Tải bản đầy đủ (.pdf) (13 trang)

[Khiem] improvement of matrix converter drive reliability by online fault detection and a fault tolerant switching strategy IEEE trans

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (2.17 MB, 13 trang )

244 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 59, NO. 1, JANUARY 2012
Improvement of Matrix Converter Drive
Reliability by Online Fault Detection and a
Fault-Tolerant Switching Strategy
Khiem Nguyen-Duy, Student Member, IEEE, Tian-Hua Liu, Senior Member, IEEE ,
Der-Fa Chen, Member, IEEE, and John Y. Hung, Fellow, IEEE
Abstract—The matrix converter system is becoming a very
promising candidate to replace the conventional two-stage ac/dc/ac
converter, but system reliability remains an open issue. The most
common reliability problem is that a bidirectional switch has an
open-switch fault during operation. In this paper, a matrix con-
verter driving a speed-controlled permanent-magnet synchronous
motor is examined under a single open-switch fault. First, a new
fault-detection method is proposed using only the motor currents.
Second, a novel fault-tolerant switching strategy is presented. By
treating the matrix converter as a two-stage rectifier/inverter, ex-
isting modulation techniques for the inverter stage can be reused,
whereas the rectifier stage is modified by control to counteract
the fault. However, the proposed techniques require no additional
hardware devices or circuit modifications to the matrix converter.
Experimental results show that the proposed method can maintain
the motor speed with a maximum ripple of 2%—a fivefold im-
provement over the uncompensated system. The proposed method
therefore offers a very economical and effective solution for the
matrix converter fault tolerance problem.
Index Terms—Fault diagnosis, fault tolerance, matrix converter,
pulsewidth modulation, reliability.
NOMENCLATURE
V
A
,V


B
,V
C
Matrix converter input voltages.
V
a
,V
b
,V
c
Matrix converter output voltages.
S
xY
Switch connecting output x to input Y .
V
dc
Virtual dc-link voltage.
i
a
,i
b
,i
c
Output phase currents.
i

a
,i

b

,i

c
Phase reference currents.
Δi
a
Phase current error i

a
− i
a
.
The modeling discussion makes references to a conventional
two-stage converter, consisting of a three-phase ac/dc bridge
rectifier and a three-phase dc/ac bridge inverter, but with no
physical dc-link storage. Rectifier switches are denoted by C,
Manuscript received September 24, 2010; revised January 22, 2011 and
March 8, 2011; accepted April 6, 2011. Date of publication May 23, 2011; date
of current version October 4, 2011. This work was supported by the National
Science Council of Taiwan under Grants NSC-99-2221-E-011-151-MY3 and
NSC 98-2811-E-011-022.
K. Nguyen-Duy and T H. Liu are with National Taiwan University of
Science and Technology, Taipei 106, Taiwan (e-mail: ;
).
D F. Chen is with the National Changhua University of Education,
Changhua 500, Taiwan (e-mail: ).
J. Y. Hung is with Auburn University, Auburn, AL 36849 USA (e-mail:
).
Color versions of one or more of the figures in this paper are available online
at .

Digital Object Identifier 10.1109/TIE.2011.2151818
and inverter switches are labeled T . Upper and lower switches
are distinguished by the s ubscripts U and L, respectively.
For example, C
UA
denotes a rectifier upper switch connected
to the source phase A, while T
cL
denotes an inverter lower
switch connected to the output phase c. Note that a virtual
dc-link voltage is denoted by plain text symbol, e.g., V
AB
,to
distinguish from matrix converter voltages.
I. B
AC KGROU N D
R
ESEARCH on matrix converter has received great at-
tention since the inception of this kind of direct ac/ac
converter [1]–[10]. Current commutation issues and modulation
techniques have been vastly explored in the research literature,
but the topic of r eliability improvement is relatively new and
has received more isolated attention. Recovering from matrix
converter system faults presents two challenges or problems
that are reviewed next.
A. Fault-Detection Methods
The first problem is how to identify when the fault occurs
and determine the exact location of the faulty device (isola-
tion). Faults that can appear within the converter include an
open-switch fault (including failure of a device driver circuit

[11]) and a shorted-switch fault. An open-circuit fault might
also occur in an output winding. For the fault detection and
protection against faults of an insulated-gate bipolar transistor,
there are two main approaches. The first approach is based on
gate-drive detection working directly with the gate-drive circuit
and usually relies on the measurement of the collector voltage
[12], the gate voltage [13], or the induced voltage across the
emitter stray inductance [14]. This kind of approach can be
used for both short- and open-circuit fault detections. Because
of its fast decision capability, gate-drive detection is particu-
larly applicable for the protection against short-circuit fault.
For protection of a matrix converter, however, this approach
may become expensive and complicated since the number of
switches is much larger than in a conventional voltage source
inverter (VSI).
The second fault-detection approach is algorithm-based de-
tection, which relies on software computation. This approach
may be applied to open-circuit fault detection so long as the
system can endure the fault for the time period required to
execute the algorithm. Abnormal behaviors of state variables
0278-0046/$26.00 © 2011 IEEE
NGUYEN-DUY et al.: IMPROVEMENT OF MATRIX CONVERTER DRIVE RELIABILITY 245
like the current and voltage in the presence of a fault are
analyzed to derive an algorithm to detect the fault location
[15]–[19].
Most algorithm-based fault-detection methods have been
devoted to VSI-based applications, and much fewer studies
have been undertaken for matrix converters. Thus, further study
should be carried out to discuss the detecting of the fault in the
matrix converter. In addition, remedial action cannot be applied

until the exact location of the faulty switch among a matrix of
switches is determined.
For safety concerns, it is recommended that, when a short-
circuit fault occurs, an autoprotecting circuit should be trig-
gered to immediately stop the system operation [20]. Another
alternative is to employ a fast fuse in series with each of the
bidirectional switches from the output phase points. Thus far
in the literature, study of the open-switch fault far outweighs
the case of the shorted-switch fault [17], [19], [21]–[26]. It
can be stated that online fault-detection ability is an essential
prerequisite for developing a solution to counteract the fault.
B. Methods for Fault-Tolerant Operation
The second challenge to improving matrix converter system
reliability is to develop a postfault solution that maintains
system operation as close as possible to normal. The solution
for fault-tolerant operation may require modification of the con-
verter topology or introduction of a different converter control
strategy [21], [22], [24].
Researchers have proposed different solutions to create either
a fault-tolerant topology or revised switching strategy in the
matrix converter system. For example, Kwak and Toliyat [25],
[26] proposed a topology with one to three additional triacs to
connect the source neutral point to the load neutral point or to
the output phase of a load. The remedial modulation technique
relied on the indirect modulation principle [27]–[29]. Under
that principle, a remedial modulation strategy in the rectifier
stage creates two equal virtual dc-link voltages with respect
to the source neutral. By restructuring the converter topology
after the fault occurrence, the remaining healthy (3 × 2) matrix
configuration, together with the remedial modulation algorithm,

can deal with the loss of an output leg of the converter, the
opened fault of a switch, or the loss of one output phase
winding.
C. Open Challenges
The aforementioned methods appear to suffer several limita-
tions. For example, the work of Kwak et al. might have greater
impact if it took into account the implementation on a practical
system such as an adjustable speed drive system instead of a
passive R–L prototype load. Furthermore, the source neutral
point and/or the load neutral point must be accessible; in
general, these points may not be available. Finally, the need for
additional triacs raises physical packaging issue and may not be
a cost-effective solution.
In terms of fault detection in matrix converter systems, recent
published papers have used the concept of output error voltage
[26], [30]–[33]. The principle of these methods is based on the
observation that, when there i s an open-switch fault, the output
error voltage signal (i.e., the difference between the output
voltage of a switch and the associated input voltage) becomes
abnormally high during the time when the faulty switch is
triggered on. Again, approaches of this kind carry another
drawback. To know the value of three-phase input and output
voltages, it is mandatory to insert voltage-measurement circuits
or to introduce soft-voltage observers, which may, in general,
lead to the need of extra space, extra cost, or computation
complexity. In [26], the author proposed both a fault-detection
method and a postfault modulation strategy in an attempt to
completely solve the fault tolerance issue. However, continuous
operation from the instant of fault detection to the onset of
fault-tolerant control was not ensured. Specifically, the fault-

detection method was verified only by computer simulation,
whereas the postfault modulation algorithm was carried out
with a separate experiment.
D. Scope of This Paper
This paper seeks to address the aforementioned problems. A
conventional 3 × 3 matrix converter driving a speed-controlled
permanent-magnet synchronous motor (PMSM) is investigated
both theoretically and experimentally. First, a system descrip-
tion is briefly presented (Section II). Next, the authors propose
an algorithmic fault-detection method for the open-switch fault,
employing only the two current sensors that are the standard
hardware elements in a conventional drive system (Section I II).
This proposed method can quickly identify the exact location of
the open-switch fault. Finally, the authors propose a postfault
(or fault-tolerant) switching strategy that relies only on the
remaining eight switches (Section IV). No extra devices are
required t o maintain balanced and nearly sinusoidal three-phase
output currents. The experimental results show the proposed
method has satisfactory performance under the operating con-
dition of a single open-switch fault (Section V). Compared
to the known prior works [25], [26], [30]–[33], the proposed
method has several key advantages. First, no extra hardware
devices are needed. Second, the entire fault detection and
postfault processing are executed by a digital signal proces-
sor (DSP) within a very short time duration. To the authors’
best knowledge, the ideas demonstrated here are the first of
their kind.
II. S
YSTEM DESCRIPTION
A. Configuration of the Drive System

The configuration of the proposed drive system is shown in
Fig. 1. The drive system hardware consists of a wye-connected
three-phase PMSM with mechanical load, a matrix converter
(10-kW rating), a complex programmable logic device (CPLD)
chip, and a DSP system. The parameters of the PMSM are
listed in Table I. An encoder (2500 pulses per revolution) is
mounted on the motor shaft. Two Hall-effect current sensors
are used to detect the stator currents of the motor. The zero-
voltage crossing detection circuit provides information of the
input voltage source angle for modulation. The DSP reads the
246 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 59, NO. 1, JANUARY 2012
Fig. 1. Configuration of the tested motor drive system.
TABLE I
P
ARAMETERS OF THE PMSM
zero-voltage crossing detection signal, the shaft angle, and
the stator currents of the motor and computes the control
algorithms. Then, the DSP outputs the nine signals (Bus) to
trigger the solid-state power switches that produce the switch-
ing patterns of the matrix converter. Finally, depending on
the sign of the motor current, a switch commutation strategy
is executed by the CPLD to achieve safe commutation. The
hardware prototype employs the minimal number of elements
necessary to ensure proper operation of a drive system.
The matrix converter is enclosed by the dashed box in Fig. 1.
It consists of nine bidirectional or ac switches, where each
of them possesses the capability of bidirectional energy flow.
Each output phase of a 3 × 3 matrix converter is connected to
three input phases by three ac switches. When one ac switch
of an output phase is an opened fault, there are two remaining

ac switches that can be used to connect to the two related
input phases. The two remaining healthy output phases can still
receive any of the possible voltage from the three input phases.
This scenario is different from an open-circuit fault introduced
by the load. In the latter case, all three ac switches connected
to the faulty output phase become useless, reforming a 3 × 2
hardware matrix with floating load neutral point. Previously
discussed approaches usually employ some supporting redun-
dant devices to either perform the following: 1) connect the
opened phase to some redundant leg or 2) connect the load
neutral point to some rational reference point [22], [24], [26].
The issue of an open-circuit fault in the load, however, is not
addressed in this paper, which focuses on a fault within the
converter.
B. Modulation Strategy
Modulation strategy is typically selected based on the appli-
cation and the underlying hardware configuration. The modu-
lation mechanism in a drive system works at the closest level
to the switching fashion of the switches. Furthermore, differ-
ent modulation strategies create different abnormal behaviors
when the hardware suffers a device failure. A detailed study
comparing the abnormal behavior (in the presence of fault)
against the normal behavior is a common approach to deriving
either a fault-detection method or a postfault solution. Thus, it
is necessary to choose a nominal modulation strategy for the
development of a fault-tolerant solution.
In this paper, the current-controlled drive is considered,
leaving the voltage-controlled drive for subsequent study. The
current-controlled drive is very common for ac servo applica-
tions where a high-bandwidth current-control loop is often de-

sirable so that a reference torque can be followed. The indirect
modulation method [27]–[29] will be carried out throughout
this paper, where later, it reveals a potential of determining the
fault (Section III), as well as a remedial modulation solution
(Section IV). A detailed description of the modulation strategy
is presented next.
The matrix converter can be modeled as a two-stage rectifier/
inverter converter with no dc-link storage, as shown in Fig. 2.
The relationship of the switching patterns between matrix
converter and the conventional two-stage converter can be
expressed as
S
xY
=

k=U,L
T
xk
C
kY
= T
xU
C
UY
+ T
xL
C
LY
(1)
NGUYEN-DUY et al.: IMPROVEMENT OF MATRIX CONVERTER DRIVE RELIABILITY 247

Fig. 2. Equivalent two-stage converter topology.
Fig. 3. Virtual dc-link voltages for a matrix converter.
where x denotes the output phase (a, b, c), Y denotes the input
phase (A, B, C), and U and L denote the upper and lower
switches, respectively.
By using the two-stage converter concept, the rectifier- and
inverter-stage conversion signals are separately generated. In
the rectifier stage, the virtual dc-link voltage is constructed by
the available line-to-line source voltage. There are six com-
binations of nonzero line-to-line voltage at any time instant,
but just the three positives among these shall be adopted to
be a virtual dc-link voltage to ensure a consistent logic for the
inverter stage. These three positive voltages can be classified
according to their relative value as the high, middle, and low
voltages. Fig. 3 shows the three-type virtual dc-link voltage for
a matrix converter. To aid in analysis, one voltage source period
has been subdivided and labeled with 12 different sectors. The
simplest way of choosing the virtual dc link is to choose the
highest line-to-line voltage in each time instant or sector. This
action is equivalent to the operation of an uncontrolled three-
phase diode-bridge rectifier, where it provides the inverter with
maximum output power.
For example, if the source angle is lying within either
sector 1 or 2, then the virtual dc-link voltage V
CB
is chosen
(study Fig. 3). By closing switches C
UC
and C
LB

in the
rectifier stage of Fig. 2, the input phase voltage V
C
is connected
to the positive dc rail, and V
B
is connected to the negative dc
rail. The merit of the selection of the highest available dc-link
voltage is to take the full advantage of the input voltage capa-
bility, particularly in the presence of a disturbance such as the
induced voltage [back electromotive force (EMF)] of a servo
drive system. If an adequately high switching frequency is used,
the ripple of the dc-link voltage can be compensated [34]. A
current-controlled drive system was successfully run under that
type of modulation, as reported in [34].
Based on the principle of indirect modulation and the min-
imal hardware system shown in Fig. 1, the block diagram
of the drive control system used in this paper is shown in
Fig. 4. The block diagram consists of two feedback loops:
speed- and current-control loops. The speed-control loop uses
a proportional–integral (PI) controller. The current-control loop
is controlled in the abc frame. The matrix converter is treated
as a t wo-stage ac/dc/ac converter or rectifier/inverter. In the
inverter stage, basic hysteresis current controllers are used.
Based on the deviation between the motor current and the
reference current, the phase current can be independently con-
trolled by manipulating either the upper or lower switch on
each leg. Using the source angle signal, a reasonable virtual
dc-link voltage in the rectifier stage is then chosen. In the
normal operation, the highest virtual dc-link voltage is chosen

in each sector, as described before. Finally, the switching states
in the rectifier and inverter stages are combined to yield the
switching logic for triggering the nine bidirectional switches.
In summary, the indirect modulation strategy is carried
through this work. In the next section, the fault-detection
method is described.
III. P
ROPOSED FAULT-DETECTION METHOD
On the basis of the modulation strategy described in
Section II, the effect of a converter open-switch fault on system
performance is first discussed by way of an example scenario
and then generalized. Following that, a fault-detection method
is proposed in this section. In this present case, an insight into
the behavior of output currents is helpful.
A. Effect of Open-Switch Fault on Converter Behavior
To better understand the effect of an open-switch fault, an
example will be discussed, and the results are then generalized.
Assume the following example conditions.
1) Open-switch fault occurs in switch S
aC
, which connects
the input phase C to the output phase a.
2) Source angle is presently at sector 4 but is entering the
range of sectors 5–8.
3) A 100-μs sampling interval is used in the current-control
loop.
Notice that, during sectors 5–8 (Fig. 3), the virtual dc links V
AC
and V
BC

are chosen in the rectifier stage, as these are the highest
available line-to-line voltages.
First, consider the case where the phase-a reference current
i

a
is negative. When Δi
a
= i

a
− i
a
is negative, the current-
control algorithm attempts to reduce i
a
during the next sam-
pling interval by connecting phase a to the negative dc rail
through lower switch T
aL
(see Fig. 2). Normally, the available
voltage in the negative dc rail is V
C
. Therefore, the controller
attempts to connect input phase C to the output phase a by
triggering switch S
aC
. Unfortunately, switch S
aC
is suffering

an open-switch fault, so normal control action leads to an
open circuit of phase a. The current i
a
discharges through
248 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 59, NO. 1, JANUARY 2012
Fig. 4. Block diagram of proposed control system.
the clamp circuit and approaches zero as time increases. The
aforementioned condition is in effect so long as i

a
is negative
and the source angle is lying within sectors 5–8. Output phase
a can remain open for a time duration equal to the overlap
between i

a
remaining negative and the time of four consecutive
sectors 5–8. In this analysis, these are called “affected sectors.”
Similar behavior occurs when i

a
is positive and V
C
is used in
the positive virtual dc rail. In the latter case, the affected sectors
are four adjacent sectors 11, 12, 1, and 2.
The aforementioned analysis can be applied to all devices in
the matrix converter. Let x denote the output phase a, b,orc.
Then, the affected sectors after fault occurrence are described
as follows.

1) For S
xC
, the affected sectors are 5–8 for i
x
< 0 and 11,
12, 1, and 2 for i
x
> 0.
2) For S
xB
, the affected sectors are 1–4 for i
x
< 0 and
(7–10) for i
x
> 0.
3) For S
xA
, the affected sectors are 9–12 for i
x
< 0 and 3–6
for i
x
> 0.
The fault condition depends on the time duration of the
output current fundamental period, denoted as T
fund
,aswell
as the duration of the affected sectors, denoted as T
aff

.The
reference current i

a
should be a sine wave that changes sign
every half of the fundamental period. The affected sector dura-
tion T
aff
is fixed to the input source period; for a 60-Hz source,
T
aff
=5.6 ms. The fault time duration is approximated by
T
fault
= min{T
fund
/2,T
aff
}. (2)
Consider an example analysis for an eight-pole PMSM being
regulated at 200 r/min. The output current fundamental period
is T
fund
=75ms. Thus, the fault time is approximately equal
to 5.6 ms, which is equivalent to 56 consecutive sampling
intervals. The smallest fault duration occurs at the highest oper-
ating speed. A maximum operating s peed of 2000-r/min speed
produces a 7.5-ms current fundamental period. According to
(2), the fault duration is 3.75 ms, equivalent to approximately
38 consecutive sampling intervals.

However, the time duration that the output current in the
faulty phase remains zero is slightly different from the fault
time. Let T
zero
denote the zero-current duration, which can be
estimated by subtracting from T
fault
the time for the current to
discharge from its initial value to zero through the clamp circuit
T
zero
= T
fault
− discharge time. (3)
The phenomenon of the affected output current remaining
zero in the faulty condition due to the open-circuit switch
can be easily distinguished from zero-current crossings of
normal operation. Due to the high-bandwidth current-control
loop operating at high switching frequency, the normal zero-
current crossing occurs in a relatively short time, less than a
few sampling intervals in the worst case.
The faulty condition for the opening of an ac switch disap-
pears when the s ource angle lies outside the affected sectors;
in this paper, such sectors are called “nonaffected sectors.”
Specifically, in the case of a single open fault occurring in
switch S
aC
, S
bC
,orS

cC
,thenonaffected sectors are sectors 3,
4, 9, and 10. Within these s ectors, all of the three-phase output
currents should be able to track their respective reference
current signals. Define the current error in the faulty output
phase as the difference between the current and its reference
signal, i.e., Δi
x
= i

x
− i
x
. Note that |Δi
x
| should be smaller
when operating under the nonaffected sectors and larger when
operating under the affected sectors. This feature will be used
later for fault diagnosis.
Based on these observations, a new fault-detection method
that relies only on the motor current signal is proposed and
explained in the next sections. The mechanism goes through
two stages.
B. First Stage of Fault Detection
In the first stage of fault detection, for each sampling interval,
the DSP determines whether the output current i
a
, i
b
,ori

c
is
equal to zero. In implementation practice, the DSP determines
whether the output phase current is inside a small dead band ,
because a practical system often contains noise in the current
NGUYEN-DUY et al.: IMPROVEMENT OF MATRIX CONVERTER DRIVE RELIABILITY 249
TABLE II
R
ECOMMENDED THRESHOLD VALUES
sensing circuit. If this condition is satisfied, the DSP continues
to check if the condition is continuously satisfied for some dura-
tion, i.e., a kind of threshold time. The threshold value is chosen
based on the zero-current duration ( 3). The smallest thresh-
old associated with the maximum operating speed (3.75 ms
or 38 sampling intervals in the example case) should be ade-
quate f or all smaller operating speeds. In addition, the discharge
time of the affected current should be also considered from the
knowledge of the load time constant.
The nature of the load being considered in this work is
given in Table I. The estimated discharge time is less than
seven sampling intervals under full load. With this knowledge,
a threshold of 30 sampling intervals is adopted for the whole
speed range in this paper. Based on the zero-current duration
(3), the recommended threshold values for different operating
speeds of an eight-pole 2.1-kW PMSM are given in Table II.
The threshold values for motors having different pole numbers
or operating at different nominal speeds can be computed using
the same principle.
If one of the three-phase output currents satisfies the con-
dition mentioned earlier, then the conclusion is that the open-

fault switch should be one among the three switches associated
with the faulty output. Thus, at the end of the first stage test,
the number of possible answers has been reduced from nine to
three. In the second stage of fault detection, the exact location
of the open switch is chosen from these three possible answers.
C. Second Stage of Fault Detection
In the second stage of the fault detection, the exact location
of the open switch among the three possible answers is iden-
tified. The summation of absolute current error |Δi
x
| in the
faulty phase during one voltage source period is computed and
categorized into three groups.
1) Group I includes sectors 7, 8, 1, and 2. These sectors are
the nonaffected sectors for the three switches S
aA
, S
bA
,
and S
cA
connected to input phase A.
2) Group II includes sectors 5, 6, 11, and 12. These sectors
are the nonaffected sectors for the three switches S
aB
,
S
bB
, and S
cB

connected to input phase B.
3) Group III includes sectors 3, 4, 9, and 10. These sectors
are the nonaffected sectors for the three switches S
aC
,
S
bC
, and S
cC
connected to input phase C.
Next, the DSP determines the smallest value among the three
groups. If group I has the smallest value, then the switch that
connects input phase A to the faulty output phase should be the
bad switch. Likewise, if group II has the smallest value, then the
switch that connects input phase B to the faulty output phase
should be the bad switch. Lastly, if group III has the smallest
TABLE III
L
OCATION OF A SINGLE OPEN-FAU LT AC SWITCH
value, then the switch that connects to input phase C should be
the faulty one.
The exact faulty switch among the three possible answers
derived from the first stage test has been identified in the second
stage test. The results are summarized in Table III. Knowing the
exact location of the open fault, further action in the control can
take place to counteract the fault.
The fault-detection algorithm executed within every sam-
pling interval is shown in Fig. 5. At the beginning of the
algorithm, the DSP reads the three-phase output currents f rom
the analog-to-digital converter and computes the three-phase

current commands. Next, hysteresis controllers are used to
provide switching states for the inverter stage. Then, the DSP
reads a fault-indicator J_faultvariable that is updated from the
previous interrupt. This variable is assigned a value at the end
of the fault-detection mechanism that it indicates the location
of the single open-fault ac switch. Therefore, this variable
can receive nine different values associated with the nine ac
switches. If a fault is detected, the DSP branches to one of
the nine subroutines of the proposed fault-tolerant switching.
Otherwise, the DSP continues to check the first condition for
three-phase output current (being inside a small dead band). If
this first stage test is satisfied, the DSP performs the second
stage test. If the second stage test fails, then normal rectifier
switching states are chosen. Otherwise, the exact location of the
bad switch is detected, where the DSP assigns a suitable value
for the fault indicator J_fault and branches to one of the nine
remedial switching strategies.
IV. F
AULT-TOLERANT MODULATION STRATEGY
Under normal conditions, the virtual dc-link voltage is ad-
justed as a function of motor speed error and output current
errors to obtain satisfactory performance of the PMSM drive
system. When an open-switch fault occurs, the high and middle
voltages at some sectors cannot be used because the faulty
switch is unable to turn on.
For example, if switch S
aC
experiences an open-switch fault,
then any attempt by the controller to trigger this switch is not
feasible. Input phase C cannot feed the output phase a. There-

fore, phase a must be modulated between input phases A and B.
Closer examination of the indirect modulation principle leads to
a solution to achieve such modulation. The two-stage ac/dc/ac
model (Fig. 2) is again used to help explain the analysis; the
case of open-switch fault at S
aC
is described next.
The switching function of the open faulted switch is de-
scribed by S
aC
= T
aU
C
UC
+ T
aL
C
LC
=0. Since the fault is
an open switch, the function value must equal zero, regardless
of switching states in either the inverter stage or rectifier stage
(study Fig. 2). So as to not to open the output phase a,
250 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 59, NO. 1, JANUARY 2012
Fig. 5. Current-control algorithm with online fault-detection mechanism.
switches T
aU
and C
UC
must not be simultaneously turned on.
This is also true for the pair T

aL
and. C
LC
.
Suppose that the current error is negative (Δi
a
=(i

a

i
a
) < 0). Within the inverter stage, the lower switch T
aL
should
be turned on so as to reduce the current i
a
. From the aforemen-
tioned switching constraints, rectifier switch C
LC
must not be
turned on. Either C
LA
or C
LB
may be chosen, which offers the
negative virtual dc rail voltages V
A
or V
B

. The positive dc rail
has no similar constraints, so any of C
UA
, C
UB
, and C
UC
can
be turned on.
Similar logic applies for the case of positive current error
(Δi
a
> 0): Inverter switch T
aU
must be turned on to increase
i
a
, but rectifier switch C
UC
must not be turned on. The alter-
native is to turn on either C
UA
or C
UB
for the positive virtual
dc rail. The negative virtual dc rail can freely adopt any one of
C
LA
, C
LB

,orC
LC
.
Summarized in Table IV are the combinations of feasible
positive and negative rail voltages that yield proper virtual
dc-link voltage for the open-switch fault of switch S
aC
.The
12 columns of the table describe rail voltages for the
12 sectors. The upper half of the table covers the case of
negative current error, while the lower half covers the positive
current error case. The selections of maximum virtual dc-link
voltage are highlighted in colors red, violet, and blue. Corre-
sponding waveforms are shown in Fig. 6.
In the implementation practice, feasible switching states in
the inverter stage are first determined, followed by selection
of switching states in the rectifier stage. Current error Δi
a
of
the affected output phase gives information on how to choose a
proper virtual dc-link voltage. Synthesis of the switching states
from the two-stage ac/dc/ac converter yields the triggering
signals for the matrix converter.
The aforementioned process is directly extended to switches
in the other output phases. For instance, when S
bA
, S
bB
,or
S

bC
has an open-switch fault, attention should be paid to
phase-b current deviation Δi
b
=(i

b
− i
b
) to choose a proper
dc-link voltage for the rectifier stage. Detailed switching rules
for all single open-switch faults are given in Table V, where x
denotes the output phase a, b,orc. The shaded cells are for the
example described earlier (fault in phase a).
V. E
XPERIMENTAL RESULTS
The experimental studies examined motor currents, motor
torque, steady-state speed error, current harmonics, load distur-
bance rejection, speed step response, and speed trajectory track-
ing. Performances were examined with and without remedial
actions.
A. Motor Currents, Torque, and Speed Error
The system operation from normal condition to a single
open-switch fault without any remedial action i s shown in
Fig. 7. Prior to the fault, the system is in steady state, with
the speed regulated at 200 r/min under a 2-N · m load. At time
t =0.1 s, switch S
aC
is opened by the program to produce the
phenomenon of an open-switch fault. The output currents i

a
,
i
b
, and i
c
are shown in Fig. 7(a)–(c), respectively. It can be
NGUYEN-DUY et al.: IMPROVEMENT OF MATRIX CONVERTER DRIVE RELIABILITY 251
TABLE IV
F
EASIBLE RAIL VOLTAGES FOR OPEN-SWITCH FAULT S
aC
Fig. 6. Virtual dc-link waveforms using the maximum available voltages for
the case of open-switch faulted S
aC
.
observed that the output current i
a
is distorted and equals zero
repetitively after time t =0.1 s. The major reason is that, in
some duration, output phase a was not connected to any input.
Fig. 7(d) shows the switching logic of T
aU
, which is used in the
two-stage converter model inverter stage to regulate i
a
.During
the fault, T
aU
stays at either zero or one to connect phase a to

the input phase C, leading to the open-circuit fault in phase a.
This phenomenon agrees with the earlier analysis in Section III.
Whenever phase a is opened, it introduces an unbalanced
two-phase machine with floating load neutral point. As a result,
i
b
and i
c
are strongly affected by the fault in phase a [see
Fig. 7(b) and (c)]. In the presence of the machine back EMF,
the system becomes more unbalanced and nonlinear. Thus, the
electromagnetic torque contains repetitive dips and ripples, as
shown in Fig. 7(f). Consequently, the speed ripple is quite high
due to the bad torque [Fig. 7(e)].
To evaluate the speed-control performance of the system,
three error measures are examined: maximum error E
max
,
mean error E
me
, and root-mean-square (rms) error E
rms
.Inthe
case of Fig. 7(e), E
me
=4.86 r/min, and E
rms
=8.08 r/min.
The largest absolute speed error is E
max

=20r/min, which is
10% of the operating speed. Fig. 7(g) shows a partial zoom of
i
a
for 75 ms after the fault, which is equal to one fundamental
period of the current. The fault time is about 5.6 ms, which
agrees with the analysis in Section III. The real zero-current
duration is just slightly smaller than 5.6 ms, since the discharge
time of current from its initial value to zero is very small for the
motor under test.
The system operation using the proposed fault detection and
fault-tolerant switching strategy is shown in Fig. 8. After the
fault occurs at 0.1 s, the system determines that current i
a
satisfies the first checking condition (i
a
=0for 30 consecutive
sampling intervals). The proposed two-stage fault-detection
mechanism is performed to identify the exact location of the
open switch. This phenomenon can be seen more clearly from
the partial zoom of i
a
in Fig. 8(g). Immediately after the
successful fault detection, the proposed fault-tolerant switching
strategy is engaged to counteract the fault, pulling the current i
a
back to its desired trajectory. It can be seen from Fig. 8(a)–(c)
that the effect of open-circuit fault is eliminated with the
proposed fault-tolerant switching technique. All of the three-
phase output currents become more balanced and are very close

to their desired sinusoidal waveforms after time t =0.1 s. As
a result, the torque response becomes much better, as shown in
Fig. 8(f). The speed response is shown in Fig. 8(e), with errors
E
me
=0.28 r/min and E
rms
=2.28 r/min. The largest absolute
speed error in steady state after the fault is E
max
=4r/min,
which is 2% of the operating speed. Table VI summarizes the
maximum, mean, and rms errors for the two experiments. Units
in the table are percentage of nominal speed.
B. Current Harmonics
The harmonic analysis is introduced in addition to the torque
and speed analyses to further reveal the impact of the fault, as
well as the efficacy of the proposed solution. Fig. 9(a) shows
the harmonic spectrum of i
a
after the fault occurs with no
remedial solution. The fundamental frequency for an eight-
pole PMSM operating at 200 r/min is 13.33 Hz. Since the
fault happens in every source period, the 60-Hz harmonic is
exceptionally large; its amplitude is about 50% of the fun-
damental amplitude. In addition, the spectrum contains many
other higher order harmonic components. The total harmonic
distortion, therefore, is expected to be very high. Such phenom-
enon contributes to the overheating of the motor winding, as
well as electromagnetic interference. The current discharges to

the clamp circuit every source period, imposing more stress on
the clamp circuit capacitor, and thus requires more attention in
clamp circuit design. The uncompensated speed ripple is not
252 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 59, NO. 1, JANUARY 2012
TABLE V
S
WITCHING RULES FOR ALL SINGLE OPEN-SWITCH FAULTS
excessively large, but the sudden drop or high deviation within
a period of the torque can be harmful for the mechanical load
coupled to the motor. Therefore, some action should be engaged
to stop the system as soon as possible for safety concerns.
Fig. 9(b) shows the harmonic spectrum of i
a
after the fault
occurs with the proposed solution. The fundamental component
of 13.33 Hz becomes bigger, outweighing the impact of other
harmonics. The 60-Hz harmonic i s considerably reduced to
about 17% of the fundamental harmonic. Furthermore, other
higher order harmonics are significantly attenuated. Thus, the
quality of the current and the torque produced are much
improved.
C. Load Disturbance Rejection, Step Response, and
Trajectory Tracking
Fig. 10 shows the disturbance rejection response to an addi-
tional 3-N · m load. It can be observed that the proposed fault-
tolerant switching strategy can also maintain some disturbance
rejection capability; the dip of speed and the recovery time were
almost the same as in the normal operation.
Fig. 11 shows the speed transient (step) response of the drive
system, for (a) the case of normal operation and (b) faulty

operation under the proposed fault-tolerant control method.
This test was taken under a no-load condition. The results show
that the proposed fault-tolerant solution can successfully start
up the drive system. This feature is noteworthy if the system is
stopped after fault detection, but has to be restarted. Without
the proposed fault-tolerant switching strategy, the imbalance
of the system might prevent the system from starting, and the
abnormal behavior is predicted to be very serious, particularly
when the torque command is large during the start-up period.
For these reasons, an experiment of faulty step response without
remedial action is not carried out in this work.
Fig. 12(a) and (b) shows the triangular speed reference track-
ing ability. The waveforms show that the drive system tracks
a varying reference, even under a fault condition. The fault-
tolerant switching strategy can still produce a proper dynamic
torque when operating under transient conditions.
In summary, all of the experimental results show that the
proposed online fault-detection method and postfault switching
strategy can operate correctly and effectively. The experimental
results also validate the theoretical analysis.
VI. D
ISCUSSION
This study has been primarily concerned on the open-circuit
fault of a single ac switch within the matrix converter system,
subject to the constraint of having neither redundant hard-
ware devices nor any hardware reconfiguration action. How-
ever, the merit of this work is recognizable. The experimental
fault-detection method and fault-tolerant switching strategy
NGUYEN-DUY et al.: IMPROVEMENT OF MATRIX CONVERTER DRIVE RELIABILITY 253
Fig. 7. Operation from normal mode to a fault without a remedial solution.

Output currents (a) i
a
,(b)i
b
,and(c)i
c
; (d) inverter-stage switching logic of
output phase a; (e) motor speed response; (f) estimated electromagnetic torque;
and (g) partial zoom of i
a
during a fault.
Fig. 8. Operation from normal mode to a fault with online fault detection
and the proposed switching strategy. Output currents (a) i
a
,(b)i
b
,and(c)i
c
;
(d) inverter-stage switching logic of output phase a; (e) motor speed response;
(f) estimated electromagnetic torque; and (g) partial zoom of i
a
during a fault.
254 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 59, NO. 1, JANUARY 2012
TABLE VI
S
PEED-CONTROL ERRORS AT 200 r/min (% SPEED)
Fig. 9. Harmonic spectrum of i
a
after the fault occurs. (a) Without a remedial

action and (b) with the proposed switching strategy.
perform correctly and well agree with the theoretical analysis.
The output currents with the proposed fault-tolerant method
become more balanced because the three outputs are affected
equally by a proper virtual dc-link voltage in each control step.
Consequently, the effect of the unbalanced outputs is elimi-
nated. Furthermore, stresses on the clamp circuit are reduced,
thus offering safer operation for the entire hardware system.
If the dc-link voltage of the clamp circuit is adopted as the
supply circuit for the control board and drive circuit [35], this
advantage becomes more noteworthy. Moreover, the proposed
fault detection and fault-tolerant switching strategy employ
only two standard current sensors, resulting in the least cost
and lowest redundant hardware development time compared to
existing approaches [25], [26], [30]–[32].
It must be acknowledged, however, that the constraint of
having no hardware redundancy introduces some limitations to
the proposed approach. First, the presented approach is a modi-
fication of a specific modulation strategy, so the fault-detection
method is not applicable if other modulation techniques are
used, such as the Alesina–Venturini modulation strategy [1]
or the space vector modulation technique [27]. A future work
will address the fault-detection issue when other modulation
strategies are used.
Fig. 10. Disturbance rejection response of 3-N · m load. (a) Normal operation
and (b) faulty operation with the proposed method.
Fig. 11. Speed transient (step) response with no load. (a) Normal operation
and (b) faulty operation with the proposed method.
NGUYEN-DUY et al.: IMPROVEMENT OF MATRIX CONVERTER DRIVE RELIABILITY 255
Fig. 12. Triangular reference tracking ability. (a) Normal operation and

(b) faulty operation with the proposed method.
Second, careful attention should be paid to the threshold
value mentioned in the fault-detection method. It is necessary
to estimate the discharge time of current to the clamp circuit,
where the peak current value and the time constant of the load
should be predetermined. According to the earlier analysis, the
smallest faulty time of about 3.7 ms is used for an eight-pole
2.1-kW PMSM. As long as the discharge time of current does
not exceed 3.7 ms, the fault-detection mechanism is applicable.
The aforementioned condition, i.e., the long discharge time, can
appear with loads that have a large time constant.
A third consideration is that the proposed fault-tolerant PWM
method results in some current-control limitations during “bad
sectors.” In the case of an open-switch fault at S
aC
, the “bad
sectors” are sectors 6, 7, 12, and 1. These are sectors where
the virtual dc-link voltages are relatively small, as shown in
Fig. 6. Within these sectors, the s mall dc-link voltage used
may not be large enough to counteract the effect of the motor
back EMF. Therefore, the current cannot be as well regulated as
in the normal operation with healthy hardware. Consequently,
the developed torque contains repetitive disturbances, occurring
whenever bad sectors are encountered. As a result, the motor
speed is affected. If the control objective is to r egulate the
speed as close to its desired value as possible, then several
other approaches might be considered. One approach is that
the system possesses sufficient inertia to smooth the torque
pulsations during the fault intervals. Another approach is to
design a more robust outer-loop speed controller instead of the

conventional PI controller. For example, modifying the q-axis
reference signal i

q
might be an alternative. Outer-loop control
modifications raise open issues for future study.
Finally, this paper focuses on the open-switch fault, but the
issue of shorted-switch fault is as well worth discussing. As
mentioned in Section I, there exist various different ways to pro-
tect against the shorted-circuit fault. For example, fast fuses can
be placed in series with the three input phase voltages. When a
shorted-circuit fault occurs, the fast fuses are anticipated to be
burned by the high amplitude currents, stopping the system for
offline fault diagnosis and maintenance.
VII. C
ONCLUSION
In this paper, the authors have investigated the fault-tolerant
control problem in a matrix converter drive system without
the help of any redundant devices. Moreover, the experimental
evaluation is on a practical adjustable speed drive system.
Experimental results prove the feasibility of the proposed
fault-detection method and its related fault-tolerant switching
strategy. Discussion on how to improve the performance with
the proposed switching strategy has been also presented. In
summary, a fully software-based fault-tolerant method requir-
ing no redundant hardware devices has been proposed and
demonstrated to provide an effective and economical solution
for the case of an open-switch-fault issue in matrix converter
drive systems. This work can be complementary to other works
in the research literature that address the short-circuit fault.

A
CKNOWLEDGMENT
The authors would like to thank the reviewers for their
constructive suggestions, which greatly improved the revised
manuscript.
R
EFERENCES
[1] A. Alesina and M. Venturini, “Analysis and design of optimum-amplitude
nine-switch direct ac–ac converters,” IEEE Trans. Power Electron.,vol.4,
no. 1, pp. 101–112, Jan. 1989.
[2] J H. Youm and B H. Kwon, “Switching technique for current-controlled
AC-to-AC converters,” IEEE Trans. Ind. Electron., vol. 46, no. 2, pp. 309–
318, Apr. 1999.
[3] J. Mahlein, J. Igney, J. Weigold, M. Braun, and O. Simon, “Matrix con-
verter commutation strategies with and without explicit input voltage sign
measurement,” IEEE Trans. Ind. Electron., vol. 49, no. 2, pp. 407–414,
Apr. 2002.
[4] R. Teichmann and J. Oyama, “ARCP soft-switching technique in matrix
converters,” IEEE Trans. Ind. Electron., vol. 49, no. 2, pp. 353–361,
Apr. 2002.
[5] F. Bradaschia, M. C. Cavalcanti, F. A. S. Neves, and H. E. P. de Souza, “A
modulation technique to reduce switching losses in matrix converters,”
IEEE Trans. Ind. Electron., vol. 56, no. 4, pp. 1186–1195, Apr. 2009.
[6] D F. Chen and T H. Liu, “Optimal controller design for a matrix con-
verter based surface mounted PMSM drive system,” IEEE Trans. Power
Electron., vol. 18, no. 4, pp. 1034–1046, Jul. 2003.
[7] T H. Liu, S H. Chen, and D F. Chen, “Design and implementation of
a matrix converter PMSM drive without a shaft sensor,” IEEE Trans.
Aerosp. Electron. Syst., vol. 39, no. 1, pp. 228–243, Jan. 2003.
[8] D F. Chen and T H. Liu, “Implementation of a novel matrix converter

PMSM drive,” IEEE Trans. Aerosp. Electron. Syst., vol. 37, no. 3,
pp. 863–875, Jul. 2001.
[9] M. Y. Lee, P. Wheeler, and C. Klumpner, “Space-vector modulated mul-
tilevel matrix converter,” IEEE Trans. Ind. Electron., vol. 57, no. 10,
pp. 3385–3394, Oct. 2010.
[10] H. Nguyen, H H. Lee, and T W. Chun, “Input power factor compensation
algorithms using a new direct-SVM method for matrix converter,” IEEE
Trans. Ind. Electron., vol. 58, no. 1, pp. 232–243, Jan. 2011.
[11] K. S. Smith, L. Ran, and J. Penman, “Real-time detection of intermittent
misfiring in a voltage-fed PWM inverter induction-motor drive,” IEEE
Trans. Ind. Electron., vol. 44, no. 4, pp. 468–476, Aug. 1997.
256 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 59, NO. 1, JANUARY 2012
[12] V. John, B S. Suh, and T. A. Lipo, “Fast-clamped short-circuit protec-
tion of IGBT’s,” IEEE Trans. Ind. Appl., vol. 35, no. 2, pp. 477–486,
Mar./Apr. 1999.
[13] M. Rodriguez, A. Claudio, D. Theilliol, and L. Velan, “A new fault de-
tection technique for IGBT based on gate voltage monitoring,” in Proc.
IEEE Power Electron. Spec. Conf., Jun. 2007, pp. 1001–1005.
[14] F. Huang and F. Flett, “IGBT fault protection based on di/dt feedback
control,” in Proc. IEEE Power Electron. Spec. Conf., Jun. 2007, pp. 1478–
1484.
[15] A. M. S. Mendes and A. J. M. Cardoso, “Voltage source inverter fault
diagnosis in variable speed ac drives, by the average current Park’s vector
approach,” in Proc. IEEE IEMDC , May 1999, pp. 704–706.
[16] R. Peuget, S. Courtine, and J P. Rognon, “Fault detection and isolation
on a PWM inverter by knowledge-based model,” IEEE Trans. Ind. Appl.,
vol. 34, no. 6, pp. 1318–1326, Nov./Dec. 1998.
[17] R. L. de Araujo Ribeiro, C. B. Jacobina, E. R. C. da Silva, and
A. M. N. Lima, “Fault detection of open-switch damage in voltage-fed
PWM motor drive systems,” IEEE Trans. Power Electron., vol. 18, no. 2,

pp. 587–593, Mar. 2003.
[18] O S. Yu, N J. Park, and D S. Hyun, “A novel fault detection scheme
for voltage fed PWM inverter,” in Proc. 32nd IEEE IECON, Nov. 2006,
pp. 2654–2659.
[19] R. L. de Araujo Ribeiro, C. B. Jacobina, E. R. C. da Silva, and
A. M. N. Lima, “Fault-tolerant voltage-fed PWM inverter AC motor
drive systems,” IEEE Trans. Ind. Electron., vol. 51, no. 2, pp. 439–446,
Apr. 2004.
[20] P. Nielsen, F. Blaabjerg, and J. K. Pedersen, “New protection issues of
a matrix converter: Design considerations for adjustable-speed drives,”
IEEE Trans. Ind. Appl., vol. 35, no. 5, pp. 1150–1161, Sep./Oct. 1999.
[21] S. Bolognani, M. Zordan, and M. Zigliotto, “Experimental fault-tolerant
control of a PMSM drive,” IEEE Trans. Ind. Electron., vol. 47, no. 5,
pp. 1134–1141, Oct. 2000.
[22] T H. Liu, J R. Fu, and T. A. Lipo, “A strategy for improving reliability of
field-oriented controlled induction motor drives,” IEEE Trans. Ind. Appl.,
vol. 29, no. 5, pp. 910–918, Sep./Oct. 1993.
[23] R. S. Chokhawala, J. Catt, and L. Kiraly, “A discussion on IGBT short-
circuit behavior and fault protection schemes,” IEEE Trans. Ind. Appl.,
vol. 31, no. 2, pp. 256–263, Mar./Apr. 1995.
[24] B. A. Welchko, T. A. Lipo, T. M. Jahns, and S. E. Schulz, “Fault tolerant
three-phase ac motor drive topologies: A comparison of features, cost, and
limitations,” IEEE Trans. Power Electron., vol. 19, no. 4, pp. 1108–1116,
Jul. 2004.
[25] S. Kwak and H. A. Toliyat, “An approach to fault-tolerant three-phase
matrix converter drives,” IEEE Trans. Energy Convers., vol. 22, no. 4,
pp. 855–863, Dec. 2007.
[26] S. Kwak, “Fault-tolerant structure and modulation strategies with fault
detection method for matrix converters,” IEEE Trans. Power Electron.,
vol. 25, no. 5, pp. 1201–1210, May 2010.

[27] L. Huber and D. Borojevic, “Space vector modulated three-phase to three-
phase matrix converter with input power factor correction,” IEEE Trans.
Ind. Appl., vol. 31, no. 6, pp. 1234–1246, Nov./Dec. 1995.
[28] C. Klumpner, F. Blaabjerg, I. Boldea, and P. Nielsen, “New modulation
method for matrix converters,” IEEE Trans. Ind. Appl., vol. 42, no. 3,
pp. 797–806, May/Jun. 2006.
[29] D. Casadei, G. Serra, A. Tani, and L. Zarri, “Matrix converter modulation
strategies: A new general approach based on space-vector representation
of the switch state,” IEEE Trans. Ind. Electron., vol. 49, no. 2, pp. 370–
381, Apr. 2002.
[30] S. M. Cruz, M. Ferreira, and A. J. M. Cardoso, “Output error voltages—A
first method to detect and locate faults in matrix converters,” in Proc. 34th
IEEE IECON, Orlando, FL, Nov. 2008, pp. 1319–1325.
[31] S. M. Cruz, M. Ferreira, and A. J. M. Cardoso, “A new method for
the detection and location of faults in matrix converters,” in Proc. IEEE
IEMDC
, Miami, FL, May 2009, pp. 165–170.
[32] S. M. A. Cruz, M. Ferreira, A. M. S. Mendes, and A. J. M. Cardoso, “Mod-
ulated error voltages for the diagnosis of faults in matrix converters,” in
Proc. IEEE ECCE, San Jose, CA, Sep. 2009, pp. 2263–2270.
[33] S. M. A. Cruz, M. Ferreira, A. M. S. Mendes, and A. J. M. Cardoso,
“Analysis and diagnosis of open-circuit faults in matrix converters,” IEEE
Trans. Ind. Electron., vol. 58, no. 5, pp. 1648–1661, May 2011.
[34] S. Kim, S K. Sul, and T. A. Lipo, “AC/AC power conversion based on
matrix converter topology with unidirectional switches,” IEEE Trans. Ind.
Appl., vol. 36, no. 1, pp. 139–145, Jan./Feb. 2000.
[35] C. Klumpner and F. Blaabjerg, “Experimental evaluation of ride-through
capabilities for a matrix converter under short power interruptions,” IEEE
Trans. Ind. Electron., vol. 49, no. 2, pp. 315–324, Apr. 2002.
Khiem Nguyen-Duy (S’09) was born in Hanoi,

Vietnam, in 1986. He received the B.S. degree (with
honors) in the field of electrical engineering from
Hanoi University of Science and Technology, Hanoi,
in 2009. He is currently working toward the M.S.
degree in the Department of Electrical Engineering,
National Taiwan University of Science and Technol-
ogy, Taipei, Taiwan.
His research interests are in power electronics,
motor drives, and the application of control theories
to these fields, specifically in active power filters,
matrix converters, and improvement of power quality.
Tian-Hua Liu (S’85–M’89–SM’99) was born in
Tao Yuan, Taiwan, on November 26, 1953. He re-
ceived the B.S., M.S., and Ph.D. degrees in electrical
engineering from National Taiwan University of Sci-
ence and Technology, Taipei, Taiwan, in 1980, 1982,
and 1989, respectively.
From August 1984 to July 1989, he was an
Instructor with the Department of Electrical En-
gineering, National Taiwan University of Science
and Technology. He was a Visiting Scholar at the
Wisconsin Electric Machines and Power Electronics
Consortium, University of Wisconsin, Madison, from September 1990 to
August 1991, and at the Center of Power Electronics Systems, Virginia Poly-
technic Institute and State University, Blacksburg, from July 1999 to January
2000. From August 1989 to January 1996, he was an Associate Professor
with the Department of Electrical Engineering, National Taiwan University
of Science and Technology, where he was the Department Chair from 2006
to 2009 and has been a Professor since February 1996. His research interests
include motor controls, power electronics, and microprocessor-based control

systems.
Der-Fa Chen (M’02) was born in Taipei, Taiwan,
on December 28, 1960. He received the M.S. de-
gree from National Chiao Tung University, Hsinchu,
Taiwan, in 1988 and the Ph.D. degree from Na-
tional Taiwan University of Science and Technology,
Taipei, in 2000.
From February 2003 to July 2006, he was an
Associate Professor with the Department of Indus-
trial Education and Technology, National Changhua
University of Education, Changhua, Taiwan, where
he has been a Professor since August 2006. His
research interests include power electronics, applications of control theory, and
motor controls.
John Y. Hung (S’79–M’80–SM’93–F’11) was born
in New York, NY, in February 28, 1959. He re-
ceived the B.S. degree in electrical engineering from
the University of Tennessee, Knoxville, in 1979,
the M.S.E. degree in electrical engineering from
Princeton University, Princeton, NJ, in 1981, and
the Ph.D. degree in electrical engineering from the
University of Illinois, Urbana, in 1989.
He was a Visiting Professor at National Taiwan
University of Science and Technology, Taipei,
Taiwan, in 2009–2010. He is currently a Professor
with the Department of Electrical and Computer Engineering, Auburn Univer-
sity, Auburn, AL.
Dr. Hung has received several awards for his teaching and research activities,
including a Best Paper Award from IEEE T
RANSACTIONS ON INDUSTRIAL

ELECTRONICS. He served as a Treasurer (in 2002–2007) and the Vice-
President for Conference Activities (in 2008–2010) in the IEEE Industrial Elec-
tronics Society and was the General Cochair for the 2008 IEEE International
Conference on Industrial Electronics, Control, and Instrumentation and the
2010 IEEE International Symposium on Industrial Electronics.

×