i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 Available online at www.sciencedirect.com ScienceDirect journal homepage: www.elsevier.com/locate/he Review Two-phase flow in GDL and reactant channels of a proton exchange membrane fuel cell Satish G. Kandlikar a,b,*, Evan J. See a, Mustafa Koz b, Preethi Gopalan b, Rupak Banerjee a,b a b Mechanical Engineering, Rochester Institute of Technology, 76 Lomb Memorial Dr., Rochester, NY 14623, USA Microsystems Engineering, Rochester Institute of Technology, 168 Lomb Memorial Dr., Rochester, NY 14623, USA article info abstract Article history: Understanding the effect of two-phase flow in the components of proton exchange Received 30 August 2013 membrane fuel cells (PEMFCs) is crucial to water management and subsequently to their Received in revised form performance. The local water saturation in the gas diffusion layer (GDL) and reactant 27 January 2014 channels influences the hydration of the membrane which has a direct effect on the PEMFC Accepted 7 February 2014 performance. Mass transport resistance includes contributions from both the GDL and Available online 7 March 2014 reactant channels, as well as the interface between the aforementioned components. Dropletechannel wall interaction, water area coverage ratio on the GDL, oxygen transport Keywords: resistance at the GDLechannel interface, and two-phase pressure drop in the channels are Two-phase flow interlinked. This study explores each factor individually and presents a comprehensive Fuel cells perspective on our current understanding of the two-phase transport characteristics in the Mass transport resistance PEMFC reactant channels. Pressure drop Copyright ª 2014, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights Flow patterns reserved. Breakthrough pressure 1. Introduction Fuel cells are considered as viable contenders as “engines” for automotive applications with almost all major manufacturers being close to the commercialization stage. As further improvements in efficiency and durability are considered, water management remains one of the outstanding areas for research. The two-phase flow issues are encountered in gas diffusion layer (GDL) and the reactant channels. The objectives for an effective water management scheme are: (i) provide adequate saturation for membrane hydration; (ii) allow for efficient passage of water and reactants through the GDL; (iii) remove water efficiently from the reactant channels. Inadequately addressing these issues may lead to higher mass transport resistances, local damage to the membrane, degradation of the fuel cell components, and loss of efficiency. This paper addresses the progress made worldwide, with specific focus on the work conducted in the authors’ laboratory in the last seven years. * Corresponding author. Mechanical Engineering, Rochester Institute of Technology, 76 Lomb Memorial Dr., Rochester, NY 14623, USA. Tel.: þ1 (585) 475 6728; fax: þ1 (585) 475 6879. E-mail addresses: [email protected], [email protected] (S.G. Kandlikar), [email protected] (E.J. See), [email protected] (M. Koz), [email protected] (P. Gopalan), [email protected] (R. Banerjee). http://dx.doi.org/10.1016/j.ijhydene.2014.02.045 0360-3199/Copyright ª 2014, Hydrogen Energy Publications, LLC. Published by Elsevier Ltd. All rights reserved. i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 Nomenclature C c DO2 eAir dh F H hM i m Nu P Pr Ru r Re Sc Sh T t u x, y, z V W W We the Chisholm parameter oxygen molar concentration oxygen diffusivity in the air hydraulic diameter Faraday’s constant channel height mass transfer coefficient current density fraction of product water transported to cathode side Nusselt number pressure Prandtl number universal gas constant droplet radius Reynolds number Schmidt number Sherwood number temperature time superficial velocity spatial coordinates voltage mass flux channel width Weber number CL DTC GDL MPL PEMFC RH USDOE catalyst layer down-the-channel gas diffusion layer microporous layer proton exchange membrane fuel cell relative humidity The United States Department of Energy Greek a b m r s q half of the wedge corner angle volumetric quality dynamic viscosity mass density surface tension contact angle 6621 Subscripts An anode B base C capillary Ca cathode F film flow regime FD fully developed G gas k element number L liquid S slug flow regime TP two-phase flow W channel wall Abbreviations Ch air channel A description of the PEMFC operation may be found in a textbook, e.g. Ref. [1]. It consists of a proton exchange membrane which allows hydrogen protons to cross freely across it, while acting as an insulator for electrons and preventing reactant cross-over. Hydrogen and air (source of oxygen) are distributed evenly with a network of reactant channels and porous GDL over the catalyst layers (CL) on either sides of the membrane. On the hydrogen side (anode), hydrogen ions and free electrons are formed at the CL. The hydrogen ions are transported from the anode CL through the membrane to the cathode CL where the reaction is completed with oxygen from the cathode flow field and electrons returning from the external load. The product water formed at the cathode CL is the primary topic of interest in the water management studies. 2. Water transport and emergence from GDL 2.1. Transport through the GDLs The study of liquid water accumulation and two-phase flow in PEMFC reactant channels have received increased attention in the recent years [2e7]. Flow in the fuel cell reactant channels differs from conventional microchannels as the fuel cell channels are bounded on one side by a porous wall which is the GDL [8e11]. The GDL typically consists of a macroporous carbon paper coated with a thin microporous layer (MPL) on one side. Water generated at the CL travels in liquid and/or gaseous form through the GDL to the reactant channels, where it is removed from the cell by the flow of reactants. This leads to continuous introduction of water along the length of the reactant flow channels. 2.1.1. Liquid water transport Liquid water has been observed to emerge into reactant channels and its presence in the GDL is well established [12e14]. A key area of research is the transport of water from the CL to the reactant channels through the MPL and the GDL. A number of researchers have proposed different mechanisms for transport from the CL to the reactant channels [4,15e20]. It has been shown by multiple investigators that the MPL improves the water management and the fuel cell performance [21,22]. The role of the MPL is still not fully understood, although the cracks appearing on its surface seem to play an important role in the vapor transport [15,23]. However, this is not yet conclusively established as the product water in the vapor form may occur through the MPL, while the condensed water in the GDL is prevented by the MPL from going back to the CL. In 2010, Lu et al. [24] designed and conducted an experimental investigation into the role of the MPL through the use of breakthrough pressure and visualization of droplet emergence. They sandwiched the GDL between two transparent 6622 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 Lexan plates with a design similar to a commercial bipolar plate [25]. The water was distributed at the MPL surface through channels providing a more uniform water introduction closely mimicking the water generation at the cathode catalyst layer. Liquid water flow rate was controlled using a syringe pump. The rate of water introduction was mapped to an equivalent current density of 1.2 A cm2. As the water breakthrough occurs from the GDL into the channels, large pressure spikes were recorded from the supply chamber in ex situ experiments. This agrees with previous observations made in literature [13,26]. Eruptive transport of liquid water emerging from the GDL into gas channels was also observed by earlier researchers [12,14]. The pressure peaks, seen in Fig. 1, correspond to the breakthrough pressures. The GDL samples are identical except for the MPL coating (SGL 25BC has an MPL while SGL 25BA does not). These GDLs have randomly distributed 5% PTFE content. When the GDL samples without MPL were tested, multiple breakthrough locations were found, corresponding to the different pressure peaks, which showed that the water emergence location from the GDL dynamically changed. This contradicted the predictions of pore network models which proposed that water Fig. 1 e Water breakthrough pressure (Pc) through a) SGL 25BA and b) SGL 25BC samples. Inset images show the visual observation corresponding to the breakthrough pressure readings. Adapted from Lu et al. [24]. breakthrough would occur from the same location. Further research on the PTFE content, its distribution, and MPL on the breakthrough behavior is warranted. Lu et al. [24] repeated the experiments for samples with the MPL. Although multiple breakthrough locations were observed, the locations did not change once established. This was contrary to the observations without the MPL and the stabilization was attributed to the presence of the MPL. The observations were explained in terms of Haines jumps, a wellestablished phenomenon seen in soil mechanics, as well as spontaneous redistribution of the water pathways within the GDL pores. 2.1.2. Water vapor (gaseous) transport Owejan et al. [15] stipulated that the majority of the water transport through the GDL is in the form of vapor, especially at elevated temperatures. The primary mode of transport is diffusion. However, the permeability of the GDL has also been shown to affect the performance of the fuel cell [27], and therefore deserves to be investigated further. Different models have been developed for diffusion in porous media in the past [28e32]. However, experimental results of diffusion in fibrous porous media such as the GDL do not follow the model predictions [30,33]. Although different authors have worked at understanding different factors that affect the effective diffusion coefficient of different gases through the GDL, none of the current models presents a good match to the experimental results. A definitive set of the experimental results as well as a diffusive model were long warranted. LaManna and Kandlikar [34] provided an exhaustive experimental study on the water vapor diffusion through the GDL materials used in PEM assemblies. Samples from three different manufacturers were utilized to test for the effects of the MPL, GDL thickness and PTFE loading. The results showed that the MPL coatings provided a significantly larger resistance to the diffusion of gases through the GDL which was attributed to the smaller pores of the MPL and lower overall porosity. Increased PTFE loading also reduced the effective diffusion coefficient, an outcome that is also due to the reduction in porosity and pore diameter. As the GDL is a heterogeneous multilayer structure, it was suggested that variation in the thickness of the GDL with similar manufacturing methods may have an effect on the effective diffusivity. However, the thickness showed a negligible effect on the effective diffusion coefficient. Permeability of a GDL represents the proportionality between the flow velocity through the GDL and the pressure drop across this layer. This parameter characterizes the ease of gas flow though the porous material. The higher the permeability is, the easier it is for the reactants and product water vapor to flow through. Numerous authors have studied the gas permeability of the GDL layers [35e40] both experimentally and numerically. Some of the studies include the effects due to Knudsen diffusion and convective transport processes (viscous and inertial transport) on permeability [41,42]. It has been shown that permeability increases with porosity and decreases with compression [40,43]. Addition of PTFE leads to a decrease in the permeability in all directions i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 [39,40] and the MPL coating decreases the through-plane permeability [35] but it does not affect the in-plane permeability [40]. Recently, Sadeghifar et al. reported that the fiber spacing also affects the transport properties of a GDL [44]. The fiber spacing has been shown to affect the air permeability and pressure drop within GDL, and further research in this area is needed to study its effect on water transport. 2.2. Channeledroplet interaction The water emergence into the reactant channel of a PEMFC can be in the liquid phase. This liquid water from the channel is removed by the gas flowing in the channel. However, due to the rough GDL surface structure, the droplets tend to get pinned on the surface and interact with the channel wall. Thus, understanding the fundamental problem of droplet detachment from the GDL surface and its interaction with the channel sidewall are critical to efficiently remove the liquid water from the channels. There have been several approaches used to visualize the liquid water in the PEMFCs to understand its behavior and removal pattern from reactant channels [12,25,45e54]. However, these techniques, such as neutron radiography, magnetic resonance imaging, and nuclear magnetic resonance are quite expensive and further restricted by their inability to easily resolve single droplets. One of the most efficient methods for single droplet level resolution is through optically transparent reactant channel materials. Bazylak et al. used an optically transparent reactant channel to visualize the droplet breakthrough and growth [13]. Similar experiments were performed by Yang et al. and they found that the droplet emerged at preferential pores and grew to a consistent size before getting detached and removed from the GDL [55]. Tuber et al. were among the first few groups to study the effect of GDL wettability using transparent PEMFCs. They expressed the water accumulation in the channel in terms of the contact angle [52]. Turhan et al. found that hydrophobic channel walls tend to form discrete droplets on the GDL surface and remove the droplets more easily [56]. However, in case of a hydrophilic surface, the liquid tends to form a film on the channel walls, a case that leads to a difficulty in water removal. Lu et al. also performed similar experiments and found that the hydrophilic channels help in uniform water distribution along the GDL surface and reduce the pressure drop in the channels [58]. Cheah et al. performed an experimental study on the effect of channel wettability, geometry and orientation on slug removal [59]. They showed that a hydrophilic square channel produced large slugs. These slugs required lower energy to be removed but they created large mass transport resistances for 6623 the reactant gases to diffuse to the catalyst layer. On the other hand, hydrophobic square channel produced smaller slugs which required higher energy for their removal but minimized the reactant transport resistance. Therefore, the reactant channel needs to be configured by taking all the studied factors into account. Extensive work has been performed by Kandlikar et al. to understand the basis for water accumulation in reactant channels using a visually accessible fuel cell [2,24,25,57,58]. They brought out the importance of superficial air velocity and water generation rate on the water hold up in the reactant channels. Apart from these and other experimental studies, there have been several papers in the literature from the recent past addressing the droplet behavior in the reactant channels numerically. Theodorakakos et al. analyzed the droplet behavior in a high aspect ratio channel geometry to avoid dropletewall interaction [60]. They correlated the critical diameter of the droplet at detachment with the superficial air velocity in the reactant channel. Similarly, Zhu et al. performed a 3D simulation of a reactant channel with different surface wettability conditions [61]. They found that a hydrophilic surface leads to the spreading of the liquid near the channel wall, forming a film and resulting in the blockage of the reactant channel. They also performed simulations to understand the droplet growth and detachment process as a function of channel geometry [62]. The droplet removal time was lower for triangular and trapezoidal channels compared to those in rectangular and upside-down trapezoidal channels. Likewise, Cai et al. numerically investigated the effect of surface wettability of the GDL on the behavior of a droplet and film [63]. They concluded that using the combination of a hydrophobic GDL and hydrophilic side channel walls leads to an efficient water removal from the channels. On the same note, Quan et al. also reported that a hydrophilic channel tends to push the water along the channel edges and facilitates the reactant gas transport to the catalyst layer [64]. The electric double layer interactions and their effect on the contact angles of droplets has been reported by Das and Mitra [66]. Liquid water flow in PEMFC reactant channels leads to different flow patterns depending on the gas and water flow rates. The flow patterns identified in the PEMFC channels are water droplet, film and slug flow as shown in Fig. 2 and discussed further in Section 4.1. The analysis of water droplet emergence, growth, interaction with the channel wall, formation of films or slugs, and removal of the water features from the channel as a function of gas flow velocity are important to understand the two-phase flow mechanism in the PEMFC channels. Fig. 2 e Flow patterns observed from the PEMFC gas channel cross section: a) Droplet, b) Slug, c) Film. Image credit [65]. 6624 2.2.1. i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 Fundamental study on dropletewall interaction There has been very little work done in the literature to understand the droplet detachment from a reactant channel from a fundamental perspective. Chang et al. studied the wetting phenomenon of a droplet in the vicinity of the corner boundary [67]. They found that the droplet follows the Young’s equation when the contact line of the droplet is away from the edge. Once the contact line reaches the edge, the droplet pinning occurs until the apparent contact angle reaches the critical angle which extends the droplet over to the new surface. Rath and Kandlikar studied the droplet formation and its accumulation that takes place in the corner of a channel with a microscopic visualization [68]. They varied the angle between the base and the sidewall (2a) to observe the droplet interaction with the wall. The effect of surface tension forces on the dropletechannel wall interaction was the main focus of this parametric study. The ConcuseFinn condition [69,70] was used to explain the behavior of the droplet near the channel corner. The ConcuseFinn condition dictates that the rise height of a fluid in a wedge shaped domain is predictable if the following condition is true: a þ q > p/2. Rath and Kandlikar Fig. 3 e a) Contact angles made by the droplet near the channel corner; b) Graphical representation of the condition for corner filling of a droplet in a channel corner. Adapted from Gopalan and Kandlikar [67]. used this formulation to predict the water filling the channel corner where the GDL and the channel wall meet [68]. Using the graphical representation in Fig. 3, they reported that the channel corner will be filled only if the contact angle data point falls inside the shaded region. They also found that the use of contact angles instantaneously measured during the water movement is the key in understanding the droplet behavior in the reactant channel as opposed to the static contact angle measurements. These instantaneously measured advancing and receding contact angles are called “instantaneous dynamic contact angles”. They showed that 2a has a transitional value at which the dropletecorner interaction switches from non-filling to filling for a given material pair. Rath and Kandlikar had studied the droplet behavior in a channel without air flow. Gopalan and Kandlikar extended the previous work to understand the droplet interaction with reactant channel walls in the presence of air flow [72]. They showed that the ConcuseFinn condition is not valid in the 2a range close to the transition angle for corner filling in presence of the air flow. It was shown that the air flow in the channel causes oscillations in the droplet which changes the dynamic contact angle and affects the droplet filling behavior near the transition angle as shown in Fig. 4. Polverino et al. also showed that the oscillations in the droplet due to the air flow affect the droplet detachment from the GDL [73]. Gopalan and Kandlikar also investigated the effect of droplet emergence location on the droplet dynamics. It was found that the droplet would fill the channel corner irrespective of any channel open angle when the droplet emerges from under the land area [71]. Gopalan and Kandlikar proposed to use grooves that are in the channel sidewall and directed towards the top channel wall to avoid the water clogging near the channel corners. Grooves strengthen the capillary forces on liquid water. Hence, the water at the corner fills in and moves to the top of the channel where the liquid water could be removed easily by the air flow [74]. However, the water accumulation near the corners is one of the issues that is still persistent in the fuel cells and needs to be addressed further. Gopalan and Kandlikar also highlighted in one of their studies that the droplet dynamics inside the gas channel is dependent on the material properties that are used for the reactant channels [75,76]. They concluded that the contact angle hysteresis (defined as the difference between the advancing and receding contact angles) of the channel wall material plays a major role in the channel corner filling behavior. Higher contact angle hysteresis of the channel wall material leads to a stronger pinning of the droplet. This consequently can allow larger oscillations to occur in the droplet without being de-pinned from the sidewall. This can prevent the contact line of the droplet from moving towards the channel corner to fill it. However, the droplet removal from the channel required a higher drag force from the air flow due to the stronger pinning of the droplet. They also found that the different GDL materials they used did not have any significant impact on the droplet dynamics inside the reactant channel. Gopalan and Kandlikar also performed similar experiments to analyze the effect of GDL deterioration due to i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 extended PEMFC operation [76]. MRC-105 GDL samples that had run for 40 and 125 h were utilized for further droplet emergence studies. The authors reported that the wettability of the GDL material decreases with the duration of operation and hence, the transition point for the corner filling of that material also decreases. Hence, the water droplet fills the channel corner at a lower open angle, 2a. 3. Oxygen interfacial transport from the air channel to GDL For accurate performance prediction of PEMFCs at high current densities, the oxygen (O2) transport resistance at the cathode side needs to be characterized. Interfacial O2 transport resistance at the GDLeair channel interface is a nonnegligible component of the total transport resistance. The two-phase flow in the channel complicates the prediction of the interfacial O2 transport resistance in two ways: 1) The GDLechannel interface is blocked by water features, such as droplets, films, and slugs. This blockage varies with location and time. 2) Water features disrupt the fully developed (FD) air velocity and O2 concentration profiles. The developing velocity and O2 concentration profiles in the wake of the water features lead to an interfacial resistance different than the fully developed value in the channel. The understanding of liquid water coverage on the GDLechannel interface enables more accurate prediction of the interfacial O2 resistance. By using the experimentally obtained water feature geometries and positions in the channel, the interfacial O2 resistance can be numerically simulated downstream of a particular water feature or set of features. 3.1. Liquid water coverage of the GDL interface As liquid water is produced and enters into air channels, the water blocks the area available for the oxygen to diffuse through the GDL to reaction sites at the catalyst layer. The ability to accurately visualize and quantify this coverage is the key to understanding the interfacial transport. A variety of imaging techniques have been experimentally used over the years to obtain a spatial distribution of water within reactant channels. Although the difficulty of gaining 6625 access to a neutron source discourages its widespread application, neutron imaging has been an effective choice [25,47,77e79]. Several groups experimented with X-rays to observe the liquid water distribution within the reactant channels [12,80e82]. X-rays have been used to obtain either a high spatial resolution or a high temporal resolution, but not both simultaneously. Direct optical imaging technique is the most advanced option in terms of spatial and temporal resolution. However, as traditional fuel cell materials are opaque in the visible spectrum, a fuel cell needs modification to allow optical visualization [78,82]. This technique is well suited to observe liquid water in the reactant channels after the modifications. Tuber et al. [52] was the first to use optical visualization in 2003 by modifying an operating PEMFC to visualize two-phase flow in the channels. Other researchers have also utilized this diagnostic technique to understand the behavior of liquid water in the channels [2,24,25,58,83e85]. Sergi and Kandlikar [86] developed a dual visualization setup to visualize the anode and cathode sides of an operating PEMFC simultaneously. A temporal resolution of 60 frames per second (higher resolution was possible, but was not warranted) was recorded along with a spatial resolution of approximately 34 mm per pixel. They defined the term area coverage ratio (ACR) which is defined as the fraction of interfacial area covered by liquid water present in the reactant channels to the total area available for gas transport in dry reactant channels. Sergi and Kandlikar [86] also developed and implemented an algorithm using MATLAB to process the image sequences captured during testing. A flow chart of the algorithm is shown in Fig. 5. The algorithm processes the test videos on a frame-by-frame basis. The test frame is first converted to grayscale and subtracted from the dry reference frame void of any water features. This subtraction highlights the water regions within the test frame. A mask is then applied to isolate the channel regions only for further processing. A threshold is applied to detect the water features based on the pixel intensities and reduces the optical noise. In the final step, morphological processing is utilized to remove the noise and smooth the detection of the water features. The morphological processes check for connected pixels and apply closing operations in order to fine tune the water feature detection. For visual inspection and spatial context of the water features, the final processed frame is added to the original video frame to create a superimposed detection area over the test frame. Fig. 4 e Droplet behavior near the transition angle (Open angle: 2a [ 50 ) for MRC-105 GDL base and polycarbonate sidewall for the superficial air velocities: a) 0.4 m sL1 e corner filling, b) 1.6 m sL1 e no corner filling [71]. 6626 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 Additionally, a second algorithm was developed to distinguish the water features present, in terms of slug and film flow regions. This provided supplementary information regarding different flow structures dominant within reactant channels at different current densities and stoichiometric ratios. The data reported in their work focused on a specific operating condition: cell temperature of 35 C and 100% inlet RH for both gas streams. 3.2. Interfacial oxygen transport resistance Interfacial oxygen transport resistance is the proportionality between the O2 molar flux and the required O2 concentration drop at the GDLechannel interface. By minimizing the concentration drop across the interface, a higher O2 concentration can be supplied through the GDL to the reaction sites, thus the performance of a PEMFC can be enhanced. The proportionality between the concentration drop and molar flux is expressed by the mass transfer coefficient (hM). This coefficient is non-dimensionalized by the channel hydraulic diameter (dh) and O2 diffusivity in the air ðDO2 eAir Þ, and the Sherwood number ðSh ¼ hM dh =DO2 eAir Þ is obtained. The interfacial resistance is reported by the Sherwood number which is specific to a flow and boundary condition, and channel aspect ratio. In the literature, simplified PEMFC models utilize the Sherwood number as a constant input to save computational resources from the simulation of forced convection in the air channels [87e98]. The constant input of Sherwood number is a valid assumption if the flow is fully developed in the reactant channels, such as the absence of liquid water features in high temperature PEMFCs [97]. However, the constant Sherwood number assumption is not always applicable. Kim et al. shared a similar anticipation that the Sherwood number might depend on specific fuel cell operating conditions and configurations [90]. Since the research on the Sherwood number in PEMFC reactant channels has not yet been very well established, 2D approaches (parallel plates) in the available literature still provided insight for the direction of the 3D studies. The results by the 2D models do not reflect realistic conditions due to fact that PEMFC reactant channels typically have an aspect ratio close to unity. However, these earlier numerical studies in 2D domains provided the following knowledge: The porous GDLechannel interface is not necessarily subjected to zero air velocity in the normal direction to the interface. The cell consumes and produces known amounts of oxygen and water respectively at a given current density. Depending on the thermal conditions of the cell, water leaves the cell in the form of liquid and vapor with a varying fraction. If the entire water flow is in the liquid form, the GDLechannel interface is subjected to air suction into the GDL. In case water is completely in the vapor phase, air injection into the channel takes place. Depending on the intensity of suction or injection, the flow may remain developing. Under developing flow conditions, the Sherwood number can vary and hence, has been a subject of investigation. 2D numerical studies by Wang et al. [99], Jeng et al. [100] and Hassanzadeh et al. [101] showed that air injection and suction have a weak or negligible effect on the Sherwood number. The aforementioned authors obtained the fully developed Sherwood number values as 5.274 (readjusted Sh with dh instead of the channel height) [99], 6.0 [100] and 5.411 [101]. These findings allow the simulations to neglect any injection or suction at the GDLechannel interface. However, these results cannot reflect the Sherwood number in 3D channels due to their 2D assumption. Moreover, they neglected the effect of liquid water present in the channel on the Sherwood number. Casalegno et al. also underlined the need for Sherwood number in two-phase flow conditions [95]. The first attempt to simulate the Sherwood number in a 3D channel which incorporated a water droplet was performed by Koz and Kandlikar [102]. The simulation domain was formed by placing a droplet-shaped obstruction into the channel. With this approach, the two-phase flow in the channel was simulated as single-phase flow. The channel cross section dimensions (W H ¼ 0.70 mm 0.40 mm) were taken from the design that meets the USDOE performance Fig. 5 e Graphical flow chart representation of the image processing algorithm developed by Sergi et al. [86]. i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 targets for automotive applications [25]. The droplet obstruction radius (r) varied between 0.10 and 0.20 mm. The superficial air flow velocity (uG) equivalent to the range of current density 0.1e1.5 A cm2 (at T ¼ 80 C and RH ¼ 100%) varied from 1.59 to 15.89 m s1. The fully developed Sherwood number was obtained to be 3.36 which was shown to significantly differ from the values provided by the 2D numerical studies: 5.274 [99], 6.0 [100] and 5.411 [101]. By placing a droplet in the fully developed region, the Sherwood number in the droplet wake was investigated. Fig. 6 shows the effect of superficial air velocity on the Sherwood number while the droplet radius was 0.15 mm. It was shown that there is a threshold value of air velocity to obtain an overshoot of Sherwood number in the droplet wake. The threshold air velocity was found to be dependent on the droplet size. The overshoot of Sherwood number in the droplet wake can be increased with superficial air velocity and droplet size. However, the air velocity and droplet size are limited by the droplet adhesion to the GDLechannel interface since the drag on the droplet increases with the aforementioned two parameters. The numerically obtained drag force on the dropletshaped obstruction was compared against the experimentally obtained force to initiate droplet sliding on an inclined surface by Das et al. [104]. Certain sets of the air velocity and droplet radius were removed from the results as their drag force exceeded the adhesion force. The distance between two droplets adhered to the GDLechannel interface can be as low as 1.00 mm which was visualized in an operative PEMFC by Yang et al. [55] and Zhang et al. [105]. The overshoot of Sherwood number in the droplet wake was shown to affect lengths which can easily cover the distance between two droplets (up to w18 mm). Hence, Koz and Kandlikar posed the possibility that droplets in a row along the flow direction may lead to an increase of Sherwood number. Moreover, this increase can be intensified in multiple stages since the visualizations by Yang et al. and Zhang et al. showed that number of droplets in a row can exceed two [55,105]. Fig. 6 e The effect of superficial air velocity (uG) on the Sherwood number (Sh) for the droplet radius r [ 0.15 mm. Adapted from Koz and Kandlikar [103]. 6627 Koz and Kandlikar extended their earlier work to multiple droplets in a row [103]. The higher resolution simulations in this recent study updated their previous fully developed Sherwood number from 3.36 [102] to 3.349. They also provided an additional aspect of validity for the use of droplet-shaped obstructions by considering the droplet deformation due to the air flow. They predicted the deformation by incorporating a correlation by Cho et al. [106,107] at a given superficial air velocity and droplet size. The calculated droplet deformations and air drag values led to sets of air velocity and droplet size that would allow the droplets to adhere to the GDLechannel interface and remain spherical. The results in Fig. 7 show eleven droplet obstructions with uniform droplet spacing of 2.00 mm, variable superficial air velocity and droplet radius. The case uG ¼ 10.59 m s1 and r ¼ 0.15 mm led to the highest average Sherwood number in the flow direction out of four sets of the aforementioned parameters. Multiple droplet obstructions were shown to have a significant effect on the Sherwood number in the channel. After the seventh obstruction, the Sherwood number follows the same pattern and the maximum average Sherwood number in between two consequent droplets is 7.466 (2.229 ShFD). The authors estimated the impact of Sherwood number increase on the PEMFC performance. They used the following equation to calculate the difference in cell voltage for a change in oxygen concentration at the catalyst layer: DV1e2 ¼ Ru TF1 c lnðc2 =c1 Þ [1]. The oxygen concentration values were calculated for the current density of 1.0 A cm2. The results showed that the exclusion of fully developed flow interfacial O2 transport resistance from a performance model would lead to a voltage prediction that is 4.8 mV higher than the case with the inclusion of the interfacial resistance. The 122.9% increase in Sherwood number can reduce this loss of voltage from 4.8 to 2.0 mV. This shows that droplets in a channel can increase the cell performance when the superficial air velocity and Fig. 7 e The effect of eleven droplets on the Sherwood number (Sh). Variable superficial air velocity (uG) and droplet radius (r). Position of the first droplet, x [ 3.00 mm and uniform droplet spacing: 2.00 mm. Adapted from Koz and Kandlikar [103]. 6628 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 droplet size are high enough. Similar to the effect multiple droplets, Heidary and Kermani proposed to increase the performance of direct methanol and proton exchange membrane fuel cells by placing indentations into the reactant channels [108]. Koz and Kandlikar also investigated the link between the heat and mass transfer analogy in a PEMFC air channel with droplets. In the literature, researchers pointed out the possible utilization of this analogy to use Nusselt number (Nu) values for Sherwood number [93e97,108]. However, there had not been any work published on how the Nusselt and Sherwood numbers are related to each other in the PEMFC air channel. Although the Sherwood and Nusselt numbers are the same in fully developed laminar flow conditions, they may differ under developing flow conditions. If the Schmidt number (Sc) is not equal to the Prandtl number (Pr), the Sherwood number differs from the Nusselt number in the developing flow regions. As the Sc Pr difference grows, so does the Sh Nu difference. Hence, for the condition Sc ¼ Pr, the Nusselt number can be mapped to the Sherwood number directly in both fully developed and developing flow conditions. In typical PEMFC operating conditions, the Schmidt and Prandtl numbers differ. Hence, the Sherwood number in the vicinity of droplets would not be same as the Nusselt number in the equivalent heat transfer problem. By numerically obtaining the difference between Sherwood-Nusselt numbers with the largest possible Sc Pr difference in PEMFC operating conditions, researchers can use the Nusselt number data to predict the Sherwood number in PEMFC air channels with a known maximum error. The same knowledge can also serve as the extent of validity for the Sherwood number data obtained for a single Sc and used in conditions at different Sc than the original. Koz and Kandlikar numerically solved for the equivalent heat transfer problem to the convective mass transport in an air channel in the presence of multiple droplets in a row [103]. The Schmidt and Prandtl numbers were 0.633 and 0.824, respectively at 80 C and fully humidified air. The numerically calculated Nusselt number was compared locally to the Sherwood number. These two values started to differ from each other with the flow disruption induced by the first droplet. The Nusselt number was always found to be larger than the Sherwood number. The difference increased in the flow direction and reached an asymptotic value. The lowest Sh/Nu ratio 0.876 was found to be at the superficial air velocity 10.59 m s1, droplet radius 0.12 mm, and uniform droplet spacing 2.00 mm. 4. Down-the-channel (DTC) transport Traditionally, DTC transport resistance is characterized through channel pressure drop. Additionally, flow maldistribution and two-phase flow maps allow for additional characterization of two-phase flow conditions and the resulting DTC transport resistance. These DTC studies fall into two primary categories: single-channel and multichannel. 4.1. Single channel transport For the study of DTC transport in a PEMFC reactant channel, single channel experiments provide a fundamental perspective that is advantageous for investigating the underlining phenomena within the channel. Through the use of a single channel the experimentation can be significantly simplified, cross-channel effects can be removed, and fewer assumptions are required. 4.1.1. Flow regimes DTC water transport in PEMFCs presents an issue due to the additional mass transport losses associated with two-phase flow: the combination of reactant gas, evaporated water, and liquid water. There are three main two-phase flow regimes which are considered in this work: slug flow, film flow, and mist flow. The key differences between these flow regimes are characterized by the frequency of their pressure drop fluctuations and their two-phase multiplier, which scales the single-phase pressure drop based on flow conditions. Slug flow is defined as large liquid plugs separated by relatively large gas pockets. These slugs completely fill the channel geometry. Slug flow typically results in a higher twophase pressure drop multiplier and low frequency fluctuations in the pressure drop of a system. Film flow is defined as liquid film on channel wall with significant gas pockets. These films are typically annular in round channels or cover only one channel wall in rectangular geometries. Film flow typically causes a relatively low twophase multiplier compared to slug flow, but induces higher frequency pressure drop fluctuations. Mist flow is defined as both phases being equally distributed throughout the channel cross section. Usually the liquid is entrained as very small droplets in the gas flow. Mist flow regularly has a two-phase multiplier close to 1 in PEMFC applications, and produces with minimal fluctuation in the pressure drop. 4.1.2. Pressure drop in single channels A primary consideration in characterizing two-pressure drop in simulated reactant channels is the method of liquid introduction into the channel. Primarily experiments utilize air and water as working fluids at near atmospheric conditions, which mimic that of a PEMFC cathode. However, it should be noted that these studies are typically adiabatic and do not consider continuous introduction of liquid throughout the channel [109]. Scaling of large tube correlations for pressure drop does not yield accurate results due to the dissimilar comparative magnitude of the gravitational, viscous, and surface tension forces between the large scale tubes and the microchannels [109]. In a separated flow model, the superficial velocity of each phase is calculated separately. A two-phase multiplier is utilized based on the flow conditions to scale the singlephase pressure drop to two-phase. Most notably, this form of modeling is based on the model by Lockhart and Martinelli [110] which suggested the two-phase multiplier as a function of the ratios of each phases fluid properties and gas fraction. i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 Chisholm [111] later provided a basis for predicting the Chisholm parameter (C) based on the liquid and gas Reynolds numbers. This parameter has become the primary basis for two-phase pressure drop correlations. In 1996, Mishima and Hibiki [112] conducted experiments with 1e4 mm hydraulic diameter channels with both circular and rectangular cross-sections. Flow was forced in a vertical upwards direction through both glass and aluminum tubes. Over a wide range of superficial velocities, turbulent flow was observed in both the liquid and gaseous phases. It was noted that the hydraulic diameter affected the two-phase multiplier. The Chisholm parameter was then correlated to hydraulic diameter. In 2006, English and Kandlikar [113] observed two-phase flow in horizontal channels while studying the effects of surfactants. Rectangular channels in Lexan with a hydraulic diameter of 1.018 mm were used. Over the range of superficial velocities studied, laminar flow conditions were observed. It was concluded that in laminar conditions the Chisholm parameter was similarly correlated to the hydraulic diameter. The correlation by Mishima and Hibiki [112] was modified to revise the two-phase multiplier for laminarelaminar flow. In 2001, Lee and Lee [114] conducted experiments while varying aspect ratio with channels of hydraulic diameter from 0.78 to 6.67 mm. The width of the channel was held constant at 20 mm while the height of the channel was varied. The effect of aspect ratio, as well as the effect of hydraulic diameter, was highlighted and a further correlation was proposed. They proposed the Chisholm parameter as a function of three non-dimensional parameters which describe the fluid properties of the liquid phase. These single channel experiments provide a fundamental basis for two-phase pressure drop prediction in PEMFCs. While most fundamental work in two-phase pressure drop focuses primarily on simultaneous air and water introduction at the inlet of the channel, this is not representative of PEMFC reactant channels. Continuous water introduction through the gas diffusion layer and condensation of water vapor from the flow stream play a primary role in changing the quality along the flow channel length. The continuous water introduction creates a variable liquid mass flow rate, which creates the potential for flow regime changes along the channel. Furthermore, mass consumption of reactant gases is neglected in most studies. These factors are more aptly examined through multichannel experiments. 4.1.3. Instantaneous flow rate Single channel experimentation as a simplified method to mimic parallel channel arrangements implies that channels will experience uniform flow distribution. However, flow maldistribution can be inherently caused by many factors. In 2009, Kandlikar et al. [57] identified two main causes of maldistribution in parallel channels: 1) Manifold design and local pressure distribution across inlet/exit; and 2) Uneven flow resistance (due to changes in channel geometry, flow length, and fluid properties or presence of twophase flow). 6629 Kandlikar et al. [57] proposed a technique for the experimental measurement of flow maldistribution in each of parallel channels to quantify flow maldistribution through a minimally invasive measurement. Through the use of the non-linear pressure drop within the entrance region, the Hornbeck equation was used to correlate the flow rate to the pressure drop. Pressure taps were located within entrance region of each channel, and known flow rates were supplied to calibrate individual flow channels. This method was then tested using 4 circular stainless steel tubes of various lengths. The measured flow rates averaged an error of 3.3% when validated against established theoretical predictions. This methodology presented by Kandlikar et al. [57] has been employed in ex situ and in situ PEMFC flow channel studies as well as flow distributor design [2,25,58,115]. 4.2. Parallel multichannel transport In order to study multichannel effects on water management, Owejan et al. [25] designed a standardized geometry 50 cm2 fuel cell that could meet the needs for in situ and ex situ experimentation. In order to ensure the robustness of the studies, the design was specified in accordance with the United States Department of Energy (USDOE) performance targets. A channel width of 0.7 mm was selected for both anode and cathode sides. A land width of 0.5 mm was selected for the cathode in order to obtain a land-to-channel ratio of 1.4. However, the anode land width was selected to be 1.5 mm for three reasons. Firstly, reducing channels would increase the hydrogen volumetric flow rate in each channel. Secondly, it would increase the contact area greatly, thus reducing the ohmic losses. Finally, it would minimize any pinching effect between cathode and anode channels. A channel depth of 0.4 mm was selected to allow for a reasonable repeat distance to meet the USDOE energy density targets. A channel length (183 mm) was back calculated from peak power density predictions. The channels were arranged as parallel straight channels, with a 15 switchback angle to prevent shearing of the GDL and the MEA. This design was employed in various water management studies through in situ [57,86] and ex situ [2,58,71] experiments. 4.2.1. Flow maldistribution The flow maldistribution seen by Kandlikar et al. [57] led to an investigation into how the GDL is compressed under channels and lands of a flow field. In a following study, Kandlikar et al. [115] investigated the GDL intrusion into the gas channel and its effect on flow. A flow field made of Lexan with the channel geometry suggested by Owejan et al. [25] was developed for visualization from two sides of the channel to measure intrusion. A confocal digital microscope (Keyence VHX-500) was used to optically image the cross-section of the channel, as well as to create 3-D scans of the GDL perpendicular to the flow direction. Additionally, flow rate was measured to quantify the effect of intrusion on the flow within the channel. The intrusion was measured at compressions ranging from 1.03 MPa to 10.34 MPa. A minimum intrusion of 0.2 mm was observed, while a maximum of 111.0 mm was observed at the highest compressive pressure. At typical operating ranges, the 6630 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 observed intrusion varied between 30 and 70 mm which can significantly constrict the flow channel. Significant variance in the intrusion between the channels was also seen to create significant flow maldistribution. Lee and Lee [114]. Grimm et al. provided transition criteria from slug to film and film to mist. For the transition from slug to film flow, the ratio between gas inertial force and surface tension is compared. 4.2.2. 1=0:594 ReL mG uG ¼ 2:808We0:216 L rG dh Multichannel effects investigation In order to investigate the combined result of the multichannel effects, Lu et al. [2] developed an ex situ PEMFC test section made of Lexan with a set of 8 parallel channels, 4 water introduction chambers, and 12 water inlet holes representing an active area. The apparatus allowed for the measurement of pressure drop and individual channel flow rates, and direct optical visualization of the flow field. The flow conditions were tested over a range of stoichiometric ratios from 1 to 50 for the equivalent current densities from 0 to 2.0 A cm2. Significant amounts of water holdup were noted at the outlet manifold. Sharp spikes in the airflow rate of an individual channel were observed. These spikes were seen to be caused by the periodic water drainage at the channel outlet. Additionally, it was found that slug residence time in the channel generally decreased with increase in the superficial air velocity. At a superficial air velocity of 4 m s1 the residence time was no longer decreased and above 7.4 m s1 the flow maldistribution and slug flow were significantly reduced. Using the pressure drop from various flow conditions, Lu et al. [2] identified pressure drop signatures indicative of slug, film and mist flow. In 2011, Lu et al. [58] investigated the effect of surface wettability, channel geometry, and orientation. Three surface wettability conditions were tested including a baseline (85 ), hydrophobic (116 ), and hydrophilic coating (11 ). The hydrophilic coating was observed to provide a more uniform water distribution with less flow maldistribution, while hydrophobic and baseline channels acted very comparable to one another. Channel geometry was varied from a rectangular shape (meant to represent laboratory studies), to sinusoidal geometry (meant to represent stamped plates), and trapezoidal shapes (meant to represent molded plates) while the hydraulic diameter was held similar to allow for comparison. The sinusoidal channels showed a predilection to forming film flow due to their small corner angles with the GDL surface. Both the rectangular and trapezoidal channels acted comparably, and formed slugs more regularly than the sinusoidal channels. 4.2.3. (1) For the transition from film to mist flow, the ratio between viscous and inertial forces is compared. (" 0:64 #1=1:116 )1=1:726 rG mL s 1:283 mG rL mG UL uG ¼ (2) Grimm et al. [9] provided a flow regime differentiated correlation for the Chisholm parameter based on the model proposed by Lee and Lee [114] for each the slug and the film flow regimes. This parameter is a function of l ¼ u2L/(rL s dh) and j ¼ mL (uG þ uL)/s. For the slug flow conditions, the Chisholm parameter was given by: l0:134 j0:421 CS ¼ 1:9087Re0:405 L 0:107 1x x (3) For flow conditions in the film flow regime, the Chisholm parameter was given by: l0:016 j1:716 CF ¼ 0:772Re0:051 L 0:034 1x x (4) For the mist flow conditions, two-phase pressure drop was given by a traditional homogeneous model where two-phase properties are given by the correlation proposed by Dukler [116]. mTP ¼ bmG þ ð1 bÞmL (5) rTP ¼ 1 1 1x þ rG rL (6) The results of these correlations are shown in Fig. 8. In the slug flow regime, the mean error of the new pressure drop Pressure drop modeling The aforementioned multichannel effects can have a significant effect on the superficial velocity within a PEMFC reactant channel. These effects must be incorporated into the twophase pressure drop models. Continuous water introduction from the GDL into the channel must be simulated experimentally to provide the changing mass quality DTC seen in operating PEMFCs. Due to the change in mass quality DTC in PEMFCs, specialized pressure drop modeling techniques must be developed. In 2012, Grimm et al. [9] utilized a test section that provided continuous water introduction in lieu of introduction at the entrance. The study provided modifications to English and Kandlikar [113] model to incorporate mass quality, as well as a three part flow regime separated model based on the work of Fig. 8 e Pressure drop during slug, film, and mist flow with variable air superficial velocity and constant water flow rate of 0.04 mL minL1. Adapted from Grimm et al. [9]. i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 model was 14%. Film flow showed a mean error of 4%. The use of a homogeneous model yielded a mean error of 6% for the mist flow regime. Averaged across all flow regimes the mean error was reported as 9%. Grimm et al. [9] also provide a single model based on the update of the English and Kandlikar [113] correlation which incorporates the changing mass quality into the modified Chisholm parameter. b 1x C¼A x (7) A ¼ 0:0856ðuL Þ1:202 (8) b ¼ 0:004ðuL Þ0:526 (9) Over the tested conditions, the mean error from experimental data was 14% for slug and film flow regimes. In the mist flow regime, a homogeneous flow model was recommended. This correlation provides a balance between the simplicity of a single correlation for all flow regimes and compromised accuracy in very low superficial velocity ranges. While the correlations proposed by Grimm et al. [9] predict the two-phase pressure drop reasonably well on the cathode side, they cannot be applied to the anode without modification to incorporate consumption of reactants. In 2012, See and Kandlikar [8] reviewed down-the-channel (DTC) two-phase pressure drop modeling within the PEMFC field. Both Grimm et al. [9] and See and Kandlikar [8] noted a lack of a comprehensive methodology and correlation to predict the two-phase pressure drop in the PEMFC field. See and Kandlikar [8] proposed three key differences between PEMFC channels and adiabatic two-phase flow literature. Firstly, the evaporation and condensation within the flow channel was reviewed. For most studies including evaporation and condensation, limit of thermal equilibrium is used. This assumption can be omitted in order to capture the effect of air flow. It was shown that with a dry inlet stream all product water could be removed at typical operating temperature (80 C), as seen in Fig. 9. Additionally, it was noted that in most studies that account for water uptake, thermal equilibrium was used. This assumption needs to be critically evaluated. Secondly, the Fig. 9 e Cathode reactant stream’s ability to remove product water. Adapted from See and Kandlikar [8]. 6631 linkage between anode and cathode water balance through electro-osmotic drag, thermo-osmosis, hydraulic permeation, and back diffusion was reviewed. As noted by Dai et al. [117] these parameters are difficult to measure individually and a lumped coefficient, such as net water drag, was suggested. See and Kandlikar [8] suggested that the fraction of product water to the cathode can be used as a representative parameter for this linkage. Lu et al. [118] reported the net water drag coefficients as a function of distance along the channel in lieu of a constant. They introduced a spatially resolved modeling approach around 10 discrete differential volumes within the PEMFC. This spatially resolved modeling approach can be extended to analyze other aspects of two-phase flow within PEMFCs. Finally, consumption of reactants in the channel was evaluated by See and Kandlikar [8]. While most studies neglect this factor, it can lead to approximately a 10% reduction in gas superficial velocity on the cathode, and a 66% reduction in gas superficial velocity on the anode as shown in Fig. 10. It was noted that a simple mass balance approach adequately accounts for the mass consumption. The stoichiometric ratio plays a critical role in the effect of consumption. As the stoichiometric ratio decreases, the rate of change in the superficial gas velocity increases significantly. Many studies have focused on the effect of material and consequently surface energy of the channel on flow pattern transitions. PEMFC reactant channels are typically formed from stainless steel, graphite, or plated alloys [119,120]. However, See and Kandlikar [8] noted that most two-phase pressure drop studies have used alternative materials such as Lexan channels, fused silica tubes, glass tubes, and aluminum channels. It was also noted that the effect of elevated temperatures has not been well investigated in the field of two-phase flow in microchannels. It has been consistently reported that PEMFCs typically have enhanced performance at higher temperature [1]; however, the majority of studies have been conducted at room temperature. A new modeling was proposed using the 1D analysis with element division in the DTC direction [8]. A step-wise marching technique allows for the identification of flow pattern transitions along the length of the channel. The use of this elemental approach was selected for the identification of Fig. 10 e Superficial gas velocity along the anode flow channel. Adapted from See and Kandlikar [8]. 6632 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 Fig. 11 e Schematic of element division. Adapted from See and Kandlikar [8]. changing flow regimes along the length of the channel. As shown in Fig. 11, fourteen elements were utilized to represent a typical PEMFC reactant channel. The authors highlighted several factors, such as the length of reactant channels, stoichiometry, and operating range that affect the number/sizing of the elements. At each element, a control volume (shown in Fig. 12) is evaluated for the mass balance using the steps summarized below: 1. Mass flux of liquid water, vapor, and reactants from the previous element, or initial condition, is evaluated. 2. Consumption of reactants is calculated using local current density and Faraday’s Law. 3. Water production is divided between anode and cathode using m (fraction of product water attributed to the cathode side flow). 4. Local saturation pressure is evaluated and water uptake is used to determine the two-phase mass quality. 5. The two-phase multiplier and pressure drop is calculated from the correlations by Grimm et al. [9], Eqs. (7)e(9). 6. The pressure drop in each element is summed and steps 1e5 are repeated for the next element. In Step 5, many of the widely available correlations for the Chisholm parameter (C) can be used in the pressure drop multiplier. However, correlations proposed by Grimm et al. [9] have been developed specifically for the PEMFC conditions. This proposed methodology fully incorporates the key considerations and operating parameters that differentiate PEMFCs from traditional adiabatic two-phase flow research. Fig. 12 e Control volume for mass flux balance. Adapted from See and Kandlikar [8]. This is accomplished through the use of an iterative control volume and step-wise marching approach. Good agreement between the model and both ex situ and in situ pressure drop data was noted [8]. 5. Concluding remarks As PEMFCs approach production in transportation sector, water management remains one of the unresolved areas for research. Understanding the effects and mechanisms of twophase flow throughout the components of PEMFCs is crucial to the water management. An in-depth review of the current status in this field is presented in this paper. Gas and liquid flow through the GDL have been investigated individually by a number of researchers. The role of the MPL in water management is probed. It has been proposed that the cracks in the MPL stabilized the flow of liquid water produced at the CL, although this interpretation requires further evaluation. Droplets which fill the corner of a channel act as a pinning site for other droplets, which are difficult to be removed and also aid in slug/film formation. Air flow in the channel causes oscillations in the droplet, which determines the dynamic contact angle the droplet makes with the channel and hence the channel corner filling condition. Channel wall material with larger contact angle hysteresis can accommodate larger oscillations due to air flow in the gas channel without de-pinning from the wall surface and fill the channel corners to form slug/film flow. A new parameter for the blockage of the reactant paths, area coverage ratio (ACR) was defined for defining the effect of liquid water on fuel cell performance. Water features on the surface of the GDL block the area available for reactant diffusion into the GDL. This adds to the interfacial resistance of reactant transport. ACR defines the ratio of area blocked for reactant transport. The interfacial O2 transport resistance in PEMFC air channels was expressed with the Sherwood Number and investigated numerically in a 3D channel. Although the fully developed Sherwood number in the channel can be predicted through the use of reported Nusselt number values in the literature, the impact of water features on the Sherwood number required the use of a numerical approach. It was demonstrated that even a single droplet can significantly affect Sh in the wake region if superficial air velocity in the channel is high enough. Droplets in a row were shown to increase Sh with a significantly higher intensity which can exceed a 100% increase. This shows the i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 importance of characterizing the Sherwood number for channels with droplets in a row to better predict PEMFC performance. Key experimental techniques, such as individual channel flow rate measurement, optically visible test apparatuses, and segmented PEMFCs, allowed for significant advancement of understanding two-phase issues within PEMFC flow channels. Significant flow maldistribution was identified through a variety of causes including manifold geometry, variation in channel dimensions, water accumulation, and intrusion of the GDL into the channel. A fundamental two-phase pressure drop model has been applied to PEMFCs and a new scheme for pressure drop predictions with constant water production and species consumption down-the-channel (DTC) has been provided. This enhanced our understanding of DTC transport and subsequently led to the development of an elemental technique which accurately represents the two-phase flow conditions in a fuel cell. Although the two-phase flow understanding of PEMFC researchers has significantly improved over the past few years, continued research in two-phase flow is needed to elevate the state of the art in PEMFC performance, longevity, and durability. [7] [8] [9] [10] [11] [12] [13] [14] Acknowledgments This work was conducted in the Thermal Analysis, Microfluidics, and Fuel Cell Laboratory in the Mechanical Engineering Department at the Rochester Institute of Technology and was supported by the US Department of Energy contract No. DE-EE0000470. The authors gratefully acknowledge the valuable input and material support from Wenbin Gu and Jeffrey Gagliardo at General Motors Electrochemical Laboratory at Pontiac, MI. [15] [16] [17] [18] [19] references [20] [1] Mench MM. Fuel cell engines. 1st ed. New Jersey: John Wiley & Sons, Inc.; 2008. [2] Lu Z, Kandlikar SG, Rath C, Grimm M, Domigan W, White AD, et al. Water management studies in PEM fuel cells, part II: ex situ investigation of flow maldistribution, pressure drop and two-phase flow pattern in gas channels. Int J Hydrogen Energy 2009;34:3445e56. [3] Anderson R, Zhang L, Ding Y, Blanco M, Bi X, Wilkinson DP. A critical review of two-phase flow in gas flow channels of proton exchange membrane fuel cells. J Power Sources 2010;195:4531e53. [4] Hui L, Yanghua T, Zhenwei W, Zheng S, Shaohong W, Datong S, et al. A review of water flooding issues in the proton exchange membrane fuel cell. J Power Sources 2008;178:103e17. [5] Misran E, Hassan NSM, Daud WRW, Majlan EH, Rosli MI. Water transport characteristics of a PEM fuel cell at various operating pressures and temperatures. Int J Hydrogen Energy 2013;38:9401e8. [6] Jithesh PK, Bansode AS, Sundararajan T, Das SK. The effect of flow distributors on the liquid water distribution and [21] [22] [23] [24] [25] 6633 performance of a PEM fuel cell. Int J Hydrogen Energy 2012;37:17158e71. Rakhshanpouri S, Rowshanzamir S. Water transport through a PEM (proton exchange membrane) fuel cell in a seven-layer model. Energy 2013;50:220e31. See EJ, Kandlikar SG. A two-phase pressure drop model incorporating local water balance and reactant consumption in PEM fuel cell gas channels. ECS Trans 2013;50:99e111. Grimm M, See EJ, Kandlikar SG. Modeling gas flow in PEMFC channels: part I e flow pattern transitions and pressure drop in a simulated ex situ channel with uniform water injection through the GDL. Int J Hydrogen Energy 2013;37:12489e503. Zhang L, Bi XT, Wilkinson DP, Anderson R, Stumper J, Wang H. Gaseliquid two-phase flow behavior in minichannels bounded with a permeable wall. Chem Eng Sci 2011;66:3377e85. Yang H, Zhao TS, Cheng P. Gaseliquid two-phase flow patterns in a miniature square channel with a gas permeable sidewall. Int J Heat Mass Transf 2004;47:5725e39. Manke I, Hartnig C, Grünerbel M, Lehnert W, Kardjilov N, Haibel A, et al. Investigation of water evolution and transport in fuel cells with high resolution synchrotron Xray radiography. Appl Phys Lett 2007;90. 174105-1e3. Bazylak A, Sinton D, Djilali N. Dynamic water transport and droplet emergence in PEMFC gas diffusion layers. J Power Sources 2008;176:240e6. Litster S, Sinton D, Djilali N. Ex situ visualization of liquid water transport in PEM fuel cell gas diffusion layers. J Power Sources 2006;154:95e105. Owejan JP, Owejan JE, Gu W, Trabold TA, Tighe TW, Mathias MF. Water transport mechanisms in PEMFC gas diffusion layers. J Electrochem Soc 2010;157:B1456e64. Zhou P, Wu CW. Liquid water transport mechanism in the gas diffusion layer. J Power Sources 2010;195:1408e15. Springer TE, Zawodzinski TA, Gottesfeld S. Polymer electrolyte fuel cell model. J Electrochem Soc 1991;138:2334e42. Nguyen TV, White RE. A water and heat management model for proton-exchange-membrane fuel cells. J Electrochem Soc 1993;140:2178e86. Sinha PK, Wang C-Y. Pore-network modeling of liquid water transport in gas diffusion layer of a polymer electrolyte fuel cell. Electrochim Acta 2007;52:7936e45. Shi W, Kurihara E, Oshima N. Effect of capillary pressure on liquid water removal in the cathode gas diffusion layer of a polymer electrolyte fuel cell. J Power Sources 2008;182:112e8. Qi Z, Kaufman A. Improvement of water management by a microporous sublayer for PEM fuel cells. J Power Sources 2002;109:38e46. Atiyeh HK, Karan K, Peppley B, Phoenix A, Halliop E, Pharoah J. Experimental investigation of the role of a microporous layer on the water transport and performance of a PEM fuel cell. J Power Sources 2007;170:111e21. Jiao K, Li X. Water transport in polymer electrolyte membrane fuel cells. Prog Energy Combust Sci 2011;37:221e91. Lu Z, Daino MM, Rath C, Kandlikar SG. Water management studies in PEM fuel cells, part III: dynamic breakthrough and intermittent drainage characteristics from GDLs with and without MPLs. Int J Hydrogen Energy 2010;35:4222e33. Owejan JP, Gagliardo JJ, Sergi JM, Kandlikar SG, Trabold TA. Water management studies in PEM fuel cells, part I: fuel cell design and in situ water distributions. Int J Hydrogen Energy 2009;34:3436e44. 6634 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 [26] Gao B, Steenhuis TS, Zevi Y, Parlange J-Y, Carter RN, Trabold TA. Visualization of unstable water flow in a fuel cell gas diffusion layer. J Power Sources 2009;190:493e8. [27] Pharoah JG. On the permeability of gas diffusion media used in PEM fuel cells. J Power Sources 2005;144:77e82. [28] Bruggeman DAG. Berechnung verschiedener physikalischer Konstanten von heterogenen Substanzen. I. Dielektrizitätskonstanten und Leitfähigkeiten der Mischkörper aus isotropen Substanzen. Ann Der Phys 1935;416:665e79. [29] Neale GH, Nader WK. Prediction of transport processes within porous media: diffusive flow processes within an homogeneous swarm of spherical particles. AIChE J 1973;19:112e9. [30] Das PK, Li X, Liu Z-S. Effective transport coefficients in PEM fuel cell catalyst and gas diffusion layers: beyond Bruggeman approximation. Appl Energy 2010;87:2785e96. [31] Zamel N, Li X, Shen J. Correlation for the effective gas diffusion coefficient in carbon paper diffusion media. Energy Fuels 2009;23:6070e8. [32] Tomadakis MM, Sotirchos SV. Effective diffusivities and conductivities of random dispersions of nonoverlapping and partially overlapping unidirectional fibers. J Chem Phys 1993;99:9820e7. [33] Zamel N, Astrath NGC, Li X, Shen J, Zhou J, Astrath FBG, et al. Experimental measurements of effective diffusion coefficient of oxygen-nitrogen mixture in PEM fuel cell diffusion media. Chem Eng Sci 2010;65:931e7. [34] LaManna JM, Kandlikar SG. Determination of effective water vapor diffusion coefficient in PEMFC gas diffusion layers. Int J Hydrogen Energy 2011;36:5021e9. [35] Gostick JT, Fowler MW, Pritzker MD, Ioannidis MA, Behra LM. In-plane and through-plane gas permeability of carbon fiber electrode backing layers. J Power Sources 2006;162:228e38. [36] Tamayol A, Bahrami M. In-plane gas permeability of polymer electrolyte membrane fuel cell gas diffusion layer. J Power Sources 2011;196:3559e64. [37] Feser JP, Prasad AK, Advani SG. On the relative influence of convection in serpentine flow fields of PEM fuel cells. J Power Sources 2006;161:404e12. [38] Ahmed DH, Sung HJ, Bae J. Effect of GDL permeability on water and thermal management in PEMFCs-I. Isotropic and anisotropic permeability. Int J Hydrogen Energy 2008;33:3767e85. [39] Park S, Lee JW, Popov BN. Effect of PTFE content in microporous layer on water management. ECS Trans 2007;11:623e8. [40] Banerjee R, Kandlikar SG. Effect of temperature on in-plane permeability of the gas diffusion layer of PEM fuel cell. ECS Trans 2011;41:489e97. [41] Pant L, Mitra SK, Secanell MA. Generalized mathematical model to study gas transport in PEMFC porous media. Int J Heat Mass Transf 2013;58:70e9. [42] Pant L, Mitra SK, Secanell MA. Absolute permeability and Knudsen diffusivity measurements in PEMFC gas diffusion layers and micro porous layers. J Power Sources 2012;206:153e60. [43] Roshandel R, Farhanieh B, Saievar-Iranizad E. The effects of porosity distribution variation on PEM fuel cell performance. Renew Energy 2005;30:1557e72. [44] Sadeghifar H, Bahrami M, Djilali N. A statistically-based thermal conductivity model for fuel cell gas diffusion layers. J Power Sources 2012;233:369e79. [45] Owejan JP, Trabold TA, Jacobson DL, Baker DR, Hussey DS, Arif M. In situ investigation of water transport in an operating PEM fuel cell using neutron radiography: part 2 e transient water accumulation in an interdigitated cathode flow field. Int J Heat Mass Transf 2006;49:4721e31. [46] Tsushima S, Nanjo T, Nishida K, Hirai S. Investigation of the lateral water distribution in a proton exchange membrane in fuel cell operation by 3D-MRI. ECS Trans 2006;1:199e205. [47] Satija R, Jacobson DL, Arif M, Werner SA. In situ neutron imaging technique for evaluation of water management systems in operating PEM fuel cells. J Power Sources 2004;129:238e45. [48] Bellows RJ. Neutron imaging technique for in situ measurement of water transport gradients within Nafion in polymer electrolyte fuel cells. J Electrochem Soc 1999;146:1099e103. [49] Chuang PA, Turhan A, Heller AK, Brenizer JS, Trabold TA, Mench MM. The nature of flooding and drying in polymer electrolyte fuel cells. In: Proceedings of the 3rd international conference on fuel cell science, engineering and technology; 2005 May 23e25. Ypsilanti, MI: ASME; 2005. pp. 31e7. [50] Tsushima S, Teranishi K, Hirai S. Magnetic resonance imaging of the water distribution within a polymer electrolyte membrane in fuel cells. Electrochem Solid-State Lett 2004;7:A269e72. [51] Feindel KW, Bergens SH, Wasylishen RE. The influence of membrane electrode assembly water content on the performance of a polymer electrolyte membrane fuel cell as investigated by 1H NMR microscopy. Phys Chem Chem Phys 2007;9:1850e7. [52] Tuber K, Pocza D, Hebling C. Visualization of water buildup in the cathode of a transparent PEM fuel cell. J Power Sources 2003;124:403e14. [53] Dunbar Z, Masel RI. Quantitative MRI study of water distribution during operation of a PEM fuel cell using Teflon flow fields. J Power Sources 2007;171:678e87. [54] Hossain M, Islam SZ, Colley-Davies A, Adom E. Water dynamics inside a cathode channel of a polymer electrolyte membrane fuel cell. Renew Energy 2013;50:763e79. [55] Yang XG, Zhang FY, Lubawy AL, Wang CY. Visualization of liquid water transport in a PEFC. Electrochem Solid-State Lett 2004;7:A408e11. [56] Turhan A, Kim S, Hatzell M, Mench MM. Impact of channel wall hydrophobicity on through-plane water distribution and flooding behavior in a polymer electrolyte fuel cell. Electrochim Acta 2010;55:2734e45. [57] Kandlikar SG, Lu Z, Domigan WE, White AD, Benedict MW. Measurement of flow maldistribution in parallel channels and its application to ex-situ and in-situ experiments in PEMFC water management studies. Int J Heat Mass Transf 2009;52:1741e52. [58] Lu Z, Rath C, Zhang G, Kandlikar SG. Water management studies in PEM fuel cells, part IV: effects of channel surface wettability, geometry and orientation on the two-phase flow in parallel gas channels. Int J Hydrogen Energy 2011;36:9864e75. [59] Cheah MJ, Kevrekidis IG, Benziger JB. Water slug formation and motion in gas flow channels: the effects of geometry, surface wettability, and gravity. Langmuir 2013;29:9918e34. [60] Theodorakakos A, Ous T, Gavaises M, Nouri JM, Nikolopoulos N, Yanagihara H. Dynamics of water droplets detached from porous surfaces of relevance to PEM fuel cells. J Colloid Interface Sci 2006;300:673e87. [61] Zhu X, Sui PC, Djilali N. Three-dimensional numerical simulations of water droplet dynamics in a PEMFC gas channel. J Power Sources 2008;181:101e15. [62] Zhu X, Liao Q, Sui PC, Djilali N. Numerical investigation of water droplet dynamics in a low-temperature fuel cell microchannel: effect of channel geometry. J Power Sources 2010;195:801e12. [63] Cai YH, Hu J, Ma HP, Yi BL, Zhang HM. Effects of hydrophilic/ hydrophobic properties on the water behavior in the micro- i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 [64] [65] [66] [67] [68] [69] [70] [71] [72] [73] [74] [75] [76] [77] [78] [79] [80] [81] [82] channels of a proton exchange membrane fuel cell. J Power Sources 2006;161:843e8. Quan P, Lai M-C. Numerical study of water management in the air flow channel of a PEM fuel cell cathode. J Power Sources 2007;164:222e37. Kandlikar SG, See EJ, Gopalan P, Koz M, Banerjee R. Twophase flow in GDL and reactant channels of a proton exchange membrane fuel cell. In: Proceedings of 8th international conference on multiphase flow; 2013 May 26e31; Jeju, Korea; 2013. Das S, Mitra SK. Electric double-layer interactions in a wedge geometry: change in contact angle for drops and bubbles. Phys Rev E 2013;88:033021. Chang F-M, Hong S-J, Sheng Y-J, Tsao H-K. Wetting invasion and retreat across a corner boundary. J Phys Chem C 2010;114:1615e21. Rath CD, Kandlikar SG. Liquid filling in a corner with a fibrous walldan application to two-phase flow in PEM fuel cell gas channels. Colloids Surf A 2011;384:653e60. Concus P, Finn R. Capillary surfaces in a wedgeddiffering contact angles. Microgravity Sci Technol 1994;7:152e5. Concus P, Finn R. Capillary wedges revisited. SIAM J Math Anal 1996;27:56e69. Gopalan P, Kandlikar SG. Droplet-sidewall dynamic interactions in PEMFC gas channels. J Electrochem Soc 2012;159:F468e75. Gopalan P, Kandlikar SG. Investigation of water droplet interaction with the sidewalls of the gas channel in a PEM fuel cell in the presence of gas flow. ECS Trans 2011;41:479e88. Polverino P, Esposito A, Pianese C. Experimental validation of a lumped model of single droplet deformation, oscillation and detachment on the GDL surface of a PEM fuel cell. Int J Hydrogen Energy 2013;38:8934e53. Gopalan P, Kandlikar SG. Contact line characteristics of liquid-gas interfaces over confined/structured surfaces. In: Proceedings of the 3rd European conference on microfluidics e microfluidics 2012; 2012 December 3e5; Heidelberg, Germany; 2012. Gopalan P, Kandlikar SG. Effect of channel materials on the behavior of water droplet emerging from GDL into PEMFC gas channels. ECS Trans 2012;50:503e12. Gopalan P, Kandlikar SG. Effect of channel material on water droplet dynamics in a PEMFC gas channel. J Electrochem Soc 2013;160:F487e95. Trabold TA, Owejan JP, Jacobson DL, Arif M, Huffman PR. In situ investigation of water transport in an operating PEM fuel cell using neutron radiography: part 1 e experimental method and serpentine flow field results. Int J Heat Mass Transf 2006;49:4712e20. Wu J, Zi Yuan X, Wang H, Blanco M, Martin JJ, Zhang J. Diagnostic tools in PEM fuel cell research: part II: physical/chemical methods. Int J Hydrogen Energy 2008;33:1747e57. Geiger AB, Tsukada A, Lehmann E, Vontobel P, Wokaun A, Scherer GG. In situ investigation of two-phase flow patterns in flow fields of PEFC’s using neutron radiography. Fuel Cells 2002;2:92e8. Lee S-J, Kim S-G, Park G-G, Kim C-S. Quantitative visualization of temporal water evolution in an operating polymer electrolyte fuel cell. Int J Hydrogen Energy 2010;35:10457e63. Markötter H, Alink R, Haußmann J, Dittmann K, Arlt T, Wieder F, et al. Visualization of the water distribution in perforated gas diffusion layers by means of synchrotron Xray radiography. Int J Hydrogen Energy 2012;37:7757e61. Bazylak A. Liquid water visualization in PEM fuel cells: a review. Int J Hydrogen Energy 2009;34:3845e57. 6635 [83] Zhan Z, Wang C, Fu W, Pan M. Visualization of water transport in a transparent PEMFC. Int J Hydrogen Energy 2012;37:1094e105. [84] Hussaini IS, Wang C-Y. Visualization and quantification of cathode channel flooding in PEM fuel cells. J Power Sources 2009;187:444e51. [85] Weng F-B, Su A, Hsu C-Y, Lee C-Y. Study of water-flooding behaviour in cathode channel of a transparent protonexchange membrane fuel cell. J Power Sources 2006;157:674e80. [86] Sergi JM, Kandlikar SG. Quantification and characterization of water coverage in PEMFC gas channels using simultaneous anode and cathode visualization and image processing. Int J Hydrogen Energy 2011;36:12381e92. [87] Gu W, Baker DR, Liu Y, Gasteiger HA. Proton exchange membrane fuel cell (PEMFC) down-the-channel performance model. In: Vielstich W, Yokokawa H, Gasteiger HA, editors. Handbook of fuel cells e fundamentals, technology and applications, vol. 6. England: John Wiley & Sons; 2009. pp. 631e57. [88] Pasaogullari U, Wang C. Liquid water transport in gas diffusion layer of polymer electrolyte fuel cells. J Electrochem Soc 2004;151:A399e406. [89] Pasaogullari U, Wang C, Chen K. Two-phase transport in polymer electrolyte fuel cells with bilayer cathode gas diffusion media. J Electrochem Soc 2005;152:A1574e82. [90] Kim GS, Sui PC, Shah AA, Djilali N. Reduced-dimensional models for straight-channel proton exchange membrane fuel cells. J Power Sources 2010;195:3240e9. [91] He J, Ahn J, Choe SY. Analysis and control of a fuel cell delivery system considering a two-phase anode model of the polymer electrolyte membrane fuel cell stack. J Power Sources 2011;196:4655e70. [92] Hussaini IS, Wang CY. Dynamic water management of polymer electrolyte membrane fuel cells using intermittent RH control. J Power Sources 2010;195:3822e9. [93] Wu H, Li X, Berg P. Numerical analysis of dynamic processes in fully humidified PEM fuel cells. Int J Hydrogen Energy 2007;32:2022e31. [94] Wu H, Berg P, Li X. Non-isothermal transient modeling of water transport in PEM fuel cells. J Power Sources 2007;165:232e43. [95] Casalegno A, Bresciani F, Groppi G, Marchesi R. Flooding of the diffusion layer in a polymer electrolyte fuel cell: experimental and modelling analysis. J Power Sources 2011;196:10632e9. [96] Darling RM, Khateeb S. The influence of gas-diffusion layer properties on elevated temperature operation of polymer-electrolyte fuel cells. J Power Sources 2013;243:328e35. [97] Park J, Min K. A quasi-three-dimensional nonisothermal dynamic model of a high-temperature proton exchange membrane fuel cell. J Power Sources 2012;216:152e61. [98] Das PK, Li X, Liu ZS. Analysis of liquid water transport in cathode catalyst layer of PEM fuel cells. Int J Hydrogen Energy 2010;35:2403e16. [99] Wang ZH, Wang CY, Chen KS. Two-phase flow and transport in the air cathode of proton exchange membrane fuel cells. J Power Sources 2001;94:40e50. [100] Jeng K-T, Wen C-Y, Anh LD. A study on mass transfer in the cathode gas channel of a proton exchange membrane fuel cell. J Mech 2007;23:275e84. [101] Hassanzadeh H, Li X, Baschuk JJ, Mansouri SH. Numerical simulation of laminar flow development with heat and mass transfer in PEM fuel cell flow channels having oxygen and hydrogen suction at one channel wall. Int J Energy Res 2011;35:670e89. 6636 i n t e r n a t i o n a l j o u r n a l o f h y d r o g e n e n e r g y 3 9 ( 2 0 1 4 ) 6 6 2 0 e6 6 3 6 [102] Koz M, Kandlikar SG. Numerical simulation of the interfacial oxygen transport resistance for a PEMFC cathode incorporating water droplet coverage. ECS Trans 2012;50:183e96. [103] Koz M, Kandlikar SG. Numerical investigation of interfacial transport resistance due to water droplets in proton exchange membrane fuel cell air channels. J Power Sources 2013;243:946e57. [104] Das PK, Grippin A, Kwong A, Weber AZ. Liquid-waterdroplet adhesion-force measurements on fresh and aged fuel-cell gas-diffusion layers. J Electrochem Soc 2012;159:B489e96. [105] Zhang F, Yang X, Wang C. Liquid water removal from a polymer electrolyte fuel cell. J Electrochem Soc 2006;153:A225e32. [106] Cho SC, Wang Y, Chen KS. Droplet dynamics in a polymer electrolyte fuel cell gas flow channel: forces, deformation, and detachment. I: theoretical and numerical analyses. J Power Sources 2012;206:119e28. [107] Cho SC, Wang Y, Chen KS. Droplet dynamics in a polymer electrolyte fuel cell gas flow channel: forces, deformation and detachment. II: comparisons of analytical solution with numerical and experimental results. J Power Sources 2012;210:191e7. [108] Heidary H, Kermani MJ. Performance enhancement of fuel cells using bipolar plate duct indentations. Int J Hydrogen Energy 2013;38:5485e96. [109] Kandlikar SG, Garimella S, Li D, Colin S, King MR. Heat transfer and fluid flow in minichannels and microchannels. 1st ed. Great Britain: Elsevier; 2006. [110] Lockhart RW, Martinelli RC. Proposed correlation of data for isothermal two phase flow, two component flow in pipes. Chem Eng Prog 1949;45:39e48. [111] Chisholm D. A theoretical basis for the Lockhart-Martinelli correlation for two-phase flow. Int J Heat Mass Transf 1967;10:1767e78. [112] Mishima K, Hibiki T. Some characteristics of air-water twophase flow in small diameter vertical tubes. Int J Multiph Flow 1996;22:703e12. [113] English NJ, Kandlikar SG. An experimental investigation into the effect of surfactants on air-water two-phase flow in minichannels. Heat Transf Eng 2006;27:99e109. [114] Lee HJ, Lee SY. Pressure drop correlations for two-phase flow within horizontal rectangular channels with small heights. Int J Multiph Flow 2001;27:783e96. [115] Kandlikar SG, Lu Z, Lin TY, Cooke D, Daino M. Uneven gas diffusion layer intrusion in gas channel arrays of proton exchange membrane fuel cell and its effects on flow distribution. J Power Sources 2009;194:328e37. [116] Dukler AE, Wicks M, Cleveland RG. Frictional pressure drop in two-phase flow. AIChE J 1964;10:44e51. [117] Dai W, Wang H, Yuan XZ, Martin JJ, Yang D, Qiao J, et al. A review on water balance in the membrane electrode assembly of proton exchange membrane fuel cells. Int J Hydrogen Energy 2009;34:9461e78. [118] Lu GQ, Liu FQ, Wang C-Y. An approach to measuring spatially resolved water crossover coefficient in a polymer electrolyte fuel cell. J Power Sources 2007;164:134e40. [119] Li X, Sabir I. Review of bipolar plates in PEM fuel cells: flowfield designs. Int J Hydrogen Energy 2005;30:359e71. [120] Dur E, Cora Ö, Koç M. Experimental investigations on the corrosion resistance characteristics of coated metallic bipolar plates for PEMFC. Int J Hydrogen Energy 2011;36:7162e73.
© Copyright 2025 Paperzz