You are on page 1of 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/223269919

Mathematical modelling of ambient air-breathing fuel cells for portable


devices

Article  in  Electrochimica Acta · March 2007


DOI: 10.1016/j.electacta.2006.11.002

CITATIONS READS

59 303

2 authors:

Shawn Litster Ned Djilali


Carnegie Mellon University University of Victoria
211 PUBLICATIONS   4,728 CITATIONS    312 PUBLICATIONS   12,437 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Wall Flows friction turbulence and roughness effects View project

Hydrogen peroxide detection in PEM Fuel cells View project

All content following this page was uploaded by Ned Djilali on 29 September 2017.

The user has requested enhancement of the downloaded file.


This article was originally published in a journal published by
Elsevier, and the attached copy is provided by Elsevier for the
author’s benefit and for the benefit of the author’s institution, for
non-commercial research and educational use including without
limitation use in instruction at your institution, sending it to specific
colleagues that you know, and providing a copy to your institution’s
administrator.
All other uses, reproduction and distribution, including without
limitation commercial reprints, selling or licensing copies or access,
or posting on open internet sites, your personal or institution’s
website or repository, are prohibited. For exceptions, permission
may be sought for such use through Elsevier’s permissions site at:

http://www.elsevier.com/locate/permissionusematerial
Electrochimica Acta 52 (2007) 3849–3862

Mathematical modelling of ambient air-breathing

py
fuel cells for portable devices
S. Litster 1 , N. Djilali ∗

co
Institute for Integrated Energy Systems, and Department of Mechanical Engineering, University of Victoria, Victoria, BC V8W 3P6, Canada
Received 27 November 2005; received in revised form 2 November 2006; accepted 5 November 2006
Available online 30 November 2006

Abstract
Competitive costs, instant recharge, and high energy density make fuel cells ideal for supplanting batteries in portable electronic devices. In this

al
study, we derive a semi-analytical model to elucidate the transport of ions, heat and mass in air-breathing fuel cells. The model includes an empirical
correlation for membrane conductivity that improves accuracy when modelling membrane dry-out. A detailed comparison with experimental
data demonstrates that the model accurately predicts fuel cell performance through detailed accounting of catalyst layer specifications, including
on
variable width, and electrochemical parameters. A comprehensive parametric study resolves the trends associated with a variety of design
specifications and operating conditions. Membrane dry-out is identified as the primary limitation on current density, and is shown to be strongly
dependent on heat transfer. The study also identifies some unique effects of coupling between ambient air temperature and humidity on the
performance of air-breathing fuel cells.
rs

© 2006 Elsevier Ltd. All rights reserved.

Keywords: Fuel cells; Analytical models; Portable devices; Convective heat transfer; Membrane dry-out
pe

1. Introduction is that even if the capacity of Li-ion batteries grows at 10%


per year, they will still not be capable of meeting the power
Techno-economic analyses have shown that portable requirements of future devices.
consumer electronics, currently powered by rechargeable The envisaged implementation of PEM fuel cells for portable
batteries, present a more accessible market for proton exchange devices is in a device integrated form. This entails placing the
r's

membrane (PEM) fuel cells than transportation applications in fuel cell either on, or within portable electronics. For example,
the immediate future, because a higher cost per unit of energy is fuel cells can be mounted on the exterior surfaces of mobile
acceptable at these smaller scales [1,2]. For example, Dyer [1] phones, as illustrated in Fig. 1, and in notebook PCs. Based
stated that the cost tolerance for fuel cells in portable equipment on the predicted planar area available on a mobile phone, the
power density requirement of the fuel cell is 60 mW cm−2 [2].
o

is two orders of magnitude greater than that for automotive


applications. Another key advantage of fuel cells for portable The cathode gas diffusion layer is fed ambient air, rather than
th

applications over batteries is longer continuous operation with conditioned air through flow channels, and the fuel is stored
almost instantaneous refueling. and distributed below the fuel cell. One advantage of an air-
An increasingly more prevalent driver of fuel cell develop- breathing cathode is that it eliminates the need for manifolding.
ment for portable devices is the ever-increasing energy density In addition, the open surface facilitates heat transfer from the
Au

requirements of new consumer electronics, such as mobile fuel cell.


phones with digital broadcast reception [2]. A key argument for A schematic of such an air-breathing fuel cell is shown in
the integration of fuel cells into next-generation mobile phones Fig. 2. This arrangement incorporates multiple methodologies
for reducing the size and number of ancillary components:
∗ Corresponding author. Tel.: +1 250 721 6034; fax: +1 250 721 6323.
E-mail address: ndjilali@uvic.ca (N. Djilali).
1. To alleviate the need for air distribution channels, along with
1 Present address: Department of Mechanical Engineering, Stanford Univer- the necessary pumps and fans, the cathode gas diffusion layer
sity, Stanford, CA 94305, USA. is in direct contact with the ambient air.

0013-4686/$ – see front matter © 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.electacta.2006.11.002
3850 S. Litster, N. Djilali / Electrochimica Acta 52 (2007) 3849–3862

flowed through the gas diffusion layer parallel to the mem-


brane. Schmitz et al.’s [4] planar micro-fuel cell was produced
using printed circuit board (PCB) technology. In this design, the
surface of the cathode gas diffusion layer was largely open to
ambient air. A plate with rectangular openings placed over the
gas diffusion layer provided improved current conduction and
allowed the passage of air to the cathode. The active area of this
cell was 10 cm2 , producing a power density of 100 mW cm−2 at
a cell voltage of 0.5 V. The second Fraunhofer planar micro-fuel

py
cell, discussed by Hahn et al. [5], was fabricated with wafer-
level patterning technologies and had a planar area of 1 cm2 . The
cell was able to operate in a stable manner at a power density of
80 mW cm−2 using dry hydrogen as the fuel. The Fabian et al. [6]

co
cell had an active area of 9 cm2 in a similar arrangement to that
of Schmitz et al. This fuel cell was fabricated using printed cir-
cuit board technology and was able to maintain a current density
of 400 mA cm−2 under most environmental conditions.
Fig. 1. Illustration of an ambient air-breathing fuel cell as a power source for a Researchers and designers are increasingly turning to fuel cell
mobile phone with video playback. models to improve fundamental understanding of the transport

al
phenomena present in PEM fuel cells and to use these models
2. External humidification systems are eliminated and the fuel for design optimization. The usefulness of such models when
cell relies on the ambient relative humidity and water produc- considering device integrated fuel cells requires an approach and
tion in the cathode for the humidification of the membrane.
on
model parameters that are not specific to overly narrow operating
3. The circulating ambient air facilitates the cooling of the fuel conditions or design parameters.
cell in lieu of a dedicated heat management system. A number of previous PEM fuel cell models have employed
4. The anode is fed dry hydrogen at roughly ambient tem- analytic or semi-analytic solution techniques [7–16]. These
perature and pressure, eliminating hydrogen conditioning
rs

models contrast those requiring discretization in conjunction


components. with advanced numerical methods such as computational fluid
5. Electrical current is distributed across the cell in an edge- dynamics [17–23]. Although analytical models typically require
collection arrangement. fairly extensive simplifying assumptions, the closed form math-
pe

ematical solution they yield can often provide significant insight


Four illustrative examples of planar fuel cell designs that into the expected physical behaviour of the system, without
passively breathe ambient air have been presented in the open requiring the time consuming and computationally intensive
literature by Wainright et al. [3], Schmitz et al. [4], Hahn et al. parametric studies that must be conducted with numerical
[5], and most recently Fabian et al [6]. The Wainright et al. [3] models. This can be particularly advantageous in a design envi-
fuel cells were assembled on ceramic substrates with thick-film ronment requiring rapid analysis and assessment cycles.
fabrication techniques. The design featured hydrogen storage
r's

In the present work we attempt, through modelling, to


in a reservoir wafer located directly below the anode; this is understand the coupling between key physical processes in air-
similar to the arrangement shown in Fig. 2. Current was dis- breathing fuel cells, and to refine design objectives and identify
tributed in porous metal gas diffusion layers in which the current constraints for such fuel cells. In the next section, we present
the derivation and validation of a semi-analytical MEA model.
o

We then combine this model with heat and mass transfer models
for air-breathing conditions, and with a membrane conductivity
th

correlation that is suitable for low humidity operation. The pre-


dictive capabilities of the model are validated against relevant
experimental data. The model is finally used to analyze the per-
Au

formance of air-breathing fuel cells for a range of heat transfer


rates, ambient conditions, and catalyst layer specifications.

2. Analytical MEA model

2.1. Introduction

In this section, we present a one-dimensional model of the


membrane electrode assembly in a PEM fuel cell suitable for
analysis and optimization of micro-fuel cells that passively
Fig. 2. Schematic of the planar air-breathing fuel cell. breathe ambient air. Passive air-breathing operation frequently
S. Litster, N. Djilali / Electrochimica Acta 52 (2007) 3849–3862 3851

following expression for the potential summation through the


MEA:
WMem
(EOC − Ecell ) − ηC − ηA − iC (ηC ) = 0 (1)
σMem
which is subject to the constraint of current conservation:
iC (ηC ) − iA (ηA ) = 0 (2)

py
The reversible potential (Er ) results from the Gibbs energy
of formation (G):
G
Er = − (3)
nF

co
where n is the moles of product per electron transferred (n = 2)
and F is Faraday’s constant. In this section, the correlation from
Parthasarathy et al. [28] is used to obtain G as function of
temperature for the formation of liquid water. In Section 3,
Fig. 3. Schematic of the one-dimensional analytical MEA model.
where an air-breathing fuel cell is considered, standard ther-
modynamic tables determine G for the formation of water
entails low-humidity conditions and lower current densities due

al
vapour. The Nernst relation accounts for the activities of oxy-
to increased electrolyte resistance [6,24,25]. Thus, we expect gen and hydrogen (aO2 and aH2 ) and provides the open-circuit
Ohmic losses in the electrolyte to have a greater relative impor- voltage:
on
tance than the mass transfer limitations. Consequently, the model  
derivation focuses on the Ohmic losses in the electrolyte rather RT 1
EOC = Er + ln(aH2 ) + ln(aO2 ) (4)
than mass transport in the cathode. nF 2

2.2.2. Catalyst layer model


rs
2.2. Model
The slow oxygen reduction reaction (ORR) distinguishes the
The model domain, shown in Fig. 3, includes both cath- cathode catalyst layer from the anode. When oxygen transport to
ode and anode catalyst layers and the membrane. As shown in the catalyst layer is insufficient, the ORR occurs predominantly
pe

the Figure, the MEA model consists of five main components. near the GDL interface, whereas when the oxygen supply is
The potential summation algorithm calculates the distribu- abundant and the electrolyte conductivity is low, the ORR con-
tion of losses through the MEA. Individual potential losses centrates by the membrane interface because of Ohmic losses in
in the cathode, anode, and membrane are calculated using a the catalyst layer. Due to a much higher electrical conductivity,
finite-thickness catalyst layer model, an interface catalyst layer the electric potential is approximately uniform in comparison
model, and a linear membrane model with uniform conductiv- [29].
ity, respectively. The finite-thickness model is implemented with In the finite-thickness catalyst layer model derived here, we
r's

parameters evaluated from a macrohomogeneous catalyst layer neglect the gradient in oxygen concentration that is due to dif-
model that correlates the required volume average quantities fusion. This is acceptable when considering low current density
with the microstructure of the catalyst layer. A novel feature operation typical of air-breathing fuel cells and in cases where
of the present macrohomogeneous model is the variable cata- Ohmic losses are dominant. Eikerling and Kornyshev [30] pre-
o

lyst layer thickness, which is determined from the catalyst layer sented a parameter that determines when these conditions prevail
specifications. based on the Tafel slope (RT/αF), ionic conductivity (σ), and a
th

diffusion parameter (Θ):


2.2.1. Potential summation algorithm αF Θ
The potential summation algorithm is similar to the voltage- g= (5)
RT σ
Au

to-current methods employed by Nguyen et al. [19] and Sivertsen


and Djilali [21]. The algorithm allocates the difference between where
open-circuit voltage and cell potential (EOC − Ecell ). This poten- 4FPair DO
CL
tial difference is the sum of the activation overpotentials in the Θ= 2
(6)
RTWCL
cathode and anode (ηC , ηA ) and the potential drop through the
membrane. We neglect the potential drop through the electrical in which DO CL is the oxygen diffusivity in the catalyst layer and
2
pathways due to its minimal contribution to the overall potential WCL is the thickness of the catalyst layer. For the parameter
drop. The potential drop through the membrane is the product values considered in this analysis g  1, and oxygen transport
of the area-specific membrane resistance (WMem /σ Mem ) and the is rapid in comparison to proton conduction. Our catalyst layer
current density (i). Taking current to be a result of the over- model is similar to that of Eikerling and Kornyshev [30] for the
potential in the cathode catalyst layer (iC (ηC )), we obtain the limit of rapid oxygen transport. It is also similar to the model
3852 S. Litster, N. Djilali / Electrochimica Acta 52 (2007) 3849–3862

derived by Gurau et al. [13] for their analytical half-cell model [13,30,33,34]. We propose a new implementation that accounts
of a PEM fuel cell, but with the important distinction that Gurau for variable catalyst layer thickness.
et al. assumed a uniform overpotential within the catalyst layer
to solve oxygen distribution. 2.2.3.1. Catalyst layer composition. The volume of the catalyst
Our derivation begins with the governing equation for proton layer is composed of four components: platinum, Nafion, car-
conduction, Ohm’s law: bon, and void space. By employing the standard catalyst layer
dφ specifications and a known catalyst layer thickness, each vol-
i = −σ (7) ume fraction can be determined [34]. The typical specifications,
dx
ranges, and units are [35]:

py
where φ is the electrolyte potential. Cast in a conservative form
with a source term for the current generation (jT ) the equation 1. Platinum loading, mPt (0.05 − 5 mg Pt cm−2 ).
becomes: 2. Platinum/carbon ratio, yPt (20 − 40% Pt/C).
3. Nafion content, yN (20 − 60 wt.%).

co
σ ∇ 2 φ = jT (8)
The rate of the ORR is locally dependent on the overpotential The volume fraction of platinum is simply the loading divided
(the difference between the potential of electrolyte and electric by the density of platinum (ρPt ) and the catalyst layer thickness
potential in the platinum η = φ − φs ). Following Perry et al. [31] (WCL ):
and many other studies, we determine the reaction rate, assuming
1
high overpotentials (η  RT/αF), with the Tafel equation: εPt = mPt (12)
 γ ρPt WCL

al
 
c̄O2 αF
jT = j0 ref exp η (9) The volume fraction occupied by carbon is a function of the
cO RT carbon loading (mC ) that results from the specified platinum
2
on
loading and the platinum/carbon ratio:
In addition, the reaction rate depends on the average oxygen
concentration in the layer (c̄O2 ). The concentration dependence 1 mPt (1 − yPt ) 1
εC = mC = (13)
exponent (γ) specifies the sensitivity of the reaction to oxygen ρC WCL yPt ρC WCL
concentration. We solve Eqs. (7)–(9) implicitly to obtain:
rs

  As shown in Eq. (14), we evaluate the volume fraction of


1 bA exp(bη0 ) Nafion from the area loading of the Nafion (mN ) divided by
η(x) = ln tan2 (x − WCL ) + . . . the Nafion density (ρN ) and the thickness of the catalyst layer.
b 2
pe

Nafion loading is a function of three commonly documented


√ 

exp(bη1 ) − exp(bη0 ) catalyst layer properties: Nafion content (weight percentage of


arctan √ + 1 + η0 the catalyst layer that is Nafion), platinum loading, and plat-
exp(bη0 )
(10) inum/carbon ratio:
1 yN mPt 1
where εN = mN = (14)
 γ ρN WCL yPt (1 − yN ) ρN WCL
j0 c̄O2 αF
r's

A= and b= , and since the volume fractions must add to unity, the void frac-
σ cOref RT
2 tion (εV ) is:
where x is the location in the layer, WCL the thickness of the layer,
εV = 1 − (εPt + εC + εN ) (15)
and η0 and η1 are the overpotentials at the GDL and membrane
o

interfaces, respectively. η1 corresponds to ηC in Eq. (1). The


derivation of Eq. (10) is presented in ref. [32]. Eq. (10) is similar, 2.2.3.2. Variable-thickness implementation. Previous studies
of catalyst layer composition optimization based on similar rela-
th

but not identical, to Eikerling and Kornyshev’s [30] expression.


The derivative of overpotential multiplied by the conductivity tions [34,36,37] have typically assumed the void fraction to vary
provides an expression for the protonic current into the cathode: with catalyst layer specifications and the catalyst layer thick-
 ness to remain constant regardless of specifications. However,
Au

2A the experimental results of Inoue et al. [38] demonstrate that the


i=σ exp(bη1 ) − exp(bη0 ) (11) Nafion content and platinum loading have a significant effect on
b
catalyst layer thickness. Their results show that the catalyst layer
2.2.3. Catalyst composition and variable thickness thickness is proportional to platinum loading and that increasing
We utilize a macrohomogeneous catalyst layer model to Nafion content from 10 to 50 wt.% results in a 100% increase in
investigate the effect of catalyst layer composition. This model thickness.
considers the catalyst layer microstructure as a homoge- In contrast with previous work [34,36,37], in our analysis
neous medium with properties that reflect the heterogeneous the thickness of the catalyst layer is calculated from the catalyst
microstructure. As noted by Eikerling and Kornyshev [30], the layer specifications and we hold the void fraction constant in
macrohomogeneous model has been around for decades and accordance with the experimental results of Inoue et al. [38].
variations in derivation and implementation continue to emerge We conjecture that the volume fraction of solvent in the catalyst
S. Litster, N. Djilali / Electrochimica Acta 52 (2007) 3849–3862 3853

py
co
Fig. 4. Schematic of the influence of changing Nafion content or platinum loading has on the microstructure and thickness of the catalyst layer.

layer prior to the evaporation step in the MEA fabrication is con- researchers [13,36,37,40], and is suitable here because of the
sistent over a range of catalyst layer specifications and produces disperse catalyst agglomerates.

al
a constant void fraction.
Fig. 4 presents a schematic of the effect of changing two 2.2.3.3. Volumetric exchange current density. When solving
different catalyst layer specifications (Nafion content and plat-
on the Tafel equation in the finite-thickness model, the main electro-
inum loading), and illustrates how increasing the Nafion content chemical parameter is the volumetric exchange current density
increases the thickness of the catalyst. In addition, for constant (j0 ). As shown in Eq. (18), the volumetric exchange current
platinum loading, an increase in Nafion content increases the density is the product of three factors. The dominant one is
mean distance between catalyst particles. Increased Nafion con- the exchange current density of the platinum/Nafion interface
rs

tent, resulting in a higher Nafion volume fraction, improves the Pt/N


(i0 ), the second is the area of platinum/Nafion interface per
ability of ions to reach the catalyst sites. However, the increased unit volume (APt ), and the third is the platinum utilization (υPt ),
thickness also increases the length scales for oxygen diffu- which is the percentage of the platinum area that is electrochem-
pe

sion and ionic conduction, and, correspondingly, reduces their ically active.
transport rates. Our hypothesis is supported by experiments doc-
Pt/N
umenting greater mass transport limitations with higher Nafion j0 = APt υPt i0 (18)
contents [35]. In contrast, when the platinum loading is increased
The platinum surface area per unit volume (APt ) is a function
there is no change in volume fractions. The only change is
of the platinum loading (mPt ), platinum surface area per unit
the thickness of the layer. Thus, the benefit of increasing the
mass (sPt ), and the thickness of the catalyst layer (WCL ):
platinum content, namely the increased total electrochemically
r's

active area (surface area of Pt × catalyst utilization), is countered mPt sPt


APt = (19)
by increased diffusion and conduction length scales. WCL
The thickness of the catalyst layer is determined by solving To account for the utilization of the total platinum surface
the previous set of volume fraction equations (Eqs. (12)–(15))
o

area, we employ an empirical relationship between the Nafion


for a constant void fraction: content in the catalyst layer and utilization. The utilization data
 
1 mPt (1 − yPt ) 1 yN mPt
th

1 1 of Sasikumar et al. [41] was obtained using the ratio of elec-


WCL = mPt + +
1 − εV ρPt yPt ρC yPt (1 − yN ) ρN trochemical surface area, measured with cyclic voltammetry, to
the theoretical value for a platinum/carbon ratio of 20% Pt/C.
(16)
A third-order polynomial fit to their platinum utilization (υPt )
Au

data is presented in Eq. (21). The optimum Nafion content (yN )


We use the volume fraction of Nafion in the catalyst layer to from a utilization standpoint is approximately 35–40 wt.%; this
calculate the layer’s effective ionic conductivity. The bulk ionic is similar to other results [35]:
conductivity is determined from the humidity and temperature
in the layer. The Bruggemann correction, which accounts for υPt = 1.833 × 10−5 yN
3
− 2.807 × 10−3 yN
2

porosity and tortuosity, determines the effective conductivity:


+ 1.332 × 10−1 yN − 1.476 (20)
σ= ε1.5
N σBulk (17)
The Bruggemann exponent of 1.5 was obtained from a study An initial increase in utilization with higher Nafion content
on the electrical conductivity of dispersions [39]. The Brugge- [41] could be a result of increased connectivity of the electrolyte.
mann correction has been employed in this form by many We hypothesize that the subsequent decline in utilization with
3854 S. Litster, N. Djilali / Electrochimica Acta 52 (2007) 3849–3862

Table 1 catalyst layer as an interface using the Butler–Volmer equation:


Catalyst layer properties and parameters    
αa F A αc F A
Property Value i = iA
0 exp η − exp − η (22)
RT RT
Air pressure, Pair 1 atm
Fuel pressure, Pf 1 atm
Fuel cell temperature, T 353 K (80 ◦ C)
2.2.5. Solution approach
Relative humidity 100% We solve the equations presented in the previous sec-
Membrane thickness (Nafion 115), WMem 125 ␮m tion without resorting to spatial numerical discretization. We
Catalyst layer void fraction, εV 55% [46] solve the one-dimensional MEA model with the bisection

py
Membrane conductivity, σ Mem, Bulk 6.94 S m−1 [46] algorithm. The function we solve is the difference between
Platinum/carbon ratio, yPt 20% Pt/C
Platinum surface area, sPt 1120 cm2 mg−1
the current calculated in the cathode and the current cal-
Platinum density, ρPt 21500 kg m−3 culated in the anode (f = iC (ηC ) − iA (ηA )). The independent
Nafion density, ρN 1900 kg m−3 variable is the cathode overpotential at the membrane inter-

co
Carbon density, ρC 2267 kg m−3 face (ηC ). The overpotential in the anode is determined from
Cathode transfer coefficient, α −1
1 [42,43] the expression ηA = (EOC − Ecell ) − ηC − WMem σMem iC (ηC ).
Anode transfer coefficient, αa and αc 1 [40]
O2 concentration, CO2 5 mol m−3
The sub-routine for solving the implicit finite-thickness cata-
H2 concentration, CH2 34.3 mol m−3 lyst layer equation employs the Newton–Raphson method. In
O2 concentration exponent, γO2 1.03 [42] this sub-routine the Newton–Rhapson function is the difference
H2 concentration exponent, γH2 0.5 [40] between the previous solution of the overpotential at the catalyst
O2 reference concentration, cO ref
113 mol m−3 [42] layer/GDL interface (x = WCL ), and that determined by Eq. (10);

al
2
H2 reference concentration, cH ref
40.88 mol m−3 [40]
2 f = η0 − η(WCL ).
Anode exchange current density, iA 0 6 × 103 A m−2 [26]
Gibbs energy of formation G◦ = −295800 − 33.5T ln T
+ 388.7T [28]
on
2.3. MEA model evaluation

To evaluate the present model, we use the experimental results


further increase in Nafion content results from reduced electrical of Sasikumar et al. [41] because of the special attention they
connectivity between the catalyst particles. paid to the catalyst layer specifications and the wide scope of
rs

We employ the exchange current density of the plat- their parametric study. In addition, Sasikumar et al. used oxygen
inum/Nafion interface and its temperature dependence as instead of air, which is better for validating the MEA model since
determined in the experimental work of Beattie et al. [42]. We this eliminates the ambiguities associated with mass transport
pe

take the interfacial exchange current density to vary exponen- limitations.


tially with temperature: Table 1 lists the operating conditions and fuel cell parameters.
The model operating conditions correspond to the experimen-
= 6.379 × 10−14 exp(6.782 × 10−2 T )
Pt/N
i0 (21)
tal ones [41]. A key parameter is the void fraction, which is
This temperature dependence, which is a consequence of the specified as 55% (based on the results of Navessin [46]). The
overall electron transfer process being an activated process, can molar density of oxygen at a pressure of 3 atm (113 mol m−3 )
be obtained in rather general form from transition state theory is used to correlate the exchange current density of Beattie et
r's

(e.g. Bockris et al. [44]). The parameters in this equation (pref- al. [42] to other gas pressures and concentration polarization. In
actor and activation energy) are treated as empirical parameters. lieu of a comprehensive mass transfer model, we set the oxy-
gen concentration in the catalyst layer to a reduced value of
2.2.4. Interface model of the anode catalyst layer 5 mol m−3 .
o

The faster kinetics of anode reaction in PEM fuel cells in con- In the following, the platinum loading varies between 0.10
junction with the low conductivity of the electrolyte drives the and 0.25 mg Pt cm−2 and the Nafion content is specified in the
th

anode reaction into a thin strip next to the membrane. Kornyshev range of 20–60 wt.%. Table 2 presents the resulting catalyst layer
and Kulikovsky’s [45] expression for the thickness of the reac- parameters for four points in this spectrum of specifications.
tive strip in a catalyst layer supports this assumption. Therefore, As discussed in the development of the model, a change in
Au

in contrast with the cathode, it is acceptable to model the anode platinum loading impacts only the catalyst layer thickness. In

Table 2
Catalyst layer properties resulting from the macrohomogeneous model
Configuration (Pt Volumetric exchange Catalyst layer Nafion volume fraction (%)
loading|Nafion content) current density (A m−3 ) thickness (␮m) (effective conductivity (S m−1 ))a

0.10 mg Pt cm−2 |40 wt.% 13355 6.9 25 (0.88)


0.25 mg Pt cm−2 |40 wt.% 13355 17.4 25 (0.88)
0.25 mg Pt cm−2 |20 wt.% 6825 13.7 12 (0.29)
0.25 mg Pt cm−2 |60 wt.% 6088 27.1 36 (1.57)
a Bulk conductivity = 6.94 S m−1 .
S. Litster, N. Djilali / Electrochimica Acta 52 (2007) 3849–3862 3855

tal observations. In the experimental data, increasing the Nafion


content from 20 to 40 wt.% significantly improves performance,
and a further increase from 40 to 60 wt.% marginally reduces
performance. The model captures this effect. Although the cat-
alyst layer conductivity is almost doubled at a Nafion content
of 60 wt.%, the fuel cell voltage drops due to reduced utiliza-
tion and a longer conduction distance. These results demonstrate
that the macrohomogeneous model with the variable-thickness
formulation can correlate the catalyst layer specifications to the

py
layer’s microstructure, and allows subsequent translation of the
microstructure to volume-average parameters for modelling.

co
3. Planar air-breathing fuel cells

3.1. Model

Having established the validity of the MEA model, we


Fig. 5. Comparison between polarization curves obtained experimentally by now expand this model to include additional phenomena for
Sasikumar et al. [41] and those calculated by the present model for platinum investigating air-breathing fuel cells. The additional phenomena
loadings of 0.10 and 0.25 mg Pt cm−2 .

al
include:

contrast, Table 2 shows that changing the Nafion content has a 1. The transport of oxygen and water vapour between the sur-
on
significant impact on all three parameters, and there is a maxi- face of the GDL and the ambient air.
mum exchange current density at approximately 40 wt.%. The 2. The diffusion of gases through the GDL.
decrease in exchange current density at higher Nafion contents 3. The diffusion of oxygen into the catalyst layer.
is due to the diminishing platinum concentration as the layer 4. The influence of temperature and humidity on the conduc-
rs

expands. tivity of the membrane.


Fig. 5 presents the polarization curves of two validation 5. The heat transfer from the fuel cell to the ambient air.
cases with platinum loadings of 0.10 and 0.25 mg Pt cm−2 ,
with a 40 wt.% Nafion content in both cases. There is excel-
pe

3.1.1. Assumptions
lent agreement with the experimental curves, especially for
In addition to the assumptions used in the derivation of the
0.10 mg Pt cm−2 , and the variable-thickness macrohomoge-
analytical MEA model, the following lists the major assumptions
neous model satisfactorily reproduces the influence of platinum
we use in developing the model.
loading.
Fig. 6 depicts the effect Nafion content on the polarization
curve. Again, the model is in good agreement with experimen- 1.Transport in the fuel cell is considered one-dimensional.
2.The fuel cell operates in steady state conditions.
r's

3.Water vapour is produced in the cathode.


4.The water exists only as vapour.
5.There is zero net transport of water through the membrane.
6.The hydrogen supply is “dead-ended” and is not recycled.
o

7.The hydrogen supply is dry (zero humidity).


8.Water does not accumulate in the hydrogen storage system.
th

9.The water activity is uniform across the electrolyte in the


membrane and catalyst layers and is equal to the water activ-
ity at the cathode interface of the gas diffusion layer and the
Au

catalyst layer.
10. Potential losses due to electron conduction are neglected.
11. The anode side of the fuel cell is perfectly insulated against
heat fluxes and heat only exits the fuel cell through the
cathode surface.

3.1.2. Gas diffusion


The flux of gases through the gas diffusion layer results from
Fig. 6. Comparison between polarization curves obtained experimentally by
the consumption of reactants and the supply of products in the
Sasikumar et al. [41] and those calculated by the present model for Nafion catalyst layer. Single-phase conditions are assumed to prevail.
contents of 20, 40 and 60 wt.% In the cathode, the consumption of oxygen is a function of the
3856 S. Litster, N. Djilali / Electrochimica Acta 52 (2007) 3849–3862

current density: In addition, we approximate the oxygen mass fraction in the


catalyst layer to capture the onset of reactant depletion, using
M O2 i
S O2 = − (23) an approach similar to Berg et al. [15]. For the purpose of
4F approximating this oxygen mass fraction, we assume a homo-
and the local source of water due to the reaction is: geneously distributed reaction rate through the catalyst layer,
MH2 O i which is reasonable in the intermediate regime when the cata-
S H2 O = (24) lyst layer thickness is close to optimal [30]. The resulting oxygen
2F
concentration is used to correct the reaction rate in the Tafel
The governing equation for gas transport in the GDL, assum- equation.

py
ing negligible advection, is Fick’s law. When expressed in terms
of a mass flux ṅA , the differential form of Fick’s law is expressed GDL/CL MO2 iw2
yO
CL
= yO2 − (32)
as: 2
32FρDCL

co
ṅA = −ρDA
eff
∇yA (25) 3.1.3. Water transport through the membrane
where DA
eff is the effective molecular diffusivity of species A in Since we consider steady state operation and a dead-ended
the porous media and yA is the mass fraction of species A. We supply of dry hydrogen, we assume zero net water transport
correct the density of air (ρ) for temperature with the ideal gas across the membrane. Kulikovsky [16] made a similar assump-
law. We also correct the diffusivity with the porosity ε and the tion. Moreover, this assumption is supported by the experimental
tortuosity factor τ, which accounts for the longer diffusion path results of Janssen and Overvelde [49] who report a very low
effective drag coefficient across the membrane (−0.01) at a

al
in the porous gas diffusion media (see [48]):
low current density (400 mA cm−2 ) with dry air and hydrogen
ε at low stoichiometries. We also assume that the water activity
DA
eff
= DA (26)
τ
on
is uniform across the membrane, which is justifiable for thin
The form of Fick’s law employed herein is typically used for membranes [32].
binary mixtures, and is justifiable given the low water concentra-
tions, and the similar molecular weights of oxygen and nitrogen. 3.1.4. Membrane conductivity
rs

We calculate the binary diffusivity as a function of temperature The potential drop through the electrolyte contributes signifi-
using the empirical method presented by Cussler [48]: cantly to the polarization of the fuel cell. We model the influence
√ of humidity on ionic conductivity with an empirical fit to the data
T 1.75 (1/MA ) + (1/MB ) presented by Sone et al. [47] for Nafion 117 without heat treat-
DAB =
pe

(27)
P 1/3
φA + φB
1/3
ment. The polynomial for ionic conductivity as a function of
activity at 303 K is:
where φA and φB represent molecular volumes [48]. The con-
vective mass transfer rates at the interface of the GDL with the σ303 K = 3.46a3 + 0.0161a2 + 1.45a − 0.175 (33)
ambient air is quantified in terms of mass transfer coefficients
(hO2 , hH2 O ), which are functions of the mass transfer Nusselt We model the influence of temperature with the expression
number (hA = NuDA /l). We calculate the interface mass fractions of Springer et al. [7]:
r's

by balancing the species production/consumption MA i/nA F with   


1 1
the convective mass transfer ρhA (yA 0 − yAmb ):
A σ = exp 1268 − σ303 K (34)
303 T
1 M O2 i
yO
0
= yO
Amb
− (28) However, the influence of temperature on ionic conductivity
o

2 2
ρhO2 4F is secondary in comparison to humidification. We also note that
1 M H2 O i almost all conductivity is lost below a relative humidity of 20%
th

yH
0
2O
= yH
Amb
2O
+ (29) when using the correlation in Eq. (33). We chose this correlation
ρhH2 O 2F
because the widely used model of Springer et al. [7] has poor
where i is the current density of the fuel cell and yA
Amb is the accuracy in low humidity conditions [50,51], which is a critical
Au

mass fraction in the ambient air. regime for air-breathing fuel cells.
Considering the GDL as a homogeneous and isotropic porous
medium, we derive the oxygen and water mass fractions at the 3.1.5. Heat transfer
GDL/CL
interface between the GDL and the catalyst layer (yA ) from Temperature assumes a critical role in the performance of
Eq. (26): PEM fuel cells because of its impact on relative humidity through
the saturation pressure and reaction kinetics. Temperature is also
GDL/CL WGDL MO2 i
yO2 = yO
0
− (30) significant for air-breathing fuel cells because they passively
2
ρDO
GDL 4F
2 dissipate the heat produced by the fuel cell without relying on
WGDL MH2 O i the cooling systems of larger PEMFC stacks. For a body with a
GDL/CL
yH2 O = yH
0
2O
+ (31) sufficiently uniform temperature distribution, a lumped system
ρDH
GDL 2F
2O analysis that considers variation of temperature with time only
S. Litster, N. Djilali / Electrochimica Acta 52 (2007) 3849–3862 3857

is adequate, and the characteristic response time of temperature Table 3


to changes in heat production is given by: Operating conditions for the planar fuel cell
Property Value
V ρc
τ= (35)
A h Heat/mass transfer length scale, l 1 cm
Air pressure, Pair 1 atm
where V and A are the volume and surface area of the body, ρ the Fuel pressure, Pf 1 atm
density of the body, c the specific heat, and h is the heat trans- Ambient air temperature, T∞ 293 K (20 ◦ C)
Ambient relative humidity 60%
fer coefficient. This shows that the heat transfer time constant
Nusselt number, Nu 10
varies proportionally with the characteristic length scale given

py
GDL thickness, WGDL 250 ␮m
by the volume to surface area ratio (V/A). Following standard GDL porosity, εGDL 0.45
heat transfer practice, a lumped system analysis is appropriate GDL tortuosity, τ GDL 2.5
when the Biot number (ratio of convective to conductive heat Platinum loading, mPt 0.25 mg cm−2
Nafion content, yN 45 wt.%
transfer) is small, i.e.

co
Catalyst layer void fraction, εV 55% [46]
hδ Platinum/carbon ratio, yPt 20% Pt/C
Bi = < 0.1 (36)
ksb
where δ is the conduction length scale and ksb is the solid body 3.1.6. Solution procedure
conductivity. Using conservative parameters [32], we estimate We solve for the current density at a given cell potential using
a Biot number of 0.03 for a 1 cm2 fuel cell, which meets the the classical bisection method. We prefer this robust method over

al
criterion for lumped system analysis. This assumption is further more rapidly converging approaches, such as Newton–Raphson,
supported by CFD modelling, which predicts minimal temper- because the system of equations is highly sensitive to tempera-
ature variation within such fuel cells [27]. ture, causing other methods to fail for some cases. The function
on
The heat generated in the fuel cell is a combination of the we solve is the difference between the bisection midpoint and the
heat of reaction and that due to the various irreversibilities. The current density calculated for the temperature and concentrations
heat of reaction is the product of the temperature and the change associated with the bisection midpoint.
in entropy (s). The sum of the irreversibilities of the fuel cell
rs

is equal to the difference between the reversible open-circuit


3.2. Results and discussion
potential and the cell potential (EOC − Ecell ). Thus, we compute
the overall heat production (qp ) from:
3.2.1. Baseline properties and parameters
pe


Table 3 lists the baseline parameters. The dimensions of
T (−s)
qp = + (EOC − Ecell ) i (37) the fuel cell are similar to that developed by Hahn et al. [5]
2F
(1 cm × 1 cm). We select an ambient air temperature of 293 K
We assume that all the heat generated by the fuel cell is (20 ◦ C) and an ambient relative humidity of 60% because of
rejected by either forced or free convection from the exterior their broad applicability. The prescribed Nusselt number of 10
surfaces of the fuel cell. The convective heat transfer to the is moderately higher than that feasible by natural convection
environment (qt ) is: alone to represent the effect of ambient air currents.
r's

AA
qt = h(T − T∞ ) (38) 3.2.2. Comparison with experiment
AS
We use the recent experimental results of Fabian et al. [6]
where AA is the electrochemically active planar area, AS the sur- to validate the model. The experimental MEA consists of a
o

face area, and T∞ is the temperature of the ambient environment. carbon cloth GDL, a catalyst layer with 1 mg Pt cm−2 , and a
For planar fuel cells with heat transfer from the cathode surface Nafion 112 membrane. We assume a typical value of 33% for
th

only, the ratio AA /AS is equal to one. Again following standard the Nafion content of the cathode catalyst layer [35]. All other
heat transfer practice, the convective heat transfer coefficient h parameters were given by Fabian et al. [6], with the exception of
is determined from an appropriate value of the Nusselt number the total heat transfer coefficient hAS , which we estimate to be
Au

(Nu): 0.0993 W K−1 from the experimental GDL temperature data. By


approximating the Nusselt number from the heat transfer coef-
Nu k
h= (39) ficient and estimated surface area, we can calculate the mass
l transfer coefficients for the GDL/air interface.
where l is the characteristic length of the system and k is the air’s Fig. 7 compares the theoretical and experimental polariza-
thermal conductivity. Herein, the characteristic length is that of tion curves. A limiting current density is evident in both set of
the shortest side of the fuel cell. We also employ this Nusselt curves, and although this presents like a reactant mass trans-
number to calculate the mass transfer coefficients at the cathode fer limitation, it is in fact the result of membrane dry-out.
surface in Eqs. (28) and (29). In this case, the same length scale This is demonstrated by the membrane conductivity data pre-
is used, but the thermal conductivity is replaced by the respective sented in Fig. 8. The Fabian et al. membrane conductivity data
diffusivities of the gases. was obtained with AC impedance spectroscopy. We estimate
3858 S. Litster, N. Djilali / Electrochimica Acta 52 (2007) 3849–3862

py
co
Fig. 7. Polarization curves obtained experimentally by Fabian et al. [6] and those Fig. 9. Fuel cell temperature vs. current density obtained experimentally by
predicted by the model for ambient temperatures of 20 and 30 ◦ C. Fabian et al. [6] and those predicted by the model for ambient temperatures of
20 and 30 ◦ C.

al
the membrane conductivity by subtracting the 11 m electrical
resistance [6] and accounting for the membrane thickness and the transport parameters, and the simplifications made in the
on
model.
the fuel cell area. The membrane dehydration phenomenon is
directly related to the rise in temperature associated with greater In noting the remarkable agreement between the experi-
current density, as shown in Fig. 9. The rate of temperature mental data and the model, it should be emphasized that the
increase outweighs water production, resulting in a decrease procedure for acquiring polarization data is critical. Fabian et
al. obtained each data point by allowing the membrane to equi-
rs

of the internal relative humidity and, consequently, a drop in


membrane conductivity. librate over a 2-h period. Long equilibration times are required
The above comparisons demonstrate the predictive capabili- to accurately depict membrane dehydration because the time
ties of the model without any a priori performance information, scales can be much larger than other relevant time scales in the
pe

fitting of electrochemical parameters or “tuning” of other model fuel cell. For example, Moxely et al. [52] show that the equili-
parameters. The model predicts quite satisfactorily the cur- bration of membrane water content can take anywhere from a
rent density that marks the transition to membrane dry-out few minutes to many hours.
and the accompanying rapid rise in fuel cell temperature. The
small deviations in predicting the exact current limitation lev- 3.2.3. Nusselt number
els is not surprising given the lack of precise data for some of We investigate the influence of the air currents surrounding
r's

the fuel cell by conducting a parametric study of the Nusselt


numbers. The variation of the Nusselt number with the magni-
tude of air current velocity can be determined from a number
of correlations for forced convection over various geometries
and flow directions [53]. We present results for a wide range of
o

Nusselt numbers to ensure broad applicability.


Fig. 10 presents the polarization curves for Nusselt numbers
th

between 10 and 40. It is evident that the rate at which heat


dissipates from the fuel cell can dramatically affect performance.
For a Nusselt number of 10 (approximating free convection) the
Au

current density is considerably restricted. An increase in the


Nusselt number from 10 to 40 increases the maximum current
density by 350%.
Although the limiting current density shown in the polariza-
tion curves of Fig. 10 resemble a reactant mass transfer limitation
pattern, the underlying process is in fact different as the change
in the oxygen concentration from the ambient air to the catalyst
layer is insignificant at these low current densities. At the max-
imum current of 520 mA cm−2 the oxygen concentration at the
Fig. 8. Membrane conductivity vs. current density obtained experimentally by
Fabian et al. [6] and those predicted by the model for ambient temperatures of catalyst layer/GDL interface was still 82% that of the ambient
20 and 30 ◦ C. air. Fifty percent of the oxygen depletion occurs in the cathode
S. Litster, N. Djilali / Electrochimica Acta 52 (2007) 3849–3862 3859

py
co
Fig. 10. Polarization curves for a range of Nusselt numbers. Fig. 12. Water vapour activity in the membrane vs. current density for a range
of Nusselt numbers.
electrode, the other 50% is depleted at the GDL/air interface.

al
Nevertheless, this reduction in concentration has a negligible 3.2.4. Ambient condition
effect on fuel cell performance. To quantify the impact of ambient air conditions, we vary
Fig. 11 depicts the influence of irreversibilities on heat pro- the temperatures from 283 K (10 ◦ C) to 303 K (30 ◦ C) and
on
duction. The increased Joule heating at the point where cell relative humidity from 30 to 90%. The resulting polarization
voltage drops sharply due to membrane dry-out is evidenced by curves, in Fig. 14, show that the highest current densities occur
the rapid rise in temperature near the end of each temperature when the ambient air is highly humid or relatively cold. The
curve. Fig. 12 presents the competing effects of water produc- improved performance at high ambient relative humidity of 90%
rs

tion and heat generation on the water activity in the membrane. increases performance directly through the membrane’s humid-
At low currents, and therefore low fuel cell temperatures, the ification requirement as in more conventional fuel cells. The
activity increases with current density because of water produc- more interesting and less intuitive result is the improved per-
pe

tion. However, the water activity decreases as the temperature formance at lower temperatures. At these lower temperatures,
rises with increasing current density. Fig. 13 serves as guideline self-humidification is enhanced because less water vapour is
for operating air-breathing fuel cells by demonstrating that for a required to saturate the air.
given heat transfer rate there is an optimal operating voltage for
which the membrane conductivity is maximized. The optimal 3.2.5. Platinum loading
voltage based on membrane conductivity ranges between 0.7 To investigate the potential for minimizing Platinum loading,
and 0.85 V and decreases with greater heat transfer coefficients. we consider loadings of 0.05, 0.25, 0.5, and 1.00 mg Pt cm−2 .
r's

As shown in Fig. 13, increasing the heat removal rate decreases Fig. 15 depicts the corresponding polarization curves. The effect
the voltage achievable without membrane dry-out.
o
th
Au

Fig. 13. Membrane conductivity vs. cell voltage over a range of fixed Nusselt
Fig. 11. Temperature vs. current density over a range of Nusselt numbers. numbers.
3860 S. Litster, N. Djilali / Electrochimica Acta 52 (2007) 3849–3862

py
co
Fig. 14. Polarization curves for five different ambient air conditions.
Fig. 16. Polarization curves for a range of Nafion contents in the cathode catalyst
layer.

al
of platinum loading is most pronounced at low current densities
where the activation of the oxygen reduction reaction is domi- 60 wt.%, there is little difference in performance. The optimum
nant. The limiting current does not change significantly because value is approximately 45 wt.%.
on
it is a result of membrane dry-out. The optimum platinum load-
ing is approximately 0.25 mg Pt cm−2 (based on diminishing 4. Conclusion
returns beyond this value). This value is in good agreement with
that found by Qi and Kaufman [54]. Fuel cells have the potential to supplant batteries in portable
rs

electronic devices because of increased cost tolerances for


3.2.6. Nafion content fuel cells at this scale, their rapid recharges, and their higher
Since ionic resistance is a major component of polarization energy densities. In this paper we presented a model of pla-
nar air-breathing PEM fuel cells for portable devices. The key
pe

in ambient air-breathing fuel cells, we expect in general an opti-


mum Nafion content for the catalyst layer [35]. We analyzed the component of our model is a 1-D semi-analytical model of the
effect of varying Nafion content from 20 to 60 wt.% as shown in membrane electrode assembly (MEA). The model incorporates
Fig. 16. The polarization curve for a Nafion content of 20 wt.% an improved membrane model correlation that is more suitable
indicates a significant drop in platinum utilization and ionic for representing membrane dry-out. We performed parametric
conductivity in the cathode catalyst layer. The reduced ionic con- studies to identify the influence of key design specifications and
ductivity in the catalyst layer is evidenced by the earlier onset operating conditions on the performance of air-breathing fuel
cells. The insights we obtain from the analysis can supplement
r's

of the membrane dry-out. For Nafion contents between 33 and


and guide the design of air-breathing fuel cells. The analysis
revealed some trends that are quite different from those observed
in fuel cells with conditioned reactant streams. Of particular
note is the elucidation of the significant effect of heat trans-
o

fer on membrane dry-out in ambient air-breathing fuel cells.


The study suggests that improved performance of air-breathing
th

fuel cells can be achieved by increasing the heat removal


rate and thus promoting higher relative humidity levels in
the GDL.
Au

Acknowledgements

The authors gratefully acknowledge Tibor Fabian of Stanford


University for providing experimental data for the validation
of the model. Funding for this research was provided by the
Natural Sciences and Engineering Research Council of Canada
(NSERC) in the form of a Canada Graduate Scholarship and
a supplement from the Canadian Space Agency to SL. This
Fig. 15. Polarization curves for a range of platinum loadings in the cathode research was also funded in part by an NSERC Discovery Grant
catalyst layer. to ND and the Canada Research Chairs Program.
S. Litster, N. Djilali / Electrochimica Acta 52 (2007) 3849–3862 3861

Appendix A. Nomenclature CL catalyst layer


eff effective value
GDL/CL GDL and catalyst layer interface
a activity H2 hydrogen
AA active area (m2 ) H2 O water
AS planar surface area (m2 ) Mem membrane
Bi Biot number N Nafion
c specific heat of solid; also concentration (J kg−1 K−1 ; O2 oxygen
mol m−3 ) OC open-circuit

py
Cp specific heat of gas (J kg−1 K−1 ) Pt platinum
D diffusion coefficient (m2 s−1 ) ref reference value
E potential
F Faraday’s constant (96485 C mol−1 )

co
References
G Gibbs energy of formation (J mol−1 )
h heat transfer coefficient (W m−2 K−1 ) [1] C.K. Dyer, J. Power Sources 106 (2002) 31.
hA mass transfer coefficient (kg m−2 s−1 ) [2] H. Kariatsumari, H. Yomogita, Nikkei Electronics Asia (February 2005).
i local current density (A m−2 ) [3] J.S. Wainright, R.F. Savinell, C.C. Liu, M. Litt, Electrochim. Acta 48 (2003)
j0 volumetric exchange current density (A m−3 ) 2869.
k thermal conductivity (W m−1 K−1 ) [4] A. Schmitz, S. Wagner, R. Hahn, H. Uzun, C. Hebling, J. Power Sources
127 (2004) 197.

al
l convection length scale (m) [5] R. Hahn, S. Wagner, A. Schmitz, H. Reichl, J. Power Sources 131 (2004)
m specific mass loading (kg m−2 ) 73.
M molar mass (kg mol−1 ) [6] T. Fabian, J.D. Posner, R. O’Hayre, S.-W. Cha, J.K. Eaton, F.B. Prinz, J.G.
specific mass flux (kg m−2 s−1 )
on
ṅ Santiago, J. Power Sources 161 (2006) 168.
[7] T.E. Springer, T.A. Zawodzinski, S. Gottesfeld, J. Electrochem. Soc. 138
Nu Nusselt number
(1991) 2334.
P pressure (Pa) [8] D.M. Bernardi, M.W. Verbrugge, J. Electrochem. Soc. 139 (1992) 2477.
qp flux of heat produced (W m−2 K−1 ) [9] T.F. Fuller, J. Newman, J. Electrochem. Soc. 140 (1993) 1218.
flux of heat transferred (W m−2 K− )1
rs

qt [10] T.V. Nguyen, R.E. White, J. Electrochem. Soc. 140 (1993) 2178.
R universal gas constant (8.3145 J mol−1 K−1 ) [11] J.C. Amphlett, R.M. Baumert, B.A. Peppley, P.R. Roberge, J. Electrochem.
sPt specific platinum surface area (m2 kg−1 ) Soc. 142 (1995) 1.
[12] J.S. Yi, T.V. Nguyen, J. Electrochem. Soc. 145 (1998) 1149.
T temperature (K)
pe

[13] V. Gurau, F. Barbir, H. Liu, J. Electrochem. Soc. 147 (2000) 2468.


V volume (m3 ) [14] R. Bradean, K. Promislow, B. Wetton, Num. Heat Transfer, A 42 (2002)
WCL catalyst layer thickness (m) 121.
y species mass fraction [15] P. Berg, K. Promislow, J. Pierre St., J. Stumper, J. Electrochem. Soc. 151
(2004) A341.
yN Nafion content
[16] A.A. Kulikovsky, Electrochem. Comm. 6 (2004) 969.
yPt platinum/carbon ratio [17] T. Berning, N. Djilali, J. Power Sources 124 (2003) 440.
[18] D. Natarajan, T.V. Nguyen, J. Power Sources 115 (2003) 66.
r's

Greek letters [19] P.T. Nguyen, T. Berning, N. Djilali, J. Power Sources 130 (2004) 149.
α transfer coefficient [20] S. Shimpalee, S. Greenway, D. Spruckler, J.W. Van Zee, J. Power Sources
135 (2004) 79.
δ heat transfer length scale (m)
[21] B. Sivertsen, N. Djilali, J. Power Sources 141 (2005) 65.
Δs entropy change (J mol−1 K−1 ) [22] S. Um, C.Y. Wang, J. Power Sources 125 (2004) 40.
ε porosity
o

[23] L. Wang, H. Liu, J. Power Sources 134 (2004) 185.


φ electrolyte phase potential (V) [24] T. Hottinen, M. Mikkola, P. Lund, J. Power Sources 129 (2004) 68.
γ concentration exponent [25] C. Ziegler, A. Schmitz, M. Tranitz, E. Fontes, J.O. Schumacher, J. Elec-
th

trochem. Soc. 151 (2004) A2028.


η overpotential (V)
[26] T. Berning, D.M. Lu, N. Djilali, J. Power Sources 106 (2002) 284.
ρ density (kg m−3 ) [27] S. Litster, J.G. Pharoah, G. McLean, N. Djilali, J. Power Sources 156 (2006)
σ conductivity (S m−1 ) 334.
Au

τ tortuosity factor [28] A. Parthasarathy, S. Srinivasan, A.J. Appleby, C.R. Martin, J. Electrochem.
υPt platinum utilization Soc. 139 (1992) 2856.
[29] C.Y. Du, P.F. Shi, X.Q. Cheng, G.P. Yin, Electrochem. Comm. 6 (2004)
435.
Subscripts/superscripts [30] M. Eikerling, A.A. Kornyshev, J. Electroanal. Chem. 453 (1998) 89.
a anodic [31] M.L. Perry, J. Newman, E.J. Cairns, J. Electrochem. Soc. 145 (1998) 5.
A anode [32] S. Litster, Mathematical modelling of fuel cells for portable devices,
air air M.A.Sc. Thesis, University of Victoria, 2005.
[33] L. Pisani, M. Valentini, G. Murgia, J. Electrochem. Soc. 150 (2003) A1549.
Amb ambient
[34] Q. Wang, M. Eikerling, D. Song, Z. Liu, T. Navessin, Z. Xie, S. Holdcroft,
c cathodic J. Electrochem. Soc. 151 (2004) A950.
C cathode [35] S. Litster, G. McLean, J. Power Sources 130 (2003) 61.
cell fuel cell [36] C. Marr, X. Li, J. Power Sources 77 (1999) 17.
3862 S. Litster, N. Djilali / Electrochimica Acta 52 (2007) 3849–3862

[37] W. Sun, B.A. Peppley, K. Karan, Electrochim. Acta 50 (2005) 3359. [47] Y. Sone, P. Ekdunge, D. Simonsson, J. Electrochem. Soc. 143 (1996)
[38] H. Inoue, H. Daiguji, E. Hihara, JSME Int. J. Ser. B 47 (2004) 228. 1254.
[39] R.E. De La Rue, C.W. Tobias, J. Electrochem. Soc. 106 (1959) 827. [48] E.L. Cussler, Diffusion–Mass Transfer in Fluid Systems, Cambridge Uni-
[40] H. Ju, C.Y. Wang, J. Electrochem. Soc. 151 (2004) A1954. versity Press, New York, NY, 1997.
[41] G. Sasikumar, J.W. Ihm, H. Ryu, Electrochim. Acta 50 (2004) 598. [49] G.J.M. Janssen, M.L.J. Overvelde, J. Power Sources 101 (2001) 117.
[42] P.D. Beattie, V.I. Basura, S. Holdcroft, J. Electroanal. Chem. 468 (1999) [50] J. Fimrite, B. Carnes, H. Struchtrup, N. Djilali, J. Electrochem. Soc. 152
180. (2005) A1815.
[43] L. Zhang, C. Ma, S. Mukerjee, J. Electroanal. Chem. 568 (2004) 273. [51] T. Thampan, S. Malhotra, H. Tang, R. Datta, J. Electrochem. Soc. 147
[44] J.O.M. Bockris, A.K.N. Reddy, M. Gamboa-Aldeco, Modern Electrochem- (2000) 3242.
istry 2A: Fundamentals of Electronics, Kluwer/Plenum, New York, 1998. [52] J.F. Moxley, S. Tulyani, J.B. Benziger, Chem. Eng. Sci. 58 (2003) 4705.

py
[45] A.A. Kornyshev, A.A. Kulikovsky, Electrochim. Acta 46 (2001) 4389. [53] N.V. Suryanarayana, Engineering Heat Transfer, West Publishing Com-
[46] T. Navessin, S. Holdcroft, Q. Wang, D. Song, Z. Liu, M. Eikerling, J. pany, 1995.
Horsfall, K. Lovell, J. Electroanal. Chem. 567 (2004) 111. [54] Z.G. Qi, A. Kaufman, J. Power Sources 113 (2003) 37.

co
al
on
rs
pe
o r's
th
Au

View publication stats

You might also like