You are on page 1of 2

CATALYSTS FOR DIRECT ETHANOL FUEL CELLS

Jonathan Mann, M. Scott Daubin, and Andrew B. Bocarsly Chemistry Department Princeton University Princeton, NJ 08544 Background Of the two existing pre-market proton exchange membrane fuel cells: hydrogen-oxygen and direct methanol, H2/O2 is suggested for applications involving high energy density, while direct methanol holds pragmatic promise for small low power application where ease of fuel handling is critical. Of equal importance, is the energy per unit volume when matching an application to a fuel, since available volume for a mobile application is often a limiting consideration. A simple review of the thermochemical parameters (provided in Table 1) indicates that ethanol has fuel characteristics similar to methanol. However, it substantially exceeds methanol with respect to volumetric energy density providing 68% of the volumetric energy density of gasoline. Thus, from an energy storage point of view, ethanol appears to be a preferred fuel over methanol. Table 1. Energy Content of Fuel Cell Fuels Assuming Complete Combustion Gasolin Thermochemical Ethanol Methanol e Parameter -1406 -658 H, kJ/mol -1324 -702 G, kJ/mol Specific Energy, MJ/kg 26.8 19.9 44.4 Energy Density, MJ/L 21.2 15.8 31.1 Likewise, a strong environmental argument can be made for ethanol over methanol. Methanol is water soluble and rather toxic. As recently seen with the withdrawal of MBTE as a gasoline additive this combination of properties represents an environmental threat. Ethanol though water soluble, is much less toxic than methanol, thus the environmental threat is dramatically attenuated. Introduction Ethanol oxidation at the anode of a fuel cell requires the development of a catalyst. The complete anode reaction (1) involves the breaking of C-H bonds and C-C bonds as well as the formation of C=O bonds. New alloys for this catalysis were investigated. CH3CH2OH + 3 H2O 2 O2 + 12 H+ + 12 e(1) Preliminary fuel cell studies at elevated temperature have shown that ethanol can be oxidized to CO2 with a yield of 96% above 140C.1 It has also been reported that the concentration of ethanol at the electrode surface has a significant effect, with ethanol concentrations much higher than 1 M resulting in catalytic poisoning by adsorbed, partially oxidized ethanol species. This is easily explained by understanding that a high ethanol concentration also means a lower water concentration. It is easy to see from reaction (1) that a decrease in water concentration at the electrode interface will result in incomplete oxidation. 2 In addition, studies3 have shown that a variety of bimetallic catalysts containing platinum and a second post-transition metal may be catalytically superior to a pure platinum electrocatalyst. Three different types of broad-based systems have been suggested: bulk alloys, surface ad-atom composites and (quite recently) intermetallic phases.4

To date, alloy base systems have received the most attention. Though similar, different trends are noticed for methanol and hydrogen oxidation than for ethanol. Specifically, Pt/Sn alloy systems have been shown to demonstrate higher activity than Pt/Ru, the preferred alloy for methanol and reformed hydrogen oxidation. This finding has been correlated with the need for a large quantity of adsorbed oxygen in the case of ethanol oxidation. Tin adopts an oxidized state while platinum and ruthenium remain in more reduced states. This added oxygen promotes the ethanol oxidation on the Pt/Sn surface over others. It is well established from electrochemical studies that bulk ethanol is oxidized only partially, while surface adsorbed ethanol undergoes complete oxidation to CO2. As might be expected, the complete oxidation of ethanol requires a higher potential than does the partial oxidation to acetaldehyde or acetic acid; surface adsorbed species do not diffuse away from the electrode before the higher potential is reached. This conclusion not only suggests a strategy for the development of new catalysts, but also indicates that electrochemical studies (which typically utilize low surface area electrodes in contact with an aqueous electrolyte) may not faithfully model the fuel cell environment that incorporates a supported nanoparticulate catalyst bed in contact with a polymer electrolyte. Experimental Electrodes. Anodes were prepared from a commercial carbon cloth with a single gas diffusion layer (ETEK). Catalysts (ETEK) were Pt binary and ternary alloys on carbon. They were suspended in a solution of iso-propanol and sprayed onto the diffusion layer of the electrode. The catalytic loading was aimed at 0.4 mg/cm2, and the exact loading was calculated by mass difference. The same procedure was used to impregnate Nafion into the newly deposited active layer of the anode electrodes. Loadings of 0.6 mg/cm2 were the aim for the Nafion. Cathodes were prepared from commercial carbon cloths with a pre-applied active layer of 0.4 mg/cm2 of Pt (ETEK). Nafion was impregnated into the active layer of the catalysts at a loading of 0.6 mg/cm2. Membranes. Nafion 115 membranes were uses after a boiling in successive bathes of 0.5 M H2SO4, H2O, 3% H2O2, and H2O twice. Membranes made from dissolved Nafion (Ion Power) combined with metal oxide particles and cast into membranes were also used. This second type of membrane underwent the same kind of treatment before being assembled into the cell. Cell Testing. The membrane and electrodes were assembled according to existing procedures.5 The membrane electrode assembly (MEA) was then inserted into a GlobeTech single cell test station where the potential on each electrode was independently cycled from 0.1 to 1.0 volts at 100 mV/s under nitrogen. The cell was then moved to a modified GlobeTech single cell test station equipped to deliver liquids to the anode. The liquid at the anode was 1.0 M ethanol flowing at 0.2 mL/min. The cathode was pure oxygen with a flow rate of 150 mL/min. Both sides of the cell were pressurized above 60 psig. Polarization curves were used to evaluate cell performance as a function of temperature and anode catalyst material. The dependence on flow rate, membrane, and pressure was also considered. Preliminary Results Polarization curves for the tested cells confirmed that the Pt:Sn, alloy (3:1 atomic ratio) is the preferred catalyst among Pt:Ru (1:1) and Pt (not shown). Both open circuit voltages and activation over potentials are more favorable for the Sn alloy. (Fig. 1.)

Prepr. Pap.-Am. Chem. Soc., Div. Fuel Chem. 2004, 49(2), 662

900 800 Cell Potential (mV) 700 600 500 400 300 200 100 0 0 20 40 60 Current Density (mA/scm) 80
Pt:Sn (3:1)
Pt:Ru (1:1)

Figure 1. Reproducible comparison of performance of Pt:Sn catalyzed cell with Pt:Ru catalyzed cell. The standard cell potential for the ethanol reaction is 1.14 V, about 90 mV less then the H2/O2 cell, while the open circuit of the ethanol cell, measured as high as 808 mV, is less than 200 mV below that of the common H2/O2 cell. This reasonably high open circuit potential is a promising number considering the relatively unresearched array of possible catalytic particles available. However, the extraordinarily low current densities demonstrate that the presently employed catalysis is insufficient at extracting power from the fuel. Summary The study of ethanol as a fuel in fuel cells is a relatively new exploration in the field. The motivation to use it over hydrogen or methanol is several-fold, while its demonstrated abilities leave something to be desired. Continued exploration into the vast array of catalytic particles will reveal the most ideal catalytic system which then can be combined with the most successful in situ conditions to produce a power source capable of competing with existing electrochemical cells.

References 1 (a) Arico, A.S., Creti, P., Antonucci, P.L. and Antonucci V., Electrochemical and Solid-State Letters, 1998, 1, 66, (b) S. Srinivasan and C. Yang, Private Communication, Princeton University 2000. Wang, J., Wasmus, S., and Savinell, R.F., Journal of the Electrochemical Soceity 1995, 142, 4218. Zhou, W.J, et al. Journal of Power Sources, 2004, 126, 16.

2 3

4
5

Casado-Rivera, E., et al. Journal of the American Chemical Society, 2004, 126(12) 4043.
Adjemian, K.T., et al. Journal of The Electrochemical Society 2002, 149, A256.

Prepr. Pap.-Am. Chem. Soc., Div. Fuel Chem. 2004, 49(2), 663

You might also like