optical fiber division of Lucent Technologies (now OFS). Subsequently, several fiber and transceiver companies, under the
auspices of the TIA and its FO-2.2.1 working group, started
to develop the specifications for this fiber. This development,
described in [2], required several years and involved simulations, round-robin measurements, and new test procedures for
the fiber.
In this paper, we describe the model that was used to aid
the development of the next-generation high-speed multimode
fiber, now standardized as 850-nm laser-optimized 50- m
multimode fiber in [5].
While experiments cannot be replaced by simulations, they
are time consuming and can probe only a small part of the available parameter space. In particular, it is difficult to determine
the worst-cases region of parameter space by experiments alone.
Ensuring a cost-effective solution requires thorough exploration
of the entire parameter space and, in particular, the marginal
zones. Since it is important to ensure that simple specifications
on the components will guarantee reliable operation under a
wide range of designs and manufacturing variations, it is necessary to use a model to explore the performance of the link over a
range of parameters describing the transmitters, fibers, and connections in the link. Extensive simulation studies will also assist
in designing and testing specifications, speed up finalizing the
specifications, and optimize the design tradeoffs with their associated costs for both the transmitter and fiber manufacturers.
Multimode fiber LAN links are complex links whose performance is affected by many physical effects [6]. Besides the
degradation due to finite rise and fall times at the transmitter and
the receiver, the performance of the link is affected by the intermodal and intramodal dispersion, as well as noises specific to
multimode fiber links (modal noise) or multimode lasers (mode
partition noise and noise due to laser mode hopping). However,
the main emphasis of our work is on the signal degradation
caused by the intermodal dispersion or the differential mode
delay in the fiber. This is due to the fact that the largest portion of the link power budget is consumed by the intersymbol
interference (ISI) penalty, caused by pulse spreading as a result
of the intermodal dispersion. An outline of the components covered in our model is presented in Fig. 1.

0733-8724/03$17.00 2003 IEEE

PEPELJUGOSKI et al.: MODELING AND SIMULATION OF NEXT-GENERATION MULTIMODE FIBER LINKS

Fig. 1.

1243

Various components included in the link model.

The components we consider (the transmitter, fiber, and connections) are the most important factors determining the 3-dB
optical bandwidth of the link, which is the parameter on which
we focused during the early stages of this work [7]. However,
it soon became clear that more relaxed specifications can be
achieved if, instead of the fiber 3-dB optical bandwidth, we use
the ISI at the output of the receiver, which directly relates to the
allocation in the link power budget and, therefore, is a more accurate system performance predictor [8].
The parameter space of component perturbations that can affect the link bandwidth is very high dimensional. To begin with,
the modal dispersion of multimode fiber is notoriously hard
to control. At the operating wavelength of 850 nm, the promultimode fiber initially supposed design of 50 m, 1%
ports 19 mode groups, each of which can potentially have its
own group velocity. This implies that the bandwidth and the
shape of the fiber transfer function may depend strongly on the
excitation conditions, which govern how much power will be
coupled into each mode group. As a consequence, the signal at
the receiver output will show strong dependence on the launch
conditions, the fiber properties, and the link configuration. The
proposed system uses VCSELs as sources, which may be either single or multimoded (though chromatic dispersion considerations limit spectral width and, hence, the number of laser
modes). The center of the laser may be offset from the center
of the fiber due to connector misalignment, and the beam spot
sizes may vary from source to source, both due to variation in
the VCSELs themselves and also due to different choices and
variation in the optics between the VCSEL and the fiber. All
these factors lead to significant variations from transmitter to
transmitter in the coupled power into the fiber modes or the
mode power distribution. Finally, the relatively loose tolerances
in multimode connections (of which, in practice, there are always at least two, in addition to those at the transmitter) lead to
redistribution of power among the modes of the two fibers.
One of the challenges in designing a robust link employing
VCSELs and multimode fiber is to define specifications for
transmitters and fibers that are simple enough to admit routine screening measurements but discriminating enough to guarantee that the system will operate if the specifications are met.
The TIA FO-2.2.1 group adopted a transmitter specification
based on an encircled-flux measurement and a fiber specification based on a differential-modal-delay measurement. One
of the primary motivations of our modeling work was to provide the capability of testing, through simulation, the efficacy
of these screening tests. The model can then be used to thor-

oughly explore the space of source/fiber/connection combinations and demonstrate that if the sources and fibers pass the proposed specifications, then the system will very rarely fail.
The outline of the rest of the paper is as follows. In Section II,
we describe the various components of the link model in detail and describe how the bandwidth and ISI penalty calculations are performed. We present an algorithm for calculating the
transfer function of any multimode link configuration, which
takes into account launch conditions and mode mixing effects
at the connectors. In Section III, we provide examples that illustrate the capabilities of the model and its usefulness in determining the optimum specifications. We conclude with discussion in Section IV.
II. LINK MODEL AND SIMULATION APPROACH
In this section, we describe the multimode-fiber-link model.
A detailed description of the signal propagation through the
driver, the laser, and the receiver, as well as the in-depth discussion of the model and mode partition noise, can be found in
[9] and [10]. For the sake of introducing the symbols and outlining the model capabilities, we give a brief description of the
laser model. Both the driver and receiver model are filter-based
models, whose rise time (or bandwidth), type, and order is supplied at the beginning of the simulation.
After the selection of the data rate, the model first generates a data pattern of given length, duty cycle distortion, and
deterministic jitter. Then, this signal is applied to the driver
and propagated further through the laser (Section II-A). The
interaction of the laser, the lens, and the fiber, as well as the
connector effects, is treated through the overlap integrals [11],
[12] to find the coupling coefficients from each laser mode into
each fiber mode (Sections II-BD). Then, the coupling coefficients into each fiber-mode group are calculated. With the
knowledge of the mode delay times, which in a simulation study
are drawn randomly from a distribution, we can calculate the
transfer function for each fiber section (Section II-E). Connector
effects are treated through the use of connector transfer matrices
(Section II-F). After calculating the transfer function of all fiber
sections (Section II-G), the signal is propagated through the receiver (Section II-H). Once the time-domain output signal is
known, we calculate the ISI penalty, the timing jitter, and available retiming window. From the fiber transfer function, we find
the fiber bandwidth of interestthe 3-dB optical, 3-dB electrical, or another bandwidth measure.
In Fig. 2, we illustrate the definition of the ISI penalty, the
retiming window, and the timing jitter. The ISI penalty in this

1244

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 21, NO. 5, MAY 2003

(2)
It should be noted that these rate equations are single mode
rate equations. Multimode rate equations are a straightforward
extension. Although almost all of the lasers used in the 850-nm
wavelength range with multimode fiber are multimode lasers,
including VCSELs, the single-mode rate equations have been
found to be a very good approximation to the large signal behavior and have the added benefit of simplicity, short computation time, and fewer convergence problems.
B. Electrical Field at VCSEL Output
Following [11], we represent the transverse modes of a
VCSEL as Gaussian beam modes. These modes, centered at
the origin and parallel to the axis, take the form
Fig. 2. Eye diagram from a typical multimode-fiber link. The figure illustrates
the definition of the ISI through the top line, baseline, and eye height, as well
as the retiming window and deterministic jitter.

(3)
paper is defined as ISI dB
, where
is
the signal amplitude (the difference between the top line and
is the eye-diagram height (the maxbaseline in Fig. 2), and
imum opening within the bit interval1 ). Using the above definition, the ISI calculation is straightforward, once the signal is
known. Similarly, we define the retiming window as the maximum width in which the eye opening drops 0.5 dB from the
center of the eye. The deterministic jitter is the peak-to-peak
difference of the zero-crossing times of the signal. The zerocrossing level can be adjusted by the user, but in our simulations,
it was assumed to be at zero. The retiming window is important
because it reflects the inability of the clock and data recovery
circuit to make the decision always at the same position.
While the individual calculations outlined have been done
before [11][14], this is the first time that an integrated multimode-link model has been used to run the entire simulation of
data signals for bit rates at 10 Gb/s or above.
A. Laser Model
The laser rate equation model is needed if one wants to perform accurate link simulations. For simulation purposes, we do
not need the temperature dependence of the laser model; therefore, we use the nonlinear laser rate equations. From the many
variations and notations in the literature, here we use the notation found in [15].
The single-mode laser rate equations are a system of two nonlinear differential equations, one for the carriers or the electron
number , given in equation (1), and the other for the photon
number , given in (2). The definition of the parameters used
is the same as in [9, Table I]. These equations include the gain
[16]
nonlinearity factor

(1)
1It

is assumed that the decision point is at the center of the eye.

and
are the radial and angular mode numwhere
is the spot size at the waist,
is the free-space
bers,
are the generalized Laguerre polynomials, and
wavenumber,

(4)
The various modes can be divided into degenerate groups indexed by their principle mode number (PMN), defined to be
. Chromatic dispersion considerations for 10-Gb/s
systems limit the VCSELs to be either single- or few-moded,
so that in the simulations below, we include the modes of the
first four mode groups in various superpositions [2]. Two representative modal fields of the six modes comprising these mode
groups are presented in Fig. 3.
Although the Gaussian beam model for a laser cavity is very
standard, its validity for VCSEL structures may be questioned;
in fact, it is safe to say that it is not a good model for some
highly multimoded VCSELs [17]. However, for the few-moded
VCSELs relevant to this work, there is some experimental evidence that the Gaussian beam model is a reasonable approximation. For instance, the beam shapes of some single-mode
VCSELs have been directly observed to be Gaussian [18]. In addition, the relationship between the splitting of VCSEL spectral
lines and the Gaussian beam spot size, predicted by Gaussian
beam theory, has been observed to hold [19]. However, additional measurements probing the validity of the Gaussian beam
model in detail are certainly desirable; in their absence, we accept the model at least as a reasonable starting point.
A potential complication that we do not address is the nonstationary character of the laser modes, i.e., the mode structure
of the laser changes as the signal evolves. Different VCSEL

PEPELJUGOSKI et al.: MODELING AND SIMULATION OF NEXT-GENERATION MULTIMODE FIBER LINKS

1245

turers. Therefore, for the TIA work (and also for the results presented in this paper), we modeled the laser as a filter with given
bandwidth. The laser model was used to simulate the behavior
of a 1-km-long next-generation fiber link operating at 20 Gb/s
in [10].
C. Electric Field in the Fiber

(a)

(b)
Fig. 3. Fields of the two of the modes comprising the lowest four Gaussian
beam mode groups (indexed by PMN). These fields, in various superpositions,
are used to model the VCSEL fields included in the simulations. The inner
dashed circle is the Gaussian beam spot size (6.25 m in this example); the
outer circle is the radius of the fiber (25 m).

modes may be excited at different times due to the fact that different laser modes may have different thresholds; therefore, they
appear at different times, and their impact on the signal shape
varies in time. While the rate equation model of the laser allows modeling of this effect, there are variations among laser
manufacturers and from device to device from the same manufacturer, which make it difficult to obtain an accurate statistical
distribution of the relevant laser parameters. Although our program implemented the laser rate equation model, at the time of
the work done by the TIA, we did not have sufficient data to
properly model such laser dynamics from all VCSEL manufac-

We will now review some relevant features of propagation


in multimode fiber and describe the assumptions on which our
propagation model is based. We will work entirely within modal
theory, based on solutions to the paraxial wave equation (see the
appendix for mathematical details). Briefly, the modes of a multimode fiber can be divided into approximately degenerate mode
groups, the modes of which nearly share a common phase velocity. If we index the mode groups by an integer
,
with the fundamental mode assigned the index of 1, it can be
shown that the degeneracy, or the number of modes including
polarization, of a mode group is twice the index of the group. In
our case, with a core diameter of 50 m and of 1%, the fiber
supports about 400 modes divided into about 20 mode groups.
Pulses launched into the various modes of the fiber may travel
at different group velocities, which are determined by the index
profile of the fiber and the properties of the doped silica constituting the fiber. In fiber with perfect translational invariance, the
various modes propagate independently of one another. In actual fibers, however, imperfections allow power to couple from
one mode to another. There is ample evidence that mode coupling between different mode groups is vanishingly small over
length scales of hundreds or even thousands of meters [20],
[21]. However, the coupling of modes within a mode group
occurs over much shorter length scales, typically hundreds of
meters [22][24]. In our model, we will assume that mode coupling between mode groups is completely absent and that coupling within a group is always complete. A consequence of
this assumption is that modal dispersion within a mode group
is absent; an entire mode group propagates as a single composite mode, traveling at a well-defined group velocity without
spreading.
In addition to mode coupling, propagating modes in a fiber
also suffer attenuation. A consequence of the complete coupling
within a mode group and lack of coupling between mode groups
is that attenuation can be completely described by assigning
each mode group its own attenuation rate , so the amplias it
tude of a pulse launched into group scales as
travels down the fiber.
There are other physical effects that influence propagation
in a multimode fiber that we will ignore. The most important
of these are the chromatic dispersion, the laser coherence and
mode selective loss resulting in modal noise, and the mode partition noise due to interaction of the chromatic dispersion and the
laser noise. Since these are included in the Ethernet link power
budgets, they will not be considered here. Similarly, the polarization-mode dispersion will not be taken into account. First,
its magnitude is masked by the much larger effect of modal
dispersion in multimode fibers. and second, due to the scalar
approximation we use throughout, both polarizations behave
identically.

1246

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 21, NO. 5, MAY 2003

Fig. 4. Various coordinate systems used in deriving the transformation of a Gaussian beam in passing through an airglass interface.
is the position of the virtual beam, as seen from inside the glass.
points out of the page.
actual launch in air, while

D. VCSELFiber Coupling
mode illuminates the end face of a
When a VCSEL
fiber, the transmitted component of the field is transformed
by the airfiber interface into a different Gaussian beam
mode, which then excites the various modes of the fiber. We
assume the launching beam is incident from air onto the end
face of the fiber. Due to the weakly guiding nature of the fiber,
it is an excellent approximation to neglect the grading in the
index of the fiber core when computing the transformed fields
and thus to consider a simple interface between air and glass for
this part of the calculation. Furthermore, in the discussion that
follows, we interpret the incident Gaussian beam as modeling
the output of any lens between the laser and the fiber. This is
safe to do if the lens is ideal, since an ideal lens will transform
the Gaussian beam modes from the laser into different Gaussian
beam modes incident on the fiber. We should note here that
lenses used in manufacturing inexpensive transceiver modules
for LAN applications can be far from ideal. Lens aberrations
can significantly change the laser electric field at the fiber input
and are difficult to model. Nevertheless, we concentrated on
the ideal case in order to give us insight into the impact of the
various laser and fiber parameters on the link performance.
The geometry we work with is illustrated in Fig. 4. Three
is defined
coordinate systems come into play.
with respect to a system with the origin at the center of the fiber
the unit vector pointing along the fiber axis.
end face, with
has as the origin the center of the beam waist
is rotated by around
of the launching beam in air, while
with respect to . The origin of the
is
of the virtual beam seen inside the glass. It is
at the center
around
with respect to .
and
will
rotated by
, and
are
be computed subsequently. Finally, all of ,
parallel. Note that, because of the circular symmetry of the fiber,
we are without loss of generality specialized to the case of the
plane.
launching beam tilted only in the
We denote the transformed Gaussian beam mode transverse
,
electric (TE) field just inside the glass by
and
and we next derive the coordinate transformation
and . We begin by considering
the transformed parameters
the special case of normal incidence. In this case, we can com-

is the position of the

pute the transformation of Gaussian beam parameters using the


formalism [11]
(5)
The
where the

theory tells us that


parameter is defined as [11]

(6)
, we find
and
.
Applying (5) and using
This result is easily explained in physical termsthe intensity profile should not change at the interface, but curvature
changes by a factor of due to Snells law acting locally on
the phase fronts. We next combine (6) with the definitions in
(4) to interpret the transformed beam with parameters and
as a virtual beam launched at position from the interface and
with beam width at waist , with
(7)
axis (see
In the presence of a launch tilted by around the
Fig. 4), we apply Snells law in combination with (7) to conand
clude that the virtual beam has apparent launch position
of
an angle
(8)
is given by
Finally, the coordinate transformation
, where is the rotation
(9)
We may now compute the coupling amplitudes
Gaussian beam mode with the
incident
mode

of the
fiber
(10)

where the integral is over the area

of the fiber end face.

PEPELJUGOSKI et al.: MODELING AND SIMULATION OF NEXT-GENERATION MULTIMODE FIBER LINKS

E. Fiber Impulse Response


Under our assumptions, the impulse response of a fiber takes
on a very simple form. We will assume that impulses from the
transmitter induce electric fields at the input end face of the fiber
of the form
(11)
are complex amplitudes. These excitations, in turn,
where
into each mode of
induce impulses of differing energies
the fiber, where

1247

therefore, the impact of parameter mismatch will be small. We


will thus assume all fibers in the link have the same core radius
and .
Having neglected the parameter mismatch and tilt, the remaining effect due to a connection is offset. The effect of an axis
offset in a connection is to re-launch light from each mode of the
input (launching) fiber into some combination of modes of the
output (receiving) fiber; light that is not collected by a guiding
mode of the receiving fiber is lost. We calculate the coupling
to output mode
amplitude from input mode
as
(15)

(12)
(we comment subsequently on the validity of this idealization).
Because of the complete coupling within mode groups, we sum
the energies of the impulses within each group to obtain the
mode group weights excited by the transmitter
(13)
is known as the mode power distribution (MPD). DeThe set
noting the group delays per-unit length (inverses of the mode
of the
group velocities) as , we obtain the response
by the transfiber at position to an impulse launched at
mitter to be
(14)
where

is the attenuation coefficient for mode .

F. Connector Impact on Fiber Electric Fields


The connector introduces another level of variability in the
link performance due to the mechanical tolerances in manufacturing for both the connector and the fiber. As a result, there
might be offset and tilt, in which the axes of the two fibers are
misaligned, and parameter mismatch, in which the core radius
and of the launching and receiving fiber are different. Each
of these effects can lead to light being lost to cladding and radiation modes or redistributed among guided modes at a connection [25], which we refer to as mode coupling. This loss and
redistribution of light can introduce distortions in analog links
[14] or cause excessive ISI penalty [26]. The tilt is considered
to be sufficiently small and is neglected in the calculation of the
connector transfer matrix.
In this paper, we will ignore parameter mismatch. The tilt is
considered to be so small in practice that its effect will be negligible. The effect of parameter mismatch has been explored in
[25], where it was shown that a reasonably large parameter mismatch leads to very little mode coupling. It is possible, however,
for light to be lost to cladding and radiation modes as a result
of parameter mismatch. This loss, however, can occur only from
the highest order mode groups and is therefore only significant if
these groups carry a substantial fraction of the power in the fiber.
In the applications we consider, launches are constrained so that
the highest order modes do not carry a large amount of power;

where is the offset, and compute the power


into output mode

coupled

(16)

is the power present in mode


at the
In (16),
end of the input fiber. Our propagation assumptions dictate that
these powers are constant across mode groups; therefore
(17)
is the degeneracy of group
. We
where
will apply these formulas to both long lengths of fiber (hundreds
of meters) and short patch cords (less than a few meters long).
For the short lengths, the propagation assumption of complete
coupling within a group do not apply; therefore, (17) does not
apply, and one might worry that (16) will yield incorrect results.
Although this is a concern, we have found empirically that the
are relatively insensitive to
results for total power in a group
the distribution of power within a group for the types of excitation conditions we consider.
These calculations assume an idealized model of a connection. The validity of the results is supported by a large literature
dating from the early days of multimode fiber [25]. In addition,
measurements specifically covering the conditions of interest to
next-generation LaserWave fiber were recently performed that
again confirm the theoretical results [27].
G. Calculation of the Fiber Transfer Function
The knowledge of the transfer function of the fiber is necessary to propagate the signal through the link described previously, as it is not always efficient or easy to perform the calculations in the time domain. Furthermore, it is necessary to find
the fiber bandwidth, which is one of the requirements that the
fiber needs to meet. In this subsection, we outline the calculation of the fiber transfer function for an arbitrary link configuration consisting of several multimode fibers and connectors, if
the mode power distribution at the fiber input is known. The formalism, which is based on vector and matrix representation of
the link elements, allows efficient treatment of the propagation
through multimode-fiber links. The formalism can be applied to
any number of sections of fiber and connectors.

1248

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 21, NO. 5, MAY 2003

The mode power distribution is represented by the vector


from (13)
(18)
and the vector of mode group delay times per-unit length is
, where the numbers
given by
denote the mode group number and the superscript denotes
,
transpose. If the signal at the input of the fiber is denoted
then the signal at the output of the fiber for each mode group (neglecting the differential mode attenuation) can be represented in
such that
time domain as a time-dependent vector

(19)

In the frequency domain, using the properties of Fourier


transform, it is straightforward to find the transfer function for
the signal propagation in each mode group in the fiber. In the
, each row corresponds to the transfer function
matrix
of one mode group

(20)

where FT denotes Fourier transform and is a continuous variable representing the angular frequency. When implemented,
the angular frequency is assigned discrete values and becomes
is of size
.
a vector of size 1 by , so the matrix
In this case, the exponential function is applied to the individual
.
depends on the
elements of each row of the matrix
desired frequency resolution and the highest frequency of interest. The frequency-domain representation of propagation in
the fiber (for each mode and arbitrary mode power distribution)
can be calculated using matrix multiplication and is given by

(21)

is a unit diagonal matrix of size .


is now
where
a complex matrix that takes into account the delays in the mode
groups and their influence on the signal. It is calculated as a right
hand matrix multiplication of the MPD vector by the identity
. This signal comes
matrix and the fiber-delays matrix
of size
,
to a connector that is described by a matrix
which represents the redistribution of power among the mode
groups as a result of offset and tilt in the connector. This matrix is easily computed using the equations in Section II-F. The
signal at the output of the connector (and input of the next fiber)
becomes
(22)
or in frequency domain, the effect of the connector is given by
(23)

Fig. 5. Link configuration consisting of a laser diode, a 1-m fiber connected


to a fiber with length L , connected to a fiber with length L , and ended with
another 1-m fiber. The connectors have random offsets, resulting in different
connector transfer matrices.

where denotes the matrix multiplication. The connector effect is taken into account with a left-hand-side multiplication by
of the frequency-domain repthe connector transfer matrix
resentation of the fiber output. Then, the system can be propagated through the second fiber, using the previously described
represents a unit vector of size
formalism. If
, then the transfer function, which is a sum of all rows in
the matrix, is given by
(24)
H. Example of Fiber Propagation With the Formalism
Let us have a link that consists of several pieces of fiber with
various lengths and several connectors. The system is schematically shown on Fig. 5. Taking into account that the propagation
through each fiber is represented by (21) and each connector is
represented by (23), we can easily find the signal at the output
of the system, as follows:

(25)
In this expression, we assumed the delays are negligible in
the 1-m fibers and that all fiber modes are coupled into the phoand
represent the fibers with
todetector. The matrices
and , respectively. Please note that if the matrix
length
multiplication by is left out, then the output signal is represented as a matrix whose number of rows is equal to the number
of modes (one gets the signal waveform into each mode group),
and the number of columns represents the frequency.
I. Receiver Model
A typical LAN receiver consists of a photodiode followed by
transimpedance amplifier and several stages of postamplifiers.
For low-input optical powers, the signal at the output of the
transimpedance amplifier is not saturated yet, and the saturation
occurs in the postamplifier chain. In this case, the relevant test
point for the performance of the link is the preamplifier output.
A detailed receiver model that can also calculate the noise
at the receiver output and is suitable for IBM receivers used in
LAN and interconnect applications was described in [9]. This
model can be used when the details of the receiver are known.
In the work of the TIA FO-2.2.1, where the links contain receivers from several component manufacturers, we need to use
more generic model. For these reasons, the receiver model used
in the multimode-link model is a filter-based model. Since in

PEPELJUGOSKI et al.: MODELING AND SIMULATION OF NEXT-GENERATION MULTIMODE FIBER LINKS

Fig. 6.

1249

(a) Laser encircled flux and (b) fiber DMD.

this paper we are mainly interested in the signal propagation


and we are not addressing the noise in the system, this model is
adequate. The receiver is approximated to be a low-pass Butterworth filter2 with given bandwidth and filter roll-off. The
roll-off factor of the transfer function was controlled through the
filter order , with higher values of generating faster decay
in the transfer function as a function of frequency. We used
between 2 and 4, which is representative for LAN receivers. In
general, its transfer function is a ratio of two polynomials whose
coefficients are determined by the filter order, filter type, and
bandwidth.
The receiver model we used in the simulations was experimentally verified on IBM receivers for short-wavelength Gigabit Ethernet links [10].
III. SIMULATION INPUTS AND RESULTS
Numerous transmission experiments have been performed on
research links showing that robust performance over lengths beyond 300 m, and data rates exceeding 10 Gb/s is indeed possible. However, it is impossible through experiment alone to
probe more than a small part of the space of possible transmitter, fiber, and connection configurations. The primary use of
the modeling tool described in this paper was to perform simulations to aid in the development of the specifications for the
next-generation 50- m multimode fiber and transmitters used in
those links. The results of those simulations are presented in [2].
In contrast with [2], the simulation results shown in this paper
apply to links 600-m long and are chosen to illustrate the use of
the modeling tool, as well as how to use the simulation results
to develop specifications for the transmitter and the fiber.
To perform the simulations, we chose distributions for the
modal delay structure of the fibers, the modal power distributions of the sources, and the connection offsets under several
different fiber configurations that are representative of what
might be used in practice. Detailed description of these distributions can be found in [2]. The link configurations included the
2The receiver is more precisely modeled as a bandpass filter with the low
frequency cutoff to be only a small fraction of the data rate. In our case, it was
set to be 0.01% of the data rate, or about 100 kHz. With this assumption, we
take into account possible dc-wander due to long run lengths in the data pattern.

cases of straight link (no connectors), one or two short jumpers


followed by a 600-m link and ending again with a short jumper,
and a short jumper followed by two medium length fibers
(300 and 300 m), again followed by a short jumper. We denote
these four link configurations as No connectors, 1-600-1,
1-1-600-1, and 1-300-300-1, where the numbers represent
the length of each of the fiber sections in the link. We tested
whether the components met the proposed specifications by
calculating the results of an encircled flux (for the transmitter)
and differential-mode delay (DMD) (for the fiber). Finally, we
calculated the 3-dB optical bandwidth of a large sample of
transmitter/fiber/connection triplets and the ISI, deterministic
jitter, and retiming window of all the link configurations.
At this point, we need to discuss the use in laser and fiber
specifications of the laser encircled-flux and fiber DMD, instead
of the laser electric field and fiber delay times. Since both the
laser electric field at fiber input and the fiber-mode group delay
times are very difficult to measure, we need to relate them to
easily measurable quantities. The laser encircled flux is a standardized measurement [28] and is the fraction of the total optical
power in the fiber contained in a given radius. The fiber DMD is
also standardized [29] and is what one measures when the fiber
is excited with a single-mode source at different radial offsets.
The single-mode source excites only a subset of fiber- mode
groups at each offset. After propagation through the fiber and
detection in the receiver, the observed signal contains several
peaks corresponding to the excited mode groups, whose heights
are proportional to the launched power, and a peak position that
corresponds to the propagation delay of the mode group. The
DMD is extracted from this signal and is defined as the maximum width from the 25% height of the signal peak value on a
leading and trailing edge over the defined range of offsets. The
details of the DMD measurements and calculation can be found
in [29]. In this paper, we arbitrarily use two DMD ranges: one
from 018 m, which we called inner DMD, and the other the
outer DMD from 023 m. Please note that these definitions are
different than those used in [5]. Fig. 6 shows simulated curves
for both the laser encircled flux and the fiber DMD.
We would also like to comment on our use of statistical distributions. One might instead prefer to compute the bandwidth or
the ISI of the worst-case scenario and require the specifications

1250

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 21, NO. 5, MAY 2003

Fig. 7. Inputs and outputs of the model. On the left are the mode power distribution and the fiber delays, which are model inputs, and on the right, the fiber
transfer function and the eye diagram at link output, which are model outputs.

to be tight enough to guarantee that it has the necessary bandwidth or ISI. However, it is not at all clear what the worst-case
triplet of transmitter/fiber/connection is. One way to view our
Monte-Carlo approach is as an attempt to find the worst-case
bandwidth for a given set of specifications. In practice, one can
glean much more than an estimate of these quantities from the
simulations; for instance, we compute the fraction of links that
fail to meet a given threshold bandwidth. The failure rates we
calculate are, of course, not estimates of actual installed system
failure rates, since we do not have accurate estimates of the actual transmitter/fiber/connection distributions that will be seen
in practice. However, by choosing sufficiently flexible models,
we can be reasonably sure that the simulations will include approximations of most configurations that will be seen in the
field. Very small failure rates in the simulations will translate to
very small failure rates in the field, as long as the actual distributions are not concentrated near the failing examples. By looking
at the failing cases in the simulations, we can judge how likely
this is to be the case. In addition, component manufacturers can
tailor their manufacturing distributions toward designs that are
far from the failure boundaries.
In the simulation study, we drew 2000 sources, 5000 fibers,
and 1000 links (250 of each of the four types) randomly from
the distributions defined in [2]. These source/fiber pairs were
simulated under each of the four connection configurations,
for a total of 40 000 links. The connection offsets for each
source/fiber/configuration triplet were also drawn randomly

[2]. We simulated DMDs from each of the delays. The DMDs


were generated by assuming a receiver with 25-GHz bandwidth, whose transfer function shape was that of a fourth-order
Butterworth filter.
The rest of the link parameters for the simulations were taken
to be the worst-case transmitter and receiver specifications for
the Draft IEEE 802.3ae Standard at the time the simulations
were performed. That meant that we used 31.5 ps for the rise/fall
time for the signal at the output of the laser and a receiver
bandwidth of 8250 MHz. To take into account the variability
of receiver designs between different transceiver manufacturers,
the simulations were repeated for several different receiver designs, whose main difference was the receiver transfer function
roll-off factor. In the simulation environment, this was achieved
by changing the filter order, while keeping the receiver bandwidth fixed. The allowed ISI was taken from the IEEE 802.3ae
draft link power budget. The data pattern was a random pseudosignal, but with pattern
random binary sequence (PRBS)
length extended to 400 b. Since the simulations were performed
for 600-m-long links, the root mean square (rms) linewidth of
the laser was set to 0.2 nm.
In Fig. 7, we illustrate the flow of information in the program
and the kind of outputs we get from it. On the left-hand side of
Fig. 7, we show the mode power distribution at the input of the
fiber and the mode group delays of the fiber, which are inputs
necessary to find the signal at the link output. On the right, we
show the fiber transfer function and, below it, the eye diagram

PEPELJUGOSKI et al.: MODELING AND SIMULATION OF NEXT-GENERATION MULTIMODE FIBER LINKS

Fig. 8.

1251

Same as for Fig. 7, but for different link inputs, resulting in much worse performance, measured by increased ISI penalty and deterministic jitter.

at the link output. To highlight the impact that the shape of the
fiber transfer function has on the performance of the link, and
the need to use the ISI as a predictor of the link performance, we
also show the same inputs and outputs for another mode power
distribution and fiber delays that result in a transfer function that
has the same 3-dB optical bandwidth but significantly different
ISI and deterministic jitter (Fig. 8).
Figs. 912 illustrate the capabilities of the model. In Fig. 9(a),
we show a two-dimensional (2-D) plot of the ISI penalty as
a function of the lasers radius of 86% encircled flux and the
fibers inner DMD (018 m).The magnitude of the ISI penalty
is represented as shades of gray, with white representing values
greater than 2.6 dB. Fig. 9(b) shows a different view of the same
data; it contains a contour plot of the failure rates at given specification limits for the radius of 86% encircled flux and fiber
inner DMD. In both plots, we required that all links have outer
fiber DMD (023 m) less than 0.3 ps/m, inner encircled flux
smaller than 30%, and an overfilled launch (OFL) bandwidth of
the fiber higher than 1500 MHz km. This figure demonstrates
the possible tradeoffs between the requirements on the laser encircled flux and the fiber DMD. It suggests that from the fiber
perspective, the optimum setting is when the encircled flux is
0.165 ps/m
from 1316 m. In this case, fiber with DMD
can be used. Similarly, from the laser perspective, the optimum
setting is when the fiber DMD is less than 0.12 ps/m, yielding
an acceptable range of the lasers radius of 86% encircled flux
from 1019 m.

To further show the capabilities of the model to arrive at


fiber and laser specifications, we calculated the percentage of
links in which the allowed ISI penalty is exceeded. The simulations were performed for 600-m-long multimode links. The
inner DMD mask was 0.15 ps/m, and the outer DMD mask
was 0.3 ps/m. The rms linewidth of the source was chosen to
be 0.2 nm. The transmitter rise time and receiver bandwidth
were as indicated previously. The results in Fig. 10 show failure
rates below 5% for realistic sources whose 86% encircled flux
is between 12 and 16 m. We also observe a rapid increase in
the failure rate for sources whose encircled flux is greater than
17 m. The cumulative failure rate, which takes into account
the overall distribution, implies that the overall failure is much
lower.
The results shown on Fig. 10 include both single-mode and
multimode lasers. Fig. 11 shows the failure rates for different
link configurations when only single-mode lasers are selected.
All other parameters are the same as in Fig. 10. We observe
higher failure rates than for the multimode devices. This might
be due to center-line defects in the fiber delay distribution. It is
possible to avoid deleterious effect of the center-line defect by
either introducing offset or designing a lens with proper magnification. Although the failure rate does not rise as dramatically as for multimode lasers when the encircled flux exceeds
17 m, it is still much higher, and the increase starts at 16 m.
We can make similar conclusions for the cumulative failure rate,
as shown in Fig. 10.

1252

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 21, NO. 5, MAY 2003

(a)

(a)

(b)
Fig. 9. (a) Two-dimensional representation of the ISI penalty and (b) failure
rate as a function of laser encircled flux and fiber DMD (018 m). The figure
illustrates the tradeoffs between the requirements on the laser encircled flux and
the fiber DMD. The outer DMD was restricted to less than 0.3 ps/m, the inner
encircled flux at 4.5 m to 30%, and the OFL bandwidth to 1500 MHz 3 km.
The figure suggests that the optimum setting for the fiber is when the encircled
flux is around 15 m (63 m). In this case, the fiber whose DMD is less than
0.15 ps/m can be used. Similarly, the optimum setting for the transmitter is a
fiber whose DMD is less than 0.12 ps/m, and the wide range of encircled flux
(1019 m) can be used.

Two features of the curves in Fig. 10 that we wish to highlight are their convex shapes with minima at moderate values
of the encircled flux (EF) radius, and the effect of mode coupling at connections, which explains the differences between
the various configurations in the figure. The first feature can be
understood by noting that the fiber DMD specification allows
poorly behaved low-order and high-order fiber modal delays but
requires those in between to be tightly controlled. Therefore,
sources that launch too much of their power near the center, or
too near the cladding, are more likely to result in link failures
than those that avoid either extreme. The shape of the DMD
specification also explains the connection effects. For multimode sources, the highest failure rates are observed in the connectionless link because, in the absence of connections, power

(b)
Fig. 10. Failure rate as a function of encircled flux for 600-m-long multimode
link configurations with both multimode and single-mode lasers included. A
mask with 0.15-ps/m inner DMD and 0.3-ps/m outer DMD was used as a screen:
(a) cumulative failure rate and (b) distribution of failure rate.

that is concentrated in the problematic low- and high-order fiber


modes must stay there, whereas the effect of connections is to
diffuse power more evenly across the fiber modes. For the case
of multimode sources, this diffusion effect almost always results
in a lowering of the failure rate, because multimode sources are
capable of concentrating power in the low- and high-order problematic modes but not in the well-behaved intermediate modes.
In the case of single-mode sources, by contrast, we see cases
where the diffusion effect of connections increases the failure
rate by coupling power out of the intermediate modes into the
low- and high-order modes.
We also considered values different from 86% for the encircled-flux threshold. The TIA adopted the radius of the 75% encircled flux as specification for the sources used at 1.25-Gb/s
and FDDI-grade 62.5- m fiber. Higher values have also been
proposed that suggested better link performance [30]. We ex-

PEPELJUGOSKI et al.: MODELING AND SIMULATION OF NEXT-GENERATION MULTIMODE FIBER LINKS

1253

configuration. We have demonstrated that such an approach


can be used to specify sufficiently high-link bandwidth and sufficiently low ISI penalty with the combination of a transmitter
specification based on the radius at which the encircled flux
attains 86% of its maximum, and a fiber specification based on
spatial and temporal limitations on its DMD.
The method we describe is quite general and applicable to a
wide variety of possible systems with different bit rate, length,
and wavelength requirements. The model described in this work
was used by the TIA FO-2.2.1 group to perform extensive link
simulations that eventually led to the final specifications of the
next-generation fiber.
APPENDIX

(a)

We work in the context of the weak guidance approximation


to obtain the guided fields [31]. Taking along the fiber axis,
we write the electric field as
(A1)
The weak guidance approximation is very good for standard
multimode fiber (MMF). In this approximation, both components of the transverse field satisfy the same scalar wave equation
(A2)
is the transverse Laplacian,
is the free-space
where
is the refractive index profile of the fiber,
wavenumber,
represents either the or component of . After
and
a separation of variables into radial and angular parts, the
eigenfunctions of (A2) may be written as
(A3)

(b)
Fig. 11.

Same as in Fig. 10, but only single-mode lasers are included.

amined the impact of the change of the threshold in the radius of encircled flux for the three values: 75%, 86%, and 95%.
The results shown in Fig. 12 were filtered with the same DMD
mask as the previous simulations. These results suggest that the
75% threshold for the encircled flux is inadequate for 50- m
next-generation fibers, having a negative impact on the resolution of the screening test.

. The integers
and
with corresponding eigenvalues
label solutions to the radial and angular parts of (A2),
(). The polarwhile labels the angular dependence () or
or . As discussed in Section II-C,
ization is denoted by
we ignore polarization effects in this paper and therefore drop
to satisfy the normalthe index throughout. We define
ization condition

(A4)

IV. CONCLUSION
There are conflicting requirements between the transmitter
and fiber specifications in the design of multimode LAN
links: a given level of performance relaxation on the transmit
side requires tightening on the fiber side, and vice versa.
The link model described in this paper enables us to perform
accurate low-level simulations to study these tradeoffs. We also
described formalism for the calculation of the fiber transfer
function, which can be easily applied to any multimode link

and corresponding propagation conThe eigenfunctions


can easily be found numerically [32].
stants
ACKNOWLEDGMENT
The authors thank their colleagues in the TIA FO-2.2.1 group
for many useful conversations; S. Golowich thanks G. D. Boyd,
L. M.F. Chirovsky, G. Giaretta, and W. A. Reed for their discussion; and P. Pepeljugoski and A. Risteski also thank D. Kuchta,

1254

JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 21, NO. 5, MAY 2003

(a)

(b)
Fig. 12. Effect on the link failure rate of changing the encircled flux threshold: (a) 75% threshold and (b) 95% threshold. The negative impact of the 75% threshold
is evident, especially for devices with larger encircled-flux radii.

J. Schaub, T. Janevski, and Z. Hadzi-Velkov for their useful


conversations.

REFERENCES
[1] J. Schlager, M. Hackert, P. Pepeljugoski, and J. Gwinn, Measurements
for enhanced bandwidth performance over 62.5-m multimode fiber in
short-wavelength local area networks, J. Lightwave Technol., vol. 21,
pp. 12761285, May 2003.
[2] P. Pepeljugoski, M. Hackert, J. S. Abbott, S. Swanson, S. Golowich,
J. Ritger, P. Kolesar, C. Chen, and J. Schlager, Development of laser
system specification for optimized 50-m multimode fiber for multigigabit short-wavelength LANs, J. Lightwave Technol., vol. 21, pp.
12561275, May 2003.
[3] R. Michalzik, G. Giaretta, K. Goossen, K. Walker, and M. Nuss, 40
Gb/s coarse WDM data transmission with 825 nm wavelength VCSELs
over 310 m multimode fiber, in Proc. ECOC, Munich, Germany, Sept.
37, 2000, pp. 3334.

[4] R. Michalzik, G. Giaretta, A. J. Ritger, and Q. L. Williams, 10 Gb/s


VCSEL based data transmission over 1.6 km of new generation 850 nm
multimode fiber, in Proc. IEEE LEOS Annu. Meeting, LEOS 99, San
Francisco, CA, Nov. 1999, Postdeadline Paper PD1.6.
[5] Detail Specification for Laser Optimized, 50 m Core Diameter/125 m
Cladding Diameter Class Ia Graded Index Multimode Optical Fibers,
TIA-492AAAC, 2002.
[6] D. G. Cunningham and W. G. Lane, Gigabit Ethernet Networking. New York: Macmillan Technical Publishing, 1999.
[7] S. E. Golowich, P. F. Kolesar, A. J. Ritger, and P. Pepeljugoski, Modeling and simulations for 10 Gb multimode optical fiber link component
specifications, presented at the OFC 2001, Anaheim, CA, Mar. 2001,
Paper WDD-57.
[8] P. Pepeljugoski and S. E. Golowich, Measurements and simulations of
intersymbol interference penalty in new high speed 50 m multimode
fiber links operating at 10 Gb/s, presented at the OFC 2001, Anaheim,
CA, Mar. 2001, Paper WDD-40.
[9] B. Whitlock, P. Pepeljugoski, D. Kuchta, J. Crow, and S. Kang, Computer modeling and simulation of the optoelectronic technology consortium (OETC) optical bus, IEEE J. Select. Areas Commun., vol. 15, pp.
717730, May 1997.

PEPELJUGOSKI et al.: MODELING AND SIMULATION OF NEXT-GENERATION MULTIMODE FIBER LINKS

[10] P. Pepeljugoski and D. Kuchta, Design of optical communication data


links, IBM J. Res. Develop., vol. 47, no. 23, Mar.May 2003.
[11] H. Kogelnik and T. Li, Laser beams and resonators, Appl. Opt., pp.
15501567, 1966.
[12] I. A. White and S. C. Mettler, Modal analysis of loss and mode mixing
in multimode parabolic index splices, Bell Syst. Tech. J., vol. 62, no. 5,
pp. 11891207, MayJune 1983.
[13] D. Marcuse, Gaussian approximation of the fundamental modes of
graded-index fibers, J. Opt. Soc. Amer., vol. 68, no. 1, pp. 103109,
Jan. 1978.
[14] K. Petermann, Nonlinear distortions and noise in optical communication systems due to fiber connectors, IEEE J. Quantum Electron., vol.
QE-16, pp. 761770, July 1980.
[15] G. P. Agrawal and N. K. Dutta, Long Wavelength Semiconductor
Lasers. New York: Van Nostrand, 1986.
[16] R. S. Tucker and D. J. Pope, Circuit modeling of the effect of diffusion
on damping in a narrow-strip semiconductor laser, IEEE J. Quantum
Electron., vol. QE-19, pp. 11791183, July 1983.
[17] K. J. Knopp, D. H. Christensen, G. vander Rhodes, J. M. Pomeroy, B.
Goldberg, and M. S. Unlu, Spatio-spectral mapping of multimode vertical cavity surface emitting lasers, J. Lightwave Technol., vol. 17, pp.
14291435, Aug. 1999.
[18] G. Giaretta, M. Y. Li, G. S. Li, W. Yuen, and C. J. Chang-Hasnain,
A novel 4 8 single-mode independently addressable oxide-isolated
VCSEL array, IEEE Photon. Technol. Lett., vol. 9, pp. 11961198,
Sept. 1997.
[19] L. M. F. Chirovsky, G. D. Boyd, W. S. Hobson, and J. Lopata, VCSEL
beam waists optical spectra, IEEE Photon. Technol. Lett., vol. 13, pp.
547549, June 2001.
[20] K. I. Kitayama, S. Seikai, and N. Uchida, Impulse response prediction
based on experimental mode coupling coefficient in a 10-km-long
graded-index fiber, IEEE J. Quantum Electron., vol. QE-16, pp.
356362, Mar. 1980.
[21] S. Shaklan, Measurement of intermodal coupling in weakly multimode
fiber optics, Electron. Lett., vol. 26, no. 24, pp. 20222024, 1990.
[22] S. Kawakami and H. Tanji, Evolution of power distribution in gradedindex fibers, Electron. Lett., vol. 19, no. 3, pp. 100102, 1983.
[23] S. Berdague and P. Facq, Mode division multiplexing in optical fibers,
Appl. Opt., vol. 21, no. 11, pp. 19501955, 1982.
[24] P. Facq and P. Fournet, Observation of tubular modes in mulitmode
graded-index optical fibers, Electron. Lett., vol. 16, pp. 648650, 1980.
[25] C. M. Miller, S. C. Mettler, and I. A. White, Optical Fiber Splices and
Connectors: Theory and Methods. New York, Dekker, 1986.
[26] P. Petar, How is fiber bandwidth affected by lateral offset of the source,
in Proc. 5th Optical Fiber Measurement Conf., Nantes, France, Sept.
2224, 1999, pp. 147152.
[27] M. Wegmuller, S. Golowich, G. Giaretta, and M. Nuss, Evolution of
the beam diameter in a multimode fiber link through offset connectors,
IEEE Photon. Technol. Lett., vol. 13, pp. 574576, June 2001.
[28] Launched Power Distribution Measurement Procedure for GradedIndex Multimode Fiber Transmitters, TIA/EIA-455-203, 2001.
[29] Differential Mode Delay Measurement of Multimode Fiber in the Time
Domain, TIA/EIA-455-220, 2001.
[30] P. Pepeljugoski, J. Abbott, and J. Tatum, Effect of launch conditions on
power penalties in gigabit links using 62.5 m core fibers operating at
short wavelength, in Symp. Optical Fiber Measurements, Boulder, CO,
Sept. 1517, 1998, pp. 119122.
[31] A. W. Snyder and D. L. John, Optical Waveguide Theory. London,
U.K.: Chapman & Hall, 1983.
[32] T. A. Lenahan, Calculation of modes in an optical fiber using the finite
element method and EISPACK, Bell Syst. Tech. J., vol. 62, no. 9, pp.
26632694, 1983.

1255

Petar Pepeljugoski (S91M92SM03) received the B.S. degree from


St. Cyril and Methodius University, Skopje, Macedonia, in 1982, the M.S.
degree from the University of Belgrade, Belgrade, Yugoslavia, in 1986, and the
Ph.D. degree from the University of California, Berkeley, in 1993.
He joined IBM in 1994, where his research work included design, modeling,
prototyping and characterization of high-speed multimode-fiber local area network (LAN) links and parallel interconnects. His modeling tools were used by
the Telecommunication Industry Association (TIA) in the development of the
next-generation fiber. He was involved in the development of the 10 Gigabit
Ethernet Standard (IEEE 802.3ae) from its inception. He is currently a Research
Staff Member in the Communication Technology Department, IBM T. J. Watson
Research Center, Yorktown Heights, NY. He is an author or a coauthor of more
than 33 technical papers.
Dr. Pepeljugoski has been awarded recognition by the 10 Gigabit Ethernet
Alliance for his contributions. He was also awarded a Certificate of Appreciation by the TIA for his contribution to the Working Group on Modal Bandwidth of Multimode Fibers and the development of U.S. standards for fiber-optic
technology.

Steven E. Golowich (M02) received the A.M. and Ph.D. degrees in physics
from Harvard University, Cambridge, MA, in 1991 and 1994, respectively, and
the A.B. degree in physics and mathematics from Cornell University, Ithaca,
NY, in 1988.
From 1994 to 1996, he was an Instructor in the Physics Department, Princeton
University, Princeton, NJ, after which he joined Bell Laboratories, Lucent Technologies, Murray Hill, NJ, where he is currently a Member of Technical Staff
in the Fundamental Mathematics Research Department. His current research
interests include waves in inhomogeneous media with applications to optical
communications.

A. John Ritger, photograph and biography not available at the time of


publication.

Paul Kolesar (M97) received the B.S.E.E. degree from Pennsylvania State
University, University Park, in 1981, and the M.S.E.E. degree from Fairleigh
Dickinson University, Madison, NJ, in 1987.
From 1981 to 2001, he was employed with Bell Laboratories, where he designed and developed PBX circuit packs and fiber-optic multiplexers and, in
1988, assumed systems engineering responsibility for optical fiber structured
cabling. He is currently a Distinguished Member of the Technical Staff for the
Connectivity Solutions Division, Avaya Laboratories, Richardson, TX. He actively contributes to the development of industry standards within the Telecommunication Industry Association (TIA) TR42 on structured cabling, TIA FO4,
and IEC SC86A on fiber optics, as well as IEEE 802.3 on Ethernet. He holds
issued and pending patents on optical patch-panel design and high-speed multimode transmission and on conceptualized 850-nm laser-optimized multimode
fiber.

Aleksandar Risteski (A97M97) was born in Gostivar, Macedonia, in 1973.


He received the B.Sc. degree in electronics and telecommunications and the
M.Sc. degree in telecommunications from St. Cyril and Methodius University,
Skopje, Macedonia, in 1996 and 2000, respectively, and is currently working
toward the Ph.D. degree in telecommunications at the university.
In 2001, he was with the IBM T. J. Watson Research Center, Yorktown
Heights, NY, where he was a Practical Trainee. He is currently with the
Department of Electrical Engineering, St. Cyril and Methodius University,
where he is a Teaching and Research Assistant. His research interests are in
optical communications and coding theory.

Das könnte Ihnen auch gefallen