You are on page 1of 13

1766

IEEE TRANSACTIONS ON POWER SYSTEMS, VOL. 28, NO. 2, MAY 2013

Compliance Analysis of PMU Algorithms and


Devices for Wide-Area Stabilizing Control
of Large Power Systems
Innocent Kamwa, Fellow, IEEE, S. R. Samantaray, Senior Member, IEEE, and Geza Joos, Fellow, IEEE

AbstractFor the first time, IEEE Std. C37.118.1-2011 now


provides metrics for PMU dynamic performance in terms of
classes P and M filter designs. This paper attempts to determine
whether fulfilling these requirements makes the PMU inherently
well suited for stability control applications such as wide-area
power system stabilizers (PSSs). In this aim, we considered two
different frequency-adaptive approaches for class-P and -M
compliance to ensure operation over a wide frequency range.
The first is based on a finite-impulse response (FIR) with no
overshoot in either the phase or the amplitude step responses,
while the second is Kalman filter-based (EKF), which allows for
a more refined out-of-band interference rejection at the cost of
a phase step response with overshoot. These two approaches are
benchmarked against Hydro-Qubecs existing PSS requirements
and the conclusion is that the total vector error-based response
time is not indicative of the phase lag within the frequency band
of interest, nor of the 3-dB bandwidth under sinusoidal amplitude/frequency modulation phenomena, which are key criteria
when specifying PSS PMUs. Using simulated and field-recorded
network fault responses, we also show that a class-M PMU is
unsatisfactory for wide-area stabilizing control, unless its performance is improved during the fault period, which is not covered
by Std. C37.118.1-2011.
Index TermsAdaptive complex bandpass filtering, changing
harmonics, IEEE Std. C371181-2011, Kalman filtering, phasor
measurement unit (PMU), power system oscillations, synchrophasor, wide-area measurement systems (WAMS), wide-area
protection and control (WAPC).

I. INTRODUCTION

HASOR measurement unit (PMU) performance has been


the subject of very intense activity recently. IEEE standard C37.118.1 [1] provides metrics for comparing dynamic
performances of various PMU brands in terms of class-P and
class-M filter designs while Std. C37.118.2 defines the communication protocols more precisely. An IEEE guide for testing
and calibrating PMUs more comprehensively and with a greater

Manuscript received May 13, 2012; revised May 27, 2012, July 18, 2012, and
August 26, 2012; accepted September 24, 2012. Date of publication November
12, 2012; date of current version April 18, 2013. Paper no. TPWRS-004942012.
I. Kamwa is with the Hydro-Qubec/IREQ, Power System and Mathematics,
Varennes QC J3X 1S1, Canada (e-mail: kamwa.innocent@ireq.ca).
S. R. Samantaray is with the School of Electrical Sciences, Indian
Institute of Technology, Bhubaneswar, Orissa-751 013, India (e-mail:
sbh_samant@yahoo.co.in).
G. Joos is with the Department of Electrical and Computer Engineering, McGill University, Montreal, QC H3A 2A7, Canada (e-mail:
geza.joos@mcgill.ca).
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org.
Digital Object Identifier 10.1109/TPWRS.2012.2221168

level of uniformity is currently under ballot [2]. Over ten years


ago [3], authors were arguing that PMU-based wide-area stabilizing control of PSS could result in a number of stability benefits ranging from markedly improved damping to less severe
post-fault voltage dips and angle shifts[4], [5]. However, despite
substantial work at the design stage to improve the controller
tuning and coordination or at the laboratory and pilot stages to
simulate the whole concept in real time or prove it in open loop
[6], [7], no significant progress towards actual implementation
has been achieved so far. The common view is that system operators and reliability regulators are reluctant to move ahead creating the extra risk posed by a centralized continuous control
bugged by communications uncertainties and delays. However,
the gaps in PMU technology and its lack of maturity are simpler explanations for the reluctance of planning and project engineers to implement new approaches to improve system stability. The objective of this paper is to determine whether the
recent developments in PMU standards and related commercial
products have significantly improved the prospects of wide-area
PSS implementation.
A literature survey on this topic revealed that some authors [8][11] have developed extensive test procedures to
benchmark commercial PMUs against Western Electricity
Coordinating Council (WECC) and North American SynchroPhasor Initiative (NASPI) PMU filtering requirements.
Novel synchrophasor algorithms claiming to meet or exceed
similar requirements are proposed in [12][14]. Specific metrics which a PMU must meet to make it suitable for wide-area
damping control are summarized in [15] while sample filters
are proposed in [16] showing that the requirements can in
fact be met. These metrics, which are defined in terms of step
response and frequency response (gain and phase) characteristics, under sinusoidal amplitude and phase modulation, are
easy to understand and are the motivation behind the proposed
work. In fact, while Std. C37.118.1 proposes specific criteria
for measurement bandwidth and step response characterization,
these are built around total vector error (TVE), which is not a
natural concept for control systems designers. In addition, out
of the five commercial PMUs tested in [17], only two met the
C37.118.1 class-M step response specification while none of
them fulfilled the frequency ramp response specification. As
observed, some work remains to be done to link the metrics in
[1] with more conventional figures of [16] and, as a result, more
work is now required by manufacturers to fully meet C37.118.1
for class-M performance [18].
This paper will impact on both fronts. On one hand, adaptive algorithms able to help manufacturers meet C37.118.1s
dynamic requirements will be identified. On the other hand,

0885-8950/$31.00 2012 IEEE

KAMWA et al.: COMPLIANCE ANALYSIS OF PMU ALGORITHMS AND DEVICES

detailed performance analysis of these algorithms will help


bridging the gaps in C37.118.1 in order to enable wide-area
stabilizing control with full confidence. Some of these gaps
are related to the deficiency of steady-state TVE metrics when
accurate magnitude is as important as angle accuracy [19][21]
and to the non-applicability of TVE as a performance metric
during fault periods when the voltage vanishes [12], [22][24].
The methodology to be followed considers a standard PMU
algorithm complying with C37.118.1 static and dynamic performance requirements and an extensive assessment is carried
out to determine whether it complies with the utilities expectations of PMUs in existing PSS and special protection schemes
(SPS) applications. Although the standard provides a PMU simulation model which can be used for this purpose, it is not implemented in a Simulink environment and therefore has difficulty
adapting to other purposes. Many other PMU algorithms have
recently been reported but with no comprehensive evidence of
complying with the latest C37.118.1 directives and thus are not
readily usable in our project without a prior audit.
In this context, we found it more realistic to consider PMU
algorithms arising from the authors own work [12], [25]. Two
radically different approaches are considered for the sake of
generality: a finite-impulse response and a Kalman filter-based
design. A PMU of the first kind has a predefined group delay
and no overshoot after step input, while the second type of
PMU shows an overshoot in its phase step response. Both are
center-frequency adaptive and therefore result in an extremely
wide frequency range of stable operation, typically from 45 to
75 Hz. However, the Kalman filter-based PMU offers the additional advantage of an unusually high out-of-band rejection,
thanks to an embedded adaptive notch filter. The paper is illustrated with comparisons of the performance of three commercial
PMUs.
II. PMU FILTERING REQUIREMENTS AT HYDRO-QUBEC
PMUs have never been used for control at Hydro-Qubec
and, as a result, no official guideline exists for selecting them
for this purpose. However, based on the requirements of existing special protection systems (SPS) and power system stabilizer (PSS) devices and systems, it is possible to baseline the
metrics to be met by PMUs for future wide-area SPS and PSS
applications.
To provide some context for this discussion, let us first recall some characteristics of the Hydro-Qubec grid [26]. It is
islanded (i.e., large frequency excursions) and characterized by
long lines with a high level of series and shunt compensation.
The system also contains a significant number of flexible AC
transmission systems (FACTS) devices such as high voltage
DC (HVDC) and static var compensator (SVC). The geographical location makes the grid sensitive to geomagnetic storms
with a strong likelihood of transformer saturation. The design
of measurement systems for WACS/WAMS has been historically driven by this specific context [27]. As an example, here
are typical requirements from the Under Frequency Load Shedding project (19982001) [28]:
Fundamental frequency range
Hz, but the device should work in the range 40-70 Hz (associated to
hydro-generator requirements).

1767

Fig. 1. Summary of Hydro-Qubecs requirements for PMU and SPS algorithms design in series-compensated network. (a) Simplified spectrum of electromagnetic phenomena of concern (fundamental frequency at 60 Hz). (b) Required filters with prescribed response times in cycles of fundamental frequency.

Rate of change of frequency (ROCOF) range


Hz/s, but
the device should work up to
Hz/s. The algorithms
should remain stable up to 25 Hz/s.
Out-of-band filtering of series damped resonance in the
range 2532 Hz (2.5% on voltage, more than 10% on
currents).
Out-of-band filtering of parallel damped resonance in the
range 70100 Hz (up to 10% on voltage).
Out-of-band filtering of sub-synchronous parallel resonances in the range
Hz (5 to 25%).
Intermodulation rejection (due to coupling of transformer
saturation and sub-synchronous parallel resonance) in the
range
and
(5% on voltage).
Harmonics rejection (5% for 2nd to 5th and 2% from
6th to 10th) at fundamental but also at offset frequencies
(5466 Hz).
Instrument transformer transients rejection.
The electromagnetic spectrum associated to these requirements is illustrated in Fig. 1(a). Hydro-Qubec does not specify
the expected accuracy under the above conditions, leaving room
for innovation. However, in SPS applications, three filters are
required for voltage and frequency with various expected timeresponses. According to Fig. 1(b), Voltage magnitude results
from the filters called V1cy, V3Cy and V12cy, with one, three
and twelve fundamental cycles response-time respectively. Frequency is obtained similarly from F1cy, F3cy and F9cy filters,
with one, three and nine cycles response time, respectively. Rate
of change of voltage (ROCOV) and ROCOF variables are mandated correspondingly.
In fast control applications (PSS type), the bandwidth of
dc-demodulated signals such as angle, frequency, electric

1768

IEEE TRANSACTIONS ON POWER SYSTEMS, VOL. 28, NO. 2, MAY 2013

Fig. 2. Frequency response of the MB-PSS frequency PMU [29].


TABLE I
TYPICAL REQUIREMENTS FOR CONTROL WHICH APPLY
TO BOTH AMPLITUDE AND PHASE MEASUREMENTS

A 50-ms time-constant PMU usually achieves a good


damping performance of the inter-area modes in
simulation.
Typically, a 50-ms pure time delay reduces the inter-area
mode damping performance only marginally (needs
100 ms to have a detrimental effect).
In addition to Hydro-Qubecs requirements, Table I includes
BPA filtering requirements for comparison. Lack of uniformity
in the selected metrics makes it difficult to compare the two sets
of criteria side-by-side. But it is also clear that they are not expressed in terms of the TVE and require interpretation to reconcile them with the C37.118.1 dynamic-performance metrics. As
a further point of comparison, [32] dealt with small-signal oscillation monitoring in the frequency range
Hz] and came
out with more relaxed metrics (20 dB attenuation at Nyquist
frequency, maximum 10 phase lag and 3 dB gain ripple in the
passband).
III. CANDIDATE ALGORITHMS FOR WIDE-AREA CONTROL
For generality, we assume that the PMU input is a single
phase sinusoidal signal
(in pu) [37], with superimposed DC,
interharmonic and fundamental harmonic components:

(1)
where
harmonic order;
highest rank of the harmonic present in the signal;
dc value;
magnitude of the
phase of the

th harmonic;

th harmonic;

fundamental frequency in Hz;


power, etc., is restricted to 10 Hz and the phase lag should not
be larger than that of a 40-ms time constant, under sinusoidal
modulation between 0 Hz and 4 Hz, with a group delay less than
1.5 cycles of the fundamental. The possibility of adding a tunable notch for within-band interference rejection is suggested.
Another source of information for control system filtering
requirements is the MB-PSS, widely used at Hydro-Qubec,
which is also known as IEEE PSS4B [29]. Fig. 2 shows that
the 3-dB bandwidth of the frequency PMU embedded in the
MBPSS is 12 Hz, with an attenuation of only
dB at 35 Hz.
The corresponding phase lag is only 30 at 5 Hz, making it a
very fast PMU.
The PMU filtering specifications retained in this work for
the Hydro-Qubec system are given in Table I. They reflect a
compromise between the stringent attenuation of interference
needed by the defense plan SPS [26][28] and the fast response
needed by the MBPSS (which takes advantage of the filtering
effect of generator inertia and field winding to relax interference
attenuation requirements). In support of the Table I metrics, it is
important to mention some findings from simulation studies of
wide-area damping control on the large-scale system:

magnitude of the interharmonic component;


frequency of the interharmonic component;
phase of the interharmonic component;
magnitude of the fundamental component;
phase of the fundamental component;
fixed sampling period;
sampling time index.
For a three-phase system, the symmetrical components can
be derived from separated analysis of the single phase a, b and
c of voltage or current signals. This additive combination of the
single phase phasors result in a positive sequence component
which is less noisy than the individual phasors. To extract the
fundamental phasor, the frequency response of the filters must
have nulls at the harmonic frequencies that are expected to be
present in the signal and a unity gain at the fundamental frequency. If the frequency is not constant, then filter parameters
have to be adapted online during frequency estimation. To provide satisfactory measurement over a wide-frequency range, it
becomes necessary to track the system frequency and apply certain corrections on the measuring algorithms and input filters.

KAMWA et al.: COMPLIANCE ANALYSIS OF PMU ALGORITHMS AND DEVICES

1769

Fig. 4. Adaptive PMU algorithm based on complex bandpass FIR filtering.

Fig. 3. Frequency response of the anti-alias filters.

This principle will guide the design of the two PMU algorithms
presented in the sequel.
We assume that an anti-alias filter with a 3-dB cut-off frequency is applied to the signal prior to any processing inside
the algorithm. Two possible choices for this filter are shown in
Fig. 3. Setting the sampling rate at
points/cycle of the
nominal fundamental frequency results in a Nyquist frequency
of 12 points/cycle, or the 12th harmonic (i.e., 720 Hz for 60 Hz
system). At this frequency, the attenuation of the input signal is
22 dB and 26 dB for a 4th-order Butterworth and Tchebycheff
design respectively. Another important point is the phase shift
introduced by these filters at the fundamental frequency (
and
, respectively) which has to be compensated in the algorithm to achieve the TVE specified in [1].
A. FIR Approach
The idea depicted in Fig. 4 is to generate a complex anacorresponding to x(t), by passing the latter
lytic signal
through a K-taps FIR bandpass filter with complex coefficients
which are expressed analytically with respect to the filter center
frequency

Fig. 5. FIR bandpass filters for center frequency adaptive PMU algorithms (cf.
[11] and [37] for more information about the selected windows).

Simply stated, the frequency is obtained as the low-pass filtered derivative of the phasor angle provided by a robust recursive DFT. The FIR filter window is typically of a three-cycle
length (50 ms at 60 Hz). For this purpose, we prefer Kay or
Taylor windows which are theoretically justified as more efficient than other windows for frequency estimation [34]. To design the bandpass filter, let us consider the following filter bank
definition which comes from the exponentially modulated (EM)
filter bank theory [12], [31]:

(2)

with reference to Fig. 4, if the incoming frequency is known


online, we can re-tune the center frequency of the bandpass filter
accordingly [12]. Note that the filtering is adaptive but not the
sampling.
The time-varying phasor is made stationary by frequency
shifting based on a local reference phasor. Interestingly, the
group delay of the filter in number of samples is given by

The adaption frequency is derived through a separate frequency estimator using the demodulation method [33], [34].

where
and
represents
the impulse response coefficients of a linear-phase low-pass FIR
prototype filter. Furthermore, the number of filter cells is selected so that
is the length of the prototype filter. If properly
chosen, the scaling factor and the center frequency of the th
filter can be located at the fundamental frequency. Assuming a
samples per cycle of the fundamental
sampling rate of
(3)
To illustrate this band-pass filter design, a few typical protoare shown in Fig. 5. For baselining
type low-pass filters
purposes, the latter include two reference designs C37P and
C37M, obtained using the P and M class filters suggested in

1770

IEEE TRANSACTIONS ON POWER SYSTEMS, VOL. 28, NO. 2, MAY 2013

at some given frequency components embedded in the input


signal:
(6)
where

and
are the
in-phase
and in-quadrature
components, respectively. For a system with
sinusoidal components, the corresponding state vector at time
instant is
(7)

with

Fig. 6. Kalman filter-based adaptive PMU algorithm. The frequency and amplitude of the interharmonic (or subsynchronous) component are denoted by
and
, respectively.

C37.118.1, in cascade with a standard DFT phasor whose amplitude is frequency-compensated as in [12]. The Kay window
is defined as follows:

Assuming that at time instant , we have available measurements


of the incoming signal
, we can setup the following state-variable representation to recursively estimate the
state variable in (7):
(8a)

(4)
..

while the Taylor window is given by

(5)

(8b)

with
(8c)
(8d)

According to Fig. 5, the Taylor window-based 4-cycle filter


has a 10-Hz bandwidth and 40 dB of attenuation at 30 Hz with a
60-dB stopband above 40 Hz. It can therefore report the phasor
at a 60-Hz reporting rate without violating the C37.118.1 requirement for Nyquist frequency attenuation. Additional metrics of FIR filters are given in Table III.
B. Kalman Filter Approach
Fig. 6 illustrates the second approach for adaptive PMU algorithm design. It relies on the fact that a Kalman filter can be configured to perform like a discrete Fourier transform [30]. More
generally, it can be reformulated as a filter bank with center frequencies positioned at arbitrary locations, which makes it suitable for tracking a fundamental frequency with its harmonics,
together with specified inter-harmonic frequencies. The aim of
the Kalman filter, therefore, is to generate analytical signals

is known, the analytic signal


associated
Once the state
to the fundamental component of
can be derived using its
definition in (6)(7).
Note that in the above equations, the number of sinusoidal
components plus the DC component is equal to
.
The scalar variable
is a zero-mean Gaussian white-noise
process with variance
and the initial state is an -mean
Gaussian random variable with covariance
. is the variance of the additive error of the measuring chain and typically,
takes the following form:

The causal estimation of the state


, using measurements
so that
is minimum, where
represents the estimate of
derived from

KAMWA et al.: COMPLIANCE ANALYSIS OF PMU ALGORITHMS AND DEVICES

1771

gain, the Kalman filter can be considered as an efficient design


tool for obtaining the frequency response template in Fig. 7. The
impact of wrong signal modeling assumptions on the filtered
phasor estimate
is easily predicted by looking at the
gain response. For instance, the maximum sidelobes is 10 dB in
both configurations while the DC rejection is much better than
that of a standard [34]. Finally, as illustrated in Fig. 6, knowing
the values of
and
in real time using separate frequency
estimators as in [12], it is possible to update the state matrix of
the filter and therefore re-tune the center lobes of the whole filter
bank on-line, without any need to also update the steady-state
Kalman filter gain K. A similar principle was recently justified and applied successfully in [36] using a bank of tunable
resonators.
IV. SALIENT FEATURES OF CANDIDATE ALGORITHMS
Fig. 7. Kalman filter gain and frequency response gain (a) without the interharmonic component and (b) with a 10-Hz interharmonic interference.

, is an optimal filtering problem whose solution is known to be the Kalman filter [30], [35]. Commonly, the
Kalman estimation process includes two steps, namely the prediction and correction phases.
Let us assume that the state estimate
is known with an
error variance
. The measured value
is then used to
update the state at instant . The additive correction of the a
priori estimated state at
is proportional to the difference
between the a priori output at instant defined as
and
the measurement
:

(9)
is the Kalman gain which guarantees the minimal
where
variance of the error
. Also, at each step
the variance
of the prediction error is calculated:
(10)
This variance matrix is then used to calculate the Kalman gain
in the next step of the recursive calculation (correction phase):
(11)
For practical purposes, the steady-state value of the Kalman
gain
a
vector, can be derived
entirely offline using simulations of recursive (9)(11) as explained in [30]. This way, the filter becomes a fixed-coefficient
state observer with predetermined stability characteristics. Also,
by assigning appropriate values to the frequency components in
(8), interharmonics and harmonics can all be tackled together
transparently, which renders the Kalman filter a more flexible
tool than the discrete Fourier transform (DFT).
For illustration purposes, the gain of the two Kalman filter
configurations used in this paper are shown in Fig. 7, along with
the resulting frequency response gain computed from the state
space (9). Solution (a) assumes a signal model with a fundamental term,
harmonics and a DC component. Solution (b) adds one interharmonic at 10 Hz-frequency to the spectral content of (a). It appears that when applied with a constant

The two families of algorithms discussed above have been


implemented in Matlab/Simulink. In each family, we have a tentative P and M class solution. The class-P FIR solution relies on
the 2.2-cycle filter while the class-M solution is based on the
4-cycle Taylor filter, both illustrated in Fig. 5. To approximate
class-P and -M behavior with the Kalman filter approach, we
use a Kalman filter with and without an adaptive notch add-on
at the interference frequency, respectively.
To demonstrate the effectiveness of the algorithms, we subjected the four PMU algorithms to basically all tests specified in
C37.118.1, following a methodology similar to [8] and [12]. We
used Simulink SimPower-Systems three-phase programmable
sources to generate high sampling rate data, including harmonics and interferences as defined in C37.118.1. These test
signals are generated at four times
the algorithm sampling rate. The 4th-order anti-alias Butterworth filter in Fig. 3 is then applied to the synthesized signals,
followed by a decimation by four, in order to set the algorithm
input sampling rate at
points per cycle of the nominal
frequency (i.e., 1440 Hz.). Therefore, the filtering is adaptive
but not the sampling.
The comparative step responses are illustrated in Fig. 8, with
some selected metrics shown in Table II. The group delay of the
algorithms, in the last line of this Table, is an important determinant of the PMU latency. As expected, the C37P and C37M
designs have predictable delays of 1 and 2 cycles (about half
the processing window size). It should be noted that the group
delays of Kalman filter-based PMUs are not integer numbers of
samples but are quite short compared to the baseline. Note that
the TVE is computed by lagging the theoretical phasor (sampled at 1440 Hz) by an amount of samples equal to the group
delay of the studied PMU shown in Table II, before taking the
difference between the theoretical and measured phasor values.
The next three figures report steady-state behavior at fundamental (Fig. 9), interference (Fig. 11), and harmonic (Fig. 10)
frequencies, respectively. Fig. 9 was obtained by simulating the
algorithms behavior when subjected to unit magnitude steady
state voltage of varying fundamental frequency ranging from 0
(
Hz offset) to 120 Hz (
offset). It can be concluded that all proposed algorithms meet the 1% TVE requirement of C37.118.1 over the range of
offset frequency,
assuming a 60 Hz fundamental frequency. But the FIR-M filter

1772

IEEE TRANSACTIONS ON POWER SYSTEMS, VOL. 28, NO. 2, MAY 2013

Fig. 8. PMU algorithm responses to 10% magnitude and 10 phase step response at fundamental frequency.

Fig. 9. PMU algorithm steady-state responses at offset frequency.

TABLE II
STEP RESPONSE CHARACTERISTICS (TIME IN SECONDS)

Fig. 10. Steady-state TVE under harmonics with varying frequency offset.

is quite special since it is entirely flat from dc to harmonics, with


a TVE below 1% throughout the whole range.
Results in Fig. 10 were obtained by subjecting the algorithms
to a steady-state unit-magnitude voltage with a single 10% harmonic added. Varying the fundamental frequency (
to 75
Hz) and harmonic rank (1 to 11) in two overlapping loops, the
TVE was computed for all combinations of these two parameters. Fig. 10 then shows the TVE as level of color (from blue
for low TVE to red for high TVE) with respect to all possible
pairs of offset frequency and harmonic rank. On the top of each
subplot which corresponds to one specific algorithm, the maximum TVE over the full range of studied parameters is shown.
These numbers confirm that the two proposed M-Class filters
are thoroughly immune against changing harmonics, i.e., harmonics at non-fundamental frequency. Their performance exceeds by a widemargin of 0.1% the C37.118.1 specification,
which mandates 1% TVE at fundamental frequency only. Interestingly, Fig. 10 shows that the EKF-P PMU (maximum TVE

of 0.11153%) is better than FIR-M PMU (maximum TVE of


0.11159%) for rejecting off-nominal harmonics.
Fig. 11 was obtained by simulating the algorithm TVE and
noise rejection gain when subjected to a steady-state unit magnitude voltage at fundamental frequency, with a 10% sinusoidal
additive interference superimposed. The interfering frequency
is varied in 1-Hz step from 1 Hz to 240 Hz. Results shown in
Fig. 11 allow to verify that the two M-filters do in fact meet
the 40-dB attenuation requirement of C37.118.1 at Nyquist frequency. In addition, the 1% TVE threshold is met, even well
above the Nyquist frequency of 30 Hz.
In fact, for the Kalman filter-based PMU, the narrow-band
magnitude attenuation is 60 dB for both P and M class designs,
which is actually the performance required for a high-sensitivity
and high-impact SPS such as that in [27]. Table II includes two
TVE-based response-time criteria set at 1% and 2%, respectively. It appears that the performance of the Kalman filter is
satisfactory when the criterion is 2%. By contrast, it is three to
four times slower than required by Std. C37.118.1 when the criterion is 1%. In general, we can conclude that the FIR-M and

KAMWA et al.: COMPLIANCE ANALYSIS OF PMU ALGORITHMS AND DEVICES

Fig. 11. PMU algorithm steady-state responses to sinusoidal interference At


); maxTVE(
).
30 or 90 Hz: maxTVE(

EKF-M algorithms meet both the BPA and Hydro-Qubecs step


response requirements.

1773

Fig. 12. PMU frequency responses to 1% amplitude sinusoidal modulation.


Gain and phase responses computed by injecting a single frequency excitation
at a time, and taking the ratio of the fast-Fourier transforms of the algorithm
input and output signals.

V. FREQUENCY RESPONSES TO SMALL-AMPLITUDE


SINUSOIDAL MODULATIONS
When performing control-tuning studies in power system
stability software, the engineer usually develops an equivalent model for the PMU. This model should be compatible
with a positive-sequence representation of the network in the
frequency range of 05 Hz, which is the range covered by
stability software. In addition, the equivalent should capture
the small-signal behavior of the PMU within 05 Hz in terms
of magnitude and phase lag, while having a 3-dB frequency
band, which is representative of the actual device.
The step responses of Fig. 8 can be used to develop these
dynamic models. However, an alternative approach, which
also allows for checking the frequency domain specification
in Table I, determines the frequency responses to amplitude
and phase modulation signals by simulation of the PMU algorithms. These tests are performed separately using sinusoids
with 0.01 pu and 0.01 rad
peak magnitude respectively;
the results are presented in Figs. 12 and 13. For easy interpretation, the phase and amplitude of the frequency responses
at critical frequency are shown in Table III, together with the
TVE computed with a 10% (and 0.1 rad) sinusoidal modulation
as required by C37.118.1 [1]. For comparison purposes, the
same metrics are included for a reference design obtained by
cascading a frequency-compensated DFT phasor with the P
and M class filters suggested in C37.118.1. The gain and phase
of the 40-ms time constant are also shown below Table III.
The main observation from Figs. 12 and 13 and Table III is
that, while FIR-based solutions have the same phase (degree)
and magnitude response, the EKF-based solutions have a flat
amplitude but a resonant phase response. It also appears that the
phase lag of the FIR-M filter closely follows that of a 40-ms time
constant in the frequency range 04 Hz. By contrast, EKF-M
does the same up to 3 Hz only for both amplitude and phase
modulations. But interestingly, at 1 Hz, the EKF-M has a 4
advantage over a pure 40-ms time constant. This increases to

Fig. 13. PMU frequency responses to 0.01 rad (or 0.57) sinusoidal phase modulation. Gain and Phase computed for each frequency at a time.

10 for the EKF-P filter, which can be very useful in certain


circumstances, such as for a power plant PSS. Finally, the TVE
metrics of phase modulation at 5 Hz shows that all algorithms
can meet the bandwidth requirements set in C37.118.1 for class
M accuracy.
VI. FAULT RESPONSE TESTS
To test the algorithms under more realistic network conditions, we picked a test case used in [27] during the design and
commission of an SPS [38] which is now a major component of
the defense plan [26]. The results are shown in Fig. 14.
The challenge here is that during a frequency ramp, the
voltage (which needs to be accurately measured in [27]) is subjected to a so-called intermediation phenomenon, manifesting
itself through fictitious and sustained voltage oscillations at
6 Hz. The EKF based algorithm with interference tracking and
notching is the only solution to remove these perturbations

1774

IEEE TRANSACTIONS ON POWER SYSTEMS, VOL. 28, NO. 2, MAY 2013

TABLE III
FREQUENCY RESPONSE CHARACTERISTICS

(PHASE IN DEGREES)

Fig. 15. Frequency responses of three commercial PMUs.

Fig. 14. Analysis of a parametric case used for designing the SPSs in [38]. Postfault response with 4-Hz/s frequency ramp, 15% sub-synchronous resonance
frequency.
with voltage magnitude intermediation at

without increasing the 95% time response of the filter up to say


200 ms. Naturally, these spurious oscillations also corrupt the
ROCOF signal, even though the frequency dynamic range does
not allow visualizing them on the frequency ramp plot. Again,
despite its good interference attenuation shown in Fig. 11, the
FIR-M still come short of polishing the ROCOF to the level
suitable for crisp decision making.
To further assess whether a C37.118.1-compliant PMU
could meet utilities requirements for wide-area control, we
tested three commercial PMUs on a laboratory setup [8], [9].

The testing was not aimed at checking their compliance with


C37.118.1 in terms of TVE-based metrics, since this would
have required a well calibrated GPS-based source. Instead,
our setup is essentially a relay-testing environment with the
goal of playing back synthesized parametric waveforms or
COMTRADE files of network events recorded in the field or
on our Hypersim real-time transient-network simulator.
The frequency responses of these PMUs are given in Fig. 15.
To facilitate comparisons, we selected for each PMU brand, the
settings closest to the M class performance level. But only one
brand (C) offered a specific M class selection. According to
Fig. 15, all three PMUs have a good frequency range of flat
magnitude around the fundamental and a good attenuation of
out-of-band interference. Since the output rate is 60 Hz, the
Nyquist frequency rejection
dB) is very close to the requirements of both the utilities and C37.118.1 in this regard. Notice however that the frequency response of PMU-A seems leftshifted. After further verification, we confirmed that this is the
correct behavior of the device. The center of its frequency range
seems to be around 50 Hz, whatever the nominal fundamental
frequency selected. Notice that, the corresponding out-of-band
interference frequency response is not available at the moment
to confirm this hypothesis.
Next, a simple startup test consisting in suddenly applying
the nominal voltage at off-nominal frequency was performed
on the three devices and their results compared with the EKF-M
filter of previous sections. Fig. 16 shows some interesting facts
about this test (obtained by playing back the same waveform
with the EKF-M Simulink model). Firstly, we observed that the
two PMUs start at zero, suggesting that their output frequency
is 60 Hz when there is no voltage while the EKF-M algorithm

KAMWA et al.: COMPLIANCE ANALYSIS OF PMU ALGORITHMS AND DEVICES

1775

Fig. 17. PMU responses to network event: fault at Nemiscau (James Bay) with
the PMU near Montreal.

Fig. 16. (a) Sinusoidal step: from zero to nominal voltage at 63 Hz fundamental
frequency. Top: initial time frame; Bottom: full time frame . Frequency response
time at 95%: A (N/A), B (75 ms), C (90 ms). (b) Sinusoidal step: from zero
to nominal voltage at 63 Hz fundamental frequency. Top: initial time frame;
Bottom: full time frame . Voltage response time at 95%: A (60 ms), B (55 ms),
C (70 ms).

and the brand C PMU produce 0 Hz output under the same conditions. Secondly, the PMU-C has a large frequency overshoot

with no amplitude overshoot whereas the PMU-B has no overshoot but the output is delayed by 0.2 s.
The behavior of PMU-A is even more delayed, as the 63-Hz
target is still way out of reach 1 s after startup (actually, a zoom
of the figure shows that the PMU-A posts a 60.5-Hz frequency
at 1.5 s but increases slowly towards the target). In other words,
except for the PMU-A, the amplitude and frequency response
times for a sinusoidal step at off-nominal frequency are close
to utilities specifications. However, since the TVE is not available, we cannot draw any conclusion about the C37.118.1 specified response times for the M-filter.
The full time-frame plots confirm that the PMU A has a significant error throughout the recorded period but the error is
slowly decreasing. Using other tests at off-nominal frequency,
we found that in fact, PMU A is just very slow in responding
when the initial frequency far from the nominal frequency. In
fact, the frequency tracking performance is very poor and well
below the 1 Hz/s required in Std. C37.118.1. This is a major
pitfall which has been communicated to the vendor who agreed
with the pattern shown in Fig. 16. The issue will be corrected in
a future release of this device.
Fig. 17 illustrates the PMU responses to a network fault event.
The only noticeable fact on the voltage response is the discrepancy in time response between the various solutions. PMU-C is
characterized by an initial undershoot of a non-minimal phase
response type while the step response overshoot of PMU-B is
evident after fault clearing. Regarding the frequency response,
the PMU-B and C exhibit spurious behavior during the fault period while PMU-A remains on idle for a long time. Interestingly,

1776

IEEE TRANSACTIONS ON POWER SYSTEMS, VOL. 28, NO. 2, MAY 2013

in the context of wide-area control is related to phase shift and


frequency tracking, during and shortly after network events. In
fact, when the voltage vanishes for any reason, the phasor is
not well defined and each PMU addresses this ambiguity differently. In the example shown, PMU-C does nothing, PMU-A
is too cautious, PMU-B offers a strategy between A and C,
while EKF-M algorithm is the best trade-off overall. Alternative
schemes to deal with this issue, not addressed in C37.118.1, can
be found in [17] and [22][24].
It has been argued that the requirements in [27] are too stringent, and counterintuitive. However, as a matter of fact, several
SPS has been implemented in the Hydro-Qubecs network
using non-synchrophasor based measurements units which met
these requirements [28], [38]. One design case used for assessing these existing SPS has been selected to benchmark the
commercial PMUs A-B-C with respect to their ability to meet
the requirements in Section II without further development. The
corresponding waveform and positive sequence voltage and
frequency measurements are shown in Fig. 18. The waveform
on the top plot was played back on the Hypersim simulator
to obtain PMU A-B-C results. It was also played back on the
EKF-M Simulink model in Matlab to compare the proposed
algorithm with the three state-of-art PMUs from major vendors.
The results in Fig. 18 show quite clearly that the three existing
PMUs are not suitable to deal with some of the phenomena in
Hydro-Qubecs SPS requirements of Section II: they display
a 5% amplitude error and between 12 and 50-mHz frequency
error in the post-fault period. By contrast, the amplitude and frequency errors ascribed to the EKF-M PMU algorithm are only
0.2% and 1-mHz, respectively.
VII. CONCLUSIONS

Fig. 18. PMU responses to a sub-synchronous response parametric case known


as design case No f0014. Steady state voltage error: A-B-C: 5%; EKF-M:
0.2%. Steady state Frequency error: A: 50 mHz; B: 12 mHz; C: 50 mHz;
EKF-M: 1 mHz.

these observations are also in line with Fig. 16. However, none
of the behavior shown by PMU-A, -B or -C is appropriate from
the wide-area control point-of-view.
In each case, the control could overreact and produce more
harm than good if it wrongly assumes that the PMU frequency
signal is correct during the fault. These laboratory tests confirm that, under dynamic conditions, most commercial PMUs
tend to perform as specified in C37.118.1 regarding the amplitude response, although excess overshoot could still be an
issue for some PMUs (B). Regarding frequency tracking, [17]
showed that none of the five PMUs tested was able to meet all
C37.118.1 dynamic-response requirements. The present test results demonstrate that one important pitfall of commercial PMU

This paper presented two families of PMU algorithms able


to meet the dynamic and static accuracy requirements of Std.
C37.118.1. Further, it highlighted possible gaps and difficulties
that may happen when reconciling the TVE-based criteria of
C37.118.1 with the utilities filtering requirements, which are
usually expressed in terms of frequency response to modulation
in amplitude, phase and frequency signals or control systembased step-response information. It appears, for instance, that a
1% TVE stabilization criterion may be too stringent compared
to a 2% stabilization response time specified by most utilities
for control system PMUs.
In addition, the laboratory testing of three PMU brands
showed that Class M PMUs from different vendors will likely
provide inconsistent measurements of dynamic activity [8].
The three PMUs A, B and C tested on the same step excitations
resulted in quite different dynamic behavior, even though their
settings were selected to be as close as possible to Class M specification according to each vendor. It was also found that the
PMU performance for frequency tracking during faults varied
a lot from brand to brand. Unfortunately, the TVE criterion
does not characterize the above inconstancies in engineering
terms, and it cannot be used to compensate them during data
alignment. A possible improvement in this regard could be to
characterize the amplitude and phase errors patterns separately
[19], [20] and further mandate PMU vendors to provide the
group delay of their PMU to facilitate data alignment at Phasor

KAMWA et al.: COMPLIANCE ANALYSIS OF PMU ALGORITHMS AND DEVICES

1777

Data Concentrator (PDC) level. A last complication with using


TVE concept is that direct TVE measurements requires GPS
synchronization, which might be unavailable to an otherwise
sufficient test facility such as a typical power utility protection
laboratory or even the world class IREQs transient network
simulator. In this context, it appears that TVE compliance
assessment will be performed by large specialized instrumentation laboratories, not by the PMU end users.
Overall, it was found that the FIR bandpass filtering PMU
was a perfect match to the current C37.118.1, as it exceeded all
of its P and M accuracy class requirements, with the advantage
of flat step responses, and linear phase. However, the adaptive
Kalman filter was evidently the best compromise in narrow band
applications with changing harmonics and torsional or sub-synchronous phenomena. It was able to address effectively transient
response to faults while meeting or exceeding other C37.118.1
requirements.

[10] J. F. Hauer, W. A. Mittelstadt, K. E. Martin, J. W. Burns, and H.


Lee, Integrated dynamic information for the Western power system:
WAMS analysis in 2005, in Power System Stability and Control
volume of the Electric Power Engineering Handbook, L. L. Grigsby,
Ed., 2nd ed. Boca Raton, FL: CRC Press, 2007, ch. 14.
[11] PMU System Testing and Calibration Guide. NASPI Report of the Performance & Standards Task Team (PSTT), Dec. 30, 2007. [Online].
Available: https://www.naspi.org/File.aspx?fileID=555.
[12] I. Kamwa, A. K. Pradhan, and G. Joos, Adaptive phasor and frequency-tracking schemes for wide-area protection and control, IEEE
Trans. Power Del., vol. 26, no. 2, pp. 744753, Apr. 2011.
[13] M. Karimi-Ghartemani, Boon-TeckOoi, and A. Bakhshai, Application of enhanced phase-locked loop system to the computation of synchrophasors, IEEE Trans. Power Del., vol. 26, no. 1, pp. 2232, Jan.
2011.
[14] R. K. Mai, L. Fu, Z. Y. Dong, B. Kirby, and Z. Bo, An adaptive dynamic phasor estimator considering DC offset for PMU applications,
IEEE Trans. Power Del., vol. 26, no. 3, pp. 17441754, Jul. 2011.
[15] D. Kosterev, Dynamic performance requirements for phasor measurement units, in Proc. NAPSI Meeting, Feb. 2010. [Online]. Available:
https://www.naspi.org/File.aspx?fileID=138.
[16] D. Trudnowski, PMU dynamic requirements for small-signal measurements, in Proc. NASPI Meeting, Chattanooga, TN, Oct. 2009.
[Online]. Available: https://www.naspi.org/File.aspx?fileID=562.
[17] J. Ren, Synchrophasor measurement using substation intelligent electronic devices: algorithms and test methodology Ph.D. dissertation,
Elect. Eng. Dept., Texas A&M Univ., College Station, 2011.
[18] A. Goldstein, PMU Model C37.118.1 NASPI, Jun. 2011. [Online].
Available: https://www.naspi.org/File.aspx?fileID=419.
[19] G. Y. Yang, K. E. Martin, and J. stergaard, Investigation of PMU
performance under TVE criterion, in Proc. 2010 5th Int. Conf. Critical
Infrastructure (CRIS), Sep. 2022, 2010, pp. 17.
[20] M. Lixia, C. Muscas, and S. Sulis, On the accuracy specifications
of phasor measurement units, in Proc. 2010 IEEE Instrum. Meas.
Technol. Conf. (I2MTC), May 36, 2010, pp. 16.
[21] P. Castello, C. Muscas, and P. A. Pegoraro, Performance comparison
of algorithms for synchrophasors measurements under dynamic conditions, in Proc. IEEE Int. Workshop Applied Measurements for Power
Syst. (AMPS), Sep. 2830, 2011, pp. 2530.
[22] J. Warichet, T. Sezi, and J. C. Maun, Considerations about synchrophasors measurements in dynamic system conditions, Int. J.
Elect. Power Energy Syst., pp. 113, 2009.
[23] A. G. Phadke and B. Kasztenny, Synchronized phasor and frequency
measurement under transient conditions, IEEE Trans. Power Del.,
vol. 24, no. 1, pp. 8995, Jan. 2009.
[24] J. Ren and M. Kezunovic, Real time power system frequency and
phasor estimation using recursive wavelet transform, IEEE Trans.
Power Del., vol. 26, no. 3, pp. 13921402, Jul. 2011.
[25] I. Kamwa, R. Grondin, and D. McNabb, Changing harmonics in
stressed power transmission systemsApplication to Hydro-Quebecs
network, IEEE Trans. Power Del., vol. 11, no. 4, pp. 20202027,
Oct. 1996.
[26] G. Trudel, J. P. Gingras, and J. R. Pierre, Designing a reliable power
system: Hydro-Quebecs integrated approach, Proc. IEEE, vol. 93, no.
5, pp. 907917, May 2005.
[27] J. Lambert, D. McNabb, and A. G. Phadke, Accurate voltage phasor
measurement in a series-compensated network, IEEE Trans. Power
Del., vol. 9, no. 1, pp. 501509, Jan. 1994.
[28] P. Cote, S.-P. Cote, and M. Lacroix, Programmable load shedding-systems-Hydro-Quebecs experience, in Proc. 2001 IEEE PES
Summer Meeting, Vancouver, BC, Canada, Jul. 1519, 2001, vol. 2,
pp. 818823.
[29] I. Kamwa, R. Grondin, and G. Trudel, IEEE PSS2B versus PSS4B:
The limits of performance of modern power system stabilizers, IEEE
Trans. Power Syst., vol. 20, no. 2, pp. 903915, May 2005.
[30] R. R. Bitmead, A. C. Tsoi, and P. J. Parker, A Kalman filtering approach to short-time Fourier analysis, IEEE Trans. Acoust., Speech,
Signal Process., vol. 34, no. 6, pp. 14931501, Dec. 1986.
[31] R. G. Lyons, Understanding Digital Signal processing, 3rd ed. Englewood Cliffs, NJ: Prentice Hall, 2011.
[32] R. Lira, D. H. Wilson, K. Hay, T. Cumming, and D. Cole, Testing synchrophasor data quality for applicability to transmission security and
optimisation tools, in Proc. 17th Power Syst. Comput. Conf., Stockholm, Sweden, Aug. 2011, pp. 2226. [Online]. Available: http://www.
pscc-central.org/uploads/tx_ethpublications/fp414_01.pdf.
[33] M. Paolone, A. Borghetti, and C. A. Nucci, A Synchrophasor
Estimation Algorithm for the Monitoring of Active Distribution Networks in Steady State and Transient Conditions, ibid, Jul. 15, 2012.
[Online]. Available: http://www.pscc-central.org/uploads/tx_ethpublications/fp221.pdf.

ACKNOWLEDGMENT
The first author would like to thank all his Hydro-Qubecs
colleagues who have contributed to this work directly and
indirectly through data and knowledge sharing. Specials
thanks are extended to C. Lafond, M. Perron, C. Cyr, and
J. Blandfrom IREQ, who performed the laboratory tests in a
rather short period of time. The leadership role of D. McNabb
for preparing numerous test cases and initiating early research projects aiming at assessing the practical feasibility of
Hydro-Qubecs requirements for SPS measurement units is
gratefully acknowledged.
REFERENCES
[1] IEEE Standard for Synchrophasor Measurements for Power Systems,
IEEE Std. C37.118.1-2011, Dec. 28, 2011.
[2] Project Authorization Request (PAR) for a New IEEE Standard: PAR
No. PC37.242, Guide for Synchronization, Calibration, Testing, and
Installation of Phasor Measurement Units (PMU) for Power System
Protection and Control, Sep. 30, 2010. [Online]. Available: https://development.standards.ieee.org/get-file/PC37.242.pdf?t=10380700003.
[3] I. Kamwa, G. Trudel, and L. Grin-Lajoie, Multi-loop power system
stabilizers using wide-area synchronous phasor measurements, in
Proc. 1998 Amer. Control Conf., Jun. 2126, 1998, pp. 29632967.
[4] P. Denys, C. Counan, L. Hossenlop, and C. Holwek, Measurement
of voltage phase for the french future defence plan against losses of
synchronism, IEEE Trans. Power Del., vol. 7, no. 1, pp. 6269, Jan.
1992.
[5] I. Kamwa, J. Bland, G. Trudel, R. Grondin, C. Lafond, and D.
McNabb, Wide-area monitoring and control at Hydro-Qubec:
past, present and future, in Proc. 2006 IEEE/PES General Meeting,
Montreal, QC, Canada, Jun. 1822, 2006.
[6] C. W. Taylor, C. Erickson, K. E. Martin, R. E. Wilson, and V.
Venkatasubramanian, WACS-wide-area stability and voltage control
system: R&D and on-line demonstration, Proc. IEEE, vol. 93, no. 5,
pp. 892906, May 2005.
[7] L. Peng, W. Xiaochen, L. Chao, S. Jinghai, H. Jiong, H. Jingbo, Z.
Yong XuAidong, and X. Aidong, Implementation of CSGs wide-area
damping control system: Overview and experience, in Proc. IEEE/
PES, Power Syst. Conf. & Expo., PSCE 09, Mar. 1518, 2009, pp.
19.
[8] Z. Huang, T. Faris, K. Martin, J. Hauer, C. Bonebrake, and J. Shaw,
Laboratory Performance Evaluation of SEL 421 Phasor Measurement
Unit, PNNL Report-16852, Pacific Northwest National Laboratory,
Dec. 2007. [Online]. Available: http://www.pnl.gov/main/publications/external/technical_reports/PNNL-16852.pdf.
[9] F. Steinhauser and Y. Ping, Testing the dynamic response of
phasor measurement units, in Proc. Int. Conf. Elect. Eng., pp. 15.
[Online]. Available: http://www.icee-con.org/papers/2009/pdf/2.
02_I9CP0546_E.Pdf.

1778

[34] I. Kamwa, M. Leclerc, and D. McNabb, Performance of demodulation-based frequency measurement algorithms used in typical pmus,
IEEE Trans. Power Del., vol. 19, no. 2, pp. 505514, Apr. 2004.
[35] D. Alves and R. Coelho, An adaptive algorithm for real-time
multi-tone estimation and frequency tracking of non-stationary signals, IEEE Trans. Nucl. Sci., vol. 58, no. 4, pp. 15821587, Aug.
2011.
[36] M. D. Kuljevi, Simultaneous frequency and harmonic magnitude
estimation using decoupled modules and multirate sampling, IEEE
Trans. Instrum. Meas., vol. 59, no. 4, pp. 954962, Apr. 2011.
[37] S. Shinnaka, A novel fast-tracking D-estimation method for
single-phase signals, IEEE Trans. Power Electron., vol. 26, no. 4, pp.
10811088, Apr. 2011.
[38] S. Bernard, G. Trudel, and G. Scott, A 735 kV shunt reactors automatic switching system for Hydro-Quebec network, IEEE Trans.
Power Syst., vol. 11, no. 4, pp. 20242030, Nov. 1998.

Innocent Kamwa (S83M88SM98F05)


received the Ph.D. degree from Laval University,
Qubec City, QC, Canada, in 1988.
He then joined Hydro-Qubecs research institute
(IREQ), where he is currently project manager for
power grid control and automation. He is also the
chief scientist for Hydro-Qubecs smart grid. He is a
P. Eng. and Adjunct Professor of Power Systems Engineering at McGill University and Laval University.
Dr. Kamwa serves on many IEEE/PES technical
committees as member and officer, including the
Fellow Evaluation, Electric Machinery, and Power System Stability committees. He is an Editor of the IEEE TRANSACTIONS ON POWER SYSTEMS and
Co-Editor-in-Chief of IET Generation, Transmission and Distribution. He was
awarded the IEEE PES Prize Paper Award in 1998, 2003, and 2009.

IEEE TRANSACTIONS ON POWER SYSTEMS, VOL. 28, NO. 2, MAY 2013

S. R. Samantaray (M08SM10) received the


B.Tech. degree in electrical engineering from UCE
Burla, India, in 1999 and the Ph.D. degree in power
system engineering from the Department of Electronics and Communication Engineering, National
Institute of Technology, Rourkela, India, in 2007.
He holds the position of Assistant Professor in
the School of Electrical Sciences, Indian Institute of
Technology Bhubaneswar, India. He visited the Department of Electrical and Computer Engineering,
McGill University, Montral, QC, Canada, as a
Post-Doctoral Research Fellow and Visiting Professor. His major research
interests include intelligent protection for transmission systems (including
FACTs) and microgrid protection with distributed generation and dynamic
security assessment in large power networks.
Dr. Samantaray is the recipient of the 2007 Orissa Bigyan Academy Young
Scientists Award, the 2008 Indian National Academy of Engineering Best Ph.D.
Thesis Award, the 2009 Institute of Engineers (India) Young Engineers Award,
the 2010 Samanta Chandra Sekhar Award, and the 2012 IEEE PES Technical
Committee Prize Paper Award. He is an Editor of IET, Generation, Transmission
& Distribution and Electric Power Components and Systems.

Geza Joos (M82SM89F06) received the


M.Eng. and Ph.D. degrees from McGill University,
Montral, QC, Canada.
He has been a Professor with McGill University
since 2001 and holds a Canada Research Chair in
power electronics applied to power systems. He is
involved in fundamental and applied research related
to the application of high-power electronics to power
conversion (including distributed generation and
wind energy) and to power systems. He was previously with ABB, the cole de technologiesuprieure,
and Concordia University. He is involved on a regular basis in consulting
activities in power electronics and power systems.
Dr. Joos is active in a number of IEEE Industry Applications Society committees as well as IEEE Power Engineering Society working groups and CIGRE
working groups. He is a Fellow of the Canadian Academy of Engineering and
of the Engineering Institute of Canada.

You might also like