Vous êtes sur la page 1sur 13

Design and optimization of a combined fuel reforming and solid oxide fuel

cell system with anode off-gas recycling


Tae Seok Lee
a
, J.N. Chung
a,
, Yen-Cho Chen
b
a
Department of Mechanical and Aerospace Engineering, University of Florida, Gainesville, FL 32611, USA
b
Department of Energy and Resource Engineering, National United University, Miaoli City 360, Taiwan
a r t i c l e i n f o
Article history:
Received 21 June 2010
Received in revised form 8 May 2011
Accepted 11 May 2011
Available online 22 June 2011
Keywords:
SOFC
Steam reforming
Fuel processor
Anode off-gas
System efciency
Dodecane
a b s t r a c t
An energy conversion and management concept for a combined system of a solid oxide fuel cell coupled
with a fuel reforming device is developed and analyzed by a thermodynamic and electrochemical model.
The model is veried by an experiment and then used to evaluate the overall system performance and to
further suggest an optimal design strategy. The unique feature of the system is the inclusion of the anode
off-gas recycle that eliminates the need of external water consumption for practical applications. The sys-
tem performance is evaluated as a function of the steam to carbon ratio, fuel cell temperature, anode off
gas recycle ratio and CO
2
adsorption percentage. For most of the operating conditions investigated, the
system efciency starts at around 70% and then monotonically decreases to the average of 50% at the
peak power density before dropping down to zero at the limiting current density point. From an engi-
neering application point of view, the proposed combined fuel reforming and SOFC system with a range
of efciency between 50% and 70% is considered very attractive. It is suggested that the optimal system is
the one where the SOFC operates around 900 C with S/C ratio higher than 3, maximum CO
2
capture, and
minimum AOG recirculation.
2011 Elsevier Ltd. All rights reserved.
1. Introduction
Solid oxide fuel cells (SOFCs) can provide clean and highly ef-
cient power for a wide spectrum of small to large-scale applica-
tions. SOFCs are expected to be around 5060% efcient at
thermal efciency. In applications designed to capture and utilize
the high-temperature waste heat through a co-generation system,
overall combined efciencies could top 8085%. SOFCs operate at
very high temperaturesaround 8001000 C and they are also
the most sulfur-resistant fuel cell type; they can tolerate several
orders of magnitude more of sulfur than other fuel cell types. In
addition, they are not poisoned by carbon monoxide (CO), which
can even be used as fuel [1,2].
Hydrogen, a very versatile fuel, can be produced from various
materials and by several methods. Industrial scale production of
hydrogen has been operational in the oil and gas industry for more
than a century and forms the base of the modern chemical indus-
try. In centralized facilities and/or distributed generation, hydro-
gen has to be supplied to stationary fuel cell applications from
nearby hydrogen mass-production processes. For mobile applica-
tions, it is difcult to store and handle hydrogen directly as an
on-board fuel. Moreover, not only the safe hydrogen storage tech-
nologies, but also its infrastructures are needed. However, the lack
of an infrastructure, storage difculties and related safety issues
can only be solved unilaterally by introducing an on-board fuel
processor which converts a commercially available fuel, such as
gasoline or diesel, into hydrogen such that it is available on de-
mand [3,4].
The development of portable fuel processors has however for
several years received signicant attention in relation to the in-
crease in worldwide fuel cell development activities. The research
and development work on fuel processors for fuel cells could be
classied according to different approaches (or methodologies)
and fuel selection. The most considered fuels are alcohols [5,6],
gasoline [7], diesel [3,4], and single or mixture of heavy-hydrocar-
bons as surrogates for (conventional) petroleum based liquid fuel
[6,810] and jet fuel [11]. There are three established methods
for reforming fuels: steam reforming (SR) [5,8,12,13], partial oxida-
tion (POX) [11] and auto-thermal reforming (ATR) [3,4,7,9,1113].
All of the mentioned methods produce a syngas mixture; however
the difference in reaction temperatures and oxidants yields differ-
ent CO concentrations in the syngas mixture. The H
2
production
generally decreases in the order of steam reforming, auto-thermal
reforming, and partial oxidation.
In this work, we investigate a combined fuel reforming and
SOFC system and present an in depth analysis and evaluation of
this system that utilizes n-dodecane as the surrogate for diesel
fuels. Most previous studies [35,12] consider low-temperature
fuel cells due to its vehicle-oriented purpose, such as the Auxiliary
0196-8904/$ - see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.enconman.2011.05.009

Corresponding author.
E-mail address: jnchung@u.edu (J.N. Chung).
Energy Conversion and Management 52 (2011) 32143226
Contents lists available at ScienceDirect
Energy Conversion and Management
j our nal homepage: www. el sevi er . com/ l ocat e/ enconman
Power Unit for heavy trucks. However, our application is not re-
stricted by the operating temperature of the fuel cell system. Car-
bon monoxide, the second most produced chemical species in the
reformate, should be either converted into hydrogen by watergas
shift (WGS) reaction or removed by a purication process for Pro-
ton exchange membrane fuel cells (PEMFC) to prevent CO poison-
ing of the anode catalyst [1]. On the contrary, it is unnecessary to
have another CO converting or purifying process for a SOFC. Carbon
monoxide can be utilized in the anode of a SOFC as a fuel by either
direct oxidation or WGS reaction to produce more hydrogen. In this
paper, we focus on the application of a SOFC in an air-decient
environment such as in the aerospace and under-water applica-
tions, rather than on lowering the operating temperatures. It is
noted that for aerospace and under-water applications, the avail-
able space is a main limiting factor. Selecting the most suitable
reforming technology depends on both the application require-
ments and the available design options. The steam reforming is
adopted here as no oxygen is required and it facilitates the highest
hydrogen production [12,13].
In this work, an analytical, parametric study is performed to
evaluate the feasibility and performance of a combined fuel
reforming and SOFC system that utilizes n-dodecane as the surro-
gate for diesel fuels. Specically the effects of adding the anode off-
gas recycling and recirculation components and the CO
2
absorbent
unit are investigated.
2. Thermodynamic and electro-chemical model
2.1. Combined fuel reforming and SOFC system
Fig. 1 shows the process ow diagram of the combined system.
The two major components, an external-reformer unit and a SOFC
unit, are coupled to convert the liquid hydrocarbon fuels to elec-
trical power. The entire reforming unit consists of a direct inter-
nal reforming component inside the SOFC, an external reformer,
four heat exchangers, a carbon dioxide adsorbent, a ue gas con-
denser and an anode off gas (AOG) recycling unit. The AOG has a
very high temperature and generally also contains high water va-
por content due to the fact that water vapor is the only product of
the electrochemical reaction which produces the electrical power
in a SOFC. The management of the AOG is of high importance to
the overall system efciency and maintaining the designed steam
to carbon ratio (S/C). With an appropriate amount of recycled
AOG, it is possible to maintain the prescribed steam to carbon ra-
tio value at the inlet of the external reformer. To provide a given
steam to carbon ratio (S/C) at the inlet of the external reformer,
the total amount of steam required is solely supplied by the recy-
cled AOG. So the S/C is closely related to AOG recycle ratio. It is
important to note that AOG recycling also accomplish the purpose
of conserving water because only when the AOG does not contain
enough water vapor, water would then be added with the fuel to
maintain the desired steam to carbon ratio. It is obvious that a
compressor is required to accomplish the AOG recycle process.
From the thermodynamics point of view, the high temperature
compression process requires much more work than a low tem-
perature process. Therefore, the AOG is cooled down before the
compressing process. Accordingly, a cooler is used for this pur-
pose. After the compressor, the temperature of the AOG is raised
back by passing it through a recuperator heat exchanger. How-
ever, a certain portion of the AOG is not recycled and is sent di-
rectly to the after-burner instead where it is burned that provides
heat to the reformer in an effort to retrieve more energy. The ue
gases from the after-burner are designed to pre-heat the fuel and
to provide more heat to the reformer.
Nomenclature
Abbreviation
AOG anode off gas
ATR autothermal reforming
LHV lower heating value
MCFC molten carbonate fuel cell
PEMFC proton exchange membrane fuel cell
POX partial oxidation
S/C steam to carbon ratio
SOFC solid oxide fuel cell
SR steam reforming
WGS watergas shift
YSZ yttria-stabilized zirconia
List of symbols
D
eff
effective gaseous diffusivity (cm
2
/s)
F Faraday constant (96485.3 C/mol)
F
i
molar ow rate of labeled i stream (mol/s)
G Gibbs free energy (J/mol)
i current density (A/cm
2
)
i
o
exchange current density (A/cm
2
)
i
s
limiting current density (A/cm
2
)
K equilibrium constant (equivalent)
L electrode thickness (lm)
NC number of carbon in the fuel ()
n
ij
number of moles changed for jth species in ith reaction
(equivalent)
P pressure (atm)
P
i
partial pressure of ith species (atm)
R gas constant (8.314 J/molK)
Rc recycle ratio ()
R
contact
contact resistance (X)
r recirculation ratio ()
T temperature (K)
U fuel utilization factor ()
V voltage (Volt)
W power (watt)
y
i
molar fraction for ith species ()
Greek
e
i
extent of ith reaction (equivalent)
e effectiveness ()
g efciency ()
m
ij
stoichiometric coefcient for jth species in ith reaction
()
q electrical resistivity (Xcm)
Subscripts
A anode
AB after-burner
act activation overpotential
C cathode
conc concentration overpotential
E electrolyte
elect electricity
f fuel
N Nernst potential
ohm Ohmic overpotential
T.S. Lee et al. / Energy Conversion and Management 52 (2011) 32143226 3215
The molar balance including electro-chemical reactions as well
as chemical equilibriumreactions is considered next. After evaluat-
ing the molar balance, an energy balance is performed for each
component.
2.2. Assumptions and Justication
The following assumptions are made in the analysis:
Steady state operation.
All gaseous phases are ideal gas.
Chemical and thermal equilibrium are achieved due to enough
residence time.
Only hydrogen is electrochemically reacted at the anode inside
the SOFC.
Pressure drop is neglected through the system.
Anode-supported SOFC
Pure oxygen provision for the cathode
No carbon deposition in the entire system.
Complete conversion of the fuel at the external reformer.
Since the current paper is basically dealing with the introduc-
tion of a conceptual energy management system design that is
intended to demonstrate the new concept, functions, and charac-
teristics of the coupled SOFC and reforming system. However, it
is not aimed at the detailed modeling of each component. Each
component is a unit and it is represented by the key characteristics.
So, we have adopted some simplifying ideal conditions as listed
above. The pure oxygen for the cathode is one of them. We believe
that these idealizations would not be likely to change the major
outcomes and conclusion in this paper if more realistic assump-
tions were used. For some special applications where air is not
readily available such as underwater vehicles, and aero and space
applications, pure oxygen is usually used. Since our external refor-
mer utilizes only steam not oxygen, it is the steam reforming of the
fuel in our case.
There are possibly some other species formed, but the key and
dominant products are CO and H
2
in our reformer. The term an-
ode-supported SOFC denotes the anode is the thickest part among
the anode, electrolyte, and cathode so that the anode can provide
enough mechanical strength for the fuel cell structure. We as-
sumed enough residence time for reactants to react completely.
This is another one of those ideal conditions. We also assumed
no carbon deposition which can be justied for the conditions of
steam to carbon ratio investigated in this work [14,15].
2.3. Molar balance and chemical equilibrium
As shown in Fig. 1, the reformer unit includes two parts, an
external reformer and an after-burner. The pre-heated mixture of
fuel and recycled AOG, mostly water, is passing through the
reforming channels of the external reformer. A reasonable way of
approaching the liquid hydrocarbon reforming modeling is by pre-
dicting the thermodynamic equilibrium composition through
Gibbs free energy minimization calculation under the presence of
catalysis, and given pressure and temperature. In the presence of
catalysis, steam reforming reaction, which is shown in Eq. (1),
takes place and the readjustment of chemical species is achieved
by chemical reaction equilibrium under the given temperature
and pressure.
C
n
H
m
n SCH
2
O !m
CO
CO m
H
2
H
2
m
CO
2
CO
2
m
CH
4
CH
4
m
H
2
O
H
2
O 1
Hydrocarbon reforming reaction, Eq. (1), is assumed irreversible
and relatively fast under the certain catalyst as well as completely
converted. In the hydrocarbon reforming chemical reaction, stoi-
chiometry coefcients for the product chemicals are arbitrary at
this time. Even though stoichiometry coefcients are arbitrary,
we are not dealing with nuclear reaction so each coefcient should
satisfy three atomic balances, i.e. carbon, hydrogen and oxygen
atomic balances. With three atomic balances, we still have an
underdetermined condition, i.e. ve unknowns and three equa-
tions. Here, chemical equilibrium assumption plays a key role.
Fig. 1. Process ow diagram.
3216 T.S. Lee et al. / Energy Conversion and Management 52 (2011) 32143226
From the classical thermodynamics, two independent chemical
reaction equations can be composed from one phase (gaseous
phase), ve chemical species (CO, H
2
, CO
2
, CH
4
, and H
2
O) and three
component atoms (C, H, and O) and those reactions are as follow,
CH
4
H
2
OCO 3H
2
2
CO H
2
OH
2
CO
2
3
The rst reaction, Eq. (2), is the well known methane steam
reforming reaction and the second one, Eq. (3), is the well known
watergas shift reaction. This pair of chemical reactions is uniquely
set for the given condition, one phase, ve chemical species, and
three atomic components. If we consider carbon deposition then
we should add solid carbon as a deposited material on the catalyst
surface. With solid carbon formation, we need one more chemical
reaction to solve the stoichiometry of steamreforming reaction, Eq.
(1). However, carbonhydrogenoxygen (CHO) species diagram
shows that carbon deposition could be eliminated by providing en-
ough water and/or oxygen into the system [14,15]. For our lowest
steam to carbon ratio operating condition the products are located
underneath the carbon deposition boundary line in the CHO
phase diagram which means that the above no carbon deposition
assumption is valid theoretically. Therefore we do not need to con-
sider carbon deposition, so we have a unique set of independent
equilibrium reactions, Eqs. (2) and (3). Let the extents of reactions,
which are dened as de
i
= dn
ij
/m
ij
, shown in Eqs. (2), (3) be e
R1
and
e
R2
, respectively. Then equilibrium molar fractions are expressed as
follows:
y
H
2
O

F
o;H
2
O
e
R1
e
R2
F
o
2e
R1
4a
y
CH
4

F
o;CH
4
e
R1
F
o
2e
R1
4b
y
H
2

F
o;H
2
3e
R1
e
R2
F
o
2e
R1
4c
y
CO

F
o;CO
e
R1
e
R2
F
o
2e
R1
4d
y
CO
2

F
o;CO
2
e
R2
F
o
2e
R1
4e
where F
o
and F
o,i
denote the inlet total molar ow rate, labeled 3b in
Fig. 1, and the inlet molar ow rate of ith component, respectively.
Once the extents of reactions are obtained, evaluation of the equi-
librium molar fraction is straightforward. Therefore, we need two
equations to be solved simultaneously to nd e
R1
, and e
R2
under
the equilibrium condition. These are the chemical equilibrium
equations corresponding to the steam reforming and watergas
shift reaction, as given below:
K
R1
T
P
P
o
_ _
2
y
3
H
2
y
CO
y
CH
4
y
H
2
O
exp
DG
R1
RT
_ _
5a
K
R2
T
P
P
o
_ _
0
y
H
2
y
CO
2
y
H
2
O
y
CO
exp
DG
R2
RT
_ _
5b
The temperature-dependent equilibrium constant is obtained
by the classical method in which the change in Gibbs free energy
of the reactions is used. After the Gibbs energy difference is ob-
tained, equilibrium molar fractions, Eq. (4), are substituted into
Eq. (5) and then the system of equations, i.e. three atomic balances
and two extent of reactions, is solved by the NewtonRahpson
method with a tolerance of 10
7
.
In this work, the direct internal reforming SOFC model is based
on the one proposed by Colpan et al. [15,16]. In this model, CO and
methane are assumed to react with the water vapor for producing
more hydrogen in the anode. However, the total amount of electri-
cal energy produced in the current system comes from watergas
shift reaction and methane steam reforming reaction in addition to
direct oxidation in the anode. Therefore, methane steam reforming
reaction, Eq. (2), watergas shift reaction, Eq. (3), and electrochem-
ical reaction, Eq. (6), occur simultaneously inside the direct inter-
nal reforming section of the SOFC.
H
2

1
2
O
2
!H
2
O 6
The extent of electrochemical reaction, e
S3
, can be expressed
using a molar balance, denition of molar fraction, and recircula-
tion ratio.
e
S3

UF
o;H
2
3e
S1
e
S2

1 r rU
7
Here, r is the recirculation ratio, U is fuel utilization, e
Si
is reac-
tion coordinates for ith reaction at the SOFC, respectively. Also, F
o
denotes the inlet of SOFC anode, labeled 5 in Fig. 1. With the above
assumptions and Eq. (7), molar fractions for all the species at the
exit of the anode of the fuel cell are given as below:
y
CH
4

F
o;CH
4
e
S1
F
o
2e
S1
8a
y
H
2
O

F
o;H
2
O
e
S1
e
S2

F
o;H
2
3e
S1
e
S2
1rrU
_ _
U
_ _
F
o
2e
S1
8b
y
H
2

F
o;H
2
3e
S1
e
S2
F
o
2e
S1
_ _
1 r1 U
1 r rU
8c
y
CO

F
o;CO
e
S1
e
S2
F
o
2e
S1
8d
y
CO
2

F
o;CO
2
e
S2
F
o
2e
S1
8e
Here, e
S1
and e
S2
are the only unknowns. With likewise chemical
equilibrium assumed in the external reformer, molar fractions at
the exit of the SOFC anode (Eq. (8)) are evaluated using Eqs. (5a)
and (5b).
2.4. SOFC model
Once the SOFC exit composition is obtained based on thermody-
namic equilibrium, calculation of the amount of energy conversion
from chemical to electrical is straightforward by means of an elec-
tro-chemical relationship. The terminal voltage of the cell can be
obtained after considering all voltage losses; V = V
N
V
ohm

V
act
V
conc
. Here, the Nernst voltage is calculated as,
V
N

DG
o
T
SOFC

2F

RT
SOFC
2F
ln
y
H
2
O;A
y
H
2
;A

y
O
2
;C
P=P
o
_
_
_
_
_
_
_ 9
Here, three types of overpotentials, e.g. ohmic, activation and
concentration, are considered and calculated using Eqs. (11),
(13), and (15). The Ohmic overpotential occurs because of resis-
tance to the ow of ions in the electrolyte and resistance to the
ow of electrons through the electrically conductive fuel cell
components.
V
ohm
iR
contact

k
q
k
L
k
10
T.S. Lee et al. / Energy Conversion and Management 52 (2011) 32143226 3217
where i is the current density, q is the electrical resistivity (inverse
of electrical conductivity), L is the thickness of electrode, and R
contact
is the contact resistance. Electrical resistivity is given as a function
of temperature for SOFC materials [17] e.g. lanthanum manganite
for cathode, YSZ for electrolyte, and Ni/ZrO
2
for anode; q
C
=
0.008114 exp(600/T) (Xcm), q
E
= 0.00294 exp(10,350/T) (X
cm), q
A
= 0.00298 exp(1392/T) (Xcm). Please note that Ref.
[15] provides the resistivity for Ni/ZrO
2
which we used for Ni/YSZ
as our anode material as the two are very close in resistivity values.
Ref. [18] also adopted the same approximation.
Activation overpotential is associated with the sluggish elec-
trode kinetics occurring at the electrodeelectrolyte interfaces.
ButlerVolmer equation with 0.5 transfer coefcient is used for
the activation overpotential.
V
act
V
act;A
V
act;C

RT
SOFC
F
sinh
1
i
2i
o;A
_ _

RT
SOFC
F
sinh
1
i
2i
o;C
_ _
11
where F is the Faraday constant, i
o,A
and i
o,C
are exchange current
density for anode and cathode, respectively. These exchange current
densities are a function of temperature, which are related to the
charge transfer resistance; i
oi
= A
i
exp(B
i
/RT). Coefcients A
i
and B
i
are determined from a curve tting of data from the litera-
ture [18]; A
a
= 1.2903 10
7
A/cm
2
, A
c
= 3.9255 10
6
A/cm
2
, B
a
=
151.532 kJ/mol and B
c
= 149.395 kJ/mol. Concentration overpoten-
tial is caused by concentration gradients due to the resistance to
mass transport through the electrodes and interfaces [19].
V
conc
V
conc;A
V
conc;C

RT
SOFC
2F
ln 1
i
i
s;A
_ _

RT
SOFC
2F
ln 1
P
H
2
i
P
H
2
O
i
s;A
_ _ _ _

RT
SOFC
4F
ln 1
i
i
s;C
_ _ _ _
12
where i
s,A
and i
s,C
are limiting current densities for anode and cath-
ode, respectively and shown below:
i
s;A

2FP
H
2
D
eff ;A
RT
SOFC
L
A
and i
s;C

4FP
O
2
D
eff ;C
PP
O
2
P
_ _
RT
SOFC
L
C
13
where D
eff
is the effective gaseous diffusivity through the electrode.
In this study, the SOFC is assumed anode supported and cell geom-
etry is provided in Table 1. As shown in Fig. 1, pure oxygen is
supplied to the cathode. It is estimated typical concentration over-
potential for the anode is several orders of magnitude larger than
that of the cathode, therefore, it is a reasonable assumption that
no mass transfer limitation for the cathode. Typical the anode effec-
tive gaseous diffusivity for an anode supported fuel cell is given by
[19]
D
eff ;A
1:3103 10
5
T
1:5
SOFC
0:26382 cm
2
=s 14
2.5. Water management and AOG recycle
As a part of the molar balance, the last step for the AOG recycle
unit design is the determination of the recycle ratio. The amount of
recycled AOG is manipulated to maintain the desired S/C ratio. The
AOG recycle ratio is dened as the recycled AOG to pre-recycled
AOG, R
c
= F
2
/F
1
. In this work, it should be noted that the steam to
carbon ratio (S/C) is dened at the inlet of the reformer instead
of at that of the SOFC anode and is given as S/C = F
2,H2O
/
(F
2,CH4
+ F
f
NC) where, F
f
and NC denote molar ow rate of freshly
added fuel, labeled f in Fig. 1, and the number of carbon in the fuel
e.g. for dodecane NC = 12. Substituting the denition of recycle ra-
tio into that of steam to carbon ratio yields,
Rc
S=CF
f
NC
F
1;H
2
O
S=CF
1;CH
4
15
The numerator of Eq. (15) denotes the required amount of
steam due to newly added fuel for the given conditions and the
denominator represents the amount of excess steam available be-
fore recycling. The range of recycle ratio for AOG is between zero
and one. It is obvious that the recycle ratio is always greater than
zero because we manipulate it to regulate the S/C and the S/C is
greater than zero. The recycle ratio should be less than one, the to-
tal recycle case, to prevent mass accumulation on the system. The
un-recycled AOG comprises mostly water vapor but also contains a
small fraction of un-reacted hydrogen as well as carbon monoxide
which can be utilized in the after-burner.
2.6. Energy balance
With the energy balance analysis, the temperature distribution
in the system can be determined. Basically, it is assumed that all
components are adiabatic with the surroundings except the SOFC,
external reformer, and afterburner. Since we are focusing on the
whole system analysis rather than individual components, a single
component temperature is assigned for a unit such as the reformer.
Therefore, a detailed model for the reformer was not needed as the
reformer is just a point (a unit) from the system point view, a rep-
resentative temperature would serve the purpose.
First, the rst law of thermodynamics is applied to the six com-
ponents (adsorbent, condenser, cooler, injector, pre-heater, and
recuperator) for energy balance calculations. e-NTU method [18]
has been applied for the pre-heater and the recuperator. For the
recuperator and the fuel pre-heater, we assume that the heat
exchangers are cross-ow type with both uids mixed, so the fol-
lowing effectiveness function is used for the e-NTU applications
[20].
e
NTU
NTU
1expNTU

Cr NTU
1expCr NTU
1
16
where e = q/q
max
= q/C
min
(T
h,i
T
c,i
), C
r
= C
min
/C
max
, and NTU = AU
o
/
C
min
. For sample calculations, we assume a representative value of
four for the NTU as the overall heat transfer coefcient, U
o
, is usu-
ally moderate for gases and the heat exchanger size is a medium
scale for aerospace and underwater applications. Also the kinetic
and potential energy terms are neglected in this work. For calcula-
tion of the enthalpy difference, specic heats for an ideal gas mix-
ture are integrated with respect to temperature. The specic heat
data is taken from the literature as a function of temperature
[21]. Regarding the chemical reaction, heats of reactions for the
SOFC, external reformer, CO
2
adsorbent, and after-burner are taken
into account as well as the sensible heat. In this work, the value of
heat of adsorption is obtained from the literature 17000 J/mol
[22]. It is assumed that the heat of combustion in the after-burner
is completely transferred to the external reformer to provide the
Table 1
SOFC unit data.
Operating temperature (T
SOFC
) C Various
Pressure of the cell (P
cell
) bar 1
Active surface area (A) m
2
1
Exchange current density of anode (i
oa
) A/cm
2
Function of T
Exchange current density of cathode (i
oc
) A/cm
2
Function of T
Effective gaseous diffusivity through anode (D
aeff
) cm
2
/s Function of T
Effective gaseous diffusivity through cathode (D
ceff
) cm
2
/s N/A
Thickness of anode (L
a
) lm 500
Thickness of electrolyte (L
e
) lm 10
Thickness of cathode (L
c
) lm 50
3218 T.S. Lee et al. / Energy Conversion and Management 52 (2011) 32143226
heat for the reforming reaction which is strongly endothermic.
Since the AOG recycle ratio is set to maintain the prescribed S/C ra-
tio, so the un-recycled AOG in the after-burner may not be able to
supply enough heat to the external reformer. To make up the energy
shortage, additional fuel is added to the after-burner to meet the
demand. Once the energy balance is set, evaluation of temperatures
is carried out by the NewtonRahpson method with a 10
7
tolerance.
3. Results and discussion
First, we compare our results with those in the open literature
in an effort to verify the current model and methodology. The clos-
est comparison we can nd in the literature is the experimental
work reported by Sasaki et al. [23]. Their system does not include
an anode off-gas recycle system but only include an external refor-
mer and a SOFC. So we have modied our simulation code to match
their system, simply by eliminating the recycle loop and adding
some minor changes (described in Table 2). The system parameters
and operating conditions used for the comparison are given in the
Table 2. The comparison between the current model and the exper-
imental work is based on the SOFC performance that is shown in
Fig. 2. We found that for the currentvoltage characteristics of
the SOFC, there is a good agreement between our model and the
experimental work by Sasaki et al. [23]. Based on the close compar-
ison, we believe that our model and the methodology are realistic
and correctly implemented.
3.1. SOFC performance
For the purpose of demonstrating the applicability and effec-
tiveness of the proposed combined fuel reforming and SOFC sys-
tem, sample calculations were performed using n-dodecane as a
representative diesel fuel and typical system conditions. As men-
tioned in the previous section, pure oxygen is assumed to be the
oxidizer and supplied to the cathode. The oxygen/fuel (n-dode-
cane) molar ratio used in our calculation is 37/2 that is the stoichi-
ometric ratio. First the basic performance characteristics of the
SOFC are examined under various system parameters. To facilitate
a parametric study, the baseline case, dened in Table 3, is used as
a reference.
Using the baseline case, the effects of varying the S/C, SOFC
operating temperature, AOG recirculation ratio and CO
2
adsorption
percentage on the SOFC performance are computed and shown in
Figs. 36 for comparison. In Fig. 3, the focus is on the effects of dif-
ferent steam to carbon ratios (S/C) while keeping all other param-
eters equal to those of the baseline case. Three different S/C ratios
of 2, 3 and 4 are evaluated and their effects on the voltage, power
density, and limiting current density are plotted. Based in Fig. 3, if
the steam to carbon ratio is increased, the terminal voltage, peak
power output, and fuel consumption rate for a constant voltage
all decrease monotonically. The same trend is also reported by
Kang et al. [8] where the authors mentioned that the measured
SOFC terminal voltage decreased from 0.733 to 0.726 and then to
0.713 as the S/C was increased from 0.625 to 1.25 and then to
2.5, respectively. As dened in the previous section, the S/C ratio
in this paper is set at the inlet of the reformer rather than at that
of the SOFC, which facilitates the involvement of the reformer on
the actual S/C ratio to the SOFC. The AOG recycle percentage is ad-
justed to supply enough water vapor for accomplishing the steam
to carbon ratio specied at the inlet of the reformer. The un-recy-
cled AOG is combusted in the after-burner to capture the un-re-
acted hydrogen and methane left in the stream. In general, as the
S/C is increased more steam is available in the reformer. So, for
the reformate at the exit of the reformer, the hydrogen molar frac-
tion would drop while that of the water vapor would increase that
causes the Nernst voltage potential to decrease according to Eq. (9).
This effect is clearly demonstrated in Fig. 2 with the highest open
circuit voltage corresponding to the lowest S/C. When the S/C ratio
is raised, more hydrogen would be produced in the reformer and
fed into the SOFC due to an increase in the amount of water vapor
available for the water gas shift reaction. Further more, according
to Eq. (13), a lower hydrogen molar fraction at the SOFC anode inlet
would result in a lower limiting current density. This explains why
the limiting current density is lowered when the S/C is increased.
As shown in Fig. 3, the terminal voltage curves are relatively close
among the three cases when the current density is less than 0.35 A/
cm
2
. While the limiting current density really sets the three cases
Table 2
System parameters used for comparison with experiment of Sasaki et al. [8].
Fuel n-Dodecane
Cathode ow Air
Fuel cell
Type SOFC
Anode Lanthanum manganite (140 lm)
Electrolyte YSZ (10 lm)
Cathode Ni-YSZ (50 lm)
Fuel cell temperature (C) 1000
Reformer Steam reforming
S/C 2.5
Reformer temperature (C) 750
Fuel utilization factor 0.1
Oxygen utilization factor 0.02
Exchange current density of anode (i
oa
)
(A/cm
2
)
0.1
Exchange current density of cathode (i
oc
)
(A/cm
2
)
0.07
Current density, i [A/cm
2
]
V
o
l
t
a
g
e
,
V

[
V
o
l
t
]
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
0.5
0.55
0.6
0.65
0.7
0.75
0.8
0.85
0.9
0.95
1
Literature (Sasaki et al.)
This work
n-dodecane fuel
S/C = 2.5
T = 1000
o
C
Fig. 2. Currentvoltage characteristics comparison with experimental results by
Sasaki et al. [8].
Table 3
System parameters and operating conditions for the baseline case.
SOFC operating temperature (C) 850
Steam to carbon ratio (S/C) () 3
AOG recirculation ratio () 0
CO
2
adsorption percentage (%) 0
External reformer temperature (C) 550
Fuel utilization factor () 0.85
SOFC active area (m
2
) 1
T.S. Lee et al. / Energy Conversion and Management 52 (2011) 32143226 3219
apart that results in quite different peak power densities. The low-
er steam to carbon ratio case has the higher peak power density
due to the higher limiting current density.
Fig. 4 shows how the SOFC operating temperature affects the
terminal voltage as well as the current density. First, we found that
the maximum power density increases with increasing SOFC oper-
ating temperature and so does the limiting current density. Similar
to the previous case (Fig. 3), the peak power density is primarily
dependent on the limiting current density that is proportional to
the SOFC operating temperature. However, the trend of open-cir-
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.40
0.45
0.0
0.2
0.4
0.6
0.8
1.0
S/C = 2
S/C = 3
S/C = 4
Voltage
Power density
V
o
l
t
a
g
e
,

V

[
v
o
l
t
s
]
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
P
o
w
e
r

d
e
n
s
i
t
y
,

[
W
/
c
m
2
]
Current density, i [A/cm
2
]
Fig. 3. Polarization curves for different steam to carbon ratios, SOFC temperature 850 C, reformer temperature 550 C, no recirculation, no CO
2
adsorption, 0.85 fuel
utilization factor, and 1 m
2
for active area.
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.40
0.45
0.0
0.2
0.4
0.6
0.8
1.0
T
SOFC
= 950 [
o
C]
T
SOFC
= 850 [
o
C]
T
SOFC
= 750 [
o
C]
Voltage
Power density
V
o
l
t
a
g
e
,

V

[
v
o
l
t
s
]
P
o
w
e
r

d
e
n
s
i
t
y
,

[
W
/
c
m
2
]
Current density, i [A/cm
2
]
Fig. 4. Polarization curves for different SOFC temperatures, S/C 3, reformer temperature 550 C, no recirculation, no CO
2
adsorption, 0.85 fuel utilization factor, and 1 m
2
for
active area.
3220 T.S. Lee et al. / Energy Conversion and Management 52 (2011) 32143226
cuit voltages differs from that of the previous case. The open-cir-
cuit voltage or the Nernst voltage, Eq. (9), is inversely proportional
to the SOFC operating temperature that results in the highest open-
circuit voltage for the lowest SOFC operating temperature. Based
on Eq. (9), we also note that all the overpotentials (voltage losses
due to irreversibilities) are also inversely proportional to the SOFC
operating temperature. This is the reason why the terminal voltage
decreases faster for a lower SOFC operating temperature which re-
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.0
0.2
0.4
0.6
0.8
1.0
r = 0.0
r = 0.1
r = 0.2
Voltage
Power density
P
o
w
e
r

d
e
n
s
i
t
y
,

[
W
/
c
m
2
]
Current density, i [A/cm
2
]
V
o
l
t
a
g
e
,

V

[
v
o
l
t
s
]
0.0 0.1 0.2 0.3 0.4 0.5 0.6
Fig. 5. Polarization curves for different AOG recirculation, S/C 3, SOFC temperature 850 C, reformer temperature 550 C, no CO
2
adsorption, 0.85 fuel utilization factor, and
1 m
2
for active area.
P
o
w
e
r

d
e
n
s
i
t
y
,

[
W
/
c
m
2
]
Current density, i [A/cm
2
]
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.40
0.45
0.0
0.2
0.4
0.6
0.8
1.0
no CO
2
Adsorption
15% CO
2
Adsorption
30% CO
2
Adsorption
Voltage
Power density
V
o
l
t
a
g
e
,

V

[
v
o
l
t
s
]
Fig. 6. Polarization curves for different CO
2
capture, S/C 3, SOFC temperature 850 C, reformer temperature 550 C, no recirculation, 0.85 fuel utilization factor, and 1 m
2
for
active area.
T.S. Lee et al. / Energy Conversion and Management 52 (2011) 32143226 3221
sults in a smaller limiting current density. For example, based on
Eq. (12), concentration overpotential is a strong function of the
limiting current density and inversely proportional to it. According
to Eqs. (16) and (17), a higher SOFC temperature yields a larger
limiting current density that results in lower concentration
overpotentials.
Fig. 5 demonstrates how the terminal voltage and power den-
sity change with the AOG recirculation ratio. It is recalled that
the recirculation of AOG puts more water vapor back to the SOFC
anode inlet and its effect is very similar to an increase of the S/C.
As the recirculation ratio is increased, terminal voltage, peak power
output, fuel consumption rate for a constant voltage, current den-
sity at peak power, and the maximum limiting current density all
decrease, that is very similar to the results shown in Fig. 3 where
the steam to carbon ratio is increased. It may be deduced that
increasing the recirculation ratio represents increasing the steam
to carbon ratio at the SOFC inlet due to high water vapor content
in recirculation stream.
The last system parameter examined is the percentage of car-
bon dioxide in the reformate captured by the adsorbent and the re-
sults are illustrated in Fig. 6. As shown in the process ow diagram,
Fig. 1, the carbon dioxide adsorption process is located between
0.20
0.25
0.30
0.35
0.40
0.45
0.50
T
SOFC
= 750
o
C
T
SOFC
= 800
o
C
T
SOFC
= 850
o
C
T
SOFC
= 900
o
C
T
SOFC
= 950
o
C
S/C = 2
S/C = 3
S/C = 4
0.3 0.4 0.5 0.6 0.7 0.8
P
e
a
k

p
o
w
e
r

d
e
n
s
i
t
y

[
W
/
c
m
2
]
Current density @ peak power, i
peak
[A/cm
2
]
Fig. 7. The peak power density for different steam to carbon ratios as well as SOFC
temperature for reformer temperature 550 C, no recirculation, no CO
2
adsorption,
0.85 fuel utilization factor, and 1 m
2
for active area.
0.3 0.4 0.5 0.6
0.20
0.25
0.30
0.35
0.40
T
SOFC
= 750
o
C
T
SOFC
= 800
o
C
T
SOFC
= 850
o
C
T
SOFC
= 900
o
C
T
SOFC
= 950
o
C
P
e
a
k

p
o
w
e
r

d
e
n
s
i
t
y

[
W
/
c
m
2
]
Current density @ peak power, i
peak
[A/cm
2
]
r = 0
r = 0.1
r = 0.2
Fig. 8. The peak power density for different recirculation ratios as well as SOFC
temperature for reformer temperature 550 C, S/C 3, no CO
2
adsorption, 0.85 fuel
utilization factor, and 1 m
2
for active area.
0.4 0.5 0.6 0.7 0.8
0.20
0.25
0.30
0.35
0.40
0.45
0.50
0.55
T
SOFC
= 750
o
C
T
SOFC
= 800
o
C
T
SOFC
= 850
o
C
T
SOFC
= 900
o
C
T
SOFC
= 950
o
C
P
e
a
k

p
o
w
e
r

d
e
n
s
i
t
y

[
W
/
c
m
2
]
Current density @ peak power, i
peak
[A/cm
2
]
no CO
2
removal
15% CO
2
removal
30% CO
2
removal
Fig. 9. The peak power density for different CO
2
removal percentage as well as
SOFC temperature for reformer temperature 550 C, S/C 3, no recirculation, 0.85 fuel
utilization factor, and 1 m
2
for active area.
750 800 850 900 950
0.1
0.2
0.3
0.4
0.5
0.6
S/C = 2
S/C = 3
S/C = 4
S/C = 5
P
e
a
k

P
o
w
e
r

d
e
n
s
i
t
y
,

[
W
/
c
m
2
]
SOFC Operating Temperature, T
SOFC
[
o
C]
Fig. 10. Peak power density with different S/C ratio for reformer temperature
550 C, no recirculation, 15% CO
2
adsorption, 0.85 fuel utilization factor, and 1 m
2
for active area.
3222 T.S. Lee et al. / Energy Conversion and Management 52 (2011) 32143226
the SOFC and the reformer unit. It is assumed that a certain fraction
of the CO
2
in reformate is removed before entering the SOFC anode
and this CO
2
capture causes a molar fraction increase for the other
chemical species in the reformate. The effect of the CO
2
capture is
opposite to that of recirculation of AOG in hydrogen molar fraction
point of view. Recirculation of AOG dilutes hydrogen concentration
at the SOFC anode inlet, while CO
2
capture enhances the hydrogen
concentration at the SOFC inlet. Considering Eq. (13), a higher
hydrogen molar fraction for the SOFC causes a higher limiting
current density that in turn results in a lower concentration over-
potential. As shown in Fig. 6, a higher limiting current density
could be achieved under a higher CO
2
capture percentage. In sum-
mary, CO
2
removal yields an opposite trend to those of AOG recir-
culation and higher reformer inlet steam to carbon ratio.
3.2. Peak power density
In engineering applications, the peak power density is of the
most concern. Therefore, here we focus on this main performance
indicator. The peak power density vs. the corresponding current
density is plotted as a function of different S/C ratios, AOG recircu-
lation ratios, CO
2
capture percentages as well as SOFC operating
temperatures in Figs. 79. For each gure, two parameters are var-
ied that results in one set of curves with different symbols. For
example, in Fig. 7 the S/C ratio and the SOFC temperature are var-
ied with all other parameters and conditions kept identical to the
baseline case. However, three curves represent different S/C ratios,
while the symbols denote the SOFC temperatures. It may be ob-
served from the gure that a relatively linear dependence between
the peak power density and the corresponding current density. It is
clear that the maximum SOFC performance is associated with the
highest SOFC operating temperature and the lowest S/C ratio. In
Fig. 8, the AOG recirculation ratio and the SOFC temperature are
varied with all other parameters and conditions kept identical to
the baseline case. Similar to Fig. 7, a linear relationship is observed.
The peak power density and the corresponding current density
move towards the upper right corner as the recirculation ratio de-
creases and/or SOFC operating temperature increases. However,
the variation of AOG recirculation ratio on the peak power density
and the corresponding current density is not as signicant compar-
ing with that of the steam to carbon ratio. The reason that an in-
crease in the AOG recirculation ratio would cause the peak
power density to decrease instead of increasing is mainly due to
the fact that AOG typically contains $90% water vapor and only
750 800 850 900 950
0.3
0.4
0.5
0.6
0.7
0.8
0.9
S/C = 2
S/C = 3
S/C = 4
S/C = 5
C
u
r
r
e
n
t

d
e
n
s
i
t
y

@

p
e
a
k

p
o
w
e
r
,

i
p
e
a
k

[
A
/
c
m
2
]
SOFC Operating Temperature, T
SOFC
[
o
C]
Fig. 11. Current density @ peak power with different S/C ratio for reformer
temperature 550 C, no recirculation, 15% CO
2
adsorption, 0.85 fuel utilization
factor, and 1 m
2
for active area.
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.40
0.45
0.50
0
10
20
30
40
50
60
70
80
T
SOFC
= 750 [
o
C]
T
SOFC
= 850 [
o
C]
T
SOFC
= 950 [
o
C]
System efficiency
Power density
P
o
w
e
r

d
e
n
s
i
t
y
,

[
W
/
c
m
2
]
Current density, i [A/cm
2
]
S
y
s
t
e
m

E
f
f
i
c
i
e
n
c
y
,

L
H
V

[
%
]

Fig. 12. System efciency based on lower heating value of the fuel and power density for different SOFC temperatures, S/C 3, pre-reformer temperature 550 C, no
recirculation, 15% CO
2
adsorption, 0.85 fuel utilization factor, and 1 m
2
for active area.
T.S. Lee et al. / Energy Conversion and Management 52 (2011) 32143226 3223
$5% hydrogen. AOG recirculation essentially creates higher steam
to carbon ratio that causes the peak power density to decrease that
is demonstrated in Fig. 10. In Fig. 9, the CO
2
removal percentage
and the SOFC temperature are varied with all other parameters
and conditions kept identical to the baseline case. Consistent with
Figs. 7 and 8, a linear trend is shown. The peak power density and
the corresponding current density approach the upper right corner
as the CO
2
removal percentage is increased and/or SOFC operating
temperature gets higher.
Figs. 10 and 11 demonstrate the steam to carbon ratio and SOFC
temperature effects on peak power density and current density at
peak power density for typical conditions with no AOG recircula-
tion and 15% CO
2
capture. Since effects of changing AOG recircula-
tion and CO
2
capture percentage have been evaluated in Figs. 8 and
9, any AOG recirculation and CO
2
capture percentage would serve
the purpose. 15% and zero CO
2
capture would be as good as any
other selection.
Both the peak power density and current density at peak power
are increasing with increasing SOFC temperature as well as with
decreasing steam to carbon ratio. Both also share the same trend
that the SOFC temperature dependency is more signicant at lower
steam to carbon ratios.
3.3. System thermal efciency
As mentioned above in the energy balance section, the un-recy-
cled AOG plays an important role in the overall energy budget for
the external reformer since the un-recycled AOG is combusted in
the after-burner and the combustion process provides heat for
the strongly endothermic steam reforming reaction in the refor-
mer. In general, the heat supplied by the combustion of the un-
recycled AOG is not enough for the need of the reformer. Based
on the molar balance and energy balance under chemical equilib-
rium and given system operating conditions, this heat decit can
be determined and is made up by consuming additional fuel in
the after-burner (second term inside the parentheses in Eq. (17)).
Therefore, this additional fuel required for the steam reforming
reaction should be included when dening the system thermal ef-
ciency as follows:
g
LHV

W
SOFC;elect
F
f
F
f ;AB
LHV
f
17
Fig. 12 demonstrates how the system efciency varies as a func-
tion of the SOFC temperature as well as the current density. From
Fig. 11, it is observed that the thermal efciency generally starts
out at a very high value (>70%) near the open circuit voltage and
then monotonically decreases to around 55% where the maximum
power density occurs before dropping steeply to zero at the limiting
750 800 850 900 950
40
42
44
46
48
50
52
54
56
58
60
S/C = 2
S/C = 3
S/C = 4
S/C = 5
S
y
s
t
e
m

E
f
f
i
c
i
e
n
c
y

@

P
e
a
k

P
o
w
e
r
,

L
H
V

[
%
]
SOFC Operating Temperature, T
SOFC
[
o
C]
Fig. 13. System efciency based on lower heating value of the fuel at peak power
density with different S/C ratio for pre-reformer temperature 550 C, no recircu-
lation, 15% CO
2
adsorption, 0.85 fuel utilization factor, and 1 m
2
for active area.
T
SOFC
= 850
o
C
S/C = 4
no recirculation
no CO
2
capture
Fuel utilization factor: 0.85 T
Reformer
= 550
o
C
T
Reformer
= 700
o
C
T
Reformer
= 550
o
C
T
Reformer
= 700
o
C
Current density, i [A/cm
2
]
P
o
w
e
r
d
e
n
s
i
t
y
,
[
W
/
c
m
2
]
V
o
l
t
a
g
e
,
V
[
V
o
l
t
s
]
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
0
0.05
0.1
0.15
0.2
0.25
0.3
0.35
0.4
0.45
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
Fig. 14. Polarization curves for S/C = 4, SOFC temperature 850 C, reformer temperature 700 C, no recirculation, no CO
2
adsorption, 0.85 fuel utilization factor, and 1 m
2
for
active area.
3224 T.S. Lee et al. / Energy Conversion and Management 52 (2011) 32143226
current density condition. Froman engineering application point of
view, the proposed combined fuel reforming and SOFC system with
a range of efciency between 50% and 70% is considered very attrac-
tive. It is noted that only for the case of 750 C SOFC temperature
under vanishingly small current densities, the efciency does not
followthe monotonic decreasing trend. As a matter of fact, it shows
a peak efciency of 76.6% at 0.0167 A/cm
2
that is due to additional
fuel consumption (F
f,AB
> 0 in Eq. (17)) at the after-burner required
for low SOFC temperature and/or low current density.
Fig. 13 demonstrates the steam to carbon ratio and SOFC tem-
perature effects on the system efciency at the peak power density
for typical system conditions with no AOG recirculation and 15%
CO
2
capture. The system efciency under the peak power condition
generally increases sharply with the SOFC temperature up to
850 C and then levels off and moves up slowly to reach the max-
imum at 900 C. After that it decreases slightly as the SOFC temper-
ature increases further to 950 C. Generally, a lower steam to
carbon ratio yields a lower system efciency but the dependence
of the system efciency on the S/C ratio is not strong except for
S/C = 2. It is interesting to note that in Figs. 10 and 11 both the peak
power and the current at peak power decrease with increasing S/C
ratio while the system efciency at peak power shows an opposite
trend with the S/C ratio. This is mainly due to the fact that in the
denition of the system efciency (Eq. (17)), the numerator and
the denominator are determined by the power density and current
density, respectively, but the dependence on the S/C ratio is rela-
tively stronger for the current density than the power density as
shown in Figs. 10 and 11.
It is noted that all the above results are based on that the refor-
mer is operating at 550 C and some questions arise on whether a
thermal dynamic equilibrium state can be reach at this tempera-
ture. For a kilowatt-scale autothermal reformer for the production
of hydrogen from heavy hydrocarbons, Liu et al. [9] reported that
reformate composition obtained from the experiment is well
matching with thermodynamic calculation at 590 C. In order to
show that the reformer temperature would not affect the major
trends and outcome, we made an additional calculation for the re-
former temperature at 700 C. The other system conditions are
kept the same as the 550 C case as follows: SOFC temperature
850 C, S/C = 4, no recirculation, no CO
2
adsorption, 0.85 fuel utili-
zation factor, and 1 m
2
for active area.
The results are given in Figs. 14 and 15. From the gures pro-
vided, we can see that the fuel cell performance (voltage and
power density) are almost identical between the 550 C and
700 C cases. While, the system efciency is lower by about 15%
for the 700 C reformer temperature that is basically due to the to-
tal amount of heat available fromthe AOG for reformer operation is
limited by the SOFC.
As mentioned above, the AOG recycle percentage is a function of
the S/C ratio. Fig. 16 provides this relationship as a function of the
SOFC operating temperature. As expected the percentage increases
with increasing S/C ratio, but the relationship is independent of the
SOFC operating temperature. Fig 16 also indicates that for the prac-
tical range of S/C ratios up to 5, there is no need to add water from
an external source.
4. Conclusions
A parametric study for a combined fuel reforming and SOFC sys-
tem with AOG is presented for thermal management applications.
The model and the methodology are veried by a favorable com-
parison with an independent experimental work from the open
literature.
Based on a parametric study, we found that in general the SOFC
terminal voltage decreases with increasing current density and
after reaching the maximum power density it sharply drops to zero
at the limiting current density condition. Whereas the SOFC power
density increases with increasing current density until the maxi-
mum is reached and after that it decreases steeply to zero at the
limiting current density point. In general, the open-circuit voltage,
the terminal voltage, the peak power density, and the limiting cur-
rent density all increase with increasing S/C ratio, increasing AOG
recirculation ratio, decreasing CO
2
adsorption percentage, and
decreasing SOFC temperature, respectively. It is also found that
the AOG recycle ratio increases with increasing S/C ratio and for
most applications, the addition of AOG recycle eliminates the need
for external water consumption.
For most of the operating conditions investigated, the system
efciency starts at around 70% and then monotonically decreases
to an average of 50% at the peak power density before dropping
down to zero at the limiting current density point. From an engi-
neering application point of view, the proposed combined fuel
T
Reformer
= 700
o
C
T
Reformer
= 550
o
C
Current density, i [A/cm
2
]
S
y
s
t
e
m
E
f
f
i
c
i
e
n
c
y
,

L
H
V
[
%
]
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
0
10
20
30
40
50
60
70
80
Fig. 15. System efciency curves for S/C = 4, SOFC temperature 850 C, reformer
temperature 700 C, no recirculation, no CO
2
adsorption, 0.85 fuel utilization factor,
and 1 m
2
for active area.
SOFC Operating Temperature, T
SOFC
[
o
C]
R
e
c
y
c
l
e

P
e
r
c
e
n
t

@

P
e
a
k

P
o
w
e
r
,

[
%
]
750 800 850 900 950
66
68
70
72
74
76
78
80
82
84
S/C = 2
S/C = 3
S/C = 4
S/C = 5
Fig. 16. AOG recycle percentage based on peak power density with different S/C
ratio for a pre-reformer temperature of 550 C, no recirculation, 15% CO
2
adsorp-
tion, 0.85 fuel utilization factor, and 1 m
2
for active area.
T.S. Lee et al. / Energy Conversion and Management 52 (2011) 32143226 3225
reforming and SOFC system with a range of efciency between 50%
and 70% is considered very attractive.
Our results point out that for the SOFC at 900 C, the steam to
carbon ratio at 5 and no AOG recirculation, the system efciency
peaks. Also the peak power density increases with decreasing
CO
2
capture. Therefore, it is suggested that the optimal system is
the one where the SOFC operates around 900 C with S/C ratio
higher than 3, maximum CO
2
capture, and minimum AOG
recirculation.
Acknowledgements
This research was supported by the NASA Hydrogen Research
for Spaceport and Space Based Applications at the University of
Florida (Grant number NAG3-2930). The support by the Andrew
H. Hines Jr./Progress Energy Endowment Fund is also acknowl-
edged. The rst author was partially supported by the Korea Sci-
ence and Engineering Foundation Grant funded by the Korea
government (Ministry of Science and Technology) (2005-215-
D00037).
References
[1] EG&G. Fuel cell handbook. 7th ed. West Virginia: U.S. Department of Energy;
2004.
[2] Singhal SC, Kendall K. High temperature solid oxide fuel cells. Oxford, United
Kingdom: Elsevier Advanced Technology Publishing; 2004.
[3] Lindstrm BJ, Karlsson AJ, Ekdunge P, De Verdier L, Hggendal B, Dawody J,
et al. Diesel fuel reformer for automotive fuel cell applications. Int J Hydrogen
Energy 2009;34:336781.
[4] Chrenko D, Couli J, Lecoq S, Pra MC, Hissel D. Static and dynamic modeling of
a diesel fuel processing unit for polymer electrolyte fuel cell supply. Int J
Hydrogen Energy 2009;34:132435.
[5] Emonts B, Bgild J, Hansen S, Jrgensen L, Hhlein B, Peters R. Compact
methanol reformer test for fuel-cell powered light-duty vehicles. J Power
Sources 1998;71:28893.
[6] Rostrup-Nielsen JR. Conversion of hydrocarbons and alcohols for fuel cells.
Phys Chem Chem Phys 2001;3:2838.
[7] Papadias D, Lee SHD, Chmielewski DJ. Autothermal reforming of gasoline for
fuel cell applications: a transient reactor model. Ind Eng Chem Res
2006;45:584158.
[8] Kang I, Kang Y, Yoon S, Bae G, Bae J. The operating characteristics of solid fuel
cells driven by diesel autothermal reformate. Int J Hydrogen Energy
2008;33:6298307.
[9] Liu DJ, Kaun TD, Liao HK, Ahmed S. Characterization of kilowatt-scale
autothermal reformer for production of hydrogen from heavy hydrocarbons.
Int J Hydrogen Energy 2004;29:103546.
[10] Shekhawat D, Berry DA, Gardner TH, Haynes DJ, Spivey JJ. Effects of fuel cell
anode recycle on catalytic fuel reforming. J Power Sources 2007;168:
47783.
[11] Shi L, Bayless DJ. Analysis of jet fuel reforming for solid oxide fuel cell
applications in auxiliary power units. Int J Hydrogen Energy 2008;33:106775.
[12] Cutillo A, Specchia S, Antonini M, Saracco G, Specchia V. Diesel fuel processor
for PEM fuel cells: two possible alternatives (ATR versus SR). J Power Sources
2006;154:37985.
[13] Ahmed S, Krumpelt M. Hydrogen from hydrocarbon fuels for fuel cells. Int J
Hydrogen Energy 2001;26:291301.
[14] Sasaki K, Teraoka Y. Equilibria in fuel cell gases. J Electrochem Soc
2003;150(7):A8858.
[15] Colpan CO, Dincer I, Hamdullahpur F. A review on macro-level modeling of
planar solid oxide fuel cells. Int J Energy Res 2008;32:33655.
[16] Colpan CO, Dincer I, Hamdullahpur F. Thermodynamic modeling of direct
internal reforming solid oxide fuel cells operating with syngas. Int J Hydrogen
Energy 2007;32:78795.
[17] Bessette NF, Wepfer WJ, Winnick J. A mathematical model of a solid oxide fuel
cell. J Electrochem Soc 1995;142:3792800.
[18] Chan SH, Low CF, Ding OL. Energy and exergy analysis of simple solid-oxide
fuel-cell power systems. J Power Sources 2002;103:188200.
[19] Kim JW, Virkar AV, Fung K-Z, Mehta K, Singhal SC. Polarization effects in
intermediate temperature, anode-supported solid oxide fuel cells. J
Electrochem Soc 1999;146:6978.
[20] Kays WM, London AL. Compact heat exchangers. New York: McGraw-Hill;
1984.
[21] Smith JM, Van Ness HC, Abbott MM. Introduction to chemical engineering
thermodynamics. 7th ed. New York: McGraw-Hill; 2005.
[22] Ding Y, Alpay E. Adsorption-enhanced steam-methane reforming. Chem Eng
Sci 2000;55:392940.
[23] Sasaki K, Watanabe K, Shiosake K, Susuki K, Teraoka Y. Multi-fuel capability of
solid oxide fuel cells. J Electroceram 2004;13:66975.
3226 T.S. Lee et al. / Energy Conversion and Management 52 (2011) 32143226

Vous aimerez peut-être aussi