797
The Chemistry of Membrane Degradation in PEM Fuel Cells

Wednesday, May 14, 2014: 08:20
Palm Beach, Ground Level (Hilton Orlando Bonnet Creek)
M. P. Rodgers, L. J. Bonville (Florida Solar Energy Center-University of Central Florida), R. Mukundan, R. L. Borup (Los Alamos National Laboratory), S. Knights (Ballard Power Systems), R. Ahluwalia (Argonne National Laboratory), P. Beattie (Ballard Power Systems), R. P. Brooker, N. Mohajeri, H. R. Kunz, J. M. Fenton, and D. K. Slattery (Florida Solar Energy Center-University of Central Florida)
The lifetime of proton exchange membrane (PEM) fuel cells depends largely on the durability of the PEM and is a critical issue in the commercialization of PEM fuel cell technology (1, 2). Perfluorosulfonated acid (PFSA) membranes, such as Nafion®, are the most widely used electrolyte material for PEM fuel cells.

Many factors can cause PEM failure, including manufacturing and design issues, material properties, and fuel cell operation conditions. Each factor results in failure through a particular degradation mode. These modes can be classified as thermal, mechanical, and chemical modes (3-7). Because fuel cells typically operate at temperatures below that at which Nafion thermally degrades, chemical and mechanical degradation are generally considered the most important modes of membrane degradation.

Although there has been a lot of progress in the area, the exact chemical membrane degradation mechanisms, the most susceptible points, and the impact of test conditions on degradation are not completely understood. The formation of radicals and their reactivity during the fuel cell operation have been confirmed as the main cause of PEM chemical degradation (8-11). These radicals are formed as a result of the reaction of crossover H2 and O2 on the surface of Pt  or other metal ion contaminants (4). In addition to the electrodes, Pt particles have been observed to migrate into the membrane, producing another site for radical formation. Radical attack leads to loss of fluorine, sulfur, and organofluorine compounds from the PEM, and PEM thinning, all resulting in loss of mechanical integrity. The radical formation mechanisms and the impact of test conditions on radical formation are also not completely understood.

It has been suggested that earlier versions of Nafion membranes chemically degrade through radical attack on carboxylic acid groups end groups (12-14) or at the C−S or O−C bonds in the side chain and main chain (12, 14-18), as shown in Figure 1. Chemically stabilized Nafion, e.g., Nafion 211, has been developed where the concentration of terminal carboxylic acid groups was decreased to negligible levels and, as a result, degradation decreased substantially (13). The use of radical scavengers such as ceria have increased the durability of PFSA membranes by as much as 1000-fold compared to unmitigated membranes (19). Despite the recent durability improvements, there is evidence that PFSA membranes are still susceptible to chemical damage.

Typical indicators for chemical degradation of Nafion membranes include increased H2crossover, decreased ion exchange capacity, decreased proton conductivity, and decreased PEM thickness. Localized crossover tests have shown that membrane damage occurs preferably at the perimeter of the active area. The quantitative method of choice used for detecting chemical degradation is typically detection of fluoride in fuel cell outlets, reported as a fluoride emission rate (FER).

Figure 1. Proposed locations of radical attack in Nafion

To improve PEM chemical durability, a true understanding of the decay mechanisms is required. Models of some of the decay modes have been developed and they show a consistent but incomplete understanding of membrane decay.  Through a combination of specific modeling and directed testing, an overall decay model can be developed.

References

1.M. P. Rodgers, L. J. Bonville, H. R. Kunz, D. K. Slattery and J. M. Fenton, Chem. Rev., 112, 6075 (2012).

2.M. P. Rodgers, L. J. Bonville, R. Mukundan, R. L. Borup, R. Ahluwalia, P. Beattie, R. P. Brooker, N. Mohajeri, H. R. Kunz, D. K. Slattery and J. M. Fenton, Meeting Abstracts, MA2013-02, 1270 (2013).

3.C. H. Paik, T. Skiba, V. Mittal, S. Motupally and T. Jarvi, ECS Meeting Abstracts, 501, 771 (2006).

4.V. O. Mittal, H. R. Kunz and J. M. Fenton, J. Electrochem. Soc., 154, B652 (2007).

5.S. Kundu, M. W. Fowler, L. C. Simon and S. Grot, J. Power Sources, 157, 650 (2006).

6.E. Endoh, S. Terazono, H. Widjaja and Y. Takimoto, Electrochem. Solid-State Lett., 7, A209 (2004).

7.M. F. Mathias, R. Makharia, H. A. Gasteiger, J. J. Conley, T. J. Fuller, C. J. Gittleman, S. S. Kocha, D. P. Miller, C. K. Mittelsteadt and T. Xie, Electrochem. Soc. Interface, 14(2005).

8.M. Danilczuk, F. D. Coms and S. Schlick, The Journal of Physical Chemistry B, 113, 8031 (2009).

9.M. Danilczuk, A. J. Perkowski and S. Schlick, Macromolecules, 43, 3352 (2010).

10.A. Panchenko, H. Dilger, E. Möller, T. Sixt and E. Roduner, J. Power Sources, 127, 325 (2004).

11.A. Panchenko, H. Dilger, J. Kerres, M. Hein, A. Ullrich, T. Kaz and E. Roduner, Phys. Chem. Chem. Phys., 6, 2891 (2004).

12.L. Ghassemzadeh, K.-D. Kreuer, J. Maier and K. Müller, The Journal of Physical Chemistry C, 114, 14635 (2010).

13.D. E. Curtin, R. D. Lousenberg, T. J. Henry, P. C. Tangeman and M. E. Tisack, J. Power Sources, 131, 41 (2004).

14.F. D. Coms, ECS Trans., 16, 235 (2008).

15.L. Ghassemzadeh, M. Marrony, R. Barrera, K. D. Kreuer, J. Maier and K. Müller, J. Power Sources, 186, 334 (2009).

16.B. Vogel, E. Aleksandrova, S. Mitov, M. Krafft, A. Dreizler, J. Kerres, M. Hein and E. Roduner, ECS Trans., 11, 1105 (2007).

17.T. H. Yu, Y. Sha, W.-G. Liu, B. V. Merinov, P. Shirvanian and W. A. Goddard, J. Am. Chem. Soc., 133, 19857 (2011).

18.L. Ghassemzadeh, K. D. Kreuer, J. Maier and K. Muller, J. Power Sources, 196, 2490 (2011).

19.B. P. Pearman, N. Mohajeri, R. P. Brooker, M. P. Rodgers, D. K. Slattery, M. D. Hampton, D. A. Cullen and S. Seal, J. Power Sources, 225, 75 (2013).