Vous êtes sur la page 1sur 9

558

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 12, NO. 3, MAY 1997

Simultaneous Design of Power Stage and


Controller for Switching Power Supplies
Cahit Gezgin, Student Member, IEEE, Bonnie S. Heck, Member, IEEE, and Richard M. Bass, Senior Member, IEEE

Abstract This paper develops a method for simultaneously


designing the power stage and controller for a switching power
supply. The method utilizes a numerical optimization procedure,
which facilitates computer-aided design. It is found that better
performance can be achieved than with a traditional two-step
design process, where the power stage and controller are designed
sequentially. Optimization and simulation results for a buck
converter are presented to illustrate the design process and
benefits.
Index Terms Computer-aided design, control-structure optimization, dcdc converters.

I. INTRODUCTION

RADITIONAL system design is a sequential process,


where the plant (structure) is designed first, with respect
to a set of structural constraints, and then, the controller
is designed for this fixed plant to yield an overall desired
dynamic response. In many cases, small changes in plant
parameters can make large changes in the performance of the
system beyond what is achievable by controller design only.
Thus, it is possible to outperform the overall response of the
traditional two-step system design by integrating the plant and
controller design into the same framework.
The idea of integrating control and structure design was
first proposed in the aerospace community for flexible space
structures [1], [2]. There are conflicting objectives in the
design of the flexible structure and controller. Reducing the
mass of space structures decreases the assembly costs, but
makes the structure more flexible and, hence, more difficult
to control to tight tolerances. A heavy, rigid structure may
cost more to build and use more control energy, but is less
sensitive to external disturbances, and, therefore, is easier to
control. Several design procedures were developed for this
problem that select the compensator parameters and structure
parameters simultaneously.
The method of integrating control and plant design can
be posed as a constrained optimization problem, often called
control-structure optimization (CSO). The cost function may
have both structural and control objectives. Similarly, the
constraints may be imposed on both plant parameters (or a
function of them, such as efficiency, losses, weight, etc.) and
certain functions of controller parameters (e.g., disturbance
Manuscript received April 18, 1996; revised November 4, 1996. This work
was supported by the National Science Foundation Contracts ECS-9 058 140
and ECS-9 412 636.
The authors are with the School of Electrical and Computer Engineering,
Georgia Institute of Technology, Atlanta, GA 30332-0250 USA.
Publisher Item Identifier S 0885-8993(97)03301-2.

rejection, constraining the maximum overshoot of transient response, etc.). The resulting problem can be solved numerically
using nonlinear programming techniques.
In power electronics, especially in the design of power
supplies, there is often as much freedom in the selection of
the plant (power circuit) parameters as there is in the controller. Moreover, both the plant and controller are designed
by the same person, which means that it is quite possible
to design them simultaneously. Thus, CSO is well-suited
to power-supply design. This type of optimal simultaneous
design has not been previously reported in power electronics
literature.
In this paper, the feasibility of applying CSO to the design
of a power supply and its control is explored. An overview of
the method is given first, and then, the method is demonstrated
on the design of a buck converter.
II. CSO
References [3][6] give different approaches to CSO, with
emphasis on mechanical systems. Since CSO is a multidomain
optimization of the system, there are generally at least two
objective functions: one is related with the structure, and
the other is related with the controller. In most cases, the
well-known quadratic cost function of the linear quadratic
regulator (LQR) theory is used as the control cost, and the total
structural mass is used as the structural cost. The optimization
parameters are cross-sectional areas of truss members and
the
and
matrices used in the optimal controller. A
number of complex constraints are defined for the multidomain optimization problem, for example, minimum buckling
strength, closed-loop damping, and robustness bounds, which
are also functions of structural and control design parameters.
Finally, the design is optimized using nonlinear programming
methods.
The LQR-based method described above has some disadvantages. First, it requires full state feedback. Second, there
is no general way to choose the weighting matrices to satisfy
both the control and structural objectives/constraints.
Another approach to the CSO problem uses parameter
optimization to select the elements of the feedback gain
matrix or coefficients of the numerator and denominator of
the compensator transfer function. A parameter-optimization
method developed by Layton and Peterson [7][9] is the
most promising for power-converter design. Laytons method
is based on covariance control theory (CCT), which was
developed by Skelton and his colleagues [10][15]. CCT is a
method for the parameterization of all-stabilizing controllers in

08858993/97$10.00 1997 IEEE

GEZGIN et al.: DESIGN OF POWER STAGE AND CONTROLLER FOR SWITCHING POWER SUPPLIES

terms of the closed-loop covariance matrix. Its advantages over


other control design methods, like
,
, etc., include
[11].
1) The controller order is selected a priori, which makes
it suitable for reduced-order dynamic compensator design and, hence, eliminates the necessity for model or
controller reduction.
2) Controller design equations are in closed form; no
Riccati, Lyapunov-type equations are needed.
3) Closed-loop stability is guaranteed. Actually, it characterizes the set of all-stabilizing controllers.
4) Closed-loop performance is guaranteed by rms values
that are assigned to the plant and controller states. This
is more advantageous than other methods, where only a
scalar performance measure is used.

559

disturbances. The set of all excitations is given by

where
is the th initial condition and
denotes the
intensity of the disturbance.
, defined in (9), is shown to be the solution of the
following Lyapunov equation [12]:
(10)
where

diag
A. An Overview of CCT
In this section, a brief introduction to CCT is presented. Its
relevance to the CSO problem will be established in Section IIB. This discussion is limited to dynamic compensators, which
are widely used for the control of power converters. Other
control schemes, like static output control or state-feedback,
are a special form of the more general dynamic compensator
case. Let the plant and compensator be described as follows:
(1)
(2)
(3)
(4)
(5)

diag

If
is a controllable pair, then
is asymptotically
stable if
solves (10) [12]. Conversely, if
is a stability matrix, i.e., all eigenvalues of
are
stable, then (10) has a unique solution for all
[10], [16].
Note that (10) has two unknowns, the covariance matrix
and the compensator matrix
The strategy is to select a
and then solve for
To be feasible,
must be
feasible
symmetric and positive definite, and it must be assignable to
(6) by some
The necessary and sufficient conditions under
which
is assignable are given in the following theorem
from [11][13].
Theorem 1: A specified
is assignable to the system
if and only if
(6) by
(11)
(12)

is the plant state vector,


is the
where
control input,
is the disturbance vector,
is
the measured output,
is the desired system output,
is the controller state vector.
and
The closed-loop system can be expressed as

(13)
denotes

are satisfied, where superscript


MoorePenrose inverse and

the

(14)

(6)
(7)

(15)
(16)

(8)

(17)
The one at-a-time covariance of the closed-loop system is
defined as

(18)
(19)

(9)

diag

diag
(20)

where
denotes the state response of (6) when only the
th excitation is applied (while others are held at zero) from a
total of
excitation events and
is the complexconjugate transpose of
These excitations are either
initial conditions on plant and controller states or impulsive

that will
There may be an infinite number of controllers
assign a given
to the closed-loop system (6). The paramthat assign
is
eterization of all covariance controllers
given in the following theorem.

560

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 12, NO. 3, MAY 1997

Theorem 2: Suppose
is assignable. Then, all the controllers, which assign
as a state covariance to the system
(6), are given by

Now, the constrained optimization problem can be posed


as follows.
Minimize

(21)

(31)

is an

subject to:
over the design space and elements of
1)
;
2)
;
3)
;
4)
.
The first three constraints for the controller are the
assignability conditions of Section II-A. They guarantee the
existence of a feedback matrix , which will assign
to
the closed-loop system. Additional structural constraints on
the plant parameters, such as upper or lower bounds, may be
included in the optimization.
The constrained nonlinear optimization problem expressed
above must be reformulated in order to apply standard nonlinear programming techniques. The matrix
must be symmet, there are actually
ric, so with
free parameters in
Also, the constraint
has to be recast in a more standard form. Let
be a Cholesky decomposition for
The matrix is an upper
triangular with
nonzero elements; the
constraint
is equivalent to requiring that the diagonal
elements of
be positive. Thus, the design parameters are
taken to be the plant parameters and the nonzero elements
of
and the constraints are the assignability constraints and
constraints
for

where
is an arbitrary skew-symmetric matrix,
arbitrary matrix of appropriate dimension, and

(22)
(23)
(24)
(25)
Proofs of these theorems are given in [13].

B. Application to CSO
When CSO, based on covariance matrix parameterization,
is used, the design variables are composed of both plant
parameters, , and controller parameters, which are the elements of the covariance matrix
One benefit of covariance
parameterization is that the control cost can be expressed as an
explicit function of
For example, the following quadratic
function is related to the control energy:
(26)
(27)

(28)
Similarly

III. SIMULTANEOUS DESIGN OF A BUCK


POWER STAGE AND CONTROLLER
The buck converter supplying a resistive load (Fig. 1) is
used to illustrate the CSO procedure and demonstrate its
feasibility for power-supply design. This circuit is taken from
the PowerExpress design package with minor modifications
[17]. The converter is operating in continous conduction mode
(CCM), and voltage-mode control is used.

(29)
A. The Plant and Controller
(30)
is related to the system performance [12], [15]. The vector
contains the error variables that are important for
the designer, and
is a positive definite weighting matrix.
The smaller the error variables, the better the performance. In
subsequent sections, (28) will be called control cost, and (30)
will be called performance cost.
In the general CSO problem, the objective function is
composed of both structural objectives (weight, price, etc.),
which are denoted as here, and controller-related objectives,
(28) and (30). The overall cost can be defined as the weighted
sum of these objectives. The boundaries of the feasible region
are determined by equality and inequality constraints. Usually,
the three assignability constraints of Theorem 1 are the only
equality constraints. Various structural objectives and closedloop stability are represented as inequality constraints.

The averaged and linearized model of the buck converter in


CCM is given as follows at the bottom of the next page.
For the circuit in Fig. 1,
V,
V,
,
m ,
m ,
m ,
m , and
V. Typically, the power-stage
(plant) design parameters are inductance
and capacitance

B. CSO Formulation
For the simultaneous optimal design of the plant and controller, different objective functions can be used. For example,
losses in the power stage are a good candidate for
structural objective. Another choice is the efficiency. Both
of these objectives can also be used as side constraints. For
example, during the design process, losses can be constrained
to be below a certain level. Equivalently, efficiency can be

GEZGIN et al.: DESIGN OF POWER STAGE AND CONTROLLER FOR SWITCHING POWER SUPPLIES

561

Fig. 1. Benchmark power supply.

kept above a predetermined percentage level by including it


in the constraint set. In this paper, losses in the circuit are kept
below a certain maximum value
The overall objective function to be minimized is the
weighted sum of
and :

which are given in (28) and (30); and are the weighting
constants. Since and are scalars, the weight matrices in
(28) and (30) are set to
For the CSO problem, the
constraints are imposed on the structural variables,
and

Duty

on the controller variables (nonzero elements of ), and some


functions of them. Specific equations for the side constraints
are developed below.
The values of
and
should not only minimize the
circuit losses, but also ensure CCM of operation and a prespecified peak-to-peak output-voltage ripple. For a given switching
frequency, power-stage components inductance
and capacitance
may be selected to meet the following criteria.
1) The value of inductance is chosen to assure CCM. To
maintain CCM down to th of rated output current ,

Duty

Duty

small-signal duty-ratio signal


steady-state values of

and

small-signal input-voltage variation


Duty steady-state duty ratio.

562

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 12, NO. 3, MAY 1997

the minimum value for

C. Design Procedure

is
(32)
Duty

For the buck converter, the CSO problem now can be cast
as follows.
Minimize

(33)
(38)

and Duty are the output current, switchwhere


ing frequency, load resistance, and steady-state duty
ratio, respectively.
2) Capacitance
is determined from maximum allowable
peak-to-peak output-ripple voltage
considerations:
(34)
Duty

Duty

and

(35)
The minimum value for

then becomes

Duty

(36)

is a time constant, which depends on the specific


where
capacitor technology. Similarly, maximum values are also
imposed on
and
, with respect to weight and price of
these components.
Another objective in power-module design is to keep steadystate losses below a certain value or to exceed a specified
efficiency. The steady-state losses in the buck converter operating in CCM are given as follows [18].
1) Inductor power loss:

2) Diode power loss:


Duty
Duty
3) Capacitor power loss:
Duty
4) MOSFET power loss:
Duty

Duty
Duty

where and
are given in MOSFET data sheets.
The total losses in the circuit become
(37)
and constraint becomes
appropriately.

, where

IV. SIMULATION RESULTS


In this section, a buck converter designed by the new
approach is compared with the benchmark circuit (Fig. 1) from
the PowerExpress design environment. Parasitic resistances,
output load, and input voltage are unchanged in all cases.
A. Benchmark Design

Duty

Duty

and the nonzero elements of


over the design space
subject to:
1)
;
2)
;
;
3)
4)
;
5)
where and are the weighting constants.
The software package MATLAB, with the optimization toolbox,1 was used to solve the problem in this paper. In particular,
the m-file constr.m was used, which utilizes a sequential
quadratic programming method. Note that this problem is not
necessarily convex so that local minima may exist. A starting
guess for
and
and the elements of is required, which
may influence the final answer. Specifically, larger numbers
give a larger cost, and smaller numbers tend to cause numerical
problems.

is chosen

The circuit parameters of the benchmark converter are given


in Fig. 1. The switching frequency is 200 kHz. Peak-to-peak
ripple voltage at steady state is 25 mV. The averaged and
linearized plant has a zero at
and two poles
at
The controller is a proportional-integral (PI) lead-type compensator, with a pole at the origin to yield zero steady-state
error, two zeros at 13.5 10 , which corresponds to 1/2 of
the resonant frequency of the plant. Another pole is located
at 314.3 10 , which is 1/4 of the switching frequency to
provide filtering at high frequencies. This is a typical procedure
to design the compensator of a buck converter.
The open-loop bode plot is given in Fig. 2(a). Fig. 3(a)
shows the impulse response of
for
of the averaged
and linearized open-loop system. Fig. 3(c) shows the closedloop impulse response of
for a disturbance
The
total steady-state loss in the circuit is 1.6234 W, which gives a
94.8% efficiency. Note that this high efficiency is due mainly
to the very small values of parasitic resistances.
B. CSO Design
The simultaneous optimal design of the plant and controller
is a nonlinear programming problem with no unique global
1 MATLAB

optimization toolbox.

GEZGIN et al.: DESIGN OF POWER STAGE AND CONTROLLER FOR SWITCHING POWER SUPPLIES

563

(a)

(b)
Fig. 2. Comparison of benchmark design and CSO design. (a) Bode plot for the linearized benchmark circuit. (b) Bode plot for the linearized CSO circuit.

minimum. Thus, a good starting point is essential. Selection of


initial values for
and
is quite straightforward. Although
the algorithm used in this paper can handle infeasible initial
and
within their limits
points quite well, choosing
improves the rate of convergence of optimization. Therefore,
initially, the inductor and capacitors set to the same values
provided by the benchmark power supply.

For controller parameters, initialization is more complicated. Since the elements of


do not have any physical
significance, a trial and error procedure is used. It is found
that choosing those elements greater than ten or smaller
than one tends to cause numerical problems. Therefore, in
are initialized to
this example, all nonzero elements in
one.

564

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 12, NO. 3, MAY 1997

EFFECT

TABLE I
INCREASING

OF

OF

TABLE II
EIGENVALUES FOR TWO CASES

(a)

(b)
Fig. 3. Comparison of benchmark design and CSO design. (a) Impulse
response of the linearized benchmark plant. (b) Impulse response of the
linearized CSO plant. (c) Closed-loop impulse response of the linearized
benchmark circuit. (d) Closed-loop impulse response of the linearized CSO
circuit.

The inequality constraints in (38) are constructed as foland


are determined from (32) and (36),
lows:
and
are chosen arbitrarily as 10 mH
respectively.
It is also possible to express these maximum
and 25
values according to some criteria like weight, price, etc. The
is set in order to
maximum allowable power loss
achieve a minimum 90% efficiency, which is comparable to
the benchmark design.
The three assignability conditions in (38) are slightly relaxed. It is observed that when these equality conditions are
forced strictly to be zero, design variables do not move from
their initial starting point, and numerical problems prevent
the optimization from converging to a solution. Note that
defines a manifold in the parameter space.

Keeping the parameters on this manifold and searching for a


minimum at the same time is a very stiff numerical problem.
Therefore, instead of making every element of , , and
identically zero, frobenius norms of the matrices are forced
to be below a specified small number
For most practical
cases, keeping about 0.1 yields a satisfactory accuracy and
convergence rate.
and
in (38) determine the
The weighting constants
and
Usually,
and
are
relative importance of
competing with each other. Decreasing one tends to increase
in order prevent
the other. It is desirable to have a small
from saturating. However, it is also
the control signal
important to have a small , which will yield a large dc
gain and, consequently, a small steady-state error ( ).
is also related to the disturbance rejection capability of the
closed-loop system.
A family of different plants and controllers are designed by
varying while keeping
An acceptable design is the
one that regulates the output voltage within 1% of the nominal
of 150 mV. Table
value, which corresponds to having an
,
, and
I shows the effect of altering on , ,
Note that increasing from 10 to 15 in Table I does not
increase
in this case. This shows that the surface defined
has a lower local minimum for
by the objective function
The result corresponding to the last row of Table
I is chosen for further analysis and comparison with the
benchmark circuit. This case will be referred to as the CSO
design. Table II shows the plant, controller, and closed-loop
system eigenvalues for benchmark and CSO designs.
The open-loop bode plot of the linearized plant and controller for the CSO design is depicted in Fig. 2(b). Fig. 3(b)
and 3(d) illustrate the open- and closed-loop impulse refor the averaged and linearized CSO circuit.
sponse of
The bandwidth of CSO design is close to the benchmark
design, and its phase and gain margins are adequate. Note
that the plant designed by CSO is more oscillatory and slower
than the averaged and linearized benchmark plant. On the
other hand, the closed-loop impulse responses show that the
linearized CSO design has less overshoot, better damping,
and is much faster. The switching circuits are then simulated
with a SABER simulator starting from a zero initial state.
ms, after the steady state has been reached, a
At

GEZGIN et al.: DESIGN OF POWER STAGE AND CONTROLLER FOR SWITCHING POWER SUPPLIES

565

(a)

(b)

(c)

(d)

Fig. 4. Vo and IL (- - benchmark design and


transient during the disturbance.

CSO design). (a) Output-voltage transient during startup and disturbance. (b) Inductor current

disturbance is applied to both converters, i.e., the load current


is increased from 2 to 5 A. The dashed curves in Fig. 4(a)
and 4(b) correspond to the benchmark circuit, and the solid
curves correspond to the CSO design. Note that the CSO
design has a much better output-voltage transient response.
The respective inductor currents are shown in Fig. 4(b). The
opamp in the compensator in Fig. 1 saturates at 15 V, which
also limits the error signal
to the comparator. Fig. 5
shows that the opamp in the CSO design does not reach
the 15-V limit. This is the benefit of minimizing control
effort in the CSO procedure, i.e., the CSO design requires
less control energy to recover from the disturbance. It has
better disturbance-rejection capability and is faster than the
benchmark design.
V. CONCLUSION
In this paper, the feasibility of simultaneously designing the
power stage and compensator of a switching power supply has
been demonstrated. The design process is cast as a nonlinear
constrained optimization problem, which allows the use of

Fig. 5. Control signal to the comparator.

existing numerical solution methods. The only data needed in


the design process are input and output voltages, load current,

566

IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 12, NO. 3, MAY 1997

and switching frequency. The compensator can be a full-order


or reduced-order controller. In most cases, the algorithm can
handle infeasible starting points.
This research also shows the interaction between the structure and controller. Small changes in plant parameters can
make drastic changes in the controllers effort and greatly affect the closed-loop system. In this example, an underdamped,
slow plant is found to be better for the controller, i.e., requires
less control energy. This performance is achieved without
violating the structural constraints. In the current method,
the overshoot observed in the inductor current during the
disturbance is not penalized. A cost function penalizing this
overshoot can be added to the overall objective function. The
current method does not guarantee the existence of a pole at the
origin in the open-loop transfer function. An open-loop pole
at the origin would give zero steady-state error. A remedy to
this problem is to augment the averaged and linearized plant
model to include an integrator. In this way, the plant will be
third-order and the controller will be first-order.
The method also reveals new directions for future research.
For a full multiobjective optimization, switching frequency
should be included in the design variables. Effects of other
cost functions associated with the power stage, like weight
and price, should also be investigated.

[14] M. Corless, G. Zhu, and R. E. Skelton, Improved robustness bounds


using convariance matrices, Proc. IEEE Conf. Decision and Control,
vol. 3, 1989, pp. 26672672.
[15] R. E. Skelton, J.-H. Xu, and K. Yasuda, Minimal energy covariance
control, Int. J. Control, vol. 59, pp. 15671578, June 1994.
[16] F. L. Lewis, Optimal Control. New York: Wiley, 1986.
[17] PowerExpress Users Guide, Release 4.0, Analogy Inc., Beaverton, OR,
1995.
[18] A. Reatti, Steady-state analysis including parasitic components and
switching losses of buck and boost dc-dc pwm converters under any
operating condition, Int. J. Electron., vol. 77, pp. 679701, May 1994.

REFERENCES
[1] J. Cheng, G. Ianculescu, C. S. Kenney, A. J. Laub, J. Ly, and P. M.
Papadopoulos, Control-structure interaction, IEEE Control Syst. Mag.,
vol. 12, pp. 413, Oct. 1992.
[2] Y.-P. Harn, G. M. Kabuli, and R. L. Kosut, Optimal simultaneous
control structure design, Proc. American Control Conf., vol. 2, 1991,
pp. 14421447.
[3] M. Milman, M. Salaman, R. E. Scheid, R. Bruno, and J. S. Gibson,
Combined control-structure optimization, Computational Mechanics,
vol. 8, pp. 18, Aug. 1991.
[4] P. G. Maghami, S. M. Joshi, and D. B. Price, An integrated controlstructures design methodology for a flexible spacecraft, Advances
Astronautical Sci., vol. 78, pp. 369385, 1992.
[5] M. Salama, J. Garba, L. Demsetz, and F. Udwadia, Simultaneous
optimization of controlled structures, Computational Mechanics, vol.
3, pp. 275282, Mar. 1988.
[6] I. Kajiwara, K. Tsujioka, and A. Nagamatsu, Approach for simultaneous optimization of a structure and control system, AIAA J., vol. 32,
pp. 866873, Apr. 1994.
[7] L. D. Peterson and J. B. Layton, Multiobjective structural control design by optimizing the closed-loop covariance, AIAA/ASME Structures,
Structural Dynamics and Materials Conf., vol. 1, 1991, pp. 534541.
[8] J. B. Layton and L. D. Peterson, Benefits of integrated control/structure
design using covariance controlA comparison of two approaches,
AIAA/ASME Structures, Structural Dynamics and Materials Conf., 1993,
pp. 31263136.
[9] J. B. Layton, Control/structure covariance optimization including performance robustness, Proc. SPIEThe Int. Society for Optical Eng.,
vol. 2442, 1995, pp. 7285.
[10] A. Hotz and R. E. Skelton, Covariance control theory, Int. J. Control,
vol. 46, 1987, pp. 1332.
[11] R. E. Skelton, Control of state and input covariances for dynamic
systems, Proc. IEEE Conf. Decision and Control, vol. 3, 1988, pp.
19021907.
[12] R. E. Skelton and M. Ikeda, Covariance controllers for linear
continuous-time systems, Int. J. Control, vol. 49, 1989, pp. 17731785.
[13] K. Yasuda and R. E. Skelton, Covariance controllers: A new parameterization of the class of all stabilizing controllers, Proc. American
Control Conf., vol. 1, 1990, pp. 824829.

Cahit Gezgin (S96) was born in Turkey in 1968.


He received the B.S. (Honors) and M.S. degrees
in electrical and electronics engineering from the
Middle East Technical University, Ankara, Turkey,
in 1990 and 1992, respectively. He is currently
pursuing the Ph.D. degree in electrical engineering
at the Georgia Institute of Technology, Atlanta.
From 1990 to 1992, he was a Research Assistant
at the Middle East Technical University. Since 1994,
he has been a Research Assistant at the Georgia Institute of Technology. His current research interests
include modeling, simulation, and computer-aided design of power converters.
Mr. Gezgin was the recipient of the Prof. Parlar Foundation Thesis Award
for his M.S. thesis Intelligent fault diagnosis in power systems using neural
networks in 1992.

Bonnie S. Heck (S86M88) was born in Michigan


City, IN, in 1960. She received the B.S. degree
in electrical engineering from the University of
Notre Dame, Notre Dame, IN, in 1981, the M.S.
degree in mechanical and aerospace engineering
from Princeton University, Princeton, NJ, in 1984,
and the Ph.D. degree in electrical engineering from
the Georgia Institute of Technology, Atlanta, in
1988.
She worked for Honeywell, Inc. as an Engineer
from 1983 to 1985. She is currently an Associate
Professor in Electrical and Computer Engineering at the Georgia Institute
of Technology. Her research interests include numerical methods, large-scale
systems, nonlinear control, power systems, measurement systems, and singular
perturbation theory.

Richard M. Bass (S82M82SM94) was born in


Jacksonville, FL, in 1959. He received the B.E.E.
and M.S.E.E. degrees from the Georgia Institute
of Technology, Atlanta, in 1982 and 1983, respectively. He received the Ph.D. degree in electrical
engineering from the University of Illinois, UrbanaChampaign, in 1990.
From 1983 to 1987, he was employed by the
Veterans Administration as an Electronics Engineer
in the Atlanta Rehabilitation Research and Development Unit. While there, he participated in the
development of an improved motor drive for powered wheelchairs and other
adaptive aids for people with disabilities. From 1987 to 1990, he was
at the University of Illinois pursuing doctoral studies in power electronic
systems analysis and design. At present, he is an Assistant Professor in
the School of Electrical and Computer Engineering, Georgia Institute of
Technology. In 1994, he was a Visiting Researcher with the Electricite de
France electric vehicle research program. His current research interests include
electric vehicle charging infrastructure, averaging and variable structure
control methods for power electronics, and power semiconductor device
modeling.
Dr. Bass is a Registered Professional Engineer in Georgia.

Vous aimerez peut-être aussi