You are on page 1of 12

Feedback-Linearization-Based Nonlinear Control

for PEM Fuel Cells


Woon Ki Na, Student Member, IEEE, and Bei Gou, Member, IEEE
AbstractThis paper presents a dynamic nonlinear model for
polymer electrolyte membrane fuel cells (PEMFCs). A nonlinear
controller is designed based on the proposed model to prolong the
stack life of the PEM fuel cells. Since it is known that large devi-
ations between hydrogen and oxygen partial pressures can cause
severe membrane damage in the fuel cell, feedback linearization
is applied to the PEM fuel cell system so that the deviation can be
kept as small as possible during disturbances or load variations. A
dynamic PEM fuel cell model is proposed as a nonlinear, multiple-
input multiple-output system so that feedback linearization can
be directly utilized. During the control design, hydrogen and oxy-
gen inlet ow rates are dened as the control variables, and the
pressures of hydrogen and oxygen are appropriately dened as the
control objectives. The details of the design of the control scheme
are provided in the paper. The proposed dynamic model was tested
by comparing the simulation results with the experimental data
previously published. The simulation results show that PEMFCs
equipped with the proposed nonlinear controls have better tran-
sient performances than those with linear controls.
Index TermsExact linearization, nonlinear dynamic model,
polymer electrolyte membrane fuel cells.
I. INTRODUCTION
A
FUEL cell is an electrochemical energy device that con-
verts the chemical energy of fuel directly into electricity
and heat with water as a by product of the reaction. As a renew-
able energy source, the fuel cell is widely regarded as one of the
most promising energy sources because of its high energy ef-
ciency, extremely low emission of oxides of nitrogen and sulfur,
and very low noise, as well as the cleanness of its energy pro-
duction. Based on the currently used types of electrolytes, fuel
cells are divided into polymer electrolyte membrane fuel cells
(PEMFCs), solid oxide fuel cells (SOFCs), phosphoric acid fuel
cells (PAFCs), molten carbonate fuel cells (MCFCs), alkaline
fuel cells (AFCs), direct methanol fuel cells (DMFCs), zinc air
fuel cells (ZAFCs), and photonic ceramic fuel cells (PCFCs) [1].
In order to generate a reliable and efcient power response and
to prevent membrane damage as well as detrimental degradation
of the stack voltage and oxygen depletion, it is necessary to de-
sign a better control scheme to achieve optimal air and hydrogen
inlet owratesi.e., controls that can performair and hydrogen
pressure regulation and heat/water management precisely based
on the current drawn from the fuel cell system [3], [4].
Manuscript received March 24, 2007; revised June 4, 2007. Paper no.
TEC-00091-2006.
The authors are with the Energy SystemResearch Center, University of Texas,
Arlington, TX 76019 USA (e-mail: bgou@uta.edu).
Color versions of one or more of the gures in this paper are available online
at http://ieeexplore.ieee.org.
Digital Object Identier 10.1109/TEC.2007.914160
Fig. 1. Polarization curve (Ballard Mark V PEMFC at 70

C) [1].
First of all, an PEMFC system must be accurately modeled in
order to apply a suitable nonlinear control scheme. Models have
been reported so far in the literature for the PEMFC, ranging
from stationary and dynamic models [3][9], [13][17], [20],
[24] for the control design applied to a fuel cell vehicle and a
distributed generation system [11], [12].
Unfortunately, those models are mainly for experimental ver-
ications other than control design [1][3] or for prediction of
the fuel cell phenomenon by analyzing an electrochemical re-
action, the thermodynamics, and the uid mechanics. Recently,
Purkrushpan et al. [3] developed a control-oriented PEMFC
model that includes ow characteristics and dynamics of the
compressor and the manifold (anode and cathode), reactant par-
tial pressures, and membrane humidity. However, because of
the nonlinear relationship between stack voltage and load cur-
rent shown in Fig. 1 [1] and the state equations [3], [20], it is
a challenge to develop a nonlinear controller for the PEMFC.
Because of the operational parametric uncertainties such as the
parametric coefcients for each cell on kinetic, thermodynamic,
and electrochemical foundations, and the resistivity of the mem-
brane for the electron ow, the linear PEMFC models proposed
by Purkrushpan and coworkers [3][5] and Chiu et al. [20]
using Jacobian linearization via a Taylor series expansion at
the nominal operating point cannot easily achieve satisfactory
dynamic performance under large disturbances. An accurate
nonlinear dynamic model needs to be developed for the fuel
cell system as well as an advanced controller design technique,
considering the nonlinearity and uncertainty that need to be
proposed.
A fuzzy control system for a boost dc/dc converter of a fuel
cell system was developed in [18]. Neural optimal control was
presented for the PEMFC by using an articial neural network
(ANN) in [19]. However, instead of controlling the PEM fuel
cell system, the neural optimal control is mainly used to derive a
new architecture to synthesize an approximated optimal control
0885-8969/$25.00 2008 IEEE
by means of the ANN, where the PEM fuel cell was chosen as
a test bed.
In this paper, feedback linearization, a well-known nonlinear
approach, is applied to design a controller, based directly on
the nonlinear dynamic fuel cell model, to achieve more robust
transient behavior. Furthermore, the fuel cell stack life can be
prolonged and stack systems can be protected by minimizing the
deviations between the hydrogen and oxygen partial pressures
[2], [4].
In the last few years, feedback linearization for nonlinear dy-
namic models has been widely used [26][29], [31]. Feedback
linearization uses a nonlinear transformation to transform an
original nonlinear dynamic model into a linear model by diffeo-
morphism mapping [26][29], [31]. An optimal control theory
is also applied to obtain a linear control that is transformed back
to the original space by using the nonlinear mapping.
In this paper, the nonlinear dynamic model developed in
[3][5] and the small signal model of an PEMFC in [20] are
considered together to obtain a new dynamic nonlinear model
that is appropriate for developing a nonlinear controller. The
proposed controller, which is expected to perform rapid tran-
sient responses under load variations, is tested in MATLAB
simulink environment.
The paper is organized as follows. Section II gives a brief
introduction to feedback linearization. The PEMFC dynamic
model based on [3][5] and [20] is proposed in Section III, and
Section IV addresses the design of a nonlinear controller for
an PEMFC. Section V provides the simulation results for the
proposed controller, and Section VI concludes the paper.
II. NONLINEAR CONTROL BY FEEDBACK LINEARIZATION
For decades, signicant progress has been made in control
designs based on nonlinear concepts. In particular, nonlinear
control theory developed from differential geometry, known as
exact linearization or feedback linearization, has become more
and more attractive for chemical process control because many
chemical processes are basically of high nonlinearity [29], [34].
Hence, one of the main motivations of utilizing feedback lin-
earization for a fuel cell system is that the operation of PEMFC
is inherently a nonlinear chemical process. In this section, feed-
back linearization of nonlinear systems is briey introduced.
More details of nonlinear control based on differential geome-
try are available in [26][28], [34].
A. Feedback or Exact Linearization
Consider a single-input single-output (SISO) nonlinear sys-
tem described by the state equation
x = f(x) + g(x)u
y = h(x) (1)
where x is an n-dimensional state vector that is assumed to be
measurable, u is a scalar input, and y is a scalar output.
The objective of feedback linearization is to create a linear
differential relation between the output y and a newly dened
input v. The notation and concepts of differential geometry are
essential to understand this approach.
The Lie derivative of a scalar function h(x) with respect to a
vector function f(x) is dened as
L
f
h(x) = hf =
h(x)
x
f(x). (2)
Repeated Lie derivatives can be dened recursively as
L
k
f
h(x) = L
f
_
L
k1
f
h
_
=
_
L
k1
f
h
_
f
L
0
f
h(x) = h(x) (3)
for k = 1, 2, . . .
Similarly, in case of another vector eld g
L
g
L
f
h = (L
f
h)g. (4)
The output needs to be differentiated for r times until it is
directly related to the input u. The number r is called the relative
degree of the system.
The system is said to have a relative r at a point x
0
if
1) L
g
L
k
f
h(x) = 0 for all x in the neighborhood of x
0
and for
k < r 1;
2) L
g
L
r1
f
h(x
0
) = 0.
Thus, according to the aforementioned condition, with a de-
ned relative degree, the r time derivatives of y are described
as
y
(k)
= L
k
f
(x)h(x), for k = 0, 1, . . . , r 1,
y
(r)
= L
r
f
(x)h(x) + L
g
L
r1
f
h(x)u. (5)
The control law is
u =
1
L
g
L
r1
f
h(x)
_
L
r
f
h(x) + v
_
(6)
where y
(r)
= v. This control law can transform the nonlinear
system into a linear one. In addition, a nonlinear transformation
of a coordinate in the state space
z = (x) (7)
is called a local diffeomorphism, in which the map between
the new input v and the output is exactly linear for all x in the
neighborhood of x
0
.
This feedback linearization theory can be used to design
multiple-input multiple-output (MIMO) nonlinear system:
x = f(x) +
m

i=1
g
i
(x)u
i
y
i
= h(x), i = 1, 2, . . . , m (8)
where x is an n-dimensional state vector and u and y are m-
dimensional input and output vectors. The system is said to
have a vector relative degree {r
1
, r
2
, . . . , r
m
} at a point x
0
if
1) L
g
j
L
k
f
h
i
(x) = 0 for all 1 j m, all 1 i m, k <
r
i
1, and for all x in the neighborhood of x
0
;
2) mm matrix
A(x) =
_

_
L
g
1
L
r
1
1
f
h
1
(x) L
g
m
L
r
1
1
f
h
1
(x)
.
.
.
.
.
.
L
g
1
L
r
m
1
f
h
m
(x) L
g
m
L
r
m
1
f
h
m
(x)
_

_
(9)
is nonsingular at x = x
0
, which is called as a decoupling matrix.
Based on the dened relative degree, the control law of an
MIMO nonlinear system is dened as
u = A
1
(x)b(x) + A
1
(x)v (10)
where
b(x) =
_

_
L
r
1
f
h
i
(x)
.
.
.
L
r
m
f
h
m
(x)
_

_ v =
_

_
v
1
.
.
.
v
m
_

_ =
_

_
y
(r
1
)
1
.
.
.
y
(r
m
)
1
_

_
.
Note that the control lawin (10) transforms the nonlinear sys-
tem into a linear one in which the aforementioned inputoutput
relation is linearized and decoupled. In this paper, feedback lin-
earization for an MIMO nonlinear system is utilized due to the
MIMO dynamic nonlinear model of the PEMFC.
III. PEM FUEL CELLS AND DYNAMIC MODEL
A. PEM Fuel Cells
An PEM fuel cell consists of a polymer electrolyte mem-
brane sandwiched between two electrodes (anode and cathode).
In the electrolyte, only ions can exit and electrons are not al-
lowed to pass through. So, the ow of electrons needs a path
like an external circuit from the anode to the cathode to pro-
duce electricity because of a potential difference between the
anode and cathode. The overall electrochemical reactions for an
PEM fuel cell fed with a hydrogen-containing anode gas and an
oxygen-containing cathode gas are as follows:
Anode: 2H
2
4H
+
+ 4e

Cathode: O
2
+ 4H
+
+ 4e

2H
2
O
Overall: 2H
2
+ O
2
2H
2
O+ electricity + heat.
In practice, a 5 kW fuel cell stack, such as a Ballard MK5-E
PEMFC stack, uses a pressurized hydrogen tank at 10 atm and
oxygen taken from atmospheric air [10], [32].
On the anode side, a fuel processor, a so-called reformer,
that generates hydrogen through reforming methane or other
fuels like natural gas, can be used instead of the pressurized
hydrogen tank. Apressure regulator and purging of the hydrogen
component are also needed. On the cathode side, an air supply
system containing a compressor, an air lter, and an air ow
controller are required to maintain the oxygen partial pressure
[1], [4], [12], [25]. On both sides, a humidier is needed to
prevent dehydration of fuel cell membrane [1], [4], [25]. In
addition, a heat exchanger, a water tank, a water separator, and
a pump may be needed for water and heat management in the
FC systems [1], [4], [25].
To produce a higher voltage, multiple cells have to be con-
nected in series. Typically, a single cell produces voltage be-
TABLE I
CELL VOLTAGE PARAMETERS [1]
tween 0 and 1 V based on the polarization IV curve, which
expresses the relationship between stack voltage and load cur-
rent [1], [25]. Fig. 1 shows that their relationship is nonlinear
and mainly depends on current density, cell temperature, reac-
tant partial pressure, and membrane humidity [1], [25].
The output stack voltage V
st
[1] is dened as a function of the
stack current, reactant partial pressures, fuel cell temperature,
and membrane humidity:
V
st
= E V
activation
V
ohmic
V
concentration
. (11)
In the aforementioned equation, E = N
o
[V
o
+ (RT/2F)
ln(P
H
2
_
P
O
2
/P
H
2
O
c
)] is the thermodynamic potential of the
cell or reversible voltage based on the Nernst equation [1],
V
activation
is the voltage loss due to the rate of reactions on
the surface of the electrodes, V
ohmic
is the ohmic voltage drop
from the resistances of proton ow in the electrolyte, and
V
concentration
is the voltage loss from the reduction in concen-
tration gases or the transport of mass of oxygen and hydrogen.
Their equations are given as follows:
V
activation
= N
RT
2F
ln
_
I
fc
+ I
n
I
o
_
(12)
V
ohm
= NI
fc
r (13)
V
concentration
= Nm exp(nI
fc
). (14)
In (11), P
H
2
, P
O
2
, and P
H
2
O
c
are the partial pressures of hy-
drogen, oxygen, and water, respectively. Subscript c means the
water partial pressure, which is vented from the cathode side.
See Table I for cell voltage parameters.
Adetailedexplanationof eachvoltage loss canbe foundin[1],
and other voltage descriptions are also reported in [3], [4], [6],
[8], [10], where the fuel cell voltage is mainly expressed by the
combination of physical and empirical relationships in which
many parametric coefcients of the membrane water content,
humidity, and temperate, as well as the reactant concentrations
are involved. In our paper, the general voltage form as given
in (11) will be used because water and temperature factors are
assumed to be constant due to their slow response time.
Fig. 2. Illustration gas ows of PEMFC.
B. State-Space PEMFC Dynamic Model
The following assumptions are made when deriving the sim-
plied dynamic nonlinear model for the PEMFC.
1) Due to a slow response time regarding the stack tempera-
ture (about 102 s [1], [23], [25], [35]), the operating stack
temperature is assumed to be constant.
2) The fuel cell is well humidied on both the anode and
cathode sides.
3) The stoichiometry for hydrogen is set as 2 to provide
exibility in operation.
4) The stoichiometriy for oxygen ranges from 2 to 2.5 for
normal operation.
5) For the faster response (like a battery), this fuel cell model
has to have a continuous supply of reactants adjusted by
the current drawn from the stack.
6) For water management, it is assumed that the liquid water
does not leave the stack and that it evaporates into the
cathode or anode gas if humidity on either side drops
below 100% [3].
7) Perfectly controlled humidier and temperature con-
troller are assumed because of their slow response time
[3].
8) The mole fractions of the inlet reactants are assumed to be
constant to build the simplied dynamic PEMFCmodel. In
other words, pure hydrogen (99.99%) is fed to the anode,
and air that is uniformly mixed with nitrogen and oxygen
by a ratio of say 21:79 is supplied to the cathode.
The ideal gas law and mole conservation rule are applied
by supposing that all gases are ideal. The partial pressures of
hydrogen and water on the anode, and the partial pressures of
oxygen, nitrogen, and water on the cathode are dened as state
variables of the PEMFC. The relationship between inlet gases
and outgases is illustrated in Fig. 2 [20].
According to the ideal gas law and the mode conservation
rule, the partial pressure of each gas is balanced by the gas inlet
ow rate minus the gas consumption and the gas outlet ow
rate. The partial pressure derivatives are given as follows.
Anode mole conservation:
dP
H
2
dt
=
RT
V
A
(H
2in
H
2used
H
2out
)
dP
H
2
O
A
dt
=
RT
V
A
(H
2
O
Ain
H
2
O
Aout
H
2
O
mbr
). (15)
Cathode mole conservation:
dP
O
2
dt
=
RT
V
C
(O
2in
O
2used
O
2out
)
dP
N
2
dt
=
RT
V
C
(N
2in
N
2out
)
dP
H
2
O
C
dt
=
RT
V
C
(H
2
O
Cin
+ H
2
O
Cproduced
H
2
O
Cout
+ H
2
O
mbr
) (16)
where H
2in
, O
2in
, N
2in
, H
2
O
Ain
, and H
2
O
Cin
are the inlet ow
rates of hydrogen, oxygen, nitrogen, the anode side water, and
the cathode side water, respectively.
In addition, H
2out
, O
2out
, N
2out
, H
2
O
Aout
, and H
2
O
Ciout
are the outlet ow rates of each gas. H
2used
, O
2used
, and
H
2
O
Ciproduced
are the usage and the production of the gases.
Normally, the membrane water inlet ow rate H
2
O
mbr
across
the membrane is a function of the stack current and the
relative humidity. However, because we assumed that the
humidity is constant with the membrane average water content

m
= 14 [3], [21] under 100%of the relative humidity, H
2
O
mbr
is dened as a function of the stack current only: H
2
O
mbr
=
1.2684(NA
fc
I
fc
/F) [3], [21], where A
fc
(inverse square
centimeter) is the fuel cell active area, N is the number of the
fuel cells, and I
fc
the cell current density. All units of ow
rates, usages, and production of gases are dened as mole per
second. V
a
and V
c
are the anode and the cathode volumes,
respectively, and their units are cubic meter. Based on the basic
electrochemical relationships, the usage and production of the
gases are a function of the stack current:
H
2used
= 2O
2used
= H
2
O
Cproduced
=
NA
fc
I
fc
2F
. (17)
For simplicity, let us dene
NA
fc
2F
= C
1
, and 1.2684
NA
fc
F
= C
2
.
With the measured inlet ow rates and the stack current, the
outlet ow rates are given by the summations of anode and
cathode inlet ow ratesthat is, Anode
in
and Cath
in
, minus
the usage and production of gases. Anode
in
is dened by H
2in
+
H
2
O
Ain
, and Cath
in
is dened by O
2in
+ N
2in
+ H
2
O
Cin
.
The outlet ow rates on the anode side are
H
2out
= (Anode
in
C
1
I
fc
)F
H
2
H
2
O
Aout
= (Anode
in
C
2
I
fc
)F
H
2
O
A
. (18)
The outlet ow rates on the cathode side are
O
2out
=
_
Cath
in

C
1
2
I
fc
_
F
O
2
N
2out
= Cath
in
F
N
2
H
2
O
Cout
= (Cath
in
+ C
1
I
fc
+ C
2
I
fc
)F
H
2
O
C
(19)
where F
H
2
, F
H
2
O
A
, F
O
2
, F
N
2
, and F
H
2
O
C
are the pressure frac-
tions of gases inside the fuel cell, given as follows [20]:
F
H
2
=
P
H
2
P
H
2
+ P
H
2
O
A
F
H
2
O
A
=
P
H
2
O
A
P
H
2
+ P
H
2
O
A
,
F
O
2
=
P
O
2
P
O
2
+ P
N
2
+ P
H
2
O
C
F
N
2
=
P
N
2
P
O
2
+ P
N
2
+ P
H
2
O
C
F
H
2
O
C
=
P
H
2
O
C
P
O
2
+ P
N
2
+ P
H
2
O
C
.
(20)
To analyze the transient behavior of fuel cells, we take into
account the pressure fraction of each gas in the Department of
Energy (DoE) model [33]. The summation of anode and cathode
pressures is constant. The fuel cell model with the pressure frac-
tions of the gases was originally proposed by Chiu et al. [20].
In [20], only three pressure fractionsF
H
2
, F
O
2
, and F
H
2
O
C

are considered. In [3], [9], and [10], the outlet ow rates were
described by the difference between upstream pressures (par-
tial pressures) and downstream pressures (regulated pressures)
with a constant ow. In those models, the ow constants and
downstreams were set at 0.065 mol s
1
(in atmosphere) and 3
(in atmosphere), respectively. However, these xed values are
not appropriate for a dynamic model of a fuel cell. In our paper,
since all pressure fractions of gases, including the partial pres-
sures of water on the anode side and nitrogen on the cathode
side, are considered, a more accurate model and transient behav-
ior of fuel cells can be obtained, as compared with [3], [9], [10],
and [20].
The state equations as shown in (21) and (22) are obtained by
substituting (17)(20) into (15)(16).
State equations on the anode side:
dP
H
2
dt
=
RT
V
A
[H
2in
C
1
I
fc
(Anode
in
C
1
I
fc
)F
H
2
]
dP
H
2
O
a
dt
=
RT
V
A
[H
2
O
Ain
(Anode
in
C
2
I
fc
)F
H
2
O
A
C
2
I
fc
].
(21)
State equations on the cathode side:
dP
O
2
dt
=
RT
V
C
_
O
2in

C
1
2
I
fc

_
Cath
in

C
1
2
I
fc
_
F
O
2
_
dP
N
2
dt
=
RT
V
C
[N
2in
Cath
in
F
N
2
]
dP
H
2
O
C
dt
=
RT
V
C
_
H
2
O
Cin
+ C
1
I
fc
(Cath
in
+ C
1
I
fc
+ C
2
I
fc
)F
H
2
O
C
+ C
2
I
fc
_
.
(22)
For the input values H
2in
, O
2in
, and N
2in
, since the initial
mole fractions Y
H
2
, Y
O
2
, and Y
N
2
are set to be 0.99, 0.21, and
0.79, respectively [1], [3], and [9], the inputs are dened by the
mole fractions, given as:
H
2in
= Y
H
2
Anode
in
O
2in
= Y
O
2
Anode
in
N
2in
= Y
N
2
Cath
in
. (23)
The water inlet ow rates on the anode and the cathode are
expressed in terms of the relative humidity, saturation pressure,
and total pressure on each side, as follows [25]
H
2
O
Ain
=

a
P
vs
P
A

a
P
vs
Anode
in
H
2
O
Cin
=

c
P
vs
P
C

c
P
vs
Cath
in
, (24)
where
a
and
c
are the relative humidities on the anode and
the cathode sides, respectively; P
A
= P
H
2
+ P
H
2
O
A
is the sum-
mation of partial pressures of the anode; and P
C
= P
O
2
+
P
N
2
+ P
H
2
O
C
is the summation of partial pressures of the
cathode. P
vs
is the saturation pressure, which can be found
in the thermodynamics tables [30]. In addition, Anode
in
and
Cath
in
are dened as the product of input control variables u
a
and u
c
and the conversion factors k
a
and k
c
[9], [10] on each
side, which are from the standard litre per minutes (SLPM) to
mole per second:
Anode
in
= u
a
k
a
Cath
in
= u
c
k
c
. (25a)
These control variables are also affected by the hydrogen
and the air stoichiometry. However, the hydrogen stoichiome-
try does not have a big impact on the cell efciency in case of
operating with pure hydrogen [1]. Then, it can be assumed to
be constant such that a simple relationship between the fuel cell
efciency and its voltage is obtained. A typical air stoichiom-
etry ranges from 2 to 2.5, and its value is set as 2, which is
kept owing through the stack [23]. Hence, both reactants are
continuously fed to the fuel cell at sufciently high ow rate.
This fuel cell control systemis mainly dependent upon the input
control variables u
a
and u
c
. The inlet ow rates, Anode
in
and
Cath
in
, can be transformed into (25b) by multiplying each con-
stant stoichiometry to them. In designing the fuel cell control,
our main focus is on the control of hydrogen and oxygen par-
tial pressures, which can avoid unwanted pressure uctuation
and prevent the membrane electrode assemblies (MEAs) from
collapsing by minimizing the pressure difference between the
anode and the cathode, as follows:
Anode
in
= u
a
k
a

H
2
Cath
in
= u
c
k
c

air
. (25b)
By substituting (23)(25) into (21)(22), we can reconstruct
the state equations as follows.
New state equations on the anode side:
dP
H
2
dt
=
RT
V
A
_
u
a
k
a
Y
H
2

H
2
C
1
I
fc
(u
a
k
a

H
2
C
1
I
fc
)F
H
2
_
dP
H
2
O
a
dt
=
RT
V
A
_
_
u
a
k
a

H
2

a
P
vs
P
H
2
+P
H
2
O
A

a
P
vs
(u
a
k
a

H
2
C
2
I
fc
)F
H
2
O
A
C
2
I
fc
_
_
. (26)
New state equations on the cathode side:
dP
O
2
dt
=
RT
V
C
_
u
c
k
c
Y
O
2

air

C
1
2
I
fc

_
u
c
k
c

air

C
1
2
I
fc
_
F
O
2
_
dP
O
2
dt
=
RT
V
C
[u
c
k
c
Y
N
2

air
u
c
k
c

air
F
N
2
]
dP
O
2
dt
=
RT
V
C
_
_
u
c
k
c

air

c
P
vs
P
O
2
+P
N
2
+P
H
2
O
C

c
P
vs
+C
1
I
fc
(u
c
k
c

air
+C
1
I
fc
+C
2
I
fc
)F
H
2
O
C
+C
2
I
fc
_
_
.
(27)
IV. NONLINEAR CONTROL OF MIMO PEMFC
In this section, an MIMO dynamic nonlinear model of
PEMFC is derived from (26) and (27), and it is then used to
design a nonlinear controller by using feedback linearization in
order to minimize the difference P between the hydrogen and
oxygen partial pressures. The main purpose of keeping P in
a certain small range is to protect the membrane from damage,
and therefore, prolong the fuel cell stack life [2], [3]. In addi-
tion, the pressures have bigger impact on the performance of
fuel cells than other parameters [2], [3]. Because the fuel cell
voltage is a function of the pressures, each pressure needs to
be appropriately controlled to avoid a detrimental degradation
of the fuel cell voltage. To achieve this goal, it is necessary
to minimize the pressure deviation between the anode and the
cathode by using precise actuators like an accurate valve con-
troller. Normally, the optimal pressure controller consists of a
pressure sensor and a solenoid ow-control valve. In this pa-
per, we focus more on developing a pressure-control algorithm
for the whole fuel cell system instead of designing the pressure
sensors and ow controllers inside the fuel cell. In addition, the
fuel cell voltage is not considered as a control output because of
the characteristics of the VI polarization curve, which makes it
difcult to perform the voltage control in a short period of time.
Therefore, only the partial pressures of hydrogen and oxygen
are chosen as the outputs. The stack current is considered as a
disturbance to the system instead of an external input [20].
Consider the following MIMO nonlinear system with a
disturbance:

X = f(X) +
m

i=1
g
i
(X)u
i
+ p(X)d, i = 1, 2, . . . , m
y
1
= h
1
(X)
.
.
.
y
m
= h
m
(X) (28)
where X R
n
is the state vector, U R
m
is the input or
control vector, y R
P
is the output vector, and f(x) and g(x),
i = 1, 2, . . . , m, are n-dimensional smooth vector elds. The d
represents the disturbance variables, and p(x) the dimensional
vector eld directly related to the disturbance.
Considering (26) and (27), the dened outputs and the dis-
turbance, the nonlinear dynamic system model of PEMFC is
rewritten as follows:

X = f(x) + g
1
(x)u
1
+ g
2
(x)u
2
+ p(x)d
_
y
1
y
2
_
=
_
x
1
x
3
_
=
_
h
1
(x)
h
2
(x)
_
(29)
where
X =
_

_
P
H
2
P
H
2
O
A
P
O
2
P
N
2
P
H
2
O
C
_

_
; U =
_
u
a
u
c
_
; Y =
_
P
H
2
P
O
2
_
;
d = I
f ct
; f(x) = 0;
g
1
(x) = RT
H
2
_

_
k
a
Y
H
2
V
A

k
a
V
A
x
1
x
1
+ x
2
k
a

a
P
vs
V
A
(x
1
+ x
2

a
P
vs
)

k
a
V
A
x
1
x
1
+ x
2
0
0
0
_

_
g
2
(x) = RT
air

_
0
0
k
c
Y
O
2
V
C

k
c
V
C
x
3
x
3
+ x
4
+ x
5
k
c
Y
N
2
V
C

k
c
V
C
x
4
x
3
+ x
4
+ x
5
k
c

c
P
vs
V
C
(x
3
+ x
4
+ x
5

c
P
vs
)

k
c
V
C
x
5
x
3
+ x
4
+ x
5
_

_
p(x) = RT

C
1
V
A
+
C
1
x
1
V
A
(x
1
+ x
2
)
C
1
x
2
V
A
(x
1
+ x
2
)

C
1
V
A

C
1
2V
C
+
C
1
x
2
2V
C
(x
3
+ x
4
+ x
5
)
0

C
1
V
C

C
1
x
5
V
C
(x
3
+ x
4
+ x
5
)

C
2
x
5
V
C
(x
3
+ x
4
+ x
5
)
+
C
2
V
C
_

_
Equation (29) implies that the inputoutput behavior of the
system is nonlinear and coupled. In order to achieve the control
objective, two steps need to follow.
1) Obtaining a nonlinear control law that not only can com-
pensate nonlinearities but also can decouple and linearize
the input and output behaviors.
2) Imposing the poles of the closed loop so that the outputs
P
H
2
and P
O
2
track asymptotically the desired trajectory 3
atm by adding a proportional integral controller.
Fig. 3. Overall control block diagram of PEMFC.
From (29), an MIMO nonlinear system is ready to develop a
nonlinear control law. Normally, the disturbance d in (29) cannot
be directly used in the control design because an additional
necessary conditionthat the disturbance can be measured and
the feedfoward action is allowedhas to be satised [26], [34].
Otherwise, the linearized map between the new input v and the
output y does not exist. The condition renders the following
control law by using the measurement of the disturbance:
U = A
1
(x)b(x) + A
1
(x)v A
1
(x)p(x)d. (30)
As shown in (29), f(x) = 0 leads to b(x) = L
r
f
h(x) = 0, and
so the control law is written as
U = A
1
(x)v A
1
(x)p(x)d. (31)
Because each control variable u shows up after the rst deriva-
tive of each y
1
= x
1
and y
2
= x
3
, the relative degree vector
[r
1
r
2
] is [1 1], and the decoupling matrix A(x) is dened as
A(x) =
_
L
g
1
h
1
(x) L
g
2
h
1
(x)
L
g
1
h
2
(x) L
g
2
h
2
(x)
_
(32)
A(x)
=
_
k
a
Y
H
2

H
2
V
A

k
a

H
2
V
A
x
1
x
1
+x
2
0
0
k
c
Y
O
2

a i r
V
C

k
c

a i r
V
C
x
3
x
3
+x
4
+x
5
_
(33)
which is nonsingular at x = x
0
. Additionally, the matrix v and
p(x) in (31) are given as follows:
v =
_
y
1
y
2
_
p(x) = RT
_

C
1
V
A
+
C
1
x
1
V
A
(x
1
+ x
2
)

C
1
2V
C
+
C
1
x
2
2V
C
(x
3
+ x
4
+ x
5
)
_

_
.
(34)
The control law given in (30) yields decoupled and linearized
inputoutput behavior (see Fig. 3):

P
H
2
= v
1

P
O
2
= v
2
(35)
The outputs P
H
2
and P
O
2
are decoupled in terms of the
new inputs v
1
and v
2
. Thus, two linear subsystems, which are
between the input v
1
and the hydrogen partial pressure y
1
=
P
H
2
, and between the input v
2
and the oxygen partial pressure
y
2
= P
O
2
, are obtained. Furthermore, note that y
1
= x
1
and
y
2
= x
3
, and so in order to ensure that y
1
and y
2
are adjusted
to the desired values 3 (in atmospher) of y
1 S
and y
2 S
, the
stabilizing controller is designed by linear control theory using
the pole-placement strategy [27]. The new control inputs are
given by
_
v
1
v
2
_
=
_
y
1 S
k
11
e
1
y
2 S
k
21
e
2
_
where e
1
= y
1
y
1 S
and e
2
= y
2
y
2 S
.
Even though the nonlinear system PEMFC is exactly lin-
earized by feedback linearization, there may exist a tracking
error in the variation of the parameters, especially when the
load changes. To eliminate this tracking error, the integral terms
are added in the closed-loop error equation as in [27] and [34]:
_
v
1
v
2
_
=
_

_
y
1 S
k
11
e
1
k
12
_
e
1
dt
y
2 S
k
21
e
2
k
22
_
e
2
dt
_

_. (36)
From (36), the error dynamics can be obtained as follows:
e
1
+ k
11
e
1
+ k
12
e
1
= 0
e
2
+ k
21
e
2
+ k
22
e
2
= 0. (37)
By appropriately choosing the roots of the characteristics
of s
2
+ k
11
s + k
12
and s
2
+ k
21
s + k
22
, asymptotic tracking
is achieved, so that P
H
2
y
1 s
and P
O
y
2 s
as t .
The overshoots also become small by choosing k
2
11
4k
12
and
k
2
21
4k
22
[27], [34].
As shown in Fig. 3, the main objective of this control scheme
is to design a nonlinear controller by appropriately dening
a transformation mapping scheme that transforms the original
nonlinear system into a linear and controllable (closed) system,
at which point a linear controller can be designed using the
pole-placement technique for tracking purposes. However, the
control law in Eq. (30) will be unobservable because the entire
dynamics has a ve order (P
H
2
, . . . , P
H
2
O
C
), whereas only a
two order (P
H
2
and P
O
2
) are observed in the outputs. So, we
may have a problem of internal dynamics. In other words, the
internal dynamics of P
H
2
O
A
, P
N
2
, and P
H
2
O
C
must be stable
so that the states of the tracking controllers in (37) are held in
a bounded region during tracking. Otherwise, with external as
well as internal dynamics, this control law cannot enhance the
overall system performance. However, it is difcult to directly
determine the internal dynamics of the system because it is non-
linear, nonautonomous, and coupled to the external closed-loop
dynamics, as seen in (29)(37). In this paper, simulation results
are used to verify whether each state remains within the rea-
sonable bounded area. The comparison between the simulation
results and experimental data in [23] was used to verify the
performance of the control law.
V. MODEL VALIDATION AND SIMULATION RESULTS
In this section, the proposed dynamic PEMFCmodel and con-
trol method are tested through simulation in a MatlabSimulink
environment. For simplicity, the fuel processor, water and heat
management, and air compressor models are not considered in
Fig. 4. PEMFC stack based on PGS-105B system [23].
the simulation, as mentioned previously in Section III. Experi-
mental data in [23] are used to justify the validity of the proposed
dynamic PEMFC model.
The fuel cell system Ballard MK5-E-based PGS105B system
in Fig. 4 is used to test the proposed nonlinear controller. This
systemhas a total of 35 cells, connected in series, with a cell sur-
face area of 232 cm
2
. The MEA consists of a graphite electrode
and a Dow membrane. The reactant gases (hydrogen and air)
are humidied inside the stack, and the hydrogen is recirculated
at the anode while the air is owing through the cathode. The
hydrogen pressure is regulated to 3 atm at the anode inlet by a
pressure regulator, and a back-pressure regulator at the air out-
let also maintains 3 atm through the Ballard fuel cell stack. The
oxidant owrate is automatically adjusted to a constant value of
4.5 l s
1
, based on a programmable load, via a mass owmeter
to ensure sufcient water removal at the cathode. Hydrogen is
replenished at the same rate as it is consumed. The stack temper-
ature measured at the air outlet is maintained between 72

Cand
75

Cto produce the maximumpower output. The simplied di-


agram of the Ballard system is depicted in Fig. 4. The PGS105
systemis more likely to be a presetting control systeminstead of
a feedback control system. In the case of outer ranges of settings,
the system automatically shuts down, so a higher level control
system is thus needed. Here, we briey outline the experimental
setup of our nonlinear control for the fuel cell for future study.
To implement our control scheme in the practical system, the
sensors to measure the states P
H
2
, . . . , P
H
2
O
c
, the stack current,
and the voltage must be installed, and a main controller such
as a digital signal processor (DSP) needs to be implemented
based on the nonlinear control law obtained from the feedback
linearization. With an DSP, the fuel cell can communicate with
the electronics rack and the PC for the user interface monitoring
shown in Fig. 5. The safety concerns for the fuel cell system
must be solved with high priority before a control scheme is
implemented. At least, the safety issues related to low cell volt-
age, stack overload, high temperature, hydrogen leaking, and
pressure differences between the anode and the cathode must
be considered in the fuel cell control system to prevent a fatal
accident and severe damage to the fuel cell system.
To control the entire fuel system, the sensors and the tempera-
ture and humidity control actuators must be added to the system.
In the simulation, since the temperature and the humidity in the
system have a very slow response time, a perfectly controlled
Fig. 5 Simplied experiment set up for the PEMFC control.
humidier and heat exchanger are assumed to be applied. In ad-
dition, an automatic-purge-controller for the hydrogen exhaust
is required, and with respect to the air and water exhaust, the
back-pressure regulator must be coordinated with the air inlet
ow-rate input as well as the hydrogen inlet ow-rate input to
maintain the pressures for hydrogen and air at the same level.
If we assume that the electrical capacitor is 1 F [1] and the
sum of the activation and concentration resistance is from 0.2 to
0.3 [37], the time constant of the fuel cell stack is approxi-
mately within 0.20.3 s. With the use of a fast actuator on the
cathode, whose performance setting time is 1090% of 50 ms,
and another actuator on the anode whose performance setting
time is 1090% of less than 20 ms, it becomes possible to main-
tain the anode and cathode pressures at a certain level. To ensure
this performance, the anode pressure controller must be three
times faster than the one on the cathode because the anode pres-
sure follows the cathode side pressure [37]. Currently, these fast
actuators are available in the market [38], yet in reality, due to the
uncertainties in the fuel cell model parameters and inaccuracy
in the measurements, we may encounter a serious obstacle in
achieving the desired responses. Moreover, the time delay of ac-
tuators, sensors, compressor, and sampling period for the DSP is
unavoidable for the design of the nonlinear control of PEMFCs.
Even though we could not exactly calculate the time delays, they
can still be compensated for by adapting a lead compensator in
practical systems. However, since the lead compensation is very
sensitive to noise, a multiple rst-order, low-pass noise lter
is recommended [37]. In our simulation, we do not consider
the time delays and other problems like uncertainty and inaccu-
racy in measurements, which will be considered in our future
investigations.
The nominal values of parameters for the simulation are
shown in the Table II.
Experimental data from Hamelin et al. [23] were used to jus-
tify the validity of the proposed dynamic model of the PEM fuel
cells. The voltage and the current, the most important variables,
were used for this comparison. In [23], a load prole with rapid
variations between 0 and 150 A was imposed on the PGS-105B
system. The corresponding stack current and voltage transients
are plotted in Fig. 6, where the experimental data [23] are indi-
cated by solid lines.
TABLE II
PEMFC PARAMETERS FOR THE SIMULATION
Fig. 6. Voltage and current under load variations.
The details of load prole are shown in Fig. 7, where the
load resistances were changed from 0.145 to 4.123 during the
simulation period. Fig. 6 shows a good agreement between the
experimental data and the simulated results, except for the time
periods between 13 and 15 s, and 25 and 29 s. The main reason
of the voltage undershoot during these periods is the internal
resistance variation of the PEM fuel cell systems [1]. The rapid
current increase can cause an immediate voltage drop across the
internal resistor of the fuel cell, which is closely related to the
temperature change that conversely affects this resistance. In our
simulation, the stack temperature was assumed to be constant
at 353 K. A small discrepancy between simulation results and
experimental data is inevitable under the rapid current changes,
so to compare the efciency of the proposed nonlinear controller,
the linear controller (PI controller) is also implemented for the
fuel cell system.
Fig. 7. Load variation prole.
Fig. 8. Fuel cell power demand under load variations.
Fig. 9. Variations of hydrogen pressure.
To achieve fewer overshoot results, the feedback gains k
11
,
k
21
, k
12
, and k
22
in Eq. (37) have to be tuned to 5 and 1,
respectively, which are the optimal values in their feasible ranges
[0.1 10] for k
11
and k
21
and [0 10] for k
12
and k
22
, and being
beyond the ranges can easily cause a violation of the Matlab
simulation limits.
For the voltage, current, and power shown in Figs. 6 and 8,
the discrepancies between the nonlinear control and the linear
control are not obvious due to the fast response times, which are
less than a few milliseconds. However, other simulation results
in Figs. 912 indicate that better transient performances are
observed when using a nonlinear controller because the linear
controller is more dependent upon the operating point, while the
nonlinear controller is independent of the operating point due
to the feedback linearization by the differential geometry [39].
Fig. 10 shows that the oxygen partial pressure has much bigger
overshoot than the hydrogen partial pressure, which implies
that the oxygen partial pressure is more sensitive to the load
variation than the hydrogen partial pressure. Fig. 11 displays
the absolute value of the difference between the hydrogen and
oxygen partial pressures. It is found in Fig. 11 that the nonlinear
Fig. 10. Variations of oxygen pressure.
Fig. 11. Variations of pressure difference of H
2
and O
2
.
Fig. 12. Variations of hydrogen ow rate.
controller has a better transient response than linear controller.
Generally, an increase in the stack current causes a decrease in
reactant pressures because more fuel consumption is required.
However, the ow rates vary with the stack current in the same
way and compensate for the increased fuel consumption.
Figs. 12 and 13 give the responses of the hydrogen and the
oxygen ow rates under the load variations. The hydrogen ow
rate varies between 0[slpm] to 5[slpm], while the oxygen ow
rate varies from 0[slpm] to 16[slpm]. It is observed that the oxy-
gen ow rate has much bigger variations than hydrogen because
the oxygen ow rate is more sensitive to the load variation than
the hydrogen owrate, as was seen with the pressure variations.
Although the nonlinear controller has slightly more overshoot
in the oxygen ow rate than the linear controller in Fig. 13,
it shows that the response time of the nonlinear controller is
faster than the linear controller. Figs. 1416 show that the states
P
H
2
O
A
, P
H
2
O
c
, and P
N
2
are stable under the load variations.
Fig. 13. Variations of oxygen ow rate.
Fig. 14. Variations of water partial pressure on the anode.
Fig. 15. Variations of water partial pressure on the cathode.
Fig. 16. Variations of nitrogen partial pressure on the cathode.
Figs. 15 and 16 show that the nonlinear controller and the
linear controller behave almost the same in the water partial
pressure on the cathode side and in the nitrogen partial pressure.
In addition, even though no tracking controllers are considered
in the design of the nonlinear and linear controls, Figs. 1416
show that P
H
2
O
A
, P
N
2
, and P
H
2
O
c
vary within bounded ranges
(0 to 4 atm) under the load variations. Hence, the internal
dynamics problem of the nonlinear tracking controller in our
design is not a concern. It is observed that the water partial
pressure on the anode side remains a little more stable than the
water partial pressure at the cathode and the nitrogen pressure.
Also, the water partial pressure on the anode side follows a
pattern similar to that of the hydrogen partial pressure because
of the high mole fraction of hydrogen, which is about 99%.
Therefore, we can conclude that the load variations have more
inuence on the cathode side than on the anode side, which
implies a more sophisticated control than the one proposed in
this paper needs to be applied on the cathode side at least.
VI. CONCLUSION
This paper proposes a dynamic nonlinear model for the PEM
fuel cells and presents a design for a nonlinear control for
PEMFCs by feedback linearization. By introducing additional
states of the partial pressure of water on the anode side and
nitrogen on the cathode, we were able to include transient be-
haviors of the PEMFC in the model. An MIMO nonlinear dy-
namic model is developed for PEMFCs that creates the platform
for the design of a nonlinear control strategy by feedback lin-
earization. The nonlinear control design was implemented in
MatlabSIMULINK. Simulation results show that the nonlin-
ear controller has better transient responses than the linear con-
troller under load variations, and because of its superb transient
behavior, the proposed nonlinear control strategy should be ex-
tended to the design of an overall control scheme for PEMFCs,
including stack voltage control and control of water and heat
management, the fuel processor, and the air compressor.
REFERENCES
[1] J. Larminie and A. Dicks, Fuel Cell Systems Explained. New York:
Wiley, 2002.
[2] W. Yang, B. Bates, N. Fletcher, and R. Pow, Control challenges and
methodologies in fuel cell vehicle development, presented at the SAE,
1998, Paper 98C054.
[3] J. Purkrushpan, A. G. Stefanopoulou, and H. Peng, Control of fuel cell
breathing, IEEE Control Syst. Mag., vol. 24, no. 2, pp. 3046, Apr. 2004.
[4] J. Purkrushpan and H. Peng, Control of Fuel Cell Power Systems: Prin-
ciple, Modeling, Analysis and Feedback Design. Berlin, Germany:
Springer-Verlag, 2004.
[5] J. Purkrushpan, A. G. Stefanopoulou, and H. Peng, Modeling and control
for PEMfuel cell stack system, in Proc. Amer. Control Conf., Anchorage,
AK, 2002, pp. 31173122.
[6] J. C. Amphlett, R. M. Baumert, R. F. Mann, B. A. Peppy, P. R. Roberge,
and A. Rodrigues, Parametric modeling of the performance of a 5-kW
proton exchange membrane fuel cell stack, J. Power Sources, vol. 49,
pp. 349356, 1994.
[7] J. C. Amphlett, R. M. Baumert, R. F. Mann, B. A. Peppy, and P. R. Roberge,
Performance modeling of the Ballard Mark IV solid polymer electrolyte
fuel cell, J. Electrochem. Soc., vol. 142, no. 1, pp. 915, 1995.
[8] R. F. Mann, J. C. Amphlett, M. A. Hooper, H. M. Jesen, B. A. Peppy, and
P. R. Roberge, Development and application of a generalized steady-state
electrochemical model for a PEM fuel cell, J. Power Sources, vol. 86,
pp. 173180, 2000.
[9] M. J. Khan and M. T. Labal, Dynamic modeling and simulation of a fuel
cell generator, Fuel Cells, vol. 5, no. 1, pp. 97104, 2005.
[10] M. J. Khan and M. T. Labal, Modeling and analysis of electro chemical,
thermal, and reactant ow dynamics for a PEM fuel cell system, Fuel
cells, vol. 5, no. 4, pp. 463475, Apr. 2005.
[11] C. J. Hatiziadoniu, A. A. Lobo, F Pourboghrat, and M. Daneshdoot, A
simplied dynamic model of grid connected fuel-cell generators, IEEE
Trans. Power Del., vol. 17, no. 2, pp. 467473, Apr. 2002.
[12] M. Y. El-Sharkh, A. Rahman, M. S. Alamm, A. A. Sakla, P. C. Byrne,
and T. Thomas, Analysis of active and reactive power control of a stand-
alone PEM fuel cell power plant, IEEE Trans. Power Del., vol. 19, no. 4,
pp. 20222028, Nov. 2004.
[13] P. Famouri and R. S. Gemmen, Electrochemical circuit model of a PEM
fuel cell, in Proc. IEEE Power Eng. Soc. Gen. Meet., Jul. 2003, vol. 3,
pp. 1317.
[14] L. You and H. Liu, A parametric study of the cathode catalyst layer of
PEM fuel cells using a pseudo homogeneous model, Int. J. Hydrogen
Energy, vol. 26, no. 9, pp. 991999, Sep. 1991.
[15] G. Maggio, V. Recupero, and L. Pino, Modeling polymer electrolyte
fuel cells: An innovative approach, J. Power Sources, vol. 101, no. 2,
pp. 275285, Oct. 2001.
[16] A. Rowe and X. Li, Mathematical modeling of proton exchange mem-
brane fuel cells, J. Power Sources, vol. 102, no. 1/2, pp. 8296, Dec.
2001.
[17] C. Wang, M. H. Nehrir, and S. R. Shaw, Dynamic model and model
validation for PEM fuel cells using electrical circuits, IEEE Trans.
Energy Convers., vol. 20, no. 2, pp. 442451, Jun. 2005.
[18] A Sakhare, A Davari, and A Feliachi, Fuzzy logic control of fuel cell for
stand-alone and grid connection, J. Power Sources, vol. 135, no. 1/2,
pp. 165176, Sep. 2004.
[19] P. E. M. Almeida and M. Godoy, Neural optimal control of PEM fuel
cells with parametric CMAC network, IEEE Trans. Ind. Appl., vol. 41,
no. 1, pp. 237245, Jan./Feb. 2005.
[20] L. Y. Chiu, B. Diong, and R. S. Gemmen, An improve small-signal mode
of the dynamic behavior of PEM fuel cells, IEEE Trans. Ind. Appl.,
vol. 40, no. 4, pp. 970977, Jul./Aug. 2004.
[21] J. Sun and V. Kolmannovsky, Load governor for fuel cell oxygen
starvation protection: A robust nonlinear reference governor approach,
IEEE Trans. Control Syst. Technol., vol. 3, no. 6, pp. 911913, Nov.
2005.
[22] R. S. Gemmen, Analysis for the effect of inverter ripple current on fuel
cell operating conditions, Trans. ASME, vol. 125, pp. 576585, May
2003.
[23] J. Hamelin, K. Abbossou, A. Laperriere, F. Laurencelle, and T. K. Bose,
Dynamic behavior of a PEM fuel cell stack for stationary application,
Int. J. Hydrogen Energy, vol. 26, pp. 625629, 2001.
[24] J. M. Correa, F. A. Farret, and L. N. Canha, An analysis of the dynamic
performance of proton exchange membrane fuel cells using an electro-
chemical model, in Proc. 27th Annu. Conf. IEEE Ind. Electron. Soc.
IECON 2001, vol. 1, pp. 141146.
[25] F. Barbir, PEM Fuel Cells: Theory and Practice. London, U.K.:
Elsevier, 2005.
[26] A. Isidori, Nonlinear Control Systems, 3rd ed. London, U.K.: Springer-
Verlag, 1995.
[27] J. J. E. Slotine and W. Li, Applied Nonlinear Control. Englewood Cliffs,
NJ: Prentice-Hall, 1991.
[28] Q. Li, Y. Sun, and S. Mei, Nonlinear Control Systems and Power System
Dynamics. Norwell, MA: Kluwer, 2001.
[29] B. W. Bequette, Nonlinear control of chemical process: A review, Ind.
Eng. Chem., vol. 30, pp. 13911413, 1991.
[30] R. E. Sonntag, C. Borgnakke, and G. J. V. Wylen, Fundamentals of
Thermodynamics. New York: Wiley, 1998.
[31] M. Bar ao, J. M. Lemos, and R. N. Silva, Reduced complexity adaptive
nonlinear control of a distributed collector solar eld, J. Process Control,
vol. 12, no. 1, pp. 131141, 2002.
[32] Ballard Power System, Inc., Burnaby, BC, Canada [Online]. Available:
http:/www.ballard.com/
[33] U.S. Department of Energy, Fuel Cell Handbook, National Energy Tech-
nology Laboratory (NETL), Morgantown, WV, Oct. 2000.
[34] M. A. Henson and D. E. Seborg, Critique of exact linearization strategies
for process control, J. Process Control, vol. 1, pp. 122139, May 1991.
[35] M. Serra, J. Auado, X. Ansede, and J Riera, Controllability analysis
of decentralized linear controllers for polymeric fuel ccells, J. Power
Sources, vol. 151, pp. 93102, 2005.
[36] A. Isidori, Nonlinear Control Systems: An Introduction, 2nd ed. New
York: Springer-Verlag, 1989.
[37] C. A. Anderson, M. O. Christensen, A. R. Korsgaad, M. Nielsen, and
P. Pederson, Design and control of fuel cells system for transport appli-
cation, Aalborg University, Aalborg, Denmark, Project report, 2002.
[38] Burkert [Online]. Available: http://www.bci,buerkert.com/product
Woon Ki Na (S00) received the B.S. and the M.S.
degrees in electrical engineering from Kwangwoon
University, Seoul, Korea, in 1995 and 1997, respec-
tively. He is currently working toward the Ph.D. de-
gree from the Electrical Engineering Department,
University of Texas, Arlington.
During 19972001, he was a Research Engineer
in the R&D center, T. H. ELEMA Company Ltd.,
Shiheung, Korea, and LG Electronics Inc., Seoul,
where he was engaged in research on inverter con-
trol systems and multimedia systems. His current re-
search interests include power electronics, power systems, and the nonlinear
control and interface design of fuel cells.
Bei Gou (M00) received the B.S. degree in electrical
engineering fromthe North China University of Elec-
tric Power, Beijing, China, in 1990, the M.E.(elect.)
degree from Shanghai JiaoTong University, Shang-
hai, China, 1993, and the Ph.D. degree from Texas
A&M University, Arlington, in 2000.
From 1993 to 1996, he was in the Department
of Electric Power Engineering, Shanghai Jiao Tong
University. During 1997, he was a Research Assis-
tant at Texas A&M University. He was with the ABB
Energy Information Systems, Santa Clara, CA, for
two years and at ISO New England for one year as a Senior Analyst. He is
currently an Assistant Professor in Energy Systems Research Center (ESRC),
University of Texas. His current research interests include power system state
estimation, renewable energy, power market operations, power quality, power
system reliability, and power system blackouts.

You might also like