Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

ARTICLE IN PRESS

Chemical Engineering Science 60 (2005) 4423 – 4430


www.elsevier.com/locate/ces

Heat transfer phenomena in a solid oxide fuel cell: An analytical approach


F.A. Coutelieris∗ , S.L. Douvartzides, P.E. Tsiakaras
Department of Mechanical and Industrial Engineering, University of Thessalia, Pedion Areos, 38334 Volos, Greece

Received 28 May 2003; received in revised form 25 January 2005; accepted 10 March 2005

Abstract
The steady-state heat transfer problem within a fuel cell is considered here. The original 2-D problem is defined while an averaging
technique is used in order to reduce its dimensions to one. The full heat transfer equations along the appropriate boundary conditions
are analytically solved for both solid and gas phase. Moreover, analytical expressions for the local as well as for the overall heat transfer
coefficients are obtained. The gas temperature increases with the distance and the current density, while the temperature of the cell is
always of so low variation that can be considered as a constant. Finally, the heat transfer coefficient presents a profile reverse to that of
gas temperature.
䉷 2005 Elsevier Ltd. All rights reserved.

Keywords: Fuel cell; Heat transfer; Conduction; Convection; Mathematical modeling; Laminar flow

1. Introduction temperature (Virkar et al., 1997). Since a SOFC stack con-


sists of many connected cells, stack models can be built
One of the most promising new technologies in the sec- upon the conventional heat balance equations but further
tor of electrical power systems is that of fuel cells, whose consideration of electrochemical reaction and its conse-
concept has been known for more than a century. Being quences should be also taken into account. Usually, the cell
electrochemical energy conversion devices, the major prod- stack models require consideration of gas flow, heat transfer,
uct of fuel cells is electrical power while their impact on mass transfer, and cell voltage–current relationship. Some
atmosphere quality is zero or near zero. Additionally, the mathematical models (e.g. molecular dynamics approach,
replacement of spontaneous combustion of the typical heat stochastic models, etc.) need several geometrical and elec-
engines with controllable electrochemical oxidation moder- trochemical parameters, which are functions of temperature
ates the production of pollutants in the fuel gases. Since the and gas composition (Dutta et al., 2001; Koh et al., 2001).
by-products of the electrochemical reaction in a fuel cell Theoretical prediction of temperature distribution is impor-
are water and a significant amount of heat, effective ther- tant for SOFC stack design and its stable operation, and
mal management is one of the important problems for all it should be achieved by solving a rigorous mathematical
the types of fuel cells (Appleby and Foulkes, 1989; Kordesh model.
and Simander, 1996). A lot of investigations concerning mathematical model-
Among the different types of fuel cells, solid oxide fuel ing have been presented in order to adequately simulate the
cells (SOFCs) that are solid state, ceramic cells operating heat transfer within SOFCs. The majority of SOFC thermal
currently at high temperatures, present special interest be- models pertain to the planar and monolithic designs (Minh
cause of the flexibility in fuel choice due to high operation and Takahashi, 1995). Ahmed et al. (1991) assumed fully
developed laminar flow presenting Nusselt number value
∗ Corresponding author at: National Center for Scientific Re- of 3 for convective heat transfer. Bassette et al. (1995)
search, “Demokritos”, 15310 Aghia Paraskevi Attikis, Greece. Tel./fax: also presented a thermal model of a tubular SOFC de-
+302106525004. sign having limited validation to one experimental datum.
E-mail address: frank@ipta.demokritos.gr (F.A. Coutelieris). Kanamura et al. (1989) and Hirano et al. (1992) presented
0009-2509/$ - see front matter 䉷 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2005.03.013
ARTICLE IN PRESS
4424 F.A. Coutelieris et al. / Chemical Engineering Science 60 (2005) 4423 – 4430

also thermal models and simulations but of limited electro- axial symmetry and fully developed incompressible lami-
chemical validation. nar flow and by ignoring heat transfer due to radiation, the
This work presents an analytical model for the thermal conduction–convection heat transport equations can be writ-
transport phenomena occurred within a SOFC stack oper- ten as follows:
ating at high temperatures. By considering the complete 
conduction–convection heat transfer equation along the cell, jTg (r, z) j2 Tg (r, z) 1 jTg (r, z)
uz (r) = f +
analytical expressions for the spatial distribution of both the jz jr 2 r jr
gas and the cell temperatures as well as for the local and 
j Tg (r, z)
2
overall heat transfer coefficient are obtained by using an av- + , (2a)
eraging technique. As the modeled fuel cell is a real cylin- jz2
drical design and the electrochemistry considered is quite  
j2 Tc (r, z) 1 jTc (r, z) j2 Tc (r, z)
complete, the basic barriers mentioned above are overcome s(z) = s + + , (2b)
and the use of these analytical formulas can be very helpful jr 2 r jr jz2
in the design and/or in the operating phase of such systems, where Tg (r, z) and Tc (r, z) are the gas and solid temperature,
as the produced results are sufficiently accurate for any prac- respectively, uz (r) is the z-component of the fluid velocity
tical application. and f and s are thermal diffusivities in the fluid and the
solid phase, respectively.
The above equation can be integrated with the following
2. Formulation of the problem boundary conditions:

A typical fuel cell stack consists of an anode and a cathode Tg (r, 0) = const.,  r , (3a)
part between which a catalyst (electrolyte) layer exists. The 
jTg (r, z)
atmospheric air flows in the cathode “area” (gas channel) = 0,  r , (3b)
while the anode is exposed to hydrogen. The present model jz z=L
studies a compatible to the Siemens–Westinghouse cylindri- 
jTg (r, z)
cal fuel cell design (George and Bassette, 1998; Hirschen- = 0, 0  z  L, (3c)
hofer et al., 1998) as it is illustrated in Fig. 1a. By consid- jr r=
ering the simplest and most studied reaction of hydrogen 
k(z) jTg (r, z)
oxidation in the presence of oxygen (air), one might have = s(z), 0  z  L, (3d)
the formation of water with heat and electricity production,
c p jr r=
as follows (Koh et al., 2001): Tc (r, 0) = const.,  r , (3e)
H2 + 1
→ H2 O + heat + electricity. 
2 O2 (1) jTc (r, z)
= 0,  r , (3f)
jz z=L
The area of interest is the cylinder ABCD of Fig. 1a,
which is presented in detail in Fig. 1b. By assuming 
jTc (r, z)
= 0, 0  z  L (3g)
jr r=

and

Tg (, z) = Tc (, z), 0  z  L. (3h)

Boundary conditions (3a) and (3e) insure a constant pro-


file for the gas and solid temperature at the inlet while Eqs.
(3b) and (3f) imposes typical von Newman conditions for
the outlet temperature. Eqs. (3c) and (3g) are proposed in
order to ensure the continuity of the temperature upon the
outer boundaries of the area of interest while Eq. (3h) im-
poses the continuity of the temperature on the gas–solid in-
terface. Finally, Eq. (3d) expresses the transport to the cath-
ode gas of the heat produced upon the electrodic material
surface due to electrochemical reaction (1).
By s(z) in Eq. (2b), is denoted the spatial distribution of
the heat sources along the cell in the cross-section z, defined
as
Fig. 1. Design of a tubular fuel cell (a) and blow up of the area of
interest (b). s(z) = j (Ut + j r eff − E(z)), (4)
ARTICLE IN PRESS
F.A. Coutelieris et al. / Chemical Engineering Science 60 (2005) 4423 – 4430 4425

where j is the current density, which is assumed to be con- information about the heat sources upon the solid–gas in-
stant, as previous investigations have shown that it presents terface. Therefore, it is necessary for the heat transfer equa-
low variability for the conditions considered here (Koh et tion to be considered in both the two different shrinking
al., 2001, Djilali and Lu, 2002). By Ut is denoted the thermo phases.
neutral voltage: As the conduction is the only available mechanism for
the heat transport within solids, the heat transfer in the solid
−HH2
Ut = ≈ 1.29 V. (5) phase of the cell is described by the following differential
2F equation:
Finally, the effective specific resistance of the cell, reff ,
includes the ohmic, the polarization anodic and the polariza- d2 Tc (z)
− s = s(z), (8)
tion cathodic resistances, according to the relation (Haynes dz2
and Wepfer, 2000):
where s denotes the thermal diffusivity within the solid
reff = rohm L + ra + rc . (6) phase.
The electromotive force at position z for a mean cell tem-
Although the above formulation is rather obvious from perature value Tc (z), averaged along the length of the cell,
transport processes point of view, the solution of the con- is given by Tsiakaras and Demin (2001)
sequent boundary value problem is difficult as it includes
sometimes circular schemes. This difficulty is the main rea- RTc (z) 0.209
E(z) = ln , (9)
son for the lack of analytical and/or numerical solutions of 4F [yH2 O /yH2 K]2
such type of formulations as well as for the mathematical
manoeuvres used in the literature (Dutta et al., 1997; Yuan where
et al., 2001; Haynes and Wepfer, 2001; Yuan and Sunden,  L
1
2004). Tc (z) = Tc (z) dz (10)
L 0

3. Problem simplification and solution or, by using dimensionless quantities,


 1
By considering a circular ring of the domain of the inter-
Tc (x) = Tc (x) dx, (11)
est, it is easy to define an averaging procedure in a cross- 0
section z, as it is presented in Fig. 2. In that case, the origi-
nal two-dimensional gas temperatures, Tg (r, z) and Tc (r, z), where x =z/L. The symbol K in Eq. (9) denotes the equilib-
become rium constant of reaction (1). The utilization factor (extent
  of reaction), f , is defined as
1
Tg (z) = Tg (r, z)2r dr, (7a) pH2 (0) − pH2 (1)
−  f = . (12)
  pH2 (0)
1
Tc (z) = Tc (r, z)2r dr. (7b) As the fuel is almost pure hydrogen, it can be highly
− 
utilized and, therefore, f is set to 0.85 as pH2 (0) = 0.98
This averaging procedure should be applied separately in and pH2 (1) = 0.15 (Demin and Tsiakaras, 2001). After in-
the solid and the gas phase, in order to avoid missing of troducing the utilization factor, a suitable expression for the

Fig. 2. Reduction of the problem’s dimensions.


ARTICLE IN PRESS
4426 F.A. Coutelieris et al. / Chemical Engineering Science 60 (2005) 4423 – 4430

electromotive force can be obtained: By concerning the above boundary conditions, the arbitrary
constants of Eq. (16) are identified as follows:
RTc (x)
E(x) = a1 ln[pH2 (0)] a1 (1 − 2f )
4F
  C1 = −
pH2 (0)(1 − f x) 2f
2 22f
× ln 0.209K 2 − 2 ln . × ln [pH2 (0)(1 − f )] − a0
1 − pH2 (0)(1 − f x)
a1 (1 − pH2 (0))2
(13) − ln[1 − pH2 (0)]
2[pH2 (0)]2 2f
After some simple mathematical manipulations, Eq. (8) (1 − pH2 (0))2 (1 − f )
can be written as + a1
2[pH2 (0)]2 2f
  × ln [1 − pH2 (0)(1 − f )]
d2 Tc (x) pH2 (0)(1 − f x) 
= a0 + a1 ln , (14) a1 pH2 (0)(1 − f )
dx 2 1 − pH2 (0)(1 − f x) − ln , (18a)
2 1 − pH2 (0)(1 − f )

where a1 ln [pH2 (0)] a1 (1 − pH2 (0))2


C2 = Tc (0) − +
 22f 2[pH2 (0)]2 2f
j × ln[1 − pH2 (0)].
a0 (Tc (x)) = − Ut + j r eff (18b)
s
 In the gas phase the governing differential equation for the
RTc (x)
− ln 0.209K 2 (15a) heat transfer is
4F
dTg (x) d2 Tg (x)
u(x) = f . (19)
and dx dx 2

j RTc (x) By involving the constant mass flow rate of the gas, m,
a1 (Tc (x)) = . (15b) through a circular surface of diameter d, and by assuming
2 s F
that the atmospheric air behaves as a perfect gas flowing
under fully developed incompressible laminar flow condi-
The general solution of Eq. (14), which contains two real tions, the velocity u(x) can be expressed by the well-known
arbitrary constants C1 and C2 , is of the form: Hagen–Poiseuille flow as (Bird et al., 1960)

(1 − 2f x) 4m
Tc (x) = a0 x + a1
2
ln [pH2 (0)(1 − f x)] u(x) = 2 . (20)
22f d
(1 − pH2 (0))2 (1 − f x) It is worth noticing that the viscosity of the atmospheric
− a1 air (gas mixture) is necessary to be constant in order to use
2[pH2 (0)]2 2f Eq. (20), but viscosity is in general an unknown function of
× ln [1 − pH2 (0)(1 − f x)] the gas temperature. However, it has been presented that the

a1 2 pH2 (0)(1 − f x) variation of air viscosity is less than 4% per 25 K (Perry and
+ x ln Green, 1997) and, thus, it can be considered as constant for
2 1 − pH2 (0)(1 − f x)
variation of gas temperature up to 100 K. On the other hand,
+ C1 x + C 2 . (16) high-temperature differences may destroy the material of the
solid electrolyte and this is another reason for keeping the
The cell surface is assumed to be quasi-isothermal temperature differences up to 100 K.
(Haynes and Wepfer, 2000; Yuan et al., 2001): The differential equation describing the heat transfer in
the gas phase becomes
Tc (1) = Tc (0) (17a) •
4 m dTg (x) d2 Tg (x)
= f . (21)
d 2 dx dx 2
while a standard Dirichlet condition is used for describ-
ing the temperature profile at inlet (Murugesamoorti et al., The general solution of the above equation can be obtained
1993): easily and is of the form:

d 2 f •
e(4 m /d f )x + C4 .
2
Tg (x) = C3 • (22)
Tc (0) = constant. (17b) 4m
ARTICLE IN PRESS
F.A. Coutelieris et al. / Chemical Engineering Science 60 (2005) 4423 – 4430 4427

By employing the boundary conditions: 1200


reff = 1Ω cm2 j = 100 mA/cm
Tg (1) − Tg (2) = Tg (23a)
1100 ∆Tg = 100K j = 300 mA/cm

gas temperature (K)


and (Demin and Tsiakaras, 2001)
1000
Tg (0) = constant and close to Tc (0) (23b)
the above-mentioned arbitrary constants are identified as 900

4Tg m
C3 = • , (24a) 800
d 2 f (1 − e(4 m /d
2
f) )
Tg 700
C4 = Tg (0) − • . (24b) 0.00 0.25 0.50 0.75 1.00
1 − e(4 m /d f )
2
dimensionless distance, x
By employing the specific flux of the heat absorbed by Fig. 3. Spatial distribution of the gas temperature for various current
the cathode gas, the temperature of the cell, Tc (x), is related densities.
to the gas temperature as follows:
j (Ut + j r eff − E(x)) = h(x)(Tc (x) − Tg (x)) (25)
because it is known that gases are bad heat conductors. Ther-
and, therefore, the local heat transfer coefficient, k(x), can mal diffusivities of atmospheric air at various temperatures
be calculated. The overall heat transfer coefficient k0 can be are taken from the appropriate published tables (Perry and
estimated as Green, 1997).
 1  1
Ut + j r eff − E(x)
h0 = h(x) dx = j dx. (26) 4.1. Gas temperature, Tg
0 0 Tc (x) − Tg (x)
Finally, the mass flow rate can be calculated by using the The distribution of the gas temperature along the dimen-
macroscopic balance for the heat absorbed by gas: sionless length of the cell is presented in Fig. 3 for two
typical values of current density (100 and 300 mA/cm2 ).
• Q dj (Ut + j r eff − E)
m= = , (27) The specific effective resistance is 1  cm2 and the temper-
cp Tg c p Tg ature difference in the gas phase, Tg , has been fixed to
where Q is the overall heat absorbed, E denotes averag- 100 K. This value for the temperature difference has been
ing of E(x) given by Eq. (13) along x, and cp is the heat chosen to be low enough to insure that the viscosity does
capacity of the atmospheric air. By assuming the standard not change significantly and, therefore, the Hagen–Poiseuille
environmental composition, cp varies in a range of about flow regime is valid. In general, the higher temperature is
31–34 J/mol of atmospheric air for a given temperature range calculated in the outlet of the cell, as the moving air ab-
of 800–1200 K. sorbs heat and, therefore, its temperature increases as x ap-
A generalized iterative scheme is necessary for the solu- proaches 1. The very high slope of the profiles presented at
tion of the problem, as follows: narrow area near inlet is due to some numerical limitations
in the calculation of electromotive force at this area and is of
• guess an initial value for Tc (x)(previous) ; senseless physical meaning. Finally, a small variable decre-
• calculate a0 and a1 from Eqs. (15a) and (15b); ment (up to 1.2%) of the gas temperature with increment of
• estimate Tc (x) from Eqs. (16), (18a) and (18b); current density is observed. Although a direct dependence

1 of Tg (x) on j is not obvious, it exists through the molar flow
• find Tc (x)(next) = 0 Tc (x) dx;

• if Tc (x)(next) = Tc (x)(previous) stop the procedure as rate m, which depends on the current density (see discussion

the solution is Tc (x) estimated at Step2; otherwise set below). The gas temperature depends on m exponentially in
Tc (x)(previous) = Tc (x)(next) and repeat Step2–Step5; a complicated manner (see Eqs. (22), (24a) and (24b)) of the
• calculate Tg (x) from Eqs. (22), (24a) and (24b); • •
form (ex−1 )m . As x  1, an increment of m corresponds to
• estimate h0 from Eq. (26). •
lower values of (ex−1 )m and, therefore, the small decrement
occurs.
4. Results and discussion Fig. 4 shows the distribution of the gas temperature along
the stack’s length for two typical values of temperature dif-
As it is senseless for a general theoretical study to fix val- ference in the gas phase (Tg = 50 and 100 K). The specific
ues for thermal diffusivities of the solid electrolyte, the ratio effective resistance is again 1  cm2 and the current density
of s /f is used instead of s and it is adjusted to 10,000 j has been fixed to 100 mA/cm2 . According to the results
ARTICLE IN PRESS
4428 F.A. Coutelieris et al. / Chemical Engineering Science 60 (2005) 4423 – 4430

1200 200

Tg(0)=600 K
1100
Tg(0)=800 K
gas temperature (K)

150
Tg(0)=1000 K
1000

reff = 1Ωcm2

∆Tg
900 100
j = 100 mA/cm2
∆T = 100K
800 ∆T = 50K
50

700
0.00 0.25 0.50 0.75 1.00
0
dimensionless distance, x

Fig. 4. Profiles of the gas temperature for various Tg .


0 2 4
.
m
6 8 10

Fig. 6. Relation between the temperature difference and the mass flowrate.
1000

stant, as it has been previously reported in the related bibli-


975 ography (Haynes and Wepfer, 2001). More precisely, there
cell temperature (K)

are some combinations of resistances and currents where


950 j = 100 mA/cm2 the cell temperature presents a really constant profile with
maximum absolute divergence from the base line less than
j = 300 mA/cm2
925 0.0025%. Finally, it should be noted that Tc (x) is indepen-
reff = 1Ωcm2 dent on Tg as it can obviously derived from Eqs. (15a),
900 (15b), (16), (18a) and (18b). It should be also stressed out
that, in general, Tg (x) = Tc (x), which can be interpreted as
a discontinuity on the gas–solid interface of the gas temper-
875
0.00 0.25 0.50 0.75 1.00 ature as it had been defined in the real problem (see Section
2). In fact, it is a reckless and false conclusion as the in-
dimensionless distance, x
terface clearly loses its physical meaning after applying the
Fig. 5. Dependence of the cell temperature on the current density. averaging procedure presented above.


4.3. Molar flow rate, m
presented in Fig. 3, the gas temperature presents its lower
values in the inlet area while it increases as x approaches the •
outlet area. Obviously, higher Tg correspond to higher the Obviously from Eq. (27), the influence of j on m is of
curves of the gas temperature. Finally, it is worth noticing parabolic form. In general, higher current densities corre-
that the second derivative of Tg (x) is of three to four orders spond to higher amounts of heat produced in the electrolyte
of magnitude lower than the first one in all cases. Addition- due to Joule phenomenon and, thus, to higher flow rates of
ally, d2 Tg (x)/dx 2 is multiplied by the small number f and the cathode gas in order to obtain complete heat transfer

the final product becomes even smaller. This underlines the from the electrolyte to gas phase. Furthermore, m depends
negligible character of conduction in the gas phase because on Tg in a reversibly analogous manner for constant j.
of the high Peclet flow of the atmospheric air in the cathode This influence is presented in Fig. 6 for j = 100 mA/cm2 ,
area. reff = 1  cm2 and various initial temperatures Tg (0). In
terms of physics, the temperature difference in the gas phase
4.2. Cell temperature, Tc represents the “flexibility” of the gas in thermal absorption
and, thus, lower values of Tg correspond to higher flow
The profile of the cell temperature is presented in Fig. 5 rates, as the total heat amount should be completely absorbed
for conditions quite analogous to those of Fig. 3 (j = 100 by gas.
and 300 mA/cm2 , reff = 1  cm2 ). As Tc (0) = Tc (1), the
spatial variation of Tc (x) is always less than 0.5% for any 4.4. Local heat transfer coefficient, h(x)
current and resistance (it corresponds to an absolute value of
5 K with respect to the 1000 K of operational temperature), Fig. 7 presents the spatial profile of the local heat trans-
which means that it could be accurately considered as con- fer coefficient, h(x), defined by Eq. (25) as a function of
ARTICLE IN PRESS
F.A. Coutelieris et al. / Chemical Engineering Science 60 (2005) 4423 – 4430 4429

E overall electromotive force


local heat transfer coefficient (W/cm2K)

0.015
reff = 1Ωcm2 F Faradays’ constant (=96, 484 J/mol V)
j = 100 mA/cm2 h(x) local heat transfer coefficient at position x
h0 overall heat transfer coefficient
0.010 j current density per unit area
k(z) thermal conductivity
K equilibrium constant of reaction (1)
L length of the cell stack
0.005 •
m cathode gas mass flow rate
pH2 (x) partial pressure of hydrogen at position x
Q overall heat absorbed by gas
0.000 r, z,  cylindrical spatial coordinates
0.00 0.25 0.50 0.75 1.00 reff effective specific resistance
dimensionless distance, x rohm ohmic resistance
ra , rc polarization anodic and cathodic resis-
Fig. 7. Spatial distribution of the heat transfer coefficient. tance
R universal gas constant (=8.314 J/mol K)
s(x) heat sources at position x
the dimensionless distance for typical values of specific ef- Tc , Tg cell and gas temperature, respectively
fective resistance and current density (reff = 1  cm2 and u(x) one-dimensional fluid velocity at position
j = 100 mA/cm2 ). In agreement to previously presented re- x
sults (Koh et al., 2001), the local heat transfer coefficient is uz (r) z-component of the fluid velocity at posi-
almost constant for a wide domain along the cell except the tion r
inlet and outlet areas, where the end-effects force the con- Ut thermoneutral voltage
vection thermal film to attain its bulk values. Additionally x dimensionless spatial coordinate(=z/L)
to this purpose, the above-mentioned numerical limitation yH2 , yH2 O molar fractions of hydrogen and water, re-
in the calculation of electromotive force at x = 0 influence spectively
significantly to the modulation of such a sharp slope in the
heat transfer coefficient profile.
Greek letters

5. Conclusions , ,  radius of the air pipe, the cathode area and


the electrolyte cylinder, respectively
The steady-state heat transfer problem from the solid elec- s , f thermal diffusivity in solid and fluid phase,
trolyte to the absorbing cathode gas within a fuel cell was respectively
considered here. After a reduction of the dimensions of the −HH2 lower heating value (enthalpy of combus-
original problem by using an averaging technique, the heat tion) of hydrogen at standard conditions
transfer equations accompanied by the appropriate bound- Tg gas temperature difference used in
ary conditions solved analytically for both the solid and gas Eq. (23a)
phases. Analytical expressions for the local and the overall f utilization factor defined by Eq. (12)
heat transfer coefficient were also obtained. It was found  density of atmospheric air
that the cell temperature varied low enough to be considered
as constant. Finally, the overall heat transfer coefficient was
found to present a profile reverse to that of gas temperature. References

Ahmed, S., McPheeters, C., Kumar, R., 1991. Thermal-hydraulic model


of a solid oxide fuel cell. Journal of the Electrochemical Society 138,
Notation 2712.
Appleby, A.J., Foulkes, F.R., 1989. Fuel Cell Handbook. Van Norstand
Reinhold, New York.
a0 , a1 constants defined by Eqs. (15a) and (15b), Bassette, N.F., Wepfer, W.J., Winnick, J., 1995. A mathematical model
respectively of a solid oxide fuel cell. Journal of the Electrochemical Society 142,
cp heat capacity at constant pressure 3792–3800.
Bird, R.B., Stewart, W.E., Lightfoot, E.N., 1960. Transport Phenomena.
C1 , C2 , C3 , C4 integration constants defined by Eqs.
Wiley, New York.
(18a), (18b), (24a) and (24b), respectively Demin, A., Tsiakaras, P., 2001. Thermodynamic analysis of a hydrogen
d diameter of the cell fed solid oxide fuel cell based on a proton conductor. International
E(x) electromotive force at position x Journal of Hydrogen Energy 26, 1103–1108.
ARTICLE IN PRESS
4430 F.A. Coutelieris et al. / Chemical Engineering Science 60 (2005) 4423 – 4430

Djilali, N., Lu, D., 2002. Influence of heat transfer on gas and water Koh, J.-H., Seo, H.-K., Yoo, Y.-S., Lim, H.C., 2001. Consideration of
transport in fuel cells. International Journal of Thermal Science 41, numerical simulation parameters and heat transfer models for a molten
29–40. carbonate fuel cell stack. Chemical Engineering Journal 3913, 1–13.
Dutta, S., Morehouse, J.H., Khan, J.A., 1997. Numerical analysis of Kordesh, J.H., Simander, G., 1996. Fuel Cells and Their Applications.
laminar flow and heat transfer in a high temperature electrolyzer. VCH Publishers, New York.
International Journal of Hydrogen Energy 26, 883–895. Minh, T.Q., Takahashi, T., 1995. Science and Technology of Ceramic Fuel
Dutta, S., Shimpalee, S., Van Zee, J.W., 2001. Numerical prediction of Cells. Elsevier, Amsterdam.
mass-exchange between cathode and anode channels in a PEM fuel Murugesamoorti, K.A., Srinivasan, S., Appleby, A.J., 1993. Research
cell. International Journal of Heat and Mass Transfer 44, 2029–2042. development and demonstration of SOFC systems. In: Blomen,
George, R.A., Bassette, N.F., 1998. Reducing the manufacturing cost of L.J.M.J., Mugerwa, M.N. (Eds.), Fuel Cell Systems. Plenum Press,
tubular SOFC technology. Journal of Power Sources 71, 131–137. New York.
Haynes, C.L., Wepfer, W.J., 2000. Design for power of a commercial grade Perry, R.H., Green, D.W., 1997. Chemical Engineers Handbook. Seventh
tubular solid oxide fuel cell. Energy Conversion and Management 41, ed. McGraw-Hill, New York.
1123–1139. Tsiakaras, P., Demin, A., 2001. Thermodynamic analysis of a solid oxide
Haynes, C.L., Wepfer, W.J., 2001. Characterizing heat transfer within a fuel cell system fuelled by ethanol. Journal of Power Sources 102,
commercial-grade tubular solid oxide fuel cell for enhanced thermal 210–217.
management. International Journal of Hydrogen Energy 26, 369–379. Virkar, A.V., Fung, K., Singhal, S.C., 1997. DOE Report, DOE/GETC/C-
Hirano, A., Suzuki, M., Ippommatsu, M., 1992. Evaluation of a new 98/7303, USA.
solid oxide fuel cell system by nonisothermal modeling. Journal of the Yuan, J., Rokni, M., Sunden, B., 2001. Simulation of fully developed
Electrochemical Society 139, 2744–2751. laminar heat and mass transfer in fuel cell ducts with different cross-
Hirschenhofer, J.H., Stauffer, D.B., Engelman, R.R., 1998. Fuel Cells: A sections. International Journal of Heat and Mass Transfer 44, 4047–
Handbook (rev. 3). US DOE Office of Fossil Energy, Morgantown. 4058.
Kanamura, K., Shoji, Y., Zen-ichiro, T., 1989. Proceedings of the Yuan, J., Sunden, B., 2004. Two-phase flow analysis in a cathode duct of
First International Symposium on SOFCs. Electrochemical Society, PEFCs. Electrochimica Acta 50, 673–679.
Pennigton.

You might also like