You are on page 1of 18

Journal of Electroanalytical Chemistry 461 (1999) 14 31

Methanol oxidation and direct methanol fuel cells: a selective review1


S. Wasmus a, A. Ku ver b,*
b

Hauptstr. 98, D -53229 Bonn, Germany Degussa AG, ZN Wolfgang, Corporate Research Functions (ZFE), PO Box 1345, D -63457 Hanau, Germany Received 5 January 1998; received in revised form 6 May 1998; accepted 8 May 1998

Abstract A status report on development of the direct methanol fuel cell (DMFC) and relevant fundamental and applied electrochemistry is the objective of this review paper. Emphasis is put on strategies and approaches rather than on individual results. The state-of-the-art in the elds of methodology, catalysis, catalyst characterization, polymer electrolytes as well as assessment and interpretation of cell and electrode performance is described. Todays cell performances are regarded critically. Problems and successes are analyzed. Directions and novel approaches for further work are also suggested. 1999 Elsevier Science S.A. All rights reserved. Keywords: Direct methanol fuel cell; Catalysis; Polymer electrolytes

1. Introduction These days, some types of fuel cells approach commercial feasibility. Fuel cell technology is now becoming applicable for a large variety of technical areas. These include, inter alia, stationary power supplies and electro-traction. For instance, a major German car manufacturer is developing fuel cell powered cars to be offered to the public around the year 2005. These considerable successes would have been impossible without the dedicated efforts of many researchers in the past decades (see Ref. [1]). Steady progress has been made in various elds, such as catalysis, electrolytes, electrode structure, theoretical understanding of gas diffusion and fuel cell engineering. This also extends to direct methanol fuel cells (DMFCs), long being considered as the most difcult fuel cell technology due to methanol crossover and catalytic inefciency.

* Corresponding author. E-mail: Dr Andreas.Kuever@Degussa.de 1 In honour of Professor W. Vielstich on the occasion of his 75th birthday and in appreciation of his contributions to electrochemistry as well as fuel cell development.

Reading the electrochemical literature in the past years dealing with methanol oxidation, oxygen reduction or DMFCs, one can conceive two different trends. A relatively large number of publications contain fundamental research data. The authors strive for a basic understanding of the methanol oxidation and its mechanisms at different catalyst systems, increasingly being done under well-characterized conditions, such as different single crystal orientations and foreign metal clusters on polycrystalline or single crystal surfaces [235]. There have been some excellent reviews on this type of study [3641]. For the purpose of this review, this type of study is called group 1. The other type of study, (group 2) [4291], deals directly with DMFC development. Therein, fuel cell performance data, sometimes together with details on the stoichiometry and pressure of the reactants and on the preparation of membrane electrode assemblies (MEAs) and details on the cells, are published. In some cases, half cell (anode or cathode) performances are also presented. The most likely type of DMFC to be commercialized in the near future seems to be the polymer electrolyte membrane direct methanol fuel cell (PEMDMFC). Consequently, a considerable amount of

0022-0728/99/$ - see front matter 1999 Elsevier Science S.A. All rights reserved. PII S0022-0728(98)00197-1

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431

15

work dealing with membrane electrolyte issues with respect to the special requirement of a DMFC has also been published. The authors feel that the results of the group 1 studies might not be entirely applicable to the development of DMFCs, even though there had been some attempts to generalize results obtained at well-characterized systems towards highly dispersed fuel cell catalysts. According to our opinion, it may be very fruitful for an accelerated progress in fuel cell technology if fundamental research could give more direct inputs into fuel cell development. Among those issues not yet claried, there are questions about the mechanism and product distribution of the methanol oxidation under conditions specic to DMFCs, i.e. temperatures between 60 and 130C, in some cases even up to 200C, various levels of humidity and highly dispersed catalysts. These conditions are frequently very different from those employed in fundamental studies. Other issues may include the mechanisms and effects of the methanol crossover on the cathode performance as well as strategies to avoid cathode performance decay and loss of fuel due to methanol crossover. The effect of catalyst preparation and morphology on the catalytic efciency may be another issue relevant to fuel cell researchers. According to our point of view, this brings into question, why certain ways of catalyst preparation lead to better results than others. Finally, the question of whether there is anything better than PtRu catalyst systems, and if this is true, why are these catalysts better than PtRu, would prove to be of central relevance for an ongoing improvement of anode performance. So far, very few studies have dealt with alternative fuels for direct oxidation fuel cells. It would be interesting to know whether fuels, other than methanol, also available in quantity and at low cost, may lead to improvements in anode performance. This review paper will start with a cursory look at the group 1 studies in which special attention is devoted to the results directly relevant for fuel cell research. It will also emphasize the role of methanol adsorption within the whole scenario of methanol anode catalysis and will clarify some typical misunderstandings. The group 2 studies are segregated into different subjects according to their main purpose. Where suitable, we compare the results of these investigations with those from the group 1 studies. Suggested directions for further work are derived from the discussion of the current state of fuel cell research with attention to the issues we feel to be most urgent, as discussed in the last paragraph. One of the main purposes of this review article is to encourage more investigators to approach the problems directly relevant for fuel cell development. Because many aspects of fuel cell development have left academia, we feel that the current situation makes it timely to collect and discuss industrys and academias viewpoints on

DMFC development in a joint paper. In this paper, we suggest some points in which an intensied co-operation between academia and industry might prove to be fruitful. Additionally, this review article contains a detailed discussion of the techniques available to fuel cell researchers including an overview table, since we consider the choice or development of suitable analytical tools a critical issue for further progress. We feel that the subject of this review article is an appropriate way to pay tribute to the contribution of the person to be honoured to the advance of direct methanol fuel cell research.

2. Methodology Many modern instrumental tools of analytical chemistry are already well established in fuel cell-related electrochemistry. Inter alia, these include various programmed potential or programmed current electrochemical methods, in-situ FTIR, ellipsometry and differential electrochemical mass spectrometry (DEMS). Studies on methanol oxidation using these methods are already too numerous to be cited within the scope of this article. Methods employing X-ray radiation have proved to be invaluable for the characterization of technical catalystseither supported or unsupportedand complete fuel cell electrodes. For instance, these methods are used in the work reported in Refs [42,4547,49 51,57,70,72,73,86,87]. X-ray diffraction (XRD) and Xray photo electron spectroscopy (XPS) are the most frequently used of these methods. Alone or in combination with other methods, such as voltammetry and chemisorption, important information about some characteristics of the catalyst can be derived. These include crystallinity, crystallite size, composition, oxidation state of species and possible interactions of the catalyst with the substrate. Extended X-ray absorption ne structure (EXAFS ) has been used for the characterization of alloyed fuel cell electrocatalysts [92] and should be extremely useful in methanol anode research because small structural changes could be detected [93]. Fourier transform infrared spectroscopy (FTIR ) [42] is among the methods being less common for the characterization of dispersed catalysts and supports. The widely used method, transmission electron microscopy (TEM ) [49,50,59,71] can give some information about particle size, surface area and shape of the metal crystallites, while FTIR is able to give an idea about the nature of the surface groups of carbon supports and on the structure of adsorbate species adsorbed on noble metal clusters. Electrochemical impedance spectroscopy (EIS) is also commonly employed with porous electrodes and can provide valuable geometric and kinetic information, especially on cathode properties. Its im-

16

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431

pact will be discussed later on. It has also to be noted that the study and comparison of the behaviour of complete gas-diffusion electrodes requires additional tools, e.g. porosimetry. These issues will not be discussed here. Generally, the tool box for the static characterization of technical catalysts, supports and electrodes or membrane electrode assemblies (MEAs) is quite complete. It is probably not too unreasonable to assert that any meaningful correlation between catalyst activity and morphology or surface state is not possible without using one or more of these methods. However, the situation is quite different as far as the study of fast dynamic phenomena is concerned. This statement does not apply only for technical catalysis, but also for some key aspects of electrode poisoning under study at smooth and well-dened surfaces. With the advent of optimized interfaces between electrochemistry and ion detection [94], DEMS is a method which is capable of following mass signals in parallel to measured currents at scan rates up to 1 V s 1. This time scale does, however, not generally apply to optical methods, such as in-situ FTIR. Exceptions for some cases could be the combination of DEMS with ellipsometry [95] and the second harmonic generation (SHG) method. It is indeed the case that methodology for the study of dynamic processes has not developed far from its status in the sixties. The toolbox for these studies remains the old repertoire of chronoamperometric and galvanostatic methods which are still capable of delivering new key results, if applied at a sophisticated level [20]. Due to their nature as current or voltage measurements they lack, however, unambiguous information on certain isolated processes or species occurring within a complex reaction scenario. We believe that many academic groups have moved to the static study of phenomena related to the oxidation of higher alcohols and other potential fuels [96] using their above listed spectroscopic methods, because the methodological gap between well-accepted ndings of static investigations and desperately lacking techniques for dynamic investigations on the methanol system could currently not be bridged. The analysis of fuel cell reaction products is another critical issue. In the late eighties and early nineties, there had been some attempts [97 99] using gas chromatography (GC) for this purpose. Unfortunately, GC is far from ideal for this purpose. The time needed for one measurement using gas chromatography is typically of the order of a few minutes. This measuring time, in effect, precludes virtually any real-time monitoring of the formation of fuel cell reaction products. Thus, only discontinuous measurements could be done without the ability to record fast changes of product formation, as being, for instance, likely to occur during voltammetry or potential step techniques. At that time, methods with a continuous product sampling and a time between

formation and detection of the order of a few seconds were needed for fuel cell research. In the mid-nineties, the Case Western Reserve University group was successful in adapting their multipurpose electrochemical mass spectrometry (MPEMS) method to the analysis of the exhaust gases emanating from an operating fuel cell [94,100102]. Using this, fuel cell reaction products formed at the anode and the cathode can be measured on-line. Recently, this was supplemented by an on-line FTIR method employing a similar set-up [103]. Both methods together allow the collection of more complete data, since mass spectrometry and infra-red spectroscopy present quite frequently complementary information. Therefore, as far as the parallel measurement of currents and mass signals is concerned, applied research is now in a similar benign situation as described above for model systems. Infrared spectroscopy is currently employed in another way to get useful insights into electrochemical reactions under fuel cell conditions. Fan et al. [104,105] developed an in-situ Fourier transform infrared diffuse reection spectroscopy (FTIRDRS) with which adsorbed species, as well as the products formed can be studied at working fuel cell electrodes. With the work described in this paragraph, fuel cell research has now reached a status comparable to that at which fundamental electrochemistry had arrived 1015 years earlier: it is possible to study the products and adsorbed intermediates of electrochemical reactions formed in operating prototype fuel cells. In fundamental electrochemistry, DEMS and, with some restrictions, in-situ FTIR and ellipsometry are serving the same purposes. However, these methods require liquid electrolytes and, in the case of in-situ FTIR, smooth electrodes also. For that reason, in-situ FTIR and DEMS were not directly applicable to fuel cell research close to application, so that fuel cell research consequently lagged behind fundamental electrochemistry as far as the study of reaction mechanisms was concerned. We feel that, considering the recent advances in methodology, more academic researchers should shift to fast on-line methods to study the effects linked to electrochemistry under fuel cell conditions. With those methods commonly applied to fundamental electrochemistry already a wealth of data had been generated, but a large amount of work is left to do for fuel cell research as far as intermediates and reaction products generated under technical conditions are concerned. Even though not directly a part of electrochemistry, the characterization of solid polymer electrolytes (SPEs) is nevertheless an important issue for fuel cell research. Consequently, a view of some important methods available for that purpose should be within the scope of this review article. Basically, the membrane electrolyte has to serve three purposes in a DMFC. (i) It has to provide a high proton conductivity; (ii) it needs to have

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431

17

a low permeability for methanol and (iii) it has to be stable under the operating temperatures employed. A number of membrane materials has been used for the hydrogen- or reformate-oxidizing PEM fuel cell so far [106109]. These materials are, in principle, also applicable to the PEMDMFC. No breakthrough has, however, been achieved concerning the minimization of methanol crossover. Using a simplied DEMS set-up, the class of sulphonate-substituted poly-(oxyphenylenes) was recently compared to commercially available membrane materials with respect to methanol crossover [110] and proved to be highly interesting as a material for membrane electrolytes because its methanol-blocking behaviour could be adjusted by adding a sufcient amount of a cross-linking compound. NMR-spectroscopy in its various forms is certainly an important tool. Naon, a peruorosulfonic acid ionomer, has been found of most signicant interest so far (see, for example, Ref. [111] and references cited therein) with one contribution concerning polybenzimidazole (PBI), a novel SPE, which has been investigated by means of solid-state NMR [112]. NMR-spectroscopy has the advantage that species (e.g. water) sorbed inside the SPE membrane, which frequently provide the pathway for protonic conductivity, can be studied without interference from the polymer matrix. Vice versa, the polymer can also be studied separately by using nuclei, such as 13C and 19F, which usually are not abundant in the species sorbed. Using solid-state NMR, even systems which are not mobile enough for standard NMRtechniques can be studied [112]. The applicability of solid-state-techniques should also extend to those systems where the polymer phase is not sufciently mobile in contrast to the polar phase and where an investigation of the former using standard techniques was consequently not possible. Another issue concerning SPEs, is thermal stability in cases where long-term operation at elevated temperatures is desired. The various techniques of thermometry are the methods of choice for that purpose. We restrict our outlook to thermogravimetric analysis/mass spectrometry (TGA/MS) [113,114] and to thermogravimetric analysis/FTIR (TGA/FTIR) [115] since we consider them to be the most elegant of these methods. Using these two methods, it is not only possible to determine the onset temperature of thermal decomposition, but also some information about the decomposition mechanisms can be extracted at the same time by identifying the decomposition products. However, investigators should be cautioned not to apply the results of thermometric methods carelessly to the problem of long-term stability. Since these methods are usually dynamic techniques, the real decomposition under long-term conditions may start at lower temperatures than those indicated by thermometry. In some cases, there might be additional electrochemical degradation so that long-

term testing under operating conditions is still unavoidable despite the methods described in this paragraph. Recent developments in the eld of membrane characterization under operating conditions of the PEM fuel cell include neutron radiography or neutron imaging [116,117] techniques for visualization of uid ow across the MEA. These methods could be of value for DMFC research in the future. The assessment of fuel cell performance seems to be a simple issue in DMFC research. One has to keep in mind, however, that running a test fuel cell does not necessarily mean that the quality of the catalysts is assessed in such an experiment. A large number of interactions of the various components of the test cell, like ow-elds and current collectors with MEA components such as carbon cloth and hydrophobic layers exists, and these interactions could cause dramatic changes in overall performance. Also, the dosing of fuel and oxygen does have an inuence on cell performance. Generally, it is difcult to attribute unequivocally the observed properties to catalyst, interface or operational parameters [118]. Therefore, performance assessment could simply be carried out by measuring currentvoltage curves of MEAs mounted in test cell hardware, but such an experiment will not provide the basis for detailed interpretations. All components and measurement parameters of the system need to be varied in a planned manner, and results have to be analyzed using statistical methods, where possible. This implies that a large amount of measured data is required as a prerequisite for statistical analysis. Further methods have to be employed to derive additional information from the current-voltage characteristics obtained. As routine methods, the measurement of the IR-drop using current interrupt or similar techniques [119,120], and the use of various forms of reference electrodes have evolved for that purpose [54,62,65,94]. The use of a membraneadapted reference electrode has provided more detailed insights into the performance of DMFC components. Meanwhile, the improvement of detail design characteristics of this so-called dynamic hydrogen electrode (DHE) [121] has allowed three electrode fuel cell cyclic voltammetry to be employed for the assessment of methanol crossover in an operating fuel cell [122]. It has to be added, nally, that the use of powerful computers has added the new tool of dynamic simulation to fuel cell research [123]. We will leave a discussion of these methods to the specialists. As of today, we see different tasks arising for industry and academia. Both believe that research using nely dispersed catalysts should be actively pursued. One of academias goals should be, however, to study mechanistic issues on technical catalysts using well-established modern tools and to perfect spectroscopical methods for the study of transient or dynamic phenomena at various catalyst surfaces. Industry is currently

18

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431

pursuing the assessment of catalysts and fuel cell systems using sophisticated test units and auxiliary methods which add further information to current voltage curves. These methods and the peripheral instrumentation of the test units are continuously worked over to achieve meaningful results under realistic conditions. Obviously, due to the proprietary nature of the work, many details of industrial achievements were not published or are obscured in patent texts. It is therefore hard to establish common ground for fuel cell testing, and it is frequently found that nowadays such a common ground is agreed upon only on the basis of a bilateral co-operation treaty or a similar regulation. We are able to conclude that many powerful methods necessary to address the problems of fuel cell development do exist. It is now up to researchers in academia and industry to add their share of methodological development and to use their tools intensively, effectively, andwhere possible in intensive collaboration. Table 1 contains a listing of the methods discussed in this section. In order to aid rapid orientation of the reader, we have included for each of the methods some typical advantages and disadvantages according to our experience.

3. Discussion

3.1. Fundamental and applied work rele6ant for fuel cell de6elopment
The mechanism of the methanol oxidation at various catalyst systems is the most important issue concerning this topic. This information is necessary to avoid a purely empirical approach for the search for better catalysts and, nally, for the improvement of the anode performance. There is a huge amount of work dealing with the methanol oxidation. Our choice of Refs [2 35] inevitably had to be selective. Nevertheless, we feel that our selection is sufciently representative, and that it covers the progress of the last few years. It has to be pointed out, that despite the large number of published articles and other sources, the mechanism of methanol oxidation at platinum alone has not yet been understood in detail. Only recently, Hamnett et al. [124] proposed a mechanistic view on the mobility of adsorbed residues on Pt terraces, but this advanced and plausible view was not fully backed up by experimental results and should therefore be, despite the authors known expertise, be taken as a further speculation on a still unknown situation. On the other hand, according to our perception, it is virtually undisputed that PtRu is a better catalyst than Pt and that the catalytic action of PtRu proceeds mainly by the so-called bifunctional mechanism. Any arguments about this are whether there is an additional

electronic or steric effect. Thus, it is worthwhile to describe the bifunctional mechanism in more detail: Ru sites adsorb oxygen-containing species at 0.20.3 V lower potentials than the pure Pt surface: the adsorbed carbonaceous species are preferentially oxidized at these sites by surface diffusion from sites where adsorption occurs. PtRu pair sites adsorb a more active form of oxygen-containing species than RuRu pairs or Ru clusters. The optimum surface composition of Ru maximizes the PtRu pair sites within the constraints of the optimum ensemble for adsorption of the molecule. In the case of CO and HCOOH, adsorption is equally facile at RuRu, PtPt and PtRu sites and the optimum surface composition is 50 at.% Ru. For methanol, the situation is quite different as its adsorption occurs through consecutive dissociative steps. In terms of geometry, the optimum adsorption site seems to be a C36 Pt ensemble, and the composition which simultaneously maximizes the number of these ensembles and PtRu pairs is 10 at.% Ru [3,8,9,11,12]. It has to be noted, however, that investigations on the methanol adsorption kinetics at different surfaces are scarce, and to the authors knowledge, this issue has never been studied dynamically for materials other than Pt [125]. Moreover, even methanol adsorption at smooth Pt has been a subject of misunderstandings in the past regarding the attribution of anodic currents to dissociative adsorption and CO2 formation or to adsorption only (see Ref. [126] and preceding work in Refs [13,14]). Therefore, some facts about methanol adsorption should be claried here: it was long expected [127], and proved by the Vielstich school using DEMS measurements [128], that positive currents are observed during a positive voltammetric scan of a platinum electrode beginning near the onset of hydrogen evolution in sulfuric acid in the presence of methanol, and that they arise from oxidative methanol dehydrogenation and subsequent adsorption of the fragments formed. The methanol system is, with respect to this current ow, different from a system known as reduced carbon dioxide [129] which yields similar adsorbate species, but does not cause any current ow during build-up. Up to ca. 450 mV versus RHE, methanol adsorption is not accompanied by the evolution of carbon dioxide [128]. It is noteworthy that, using only well-established electrochemical methods, one could develop some basic ideas about adsorbate stoichiometry and structure without the need to look at the surface using spectroscopic methods. It was recently found in work [130], encouraged by the person to be honoured here, that an averaged binding degree as a function of the potential could be obtained using potential steps from 50 mV versus RHE to several potentials and using the Pt surface area obtained employing the method of Biegler et al. [131] for calibration purposes: Upon integration of chronoamperometric data of

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431

19

Table 1 Summary of techniques used for general electrochemical catalysis research and for the characterization of fuel cell components (this list takes into account only applications which are relevant to the subject of this review paper) Technique In-situ FTIR Subject of investigation Products and adsorbed species formed at smooth electrodes Advantages Detection of volatile and non-volatile products Identication of adsorbed species Disadvantages Restriction to smooth electrodes with sufcient reectivity and to liquid electrolytes Sometimes signicant IR -drop in experimental setup due to thin layer techniques Potential sweep techniques are difcult

Separation of adsorbed and non-adsorbed species using IR-light of different polarization

Differential Electrochemical Mass Spectrometry (DEMS)

Volatile products formed at sputtered, lacquer-type and, in some cases, technical electrodes

X-ray radiation methods (e.g. XRD, XPS and EXAFS)

Characterization of technical electrodes

Difcult adjustment of electrochemical cell and optical system Electrodes need frequent polishing Detection of volatile products using Restriction to liquid electrolytes galvanostatic, potentiostatic and potential sweep methods Separation of overlapping electrode No detection of non-volatile processes and reaction pathways products Sweep rates of less than 50 mV s1 Simultaneous formation of a larger number of products may lead to overlapping of mass signals No ready-to-use setup is available on the market for analytical instrumentation Qualitative and, under certain Quantitative analysis needs careful conditions, quantitative analysis of calibration catalyst or support surface composition Different elements can be studied Sensitivity to contamination separately from each other In some cases, the Sometimes difcult interpretation oxidation/bonding state of elements (especially EXAFS, see Refs can be determined [177,178]) Determination of particle size and shape Nature of surface groups can be determined Electrode structure and kinetic parameters can be determined Fully automated setups including complete software package are readily available Substrate must be electronically conductive Sensitive to electrostatic charge Weak sensitivity due to strong IR-absorption of carbon materials Not suited to all types of carbon Equilibrium conditions are required Sensitive to artefacts

Transmission Electron Microscopy (TEM) FTIR

Characterization of technical electrodes Characterization of technical electrodes Characterization of technical electrodes

Electrochemical Impedance Spectroscopy

Multi-Purpose Electrochemical Mass Spectrometry (MPEMS)

Volatile products formed at sputtered, lacquer-type and technical electrodes including fuel cells

DEMS mode with sweep rates of up to 1000 mV s1

Detection of volatile fuel cell reaction products Setup can be adapted to other non-electrochemical applications with minimum reconstruction

Frequently used concept of equivalent circuits may not always work for data analysis Simultaneous formation of a larger number of products may lead to overlapping of mass signals No ready-to-use setup is available on the market for analytical instrumentation

(continued o6erleaf)

20 Table 1 (continued ) Technique On-line FTIR

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431

Subject of investigation Volatile products formed in fuel cells

Advantages Detection of volatile fuel cell reaction products Complementary to MPEMS due to the possibility of product identication in cases where MPEMS shows problems with fragment overlapping Volatile products and adsorbed intermediates can be studied under fuel cell conditions Reactions within the polymer matrix can be studied Identity and behavior of sorbed species can be investigated Sorbed species and polymer matrix can be studied separately In case of solid-state-techniques, even non-mobile phases can be investigated Thermal stability of polymer electrolytes can be assessed In case of TGA/MS and TGA/FTIR, even a thermal degradation mechanism can be established Relatively simple techniques suitable for routine investigations Instrumentation readily available

Disadvantages No ready-to-use setup is available on the market for analytical instrumentation

FTIR Diffuse Reection Spectroscopy NMR-spectroscopy

Volatile products and adsorbates formed in fuel cells Polymer electrolytes

Sample preparation is critical for reproducible results Complex and expensive instrumentation in the case of solid-state-techniques

Thermogravimetric Analysis (TGA) and related techniques (TGA/MS and TGA/FTIR)

Polymer electrolytes

Care has be exercised for applying the results of TGA measurements to long term stability

methanol adsorption in the hydrogen region, one can nd the amounts of charge released during build-up of various equilibrium coverages with organic residues. The potential program and current-time transients thus obtained are given in Fig. 1. It is important to note that the start potential of the potential step necessarily has to be 50 mV versus RHE or less (40 mV was employed here), as methanol adsorption does not occur below 50 mV (see also Ref. [128]). The charges obtained through integration of the transients are plotted versus the step potential in Fig. 2. They are normalized on the basis of a full PtH coverage and are thus given as fractions of an intact PtH monolayer. It is surprising to see that the relation of charge versus potential shows a very steep rise above 250 mV versus RHE, indicating that, for polycrystalline electrodes, a lot more Pt sites are accessible for methanol adsorption at and above this potential. Based on the generalized chemisorption scheme neglecting structural inuences of Pt surface heterogeneities given in Fig. 3, one can conceive that three adjacent Pt sites are initially needed for dissociative methanol adsorption to COHads, but up to two of these sites are liberated again during subsequent reaction steps yielding either linear bonded CO or bridge-bonded CO species. A net ow of four electrons has then occurred per adsorbed CO species.

This implies that a net charge balance of far more than one electron per Pt site (eps) has to be expected for a Pt surface highly covered by CO, and that no correlation between the net ow of charge and the dominating adsorbate species could be drawn for a Pt surface being partially covered with organic residues. A second implication is that methanol adsorption could follow a selffeeding mechanism involving surface diffusion processes in inter-adsorbate conversion processes. This is because three sites only are needed for the rst step to take place, but the release of two non-adjacent Pt sites during the course of the reaction does not provide the C36 symmetry of Pt sites necessary for the next CH3OH COHad adsorption step. It is therefore quite clear that methanol anode electrocatalysis has a lot to do with understanding the methanol adsorption process in detail, e.g. knowing the steric, kinetic and energetic details of all reaction, dissociation and diffusion steps involved. It was shown that methanol adsorption is sensitive to the surface geometry [4,1315]. Therefore, more studies employing single crystal electrodes are encouraged here in order to gain insight into more details of the process. Promising attempts to resolve methanol adsorption dynamically using the new optical second harmonic generation (SHG) method have recently appeared in the literature [132].

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431

21

Fig. 1. Current transients obtained during potential steps from 40 to 350 mV versus RHE in 0.5 M sulfuric acid solution at a polycrystalline Pt foil electrode. The potential program is given in the insert. The methanol concentration was 1 mol l 1.

Having resolved the outstanding task of understanding methanol adsorption, one could further investigate the adsorption properties of alloy or adsorbed metalmodied catalysts and derive conclusions as to whether adsorption kinetics found with these materials arise from altered steric and/or geometric conditions only or from

Fig. 2. Plot of released anodic charge (normalized in terms of PtH coverages) versus potential step width for a Pt foil electrode in 1 M methanol solution. Start potential: 40 mV versus RHE.

changes in surface energies of the multi-component catalysts as well. Some aspects of methanol adsorption at alloy catalysts were recently discussed by Hamnett [133]. Due to the assumed self-feeding mechanism and experimental difculties, rather simple kinetic models as presented in [1315] are to be worked over and rened for multi-component catalysts. First attempts do exist for various PtRu surfaces [3,130]. Understanding the mechanisms behind the formation of oxygenating species is of equal importance. The investigations on oxide layers on noble metals are, however, too numerous to be looked at here. In general, Burke and co-workers have contributed signicantly to this eld [134]. Trying to summarize work on mechanistic issues under realistic conditions, one will nd that relevant publications are scarce. Outstanding is the work of Lin et al. [64], who deal with methanol molecular orientations in the presence of varying amounts of water (see Section 3 on composition of the fuel cell exhaust gas). An interesting approach to investigate the anodic oxidation of methanol using EIS was developed by Horvat et al. [135]. The method is, in principle, capable of delivering data on all time scales. It was found that the reaction rate depended on methanol concentration rather than on concentration of adsorbed intermediates, as for pure Pt. Some criticism has to be made of the use of EIS for the study of methanol anode processes: EIS is dened as a small perturbation imposed on a system in equilibrium. The equilibrium state is therefore an absolutely necessary prerequisite for EIS measurements, and the method cannot be used to investigate multi-step transient processes. It could, however, be applied to investigate pseudo-stationary processes under certain circumstances. The validity of this method has to be examined critically in further studies. On the other hand, EIS is of undoubted use for the characterization of porous electrodes, especially oxygen reducing cathodes. Holze and Vielstich did some pioneering work [136,137]. The work of Arico ` et al. [138] is of importance here as a discussion of impedance spectra provided by them shows quite clearly that the oxygen reduction activity is not severely affected by methanol crossover when working in the PtO potential regime. The authors prove that cathode deactivation in this regime is mainly due to a minor site blocking effect of the Pt sites by organic residues. Theoretical work [139] has supported the view that foreign metals alloyed with Pt activate the formation of OHads which is believed to be one of the oxygen-containing species mentioned above. Considering their calculations, the authors suggest a number of alloying metals which promise to be as active or even more active as Ru alloyed with Pt. We think that this may serve as a direction for further experimental work. However, some of these metals are non-noble. Therefore, we are worried (i) that their alloys with Pt might not show long-term

22

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431

Fig. 3. Scheme of the consecutive dissociative electrosorption of methanol at a Pt electrode.

stability and (ii) that the active surface area of these alloys is uncertain, possibly even changing with time due to a leaching-out effect, which in turn makes a comparison of different catalysts really difcult. For a more detailed discussion of this issue, we refer to the sub-section catalyst morphology. The bifunctional mechanism of PtRu has also been supported by work using fuel cell-type electrodes [72,78]. Generally, it is reported that the onset of methanol oxidation is between 0.2 and 0.25 V versus RHE [55,100,130]. In agreement with this, it was found using ellipsometry that PtRu forms chemisorbed oxygen species at potentials of ca. 0.25 V versus RHE [24]. Therefore, we agree with Hamnett [133] that the limit of the optimization of PtRu catalysts is reached as far as the diminution of the methanol oxidation overpotential is concerned. However, this has nothing to do with the still necessary maximization of the active surface area where a large amount of work is left to do and which is also discussed in the sub-section catalyst morphology. It has also to be noted that the above noted potentials are reported for experiments carried out at or slightly above room temperature. The role of methanol and oxygen chemisorption needs to be thoroughly studied at elevated temperatures to dene clearly the catalytic limits of PtRu. Experimental hints regarding the activity of dispersed catalysts could also be derived from single cell tests (see below). PtSn is another catalyst system which has found considerable interest, probably only second to that in PtRu. Theoretical studies suggest that Sn alloyed into Pt is inactive in generating OHads [140]. In agreement with this, it was found experimentally that PtSn alloys are not active [8,9], while electrosorbed or electrodeposited Sn on Pt is a reasonably good catalyst for methanol oxidation [8,18,25,26,81,141]. However, the conclusion that PtSn alloys are inactive for methanol oxidation is disputed [87], even though the majority of experimental and theoretical evidence suggests otherwise. According to our perception, this contradiction might well be an apparent one since (i) it is sometimes unclear whether a binary system is an alloy

or just a mixture of two metals, and (ii) Sn may leach out under acid conditions being in turn readsorbed electrolytically at Pt sites. Under less well controlled experimental conditions, the latter effect may simulate a catalytic activity of PtSn mixtures or alloys. Generally, it is thought that adsorbed Sn acts in a similar fashion as Ru alloyed into Pt by activation of adsorbed water [8,26,81]. The fact that adsorbed Sn on Pt is active while this is not the case for PtSn alloys might be due to the ionic nature of adsorbed tin [26]. We feel that neither chemisorbed nor electrosorbed foreign metals on Pt are a practical way for fuel cell catalysts since a load variation during fuel cell operation may lead to a change of the anode potential, resulting in desorption of the foreign metal so that its ions may diffuse into the electrolyte, and in turn to the cathode, which is a highly undesirable effect. There are some studies on catalysts for methanol oxidation other than PtRu and PtSn. These catalysts include, inter alia, carbon-supported and unsupported PtWO3x [45,142], PtRuSnW [46], PtRuSnW/C [49], PtRuOs [143] and PtRe [25,76]. Although some of these catalysts exhibit attractive properties, this approach seems to be somewhat arbitrary, particularly since today even binary systems are not completely understood, let alone ternary and quaternary systems with possible metallic interactions which are a multitude of those likely in binary systems. We feel that a clear-cut strategy is necessary in order to seek new catalysts so that the currently used empirical approach can be avoided. This may consist of theoretical work (see for instance Refs [139,140]) for sorting out the unpromising candidates, with the remainder then being evaluated by means of combinatorial analysis [144] to improve screening efciency. We feel that the choice of a common basis for the comparison of different catalysts is a key issue. In the case of smooth and rough electrodesmade from pure Pt, the method of Biegler et al. [131] has proved itself again and again as such a common basis in electrochemical research. The methods proposed by Gilman [145], Tran and Langer [146] or Binder et al. [147] serve a similar purpose. Alloyed catalysts like PtRu are difcult to characterize. As yet,

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431

23

no electrochemical method is commonly accepted for their characterization. In the case of fuel cell catalysts, the situation is far more complex. It is not only the electrochemically active surface area which determines the activity, but also the preparation method, the kind of supporting material and the pretreatment are inuencing catalyst performance. We understand that the latter three may be specic to individual catalyst systems. Together with a commonly agreed method for surface area determination, meaningful comparisons between different catalysts can be made only if the optimum means of preparation and pretreatment for each of the two catalysts (which have to be based on the same carbon support) has been identied. Therefore, it is essential for catalyst research that all investigators apply those methods once they have been found and generally agreed upon. Such an agreement does also have an impact on catalyst or MEA commercialization, as is reected by the patent situation. Details on electrochemical catalyst characterization are discussed in the sub-section catalyst morphology. Due to the increased attention to binary, ternary and quaternary catalyst systems, the interest in methanol oxidation at smooth Pt-only electrodes has faded away slightly. Nevertheless, there are some studies of this kind [47,13,15,20,126] which are worthwhile mentioning within the context of this review article. Their main objective is to understand methanol oxidation at different single crystal orientations and to elucidate the precise nature of adsorbed methanol (e.g. adsorbed CO, COH or similar species). We feel that single crystal research may have laid the foundations for some basic understanding of methanol electrocatalysis. Some authors [15,148] claim that results obtained using single crystals allow conclusions for fuel cell type electrodes. Even though this may be true, we would like to see a more vivid discussion on this subject. Single crystal research has already created a wealth of data from which technical catalysis should prot. The understanding of anion adsorption and surface blocking by adsorbed anions is a typical example [149], which is important as a foundation for the design of components for the phosphoric acid fuel cell (PAFC). The role of adsorption in polymer electrolyte membranes was estimated [150]. In order to transfer results and conclusions from single crystal to technical electrocatalysis, it is essential to establish some basic rules on how this might be done. Issues to be addressed should include the question of whether the behaviour of a polycrystalline material can be understood as the sum of its single crystal-like crystallites. Another question interconnected with the latter is whether neighbouring crystallites show a mutual inuence, thus possibly creating behavior quite different from that of the individual surfaces. Answering these questions would be an extremely useful input into technical electrocatalysis from

those researchers who concentrate on single crystal work.

3.2. Composition of the fuel cell exhaust gas


Until recently, the analysis of fuel cell exhaust gases was a somewhat neglected aspect of fuel cell development. This is the more surprising if one realizes that this has an immediate inuence on fuel cell operation since (i) it is desirable to achieve 100% fuel utilization (i.e. all methanol should be converted to CO2), and (ii) emission of toxic products should be avoided because one of the main sales arguments of fuel cells is that they are an environmentally friendly technology. About 610 years ago, results obtained under conditions similar to those prevailing in operating fuel cells suggested that compounds other than the desired CO2 may well be formed as the main products [9799,151 153]. Formic acid, methylformate, formaldehyde and methylal were detected as major and even main products. Wasmus et al. [100] and Lin et al. [103] were able to identify conditions under which the methanol oxidation in an operating fuel cell yields CO2 as the main product (ca. 98%), i.e. an excess of water and PtRu as the catalyst, while under water-starved conditions, methylal was obtained as the main product with a CO2 yield of the order of 20% or less. No formaldehyde and no formic acid were detected under these conditions [103]. Fan et al. [104,105] essentially conrmed these results with the notable exception that formaldehyde was detected under water-starved conditions. This difference may be explained by the fact that Wasmus et al. [100] and Lin et al. [103] employed methanol in a large excess with phosphoric acid being present in the electrode structure. Thus, all formaldehyde formed electrochemically is captured in this form. The facts that PtRu is more selective towards the formation of CO2 than Pt [100,103] and that PtRu not only catalyzes the formation of CO2, as suggested by the bifunctional mechanism, but also that of methylal (formaldehyde) [100] may indicate that catalysis under fuel cell conditions is different from that observed by fundamental studies. The results of Lin et al. [64] provide further evidence for that perception pointing out that the orientation of the methanol molecules to the electrode surface or to the water molecules is a factor which inuences catalysis and which is different in the cases of liquid-fed and gas-fed electrodes. We feel that the mechanisms of methanol oxidation under fuel cell conditions are less well understood than those at smooth electrodes. Consequently, this requires more work performed under fuel cell conditions. With the methods now available, we would like to see more researchers taking on the challenge. There have been surprisingly few studies on alternative fuels. For instance, Wang et al. [102] investigated

24

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431

ethanol, 1-propanol and 2-propanol under conditions identical to those applied for methanol [100]. While 1-propanol and 2-propanol could be ruled out as candidates due to a low activity and prevalent formation of aldehydes and ketones, ethanol has an activity comparable to that of methanol. In reference [102] the authors showed a cell voltage for a fuel cell fed with ethanol which is at a current density of 300 mA cm 2 just about 50 mV lower than that obtained with a methanol fed fuel cell under otherwise identical conditions. However, the problem of dominant aldehyde formation under all conditions investigated extends also to ethanol. Nevertheless, we consider ethanol an attractive candidate for direct oxidation fuel cells, especially for countries with an already well established infrastructure for ethanol (e.g. Brazil), provided that a catalyst is found which is able to convert ethanol mainly into CO2. Recently, alternative fuels, such as trimethoxymethane, dimethoxymethane and trioxane were also contemplated as alternative fuels [83 85], the patent [83] claiming, for instance, that trimethoxymethane is more active than methanol. Wang et al. [154] found that trimethoxymethane is being hydrolyzed at high temperatures ( \ 120C) and that the resulting mixture of methanol and formic acid is, in effect, oxidized at the anode which explains the better performance of trimethoxymethane. Since formic acid is electrochemically more active than methanol [101], trimethoxymethane may serve as a vehicle to supply formic acid in-situ to the anode, thus, avoiding the corrosive effects of formic acid, even though the full anode performance enhancement of formic acid is not obtained that way. According to our viewpoint, the search for more active fuels than methanol may provide a useful supplement to anode catalyst improvement, provided that these fuels are also available in a sufcient amount and at a reasonably low cost. The methanol crossover is another issue, though well known to affect cathode performance [54], whose mechanism has surprisingly not been investigated in detail until very recently. Wang et al. [94] found that methanol is oxidized heterogeneously in the presence of oxygen and a suitable catalyst (e.g. Pt) leading to a so-called mixed potential effect, and that the cathode is additionally poisoned by methanol. The impact of methanol crossover on the current voltage-characteristics of prototype DMFCs was investigated in numerous studies [62,65,77,122], and was estimated to equal a loss of at least 50 mA cm 2, depending on the conditions. As an aside, it is noteworthy that the effect responsible for cathode performance decay was also used to diminish anode poisoning [66,155]: oxygen oxidizes adsorbed species derived from methanol heterogeneously. However, part of the anode performance gain is compensated by the fact that also at the anode a mixed potential is established.

Two options do exist to overcome cathode performance losses due to methanol crossover: (i) polymer electrolyte membranes have to be used which allow less methanol permeation (see Section 3.4), or (ii) a cathode catalyst has to be found which neither catalyzes the heterogeneous oxidation of methanol nor is poisoned by methanol, such as metal-containing porphyrins [156]. With the recent advances and the prospects of on-going polymer electrolyte improvement, approach (i) seems to be more promising as of today, since approach (ii) may work, but its perspectives concerning cathode catalyst cost, performance and long-term stability are yet uncertain. Furthermore, approach (ii) does not solve the problem of a signicant loss of fuel due to crossover.

3.3. Catalyst morphology


This issue contains several different directions which, however, merge, as far as the objective is concerned: how to extract the best possible performance out of a given amount of noble metals incorporated in a particular catalyst? It can be broken down into catalyst preparation, pretreatment and characterization. Other topics, such as possible catalyst-support interactions and the choice of a suitable carbon support, are involved indirectly. Generally, the majority of investigators agrees that catalysis by small particles is different from that at smooth electrodes [28,157,158]. Especially, Christensen et al. [157] express it very clearly by stating that extrapolation from bulk to particle electrodes is very dangerous. This, again, emphasizes our remarks above, that some rules for the application of single-crystal electrochemistry to technical electrodes have to be established before meaningful conclusions can be drawn. Several routes of catalyst preparation have been attempted [63,68,73,86]. Especially from research for the phosphoric acid fuel cell (PAFC) it is known that stable multi-metal catalysts can be manufactured [159,160]. It is generally agreed that the preparation has an important inuence on catalyst performance. In these studies, the authors found an optimum route, by comparing different ways of catalyst preparation. We consider this an adequate approach, wishing that there might be more co-operation between researchers active in this eld, so that a larger numerical basis for the comparison of different preparation routes with standardized methods is obtained. As of today, comparison of preparation routes reported in different studies is not always simple. Recently, colloid systems were proposed as promising precursors for supported fuel cell catalysts consisting of ultra-ne metal particles showing a mean diameter of only 1.7 9 0.5 nm [161]. Due to the ultralow particle diameter, such systems represent a challenge for comparative studies and their activity needs to

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431

25

be assessed relative to conventional systems in order to be able to judge the progress provided by this novel approach. Catalyst pretreatment (activation) is another factor having great inuence on catalyst performance [67]. It was found that heating a PtRu catalyst in air leads to better results than that in hydrogen due to a surface enrichment of Pt in the presence of hydrogen. Thus, surface enrichment/depletion effects have to be taken into account, not only during the activation procedure, but also during the lifetime of the catalyst. The choice of a suitable carbon support is a factor in addition to those mentioned above which may affect the performance of supported catalysts. Interactions between the catalyst and the carbon support have been identied which modify the catalyst activity [59,61,71]. These interactions are dependent of the nature of the functional groups of the carbon support. For instance, it was found that carbons with lower concentration of acid/base groups [59] and carbons with sulfur- or nitrogen-based functionalities [71] have an enhanced catalytic activity. Therefore, we think it is necessary to agree on an identical carbon support if meaningful comparisons between supported catalysts prepared by different researchers are to be made. The determination of the electrochemically active surface area is a key issue for a reliable and meaningful comparison of different catalysts. Mass normalization, as used commonly so far, has a limited value since it does not take into account the area in contact with the electrolyte. It does, however, represent the economic situation behind fuel cell development. Several approaches do exist to overcome this problem. Liu et al. [162] employed a coulometric analysis of hydrogen desorption waves and applied it to the normalization of methanol oxidation currents. By combining this with mass normalization, they were able to estimate the catalyst utilization. Another possibility is the measurement of the double layer capacitance since it is usually proportional to the active surface area [163]. However, although very useful for non-supported catalysts, both methods are likely to fail in the case of supported catalysts: hydrogen desorption may be masked by the redox currents of the carbon support, and double layer capacitance measurements cannot distinguish between the metal and carbon in contact with the electrolyte. In that case, a combination of particle size analysis by TEM and double layer capacitance measurements by impedance spectroscopy might be a workable expedient. The total area in contact with the electrolyte can be derived from the double layer capacitance, and the portion due to the metal can be estimated from the particle size analysis, since the metal is the catalytically active ingredient so that the carbon area in contact with the electrolyte is irrelevant. Once methods for the determination of the active surface area have been established and agreed upon, it is necessary that they are

applied throughout the work reported in the literature. We understand that different methods are required for supported and unsupported catalysts.

3.4. Electrolytes
The vast majority of polymer electrolyte membrane fuel cell work has been performed using Naon as the electrolyte. As far as DMFCs are concerned, we feel it is time to rethink the adherence to Naon117. Naon117 has, in our view, signicant disadvantages for this application. Its methanol crossover rate of ca. 100 mA cm 2 and the resulting cathode performance decay as well as the loss of fuel are simply not acceptable. Furthermore, in cases where operation beyond 100C is desired, Naon117 neither provides sufcient conductivity nor is there a comfortable thermal stability margin [114], especially in view of possible accidental overheating during operation. Attempts to ensure sufcient conductivity beyond 100C by sorption of phosphoric acid into Naon117 have failed [111]. Thin samples of Naon show advantages with respect to their ionic conductivity. As they only represent a very thin diffusion barrier for methanol, they were initially supposed to increase methanol crossover to unacceptably high levels, but this issue was resolved [60]. As remaining drawbacks, key issues like thermal stability remain disadvantageous. In addition, Naon is currently relatively expensive due to its uorine-based synthesis. For hydrogenoxygen fuel cells, however, it was recently announced that the cost could be brought down to US$ 10 per installed kW [164]. Such a drastic change in cost, combined with the improvement of Naons properties for DMFC applications, could change the commercial situation. There have been several attempts to explore other membrane systems without the disadvantages of Naon for use in DMFCs [75,79]. One promising approach so far is to use basic polymers (polybenzimidazole and polyacrylamide) doped with inorganic acids [79]. In addition to a tenfold decrease in the methanol crossover rate as compared to Naon, polybenzimidazole also shows a very satisfactory thermal stability [113]. These types of polymers are also cheaper than Naon. Therefore, we consider basic polymers doped with inorganic acids as a promising route to obtain new membrane systems for DMFCs with tailor-made properties and with reasonable costs. A method for the preparation of a three-layer structure including a metallic methanol-blocking layer was presented by Cong Pu et al. [75]. As nobody has ever adopted this technique, we consider that this method might not be feasible to produce DMFC membranes. Another approach is to use electropolymerized sulfonated phenols, of which prototypes have already been characterized [110]. It is necessary to investigate the physicochemical properties

26

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431

of all of these polymers, including feasibility studies for MEA manufacture. If the remaining issues of membrane preparation and design of the electrolyte catalyst interface could be resolved, poly-(oxyphenylenes) and other materials could open a new era for crossoverfree DMFCs. The phenomenon of fuel crossover is somehow linked with the thickness of the membrane: if an excess of methanol is present at the anode side, the membrane acts as a diffusion barrier according to Ficks law, meaning that the methanol ux becomes proportional to the membrane thickness. This made researchers initially feel reluctant to employ thin commercial (sometimes prototype) membranes like inter alia Naon 112, Dow XUS and Gore Select. It was shown [60], however, that thin membranes also show potential for DMFC applications. It is currently believed that cathodes with low catalyst utilization for oxygen reduction are capable of oxdizing methanol in a heterogeneously catalyzed gas phase reaction (see above). Therefore, the absolute amount of methanol crossover does not signicantly affect the mixed potential which is established under crossover conditions, as long as electrochemically inaccessible catalyst sites are available for the heterogeneous reactions. Thus, the DMFC could also benet from the reduced internal resistances with thin membranes. The eld of DMFC membrane development is currently very lively and recent achievements using layered structures have been patented [165].

3.5. Components, electrode preparation and structural characterization


A variety of methods for MEA preparation is known from the scientic [166] and patent [167,168] literature. Srinivasan and co-workers prepared Naon-impregnated electrodes showing an extended three-phase boundary. Their invention made it possible to run hydrogen-oxygen fuel cells with noble metal loadings below 1 mg cm 2. The approach to incorporate proton-conducting polymer into the electrode structure was adopted by various groups, e.g. by preparing their electrodes using a paste made of catalyst, binder and Naon solution [52,62]. Gottesfeld et al. [166] then presented the so-called ink method according to which a catalyst suspension is applied directly to the membrane. Up to now, this method has found many variations and seems to be universal for the preparation of large electrodes. Using a catalyst ink makes it possible either to print, brush or spray-coat the catalyst onto the membrane. Some of these methods can be scaled up easily and have led to the manufacture of electrode sizes of at least 25 cm2 [169]. As the mechanical and transport properties of the catalytic layer depend on the ink composition and the application method, various

contact methods were developed, e.g. by using variously impregnated and further treated carbon papers or carbon cloths. There are two reasons why methanol PEM anodes need to have internal structures different from hydrogen PEM anodes: (i) the wetting nature of methanol and its larger molecular size (compared to hydrogen) and (ii) the fact that the gaseous product CO2 has to be removed from the catalyst sites and the electrode pore system. Consequently, methanol anodes have to have a more open structure for both reasons [80,130]. As already mentioned, EIS is capable of delivering information on geometric properties of porous electrodes. Therefore, it is recommended that EIS data of DMFC anodes is accumulated systematically in order to nd a correlation between EIS spectral ngerprint features, DMFC anode pore systems, and anode performance. The EIS experiment could be carried out in the absence of methanol, since only geometric features are to be evaluated. Closely related to the structure of the gas diffusion electrodes is the design of the remaining cell components, such as ow elds and current collectors. The ow eld and the diffusion layer have to provide fuel or oxygen to the electrodes in a uniform manner and both also have to help drain-off of reaction products. As far as the hydrogenoxygen PEM fuel cell is concerned, the design of the ow eld was identied as one of the key technologies in fuel cell engineering, for which fundamental theoretical relations were laid down [170] and technically feasible designs were patented [171 173]. Design principles providing an even fuel distribution and an effective drain-off of liquids are applicable to the DMFC as well. It is therefore expected that DMFC engineering could vastly prot from well-established PEM fuel cell technology and DMFC commercialization could start immediately after specic membrane and catalytic issues have been resolved.

3.6. Analysis of the state -of -the -art in single cell DMFC performance
Besides the technical difculties of DMFC testing outlined above, interpretation of published performance data remains a critical issue, too. The DMFC will possibly become the preferred development target as a power source for fuel cell-powered electric vehicles if a performance target of 200250 mW cm 2 (proposed in Refs [52,82,175]) can be met by experimental systems operating on methanol and air below 100C. Hogarth and Hards [174] recently presented a compilation of single cell data achieved by academic and industrial groups. It is agreed here that the results of all leading groups are promising; nevertheless, the performances shown need to be demonstrated using lower noble metal loadings and less drastic reaction condi-

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431

27

tions. Most of these groups have achieved voltages between 400 and almost 600 mV at a current density of 400 mA cm 2, equal to power densities between 166 and 240 mW cm 2. In all cases reported [52,60,82,91], PtRu was used as the anode catalyst. Therefore, the question already stated earlier in this paper, as to where the catalytic limits (in W cm 2 at moderate loadings signicantly below 4 mg cm 2) of the currently preferred catalyst system PtRu are, should be asked again in the light of the data presented by leading groups. Without going too deeply into details, one could see that all these groups except Newcastle University achieved their high performances using off-the-shelf supported or noble metal black catalysts. The achievements cited were possible because rather drastic reaction conditions were used in all cases, i.e. pressurized oxygen at the cathode side (with one exception) and high temperatures between 100 and 130C. Also, very high noble metal loadings with perhaps poor catalyst utilization were employed. It is therefore obvious that the true performance limits of PtRu are obscured behind low catalyst utilizations and temperature-accelerated kinetics. It is quite realistic to conclude that those performances reported are not valuable for a ranking of anodes because of a leveling by the conditions explained above. It is, however, possible to state some general assumptions on the relative performances of the single electrodes: it is well-known (Degussa fuel cell team, unpublished results) that cathode performances of pressurized hydrogen-oxygen PEM fuel cells are usually in the region of 800 mV at 400 mA cm 2 when operated at 8090C. Assuming a crossover-rate leading to a voltage loss of ca. 100 mV, measured DMFC performances of 500 mV at 400 mA cm 2 imply that the anode potentials should be in the range of 150 250 mV under these conditions, depending on the actual cathode potential. These values seem to contradict the earlier-stated value of 250 mV, but it has to be noted here that we are now referring to DMFC operating temperatures, not room temperature. Although contributions boasting of remarkable DMFC performances have appeared almost ad nauseam, no precise statement could be provided on the performances of the respective anodes. A typical example is the paper published by Surampudi et al. [82] where interesting cell performances are shown, but anode performance data are presented which were not measured in the PEM fuel cell system and which are additionally too poor to explain the good system performances shown in the paper. It has to be ascertained here that DMFC development goals had been raised by various authors without having dened the particular target levels for anode and cathode performances. Certain methods for the distinct analysis of current

voltage curves do exist in industrial laboratories; e.g. it is possible to judge upon the relative progress in anode development by using experimental anodes in MEAs, combining them always with the same cathode materials and employing always the same measurement conditions (p, T etc.). But the absolute status and targets have seldom been stated by the major developers. This is the case for Siemens [91], which has, in terms of DMFC performance, for a long time been the worlds leading organization. The whole scientic community was never aware of the quality of their single electrodes, because they were the only group working at 130C, and they never published single electrode data. Moreover, it is absolutely unknown whether the oxygen cathodes potential at 130C was shifted far away from the working potentials stated above, or if the excellent performance was caused by a performance gain due to the acceleration of methanol oxidation kinetics at high temperatures (for an illustration of the inuence of temperature see [82]). We would therefore strongly recommend the use of reference electrodes when data are to be published. As todays basis for systems comparisons [175] between the reformateair fuel cell (RAFC) and the DMFC consists of working temperatures between 70 and 100C, the ranking list of DMFC performances should be re-considered critically; the top performer in single cell DMFC performance could certainly be found in the list published by Hogarth and Hards, but this title should be assigned to those developers achieving the economically most attractive performances as regards the operating conditions and the noble metal loadings, rather than to those showing the best currentvoltage curves.

4. Conclusions We feel that there is a need for further work in the following areas:

4.1. Catalysis
It has to be evaluated whether there are any better catalysts than PtRu. The best strategy for this search is a combined approach of theoretical and fundamental electrochemical studies and a modern screening method like combinatorial analysis. A generally agreed method for active surface area normalization is an indispensable prerequisite for this. It is desirable that catalyst manufacture and physical data be well-described in the open literature in order to make inter-catalyst comparisons possible. The methodological gap outlined earlier in this paper should be closed in the near future. With all necessary methods available, the mechanisms of the methanol oxidation under real fuel cell

28

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431

conditions have to be studied. This involves catalyst performance under these specic conditions, fuel utilization and the avoidance of toxic product emission. In order to reduce the usage of precious metals and, thus, cost, the best possible performance has to be extracted from a given amount of catalyst. This involves mainly catalyst preparation, pretreatment, electrode preparation and, in the case of supported catalysts, also the choice of a suitable support. We feel that more systematic and better co-ordinated efforts than before could be extremely helpful to reach this goal. The rst step in achieving this is jointly to nd a clear denition of the current state-of-the-art in anode catalysis under realistic conditions. Then, electrodes capable of delivering high performances should be monitored in long-term studies and their morphology and chemical composition should also be followed as a function of time in order to make sure that catalyst compositions found to perform well are able to sustain prolonged operation. At this stage, it will be necessary to monitor the components of the anode exhaust gas as well. It is worthwhile to note that electronic devices could take the role of bifunctional catalysts in the future: it was recently claimed [176] that platinum anode poisoning by CO in hydrogen-oxidizing fuel cells could be circumvented by using very short current pulses capable of removing poisoning residues and restoring the initial catalytic activity. Similar approaches are likely to be effective in DMFCs as well. Therefore, DMFC research should be target-oriented rather than sticking to established approaches.

able cell (or stack) components and materials suitable to build maintenance-free and cost-effective DMFCdriven systems.

References
[1] W. Vielstich, Fuel Cells, Wiley, London, 1970. [2] H.A. Gasteiger, N. Markovic, P.N. Ross, E.J. Cairns, J. Phys. Chem. 97 (1993) 12020. [3] H.A. Gasteiger, N. Markovic, P.N. Ross, E.J. Cairns, J. Electrochem. Soc. 141 (1994) 1795. [4] X.H. Xia, T. Iwasita, F. Ge, W. Vielstich, Electrochim. Acta 41 (1996) 711. [5] T. Iwasita, W. Vielstich, E. Santos, J. Electroanal. Chem. 229 (1989) 367. [6] S. Wilhelm, T. Iwasita, W. Vielstich, J. Electroanal. Chem. 238 (1987) 383. [7] B. Bittins-Cattaneo, E. Santos, W. Vielstich, U. Linke, Electrochim. Acta 33 (1988) 1499. [8] P.N. Ross Jr., H.A. Gasteiger, Proceedings of the rst international symposium on new materials for fuel cell systems, Montreal, July 9 13, 1995. [9] K. Wang, H.A. Gasteiger, N.M. Markovic, P.N. Ross Jr., Electrochim. Acta 41 (1996) 2587. [10] H.A. Gasteiger, N.M. Markovic, P.N. Ross Jr., Catal. Lett. 36 (1996) 1. [11] N.M. Markovic, H.A. Gasteiger, P.N. Ross Jr., X. Jiang, I. Villegas, M.J. Weaver, Electrochim. Acta 40 (1995) 91. [12] H.A. Gasteiger, N.M. Markovic, P.N. Ross Jr., E.J. Cairns, Electrochim. Acta 39 (1994) 1825. [13] K. Franasczuk, E. Herrero, P. Zelenay, A. Wieckowski, J. Wang, R.I. Masel, J. Phys. Chem. 96 (1992) 8509. [14] E. Herrero, K. Franasczuk, A. Wieckowski, J. Phys. Chem. 98 (1994) 5074. [15] A. Wieckowski, W. Chrzanowski, E. Herrero, Proceedings of the rst international symposium on new materials for fuel cell systems, Montreal, July 9 13, 1995, p. 326. [16] R. Ortiz, O.P. Marquez, J. Marquez, C. Gutierrez, J. Phys. Chem. 100 (1996) 8389. [17] M.M. Hefny, S. Abdel-Wanees, Electrochim. Acta 41 (1996) 1419. [18] B. Bittins-Cattaneo, T. Iwasita, J. Electroanal. Chem. 238 (1987) 151. [19] M.M.P. Janssen, J. Moolhuysen, Electrochim. Acta 21 (1976) 869. [20] T.D. Jarvi, S. Sriramulu, E.M. Stuve, J. Phys. Chem. B 101 (1997) 3649. [21] M. Watanabe, Y. Genjima, K. Turumi, J. Electrochem. Soc. 144 (1997) 423. [22] H. Matsui, A. Kunugi, J. Electroanal. Chem. 292 (1990) 103. [23] K.-I. Machida, A. Fukuoka, M. Ichikawa, M. Enyo, J. Electrochem. Soc. 138 (1991) 1958. [24] E. Ticianelli, J.G. Beery, M.T. Paffett, S. Gottesfeld, J. Electroanal. Chem. 258 (1989) 61. [25] K.J. Cathro, J. Electrochem. Soc. 116 (1969) 1609. [26] J. Sobkowski, K. Franasczuk, A. Piasecki, J. Electroanal. Chem. 196 (1985) 145. [27] H. Ogasawara, M. Ito, Chem. Phys. Lett. 245 (1995) 304. [28] P.A. Christensen, A. Hamnett, G.L. Troughton, J. Electroanal. Chem. 362 (1993) 207. [29] H.A. Gasteiger, N.M. Markovic, P.N. Ross, E.J. Cairns, Meeting of the Electrochemical Society, Honolulu, Hawaii, 1993, Abstract 1747, p. 2385.

4.2. Electrolytes
New electrolytes have to be found whose properties are better suited for DMFCs, having e.g. drastically diminished methanol crossover rates, and which are available at low costs. Experimental membranes released for hydrogen-oxygen fuel cell testing should also be immediately evaluated in DMFCs, but we believe that a membrane exclusively developed for the DMFC will be the commercially viable product. An excellent guideline for membrane investigations is given in Ref. [106].The need for long-term studies (see above) also applies to membrane testing.

4.3. Components
Provided that success in the elds mentioned above is achieved, we see a good chance for DMFCs on the emerging future markets for mobile electric power sources. Coming closer to realization, all research work should be performed with costs in mind, since this issue will be decisive for the marketing efforts within the next decade. Therefore, it is high time to think about suit-

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431 [30] T. Iwasita, F.C. Nart, B. Lo pez, W. Vielstich, Electrochim. Acta 37 (1992) 2361. [31] R. Ianniello, V.M. Schmidt, U. Stimming, J. Stumper, A. Wallau, Electrochim. Acta 39 (1994) 1863. [32] E.P.M. Leiva, M.C. Giordano, J. Electrochem. Soc. 130 (1983) 1305. [33] M. Castro Luna, M.C. Giordano, A.J. Arv a, J. Electroanal. Chem. 259 (1989) 173. [34] A.V. Tripkovic, K.D.J. Popovic, Electrochim. Acta 41 (1996) 2385. [35] N. Markovic, P.N. Ross, J. Electroanal. Chem. 330 (1992) 499. [36] N.A. Hampson, M.J. Willars, B.D. McNicol, J. Power Sources 4 (1979) 191. [37] R. Parsons, T.J. VanderNoot, J. Electroanal. Chem. 257 (1988) 9. [38] T. Iwasita-Vielstich, in: C. Tobias, H. Gerischer, (Eds.), Advances in Electrochemical Science and Engineering, Vol. 1, VCH, Weinheim, 1990, p. 127. [39] D.S. Cameron, G.A. Hards, D. Thompsett, Proc. Electrochem. Soc. 1992, p. 10. [40] W. Vielstich, T. Iwasita, in: G. Ertl, H. Kno zinger, J. Weitkamp (Eds.), Handbook of Heterogeneous Catalysis Vol. 4, Wiley, Chichester, 1997. [41] K. Kordesch, G. Simader, Fuel Cells and Their Applications, VCH, Weinheim, 1996. [42] T. Torre, A.S. Arico ` , V. Alderucci, V. Antonucci, N. Giordano, Appl. Catal. 114 (1994) 257. [43] N. Giordano, E. Passalacqua, L. Pino, A.S. Arico ` , V. Antonucci, M. Vivaldi, K. Kinoshita, Electrochim. Acta 36 (1991) 13. [44] M.P. Hogarth, J. Munk, A.K. Shukla, A. Hamnett, J. Appl. Electrochem. 24 (1994) 85. [45] A.K. Shukla, M.K. Ravikumar, A.S. Arico ` , G. Candiano, V. Antonucci, N. Giordano, A. Hamnett, J. Appl. Electrochem. 25 (1995) 528. [46] A.S. Arico ` , Z. Poltarzewski, H. Kim, A. Morana, N. Giordano, V. Antonucci, J. Power Sources 55 (1995) 159. [47] A.K. Shukla, K.V. Ramesh, R. Manoharan, P.R. Sarode, S. Vasudevan, Ber. Bunsenges. Phys. Chem. 89 (1985) 1261. [48] N. Giordano, P. Staiti, S. Hocevar, A.S. Arico ` , Electrochim. Acta. 41 (1996) 397. [49] A.S. Arico ` , P. Creti, N. Giordano, V. Antonucci, P.L. Antonucci, A. Chuvilin, J. Appl. Electrochem. 26 (1996) 959. [50] A.S. Arico ` , P. Creti, H. Kim, R. Mantegna, N. Giordano, V. Antonucci, J. Electrochem. Soc. 143 (1996) 3838. [51] A.K. Shukla, M.K. Ravikumar, A. Roy, S.R. Barman, D.D. Sarma, A.S. Arico ` , V. Antonucci, L. Pino, N. Giordano, J. Electrochem. Soc. 141 (1994) 1517. [52] A.K Shukla, P.A. Christensen, A. Hamnett, M.P. Hogarth, J. Power Sources 55 (1995) 87. [53] W. Vielstich, A. Ku ver, M. Krausa, A.C. Ferreira, K. Petrov, S. Srinivasan, Proceedings on the 183rd Meeting of the Electrochemical Society, 93 8 (1993) p. 269. [54] M.K. Ravikumar, A.K. Shukla, J. Electrochem. Soc. 143 (1996) 2601. [55] S. Wasmus, W. Vielstich, J. Appl. Electrochem. 23 (1993) 120. [56] A.S. Arico ` , P. Creti, R. Mantegna, P.L. Antonucci, N. Giordano, V. Antonucci, 1996 fuel cell Seminar, November 1720, Orlando, Florida, Abstracts, p. 678. [57] A.S. Arico ` , V. Antonucci, N. Giordano, A.K. Shukla, M.K. Ravikumar, A. Roy, S.R. Barman, D.D. Sarma, J. Power Sources 50 (1994) 295. [58] T. Frelink, W. Visscher, J.A.R. van Veen, J. Electroanal. Chem. 382 (1995) 65. [59] M. Uchida, Y. Aoyama, N. Tanabe, N. Yanagihara, N. Eda, A. Ohta, J. Electrochem. Soc. 142 (1995) 2572.

29

[60] X. Ren, M.S. Wilson, S. Gottesfeld, J. Electrochem. Soc. 143 (1996) 12. [61] A. Aramata, T. Kodera, M. Masuda, J. Appl. Electrochem. 18 (1988) 577. [62] A. Ku ver, I. Vogel, W. Vielstich, J. Power Sources 52 (1994) 77. [63] M.P. Hogarth, P.A. Christensen, A. Hamnett, Proceedings of the rst international symposium on new materials for fuel cell systems, Montreal, July 9 13, 1995. [64] A.S. Lin, A.D. Kowalak, W.E. OGrady, J. Power Sources 58 (1996) 67. [65] A. Ku ver, I. Vogel, W. Vielstich, 45th Meeting of the International Society of Electrochemistry, Porto 1994, Abstract OIV-4. [66] H. Tobias, M.T. Paffett, P.A. Pappin, J. Valerio, S. Gottesfeld, Proceedings of the Direct Methanol/Air Fuel Cell Workshop Washington, D.C., May 1990, Electrochemical Society Softbound Series. [67] B.D. McNicol, R.T. Short, J. Electroanal. Chem. 81 (1977) 249. [68] M. Watanabe, M. Uchida, S. Motoo, J. Electroanal. Chem. 229 (1987) 395. [69] D.R. Shin, D.R. Jung, C.H. Lee, Y.G. Chun, 1996 fuel cell Seminar, November 17 20, Orlando, Florida, Abstracts, p. 651. [70] A.S. Arico ` , P. Creti, Z. Poltarzewski, R. Mantegna, H. Kim, N. Giordano, V. Antonucci, Mater. Chem. Phys. 47 (1997) 257. [71] S.C. Roy, P.A. Christensen, A. Hamnett, K.M. Thomas, V. Trapp, J. Electrochem. Soc. 143 (1996) 3073. [72] J.B. Goodenough, A. Hamnett, B.J. Kennedy, R. Manoharan, S.A. Weeks, J. Electroanal. Chem. 240 (1988) 133. [73] J.B. Goodenough, A. Hamnett, B.J. Kennedy, R. Manoharan, S.A. Weeks, Electrochim. Acta 35 (1990) 199. [74] J.S. Wainright, J.-T. Wang, D. Weng, R.F. Savinell, M. Litt, J. Electrochem. Soc. 142 (1995) 121. [75] C. Pu, W. Huang, K.L. Ley, E.S. Smotkin, J. Electrochem. Soc. 142 (1995) 119. [76] K.J. Cathro, J. Electrochem. Soc. 5 (1967) 441. [77] D.L. Maricle, B.L. Murach, L.L. Van Dine, 36th Power Sources Conference, Cherry Hill, NJ, 1994, p. 99. [78] S. Swathirajan, Y.M. Mikhail, J. Electrochem. Soc. 138 (1991) 1321. [79] J.-T. Wang, J.S. Wainright, R.F. Savinell, M. Litt, J. Appl. Electrochem. 26 (1996) 751. [80] M. Watanabe, M. Uchida, S. Motoo, J. Electroanal. Chem. 199 (1986) 311. [81] T. Frelink, W. Visscher, J.A.R. van Veen, Electrochim. Acta 39 (1994) 1871. [82] S. Surampudi, S.R. Narayanan, E. Vamos, H.A. Frank, G. Halpert, A. LaConti, J. Kosek, G.K. Surya Prakash, G.A. Olah, J. Power Sources 47 (1994) 377. [83] S. Surampudi, S.R. Narayanan, E. Vamos, H.A. Frank, G. Halpert, G.A. Olah, G.K. Surya Prakash, U.S. Pat. 5,599,638. [84] S.R. Narayanan, E. Vamos, S. Surampudi, et al., J. Electrochem. Soc. 144 (1997) 4195. [85] S.R. Narayanan, W. Chun, T.I. Valdez et al., 1996 fuel cell Seminar, November 17 20, Orlando, Florida, Abstracts, p. 525. [86] R. Pattabiraman, Proceedings of the rst international symposium on new materials for fuel cell systems, Montreal 1996, 362. [87] Z. Wei, H. Guo, Z. Tang, J. Power Sources 58 (1996) 239. [88] L.L. Van Dine, D.L. Maricle, U.S. Pat. 5,573,866. [89] J.A. Kosek, C.C. Cropley, A.B. LaConti, U.S. Pat. 5,523,177. [90] M.H. Litt, R. Savinell, WO PCT 96/13872. [91] M. Waidhas, W. Drenckhahn, W. Preidel, H. Landes, J. Power Sources 61 (1996) 91. [92] L. Tro ger, A. Freund, P. Albers, K. Seibold, G. Prescher, Ber. Bunsenges. Phys. Chem. 101 (1997) 851. [93] P.G. Allen, S.D. Conradson, M.S. Wilson, S. Gottesfeld, I.D. Raistrick, J. Valerio, M. Lovato, Electrochim. Acta 39 (1994) 2415.

30

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431 [127] V.S. Bagotzky, Y. Vassiliev, O.A. Khazova, J. Electroanal. Chem. 81 (1977) 229. [128] M. Krausa, W. Vielstich, J. Electroanal. Chem. 379 (1994) 307. [129] A. Ku ver, W. Vielstich, D. Kitzelmann, J. Electroanal.Chem. 353 (1993) 255. [130] A. Ku ver, Doctoral Thesis, Bonn University, 1996. [131] T. Biegler, D.A.J. Rand, R. Woods, J. Electroanal. Chem. 29 (1971) 269. [132] I.T. Bae, J. Phys. Chem. 100 (1996) 14081. [133] A. Hamnett, Catalysis Today 38 (1997) 445. [134] L.D. Burke, J.K. Casey, J.A. Morrissey, J.F. OSullivan, J. Appl. Electrochem. 24 (1994) 30. [135] R. Horvat, M. Metikos-Hukovic, P. Pervan, M. Milun, 47th ISE meeting, Balatonfu red 1996, Abstract P7b-14. [136] R. Holze, W. Vielstich, J. Electrochem. Soc. 131 (1984) 2298. [137] R. Holze, W. Vielstich, Electrochim. Acta 29 (1984) 607. [138] A.S. Arico ` , V. Antonucci, V. Alderucci, E. Modica, N. Giordano, J. Appl. Electrochem. 23 (1993) 1107. [139] A.B. Anderson, E. Grantscharova, S. Seong, J. Electrochem. Soc. 143 (1996) 2075. [140] A.B. Anderson, E. Grantscharova, P. Shiller, J. Electrochem. Soc. 142 (1995) 1880. [141] M. Watanabe, Y. Furuuchi, S. Motoo, J. Electroanal. Chem. 191 (1985) 367. [142] P.K. Shen, A.C.C. Tseung, J. Electrochem. Soc. 141 (1994) 3082. [143] K.L. Ley, R.X. Liu, C. Pu, Q.B. Fan, N. Leyarovska, C. Segre, E.S. Smotkin, J. Electrochem. Soc. 144 (1997) 1543. [144] T.E. Mallouk, E. Reddington, C. Pu, K.L. Ley, E.S. Smotkin, 1996 fuel cell Seminar, November 17 20, Orlando, Florida, Abstracts, p. 686. [145] S. Gilman, J. Electroanal. Chem. 7 (1964) 382. [146] T.D. Tran, S.H. Langer, Anal. Chem. 65 (1993) 1805. [147] H. Binder, A. Ko hling, K. Metzelthin, G. Sandstede, M.-L. Schrecker, Chem. Ing. Tech. 40 (1968) 586. [148] N. Markovic, H. Gasteiger, P.N. Ross, J. Electrochem. Soc. 144 (1997) 1591. [149] M. Weber, F.C. Nart, I.R. de Mora es, T. Iwasita, J. Phys. Chem. 100 (1996) 19933. [150] A. Parthasarathy, C.R. Martin, S. Srinivasan, J. Electrochem. Soc. 138 (1991) 916. [151] K.-I. Ota, Y. Nakagawa, M. Takahashi, J. Electroanal Chem. 179 (1984) 179. [152] I. Yamanaka, K. Otsuka, Chem. Lett. 753 (1988) 756. [153] H. Nakajima, H. Kita, Electrochim. Acta 33 (1988) 521. [154] J.T. Wang, W.F. Lin, M. Weber, S. Wasmus, R.F. Savinell, Electrochim. Acta, submitted. [155] B. Bittins-Cattaneo, S. Wasmus, B. Lo ` pez-Mishima, W. Vielstich, J. Appl. Electrochem. 23 (1993) 625. [156] K. Sundmacher, O. Nowitzki, U. Hoffmann, Chem. Ing. Tech. 69 (1997) 8197. [157] P.A. Christensen, A. Hamnett, J. Munk, G.L. Troughton, J. Electroanal. Chem. 370 (1994) 251. [158] T. Frelink, W. Visscher, J.A.R. van Veen, J. Electroanal. Chem. 382 (1995) 65. [159] A. Freund, J. Lang, Th. Lehmann, K.A. Starz, Catalysis Today 27 (1996) 279. [160] K.E. Deluca, F.J. Luczak, PCT, WO 94/10715 (to International Fuel Cells). [161] T.J. Schmidt, M. Noeske, H.A. Gasteiger, R.J. Behm, P. Britz, H. Bo nnemann, J. Electrochem. Soc. 145 (1998) 925. [162] R.X. Liu, K. Triantallou, L. Liu, C. Pu, C. Smith, E.S. Smotkin, J. Electrochem. Soc. 144 (1997) 148. [163] S. Wasmus, Doctoral Thesis, Bonn University, 1992. [164] Hydrogen and Fuel Cell Letters XIII, No. 4, April 1998, p. 1. [165] PCT, WO 97/14189 (to DuPont). [166] M.S. Wilson, S. Gottesfeld, J. Electrochem. Soc. 139 (1992) 28.

[94] J.-T. Wang, S. Wasmus, R. Savinell, J. Electrochem. Soc. 143 (1996) 1233. [95] Z. Jusys, H. Baltruschat, Joint meeting of the Electrochemical Society and the International Society of Electrochemistry, Paris 1997, Abstract No. 909. [96] T. Iwasita, E. Pastor, Electrochim. Acta 39 (1994) 531. [97] I. Yamanaka, K. Otsuka, Electrochim. Acta 34 (1989) 211. [98] H. Nakajima, H. Kita, Electrochim. Acta 33 (1988) 521. [99] R. Liu, P.S. Fedkiw, J. Electrochem. Soc. 139 (1992) 3514. [100] S. Wasmus, J.-T. Wang, R.F. Savinell, J. Electrochem. Soc. 142 (1995) 3825. [101] M. Weber, J.-T. Wang, S. Wasmus, R.F. Savinell, J. Electrochem. Soc. 143 (1996) 158. [102] J. Wang, S. Wasmus, R.F. Savinell, J. Electrochem. Soc. 142 (1995) 4218. [103] W.F. Lin, J.T. Wang, R.F. Savinell, J. Electrochem. Soc. 144 (1997) 1917. [104] Q. Fan, C. Pu, K.L. Ley, E.S. Smotkin, J. Electrochem. Soc. 143 (1996) 21. [105] Q. Fan, C. Pu, E.S. Smotkin, J. Electrochem. Soc. 143 (1996) 3053. [106] C. Heitner-Wirguin, J. Membr. Sci. 120 (1996) 1. [107] J.T. Hinatsu, M. Mizuhata, H. Takenaka, J. Electrochem. Soc. 141 (1994) 1493. [108] G.S. Kumar, M. Raja, S. Parthasarathy, Electrochim. Acta 40 (1995) 285. [109] M. Uchida, Y. Aoyama, N. Eda, A. Ohta, J. Electrochem. Soc. 142 (1995) 463. [110] A. Ku ver, K. Potje-Kamloth, Electrochim. Acta, in print. [111] S. Wasmus, A. Valeriu, G.D. Mateescu, D.A. Tryk, R.F. Savinell, Solid State Ionics 80 (1995) 87. [112] S. Wasmus, R.F. Savinell, H. Moaddel, M.H. Litt, C. Rogers, A. Valeriu, G.D. Mateescu, D.A. Tryk, W.A. Daunch, P.L. Rinaldi, 187th Meeting of the Electrochemical Society, Reno/ Nevada, May 21 26 1995. [113] S.R. Samms, S. Wasmus, R.F. Savinell, J Electrochem. Soc. 143 (1996) 1225. [114] S.R. Samms, S. Wasmus, R.F. Savinell, J. Electrochem. Soc. 143 (1996) 1498. [115] B. Gupta, J.G. Higheld, G.G. Scherer, J. Appl. Polym. Sci. 51 (1994) 1659. [116] O.A. Velev, S. Srinivasan, A.J. Appleby, I. Carron, S. Gutner, A. Sanchez, Joint meeting of the Electrochemical Society and the International Society of Electrochemistry, Paris 1997, Abstract No. 193. [117] R.J. Bellows, M.Y. Lin, M. Arif, D. Jacobson, Joint meeting of the Electrochemical Society and the International Society of Electrochemistry, Paris 1997, Abstract No. 962. [118] A.S. Arico ` , P. Creti, H. Kim, E. Modica, P.L. Antonucci, V. Antonucci, Joint meeting of the Electrochemical Society and the International Society of Electrochemistry, Paris 1997, Abstract No. 70. [119] F. Bu chi, A. Marek, G.G. Scherer, J. Electrochem. Soc. 142 (1995) 1895. [120] T.A. Zawodzinski Jr., T.E. Springer, J.D.R. Jestel, C. Lopez, J. Valerio, S. Gottesfeld, J. Electrochem. Soc. 140 (1993) 1981. [121] J. Giner, J. Electrochem. Soc. 111 (1964) 376. [122] A. Ku ver, W. Vielstich, J. Power Sources, in print. [123] K. Scott, W. Taama, J. Cruickshank, J. Power Sources 65 (1997) 159. [124] A. Hamnett, P.A. Christensen, G. Troughton, M.P. Hogarth, Joint meeting of the Electrochemical Society and the International Society of Electrochemistry, Paris 1997, Abstract No. 1079. [125] A. Papoutsis, J.-M. Le ger, C. Lamy, J. Electroanal. Chem. 234 (1987) 315. [126] W. Vielstich, X.H. Xia, J. Phys. Chem. 99 (1995) 10421.

S. Wasmus, A. Ku 6er / Journal of Electroanalytical Chemistry 461 (1999) 1431 [167] M.S. Wilson, PCT, WO 92/15121. [168] J. Denton, J.M. Gascoyne, R.J. Potter, EP 0 791 974 A1. [169] A. Fischer, H. Wendt, R. Zuber, EP 0 797 265 A2 (applicant: Degussa AG). [170] A.C. West, T.F. Fuller, J. Appl. Electrochem. 26 (1996) 557. [171] D.S. Watkins, K.W. Dircks, D.G. Epp, U.S. Patent 5,108,849. [172] C. Cavalca, S.T. Homeyer, E. Walsworth, U.S. Patent 5,686,199 (to Allied Signal). [173] E. Ramunni, G. Fleba, M. Brambilla, EP 0 817 297 A2 (applicant: De Nora SPA).

31

[174] M.P. Hogarth, G.A. Hards, Platinum Metals Rev. 40 (1996) 150. [175] S. Gottesfeld, S.J.C. Cleghorn, X. Ren, T.E. Springer, M.S. Wilson, 1996 fuel cell Seminar, November 17 20, Orlando, Florida, Abstracts, p. 521. [176] U. Stimming, K.A. Friedrich, W. Unkauf, DE 197 10 819 C1 (to Forschungszentrum Ju lich). [177] M. Vaarkamp, Catalysis Today 39 (1997) 271. [178] J.J. Rehr, A. Ankudinov, S.I. Zabinsky, Catalysis Today 39 (1997) 263.

You might also like