A General Model of Proton Exchange Membrane Fuel Cell
A General Model of Proton Exchange Membrane Fuel Cell
A General Model of Proton Exchange Membrane Fuel Cell
a r t i c l e i n f o a b s t r a c t
Article history: In this study, a general model of proton exchange membrane fuel cell (PEMFC) was constructed, imple-
Received 18 January 2008 mented and employed to simulate the fluid flow, heat transfer, species transport, electrochemical reaction,
Received in revised form 20 March 2008 and current density distribution, especially focusing on liquid water effects on PEMFC performance. The
Accepted 20 March 2008
model is a three-dimensional and unsteady one with detailed thermo-electrochemistry, multi-species,
Available online 27 March 2008
and two-phase interaction with explicit gas–liquid interface tracking by using the volume-of-fluid (VOF)
method. The general model was implemented into the commercial computational fluid dynamics (CFD)
Keywords:
software package FLUENT® v6.2, with its user-defined functions (UDFs). A complete PEMFC was consid-
PEM fuel cell
UDF
ered, including membrane, gas diffusion layers (GDLs), catalyst layers, gas flow channels, and current
FLUENT® collectors. The effects of liquid water on PEMFC with serpentine channels were investigated. The results
3D showed that this general model of PEMFC can be a very useful tool for the optimization of practical
Multi-phase engineering designs of PEMFC.
VOF © 2008 Elsevier B.V. All rights reserved.
0378-7753/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.jpowsour.2008.03.047
198 A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222
Nomenclature
e electrochemical reaction
a water activity eff effective
As heat transfer surface area, m2 g gas phase
Asurf reactive surface area, m2 H2 hydrogen
cf concentration of sulfonic acid ions (HSO3 − ), H2 O water
kmol m−3 i species i
cp specific heat capacity, J kg−1 K−1 in inlet
Ci species concentration i, kmol m−3 l liquid phase
Di diffusion coefficient of species i in gas mixture, m membrane phase
m2 s−1 out outlet
F Faraday constant, 9.6487 × 107 C kmol−1 N2 nitrogen
h convective heat transfer, W m−2 K−1 O2 oxygen
hH2 O the enthalpy of formation of water vapor, N m kg−1 ref reference
I current density, A m−2 s solid phase
Iave average current density, A m−2 sat saturated
J mass flux, kg m−2 s−1 surf surface
keff effective thermal conductivity, W m−1 K−1 w water vapor
Mi molecular weight of species i in gas mixture,
kg kmol−1
nd electro-osmotic drag coefficient transport phenomena in micro-scale parallel flow channels was
nf charge number of the sulfonic acid ion conducted by Cha et al. [4] in which oxygen concentration along
P pressure, Pa a single gas flow channel and other flow patterns that may affect
Pi partial pressure of species i, Pa fuel cell performance were discussed. Similarly, gas concentration
R universal gas constant, 8314 J kmol−1 K−1 of a steady-state flow along fuel cell flow channels was obtained
Q̇ heat rate, W numerically by Kulikovsky [5]. However, in all the studies men-
Rcat , Ran volumetric current density, A m−3 tioned above, the effects of liquid water were neglected. Yi et al. [6]
S source term pointed out that water vapor condensation was inevitable on both
t time, s the anode and cathode sides of a PEMFC, and they discussed a liquid
T temperature, K water removal technique that used a water transport plate to lead
u, v, w velocities in X, Y, and Z directions, respectively, m s−1 excess liquid water to the coolant flow channels by a pressure dif-
Voc open-circuit potential, V ference. Wang et al. [7] conducted a two-phase model on PEMFC
Vcell cell potential, V cathode to address the liquid water saturation. You and Liu [8]
Vref reference potential, V also considered liquid water saturation in a straight channel on the
volume, m3 cathode side. Both the Refs. [7,8] showed the importance for consid-
Xi mole fraction of species i ering liquid water in numerical modeling of PEMFCs. A 3D model
Yi mass fraction of species i by Natarajan and Nguyen [9] introduced the effects of flow rate,
inlet humidity and temperature on the liquid water flooding in the
Greek symbols
PEMFC cathode. The numerical results of this model were validated
˛ transfer coefficient
with experimental data. Berning and Djilali [10] presented a multi-
ˇ the factor accounts for energy release
phase, multi-component 3D model with heat and mass transfer,
concentration dependence
where liquid water transport in GDLs was numerically modeled by
p, T exponent factors
using viscous and capillary effects. This method was also used in
ε porosity
a 3D model by Mazumder and Cole [11]. Other approach to deal
w contact angle, ◦
with two-phase flow in GDLs based on thermodynamic equilib-
overpotential, V
rium conditions was given by Vynnycky and Vynnycky [12]. In this
surface curvature
approach, the location of interface between one phase and the other
p hydraulic permeability, m2
phase was pointed out from its numerical results. In recent years,
electrokinetic permeability, m2
more two-phase models have been published [13–15], these sim-
water content
ulations predicted water flooding inside PEMFCs, and the liquid
dynamic viscosity, kg m−2 s−1
water effects on PEMFC performances. Large-scale simulations for
density of gas mixture, kg m−3
complex flow field were also performed with experimental valida-
i density of species i, kg m−3
tions [15–18].
phase conductivity, −1 m−1
All the above research papers have significantly contributed to
gaseous permeability, m2
PEMFC research and development. However, by far, to the authors’
reaction rate, kmol m2 s−1
knowledge, most of the two-phase numerical models have not con-
phase potential, V
sidered the interface tracking between liquid water and gas. The
ϕ relative water content
detailed behaviors of liquid water transport inside the PEMFCs were
surface tension coefficient, N m−1
rarely discussed except for the present authors’ previous study [19],
ω excess coefficient
which only dealt with part of serpentine channels—the single U-
shaped channel. Recently, Zhou et al. [20–22] also conducted two
Subscripts and superscripts
more studies that dealt with liquid water in serpentine and straight
an anode
parallel fuel cell stacks. The results showed that different designs of
cat cathode
gas diffusion layers (GDLs) and flow channels will affect the liquid
A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222 199
4. The fuel cell cooling was controlled by forced convection heat Table 2
The source terms of governing equations
transfer.
Governing equation Volumetric source terms and location of
2.2. Governing equations application
For membrane:
= sl l + (1 − sl ) g (2) 2 ∇s
Sv = g − ε2 v + ( + l ) + p cf nf F∇m
l g
sl + sg = 1 (3)
I2
Conservation of energy For current collectors: ST = s
MH
For a mixture of liquid and gas phases, the local mass average Hydrogen transport For anode catalyst layer: SH2 = − 2F
2
Ran
velocity v is defined as [33,34]:
MO
sl For cathode catalyst layer: SO2 = − Rcat
l vl+ sg g vg
2
Oxygen transport 4F
v= (6)
sl l + sg g
Water vapor transport For anodecatalyst layer:
The source term for the momentum equation used in the model nd M H O
SH2 O = − F
2
Ran − rw
describes the flow of the fluid through a porous media by using
For cathode catalystlayer:
Darcy’s drag force. The gravity and surface tension forces were also MH O n d MH O
considered in the momentum equation source term. The source SH2 O = 2
2F
Rcat + F
2
Rcat − rw
terms for different regions of the fuel cell are given in Table 2.
Conservation of Charge For anode catalyst layer:
Ses = −Ran ; Sem = Ran
2.2.3. Volume fraction equation
For cathode catalyst layer:
The tracking of the interface between the phases was accom-
Ses = Rcat ; Sem = −Rcat
plished by the solution of a continuity equation for the volume
fraction of one (or more) of the phases. The motion of the interface For other parts: Ses = 0; Sem = 0
∂
(εsl l )
l ) + ∇(sl l v = Ss (7) 2.2.5. Species transport equations
∂t
The model predicts the local mass fraction of each species, Yi ,
2.2.4. Energy conservation equation through the solution of a convection–diffusion equation for the
In this model, the energy balance in terms of temperature ith species. The species transport equations are generally in the
change was also considered. In the multi-phase model, the energy following form:
equation is also shared among the phases [33,35]:
∂
(ε Yi ) + ∇ · (ε v Yi ) = Di,m ∇ 2 ( Yi ) + Si (12)
∂T ∂t
( cp )eff + ( cp )eff (v ∇T ) = ∇(keff ∇T ) + ST (8)
∂t
where Di,m is the diffusion coefficient for species i in the mixture
where T is the temperature (K) that is defined as [34].
sl l Tl + (1 − sl ) g Tg P p T T
T= (9) Di,m = ε1.5 (1 − sl )rs Di,m
0 0
(13)
sl l + (1 − sl ) g P T0
To specify these parameters in porous media, the effective prop- 0
where Di,m is the diffusion coefficient for species i in the mix-
erties were determined:
ture at reference temperature and pressure. The source term in the
( cp )eff = ε f cp,f + (1 − ε) s cp,s (10) transport equation is shown in Table 2.
During the operation, the H+ move from the anode to the cath-
keff = εks + (1 − ε)kf (11)
ode and also pull water molecules with them, this is known as
The source term of energy equation for two-phase model may the electro-osmotic drag effect. Physically, the water transport rate
consist of heat from electrochemical reactions, heat of formation through the membrane from anode to cathode by electro-osmotic
A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222 201
drag is computed as 2.3. Water transport and its effect on the properties of porous
media
nd MH2 O
ṁH2 O = Rcat (14)
F
Water is formed in the cathode catalyst layer by electrochem-
where nd is the drag coefficient and is proposed by Springer et al. ical reactions—an amount of oxygen is consumed and an amount
[36] for Nafion membrane: of water is produced. Due to the proton movement from the anode
2.5 to the cathode through the membrane, water molecules are pulled
nd = (15) with the protons by a force called electro-osmotic drag [30]. Addi-
22
tionally, water may diffuse through the membrane due to the
where is the water content inside the polymer membrane [36]. concentration differences.
In the present study, the membrane water diffusivity and mem-
2.2.6. Conservation of charge brane ionic conductivity were calculated by the expressions given
The current transport of electrons through the solid phase and by Springer et al. [36]. It is almost a standard model that has been
ions through the membrane phase was represented by the follow- well explained in the recent textbook [37].
ing equations [33,35]:
∇ ·( m ∇m ) = Sem (17) The VOF technique was implemented in the channels and porous
media (including GDLs and catalyst layers) [33]. Interface between
The volumetric source terms Ses and Sem are defined as volu-
gas and liquid (two-phase flow) is tracked by the volume fraction of
metric transfer currents. The source term in the equation of charge
liquid water in the computational cell volume. In the VOF approach,
is shown in Table 2.
the source term of continuity and momentum equations used in
The source terms representing transfer current were calculated
porous media include the effects of surface tension and wall adhe-
by using Butler–Volmer equation [33]:
sion and capillary water transport phenomenon, therefore, was also
an
CH2
˛ F ˛ F considered.
ref an an cat cat
Ran = Ran exp − exp − (18)
ref
CH RT RT
2 2.4.1. Geometric reconstruction scheme
cat The geometric reconstruction PLIC scheme (piecewise linear
CO2
˛ F ˛ F
ref
Rcat = Rcat exp −
cat cat
− exp
an an
(19) interface construction) was employed because of its accuracy and
COref RT RT applicability for general unstructured meshes, compared to other
2
methods such as the donor–acceptor, Euler explicit, and implicit
The relation between species concentration and species mass schemes. A VOF geometric reconstruction scheme is divided into
fraction is two parts: a reconstruction step and a propagation step. Details
Yi can be found in Refs. [33,38].
Ci = (20)
Mi
2.4.2. Implementation of surface tension
The volumetric transfer current R is driven by the activation
The addition of surface tension to the VOF method is modeled
overpotential , which is the potential difference between solid and
by a source term in the momentum equation. The pressure drop
membrane phases [33,35]:
across the surface depends upon the surface tension coefficient
= s − m − Vref (21) [33,38]:
1 1
The reference potential of the electrode Vref is 0 on the anode p = + (26)
side and is equal to the open circuit voltage on the cathode side: R1 R2
Vref = 0, on anode side; Vref = Voc , on cathode side (22) where R1 and R2 are the two radii, in orthogonal directions, to
measure the surface curvature.
The cell potential is then the difference between the cathode The surface tension can be written in terms of the pressure jump
and anode solid phase at the terminal collectors (two ends of the across the interface, which is expressed as a volumetric force F
cell collectors which are connected to the external electric circuit added to the momentum equation:
from the both electrodes).
∇sl
F = (27)
Vcell = s,cat − s,an (23) ( l + g )/2
In order to satisfy the conservation of charge, the total current where denotes the surface tension coefficient and is the surface
of either electrons or protons coming out from the anode catalyst curvature. The source terms for different regions of the fuel cell are
layer must be equal to the total current coming into the cathode given in Table 2.
catalyst layer and must be equal to the total current caused by the The surface curvature in Eq. (27) can be defined in terms of the
proton movement through the membrane [35]: divergence of the normal unit vector of the interface n̂:
Iave Asurf
UO2 ,in = ωO2 MO2 (32)
4F O2 Ain
To close the equation systems including conservation equa- where h is the convection heat transfer coefficient (W m−2 K−1 ), As
tions of continuity, momentum, energy, species, and charge with is heat transfer surface area (m2 ), Ts and T∞ are the temperatures
unknowns: u, v, w, p, T, Yi and , the boundary conditions are of the surfaces and the free stream, respectively.
required. The summary of boundary conditions is listed in Table 3. Noted, heat is produced when the fuel cell operates. For the case
Below a brief description is provided. in which water finally ends in vapor form, this heat generation rate
is defined as [39]:
2.6.1. Inlet of flow channels Q̇generated = I(1.25 − Vcell ) (W) (34)
Inlet velocities, fuel and oxidant temperatures and mass concen-
tration of species were set as given parameters listed in Table 3. Inlet In order to ensure that the forced convection dissipates the heat
velocities can be preliminarily calculated by the inlet flow rates converted from electricity, the following condition must be satis-
based on the average current density (Iavg ) and excess coefficient ω fied:
that is defined as
mH2 ,supply mO2 ,supply Q̇convection = Q̇generated (35)
ωH2 = ; ωO2 = (30)
mH2 ,consumption mO2 ,consumption then the convective coefficient h was chosen as
Then, inlet velocity expression is given by the following equations: I(1.25 − Vcell )
h= (36)
Iave Asurf As (Ts − T∞ )
UH2 ,in = ωH2 MH2 (31)
2F H2 Ain
2.7. Solution procedure
Table 3
Boundary conditions
The above-coupled set of governing equations and relative equa-
tions were implemented into FLUENT® 6.2 by developing our own
Locations of application Boundary conditions user-defined-functions (UDFs) based on the general FLUENT® pack-
Inlet of the anode flow channel u = Uan,in ; YH2 = YH2 ,in ; YH2 O age that does not include the PEMFC module developed by Fluent
= YH2 O,an,in ; Tan,in = TH2 ,in Inc. The developed own UDFs are written in C language with
Inlet of the cathode flow channel u = Ucat,in ; YO2 = YO2 ,in ; YH2 O about 2000 statements. There were totally 279,500 grid cells in
= YH2 O,cat,in ; Tcat,in = Tair/O2 ,in the computation domain used to simulate the physical and elec-
∂uout,an ∂YH ∂YH O,an trochemical phenomena in the fuel cell. The solution procedure for
Outlet of the anode flow channel ∂x
= 0; ∂x2 = 0; 2
∂x
∂Tan,out
pressure–velocity coupling was based on PISO algorithm [40].
= 0; =0
∂x In order to save the calculation time, an unsteady single-phase
∂uout,cat ∂YO ∂YH O,cat
Outlet of the cathode flow channel = 0; ∂x2 = 0; 2 model was simulated from the time t = 0 s to 0.5 s. At the time
∂x ∂x
= 0;
∂Tcat,out
=0 t = 0.5 s, an initialization of a series of liquid water droplets was
∂x
initially suspended on the cathode channel at the time t = 0.50 s
∂m
The anode terminal s = 0; ∂y
=0 and a general model of PEMFC was called to further investigate
∂m two-phase flow behavior, especially liquid water behavior across
The cathode terminal s = Vcell ; ∂y
=0
∂s ∂s ∂m ∂m
porous media, together with electrochemical reaction, heat and
External boundaries = 0; = 0; = 0; =0
∂x ∂z ∂x ∂z mass transfer.
A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222 203
Fig. 6. The motion and deformation of droplets vs. time on the cathode channel: (a) t = 0.500 s; (b) t = 0.50025 s; (c) t = 0.5005 s; (d) t = 0.50075 s; (e) the velocity distribution
and the volume-of-fluid of liquid water in the X–Z plane at Y = 0.002 m at t = 0.50075 s; (f) the pressure distribution and the volume-of-fluid of liquid water in the X–Z plane at
Y = 0.002 m at t = 0.50075 s; (g) the water and velocity distribution in the different cross-sections along the main flow direction at t = 0.50075 s, corresponding to (a) X = 0.001 m,
(b) X = 0.006 m, (c) X = 0.014 m and (d) X = 0.019 m, respectively.
A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222 205
Fig. 6. (Continued ).
outlet was more concentrated in the cathode GDL, especially the droplets that enter the turning area from the straight channels hit
GDL above the land area. This is consistent with the generation of the turning-wall surface. As a result, the deformed droplets were
water at the cathode and its removal by the gas flow in the channels. further elongated and the strong “turning” gas flow breaks these
The water generated above the land area needs to diffuse out later- droplets into even smaller ones sticking on the turning-wall sur-
ally to the channel before it can be removed through the flow fields. face. Due to the effect of wall adhesion and surface tension, the
The numerical result in Fig. 4a qualitatively shows good agreement smaller water droplets sticking on the turning-wall would slowly
with the experimental data in Fig. 4b made by Davey et al. [46]. In move forward while the other water droplets from the channel is
addition, more liquid water concentrated in the turning area of ser- continuously coming to the turning, and further broken into the
pentine channel fuel cell is also presented by the computed model other smaller ones. Such these water droplets would coalesce, then
in this study shown in Fig. 5a and observed in Fig. 5b that was cap- expanding and deforming into “water bands” due to the shear stress
tured by using neutron tomography method in the experiments of from “turning” gas flow and pressure drop, as shown in Fig. 6c and
Manke et al. [45]. d. It is noted that since the coalescence of water droplets is formed
In summary, by comparing the numerical results with avail- in the turning area resulting in a high concentration of liquid water
able experimental results, we can conclude that the present general volume, a substantial blockage in the channel would take place. This
model is qualitatively in good agreement with the available experi- blockage induces a decrease of the gap between the water droplets
mental results. For this research, qualitative comparisons could not and the channel wall. In addition, the water droplets hitting the
be conducted at the present time due to the sizes of the fuel cell in turning wall would stick onto it due to the wall adhesion. Fig. 6e
the cited experimental research are not known. The present authors shows the velocity distribution and the volume fraction of liquid
are currently developing a bench experimental set-up to further water in the X–Z plane at Y = 0.002 m at t = 0.50075 s. The velocity
validate the general model presented here and detailed results will field of airflow is accelerated through the gaps that were formed by
be published in future. the blockage of liquid water in the turning area, and then is decel-
erated after leaving the turning area in which the blockage occurs.
3.2. The motion and deformation of liquid water inside the In the other views shown in Fig. 6g, one can see that the water
cathode channel droplets hit the turning wall and expanded to the surroundings of
the landing points. The droplets could induce high-pressure zones
Fig. 6 shows the motion and deformation of droplets versus time in the channel, as shown in Fig. 6f, especially in the turning area.
on the cathode channel of the PEMFC. At t = 0.5 s, the droplet was As the time progressed to t = 0.501 s, as shown in Fig. 7a, the
placed at the channel inlet and was in its initial spherical shape water bands in the turning walls keep moving forwards along
(Fig. 6a). As time progressed, droplet deformation (elongation) in the main flow direction. One can see in Fig. 7a and b that, the
the flow direction, which is due to a combination of droplet sur- water bands have a tendency of elongating without breaking away
face tension and shear stress from surrounding airflow, could be under the impact of shear force from the gas flow. However, the
observed. In other words, the enlargement of droplet surface area water bands were broken again into smaller droplets distributed
increased the surface tension such that the force balance on the over the straight channel wall since these bands left the turn-
droplet could be achieved. Fig. 6b indicates that, after hitting the ing area. Subsequently, the formed small water drops are easily
turning surface, these deformed droplets had the tendency of frag- removed and distributed everywhere on the straight channel walls,
menting and then entering the central airflow due to the dragging as it can be seen in Fig. 7c and d at t = 0.504 s and 0.506 s. Fig. 7e
effect from the turning flow. Obviously, in the turning area, the shows that the airflow is rapidly accelerated at the turning area
shear force from the airflow is much stronger than the droplet due to the water blockage as earlier mentioned. Hence, the shear
surface tension, thus together with the inertia force, resulting in stress from the airflow in the straight channel is significant to
significant droplet deformation and complex liquid water distri- break the water bands into small, discrete liquid droplets rather
bution. As the time progressed after milliseconds, the deformed than deforming or elongating such water bands. This also results
206 A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222
Fig. 7. The motion and deformation of droplets vs. time on the cathode channel: (a) t = 0.501 s; (b) t = 0.502 s; (c) t = 0.504 s; (d) t = 0.506 s; (e) the velocity distribution and
the volume-of-fluid of liquid water in the X–Z plane at Y = 0.002 m at t = 0.506 s; (f) the pressure distribution and the volume-of-fluid of liquid water in the X–Z plane at
Y = 0.002 m at t = 0.506 s.
A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222 207
Fig. 8. The motion and deformation of droplets vs. time on the cathode channel: (a) t = 0.520 s; (b) t = 0.522 s; (c) t = 0.524 s; (d) t = 0.526 s; (e) the velocity distribution and
the volume-of-fluid of liquid water in the X–Z plane at Y = 0.002 m at t = 0.526 s; (f) the pressure distribution and the volume-of-fluid of liquid water in the X–Z plane at
Y = 0.002 m at t = 0.526 s.
208 A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222
Fig. 9. The motion and deformation of droplets vs. time on the cathode channel: (a) t = 0.560 s; (b) t = 0.562 s; (c) t = 0.564 s; (d) t = 0.566 s; (e) the velocity distribution and
the volume-of-fluid of liquid water in the X–Z plane at Y = 0.002 m at t = 0.566 s; (f) the pressure distribution and the volume-of-fluid of liquid water in the X–Z plane at
Y = 0.002 m at t = 0.566 s; (g) the liquid water in the Y–Z plane at X = 0.001 m, 0.002 m, 0.018 m and 0.019 m.
A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222 209
Fig. 9. (Continued ).
in a decrease of pressure drop in comparison with the value results in an evident decrease in pressure drop distribution which
at t = 0.50075 s. is presented in Fig. 8f.
At t = 0.52 s as shown in Fig. 8a, a portion of water droplets was Most of the water droplets were moved out of the channel by the
removed from the cathode channel outlet by the gas flow. The small gas flow at t = 0.56 s. As the time proceeded to t = 0.566 s, it could
droplets tend to scatter everywhere in the channel rather than coa- be observed that there were only few small droplets remaining
lescing or forming water bands. Even in the turning area, it can in the exit channel. However, an amount of liquid water pene-
be seen that the water bands either are tiny or are not formed at trated through the gas diffusion layer seemed to be retained at
all. The wall adhesion of the small droplets previously coalesced the same position in the layer regardless the droplets in the chan-
at the turning areas is insufficient to resist the shear stress from nel continued to be removed outwards. It can be obviously seen
the gas flow. In addition, one can see that there are some droplets that the water amount kept in the diffusion layer is marked by cir-
penetrating to the gas diffusion layer. It can be realized that this cles as shown in Fig. 9a–d. Noticeably, the marked circles in Fig. 9
phenomenon takes place mostly in the turning areas where the evoke that the penetrated water usually occur in the turning area
water is not easy to be removed rather than in the straight chan- in which the liquid water is more difficult to be removed and have
nel as previously discussed. As shown in Fig. 8b–d that as the time a tendency of passing through the gas diffusion layer instead of
proceeds to t = 0.526 s, the water continues to be taken away from moving along the channel to the outlet. The liquid water in the
the channel outlet except the water amount remaining in the gas gas diffusion layer is also clearly illustrated in Fig. 9g since the
diffusion layer. At t = 0.526 s, as shown in Fig. 8e, the velocity field cross-sectional planes along X-direction at X = 0.001 m, 0.002 m,
is observed to be “more stable” since a large amount of liquid water 0.018 m and 0.019 m are extracted. As time progresses, the more
was taken away and then the blockage caused by the small droplets liquid water gradually move out of the channel, the less the liquid
less influences on the gas flow movement. Consequently, this also water blocks the flow field in the channel. Therefore, the block-
210 A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222
Fig. 10. (a) The distribution of volume fraction of liquid water in the cross-sectional planes (Y–Z plane) at X = 0.0015 m, 0.010 m and 0.0185 m (Case 2); (b) distribution of
volume fraction of liquid water in the middle-plane of the cathode flow channel (Y = 0.0015 m) and catalyst layer (Y = 0.002305 m) (Case 2).
A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222 211
Fig. 11. Velocity vectors in the channel: on the middle-plane of cathode channel (Y = 0.0015 m) and the middle-plane of anode channel (Y = 0.00317 m): (a) Case 1 and (b)
Case 2; on the cross-sectional planes (Y–Z plane) at X = 0.0015 m, 0.010 m and 0.0185 m: (c) Case 1 and (d) Case 2.
212 A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222
Fig. 12. Velocity vectors in the porous media on the cross-sectional planes (Y–Z plane) at X = 0.0015 m, 0.010 m and 0.0185 m: (a) Case 1 and (b) Case 2.
A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222 213
age insufficiently influences on the velocity and pressure fields in 3.3. Effects of liquid water on the fuel cell characteristics
the fuel cell channel. It can be seen in Fig. 9e and f, the velocity
and pressure fields at t = 0.566 s are more stable, have no sudden In order to investigate the influences of liquid water on the cell
changes and have a smooth distribution along the channel, similar characteristics and operating condition, two cases (Cases 1 and 2)
to those when compared with the case without liquid water in the are compared in this study. The first case would be considered
channel. at the time instant t = 0.53 s when the liquid droplets were not
Fig. 13. Velocity vectors and pressure distributions on the middle-plane of catalyst layers (0.002305 m): (a) Case 1 and (b) Case 2.
214 A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222
introduced in the cathode channel and therefore the fuel cell oper- effects. In Case 2, fuel cell characteristics such as velocity, pressure,
ates in the condition without liquid water. The second case would temperature, species concentrations and the fuel cell performance,
be presented at the time instant t = 0.53 s since a series of water etc. are considered as taking the appearance of water droplets into
droplets had been introduced at t = 0.5 s into the cathode channel account. Hence, the numerical model at the first case was consid-
and evolved for 0.03 s that can be used to show the liquid water ered as single-phase model and the second case was considered
Fig. 14. Velocity vectors and pressure distributions on the middle-plane of anode channel (Y = 0.00317 m) and catalyst layers (0.002365 m): (a) Case 1 and (b) Case 2.
A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222 215
as the general model that includes two-phase interface tracking X = 0.0015 m, 0.010 m and 0.0185 m and Fig. 10b depicts in the
together with all other detailed sub-models. middle-plane of cathode flow channel and catalyst layer, respec-
tively. Note that the liquid water was not taken into account in
3.3.1. Volume fraction of liquid water single-phase flow that was applied for the fuel cell prior to t = 0.50 s.
In Case 2, Fig. 10a shows the distribution of volume frac- Therefore, the presence of volume fraction of liquid (liquid water
tion of liquid water in the cross-sectional planes (Y–Z plane) at distribution) is not applicable to Case 1. It could be observed that
Fig. 15. Mass fraction distributions of oxygen on the middle-plane of cathode channel (Y = 0.0015 m) and catalyst layers (0.002305 m): (a) Case 1 and (b) Case 2.
216 A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222
the liquid droplets distributed in the fuel cell at a specified time gressed after 0.03 s at t = 0.53 s. The change of the water occupation
were presented via volume fraction of water. To investigate the fraction in the channel and liquid water transport from the cath-
motion and interaction of liquid droplets inside the fuel cell and ode channel to the porous media were studied. By tracking the
the influence of liquid phase on the fuel cell operating condition, presence of volume fraction of water, one can determine where
the results of this simulation would be analyzed as the time pro- the liquid water occurs and the effects of presence of liquid water
Fig. 16. Mass fraction distributions of hydrogen in the middle-planes of anode channel (Y = 0.00317 m) and catalyst layers (0.002365 m): (a) Case 1 and (b) Case 2.
A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222 217
on the physical parameters and characteristics of the fuel cell. 3.3.2. Liquid water effects on velocity field and pressure
Fig. 10a and b shows that most of the liquid water concentrated distribution
in the channel and moved through porous media is in the turning The velocities at the cathode and anode inlets are 7.0 m s−1 and
area of the serpentine-shaped flow channel. As a result, this would 4.5 m s−1 , respectively. Fig. 11a and b shows that the gases (air
especially influence the other flow parameters of the turning area and water vapor at cathode; and hydrogen and water vapor at
also. anode) are supplied from the inlet, moving along each branch of the
Fig. 17. Mass fraction distributions of water vapor on the middle-plane of cathode channel (Y = 0.0015 m) and catalyst layers (0.002305 m): (a) Case 1 and (b) Case 2.
218 A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222
serpentine-shaped channel for Case 1 and Case 2, respectively, dra- media had no change on the same cross-section for every channel
matically varied their direction when approaching the turn, then branch.
go to the next branches. In Case 1 (Fig. 11c), the velocity vectors on Fig. 13a and b shows velocity vectors and pressure distribu-
Y–Z planes on cross-sectional planes of channel branches shown in tions on the middle-plane of catalyst layers (0.002305 m) for Cases
Fig. 11c seem to have similar variation to each other. In other words, 1 and 2, respectively. Now, the flow field on X–Z planes could be
these velocity distributions on the same cross-section through all observed. In both cases, there remained alternate velocity distribu-
channel branches are the similar. This is because that liquid water tions having “triangle-shaped” on the porous media in which the
was not considered in Case 1 and therefore there are no liquid magnitude of velocity vectors increased or decreased along the X-
droplets effects on the flow field. In Case 2 (Fig. 11d), the velocity direction. This is the nature of fluid flow. It could be noticed that
and pressure distributions are quite different. Due to the presence the high velocity vectors were located in the region in the land
and blockage of water droplets over the channel as previously men- area between the two corners of two sequent branches, also having
tioned, the flow field was not similar to each other depending on the largest pressure difference. The flow in the corner of a branch,
where the droplets located. It is noticeable that the pressure drop therefore, tends to move towards the corner of the next branch at
in Case 2 was observed to be higher than that in Case 1, especially the same side through the porous media instead of moving along
around the location of the droplets, the pressure drop reaches high- the channel. This is the shortest way and easiest way as well. The
est values due to the momentum resistance caused by liquid water magnitude of such velocity vectors decreases along the channel due
in airflow. to the decrease of pressure difference and it is lowest at the corner
The velocity in the gas channels on both sides shown in Fig. 10a at opposite side where the flow freely moves in the channel turn.
and b were very high (the order of magnitude is 1) by comparing Moreover, the chaos of velocity distribution caused by the effect of
to the low velocity in the porous media shown in Fig. 12a and b liquid water corresponding to the location of droplets in the porous
(the order of magnitude is 10−2 ) due to the flow resistance in the media is more realized as shown in Fig. 13b.
porous media (including GDLs and catalyst layers). Although the Fig. 14a and b shows velocity vectors and pressure distributions
porous media is solid, there still exists the flow field due to pres- on the middle-plane of anode channel (Y = 0.00317 m) and catalyst
sure gradient distribution in the void of such regions where reactant layers (0.002365 m) for Cases 1 and 2. Note that the flow field in
gases and water can move through. It can be seen in Fig. 12a and b the fuel cell may vary at the different times. Because the flow con-
for both cases that the flow in Y–Z planes had a tendency of moving ditions at the anode side reached the steady state at Case 1 and the
through the porous media from one channel branch to the next, time period between Case 1 (at t = 0.50 s) and Case 2 (at t = 0.53 s)
especially in the land area (the porous zones underneath the col- is almost negligible, then the velocity and pressure distributions
lector ribs). The difference of pressure on the same cross-section were almost the same for the both cases due to the absence of liquid
in all branches is the driving force to such flow motion. Similarly, water at the anode side.
it can be clearly seen in Fig. 12b how the blockage of liquid water
in the porous media influenced the velocity fields. On the contrary, 3.3.3. Liquid water effects on concentrations
for the case where liquid water was absent in the fuel cell as shown In the both cases, it could be observed that the oxygen fraction
in Fig. 12a, it is most likely that the flow in Y–Z planes in the porous decreased along the cathode channel from the inlet to the outlet
Fig. 18. The distribution of local current density at the catalyst/GDL interfaces at the cathode and anode: (a) Case 1 and (b) Case 2.
A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222 219
and from the channel to the catalyst layer due to the oxygen con- oxygen mass fraction distribution in the porous media underneath
sumption on the catalyst layer. Mass transport of oxygen from the the turning area could be seen to have a uniform, smooth U-shaped
channel to the GDL was mainly driven by its concentration gradi- as shown in Fig. 15a. Differently in Case 2, the shape was various,
ent (diffusion) and this effect was also partially influenced by the unsmoothed and arbitrary. This is because the liquid water occu-
velocity component (convection) inside the porous media. The flow pied the void in which oxygen was supposed to occupy. The higher
in cathode channel enhanced the gas transport in the region under volume fraction of liquid water occupies, the lower the oxygen mass
the channel, rather than in the region under the rib (the land area). fraction could be observed.
As shown in Fig. 15a and b, high oxygen region under the cath- Similar to the oxygen mass fraction distribution in the cathode,
ode channel was observed. On the other hand, the land area had a the hydrogen gradually decreased from the inlet to the outlet in
low amount of oxygen. There are two reasons accounting for this; both the anode flow channel and porous media as shown in Fig. 16a
the first is that mass transport of oxygen in such area was low due and b. However, the decrease of hydrogen mass fraction along the
to weak influence of convection and diffusion. The second is that anode channel was insignificant due to absence of liquid water.
a large amount of oxygen was consumed due to a high reaction For the both cases, the water vapor mass fraction increased along
rate—the chemical reaction had strongly taken place in the land the cathode channel by two sources: by-product water was gen-
area which was directly connected to the collector ribs that the erated from the electrochemical reaction at the cathode and the
current density vectors mainly went through. This phenomenon migration of water from the anode to the cathode due to the electro-
will be discussed in later section. The presence of liquid water sig- osmotic drag. As shown in Fig. 17 that there was a high water vapor
nificantly influences the oxygen concentration in the channel and content remaining in the land area at the cathode as a result of
porous media as well. This is very important in water management water build-up due to the higher reaction rate. Furthermore, due to
of a fuel cell. As mentioned in many studies, liquid water occurring the small velocity field in the land area nearby the edge (shown in
in the fuel cell causes the flooding phenomenon and further blocks Figs. 11 and 12), the accumulated water in such area was not easy to
the oxygen transport through the GDL and catalyst layer, then sig- be removed. Obviously, water concentration is highest in the edge
nificantly influences the fuel cell performance. In Case 2, there was adjacent to the outlet channel branch due to water build-up and
a noticeable blockage of liquid water in oxygen concentration, espe- low convection effect. In Case 2, the presence of liquid droplets also
cially in the land region underneath the turning area. In Case 1, the influenced the distribution of water vapor mass fraction, similar to
Fig. 19. (a) Temperature distribution in the cross-sectional planes (Y–Z plane) at X = 0.0015 m, 0.010 m and 0.0185 m (Case 2); temperature distributions on the middle-plane
of cathode channel (Y = 0.0015 m) and catalyst layers (0.002305 m): (b) Case 1 and (c) Case 2.
220 A.D. Le, B. Zhou / Journal of Power Sources 182 (2008) 197–222
that on oxygen transport. As can be seen in Fig. 17b, the higher the
volume fraction of liquid water occupied, the larger the water vapor
mass fraction observed.
PEMFC and coupled the fluid flow (momentum transport), species perature of the channel and porous media with the presence of
transport, energy transport, electron and proton transport, with liquid water was lower than that without liquid water.
volumetric electrochemical reactions. The membrane, gas diffusion
layer and catalyst layers were treated as volumetric domain with Acknowledgments
grid points, and the electrochemical reactions in the catalyst layers
as volumetric reactions. Especially, by using VOF interface tracking The authors are grateful for the support of this work by the
algorithm, the flow behaviors of liquid water and temperature field Auto21TM Networks of Centers of Excellence (Grant D07-DFC),
were also investigated. The main interpretations and conclusion of the Natural Sciences and Engineering Research Council of Canada
this study are summarized as follows: (NSERC), the Canada Foundation for Innovation (CFI), the Ontario
Innovation Trust (OIT), and the University of Windsor.
1. The motion and behaviors of liquid water in a full 3D, unsteady
PEMFC was completely described by using VOF interface tracking
References
algorithm combined with solving the fluid flow, species trans-
port, energy transport and electrochemical governing equations. [1] S. Um, C.Y. Wang, K.S. Chen, J. Electrochem. Soc. 147 (2000) 4485.
2. The motion of droplets in the PEMFC model was illustrated at [2] S. Dutta, S. Shimpalee, J.W. Van Zee, J. Appl. Electrochem. 30 (2000) 135.
different time steps in numerical simulation. Firstly, the droplets [3] E. Hontanon, M.J. Escudero, C. Bautista, P.L. Garcia-Ybarra, L. Daza, J. Power
Sources 86 (2000) 363.
freely moved in the channel then deformed (elongation) in the [4] S.W. Cha, R. O’Hayre, Y. Saito, F.B. Prinz, J. Power Sources 134 (2004) 57.
flow direction, which was due to a combination of droplet surface [5] A.A. Kulikovsky, Electrochem. Commun. 3 (2001) 460.
tension and shear stress from surrounding airflow. After hitting [6] J.S. Yi, J.D. Yang, C. King, AIChE J. 50 (2004) 2594.
[7] H. Ju, G. Luo, C.Y. Wang, J. Electrochem. Soc. 154 (2007) B218.
the turning surface, these deformed droplets had the tendency [8] L. You, H. Liu, Int. J. Heat Mass Transfer 45 (2002) 2277.
of fragmenting and then entering the central airflow due to the [9] D. Natarajan, T.V. Nguyen, J. Power Sources 115 (2003) 66.
dragging effect from the turning flow. Since the coalescence of [10] T. Berning, N. Djilali, J. Electrochem. Soc. 150 (2003) A1589.
[11] S. Mazumder, J.V. Cole, J. Electrochem. Soc. 150 (2003) A1510.
water droplets was formed in the turning area resulting in a high [12] M. Vynnycky, M. Vynnycky, Appl. Math. Comput. 189 (2007) 1560.
concentration of liquid water volume, a substantial blockage in [13] U. Pasaogullari, C.Y. Wang, J. Electrochem. Soc. 152 (2005) A380.
the channel occurred. After leaving the turning area, a significant [14] H. Meng, C.Y. Wang, J. Electrochem. Soc. 152 (2005) A1733.
[15] Y. Wang, C.Y. Wang, J. Electrochem. Soc. 153 (2006) A1193.
shear stress from the airflow in the straight channel broke the
[16] H. Meng, C.Y. Wang, Chem. Eng. Sci. 59 (2004) 3331.
coalescent droplets into smaller, discrete liquid droplets rather [17] H. Ju, C.Y. Wang, J. Electrochem. Soc. 151 (2004) A1954.
than deforming or elongating such water bands. Finally, liquid [18] H. Ju, C.Y. Wang, S. Cleghorn, U. Beuscher, J. Electrochem. Soc. 153 (2006) A249.
droplets were gradually removed out of the channel under a [19] Y. Wang, C.Y. Wang, J. Power Sources 153 (2006) 130.
[20] P. Quan, B. Zhou, A. Sobiesiak, Z. Liu, J. Power Sources 152 (2006) 131.
strong flow field. [21] K. Jiao, B. Zhou, P. Quan, J. Power Sources 154 (2006) 124.
3. The presence of liquid water in the channel significantly influ- [22] K. Jiao, B. Zhou, P. Quan, J. Power Sources 157 (2006) 226.
ences the flow field parameter. Namely, due to the blockage of [23] X. Zhu, P.C. Sui, N. Djilali, J. Power Sources 172 (2007) 287.
[24] J.H. Nam, M. Kaviany, Int. J. Heat Mass Transfer 46 (2003) 4595.
water droplets, the gas flow would become unevenly distributed, [25] B. Markicevic, A. Bazylak, N. Djilali, J. Power Sources 171 (2) (2007) 706.
the high pressure-drop regions would occur around the location [26] J.I. Gostick, M.A. Ioannidis, M.W. Fowler, M.D. Pritzker, J. Power Sources 173 (1)
of liquid droplets. (2007) 277.
[27] P.K. Sinha, C.Y. Wang, J. Electrochem. Acta 52 (28) (2007) 7936.
4. The structure of the channel will significantly influence the [28] H. Dohle, R. Jung, N. Kimiaie, J. Mergel, M. Muller, J. Power Sources 124 (2003)
distribution of liquid water inside the cell. As simulated in a 371.
serpentine channel, the liquid water is difficult to be removed [29] U. Pasaogullari, C. Wang, Electrochem. Acta 49 (2004) 4359.
[30] S. Freni, G. Maggio, E. Passalacqua, Mater. Chem. Phys. 48 (1997) 199.
in the turning area. Therefore, a portion of liquid water tends [31] K. Jiao, B. Zhou, J. Power Sources 169 (2007) 296.
to move through the gas diffusion layer in the porous media. [32] K. Jiao, B. Zhou, J. Power Sources 175 (2008) 106.
By simulating different types of the channels, the character- [33] Fluent® 6.2 Documentation, Fluent Inc., 2005.
[34] R.B. Bird, W.E. Steward, E.N. Lightfoot, Transport Phenomena, John Wiley & Sons,
istics of liquid water removal for different designs could be
New York, 1960.
investigated. [35] F. Barbir, PEM Fuel Cells, Elsevier Academic Press, 2005.
5. By quantitatively and qualitatively analyzing the behaviors of liq- [36] T. Springer, T. Zawodzinski, S. Gosttesfeld, J. Electrochem. Soc. 138 (1991) 2334.
uid water and the average current densities at different periods [37] R. O’Hayre, S.-W. Cha, W. Colella, F.B. Prinz, Fuel Cell Fundamentals, John Wiley
& Sons, New York, 2006.
of time, it could be observed that the liquid water would hinder [38] C.S. Peskin, J. Comput. Phys. 25 (1997) 220.
the gas transport in the fuel cell, resulting in a high concentration [39] J. Larminie, A. Dicks, Fuel Cell Systems Explained, 2nd ed., John Wiley & Sons,
loss and directly decreasing the current density—in other words, England, 2003.
[40] J.H. Ferziger, M. Peric, Computational Methods for Fluid Dynamics, 3rd ed.,
severely affect the fuel cell performance. Springer, 2002.
6. The local current density contours are also presented, show- [41] X.G. Yang, F.Y. Zhang, A. Lubawy, C.Y. Wang, Electrochem. Solid-State Lett. 7
ing that the high magnitude of the vectors mostly distribute in (2004) A408.
[42] F.Y. Zhang, X.G. Yang, C.Y. Wang, J. Electrochem. Soc. 153 (2006) A225.
the region underneath the channel where high reaction rates [43] X. Liu, H. Guo, C. Ma, J. Power Sources 156 (2006) 267.
exist. [44] K.W. Feindel, L.P.-A. LaRocque, D. Starke, S.H. Bergens, R.E. Wasylishen, J. Am.
7. By solving the energy equation, the temperature distribution Chem. Soc., Commun. 126 (37) (2004) 11436.
[45] I. Manke, Ch. Hartnig, M. Grünerbel, J. Kaczerowski, W. Lehnert, N. Kardjilov, A.
over the fuel cell was described. The temperature was controlled Hilger, W. Treimer, M. Strobl, J. Banhart, Appl. Phys. Lett. 90 (2007) 184101.
at around 333 K by using a suitable forced convective coefficient [46] J.R. Davey, R. Mukundan, T. Rockward, J.S. Spendelow, D.S. Hussey, D.L. Jacobson,
applied on the fuel cell. The model also showed that the tem- M. Arif, R.L. Borup, 212th ECS Meeting, Washington, October 2007.