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

Electrochimica Acta 53 (2008) 7630–7637

Contents lists available at ScienceDirect

Electrochimica Acta
journal homepage: www.elsevier.com/locate/electacta

Numerical and microfluidic pore networks: Towards designs for directed


water transport in GDLs
A. Bazylak, V. Berejnov, B. Markicevic, D. Sinton, N. Djilali ∗
Institute for Integrated Energy Systems and Department of Mechanical Engineering, University of Victoria, Victoria, BC, Canada V8W 3P6

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

Article history: Based on a pore network representation of porous media, numerical and experimental methods are
Received 10 January 2008 employed to explore the design of gas diffusion media in order to control and direct water transport.
Received in revised form 28 March 2008 Randomized pore networks with structured biasing in both diagonal and radial directions show a marked
Accepted 29 March 2008
influence on liquid water transport through the two-dimensional media. Radial biasing is shown to have
Available online 11 April 2008
the beneficial effect of decreasing the network saturation, which is a desirable quality in fuel cell gas dif-
fusion layers. Similar flow patterns are obtained from both numerical simulations and experiments, and
Keywords:
it is found that relatively small radial biasing can yield a 43% decrease in average saturation compared to
Gas diffusion layer
Porous media
pore networks without a prescribed radial gradient.
Two-phase flow © 2008 Elsevier Ltd. All rights reserved.
Microfluidic
Polymer electrolyte membrane fuel cell
Biased networks

1. Introduction Since the early development of pore network modelling [10–13],


models with spatially homogeneous pores and throats have been
In polymer electrolyte membrane (PEM) fuel cells, the gas dif- employed to study the displacement processes in idealized porous
fusion layer (GDL) provides pathways for gaseous fuel to reach the media [14]. Over the past two decades, network models have
catalyst sites, for electrons generated at the catalyst sites to reach been applied to a multitude of areas ranging from oil recov-
the current collector, and very importantly, for excess liquid water ery [15–20], ground water hydrology [21], textile characterization
to exit the system. Conventional GDLs are ineffective for water [22], etc. Blunt [23] provides a broad review of recent pore net-
removal and organized water transport due in part to its arbitrary work model developments. The pore network modelling approach
structure [1], and excess water can reduces transport of reactant has been very recently extended to the gas diffusion layer by
gases [2] as well as exacerbate degradation [3]. Significant improve- Gostick et al. [24], Markicevic et al. [25] and Sinha and Wang
ments in the PEM fuel cell are contingent on advancements in water [26]. Marckicevic et al. [25] used network simulations to inves-
management, particularly in the GDL. Recently, Litster et al. [4] and tigate and establish the functional dependence of the relative
Bazylak et al. [5,6] investigated liquid water transport through PEM permeability and capillary pressure on liquid water saturation,
fuel cell GDLs at the microscale level by employing fluorescence two of the key parameters required for two-phase PEM fuel cell
microscopy. New methods for enhanced water transport have been models [2]. Recently, Chapuis et al. [27] employed pore net-
demonstrated via the active removal of water through the use of works composed of randomly distributed disks to simulate the
transport plates [7] and electro-osmotic pumping [8,9]. One poten- anisotropic arrangement of GDL fibers, and they found good
tial avenue for improving water management is the development agreement between their numerical and experimental flow pat-
of GDL materials engineered for improved water management at terns.
the microscale. Isichenko [28] introduced correlation as a useful modelling tech-
nique for porous media, whereby the liquid occupation probability
of pores and throats are considered. The method of correlation
[29,30] has been used to study the large-scale effects of two-phase
flow in porous media [31–34]. These correlated or long-range cor-
∗ Corresponding author at: Institute for Integrated Energy Systems and Depart-
related porous media differ from those employed in this work.
ment of Mechanical Engineering, University of Victoria, PO Box 3055, Victoria, BC,
Canada V8W 3P6. Tel.: +1 250 721 6034; fax: +1 250 721 6323. Here, pore network obstacles are arranged in a biased manner with
E-mail address: ndjilali@uvic.ca (N. Djilali). respect to the entire occupied volume. To distinguish our networks

0013-4686/$ – see front matter © 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.electacta.2008.03.078
A. Bazylak et al. / Electrochimica Acta 53 (2008) 7630–7637 7631

from the correlated networks, we refer to our networks as biased 2.1. Pore network modelling
networks.
Experimental validation has been provided for theoretical pore Pore network models provide an alternative to continuum phase
network models [35–45]. Recently, Berejnov et al. [46] presented models to elucidate transport phenomena and determine trans-
a lab-on-chip approach to study two-phase transport in porous port parameters in porous media. Continuum modelling of porous
media, employing microfluidic rapid prototyping fabrication meth- media becomes impractical when the porous media has a com-
ods in conjunction with fluorescence microscopy to study water plex three-dimensional heterogeneous and disordered geometry,
transport in pore networks. Experimental network results have also such as the gas diffusion layer shown in Fig. 1(c). Numerically track-
been used in other porous media applications to guide and validate ing phase interfaces and calculating surface tension forces becomes
theoretical modelling [47,48] or as inputs to existing models [49]. computationally expensive. In contrast, pore network modelling is
Work to date in the area of pore network and experimental an intermediate approach where the disordered porous media is
modelling of porous media has focused on characterizing existing represented by an equivalent network of pores and throats, and the
media. Our goal is to demonstrate a new approach to design- throats are assigned varying hydraulic conductivities.
ing porous networks for directing water transport by applying The invasion percolation algorithm with trapping [14] was used
long-scale correlations (biasing) to an initially homogeneously ran- to investigate deterministic realizations of pore network filling. The
dom 2D porous network. This approach enables us to qualitatively two-dimensional model accounts for capillary dominated drainage
demonstrate the potential use of biasing for directing water trans- and cluster formation in a regular square capillary network of size
port, which is of specific interest for PEMFC gas diffusion layers. nL × nL pores. The porous media is represented by a matrix of
This novel technique is demonstrated with both numerical and pores and connecting throats. The invading phase (water) enters
experimental networks that clearly illustrate the effect of biasing the network along the top side (inlet), and advances through the
on directing water transport. network in a series of quasi-static steps by occupying available
throats having the largest throat radius. The fluid exits through
2. Methodology the opposite side (outlet) to the entrance, and the invasion pro-
cess is stopped once breakthrough occurs, where breakthrough to
Pore network modelling provided a method to evaluate the liq- the state when the first throat at the outlet is filled with the invad-
uid water transport through the gas diffusion layer in a PEM fuel ing fluid. In [25], this model was applied to determine multiphase
cell, as indicated in Fig. 1. In this work, pore networks with pore size parameter such as relative permeability and capillary pressure and
characteristics of the gas diffusion layer [50] were designed with a their dependence on the invading phase saturation. Since on any
software package developed in-house, and the invasion percolation given network, each simulation is deterministic, the analysis in [25]
with trapping algorithm developed by Markicevic et al. [25] was required averaging of simulations performed with several stochas-
used to evaluate the drainage process. Once a photomask design tically generated networks having the same global properties. It
was chosen, the photomask was generated and printed with a high should be noted that the two-dimensional pore networks consid-
resolution printer. The photomask was employed during the pho- ered here have limitations as pointed out by Chatzis and Dullien
tolithographic microfabrication process, where a microfluidic pore [51], as well as by Chapuis et al. [27] in their very recent GDL
network was generated. The spread of the water into the microflu- study. In particular 2D and 3D simulations yield different percola-
idic network was monitored using fluorescence microscopy, and tion thresholds for a given coordination number. Notwithstanding
flow patterns and time series evolution of liquid water transport these limitations, the 2D network systems are expected to be quali-
through the network were recorded. The pore network design tatively representative and allow acceptable simulation turnaround
process is illustrated in Fig. 2. The experimental breakthrough pat- times to explore the concept of biasing for the purpose of control-
tern were then compared to pore network simulations based on ling water transport. The deterministic nature of the pore network
Wilkinson and Willemsen’s algorithm for invasion percolation with invasion process provides is a useful feature allowing compari-
trapping [14,25]. son with experimental results. Fig. 1 shows a sample numerical

Fig. 1. Schematic showing pore network modelling results with respect to PEMFC geometry: (a) diagram of a PEMFC, (b) sample numerical pore network breakthrough
pattern of a simulated gas diffusion layer, and (c) scanning electron microscope image of Toray TGP-H-060 GDL (length bar represents 200 ␮m).
7632 A. Bazylak et al. / Electrochimica Acta 53 (2008) 7630–7637

breakthrough pattern in relation to a fuel cell schematic. For fur-


ther details on the pore network model, readers are directed to
Markicevic et al. [25].

2.2. Microfluidic pore networks

Using Matlab 7.0, a pore network was designed, and the


coordinates were then imported into Autocad 2007, where the
photomasks were drafted. All photomasks were then printed by
CAD/Art Services, Inc. (California, USA), with a maximum resolution
of 20,000 dpi. The microfabrication process begun with pattern-
ing a photoresist resin layer on a silicon substrate. A thin layer of
SU8-25 photoresist (25 ␮m thick) was distributed over the silicon
substrate through the use of a spin coater. A 25 ␮m height was
chosen for the experimental pore network fabrication due to the
manufacturing constraints in the laboratory. The photoresist layer

Fig. 2. Flow diagram illustrating the pore network design and fabrication procedure.
Fig. 3. Schematic illustrating various methods to construct a microfluidic pore net-
work: (a) randomly perturbing a regular lattice of square obstacles, (b) sequentially
placing square obstacles in random locations, and (c) sequentially placing square
obstacles of varying sizes in random locations.
A. Bazylak et al. / Electrochimica Acta 53 (2008) 7630–7637 7633

was then soft baked to evaporate solvent and condense the photore- 3. Pore network design
sist. The photoresist layer was exposed to a near-UV light source
(350–400 nm), which allowed for photoresist cross-linkage. The There are several possible approaches to biasing a network.
substrates were baked to complete the photoresist cross-linking In the present study, two biasing functions were applied to the
reaction. Any SU8-25 photoresist substrate that was not exposed pore (throat) distribution: a random function and a determinis-
to UV light was removed with chemical etching. The pore network tic function. From an engineering point of view, introducing the
was then fabricated by curing a polymer on the patterned template, deterministic function allows networks that are intrinsically ran-
removing the polymer from the template and sealing the template dom to be imparted with new properties for the porous medium.
to a planar surface. Through this rapid prototyping technique, sev- The random function defines the local random properties of the
eral microfluidic pore networks were generated and tested within network, while the deterministic function defines the deviation of
a short period of time. The reader is directed to recent work by these properties on the scale of the entire network. Generating a
Berejnov et al. [46] for specific microfabrication details. Typically, network using the combination of these functions yields a random
in PEM fuel cells the capillary number, Ca , is on the order of 10−8 network with a biased pattern. If the dominance of the random
for two-phase flow in the GDL. In this work, the employed flow and deterministic functions are comparable, then the preferential
rates can be characterized by Ca on the order of 10−7 to 10−8 , which pathway for water cannot be determined a priori, but rather the
reasonably represents the flow conditions in a GDL of an operating pattern needs to be determined according to the network laws. This
PEMFC. approach of combining deterministic and random biasing functions

Fig. 4. Perturbation procedure for generating photomask: (a) regular square lattice, (b) translation of regular lattice in both the horizontal and vertical directions, (c)
uniform distribution of horizontal and vertical translations, (d) resulting arrangement of random pore network, (e) throat width distribution, and (f) hydraulic throat radius
distribution.
7634 A. Bazylak et al. / Electrochimica Acta 53 (2008) 7630–7637

also provides a systematic way of characterizing and modelling a


network/porous medium under realistic conditions. For instance
when considering a purely random network under some deforma-
tion, such as expansion or compression, the characteristics of the
medium may be related to a deterministic function of a similar
nature.

3.1. Pore network structure

A random pore network can be generated in a number of ways.


Starting with a blank two-dimensional domain, as shown in Fig. 3,
square obstacles can be arranged randomly to fill the domain, while
overlapping is prohibited. The spaces that are not covered by these
square obstacles will constitute the throats and pores. In Fig. 3(a),
the network is generated by randomly perturbing a regular lattice
of square obstacles. In Fig. 3(b), the domain is filled by adding one
square obstacle at a time, until the maximum number of obstacles
is reached. Fig. 3(c) shows a randomly arranged network, which is
generated by first filling the network with square obstacles (obsta-
cle 1). Once the maximum number of obstacles is placed in the
domain, the remaining space is filled as tightly as possible with-
out over lapping with obstacles that are half the size as obstacle 1
(obstacle 2). Again, once the maximum number of obstacles of type
2 are placed, smaller type obstacles can be used to fill the domain.
As shown in Fig. 3, the benefit of utilizing method (a) is that a higher
packing density can be obtained, while maintaining a constant con-
nectivity of 4. In method (b), the porosity of the media is very high
and difficult to control. In method (c), the obstacle size variation
provides a high packing density, but at the cost of varying connec-
tivity and network complexity. All pore networks in this work were
generated with the first method (Fig. 3(a)), whereby a regular lat-
tice of square obstacles is randomly perturbed to generate a random
network.

3.2. Randomization

The pore network generation procedure begins with a regular


lattice of square obstacles. Fig. 4(a) shows a schematic of the regu-
lar lattice, where w is the throat width, and L is the obstacle square
side length. A random translational perturbation is then applied
to each obstacle, where X refers to the horizontal translation,
and Y refers to the vertical translation of the original obstacle.
The uniform random distribution of translational perturbations is
shown in Fig. 4(c). Fig. 4(d) is the resulting schematic of a portion
of the random pore network. The resulting throat width distribu-
tion is shown in Fig. 4(e). For a uniform channel height of 25 ␮m, the
hydraulic throat radii distribution is shown in Fig. 4(f). All hydraulic
throat radii employed in this work are characteristic of Toray TGP-
H-060 gas diffusion layers [50]. The hydraulic radius is calculated
with the following formula: rh = wt/(w + t), where w is the throat
width, and t is the throat thickness.

3.2.1. Isotropic vs. diagonal biasing


To generate an isotropic random pore network, the transla-
tion of the regular lattice in the horizontal direction X and the
vertical direction Y must be independent. It is also possible to
generate a random pore network with a diagonal bias by coupling
the translation of the regular lattice in the horizontal and vertical
directions. To implement a diagonal bias of bias = 45◦ , for each
square obstacle, Y = X. The ratio of vertical to horizontal obsta-
cle translations can be adjusted to implement the desired diagonal
bias for 0◦ < bias < 90◦ , as shown in
Fig. 5. Procedure for generating photomask for a pore network with a prescribed
Y radial gradient: (a) regular square lattice, (b) radial translation of individual square
tan bias = (1) obstacles, (c) network with radial gradient, and (d) random perturbations applied
X to the network with a radial gradient.
A. Bazylak et al. / Electrochimica Acta 53 (2008) 7630–7637 7635

3.2.2. Radial gradients square obstacle from the central coordinate axis is labeled r. The
The procedure to generate a pore network with a radial gradient square obstacle is then translated in the radial direction to a new
is illustrated in Fig. 5. The first step is to begin with a regular lat- radial distance, R, from the origin, as shown in Fig. 5(b). The new
tice of square obstacles, as shown in Fig. 5(a). The distance of each radial distance, R, is transformed by the following expression:

Fig. 6. Comparison of numerical and experimental random pore network flow patterns at breakthrough: (a) isotropic perturbations, (b) diagonally biased perturbations, (c)
radial gradient, and (d) checkerboard pattern with multiple radial gradient patches. Each numerical and experimental result comparison is accompanied by a photomask
schematic with a reduced number of obstacles for clarity (along left column).
7636 A. Bazylak et al. / Electrochimica Acta 53 (2008) 7630–7637

˛r 2 A larger pore network with several patches with independent


R→r+ (2)
rmax radial gradients is shown in Fig. 6(d). The individual patches have
been highlighted with gray and black to indicate their size and loca-
where, ˛ is the coefficient that is employed to vary the radial gradi- tion. The percolation breakthrough pattern associated with that
ent of the network. The radial gradient, G, is the increase in throat checkerboard pattern illustrates the preferential pathways created
width per obstacle, beginning from the central coordinate axis. from the application of radial gradients.
After a regular network (with radial distance, r) is transformed For all the trials shown in Fig. 6, the experimental results agree
into a new network (with radial distance, R), the radial gradient, well with the numerical results, especially when considering that
G, was numerically measured. Numerical tests yield the following the dynamic effects and contact angle hysteresis that exist in the
expression between G, ˛, rmax , and N: microfluidic pore networks were not considered in the model.
The agreement between the experimental and numerical results
4rmax
G= ײ (3) is especially highlighted by the trials shown in Fig. 6(c) and (d),
N2
where the percolation pathways leading to breakthrough occur in
Fig. 5(c) shows the resulting pore network design when a radial the same place. The small deviations at the small (nodal)-scale level
gradient has been applied to the regular lattice of square obstacles. between experiments and simulations are expected and inherent
The final step is to apply a random translation with independent to the fabrication technique which results in micro-feature vari-
horizontal (X) and vertical (Y ) perturbations to each obsta- ations between individual microchip channels. It is interesting to
cle. An example of the resulting pore network pattern is shown note that when comparing the numerical and experimental flow
in Fig. 5(d). To quantify the effect of an applied radial gradient on a patterns of deterministic network designs, the overall water-air
random network, the radial gradient is normalized with respect to flow pattern is of more importance than the exact saturation val-
the average throat width of a network without a radial gradient. ues. For example, the biased networks shown in Fig. 6 exhibit very
The network saturation is calculated to further quantify the distinct flow structures that correspond well between the numer-
effect of liquid transport in a network with a radial gradient. The ical and experimental results, while their specific saturation value
saturation of the pore network is defined as the number of filled could correspond to any number of arbitrary flow patterns.
throats divided by the total number of throats in the system, as Of interest from a GDL design perspective is the relationship
described in between the degree of geometric perturbation of the network and
its influence on water transport. In an operating fuel cell, it is
Nfilled
s= (4) desirable to optimize the access of reactants to the catalyst layer.
Ntotal Reactant transport takes place via a combination of convection and
diffusion. From the view point of convective transport, no defi-
4. Results and discussion nite relationship between relative permeability and saturation has
yet been established for GDL media, but it is generally accepted
Numerical and experimental pore networks with identical net- that the functional relationship is of the form krg = (1 − s)n , i.e.
work designs were invaded with liquid water. Comparison of higher relative permeabilities are obtained at lower saturations.
the flow patterns and associated characteristics of the numeri- The dependence of saturation on increasing radial gradient was
cal and experimental networks, which have nominally identical studied for a 100 × 100 pore network, as shown in Fig. 7. We
microstructures, confirms the applicability of the pore network note that the patterns obtained with the larger network are iden-
rules and dominant mechanisms in the processes considered. tical to those in Fig. 6, which were performed with a 48 × 48
Fig. 6 shows the numerically and experimentally obtained pat- pore network. Fig. 7 shows that as the normalized radial gradient
terns at breakthrough for 48 × 48 pore networks with (a) isotropic increases, the saturation of the network decreases. This is attributed
perturbations, (b) diagonally biased perturbations, (c) a radial gra- to the increasing localization and channeling of the invading phase
dient, and (d) multiple radial gradients. For each network design clearly shown by the results. The saturation of individual deter-
in Fig. 6 shown from left to right are: a scaled down version of
the photomask, the numerical breakthrough and the experimen-
tal breakthrough patterns. By comparing the percolation patterns
shown in Fig. 6(a) and (b), the effect of a diagonal bias on the pore
network compared to the isotropically random pore network is
evident, even though the difference cannot be detected when com-
paring their respective photomasks alone. Fig. 6(c) illustrates the
effect of a radial gradient on the liquid water transport. With a pos-
itive radial gradient, larger throat sizes are located around the outer
edges of the pore network. As expected in a hydrophobic medium,
the liquid penetrates the throats with the lowest threshold capillary
pressure, which are the largest throats. With an applied radial gra-
dient, the liquid saturation is zero at the center of the network and
increases with radial distance from the center. With this radial gra-
dient design, liquid water percolates along the outer left and right
sides of the network, until breakthrough is achieved. The radial gra-
dient effectively provides directed water transport along the sides
of the media. The percolation thresholds for 2D and 3D systems
are expected to be different in general, and though the results pre- Fig. 7. Dependence of saturation on increasing radial gradient for a 100 × 100
sented here should be interpreted with caution in the context of numerical pore network. The radial gradient is normalized by the average throat
width (30 ␮m). Filled squares represent single realizations, where empty squares
an actual GDL, they are expected to provide the right trends and
represent an average value over 100 stochastic realizations with the standard devia-
be qualitatively useful for guiding the design of novel media for tion marked with error bars. The shaded region of the plot shows the expected range
directed water transport. of saturation values.
A. Bazylak et al. / Electrochimica Acta 53 (2008) 7630–7637 7637

ministic simulations is plotted and shown with solid black squares. [9] S. Litster, C. Buie, T. Fabian, J. Eaton, J. Santiago, Journal of the Electrochemical
Since pore network modelling is stochastic by nature, 100 ran- Society 154 (10) (2007) B1049.
[10] I. Fatt, The network model of porous media I. Capillary pressure characteristics,
dom simulations were performed for each of four points (shown Petroleum Transactions, the American Institute of Mining, Metallurgical, and
with non-filled black squares), with the average value and standard Petroleum Engineers (AIME) 207 (1956) 144.
deviation marked with error bars. The largest standard devia- [11] I. Fatt, The American Institute of Mining, Metallurgical, and Petroleum Engi-
neers (AIME) 207 (1956) 160.
tion corresponds to pore networks without a radial gradient, and [12] I. Fatt, The American Institute of Mining, Metallurgical, and Petroleum Engi-
decreases with increasing radial gradient. With a radial gradient neers (AIME) 207 (1956) 164.
of 0.016/obstacle (i.e. the throat width increases by 1.6%/obstacle [13] S. Broadbent, J. Hammersley, Proceedings of the Cambridge Philosophical Soci-
ety 53 (1957) 629.
in the radial direction), the average saturation decreased by 43% [14] D. Wilkinson, J. Willemsen, Journal of Physics A: Mathematical and General 16
compared to the average pore network without a radial gradient. (1983) 3365.
[15] R. Chandler, J. Koplik, K. Lerman, J. Willemsen, Journal of Fluid Mechanics 119
(1982) 249.
5. Conclusions [16] A. Payatakes, Annual Review of Fluid Mechanics 14 (1982) 365.
[17] J. Koplik, T. Lasseter, Society of Petroleum Engineers Journal 25 (1) (1985) 89.
Numerical and experimental pore networks were employed to [18] A.B. Dixit, S. McDougall, K. Sorbie, J. Buckley, SPE Reservoir Evaluation and
Engineering 2 (1) (1999) 25.
explore the design of gas diffusion media for directed water trans-
[19] M. Valavanides, A. Payatakes, Advances in Water Resources 24 (2001) 385.
port through the use of pore network modelling and microfluidic [20] L. Paterson, Physical Review E 66 (2002) 1.
fabrication techniques. The concept of biasing was shown to be [21] M. Bernadiner, Transport in Porous Media 30 (1998) 251.
effective from the 2D numerical simulations and planar experi- [22] K. Thompson, American Institute of Chemical Engineers Journal 48 (7) (2002)
1369.
mental networks. These 2D results can guide more representative [23] M. Blunt, Current Opinion in Colloid & Interface Science 6 (2001) 197.
and complex 3D network simulations that need to be performed [24] J. Gostick, M. Ioannidis, M. Fowler, M. Pritzker, Journal of Power Sources 173 (1)
in the future for a full quantification of the patterns and transport (2007) 277.
[25] B. Markicevic, A. Bazylak, N. Djilali, Journal of Power Sources 171 (2) (2007) 706.
processes. [26] P. Sinha, C.-Y. Wang, Electrochimica Acta 52 (2007) 7936.
Randomized pore networks with diagonal and radial biasing [27] O. Chapuis, M. Prat, M. Quintard, E. Chane-Kane, O. Guillot, N. Mayer, Journal of
show a marked influence on liquid water transport through the Power Sources 178 (2008) 258.
[28] M.B. Isichenko, Review of Modern Physics 64 (4) (1992) 961.
two-dimensional media. Unidirectional water transport is achieved [29] M. Sahimi, Journal De Physique I France 4 (9) (1994) 1263.
through the use of diagonally biased networks. Radial biasing is [30] S. Cordero, I. Kornhauser, A. Domı́nguez, C. Felipe, F. Esparza, R.H. López, A.M.
shown to have the beneficial effect of decreasing the network sat- Vidales, J.L. Riccardo, G. Zgrablich, Particle & Particle Systems Characteristics 21
(2004) 101.
uration, which is a desirable quality in fuel cell gas diffusion layers. [31] A. Vidales, R. Faccio, J. Riccardo, E. Miranda, G. Zgrablich, Physica A 218 (1995)
It was found that with a radial gradient of 1.6% per obstacle, the 19.
average saturation decreased by 43% compared to the average pore [32] M. Sahimi, S. Mukhopadhyay, Physical Review E 54 (4) (1996) 3870.
[33] M. Knackstedt, A. Sheppard, M. Sahimi, Advances in Water Resources 24 (2001)
network without a radial gradient. Both the numerical and exper-
257.
imental pore network liquid breakthrough patterns were in good [34] R. López, A. Vidales, A.D. Ortiz, G. Zgrablich, Colloids and Surfaces A: Physico-
agreement. The results from this paper indicate that porous media chemical and Engineering Aspects 300 (2007) 122.
can be designed to have these directed water transport features, [35] R. Lenormand, C. Zarcone, Physical Review Letters 54 (20) (1985) 2226.
[36] R. Lenormand, E. Touboul, C. Zarcone, Journal of Fluid Mechanics 189 (1988)
which can be applied to PEM GDL designs. 165.
[37] R. Lenormand, C. Zarcone, Transport in Porous Media 4 (1989) 599.
Acknowledgments [38] A. Birovljev, L. Furuberg, F. Feder, T. Jossang, A. Maloy, A. Aharony, Physical
Review Letters 67 (5) (1991) 584.
[39] O. Vizika, D. Avraam, A. Payatakes, Journal of Colloid and Interface Science 165
The authors are grateful for the financial support of the Natural (1994) 386.
Sciences and Engineering Research Council (NSERC) of Canada, the [40] B. Xu, Y. Yortsos, Physical Review E 57 (1) (1998) 739.
[41] C. Tsakiroglou, M. Theodoropoulou, V. Karoutsos, D. Papanicolaou, V. Sygouni,
Canada Research Chairs Program, and the National Research Council Journal of Colloid Interface Science 267 (2003) 217.
(NRC). The authors would also like to acknowledge the support of [42] E. Kossel, M. Weber, R. Kimmich, Solid State Nuclear Magnetic Resonance 25
the Canadian Foundation for Innovation (CFI). (2004) 28.
[43] C. Perrin, P. Tardy, K. Sorbie, J. Crawshaw, Journal of Colloid and Interface Science
295 (2006) 542.
References [44] M. van Dijke, K. Sorbie, A. Danesh, SPE Journal 9 (1) (2004) 57.
[45] M. van Dijke, K. Sorbie, M. Sohrabi, A. Danesh, Journal of Petroleum Science and
[1] K. Jiao, B. Zhou, Journal of Power Sources 169 (2007) 296. Engineering 52 (2006) 71.
[2] T. Berning, N. Djilali, Journal of the Electrochemical Society 150 (12) (2003) [46] V. Berejnov, N. Djilali, D. Sinton, Lab Chip 8 (2008) 689.
A1589. [47] R. Lenormand, C. Zarcone, Proceedings of the 59th Annual Technical Conference
[3] A. Taniguchi, T. Akita, K. Yasuda, Y. Miyazaki, Journal of Power Sources 130 and Exhibition of SPE AIME, SPE, Houston, Texas, 1984, p. 13264.
(2004) 42. [48] U. Oxaal, M. Murat, F. Bojer, A. Aharony, J. Feder, T. Jossang, Nature 329 (1987)
[4] S. Litster, D. Sinton, N. Djilali, Journal of Power Sources 154 (2006) 95. 32.
[5] A. Bazylak, D. Sinton, Z.-S. Liu, N. Djilali, Journal of Power Sources 163 (2007) [49] M. Theodoropoulou, V. Sygouni, V. Karoutsos, C. Tsakiroglou, International Jour-
784. nal of Multiphase Flow 31 (2005) 1155.
[6] A. Bazylak, D. Sinton, N. Djilali, Journal of Power Sources 176 (1) (2007) 240. [50] M. Mathias, J. Roth, J. Fleming, W. Lehnert, Handbook of Fuel Cells, John Wiley
[7] A. Weber, R. Darling, Journal of Power Sources 168 (2007) 191. & Sons, Ltd., New York, 2003, pp. 1–21.
[8] C. Buie, J. Posner, T. Fabian, S.-W. Cha, D. Kim, F. Prinz, J. Eaton, J. Santiago, Journal [51] I. Chatzis, F. Dullien, Canadian Journal of Petroleum Technology (January–
of Power Sources 161 (2006) 191. March, 1977) 97.

You might also like