You are on page 1of 11

Journal of Power Sources 326 (2016) 417e427

Contents lists available at ScienceDirect

Journal of Power Sources


journal homepage: www.elsevier.com/locate/jpowsour

Proton exchange membrane fuel cell model for aging predictions:


Simulated equivalent active surface area loss and comparisons with
durability tests
rard a, *, M. Quinaud a, J. d’Arbigny d, Y. Bultel b, c
C. Robin a, b, c, d, M. Ge
a
Univ. Grenoble Alpes, CEA, LITEN, F-38054, Grenoble, France
b
Univ. Grenoble Alpes, LEPMI, F-38000, Grenoble, France
c
CNRS, LEPMI, F-38000, Grenoble, France
d
Zodiac Aerospace, F-78370, Plaisir, France

h i g h l i g h t s

 Prediction by simulation of PEM fuel cell lifetime.


 Multiscale modelling of catalyst dissolution and global loss of performance.
 Reconstruction of active surface area loss by model inversion.
 Prediction and validation of current density evolution during aging.
 Validation by two 2000-h experimental stack aging tests.

a r t i c l e i n f o a b s t r a c t

Article history: The prediction of Proton Exchange Membrane Fuel Cell (PEMFC) lifetime is one of the major challenges to
Received 17 February 2016 optimize both material properties and dynamic control of the fuel cell system. In this study, by a mul-
Received in revised form tiscale modeling approach, a mechanistic catalyst dissolution model is coupled to a dynamic PEMFC cell
3 July 2016
model to predict the performance loss of the PEMFC. Results are compared to two 2000-h experimental
Accepted 5 July 2016
aging tests. More precisely, an original approach is introduced to estimate the loss of an equivalent active
Available online 15 July 2016
surface area during an aging test. Indeed, when the computed Electrochemical Catalyst Surface Area
profile is fitted on the experimental measures from Cyclic Voltammetry, the computed performance loss
Keywords:
PEM fuel cell
of the PEMFC is underestimated. To be able to predict the performance loss measured by polarization
Durability test curves during the aging test, an equivalent active surface area is obtained by a model inversion. This
Modeling methodology enables to successfully find back the experimental cell voltage decay during time. The
ECSA loss model parameters are fitted from the polarization curves so that they include the global degradation.
Lifetime prediction Moreover, the model captures the aging heterogeneities along the surface of the cell observed experi-
Degradation mentally. Finally, a second 2000-h durability test in dynamic operating conditions validates the
approach.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction cleaner and less noisy energy conversion [1]. However, two main
barriers are still remaining for a widespread commercialization of
Fuel cells are a promising technology to address energy issues in PEMFC: high components’ costs and moderate durability [2]. Fuel
various fields. For transportation, Proton Exchange Membrane Fuel cells aging concerns for transportation may be one of the most
Cells (PEMFC) offer many advantages compared to classic thermal challenging issue to be dealt with. Indeed, the strong dynamics of
engines, such as a better efficiency, a good dynamic response, a the power profiles together with constraining operating conditions
are most of the time incompatible with a high lifetime of PEMFC.
Due to the variety of materials used inside a fuel cell, various
* Corresponding author. degradation mechanisms can occur [3e5]. These mechanisms can
rard).
E-mail address: mathias.gerard@cea.fr (M. Ge be reversible (temporary performance loss) or irreversible. Among

http://dx.doi.org/10.1016/j.jpowsour.2016.07.018
0378-7753/© 2016 Elsevier B.V. All rights reserved.
418 C. Robin et al. / Journal of Power Sources 326 (2016) 417e427

Nomenclature t time (s)


Ucell cell voltage (V)
Urev reversible cell voltage (V)
Acronyms v surface chemical velocity (mol m2 s1)
BoL Beginning of Life vdiss dissolution rate of the platinum particle (mol m2 s1)
ECSA Electrochemical Catalyst Surface Area
EoT End of Test Greek letters
GDL Gas Diffusion Layer a reaction transfer coefficient ()
MEA Membrane Electrode Assembly bk semi-empirical electrochemical reaction coefficients
PEMFC Proton Exchange Membrane Fuel Cell (V or V K1 or U.m2 or m)
STEM Scanning Transmission Electron Microscopy DG variation of free Gibbs energy (J)
D Gs variation of free enthalpy required for a platinum atom
Symbols extraction (J)
Cj concentration of the species j (mol m3) DGelec variation of free enthalpy required for a platinum atom
dnk variation of the platinum particles number of the oxidation (J)
population k () DGdes variation of free enthalpy required for leaching of Pt2þ
EGT Gibbs-Thomson energy (J mol1) (J)
F Faraday constant (C mol1) DNk ðtÞ cumulated variation of the particles number in the
i local current density (A m2) population k ()
Icell fuel cell current (A) Dc local potential at the inner layer (V)
k direct reaction constant (mol m2 s1) g surface tension (J m2)
Mj molar mass for the species j (kg mol1) h over potential (V)
mj mass for the species j (kg) rj density for the species j (kg m3)
n number of exchanged electrons () s protonic conductivity (S m1)
P0 standard pressure (Pa) tdegr degradation rate of the Electrochemical Surface Area
Pj partial pressure for the species j (Pa) ()
r radius (m)
R gas constant () Subscripts
Rohm specific ohmic resistance (ohm.m2) a anode
S electrode surface (m2) c cathode
T temperature (K) m membrane

the latter, the polymer membrane degradation is likely to be the information neither on the degradations mechanisms that occur
one mostly responsible for the failure of PEMFC [6]. Mechanical and nor on their locations at the cell surface. As an original modeling
thermal membrane degradations could be modeled [7], however approach, Jouin et al. [11e12], have used neural networks to predict
they have little effects on the fuel cell performances evolution the fuel cell lifetime after a learning phase during the first hundred
during operation, except at the very end of the cell lifetime when its hours of a test. Conversely to these approaches, Kulikovsky [13] has
permeation current increases dramatically. It is a very sudden built a purely theoretical model to study catalyst and GDL degra-
phenomenon which could be correlated also with the Membrane- dation effects on the fuel cell performance and especially on the
Electrode Assembly (MEA) initial fabrication defects and the bipo- current density distribution. Pohl et al. [14] have used differential
lar plate mechanical constraints. Therefore, the mechanical and equations to take into account the membrane degradation and the
chemical membrane degradations with the estimation of a sudden evolutions of the current density distribution with time. The au-
failure are beyond the scope of this paper. Another critical degra- thors couple this approach with the dual time theory in order to
dation mechanism in PEMFC concerns the catalyst layers and dissociate the time scales proper to the different degradation
especially the active surface area loss during fuel cell operation. For mechanisms. Finally, Dhanushkodi et al. [15] have modeled a
instance, the irreversible losses can be due to carbon support simplified carbon corrosion mechanism which have been success-
corrosion, catalyst dissolution/redeposition and catalyst layer fully validated on several Accelerated Stress Tests. In all these
micro-porous structure modification [8]. Conversely to the mem- studies however, the simulated data concerning aging phenomena
brane degradation, these mechanisms affect directly the fuel cell are scarce. No results are given on the location of the various
performances with time but not directly a fuel cell failure. Degra- degradations on the cell surface.
dation of Gas Diffusion Layers (GDL) and bipolar plates can also be The objectives of this work are to build a PEM fuel cell predictive
observed but they seem to be of less value regarding fuel cells model able to simulate the fuel cell aging with time and to explain
durability [7]. the behavioral evolutions with local degradation mechanisms. A
The understanding of degradation mechanisms by modeling has 2000-h durability test is used in this study to help building the
been conducted intensively with very different approaches [7]. degradation module and to validate it.
However, the prediction of the fuel cell lifetime under dynamic
cycles is less described even if some studies focus on it. In partic- 2. Experimental part
ular, Fowler et al. [9] have introduced a parameter in the semi-
empirical electrochemical response of their model to take into ac- 2.1. Description of the 2000-h durability test with steady operating
count an Electrochemical Surface Area (ECSA) loss. Soltani et al. [10] conditions
have used some durability tests to build empirical laws giving the
evolution of global parameters (potential, temperature, etc.) with A stack equipped with 30 large-area cells has been operated
time. However, this approach is not able to give any physical during 2000 h with steady operating conditions. The 220 cm2 MEAs
C. Robin et al. / Journal of Power Sources 326 (2016) 417e427 419

are in-house from the CEA-LITEN and their properties are listed in the CV and shown in Fig. 2b, which follows the same trend. The
Table 1. The bipolar plates are in steel SS316L with a serpentine flow surface loss intervenes mainly during the first 600 h of the test,
field. An in-situ device provided by the Sþþ company [16] [17], is then stabilizes with a much smaller degradation rate.
inserted in the middle of the stack to measure current density and This first durability test is used to fit the aging module of the
temperature distributions at the cell surface. model developed in this work.
Humidified pure hydrogen and air are fed to the anode and
cathode respectively. Before the test, the stack is conditioned dur- 2.2. Description of the 2000-h durability test with dynamic
ing 24 h on bench with highly humidified gases (80% Relative operating conditions
Humidity). The 30-cell stack has then been operated at a steady
current density of 0.4 A cm2 while keeping constant the operating A second 30-cell stack with the same components’ properties as
temperature (80  C) and the Relative Humidity (Air/H2 ¼ 30%/50%), the first one has been tested, using dynamic operating conditions.
as well as the gases stoichiometry (Air/H2 ¼ 2/1.5) and the gases More precisely, all these conditions remain identical to the first test
inlet pressure (1.5 bar). (cf. Table 1) except the fuel cell current profile and the cathode
Voltage-current polarization curves (VeI curves) are recorded Relative Humidity. The current profile is adapted from the New
every 200 h during the durability test with the same operating European Driving Cycle (NEDC), which is a simplified cycle adapted
conditions. They are obtained by measuring the voltage with a by the Motor Vehicle Emissions Group. NEDC considers the driving
current stepwise increment of 0.0045 A cm2 (between 0 and cycle as a combination between an urban cycle (800 s) and an
0.05 A cm2) and 0.05 A cm2 (between 0.05 and 2 A cm2). Each extra-urban cycle (400 s) with constant speeds, accelerations and
current plateau is held during 2 min to get a more stable measure. decelerations. However, it suffers some criticism for not consid-
Cyclic Voltammetries (CV) are recorded as well after each VeI ering the real accelerations of a light-duty vehicle. The minimum
curve in order to assess the remaining ECSA at the cathode. The and maximum currents are set respectively to 0.1 A cm2 and
measurement is conducted for cells n 1, 7, 15, 20, 25 and 30 with a 0.8 A cm2 in this work. The cathode RH is set to 50% during the
potentiostat using a 200 mV s1 voltage sweep rate. During the major part of the cycle but is set to decrease to 30% during the high
measurement, nitrogen is fed to the cathode and the operating power phase, in order to reproduce the probable behavior of an on-
conditions are changed to 60  C and 50%/50% RH to prevent the board fuel cell system. After approximately 1400 h, this second
membranes from drying since no water is produced in the fuel cell durability test had to be stopped to remove the cell n 15, whose
in the meantime. voltage collapsed due to a severe pinhole. The stack was then
Linear Sweep Voltammetries (LSV) and Electrochemical reassembled with the original 29 cells and operated again to reach
Impedance Spectroscopies (EIS) have also been conducted every the 2000 h of test (more than 6000 cycles). The performances loss
200 h during the durability test for several cells of the stack. between 200 h and the EoT gave an average degradation rate of
Resulting measures are not shown in this paper for sake of clarity 25 mV h1 at 0.5 A cm2, slightly higher than the first test’s one.
since no clear evolutions of the permeation current nor of the This second durability test is used to confirm that the model’s
impedance have been reported. ECSA aging module with parameters fitted on the first test is able to
The evolution of the 30 cells’ voltages during the durability test predict the experimental performances loss, even with a different
is shown in Fig. 1. The presence of both reversible and irreversible operating profile.
degradations can be observed on the graph. Indeed, a voltage re-
covery occurs after each characterization phase every 200 h, which 2.3. Description of the post-mortem analyses
highlights the reversible phenomena. This recovery is mainly due
to the cathode wash-out with N2 and OCV voltage. Meanwhile, an Two post-mortem devices are exploited in this study to acquire
overall voltage decrease can be noted between the Beginning of Life more data on the ECSA degradation locations, a mini-cell and a
(BoL) and the End of Test (EoT), sign of irreversible degradations. microscope.
The behavior of the cell n 1 and especially the voltage drop For the mini-cell tests, 2 cm2 circular samples are cut in several
occurring from approximately 1400 h of test seems isolated aged MEA in 3 locations: the air inlet, the cell center and the air
compared to the other cells. It could be attributed to the formation outlet. The samples are inserted in the mini-cell device for
of a pinhole on the cell n 1 membrane, which could come from an measuring their ECSA using the Cyclic Voltammetry. To ensure that
initial defect of the MEA. all samples are in the same hydration state before the measure-
Fig. 2 presents both the polarization curves and the percentage ment, a 1-h conditioning phase with a steady current density of
of ECSA variation during this test. The performances evolution 0.2 A cm2 at ambient operating conditions is applied. Then the CVs
shown in Fig. 2a enables to assess the total degradation occurring are measured using the same protocol as described for the mea-
during the 2000-h durability test. The non-linear evolution of the surements in stack.
performances can be pointed out. An average degradation rate of The cathode catalyst layer’s particles of several aged MEA have
14 mV h1 from 200 h to the end of the test can be obtained at been observed using a commercial field emission gun scanning
0.5 A cm2 with the test operating conditions, while an average electron microscope (LEO 1530 Gemini) for the same samples used
degradation rate of 250 mV h1 from the BoL to 200 h is observed. for the mini-cell tests.
Therefore, a higher loss occurs at the BoL (0e200hr period)
whereas the loss is much smaller for the rest of the test. This 3. Modeling
behavior could be related to the evolution of the ECSA obtained by
3.1. 2Dþ1D fuel cell model
Table 1
MEAs’ properties for the 2000-h durability test with steady operating conditions.
The fuel cell model used in this study is fully described in a
previous work [18]. It will be referred to as “PSþþ model” in this
Anode Cathode Membrane
paper. It is based on the resolution of Ordinary Differential Equa-
Catalyst nature Pt Pt3Co e tions (ODE) through a Bond-Graph representation. The complex in-
Catalyst loading (mgPt/cm2) 0.1 0.4 e plane serpentine flow field of the bipolar plates is modeled in 2D.
Thickness (mm) 5 10 25
The through-plane species transports are modeled in 1D (no in-
420 C. Robin et al. / Journal of Power Sources 326 (2016) 417e427

Fig. 1. Evolution of the 30 cells’ voltages during the 2000-h durability test with steady operating conditions (cf. Table 1).

to b9 coefficients are added to compensate some assumptions of the


model, such as the absence of thickness for the active layer
(interface).
8
>
>
>
> Ucell ¼ Urev þ h  Rohm i
>
>
>
> !1 !1
>
> 0
>
> PO*
2
PH*
>
>
>
2 2
>
> B P0 P0 C
>
> RT B ! C
>
> with Urev ¼ U 0 þ logB C
>
> nF @ PH* A
>
< 2O
P0
i¼0
>
> ! !
>
>
>
> PO* 2 *
PH
>
> and h ¼ b1 þ b2 T þ b3 TlnðiÞ þ b4 Tln þ b5 Tln 2
>
>
>
> P0 P0
>
> " ! #
>
>
>
> *
PH i i
>
> þ b6 Tln 2O
þ b7 þ b8 þ b9 i
>
>
>
> P 0 s c s a
:

(1)

where Ucell is the cell voltage (V), Urev is the reversible cell voltage
from the thermodynamic equilibrium (V), h is the overpotential (V)
with h  0, Rohm is the specific ohmic resistance of the cell
(ohm.m2), U0 is the standard voltage (V), R is the gas constant
(J mol1 K1), T is the temperature (K), n is the number of
Fig. 2. a) Performances evolution through VeI curves and b) average remaining ECSA
exchanged electrons (), F is the Faraday constant (C mol1), Pk* is
obtained from the Cyclic Voltammetries for the 2000-h durability test. the partial pressure of the species k at the active layer interface (Pa),
P 0 is the standard pressure (Pa), i is the local current density
(A m2), sk is the protonic conductivity of the active layer at the
plane transport in the GDL and catalyst layer) using Maxwell-Stefan
electrode k (S m1) and bk are semi-empirical coefficients (V or
equations for the gases and adding the formalism from Young et al.
V K1 or U m2 or m).
[19]. The through-plane liquid water transport is modeled using
A degradation module is then added to the PSþþ model by a
Darcy equations in the GDL. Mass balance and enthalpy balance are
coupling with a nano/micro scale model of the active layer,
solved in each discretized layer. The water transport in the mem-
following the approach described in the next sub-section.
brane is modeled by solving the electro-osmosis and diffusion
mechanisms equations. Thermal aspects are considered in the
PSþþ model by solving enthalpy balances in every physical layer. In 3.2. Nano/micro scale model description and coupling method
the active layer, which is considered only as an interface in the
model, the electrochemical response is calculated from a semi- A nano/micro scale model of PEM fuel cells has been built at CEA
empirical relation derived from the Butler-Volmer expression of Grenoble for a decade. This model, previously described in
the exchanged current density. The cell voltage is then given by the Refs. [20], will be referred to as “EDMOND model” in this paper.
system of equation (1), as a function of the local conditions of the Detailed descriptions of this model can be found in the literature
fuel cell which are dynamically solved by the model and of bk co- [20e23]. Its main features are the accurate modeling of the elec-
efficients which need to be adjusted on experimental data. The b6 trochemical double layer using non-equilibrium thermodynamics,
C. Robin et al. / Journal of Power Sources 326 (2016) 417e427 421

along with the inclusion of several degradation mechanisms Table 2


(platinum dissolution, carbon corrosion). Reactions constants values used in the Edmond model for the Pt dissolution
mechanism.
In particular, the platinum dissolution through equation (2) is
taken into account in the EDMOND model based on [24]: Parameter Value Unit

þ
k 2:2*106 mol:m2 :s1
Pt/Pt 2 þ 2e (2) DGs 75000 J
g 2:4 J:m2
The dissolution rate is known to depend on the radius of the a 0:5 ðÞ
particle. Indeed, small particles are less stable than big ones. In this MPt 0:1951 kg:mol1
paper, the Pt dissolution is assumed to be a 3-step process at the rPt 21470 kg:m3
atomistic scale: i/Pt sintering (removal of a Pt atom from the crystal
lattice and positioning of this adatom on a reactive site). An adatom
can be considered as a Pt atom adsorbed at the surface of the Pt initial number of particles. Therefore, the degradation rates of the
crystal: it shares one single bond with the Pt crystal (DGs is asso- two populations exposed to the same local conditions are different.
ciated to this step); ii/electrochemical reaction: Pt oxidation into Also, the local electrode potential at the nanoscale, which is one of
Pt2þ (DGelec is associated to this step); iii/desorption of Pt2þ (DGdes the inputs of each 4-D table, is approximated by (Erev þ h) at the
is associated to this step). Thus, the total variation of free Gibbs microscale, Erev and h being respectively the reversible potential
energy is given by (3): and the overpotential calculated in the PSþþ model according to
equation (1). This means that the potential loss by ionic transport
DG ¼ DGs þ DGelec þ DGdes (3) through the electrochemical diffusion layer is neglected. A
dimensionless effective surface area degradation rate, tdegr ðtÞ, is
with DGs the free enthalpy required for a platinum atom extraction then deduced from the catalyst degradation model.
calculated by DFT; DGelec ¼ 2aF Dc is the free enthalpy corre- The loss of Pt mass due to dissolution should be represented as
sponding to the oxidation of the Pt atom (with a the transfer co- the decrease of the radius of the particles (the number of the par-
efficient and Dc the local potential in the inner layer calculated by ticles in each population being kept constant). For sake of
the EDMOND model) and DGdes ¼ b EGT ðrPt Þ with 0 < b < 1 the simplicity, the loss of Pt mass will be described in this paper as the
free enthalpy required for leaching of the Pt 2þ (where EGT ¼ gM2r
Pt =rPt
decrease of the number of particles in each population (their radius
is the Gibbs Thomson energy that depends on the radius of the being kept constant).
platinum particle). This term accounts for the nanoparticle radius Even if the first representation is closer to reality, the second one
dependency on the dissolution rate. This means that the smallest can be seen as a translation of the variation of the radius into a
particles will dissolve the fastest. variation of the number of particles. However, this second repre-
The dissolution rate of the platinum particles (vdiss in sentation is interesting since the simulated results correspond
mol:m2 :s1 ) is given by equation (4). In this work, only Pt disso- directly to the post-mortem characterization by Transmission
lution is considered. Therefore, the contribution of the reverse re- Electron Microscopy.
action, which corresponds to Pt precipitation, is not taken into To translate the radius decrease into a decrease of the particles
account in this equation: number along with the definition of the catalyst degradation rate,
the variation of particles number dnk of the population k at the
vdiss ¼ keRT ð2aF DcþbEGT DGs Þ
1
(4) time t is given by the equation (6):

with k the rate constant (mol:m2 :s1 ).


A catalyst degradation rate dr (m s1) is then dynamically
dnk 4prk2 rPt drk
dt ¼ (6)
calculated, following the equation (5): dt mk dt

dr M where drk is the variation of the average particles’ radius for the
¼ vdiss Pt (5) population k given by the lookup table, rk is the initial average
dt rPt
particles’ radius for the population k, rPt the platinum density and
where r is the average particles’ radius (m), MPt the platinum molar mk the mass of a particle of the population k. The cumulated vari-
mass (kg mol1), rPt the platinum density (kg m3). The values used ation of particles number in the population k at t, DNk ðtÞ, is given by
for the platinum dissolution are given in Table 2. the equation (7):
In order to import an ECSA degradation mechanism in the PSþþ
model from the Edmond one, an indirect coupling procedure is Zt
conducted. The approach is similar to the one described in DNk ðtÞ ¼ dnk dt (7)
Refs. [20], but has been slightly improved since then. As a pre- 0
liminary step, a 5-D table is built by performing thousands of
simulations with the EDMOND model to calculate the particles’ Therefore, a global degradation tdegr is computed by the
radius degradation rate for each set of values related to 5 input equation (8):
parameters: the local potential, the initial average particles’ radius, P
the temperature, and the oxygen and vapor molar fractions in the k Sk
tdegr ðtÞ ¼ P (8)
cathode active layer. These parameters have been selected thanks k Sk;ini
to a sensitivity study because they affect the catalyst degradation.
As a second step, two 4-D tables created from the 5-D one are where Sk ðtÞ ¼ Sk;ini ðNk;ini þ DNk Þ is the surface of the population k at
integrated in the PSþþ model’s active layer part, with the 4 input t with Nk;ini being the initial number of particles in the population k
parameters being dynamically calculated by the model (local and Sk;ini being the initial surface of the population k.
electrode potential, local temperature, local oxygen molar fraction, Finally, this degradation rate is included in the expression of the
local vapor molar fraction). Each 4-D table is associated to a specific computed cell voltage to update the local current density i with the
particle population, with a given initial average radius and a given evolving ECSA, as shown in equation (9).
422 C. Robin et al. / Journal of Power Sources 326 (2016) 417e427

" ! ! 2000-h durability test with the PSþþ model using the same
i PO* 2 operating conditions. As previously mentioned, there are two pa-
Ucell ¼ Urev  Rohm i þ b1 þ b2 T þ b3 Tln þ b4 Tln
tdegr P0 rameters to be fitted for the ECSA calculation:
!# " *
!
*
PH PH i - The initial average particles’ radius for each population, r1;ini and
þ b5 Tln 2
þ b6 Tln 2O
þ b7
P0 P0 sc *tdegr r2; ini .
# - The initial number of particles in each population, N1;ini and
i i
þ b8 þ b9 N2;ini .
sa *tdegr tdegr
(9) Besides, the kinetic constants involved in Pt dissolution have
been fitted so that the simulated ECSA decay is in good agreement
Therefore, the PSþþ model is now endowed with an ECSA with the experimental data (previous work to be published).
degradation module. Some of its parameters still need to be The values for these parameters after the fitting procedure are
adjusted. This operation along with first computational results are listed in Table 3. A good agreement between the measured ESCA in
given in the next sub-section. the present work and the simulated one is found (Fig. 3a).
The PSþþ model with the fitted aging module is then computed
4. Algorithm for the aging module’s parameters fitting to simulate the 2000-h durability test. The average cell voltage
evolution monitored during the test is compared with the simu-
This section aims to propose and compare two algorithms to fit lation results in Fig. 3b. A large discrepancy can be observed be-
the parameters of the catalyst layer aging module. The first one tween the two curves. The computed cell voltage decrease is much
relies on CV measurements that are usually considered for the lower than the experimental one, despite the computed ECSA loss
measurement of the ECSA. The performance losses are then being virtually identical to the ECSA loss measured from the Cyclic
computed from the ECSA losses. The second one relies on the global Voltammetry and shown in Fig. 3a. Interestingly, the voltage decay
experimental performance losses from the polarization curves. It predicted from the model seems to match all the upper points of
requires inverting the performance model to get access to an the experimental curve, i.e. the cell voltage level obtained just after
equivalent active surface area degradation that actually integrates the characterization phase. This could correspond to the irrevers-
various degradation mechanisms, including reversible ones. ible losses only, whereas the experimental data during operation
For both approaches, a major assumption is that for modeling integrate both reversible and irreversible losses. This result is
the performance losses of the stack with aging, only catalyst neither surprising nor inconsistent, owing to the CV measurement
degradation through Pt dissolution can be considered as a degra- procedure that implies to rinse the cathode with nitrogen before
dation mechanism. The carbon support corrosion is not taken into the ESCA measurement (see section 2). This latter operation may
account because neither startups nor shutdowns of the fuel cell are actually lead to the elimination of the majority of the reversible
modeled or tested in this study. These operating conditions remain catalyst layer aging phenomena occurring inside the fuel cell. The
the main causes of the carbon support corrosion [25]. Neither the lower experimental performances are certainly related to the lower
degradations of the ionomer nor of the membrane nor of the GDL catalyst utilization rate induced either by the catalyst flooding or by
nor of the bipolar plates are taken into account as well. However, the membrane drying or by the platinum oxidation.
the loss of the equivalent surface area, which modifies locally the Therefore, keeping in mind the willingness to build a simple
current density of the electrochemical response, impacts also the equivalent model of catalyst degradation through Pt dissolution to
ohmic transport of the active layer (both ionic and electronic). model the stack performance decay, fitting the parameters of the
These assumptions are supported by the experimental results degradation module with the ECSA profile measured from the CV is
from the two 2000-h durability tests: not relevant. To simulate the global catalyst losses
(irreversible þ reversible phenomena), the global experimental
- A little internal resistance variation has been measured along performance losses from the polarization curves become our
the aging tests by EIS. The minor evolution of the internal reference data. This leads to the second algorithm proposed in the
resistance is indirectly taken into account by the modification of next sub-section.
the local current density and the modification of the ohmic
losses in the active layer;
- The permeation current remains constant for most of the cells 4.2. Fitting procedure of the aging module using the experimental
(measured by LSV); performance losses from the polarization curves
- Post-mortem tests on the GDL have not shown any loss of hy-
drophobia (observed by water droplets deposition on the Since the first fitting procedure is only able to account for the
surface); irreversible catalyst layer losses, a second one is proposed to ac-
- No trace of bipolar plates corrosion has been noticed during count for both irreversible and reversible losses. This second pro-
post-mortem observations. cedure consists in fitting the aging parameters based on the
polarization curves. Even though the aging module has been built
The model parameters are first fitted using experimental data considering only the catalyst degradation, the second fitting pro-
from the stationary durability test, then the degradation module cedure considers the equivalent loss of active surface area by the
obtained from the second algorithm is validated using data from different irreversible (mainly Pt dissolution) and reversible (such as
the dynamic durability test.
Table 3
4.1. Fitting procedure of the aging module using the CV Resulting values of the ECSA module’s parameters after the fitting procedure with
measurements the experimental ECSA loss (CV).

Parameter r1;ini r2; ini N1;ini N2;ini


This first algorithm for the fitting procedure aims to be able to
Value 0.95 nm 4 nm 1.95  1017 (2/3) 9.73  1016 (1/3)
find back the experimental ECSA loss (cf. Fig. 2b) by simulating a
C. Robin et al. / Journal of Power Sources 326 (2016) 417e427 423

 
i i
b3; degr ðtÞ T lnðiÞ þ b7;degr ðtÞ þ b8;degr ðtÞ þ b9;degr ðtÞ i
sc sa
" !
i i i
¼ b3;ini T ln þ b7;ini þ b8;ini
tdegr ðtÞ sc *tdegr ðtÞ sa *tdegr ðtÞ
#
i
þ b9;ini
tdegr ðtÞ
(10)

where b3; degr ðtÞ, b7; degr ðtÞ, b8; degr ðtÞ and b9; degr ðtÞ are the values
for b3, b7, b8 and b9 coefficients that are fitted every 200 h to match
the experimental performance losses and where b3; ini , b7;ini , b8;ini
and b9;ini are the b3, b7, b8 and b9 coefficients that are fitted at the
beginning of life.
The resulting evolution of the ECSA degradation rate tdegr is
plotted in Fig. 4a and compared to the evolution obtained from the
Cyclic Voltammetry. It can be noticed that the equivalent loss of
active surface extrapolated from the b3, b7, b8 and b9 coefficients is
around 40%, which is much larger than the one inferred from the CV
measurements that is around 13%. We propose that it can be seen as
an equivalent loss of active surface area that would be represen-
tative of both reversible and irreversible losses in the simple
equivalent catalyst degradation module proposed here. Therefore,
Fig. 3. (a): ECSA loss obtained experimentally from the CV measurements (crosses) the equivalent loss of active surface that integrates all the degra-
and fitted ECSA profile from the model (dash line); (b): Comparison of the average cell dation mechanisms is about three times larger than the ECSA losses
voltage evolution between the stationary test (dark) and the model (light). that integrate only irreversible degradations.
The extrapolated loss of active surface can now be used as the
new reference profile to fit the parameters from the degradation
Pt oxidation [25] and ionomer reversible degradations) degradation
mechanisms.
To simulate the performance losses with our aging model, the bk
coefficients of the electrochemical response (cf. equation (1)) have
to be fitted from the polarization curves recorded during the aging
test (see Fig. 2a). Indeed, these coefficients may be extracted from
each 200-h polarization curve recorded during the test.
However, the expressions of the bk coefficients depend only on
constants and therefore should not vary with aging [9,26]. The
ECSA degradation rate tdegr is assumed to be only involved in the
terms depending on the surface area and hence on the current
density (cf. equation (7)). Therefore, the new strategy is to include
the effect of the variation of tdegr in the b3,degr(t), b7,degr(t), b8,degr(t)
and b9,degr(t) coefficients that are therefore varying with aging.
Then tdegr can be calculated and finally the equivalent active sur-
face area can be deduced.
More precisely, the first step of this second algorithm is to
determine all the values for bk,ini coefficients (b1,ini to b9,ini) at the
BoL for the stationary 2000-h durability test (see Appendix A) by
fitting the computed electrochemical response with the initial
polarization curves measured in various operating conditions.
The second step consists in determining new values for b3,degr(t),
b7,degr(t), b8,degr(t) and b9,degr(t) every 200 h (while keeping con-
stant the other bk coefficients) by fitting the electrochemical
response with the polarization curves recorded during the char-
acterization phases. Along the process, the comparison between
the simulated performances with the new values for b3,degr(t),
b7,degr(t), b8,degr(t) and b9,degr(t) and the experimental polarization
curves is carried out to ensure the validity of the approach.
Besides, it is possible to go further with this analysis by using the
values obtained for the b3,degr(t), b7,degr(t), b8,degr(t) and b9,degr(t)
Fig. 4. a) ECSA loss obtained experimentally (CV) and numerically (bi coefficients)
coefficients every 200 h. Their evolutions can actually be directly
with the corresponding computed profiles after fitting the ECSA degradation module’s
exploited to extrapolate an equivalent active surface area degra- parameters; b) Simulated average cell voltage evolution after the different fitting
dation rate tdegr , by mathematically solving the equation (10). methods and comparison to the experimental profile.
424 C. Robin et al. / Journal of Power Sources 326 (2016) 417e427

Table 4 5. Model validation and additional results


Resulting values of the ECSA module’s parameters after the fitting procedure with
the extrapolated ECSA loss.
5.1. Confrontation with the in-situ measuring device (Sþþ)
Parameter r1;ini r2; ini N1;ini N2;ini

Value 1 nm 2 nm 1.95  1017 (2/3) 9.73  1016 (1/3) Thanks to the Sþþ electronic card which has been inserted in
the middle of the 30-cell stack ([16], [18]), it is possible to monitor
the evolution of the current density distribution during the 2000-h
stationary durability test (Fig. 5). The resulting color maps (Fig. 5a)
module of the PSþþ model. These new parameters’ values are
can be compared to the current density distributions obtained after
listed in Table 4.
running the PSþþ model in the test conditions (Fig. 5b).
To check our assumption that the performance losses due to
Since the stack is operated in dry conditions, the BoL current
aging can be described as losses due to catalyst degradation by Pt
density is higher near the air outlet (bottom right hand corner of
dissolution only, the 2000-h stationary durability test is simulated
the color maps) where the membrane is better humidified. Indeed,
again with the model (including the degradation module whose
the protonic conductivity close to the air inlet is too low because of
parameters are fitted according to the second algorithm) in the
the low humidified gases, resulting in a lower current located in
same operating conditions as for the experimental one. The
this area compared with the air outlet [18] [28], [29]. Moreover, the
computed loss of equivalent active surface area is consistent with
experimental and the computed evolutions of the current density
the data extrapolated from the variations of the b degr coefficients
with aging show similar trends. In particular, the BoL heterogene-
(cf. Fig. 4a). The resulting computed average cell voltage decay is
ities seem to be amplified by the long-time operation with steady
plotted in Fig. 4b and compared to the experimental one.
operating conditions. Indeed, the relative current density distri-
The simulated fuel cell aging behavior matches the experi-
bution obtained by subtracting the EoT map from the BoL one (cf.
mental profile with very good agreement, which means that our
Fig. 5) indicates a slight shift of the current density from the air inlet
assumption is relevant. Introducing the equivalent loss of active
to the air outlet during the 2000-h durability test.
surface area seems therefore sufficient enough to simulate the
Therefore, the equivalent active surface degradation module
performance decay due to the complex aging phenomena occurring
included in the PSþþ model enables to find back the evolution of
during a 2000-h fuel cell operation with steady operating condi-
the current density distributions with aging, which further
tions. For our stationary test, this assumption is also supported by
strengthen the model validity.
post-mortem analyses that have shown that the losses occurring
during the test are mainly due to the catalyst layer aging, no major
membrane degradation being observed [27]. 5.2. Confrontation with a local ECSA assessment using a post-
The values for the parameters that are now used in the PSþþ mortem mini-cell device
model are listed in Table 4 and the validity of the aging module is
further examined in the next part. As described in the modeling part, the PSþþ model is well

Fig. 5. Current density distributions measured by the Sþþ device (upper line) and predicted by the PSþþ model (lower line). The maps of the right column are obtained after
subtracting the EoT map from the BoL one.
C. Robin et al. / Journal of Power Sources 326 (2016) 417e427 425

Fig. 6. Distribution of the ECSA loss (a) and cathode Relative Humidity (b) computed by the PSþþ model and corresponding to the 2000-h durability test with steady operating
conditions.

adapted to study the distribution of the local conditions on the cell conducted. After the end of the 2000-h stationary durability test,
surface. In particular, the map of the aged equivalent surface area several MEAs are selected to be further examined using a mini-cell
corresponding to the voltage decrease shown in Fig. 4b can be experiment. The coulometry curves obtained for the 3 regions of
obtained. The distribution observed in Fig. 6a indicates a larger loss the cell n 10 are plotted in Fig. 7b and compared to the ones of a
of equivalent surface area near the air outlet, even if the degrada- fresh MEA sample (Fig. 7a). Two peaks can be observed on the fresh
tion seems to be relatively homogeneous. After examining all the MEA coulometry curves, which is typical for a fresh Pt3Co catalyst.
local conditions thanks to the PSþþ model, it seems that in the For the aged cell, the second peak of the air outlet sample is much
areas of the cell surface where the RH is higher, the loss of equiv- more pronounced than for the other samples. This would mean that
alent surface area is also stronger (Fig. 6b). This could be explained the air outlet is more prone to catalyst structural modifications such
by the higher catalyst oxidation in well-humidified zones. as facets reorganizations, which could lead to a more severe ECSA
To confirm the heterogeneities in the loss of equivalent surface loss in this region. It is important to highlight here that these trends
area predicted by the PSþþ model, post-mortem analyses are obtained for the cell n 10 are reproducible and have been observed

Fig. 7. (a): local CV result for a fresh MEA; (b): local CV result for an aged MEA (cell n 10 of the 30-cell stack); (c): cathode catalyst layer’s particles size distribution resulting from
the STEM observations.
426 C. Robin et al. / Journal of Power Sources 326 (2016) 417e427

conditions is performed (see the Experimental part).


The average cell voltage decay corresponding to this second test
is simulated with the PSþþ model, with the same parameters for
the degradation module than the ones set thanks to the first test
(see Table 4). The results are shown in Fig. 8 and compared with the
experimental cell voltage decay. The comparison is done during the
cycle plateaus of 0.4 A cm2 for sake of clarity and to be consistent
with the current density of the first durability test. The equivalent
loss of active surface area enables to get a computed electro-
chemical response which is in a relatively good agreement with the
experimental data. The computed performance decay is however
slightly lower than the experimental one. This could be explained
by degradation mechanisms that could be very different in dynamic
conditions compared to stationary ones. Refinement of the equiv-
Fig. 8. Application of the equivalent ECSA loss approach developed thanks to the first
test during the 0.4 A cm2 plateaus of the current profile. alent degradation model to better fit to aging in dynamic conditions
is needed to increase the predictability of the model.

on the other aged cells from the 30-cell stack.


6. Conclusion
Scanning Transmission Electron Microscopy (STEM) has also
been used in this study to complete the mini-cell results. The mi-
In this work, an innovative numerical method has been intro-
croscopy pictures are not shown in this paper. As it can be shown in
duced to model an equivalent loss of active surface area during PEM
Fig. 7c, the count of the particles and their diameters in several
fuel cells long-time operation. This method leads to simulated
pictures indicate that the air outlet region contains bigger particles
performance losses that are similar to the experimental ones. To do
than the air inlet. This confirms that the ECSA loss at the air outlet is
so, an experimental 2000-h durability test has been performed in
more severe than for the air inlet.
steady operating conditions. The Cyclic Voltammetries measured
The mini-cell test results and the STEM observations therefore
regularly during the test are first used to fit the parameters of a
validate the ECSA degradation location predictions by the PSþþ
physical ECSA aging module included in a 2Dþ1D fuel cell model.
model. It is important to highlight here that the shift of the current
This first approach revealed an underestimation of the cell voltage.
density is not expected to be neither a cause nor a consequence of
Interestingly, it was shown to simulate the irreversible degrada-
the location of the loss of ECSA.
tions correctly. This is probably due to the CV measurement pro-
tocol, which does not enable to observe the reversible degradation
5.3. Confrontation with the 2000-h durability test in dynamic effects on the ECSA. To consider realistic degradations while
operating conditions keeping a simple equivalent degradation module whose equations
correspond to catalyst degradation only, a procedure is built to fit
Since the degradation module of the PSþþ model has been built the parameters of the degradation module on an equivalent loss of
and adjusted thanks to the 2000-h durability test in steady con- active surface area profile, which is extrapolated from the decrease
ditions, it is important to validate the model with another dura- of the experimental performances with aging. The results show
bility test. To do so, a 2000-h durability test in dynamic operating that the computed cell voltage evolution using this method is in

Table 5
Values of the bk coefficients determined by fitting the performances evolution for the first durability test (steady operating conditions)

0 h (BoT) 200 h 400 h 600 h 800 h 1000 h 1200 h 1400 h 1600 h 1800 h 2000 h

b1 0.704 Coefficient kept constant.


b2 1.75 105 Coefficient kept constant.
b3 3.44 105 4.73 105 4.94 105 5.10 105 5.22 105 5.03 105 4.84 105 5.60 105 5.30 105 5.47 105 5.26 105
b4 1.23 104 Coefficient kept constant.
b5 1.32 105 Coefficient kept constant.
b6 2.70 107 Coefficient kept constant.
b7 3.30 105 6.78 105 8.36 105 6.55 105 8.75 105 5.58 105 8.36 105 5.93 105 9.05 105 9.46 105 6.42 105
b8 1.10 105 4.47 106 9.52 106 1.84 105 4.02 107 6.78 106 1.25 105 2.65 105 2.88 106 3.78 109 4.40 106
b9 9.62 106 1.07 106 4.54 107 7.42 107 1.98 108 6.51 106 7.49 1011 1.15 109 4.62 108 8.07 1011 6.96 106

Table 6
Values of the bk coefficients determined by fitting the performances evolution for the second durability test (dynamic operating conditions)

0 h (BoT) 200 h 400 h 600 h 860 h 1040 h 1270 h 1430 h 1710 h 1970 h

b1 0.739 Coefficient kept constant.


b2 4.34 105 Coefficient kept constant.
b3 2.94 105 3.95 105 4.11 105 4.39 105 4.53 105 4.72 105 4.77 105 4.63 105 4.40 105 4.17 105
b4 1.37 104 Coefficient kept constant.
b5 2.06 106 Coefficient kept constant.
b6 4.67 107 Coefficient kept constant.
b7 2.83 105 4.67 105 5.87 105 9.20 105 8.71 105 9.05 105 9.35 105 8.29 105 8.77 105 2.52 105
b8 8.29 106 3.38 105 2.44 105 8.06 108 1.58 106 4.61 106 5.68 108 1.43 105 3.73 106 2.19 105
b9 1.34 105 2.13 106 1.53 106 5.21 107 1.73 106 1.05 106 1.11 106 9.98 107 5.42 106 1.60 105
C. Robin et al. / Journal of Power Sources 326 (2016) 417e427 427

very good agreement with the experimental one, which validates [5] J. Wu, X.Z. Yuan, J.J. Martin, H. Wang, J. Zhang, J. Shen, S. Wu, W. Merida,
A review of PEM fuel cell durability: degradation mechanisms and mitigation
the assumption that the performance losses due to the complex
strategies, J. Power Sources 184 (1) (2008) 104e119.
degradation phenomena occurring in the stack can be simulated by [6] L. Dubau, L. Castanheira, M. Chatenet, F. Maillard, Ja ~© Dillet, G. Maranzana,
a simple equivalent degradation module. The approach is then S. Abbou, O. Lottin, G.D. Moor, A.E. Kaddouri, C. Bas, L. Flandin, E. Rossinot,
N. CaquA©,~ Carbon corrosion induced by membrane failure: the weak link of
further supported by exploiting other data from the first durability
{PEMFC} long-term performance, Int. J. Hydrogen Energy 39 (36) (2014)
test (current density distributions evolution from the Sþþ in-situ 21902e21914.
device, ECSA loss main location from a post-mortem local charac- [7] T. Jahnke, G. Futter, A. Latz, T. Malkow, G. Papakonstantinou, G. Tsotridis,
terization using a mini-cell device and STEM observations). The P. Schott, M. Gerard, M. Quinaud, M. Quiroga, A.A. Franco, K. Malek, F. Calle-
Vallejo, R.F. de Morais, T. Kerber, P. Sautet, D. Loffreda, S. Strahl, M. Serra,
aging model is finally validated by comparing its simulated results P. Polverino, C. Pianese, M. Mayur, W.G. Bessler, C. Kompis, Performance and
to the experimental results from a second 2000-h durability test in degradation of proton exchange membrane fuel cells: state of the art in
dynamic conditions. modeling from atomistic to system scale, J. Power Sources 304 (2016)
207e233.
This study aims at modeling the experimental performances [8] A.V. Virkar, Y. Zhou, Mechanism of catalyst degradation in proton exchange
losses of a PEMFC stack. Interestingly, an equivalent catalyst membrane fuel cells, J. power sources 154 (2007) B540eB547.
degradation module whose parameters are fitted according to the [9] M.W. Fowler, W.F. Mann, J.C. Amphlett, B.A. Peppley, P.R. Roberge, Incorpo-
ration of voltage degradation into a generalised steady state electrochemical
methodology proposed here was shown to lead to similar perfor- model for a PEM fuel cell, J. Power Sources 106 (2002) 274e283.
mance losses. However, to predict the fuel cell lifetime, more [10] M. Soltani, S.M.T. Bathaee, Development of an empirical dynamic model for a
degradation mechanisms should be added to the fuel cell model, Nexa PEM fuel cell power module, Energy Convers. Manag. 51 (2010)
2492e2500.
with a special need for a membrane degradation module. Indeed, as
[11] M. Jouin, R. Gouriveau, D. Hissel, M.-C. Pera, N. Zerhouni, Prognostics of PEM
illustrated in this study, the Electrochemical Catalyst Surface Area fuel cells under a combined heat and power profile, IFAC 48 (3) (2015) 26e31.
degradation seems to be the main aging mechanism decreasing the [12] M. Jouin, R. Gouriveau, D. Hissel, M.-C. Pera, N. Zerhouni, Prognostics of PEM
cell performances, but it lowers with time and is not responsible for fuel cell in a particle filtering framework, Int. J. Hydrogen Energy 39 (2014)
481e494.
the cell failure. Conversely, the membrane degradation effect on the [13] A.A. Kulikovsky, The effect of noneuniform aging of a polymer electrolyte fuel
fuel cell voltage evolution remains weak, except when the cell is cell on the polarization curve: a modeling study, Electrochimica Acta 123
very close to failure which entails a dramatic permeation current (2014) 542e550.
[14] E. Pohl, M. Maximini, A. Bauschulte, J. vom Schloss, R.T.E. Hermanns, Degra-
increase and cell voltage collapse. Dedicated works should address dation modeling of high temperature proton exchange membrane fuel cells
the membrane failure modeling in order to improve numerical using dual time scale simulation, J. Power Sources 275 (2015) 777e784.
lifetime predictions. In particular, a statistical study could be per- [15] S.R. Dhanushkodi, S. Kundu, M.W. Fowler, M.D. Pritzker, Use of mechanistic
carbon corrosion model to predict performance loss in Polymer Electrolyte
formed to try establishing a link between the membrane failure and Membrane fuel cells, J. Power Sources 267 (2014) 171e181.
a possible initial defect on the surface of the MEA. Other fuel cell [16] L. Jabbour, C. Robin, F. Nandjou, R. Vincent, F. Micoud, J.-P. Poirot-Crouvezier,
operating profiles could also be examined, like start/stop cycles J. d Arbigny, M. Gerard, Feasibility of in-plane GDL structuration: impact on
current density distribution in large-area proton exchange membrane fuel
which can be numerous for a fuel cell use in transportation. In cells, J. Power Sources 299 (2015) 380e390.
particular, their effects on the ECSA loss should be assessed by [17] Sþþ Simulation Services, 2015. www.splusplus.com.
modeling. [18] C. Robin, M. Gerard, J. d Arbigny, P. Schott, L. Jabbour, Y. Bultel, Development
and experimental validation of a PEM Fuel Cell 2D-model to study hetero-
geneities effects along large-area cell surface, Int. J. Hydrogen Energy 40
Acknowledgement (2015) 10211e10230.
[19] J.B. Young, B. Todd, Modelling of multi-component gas flows in capillaries and
The authors would like to thank Zodiac Aerospace for their porous solids, Int. J. Heat Mass Transf. 48 (25e26) (2005) 5338e5353.
[20] C. Robin, M. Gerard, A.A. Franco, P. Schott, Multi-scale coupling between two
financial support. The authors are also grateful to Irina Profalitova dynamical models for PEMFC aging prediction, Int. J. Hydrogen Energy 38 (11)
and Marion Scohy for their help concerning the mini-cell device. (2013) 4675e4688.
The authors finally thank Laure Guetaz for the STEM observations. [21] A.A. Franco, Chapter 11-Modelling and analysis of degradation phenomena in
polymer electrolyte membrane fuel cells, in: C. Hartnig, C. Roth (Eds.), Poly-
mer Electrolyte Membrane and Direct Methanol Fuel Cell Technology, vol. 1,
Appendix A. bk coefficients’ values Woodhead Publishing, 2012, pp. 291e367.
[22] A.A. Franco, P. Schott, C. Jallut, B. Maschke, A multi-scale dynamic mechanistic
model for the transient analysis of PEFCs, Fuel Cells 2 (2007) 99e117.
The bk coefficients of the PSþþ model’s electrochemical [23] A.A. Franco, M. Gerard, Multiscale model of carbon corrosion in a PEFC:
response have been fitted on the experimental polarization curves coupling with electrocatalysis and impact on performance degradation,
using a Matlab® linear optimization script. The convergence crite- J. Electrochem. Soc. 155 (4) (2008) B367eB384.
[24] E.F. Holby, W. Sheng, Y. Shao-Horn, D. Morgan, Pt nanoparticle stability in
rion has been set to 20 mV of gap between the experimental per- PEM fuel cells: influence of particle size distribution and crossover oxygen,
formances and the computed ones. The values of the bk coefficients Energy Environ. Sci. 2 (2009) 865e871.
for the first 2000-h durability test (steady operating conditions) are [25] G. Maranzana, A. Lamibrac, J. Dillet, S. Abbou, S. Didierjean, O. Lottin, Startup
(and shutdown) model for polymer electrolyte membrane fuel cells,
listed in Table 5. The values of the bk coefficients for the second
J. Electrochem. Soc. 162 (7) (2015) F694eF706.
2000-h durability test (dynamic operating conditions) are listed in [26] J.C. Amphlett, R.M. Baumert, R.F. Mann, B.A. Peppley, P.R. Roberge, T.J. Harris,
Table 6. Performance modeling of the Ballard Mark IV solid polymer electrolyte fuel
cell, J. Electrochem. Soc. 142 (1) (Jan. 1995) 9e15.
[27] C. Robin, De veloppement d’un mode le pre
dictif de dure
e de vie d’une pile
PEMFC pour une application ae ronautique: e tude des interactions entre le
References cœur de pile et les conditions d’ope ration du systeme, Universite de Grenoble,
2015.
[1] V. Ramani, Fuel cells, Electrochem. Soc. Interface 15 (1) (2006) 41e44. [28] I. Alaefour, G. Karimi, K. Jiao, X. Li, Measurement of current distribution in a
[2] Fuel Cell Technical Team Roadmap, U.S. DRIVE Partnership, 2013. proton exchange membrane fuel cell with various flow arrangements: a
[3] F.A. de Bruijn, V. a. T. Dam, G.J.M. Janssen, Review: durability and degradation parametric study, Appl. Energy 93 (2012) 80e89.
issues of PEM fuel cell components, Fuel Cells 8 (1) (2008) 3e22. [29] M. Schulze, E. Gulzow, S. Schonbauer, T. Knori, R. Reissner, Segmented cells as
[4] N. Yousfi-Steiner, Ph. Mocoteguy, D. Candusso, D. Hissel, A review on PEM tool for development of fuel cells and error prevention/prediagnostic in fuel
voltage degradation associated with water management: impacts, influent cell stacks, J. Power Sources 173 (2007) 19e27.
factors and characterization, J. Power Sources 183 (2008) 260e274.

You might also like