You are on page 1of 22

Energy Conversion and Management 45 (2004) 1861–1882

www.elsevier.com/locate/enconman

Three dimensional, two phase flow mathematical model


for PEM fuel cell: Part I. Model development
a,b,*
Mingruo Hu , Anzhong Gu a, Minghua Wang b, Xinjian Zhu b, Lijun Yu c

a
Institute of Refrigeration and Cryogenics, Shanghai Jiao Tong University, 1954 Huashan Road,
Shanghai 200030, PR China
b
Institute of Fuel Cell, Shanghai Jiao Tong University, 1954 Huashan Road, Shanghai, 200030, PR China
c
Department of Energy, School of Mechanical Engineering, Shanghai Jiao Tong University, Shanghai, 200240, PR China
Received 2 December 2002; received in revised form 2 June 2003; accepted 12 September 2003

Abstract
A full three dimensional PEM fuel cell model is developed, which considers not only the rib resistance to
the species but both the single and two phase flow and transport in the gas channels and diffusers at both
the anode and cathode sides of the PEM fuel cell. Two sets of boundary conditions, one for a conventional
flow field and the other for an interdigitated one, are presented. A detailed discussion of the numerical
techniques for the PEM fuel cell model is given with a flow diagram to provide an overview of the solution
procedure using the FORTRAN language. A rigorous validation method is used to show good agreement
between our predicted results and the experimental data.
 2003 Elsevier Ltd. All rights reserved.

Keywords: PEM fuel cell; Two phase flow; Model; Conventional flow field; Interdigitated flow field; SIMPLE algorithm

1. Introduction

As an important method to study proton exchange membrane (PEM) fuel cells, mathematical
models have received strong emphasis in resent years. A good PEM fuel cell model can not only
help to understand the internal mechanisms, such as heat and mass transfer, but also, it can help
to improve the efficiency and save much money when designing and experimenting with fuel cells.
Referring to PEM fuel cell models developed in the most recent ten years, they can be classified

*
Corresponding author. Tel./fax: +86-21-6293-2602.
E-mail address: mingruohu@sh163.net (M. Hu).

0196-8904/$ - see front matter  2003 Elsevier Ltd. All rights reserved.
doi:10.1016/j.enconman.2003.09.022
1862 M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882

Nomenclature

a water activity
Av effective catalyst area per unit volume (m1 )
cf fixed charge concentration in proton exchange membrane (mol m3 )
C mass concentration fraction in gas channel and gas diffuser or mole concentration in
catalyst layer (mol m3 )
Cref reference mole concentration (mol m3 )
D diffusivity (m2 s1 )
F Faraday constant (96,487 C mol1 )
g gravitational acceleration (m s2 )
io;ref reference exchange current density (A m2 )
I current density (A m2 )
j diffusive mass flux (kg m2 s1 )
J capillary pressure function
k conductivity (X m1 )
kp permeability of proton exchange membrane (m2 )
kr relative permeability
K permeability of gas diffuser (m2 )
M molecular weight (kg mol1 )
nd electro-osmotic drag coefficient
N net mass flux in gas channel and gas diffuser (kg cm2 s1 ) or net mole flux in catalyst
layer and membrane (mol cm2 s1 )
P total pressure (Pa)
Pc capillary pressure (Pa)
p partial pressure of gas (Pa)
R gas constant (J mol1 K1 )
Rm membrane resistance (X m2 )
s liquid water saturation
T temperature (K)
t time (s)
u intrinsic velocity vector (m s1 ) or intrinsic velocity in x direction (m s1 )
U cell working voltage (V)
UO thermodynamic open circuit potential (V)
v intrinsic velocity in y direction (m s1 )
Vp total pore volume in gas diffuser (m3 )
Vw liquid water volume in gas diffuser (m3 )
w intrinsic velocity in z direction (m s1 )
x x direction coordinate (m)
y y direction coordinate (m)
z z direction coordinate (m)
M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882 1863

Greek symbols
e porosity
q density (kg m3 )
l dynamic viscosity (kg m1 s1 )
c advection correction factor
t kinematic viscosity (m2 s1 )
/ potential (V)
k relative mobility or water content in proton exchange membrane (mol H2 O/equivalent
SO1
3 )
r surface tension (N m1 ) or local conductivity of proton exchange membrane (X m1 )
a charge transfer coefficient
g overpotential (V)
d coefficient thickness (m)
Subscripts and superscripts
a anode
avg average
c cathode or catalyst layer
eff effective
g gas
in entrance of gas channel
k phase
l liquid
m proton exchange membrane or Naifon phase in catalyst layer
s solid phase
sat saturation
w water
a species

into three categories, i.e. one dimensional models, two dimensional models and three dimensional
models.
As a first step to simulate a PEM fuel cell, Bernardi and Verbrugge [1] and Springer et al. [2]
developed one dimensional models, that is, they only considered the changes across the mem-
brane. These models not only presented some basic results for understanding the internal
mechanisms of a working PEM fuel cell, but more importantly, they also provided a funda-
mental framework to build the multidimensional models that followed. Fuller and Newman [3],
Nguyen and White [4] and Yi and Nguyen [5] started to develop two dimensional models that
considered both the changes across the membrane and in the direction of the bulk flow. In these
models, the authors demonstrated the important roles played by water and thermal management
in maintaining high performance of PEM fuel cells. Furthermore, Gurau et al. [6] and Um et al.
[7] developed two dimensional models using the computational fluid dynamics (CFD) approach.
However, all the models mentioned above neglected the ribs between flow channels. West [8]
1864 M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882

developed a two dimensional model mainly considering the influence of rib spacing in a PEM
fuel cell and concluded that the ribs would restrict the access of fuel and oxidant gases to the
catalyst layer and the movement of water under the ribs. Recently, Shimpalee et al. [9] developed
a three dimensional model. This model used the commercially available computational fluids
code, Fluent, as a solver and presented such results as velocity, mixture density, pressure and
local current contours under the ribs, which could not be understood in a two dimensional
model.
However, all the models mentioned above assumed single phase flow in the gas channels and
gas diffusers and are only valid in the absence of liquid water. During the fuel cell operation,
especially at high current densities, liquid water is likely to appear in the cathode side, resulting in
two phase transport phenomena, and also liquid water will appear in the anode side if the
humidification temperature is higher than that of the PEM fuel cell or liquid is injected directly
into the anode gas channel [4,10]. Baschuk and Li [11] noted that the fraction of liquid water
existing in a gas diffuser was small, having a magnitude of 104 when air was used as oxidant.
However, a small amount of liquid water could block some pores of the gas diffuser and influence
performance at high current densities [11]. Predicting the formation of liquid water and mini-
mizing its detrimental effect on fuel cell performance are important issues in the design and
operation of PEM fuel cells. You et al. [12] presented a two dimensional, two phase flow model to

Fig. 1. Schematic diagram of PEM fuel cells with conventional and interdigitated flow fields.
M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882 1865

consider the transport mechanisms in the two phase flow area. However, this model, which was a
half cell model and only included the gas channel, gas diffuser and catalyst layer in the cathode
side of a PEM fuel cell, could not explain the two phase transport phenomena in the anode side of
a PEM fuel cell, could not find the water distributions in the membrane to direct the humidifi-
cation of a PEM fuel cell and could not know what level of flooding happened inside a gas diffuser
near the rib region.
Recently, some special focuses have been put on a new flow field called an interdigitated flow
field [13]. As shown in Fig. 1, by making the inlet and outlet gas channels dead-ended, the reactant
gases in an interdigitated flow field are forced to flow into the electrode as compared to flowing
over the surface of the electrode in a conventional parallel channel flow field. This design can
convert the transport of the reactant/product gases to/from the catalyst layers from a diffusion
mechanism to a forced convection mechanism with a much reduced gas diffusion boundary layer
over the catalyst sites. Since forced convection is much faster than diffusion, the reaction rates at
the catalyst sites can be significantly enhanced. In addition, the shear force of the gas flow helps
remove most of the liquid water that is entrapped in the inner layer of the electrode, thereby
significantly reducing the electrode flooding problem. Yi [14] developed a cathode half cell model
to compare the performance of PEM fuel cells containing interdigitated and conventional flow
fields. However, this two dimensional model could only describe the cross flow section of PEM
fuel cells, so the boundary conditions at the interfaces between the gas channels and gas diffusers
must be assumed, which actually are not known a priori and are not constant in the main flow
direction. Furthermore, this model neglected the existence of liquid water and did not explain
quantitatively how well an interdigitated flow field helped to remove the liquid water as compared
to a conventional flow field.
This paper forms the first part of a two part study on a PEM fuel cell model. In this paper, a
three dimensional numerical model is developed, which considers not only the rib resistance to the
species but also both the single and two phase flow and transport in the gas channels and diffusers
at both the anode and cathode sides of the PEM fuel cell. Two sets of boundary conditions, one
for a conventional flow field and the other for an interdigitated one, are presented. A detailed
discussion of the numerical techniques for the PEM fuel cell model is given with a flow diagram to
provide an overview of the solution procedure using the FORTRAN language. At last, a
parameter based validation method is used, and the model shows good agreement between
numerically predicted polarization curves and experimental data.

2. Model development

Fig. 2 shows a schematic of a three dimensional PEM fuel cell model. The model regions, which
are symmetrically divided through a PEM fuel cell, consist of half gas channels separated by the
membrane and electrode assembly (MEA) that includes two diffusion layers and two catalyst
layers.
The numerical model assumes that

1. The PEM fuel cell operates under steady state condition. This is because the startup or stop or
any transient process of a fuel cell are not considered in this model.
1866 M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882

Fig. 2. Schematic of a PEM fuel cell model.

2. There are no temperature changes in the PEM fuel cell, i.e. isothermal operation, This is
because, when a single cell is tested, the cell temperature is usually controlled using heating
rods and thermocouples. As a result, the operation temperature of a single fuel cell can be kept
almost constant in all the regions of the working fuel cell.
3. Laminar flow exists everywhere in the gas channels and the flow is fully developed at the exits
of the gas channels. The assumption of laminar flow is based on the small flow velocities and
the small pressure gradients through a single cell [7]. After the gases flow out of the exits of the
gas channels, there are no further chemical reactions, so there are no further changes in the con-
centration of the gases, such as O2 or H2 . As a result, the assumption means that, at the exits of
the gas channels, the gradients of the gas flow at the outlets are zero [7].
4. Isotropic porous media exists in the gas diffusers, catalyst layers and membrane. This assump-
tion asserts that the porosity is a constant in the whole region of the gas diffusers, and the vol-
ume fraction of Nafion in the catalyst layer is also constant.
5. H2 , O2 and N2 are insoluble in the liquid water probably existing in the anode and cathode gas
channels and diffusers. This is an assumption for our two phase flow model. In the gas channel
and gas diffuser, there usually co-exists a gas phase and a liquid water phase. However, the sol-
ubilities of H2 , O2 and N2 in liquid water are quite small, and the diffusivities of H2 , O2 and N2
in liquid water are also quite small compared with the diffusivities of H2 , O2 and N2 in the gas
phase. You [12] also considered O2 and N2 insoluble in liquid water to neglect the extremely
small amount of gas diffusion in the liquid water.

Actually, assumptions 1–4 are usually considered as standard assumptions in many papers [1–
3,6–9,11,12,14,18–20]. Assumption 5 is based on Ref. [12].
In this paper, a ‘‘multiple phase mixture’’ based two phase flow and transport model is developed
to study the two phase flow in the gas diffusers and the gas channels in both the anode and cathode
sides. Wang and Cheng [15] first proposed a detailed multiphase mixture mathematical model for
multiphase flow in the porous media. The key idea in the multiphase mixture model is to focus on
the level of a multiphase mixture, rather than on the levels of the separate phases. The multiple
phases are regarded as constituents of a multiphase mixture in the model. A mass averaged mixture
velocity and a diffusive flux representing the difference between the mixture velocity and an indi-
M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882 1867

vidual phase velocity can describe the multiphase flow. In this definition, the phases are assumed to
be distinct and separable components with non-zero interfacial areas, and their mixture represents
a single fluid with smoothly varying phase compositions [15]. The multiphase mixture model can be
derived from the classical multiphase approach without any additional assumptions [15].

2.1. Model equations

2.1.1. Governing equations in gas channels and gas diffusers


Conservation of mass
oeq
þ r  ðequÞ ¼ 0 ð1Þ
ot
Momentum transport
oðequÞ l
þ r  ðequuÞ ¼ erP þ r  ðelruÞ þ eqg  e2 u ð2Þ
ot K
where the last term represents the drag of the porous solid on the liquid [16].
Species transport [15]
" #
o X
a a a a a a a
e ðqC Þ þ r  ðeca quC Þ ¼ r  ðeqD rC Þ þ r  e ½qk sk Dk ðrCk  rC Þ
ot k
" #
X
r Cka jk ð3Þ
k

Eq. (3) reduces to the following for a two phase (gas and liquid) system
o
e ðqC a Þ þ r  ðeca quC a Þ ¼ r  ðeqDa rC a Þ þ r  fe½q1 sDal rCla þ qg ð1  sÞDag rCga  qDa rC a g
ot
 r  ½ðCla  Cga Þjl  ð4Þ
The definitions of the variables in the above equations are defined in Table 1. The velocity vector u
in Eqs. (1)–(4) is the intrinsic velocity vector based on the open pore area. Eqs. (1), (2) and (4) can
be used to calculate the velocity and concentration fraction distributions in both the single and
two phase zones in the gas channels and gas diffusers. When the liquid saturation s equals zero,
Eq. (4) reduces to the species equation in the single phase zone. When the porosity equals unity
and the permeability equals infinity, Eq. (2) reduces to the momentum equation in the gas
channel.
Eq. (4) can be rewritten for the species H2 , O2 and N2 at the anode and cathode sides as follows.
At the anode side, we get
" #
H2 H2 H2 H2 H2 q
r  ðecH2 quC Þ ¼ r  ðeqDg rC Þ þ r  eð1  sÞqg Dg C r
qg ð1  sÞ
" #
qC H2
þr jl ð5Þ
qg ð1  sÞ
1868 M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882

Table 1
Variables used to describe the two phase flow and transport
Variables Definitions
Liquid water saturation s Vw

Vp
Velocity vector u qu ¼ ql ul þ qg ug
Advection correction factor ca q½kl Cla þ kg Cga 
ca ¼
ql sCla þ qg ð1  sÞCga
Mass concentration fraction for species a qal qag
Cla ¼ Cga ¼
in liquid or gas Cla , Cga q qg
Mass concentration fraction for species a, C a qa
Ca ¼ or qC a ¼ ql sCla þ qg ð1  sÞCga
q
Diffusion coefficient for species a, Da qDa ¼ ql sDal þ qg ð1  sÞDag
Diffusive mass flux of liquid and gas within jg ¼ qg ug  kg qu; jl ¼ ql ul  kl qu or
the two phase mixture jg , jl Kkl kg
jl ¼ ½rPc þ ðql  qg Þg and jl þ jg ¼ 0
m
The individual intrinsic phase velocities eql ul ¼ jl þ kl equ; eqg ug ¼ jl þ kg equ
for liquid and gas ul , ug
 e 12
Capillary pressure between liquid and gas Pc
Pc ¼ Pg  Pl ¼ r J ðsÞ;
K
J ðsÞ ¼ 1:417ð1  sÞ  2:120ð1  sÞ2 þ 1:263ð1  sÞ3
Density q q ¼ ql s þ qg ð1  sÞ
Viscosity l q

krl =ml þ krg =mg
Relative mobility for liquid and gas kl , kg krl =ml krg =mg
kl ¼ kg ¼
krl =ml þ krg =mg krl =ml þ krg =mg
Relative permeability for liquid and gas krl , krg krl ¼ s3 krg ¼ ð1  sÞ3

The water mass concentration fraction in the anode side can be obtained by

C H2 O ¼ 1  C H2 ð6Þ

At the cathode side, we get

" #
q
r  ðecO2 quC O2 Þ ¼ r  ðeqDO O2 O2 O2
g rC Þ þ r  eð1  sÞqg Dg C r
2
qg ð1  sÞ
" #
qC O2
þr jl ð7Þ
qg ð1  sÞ
M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882 1869
" #
q
r  ðecN2 quC N2 Þ ¼ r  ðeqDN N2 N2 N2
g rC Þ þ r  eð1  sÞqg Dg C r
2
qg ð1  sÞ
" #
qC N2
þr jl ð8Þ
qg ð1  sÞ
The water mass concentration fraction in the cathode side can be obtained by
C H 2 O ¼ 1  C O2  C N 2 ð9Þ
In Eqs. (5), (7) and (8), we neglect the transient terms and consider the mass concentration
fractions of hydrogen, oxygen and nitrogen ClH2 , ClO2 and ClN2 in the liquid water equal to zero
according to the assumption above.

2.1.2. Governing equations in the catalyst layers


Fig. 3 shows the construction of a catalyst layer. In this layer, the oxygen or hydrogen dissolves
into the Nafion and diffuses to the surface of the Pt/C where it reacts. The electro-chemical
reactions at both the anode and the cathode sides can be expressed by the following equations.
At the anode side
2H2 ! 4Hþ þ 4e ð10Þ
At the cathode side
4Hþ þ 4e þ O2 ! 2H2 O ð11Þ
From the mass-current balance, we can get
dN 1 dI
¼ ð12Þ
dy nF dy
where n ¼ 2 for H2 and n ¼ 4 for O2 .
A Butler–Volmer expression is used to characterize the potential dependence of the rates of
Eqs. (10) and (11) as follows [17]:
 n     
dI Cm ac F g aa F g
¼ Av io;ref exp  exp  ð13Þ
dy Cref RT RT

0
y δc
Gas pore Pt/C
Nafion

Gas diffuser Membran


O2 e +
H
or H2 H
+

Catalyst layer

Fig. 3. Schematic of a catalyst layer.


1870 M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882

where n ¼ 0:5 for H2 , n ¼ 1 for O2 and g is the total electrode overpotential [6] expressed as
follows [18]:
g ¼ /s  /m þ constant ð14Þ
where /s and /m stand for the potentials of the matrix of the solid phase and the Nafion phase,
respectively, in the catalyst layer.
Assuming that the matrix of the solid phase is equipotential [19], we get the expression for the
overpotential
dg I
¼ eff ð15Þ
dy km
The hydrogen or oxygen diffusion equation can be described as follows:
dCm
N ¼ Deff
m ð16Þ
dy
where kmeff ¼ e1:5 eff 1:5
mc km , Dm ¼ emc Dm [7] in Eqs. (15) and (16).
At the interface between a gas diffuser and a catalyst layer, the hydrogen or oxygen concen-
tration fraction is related with HenryÕs law
p
Cm ¼ ð17Þ
H
where p is the partial pressure of hydrogen or oxygen on the gas phase side, and H is HenryÕs
constant.

2.1.3. Governing equations in proton exchange membrane


The water is transported by three mechanisms in a proton exchange membrane, that is, the
electro-osmotic drag, diffusion and permeation, which are induced separately by the moving
protons, the water concentration difference and the pressure difference between the two sides of
the membrane as shown in Fig. 4. The net water flux through the membrane is the sum of these
three water fluxes and is expressed by the following equation.

Fig. 4. Schematic of water transport phenomena in a proton exchange membrane.


M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882 1871

I kp;m w dp dC w
N w ¼ nd  Cm  Dwm m ð18Þ
F l dy dy
where the electro-osmotic drag coefficient nd is expressed by the following equation [2]:
2:5
nd ¼ k ð19Þ
22
the water mole concentration Cmw in the membrane has a form as [20]
Cmw ¼ kcf ð20Þ
and the water diffusion coefficient Dwm in the membrane can be expressed by the following equation
[9]
  
1 1
Dwm ¼ nd 5:5  1011 exp 2416  ð21Þ
303 T
Substituting Eqs. (19) and (20) into Eq. (18), we get
I kp;m dp dk
N w ¼ 0:1136k  kcf  Dwm cf ð22Þ
F l dy dy
where cf means the fixed charge concentration in the membrane, and k stands for the water
content (mol H2 O/equivalent SO1
3 ). The water content k can be expressed by [2]

k ¼ 0:043 þ 17:81a  39:85a2 þ 36:0a3 ð0 < a 6 1Þ ð23Þ

k ¼ 14:0 þ 1:4ða  1Þ ð1 < a 6 3Þ ð24Þ

k ¼ 16:8 ða P 3Þ ð25Þ
a being the water vapor activity given by
xw P
a¼ ð26Þ
pwSat
where xw is the mole fraction of water.

2.1.4. Fuel cell voltage


The fuel cell voltage is given by
U ¼ UO  ga  gc  Iavg Rm ð27Þ
where UO is the thermodynamic open circuit potential given by [1]
RT 2
UO ¼ 1:23  0:9  103 ðT  298Þ þ 2:3 ðp pO Þ ð28Þ
4F H2 2
The average current density is determined by
R
Iðx; zÞ
Iavg ¼ A ð29Þ
A
where A is the reaction area. The membrane resistance Rm in X m2 is given by
1872 M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882
Z dm
dy
Rm ¼ ð30Þ
0 rðkÞ
where rðkÞ is the local conductivity of the membrane. Springer et al. [2] presented the experi-
mental fit of conductivity for Nafion117 as follows:
  
1 1
rðkÞ ¼ exp 1268  ð0:5139k  0:326Þ ð31Þ
303 T

3. Boundary conditions

The computation domains, as shown in Fig. 5, are composed of half gas channels, gas diffusers
and catalyst layers at each side of the PEM fuel cell and proton exchange membrane.

3.1. Boundary conditions for gas channels and gas diffusers

At the entrances of the gas channels, constant flow rate and mass concentration fraction of each
species are specified, and at the exits, both the velocity and concentration fraction distributions
are assumed to be fully developed. The boundary conditions for the gas channels and gas diffusers
in the two PEM fuel cells with conventional and interdigitated flow fields are given separately as
follows.

3.1.1. For the conventional flow field


At faces F0 M0 N0 E0 , FMNE, B0 O0 P0 A0 and BOPA (the entrances of the half gas channels)
u ¼ uin ; v ¼ 0; w ¼ 0; C a ¼ Cina ð32Þ
at faces G0 Q0 R0 H0 , GQRH, K0 S0 T0 L0 and KSTL (the exits of the half gas channels)
ou ov ow oC a
¼ 0; ¼ 0; ¼ 0; ¼0 ð33Þ
ox ox ox ox

Fig. 5. Schematic of the computation domains.


M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882 1873

at faces EE0 D0 D, BB0 C0 C, HH0 I0 I and KK0 J0 J (the front and end faces of the diffusion layers)
oC a
u ¼ 0; v ¼ 0; w ¼ 0; ¼0 ð34Þ
ox
at faces FGID, F0 G0 I0 D0 , CJLA and C0 J0 L0 A0 (the side faces of the half gas channels and diffusers),
having the symmetrical conditions
ou ov ow oC a
¼ 0; ¼ 0; ¼ 0; ¼0 ð35Þ
oz oz oz oz
at faces FMQG, F0 M0 Q0 G0 , APTL and A0 P0 T0 L0 (the top and bottom faces of the gas channels)
oC a
u ¼ 0; v ¼ 0; w ¼ 0; ¼0 ð36Þ
oy
at the interface between the cathode gas diffuser and catalyst layer
N O2 þ N H2 O
u ¼ 0; v¼ ; w¼0
eq
h i
q
oCO2 o qg ð1sÞ qC O2
N O2 ¼ ecO2 qvC O2  eqDO
g
2
 eð1  sÞqg DO
g C
2 O2
 jl;y
oy oy qg ð1  sÞ
h i ð37Þ
q
oC N2 o qg ð1sÞ qC N2
N N2 ¼ ecN2 qvC N2  eqDN
g
2
 eð1  sÞq DN2 N2
g g C  jl;y
oy oy qg ð1  sÞ
at the interface between the anode gas diffuser and catalyst layer
N H2 þ N H2 O
u ¼ 0; v¼ ; w¼0
eq
h i
q
oCH2 o qg ð1sÞ qC H2
N H2 ¼ ecH2 qvC H2  eqDH
g
2
 eð1  sÞqg DH
g C
2 H2
 jl;y ð38Þ
oy oy qg ð1  sÞ

3.1.2. For the interdigitated flow field


At faces F0 M0 N0 E0 and B0 O0 P0 A0 (the entrances of the inlet gas channels)
u ¼ uin ; v ¼ 0; w ¼ 0; C a ¼ Cina ð39Þ
at faces FMNE and BOPA (the dead ends of the outlet gas channels)
oC a
u ¼ 0; v ¼ 0; w ¼ 0; ¼0 ð40Þ
ox
at faces GQRH and KSTL (the exits of the outlet gas channels)

ou ov ow oC a
¼ 0; ¼ 0; ¼ 0; ¼0 ð41Þ
ox ox ox ox
1874 M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882

at faces G0 Q0 R0 H0 and K0 S0 T0 L0 (the dead ends of the inlet gas channels)


oC a
u ¼ 0; v ¼ 0; w ¼ 0; ¼0 ð42Þ
ox
The other boundary conditions for the gas channels and gas diffusers in a PEM fuel cell with
interdigitated flow fields are the same as Eqs. (34)–(38).
The net mass fluxes of species at the interface between the gas diffusers and the catalyst layers in
Eqs. (37) and (38) are given by
MH2 I
N H2 ¼ ; NaH2 O ¼ MH2 O N w ð43Þ
2F
MO2 I MH2 O I
N O2 ¼  ; N N2 ¼ 0; NcH2 O ¼ þ NaH2 O ð44Þ
4F 2F
Here, we do not present the boundary conditions at the surfaces of ribs because they are closely
related with the solution techniques introduced in the next section.

3.2. Boundary conditions for the catalyst layer and proton exchange membrane

As the current density distribution at the interface between a catalyst layer and a membrane
and the reactant concentration fraction distributions at the interface between a gas diffuser and a
catalyst layer are not known a priori, the total electrode overpotential in the catalyst layer pre-
sented in Eq. (13) is initially assumed [6].
Referring to Fig. 3, the boundary conditions for a membrane are given by
at y ¼ 0; k ¼ kð0Þ and at y ¼ dm ; k ¼ kðdm Þ ð45Þ

4. Numerical techniques and procedures

The model equations developed above are coupled closely, so the whole set of equations must
be solved simultaneously and iteratively. In the gas channels and gas diffusers, the model equa-
tions, including the continuity, momentum and species equations, are extremely similar to each
other as mentioned above, so these two regions can be solved together, and the boundary con-
ditions between these two layers are omitted. The SIMPLE algorithm [21] is used to solve the
transport equations in the gas channels and gas diffusers. However, the ribs between two
neighboring channels, as shown in Fig. 5, cause some difficulties when using the SMPLE algo-
rithm to solve such non-rectangular boundaries, excluding the ribs.
In order to solve the above difficulties, we consider the ribs as a special fluid, which has an
extremely large viscosity (for example, l ¼ 1030 ) and unity mass concentration fraction. First, we
consider the region F0 G0 GFDII0 D0 or A0 L0 LACJJ0 C0 in Fig. 5, including a gas diffuser, two half
gas channels and a rib, as an integrated domain for calculation. Second, we must give the
additional boundary conditions not mentioned in the above section, that is, at faces MQQ0 M0 and
PTT0 P0 shown in Fig. 5, that is, the interfaces between the ribs and the plates
u ¼ 0; v ¼ 0; w ¼ 0; Ca ¼ 0 ða ¼ O2 ; N2 ; H2 Þ; C Rib ¼ 1 ð46Þ
M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882 1875

at faces MM0 N0 N, OO0 P0 P, QQ0 R0 R and SS0 T0 T, the front and end faces of the anode and cathode
ribs

u ¼ 0; v ¼ 0; w ¼ 0; Ca ¼ 0 ða ¼ O2 ; N2 ; H2 Þ; C Rib ¼ 1 ð47Þ

The large viscosity set purposely in a rib region and the zero velocities set on the three surfaces
of the rib make this region static. At last, the large source term [21] is used to freeze the reactant
concentration fractions at zero in the whole rib region, and the rib concentration fraction at unity
in the rib region and at zero in the gas channels and gas diffusers. Hence, the mass concentration
fractions are coupled automatically at the interfaces between the gas channels and the rib, and so,
it is no use presenting the boundary conditions for faces MNRQ, M0 N0 R0 Q0 , OPTS and O0 P0 T0 S0
in Fig. 5.
The shooting method and fourth order Runge–Kutta method are used separately to solve the
equations in the catalyst layers and equation in the membrane.

Begin

Set initial values at anode and cathode


gas channels and gas diffusers

Solve the catalyst subroutine and find


local current density

Solve the membrane subroutine and


find net water flux

Solve anode x, y, z-Momentum, correct Solve cathode x, y, z-Momentum, correct


pressure and find x, y, z velocity pressure and find x, y, z velocity

Solve H2-transport Solve O2 and N2-transport

No
ξ < exp ?

Yes
Output results

End

Fig. 6. Flow diagram of the solution procedure.


1876 M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882

Fig. 6 presents the coupled algorithms to solve the PEM fuel cell model. The solution is
considered to be convergent when the relative error between two consecutive iterations is less than
a small value.

5. Model validation

Because of the small size of the gas channels and the small thickness of the gas diffuser, catalyst
layer and membrane, it is very difficult to measure the flow field and concentration field in the gas
channel and gas diffuser, and it is also very difficult to measure the concentration field and local
current density distributions in the catalyst layer and in the membrane [22]. As a result, few papers
can be found to address measurement of such fields or distributions. However, voltage vs. current
density curves, which are highly related to the distributions in a PEM fuel cell, can be measured
directly and accurately in experiments [23]. As a result, all PEM fuel cell models published pre-
viously have used experimental voltage vs. current curves to validate the reliability of these
models, with some parameters supplied by the experiments and others by the models [1,6,7,22,25].
To validate the current three dimensional, two phase flow model, comparisons are made to the
experimental data of Ticianelli et al. [23] for a single cell. The schematic of the computation
domains are presented in Fig. 5, and the basic parameters used in the model validation are given
in Table 2. The grid numbers of 42 · 42 · 42, used separately in the anode and cathode gas
channels and gas diffusers, provides satisfactory results.
In Table 2, the parameters for fuel cell operating conditions, such as cell temperature and
pressure, were supplied in Ref [23]. However, the parameters, such as the exact fuel cell geometry
and stoichiometric flow ratio are unknown in Ref. [23]. Usually, papers presenting experimental
data do not supply the fuel cell geometries. However, this is not a problem for using these
experimental data to validate a new model through the following steps. The common method for
validation is as follow [1,6,7,22,25]:
First, supply a set of geometry parameters, stoichiometric flow ratios or other operating con-
ditions for the PEM fuel cell model if such parameters are not given with the experimental data.
As a result, Av iref ref
o;c and Av io;a are the only two adjustable parameters.
Second, adjust Av iref ref
o;c and Av io;a to fit the numerically predicted curve to its corresponding
experimental voltage vs. current density curve for the PEM fuel cell.
Usually, based on the above two steps, a model can be proved to be a good one by only fitting
one predicted curve to its corresponding experimental voltage vs. current density curve through
adjusting Av iref ref
o;c and Av io;a [1,6,22,25]. However, here, a third step is used to eliminate the possible
inaccuracy of model validation, which is derived from the supplied parameters in step one. For
example, through a special choice for the geometry parameters and adjusting Av iref ref
o;c and Av io;a , a
bad model possibly can show good agreement between predicted results and experimental data,
but such parameters and Av iref ref
o;c and Av io;a are closely related to this experimental voltage vs. current
density curve and cannot be used in other experimental voltage vs. current density curves for the
same fuel cell.
Third, choose another experimental voltage vs. current density for the same fuel cell, for
example, choosing an experimental voltage vs. current density at a different operating pressure. If
the same value for Av iref ref
o;c and Av io;a , which is determined in step 2, can make the numerically
M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882 1877

Table 2
Basic parameters for PEM fuel cells
Quantity Value Source
2
Gas channel length FG, BK 7.11 · 10 m [6,7]
Gas channel height FE, AB 7.62 · 104 m [6,7]
Half gas channel width F0 M0 , MF or B0 O0 , OB 5 · 104 m –
Rib width M0 M or P0 P 1 · 103 m –
Gas diffuser height ED, BC 2.54 · 104 m [6,7]
Catalyst layer thickness dc 2.87 · 105 m [6,7]
Membrane thickness for Nafion 117 dm 1.75 · 104 m [23]
Gas diffuser porosity e 0.4 [6,7]
Volume fraction of Nafion in the catalyst layer em;c 0.4 [6,7]
Cell Temperature T 353 K [23]
Anode and cathode side inlet pressure Pa =Pc 3.03 · 105 /5.05 · 105 Pa [23]
Anode fuel stoichiometric flow ratio 3 –
Cathode oxidant stoichiometric flow ratio 4 –
Humidification temperature for anode inlet fuel 358 K [7]
Relative humidity of anode inlet fuel 100% [7]
Relative humidity of cathode inlet air 0% [7]
Fixed charge concentration in the membrane cf 1.2 · 103 mol/m3 [1,6,7]
Oxygen diffusivity in gas DO g
2 5.22 · 106 m2 /s [24]
Nitrogen diffusivity in gas DN g
2 2.2 · 106 m2 /s [24]
Hydrogen diffusivity in gas DH g
2 3.76 · 106 m2 /s [24]
Oxygen diffusivity in Nafion DO m
2 2.0 · 108 m2 /s [7]
Hydrogen diffusivity in Nafion DH m
2 2.59 · 1010 m2 /s [7]
Ionic conductivity of Nafion in catalyst layer km 17 S/m [18]
Permeability of gas diffuser K 1.76 · 1011 m2 [1,6,7]
Permeability of membrane kp 1.8 · 1018 m2 [1,6,7]
Anodic transfer coefficient aa 2 [1,6,7]
Cathodic transfer coefficient ac 2 [1,6,7]
Reference exchange current density multiplied by area of anode Av iref
o;a 5.2 · 108 A/m3 –
Reference exchange current density multiplied by area of cathode Av iref
o;c 110 A/m3 –

predicted curve show good agreement with the newly selected experimental voltage vs. current
density curve for the same fuel cell, then the model is proved to be an accurate model, and the
possibility of inaccuracy of model validation, which is derived from the supplied parameters, is
eliminated. Furthermore, because irefo;c is related to temperature [26], when choosing an experi-
mental voltage vs. current density at a different operating temperature, a further step should be
done as follows: While Av is a constant for a certain electrode [18], Av iref
o;c for other temperatures
should be calculated based on Eq. (48), which is a correlation of the exchange current density and
the fuel cell temperature presented in Ref. [26]. Then, if the newly calculated Av iref
o;c can make the
numerically predicted curve show good agreement with the newly selected experimental voltage
vs. current density at the different operating temperature, then the model is proved to be an
accurate model.

4001
log10 ðiref
o;c Þ ¼ 3:507  ð48Þ
T
1878 M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882

As a result, the probably curve related inaccuracy, which is derived from the selected para-
meters, is omitted using the third step. The same method was used in Ref. [7] to present an
accurate model.
The parameters that are additionally required to apply our model when using the experimental
data presented in Ref. [23] are the fuel cell geometry and stoichiometric flow ratio, leaving the
Av iref ref
o;c and Av io;a to be adjusted. In order to compare the current model with Ref. [7], which also
used the same experimental data presented in Ref. [23], our model shares parameters for the
geometry and stoichiometric flow ratio with Ref. [7] and presents half gas channel width and rib
width for the three dimensional property of our model. Based on the three steps mentioned above,
a good agreement is reached between the predicted results and the experimental data at two
operating temperatures, but with larger values for Av iref o;c than those presented in Ref. [7]. When
considering the half gas channel width and rib width as the only differences between the supplied
parameters listed in Table 2 and those presented in Ref. [7], it is deduced that the increased value
of Av iref
o;c is mainly derived from the addition of such three dimensional geometry parameters. In
consequence, the cathode oxidant stoichiometric flow ratio is increased to 4, an increase of 33%
compared with the 3 presented in Ref. [7], and then, the new value found for Av iref o;c at 353 K is
almost the same as that presented in Ref. [7]. The newly predicted curves with an oxidant stoi-
chiometric flow ratio of 4 show good agreement with the experimental data at two operating
temperatures, as shown in Fig. 7. The main reason for the above operation can be explained as
follows: First, referring to Eq. (13), it is found that the oxygen mole concentration Cm and Av iref o;c
are in direct proportion to the current density. Second, the model presented in Ref. [7] is a two
dimensional model and neglects oxygen transport in the z direction, as shown in Fig. 2. As a
result, the oxygen mole concentration Cm at the interface between the catalyst layer and the gas
diffusion layer is always over predicted because a two dimensional model considers a uniform
distribution of oxygen in the z direction [6,7], and then, a smaller value for Av iref
o;c is acquired when
fitting the modeling results to experimental data because of the relation in Eq. (13), while with the
same stoichiometric flow ratio of 3, in order to fit our model results to the same experimental data,
a larger value for Av iref
o;c is acquired in the current model because of the decreasing value of Cm in
the z direction as shown in Ref. [22]. As a result, an increase of stoichiometric flow ratio, which
can supply much more oxygen to the region above the rib as shown in Fig. 2, increases the value

Fig. 7. Comparison of numerically predicted polarization curves with experimental data.


M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882 1879

for Cm , and, hence, decreases the adjustable value for Av iref


o;c in our model to show good agreement
with the same experimental data in Fig. 7.
Furthermore, Fig. 8 shows comparisons of the predicted polarization curves in Fig. 7 with the
predicted curves that are calculated by the current model but using the same oxidant stoichio-
metric flow ratio taken from Ref. [7]. As shown in Fig. 8, when the smaller stoichiometric flow
ratio presented in Ref. [7] is used in our three dimensional model (the other parameters in Table 2
are not changed), the current model under predicts the fuel cell performance. Based on the reasons
mentioned above, it can be concluded that the two dimensional, single phase model presented in
Ref. [7], which did not take into account the rib resistance to the cross diffusion of oxygen and did
not take into account the liquid water partial flooding in the gas diffuser, always over predicted
the performance of the PEM fuel cell with a smaller inlet velocity. On the contrary, with such a
smaller inlet velocity, the fuel cell performance presented in Ref. [23] cannot be reached because of
the lack of enough oxygen in the region above the rib, as shown in Fig. 2, where the current
density is lowest for a conventional flow field as shown in Ref. [22]. As a result, the three
dimensional, two phase flow model developed above is superior to the two dimensional, single
phase models developed before [6,7].
The slight differences between the predicted curves and the experimental data that are shown
in Fig. 7 and can also be seen in Ref. [7] are mainly derived form Eq. (48). Actually, based on
the first two steps mentioned above, a predicted curve in our model can show better agreement
with its corresponding experimental voltage vs. current density at an operating temperature by
adjusting Av iref ref
o;c and Av io;a . In order to further validate the accuracy of the PEM fuel cell model
at different temperature, Eq. (48) should be used after acquiring Av iref o;c at a temperature to
ref
calculate another Av io;c at another temperature. Although Eq. (48) was proved to be accurate
based on the experimental data of Parthasarathy [26], it probably shows a little error when used
in other cases [7]. In consequence, a little accuracy for Av iref o;c at a temperature should be sac-
rificed in order to acquire more accuracy for Av iref o;c at another temperature to get totally good
agreement between the predicted curves with the experimental data at two operating temper-
atures. As a result, the slight differences shown in Fig. 7 are derived from Eq. (48), instead of
the model itself.

Fig. 8. Comparisons of predicted polarization curves in Fig. 7 with predicted curves using 2D parameters taken from
Ref. [7].
1880 M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882

Fig. 9. Comparison of numerically predicted polarization curves for PEM fuel cells using conventional and inter-
giditated flow fields without cathode humidification at an operating temperature of 353 K.

At last, based on the validation result at 353 K, as shown in Fig. 7, a comparison between the
predicted performances of PEM fuel cells with a conventional flow field and an interdigitated one
is presented in Fig. 9. The parameters presented in Table 2 are also used in modeling the PEM fuel
cell with the interdigitated flow field. As a result, the performance difference between these two
fuel cells is only derived from the different structures of these two flow fields. As shown in Fig. 9,
the performance of the PEM fuel cell with interdigitated flow field is poorer than that of a PEM
fuel cell with conventional flow field when no water is added to humidify the cathode inlet air
(base case shown in Table 2), a case that was verified by the experiment of Wood et al. [10]. The
reason for this phenomenon is mainly derived from a larger ohmic overpotential in the membrane
of a PEM fuel cell with interdigitated flow field than that in the membrane of a PEM fuel cell with
conventional flow field, and it will be discussed in detail in Part II of this series [27].

6. Conclusion

A new PEM fuel cell model has been developed, which considers species transports or reactions
in both the anode and cathode gas channels, gas diffusers and catalyst layers and in the proton
exchange membrane, that considers the impact of ribs on the species transport to extend this
model to a three dimensional one and considers two phase flow and transport mechanisms of the
species in both the anode and cathode gas channels and gas diffusers. Two sets of boundary
conditions, one for a conventional flow field and the other for an interdigitated one, have been
presented. A SMPLE algorithm coupled with the shooting method and fourth order Runge–
Kutta method has been used to solve the equations and show a good agreement between the
predicted results and experimental data. Furthermore, when applying the same oxidant stoichi-
ometric flow ratio taken from Ref. [7] to the current model, it is found that a two dimensional,
single phase model always over predicts the performance of the same PEM fuel cell. A comparison
between the predicted performances of PEM fuel cells with a conventional flow field and an
interdigitated one shows the performance of a PEM fuel cell with interdigitated flow field is poorer
than that of a PEM fuel cell with conventional flow field when no water is added to humidify the
cathode inlet air.
M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882 1881

The internal transport mechanisms in PEM fuel cells, which are derived from the above cal-
culation as shown in Fig. 9, will be discussed in Part II of this series [27].

Acknowledgements

This work was supported by the ‘‘985’’ Funds of Shanghai Jiao Tong University, Shanghai,
China and by National Nature Science Foundation of China (No. 50206012). The authors would
like to thank Dr. Hongtan Liu, Director of Dorgan Solar Energy and Fuel Cell Laboratories,
University of Miami, USA, for his constructive advices for revising this paper and his kindness
and patience.

References

[1] Bernardi DM, Verbrugge MW. A mathematical model of the solid-polymer-electrolyte fuel cell. J Electrochem Soc
1992;139(9):2477–91.
[2] Springer TE, Zawodzinski TA, Gottesfeld S. Polymer electrolyte fuel cell model. J Electrochem Soc
1991;138(8):2334–42.
[3] Fuller TF, Newman J. Water and thermal management in solid-polymer-electrolyte fuel cells. J Electrochem Soc
1993;140(5):1218–24.
[4] Nguyen TV, White RE. A water and heat management model for proton-exchange-membrane fuel cells.
J Electrochem Soc 1993;140(8):2178–86.
[5] Yi JS, Nguyen TV. An along-the-channel model for proton exchange membrane fuel cells. J Electrochem Soc
1998;145(4):1149–59.
[6] Gurau V, Liu H, Kakac S. Two-dimensional model for proton exchange membrane fuel cells. AIChE
J 1998;44(11):2410–22.
[7] Um S, Wang CY, Chen KS. Computational fluid dynamics modeling of proton exchange membrane fuel cells.
J Electrochem Soc 2000;147(12):4485–93.
[8] West AC. Influence of rib spacing in proton-exchange membrane electrode assemblies. J Appl Electrochem
1996;26:557–65.
[9] Dutta S, Shimpalee S, Van Zee JW. Three-dimensional numerical simulation of straight channel PEM fuel cells.
J Appl Electrochem 2000;30:135–46.
[10] Wood III DL, Yi JS, Nguyen TV. Effect of direct liquid water injection and interdigitated flow field on the
performance of proton exchange membrane fuel cells. Electrochim Acta 1998;43(24):3795–809.
[11] Baschuk JJ, Li X. Modelling of polymer electrolyte membrane fuel cells with variable degrees of water flooding.
J Power Source 2000;86(1):181–96.
[12] You L, Liu H. A two-phase flow and transport model for the cathode of PEM fuel cells. Int J Heat Mass Transfer
2002;(45):2277–87.
[13] Nguyen TV. A gas distributor design for proton-exchange-membrane fuel cells. J Electrochem Soc 1996;
143(5):L103–5.
[14] Yi JS, Van Nguyen T. Multicomponent transport in porous electrode of proton exchange membrane fuel cells
using the interdigitated gas distributors. J Electrochem Soc 1999;146(1):38–45.
[15] Wang CY, Cheng P. A multiphase mixture model of multiphase, multicomponent transport in capillary porous
media––I. Model development. Int J Heat Mass Transfer 1996;39(10):3607–18.
[16] Wang CY, Gu WB. Micro-macroscopic coupled modeling of batteries and fuel cells––I. Model development.
J Electrochem Soc 1998;145(10):3407–17.
[17] Weisbrod KR, Grot SA, Vanderborgh NE. Through-the-electrode model of a proton exchange membrane fuel cell.
Electrochem Soc Proc 1995;95–23:152.
1882 M. Hu et al. / Energy Conversion and Management 45 (2004) 1861–1882

[18] Marr C, Li X. Composition and performance modeling of catalyst layer in a proton exchange membrane fuel cell.
J Power Source 1999;77(1):17–27.
[19] You L, Liu H. A parametric study of the cathode catalyst layer of PEM fuel cells using a pseudo-homogeneous
model. Int J Hydrogen Energy 2001;26(9):991–9.
[20] Ge S, Yi B, Xu H. Model of water transport for proton-exchange membrane fuel cell (PEMFC). J Chem Ind Eng
(China) 1999;50(1):39–47.
[21] Patankar SV. Numerical heat transfer and fliud flow. New York: Hemisphere; 1980.
[22] Berning T, Lu DM, Djilali N. Three-dimensional computational analysis of transport phenomena in a PEM fuel
cell. J Power Source 2002;106:284–94.
[23] Ticianelli EA, Derouin CR, Srinivasan S. Localization of platimun in low catalyst loading electrodes to attain high
power densities in SPE fuel cells. J Electroanal Chem 1988;251:275–95.
[24] Cussler EL. Diffusion mass transfer in fluid systems. New York: Cambridge University Press; 1984.
[25] Rowe A, Li X. Mathematical modeling of proton exchange membrane fuel cells. J Power Source 2001;102(1–2):
82–96.
[26] Parthasarathy A, Srinivasan S, Appleby J. Temperature dependence of the electrode kinetics of oxygen reduction at
the platinum/Nafion interface––a microelectrode investigation. J Electrochem Soc 1992;(139):2530.
[27] Hu M, Zhu X, Wang M, Gu A. Three dimensional, two phase flow mathematical model for PEM fuel cell: Part II.
Analysis and discussion of the internal transport mechanisms. Energy Convers Manage, doi:10.1016/j.encon-
man.2003.09.023.

You might also like