Table of contents

Volume MA2016-02

Previous issue Next issue

I01-Polymer Electrolyte Fuel Cells 16 (PEFC 16)

A-01 Cell Performance 2 - Oct 2 2016 7:40AM

2350

, , , , and

The development of complex mesoscale (nm - µm) materials used for electrochemical applications requires comparable progress in the analytical instruments and techniques in order to understand the physical and chemical structure-property relationships underlying their performance. Conventional "hard" X-ray (i.e. > 10 keV) scattering has received considerable attention due to the fact that it is a high-resolution nondestructive structural probe that can interrogate a statistically significant 3-dimensional sample area. The non-resonant nature of this scattering process limits its applicability to materials that possess significantly different electron densities. Unfortunately, the performance of many electrochemical materials hinges on subtle heterogeneities that do not possess a high electron density contrast such as interfacial nanostructures, impurities, and chemical composition gradients. To help address this challenge, resonant soft X-ray scattering (RSoXS) uses tunable "soft" X-rays (100 - 2000 eV) to dramatically enhance the scattering cross sections from heterogeneous materials when the X-ray photon energy is judiciously chosen to coincide with favored transitions near a material's absorption edges. The RSoXS results in Fig. 1a show an example of how we used the resonance-enhanced scattering signals at selected photon energies to isolate the scattering contribution from different polymers in a phase separated block copolymer in order to unambiguously define the complex morphology of a triblock copolymer sample with both chemical and nm-scale spatial sensitivity.1

In this presentation, we reveal how operando RSoXS can be a powerful reciprocal space probe for mesoscale electrochemistry due to its chemical sensitivity, large accessible size scale, and polarization control.2, 3 We will convey how this technique can be applied under operando conditions to study pores, surfaces,4 and buried interfaces5 of low-Z element materials6 including many transition metals; the practical considerations of conducting such experiments will also be discussed. We will explain how the intrinsic combination of scattering and spectroscopy allows us to monitor spatio-chemical changes at a specific location by detecting the change in intensity at a specific scattering vector, q, for X-ray energies that are both ON and OFF resonance with the species of interest (Fig. 1b).

As an example of the utility of RSoXS to electrochemical applications, we present recent results on Nafion, a perfluorinated sulfonic acid (PFSA) membrane material that is considered to be a critical cost and performance-limiting component in many devices including fuel cells, electrolyzers, and redox-flow batteries. Recent RSoXS results acquired with a wet sample cell interrogated the Nafion films' partially orientated molecules inside ionomer domains. Using polarized X-rays with a photon energy tuned to the fluorine absorption edge (~690 eV), we observed a surprisingly strong scattering anisotropy that indicated preferred local crystalline grain orientation at the interface between different phases, an effect which is not visible when the X-ray photon energy is off-resonance with the fluorinated ionomers (Fig. 1c). These results enable us to develop a full electron density map that helps us understand why the pore structure of Nafion works so well, but may also yield insights into whether the development of porous separators as alternatives to PFSAs require pore sizes that are comparable to the hydrophilic channels in PFSAs (e.g., ≤ 3 nm).7 We will then expand on how combining such operando RSoXS data with electrochemical analytical methods could uncover important dynamic structure-property relationships underlying the interplay of various factors such as migration and electro-osmosis, chemical/physical stability, water uptake, permeability, etc.8-10

References

1. C. Wang, D. H. Lee, A. Hexemer, M. I. Kim, W. Zhao, H. Hasegawa, H. Ade and T. P. Russell, Nano Lett, 2011, 11, 3906-3911.

2. B. A. Collins, J. E. Cochran, H. Yan, E. Gann, C. Hub, R. Fink, C. Wang, T. Schuettfort, C. R. McNeill, M. L. Chabinyc and H. Ade, Nat Mater, 2012, 11, 536-543.

3. S. C. B. Mannsfeld, Nat Mater, 2012, 11, 489-490.

4. J. Schlappa, C. F. Chang, Z. Hu, E. Schierle, H. Ott, E. Weschke, G. Kaindl, M. Huijben, G. Rijnders, D. H. A. Blank, L. H. Tjeng and C. Schussler-Langeheine, J Phys-Condens Mat, 2012, 24.

5. M. Nayak, P. C. Pradhan and G. S. Lodha, Sci Rep-Uk, 2015, 5.

6. C. Wang, T. Araki and H. Ade, Appl Phys Lett, 2005, 87.

7. M. L. Perry and A. Z. Weber, J Electrochem Soc, 2016, 163, A5064-A5067.

8. R. M. Darling, A. Z. Weber, M. C. Tucker and M. L. Perry, J Electrochem Soc, 2016, 163, A5014-A5022.

9. X. L. Wei, B. Li and W. Wang, Polym Rev, 2015, 55, 247-272.

10. Y. S. Kim and K. S. Lee, Polym Rev, 2015, 55, 330-370.

Figure 1

2351

, and

Land-channel geometry is necessary for proton exchange membrane fuel cells to transport electron and at the same time to transport reactants/products. However, this configuration causes the difference in transport distance between the flow channel to the catalyst layer, and results in the non-uniform distribution of various factors, such as species concentration, current generation, and rate of degradation. In order to investigate the distributions of various parameters in the land-channel direction, a small-scale segmented cell with about 300-micron resolution was developed by the authors' group, and successfully measured the current and high-frequency resistance distribution in the land-channel direction. Distribution of oxygen transport resistance from the flow channel to the catalyst layer was also measured using limiting current technique.

The oxygen transport mechanism is completely different in the conventional flow field and in the interdigitated flow field. In the conventional flow field, the oxygen transport is mostly driven by the duffusion due to the concentration gradient. In the interdigitated flow field, on the other hand, the convective flow through gas diffusion layer (GDL) makes the diffusion distance shorter than the conventional flow field case, and therefore the lower oxygen transport resistance is expected.

In this study the oxygen transport resistance distribution in land-channel direction is measured in two different flow field configurations, that is, conventional flow field and interdigitated flow field. The comparison of the results from these two flow fields reveals the impact of the convective transport through GDL on the oxygen transport resistance and on the distribution.

2352

, and

Since most of the PEMFCs have land-channel geometry in their flow field, the distance between the flow channel to the catalyst layer is not uniform. The distance is longer from the flow channel to the catalyst layer under the land area than to the catalyst layer under the channel area. This difference causes the non-uniform current density in land-channel direction, and the distribution largely depends on the operating conditions. As the inlet gas humidity in the operating condition decreases, the trend in the current density distribution changes from high local current under the channel to relatively uniform, and to higher local current under the land (Figure 1). This trend is due to the combined effects of (1) non-uniform oxygen transport resistance, (2) non-uniform liquid water transport distance causing the local flooding under the land area at wet condition, and (3) the tendency of faster membrane dryout under the channel area at dry condition.

In order to investigate the distributions of various parameters in the land-channel direction, a small-scale segmented cell with about 300-micron resolution was developed by the authors' group, and successfully measured the current and high-frequency resistance distribution in the land-channel direction. Distribution of oxygen transport resistance from the flow channel to the catalyst layer was also measured using limiting current technique.

In this study a new segmented cell is developed. Using the transparent cathode plate the behavior of liquid water in cathode flow channel can be observed, and simultaneously the current density distribution is measured in land-channel direction. With this cell the relation between the current density distribution in the land-channel direction and the presence of liquid water is investigated. First, it is studied that at what gas humidity and at what overall current density the liquid water starts emerging from the gas diffusion layer surface to the flow channel. This provides us the information about the dependency of the liquid water emergence location on the current density distribution is studied. Then, how the presence of liquid water in the flow channel changes the current distribution is studied.

Figure 1

2353

, , , , and

In recent years phosphoric acid doped PBI-type fuel cells have drawn much attention as a promising candidate for energy storage and conversion applications at relatively high temperatures (100 – 200 °C). The elevated operating temperature of HT-PEMFCs (high-temperature polymer electrolyte membrane fuel cells) simplifies water and thermal management. In addition, it greatly enhances the fuel cell's tolerance against impurities (e.g. CO) in hydrogen, which is critical for operation with reformate hydrogen. However, durability and stability of high-temperature PEM-MEAs still need to improve for widespread commercialization [1], and better tools for identifying performance-loss related mechanisms are desirable. A very useful in-situ technique for performance analysis of HT-PEMFCs is electrochemical impedance spectroscopy (EIS). As previously reported, EIS can be used to characterize the impact of different parameters (e.g. stoichiometry, temperature, etc.) on cell kinetics [2, 3]. Equivalent circuit models are usually used to analyze the EIS data and identify the performance-loss related mechanisms.

A major drawback of this approach is that the assumptions for the model equivalent circuits are sometimes ambiguous or even misleading due to the lack of priori knowledge about the electrochemical system under study. Furthermore, a clear separation of various physiochemical processes is difficult if the time constants of these processes overlap significantly in the frequency domain. To overcome these drawbacks, advanced mathematical methods such as Distribution of Relaxation Times (DRT) can be applied. DRT relies on the representation of the polarization impedance by its characteristic time constants and is numerically approximated by a discrete distribution function. This method has been successfully demonstrated for process identification and separation in solid oxide fuel cells (SOFC) [4].

In this study, we applied this technique on high temperature PEM-MEAs for the first time. Electrochemical processes were identified and analyzed by varying cell parameters such as temperature and stoichiometry. The results offer a refined understanding of loss mechanisms and provide valuable guidance for fuel cell improvement and optimization.

[1] A. Chandan, M. Hattenberger, A. El-kharouf, S. Du, A. Dhir, V. Self, B.G. Pollet, A. Ingram, W. Bujalski, J. Power Sources 231 (2013) 264

[2] J. L. Jesperesen, E. Schaltz, S.K. Kær, J. of Power Sources 191 (2009) 289-296

[3] F. Mack, R. Laukenmann, S. Galbiati, J. A. Kerres, R. Zeis, ECS Transactions 69 (17), 1075-1087 (2015)

[4] A. Leonide, V. Sonn, A. Weber, and E. Ivers-Tiffée, Journal of the Electrochemical Society 155 (1) B36-B41 (2008)

Figure 1 Typical Nyquist plot of a phosphoric acid doped PBI-type HT-PEMFC at different current densities (a) and the corresponding distribution function (b).

Figure 1

2354

, and

A conventional polymer electrolyte fuel cell (PEFC) incorporates a membrane electrode assembly (MEA) which is comprised of the anode catalyst layer-polymer electrolyte -cathode catalyst layer sandwiched between two gas diffusion layers (GDLs), and bipolar plates. Conventional GDL is typically comprised of highly porous carbon paper or cloth to help diffuse the reactant gases onto the electrode, which is coated with a thin micro-porous layer (MPL). At high current densities, mass transport limitations of fuel or oxidizer in the PEFC occur in porous structures of the GDL, particularly at the cathode, which result in a sharp drop in the output voltage. The key to minimizing mass-transport losses is effective water management in the cell. Liquid water in macro-pores of GDL decreases the fuel cell performance at high current density due to the lack of oxygen reaching the catalyst layer.

A new MEA structure is recently introduced, where the carbon paper backing layer is eliminated and the entire gas diffusion layer consists of only the MPL [1]. We have further improved on this concept by directly depositing the MPL onto the CCM, resulting in an improvement in the interfacial contact between the MPL and the catalyst layer, as well as a simplified fabrication and assembly process. Spray deposition method was used for depositing this MPL onto a commercial CCM [2]. This concept was proven to work and perform better than a conventional cell with a micro-channel flow field used to provide the desired pathway for the reactant gases throughout the cell substituting for the macro porous carbon paper and the millimeter sized flow field.

In this work, a porous foam is utilized as the flow field to distribute the reactant gases over the micro-porous layer instead of the micro-channel flow field. Various pore sizes of the foam, i.e. 60-100 PPI, are used to compare with the conventional micro channel flow field. Contact resistance, permeability, electrochemical impedance spectroscopy, and mechanical properties are used to further characterize the new MEA structure incorporating the porous foam flow field. This method can be an effective way of enabling very high power density operation by reducing the mass transport limitations, and provides an improvement in the interfacial contact between the MEA and the flow field as well as improved mechanical properties.

[1] T. Kotaka, Y. Tabuchi, U. Pasaogullari, and C. Y. Wang, Electrochim. Acta, 146, 618 (2014).

[2] J. Park, U. Pasaogullari, L. J. Bonville, ECS Transactions, 69, 1355-1362 (2015).

2355

, and

There is much work in the literature related to the mathematical optimization of low-temperature polymer electrolyte membrane fuel cells (PEMFCs). In general, the optimization of PEMFCs involves (1) geometry optimization, which focuses on finding the optimum channel geometry and dimensions; and (2) microstructure optimization, which focuses on finding the optimum distributions of the porosity, catalyst, and electrolyte that maximize the power density of the cell. Depending on the complexity of the mathematical model used to describe the PEMFC, the number of variables that need to be optimized can become very large. For instance, let us consider a fuel cell that is discretized using a 3-D finite element mesh with 106 nodes. If we would like to find the optimum values of the porosity, catalyst concentration, and electrolyte density at each mesh point, the total number of design parameters is 3 ×106. Because the number of design parameters is so large, traditional heuristic techniques such as genetic algorithms, swarm-optimization, or other evolutionary algorithms are practically impossible to use even when implemented on massive computer clusters. (Notice that the number of optimization parameters in non-heuristic techniques is usually between 1-10, which is a few orders of magnitude smaller than the total number of design parameters for the PEMFC presented above.) Hence, so far, to make the problem computationally tractable, heuristic optimization techniques have been applied to optimize only a few number of PEMFC parameters.

In this presentation we develop a gradient-based technique for the optimization of PEMFCs, when the number of degrees of freedom, which is defined as the number of optimization parameters, is very large (the number of degrees of freedom in our work should not be confused with the number of nodes or elements in the discretization of the fuel cell) [1]. The technique is based on the computation of the sensitivity functions of the parameters that need to be optimized and is using these sensitivity functions to optimize the cell. The sensitivity functions of the parameters of interest are computed using an adjoint space method initially developed by the applied mathematics community to solve 1-D optimization problems in fluid dynamics, climate, and heat transfer problems [2]. Our group has also used a similar adjoint space technique to analyze the variability and optimize the doping profiles in 2-D and 3-D semiconductor devices [3]. The computational cost required to compute the sensitivity functions using the adjoint space method is relatively small since this method requires solving only one sparse system of linear equations instead of performing multiple fuel cell simulations.

Using the proposed technique we are able to predict the optimum 3-D distribution of platinum particles and porosity profile that maximizes the power density of the cell at different operating current densities. The optimum distributions and porosity profiles depends on the positions of the landings and openings, and on the geometry and dimensions of the layers. In agreement with existing experimental data and previous theoretical estimations [4], we obtain that, at large current densities, the catalyst density should be distributed non-uniformly inside the cell in order to increase the power density of the cell. At low operating current densities the optimum catalyst distribution is more or less uniform distributed inside the catalyst layer. In the case of the porosity distribution, at large operating current, it is more efficient to increase the porosity of the catalyst layer towards the gas diffusion layer side in order to increase the flow of the oxygen and water vapors. At low operating current densities the optimum porosity is uniformly distributed inside the catalyst and gas diffusion layers. More details about the technique, the numerical implementation, and a number of fully optimized structures will be presented at the meeting.

[1] P. Andrei and M. Mehta, "Large-scale optimization of polymer electrolyte membrane fuel cells," 227th ECS Meeting, Chicago, IL, 2015.

[2] D. Cacuci, Sensitivity and Uncertainty Analysis, Volume 1: Theory, Chapman & Hall/CRC, 2003.

[3] P. Andrei, I. Mayergoyz, "Quantum mechanical effects on random oxide thickness and random doping induced fluctuations in ultrasmall semiconductor devices", Journal of Applied Physics, 94 (2003) 7163-7172.

[4] P. Andrei, G. Mixon, M. Mehta, and V. Bevara, "Design of the catalyst layers in PEMFCs using an adjoint sensitivity analysis approach," 227th ECS Meeting, Chicago, IL, 2015.

2356

and

Additional cost reductions are needed for proton exchange membrane fuel cells (PEMFC) to effectively compete with internal combustion engine powered vehicles. Recent trends have focused on the use of a high current density and a lower Pt catalyst loading to reduce the quantity of expensive materials. However, such strategies have led to higher mass transfer overpotentials (1). These performance losses are advantageously characterized by the mass transfer coefficient. Its separation into fundamental contributions will focus research efforts to improve cell performance. A separation method previously proposed (2) was extended, which provides information for a wider cell voltage range of relevance to duty cycling operation.

The extended model yields a closed form and implicit current distribution expression (Eq. 1) where X represents the dimensionless Cartesian coordinate along the flow field channel, ie the inlet oxygen flow rate equivalent current density, ik the kinetic current density, iL,0 the inlet limiting current density and i the current density. Experimental data obtained with a segmented single cell of 100 cm2 active area (Gore catalyst coated membrane, 18 μm thick membrane, 0.2 and 0.1 mg Pt cm-2 catalyst loading, segmented and unsegmented Sigracet 25BC gas diffusion layer for the cathode and anode, respectively) validated Eq. 1 (Fig. 1a) and led to overall oxygen mass transfer coefficients k (derived from iL,0 and operating conditions). A plot of 1/k versus M, the diluent molecular weight, yields a linear relation (Eq. 2) enabling the separation of the overall mass transfer coefficient into 2 contributions where ke+K represents the oxygen mass transfer coefficient in the electrolyte and the gas phase (Knudsen diffusion), km the oxygen mass transfer coefficient in the gas phase (molecular diffusion) for the specified diluent (subscript He for helium, N for N2, CF for C3F8) and b a parameter. The molecular diffusion mass transfer coefficient km is larger with a lighter diluent (Fig. 1b). Each mass transfer coefficient increases as the cell voltage is decreased. However, the increase in the mass transfer coefficient in the electrolyte and the gas phase (Knudsen diffusion) ke+K is more substantial than for the other contribution. As a result, the k/ke+K ratio decreases for lower cell voltages. Specifically for N2, the mass transfer rate limiting step is associated with the electrolyte and the gas phase (Knudsen diffusion), which suggests a focus on the micro-porous and catalyst layers to increase performance. Two reasons are proposed to explain the cell voltage dependency observed in Fig. 1b. First, the local change in gas phase composition promotes convection. This effect is particularly important before reaching the mass transfer control regime, in the mixed kinetic and mass transfer control region, as the interfacial oxygen concentration changes from its bulk flow field channel value to near 0 in the 0.75 to 0.6 V cell potential range creating a diluent enriched layer. Second, the local temperature in the gas diffusion electrode (through plane direction) is also higher at a lower cell voltage owing to additional heat production. The larger local temperatures in turn promote larger diffusion coefficients in the gas phase (molecular and Knudsen) and permeabilities in the ionomer phase. A comparison with other mass transfer coefficient values derived from impedance data (3) and the model limits of validity will also be discussed.

Acknowledgments

We gratefully acknowledge funding from the Army Research Office (W911NF-15-1-0188). The authors are grateful to the Hawaiian Electric Company for their ongoing support of the operations of the Hawaii Sustainable Energy Research Facility.

References

1. A. Z. Weber, A. Kusoglu, J. Mater. Chem. A, 2, 17207 (2014).

2. T. V. Reshetenko, J. St-Pierre,J. Electrochem. Soc., 161, F1089 (2014).

3. S. Arisetty, X. Wang, R. K. Ahluwalia, R. Mukundan, R. Borup, J. Davey, D. Langlois, F. Gambini, O. Polevaya, S. Blanchet, J. Electrochem. Soc., 159, B455 (2012).

Fig. 1. a) Selected comparisons between current density distributions along the dimensionless flow field channel length obtained with He diluent and calculated model curves (Eq. 1), b) Gas phase molecular diffusion, and, gas phase Knudsen diffusion and ionomer phase diffusion oxygen mass transfer coefficients for different cell voltage values and gas diluents. 60 °C, 48.3/48.3 kPag, 100/100 % relative humidity and 5 % O2 + diluents/H2 for the cathode and anode respectively. 50 mL min-1 O2 + 958 mL min-1 He or N2/424 mL min-1 H2 and 25 mL min-1 O2 + 479 mL min-1 C3F8/212 mL min-1 H2.

Figure 1

2357

, and

Cold-start (subzero startup) capability of polymer electrolyte fuel cell (PEFC) is of great importance. In this paper, the effects of the micro-porous-layer (MPL), varying start-up temperatures, start-up current density variation on the cold-start operation are investigated. We found PEFC with anode micro-porous-layer (AMPL), compared with one with CMPL, has higher possibility of successful cold start. By analyzing the anode and cathode pressure revolution of PEFC, the effect of MPL on the super cooled water removal and ice formation at different temperatures (-7,-10,-15 and-20℃) are discussed . In addition, we investigate the ice distribution in MEA through the X-ray device, by comparing the initial image when fuel cells shut down (before the produce the water) and final image after failed cold-start. We also explore the negative voltage phenomenon in the initial process of cold-start operation.

2358

, , , and

Cost has been one of the primary barriers to the commercialization of polymer electrolyte membrane fuel cells (PEMFCs), clean energy conversion devices used primarily for power generation applications. While system design improvements and the development of novel electrode materials with increased activity have helped lower costs, the primary driver of cost reduction will be manufacturing membrane electrode assemblies (MEAs) at high volume [1]. Recently developed in-line diagnostic tools to monitor quality have successfully detected a variety of electrode coating variations [2]. This work studies a subset of these electrode irregularities fabricated onto MEAs operating under accelerated stress tests (ASTs) to simulate stresses of real-world driving conditions while spatially monitoring the onset and development of failure points with a thermal camera. An understanding of the impact that electrode irregularities have on MEA lifetime will subsequently allow for (i) the potential classification of these irregularities as defects and (ii) the development of threshold detection limits for in-line quality control diagnostics.

Catalyst-coated membranes (CCMs) and gas-diffusion electrodes (GDEs) were fabricated with an ExactaCoat Sono-Tek ultrasonic spray system. The active area and nominal loading of these Pt/C electrodes were 50 cm2 and 0.2/0.2 mgPt/cm2, respectively. Membrane material was NRE-212. These layers were laminated into edge-protected MEA structures using a hot-pressing process. The MEA structures were exposed to a combined chemical and mechanical AST by operating the fuel cell at 80°C and 0.5/0.5 SLM H2/Air under open circuit potential (OCV), and cycling between 0 and 80% relative humidity for a 30 second duty cycle. The hydrogen crossover limiting current and the OCV were indicators for a developing failure and the presence of a failure, respectively. A novel cell hardware shown in figure (a) was used to remove the cathode flow-field and conduct a spatial hydrogen crossover measurement. Any hydrogen crossing from the anode to the cathode exothermically reacts at the cathode catalyst with ambient air to produce a heat signature that is detected with the thermal camera [3]. Higher crossover rates at for example discrete failure points become apparent through a significant local temperature rise.

Sample results of a pristine MEA tested as described above are shown in the figures. The graph (b) shows the hydrogen crossover limiting current density rapidly increasing after the OCV declined with an increasing rate. This indicated that the MEA developed a failure point. The images show the thermal response during the spatial hydrogen crossover measurements for the pristine MEA (c) prior to and (d) subsequent to aging. The heat signature indicated a failure point developed near the inlet of the cell. Results will include studies on MEA structures with intentionally and systematically introduced coating variations of various sizes and severities.

References

1) S. Kamaruzzaman, and W.R. Daud. Renewable Energy 31.5 (2006).

2) M. Ulsh and B. Sopori and N.V. Aieta and G. Bender. ECS Transactions 50.2 (2013).

3) G. Bender and W. Felt, and M. Ulsh. Journal of Power Sources 253 (2014).

Figure 1

2359

, , , , , , , , , et al

The polymer electrolyte membrane fuel cell (PEMFC) continues to develop as a viable alternative to the combustion engine in automotive applications. As this technology advances, it is critical that the PEMFC is capable of maintaining performance for long term operation. In an effort to understand the degradation of fuel cell performance over time, a study into the development of an accelerated gas diffusion layer degradation protocol was completed. As part of an effort to quantify the effects of this accelerated degradation, in situ synchrotron radiography imaging techniques were coupled with performance testing in order to quantify the effects of this degradation on water saturation profiles, transport resistance, and limiting current in an operating fuel cell.

Gas diffusion layer samples of SGL 25 BC and SGL 29 BC were artificially aged in a concentrated solution of hydrogen peroxide (30% wt.) for a period of 12 hours at an elevated temperature of 90 degrees Celsius. Hydrogen peroxide facilitates chemical corrosion of the carbon material in the gas diffusion layer and was found to significantly affect the wettability of the degraded samples. Hydrogen peroxide was selected to facilitate the degradation mechanism due to the fact that it is a recognized chemical species found in operating fuel cells and produced at the catalyst layer (1, 2).

The accelerated degradation procedure that was used in this study was found to primarily affect the wettability of the tested gas diffusion layers. Consequently, it was expected that the most significant impact of this degradation mechanism would be in the mass transport losses at high current densities. For this reason, a limiting current investigation was performed. Unlike other studies of limiting current in which the cathode oxygen concentration is varied (3), a limiting current study based on varying the relative humidity of reactant gases was performed. In this study, limiting current was measured for relative humidities of 0%, 50%, 80%, 90%, and 100% for several fuel cells with fresh and degraded GDLs. Simultaneous synchrotron imaging was used to quantify the distribution of liquid water in the anode and cathode of the operating cell during these limiting current studies.

Trends with respect to reactant transport resistance, water saturation profiles, and limiting current were quantified as a function of relative humidity and degree of degradation. For all tested samples, as the relative humidity was reduced, the corresponding limiting current increased. It was also found that the limiting current liquid water saturation profile was independent of the reactant gas relative humidity. Additionally, artificial degradation led to increased liquid water saturation levels in the operating fuel cell. This work illustrates the potential impacts of long term fuel cell operation on the water management of PEMFC gas diffusion layers.

References

 

1. C. Chen and T. Fuller, ECS Transactions., 11, 1 (2007).

2. W. Liu and D. Zuckerbrod, J.Electrochem.Soc., 152, 6 (2005).

3. D. R. Baker, D. A. Caulk, K. C. Neyerlin and M. W. Murphy, J.Electrochem.Soc., 156, 9 (2009).

2360

, and

R&D on Proton exchange membrane fuel cells (PEMFC's) technology has been accelerated in the last few years to reduce the dependency on fossil fuel. PEMFC's operate by internally combining oxygen from air with hydrogen to form water and generate electricity and waste heat. The most common hurdle for enhanced PEMFC durability and performance is still the water management: the proton exchange membrane in the center of these fuel cells has to be hydrated in order to keep its ability to conduct protons from anode to cathode side while on the other hand excessive liquid water can lead to cell flooding and increased degradation rates of the cell. Thus, a detailed understanding of all aspects of water management in PEMFC is important. This includes the fuel cell water balance, i.e. the question which fraction of the product water leaves the fuel cell via the anode channels versus the cathode channel. Our research group is currently developing a state of art technology to obtain an ad-hoc and real time electrical signal of the fuel cell water balance by employing a constant temperature hot wire anemometry [1]. The hot wire sensor is placed in the anode outlet of PEMFC, and the voltage signal received gives valuable insight into heat and mass transfer phenomena which can be interpreted directly to real time PEMFC water balance. So far, ex-situ experiments have been conducted to measure the voltage signal of the hot wire anemometer for a known gas stream that contains a binary mixture of hydrogen and water vapour as would ideally leave the fuel cell anode. However, in an operating fuel cell there are two main uncertainties in this method: (i) there is some internal hydrogen crossover that does not load to an external current, thus the exact amount of hydrogen leaving the fuel cell anode is not known, and (ii) that there is nitrogen crossing over from the cathode side to the anode side, and this nitrogen does slighty falsify the voltage signal that the hot wire yields. The effects of nitrogen-cross over on the hot wire voltage signal at the anode outlet and consequently on the measured water balance (rd) is studied in this work.

In our previous work it was shown that the only unknown in the determination of the hot wire voltage signal is the equation that determines the heat transfer around the hot wire, and we have shown that it is necessary to employ a power-law equation as suggested by Hilpert:

Nu = C Pr0.33 Rem

where Re is the Reynolds number and Pr is the Prandtl number are calculated at film temperature. The constants C and m usually depend on the Re, and they will be determined out of the experimental data for the gas mixture, by plotting the measured (Nu/ Pr0.333) versus the Re number of the gas mixture stream. The missing C and m from Hilpert equation can then be easily obtained from a power-fit of the general form Y = CXm. Also, it was shown somewhere else that in fact only the exponent to the Reynolds number, of the gas mixture stream is important when determining the fuel cell water balance out of the hot wire signal [1].

 The nitrogen-cross over is experimentally demonstrated by introducing 1% of nitrogen to the dry hydrogen molar flow. The 99%H2+1%N2 of the dry mixture is humidified with water vapor by controlling the relative humidity (RH) of the dry mixture with in the range of (0-100)%RH, simulating the PEMFC anode outlet. The hot wire voltage is measured with and without nitrogen and it was slightly lower with the presence of 1 % nitrogen in the flow. The effect of the voltage reduction on the measured water balance can be neglected. This because the effect of 1% nitrogen on power law constant's m which is used in determining the water balance as explained somewhere else is extremely low. It might be concluded that, the hot wire technique for measuring the water balance is still accurate and can be used for ad-hoc and real time water balance measurements and the nitrogen cross-over affect can be neglected, knowing that the 1% N2 is one order of magnitude higher than the actual concertation of 0.02 m2 active area cell used for lab testing.

 

[1] Berning, T., & Al Shakhshir, S., Applying hot wire anemometry to directly measure the water balance in a proton exchange membrane fuel cell–Part 1: Theory. International Journal of Hydrogen Energy, 40(36) (2015), 12400-12412.

2361

, , , , and

In PEM fuel cells, electrocatalyst interfaces including Pt-C, Pt-ionomer, C-ionomer interfaces, play a critical role for both performance and durability.1-3 Pt-C interface influences the stability and activity of Pt dramatically; electrocatalyst/ionomer interfaces4 in fuel cell catalyst layers determine the electrochemical surface area (ECSA), local transport, and structural stability of catalyst layer. Delineating the structure evolution of the interfaces in response to PEM fuel cell operating environment will provide the science foundation for new electrocatalyst/catalyst layer design therefore improve the performance and durability of PEM fuel cells.

We have developed an aberration corrected environmental transmission electron microscopy (ETEM), which allows unprecedented spatial and temporal resolution in the relevant gas environment, such as O2, H2, H2O, CO, CH4, and CO2. With customized intelligent gas delivery system, the gas environment relevant to PEM fuel cell anode and cathode (H2/H2O, O2/H2O) can be simulated and precisely controlled through the residual gas analyzer (RGA). The ETEM opens new avenue to study the dynamic changes of electrocatalyst interfaces.

In order to study the surface chemistry of catalysts under different environment, a specially-designed pre-treatment system is integrated into X-ray photoelectron spectroscopy (XPS), which allows us to move catalysts seamlessly between the pre-treatment system and XPS chamber so that surface contamination from external environment will be eliminated.

In this talk, we will present our recent progress on identifying both structure and surface chemistry evolution and their correlation in electrocatalysts under PEM fuel cell relevant environment through the combination of ETEM and XPS capabilities.

1. Y. Y. Shao, G. P. Yin and Y. Z. Gao, J. Power Sources, 2007, 171, 558-566.

2. S. Park, Y. Y. Shao, H. Y. Wan, V. V. Viswanathan, S. A. Towne, P. C. Rieke, J. Liu and Y. Wang, J. Phys. Chem. C, 2011, 115, 22633-22639.

3. R. Kou, Y. Y. Shao, D. H. Mei, Z. M. Nie, D. H. Wang, C. M. Wang, V. V. Viswanathan, S. Park, I. A. Aksay, Y. H. Lin, Y. Wang and J. Liu, J. Am. Chem. Soc., 2011, 133, 2541-2547.

4. A. Z. Weber and A. Kusoglu, J. Mater. Chem. A, 2014, 2, 17207-17211.

F-01 Alkaline Electrolysis/Hydrogen Evolution - Oct 2 2016 7:40AM

2362

, and

Hydrogen is regarded as a valuable high energy density, zero-emission fuel for the future, with the potential to reduce or eliminate fossil fuels. H2 also has potential as an energy carrier or storage molecule: i.e. being generated, stored, and consumed to bridge the gap of high and low production cycles of renewable energy sources, such as solar cells and wind turbines. Because molecular hydrogen is not naturally available, the complications of generation, storage and transportation must be addressed. On Earth, hydrogen (H) mainly occurs in combination with oxygen, i.e. in water. One possible technology option for clean H2 generation is low temperature water electrolysis, with the two half-reactions being the oxygen evolution reaction (OER) and hydrogen evolution reaction (HER). These reactions, particularly the OER, are mechanistically complex and dependent on the surface energy of multiple intermediate species, leading to high overpotentials and low efficiency. Precious metal-based catalyst materials are ultimately not viable, due to rarity and cost, and typically provide high catalytic activity for the OER or HER, but not both. More effective catalysts with the ability to catalyze both the OER and HER (i.e. OER/HER bi-functional behavior) could improve current electrolyzers and enable future technologies. Thus, there is a desire to design electrocatalysts based on earth- abundant, low-cost materials for overall water splitting.

In this talk we will present on the synthesis and electrochemical characterization cobalt phosphide-based materials. CoP-based electrocatalysts in both thin film and nanoparticle form have been prepared from their respective Co3O4 spinel precursors using phosphine gas, generated in situ from the thermal decomposition of sodium hypophosphite. These electrocatalysts were found to have performance comparable to that of the commercial precious metal benchmarks when examined in alkaline electrolyte. For example, ΔE(OER–HER) values were calculated by adding the overpotentials at which jgeo = 10 mA cm-2 and jgeo = -10 mA cm-2were attained for the OER and HER, respectively, to the thermodynamic OER–HER potential difference of 1.23 V. Values of ΔE(OER–HER) = 1.77 V for cobalt phosphide catalysts are within 70 mV of the idealized water electrolysis with 20% Ir/C (anode):20% Pt/C (cathode) where ΔE(OER–HER) = 1.70 V. Various aspects of this work will be presented.

This work was supported by the Laboratory Directed Research and Development program at Sandia National Laboratories, a multi-program laboratory managed and operated by Sandia Corporation, a wholly owned subsidiary of Lockheed Martin Corporation, for the U.S. Department of Energy's National Nuclear Security Administration under contract DE-AC04-94AL85000.

2363

, , , and

Introduction

Proton Exchange Membrane (PEM) water electrolysis provides a sustainable solution for the production of hydrogen, and is well suited to couple with the intermittent nature of energy sources such as wind and solar. To spur large-scale commercial application of PEM water electrolysis, more efficient and less expensive non-noble metal electrocatalysts towards the hydrogen evolution reaction (HER) are required. Recently, transition metal phosphides (TMPs) have become the typical representatives of the burgeoning non-noble metal HER electrocatalysts due to that their intrinsic structures meet the criteria of outstanding electrocatalysts. It should be noted that the currently studied TMPs are mainly focusing on improving their HER performance by a series of structural engineering methods [1-3]. To the best of our knowledge, there are very few reports on promoting the intrinsic activity of the TMPs in water electrolysis.

In this communication, we reported the scalable synthesis of nickel phosphides with a mixed crystalline structure (referred to as Ni2P&Ni3P) toward HER by a solution-phase reaction. To elucidate the superior catalytic activity of Ni2P&Ni3P, we also explored the preparation of the nickel phosphides in a single crystalline state (referred to as Ni2P) and that in a single amorphous state (referred to as Ni-P) for comparison.

Result and Discussion

X-ray diffraction patterns of Ni2P&Ni3P、Ni2P and Ni-P are shown in Fig 1a. The analysis of the diffraction data reveals that both tiny Ni2P nanocrystallines and Ni3P nanocrystallines are observed for Ni2P&Ni3P. As for Ni2P, the Ni3P tiny nanocrystallines are vanished and only the intensive diffraction peaks corresponding to Ni2P are observed. In comparison with the crystalline state nickel phosphides, the Ni-P shows poor crystalline order with a very broad peak. The crystalline structures could take effects on the intermediate reaction of HER. In order to compare the hydrogen desorption ability of the three electrocatalysts, the hydrogen oxidation reaction (HOR) was investigated by LSV (Linear scan voltammetry) at 20 mA cm-2 for 5 min immediately after the hydrogen evolution reaction (HER). As it is shown in Fig 1b, the characteristic mixed crystalline structure is expected to promote the hydrogen desorption ability. Thus, the largest peak current density is obtained with Ni2P&Ni3P. Furthermore, steady state polarization curves of different electrocatalysts were shown in Fig 1c. As expected, the Ni2P&Ni3P shows the best HER activity, and the overpotentials required for Ni2P&Ni3P、Ni2P and Ni-P to produce cathodic current densities of 20 mA cm-2 are 154 mV, 221 mV, 262 mV, respectively.

Fig 1 (a) XRD patterns of Ni2P&Ni3P, Ni2P and Ni-P; (b) LSV and (c) Steady-state polarization curves of the three types of prepared catalysts at a scan rate of 1 mV s-1. All the electrochemical characterizations are tested in the mixture of 0.1 M H2SO4 and 0.5 M Na2SO4 at 25 C and atmosphere pressure.

Reference

[1] Liu P., Rodriguez J. A. Catalysts for Hydrogen Evolution from the [NiFe] Hydrogenase to the Ni2P (001) Surface: The Importance of Ensemble Effect [J]. Journal of the American Chemical Society, 2005,127(42):14871-14878.

[2] Popczun E. J., Read C. G., Roske C. W., Lewis S., Schaak R. E. Highly Active Electrocatalysis of the Hydrogen Evolution Reaction by Cobalt Phosphide Nanoparticles [J]. Angewandte Chemie International Edition, 2014,53(21):5427-5430.

[3] Liu Q., Pu Z. H., Asiri A. M., Sun X. P. Nitrogen-doped carbon nanotube supported iron phosphide nanocomposites for highly active electrocatalysis of the hydrogen evolution reaction [J]. Electrochim Acta, 2014,149(10):324-329.

Figure 1

2364

, , , , , and

Hydrogen is an ideal energy carrier with an energy density of 140 MJ kg-1 and water as its final product, which is considered as an ideal candidate for the replacement of fossil fuels.1 To date, most H2 is produced today through nickel-catalyzed conversion of CH4 to H2 and CO followed by a water gas shift reaction to yield H2 and CO2. However, this approach is based on unsustainable fossil energy and discharges greenhouse gas. Water is an ideal source for H2 production as it is carbon-free, plentiful and almost costless. Water electrolysis, with only H2 and O2 as the final product, is a highly promising route in environmental benign H2 generation.2, 3 Although the ultimate goal is to couple hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) to an integrated efficient device for whole water electrolysis, one of the most crucial issues that must be solved is the development of highly efficient and durable catalysts for water oxidation and reduction.3

In this study, monocrystalline Ni12P5 hollow spheres with ultrahigh specific surface area (222.5 m-2 g-1) were prepared by a water-in-oil microemulsion method (Figure 1). The over potential required for 10, 20 50 and 100 mA cm-2 current densities are 144, 173, 225 and 277 mV in strong acidic solution with a Tafel slope of only 45 mV dec-1, ranking among the most active nonprecious HER catalysts. Moreover, the monocrystalline Ni12P5 hollow spheres catalyst exhibited a highly active and stable performance during a 3,000 cycling tests through cyclic voltammetry and a 12 hour duration chronopotentiometry test. The results indicate the highly promising application potential of the monocrystalline Ni12P5 hollow spheres catalysts in hydrogen generation to replace Pt. These promising features make Ni12P5 hollow spheres ideal candidates for the next generation HER catalysts. These catalysts with novel monocrystalline structure can be extended to electrochemical applications in various fields, such as supercapacitors, fuel cells, batteries and electrochemical sensors, and so on.

Figure 1. The illustration of synthesis procedure for the Ni12P5 hollow spheres and the water electrolysis behavior of the catalysts.

Acknowledgments 

This work is supported by the Strategic priority research program of CAS (Grant No. XDA09030104), the Jilin Province Science and Technology Development Program (Grant No. 20130206068GX, 20140203012SF, 20160622037JC) and the Recruitment Program of Foreign Experts (Grant No. WQ20122200077).

References

1. X. Zou and Y. Zhang, Chem. Soc. Rev., 2015, 44, 5148-5180.

2. M. G. Walter, E. L. Warren, J. R. McKone, S. W. Boettcher, E. A. S. Qixi Mi and N. S. Lewis, Chem. Rev., 2010, 110, 6446–6473.

3. S. D. Ebbesen, S. H. Jensen, A. Hauch and M. B. Mogensen, Chem. Rev., 2014, 114, 10697-10734.

Figure 1

2365

, , , and

There is an ongoing effort to develop alkaline stable anion exchange membranes to be employed as solid-electrolytes in alkaline membrane fuel cells (AMFCs) and solid-state alkaline water electrolyzers so as to take advantage of the distinct and attractive advantages of operation at high pH. These advantages, which are not accessible to acidic proton-exchange membrane fuel cells and water electrolyzers, include the possibility of using stainless steel flow fields and endplates (with the subsequent reduction in costs), more flexibility in the choice of fuels, and the possibility of using cheaper non-platinum-group-metal electrocatalysts for both the hydrogen evolution/oxidation and the oxygen evolution/reduction reactions (1).

However alkaline operation has an important drawback. The hydrogen evolution reaction (HER) is much more sluggish because of the low concentration of protons in alkaline media and the lower activity of noble metals for this reaction. Pt electrocatalysts show considerable lower HER activity in alkaline conditions than in acidic conditions.

The following mechanism for water electrolysis in alkaline conditions has been proposed:

a) Volmer: H2O+e- +* → Had + OH[1]

b1) Tafel: Had + Had → Had + 2OH-+2* [2]

b2) Heyrovsky: Had + H2O+e- → H2 + OH-+ * [3]

 It is generally accepted that the Volmer step is the rate determining step (rds) for the overall reaction. This assumption implies that the presence of a bi-functional catalyst with hydrophilic domains able to stabilize the water in the transition complex will speed up the overall reaction, as observed by several authors for platinum, iridium, ruthenium and nickel electrocatalysts modified by deposition of transition metal hydroxides (i.e. nickel hydroxide) (2, 3).

In this work, we investigated the HOR/HER kinetics on Pt/C and Pt/C/Ni(OH)2 bi-functional electrocatalysts. Our objective was to determine the optimum concentration of Ni(OH)2 required for the best performance of the electrocatalyst for the HER under alkaline conditions. The catalysts were prepared by mixing the required amounts of colloidal dispersions of Ni(OH)2 with commercial Pt/C catalyst previously dispersed (using ultrasonication) in water. The RDE measurements were done in 0.1M KOH, at temperatures ranging from 0°C to 30°C to extract useful kinetic information such as exchange current densities and activation energies for each catalyst. Kinetic currents were obtained after IR and mass transport corrections. The HOR/HER kinetic current densities (see Figure 1) were fitted using the Butler–Volmer equation to estimate the exchange current densities for each catalyst at each temperature. Arrhenius behavior was observed. The activation energy was the same for all Pt/C and the bi-functional catalysts, at 37.2 kJ/mol, which was in good agreement with previously published value for Pt (4).

The highest exchange current densities were obtained for the catalyst containing 10wt% Ni(OH)2. Higher Ni(OH)2 concentrations resulted in a poorer performance due to the coverage of the Pt active sites; a reduction in ECSA was observed.

The bi-functional catalysts were also tested in a solid-state alkaline water electrolyzer operated with ultrapure water. The membrane electrode assemblies were fabricated with a commercial anion exchange membrane The anode catalyst was iridium oxide (2.5 mg/cm2). The optimal bi-functional electrocatalyst outperformed Pt/C by 0.1-0.2 V at relevant current densities.

References

1. N. W. Li and M. D. Guiver, Macromolecules, 47, 2175 (2014).

2. N. Danilovic, R. Subbaraman, D. Strmcnik, K. C. Chang, A. P. Paulikas, V. R. Stamenkovic and N. M. Markovic, Angewandte Chemie, 51, 12495 (2012).

3. D. Strmcnik, M. Uchimura, C. Wang, R. Subbaraman, N. Danilovic, D. van der Vliet, A. P. Paulikas, V. R. Stamenkovic and N. M. Markovic, Nature chemistry, 5, 300 (2013).

4. W. Sheng, H. A. Gasteiger and Y. Shao-Horn, Journal of The Electrochemical Society, 157, B1529 (2010).

Figure 1

2366

, , and

Hydrogen is poised to be a key energy carrier in a sustainable energy economy (1). Among the different approaches to produce hydrogen, water electrolysis has received special attention due to its unique advantages including reactant availability, safety, stable output and product purity (2). Alkaline water electrolysis offers additional advantages, including the possibility of using stainless steel flow fields and endplates (with the subsequent reduction in costs) and the possibility of using cheaper non-platinum-group-metal electrocatalysts for the oxygen evolution reaction. However, the slow kinetics for the hydrogen evolution reaction (HER) in alkaline media is currently a key drawback.

In this work we evaluated a mixed-metal-oxide composed of titanium dioxide and ruthenium dioxide as a support for Pt and measured the catalytic activity of this supported electrocatalys for the HER. The mixed-metal-oxide (TiO2-RuO2, RTO) was synthesized using a wet chemical synthesis procedure, and the Pt/RTO was synthesized by reduction of Pt precursor onto the support using an impregnation-reduction method. These materials were characterized by XRD, TEM and BET. The activity of Pt/RTO towards HER the was compared with a benchmark Pt/C catalyst (46%Pt; Tanaka, K. K.). Rotating disk electrode (RDE) measurements showed Pt/RTO outperforming (under all the conditions) the benchmark catalyst (Pt/C). The specific exchange current density for Pt/RTO was 2.51mA/cm2Pt (295 K, 1600rpm, H2-saturated 0.1M KOH), more than five times that of Pt/C. The catalysts were also evaluated in a solid-state water electrolyzer operated with ultrapure water at 50°C, wherein MEAs were prepared with IrO2 as a common anode electrocatalyst. The MEAs fabricated with Pt/RTO as the cathode catalyst outperformed the benchmark catalyst by 0.1-0.2 V across relevant current densities (see Figure 1).

1. M. S. Dresselhaus and I. L. Thomas, Nature, 414, 332 (2001).

2. H. Yin, S. Zhao, K. Zhao, A. Muqsit, H. Tang, L. Chang, H. Zhao, Y. Gao and Z. Tang, Nat Commun, 6, 6430 (2015).

Figure 1

2367

, , , , and

N

Introduction

Fuel cell technology has been extensively developed in order to create new clean sources of energy. In order to have a wide implementation of fuel cell devices powered by hydrogen (H2), an inexpensive source of clean H2 should be created. One of the technologies that produces hydrogen gas is the splitting of water through electrolysis. The majority of electrolyzers available on the market use proton exchange membrane (PEM) technology, which requires catalysts consisting of expensive and rare platinum group metals (PGMs). To increase penetration into the water electrolyzer and fuel cell markets, abundant and inexpensive materials should be incorporated. Alkaline exchange membrane water electrolysis (AEMWE) is an alternative technology that does not require PGM catalysts and, therefore, enables the use of base metals, alloys and oxides.

The goal of this research has been focused on the creation of a catalyst that is stable in alkaline media, has a high surface area, is easy to synthesize, and can be produced at a lower cost while maintaining a high oxygen evolution reaction (OER) activity.

Experimental

Unsupported ternary Ni-Mo-Cu materials were synthesized by the Sacrificial Support Method (SSM) using a high surface area silica [1-3]. The following process was followed, and produced a NiMoCo OER catalyst achieving a surface area (SA) of 25 m2 g-1.

15g of EH-5 (SA=400 m2 g-1) silica, 2.5g of nickel nitrate, 2.5g of copper nitrate and 3g of ammonium molybdate tetrahydrate were mixed by grinding them in a porcelain crucible. The crucible was placed into a tube furnace and heated to 100 °C with a ramp rate of 10ºC/min. The 100°C temperature was held for 50 minutes before an additional temperature ramp was conducted at an interval of 5ºC/min until 550°C was reached and held for four hours.. During this time, a reducing atmosphere was maintained using 7% H2. The silica was later removed by rinsing with 7M potassium hydroxide (KOH) for 24 hours, followed by washing and drying.

Cell testing was performed using a Fuel Cell Technologies 25 cm2 fuel cell stack, modified for electrolysis operation. The carbon anode flow field was replaced with an anode piece fabricated at Proton OnSite, based on 316L SS, which was cleaned and passivated in a nitric solution prior test. ASTM Type II DI water was fed to the anode side of the cell via a diaphragm pump. A submersible heater, placed in the water reservoir, was used to control stack temperature at 50ºC. An image of the test station is shown below in Figure 1.

Figure 1. 25 cm2 Test Stand

 Results and Discussion

SEM images show that the NiMoCu material has a well-developed porous structure, as well as primary particles on the nanometer scale (Figure 2). The surface area was measured at 25±3 m2 g-1 via BET analysis. Based on these results, the team decided to proceed with high surface area silica as the sacrificial support.

Figure 2. SEM on Ni-Mo-Cu catalyst Electrochemical activity of the OER is shown in Figure 3.

Figure 3. ORR activity of different Ni-Mo-Cu electrocatalysts in alkaline media.

Operational testing was conducted on the NiMoCo OER catalyst provided by UNM in the 25 cm2 cell previously described. As shown below in Figure 4, the UNM supplied material experienced relative stability at the steady-state current density of 200 mA/cm2, as compared to the PGM based reference plot. Both tests used the same membrane and electrode binders.

Figure 4. Steady-state operation graph of non-PGM anode versus PGM reference

Conclusion

Synthesis of spinel-based catalysts by SSM was investigated. Upon examination of the physical characteristics, it can be assumed that further modification to the design of the electrocatalysts is necessary. The electrochemical characteristics from the RDE experimentation proved that the catalysts are active in OER and can act as a replacement for conventional Pt-group catalysts in the appropriate conditions. Short duration operational testing has shown some translation of the RDE measurements to in-cell performance. Additional testing is planned for evaluation of repeatability and longer term stability.

References

[1] U. Martinez, A. Serov, M. Padilla, P. Atanassov ChemSusChem 7(8) (2014) 2351–2357.

[2] A. Serov, K. Artyushkova, P. Atanassov Adv. Energy Mater. 4: 1301735 (2014) doi: 10.1002/aenm.201301735.

[3] A. Serov, U. Tylus, K. Artyushkova, S. Mukerjee, P. Appl. Catal. B: Environmental 150 (2014) 179-186.

Figure 1

2368

, , , and

The production of hydrogen via water electrolysis is currently recognized as the only option to store gigawatt electrical energy coming from renewable energy sources such as wind and sunlight. However, commercially available alkaline electrolyzers are still limited to low current densities (around 0.5 Acm-2), primarily due to its internal resistance losses. In order to reach the excellent current densities of acidic polymer electrolyte membrane (PEM) water electrolysis (ranging between 2 and 4 Acm-2), it is crucial to tailor electrolyte construction and cell design in order to overcome the high ohmic losses of conventional diaphragms and archaic cell designs used in alkaline electrolysis. Here, we demonstrate for the first time alkaline electrolysis achieving current densities as high as 2 A.cm-2 at 2 volts, not only well overcoming the performances of lab-scale and commercial alkaline electrolyzers, but also closely matching the performance of state-of-the-art PEM water electrolysis.

2369

, and

This paper will describe the performance of alkaline water electrolyzers using newly developed Sustainion™-X5 membranes.

Sustainion™ membranes have a styrene base, and mixtures of other components to produce a high conductivity in alkaline systems. The membranes are stable when soaked in 1 M KOH at room temperature for months. A CO2electrolyzer has run for over 4000 hours with no loss in performance. Initial tests also show stability in 1 M KOH at 60°C, See Figure 1, although no long term tests have yet been completed.

The work here considers the performance of the membranes in model alkaline membrane electrolyzers. We manufactured a small MEA using platinum coated carbon paper as a cathode and a IrO2 coated carbon paper as an anode and a Sustainion™-X5 membrane. The MEA was mounted into Fuel Cell Technologies 5 cm² cell hardware with carbon flowfields. The cell was heated to 60 °C and 1 M KOH was fed into the anode and cathode.

Figure 2 shows how the cell voltage varied with time at a fixed current of 0.8 A/cm² and 1 A/cm². The data were taken on a single cell, at a fixed current of 1 A/cm² during the day and 0.8 A/cm² at night. Initially the cell produced 0.8 A/cm² at 1.74 V and 1 A/cm² at 1.81 V, but the voltage rose to nearly 2V after 150 hours of operation. Disassembly of the cell showed that the carbon paper on the anode was corroding leading to a loss of anode catalyst.

We have also started work on nickel catalysts. The initial results were modest, but additional results will be presented at the meeting.

Acknowledgement

Parts of this work were supported by ARPA-E under contract DE-AR0000684 and by 3M. The opinions here are those of the authors and may not reflect the opinions of ARPA-E or 3M.

Figure 1

2370

, and

Polymer electrolyte membrane (PEM) fuel cells are an attractive energy conversion device because of the potential for extremely high efficiency and low emissions. Pure hydrogen is an ideal fuel for PEM fuel cells but it is energy intensive to produce. Currently, the main method of producing hydrogen is steam reforming of natural gas. This is a break from the renewable energy future promised by hydrogen fuel cells. A greener approach to producing hydrogen is electrochemical reforming, the most obvious feedstock is water. Water electrolysis occurs at a relatively high potential (1.23 V), which makes it expensive both in terms of materials required and electricity. To fulfill the promise of fuel cells, more efficient methods of producing hydrogen must be found.

One way to reduce the energy required is to depolarize the PEM electrolyzer anode by using an alternate feed. This feed must have a lower oxidation potential than water and be generated renewably. Two alternative feeds we have looked at are methanol and phosphomolybdic acid.

Oxidation of methanol-water solutions to hydrogen takes place at a theoretical voltage of 0.03 V. Methanol fuel cells are a well-studied problem and the issues are known. The main issue is crossover of methanol to the cathode, lowering the total cell potential. We looked to see if this would be as problematic in a hydrogen pumping environment since there would be no oxygen at the cathode for the methanol to react with.

The other feed we examined was phosphomolybdic acid (PMA). PMA is a keggin ion which is readily reduced by biomass substrates in the presence of sunlight or heat.1Our idea was to apply the concept of a depolarized anode electrolyzer to a mediated electrochemical system. This system has the PMA in solution as a combination catalyst and charge carrier. The PMA is reduced by biomass and circulated to an anode where it is oxidized to release the hydrogen removed from the biomass. This gives many of the advantages of a flow battery, where power and capacity scale separately. Our focus has been on studying the anode performance characteristics of PMA.

 References.

1. Liu W, Mu W, Liu M, Zhang X, Cai H, Deng Y. Solar-induced direct biomass-to-electricity hybrid fuel cell using polyoxometalates as photocatalyst and charge carrier. Nat Commun. 2014;5:3208. doi:10.1038/ncomms4208.

2371

, , , and

An electrochemical hydrogen pump (EHP) is a device employed for compressing hydrogen gas. The compression efficiency of the EHP, which operates by an isothermal process in principle, is theoretically higher than that of a conventional mechanical compressor, which operates by an adiabatic process. The structure of an EHP cell is similar to that of polymer electrolyte membrane fuel cell (PEMFC). In the EHP, compressing process is realized by imposing a potential difference to catalyst coated membrane (CCM) and completed by 3 steps: hydrogen oxidation reaction (HOR) at anode, conduction of hydrogen protons through membrane and hydrogen evolution reaction (HER) at cathode.

Similar with PEMFC, overpotentials involved in EHP are categorized into two components: ohmic overpotential, mainly attributed to the conduction of protons through the membrane, and non-ohmic overpotential, which involves activation and concentration overpotentials. Due to the fast speed of hydrogen electrode reaction, ohmic overpotential is the main overpotential in the EHP. In order to reduce membrane resistance, bubbler humidifier same as that used in PEMFC system is conventionally utilized for humidifying membrane. However, different with PEMFC there is no water generation reaction occurs in the cathode side of EHP, the water at cathode dragged by hydrogen protons may be not enough for humidifying membrane at low current and high temperature conditions. Moreover, the effects of compression ratio (pressure of hydrogen at cathode) on overpotentials are important but not well discussed yet.

In order to solve the problems which mentioned above, we supposed a different humidification method and employed a reference electrode for separating overpotentials. An internal humidification method is employed which is realized by storing liquid water in the cathode end plate shown in Fig.1. The liquid water directly humidifies membrane, leading to higher conductivity of membrane than that using a conventional bubbler humidifier [1]. This method is expected to balance water transport through membrane and keep membrane hydrated at high temperature. Electrochemical impedance spectroscopy (EIS) is introduced to clarify the effects of compression ratio on ohmic overpotential and non-ohmic overpotential.

The compression ratio effects on overpotentials evaluated by EIS show following characteristics: For the part of ohmic overpotential, experimental results don't show any dependence on compression ratio which suggests that water transport in EHP is more like diffusion process rather than convection process. On the other hand, non-ohmic overpotential decreases with increasing of hydrogen pressure of cathode. Separation results with reference electrode clearly shows that decreasing of non-ohmic overpotential occurs at cathode and it is attributed to the increasing of exchange current density of HER by increasing hydrogen pressure.

Fitting I-V curves with a Volmer-Heyrovsy-Tafel expression [2] shows that the kinetic of HER and HOR are limited by diffusion process to/from electrode at anode/cathode. This is also confirmed by electrochemical impedance spectra data obtained at low frequency region. The concentration overpotential at anode is attributed to hydrogen diffusion through gas diffusion layer (GDL). In the EHP, in order to decrease hydrogen leakage, compression pressure supplied by end plates is set to 5MPa, which is much higher than common value (1~2MPa) used in PEMFC cell. At such a high compression pressure, GDL is compressed extremely, leading to small value of efficient gas diffusivity of hydrogen. Limiting current of anode calculated by hydrogen gas diffusion through GDL is less than 2A/cm2which agrees with the fitting results generally. The concentration overpotential at cathode is larger than that at anode. Frist, we suspected that water layer built at cathode may work as a mass transfer barrier for hydrogen gas and lead to large mass transfer resistance. However, an additional experiment carried out with bubbler humidification showed even larger mass transfer resistance at cathode. It is clear that mass transfer of hydrogen at cathode is not limited by water layer. It is possible that hydrogen diffusion in catalyst layer leads to the high concentration overpotential. The structure of catalyst layer is assumed to consist of gas pores and aggregates of Pt/C covered by ionomer layers [3]. Hydrogen that produced at reaction sites of cathode need to diffuse through this ionomer layer. With the liquid water humidification, water content of ionomer layer increases [4] which enhances the permeability of hydrogen and decreases mass transfer resistance.

Reference:

[1] T. A. Zawodzinski Jr., T. E. Springer, and S. Gottesfeld. Journal of The Electrochemical Society, 1993, 140(7) 1981-1985.

[2] P.M. Quaino, A.C. Chialvo, Electrochimica Acta, 2007(52) 7396–7403.

[3] Firat C, Ajay K. Prasad, Journal of The Electrochemical Society, 2014,161 (6) F803-F813.

[4] T. Sakai, E. Torikai. Journal of The Electrochemical Society, 1985, 132(6) 1328-1332.

Figure 1

2372

, , , , and

New trends in PEM water electrolysis systems development open up new technology gaps and requirements that have not been discussed before with respect to PEM water electrolysis. For example, hydrogen is considered as one of the best solutions for large-scale energy storage that comes from renewable and intermittent power sources such as wind and solar electricity [1], therefore, new megawatt PEM electrolysers are required. These trends are evident through new large-scale recent installations, especially for Power-to-Gas projects in Europe and plan to treat contaminated water at Fukushima Nuclear Power Plant.

Even though PEM electrolysis has in fact been used for quite a few years now without undergoing substantial improvements over those years, however, with the new focus on hydrogen as the energy carrier there is much more interest in low-cost and high-efficiency H2 production. There are two main ways to lower the cost of hydrogen production via PEM water electrolysis: to lower the capital expenses (CAPEX) and/or to lower the operating expenses (OPEX). 3M has recently demonstrated a very effective way to address reducing the high CAPEX by increasing the range of current densities where electrolyzers can operate from a maximum of about state-of-the-art 2.0 A/cm2, to as much as 20 A/cm2 by employing a novel 3M's proprietary Nano Structured Thin Film (NSTF) catalysts and more conductive 3M PFSA based electrolytes in the electrolyzer MEA [2].

What the cited work does not however addresses is an inherent limitation associated with 3M's Ir-NSTF based anode catalysts, namely their near complete inertness to HOR. An additional tradeoff between performance and gas-crossover also exists when thinner PEM membranes are used. Resulting high hydrogen cross-over creates a gap in the otherwise complete list of requirements for catalyst coated membrane (CCM) that has to be met to become successfully employed in PEM water electrolyzer [3-4]. This was plainly demonstrated in the early 3M catalyst trials where fuel cell derived Pt alloy based NSTF catalysts were used for water electrolysis. The reported durability of such electrolyzer catalysts was excellent, with performance however leaving room for improvement [5]. The performance aspect of 3M electrolyzer catalysts has since been addressed by switching to electrolyzer specific (and Ir based) catalyst compositions. Such approach has left a utility gap by eliminating hydrogen-crossover mitigating components out of the catalyst. To address this problem alternative means for mitigation strategies have to be devised.

In this work we will present some data, such as permeability (gas cross-over) of oxygen and hydrogen as a function of current density and other operational variables, aimed at establishing baselines for un-mitigated hydrogen cross-over of 3M electrolyzer MEAs based on 3M NSTF low-PGM loading catalyst and several types of PerFluoro Sulfonic Acid (PFSA) based PEM membranes, both widely commercially available (such as Nafion™ membranes) and their counterparts made by 3M. Dependence of such unmitigated hydrogen cross-over on applied pressure, pressure differentials, and temperature will be discussed and analyzed with reference to existing models of gas crossover [6-8](. In addition, we will also discuss challenges of in-situ and ex-situ gas –cross over measurements and also intend to present results of our initial attempts to employ alternative (to alloying Ir with Pt) mitigation strategies and gauge their respective effectiveness at various electrolyzer operating conditions and time frames. All proposed X-over mitigation strategies are selected with strong emphasis on compatibility of these candidate solutions with high speed/low cost roll-to-roll manufacturing process that will not be negatively affected by their properties and/or needed modifications.

  • In: D. Bessarabov, H. Wang, H. Li, and N. Zhao (Eds): "PEM Electrolysis for Hydrogen Production: Principles and Applications", CRC Press., 2015.

  • K. A. Lewinski, S. M. Luopa, "High Power Water Electrolysis as a New Paradigm for Operation of PEM Electrolyzer", Spring ECS Meeting, Chicago, IL, May 2015.

  • K. A. Lewinski, et al., "NSTF Advances for PEM Electrolysis - the Effect of Alloying on Activity of NSTF Electrolyzer Catalysts and Performance of NSTF Based PEM Electrolyzers", Fall ECS Meeting, Phoenix, AZ, Oct 2015.

  • K. A. Lewinski, et al., ECS Transactions, 69 (17) , p. 893-917 (2015).

  • M.K. Debe, et al., Journal of The Electrochemical Society, 159 (6), (2012), p. K165-K176.

  • C. Mittelsteadt, M. Umbrell, 207th ECS Meeting, Abstract #770

  • D. Bessarabov, "Gas Permeability of Proton Exchange Membranes", Chapter 21, in: PEM Fuel Cell Diagnostic Tools , Editor(s): H. Wang, Xiao-Zi Yuan, Hui Li, CRC Press, Pages: 443-473, 2011

  • M. Schalenbach at al., J. Phys. Chem. C 2015, 119, 25156−25169

Figure 1

2373

and

Porous electrodes for PEM electrolyser and fuel cell technologies are optimized primarily to provide a high specific surface area for the desired electrochemical reactions. However, the increase in surface area comes at the cost of increased voltage losses due to the transport of reactant and product species through the porous medium. In gas evolving electrolyzer electrodes, the formation and transport of gas bubbles plays an important role: on the one hand, gas in the pores replaces the electrolyte phase, thus hindering ion transport and reducing the effective ionic conductivity in the porous electrode; on the other hand, gas bubbles cover and deactivate a portion of the internal surface area.

The impacts of gas formation and removal must be accounted for in the design of electrolyzer electrodes. The focus of the presented work is on the fundamental understanding of the relationships between structure, properties and performance of porous gas-evolving electrodes. The approach combines a macro-scale performance model of the electrode with a micro-scale model of gas evolution. At the macro-scale, a classic porous electrode model describes the transport of ions, electrons and produced oxygen using effective medium theory and accounting for the gas phase volume and distribution.

The micro-scale model describes the formation and growth of gas bubbles based on chemical energy considerations. It can explain the experimentally found high oversaturation that is necessary to nucleate bubbles [1]. Additionally, the size of the bubble nucleus in the experiment was estimated.

The model was used to study the influence of structural parameters on the operation of porous electrodes and to establish guidelines for an enhanced electrode design. It was found that the transport regime of the dissolved gas, viz. diffusion control vs. transfer control at the liquid-gas interface, determines the bubble growth law. Applications of the model to gas evolving porous electrodes in electrolyzers with liquid electrolyte as well as polymer electrolyte electrolyzers (PEEC) will be discussed.

[1] Chen, Q., Luo, L. and White, H. S. Electrochemical generation of a hydrogen bubble at a recessed platinum nanopore electrode. Langmuir 31, 4573–4581 (2015).

Figure 1: Bubble growth on an artificial nucleation site at 20 mA/cm2. Comparison of model results with experimental data from C. Brussieux et al., Electrochimica Acta 56 (2011) 7194-7201. Deviation at large bubble radii due to deformation of the bubbles before detachment, which is neglected in the model.

Figure 1

2374

, , and

Hydrogen has been identified as an attractive energy carrier for the future in terms of on-site energy generation using for example fuel cells. The production of hydrogen and oxygen (in this case the by-product) from water electrolysis has been commercialised, especially for on-site H2 generation. For high volume H2 production using electrolysis, a more intricate system has been proposed in the form of the sulfur dioxide depolarised electrolyser. The SO2 electrolyser, consisting of a platinum anode and cathode separated by a proton exchange membrane, produces both H2 and H2SO4 from H2O and SO2according to equation 1 (anode reaction) and equation 2 (cathode reaction).

SO2 (g)+ 2H2O (l) = 2H+ (aq.) + 2e- + H2SO4 (aq.) Eo = 0.158 VSHE(1)

2H+ (aq.) + 2e- = H2 (g) Eo = 0 VSHE(2)

The anode is first depolarised by SO2 to facilitate the electrochemical decomposition of water to protons and electrons. Recombination of the protons (permeating through the PEM) and electrons produces hydrogen at the cathode while H2SO4 is formed at the anode. The primary advantage of the SO2 electrolyser over conventional water electrolysis is based on the theoretical potential needed to drive the reaction. By polarising the anode catalyst with SO2, the voltage is reduced to E0 = 0.158 VSHE, significantly less than the 1.23 VSHErequired for normal water electrolysis.

This paper will review the current literature on the operating methods for the SO2 electrolyser while including an in-depth study performed at HySA Infrastructure CoC, such as MEA manufacturing parameters, effect of H2S on cell performance. Additionally, the use of PBI based membranes [1] (manufactured by University of Stuttgart through a joint collaboration) in the electrolyser is evaluated by focussing on membrane stability, MEA acid doping and cell performance.

Initially the SO2 assisted water electrolyser was developed as part of the Hybrid Sulfur (HyS) cycle in the 1970's by the Westinghouse Corporation [2]. The SO2 electrolyser was operated, for the HyS cycle, using SO2saturated sulfuric acid as anode and clean sulfuric acid as cathode. The produced sulfuric acid is then returned to the decomposition step.

Sivasumbramanian et al. showed that the electrolyser can also be operated using a gaseous SO2 anode and a liquid cathode with a proton exchange membrane (PEM) [3] as separator. By this configuration the cell performance could be increased significantly. Staser et al. [4] further investigated this operating method in-depth with regards to the influence of water transport across the PEM on cell performance and acid concentration produced. This operating method was also shown for the application of SO2 usage in the flue gas compositions where H2S gas is present. Although a reduced performance was observed it was significantly less than expected when compared to fuel cells for example.

The development of high temperature membranes, such as PBI based polymers[1,5], for operating the electrolyser in the 120 – 160 ᵒC range has been shown to perform either better (at 80 ᵒC) or comparable to that of commercial PFSA membranes in the 80 - 90ᵒC range [5]. However, at this stage, only one article is available REF where humidified SO2was used as the anode feed using a sulfonated PBI membrane although low temperature (80 and 90 ᵒC) was used [6].

References

[1] Krüger AJ, Kerres J, Bessarabov D, Krieg HM. Evaluation of covalently and ionically cross-linked PBI-excess blends for application in SO2 electrolysis. Int J Hydrogen Energy 2015;40:8788–96. doi:10.1016/j.ijhydene.2015.05.063.

[2] Brecher LE, Spewock S, Warde CJ. The Westinghouse Sulfur Cycle for the thermochemical decomposition of water. Int J Hydrogen Energy 1977;2:7–15. doi:http://dx.doi.org/10.1016/0360-3199(77)90061-1.

[3] Sivasubramanian P, Ramasamy RP, Freire FJ, Holland CE, Weidner JW. Electrochemical hydrogen production from thermochemical cycles using proton exchange membrane electrolyzer. Int J Hydrogen Energy 2007;32:463–8.

[4] Staser JA, Weidner JW. Effect of Water Transport on the Production of Hydrogen and Sulfuric Acid in a PEM Electrolyzer. J Electrochem Soc 2009;156:B16–21.

[5] Kruger AJ, Cichon P, Kerres J, Bessarabov D, Krieg HM. Characterisation of a polaromatic PBI blend membrane for SO2 electrolysis. Int J Hydrogen Energy 2015;40:3122–33. doi:http://dx.doi.org/10.1016/j.ijhydene.2014.12.081.

[6] Jayakumar JV, Gulledge A, Staser JA, Kim C-H, Brian C B, Weidner JW. Polybenzimidazole Membranes for hydrogen and sulfuric acid production in the Hybrid Sulfur Electrolyzer. ECS Electrochem Lett 2012;1:F44–8.

D-01 Pt and Pt-alloy Cathode Catalysts - Oct 2 2016 8:40AM

2375

, and

In electrochemical systems, metal surface charging phenomena dictate the strength of electrostatic interactions between the electrified electrode and ions in solution. Historically, the potential of zero charge (pzc) of the metal has been employed to parameterize the surface charging relation [1][2].

However, for a Pt electrode, classical radiotracer experiments of Frumkin and Petrii revealed a non-monotonic charging behavior upon increasing the metal phase potential, with consecutive transitions from a negative to a positive surface charging region and further to another negative surface charging region [3]. Garcia-Araez et al. later reported a similar 'negative-positive-negative' charging relation derived from laser-pulsed experiments [4]. This alternating charging behavior defies the role of the pzc as a sufficient descriptor of electrode charging phenomena. The situation calls for an overhaul of the charging relation that must be derived from first principles. This study presents a refined structural model for electrified interfaces, as shown on the left side of Figure 1. The model accounts explicitly for charging effects caused by the formation of a surface oxide layer and a layer of oriented water molecules. An analytical expression for the charging relation is derived.

The charging relation is parameterized using experimental data for a Pt surface as well as DFT calculations. It exhibits the three different charging regions observed in the experiments cited above. To illustrate the impact of the modified charging relation, it is employed to study the oxygen reduction reaction in nanoporous electrodes of polymer electrolyte fuel cells.

References

[1]. Frumkin, A.; Gorodetzkaya A. Z. Phys. Chem. 1928, 136, 451.

[2]. Petrii, O. A. Zero charge potentials of platinum metals and electron work functions (Review). Russ. J. Electrochem. 2013, 49, 401.

[3]. Frumkin, A. N.; Petrii, O. A. Potentials of zero total and zero free charge of platinum group metals. Electrochim. Acta. 1975, 20, 347.

[4]. Garcia-Araez, N.; Climent, V.; Feliu, J. Potential-dependent water orientation on Pt (111), Pt (100), and Pt (110), as inferred from laser-pulsed experiments. Electrostatic and chemical effects. J. Phys. Chem. C. 2009, 113, 9290.

Figure 1

2376

and

While first-principles investigations are suitable for exploring oxygen reduction reaction (ORR), they are not straightforward in estimating the final performance of a fuel cell. We are interested in how the important properties of ORR can be extracted and bridged to fuel cell performance. We have calculated the redox potentials and activation energies of the elementary electrochemical reactions of ORR, and evaluated the cyclic and linear sweep volatmograms from rate and diffusion equations. Our results demonstrate how ORR properties affect fuel cell performance and indicate a future direction for fuel cell investigations. The adsorption energy of the intermediates must not only be tuned (a point already made by many researchers from what is referred to as the volcano plot), but the activation energy of the elementary reactions must also be decreased to obtain much higher fuel cell performance beyond the volcano top performance. We took first-principles molecular dynamics (FPMD) to simulate two elementary steps of the electrochemical oxygen reduction reaction on Pt(111);

O(ad) + H+ + e--> OH(ad)

and

OH(ad) + H+ + e--> H2O

In order to control the electrode potential, i.e. Fermi energy, we used constant-µ scheme [1] of the effective screening medium (ESM) method [2]. The redox potentials of these reactions were determined by changing the OH(ad) and O(ad) coverages. The activation energies of the reactions were determined using the blue-moon ensemble (bond constrained) method [3]. Using the parameters obtained, cyclic and linear sweep voltammograms were obtained.

 We found that it is necessary not only to tune the adsorption energy of the intermediates, O(ad) and OH(ad), which are the origin of the so-called volcano plot, but also to reduce the activation energy of the second elementary step.

 FPMD calculations were performed with "STATE" code [4], which uses density functional theory for electronic structure calculation with plane wave basis, ultrasoft pseudo potentials, and the GGA-PBE correlation functional. A unit cell of 8.38Å×9.67Å×26.04Å in size was established under the periodic boundary conditions, consisting of a slab of three layers of the Pt(111) plane, each containing 12 Pt atoms and about 32 water molecules on it. The details of the calculations are given elsewhere [5].

Computations were performed using the supercomputers at HPCC in Hokkaido University, IMR in Tohoku University, and ITC in the University of Tokyo. This work was partly supported by the NEDO (New Energy and Industrial Technology Development Organization) projects.

[1] N. Bonnet, T. Morishita, O. Sugino, and M. Otani, Phys. Rev. Lett. 2012, 109, 266101.

[2] M. Otani and O. Sugino, Phys. Rev. B 2006, 73, 115407.

[3] M. Sprik and G. Ciccotti, J. Chem. Phys. 1998109, 7737.

[4] Y. Morikawa, K. Iwata, and K. Terakura, Appl. Surf. Sci. 2000. 169-170, 11.

[5] O. Sugino, I. Hamada, M. Otani, Y. Morikawa, T. Ikeshoji, and Y. Okamoto, Surf. Sci. 2007, 601, 5237; M. Otani, I. Hamada, O. Sugino, Y. Morikawa, Y. Okamoto, and T. Ikeshoji, J. Phys. Soc. Jpn. 2008 77, 024802; T. Ikeshoji, M. Otani, I. Hamada, O. Sugino, Y. Morikawa, Y. Okamoto, Y. Qian, and I. Yagi, AIP Adv. 2012, 2, 032182.

2377

, , and

Certain aspects of the interpretation of rotating ring-disk electrode, RRDE, data for the oxygen reduction reaction, ORR, in aqueous electrolytes under forced convection have been examined. This presentation will highlight three specific topics:

  • • The correct interpretation of the fraction of the oxygen consumption that generates H2O2(aq), as extracted from the ratio of ring current to disk current denoted as XH2O2, A plot of this ratio vs ω-1/2, where ω is the rotation rate of the disk, should be linear and, with a slope proportional to k3/XH2O2 where k3 is the rate constant for the reduction of solution-phase hydrogen peroxide1

  • • Changes in the electrocatalytic activity of Pt for the oxygen and hydrogen peroxide reduction reactions (ORR, and HPRR, respectively) in an aqueous acidic electrolyte induced by the adsorption of bromide, as a model impurity, have been investigated using chronoamperometric techniques. Significant drops in the diffusion-limited currents for O2 reduction, as well as an increase in the measured H2O2(aq) at the ring were induced by the presence of 10 M KBr, and could only be observed for Br- > 0.25 regardless of Estep. This behavior was found to be consistent with a simple blockage of surface active sites by adsorbed bromide, as predicted by the attenuation model proposed by Levart.

  • • An analysis of O2- generation at the disk and subsequent oxidation at the ring (see Figure 1) of a RRDE, which undergoes a subsequent second-order dismutation2

Acknowledgements: This research was supported by a grant from NSF: CHE-1412060

References

1. N. S. Georgescu, A. J. J. Jebaraj and D. Scherson, ECS Electrochemistry Letters, 4, F39 (2015).

2. Z. Feng, N. S. Georgescu and D. A. Scherson, Analytical Chemistry, 88, 1088 (2016).

Figure 1

2378

, , , , and

In order for hydrogen polymer electrolyte membrane fuel cell (PEMFC) technology to be economically competitive, overall system costs still need to be lowered. A key component of cost is the amount (grams) of Pt used in the membrane electrode assembly (MEA). Significant advances in electrocatalyst research, especially focused on the oxygen reduction reaction (ORR) for PEMFC cathodes, have thus been made and have led to the development of highly active catalysts capable of delivering target power at reduced Pt loadings.1 Unfortunately, as the Pt loading i.e. electrode roughness factor, decreases, an additional resistance emerges leading to voltage loss. This resistance, associated with phenomena at or near the Pt surface, is poorly understood and has been shown to have the greatest impact on low loaded Pt electrodes (<0.1 mgPt/cm2) operating at high current densities (>1.0 A/cm2). Therefore, an understanding of the kinetic parameters governing the electrochemical reactions, specifically at the cathode, is essential in order to distinguish between voltage losses that are kinetically derived from those originating from other sources. An understanding of the nature and mechanisms of performance losses in PEMFCs is essential in order for industry to best direct R&D resources.

Discrepancies observed in theoretical kinetic predictions and real performance polarization curves for low Pt loaded MEAs (>0.1 mg Pt/cm2) have led to alterations in the simple Tafel kinetic model2 by including an additional term related to oxide coverage.While the operating cell potential inherently drops as the catalyst loading decreases, oxide-coverage kinetics become increasingly important in modeling performance especially when transitioning between catalyst surfaces with little to no oxide coverage to those with high oxide coverage. Since some kinetic parameter values change considerably when incorporating an oxide coverage term into the simple tafel model, it is important to know how dependent these changes are in regards to catalyst type and support, as well as in how oxide coverage is defined. Because specific oxide species generated at the catalyst surface can differ with changing conditions e.g. Pt-OH (1e- process), Pt-O (2e- process), it is difficult to obtain a unified model which accurately describes oxide formation. Subsurface oxide generation has also been shown to occur,which can skew interpretations of oxide measurements. Additionally, oxide formation has been shown to differ substantially among different catalyst surfaces and supports.

This work attempts to clarify the impact that oxide species have on kinetic parameters (exchange current density (ios), reaction order (γ), activation energy (EA) etc.) for Pt-based catalysts. Experiments were performed under pure oxygen using a 5cm2 differential cell with MEAs consisting of Pt/V, Pt/HSC, and Pt alloy/HSC at loadings of 0.2 – 0.05 mg Pt/cm2. Sub-ambient system pressures were employed by means of a vacuum panel in order to lower pure oxygen partial pressure while minimizing gas transport effects. An example measurement is shown in Figure 1, with the oxide and transport free surface potential and current region highlighted. Interpretations of oxide measurement techniques as well as the impact of different catalyst surfaces and supports are discussed. As catalyst activity continues to improve, future implications and significance of oxide kinetics are also considered.

Figure 1. iR-free potential vs log current density as a function of oxygen partial pressure for Pt/Vu electrocatalyst.

The authors would like to acknowledge funding from the U.S. Department of Energy under CRADA #CRD-14-539.

References

1. A. Kongkanand and M. F. Mathias, The Journal of Physical Chemistry Letters, 7, 1127 (2016).

2. K. C. Neyerlin, W. Gu, J. Jorne and H. A. Gasteiger, Journal of the Electrochemical Society, 153, A1955 (2006).

3. N. Subramanian, T. Greszler, J. Zhang, W. Gu and R. Makharia, Journal of The Electrochemical Society, 159, B531 (2012).

4. B. Conway, B. Barnett, H. Angerstein‐Kozlowska and B. Tilak, The Journal of chemical physics, 93, 8361 (1990).

Figure 1

2379

, , and

H2-fed proton exchange membrane fuel cells (PEMFCs) represent the most advanced fuel cell technology and have a great deal of potential applications in low/zero-emission electric vehicles, distributed home power generators, and power sources for small and portable electronics. However, the commercialization of PEMFC technology has been greatly hindered by some challenges, mainly sluggish kinetics of the oxygen reduction reaction (ORR) at the cathode and the high cost of the noble metal Pt (1, 2).

Alloying platinum with non-noble metals such as Co is an effective approach to improve the catalytic performance and reduce the usage of Pt. In addition, through transferring Pt nanoparticles (NPs) into a hollow nanostructure, electrocatalytic performance can be improved greatly due to its relatively lower density and higher surface area-to-volume ratio than its solid counterpart (i.e., NPs).

In this work, Pt hollow nanospheres (HNSs) were fabricated through a replacement reaction of Co atoms by PtCl62- ions to reduce Pt usage and improve the catalytic activity towards the ORR. Carbon black Vulcan XC-72R (VC) was introduced into a solution prior to the addition of Co(II) and the formation of Co NPs and the replacement of Co by PtCl62-ions for a uniform dispersion of Co NPs and the Pt HNSs on the carbon support. Some Co atoms have alloyed Pt in the synthesis and exist in the Pt HNSs (3).

The hollow mesoporous core in the Pt HNSs can be utilized as an electrolyte solution buffering reservoir to minimize the diffusion distance to the interior surface of the porous shell of Pt crystallites while the porous nanochannels (i.e., the micropores between the Pt crystallites) in the shell open to the mesoporous hollow core form fast mass transport networks providing more accessible sites for oxygen transfer. In addition, triple phase boundaries (i.e., gas-electrolyte-Pt NPs) can be developed more easily in the Pt HNS enabling individual Pt crystallites in the shell to be accessible to electrolyte ions and oxygen, and thus catalytically active. Furthermore, alloying Pt with Co might result in a significant lattice shrinking because of the change in Pt-Pt bond distance which also contributes to the improved electrocatalytic activity. In contrast, for the state-of-the-art Pt NPs/VC catalyst, the Pt NPs can agglomerate more easily to form larger particles, resulting in reduced active sites. Besides, the interior voids between Pt NPs may not be accessible to electrolyte ions and do not contribute to the ORR activity due to the lack of triple phase boundaries. As a result, the as-developed Pt(20 wt%) HNS/VC catalyst outperforms significantly the state-of-the-art Pt(20 wt%)NP/VC catalyst.

References

1. B. Fang, M. Kim, J. Kim, M. Song, Y. Wang, H. Wang, D. Wilkinson and J.-S. Yu, J. Mater. Chem., 21, 8066 (2011).

2. B. Fang, N. Chaudhari, M. Kim, J. Kim and J.-S. Yu, J. Am. Chem. Soc., 131,15330 (2009).

3. B. Fang, B. Pinaud and D. Wilkinson, Electrocatalysis, DOI: 10.1007/s12678-016-0311-4 (2016).

Figure 1

2380

, , and

One of the strategies to further increase PEFC performance is elevation of the operation temperature, reducing the reaction overvoltage of oxygen reduction reactions. On the other hand, durability of materials used in PEFC is assumed to be lowered at the higher temperature condition. In this study, understanding degradation phenomena of electrocatalysts at the higher temperature is particularly focused on in order to develop durable electrocatalysts at the elevated temperature.

MEAs were prepared using the standard catalyst and electrolyte, Pt/Ketjen black catalyst (TEC10E50E, TKK Corp.) and Nafion, respectively. Three different conditions of operation temperature and humidity (80 °C/RH100%, 105 °C/RH57%, and 80 °C/RH57%) were applied. Durability of MEAs in three conditions was studied through the potential cycles between 0.6 and 1.0 V assuming "load cycles" of FCVs.[1], [2] 

Since Pt dissolution/re-precipitation is assumed to occur during the cycles, the changes in activation overvoltage and ECSA were carefully examined under three conditions. After 20000 potential cycles, activation overvoltage and ECSA increased and decreased, respectively, the most at 105 °C/RH57%. On the other hand, such degradation was much suppressed under the operation at 80 °C/RH57%. In order to understand the difference in degradation phenomena, the amount of water molecules was considered. With the presence of a large amount of water molecules, the rate of reducing ionized Pt (Pt2+) into Pt metal is assumed to be smaller. Then, in the case of 80 °C/RH57%, the amount of water molecules was smallest among three conditions, leading to fast and reversible reduction of Pt2+to Pt metal. Therefore, Pt particle growth was believed to be suppressed. However, if two factors, the amount of water and temperature, are considered, the temperature factor most likely influences more on Pt particle growth. 

Reference

[1] A. Ohma et al., ECS Trans, 41 (2011) 755.

[2] M. Kitamura et al., ECS Trans, 69 (2015) 701.

2381

, and

The durability of polymer electrolyte fuel cells (PEFCs) is a big concern for its applications like FCVs to be commercialized on a large scale. In particular, the instability of Pt-based cathode catalyst has been discussed intensively. Platinum (Pt) dissolution is recognized as one of the degradation mechanisms for carbon supported Pt catalysts (Pt/C) in PEFCs. It has been found that repeated potential cycles accelerate Pt dissolution.[1-3] Fundamental studies in aqueous media revealed preliminary behaviors of Pt dissolution during one potential cycle. The up-to-date data were mostly obtained near room temperature, and a dominant cathodic dissolution during reduction of Pt oxide was understood. As reported by Fuller et al.,[4] however, degradation of Pt/Cs at an elevated temperature near 60 to 80 °C, whereat a PEFC normally operates, was severer than that at room temperatures. Thus, it is necessary to evaluate Pt dissolution from Pt/Cs at such temperatures.

In prior studies,[5-6] we compared Pt dissolution at 25 °C and 65 °C by inductively coupled plasma mass spectrometry (ICP-MS) and a channel flow double electrode (CFDE). The experiments had some difficulties, but the results showed a ca. 5-time enhancement from 25 to 65 °C. The cathodic dissolution was enhanced because larger amount of Pt oxide was formed at 65 °C. The effect of temperature on the anodic dissolutions and the place exchange process remains unclear yet.

In this study, we evaluate Pt dissolution under temperatures ranging from 20 to 80 °C. We test a series of commercial Pt/C catalysts produced by TKK that have Pt loadings from 20 to 70 wt% and average Pt particle size from ca. 1.5 to 5 nm. We limited the potential region between 0.6 and 1.4 V, which is closely related to the operation region of a PEFC cathode. We use ICP-MS to evaluate to overall dissolutions and a CFDE to analyze the dissolution behavior in one potential cycle. The attached Figure showed a typical CFDE data presenting the current on a working electrode loading a 30wt% Pt/C (TEC10E30A) under potential cycling (IWE) and the current recorded on a collector electrode (ICE) at 80 °C in 0.5 M H2SO4. The ICE helps to understand the dissolution of Pt4+ and Pt2+ complex during one potential cycle. Similar analysis will be discussed at other temperatures like 20, 40, and 80 °C in the presentation.

Reference

  • Z. Wang, E. Tada, and A. Nishikata, J. Electrochem. Soc., 161, F380 (2014).

  • A.A. Topalov, S. Cherevko, A.R. Zeradjanin, J.C. Meier, I. Katsounaros, and K.J.J. Mayrhofer, Chem. Sci., 5, 631 (2014).

  • P. Jovanovič, A. Pavlišič, V.S. Šelih, M. Šala, N. Hodnik, M. Bele, S. Hočevar, and M. Gaberšček, ChemCatChem, 6, 449 (2014).

  • W. Bi and T.F. Fuller, J. Electrochem. Soc., 155, B215 (2008).

  • Z. Wang, E. Tada, and A. Nishikata, J. Electrochem. Soc., 163, F1-F3 (2016).

  • Z. Wang, E. Tada, and A. Nishikata, ECS Trans., 69, 255 (2015).

Figure 1

2382

, and

Introduction

Pt-M alloys catalysts have attracted attention over the years and expected to be used in proton exchange membrane fuel cells (PEMFCs) cathode catalysts due to their high oxygen reduction activity than pure Pt and reduction of Pt usage in PEMFCs [1]. However, Pt-M catalysts is easily degraded by both Pt and M dissolution from Pt-M alloys during on-off cycle and load cycle, which leads the performance loss of PEMFCs [2]. Thus, to understand the nature of dissolution mechanism of both Pt and M from Pt–M alloy catalysts is very significant to advance durability of PEMFCs under its operating conditions. In this study, we investigated that the dissolution mechanism of Pt-50 at% Fe alloy (Pt50-Fe50) and Pt50-Cu50 under potential cycling with a channel flow multi electrodes (CFME), and elucidate the effect of additive element to the dissolution behavior of Pt50-M50.

Experimental

Pt50-Fe50 and Pt50-Cu50 were subjected to potential cycling tests at 298 K using Ar-purged 0.5 M H2SO4 solution with a CFME. A double junction KCl-saturated Ag / Ag-Cl electrode was used as reference electrode, and Au coil was used as the counter electrode. Flow rate of the electrolyte was 10 cm s-1 in order to establish laminar flow condition. Potential cycling tests were employed between 0.05 V and 1.4 V vs. SHE at 20 mV s-1. Before potential cycling, Pt50-M50 was set at 0.45 V vs. SHE in order to remove Pt oxide layer formed on Pt–M alloy surface in ambient air and to measure baseline currents of collector electrodes (CEs). Reactions and potentials for detecting the dissolved Fe2+, Fe3+, and Cu2+ on Au-CEs are as follows,

Fe2+ → Fe3+ + e- (Ec = 1.0 V vs. SHE) (1)

Fe3+ + e-→ Fe2+ (Ec = 0 V vs. SHE) (2)

Cu2+ + 2e-→ Cu (Ec = 0 V vs. SHE) (3)

Dissolved Fe2+, Fe3+, and Cu2+ from Pt50-M50 are monitored by the current change of CEs.

Results and discussion

 Figure 1 shows M detection current with a CFME during potential cycling, and red and blue line show the results of Pt50-Fe50 and Pt50-Cu50, respectively. As shown in inset, Fe dissolution from Pt50-Fe50 starts from around 0.3 V in anodic scan. Then Fe dissolution is clearly enhanced between 0.6 and 1.4 V in both anodic and cathodic scan, and finally terminate at 0.3 V in cathodic scan. Wang et al reported that Pt dissolution occur above approximately 0.6 V [3], accordingly major Fe dissolution is enhanced by Pt dissolution. On the contrary, minor Fe dissolution between 0.3 and 0.6 V in both anodic and cathodic scan is contributed by Pt surface diffusion, because Pt atoms are easily diffuse on Pt-M alloy surface in this potential range due to no adsorbed species on its surface [4]. Cu dissolution from Pt50-Cu50 only appears above 0.6 V, whereas Fe dissolution from Pt50-Fe50 starts from 0.3 V. Moreover, the amount of Cu dissolved from Pt50-Cu50 is smaller than that from Pt50-Fe50, which indicate that Cu is more durable elements than Fe. This is explained by the difference of standard electrode potentials between Cu and Fe. Pt-enriched layer formed on Pt50-Cu50 is denser and flatter than that on Pt50-Fe50, because the amount of dissolved Cu is small due to more noble standard electrode potential than Fe. Flat Pt-enriched layer inhibits Pt surface diffusion between 0.3 and 0.6 V, and Pt dissolution above 0.6 V.

Reference

[1] V. R. Stamenkovic, B. S. Mun, M. Arenz, K. J. J. Mayrhofer, C. A.Lucas, G. Wang, P. N. Ross, and N. M. Markovic: Nat. Mater.6 (2007) 241–247.

[2] H. A. Gasteiger, S. S. Kocha, B. Sompalli, and F. T. Wagner: Appl. Catal. B: Environ.56 (2005) 9–35.

[3] Z. Wang, E. Tada, and A. Nishikata: J. Electrochem. Soc.161 (4) (2014) F380–F385

[4] Q. Xu, E. Kreidler, and T. He: Electrochim. Acta55 (2010) 7551–7557.

Figure 1

2383

, , , , , , , and

We developed connected carbon-free platinum-iron (PtFe) nanoparticle catalysts with porous hollow capsule structure (Fig. 1) as oxygen-reduction-reaction (ORR) electrocatalysts for polymer electrolyte fuel cells (PEFCs).[1] This catalysts consist of a beaded network by connected PtFe nanoparticles with a crystallite size of 6~7 nm and a chemically-ordered (face centered tetragonal) superlattice structure. The beaded network is electrically conductive; thus, carbon-supports can be removed from catalyst layers in PEFCs. We have demonstrated that an MEA prepared using a carbon-free cathode of connected PtFe-nanoparticle catalysts was highly durable against start-up/shut-down operations because of the elimination of carbon-corrosion problems. Moreover, an ORR specific activity of a connected PtFe catalyst is about 9 times higher than that of a commercial Pt-nanoparticle catalyst supported on carbon black (Pt/C).

In this study, the structural effects of a connected PtFe catalyst on an ORR activity were investigated in order to elucidate the factors of its high ORR activity. Note that the connected PtFe catalyst exhibited about twice higher ORR specific activity than the PtFe-nanoparticle catalyst (just nanoparticles without any connection between them) on carbon black. In addition, the ORR specific activity of the connected PtFe nanoparticles with hollow structure (with no supports) was 1.5 times higher than that of the connected PtFe nanoparticles on the electroconductive carbon support. These results indicate that the unique structures of connected nanoparticle catalysts, such as connected beaded network and hollow structure, enhance an ORR activity.

We also have performed the refined structural analyses of a connected PtFe catalyst by using in-situ X-ray absorption spectroscopy (XAS) combined with Rietveld structure refinement with powder X-ray diffraction data, and spherical aberration corrected scanning transmission electron microscope (Cs-corrected STEM). The in-situ XAS analysis disclosed shorter Pt‒Pt bond distance and more Pt 5d vacancy of the connected PtFe catalyst compared with the commercial Pt/C. In addition, the Cs-corrected STEM observations revealed that the beaded network formed by connected PtFe nanoparticles possessed polycrystalline structure. In this presentation, the effects of the geometric and electronic structures of a connected PtFe catalyst on an ORR activity will be discussed in details. The knowledge obtained in this study provides guidelines for the design of ORR catalysts with enhanced ORR activities to achieve a high performance PEFC.

[1] T. Tamaki, H. Kuroki, S. Ogura, T. Fuchigami, Y. Kitamoto, and T. Yamaguchi, "Connected nanoparticle catalysts possessing a porous, hollow capsule structure as carbon-free electrocatalysts for oxygen reduction in polymer electrolyte fuel cells", Energy Environ. Sci., 2015, 8, 3545-3549.

Figure 1

2384

, , , , , and

The widespread use of polymer-electrolyte fuel cells (PEFCs) requires more active, stable, and low cost electrocatalysts than platinum. Bimetallic electrocatalysts such as platinum-iron (PtFe) have been attracting a great deal of attention as an oxygen reduction reaction (ORR) catalyst since they have demonstrated both higher oxygen reduction activity and improved stability with much smaller amounts of platinum. Connected PtFe-nanoparticle catalyst with a beaded network by connected PtFe nanoparticles and a superlattice structure (a chemically-ordered face-centered tetragonal (fct) structure) exhibit specific ORR activity that is higher than those of the commercial Pt catalyst supported on carbon black (Pt/C) and even fct-PtFe nanoparticles (without any beaded networks) supported on carbon black. [1] Thus, the unique structures of connected PtFe-nanoparticle catalysts, such as connected beaded network and hollow structure, enhance an ORR activity.

Electrochemical x-ray photoelectron spectroscopy (EC-XPS) applied for the reaction mechanism analyses of PEFC catalysts. Electrochemical reactions occur at catalyst electrode surface, and then, the electrode is transferred to an UHV chamber immediately. Finally the electrode surface involving reaction intermediates is analyzed by using XPS without exposing to the air. Since structure of electrochemical-double layers is maintained during the measurement process, we can identify the intermediate species, and evaluate adsorption strength, etc., namely, we can analyze reaction mechanism. In this study, we performed electrochemical x-ray photoelectron spectroscopy (EC-XPS) measurements in order to understand the factors that influence the high ORR activity of the connected PtFe-nanoparticle catalyst. The EC-XPS results for connected PtFe-nanoparticle catalysts under ORR conditions showed the adsorption of oxygen species on the catalyst surfaces, indicating that ORR intermediates were detected. We could observe ORR processes of connected PtFe-nanoparticle catalysts. In this presentation, we will also show the EC-XPS measuring data for other platinum catalysts and discuss the factors that influence the high ORR activity of the connected PtFe-nanoparticle catalyst in detail.

[1] T. Tamaki, H. Kuroki, S. Ogura, T. Fuchigami, Y. Kitamoto, and T. Yamaguchi, "Connected nanoparticle catalysts possessing a porous, hollow capsule structure as carbon-free electrocatalysts for oxygen reduction in polymer electrolyte fuel cells", Energy Environ. Sci., 2015, 8, 3545-3549.

A-02 Modeling - Oct 2 2016 1:00PM

2385

, and

The current-voltage performance of polymer electrolyte fuel cells (PEFCs) in the intermediate voltage region has received attention from the standpoint of increasing the maximum power of the cell. The performance-limiting factors in this kinetic-mass transport mixed control region were first considered as a protonic resistance in the polymer electrolyte, an oxygen diffusion resistance in the pores of the catalyst layer (CL) and gas diffusion layer (GDL), and a diffusion resistance of the dissolved oxygen in the polymer electrolyte that covers the catalyst. These three factors were taken into consideration in mathematical models that predict the performance, which resulted in the predictions much better than experimental results. To bridge the gap, the diffusion resistance in the catalyst agglomerates, which are composed of carbon-supported catalysts and an ionomer, was proposed. However, huge agglomerates that the models assumed were not observed in the scanning electron micrographs of the cross sections of the catalyst layers. In the meanwhile, Kudo and Morimoto measured the oxygen diffusion resistance using Nafion thin films and concluded that the interfacial resistance at the Nafion-Pt or Nafion-gas interface gave rise to the diffusion resistance [1]. The existence of the resistance at the Nafion-Pt interface was supported by molecular dynamics simulations [2]. In this work, the interfacial resistance estimated from the measurement of limiting current densities was incorporated into the performance model and the model predictions were compared with experimental results. The discrepancy between them was assumed to stem from a dependence of diffusion resistance on the electrode potential and then the dependency was analyzed.

In the experiments, the cell was operated at 65 °C with a large excess flow rate of hydrogen and 1% oxygen balanced with nitrogen, both humidified to 80% relative humidity, under atmospheric pressure. The condition was selected to reduce the overpotential variation in the through-plane direction caused by the protonic resistance and to avoid the effect of the produced water and generated heat, and thus to emphasize the effect of oxygen diffusion. Following the seminal studies of Mashio et al. [3], we decomposed the diffusion resistance into two resistances in series, namely, the one caused by molecular diffusion (Rmolec) and the remainder (Rother). Rmolec was estimated with the aid of limiting-current density measurements using 1% oxygen balanced with helium instead of 1% O2-N2. Rother was calculated by subtracting Rmolec from the overall diffusion resistance. We well call Rother determined by this procedure as Rotherref.

The main features of the model to analyze the experimental results are as follows: (1) One-dimensional through-plane distribution is considered; (2) The anode CL and GDL are neglected; (3) Overpotential distribution and hence the reaction rate distribution is taken into account; (4) The oxygen reduction reaction rate is expressed by Tafel equation and is proportional to the oxygen concentration at the catalyst surface; (5) Oxygen concentration at the catalyst surface is calculated using the diffusion resistance and reaction rate.

In Figure 1a, an experimental result (dotted line) and a model prediction (solid line) of the performance of a cell are compared. In the model, constants in Tafel equation are selected so that the prediction fits to the experiment in the activation-controlled region; the membrane and CL resistances were the values measured by impedance spectroscopy; Rmolec and Rotherref were used as the diffusion resistances. The model prediction gives larger current density than the experiment in the intermediate cathode potential region. We assumed that the discrepancy stemmed from the constant Rother (= Rotherref), and therefore hypothesized that Rother is a function of electrode potential. We therefore increased Rother from Rotherref at each potential to find the value that the current density from model prediction matched the experimental value. The resulting Rother is shown in Fig. 1b as a function of cathode potential (Rotherref is shown as a dashed line). Rother increased with potential except for high potential region, where the effect of diffusion resistance on the performance is less significant and hence the accuracy of the fitting is deteriorated. The effect of catalyst loading will be discussed and a mathematical formulation of Rother will be presented.

[1] K. Kudo, Y. Morimoto, ECS Trans., 50, 1487 (2012)

[2] R. Jinnouchi, K. Kudo, N. Kitano, Y. Morimoto, Electrochim. Acta, 188, 767 (2016)

[3] T. Mashio, A. Ohma, S. Yamamoto, K. Shinohara, ECS Trans., 11, 529 (2007)

Figure 1

2386

, and

Polymer electrolyte fuel cells (PEFCs) are expected as a next-generation power sources due to their low emissions and high efficiencies. However, it is required to improve the power density at lower Pt loading. It is known that the losses of the power density are caused by activation losses, ohmic losses and mass transport losses, and that the mass transport losses are dominant at higher current density. Moreover, it is suggested that one of the causes of the mass transport losses is the shortage of oxygen in the cathode side of PEFC. In the cathodic catalyst layer, there are Pt catalysts on supported carbon microparticles, and those particles are covered with ionomer films which are composed of polymer electrolytes and water molecules. The ionomer has two properties for the water-generating reactions in the cathodic catalyst layer: the proton conductivity and the oxygen permeability. In particular, the dependence of the oxygen permeability on water content has not been clear.

In this study, we analyzed water content dependence of oxygen permeation properties in ionomer on Pt surface using molecular dynamics simulations. Regarding the calculation system, Nafion was adopted to the polymer electrolyte in the ionomer, and water content, λ, was defined as the ratio of the number of sulfo groups in Nafion to the number of water molecules. We set λ = 3, 7, 11 and constructed a system of an equilibrium state of ionomer on Pt surface at each water content. Then oxygen molecules were inserted above the ionomer and permeate the ionomer along the thickness direction at steady state.

Firstly, the density distributions were obtained to evaluate the structural properties of the ionomer. As a result, the region of the ionomer is divided into three region: ionomer/gas interface, bulk region, ionomer/Pt interface. Next, the number of permeated oxygen molecules through the ionomer was counted, and the oxygen permeability of the ionomer was estimated at different water content. As a result, the oxygen permeability decreases with increasing water content. Gas permeability is estimated using the permeation coefficient which is obtained by the product of the diffusion coefficient and solubility coefficient of gas molecules, indicating that the diffusivity and solubility are governing factors of the permeability. Therefore, the oxygen diffusivity and solubility was evaluated to discuss the oxygen permeation mechanism in the ionomer. The oxygen diffusion coefficient was obtained by the density distribution of the oxygen molecules along the thickness direction using the Fick's law. As a result, the oxygen diffusivity in each region of the ionomer decreases at lower water content while slightly increases at higher water content, with the increase in water content. On the other hand, the oxygen solubility coefficient was calculated using the test-particle-insertion method, and the solubility in each region of the ionomer decreases with increasing water content. Finally, the oxygen permeability was evaluated using the diffusivity and solubility in each region. Consequently, the oxygen permeability decreases with increasing water content in each region, and the permeability in the ionomer/Pt interface is the smallest, which indicates that the oxygen dissolution in the ionomer/Pt interface is dominant in the oxygen permeation through the ionomer.

2387

, , , , , , and

The objective of this work is to gain the mechanical understanding of the water management in the proton exchange membrane fuel cell (PEMFC). Water is a by-product of the fuel cell reaction and its amount is proportion to the current of fuel cell output. Water is used to retain the proper hydration level in the PEM. At the same time, condensed water in the flowfield and the gas diffusion layer (GDL) reduces oxygen transport to the oxygen reduction reaction (ORR) area. Therefore, excess accumulation of condensed water causes to lower the performance, particularly at the higher current densities. Detailed simulation of two-phase water in the GDL is significantly important to understand the water management.

This work shows the successful in development of a multi-scale calculation technique that incorporates various scales of numerical model in a fuel cell and simultaneously performed a prediction. The outcomes of this work are, (i) development of two-phase fluid model in the micro-scale porous structure of GDL, and (ii) integration of this micro-scale GDL fluid model with a state-of-the-art macro-scale flowfield fluid model to predict two-phase water in the fuel cell. Figure 1 gives the concept of modeling approach at each component of the fuel cell. The flowfield in the bipolar plate and MEA models are calculated using traditional computational fluid dynamics (CFD) method with existing PEMFC model [1-3]. The two-phase fluid in the micro-scale porous structure of GDL is numerically predicted by Lattice Boltzmann Method (LBM) [4-5]. The solution of each iteration from CFD and LBM are required to simultaneously exchange for the next iteration until all solutions are converged, especially at the interfaces. In this figure, it shows the flowfield is transformed into the macro-scale CFD model and the image of detail structured GDL is converted into micro-scale LBM model. Actual porous structure imaging of GDL is obtained by the nano-scale high special resolution x-ray computed tomography. Figure 2 shows an example of the predictions from the simulation when the macroscopic flowfield model is combined with the microscopic GDL model while performing the calculation. The predictions give the detail distributions of variables in every components of the multiscale model. In this figure, the current density distribution on MEA surface, the liquid water transport inside the GDL, and the temperature of solid and fluid phases inside the GDL are presented.

References:

1. S. Shimpalee, D. Spuckler, and J. W. Van Zee J. Power Sources, 167 (2007) 130-138.

2. S. Shimpalee, M. Ohashi, C. Ziegler, C. Stoeckmann, C. Sadeler, C. Hebling, J. W. Van Zee, Electrochemica Acta, 54 (2009) 2899-2911.

3. S. Shimpalee, J. of Electrochem. Soc., 161 (2014) E3138-E3148.

4. U. Frisch, B. Hasslacher, Y. Pomeau, Physical review letters 56 (14) (1986) 1505-1508.

5. G. R. McNamara, G. Zanetti, Physical Review Letters 61 (1988) 2332-2335.

Figure 1

2388

, , , and

Polymer electrolyte fuel cells (PEFCs) have been developed for automotive, residential, and portable power applications. Typically, PEFC stacks use platinum (Pt) or Pt alloys as the oxygen reduction reaction (ORR) catalyst in the cathode. Although Pt performs well, its raw material cost is significant and limits future cost reductions. As an alternative, platinum group metal-free (PGM-free) catalyst may have the potential to significantly reduce PEFC costs. Despite recent advances in PGM-free catalyst ORR activity,1 the volumetric activity remains well below that of Pt cathodes and results in thick electrodes with significant transport losses.

To advance PGM-free catalyst electrodes, it is important to understand the coupling between electrochemical reactions, transport phenomena, and microstructure in the catalyst layer. Here, we expand this understanding using a combination of nano-scale X-ray computed tomography (nano-CT) and direct simulations on the CT images. Nano-CT imaging was done for cyanamide-polyaniline-iron PGM-free catalyst cathodes with three different Nafion® loadings (35, 50 and 60 wt%). Using Zernike phase contrast we differentiated the solids (Nafion®, carbon, and iron) and the pore phase,2 and by staining the electrode with cesium, we distinguished the Nafion® phase by absorption contrast in a second scan. The two data sets were overlaid to complete the structure of the whole cathode.3 In other words, at each voxel we have the volume fractions of pore, Nafion®, and carbon catalyst phases. To illustrate, Figure 1a shows the Nafion® fraction over the simulation domain where the Nafion® had agglomerated and was heterogeneously distributed. The spatially-resolved volume fraction data is then input to the simulation of oxygen, proton, and electron transport as well as the local ORR volumetric current. Figure 1b shows the ionic potential and current streamlines in Nafion® phase resulting from the distributed reaction.

Acknowledgement

The authors gratefully acknowledge the support the support of the technology development manager Nancy Garland and funding from the US Department of Energy, the Office of Energy Efficiency and Renewable Energy, Office of Fuel Cell Technologies. The nano-CT instrument was acquired with the support of a Major Research Infrastructure award from the National Science Foundation under Grant No. 1229090.

References

1. G. Wu, K. L. More, C. M. Johnston, and P. Zelenay, Science, 332, 443–447 (2011).

2. S. Komini Babu, H. T. Chung, G. Wu, P. Zelenay, and S. Litster, ECS Trans., 64, 281–292 (2014).

3. S. Komini Babu, H. T. Chung, P. Zelenay, and S. Litster, ECS Trans., 69, 22–33 (2015).

Figure 1

2389

, , , and

INTRODUCTION

Although extensive studies have been conducted on PEFC behavior and a lot of knowledge has been published, understanding the PEFC is still not easy since the system is considerably complicated. By investigating a dimensionless form of equations describing the cathode catalyst layer (CL), the cathode model was simplified and a few intrinsic moduli which control the dimensionless rate profile were proposed (1). The impact of convection on the oxygen reduction reaction (ORR) rate and some case studies are demonstrated.

DIMENSIONLESS MODEL

The ORR rate is proportional to the oxygen partial pressure at fixed cathode emf. The emf dependency follows a Tafel equation. Oxygen balance in cathode catalyst layer is expressed as follows:

d(NgyOCgDeOdyO/dz) / dz = –rvc[1]

where Ng is the total gas flux, yO is the oxygen mole fraction, and DeOis the effective oxygen diffusivity (2). Proton flux can be calculated by the reaction stoichiometry. Proton potential follows the Ohm's law.

With the 1D model of cathode CL, 12 variables including boundary conditions have to be specified so as to determine the profiles of partial pressure, emf, reaction rate, etc. Dimensionless forms of model equations were derived, which include the following dimensionless moduli we propose:

MO(C)m δ(C) (kvcm RT / DeO)1/2[2]

Mp(C)m =  δ(C) {4F kvcm pOc / (σepbc)}1/2 [3]

where σep is the effective proton conductivity, bc is the Tafel slope, δ(C) is the CL thickness, kvcm is the ORR rate constant per unit volume of cathode CL at the PEM–CL boundary and pOc is oxygen partial pressure at the CL–GDL boundary. kvcm is the greatest rate constant and pOc is the highest partial pressure in the CL. Therefore, kvcmpOc represents the intrinsic ORR rate without transport resistance. MO(C)m and Mp(C)mrespectively represent the ratio of reaction performance to oxygen transport performance and the ratio of reaction to proton transport.

A Peclet number which represents the ratio of convection to diffusion of oxygen is also required for formulation:

PO(C)m =  δ(C)NA(M) /(Cg DeO) [4]

where NA(M)is the net water flux through the PEM, which equals the total gas flux at the PEM–CL boundary.

The dimensionless ORR rate profile is determined by only 4 dimensionless parameters, MO(C)m, Mp(C)m, PO(C)m, and yOc. The dimensionless system greatly reduces the complication of system.

RESULTS AND DISCUSSION

Define the cathode effectiveness factor Fec as a ratio of the ORR rate to the intrinsic ORR rate without transport resistance. Figure 1 shows that increase in transport resistance reduces the effectiveness factor. Typical values of MO(C)m and Mp(C)m in usual cells are around 1. When Mp(C)m is high, proton transport is slow, which increases the cathode emf and reduces the ORR rate. If MO(C)m >> Mp(C)m, oxygen transport is suppressed and ORR takes place near the oxygen inlet, the GDL side.

As the convective flow from PEM to GDL increases, oxygen distribution is shifted towards the GDL, reducing the effectiveness factor as shown in Fig. 2. Even at quite low MO(C)m and Mp(C)m of 0.01, the convection can reduce the effectiveness factor. When MO(C)m and Mp(C)m are high, the impact of convection is not remarkable since the reaction takes place only near the boundaries. When back diffusion of water in the PEM is dominant, PO(C)m can be negative. Fec reaches 1 under reaction control.

Figure 3 shows a case study in which the CL thickness δ(C) is varied with a fixed Pt loading per volume. As δ(C) increases, the Pt amount increases proportionally but the transport resistance increases MO(C)m and Mp(C)m and the effectiveness factor decreases as shown in Fig. 1. As a result, the increase in the current density is limited. Since the increase in δ(C)raises both moduli, it overtakes the increase of Pt; a maximum current appears in Fig. 3.

Employing the proposed dimensionless moduli, the effects of dimensions and operating conditions can be estimated quantitatively without expensive simulation.

Acknowledgement This work is a part of the project by the New Energy and Industrial Technology Development Organization (NEDO), Japan.

REFERENCES

1. M. Kawase et al., 24th Int. Symp. Chem. React Eng. (Minneapolis, June 2016).

2. M. Kawase et al., ECS Trans. 16(2), 563–573 (2008).

Figure 1

2390

, , , and

Abstract: The polymer electrolyte membrane fuel cell (PEMFC) possesses many advantages for both automotive and stationary application, including high energy efficiency, low operating temperature, zero emission, and so on. Aside from the fact that great improvement should be made on its key materials, i.e., the ORR electrocatalysts and proton exchange membrane, it is believed that the PEM fuel cell performance is greatly affected by the operating temperature, gas inlet humidity as well as the flow pattern, and so on [1,2]. Thus, in this work, a 3D steady state model is established to investigate detailedly the effects of the flow pattern (co-flow and counter-flow), the anode and cathode gas inlet relative humidity (RH) on the cell performance. The governing equations of the 3D model result from careful analysis on the electro-chemical reactions, current conservation, membrane proton migration, membrane water transport and water-vapor phase transition. The membrane water transport takes into accounts the electro-osmatic drag and water back diffusion, and the conservation of momentum, species and energy is applied to all components of the PEM fuel cell. Experimental validation is also performed and fits very well with the simulation. Figure 1 shows the current density distribution for different flow patterns: (a), co-flow and (b), counter-flow. The corresponding achievement will offer an efficient guide on the design and performance optimization of PEMFC.

Acknowledgements

This work was supported in part by National Natural Science Foundation of China (Grant No. 21373135 and 21533005) and Science Foundation of Ministry of Education of China ( Grant No. 413064).

References

  • K. Dannenberg, P. Ekdunge, G. Lindbergh, Mathematical model of the PEMFC, Journal of Applied Electrochemistry, 2000, 30: 1377-1387.

  • S.H. Ge, B.L. Yi, A mathematical model for PEMFC in different flow modes, Journal of power sources, 2003, 124: 1-11.

Figure 1

2391

and

Building a catalyst layer includes many elements that make understanding and studying it a complex problem. These systems have different elements with multiscale, multiphase and multicomponent issues, as Figure 1 illustrates. Typical PEMFC catalyst layers are composed of three elements: porous carbon to provide electron and gas/liquid transport, ionomer to provide protonic transport and catalyst to facilitate the reduction or oxidation reactions.

The replacement of Pt with non-precious-metal-based catalysts (NPMCs) for the ORR is considered to be one key to cost-effective power generation in PEMFCs. Accordingly, developing NPMCs with high performance for ORR (1-4) has been a focal point of research. A major complication in analysis of the NPMC systems under study lies in the unknown nature of the active site. It is difficult to identify structural and electronic parameters based on most characterization techniques. Many questions arise when compared with Pt-based systems such as: Which elements do Pt and NPMC catalyst layers have in common? What makes one metal more suitable than another in the NPMC? What specific catalyst layer composition and structure are optimal? How do we (or can we) correlate energetics, structure and transport issues?

Modeling catalyst layers is a complicated task because we need to take into account the phenomena occurring over different time and size scales as represent in Figure 1: (i) heterogeneous electrochemical reactions in the microscale; (ii) proton, electron and gases transport in the mesoscale; and (iii) gases and water transport through membranes and porous media in the macroscale (5-10).

The first steps are to identify and classify the main macroscopic parameters that affect the performance by using complex and adequate experimental data sets and to reduce the degrees of freedom for fitting. Once the parameters are identified, we can work at the right scale, with the appropriate physics and computational methodology that the macroscopic results point out in order to complete the picture and resolve the unknowns.

The aim of this study is to extend previous results from a 1D model in which we identified the active area and permeability as the most distinguishing elements between Pt-based and NPM-based catalyst layers. On one hand, the active area enters into macroscopic models convolutes with the intrinsic activity of the catalyst site. The latter is related to the specific catalytic properties of the active site; understanding it requires a deep study of the catalyst nature, which is not well known due to the pyrolysis of NPM materials. In this case, a coupled modeling approach using DFT (Density Functional Theory) and MD (Molecular Dynamics) techniques could give us insight in the active site behavior. On the other hand, transport properties and differences can be study by using CFD (Computational Fluid Dynamics) tools to compare fluid transport and local distributions within the catalyst layers with parameters obtained via DFT-MD simulations. This contribution will summarize the previous modeling of transport and describe calculations to probe the final configuration after high temperature treatment of porphyrin and iron porphyrin carbon structures. We aspire to describe interactions of the porphyrin with the carbon substrate during the heating process and give insight into the active site final configuration from a theoretical perspective by combining MD with DFT calculations.

Acknowledgement

 

We gratefully acknowledge the support of the NSF EPSCoR program and Colciencias for support of this work.

 

References

 

1. F. Jaouen, E. Proietti, M. Lefevre, R. Chenitz, J.-P. Dodelet, G. Wu, H. T. Chung, C. M. Johnston and P. Zelenay, Energy & Environmental Science, 4, 114 (2011).

2. G. A. Goenaga, J. Brooksbank, C. Dabke, A. Belapure, A. Papandrew, S. Foister and T. A. Zawodzinski, ECS Transactions, 50, 1749 (2013).

3. M. Ferrandon, X. Wang, A. J. Kropf, D. J. Myers, G. Wu, C. M. Johnston and P. Zelenay, Electrochimica Acta, 110, 282 (2013).

4. F. Jaouen, V. Goellner, M. Lefèvre, J. Herranz, E. Proietti and J. P. Dodelet, Electrochimica Acta, 87, 619 (2013).

5. A. Z. Weber and J. Newman, Chemical Reviews, 104, 4679 (2004).

6. T. A. Zawodzinski Jr, T. E. Springer, F. Uribe and S. Gottesfeld, Solid State Ionics, 60, 199 (1993).

7. A. Ohma, T. Mashio, K. Sato, H. Iden, Y. Ono, K. Sakai, K. Akizuki, S. Takaichi and K. Shinohara, Electrochimica Acta, 56, 10832 (2011).

8. T. S. Olson, K. Chapman and P. Atanassov, Journal of Power Sources, 183, 557 (2008).

9. C. Song and J. Zhang, in PEM Fuel Cell Electrocatalysts and Catalyst Layers, J. Zhang Editor, p. 89, Springer London (2008).

10. F. Charreteur, F. Jaouen, S. Ruggeri and J.-P. Dodelet, Electrochimica Acta, 53, 2925 (2008).

 

Figure 1

2392

, , , , and

Depending on the operating conditions, the overall performance and durability of polymer electrolyte membrane fuel cells (PEMFC) can be strongly influenced by liquid water saturation of pores within the gas diffusion layer (GDL). Water agglomerations in the GDL may occur due to material morphology, surface characteristics and local boundary conditions. In the catalyst layer (CL) and microporous layer (MPL) these parameters are considerably different than those of the fibre substrate. The MPL and CL feature by orders of magnitude smaller pores than the substrate and the boundary conditions, e.g. temperature and relative humidity, are different than those in the substrate near the bipolar plate. Thus, different physical behavior within each layer can be observed, e.g. local liquid sorption. Therefore, finding an optimal strategy regarding water management requires knowledge on virtually all the factors involved.

To approach this target, we are using and further developing a 3D Monte Carlo (MC) model which simulates water distribution within the GDL, employing the real GDL structure and operating conditions, exploited respectively from tomography and computational fluid dynamics (CFD) studies1. The 3D images are obtained from an assembled fuel cell and therefore, the results include compression and inhomogeneity information of the GDL. Furthermore, recent progress of the model has enabled simulations on the length scales relevant for the MPL and CL in addition to the fibre substrate.

GDL substrates usually have an average pore size in the micrometer scale, whereas this value for the MPLs and CLs may go down to the nanometer scale. Therefore, a complete simulation of the water distribution in GDL includes working on at least two scales. Although obtaining accurate results on all three sub-layers is equally important, MPL and CL are the more challenging ones from the point of view of both structure and surface properties acquisition. Furthermore, simulating relevant domains within the material requires expensive computational effort. In this study, we report first results on MC simulations of the water distribution within the MPL and CL pores for different operating conditions.

The 3D reconstruction of the material is obtained from the focused ion beam - scanning electron microscope (FIB-SEM) tomography2. By using atomic layer deposition (ALD) of ZnO, a good material contrast between pores and solid particles is achieved and therefore a reliable binarization can be obtained3. The voxel size is set to 45 nm in order to achieve both, an accurate computation with respect to the applied equations and an appropriately sized simulation domain.

Our MC model provides the water distribution within the actual MPL and CL structures based on the local boundary conditions. The results show how the water distribution depends on parameters such as contact angle, porosity, local temperature and relative humidity. An example of simulation results is illustrated in the Figure 1.

References

1. K. Seidenberger et al., J. Power Sources, 239, 628–641 (2013)

2. S. Thiele et al., J. Power Sources, 228, 185–192 (2013)

3. S. Vierrath et al., J. Power Sources, 285, 413–417 (2015)

Figures

Figure 1

2393

and

The work elucidates the effect of micro-morphology parameter. i.e. catalyst distribution, on macroscopic mass transport properties of a catalyst electrode of a PEM fuel cell. Indeed, the micro-morphology parameters have significant effect on transport properties, such as permeability, tortuosity. Pore-scale analysis of fluid flow has been done using Lattice Boltzmann method. Besides other morphological parameters, like porosity, clusterization, geometry those are common for a porous medium, the mass transport properties of a porous electrode of a fuel cell is governed by catalyst particle deposition distribution. This fact enables to control flow field inside of the most vulnerable membrane electrode assembly compartment. Redistribution of catalyst might be required to design novel MEA more robust prone to degradation processes, especially for transport application. In particular, the influence of a catalyst loading on permeability has been studied. It is explored that increasing the catalyst loading definitely improves transport parameter, but this improvement has maximum effect up to certain (ωPt≈0.2) value of catalyst loading. Moreover, how permeability changes depending on a distribution of the same amount catalyst towards membrane has been considered. Pore-scale simulation of different catalyst layer configuration with regards to catalyst deposition shows that redistribution of catalyst in exponential decaying towards membrane allows three times improve permeability of catalyst electrode compared to conventional/homogeneous distribution.

E-02 Alkaline and DF Cells - Oct 2 2016 2:00PM

2394

, , , , , , , and

Over the last decade, research team from Daihatsu Motor Co. has introduced the concept of anion-exchange membrane fuel cell with liquid fuels for automotive applications. Switching from more commonly used proton exchange membranes (PEM) to alkaline, anion-exchange membranes (AEM) has several kinetic and materials benefits and opens the possibility of developing Fuel Cell Vehicle (FCV) entirely based on Platinum Group Metal-free (PGM-free) technology.1 This has been made possible by the adoption of hydrazine hydrate as a liquid fuel, which affords technology deployment while using the existing fuel distribution and dispensing technology. This concept pioneered the use of AEM Fuel Cells for automotive applications worldwide and created the state of the art AEM Fuel Cell stack technology2.

UNM team has been a long-term participant in the multi-institutional program led by Daihatsu Motor Co., focusing on the development of Oxygen Reduction Reaction (ORR) catalysts and selective hydrazine electro-oxidation catalysts used as cathode and anode material in AEM membrane electrode assembly (MEA)3. The theoretical electromotive force of such direct hydrazine fuel cell is 1.56V and hydrazine hydrate as a fuel can be oxidized by number of catalysts using exclusively Earth-abundant metals, such as NiZn unsupported catalysts developed at UNM4. This line of research has been continued in NiLa5 and supported Ni-allowed catalysts6 aiming not only high activity, but also extreme selectivity of hydrazine electro-oxidation to N2 and H2O only. Elucidating the mechanism of hydrazine oxidation included EXAFS/XANES study7 and let to formulating of the hypothesis of the reaction mechanism8 that involves surface hydroxyl groups on Ni as critical participants in the hydrazine dehydrogenation reaction step (see Fig. 1). The role of the nickel, secondary metal phase and the carbon support in the activity and durability of Ni-allow catalysts will be discussed.

ORR catalysts developed under this program included initially Co-polypyrrole materials system9. The mechanism of oxygen reduction and the structure-to-property relationships have been described in a recent EXAFS/XANES study10. This paper will present a family of Fe-Nitrogen-Carbon ORR catalysts that were used in the successful stack development and Daihatsu FCV prototype tests. This new generation Fe-containing PGM-free catalysts are synthesized by UNM Sacrificial Support Method (SSM) a type of for templated synthesis of hierarchically structured electrocatalysts materials.11-13 In this method the catalysts precursors are being absorbed on, impregnated within or mechanically mixed with the support (usually mono-dispersed or meso-structured structured silica), thermally processed (pyrolyzed) and then the silica support is removed by etching (in KOH or HF) to live the open frame structure of a "self-supported" material that consists of the catalysts only. Such hierarchical structures are advantageous in enhancement of the fuel cell performance since they correspond to the different levels of transport in the corrugated electrode matrixes. A wide variety of materials can be made by these methods in which not only the composition but also the microstructure can be varied. It is the combination of these attributes - control over microstructure at a number of different length scales and composition, simultaneously - that is extremely important to the performance of the electrocatalyst materials in a fuel cells.

The high activity of these Fe-N-C cathode catalysts was confirmed in both RDE and MEA tests. As synthesized materials were extensively studied by XPS, SEM, TEM, BET and other methods, in order to elucidate the structure-to-properties correlations. This paper describes a new class of templated, self-supported PGM-free catalysts derived by sacrificial support method (SSM) and their activity in free alkaline electrolyte (KOH) and in anion-exchange MEA.

  • P. Atanassov and A. Serov, Japanese Society of Automotive Engineers, 67 (2013) 68-71

  • J. Varcoe, P. Atanassov, et al., Energy & Environmental Science, 7 (2014) 3135-3191

  • A. Serov, M. Padilla, et al., AngewandteChemie Intern. Ed., 53 (2014) 10336 –10339

  • U. Martinez, K. Asazawa, et al., Physical Chemistry & Chemical Physics, 14 (2012) 5512-5517

  • T. Sakamoto, K. Asazawa, et al., J. Power Sources, 234 (2013) 252-259

  • T. Sakamoto, K. Asazawa, et al., J. Power Sources, 247 (2014) 605-611

  • T. Sakamoto, D. Matsumura, et al., Electrochimica Acta, 163 (2015) 116-122

  • T. Sakamoto, H. Kishi, et al., Electrochimica Acta, (2016) in press

  • T.S. Olson, S. Pylypenko, et al., J. Phys. Chem. C., 114 (2010) 5049–5059

  • K. Asazawa, H. Kishi, et al., Journal of Physical Chemistry - C, 118 (2014) 25480–25486

  • S. Pylypenko et al., Electrochimica Acta, 53 (2008) 7875-7883

  • A. Serov et al., Electrochem. Comm., 22 (2012) 53-56

  • Serov et al., Advanced Energy Materials, 4 (2014) DOI: 10.1002/aenm.201301735

  • Serov, et al., Nano Energy, 16 (2015) 293-300

Figure 1

2395

and

Among the existing fuel-cell types, alkaline-exchange-membrane fuel cells (AEMFC) have intriguing features as compared to proton-exchange-membrane fuel cells (PEMFC). Their major advantage is the possibility of using non-noble catalysts due to faster oxygen-reduction reaction (ORR) kinetics in alkaline than in acidic media. However, water management is a more serious concern in AEMFCs because OH- conductivity is more highly dependent on water content and the ORR consumes water. Compared to PEMFCs, the lower performance of AEMFCs is mostly caused by extremely nonuniform distribution of water in the ionomer phase between the anode and cathode as well as the increased overpotential for the hydrogen oxidation reaction. In this presentation, we will discuss the performance-limiting mechanisms specific to different operating conditions (e.g. varying inlet relative humidity (RH)) based on a cell-level mathematical model. For example, anode flooding can be a critical issue at 100% RH, whereas at lower RH high ionic transport resistance in the cathode dominates due to dehydrated ionomer phase at high current, where the impact of water-consuming ORR kinetic polarization is also critical. Low AEMFC performance at high current is not simply due to mass transport issues with vapor/membrane water, but a consequence of poor water distribution leading to sluggish OH- conduction and ORR kinetics. A sensitivity analysis of design parameters including the humidifying condition and membrane property is performed to identify the most significant factors controlling performance. Overall, water management and ionic/mass transport characteristics of an AEMFC assembly are discussed in detailed. The developed model will be used to examine and elucidate performance bottlenecks and enable strategies to overcome them, significantly increasing the possibility of AEMFC commercialization.

Acknowledgements

This work was funded by the Fuel Cell Technologies Office, Office of Energy Efficiency and Renewable Energy, of the U. S. Department of Energy under contract number DE-AC02-05CH11231, program manager David Peterson.

2396

, , , and

In searching for an exemplary carbon neutral fuel, dimethyl ether (DME) may be the most appealing. This simplest of the ethers can be readily produced from renewably sourced hydrogen and CO2, making it essentially a hydrogen carrier (1). Both nontoxic and easy to liquefy under moderate pressure, DME closely matches diesel and has been run in trucks (2). The ubiquitous propane grill tanks can store DME without modification. Recently, Los Alamos National Laboratory (LANL) demonstrated the potential for direct oxidation of DME in a fuel cell.The breakthrough of our work is LANL's highly active catalyst for direct oxidation of DME that in the early phase of development shows similar performance to direct methanol when using typical low-temperature membranes. However, the output is not sufficient to approach commercial acceptance targets for higher power applications or precious metal cost. High-temperature membrane electrode assemblies (HT-MEAs), based on phosphoric-acid-imbibed membranes, operate at 160 °C to 180 °C without additional water and are highly tolerant to carbon monoxide – an intermediate of DME oxidation. This work exploits a novel ternary LANL anode catalyst with the features of high-temperature operation to produce high-power, low-cost direct DME MEAs.

In support of direct DME oxidation, scientists at LANL have recently shown a direct liquid fuel cell operating with DME as part of a program to develop technology for direct methanol (MeOH), ethanol (EtOH), and DME fuel cells (3). In this work, it has been shown that DME can be directly oxidized via a PtRuPd ternary catalyst that not only activates the ether bonds but facilitates the removal of CO as a reaction by product (Figure 1). Even at this early stage, the team led by LANL has shown results that approach a highly developed DMFC system with regards to performance and precious metal loading (3). One highly relevant aspect of this work is the sharp temperature dependency of the current-voltage behavior, with an increase in the cell temperature from 80 °C to 90 °C leading to an increase in the current density at 0.5 V by ca. 50%, from 140 mA cm-2 to 220 mA cm-2

High-temperature PEM MEAs developed by BASF for the polybenzylimide (PBI) system or pyridine polymer based TPS®system of Advent Technologies, Inc. have demonstrated utility and robustness when operating with highly impure reformates of around 1-3% CO (4). Similarly, cathode sensitivity to air contaminants is also ameliorated. The electrolyte in both systems is phosphoric acid imbibed into a stable and inert polymer matrix. Phosphoric acid is notable in that it does not need water to conduct protons, and can do so over a wide temperature range from ~120 °C to more than 200 °C. Advent is a licensee of BASF's MEA and gas-diffusion-electrode technology.

This presentation will summarize our on-going efforts at facilitating direct oxidation of DME by using high temperature PEM MEAs with both PBI and TPS combined with the LANL catalyst operating at 160 oC to 180 oC.

  • Olah, G.A., Goeppert, A. and Prakash, G.K. Surya in J. Org. Chem., 2009, 74, 487-498.

  • Hutchinson, H., 2013 in ASME web article www.asme.org/engineering-topics/articles/transportation/diesel-alternative-hits-the-road.

  • Advanced Materials and Concepts for Portable Power Fuel Cells, 2014 Hydrogen and Fuel Cells Program AMR presentation.

  • Technical Brochure, 2014 Advent Technologies, Inc. at www.Advent-Energy.com.

Figure 1

2397

, , , , , and

An anion exchange membrane fuel cell (AEMFC) using hydrazine hydrate liquid fuel has been investigated1 and the concept vehicle was exhibited at the 42nd and 43rdTokyo Motor Show.

For the improvement of the AEMFCs, technological breakthrough is necessary for anion exchange membranes. Especially, it is important to overcome the counteracting relationship of high ion conductance and low fuel permeability.

In addition, technology to optimize triple-phase interface formed by catalyst and ionomer is important to improve the performance.

Through many efforts to control the triple-phase interface, state-of-the-art proton exchange membrane fuel cell electrode that consists of platinum group electrocatalysts and perfluorinated ionomer has been investigated. However, the catalyst utilization is still insufficient. In addition, it is more challenging to control the triple phase interface for PGM-free electrocatalysts and anion exchange ionomers.

Our approach to develop the advanced AEMFC is using anion conductive quaternized aromatic multiblock copolymers, poly(arylene ether)s (QPEs)2as shown in Fig.1 and controling the interface and PGM-free electrocatalysts both for anode and cathode.

In this preliminary study, we focus on ionomer layer design for anode and cathode evaluated by ionomer thickness covered on the electrocatalysts. (TEM image shown in Fig.2)

MEA consisting of the QPE membranes and ionomers was fabricated with PGM-free electrocatalysts and subjected to performance evaluation. The MEA performance of direct hydrazine fuel cells will be shown and discussed in the meeting.

Figure captions

Fig. 1. Representative structure of QPEs.3

Fig. 2. TEM Image of ionomer layer (red line) on the cathode catalysts.

References

[1] K. Asazawa, K. Yamada, H. Tanaka, A. Oka, M. Taniguchi, and T. Kobayashi, Angew. Chem. Int. Ed., 46, 8024, 2007.

[2] M. Tanaka, K. Fukasawa, E. Nishino, S. Yamaguchi, K. Yamada, H. Tanaka, B. Bae, K. Miyatake, and M. Watanabe, J. Am. Chem. Soc., 133, 10646–10654, 2011

[3] R. Akiyama, N. Yokota, E. Nishino, K. Asazawa, and K. Miyatake, submitted.

Acknowledgements

This work was supported by CREST, JST.

Figure 1

2398

, , , and

Water management in anion exchange membrane fuel cells (AEMFCs) is more complex than proton exchange membrane fuel cells (PEMFCs). In the PEM system, water is generated at the cathode and otherwise only serves as the membrane proton transport medium. In the hydroxide AEMFC, water is generated at the anode, consumed at the cathode, and is greatly needed for transport of the larger hydroxide ion. The challenge this introduces is the need to provide adequate water to maintain the membrane humidity without flooding the catalyst or gas diffusion layers.1 One method being used to achieve the necessary high membrane water content is utilizing hydrophilic gas diffusion layers without microporous layers.2 Others have used gas diffusion layers with a hydrophobic coated microporous layer while raising the temperature of the humidifiers well above the cell temperature.3 Gas stream dew points above 100% relative humidity and back pressure are common approaches in increasing the membrane water content. However there is instability in the cell performance and conditions that generate high performance levels are not easily repeatable. The problem is the line between proper membrane hydration and flooded catalysts layers is very thin, and non-existent in some cases.

In this study, the influence of the membrane, ionomer and gas diffusion layer as well as the flow rate and dew points of the anode and cathode gases on AEMFC performance were explored. Tokuyama A201 membranes and AS-4 ionomer were investigated alongside quaternary-ammonium-functionalized radiation-grafted ETFE alkaline anion-exchange membranes and ionomers.4 Using a hydrophilic gas diffusion layer without a microporous layer increased membrane hydration, but as expected increased the potential for flooding. Manipulating the dew points led to the counter-intuitive discovery that the cell performs better with the humidity higher at the anode than the cathode, despite water generation and electro-osmotic drag towards that electrode. In fact, removing too much water from the anode caused instability in the cell, while increasing the water at the anode decreased the membrane resistivity. Back diffusion likely plays an important role in membrane hydration and hydroxide transport through the membrane. A very high flow rate (1.0 L/min) also increased cell performance, despite being several orders of magnitude above the stoichiometric need.

Combining all areas of improvement resulted in a very high performing AEMFC with a maximum current density of 2.2 A/cm2 (at 0.1 V) and max power density of 670 mW/cm2 (880 mW/cm2 iR-corrected) with a membrane resistivity of 75 mOhms*cm2 (Figure 1a). Only a minor drop in the current was observed using air at the cathode with the same 1.0 L/min flow rate as oxygen, giving max current density of 1.7 A/cm2 and a max power density of 580 mW/cm2 with a resistivity of 74 mOhms*cm2 (Figure 1b). This near identical behavior confirms that the amount of reactant present supplied by the higher flow rate is not necessary, but the volumetric flow rate is needed for water management. It is likely that the pressure drop along the single pass cell hardware allows the gas to "jump the bar" only at very high flowrates, which results in better water removal and limits cell flooding in the cell.

References:

1. T. D. Myles, A. M. Kiss, K. N. Grew, A. A. Peracchio, G. J. Nelson and W. K. Chiu, J.Electrochem.Soc., 158, 7 (2011).

2. R. B. Kaspar, M. P. Letterio, J. A. Wittkopf, K. Gong, S. Gu and Y. Yan, J.Electrochem.Soc., 162, 6 (2015).

3. M. Mamlouk, J. Horsfall, C. Williams and K. Scott, Int J Hydrogen Energy., 37, 16 (2012).

4. S. D. Poynton, R. C. Slade, T. J. Omasta, W. E. Mustain, R. Escudero-Cid, P. Ocón and J. R. Varcoe, Journal of Materials Chemistry A., 2, 14 (2014).

Figure 1

D-02 Pt-alloy Cathode Catalysts 1 - Oct 2 2016 2:00PM

2399

, , and

Catalyst for polymer electrolyte fuel cell (PEFC) remains a central issue toward a large-scale penetration of technologies especially for automobile applications. Along with the development of synthesis as well as analytical methodologies, a few nano-meter scale nano-particle with manipulated structure or compositions, such as core-shell or skin-layer alloy catalysts, have been intensively developed (1). Because of the complexity of catalytic reactions as well as the porous microstructure of catalyst layer with multi-components, theoretical methods are also intensively applied (2). While majority of theoretical papers adopts slab models under periodic boundary condition for simpler computation, properties of nano-particles such as strain, stability, or electronic structures should be discussed by using a cluster model. In discussing electronic structure of nano-particles, it is reported that small cluster models show discrete electronic structure, which may lead to a molecular properties rather than metallic properties (3). Thus, it is necessary to adopt models with a realistic diameter. While we can find a few papers adopting models over 2 nm (4), we often find ca. 3 nm particles as PEFC electrocatalyst (5). In our preceding study, we have discussed the stability and electronic structures of Pd-Pt alloys with different atomic configurations by using ca. 3 nm model consisting of 711 atoms (6). In this study, we extend our efforts to Pt-based electrocatalyst with skin-layer configuration.

For density functional theory calculations, we used Vienna ab initio simulation package (VASP) (7) with the projector augmented wave method (8). The Perdew-Burke-Ernzerhof under generalized gradient approximations (9) was used as the exchange and correlation functional. The cutoff energy was set to be 400 eV. Pt-Co alloy nano-particle models with solid solution and Pt-skin configurations are prepared. To discuss the stability of nano-particles with different configuration, excess energy (10) are evaluated.

By comparing the excess energies of the Pt-Co models with solid solution and Pt-skin configurations, we found that the Pt-skin configuration is more stable than solid solution configuration at the same Pt-Co composition. To have insight on the activity toward oxygen reduction reaction, we have analyzed interatomic distances to find shorter distances for Pt-skin configuration. Details will be discussed in the presentation.

  • For example, J. Wu, H. Yang, Acc. Chem. Res., 46, 1848 (2013); S. Guo, S. Zhang, S. Sun, Angew. Chem. Int. Ed., 52, 8526 (2013); M. Watanabe, D. A. Tryk, M. Wakisaka, H. Yano, H. Uchida, Electrochim. Acta, 84, 187 (2012).

  • For example, J. R. De Lile, S. Zhou, Electrochim. Acta, 177, 4 (2015); F. Calle-Vallejo, M. T. M. Koper, Electrochim. Acta, 84, 3 (2012); Z. Shi, J. Zhang, Z.-S. Liu, H. Wang, D. P. Wilkinson, Electrochim. Acta, 51, 1905 (2006); J. Greeley, M. Mavrikakis, Nature Mat., 3, 810 (2004).

  • L. Li, A. H. Larsen, N. A. Romero, V. A. Morozov, C. Glinsvad, F. Abild-Pedersen, J. Greeley, K. W. Jacobsen, J. K. Nørskov, J. Phys. Chem. Lett., 4, 222 (2013).

  • W. An, P. Liu, ACS Catal., 5, 6328 (2015); S. H. Noh, B. Han, T. Ohsaka, Nano Res., 8, 3394 (2015).

  • M. Watanabe, H. Yano, D. A. Tryk, H. Uchida, J. Electrochem. Soc., 163, F455 (2016); M. Chiwata, H. Yano, S. Ogawa, M. Watanabe, A. Iiyama, H. Uchida, Electrochem., 84, 133 (2016).

  • T. Ishimoto, M. Koyama, J. Phys. Chem. Lett., 7, 736 (2016).

  • G. Kresse, J. Furthmüller, Phys. Rev. B, 54, 11169 (1996); G. Kresse, J. Hafner, Phys. Rev. B, 47, 558 (1993).

  • G. Kresse, J. Joubert, Phys. Rev. B, 59, 1758 (1999); P. E. Blöchl, Phys. Rev. B, 50, 17953 (1994).

  • J. P. Perdew, K. Burke, M. Ernzerhof, Phys. Rev. Lett., 77, 3865 (1996).

  • R. Ferrando, A. Fortunelli, G. Rossi, Phys. Rev. B, 72, 085449 (2005).

2400

, , , and

Binary and ternary alloys often show improved activity, selectivity or durability as electrocatalysts compared to elemental catalysts. For practical applications, nanoparticle catalysts are required to maximize utilization of precious metals such as platinum. Theoretical simulations based on density functional theory (DFT) have helped rationalize experimental observations and suggested new design directions for experimentalists.Commonly, theoretical studies are, however, limited to small clusters or extended surfaces under periodic boundary conditions. Although DFT simulations of nanoparticles with hundreds of atoms have become possible in recent years,prediction of properties such as morphology, atomic distribution and catalytic activity of nanoalloys remains a challenge.

Here, we report recent theoretical work on the global optimization of atomic distribution within nanoalloy catalysts using genetic algorithms and we predict their activity using density functional theory. We investigate the activity of Pt-Pd-Au binary and ternary alloys for the oxygen reduction reaction (ORR) and the stability of these ternary alloys under ORR conditions. Our findings suggest subsurface Pd, in binary Pt-Pd films, results in improved activity over pure Pt, while addition of small amounts of Au will hinder Pt dissolution and might also enhance activity. We study the ORR activity of 2 nm ternary Pt-Au-M Mackay icosahedral core-shell nano-particles in order to identify core elements, which increase activity and particle stability through strain and ligand effects.

Acknowledgements

Financial support from the European Commission under the FP7 Fuel Cells and Hydrogen Joint Technology Initiative grant agreement FP7-2012-JTI-FCH-325327 for the SMARTcat project is gratefully acknowledged.

References

1. I. E. L. Stephens, A. S. Bondarenko, U. Grønbjerg, J. Rossmeisl and I. Chorkendorff, Energy Environ. Sci., 2012, 5, 6744.

2. L. Li, A. H. Larsen, N. A. Romero, V. A. Morozov, C. Glinsvad, F. Abild-Pedersen, J. Greeley, K. W. Jacobsen and J. K. Nørskov, J. Phys. Chem. Lett., 2013, 4, 222–226.

2401

Fuel cells are expected to be a key next-generation energy source used for vehicles and homes, offering high energy conversion efficiency and minimal pollutant emissions. The high efficiency arises from the fact that fuel cells convert chemical energy directly into electrical energy without the Carnot limitation of thermal engines. However, due to the extreme operating environment, Pt was almost exclusively the only practical catalyst for molecular oxygen electroreduction in PEM fuel cells. Recently, considerable efforts have been made to investigate synergy effects of platinum alloyed with base metals to improve the sluggish kinetics of oxygen reduction reaction of Pt catalyst. The more active Pt-alloy catalysts, in return, may arise the durability issue under the extreme environment of PEM fuel cell operation.

In this presentation, results of a more active yet durable PtTi alloy catalyst will be discussed for oxygen reduction reaction in acidic media, including (i) combinatorial high-throughput discovery of PtTi alloy composition; (ii) synthesis and characterization of nano-scale PtTi particles with controllable size, phase, and individual particle composition; and (iii) electrochemical STM and ICP characterization of coarsening and corrosion of PtTi alloy catalyst in comparison with Pt and PtCo catalysts.

A novel combinatorial high-throughput electrocatalyst discovery workflow and associated tools were developed [1-4], where thin films of alloys can be made through a Multi-sources Physical Vapor Deposition system [5], and the resulted materials can be electrochemically screened by a Multi-channel Rotating Disk Electrode system [6]. PtTi binary system has been screened using this methodology for activity-stability-composition relationship and the best alloy composition was identified.

In addition to the discovery of alloy compositions, engineering the PtTi alloy particles in nanoscale has been a challenge. Several synthesis technologies were studied and developed to achieve capabilities of controlling particle size and particle microcomposition [7]. The results show that by careful engineering the particle size and microcomposition in nanoscale, it is able to achieve the superior electrocatalytic activity predicted by the combinatorial high-throughput discovery.

Electrochemical STM technique, in combination with electrochemical potentiostating, potential cycling, and ICP analysis, was used to monitor the coarsening and base metal corrosion under dynamic fuel cell operation conditions and compared with phenomena observed from Pt and PtCo catalysts [8-10]. The results show that PtTi not only is more active than Pt and much stable than PtCo, but also can mitigate the catalyst coarsening.

Reference

[1] T. He and E. Kreidler, Phys. Chem. Chem. Phys. 10, 3731 (2008). "Combinatorial screening of PtTiMe ternary alloys for oxygen electroreduction".

[2] E. Kreidler and T. He, in Catalysts for Oxygen Electroreduction – Recent Developments and New Directions (Ed. T. He, Transworld Research Network) Chapter 4, p.63 (2009). "Oxygen reduction on Pt alloys: high throughput combinatorial screening of Pt binary alloys".

[3] T. He, E. Kreidler, L. Xiong and E. Ding, J. Power Sources 165, 87 (2007). "Combinatorial screening and nano-synthesis of platinum binary alloys for oxygen electroreduction".

[4] T. He, E. Kreidler, L. Xiong, J. Luo and C.J. Zhong, J. Electrochem. Soc. 153, A1637 (2006). "Alloy electrocatalysts: combinatorial discovery and nano-synthesis".

[5] T. He, E.R. Kreidler and T. Nomura, US 8,944,002 (2015). "High throughput physical vapor deposition system for material combinatorial studies".

[6] T. He, US 7,077,946 (2006). "High throughput multi-channel rotating disk or ring-disk electrode assembly and method".

[7] E. Ding, K.L. More and T. He, J. Power Sources 175, 794 (2008). "Preparation and characterization of carbon-supported PtTi alloy electrocatalysts".

[8] L. Tang, B. Han, K. Persson, C. Friesen, T. He, K. Sieradzki and G. Ceder, J. Am. Chem. Soc. 132, 596 (2010). "Electrochemical stability of nanometer-scale Pt particles in acidic environments".

[9] Q. Xu, E. Kreidler, D.O. Wipf and T. He, J. Electrochem. Soc. 155, B228 (2008). "An in-situ electrochemical STM study of potential-induced coarsening and corrosion of platinum nano crystals".

[10] Q. Xu, T. He and D. Wipf, Langmuir 32, 9098 (2007). "An in-situ electrochemical STM study of the coarsening of platinum islands at double-layer potentials".

2402

, , , , and

In current polymer electrolyte fuel cells (PEMFC), high amounts of platinum catalysts is required due to the sluggish kinetics of the oxygen reduction reaction (ORR). As such it is important to find ways to increase the activity of the ORR catalyst. Previous studies have shown that alloying platinum with yttrium results in a catalyst which is more than six times as active as sole Pt in half cell rotating disk electrode measurements [1].

The model electrodes in this study are prepared by sputtering a well-defined layer of alloy catalyst onto a gas diffusion layer (GDL). In this manner the composition and thickness of the film can be controlled. These electrodes are studied with single cell PEMFC measurements to characterize the materials, similar to previous work done with these types of electrodes [2]. Using these methods the electrochemical surface area (ECSA), ORR activity, and stability are evaluated. The effects of different MEA preparations will also be investigated.

Scanning electron microscopy (SEM) with energy dispersive X-ray (EDX) analysis as well and X-ray photoelectron spectroscopy (XPS) are used to characterize the structure of the alloy catalyst as well as the bulk and surface compositions, i.e. platinum to yttrium ratio. This characterization is performed on both as-prepared samples and electrodes after being used in the PEMFC.

The main focus is to determine the role of yttrium in platinum thin-film platinum electrodes, and if the excellent activity in half cell experiments will also be seen in real PEMFC measurements.

References

[1] J. Greeley, I. E. L. Stephens, A. S. Bondarenko, T. P. Johansson, H. A. Hansen, T. F. Jaramillo, J. Rossmeisl, I. Chorkendorff and J. K. Nørskov , Nature Chemistry, 2009, 1, 552-556

[2] M. Wesselmark, B.Wickman,C.Lagergren,G.Lindbergh, Electrochimica Acta, 2013, 111, 152-159

2403

, , , , , , , , and

The high platinum loadings required to compensate for the slow kinetics of the oxygen reduction reaction (ORR) impede the widespread uptake of polymer electrolyte membrane fuel cells. In order to improve the ORR kinetics and reduce the Pt loading, we can tailor the electronic properties of the Pt surface atoms by means of alloying Pt with other metals. Researchers have intensively studied alloys of Pt with late transition metals such as Ni and Co during the last decades. However, these compounds typically degrade under fuel cell reaction conditions, due to dealloying. In contrast, alloys of Pt and lanthanides present very negative enthalpy of formation [1,2], which should increase their resistance to degradation.

Herein we present eight novel Pt-lanthanide and Pt-alkaline earth ORR electrocatalysts: Pt5La, Pt5Ce, Pt5Sm, Pt5Gd, Pt5Tb, Pt5Dy, Pt5Tm and Pt5Ca [3]. All the materials are highly active, presenting a 3 to 6-fold activity enhancement over Pt. Pt5Tb is the most active polycrystalline Pt-based catalyst reported in the literature. A Pt overlayer with a thickness of few Pt layers is formed onto the bulk alloys by acid leaching [1-3]. Notably, the experimental ORR activity as a function of the bulk lattice parameter and the Pt-Pt distance follows a "volcano" relation [3], with Pt5Tb presenting the highest initial activity while Pt5Gd is the most active after 10 000 cycles stability test between 0.6 and 1.0 V versus the reversible hydrogen electrode. We use the lanthanide contraction to control strain effects and tune the electrocatalytic activity, stability and reactivity of Pt [3].

References

[1] M. Escudero-Escribano, A. Verdaguer-Casadevall, P. Malacrida, U. Grønbjerg, B.P. Knudsen, A.K. Jepsen, J. Rossmeisl, I.E.L. Stephens, I. Chorkendorff, J. Am. Chem. Soc. 2012, 130, 16476.

[2] P. Malacrida, M. Escudero-Escribano, A. Verdaguer-Casadevall, I.E.L. Stephens, I. Chorkendorff, J. Mater. Chem. A2014, 2, 4234.

[3] M. Escudero-Escribano, P. Malacrida, M.H. Hansen, U.G. Vej-Hansen, A. Velázquez-Palenzuela, V. Tripkovic, J. Schiøtz, J. Rossmeisl, I.E.L. Stephens, I. Chorkendorff, Science2016, 352, 73.

F-02 Electrolysis Cells and Systems - Oct 2 2016 2:00PM

2404

and

The conversion of electric energy into pressurized hydrogen by means of water electrolysis is currently discussed as promising option to handle intermittent electricity supply that occurs with the increasing use of renewable energy sources. Subsequently, this hydrogen could be (1) either reconverted into electricity, or (2) converted into other energetic molecules, such as methane (power to gas) or liquid fuels, or (3) utilized as a feed stock for the chemical industry.

Besides conventional alkaline water electrolysis, a water electrolysis technology based on a Polymer Electrolyte Membrane (PEM) is discussed since several years, which is the focus of the present contribution. The particular advantages of PEM water electrolysis are higher attainable current densities, higher voltage efficiency, and higher load flexibility. Current disadvantages include far lower technical maturity, shorter life time, and higher specific costs compared to alkaline water electrolysis.

The formulation of appropriate physico-chemical models for solid polymer electrolysis can help to contribute to the solution of the above mentioned problems. In the present contribution state of the art of modelling of solid polymer electrolyzers is summarized and discussed.

2405

, , , and

Proton exchange membrane (PEM) water electrolysis has become increasingly attractive due to the penetration of renewable energy (e.g, solar and wind). Hydrogen production from PEM water electrolysis is advantageous over other technologies due to its simple and clean nature. Membrane and electrode assemblies (MEAs) of PEM electrolyzers typically use iridium (Ir) as an anode catalyst and Pt as a cathode catalyst.

Performance and durability of the MEAs play an essential role for the cost and viable commercialization of PEM water electrolysis. However, unlike the well-established MEA benchmarks of PEM fuel cells, the performance and durability of PEM electrolyzer MEAs have not been thoroughly studied. The objective of this work is to establish benchmark MEA performance and durability for PEM water electrolysis. For this purpose, a series of oxygen evolution reaction (OER) catalysts, which includes commercial Ir black and various Ir nanostructures, has been evaluated under test protocols established at Giner Inc. These approaches include high-voltage hold (>1.8 V), accelerated stress test (e.g., voltage cycling from 1.4 to 2.0 V), and constant low-current operations. The polarization curves of the MEAs will be obtained after each test. The morphology and structure of MEAs after durability tests will be characterized to correlate to their performance and durability.

The established performance and durability may provide metrics and guidance to the community of PEM water electrolysis.

Acknowledgement: The financial support is from the Department of Energy under the Contract Grant DE-SC0007471.

2406

, , , and

Water electrolyzers are predominantly operated at elevated pressures. This is especially favorable in the context of the subsequent utilization of the produced hydrogen e.g. due to the more efficient, electrochemical and isothermal hydrogen compression or less afford for hydrogen drying. However, higher pressures lead to higher cross-permeation of the product gases in electrolysis cells.

In the present contribution, hydrogen permeation across a fumea EF-40 membrane is systematically investigated for pressure differences between 5-35 bar and for four different temperatures between 22 °C and 60 °C. The measured permeation fluxes show a quadratic dependency on the pressure difference. A permeation model combining a diffusive and convective transport can describe the experimental data quantitatively. For the investigated membrane, the diffusive as well as the proposed convective permeability coefficient and their temperature dependencies are obtained.

The diffusive permeability coefficient and its temperature dependency agrees very well with published data for Nafion membranes. The slightly lower value for the EF-40 can be explained by the lower water volume fraction. The obtained convective permeability coefficient indicates a high hydraulic permeability, especially in comparison to recently reported values for Nafion membranes. However, the reported values of the hydraulic permeability show high variations of about two orders of magnitude, with strong dependence on investigated materials, that show the importance of this parameter with regard to convective hydrogen permeation.

The present work discusses two potential reasons for the high obtained hydraulic permeability. The first one is related to the membrane properties. In particular, the lower equivalent weight of the fumea EF-40 membrane could cause the higher hydraulic permeability. Second the electrolysis conditions, in particular the applied small current, might increase the hydraulic permeability due to wider membrane water channels and a change to a hydrophilic membrane surface.

Figure: Parameterized equation in comparison to the experimental results.

Figure 1

2407

, and

To store or transport hydrogen (and optionally oxygen) in high pressure tanks or pipelines pressures up to 1000 bar are needed. Pressurized polymer electrolyte electrolyzers are typically operated at around 30 bar to reduce downstream mechanical compression and gas drying stages. Interestingly, often the cell voltage does not show the expected thermodynamic voltage increase [1,2]. Consequently beneficial influences are compensating the compression work, at the expense of higher faradaic losses due to increased gas crossover.

In this work we analyze cell voltages at gas pressures up to 100 bar as function of temperature up to 70 °C and current densities up to 4 A/cm2, considering the three main overpotentials (ohmic, kinetic, mass transport) using a zero-order analysis based on the Tafel kinetics model.

As expected, the ohmic overpotential is mainly a function of temperature and almost independent from pressure. Therefore the overpotential gains with increasing pressure, observed for current densities above about 0.8 A/cm2, stem from the kinetics and/or the mass transport. Both overpotentials decrease with increasing pressure, but over a wide current density region kinetics contribute about two thirds to the voltage gain. The improved kinetics can be attributed to an increased apparent exchange current density, whereas the increased mass transport might be related to smaller gas volumes or velocities resulting in an improved two-phase flow in the porous structures.

Literature:

[1] S.A. Grigoriev, V.I. Porembsky, V.N. Fateev, Pure hydrogen production by PEM electrolysis for hydrogen energy, Int J of Hydrogen Energy, 31 (2006) 171-175.

[2] P. Millet, R. Ngameni, S.A. Grigoriev, N. Mbemba, F. Brisset, A. Ranjbari, C. Etiévant, PEM water electrolyzers: From electrocatalysis to stack development, Int J of Hydrogen Energy, 35 (2010) 5043-5052.

2408

, , , and

This conference contribution touches upon electrochemical characterization of operating polymer electrolyte membrane electrolysis cells (PEMECs) by the application of electrochemical impedance spectroscopy (EIS). Analysis of differences in impedance spectra (ADIS) (Jensen et al., 2007) can be applied to gain insight into the relative magnitudes of resistance contributions from the two electrodes, the electrolyte and of mass transfer limitations and can help identifying the time scale of the respective processes. The gained knowledge may facilitate further development of the PEMECs.

A state-of-the-art PEMEC obtained from IRD with a 2.9 cm2 active electrode area and loaded with 0.3 mg/cm2 IrO2 on the anode and 0.5 mg/cm2 Pt on the cathode was operated at current densities ranging from 0.07 to 1 A/cm2 while examined with EIS. The cell temperature at operation was 61 ˚C at all current densities, and the cell was conditioned for 20 minutes at each current step prior to EIS measurements. A Solartron 1260 was used for the data acquisition, the frequency range was 100 kHz - 0.01 Hz, the amplitude was 17.3 A/cm2 and 12 points were measured per decade. Nyquist plots of the EIS results are shown in the upper graph in the figure. One arc is evident in the spectrum at 0.07 A/cm2, whereas the spectra at 0.35 and 0.69 A/cm2 include two arcs, and two overlapping arcs are observed at 1 A/cm2. ADIS plots, with the impedance measured at 0.07 A/cm2 subtracted, are shown in the lower graph in the figure. Based on the ADIS plots it can be concluded that the impedance is independent of the current density at frequencies above 100 Hz, whereas two processes observed at approximately 0.4 and 10 Hz are current density dependent, the one increasing and the other decreasing in size with increasing current density.

Acknowledgement

The work is part of the research project e-STORE funded by the Innovation Fund Denmark.

Reference

S. H. Jensen, A. Hauch, P. V. Hendriksen, M. Mogensen, N. Bonanos, T. Jacobsen. (2007). A Method to Separate Process Contributions in Impedance Spectra by Variation of Test Conditions. Journal of The Electrochemical Society, 154 (12), B1325-B1330.

Figure

Nyquist plots (upper graph) and ADIS plots (lower graph) of EIS measured on a PEMEC operating at 0.07, 0.35, 0.69 and 1 A/cm2. The ADIS plots represent the difference between the differentiated real part of the impedance at each current density and the differentiated real part of the impedance collected at 0.07 A/cm2.

Figure 1

I01 Plenary Session I - Oct 3 2016 8:15AM

2409

Honda launched new fuel cell vehicle "CLARITY FUEL CELL" in March 10, 2016. In this presentation, I explain the fuel cell vehicle (FCV) development status of Honda including the various main technologies of "CLARITY FUEL CELL". Honda does not only develop the FCV but also small hydrogen station and mobile inverter for power output from FCV. Honda has efforts toward the hydrogen society with concept of "Produce" "use" and "Connect". I explain the small hydrogen station named Smart Hydrogen Stration (SHS) and mobile inverter named "Power Exportor 9000". This explanation will be explained technical conceopts and actual demonstration test resylts. In terms of FCV development, I will explain the FCV development history and future technology direction. Of courese, FCV is not expanded without prepareation of hydrogen infrastructure (Hydrogen Refueling Station : HRS). So this presentation will be explained the situration of HRS preparation and recent direction of movement. Especially, I will introduce the most aggressive Japanese situation. And I expain the future strategy of Honda.

2410

, , , , and

Renewable hydrogen is becoming an increasingly important component of the transition away from fossil fuel use and towards reduction in carbon dioxide production. Hydrogen is the intermediary between primary energy sources and end products in many chemical processes such as ammonia generation, refining, and biogas processing, and is currently mainly produced by reforming of natural gas. Hydrogen from electrolysis can both make a strong environmental impact on these industries and also improve utilization of intermittent renewable energy sources such as wind and solar by leveraging otherwise stranded resources. Hydrogen can also enable improved integration of the electrical grid with the transportation sector and industrial processes, by serving as a storage medium until needed (Figure 1). Because of hydrogen's end use flexibility, electrolysis is a key component of Europe's strategy for energy storage and management of these renewable energy sources on the electrical grid. Proton exchange membrane (PEM) electrolysis is especially well suited to energy capture because of the dynamic range and ability to quickly ramp up and down from near zero output to full capacity.

PEM electrolysis was largely developed in the 70's by GE for life support applications where reliability and reproducibility were key. As the technology was translated to commercial markets it started to compete with the incumbent alkaline water electrolysis on cost, safety (no mixing of O2 and H2), ease of maintenance (no caustic) and operational flexibility (differential pressure operation, idling). While the resulting commercial success of PEM electrolysis suggests that the current state of the technology is sufficiently mature, the truth is that while PEM electrolyzers have demonstrated > 50,000 hrs lifetimes and > 99% reliability in the field, the technology and manufacturing techniques have not changed significantly from those early days. In traditional industrial applications the customer rarely cares about efficiency as much as reliability. At the same time, the robustness and manufacturability of the technology (cell, stack and system engineering), have allowed the systems to be scaled from laboratory scale (0.01kg/day) to megawatt scale (100kg/day) with no loss in performance or reliability. To date the research community has mostly focused on advancing PEM fuel cell technology, producing staggering PGM reduction, improvement in efficiency and translation to high throughput manufacturing. Thus electrolysis technology development has lagged behind PEM fuel cells, providing significant opportunities for continued improvements. Thinner membranes, reduced PGM usage, and improved manufacturing techniques have all demonstrated viability for electrolysis, and have the potential to make higher cost and performance impact.

As the application areas for PEM electrolysis start to shift from traditional markets to fueling and renewable energy capture, efficiency, capital cost and ethical resource use become significant factors. Thus advancing the state of the art, translating laboratory advances to commercial PEM stacks and then manufacturing them at scale, while maintaining reliability becomes paramount. Questions that arise are: How do you test reliability, without testing for 50,000 hrs? How do you pick research "winners" at the rotating disk electrode (RDE) level and develop them to large scale PEM membrane electrode assemblies (MEAs). Where is the gap in research at the laboratory scale based on commercial failures? This talk will discuss the challenges in continued scale up, translating laboratory scale findings to commercial systems as well as some of recent advancements and impact on cost.

Figure 1

2411

Transition metal oxides (TMOs) are attracting an increasing attention as promising materials for oxygen electrode in polymer electrolyte fuel cells (PEMFCs) and electrolyzers (PEMEC). Despite significant number of publications oxygen electrocatalysis over TMOs is still insufficiently understood. Pending questions relate to the reaction mechanisms, the state of the surface during the oxygen reduction (ORR) and oxygen evolution reaction (OER), structure-activity relationships, and degradation phenomena which are particularly detrimental during the OER. This slows down development of potent and durable materials for the oxygen electrode of PEMFCs and PEMECs.

This presentation will consist of two parts. The first deals with the ORR on noble-metal free single and complex Mn, Co and other TMOs in view of their application for PEMFCs based on anion-exchange membranes. In this work we seek to understand the relations between the structure and the composition of TMOs oxides on the one hand, and the kinetics and the mechanism of the electrocatalytic ORR on the other hand. An experimental rotating ring disc electrode investigation of the ORR is complemented with the rotating disc electrode study of the oxidation/reduction of the stable ORR intermediate, hydrogen peroxide, and combined with microkinetic modeling in order to arrive at a self-consistent model of the ORR on TMOs oxides. For Mn oxides we conclude that the potential of the Mn(III/IV) red-ox transition of surface cations can be used as a descriptor of the catalytic activity of Mn oxides in the ORR. We further note that the reaction mechanism and the selectivity for the 4e ORR depend on the structure and composition of Mn oxides. Finally, we propose a tentative explanation for the discovered relationship between the catalytic activity and the crystalline structure.

The second part of the presentation is related to the OER on mixed IrxRu1-xO2 oxide in acid medium and on noble-metal free TMOs in alkaline media. We apply synchrotron-radiation-based near-ambient pressure X-ray photoelectron spectroscopy (NAP-XPS) for studying OER on IrxRu1-xO2 anodes of PEM electrolyzers with the objective to shed light on the long time known but still insufficiently understood stabilization effect of Ir. We conclude that various forms of Ru coexist in the anode during the OER, and their relative contributions strongly depend on the electrode potential and on the presence of Ir. Based on the in situ spectroscopic data we propose a tentative mechanism of the OER on mixed Ir(Ru) anodes and the stabilization effect of Ir. Finally, we will discuss stability of various oxide materials in alkaline and in acid media during the OER.

Acknowledgements The author is indebted to Anna S. Ryabova, Ivan S. Filimonenkov, Galina A. Tsirlina, Sergey Y. Istomin, Evgeny V. Antipov, Filipp S. Napolskiy, Artem M. Abakumov, Tiphaine Poux, Antoine Bonnefont, Gwenaelle Kerangueven, Spyridon Zafeiratos, Viktoriia A. Saveleva, Li Wang, Aldo S. Gago, K. Andreas Friedrich, M. Haevecker and A. Knop-Gericke for their contribution to the work. Financial support received in the framework of ERA.NET.RUS.PLUS (project ID # 270 – NANO-MORF) and European Union's Seventh Framework Programme (FP7/2007-2013) for Fuel Cell and Hydrogen Joint Technology Initiative under Grant No. 621237 (INSIDE) is gratefully appreciated.

2412

Me-N-C catalysts are today's most active non-precious metal catalysts for the oxygen reduction reaction (ORR) in proton exchange membrane fuel cells (PEMFC). The application of these catalysts dates back to 1964 when Jasinski was the first demonstrating ORR activity of different phthalocyanines in alkaline electrolyte. The most important steps in improvement of ORR activity and stability were made by the heat-treatment of MeN4-macrocycles in 1976 and the pyrolysis of independent metal, nitrogen and carbon sources in 1989 as shown by Jahnke et al. respectively Gupta et al..

However, especially during the last decade important progress in the development and fundamental understanding of these catalysts was gained. Based on these findings the ORR activity was boosted significantly. This motivated researchers worldwide to contribute to the development of Me-N-C catalysts and to the controversy of the origin of ORR activity.

In this contribution the most important milestones in the development of Me-N-C catalysts will be summarized. Based on this rational criteria can be defined for the further optimization of these catalysts with respect to a final implementation in the car.

2413

, , , , and

Polymer electrolyte membrane (PEM) fuel cell performance and materials degradation, particularly associated with the cathode catalyst layer (CCL), can be directly attributed to the structure and chemistry of individual material components, as well as their uniformity/homogeneity within a CCL. The individual material constituents used to form the CCL within the membrane electrode assemblies (MEAs), which include the electrocatalyst, catalyst support, and ionomer films, and especially the critical interfaces that are formed between these various constituents, are critically importance in controlling fuel cell perfomance. Understanding the specific microstructural characteristics of the individual materials within the CCL, and how the materials interact to "form" the CCL, is important for identifying materials optimization parameters that can significantly enhance performance and durability. Materials in several states/conditions, e.g., prior to incorporation in the CCL (as-synthesized), after MEA preparation (CCL), and after fuel cell testing, are beig evaluated and quantified using a combination of advanced electron microscopy methods, which are used to interrogate the materials constituents and interfacial structures and chemistries from the μm- to the Å-level. The as-processed (prior to and following incorporation into a CCL) microstructural evidence is directly correlated with observations of materials-specific degradation mechanisms that contribute to fuel cell performance loss, and are then used to identify potential processing variables (materials-based mitigation strategies) to improve the microstructure and compositional homogeneity within the electrode structure, and enhance MEA durability and stability during fuel cell operation.

Research efforts at Oak Ridge National Laboratory are focused on the high-resolution microstructural and microchemical characterization of MEAs fabricated using different electocatalysts (typically Pt-based) and catalyst loadings, carbon-based support materials, and ionomer solutions, as well as the same MEAs subjected to accelerated stress tests (ASTs) designed to degrade specific MEA components. While a significant microscopy effort has been aimed towards understanding catalyst degradation (e.g., coarsening, de-alloying), recently, high-resolution analytical microscopy methods have been used to directly image/map the distribution and chemistry of the ionomer films/layers within CCLs, results of which are being combined with high-resolution imaging and 3-D tomography data of powder materials and MEAs, to provide unprecedented insight into the MEA architecture and interfaces (ionomer/support, ionomer/catalyst, catalyst/support, ionomer/pore). This presentation will focus on understanding materials distributions within CCLs as a function of materials used and ink/MEA processing variables, e.g., initial ionomer and/or ink chemistry, electrocatalyst (type, content, and dispersion), and the type of carbon support used. Additionally, the stability of the ionomer films, electrocatalysts, and support structures in CCLs following ASTs designed specifically for either catalyst degradation or carbon corrosion, will be evaluated.

 ________________________________________________________________________

Research sponsored by (1) the Fuel Cell Technologies Office, Office of Energy Efficiency and Renewable Energy, U.S. Department of Energy and (2) Oak Ridge National Laboratory's Center for Nanophase Materials Sciences (CNMS), which is a U.S. Department of Energy, Office of Science User Facility.

I01 Plenary Session II - Oct 3 2016 1:20PM

2414

, , , , , and

Widespread of fuel cells are strongly desired from the point of reducing the GHG and for clean environment. In order to achieve these, reducing the cost of the fuel cell vehicle is needed and from this point, higher operating temperatures of PEFC can contribute for designing compact and simple cooling system. We have been developing variety of new Perfluorosulfonic acid (PFSA) ionomers having high ion exchange capacity (IEC) and other properties required for each the membrane and the electrolyte in the electrodes. The ionomer with high proton conductivity was tested by copolymerization of tetrafluoroethylene with a new bifunctional fluorosulfonyl monomer 1. The ion exchange capacity was 1.58 meq/g and the water uptake after immersion in hot water at 80 °C for 16 hours was much lower compared to that of conventional ionomer.

Reduction in Pt loading in the cathode is another requirement for PEFC to reduce the cost. Especially at higher temperature and lower RH conditions under which the flux of oxygen to the Pt surface through the ionomer in the electrode could be the rate limiting factor for the oxygen reduction reaction (ORR). In order to increase the oxygen flux to Pt surface, the use of ionomers having higher oxygen permeability can be a solution. We have designed and synthesized various ionomers based on a general rule that the lower the density of a fluorinated polymer, the higher the oxygen permeability. One of the newly prepared ionomer having lower density showed more than 2 times higher the oxygen permeability 2 at low RH conditions. The use of this high oxygen permeable ionomer at cathode showed the possibility of achieving the same cell voltage with half the amount of Pt loading.

We see that the ionomer plays an important row in developing PEFC achieving the requirements from the market. The ionomer can also be used in the micro porous layers (MPLs) between the catalyst layers and the Gas diffusion substrate to create hydrophilic MPLs. Conventionally, hydrophobic MPLs were applied between the GDM and the catalyst layer to prevent flooding due to excess water in the electrode. But the hydrophobicity of the MPL is not favorable when the cell is operating under a dry condition. We have been reporting the use of hydrophilic MPL between the GDM and the catalyst layer 3. It can control the water balance in the MEA, achieving higher performance under low RH conditions without facing the low performance due to mass transfer issue at wet conditions. This hydrophilic MPL was used in the cathode in combination with a high proton conductive membrane, and the MEA showed excellent performance under a very dry condition.

We are continuously making efforts developing PEFC materials getting into the next stage.

References

1. A.Watakabe, H Yamamoto, M Iwaya, S Saito, K Yamada, S Hommura and T Shimohira, Fluoropolymer (2012)

2. K. Yamada, S. Hommura, and T. Shimohira, ECS Trans., 50 (2), 1495 (2013).

3. T. Tanuma and S. Kinoshita, J. Electrochem. Soc., 161 (1), F94 (2014).

2415

Background and objective

One of the important challenges in the current fuel cell research is to develop an alternative membrane to state-of-the-art perfluorinated sulfonic acid (PFSA) ionomers. While the PFSA ionomer membranes are highly proton conductive and chemically and physically stable, high gas permeability, high cost, and environmental incompatibility of the fluorinated materials are drawbacks for the wide-spread dissemination of fuel cells. In the past couple of decades, a variety of proton conductive materials have been proposed as alternative membranes. Nonfluorinated hydrocarbon polymers, acid-doped polymers, inorganic/organic nanohybrids, solid acids with superprotonic phase transition, and acid/base ionic liquids fall into this category. Some of them are claimed to exhibit high proton conductivity, very low gas permeability, and reasonable stability. However, none of them can compete with PFSAs under a wide range of fuel cell operating conditions. The most critical issues of these alternative membranes are insufficient durability and significant dependence of the proton conductivity upon humidity. In addition, interfaces between these emerging proton conductive materials and the catalyst layers have not been well explored. In this communication, current issues (in particular, chemical and mechanical stabilities and interfacial problems) and future possibilities of aromatic ionomer membranes will be discussed.

Stability issues

In pursuit of better performing and more stable aromatic ionomer membranes, we have investigated the effect of a simple structure, sulfo-1,4-phenylene unit, as a hydrophilic component on the membrane properties. Unlike typical aromatic ionomer membranes, the newly designed ionomer membranes exhibited reduced water uptake and excellent mechanical stability under humidified conditions due to the absence of polar groups such as ether, ketone, or sulfone groups in the hydrophilic component. The membrane has survived several tens of thousands cycles in humidity cycling test of US DOE protocol. The high local ion concentration in the hydrophilic segments contributed to an increase in the proton conductivity especially at low humidity. Consequently, the high ion exchange capacity (IEC = 2.67 meq/g) membrane exhibited high proton conductivity under low humidity (20% RH, 7.3 mS/cm) conditions at 80 °C, which are one of the highest values reported thus far for aromatic ionomer membrane.1)

Interfacial issues

In the literature, a number of hydrocarbon ionomer membranes that exhibit comparable proton conductivity to PFSA membranes have been reported, however, they underperformed in operating fuel cells compared to PFSA membranes in most cases. This is partly because of the cathode performance was lower even when the same cathode catalyst layers were used. We have revealed that the double ionomer membrane composed of thin layer Nafion attached on our aromatic ionomer membrane showed higher fuel cell performance than the parent (single layer) ionomer membrane.2,3) The Nafion inter-layer improved the interfacial contact between the aromatic ionomer membrane and the catalyst layer. Based on the results, we have then designed and synthesized novel ionomer membranes composed of perfluoroalkyl and the sulfonated phenylene groups. The resulting membranes could support good electrocatalytic performance of the catalyst layer at the membrane-electrode interface due to the well-controlled finely phase-separated morphology.

Acknowledgement

This work was partly supported by the New Energy and Industrial Technology Development Organization (NEDO) through the SPer-FC Project.

References

1) J. Miyake, T. Mochizuki, K. Miyatake, ACS Macro Lett., 4, 750-754 (2015).

2) T. Mochizuki, K. Kakinuma, M. Uchida, S. Deki, M. Watanabe, K. Miyatake, ChemSusChem, 7, 729-733 (2014).

3) T. Mochizuki, M. Uchida, H. Uchida, M. Watanabe, K. Miyatake, ACS Appl. Mater. Interfaces, 6, 13894-13899 (2014).

2416

, , , , and

Proton exchange membrane (PEM) fuel cells are being developed as alternative energy sources for both residential and automotive application. In order for this technology to become fully commercial, the reduction of cost and improvements in performance and durability of PEM fuel cells membrane electrode assemblies (MEAs) are still required. To address the requirement for further cost reduction the Pt loading of the cathode catalyst layer (CCL) needs to be reduced to 0.2- 0.1 mg/cm2 while maintaining high efficiency of Pt-utilization at high power densities. Consequently, it becomes increasingly important to not only develop new catalyst materials, but also optimize the 3D structural arrangement of the CCL components such as catalyst, ionomer and void space so that all critical functionalities can be achieved simultaneously and optimize PGM utilization. In general terms this entails to provide sites catalytically active for ORR, and to further provide transport to/these sites for the reactants O2, protons, electrons and products H2O and heat, respectively. In order to be able to design CCL structures that meet the performance and durability requirements, it is necessary to obtain a better understanding of structure versus performance relationships. This requires the capability to fabricate different CCL structures, to characterize the spatial distribution of all components within the catalyst layer1,2 (carbon, Pt, ionomer and void), to measure the physico-chemical properties (both ex-situ and in-situ) and finally to use these experimental data as inputs for the development a model based understanding of the relationship between CCL structure and CCL performance and durability3.

 

References

  • A. P. Hitchcock, V. Berejnov, V. Lee, D. Susac and J. Stumper, "In situ methods for analysis of polymer electrolyte membrane fuel cell materials by soft X-ray scanning transmission x-ray microscopy", Microscopy & Microanalysis 20(S3) 1532 (2014)

  • V. Lee, V. Berejnov, M. West, S. Kundu, D. Susac, J. Stumper, R.T. Atanasoski, M. Debe and A.P. Hitchcock, "Scanning Transmission X-ray Microscopy of Nano Structured Thin Film Catalysts for Proton-Exchange-Membrane Fuel Cells" J. Power Sources 263 163 (2014)

  • S. G. Rinaldo, W. Lee, J Stumper, M. Eikerling: "Nonmonotonic dynamics in Lifshitz-Slyozov-Wagner theory: Ostwald ripening in nanoparticle catalysts" Physical Review E 86, 041601 (2012)

2417

, , , and

One key challenge to enable a wide spread adoption of proton-exchange membrane fuel cell vehicles is the reduction of Pt use to a level comparable to the incumbent internal combustion engine vehicles (<5 gPM/vehicle). Great progress in recent years in improving the activity and durability of Pt-based catalysts, particularly Pt alloy catalysts, has enabled a reduction of Pt loading in the cathode approaching 0.15 mgPt/cm2 or about 15 gPt/vehicle. However, as we attempt to go below 10-15 g/vehicle, a large performance loss is often observed at high-current density (high power).

In the cathode, where oxygen molecules meet protons and electrons to produce water, one must create as much interface as possible in order to minimize the transport resistances. As Pt loadings are reduced, performance losses due to these transport phenomena become more localized to the Pt and ionomer interface, posing new challenges to diagnose and understand these electrodes.

In this talk, we will discuss how we use in-situ electrochemical diagnostics and modeling to understand the performance of low-Pt electrodes, identify their key limiting factors, and guide our catalyst layer development. Diagnostics used to quantify the kinetic terms, proton conduction, oxygen transport resistance, and Pt-ionomer interaction will be introduced. Moreover, because non-noble metals leach out from Pt alloy catalysts and can replace protons in the ionomer, we will also consider the resulting impact on low-Pt electrode performance.

This work was partially supported by the U.S. Department of Energy, Office of Energy Efficiency and Renewable Energy under grant DE-EE0007271.

F-21 Oxygen Evolution Reaction - Oct 4 2016 7:40AM

2418

and

The United Nations Framework Convention on Climate Change (COP21) in Paris last year has underlined the urgency of a fast and consequent transformation of the various energy systems from a fossil based energy feedstock towards renewable energies like photovoltaics, wind- and hydropower, in order to drastically decrease the greenhouse gas emissions (GHG). Such a transformation requires a thorough restructuring of the energy systems with all energy carriers and consumer sectors including its intersectoral dependencies, grid structures and energy storage needs.

At Fraunhofer ISE we performed time-resolved studies on how to reach the main goal of the German energy transformation to decrease its GHG emissions by at least 80% until 2050 and thus, to widely decarbonize all energy related sectors. Potential transformation pathways based on various scenarios were studied and the required costs were modelled for each scenario. The results of this modelling made very obvious that the ambitious GHG reduction targets can only be fulfilled by a strong rise in electricity generation by fluctuating renewable energies and on the other hand by the installation of large plants for producing synthetic energy carriers from renewable energies.

The most versatile energy carrier for such large amounts of energy is hydrogen produced from water electrolysis. Hydrogen can be stored in large amounts in gas tube fields or salt caverns and can be used directly either in fuel cell cars and buses as a fuel or in stationary combined heat and power fuel cell systems. Furthermore the hydrogen can be used in order to hydrogenate CO2 and thus to generate green liquid fuels such as oxymehylenether (OME) or other chemicals in order to replace oil based products. The conversion of electricity into hydrogen is called Power-to-Gas, the further processing of the hydrogen and CO2 towards liquid energy carriers and chemicals is called Power-to-Liquid.

With the advent of various Power-to-Gas demonstration projects it becomes clear how important this technology will be in the future energy system. Distributed water electrolyzers in the Megawatt power range will stabilize the grid frequency and voltage or will deliver other ancillary services and thus, will contribute to the secondary balancing market.

Alkaline water electrolyzers are applied for more than 100 years in industrial applications and are nowadays complemented by proton exchange membrane (PEM) electrolysers in order to cover the needs for a highly dynamic load in the grid. However, main challenges of the PEM technology are high costs associated with expensive materials and the durability of cells and stacks. With ongoing technological development it is expected that PEM electrolysis systems become a competitive alternative to alkaline systems in the next years.

In this talk, the technological base of water electrolyzers will be described and the related performance data, life-time expectation and cost reduction potential will be discussed. Furthermore the impact of distributed water electrolyzers in the Megawatt power range as part of an increasingly renewable energy system in terms of Power-to-Gas modelling scenarios will be given.

Figure 1

2419

and

 PEM water electrolysis is one of the promising technologies for hydrogen production from renewable energy sources. Iridium and their oxides are recognized as the most suitable catalyst showing high activity and stability for oxygen evolution reaction (OER) in acidic media [1], however, overpotential is still high and its scarcity and price are also drawbacks. To overcome these problems, a number of studies concerning particle size reduction, catalyst compositions, and utilization of catalyst support have been reported [2]. Among them, use of corrosion-resistant and highly electro-conductive oxides, especially Magneli phase Ti4O7 as catalyst support would be beneficial for reducing Ir loading and increasing durability. We have developed Ti4O7-supported Ir catalyst and examined activity and durability as an anode catalyst of PEM electrolyzer. 

Titanium oxide support was prepared by UV laser technique [3]. Ir/Ti4O7 catalysts obtained by using H2IrCl6 precursor showed conductivity comparable with that of commercial Pt/C. MEAs with the Ir/Ti4O7 anode can be operated up to 4 A cm-2 at low cell voltage < 2.0 V (with DuPont NR212 membrane) with very low Ir loadings (0.25 mg-Ir cm-2). Considering fluctuating nature of the renewable energy inputs, catalyst stability against voltage change should be another important issue. To simulate and accelerate such degradation modes, potential sweep between 1.0 V and 2.0 V at a sweep rate of 500 mV s-1 was applied up to 10,000 cycles. The Ir/Ti4O7 MEA shows limited degradation even after 10,000 cycles. Stability of the cell performance and the catalyst structure will be discussed at the Meeting.

References

1. M. H. Miles and M. A. Thomason, J. Electrochem. Soc.,123, 1459 (1976).

2. M. Carmo, D. L. Fritz, J. Mergel, D. Stolten, Int. J. Hydrogen Energy, 38, 4901 (2013).

3. T. Ioroi, H. Kageyama, T. Akita, K. Yasuda, Phys. Chem. Chem. Phys., 12, 7529 (2010).

2420

, and

1. Introduction

Renewable energy is an alternative choice to solve the universal problem of energy depletion and environmental pollution. Among various energy carriers, hydrogen is a most competitive one, because it is friendly to environment and can be produced from renewable energy through water electrolysis. Among types of water electrolysis, polymer electrolyte membrane water electrolysis (PEMWE) attracts most attention for its advantages because of simple structure and high energy conversion efficiency.

Usually, water electrolysis needs higher voltage supply than theoretical value, due to high overvoltage in oxygen evolution reaction at anode. To minimize the anodic overvoltage, it is recently suggested that annealing catalyst powder at certain temperature and time can decrease both the HFR and activation overvoltage, and the annealing effects are proven in the conventional PEMWE, where operation temperature is around 80℃ [1-4]. The current mechanisms about how annealing impacts electrolysis performance are as followings: (i) annealing can largely decrease electrolysis voltage by reducing HFR [1]; (ii) annealing will enlarge the specific surface area [2]; (iii) annealing increases the catalyst activity by changing the structure of catalyst particle [3~4]. In order to examine the above three mechanisms, we evaluated I-V and I-HFR characteristics of IrO2 samples which have different specific surface area and were annealed under different temperatures and time. This study also challenges to find the optimum condition of annealing time and temperature, to minimize electrolysis voltage under rather high temperature operation of 100℃.

2. Experiment

PEMWE cell used in this study consists of a homemade CCM (with electro-catalyst area of 4cm2), porous current collectors, and separators with flow channels. For the CCM fabrication, anode and cathode catalytic electrodes are hot pressed to a membrane (Nafion117). Iridium oxide powder (IrO2 powder, type IV, Tokuriki Co., Japan) as for anode catalyst is annealed in air at different time and temperatures shown in table 1. Rasten et al [1] suggests that annealing at 490℃ and 8hrs results in the lowest electrolysis voltage. Siracusano et al [3~4] shows that the optimum annealing condition should be 350℃ and 1hr. Based on these, Annealing condition of the powder is at 350℃ and 490℃ from 1 to 8 hrs. To examine the effect of enlarging specific surface area, IrO2 samples with two different specific surface area were selected: sample 1 and sample 2 originally has the specific surface area of 42m2/g and 84m2/g respectively. A commercial 46% Pt/C (Tanaka Kikinzoku Japan) catalyst is used for cathode electrode. Ti mesh (Nikko Techno Co., Japan, fiber diameter 20μm) is used for anode current collector. As for cathode current corrector, carbon paper (hydrophobic, SGL Co., Germany, 34BC) is chosen for comparing with each other. Water at room temperature is fed into only anode channel at a flow rate of 1mL/min, which corresponds to the water utilization of 2% at 1A/cm2. All the experiments are conducted at atmospheric pressure and 100℃.

3. Result and discussion

Fig. 1 (a) and (b) shows how specific surface area and annealing conditions impact on I-V characteristics and high frequency resistance at 10 kHz. The IV and I-HFR characteristics revealed that: (i) enlarging specific surface area to two times has little impact on electrolysis performance; (ii) annealing catalyst powder at 350℃ and 1hr introduced the best electrolysis performance and annealing at 490℃ and 8hrs introduced the highest electrolysis voltage. When annealing temperature is 350℃, long annealing time will raise the electrolysis voltage. But when annealing temperature is 490℃, annealing time makes a limited impact on electrolysis voltage. It also should be noticed that the OCV changes with annealing temperature, annealing catalyst annealed at 490℃ generates much higher OCV (about 80mV) than that annealed at 350℃. Such phenomenon means annealing at 490℃ will decrease the catalyst activity. Annealing catalyst at certain temperature can decrease the HFR, but the decreased ohmic overvoltage can't explain the total electrolysis voltage difference. Therefore, among the above three main mechanisms, annealing method should impact the electrolysis performance by increasing the catalyst activity via changing the structure of catalyst particle.

References

[1]. Egil Rasten, Georg Hagen, Reidar Tunold. Electrochimica Acta (2003) 3945-3952.

[2]. J. C. Cruz, V. Baglio, S. Siracusano, R. Ornelas, L. Oritiz-Frade, L. G. Arriaga, V. Antonucci, A.S. Aricò. J Nanopart Res (2011) 13: 1639-1646.

[3]. S. Siracusano, V. Baglio, A. Stassi, R. Ornelas, V. Antonucci, A.S. Aricò. International journal of hydrogen energy 36 (2011) 7822-7831.

[4]. S. Siracusano, S. Siracusano, A. Di Blasi, N. Briguglio, A. Stassi, R. Ornelas, E. Trifoni, V. Antonucci, A.S. Aricò. International journal of hydrogen energy 35 (2010) 5558-5568.

Figure 1

2421

, , , and

The development of highly active and stable electrocatalysts for OER is urgent due to the slow kinetics and high overpotential of oxygen evolution reaction (OER) [1]. To date, the most promising catalysts for OER are IrO2 based catalysts with ordered porous structures, where a templating method was generally used to ensure the high utilization of the catalysts. However, the synthesis route for such catalysts is generally of high technique complexation, where the templates need to be effectively removed after synthesis [2,3]. Hence, a simple template-free method that leads to the formation of highly porous IrO2 catalysts with high surface area is highly attractive.

Herein, a satisfactory template-free ammoniating method was developed in which NH3 ligand was successfully kept until pyrolysis treatment, thus acting as the pore-forming agent to achieve highly porous IrO2. The specific and electrochemical surface areas of as-prepared IrO2 catalysts increased with the increasing quantity of NH3·H2O additive. Specifically, IrO2 (1:100)-450 °C catalyst synthesized with the highest ratio of NH3 ligand exhibited high specific surface area of 363.3 m2 g-1 as metal oxide. Furthermore, IrO2 (1:100)-450 °C showed remarkable electrocatalytic activity for acidic OER in the whole potential window, which can be ascribed to the high catalysts utilization. As shown in Figure 1, the over-potential for the IrO2 (1:100)-450 °C catalysts to attain current at 10 mA cm-2 was only 282 mV, which was shifted negatively by 31 mV and 46 mV in comparison to the home-made IrO2 (without NH3) and commercial IrO2 (CM) catalysts, respectively. Thus, the NH3 mediating method was found a superior method for increasing catalysts utilization, where the IrO2can be conveniently synthesized in large scale.

Figure 1. CV curves (Left, the inset shows the specific surface area) and LSV curves (Right, the inset shows the TEM of IrO2 (1:100)-450°C) of IrO2 (1:100)-450°C, IrO2 (1:10)-450°C, IrO2-450°C, and IrO2(CM) catalysts.

Acknowledgments

This work is supported by National Basic Research Program of China (973 Program, 2012CB932800), National Natural Science Foundation of China (21433003, 21373199), the Science & Technology Research Programs of Jilin Province (20150101066JC, 20160622037JC).

References

[1] S. Park, Y. Shao, J. Liu, and Y. Wang, Energy Environ. Sci., 5, 9331 (2012).

[2] W. Hu, Y. Wang, X. Hu, Y. Zhou, and S. Chen, J. Mater. Chem., 22, 6010 (2012).

[3] E. Ortel, T. Reier, P. Strasser, and R. Kraehnert, Chem. Mater., 23, 3201 (2011).

Figure 1

2422

and

A bifunctional catalyst for both oxygen reduction reaction (ORR) and oxygen evolution reaction (OER) could be employed in a unitized regenerative fuel cell, an energy storage device that can be coupled to intermittent renewable energy such as wind or solar to peak-shift electricity to the grid. Though significant work has been done over the past few decades, no catalyst for both ORR and OER with high activity for both reactions have been discovered. It is well know that platinum or platinum based alloys exhibit the highest activities for ORR among all other catalyst, however they display poor activity for OER. In contrary, transition metal (hydro)oxides, such as iron, nickel, or cobalt (hydro)oxides, always possess highly activities for OER but poor activity for ORR. Due to "migration-effect" of transition metal in ORR, in which the free transition metal (hydro)oxides will move to the surfaces of platinum during electrochemical process and results in the dramatically decrease of the ORR activity of Pt eventually, it is impossible to achieve high activity for both ORR and OER by simply mix platinum or platinum based alloy with transition metal (hydro)oxides together. Inspired by this, we developed a novel method to prepare 7nm platinum nanoparticles doped with atomic sized cobalt oxides for the applications of ORR and OER. Our primary results showed that these hetero-structured materials exhibited higher activity than state-of-the-art commercial platinum for ORR, but also 10 times higher activity for OER than Co3O4. Most importantly, our catalysts also demonstrated excellent stability. Apparently, the interfaces must play a dominated role in both ORR and OER. Investigation of the interfaces between platinum and Co3O4 of our catalyst in atomic scale is crucial to understand how the interfaces worked and what kinds of interfaces are better for both ORR and OER. Our work will advanced the fundamental understanding of the metal-oxide interface in enhancing the ORR and OER, which is envisioned to shed light on the design of advanced catalysts for catalyzing complex chemical processes.

2423

A PEM (Polymer Electrolyte Membrane) water electrolysis cell is a concept of zero-gap cell that uses a thin (100-200 micrometers thick) film of proton-conducting polymer as solid electrolyte [1]. Elevated operating current densities (in the multi A.cm-2 range) can be reached efficiently. Most popular polymer electrolytes operating between 40 and 80°C are perfluoro-sulfonated materials (PFSA like Nafion® or Aquivion® [2]). Attempts to use phosphoric acid doped PBI (polybenzimidazole) electrolytes operating at higher (150-200°C) temperature have also been reported [3]. Platinum group metals (PGM) electrocatalysts are extensively used because only them can sustain the highly acidic environments [4]. Regarding electrocatalysis, conventional PEM water electrolysis cells are using platinum black (unsupported Pt particles) at the cathode for the hydrogen evolution reaction (HER) and unsupported iridium dioxide at the anode for the oxygen evolution reaction (OER). Typical PGM loadings vary from 0.5 – 1.0 mg.cm-2 at the cathode to 1-2 mg.cm-2 at the anode. Whereas the relative cost of platinum group metal (PGM) electrocatalysts in industrial PEM systems is limited to a few percent, costs constraints (especially in view of the development of electrolysers at the multi MW scale, for example for energy storage applications) are calling for cheaper solutions. A first option is to reduce PGM loading. A second option is to develop non-PGM electrocatalysts.

Regarding the reduction of PGMs, carbon-supported Pt nano-particles can be advantageously used to reduce Pt loadings from 1 down to 0.1 mg.cm-2 at the cathode. Reduced IrO2 loadings (down to 0.5 – 1.0 mg.cm-2) have been demonstrated [5] but addition of an inert metal [6] or alloying (to form ternary or quaternary mixed oxides [7]) are also viable alternatives. Regarding the development of non-PGM materials, most advances concern the cathode. Ideally, catalyst containing transition metals (Ni, Co, Fe) that maximize the exchange current density as a function of the M-H bond strength should be used [8]. They are significantly less expensive than PGMs and still adequately electro-active. Whereas in alkaline water electrolysis technology, Ni and Co bulk particles are commonly used, their implementation at the cathode of PEM water electrolyzers is not a straightforward task because they are rapidly corroded. New approaches based on molecular chemistry have been developed to bypass the problem. For example, metallic molecular complexes such as Co, Ni, Fe clathrochelates have been successfully implemented at the surface carbonaceous substrates (carbon powder and fibers) and implemented in PEM water electrolyzers [9]. They offer several advantages compared to nanoparticles: (i) non noble metals can be used as active centres; (ii) only limited amounts of metals are required; (iii) redox properties of active centres can be tuned to desirable values by selecting appropriate organic ligands; (iv) ligands can be used as chemical linkers for efficient surface functionalization.

The purpose of this communication is to review existing and innovative electrocatalysts for PEM water electrolysis applications and to compare their efficiency in relation with microstructural aspects.

[1] W.T. Grubb, L.W. Niedrach, Batteries with Solid Ion-Exchange Membane Electrolytes. II. Low Temperature H2-O2 Fuel Cells, J. Electrochem. Soc., 107(2) (1960) 131 – 134.

[2] K.A. Mauritz, R.B. Moore, 'State of understanding of Nafion', Chemical Reviews, 104 (2004) 4535 – 4585.

[3] D. Aili, M.K. Hansen, C. Pan, Q. Li, E. Christensen, J.O. Jensen, N. Bjerrum, Phosphoric acid doped membranes based on Nafion, PBI and their blends, Int. J. Hydrogen Energy, 36 (2011) 6985 – 6993

[4] P. Millet, 'PEM water Electrolysis for hydrogen production: Principles and Applications', chapter 10, PEM electrolyzer characterization tools, D. Bessarabov, H. Wand, H. Li, N. Zhao, CRC Press (2015).

[5] C. Rozain, E. Mayousse, N. Guillet, P. Millet, Influence of iridium oxide loadings on the performance of PEM water electrolysis cells: Part I – Pure IrO2-based anodes, J. Appl. Catalysis B: Environmental, 182 (2016) 153 – 160.

[6] C. Rozain, N. Guillet, E. Mayousse, P. Millet, Influence of iridium oxide loadings on the performance of PEM water electrolysis cells: Part II – Advanced anodic electrodes, J. Appl. Catalysis B: Environmental, 182 (2016) 123 – 131.

[7] A.T. Marshall, S. Sunde, M. Tsypkin, R. Tunold, Performances of a PEM water electrolysis cell using IrxRuyTazO2 electrocatalysts for the oxygen evolution electrode, Int. J. Hydrogen Energy, 32(13) (2007) 2320-2324.

[8] S. Trasatti, Work function, electronegativity, and electrochemical behaviour of metals: III. Electrolytic hydrogen evolution in acid solutions, J. Electroanal. Chem. 39 (1972) 163-184.

[9] M-T. Dinh Nguyen, A. Ranjbari, L. Catala, F. Brisset, P. Millet and A. Aukauloo, Implementing Molecular Catalysts for Hydrogen Production in Proton Exchange Membrane Water Electrolysers, Coord. Chem. Review, 256 (2012) 2435 – 2444.

2424

, and

The interest in harnessing energy from renewable sources and a push to achieve environmental cleanliness in the energy industry keeps gaining momentum. The need to be able to convert and store all that renewable energy has rekindled interest in hydrogen as a clean and environmentally benign energy carrier. In fact, hydrogen seems to be the only energy storage medium capable of handling the vast grid-scale energy storage in a reasonable manner. Likewise, recent accomplishments in auto-motive fuel cells adds further impetus to advance simultaneously renewable and economical hydrogen generation technology, which will be the ultimate key to success for hydrogen as fuel in societal everyday transportation solutions. All this is done with the ultimate goal of supplanting the current hydrocarbon based economy by hydrogen economy as means to achieve globally environmentally friendly energy industry and regional/local energy independence.

This recent increase in significance placed on the age-old water electrolysis process gave rise to new approaches to generate hydrogen with increased efficiency and lower cost. Polymer Electrolyte Membrane (PEM) water electrolysis emerged as one of the best matched choices today for the highly variable and unpredictable nature of the renewable energy.

There are two main ways to lower the cost of hydrogen production via PEM water electrolysis: to lower the capital expenses (CAPEX) and/or to lower the operating expenses (OPEX). We at 3M have recently showed a way to address reducing the first (high CAPEX ) by successfully demonstrating the ability to widen the range of current densities where electrolyzers can operate from a previous maximum of about 2.0 A/cm2 as used today in commercial electrolyzers, to as much as 20 A/cm2 in our novel constructions. While the research on high power density operation of PEM water electrolysis still continues and many questions remain as yet unanswered, the 3M's proprietary Nano Structured Thin Film (NSTF) catalyst with its highly conductive, compact, and hydrophilic catalyst layer seems uniquely fitting to the task. Indeed, its hydrophilicity, which is perhaps a major challenge in fuel cells, is a tremendous strength and asset in PEM water electrolysis and one of the features enabling the extremely high virtually mass transport free current densities we reported previously1.

In this work, we intend to present results from our attempts to tackle the second means to reduce the cost of hydrogen production (the OPEX), by increasing the intrinsic activity of the catalyst as means to increase the kinetics of oxygen evolution and hence the efficiency of water electrolysis. In the preceding work we have explored the Pt/Ir compositional parameter space (in the NSTF alloy catalyst format) and the compositional effects on fundamental catalyst activity, as determined by Rotating Disk Electrode (RDE) measurements and the evaluation of single electrolyzer cells2,3. That work has shown, rather unexpectedly we must add, quite detrimental effects of alloying basic Ir-NSTF catalyst with Pt in RDE tests. The results were confirmed, and perhaps even more drastic, when catalyst compositions were tested for performance in the actual electrolyzer hardware. Now, we intent to show the effect of alloying Ir with sample of non-PGM transition metals (again, using 3M's proprietary Ir-NSTF catalyst format) and the effects of such alloys on the basic catalytic activity. Subsequently, we will correlate this to the in-situ performance measurements in a PEM electrolysis cell. Finally, we will discuss the possibilities in other promising alloying compositions and structures.

1.    Krzysztof A. Lewinski, Sean M. Luopa, (invited) "High Power Water Electrolysis as a New Paradigm for Operation of PEM Electrolyzer" (abstract 1948), Spring ECS Meeting, Chicago, IL, May 2015.

2.    Krzysztof A. Lewinski, Dennis van der Vliet, and Sean M. Luopa, "NSTF Advances for PEM Electrolysis - the Effect of Alloying on Activity of NSTF Electrolyzer Catalysts and Performance of NSTF Based PEM Electrolyzers" (Abstract 1457), Fall ECS Meeting, Phoenix, AZ, Oct 2015.

3.    Krzysztof A. Lewinski, Dennis van der Vliet, and Sean M. Luopa, "NSTF Advances for PEM Electrolysis - the Effect of Alloying on Activity of NSTF Electrolyzer Catalysts and Performance of NSTF Based PEM Electrolyzers", (10.1149/06917.0893ecst), ECS Transactions69 (17) , p. 893-917 (2015).

2425

, , , , , , and

The development of cost-effective polymer electrolyte membrane (PEM) electrolyzers is key to the implementation of a hydrogen-based infrastructure. It cannot be denied that there is an ever-increasing demand to shift away from the use of nuclear power and the traditional burning of fossil fuels as primary energy systems. The only foreseeable long-term solution has been increasingly focused on the implementation of clean, renewable power supplies, such as wind farms and solar power stations. In this regard, the electrochemical conversion of water to hydrogen is expected to play a key role in the development of scalable energy storage that is required for such intermittent power supplies(1).

The cathodic generation of hydrogen from water splitting is known to be extremely facile on Pt-based catalysts(2). The simultaneous oxygen evolution reaction (OER) occurring at the anode, however, is limited by sluggish kinetics and requires a considerable overpotential to achieve modest current densities. Moreover, the harsh acidic environment and high anodic operating potentials limits the choice of stable electrocatalyst materials to those of the noble metal oxides. Reduction of the noble metal loading at the anode and enhancing catalyst stability for OER in PEM electrolyzers remains a challenge. Perhaps the most widely implemented approach for reducing the noble metal content is that which considers reducing the catalyst particle size(3, 4). A major issue that arises from this approach, however, is that it becomes increasingly difficult to establish structure-activity relationships for the OER due to the transformation processes, e.g. changes in microstructure and crystallinity, that commonly occur during the preparation of the materials.

The research reported herein is focused on expanding the fundamental understanding of the influence of crystallinity, particle size, and microstructure on the electrochemical OER activity of nanocrystalline IrO2 with particular emphasis pointed towards formation of a hydrous surface layer. Chlorine−free iridium oxide nanoparticles are synthesized using the modified Adams fusion method(5), which is capable of producing spherical 1.7 ± 0.4 nm particles with a specific surface area of 150 m2/g using a low temperature synthesis (350 °C). Increasing the synthesis temperature to 600 °C results in the formation of larger, rod−shaped particles mostly terminated by highly ordered non−defective (110) surfaces. X-ray absorption spectroscopy (XAS), X-ray photoelectron spectroscopy (XPS), and electrochemical studies indicate the presence of a hydrous surface layer, i.e. Ir(O)OH, that leads to an enhanced OER activity. We report that it is possible to create a larger hydrous layer on smaller nanoparticles and thus increase the specific OER activity.

References:

1. L. Bertuccioli, A. Chan, D. Hart, F. Lehner, B. Madden and E. Standen, Development of Water Electrolysis in the European Union, in, p. 83, Fuel Cells and Hydrogen Joint Undertaking (2014).

2. W. Sheng, H. A. Gasteiger and Y. Shao-Horn, Journal of The Electrochemical Society, 157, B1529 (2010).

3. J. C. Cruz, V. Baglio, S. Siracusano, R. Ornelas, L. Ortiz-Frade, L. G. Arriaga, V. Antonucci and A. S. Aricò, J. Nanopart. Res., 13, 1639 (2011).

4. Y. Lee, J. Suntivich, K. J. May, E. E. Perry and Y. Shao-Horn, The Journal of Physical Chemistry Letters, 3, 399 (2012).

5. E. Oakton, D. Lebedev, A. Fedorov, F. Krumeich, J. Tillier, O. Sereda, T. J. Schmidt and C. Coperet, New Journal of Chemistry, 40, 1834 (2016).

2426

, and

Within the United States, hydrogen is a major chemical commodity for the production of ammonia in agriculture and the upgrading of crude oil in transportation. Hydrogen in the US is produced largely from natural gas by steam methane reformation.[1] Although electrochemical water splitting currently accounts for a small amount of hydrogen production, it is expected to have a larger role in the future, particularly when using renewable energy sources as input. Studies in proton exchange membrane (PEM) electrolysis are usually focused on catalysts in the oxygen evolution reaction, due to the high overpotential relative to the hydrogen evolution reaction. PEM electrolyzers typically use iridium (Ir) or Ir oxide as the anode catalyst, due to reasonable activity and stability. Although platinum (Pt) and ruthenium (Ru) are often examined as potential alternatives, Pt requires a higher overpotential and Ru is prone to dissolution at elevated potential.[2, 3]

Determining the electrochemical surface area (ECA) of Ir allows for the quantification of the number of sites available to participate in oxygen evolution. It also allows for the evaluation of site-specific activity, or site quality. ECAs are of interest in Ir catalyst development, to compare different catalyst types and direct future research. ECAs are also of interest in durability studies to evaluate the modes of Ir degradation, and to determine whether activity losses are due to deteriorating surface area or site quality.

The ECAs of Ir catalysts have typically been evaluated using hydrogen underpotential deposition, carbon monoxide oxidation, or capacitance. These methods are generally limited to predurability measurements and catalyst type: hydrogen underpotential deposition and carbon monoxide for Ir metals; capacitance for Ir oxides. Hydrogen underpotential deposition and carbon monoxide oxidation can be used to determine the ECA of Ir metals, but not Ir oxides, and are sensitive to surface oxides formed on Ir metals following oxygen evolution characterization and durability testing. To date, no method is available to determine the ECAs of Ir and Ir oxide, prior to and following durability testing. Mercury underpotential deposition is presented in this study as an alternative, able to produce reasonable ECAs on Ir and Ir oxide nanoparticles, and able to produce similar ECAs prior to and following characterization in oxygen evolution. This method was previously developed for the study of polycrystalline Ir, and expanded in this study to include nanoparticle catalysts, oxides, and oxygen evolution relevant testing condition.[4]

Figure 1. Cyclic voltammograms of (a) Ir nanoparticles and (b) Ir oxide nanoparticles in 0.1 m perchloric acid (blue) and 0.1 m perchloric acid containing 1 mm mercury nitrate (red).

[1] A. Milbrandt, M. Mann, in: U.S. Department of Energy (Ed.), http://www.nrel.gov/docs/fy09osti/42773.pdf, 2009.

[2] T. Reier, M. Oezaslan, P. Strasser, ACS Catalysis, 2 (2012) 1765-1772.

[3] M. Pourbaix, Atlas of electrochemical equilibria in aqueous solutions, National Association of Corrosion Engineers, Houston, Texas, 1974.

[4] S.P. Kounaves, J. Buffle, Journal of The Electrochemical Society, 133 (1986) 2495-2498.

Figure 1

2427

, , and

Renewable H2 production is a prerequisite for successfully establishing fuel cell-based electromobility and hydrogen infrastructure. In this respect, proton exchange membrane water electrolysis (PEM-WE) has attracted much interest, not least due to the enormous power densities possible in these devices (1). Research is mainly focused on the anode side, where the oxygen evolution reaction (OER) takes place, causing the vast majority of kinetic losses.

OER catalysts of choice are usually iridium oxide (IrOx) based, owing to its decent activity and acceptable stability (2,3). Recent studies showed a strong dependence of the OER activity and dissolution resistance of IrOx on its surface morphology and hydration state, which were controlled by the calcination temperature during the synthesis. They found that crystalline, thermal IrO2is the most stable but least active species (4,5).

TGA-analysis from our lab reveals that IrOx can easily be reduced to metallic Ir in dilute H2 at a typical PEM-WE operation temperature of 80 °C. This is also observed for IrOx based membrane electrode assemblies (MEAs) held at OCV in a PEM-WE, where crossover H2 from the cathode side reduces the surface of the IrOx catalyst at the anode within hours, as evident from the formation of H-UPD features in the cyclic voltammogram (see Figure 1b, red vs. blue CV). At the same time, polarization curves after this reduction show significantly decreased cell voltage at slightly reduced Tafel slopes (Figure 1a, red vs. blue curves), corresponding to an increased OER activity. However, a subsequent CV following these polarization curves (see Figure 1b, black CV) reveals that the H-UPD features have disappeared again and that the catalyst surface properties have transformed to a state closer to the less crystalline, hydrous IrOx reported in ref. 4.

This suggests that partial IrOx reduction and re-oxidation into (hydrous) IrOx can occur during cycles of extended OCV periods and electrolyzer operation. In analogy to voltage cycling degradation observed in fuel cells, the here described reduction/oxidation cycles might also lead to iridium dissolution. Therefore, we will study the effect of transient operation conditions in PEM-WEs on the OER activity, surface properties, and stability of IrOx based anodes. We will also provide a systematic analysis of OER kinetic parameters such as exchange current density and activation energy as well as their dependence on relative humidity for a commercial iridium oxide catalyst.

References

(1) K. A. Lewinski, D. F. van der Vliet, and S. M. Luopa, ECS Transactions, 69, 893 (2015).

(2) C. Rozain, E. Mayousse, N. Guillet, and P. Millet, Appl. Catal. B, 182, 123 (2016).

(3) E. Fabbri, A. Habereder, K. Waltar, R. Kötz, and T. J. Schmidt, Catal. Sci. Technol., 4, 3800 (2014).

(4) T. Reier, D. Teschner, T. Lunkenbein, A. Bergmann, S. Selve, R. Kraehnert, R. Schlögl, and P. Strasser, J. Electrochem. Soc., 161, F876 (2014).

(5) S. Cherevko, T. Reier, A. R. Zeradjanin, Z. Pawolek, P. Strasser, and K. J. J. Mayrhofer, Electrochem. Commun., 48, 81 (2014).

Figure 1a. PEM-WE polarization curves recorded at 80 °C under dynamic O2 (anode)/H2 (cathode) at 147 kPaabs and 200 % RH (inlet). Blue curves were taken after a 12 h conditioning at 1 Acm-2 while red curves signify the status after a 15 h in-situ reduction of the anode catalyst in an H2 atmosphere. Hollow circles and crosses are subsequent repetitions.

Figure 1b. Anode CVs recorded under dry N2 at 100mVs-1. Blue CV: before the polarization curves in blue, red and black CVs: before and after the polarization curves in red.

The anode was loaded with 0.66 mgIrcm-2 of a commercial IrOx/TiO2 using 12 wt-% of ionomer in the catalyst layer, cathode was loaded with Pt/C at 0.35 mgPtcm-2 and a Nafion® XL membrane was used. MEAs with 5 cm2 active area were tested in a house-made single cell hardware with gold-plated titanium flowfields using porous Ti sheets and carbon fiber paper as GDLs on the anode and cathode side, respectively.

Acknowledgements:

P. J. Rheinländer would like to acknowledge financial support from Greenerity GmbH. M. Bernt would like to acknowledge funding from the Bavarian Ministry of Economic Affairs and Media, Energy and Technology through the project ZAE-ST (storage technologies).

Figure 1

2428

, , , and

Water electrolysis technology has attracted much attention because it serves as the key part of a carbon-neutral energy storage system as follows: electricity generated from renewable resources such as sunlight, water power, and wind, is used to produce hydrogen by water electrolysis. Then, the hydrogen is supplied to fuel cell system for power generation, which only generates electricity and water. Proton exchange membrane water electrolyzers (PEMWEs) are currently the most promising candidate because polymer electrolyte membrane provides higher ionic conductivity, better load flexibility, and purer hydrogen than alkaline electrolyte solution, anion exchange membrane, and solid oxide electrolyte, etc.

The key challenge for PEMWEs is to develop efficient anode electrocatalyst because of the high catalytic efficiency loss related with the oxygen evolution reaction (OER) on anode. In addition, the use of PEM requires the anode material durable under long-term anodic polarization in acid environment, which significantly limits the options of materials. At present stage, iridium oxide (IrO) is widely regarded as the most promising anode electrocatalyst for PEMWEs owing to its high activity for OER and chemical stability in strong acid environment. Considering the price and scarcity of Ir, however, it is highly desired to lower the amount of Ir in the anode. In view of this, many approaches have focused on enhancing the mass activity of Ir-based catalysts towards OER by designing novel nanostructures. Another effective way is to disperse IrO2 on electronic conductive support with large surface area. Unfortunately, the most common support material for electrocatalysts, carbon materials, are regarded as not appropriate for PEMWEs because it would encounter severe degradation at high oxidizing potential region. Instead, metal oxides with large surface area and resistance to acid are used. However, their low electrical conductivity limits the electrochemical performance for OER.

In previous studies, our group wrapped pristine multi-wall carbon nanotubes (MWNTs) with polybenzimidazole (PBI). The thin layer of PBI (thickness<1 nm) provides abundant binding sites for cations. As a result, uniform Pt nanoparticles could be deposited homogeneously on the PBI-wrapped pristine MWNTs. The resultant composite catalyst exhibited ehnhanced electrocatalytic activity and extraordinary long-term durability even polarized at high potential region of 1.0-1.5 V vs. RHE because (1) the aggregation of Pt nanoparticles was significantly suppressed due to the constraint effect of polymer and (2) the highly crystallized graphitic surface of pristine MWNTs refrained the electro-oxidation of carbon. The results provide a new insight into carbon nanotubes as an excellent support material with high corrosion-resistance under anodic polarization in acid environment.

In present work, a composite catalyst of pristine MWNT, PBI, and IrO2 prepared by similar strategy. For comparison, Ir was also deposited on carbon black (CB) and PBI-wrapped CB. The products were characterized by thermogravimetry (TG), X-ray diffraction (XRD) analysis, X-ray photoelectron spectroscopy (XPS), Raman spectroscopy, and scanning transmission electron microscope (STEM). The electrochemical properties of the catalysts were investigated by half-cell tests.

E-11 Alkaline & DFC Electrocatalysis 1 - Oct 4 2016 8:00AM

2429

, and

Alkaline environment offers the increased electrocatalystic activity and stability of non-precious metal group catalysts for the oxygen reduction reaction. However, sluggish hydrogen oxidation reaction (HOR) activity of electrocatalysts under alkaline environment has remained an issue to overcome. In this presentation, we report the hydrogen oxidation inhibition by organic cation adsoprtion to expalin the low HOR activity of Pt based electrocatalyst. The HOR activity of Pt based electrocatalyst in several organic cationic solutions including tetramethylammonium, tetrabuthylammonium, and tetrabutylphosphonium is investigated using rotating disk electrodes. Time-dependent and potential driven adsorption behaviors of the cationic groups suggest that this adsorption has rather complex chemisorption nature rather than just simple adsorption. The new proposed inhibition mechanism may give useful insites on how to design ionomeric binders and electrocatalysts for advanced alkaline membrane fuel cells.

2430

, , and

It has long been recognized that the reaction rates of the hydrogen oxidation and hydrogen evolution reactions (HOR and HER) are slower in basic than acidic electrolytes, even though the surface intermediate of adsorbed hydrogen is independent of solution pH. Understanding the root of this observation is critical to designing catalysts for a multitude of electrochemical reactions with relevance to energy conversion and storage. In this work, we undertake a fundamental investigation relating interfacial water structure to the reaction barrier for hydrogen underpotential adsorption and to the HER/HOR kinetics. The pH-dependence of HOR/HER kinetics has been attributed to a variety of reasons. The research groups of Gasteiger and Yan have both suggested that in base, hydroxide ions stabilize the Pt-H bond for stronger binding and slower catalysis1,2. Central to this hypothesis is an experimentally measured shift with pH in the peak potential for hydrogen underpotential deposition on the (100) and (110) steps on the Pt surface. Although it is unclear why pH would stabilize adsorbed hydrogen, stronger binding would push platinum further to the right of the volcano peak and decrease the overall activity. Markovic et al proposed instead that in base, the water recombination/dissociation step of HOR/HER requires specific adsorption of hydroxide ions and therefore optimal binding to two adsorbates3. One observation that has been less discussed than either hydrogen or hydroxide adsorption is the effect of pH not only on the apparent hydrogen adsorption energy, but the kinetic barrier to adsorption. Koper et al attributed this observation to a pH-dependent water orientation at the surface, and hypothesized that "H-down" in acid vs "H-up" in base resulted in slower transfer of hydrogen to and from the surface4. The same pH-dependence of water orientation was recently supported computationally by Rossmeisl's group5.

 In this work, we examine specifically the hypothesis that water orientation governs the rate of hydrogen adsorption and thus the overall HER/HOR kinetics. The native oxide formed on chromium metal is known from colloidal literature to have a potential of zero charge (Epzc) of 1.22V vs. RHE6. We hypothesize that if water orientation truly governs the alkaline HOR/HER rate, the negative oxide surface charge on chromium oxide will induce an H-down water orientation adjacent to platinum nanoparticles supported on the surface. This will result in faster hydrogen adsorption kinetics and thus, higher activity for alkaline HOR/HER than extended platinum surfaces. To test this hypothesis, platinum nanoparticles were synthesized via pulsed electrodeposition on chromium oxide. The nanoparticle loading was controlled by the duration of deposition pulse, with longer times corresponding to higher coverage as well as larger particles. The hydrogen adsorption kinetics were measured by cyclic voltammetry in the hydrogen underpotential deposition region. Varying the sweep rate results in greater peak separation that can be used to extract rate constants for the adsorption reaction via traditional electroanalysis7. A typical voltammogram of Pt/CrOx in 0.1M KOH is shown in Fig. 1a. As the scan rate increases, kinetic barriers cause greater peak separation (dEp) for hydrogen adsorption on the 100 and 110 facets of platinum. For lower loadings with shorter pulse periods, both smaller currents and smaller peak separations are observed due to the lower Pt surface area. The effects of surface area can be accounted for by normalizing current to Pt surface area as measured by Hupd coulometry. The resulting surface-adsorption Tafel plot is shown in Fig 1b. The peak potential separation dEp is much greater for the larger nanoparticles with higher loading. This difference suggests that smaller particles have qualitatively more rapid adsorption kinetics due to the advantageous water orientation induced by negatively charged CrOx. The results of this study contribute to resolving a long-standing paradox in electrocatalysis and surface science by highlighting the importance of kinetic barriers, as well as adsorption energies, and suggest the ability to modulate alkaline hydrogen activity through metal-support interactions. Future work will discuss the relation of hydrogen underpotential deposition kinetics to HER/HOR activity and the effects of different oxide supports.

(1) Durst et al. Energy Environ. Sci.2014, 2255.

(2) Sheng et al. Nat. Commun.2015, 5848.

(3) Strmcnik et al, Nat. Chem.2013, 1.

(4) Van Der Niet et al. Catal. Today2013, 105.

(5) Rossmeis et al. Phys. Chem. Chem. Phys.2013, 10321.

(6) McCafferty, E. Electrochim. Acta2010, 1630.

(7) Angerstein-Kozlowska et al. J. Electroanal. Chem. Interfacial Electrochem.1979, 1.

Figure 1

2431

, , , and

Anion Exchange Membrane Fuel Cells (AEMFCs) are fast growing alternative to Proton Exchange Fuel Cell (PEMFC) technology. It was shown that several classes of non-platinum group metal (non-PGM) catalysts outperformed platinum in the oxygen reduction in alkaline media.[1] In contrast the published data on alternative to platinum anodes for hydrogen oxidation reaction in alkaline media is scarce.[2]One of the possible non-PGM catalysts for the HOR in alkaline media are Ni- and Pd-based materials. We have prepared Ni-Mo-Cu catalysts that shows HOR activities in the rotating disk electrode (Figure 1). The preliminary results on MEA tests show that the Ni-based alloys as an anode have shown extremely high open-circuit voltages (OCVs) in both oxygen and air (up to 1.1V), which is significantly higher compared to PGM-free anodes in PEMFCs. However, the effort in integration of those materials into MEA with most appropriate ionomer and MEA fabrication technique is needed. In this talk, we will discuss the preparation and characterization of the novel electrocatalysts for HOR in alkaline media.

Figure 1. HOR on Ni-Mo-Cu catalysts in 1M KOH saturated with H2, 1600RPM, 10 mV s-1.

Acknowledgements: Department of Energy, Hydrogen Oxidation Reaction in Alkaline Media, Control Number: 0966-1624, Award Number: DE-EE0006962 (PI A. Serov).

[1] M. H. Robson, A. Serov, K. Artyushkova, P. Atanassov, Electrochim. Acta, 90, 2013, 656–665.

[2] A. Serov, C. Kwak, Applied Catalysis B: Environmental, 91, 2009, 1–10.

Figure 1

2432

, and

Electrochemical energy productions are considered as an alternative energy sources because those energy consumptions are recognized to be more environmentally friendly and sustainable. The fuel cells, especially the direct methanol fuel cells (DMFCs), have received a lot of attention due to the higher energy density (5.04 KWh/L), easy storage and transportation compared with the hydrogen fuel cells (0.53 KWh/L). Also the direct conversion of methanol has a voltage similar (2CH3OH+3O2→2CO2+4H2O+6e-, E=1.19 V) to that of hydrogen (E=1.23 V). However, the DMFCs still suffer from three problems on the anode side; namely, i) the carbon monoxide (CO) poisoning of the platinum nanoparticles (Pt-NPs), which is generated from the uncompleted methanol oxidation reaction (MOR), ii) the sluggish methanol oxidation reaction (MOR) compared to the hydrogen oxidation reaction (HOR) and iii) the low durability of the electrocatalyst in terms of Pt stability and carbon corrosion, which degrades the fuel cell performance.

To commercialize the direct methanol fuel cells (DMFCs), the durability of the anodic electrocatalyst needs to be highly considered, especially under high temperature and methanol concentration conditions. The low durability caused by the carbon corrosion as well as the carbon monoxide (CO) poisoning of the platinum nanoparticles (Pt-NP) leading to the decrease of the active Pt-NPs and increase in the inactive Pt-NPs covered by CO species. We previously reported that poly(vinylphosphonic acid) (PVPA) plays an important role in enhanced CO tolerance of the electrocatalyst due to the acceleration of the reaction between Pt and H2O to form the Pt(OH)ads that consumes the CO poisoned Pt, namely, Pt(CO)ads. 1 In this study, we deposited Pt-NPs on poly[2,2'-(2,6-pyridine)-5,5'-bibenzimidazole] (PyPBI)-wrapped nanoporous carbon (NanoPC)2,3) and coated the as-synthesized electrocatalyst with poly(vinylphosphonic acid) (PVPA). The durability of the as-synthesized NanoPC/PyPBI/Pt/PVPA was tested in 0.1M HClO4 electrolyte at 60 oC by cycling the potential from 1.0 to 1.5 V vs. RHE, and the results indicated that the NanoPC/PyPBI/Pt/PVPA showed an ~5 times better durability by comparison with that the commercial CB/Pt. The methanol oxidation reaction (MOR) of the electrocatalyst was tested before and after the potential cycling in the presence of 4M or 8M methanol at 60 oC and found that the CO-tolerance of the electrocatalyst was ~3 times higher than that of the commercial CB/Pt. Such a higher CO tolerance is due to the coating of the PVPA, which was proved by an EDX mapping measurement. The NanoPC/PyPBI/Pt/PVPA showed a high durability and CO tolerance under high temperature and high methanol concentration conditions indicating that the electrocatalyst would be used in real fuel applications.

References

1) N. Nakashima et al., J. Mater. Chem. A 2014,2, 18875-18880.

2) N. Nakashima et al., ACS Appl. Mater. Interfaces, 2016, in press.

3) N. Nakashima et al., ACS Appl. Mater. Interfaces, 2015, 7, 9800-9806.

2433

, and

Alkaline fuel cells have some advantages as compared with conventional acid fuel cells. Our research group has investigated the possibility of alkaline fuel cells using anion-exchange membrane as an electrolyte. In the presentation, I would like to introduce and discuss about some topics relating to alkaline fuel cells as follows:

1) Direct oxidation fuels: alcohols

We have been focusing on ethylene glycol as a direct fuel in anion-exchange membrane fuel cells (AEMFCs). Ethylene glycol has superior energy density (7.56 kWh dm−3) and higher boiling point (471 K) than some typical alcohol fuels such as methanol and ethanol. The oxidation of ethylene glycol in alkaline medium is faster than that in acid medium. And, surprisingly, ethylene glycol provides larger oxidation currents than methanol and polyols such as glycerol, erythritol, and xylitol in alkaline solutions. Thus, ethylene glycol is a promising fuel for AEMFCs, which overcomes a conventional direct methanol fuel cell (DMFC).

Ethylene glycol has the above-mentioned advantageous features as DAFCs' fuel, but it has a crucial problem derived from its inherent molecular structure. Ethylene glycol is a C2 molecule having a C−C bond in its structure. C−C bond is a stumbling block for electroorganic chemists since thus C−C bond is comparatively strong and cannot be easily cleaved. As well as ethanol oxidation, ordinal Pt catalyst cannot achieve the complete oxidation of ethylene glycol, and leaves some intermediate products. Therefore, in order to increase the efficiency of fuel utilization, active catalysts, having sufficient ability to break C−C bond in ethylene glycol, are strongly required. To our best knowledge, there is no available literature concerning the oxidation of ethylene glycol to CO2from an electrocatalytical view point. Here we introduce an effective C−C bond cleavage electrocatalyst for ethylene glycol.

2) Oxygen reduction catalysts: non-precious metal catalysts

Electrochemically reduction of oxygen is an essential reaction involving in fuel cells. Among the various kinds of electrocatalysts, Pt, Pt-alloy, Ag are known to be fairly active for oxygen reduction in alkaline solutions. Precious metals, however, have some serious problems to be overcome before worldwide spread of fuel cells, such as economical cost, reproducibility and limited resources. These problems motivated many researchers to study for alternative catalysts besides a standard catalyst of Pt/C. Among proposed alternative catalysts, perovskite-type oxide is one of the promising candidates as a non-precious metal catalyst. Since Meadowcroft pointed out the catalytic activity of lanthanum-cobalt oxide for oxygen reduction, perovskite-type oxides become attractive materials as electrocatalysts for many electrochemists. However, the detailed mechanism for oxygen reduction on perovskite-type oxide electrodes has not been fully clear yet. We investigated the oxygen reduction on perovskite-type oxides using thin film electrodes and single crystal electrodes, and shed lights on the catalytic roles of oxides and carbon additives.

2434

, and

Small alcohol molecules such as methanol and ethanol represent promising fuels to power fuel cells for mobile electronic devices, transportation and stationary applications. Ethanol is of particular interest due to its renewability by generation from biomass feedstocks, large energy density and compatibility with existing infrastructures for fuel storage and delivery. Although substantial progress has been made in the recent years for the development of direct ethanol fuel cell (DEFC) technology, obstacles are still present for the practical implication, largely in the lack of efficient and selective electrocatalysts for complete oxidation of ethanol to CO2.

Platinum (Pt) is commonly used as electrocatalyst in fuel cells, including DEFCs. However, ethanol oxidation on Pt is found to be incomplete and produces undesired partial oxidation products such as acetaldehyde and acetic acid. It is thereby considered that Pt itself cannot cleave the C-C bond in ethanol molecules and a second transition metal such as Rh and Ir is needed to facilitate this rate- and selectivity-limiting step. However, evidences are also present in the literature for the formation of strongly binding adsorbates, such as -CO and -CHx, at low potentials (<0.5 V vs. RHE), from both in situ molecular spectroscopy (e.g., Infrared (IR) and surface enhanced Raman spectroscopy (SERS)) and differential electrochemical mass spectroscopy (DEMS) studies. These results suggest that oxidative removal of these C1 species, instead of C-C bond cleavage, is the sole rate-limiting factor on Pt catalysts.

In attempt to resolve this controversy, we have performed systematic studies of methanol, ethanol and ethylene glycol oxidation on polycrystalline Pt electrodes. In particular, surface-specific sum frequency generation (SFG) spectroscopy is combined with electrochemistry to determine the reaction intermediates and evaluate their dependence on electrode potentials. It is found that C1 adsorbates form at low potentials, and the coverages increase as the electrode potential is raised from 0.1 V to ~0.4 V. While the -CO feature drops above 0.4 V for methanol and ethylene glycol oxidation, it persists up to 0.5 V and then drops for ethanol oxidation. Our results confirm the presence of active Pt sites for C-C bond cleavage and underline the role of -CHx species generated from the β-carbon in governing the kinetics of alcohol oxidation reactions. It is proposed that the slow transition from -CHx to -CO gives rise to the higher overpotentials for ethanol oxidation compared to methanol and ethylene glycol oxidation.

2435

, , , , , and

In this work, novel Pd-(metal oxide)/C electrocatalysts were synthesized and evaluated as anodes for the oxidation of ethanol, methanol, ethylene glycol, glycerol and isopropanol in alkaline electrolyte. The metal oxides proposed to enhance the catalytic activity of Pd nanoparticles were Fe3O4, Fe2O3 and cerium oxide nanorods (CeO2-NR) pyrolized at 200 and 400 °C. The Pd-(metal oxide)/C anodes, as well as a Pd/C electrocatalyst as comparison, were synthesized using NaBH4 as reducing agent and Vulcan XC-72 as the support. The electrochemical characterization revealed a higher performance of the Pd-(metal oxide)/C electrocatalysts compared to Pd/C for all the oxidation reactions, at several concentrations of the fuels. Among the novel electrocatalysts, that containing CeO2-NR thermally treated at 400 °C (Pd-CeO2-NR400/C) showed the highest catalytic activity. Only in the case of the Ethylene Glycol Oxidation Reaction, Pd-Fe2O3/C showed a performance as high as that of Pd-CeO2-NR400/C.

2436

, , , , and

It is of fundamental importance to understand the factors controlling trends in activity for electrocatalytic reactions as a function of pH. In the case of the oxygen reduction reaction, numerous reports suggest significant divergences between noble metals surface catalytic performances in acid and base.[1,2]

In our earlier studies, we mapped out the experimental Sabatier volcano for the oxygen reduction reaction in 0.1 M HClO4 using the Cu/Pt(111) near-surface alloy system, see Figure 1 for near-surface alloy schematic.[3,4]

In this study, as those of [3,4], we found that by changing the subsurface coverage of Cu we could tune the surface binding of the key reaction intermediate, OH; we thus monitored the OH binding energy shift through the observable shifts in the base voltammograms in both acidic and alkaline media.

Further, we elucidate the experimental oxygen reduction volcano in 0.1 M KOH for the Cu/Pt(111) near-surface alloy system. Remarkably, we observe that the same trend persists between OH binding shifts and Cu/Pt(111) oxygen reduction activities between acid and alkaline electrolyte, with the optimum catalyst in alkaline exhibiting an 8-fold improvement in activity, relative to Pt(111). However, all surfaces show a ~4 fold improvement in activity in 0.1 M KOH, relative to the same surface in 0.1 M HClO4. At the peak of the volcano the surface exhibits an exceptionally high specific activity of 90 mA/cm2 at 0.9 V with respect to the reversible hydrogen electrode. Thus, our results confirm that OH binding energy is the key descriptor in both alkaline and acid electrolytes.

[1] R. Rizo, E. Herrero, J. M. Feliu, Phys. Chem. Chem. Phys. 201315, 15416–25.

[2] J. Staszak-Jirkovský, R. Subbaraman, D. Strmcnik, K. L. Harrison, C. E. Diesendruck, R. Assary, O. Frank, L. Kobr, G. K. H. Wiberg, B. Genorio, et al., ACS Catal. 20155, 6600–6607.

[3] I. E. L. Stephens, A. S. Bondarenko, F. J. Perez-alonso, F. Calle-vallejo, L. Bech, T. P. Johansson, A. K. Jepsen, R. Frydendal, B. P. Knudsen, J. Rossmeisl, et al., 2011, 5485–5491.

[4] A. S. Bondarenko, I. E. L. Stephens, I. Chorkendorff, Electrochem. commun. 201223, 33–36.

Figure 1

2437

, , , and

a, Key Laboratory of Material Chemistry for Energy Conversion and Storage (Huazhong University of Science and Technology), Ministry of Education, Hubei Key Laboratory of Material Chemistry and Service Failure, School of Chemistry and Chemical Engineering, Huazhong University of Science and Technology, Wuhan, 430074, P.R. China. *Email: wangdl81125@hust.edu.cn

b, Center for Functional Nanomaterials, Brookhaven National Laboratory, Upton, NY, USA

Graphene, a single-atom-thick hexagonally arrayed sp2 carbon atoms bonded sheet, has attracted tremendous interest in renewable energy conversion and storage devices due to its extraordinary properties, especially ultrahigh specific surface area and predominant electron conductivity1. However, graphene based two-dimensional catalysts suffer from an irreversible re-stacking of the nanosheets due to the strong π-interaction during thermal annealing process and electrochemical measurements which would result in a lower surface area, limiting the mass transfer rate and even decreasing the catalytic activity as well as electrochemical stability.

Figure 1 (a) TEM image of NSGCB; (b) ORR activity of NSGCB and the corresponding comparative catalysts; (c) TEM image of NSCNT-3; (d) ORR activity of NSCNT-3 and the corresponding comparative catalysts.

Herein, we report two strategies for the synthesis of nitrogen and sulfur co-doped three-dimensional graphene intercalated structure nanomaterials as electro-catalysts for ORR in alkaline medium. One is the insertion of Vulcan XC-72 carbon spheres into the graphene layer, forming a graphene-Vulcan XC-72- grapheme "sandwich-like" structure2. The other way is partially unzipping of multi-walled carbon nanotubes to form a graphene nanoribbon-carbon nanotube-graphene nanoribbon structure. Both of the prepared three-dimensional structured nanocomposites were experienced to nitrogen and sulfur co-doping process, and moreover, the resulting doped electro-catalysts exhibited outstanding ORR performance.

Acknowledgements

This work was supported by the National Natural Science Foundation (21306060, 21573083), the Program for New Century Excellent Talents in Universities of China (NCET-13-0237), the Doctoral Fund of Ministry of Education of China (20130142120039).

References

(1) Li, Y., Zhao, Y., Cheng, H., Hu, Y., Shi, G., Dai, L., & Qu, L. Nitrogen-doped graphene quantum dots with oxygen-rich functional groups. Journal of the American Chemical Society, 2011, 134, 15-18.

(2) Wu, M., Wang, J., Wu, Z., Xin, H. L., & Wang, D. Synergistic enhancement of nitrogen and sulfur co-doped graphene with carbon nanosphere insertion for the electrocatalytic oxygen reduction reaction. Journal of Materials Chemistry A, 2015, 3, 7727-7731.

Figure 1

2438

, , , and

Carbon-based electrodes are known to undergo corrosion in acidic and alkaline media during the oxygen evolution reaction (OER), yet studies on the efficacy of OER catalysts frequently employ carbon as catalyst support. Carbon remains a material of interest in the development of bifunctional air electrodes, particularly due to its utility in supporting efficient oxygen reduction reaction (ORR) catalysis. Methods to quantify carbon corrosion rates are needed in order to assess the practical utility of carbonaceous materials in bifunctional air electrodes and other OER electrodes. In-situ electrochemical mass spectrometry enables direct and precise assessment of carbon corrosion in acidic media by quantifying gaseous CO2 that is generated as a product of the corrosion reaction. For alkaline media, this in-situ approach is unsuitable since CO2 generated during corrosion reacts with and is sequestered within the alkaline electrolyte as carbonate.

In this presentation, we describe a method to precisely quantify carbon corrosion in alkaline media. The method is direct and sufficiently sensitive to provide rapid (1-2 days) characterization of carbon corrosion, whereas methods based on electrochemical characterization or mass measurement may yield ambiguous results and take months. We report on the use of this method to quantitatively compare the rate of corrosion using different types of carbon, including reduced graphene oxide, doped carbon and carbons with various coatings and catalysts. A critical assessment of the use of carbon in bifunctional air electrodes will be provided.

D-11 Pt-alloy Cathode Catalysts 2 - Oct 4 2016 8:00AM

2439

, , and

In Pt-transition metal (TM) alloy nanoparticles for proton exchange membrane fuel cells (PEMFCs), the TM surface can be easily oxidized by air, water, and acidic electrolytes. Accordingly, the electron transfer from the TM to Pt is retarded owing to the electronegative O species on the TM surface, and the dissolution of the surface TM oxide is accelerated during the fuel cell operation, which results in catalyst degradation. Therefore, we propose a novel hybrid catalyst concept that selectively modifies the surface Co atoms with N-containing polymers, resulting in highly active and durable PtCo nanoparticles useful for the oxygen reduction reaction (ORR). The N-containing polymer functionalized on carbon black selectively interacts with the Co precursor, resulting in Co-N bond formation on the PtCo nanoparticle surface. Electron transfer from Co to Pt in the PtCo nanoparticles is therefore enhanced by the increase in the difference in electronegativity between Pt and Co. In addition, the dissolution of Co and Pt is prevented by the selective passivation of surface Co atoms and the decrease in the O binding energy of surface Pt atoms, respectively. As a result, the activity and durability of organic-inorganic hybrid PtCo nanocatalysts for the ORR are significantly enhanced by the electronic ensemble effects.

2440

and

Demand on the practical synthetic approach to the high performance electrocatalyst is rapidly increasing for the fuel cell commercialization. However, sluggish kinetics hindered high performance. Therefore methods to improve the activity of catalysts that facilitate reaction kinetic have been and continue to be a popular area of research. Since electrocatalytic reactions occur on the catalyst surface, nanoparticles (NPs) with high surface area are promising candidates to overcome problems. However, their low stability at harsh operating condition impedes nanoparticle utilization. Here we present a synthesis of highly durable and active nanoparticle based electrocatalyst for fuel cell electrocatalyst. One example is ordered intermetallic face-centered tetragonal (fct)-PtFe NPs encapsulated by thin-layer carbon shell. Only few nanometers of carbon shells which is in situ formed from dopamine coating could effectively prevent the coalescence of NPs. This carbon shell also protects the NPs from detachment and agglomeration as well as dissolution throughout the harsh fuel cell operating conditions. The other is carbon encapsulated metal phosphide NPs for robust hydrogen evolution reaction. Anti-oxidative carbon shell prevent surface oxidation of NPs, leading exceptionally long-term durability. We believe that thin-layer carbon shell encapsulation of NPs can open a new possibility for the development of highly active and durable electrocatalyst in near future.

2441

, , , , , and

The progress in the ORR electrocatalysis over the last two decades has been achieved primarily through the fundamental understanding of processes that are taking place at well-defined surfaces. For instance, the nature of active sites is a common topic in the literature, however, despite decades of persistent studies aimed to reveal the fundamentals, insight at atomic level is still lacking. Properties such as surface crystallographic orientation, morphology, composition and defects are yet to be assigned to the correlation between the atomic structure and catalyst activity. For the first time, we report on the atomic structure that has been investigated in combination with durability. The ultimate precision that goes beyond a part per million of a single atomic layer has been achieved in determining electrocatalytic properties of the ORR catalysts. Obtained knowledge is of paramount importance in design of advanced highly functional nanoscale materials. Surfaces of Pt-based materials have been characterized by AES, LEED and UPS, which was followed by controlled transfer to electrochemical, in-situ FTIR and STM cells. These findings have been further used to optimize a unique RDE coupled ICP-MS system. Such effort has led towards the design and synthesis of nanoscale materials with superior electrocatalytic properties. Fine tuning of the surface properties induced unprecedented improvements in functionality of real world catalyst for the ORR in fuel cells that will be reported for the first time.

2442

, , , , , and

In case of polymer electrolyte fuel cells (PEFCs) Platinum is still the most used catalyst for the oxygen reduction reaction (ORR) at the cathode – where PtNi alloys are the state of the art [1, 2]. Following a study by Stamenkovic et al. in 2007 [3] several groups dealt with octahedral shaped nanoparticles (NPs) during the last decade to enhance both specific and mass activity [4]. Well-defined octahedral shaped NPs can be obtained e.g. by solvothermal synthesis using DMF as solvent and reduction agent [5]. Unfortunately, such PtNi octahedra show a major disadvantage in lacking of morphological stability during electrochemical cycling [6]. In the present work, we combine the DMF-based synthesis with a thermal post-treatment as annealing step toward an improved PtNi alloy structure with higher stability. The study is primarily based on the morphological investigations by transmission electron microscopy (TEM) and scanning transmission electron microscopy (STEM) combined with energy dispersive X-ray analysis (EDX), both ex situ as well as in situ mode, to gain insights into the structure-activity-stability relationship.

In detail, based on a previous work [7], we exposed octahedral PtNi particles (supported on carbon) to a post-treatment in hydrogen using a tubular furnace. First off all, an increased annealing temperature results in a better defined PtNi alloy – proven by XRD. Whereas after 300°C mainly octahedral NPs were present, a complete loss of shape was observed after annealing at 500°C, where nearly spherical particles resulted. Additionally, the morphological changes during heating were also observed in situ using a TEM heating nano-chip (i.e. heating in vacuum). Both annealing in hydrogen as well as in vacuum show indications for Pt diffusion to (111) facets before the octahedral shape collapsed. Indeed, a rather mild treatment in hydrogen can improve the alloy structure and, thus, lead to higher activity in ORR, whereas a complete PtNi alloying at higher temperatures is diminishing the ORR activity due to the loss of octahedral shape and, accordingly, (111) facets.

References

[1] Y. Bing et al., Chem. Soc. Rev 39 (2010) 2184-2202.

[2] H. A. Gasteiger, N. M. Marković, Science 324 (2009) 3791.

[3] V. R. Stamenkovic et al., Science 315 (2007) 493.

[4] P. Strasser, Science349, 379-380 (2015).

[5] M. K. Carpenter et al., J. Am. Chem. Soc. 134 (2012) 8535-8542.

[6] L. Gan et al., Science 346 (6216) (2014) 1502-1506.

[7] M. Ahmadi et al., ACS Nano 7 (2013) 9195-9204.

2443

, , , , and

The oxygen reduction reaction (ORR) is a corner stone for energy conversion techniques such as fuel cells and metal-air batteries[1]. The use of Pt unfortunately generate a number of obstacles including the high cost, scarcity and readily poisoning by impurity and the cross-over of fuel9. Shaped bimetallic nanocrystals (NCs) with precisely controlled structure and composition at the atomic level, e.g. PtxNi1−x alloy NCs[2], intermetallic PtCo core–shell nanoparticles[3] and Pt3Ni nanoframes[4], represent a class of exciting and promising ORR catalysts.

Herein, we demonstrate a one-pot protocol for the controlled synthesis of icosahedral Pt3M nanocrystal with twenty active Pt3M (111) facets using glucosamine as both reducing and structure-directing agent. Among the investigated catalysts, the Pt3Ni icosahedral with nano-segregated Pt-skin displays the unexpected 32-fold enhancement in specific activity and 12-fold improvement in mass activity relative to state-of-the-art Pt/C catalyst. Besides, the catalyst was found ultra-robust during 20,000 potential cycling tests. The Pt3Ni icosahedral thus represents one class of the most efficient ORR electrocatalysts ever reported and demonstrates an effect technique in mimicing the active extended surface in nano-scale.

Fig.1 (a) TEM image of the typical icosahedral Pt3Ni nanocrystal, (b) HRTEM micrograph showing the twin boundaries of five-fold symmetry, (c) the (111) and (200) planes, (d)cyclic voltammetry curves of Pt/C, Pt NI/C and Pt3Ni/C catalysts, (e) ORR polarization curves for Pt/C, Pt NI/C and Pt3Ni/C catalysts in O2-saturated 0.1M HClO4, (f) comparative ORR activities of Pt/C and Pt3Ni/C catalysts before and after 20,000 potential cycles.

Acknowledgments

The study was financed by the National Basic Research Program of China (973 Program, 2012CB215500), the National Natural Science Foundation of China (21373199, 21433003) and the Strategic Priority Research Program of the Chinese Academy of Sciences (XDA09030104). 

References

[1] Stamenkovic, V.R., B. Fowler, B.S. Mun, G. Wang, P.N. Ross, C.A. Lucas, and N.M. Marković, Science, 315(5811), 493-497,(2007).

[2] Zhang, C., S.Y. Hwang, A. Trout, and Z. Peng, Journal of the American Chemical Society, 136(22), 7805-7808,(2014).

[3] Wang, D., H.L. Xin, R. Hovden, H. Wang, Y. Yu, D.A. Muller, F.J. DiSalvo, and H.D. Abruña, Nat Mater, 12(1), 81-87,(2013).

[4] Chen, C., Y. Kang, Z. Huo, Z. Zhu, W. Huang, H.L. Xin, J.D. Snyder, D. Li, J.A. Herron, M. Mavrikakis, M. Chi, K.L. More, Y. Li, N.M. Markovic, G.A. Somorjai, P. Yang, and V.R. Stamenkovic, Science, 343(6177), 1339-1343,(2014).

 

Figure 1

2444

, , , , and

Improving design and/or reducing noble metal content in electrocatalysts for fuel cell electrodes while maintaining and/or increasing proton exchange membrane fuel cell (PEMFC) performance in terms of durability and power density are crucial challenges for PEMFC mass market applications. A possible way consists in combining noble metals (Pt, Pd, Au ...) and non-noble metals for preparing binary and ternary nanocatalysts.

The formation of platinum alloys with transition metals such as Ni, Co, Fe, Cr represents a promising way for improving the activity of Pt-based catalysts [1-3].  Pt based alloys have indeed often demonstrated higher electrocatalytic activity towards ORR than pure platinum.

However, while non-noble metals are interesting from a cost reduction point of view, their presence may involve lower stability of the catalyst than that of pure platinum due to their dissolution [4.5], which leads to a loss of performances in PEMFC. Au addition was proposed to improve durability of the nanocatalysts [6-9].

On the other hand, the atomic structure and morphology [10] also play very important roles in the electrocatalytic efficiency for ORR.

In this context, the present study aims at systematically comparing the catalytic activity and the selectivity towards the ORR of binary and ternary catalysts based on Pt, Ni, Co and Au in order to determine the best atomic ratio for this reaction, in terms of kinetics current density and of number of exchanged electrons per oxygen molecule reduced. For this purpose, monometallic (Pt/C), binary (PtxNi10-x/C and PtxCo10-x/C) and ternary (PtxAuyNiz/C and PtxAuyCoz/C) nanocatalysts supported on carbon Vulcan XC-72 have been first synthesized by a wet chemical method and comprehensively characterized. The morphologies, compositions and structures of the particles were characterized by physical methods (transmission electron microscopy, X-ray diffraction and atomic absorption), whereas metal loadings on the carbon support was determined by thermogravimetric analysis. Electrochemical active surface areas and surface compositions were estimated by cyclic voltammetry. Electrocatalytic activity, selectivity and durability of catalysts were studied by the rotating disc (Figure 1) and rotating ring disc electrodes.

After having determined the best atomic ratios for the different metals, i. e. Pt7Ni3, Pt6Ni2Au2/C and Pt5Ni1.7Au3.3/C, plasma sputtering method was use to prepare alloyed PtxAuyNiz/C and Auy@PtxNiz core-shell electrodes with ultra-low metal loadings (10 µgmetal cm-2). This method is indeed very convenient for controlling the materials loading, composition and structure by controlling the physical parameters of deposition. The electrocatalytic behavior of these different electrode structures was evaluated and compared to that of chemical catalysts.

Acknowledgement:

The research leading to these results has received funding from the European Union's Seventh Framework Programme (FP7/2007-2013) for the Fuel Cells and Hydrogen Joint Technology Initiative under grant agreement #325327 (SMARTCat project). 

Reference

[1] M.T. Paffet, J.G. Beery, S. Gottesfeld, J. Electrochem. Soc. 135 (1988) 1431–1436.

[2] C. Wang, M. Chi, D. Li, D. van der Vliet, G. Wang, Q. Lin, J. F. Mitchell, K. L. More, N. M. Markovic, V. R. Stamenkovic, ACS Catal. 1 (2011) 1355–1359.

[3] T. Toda, H. Igarashi, H. Uchida, M. Watanabe, J. Electrochem. Soc. 146 (1999) 3750–3756.

[4] M. Watanabe, K. Tsurumi, T. Mizukami, T. Nakamura, P. Stonehart, J. Electrochem. Soc. 141 (1994) 2659–2668.

[5] C.F. Yu, S. Koh, J.E. Leisch, M.F. Toney, P. Strasser, Faraday Discuss. 140 (2009)

283–296.

[6] D.F. Yancey, E.V. Carino, R.M. Crooks, J. Am. Chem. Soc. 132 (2010) 10988–10989.

[7] J. Zhang, K. Sasaki, E. Sutter, R.R. Adzic, Science 315 (2007) 220–222.

[8] K. Sasaki, H. Naohara, Y. Cai, Y.M. Choi, P. Liu, M.B. Vukmirovic, J.X. Wang, R.R.

Adzic, Angew. Chem. Int. Ed. 49 (2010) 8602–8607.

[9] V. Tripkovic, H.A. Hansen, J. Rossmeisl, Tejs Vegge, Phys. Chem. Chem. Phys. 17 (2015) 11647–11657.

[10] D. Alloyeau, C. Mottet, C. Ricolleau, Nanoalloys, London: Springer-Verlag; 2012.

Figure 1

2445

, , , , , , and

The amount of platinum (Pt) in the catalyst layer accounts for a significant portion of fuel cell cost and limits the commercial deployment of proton exchange membrane fuel cells.[1] Catalyst studies typically focus on the oxygen reduction reaction (ORR), since the reaction is orders of magnitude slower kinetically than hydrogen oxidation. Extended surface nanomaterials offer key advantages in ORR, including an order of magnitude higher site-specific activity, long range conductivity, and long term durability.[2] Traditionally, however, the catalyst type has been limited by low surface areas.

Spontaneous galvanic displacement occurs when a more noble metal cation contacts a less noble metal template, and combines aspects of corrosion and electrodeposition. In ORR, catalysts formed by spontaneous galvanic displacement are ideally situated, taking advantage of the specific activities generally associated with the catalyst type while significantly improving upon their surface area.[3] Galvanic displacement has been used to deposit small amounts of Pt onto extended templates, and in the case of Pt-nickel (Ni) nanowires, has produced materials with surface areas in excess of 90 m2 gPt‒1.[4] Post-synthesis processing has been used to improve the activity and durability of Pt-Ni nanowires in rotating disk electrode (RDE) half-cells. Thermal annealing integrated Pt-rich and Ni-rich zones, compressing the Pt lattice and improving ORR activity.[5] Acid leaching and oxidation of the Pt-Ni nanowires also removed surface Ni and formed Ni oxides near the nanowire surface, improving catalyst durability and reducing Ni dissolution in durability tests (30,000 cycles, 0.6‒1.0 V).

Recently, Pt-Ni nanowires were studied for their ORR activity in RDE half-cells using updated testing protocols. A rotational air drying method formed thin, uniform coatings onto the RDE working electrodes.[6] Thin catalyst ink dispersions also allowed for the ORR diffusion-limited current to be reached while maintaining a low Pt loading. Using these methods improved Pt-Ni nanowire activity in RDE half-cells by 60%. The optimized Pt-Ni nanowires produced an ORR mass activity 9 times greater than Pt/HSC, while losing less than 3% of that activity and less than 0.5% catalyst mass to dissolution in durability testing.

Atomic layer deposition (ALD) has also been used in an effort to replicate the Pt-Ni nanowire synthesis by galvanic displacement. Works has included different chemistries (oxygen and hydrogen routes), as well as efforts to disperse the nanowire template without the benefit of a liquid medium. Successful use of ALD can potentially produce catalysts on a much larger, more commercially viable, scale.

Figure 1. Surface areas (x-axis) and site-specific oxygen reduction activities at 0.95 V (y-axis) of Pt/HSC and Pt-Ni nanowires, as-synthesized and modified by post-synthesis processing. The solid black line denote constant mass activities of 100, 400, and 800 mA mgPt1.

[1] D. Papageorgopoulos, in: U.S. Department of Energy (Ed.), http://www.hydrogen.energy.gov/pdfs/review14/fc000_papageorgopoulos_2014_o.pdf, 2014.

[2] M. Debe, in: U.S. Department of Energy (Ed.), http://www.hydrogen.energy.gov/pdfs/review09/fc_17_debe.pdf, 2009.

[3] S.M. Alia, Y.S. Yan, B.S. Pivovar, Catalysis Science & Technology, 4 (2014) 3589-3600.

[4] S.M. Alia, B.A. Larsen, S. Pylypenko, D.A. Cullen, D.R. Diercks, K.C. Neyerlin, S.S. Kocha, B.S. Pivovar, ACS Catalysis, 4 (2014) 1114-1119.

[5] B. Pivovar, Extended, Continuous Pt Nanostructures in Thick, Dispersed Electrodes, in: U.S. Department of Energy (Ed.), http://www.hydrogen.energy.gov/pdfs/review14/fc007_pivovar_2014_o.pdf, 2014.

[6] K. Shinozaki, J.W. Zack, R.M. Richards, B.S. Pivovar, S.S. Kocha, Journal of The Electrochemical Society, 162 (2015) F1144-F1158.

Figure 1

2446

, , , , and

State-of-the-art polymer electrolyte fuel cells (PEFCs) require large amounts of carbon-supported platinum nanoparticle (Pt/C) catalysts (~ 0.4 mgPt/cm2geometric) to account for the large overpotential of the oxygen reduction reaction (ORR).1 These excessive Pt-loadings that impede widespread commercialization of PEFCs can be mitigated by increasing the catalysts' ORR activity, e.g. by alloying platinum with other metals like Ni, Cu and Co, to form materials which show up to one order of magnitude higher mass-specific activity than commercial Pt/C catalysts.2 On the other hand, state-of-the art carbon-supported materials suffer from significant carbon and Pt corrosion during the normal operation of PEFCs, gradually compromising their performance. To partially overcome these stability issues, a lot of research effort is dedicated to the development of unsupported ORR catalysts. Among these materials, bimetallic alloy aerogels consisting of nanoparticles interconnected to nanochains3 present an interesting option, since their extended 3D structure should facilitate transfer to actual PEFC cathodes.

With this motivation in mind, we have modified a previously published synthesis3 in aqueous environment to prepare 'clean' bimetallic Pt-Ni aerogels with different stoichiometries, among which Pt3Ni and Pt1.5Ni were characterized in depth and tested for ORR activity. The alloy formation and extended 3D nanochain structure were confirmed by X-ray diffraction and transmission electron miscroscopy, respectively. Moreover, X-ray photoelectron and X-ray absorption spectroscopy indicated the existence of a separate Ni-(hydr)oxide phase in the Pt1.5Ni aerogel that was not present in the Pt3Ni sample. This hypothesis was further verified by complementing electrochemical measurements in alkaline electrolyte. Despite this difference in composition, both Pt-Ni aerogels showed comparable ORR activities in rotating disk electrode measurements in 0.1 M HClO4 electrolyte, reaching the activity target of 440 A/gPt at 0.9 VRHE set by the U.S. Department of Energy4 for automotive PEFCs. However, accelerated stress tests5 using the above mentioned RDE setup did not reveal significant durability improvements with respect to a commercial Pt/C benchmark catalyst. Consequently, these results were compared to those obtained in a differential PEFC setup to better understand the degradation behavior in an environment closer to real application.

In summary, this contribution reports the detailed investigation of surface and bulk properties of bimetallic Pt-Ni aerogels alongside with performance and stability assessments in aqueous electrolyte and a differential PEFC setup.

Acknowledgement

Funding from the Swiss National Science Foundation (20001E_151122/1), the German Research Foundation (EY 16/18-1) and the European Research Council (ERC AdG 2013 AEROCAT) is greatly acknowledged.

References

1. F. T. Wagner, B. Lakshmanan and M. F. Mathias, J. Phys. Chem. Lett., 1, 2204 (2010).

2. C. Wang, M. Chi, D. Li, D. Strmcnik, D. van der Vliet, G. Wang, V. Komanicky, K. C. Chang, A. P. Paulikas, D. Tripkovic, J. Pearson, K. L. More, N. M. Markovic and V. R. Stamenkovic, J. Am. Chem. Soc., 133, 14396 (2011).

3.  W. Liu, P. Rodriguez, L. Borchardt, A. Foelske, J. Yuan, A. K. Herrmann, D. Geiger, Z. Zheng, S. Kaskel, N. Gaponik, R. Kotz, T. J. Schmidt and A. Eychmüller, Angew. Chem. Int. Ed., 52, 9849 (2013).

4. D. Papageorgopoulos, http://www.hydrogen.energy.gov/pdfs/review15/fc000_papageorgopoulos_2015_o.pdf (accessed 09.03.2016).

5. F. Hasché, M. Oezaslan and P. Strasser, ChemCatChem, 3, 1805 (2011).

2447

, , , , , , , and

Platinum-coated nickel nanowire (PtNiNW) catalysts have shown high mass activity, specific activity, and exceptionally high surface areas in oxygen-reduction reaction (ORR) rotating disc-electrode experiments, all several times higher than platinum on carbon and well above DOE targets.1,2 However, their subsequent incorporation into proton exchange membrane fuel cell (PEMFC) membrane electrode assemblies (MEA) has been exceedingly difficult, resulting in substantially lower mass and specific activities and lower than anticipated electrochemical areas (ECAs).2 This presentation will focus on the challenges associated with the incorporation of extended surface Pt alloys in PEMFC MEAs, the ensuing conditioning protocols and the large discrepancy between RDE and MEA data. It is found that soaking the MEA in dilute sulfuric acid to remove nickel and increase the platinum weight fraction increases cell performance metrics such as open-circuit voltage, mass activity, and specific activity, as shown in Figure 1. Decreases in membrane resistance are also observed. The effects of ionomer, carbon and unsupported electrocatalyst ratios on MEA performance will also be discussed. Scanning transmission electron microscopy (STEM) and transmission x-ray microscopy (TXM) tomography will be examined as tools to elucidate the relationship between electrocatalyst/electrode structure and MEA performance.

1. Alia, S. M. et al. Platinum-Coated Nickel Nanowires as Oxygen-Reducing Electrocatalysts. ACS Catal.4,1114–1119 (2014).

2. Pivovar, B. S. Extended, Continuous Pt Nanostructures in Thick, Dispersed Electrodes. 2015 DOE Hydrogen and Fuel Cells Program Review (2015).

Figure 1

2448

, , , , , , , and

Proton Exchange Membrane Fuel Cell (PEMFC) technologies have focused much attention over the last decades as energy sources for both stationary and transportation applications. Their electrical performances have now reached a sufficient level for the deployment in specific markets. However, higher power densities are needed to decrease the size of the stack, and to reduce the total cost and/or to take into account the high degradation of the electrical performance during operation [1]. PEMFC performances are usually ruled out by the cathode side where the sluggish Oxygen Reduction Reaction (ORR) is processed. Moreover, the stability of the nanomaterials is also a concern in the harsh operating conditions of a PEMFC cathode [1].

The formation of hollow Pt/C particles was reported after degradation of Pt3Co/C alloys in PEMFC on-site operation [2]. Having unveiled the enhanced ORR activity of such nanostructures, LEPMI recently developed a simple "one pot" method to synthesize hollow PtNi nanoparticles supported on high surface area carbon (PtNi/C). The best porous hollow PtNi/C nanocatalysts achieved 6-fold and 9-fold enhancement in mass and specific activity for the ORR, respectively over standard solid Pt/C nanocrystallites of the same size [3] (liquid electrolyte).

Upscaling the synthesis process is now a key-step to integrate these hollow nanoparticles into Membrane Electrode Assemblies (MEAs). Step-by-step development of the synthesis enabled to reach a 10-fold upscale of the synthesis corresponding to ca. 4 g of catalyst per batch (Figure 1). Combined physical and electrochemical characterizations were performed to ensure that the initial structure and the intrinsic properties of the hollow PtNi/C nanoparticles were maintained (Figure2). In parallel, different experimental conditions were also experimented to optimize the overall manufacturing process. This material upscaling also rendered possible the use of classical electrodes manufacturing processes for MEAs such as bar coating and screen printing, similarly to commercial catalysts. The first measurements performed in 25 cm² single cells demonstrated promising results for MEA using hollow PtNi/C at the cathode (Figure 3). Moreover, after an accelerated stress test designed to test the robustness of the metal nanoparticles [4], the MEA performance became higher than that of the Pt/C-based reference MEA. This is particularly true at high efficiency (up to medium current densities), which tends to confirm the interest of such nanostructures on the long-term. Complementary characterizations revealed that the utilization factor of the hollow catalyst is only 50%, and can be improved by optimizing the cathode catalyst layer formulation before integrating these new materials in larger MEAs and finally into representative PEMFC stacks.

References :

[1]: L. Dubau, L. Castanheira, F. Maillard, M. Chatenet, O. Lottin, G. Maranzana, J. Dillet, A. Lamibrac, J.-C. Perrin, E. Moukheiber, A. Elkaddouri, G. De Moor, C. Bas, L. Flandin, N. Caqué., Wiley Interdisciplinary Reviews: Energy and Environment,2014, 3, 540-560.

[2]: L. Dubau, J. Durst, F. Maillard, L. Guétaz, M. Chatenet, J. André, E. Rossinot, Electrochim. Acta 2011, 56, 10658-10667.

[3]: L. Dubau, T. Asset, R. Chattot, C. Bonnaud, V. Vanpeene, J. Nelayah, F. Maillard, ACS Catal. 2015, 5, 5333-5341

[4]: L. Castanheira, W.O. Silva, F.H. Lima, A. Crisci, L. Dubau, F. Maillard, ACS Catal. 2015, 5, 2184-2194

Figure 1

2449

, , , and

Proton exchange membrane fuel cells (PEMFCs) are characterized with many advantages including high power density, low operating temperature (~80 oC), quick start-up and quick match to shifting demands for power, and regarded as a very promising technology for powering transportation and stationary applications1. However, the kinetics of the cathodic oxygen reduction reaction (ORR) in PEMFCs is sluggish and the most efficient material is platinum (Pt) that is rare on the earth and has a prohibitive cost. To make PEMFCs economically available, it is imperative to maximize the utilization of Pt or rather noble metal by improving catalyst layer's mass transport of reactants and products, and by increasing noble-metal-mass activities toward the ORR in cathode.

One of the strategies to improve mass transport in cathode is to thin the catalyst layer through developing high-content and high-dispersion Pt electrocatalysts. However, it is difficult to prepare Pt supported catalysts with a high Pt content because of the agglomeration of Pt nanoparticles. It was reported that the commercial Pt/VC catalysts had a particle size of 2.0 nm in Pt(10 wt%)/VC, 3.2 nm in Pt(30 wt%)/VC, and 8.8 nm in Pt(60 wt%)/VC2. Although great efforts have been made to improve traditional methods, such as impregnation method and colloidal method, it remains very challenging to find an economical and controllable way to synthesize Pt/VC electrocatalysts with a high Pt content for the ORR in PEMFCs. This can be attributed to several facts, including the poor size control, the contaminants from stabilizer residue, tedious procedures and so on. In this work, we propose a simple and controllable approach to deposit highly-dispersed Pt nanoparticles on Vulcan XC-72 carbon black. The as-synthesized Pt/VC possesses a high Pt content as well as high-dispersion Pt NPs via controlling the nucleation and growth process, and was investigated as the cathodic catalyst in PEMFC.

Ideally highest Pt utilization with an extraordinary Pt-mass activity has been achieved by the Pt-monolayer-shell electrocatalysts which possess monoatomic thick Pt shells on noble-metal substrates3. However, the noble-metal-mass activities for these Pt-monolayer-shell electrocatalysts were also compromised, since the noble-metal substrates are quite essential for fabricating such electrocatalysts. In this regard, with the highest Pt utilization, improving the non-Pt noble-metal utilization in the substrates is critical for developing the Pt-monolayer-shell electrocatalysts. In this work, a surfactant-based, composition- and size-tunable method has been demonstrated to synthesize the monodispersed Pd-Ni as the substrates for fabricating the carbon-supported Pd-Ni@Pt nanospheres with Pt monolayer shells. And an excellent balance among activity, durability and noble-metal utilization was acquired.

Acknowledgements

This work was supported in part by National Natural Science Foundation of China (Grant No. 21373135 and 21533005) and Science Foundation of Ministry of Education of China ( Grant No. 413064).

Reference

1. M.L. Perry, F.T. Fuller, J Electrochem Soc, 149 S59-S67 (2002).

2. B. L. Gratiet, H. Remita, G. Picq, M. Delcourt, J Catal, 164, 36 (1996).

3. (a) J. Zhang, M. B. Vukmirovic, Y. Xu, M. Mavrikakis, R. R. Adzic, Angewandte Chemie International Edition, 44 2132-2135, (2005); (b) M. Shao, K. Shoemaker, A. Peles, K. Kaneko, L. Protsailo, J Am Chem Soc, 132 9253-9255, (2010). (c) J. X. Wang, H. Inada, L. Wu, Y. Zhu, Y. Choi, P. Liu, W. P. Zhou, R. R. Adzic, J Am Chem Soc, 131,17298( 2009) .

B-21 Fuel Cell Systems - Oct 4 2016 8:10AM

2450

, , , , , and

High-temperature polymer electrolyte fuel cells (HT-PEFCs) are a suitable technology for decentralized small scale electricity and heat production. HT-PEFCs do not require hydrogen infrastructure and are characterized by a simpler and therefore less expensive system compared to available fuel cell systems. In combination with a reformer unit, HT-PEFC systems offer an efficiency gain over conventional combustion of hydrocarbons, such as natural gas. In near future, HT-PEFCs will be even more effective and efficient when renewable biofuels and hydrogen are widespread available.

We present an efficient HT-PEFC based combined heat and power (µ-CHP) system for the provision of electrical energy and hot water in single family households (see Table 1). Due to the optimized design and layout of the fuel cell based µ-CHP system and the respective manufacturing processes of the catalysts and the membrane electrode assemblies (MEA), the demand for cost effective and greenhouse gas efficient energy at customer level is addressed.

Major efforts were devoted to the establishment of scalable catalyst deposition methods, which enable a loss-free utilization of precious metals. By using appropriate multimetallic catalyst systems at anode and cathode, the precious metal loading was reduced by approx. 20% in comparison to commercially available electrodes without compromising performance (see Figure 1) [1]. Furthermore, by introducing post-preparation treatments, the stability of the catalysts was enhanced over commercial Pt/C. The activity and stability of the catalyst systems were evaluated ex situ by means of cyclic voltammetry and accelerated stress tests using a rotating disk electrode (RDE) setup. Furthermore, the catalyst systems were characterized in situ by means of polarization curves, continuous operation, accelerated stress tests and electrochemical impedance spectroscopy measurements at single cell, stack and at system level (see Figure 1).

Acknowledgment

Financial support was provided by The Climate and Energy Fund of the Austrian Federal Government and The Austrian Research Promotion Agency (FFG) through the program Energieforschung (e!Mission).

[1] A. Schenk, C. Grimmer, M. Perchthaler, S. Weinberger, B. Pichler, C. Heinzl, C. Scheu, F.-A. Mautner, B. Bitschnau, V. Hacker, Platinum–cobalt catalysts for the oxygen reduction reaction in high temperature proton exchange membrane fuel cells – Long term behavior under ex-situ and in-situ conditions, J. Power Sources. 266 (2014) 313–322.

Figure 1

2451

and

A hydrogen/bromine regenerative fuel cell involves complex water and heat transport phenomena during charge and discharge that significantly affect the degree of electrolyte dehydration, electrochemical reactions, proton transport, the concentrations of bromide complexes at the hydrogen electrode, etc. Therefore, developing innovative water and heat management schemes is critical to improve the uniformity of key distributions inside a cell as well as to enhance cell performance and durability. In this work, a three-dimensional, transient, two-phase, non-isothermal hydrogen/bromine fuel cell model is developed by rigorously accounting for two-phase transport and various heat generation mechanisms including irreversible heat, entropic heat, and Joule heating. The model is applied to a 25 cm2 cell in order to precisely investigate water transport and thermal aspects under various charge and discharge conditions. A parallel computing methodology is employed to handle large-scale simulations involving millions of grid points. The large-scale simulations are able to provide extensive multidimensional contours of species concentration, temperature, and current density, assisting in identifying optimal water and thermal management strategies based on a typical geometry of a real-scale hydrogen/bromine cell.

2452

, , , and

One of the most important degradation mechanisms, which cause a reduction of lifetime and performance of high temperature (HT) polymer electrolyte membrane (PEM) fuel cells, is phosphoric acid loss. This paper presents the effect of different operation strategies on acid loss of HT-PEM FCs compared with fresh membrane electrode assemblies (MEA).

These investigations have been executed during the run-time of the European Project CISTEM (Construction of Improved HT-PEM MEAs and Stacks for Long Term Stable Modular CHP Units, GA-No. 325262). The vision of this project is the development of a new HT-PEM FC based CHP (combined heat and power) technology with high efficiency and long lifetime.

Degradation investigations of single components, MEAs, FC stacks and complete CHP units are main objectives within the project. With this paper we provide a more detailed insight to the phosphoric acid loss driven degradation process which causes shorter lifetimes and lower performances [1-2].

In dependence on the typical operational occurrences of the HT-PEM CHP-systems, the following operational strategies have been tested:

  • Fuel switching between pure H2 and synthetic reformate (Test 1 - grey)

  • Start-stop-cycling with an idling temperature of 25 °C (Test 2 - red)

  • Long term tests at constant current density (0.3 A/cm²) with different reactant gas compositions:

    • Synthetic reformate/pure O2 (Test 3 - blue)

    • Synthetic wet reformate/O2 enriched air (Test 4 - black)

All tests have been performed with Dapozol®-G55 MEAs, which consist of thermally cured polybenzimidazole (PBI) membranes and Pt/C based electrodes. During operation, the MEAs have been electrochemically characterized in-situ at Beginning of Life (BoL), once per week and End of Test (EoT) via polarization curves, electrochemical impedance spectroscopy (EIS), cyclic voltammetry (CV) and linear sweep voltammetry (LSV).

In addition, product water samples have been weekly collected from both cathode and anode gas outlets, and their acid content has been determined with ion chromatography (IC). The results of the total acid leaching in product water for every operation strategy are shown in Figure 1. After EoT, the remaining H3PO4within the tested MEAs was identified by titration with sodium hydroxide [3].

By ex-situ ante- and post-mortem micro-computed tomography (µ-CT) investigations, mechanical transformations can be visualized and the thickness changes of all layers can be verified.

Fresh MEAs are characterized by mirror symmetry through all layers, while MEAs, operated under dry reactant gases, disclose a reduction of catalyst layer thicknesses on anode sides. The operation with pure O2results into increased water production, which therefore causes higher acid-leaching. The MEA tested with synthetic reformate and pure oxygen has revealed the highest phosphoric acid loss (Figure 1) on one hand and lowest degradation rates (-6.4 µV/h) on the other hand.

Figure 1: Total H3PO4 loss in product water after EoT and degradation rates (DR) of each operation strategy

References:

1. S. Yu, L. Xiao, B.C. Benicewicz, Fuel Cells, 08, 3-4, 165-174 (2008).

2. Y. Zhai, H. Zhang, G. Liu, J. Hu and B. Yi, J. Electrochem. Soc., 154, B72 (2007).

3. N. Pilinski, M. Rastedt, P. Wagner, ECS Trans., 69, 323-335 (2015)

Figure 1

2453

, , and

Alkaline anion exchange membrane fuel cells have become a topic of substantial interest in recent years, opening up a new electrochemical environment for hydrogen fuel cells. Facile kinetics for the oxygen reduction reaction open the promise of non-PGM or even non-metal AEMFCs, and radical stability. The most challenging aspect for the field, as defined by the 2016 DOE AMFC III Workshop, is membrane and ionomer stability in alkaline conditions at elevated operating temperatures. Few papers report endurance data, and best practices for in situ fuel cell conditioning and electrochemical characterization are still being developed by the community.

Here, we report that HMT-PMBI exhibits membrane and ionomer stability in situ as AEMFCs in relevant conditions for device operation. These fuel cells demonstrate re-equilibration from extensive carbonation and complete re-conditioning in a shut-down / start-up cycle. We further report operation in typically challenging conditions, e.g. increased temperature and reduced humidity. Finally, we report on our attempts to define best-practices for electrochemical characterization.

2454

, , and

With the advancement of portable technology, suitable power sources must be developed. Alternatives to standard battery technology, such as fuel cells, have shown promise; however such fuel cells are in their infancy with regards to industrial and consumer adoption [1-3]. In order for fuel cells to become a feasible alternative to traditional dry-cell alkaline and lithium-ion batteries, their performance must be optimized for their intended application via manipulation of parameters such as fuel molarity, flow rate, temperature, and electrode geometry. Here, we present a mathematical model of a microscale direct methanol fuel cell (DMFC), validated with experimental data. In the past, an iterative fabrication approach was undertaken to optimize DMFC performance, involving the fabrication and testing of many individual components to assess the effects of individual parameters. This model provides a means to automate the design process and remove the fabrication requirement.

Prior parametric studies by Thorson et al. [4] show that shorter and wider electrodes yield higher current densities, however a recent review by Goulet et al. [5] highlights the need for a combined modeling and experimental approach to evaluating electrode design in a microscale fuel cell. Here we present a mathematical model to analyze the fuel flow pattern and performance of a direct methanol microscale fuel cell as a function of electrode geometry. Localized current densities are calculated over the electrode surface to elucidate the differences between electrode geometries. In the present work, the dimensions of the fuel channel are constant (0.5 cm (L) x 1.0 cm (W) x 0.125 cm (H)), while the length and width of the catalyst deposition region are varied. The model is validated with experimental data taken as a function of electrode geometry, fuel concentration, fuel temperature, and fuel flow rate. The performance of our experimental fuel cell is consistent with our modeling studies, achieving a maximum power density greater than 25 mW/cm2 at room temperature with 1 M methanol. The model presented here, in conjunction with the supporting experimental data, expands prior parametric studies [4] to predict optimal electrode design with regard to methanol concentration, temperature, and flow rate.

[1] C.K. Dyer, Fuel Cells Bulletin, 2002 (2002) 8-9.

[2] E. Kjeang, N. Djilali, D. Sinton, Journal of Power Sources, 186 (2009) 353-369.

[3] A.S. Hollinger, P.J.A. Kenis, Journal of Power Sources, 240 (2013) 486-493.

[4] M.R. Thorson, F.R. Brushett, C.J. Timberg, P.J.A. Kenis, Journal of Power Sources, 218 (2012) 28-33.

[5] M.A. Goulet, E. Kjeang, Journal of Power Sources, 260 (2014) 186-196.

Figure 1

A-11 Cell Performance - Oct 4 2016 8:40AM

2455

, and

An alternative approach to the rotating disk electrode (RDE) for characterising fuel cell electrocatalysts has been recently demonstrated1. The approach combines high mass transport with a flat, uniform, and homogeneous catalyst deposition process, well suited for studying intrinsic catalyst properties at realistic operating conditions of a polymer electrolyte fuel cell (PEFC). Uniform catalyst layers were produced with loadings as low as 0.16 μgPt cm-2 and thicknesses as low as 200 nm. Such ultra thin catalyst layers are considered advantageous to minimize internal resistances and mass transport limitations leading to very high performance at low platinum loadings. Modelling of the associated diffusion field suggests that such high performance is enabled by fast lateral diffusion within the electrode. The electrodes operate over a wide potential range with insignificant mass transport losses, allowing the study of the hydrogen oxidation reaction (HOR) and hydrogen evolution reaction (HER) at high overpotentials. For the HOR, geometric current densities as high as 5.7 A cm-2Geo were experimentally achieved at a loading of 10.15 μgPt cm-2 at room temperature (561 A mgPt-1), which is three orders of magnitude higher than current densities achievable with the RDE. For the HER specific current densities greater than 5 A cm-2Specific  have been achieved. The latter corresponds to a turnover frequency of almost 20,000 hydrogen molecules per surface platinum site per second.

We have taken these results and applied a new microkinetic model for hydrogen evolution/oxidation based around the classical Heyrovsky-Tafel-Volmer mechanistic steps. This new mechanistic formalism is easy to implement and provides insights into the HOR/HER. For each of the different mechanisms, an "atlas" of Hads coverage with overpotential and corresponding current density is provided, allowing an understanding of all possible responses depending on the dimensionless parameters. Analysis of these mechanisms provides the limiting reaction orders of the exchange current density for protons and bimolecular hydrogen for each of the different mechanisms, as well as the possible Tafel slopes as a function of the molecular symmetry factor, b. Only the HV mechanism is influenced by pH whereas the TV,HT, and HTV mechanisms are not. The cases where the equations simplify to limiting forms are discussed. Analysis of the exchange current density from experimental data is discussed, and it is shown that fitting the linear region around the equilibrium potential underestimates the true exchange current density for all of the mechanisms studied. Furthermore, estimates of exchange current density via back-extrapolation from large overpotentials is also shown to be highly inaccurate. Analysis of Tafel slopes is discussed along with the mechanistic information which can and cannot be determined.

Finally, some interesting effects are discussed which show that the HER can be accelerated under certain conditions.

2456

Ohmic shorting through the membrane has been identified as one of the major failure modes in polymer electrolyte fuel cells (PEFC) [1]. Shorting occurs when electrons flow directly from the anode to the cathode instead of through the device being powered. Ohmic shorting not only can reduce the fuel cell performance, but also can lead to local heat generation in the vicinity of the short, causing membrane damage that can ultimately result in gas crossover failure. Two types of membrane shorts have been observed during the fuel cell testing: soft shorts and hard shorts. A soft short is a sub-critical short that results from the conductive carbon fibers or debris penetrating through the membrane. The soft shorts do not immediately lead to fuel cell failure; however, significant accumulation of soft shorts can reduce the overall cell resistance and compromise fuel cell durability through cell voltage degradation. A hard short is a critical short that is the result of significant heat release from an existing soft short. A hard short can directly lead to membrane crossover and cell failure. Hard shorts occur suddenly in an operating fuel cell stack where one cell develops a cell voltage reversal well below -1 V [2].

The present study is composed of two parts. The first is on the development of a test method to determine the density of soft shorts and their ohmic resistance that may lead to hard shorts when the adverse fuel cell operating conditions are met. The second part of the study is to investigate these adverse fuel cell operating conditions that lead to cell reversal and the potential hard shorts. The ultimate objective of the study is to prevent the fuel cell shorting failure through better material selection/designs and fuel cell operation controls.

In the first part of this study, we use a current distribution circuit board [3] in a fixture that induces uniform compression over a 32 cm2 area of a sample that consists of a piece of proton exchange membrane or a membrane electrode assembly sandwiched between two gas diffusion layers. By incrementally increasing the compressive pressure and subjecting the sample to a voltage of 0.6 V, we can measure increasing shorting currents in some of the 0.5 cm2 current distribution segments. By de-convoluting the current density distribution measured in this experiment, the number and severity of each short can be identified and the robustness of various material sets against soft shorts can be compared.

In the second part of the study, we stepwise dried out a single large-active-area fuel cell by decreasing the inlet gas dew points while under galvanostatic control. At a critical dew point, the cell potential dropped rapidly to reach a negative voltage that was limited to -1 V. Tests were conducted at various cell current densities and inlet temperatures, which showed that the corresponding cell resistance at the onset of rapid cell voltage drop decreases as the current density increases or as the temperature decreases. Good correlation is found between the cell resistance and the onset of cell reversal at various current densities. When the cell reversal limit of -1V was removed, we found that hard shorts can develop at a cell voltage around -3V, confirming the importance of operating the fuel cell within the operating window developed in this study.

  • A.B. LaConti, M. Hamdan, R.C. McDonald, Mechanism of Membrane Degradation for PEMFCs, in Mechanisms of Membrane Degradation for PEMFCs, in: W. Vielstich, A. Lamn, H.A. Gasteiger (Eds), Handbook of Fuel Cells: Fundamentals, Technology and Applications, Vol. 3, John Wiley & Sons, New York, 2003, pp. 647–662.

  • Gittleman, C. S., Coms, F. D., and Lai, Y. H., "Chapter 2: Membrane Durability: Physical and Chemical Degradation", Polymer Electrolyte Fuel Cell Degradation, Editors: Matthew M. Mench, E. Caglan Kumbur, T. Nejat Veziroglu, Elsevier (2012).

  • J.J. Gagliardo, J.P. Owejan, T.A.Trabold, and T.W.Tighe, Nucl. Instrum. Meth. A, 605 (2009) 115-118.

2457

, and

The successful commercialization of Polymer Electrolyte Fuel Cells (PEFCs) system requires highly active oxygen-reduction reaction (ORR) electrocatalysts to meet the cost, performance, and durability targets for automotive applications. Although significant progress has been made in the past decade, Pt/C catalyst has the limitation of high material cost and insufficient ORR activity. Platinum-nickel alloy catalysts with core-shell structure have been demonstrated by multiple research groups to have higher intrinsic ORR activity, which can exceed the U.S. DOE 2015 ORR activity target, and also have promising stability in PEFC conditions [1, 2]. Additionally, Membrane Electrode Assemblies (MEAs) with low loadings of highly active Pt-group metal (PGM) catalysts have already been demonstrated to exceed the high-efficiency target of > 0.3 A/cm2 at 0.8 V [3]; however, the rated-power target of 1 W/cm2 cannot be met with these MEAs due to transport losses that are unique to MEAs with ultra-low catalyst loadings [4, 5].

United Technologies Research Center (UTRC) is a part of a DOE-supported research project, led by Argonne National Laboratory (ANL), focused on improving the understanding, performance, and durability of advanced Pt-alloy catalysts operating in complete PEFCs. In the current study, advanced MEAs with high ORR activity Pt-Ni alloy catalysts and ultra-low Pt loading (total ~ 0.13 mg-Pt/cm2) are prepared by Johnson Matthey Fuel Cells (JMFC) and tested by UTRC.

The focus of this talk will be on using a variety of in-cell diagnostics and analysis methods to investigate both the initial performance and durability of these state-of-the-art MEAs. One example diagnostic that will be utilized are Polarization Change Curves (PCCs), which were originally developed for durability studies [6], but can also be used to investigate changes in performance due to operating conditions or MEA composition. For example, Fig 1a is a PCC comparison of the initial performance of the PtNi MEA and a Pt-only MEA, which have similar average catalyst particle size and the same carbon support, metal/carbon ratio, and ionomer/carbon ratio. Clearly, in addition to the differences in catalyst activity, there are also differences in the transport losses between these two different electrocatalysts. The durability of ultra-low Pt loading MEAs will also be discussed. Fig. 1b is a durability example of a PCC result, which is after applying a recent DOE-suggested AST protocol (0.6V-0.95V trapezoid potential cycle, with 700mV/s ramp rate and 6s/cycle). Other diagnostics and analysis beyond PCCs will also be included. The results obtained from these in-cell diagnostics will be used to hypothesize the potential mechanisms responsible for the differences in polarization curves, as well as project possible pathways to future performance and durability improvements.

Acknowledgements

This work is partially funded by the U. S. Department of Energy (DOE) under contract number DE-AC02-06CH11357. The authors would like to thank their DOE-project colleagues, especially those at ANL and JMFC.

References

1. V. R. Stamenkovic, et al. Science, 315, 493 (2007).

2. B. Han, et al, Energy Environ. Sci., 8, 258, (2015)

3. D. Myers, et. al., DOE AMR, ID# FC106 (2015).

4. W. Yoon & A. Z. Weber, J. Electrochem. Soc., 158, B1007 (2011).

5. A. Kongkanand, et al., ACS Catalysis, 6, 1578 (2016).

6. R. Darling, et al., Modern Topics in Polymer Electrolyte Fuel Cell Degradation, Chap. 7, M. Mench, et.al., Ed.; Elsevier, Denmark (2011).

Figure 1

2458

, , and

It is important to ensure that the water content of a catalyst coated membrane (CCM) has equilibrated before it is used in the manufacturing of a proton exchange membrane (PEM) fuel cell stack. Equilibration is achieved when the water content in the CCM reaches equilibrium with controlled relative humidity (RH) and temperature conditions. A CCM consists of a membrane coated with a catalyst layer (CL) electrode on both sides. The ionomer, which is a component of both the membrane and the CLs, absorbs/desorbs water as the CCM equilibrates, and as such, the ionic conductivity of the CCM is directly impacted. The electronic conductivity is indirectly impacted as dimensional changes in the ionomer result in changes to the network of electronically conducting materials in the CLs. Alternating current (AC) and direct current (DC) methods were investigated to measure the through-plane ionic resistance and in-plane electronic resistance as the CCM equilibrated.

When CCM is manufactured it is wound around a plastic core with an impermeable leader and trailer portion protecting the surface of the CCM from the core and from the environment. As such, only the two edges of the roll are exposed to environmental conditions. It is in this state that the CCM must equilibrate to the manufacturing environment. To simulate this, discrete 4 samples of a commercially available CCM (Ion Power) were tested in a sample holder with pressure applied to replicate the roll winding tension. The sample holder covered the top, bottom, and two sides (roll direction) of the CCM, leaving the other two sides exposed to controlled conditions in an environmental chamber. The measurements described below were taken as the samples equilibrated from 50% RH to 45% RH at a constant temperature.

The first method is an AC technique used to measure the through-plane sample impedance. The correlation between the RH of ionomers and their ionic conductivity is well established in literature [Cooper 2009, Maréchal 2007, Anantaraman 1996]. To measure the ionic conductivity of the sample, a sinusoidal alternating voltage was applied to silver foil strips on the top and bottom of the sample. This results in an alternating current passing through the sample. Electrochemical impedance spectroscopy (EIS) was used to measure the impedance as a function of frequency. From this data it was possible to extract the conductivity of the membrane and the equilibration time was investigated by tracking the impedance of the samples at 1 kHz.

The second method is a DC technique, which employs more readily available equipment and reduces the data analysis complexity compared to the AC technique. The DC setup used four probes made of gold plated foil strips contacting each corners of the sample: two probes to measure a voltage difference and two to apply current to the sample. Current flows through the CL via a carbon/platinum network, which is only indirectly affected by water content as the ionomer expands or contracts due to water absorption or desorption [Morris 2014]. A constant current was applied to the sample undergoing equilibration and the voltage difference was monitored to determine the equilibration time.

Both methods showed sensitivity to changing RH and exhibited reasonably constant results after a sufficient time, indicating that the samples had equilibrated to the environment. Figure 1 shows the impedance (1 kHz) of a CCM initially equilibrated to 50% RH as it equilibrates to 45% RH. The impedance increased rapidly for the first few hours and then slowed down before eventually reaching an approximately constant value, indicating equilibrium with the environment was achieved after approximately 18 h. The results of the DC method were consistent with those of the AC method.

The novelty of this research is in providing quantitative methods to determine when a CCM roll has equilibrated with its environment. Once equilibrated, the CCM roll can be used in the manufacturing of fuel cell stacks. This research will help increase the efficiency of the manufacturing process and improve the quality assurance of the final product.

Research funding from both MITACS (Accelerate program) and NSERC (Engage Plus program) is gratefully acknowledged, as is early and significant work by Anne Moore and Jeff Laflamme.

References

Anantamaran, A. V.; Gardner, C. L. J. Electroanal. Chem., 414, 115 (1996)

Cooper K. R. ECS Transactions, 25, 995 (2009)

Maréchal, M; Souquet, J. -L.; Guindet, J; Sanchez, J. -Y. Electrochem. Comm., 9, 1023 (2007)

Morris, D.; Liu, S.; Gonzalez, D.; & Gostick, J. ACS Applied Materials & Interfaces, 18609 (2014)

Figure 1

2459

, and

Proton-exchange-membrane fuel cells (PEMFC) have been greatly advanced towards commercially viable hydrogen-powered fuel cell systems for automotive application in recent years. Significant reduction in cost has been achieved by reducing the loading of Pt catalyst in the electrodes. Unfortunately lower Pt catalyst loadings accompany higher losses in cell voltage at high power. To quantify various voltage loss terms and thus identify the opportunity for improvement, mathematical models with parameters extracted from experiments are commonly employed. A differential cell operates at high stoichiometry flow with minimal pressure drop down the channel, making it almost one-dimensional across the cell. Hence the experiments are conducted on it to extract model parameters. The cell performance can be described by:

        Ecell = Erev - iRΩ - ηHOR - ηORR - i(RH+,a +RH+,c) - ηtx O2           (1)

 where Ecell is the cell voltage; Erev is the reversible cell voltage, i is the current density; RΩ is the sum of the Ohmic resistances of proton conduction through the membrane and of electron conduction (commonly referred to as high frequency resistance or HFR); ηHOR and ηORR are the charge transfer overpotentials for the hydrogen oxidation reaction (HOR) and the oxygen reduction reaction (ORR), respectively;  is the effective resistance to proton conduction in an electrode with subscripts a and c denoting anode and cathode, respectively; and lastly, is the oxygen transport loss which consists of two parts: one results from oxygen transport through gas phase, and the other is associated with the oxygen transport resistance local to the Pt surface, namely

 ηtx O2 = ηtx O2 (gas) + ηtx O2 (Pt-local)  (2)

 Limiting current densities are commonly used to evaluate gas phase and local oxygen transport resistances. Measurements with oxygen pressure and concentration variation can be used to evaluate oxygen transport resistance through the bulk phase, 1 and limiting current density measured on MEAs with varying Pt loading can be used to estimate local transport resistance.2 While both gas phase and local transport resistances are coupled, models with 1D transport lengths (across the gas diffusion layer (GDL), catalyst layer etc.) are used to de-convolute the resistances. The differential cells used are usually land-channel configuration which also has second diffusion length under the land. Transport resistance due to oxygen under the land across the GDL can dominate the total transport resistance. The pO2 is also largely decreased under the land and in an actual (non-differential) cell its impact on local oxygen transport resistance is enhanced down the channel where oxygen concentrations are lower. Therefore an accurate description of transport resistance under the land is needed to decouple the transport resistances. Differential cell tests with an active area of 5 cm2 were designed to study the effects of O2 diffusivity under the land by varying land/channel dimensions or flow field pitch. Cathode land/channel dimensions (mm/mm) of 0.2/0.2, 0.5/0.5, 0.7/0.7 and 1.0/1.0 were studied. In addition to the flow field dimensions, membrane electrode assemblies (MEAs) with different carbon supports and various Pt loadings were fabricated and tested. Catalyst such as Pt/HSC (high surface area carbon) and Pt/Vulcan at different Pt wt% were chosen to provide wide contrast in local oxygen transport resistance owing to the Pt particle distribution and location.  Electrochemical performance of the electrodes were characterized with electrochemical surface area, mass activity and polarization curve measurements. Additionally, protonic resistance of the catalyst layer were determined with electrochemical impedance spectroscopy in H­2-N2.

 An electrochemical model coupled with transport across the cell was developed to interpret the differential cell data. The oxygen transport resistance local to the Pt surface, hypothesized to originate from ionomer-Pt interaction, is also evaluated by model-data comparison. The model parameters perceived to be critical for accurate voltage-loss and cell performance prediction will be examined by sensitivity analysis to design and operating variations

 References

1) D. R. Baker, D. A. Caulk, K. C. Neyerlin, and M. W. Murphy, J. Electrochem. Soc., 156, B991 (2009).

2) T.A. Greszler, D.A. Caulk, P. Sinha, J. Electrochem. Soc., 159, F831 (2012).

2460

and

The durability of fuel cells is closely related to water content of the catalyst coated membrane (CCM). When the water content is high, the electrode catalysts become swollen due to dissolution, agglomeration, and redeposition, and the decline in the catalyst surface area accelerates. When the water content is low, impurities and hydrogen peroxide accumulate in the PEM, promoting chemical degradation and reducing the thickness of the PEM.

To help ensure durability of automotive fuel cells, it is therefore necessary to identify the in-plane distribution of the water content of CCM in the cells and to control it within the proper range. The in-plane distribution of water content is also affected by the operating condition parameters of the cells. In addition to the fact that a large number of combinations of operating condition parameters are possible, measurements also take time, and it has been challenging to determine parameters with an adequate level of accuracy.

 The development of the fuel cell stack for the 2016 model FCV sought to realize increased fuel cell stack durability by clarifying the distribution of the water content of the CCM in the cells through the development of a multi-segment impedance sensor system and the prediction of the optimum water content distribution using simulations.

 There are various measurement methods of water content inside fuel cell. For example nuclear magnetic resonance (NMR), neutron radiography, X-ray and so on. In order to clarify CCM water content in an actual vehicle, we need a method that would make it possible to obtain measurements without influencing the in-plane humidity distribution and other such factors during electric generation.

 Based on the results of examination, the impedance method was selected to measure in-plane distribution of water content during actual vehicle operation and development carried out.  The sensor is intended for incorporation in a fuel cell stack and was therefore designed with the same shape as the MEA. As a result, it becomes possible to obtain CCM water content measurements without any need to prepare special separators or seals. The surface on either side of the sensor has 75 square sensor pads which used to make impedance measurements of the CCM arranged. With measurement using the sensor, the distribution of CCM water content could be measured.

 In order to develop a method of simulation of the in-plane water content of CCM in cells, original Honda functional modifications were made to commercially available polymer electrolyte fuel cell simulation software. Simulation results correspond well with results of measurements of multi-segment impedance sensor system and the average predictive accuracy for the 75 segments was within 5.0%. Following the optimization of the operating condition parameters with the simulation, variations in the distribution of water content in the CCM were reduced by about 30%. In addition, the use of the developed simulation made it possible to determine operating conditions in several days that would previously have necessitated several weeks using tests.

The in-plane water content distribution of the CCM varies due to instantaneous changes in operating conditions, considerably during running mode, increasing during acceleration, and decreasing during deceleration. Because of this, it was necessary to verify whether the in-plane water content distribution of the CCM, which was optimized by simulation above, was maintained within the scope of the upper and lower limit values during transient current generation in an actual on-board generation system. During acceleration, In-plane water content of the CCM reaches its maximum value in the center of the cell in the direction of gas flow but remains below the upper limit value. In the transition from acceleration to deceleration, the in-plane water content of the CCM at the cathode inlet declines to close to the lower limit value. It was verified, however, that the in-plane water content of the CCM during deceleration remained within the upper and lower limit values.

2461

and

The effect of the microporous layer (MPL) on membrane electrode assemblies (MEAs) has been investigated. The performance of an MEA employing an MPL with a larger pore volume was better under 100%RH condition, which we reported in 2015 (1). Under 30%RH, no distinct relationship was found between the pore volume of the MPL and the MEA performance. However, the MEA employing an MPL with a larger median pore diameter showed higher cell voltage.

 

We reported that the MEA using a hydrophilic MPL showed much better performance in a wide range of pressure and humidity conditions than that using a conventional hydrophobic MPL. We applied our technique of preparing a carbon fiber dispersion to the fabrication of MPLs using carbon particles and carbon fibers with various fiber diameters. Properties of carbon materials and MPLs are shown in Table 1. The pore size distribution (Fig. 1) was measured using mercury intrusion porosimetry.

Figure 2 shows the surface SEM images of the MPLs: The MPL made of carbon black (Vulcan XC) has the smallest pore diameter (0.068 μm) and that made of carbon fiber (VGCF-H) has the largest pore diameter (0.77 μm). Although carbon fiber CF-X has the largest fiber diameter of 250 nm, the MPL made of it has the smallest pore volume. This is probably because fibers of CF-X are straight and longer than other carbon fibers, tending not to become tangled and form larger pores.

 

MEAs having an MPL made of four different carbon materials (Vulcan, 24PS, CF-X, and VGCF-H) were prepared, and polarization curves were measured under the conditions of 80 ºC, 30%RH. The IR free cell voltage (@0.1, 0.5, 1.0 and 2.0 A/cm2) for the four MEAs are plotted against the median pore diameter of the MPL in Fig. 3. The larger the pore diameter of the MPL, the higher the cell voltage becomes. The MEA employing an MPL made of VGCF-H, which has the largest median pore diameter, shows the best performance under the dry condition of 30%RH. This is probably because there is little liquid water in the MPL under the dry condition, and the cell voltage is largely dependent on the oxygen diffusivity of the MPL: The larger median pore diameter of the MPL, the better oxygen diffusivity and the higher the cell voltage.

 

References

1. T. Tanuma and M. Kawamoto, ECS Transactions, 69, 1323 (2015).

Figure 1

2462

, , , , and

Manufacturing methods for fuel cell catalyst layers are critically important for achieving high material utilization and performance. Typically, electrode systems are cast from an 'ink' or 'colloidal' state, where the solid material is suspended in a solution containing a solvent and a polymer. For polymer-electrolyte fuel cells, a perfluorinated sulfonic-acid (PFSA) ionomer, typically Nafion, is used in the ink and catalyst layer, where its roles are to act as a binding agent and conduit for proton conduction to the reaction site.1 Recently, it has been suggested that the ionomer film that surrounds the platinum reaction site may contribute to increased resistances that limit performance at low Pt loadings.1,2 Thus, it is of great interest to understand the factors controlling the film formation in catalyst layers to optimize material properties and cell performance. The nature of the dispersion and casting process could have significant influence on the morphology and properties of the dispersion-cast membranes.3-5 To isolate these effects, the morphology of PFSA dispersed in various solvents (glycerol, water/2-propanol, and NMP) was reported previously using small angle neutron scattering (SANS), where catalyst layers from the different solvents exhibit different performance.6 The understanding of local'ink' structures as a function of active material (polymer and solid) volume fraction (Φ) and across solvent space and length scales can aid in understanding bulk properties and the impact of ink formulation in cast systems. In this talk, a direct observation of the colloidal suspension of the ionomer in different solvents will be described using confocal laser scanning microscopy (CLSM). CLSM has long been utilized as a tool for characterizing the structure of biphasic colloidal gels. We combine dynamic light scattering, ultra-small angle x-ray scattering, and confocal microscopy to understand PFSA size and structure in dilute (1,3, and 5 wt%) ionomer solutions. The structural formation of nafion membranes are observed in-situ with small angle x-ray scattering. Finally, we contrive a model 'ink' composed of attractive and repulsive silica spheres and PFSA suspended in different solvents (water, NMP, and isopropyl-alcohol) and use CLSM to gain a greater understanding about solid|PFSA interactions in model inks.

1. A. Z. Weber and A. Kusoglu, Journal of Materials Chemistry A, 2, 17207 (2014).

2. A. Kongkanand and M. F. Mathias, The Journal of Physical Chemistry Letters, 1127 (2016).

3. T. T. Ngo, T. L. Yu and H.-L. Lin, Journal of Power Sources, 225, 293 (2013).

4. C.-H. Ma, T. L. Yu, H.-L. Lin, Y.-T. Huang, Y.-L. Chen, U. S. Jeng, Y.-H. Lai and Y.-S. Sun, Polymer, 50, 1764 (2009).

5. Y. S. Kim, C. F. Welch, R. P. Hjelm, N. H. Mack, A. Labouriau and E. B. Orler, Macromolecules, 48, 2161 (2015).

6. C. Welch, A. Labouriau, R. Hjelm, B. Orler, C. Johnston and Y. S. Kim, Acs Macro Letters, 1, 1403 (2012).

B-21 Fuel Cells in the Air and Undersea - Oct 4 2016 10:20AM

2463

, , , , and

Aviation industry and governing bodies have set out to significantly reduce the environmental footprint of aviation. The visions brought forward request such low emissions that continuous optimization and advancement of current technology alone are not sufficient. This has led research groups to explore disruptive paths of innovation for aircraft and a trend to the electrification of onboard systems in more electric aircraft. Providing the required electric power using fuel cells and batteries not only promises the reduction of carbon emissions and noise but also allows for system level optimizations using synergy effects from a multi-functional system-integration that aircraft design could benefit from. At the same time using fuel cells in aircraft environments provides great challenges towards system design. System weight and size constraints require the use of air breathing fuel cells in commercial aircraft and airliners. This results in operation at low pressure and temperature which in term confronts the system design with two problems. First the low air density makes cooling of power systems difficult especially considering the low operating temperatures of PEM Fuel Cells. Second the low ambient temperature means the ingested air will heat up much more than during ground operation and in term will remove more water from the fuel cells, making water management and fuel cell operation strategies more important and complicated than in other applications. Also ambient conditions very rapidly change in air, adding even more complexity to the application.

Taking into consideration possibilities of advanced aerodynamic concepts like distributed propulsion and propulsion/wing vortex interaction, that electric propulsion could enable, Fuel cells could be of even larger benefit for aviation when considering a larger time-frame. All-electric aircraft could potentially be emissions-free and very low noise, thus enabling sustainable, environmentally friendly aviation. While not yet feasible for large passenger aircraft, application of such systems is currently being researched in aircraft like the HY4 enroute to this goal. This presentation will discuss the challenges of fuel cell operation in-flight, as well as introduce the concepts for multifunctional fuel cell APU replacements for A320 type aircraft and drive-trains for all-electric aircraft. Finally results from ground- and flight-testing of the all-electric fuel cell powered HY4 aircraft will be discussed.

2464

and

Hydrogen fuel cells materials and systems have attracted billions of dollars of government and commercial research such that they are now reliable and cost effective enough to be commercialized as the power source for automobile propulsion and for materials handling. Also known as polymer electrolyte membrane fuel cells (PEMFCs), hydrogen fuel cells are attractive because they convert high-energy hydrogen fuel and oxygen (typically from air) efficiently to electricity and water through electrochemical processes, and produce relatively little heat and noise. They also respond rapidly to changes in load. The result is that they can have higher specific energy and energy density than battery systems for certain applications. One such application for commercial automotive hydrogen fuel cells is for large unmanned undersea vehicles (UUVs). This paper will describe an Office of Naval Research effort by the U. S. Naval Research Laboratory, Naval Surface Warfare Center Carderock Division, University of Hawaii, and General Motors (GM) to integrate an automotive fuel cell from GM into an unmanned undersea vehicle. The advantages and challenges with taking commercial fuel cell technology and adapting it to undersea use will be discussed.

2465

, , , and

In this work a design process for the development of an air independent fuel cell system operating a commercial fuel cell stack as a range extender in an AUV is presented. The development follows an optimization strategy which has been derived from the universal fuel cell development plan proposed by Ang et al.[1] The first step is defining the constraints and the objectives in the design process. While constraints set the boundary conditions of the system the identified objectives are used as optimization parameters. The design plan follows the schematic in figure1:

Figure 1: design concept for an optimized system used in an AUV

The development of fuel cell systems used in a submarine environment will need to consider different boundary conditions than the development of a fuel cell system for automobiles. These constraints result from the sea water environment which can't be used as an oxygen source and additionally hinders mass exchange with the fuel cell system because of pressure differences. As a result a closed fuel cell system relying on a pure oxygen supply has to be developed. The fuel cell is powering the AUV with a stationary output while a battery is buffering the power differential between the output und the consumption of the AUV. The AUV is used as a testing platform by the German navy and thereby sets very unique constraints on the system development. Additionally the commercial fuel cell stack is setting boundary conditions for the system development, which has to meet the stacks individual operating conditions. After identifying the boundary conditions the system size was identified as the optimization parameter.

To optimize the system size different setups for the cooling system as well as the gas supply system found in the literature are evaluated. Setups meeting the boundary conditions are validated in a test bench with hardware-in-the-loop tests. Based on these measurements a volume-optimized system can be found. Challenges which deviate from this setup are identified and can be tackled in the subsequent design step.

The next step is a discussion of the different concepts and an evaluation of their impact on the system operation. While the discussed solutions are still meeting the maximum system size, an evaluation of their influence on other design parameters can be made. From this different system designs are derived and the resulting operation strategies are compared to further improve the volume-optimized system. Also solutions targeting specific challenges like standby mode and system shutdown are evaluated and presented.

In a last step the qualitative connections between the found system designs and the boundary conditions in this system are discussed leading to a more general approach for the development of air independent fuel cell systems for AUVs.

[1] S. M. C. Ang, E. S. Fraga, N. P. Brandon, N. J. Samsatli, and D. J. L. Brett, "Fuel cell systems optimisation – Methods and strategies," Int. J. Hydrogen Energy, vol. 36, no. 22, pp. 14678–14703, Nov. 2011.

Figure 1

B-21 Fuel Cell Contamination - Oct 4 2016 11:20AM

2466

, , , , , and

Lower cost materials for stack hardware and system components help reduce the overall cost of the automotive and stationary fuel cell systems and make these competitive in the market. However, low-cost system component materials need to provide similar function, performance and durability. Intelligently selecting low cost materials for application in polymer electrolyte membrane fuel cell (PEMFC) systems requires understanding the potential adverse effects that system contaminants may have on the fuel cell performance and durability.

There are many prospective balance of plant (BOP) materials that can be used in fuel cell systems. Families of material, based on input from OEMs, fuel cell system manufacturers and other attributes like cost, physical properties, have been studied. These include structural materials, elastomers for seals and (sub)gaskets, and assembly aids (adhesives, lubricants).

The contaminants from the BOP materials were leached out via an accelerated aging procedure. The leachates obtained from these plastics were a mixture of organics, inorganics, and ions. The sulfate anion was one of the anions identified in the leachates and was chosen as the model compound for further studies. Sulfate anions can also come from membrane degradation. Ex-situ electrochemical measurements were carried out to understand the impact of sulfate anion concentration on oxygen reduction reaction (ORR). The ORR scan rate and direction (low to high potential and vice versa) were studied to determine the effect of the initial state of the Pt surface on sulfate contamination. In-situ fuel cell experiments were also carried out. Sulfate anion was introduced to a working fuel cell to determine their effect on fuel cells performance. Our previous work showed that sulfate anion did not result in performance loss when it was infused into the cathode at 0.2 A/cm2. The cathode potential was greater than 0.8 V at this current density. It is postulated that sulfate contamination is potential dependent and that sulfate adsorption only occurs when Pt is in its metallic state. In the current study, the effect of potential on sulfate contamination will be studied by infusing sulfate into the cathode at several constant current densities. Several in-situ diagnostics such as infusion, cyclic voltammetry, electrochemical impedance spectroscopy, and I-V curves were carried out to better characterize the contaminant effects of sulfate anion. The goal is to better understand the contamination mechanisms of specific species so that mitigation strategies can be determined.

The authors would like to acknowledge funding from the U.S. Department of Energy EERE Fuel Cell Technologies Office, under Contract No. AC36-08GO28308 with the National Renewable Energy Laboratory and collaborations with colleagues at GM. Structural plastic materials and leachates were provided by GM for this study.

2467

, and

Research has shown that contaminants in the hydrogen fuel stream have an impact on Polymer Electrolyte Membrane Fuel Cells (PEMFCs) performance that is influenced by the platinum loading, membrane thickness, the impurity, its concentration and PEMFC operating conditions such as cell temperature, relative humidity, and back pressure. A significant portion of previous reported results were conducted with the fuel cell being operated in single-pass mode where the exhaust hydrogen is vented from the fuel cell. While these findings have proven useful in improving the fundamental understanding of the poisoning mechanisms of non-hydrogen species (contaminants) in operating single fuel cells; the question remains whether or not these findings correlate to PEMFC systems that use a re-circulating fuel stream as being designed by the OEMs. Some advantages of re-circulating the anode exhaust gas back into the anode inlet are: 1) the anode outlet water is returned into the dry H2 fuel stream to obtain proper humidification and also 2) the excess H2 fuel is returned to enhance fuel utilization. While these advantages are attractive, there are some inherent challenges that are introduced in fuel re-circulation mode. Two of the more notable drawbacks involve inert gas build-up due to crossover and the accumulation of contaminants in the anode. In this study, we focus our efforts on correlating the impact of CO and H2S on fuel cell performance in H2 single-pass versus H2 re-circulation mode using the SAE J27191 and ISO 14687-2 Hydrogen Fuel Product Specification2 limits (4 ppb H2S and 200 ppb CO). This study was conducted on Membrane Electrode Assemblies (MEAs) at the 2015 DOE target loadings for platinum (anode: 0.05 mg/cm2 and cathode: 0.1 mg/cm2).

References:

  • SAE J2719: Hydrogen Fuel Quality for Fuel Cell Vehicles, www.sae.org

  • ISO 14687-2, Hydrogen Fuel – Product Specification, Part 2: PEM fuel cell applications for road vehicles, http://www.iso.org/iso/catalogue_detail.htm?csnumber=55083

Acknowledgements:

The authors gratefully acknowledge the financial support of the DOE Fuel Cell Technologies Office and the support of Technology Development Manager, Charles (Will) James, Jr.

B-22 Hydrogen Sources and Systems - Oct 4 2016 2:00PM

2468

, , , , and

In Hawaii, very high penetration levels of intermittent renewable energy sources such as wind and solar are causing challenges in regulation of grid frequency, and as the percentage of these intermittent resources increases can result in their curtailment. It has been postulated that, similar to a battery, the optimized use of an electrolyzer as a variable load may help regulate grid frequency and increase the penetration of renewable energy resources on the grid. The use of an electrolyzer in this way could provide an "ancillary service" to the grid that can be assigned a monetary value. This monetary value can be used to offset the cost of hydrogen production. The hydrogen in turn can be used in high value applications such as a transportation fuel. The Hawai'i Natural Energy Institute (HNEI) is conducting research to assess the technical potential and economic value of using an electrolyzer-based hydrogen production and storage system as a demand response tool for grid management.   A 65 kg/day hydrogen energy system (H2ES) consisting of a PEM electrolyzer, 35 bar buffer tank, 450 bar compressor, and associated chiller systems, has been purchased and installed at the Hawaii Natural Energy Laboratory Hawaii Authority (NELHA) to demonstrate long-term durability of the electrolyzer under cyclic operation required for frequency regulation on an island grid system. A secondary objective is to supply hydrogen for three fuel-cell buses to be operated at Hawai'i Volcanoes National Park (HAVO) and by the County of Hawai'i Mass Transit Authority. The hydrogen will be transported from the production site by tube trailers to the HAVO dispensing site. A second dispensing site is located at the NELHA production site.     

A comprehensive test plan has been developed to characterize the performance and the durability of the electrolyzer under dynamic load conditions and initial testing was conducted at Powertech Labs in Vancouver Canada. A main objective of the test plan is to determine the operating envelope and dynamic limits of the electrolyzer and the overall H2ES.  A long-term study shall evaluate the H2ES performance under a load profile developed from a fast acting Battery Energy Storage System (BESS) that is currently in use as a grid management tool on the Big Island grid.  These tests shall demonstrate whether the electrolyzer has the capability to effectively mitigate the impacts of intermittent solar and wind power on the grid, while continuously generating 90-80% of its designed hydrogen productioncapacity.

The presentation will describe the H2ES configuration, plans for operation in the field, and results of initial tests to determine the dynamic response of the eletrolyzer and overall system

2469

These are exciting times to be pioneering Electrochemical Hydrogen Compression (EHC) as the call for linking renewable electrical energy and chemical energy storage is becoming stronger. Hydrogen is ideally suited for electrochemical conversion, thanks to its relative abundance in compounds on earth, extremely fast redox reaction and highest energy content by weight. Although we do not produce hydrogen, we are utilising the fast redox reaction to convert hydrogen gas at the membrane interface to protons, and back to hydrogen gas again.

The principles of Electrochemical Hydrogen Compression have been disseminated before and demonstrated the capability to achieve high pressures of 100 MPa single stage (14,503 psi), while simultaneously purifying the hydrogen gas. The company HyET is developing a product that will be suitable for multiple applications, ranging from (home) refuelling hydrogen vehicles to purification and recycling of industrial hydrogen (bio)gas.

Here we present the outcome of the project PHAEDRUS and DONQUICHOTE, where EHC technology was developed and validated as compressor for a Hydrogen Refuelling Station, further compressing the hydrogen gas supplied directly from a PEM electrolyser. Both technologies exhibit dynamic operation, fast response and modularity making them well compatible. The most effective system configuration to capture renewable wind and solar energy and deliver high pressure hydrogen has been investigated.

In addition, we review the purification ability of the EHC technology, extracting hydrogen selectively from mixed gasses in low and high concentrations. Hydrogen gas is typically produced from compounds (e.g. methane), its value is highly determined by its quality and any residual contaminations. The solid membrane was design to hold back high pressure hydrogen, and therefore shows superior permselectivity limiting passage of larger molecules. As for the active hydrogen extraction process, the membrane selection should be viewed in tandem with the catalyst system, which may be affected by specific contaminations.

Technical working principles are elaborated on in this presentation, pointing out key benefits and scientific challenges that were encountered and to some degree resolved, finally discussing the roadmap ahead leading towards very attractive applications.

Figure 1

2470

, , and

One of the main concerns in the commercialization of fuel cell powered electrical vehicles is the on-board storage of hydrogen. Although pressurized vessels used in conventional systems provide higher hydrogen feeding rates, they have disadvantageous due to their extra weight and higher volumes in the systems. In this study, we report direct hydrogen generation from chemical hydrides, which does not require pressurized vessels. Self-dehydrogenation of chemical hydrides is a very slow process. In order to increase hydrogen generation kinetics, highly active and stable catalysts should be used. Although precious catalysts such as platinum and ruthenium meet these demands, their higher cost and recycling problems make them less useful in the fuel cell area.

A novel approach will be presented here for the fast, safe and stable hydrogen generation for fuel cells. This was achieved by circulating the acid and the chemical hydride solutions which were separated by the disulfonated poly(arylene ether sulfone) copolymer membrane. Since the rate of proton transfer is a function of ion exchange capacity (IEC) and scales with the degree of disulfonation of membranes, tailoring the IEC of membranes is crucial during the hydrogen generation. Therefore, membranes with varying sulfonation degrees (25-45 molar %) have been prepared and their hydrogen generation rates have been evaluated. Additionally, it was found that the type of acid and acid concentration influenced the hydrogen generation rates.

The effect of temperature on hydrogen generation performances was also investigated. When temperature exceeded the 80 °C, the hydrogen generation performance of N212 (the state of art membrane, Nafion™) was decreased due to the morphological changes occurred at this temperature. However the performances of our membranes without morphological relaxations continued to increase at higher temperatures. Furthermore, we run a proton exchange membrane fuel cell with our hydrogen generator and demonstrated continuous hydrogen generation more than 400 h. As a result, highly efficient on-board hydrogen generator for electrical vehicles powered by fuel cells presented here completely eliminates the need for separate facilities for producing hydrogen.

2471

The properties of nanomaterials are known to be size depend. Such size depend effects could offer powerful means to finally control both the thermodynamic and kinetic properties of hydride materials at the molecular level [1] and thus enable the practical design of hydrogen storage materials from these effects. However, investigations through such an approach require new strategies to both synthesise and stabilise nanoparticles of highly reactive hydride materials and the understanding to control the key parameters that will allow reversibility with high storage capacity. The use of porous host structures offers such a route, however the storage capacity usually remains limited by the intrinsic difficulty of entirely fill the porosity.

Herein, the potential of a new core-shell confinement approach and recent progress we have made through this nanosizing method will be discussed [2]. Since, complex hydrides still undergo phase transitions including melting at the nanoscale, this core-shell approach has proven to be an very effective strategy to stabilise and simultaneously catalyse the reversible storage of hydrogen with hydrides including borohydrides and alanates (Figure 1). It also provides us with new means to stabilise against oxidation highly reactive nanoparticles of elements such as lithium.

References

[1] K.-F. Aguey-Zinsou and J.-R. Ares-Fernandez, Energy Environ. Sci., 3 (2010) 526.

[2] M. Christian and K.F. Aguey-Zinsou, ACS Nano, 6, 9, (2012) 7739.

Figure 1. (a) Approach for the synthesis of core-shell borohydrides and (b) TEM and associated elemental mapping of NBH4@Ni, and (c) corresponding hydrogen cycling.

Figure 1

2472

, , , , , and

In rural areas, networkable environmental monitoring equipment is often used to provide continuous, real-time information without the need for a dedicated operator for maintenance or monitoring. However, power and energy availability are often controlling factors for operational lifetime when these systems are put on location. In these situations, batteries are often used as the main energy storage medium and require servicing or replacement when exhausted. This can lead high costs related to man-power and logistics to return and maintain the equipment at those locations.

Sodium borohydride hydrolysis was once looked at as a promising way to store hydrogen for use in PEM fuel cells. Concerns regarding hydrogen storage capacity and recyclability of reaction products have prevented its use in larger transportation or energy storage applications. However, the energy density of sodium borohydride still makes the reaction suitable for specific applications where reuse is not needed. As such, present here a system utilizing solid sodium borohydride as a storage medium to meet on demand power needs of these types of systems. The system utilizes a PEM fuel cell integrated with an electro-mechanical system to control chemical metering and power management.

We have previously shown results from our initial testing with solid sodium borohydride using commercial-off-the-shelf tablets for generating hydrogen when added to a reaction vessel containing water. Conceptually, a tablet would be introduced to the reaction vessel whenever additional hydrogen gas would be required. Operating in this manner would require good stability and reproducibility over multiple tablet reaction events.

Initial testing showed good reaction rates when a cobalt catalyst was employed, either pre-doped into the solid tablet mixture or when the catalyst was pre-mixed with the water in the reaction vessel. Analysis of the gas mixture showed high quality hydrogen gas flows could be produced and sent to the PEM fuel cell for power production. However, concerns with the stability, safety, and handling of the catalyst have led to a need to identify a more green chemistry approach.

We present here an alternative approach using common mineral and organic acids as hydrolysis accelerators that could be considered a greener approach to hydrogen generation for remote power applications. These include hydrochloric acid, phosphoric acid, citric acid, and acetic acid. Initial testing shows good reproducibility over multiple hydrogen generation events and reasonable reaction control. A comparison of hydrogen yield and gas composition is also shown. Finally, cost comparison between these accelerators is performed in order to assess their suitability in practical application.

E-12 Alkaline & DFC Electrocatalysis 2 - Oct 4 2016 2:00PM

2473

, , , and

The present study refers to a novel and unique approach of fabrication and deposition of gold nanoparticles on the surfaces of reduced graphene oxide and multi-walled carbon nanotubes. Among important issues is application of inorganic (rather than organic) capping ligand, heteropolymolybdate, to modify and stabilize (as well as probably also to link with the oxygen or hydroxyl groups on graphene surfaces) gold nanoparticles. During operation in alkaline medium or neutral media, polyoxometallates disappear but catalytically highly active gold remains and it exhibits excellent stability. The resulting material has occurred to show highly potent electrocatalytic properties toward electroreductions of carbon dioxide in neutral medium and oxygen in alkaline solution. What is even more important is that both carbon nanotubes and graphene have occurred to act effectively as carriers for gold nanostructures. Mutual activating interactions are feasible. The conclusions are reached on the basis of diagnostic electrochemical (e.g. rotating ring disk voltammetry), spectroscopic (FTIR) and microscopic (SEM, TEM) experiments. A series of comparative experiments with different carbon carriers and model catalytic materials (e.g. Vulcan-supported platinum) have also been performed. With respect to oxygen reduction, our diagnostic experiments at different concentrations of H2O2, support a view that the effect of the fast following chemical (H2O2-reductive-decomposition) reaction could be the dominating factor in explaining the observed positive potential shift observed during the oxygen reduction. The fact, that the optimum graphene-based catalytic system produced the oxygen reduction peak current comparable to that observed at the model platinum containing catalyst, would imply the efficient four-electron-type reduction mechanism.

Graphene and other distinct carbon nanostructures are explored as supports (carriers) for dispersed metallic silver nanocenters which exhibit electrocatalytic activity toward reduction of oxygen in alkaline media. To facilitate fabrication, immobilization and distribution of Ag(0) nanoparticles, various types of graphene (e.g. reduced graphene oxide, RGO, and CFx graphene with carbon black) have been modified with the silver analogue of polynuclear Prussian Blue, namely with ultra-thin silver(I) hexacyanoferrate(II,III) layers. Following the heat-treatment step at temperatures as high as 400-600oC, some thermal decomposition of the cyanometallate network occurs and, consequently, metallic silver sites are generated. Their formation and distribution are facilitated by the voltammetric potential cycling in KOH electrolyte. The most promising electrocatalytic results with respect to the reduction of oxygen (the highest currents and the most positive electroreduction potentials) have been obtained when graphene nanostructures are combined or intermixed with Vulcan XC-72R nanoparticles. What is even more important that, due to the presence of the polynuclear cyanoferrate modifier or linker, the amounts of the undesirable hydrogen peroxide intermediate are significantly decreased.

2474

, , , , , and

Understanding of the catalyst-ionomer interfacial behavior is critically important for fuel cell research. In fact, alkaline membrane fuel cells (AMFCs) have shown limited performance compared to proton exchange membrane fuel cells due to relatively poor hydrogen oxidation reaction (HOR) kinetics [1]. Previous work has shown that cationic adsorption at the surface of the catalyst was closely related to the HOR activity [2, 3]. Here, we present the results of a study of the benzyltrimethyl ammonium electrolyte-Pt interface in hydrogen oxidation reaction (HOR). More specifically, we will discuss the adsorption of the cationic group that may inhibit the HOR activity of Pt electrocatalysts. Microelectrode, Infrared Reflection Adsorption Spectroscopy (IRRAS) and Neutron Reflectometry are complementary tools to investigate the catalyst-ionomer interface.

Acknowledgement

US department of Energy Fuel Cell Technologies Program (Program Manager: Nancy Garland) supports this research through Fuel Cell Incubator Project. We also wish to acknowledge Oak Ridge National Laboratory Spallation Neutron Source, the NIST Center for Neutron Research and the Los Alamos Neutron Science Center.

References

[1] W.C. Sheng, H.A. Gasteiger, Y. Shao-Horn, "Hydrogen Oxidation and Evolution Reaction Kinetics on Platinum: Acid vs Alkaline Electrolytes," J. Electrochem. Soc., 157, B1529 (2010).

[2] H. T. Chung, U. Martinez, J. Chlistunoff, Y. S. Kim, Y. K. Choe, and I. Matanovic, "Impact of Organic Cation Adsorption on the Hydrogen Oxidation Reaction of Pt in Alkaline Fuel Cells", 2015, ECS Meeting Abstracts, 1510-1510

[3] Sung-Dae Yim, Hoon T. Chung, Jerzy Chlistunoff, Dae-Sik Kim, Cy Fujimoto, Tae-Hyun Yang, Yu Seung Kim, "A Microelectrode Study of Interfacial Reactions at the Platinum-Alkaline Polymer Interface," J. Electrochem. Soc., 162, F499 (2015).

2475

, , , , , , , and

Non-precious metal catalysts (NPMCs) for the oxygen reduction reaction (ORR) present an opportunity to improve the industrial feasibility of low temperature fuel cells by replacing expensive precious metal catalysts. However, improvements in performance of NPMCs are needed before they can realize widespread commercial adoption. Since the catalyst is projected to be the largest single contributor to the overall cost of the fuel cell, this area has seen heavy research activity. Many NPMC chemistries have been studied for the ORR in alkaline media, with perovskite oxides (of the form ABO3) being of particular interest due to the large variety of possible compositions that form perovskite structures, and the ability to tune material properties through doping of the A and B sites.1 Optimization of the bulk and surface chemistry to maximize intrinsic catalytic activity is of major importance to drive the development of perovskite oxide NPMCs. Synthesis strategies that balance surface chemistry and surface area optimization are necessary to improve the active site density while maintaining high catalytic activity. In addition, formation of effective perovskite oxide/carbon composites is essential to catalyze the ORR by a 2 step, 4 electron (2 per step) mechanism.2

This work focuses on the development of high performance perovskite/carbon composites by tuning surface chemistry, surface area, and optimizing the ratio of perovskite oxide to carbon. Through an aerogel synthesis process, a high surface area Ca0.9La0.1Al0.1Mn0.9O3-δ perovskite oxide catalyst was produced. Calcination temperature was varied between 500 and 1000C to study the interplay between phase purity and the catalyst surface area. Rotating disk electrode (RDE) studies showed that the optimum balance between phase, purity, surface composition and surface area is obtained using calcination at 800C. Rotating ring disk electrode studies (RRDE) were used to evaluate the effect of perovskite oxide to carbon ratio with further investigations in membrane electrode assemblies (MEAs) also demonstrating a significant impact on the performance (Figure 1). Additional improvements in performance were realized by varying the catalyst loading and using carbon functionalized with nitrogen.

References

1. Hardin, W. G., Mefford, J. T., Slanac, D. A., Patel, B. B., Wang, X., Dai, S., ... Stevenson, K. J. (2014). Tuning the electrocatalytic activity of perovskites through active site variation and support interactions. Chemistry of Materials, 26(11), 3368–3376. http://doi.org/10.1021/cm403785q

2. Poux, T., Napolskiy, F. S., Dintzer, T., Kéranguéven, G., Istomin, S. Y., Tsirlina, G. A., ... Savinova, E. R. (2012). Dual role of carbon in the catalytic layers of perovskite/carbon composites for the electrocatalytic oxygen reduction reaction. Catalysis Today, 189(1), 83–92. http://doi.org/10.1016/j.cattod.2012.04.046

Figure 1

2476

, and

Introduction

Alkaline fuel cells (AFC) have attracted much attention because non-Pt-based catalysts can be potentially applied.

Metallocomplex-based catalysts are the one of the most prospective candidates of non-Pt-based oxygen reduction reaction (ORR) catalysts. These catalysts have several advantages over Pt catalysts. First, since active sites consist of isolated metal atoms in these catalysts, metal resource can be used economically. Second, the catalytic activity can be improved by modifying structure of the ligand. Among these metallocomplexes, iron complexes-based catalysts are intensively studied because of low cost and relatively high activity.

We have been particularly interested in tetraazaannulene (TAA) iron complexes. The TAA complex is one of the N4-type macrocyclic complexes, as shown in Fig. 1. The main advantage of TAA ligands is that the molecular size is smaller than those of other conventional N4macrocyclic ligands such as porphyrins and phthalocyanines. The density of small molecules on an electrode can be increased. Furthermore, the substituents would have strong electronic effects on the active sites due to the small molecular size. Especially, the introduction of a substituent into α- or β- carbon located at diketiminate motif (Fig. 1) is believed to give large electronic effects on the active center. This indicates that the ORR activity of TAA complexes might be enhanced by the introduction of suitable substituents.

Experimental studies on ORR activities of iron complexes of TAA have not been conducted to date, while computational studies reported that iron complexes of TAA and its derivatives have good ORR activity1. In this work, we investigate the redox property and ORR of the iron complexes ligated by TAA (Fig. 1 (a)) and TAA-NO2 (an electron withdrawing derivative, Fig. 1 (b)) in alkaline conditions, and discuss the effect of the ligand structure on the electrochemical behavior.

Experiments

The ligands of TAA2 and TAA-NO23 were synthesized according to the reported procedures. The iron complexes abbreviated as Fe-TAA4 and Fe-TAA-NO2 were synthesized by refluxing the ligand with FeCl2・4H2O in an organic solvent in an inert atmosphere.

The iron complex-modified carbon materials were prepared as follows. Ketjenblack (KB) was added to the solution of the iron complexes. The mixture was evaporated to dryness.

The hydrodynamic voltammetry using a rotating ring(Pt)-disk(GC) electrode was performed in 0.1 M NaOH under 25 °C. A Pt coil and a reversible hydrogen electrode (RHE) were used as a counter electrode and a reference electrode, respectively. The iron complex-modified KB (Fe-TAA/KB or Fe-TAA-NO2/KB) was immobilized on a glassy carbon (GC) electrode.

Results and discussion

The ORR activities of the Fe-TAA/KB and Fe-TAA-NO2 were evaluated in the alkaline aqueous solution. Strong activities were observed; the onset potential, at which the value of the catalytic current is 0.2 mA cm-2, reached 0.94 V (Fe-TAA) and 0.97 V (Fe-TAA-NO2) vs. RHE. The onset potential was increased by the introduction of the -NO2 substituents. In the both of the cases, few H2O2was detected at ring electrode. This indicates that the numbers of electron transfer number of ORR by these catalysts are close to 4. The correlation between the onset potentials and redox potentials will also be discussed.

Acknowledgement

The authors are grateful to Tokuyama Corporation for providing the anion exchange resin (AS-4).

References

1) A.L. Pereira Silva, L.F. de Almeida, A.L. Brandes Marques, H.R. Costa, A.A. Tanaka, A.B. Ferreira da Silva, and J.d.J. Gomes Varela Junior, Journal of Molecular Modeling, 20, 2131-2140 (2014).

2) Y.G. Yatluk, and A.L. Suvorov, Chemistry of Heterocyclic Compounds, 23, 316-320 (1987).

3) N. Nishiwaki, T. Ogihara, T. Takami, M. Tamura, and M. Ariga, J. Org. Chem., 69, 8382-8386 (2004)

4) Sustmann, R., Korth, H. G., Kobus, D., Baute, J., Seiffert, K. H., Verheggen, E., Bill, E., Kirsch, M., de Groot, H. Inorg. Chem., 46, 11416-11430 (2007)

Figure 1

2477

, , and

Introduction

Biodiesel fuel (BDF) attracts attention as a carbon-neutral fuel as well as bioethanol. As the production of BDF increases, the production of the glycerol byproduct is also increasing. It is important to develop new uses for glycerol to cope with the mass production of BDF. The use of glycerol as a fuel for direct alcohol fuel cells is promising, because anodic oxidation of glycerol produced by bioprocesses occurs in carbon-neutral cycles and direct glycerol fuel cells are expected to produce electricity with a low environmental load and to be energy efficient.

It is well-known that Pd is cheaper than Pt and Au, and Pd-based alloy electrodes are highly active for the oxidation of alcohols in alkaline media. We also have reported the PdAg alloys-loaded carbon exhibited higher glycerol oxidation reaction (GOR) activity and tolerance to poisoning species than Pd.1 The mechanism for GOR, however, is not so clear. In this study we prepared an Ag atomic layer-modified Pd model electrode, and investigated the potential dependence of GOR products by in situ infrared reflectance-absorption spectroscopy (IRAS). In addition, we discussed the GOR mechanism based on these data.

Experimental

The Ag atomic layer-loaded Pd (Ag/Pd) electrode was prepared by underpotential deposition of Cu (Cu-upd) on a Pt polycrystalline substrate and the following galvanic replacement with Ag. The coverage of Ag (θAg) on the Pd substrate was evaluated to be 0.5 from the charges for Cu-upd before and after the Ag modification. 1 M KOH and (1 M KOH + 0.5 M glycerol) solution were used for electrochemical measurements. In situ infrared reflection-absorption spectroscopy (IRAS) was used to qualitatively analyze the products of GOR on the Pd and Ag/Pd electrodes in alkaline solution. All electrochemical measurements were performed at room temperature.

Results and Discussion

The cyclic voltammograms of the Pd and Ag/Pd electrodes in a (1 M KOH + 0.5 M glycerol) solution is shown in Fig. 1. The Ag/Pd electrode had higher oxidation current and more negative onset potential of the oxidation current than the Pd electrode, clearly indicating that the GOR activity was enhanced by the modification of the Ag atomic layer. In addition, the peak potential of the oxidation current for the Ag/Pd electrode was more positive than that for the Pd electrode, suggesting that the former was superior in tolerance to poisoning species to the latter.

IRAS spectra for the Pt electrode exhibited that the absorption peak at 1335 cm-1 assigned to the formation of dihydroxyacetone was mainly increased at lower potentials, but the absorption peak at 1310 cm-1 assigned to the formation of glycerate was increased at higher potentials. This suggests that the secondary OH group in a glycerol molecule is preferentially oxidized at smaller overpotentials, and the primary OH group is greatly oxidized at larger overpotentials. For the Ag/Pd electrode, the oxidation of the primary OH group occurred even at smaller overpotentials, and hydroxypyruvate was also formed at larger overpotentials. 

Acknowledgement

This work was partially supported by JSPS KAKENHI Grant Number 15H04162.

Reference

1 B. T. X. Lam, M. Chiku, E. Higuchi, H. Inoue, J. Power Sources, 297, 149 (2015).

Figure 1

2478

, and

The rapidly increasing energy demand for human activities stimulates the lasting research interests to develop renewable energy alternatives worldwide. The electrochemical reduction of oxygen is one of key steps in controlling the performance of various next-generation energy conversion and storage devices, such as fuel cells, metal-air batteries. The electrochemical reduction of oxygen is one of key steps in controlling the performance of various next-generation energy conversion and storage devices, such as fuel cells, metal-air batteries. The commercialization of these technologies prominently depends on the development of low-cost high-performance electrocatalysts for oxygen reduction reaction (ORR) to replace the precious metal-based catalysts.

In this presentation, several reasonable ways for designing new low-cost nanocatalysts with high electrocatalytic activities and superior stability for ORR will be discussed. By focusing on the creation and the enrichment of highly active sites for ORR and simultaneously considering the mass transfer and electron transportation, we have developed several efficient ORR nanocatalysts.1-7 The further improvement of the performance can be achieved by introducing transition metal or nanostructures into these nanocatalysts. Furthermore, understanding the origin of high activity of these electrocatalysts in ORR is also critical for developing efficient non-precious metal catalysts but still challenging. We developed a new highly active Fe-N-C ORR catalyst containing Fe-Nx coordination sites and Fe/Fe3C nanocrystals, and revealed the origin of its activity by intensively investigating the composition and the structure of the catalyst and their correlations with the electrochemical performance. Based on our experimental and theoretical results, it can be concluded that the high ORR activity in this type of Fe-N-C catalysts should be ascribed to that Fe/Fe3C nanocrystals boost the activity of Fe-Nx. These new findings open an avenue for the rational design and bottom-up synthesis of low-cost highly active ORR electrocatalysts.

References

[1] W. Ding, Z.D. Wei, S.G. Chen, X.Q., T. Yang, J.S. Hu, D. Wang, L.J. Wan, S.F. Alvi, L. Li, Angew. Chem. Int. Ed., 2013, 52, 11755.

[2] Y.P. Xiao, W.J. Jiang, S. Wan, X. Zhang, J.S. Hu, Z.D. Wei, L.J. Wan, J. Mater. Chem. A.2013, 1, 7463.

[3] W.J. Jiang, J.S. Hu, X. Zhang, Y. Jiang, B.B. Yu, Z.D. Wei, L.J. Wan, J. Mater. Chem. A.2014, 2, 10154.

[4] R.J. Huo, W.J. Jiang, S.L. Xu, F.Z. Zhang, J.S. Hu, Nanoscale, 2014, 6, 203.

[5] Y. Zhang, W.J. Jiang, X. Zhang, L. Guo, J.S. Hu, Z.D. Wei, L.J. Wan, Phys. Chem. Chem. Phys., 2014, 6, 13605.

[6] L. Guo, W.J. Jiang, Y. Zhang, J.S. Hu, Z.D. Wei, L.J. Wan, ACS Catal.,2015, 5, 2903.

[7] Y. Zhang, W.J. Jiang, X. Zhang, L. Guo, J.S. Hu, Z.D. Wei, L.J. Wan, ACS Appl. Mater. Interfaces,2015, 7, 11508.

[8] Y. Zhang, W.J. Jiang, X. Zhang, L. Guo, J.S. Hu, Z.D. Wei, L.J. Wan, J. Mater. Chem. A.2016, DOI: 10.1039/C6TA01655C.

[9] W.J. Jiang, L. Gu, L. Li, Y. Zhang, X. Zhang, L.J. Zhang, J.Q. Wang, J.S. Hu, Z.D. Wei, L.J. Wan, J. Am. Chem. Soc., 2016, 138, 3570.

2479

, , , , , , and

Fuel cells are power sources converting chemical energy into electrical energy through electrochemical reactions occurring on catalyst surfaces, in general, on Pt or Pt alloy surfaces. Hydrogen is mostly and widely used as the fuel, but other fuels such as methanol, ethanol and formic acid are investigated as alternatives because of their higher energy density as well as easy transport in our society. Among these alternative systems, direct methanol fuel cells (DMFCs) have been developed and considered for applications in electronics because of their high theoretical energy density, simple systems and ease of handling. In contrast to these merits even over the hydrogen-based systems, the complicated and slow electrooxidation reactions of methanol have hampered their commercialization. In this light, a variety of electrocatalysts for methanol oxidation reactions have been synthesized and evaluated in order to enhance the sluggish reactions, in other words, to reduce a fuel cell stack cost by decreasing the amount of Pt usage (e.g., alloying with other metals, making Pt monolayer on other metals, making composites with other substances to have synergy effects, and controlling the shapes to utilize more active surface structures).

The carbon supported Pt nanoparticles (Pt/C) have been recognized as the most reasonable starting point for each fundamental research, though alloys are widely used in the research and commercial products for achieving better performance. Pt and its alloys have the best activities in methanol oxidation reactions (MOR) and oxygen reduction reactions (ORR), and, in order to utilize Pt surface effectively, Pt/C catalysts have been generally prepared, e.g., by an impregnation method involving a wet-chemical reduction (i.e., impregnation of the carbon supports withsolutions containing the metal salts to be deposited). This method is quite simple and is easy for large-scale production. Very recently, we have proposed a new approach for preparation of Pt/C catalysts by arc plasma deposition (APD) method, where Pt nanoparticles can be deposited on carbon in vacuum and has been applied for various fields other than fuel cell catalyst. Herein, we focus on a preparation of new MOR and ORR catalysts by using our coaxial pulse APD method, where two different metal sources (Pt and Ti) can be deposited on carbon support through vacuum processes.

The morphology of the obtained sample was characterized by the scanning electron microscope (SEM) and the transmission electron microscopy (TEM). The SEM images show a typical morphology of an activated carbon, and the TEM images show that the deposited nanoparticles are well dispersed on the carbon support and have an average size of ~3 nm. The elemental mapping data show that the Pt and Ti are homogeneously distributed on the entire of carbon support. X-ray photoelectron spectroscopy (XPS) measurement is carried out to ascertain the electronic state of Pt and Ti, and the Pt 4f7/2 and Pt 4f5/2 doublet peaks are observed at ~71.5 and ~74.5 eV, which mean the Pt is pure metallic Pt. Regarding core-level Ti XPS, the peaks located at 458.9 eV and 464.8 eV can be assigned to Ti(IV) 2p3/2 and Ti(IV) 2p1/2, respectively. Another peak located at 455 eV is probably attributed to the Ti3+. Considering the fact that the XRD pattern does not show any crystalline TiOx phases, the Ti should form amorphous TiOx phases. Thus, the materials are found to be Pt nanoparticles (~3 nm) and amorphous TiOx phases on carbon black.

Compared with other Pt-based catalysts, our catalysts show superior electrochemical performances in MOR and ORR, probably owing to the presence of the secondary metal oxide phase. The cyclic voltammograms (CVs) of methanol oxidation reaction in 0.5 M H2SO4 containing 0.5 M methanol solution were scanned for the synthesized catalyst, compared with pure Pt catalyst on carbon support prepared by the APD method, commercially available Pt/C catalyst, and Pt black. All the current densities were normalized by Pt mass (i.e., Pt mass activity), and our catalyst exhibits higher activities than the other catalysts, probably because of bifunctional mechanism as found in PtRu or other systems, where the metal oxide surface provides active oxygen for removal of intermediates, such as CO, on the Pt surface effectively. The ORR performance of our catalysts were preliminarily investigated by linear sweep voltammetry (LSV) using a rotating disk electrode in O2-saturated 0.1 M HClO4 solution at a scan rate of 10 mV s-1 and 1600 rpm. The synthesized PtTi/C and the synthesized Pt/C show almost the same onset potentials, and show the better mass activity than the commercially available products. More details will be presented on the day.

2480

, , , , and

With the dramatically increased market demands of new energy materials due to the development of a large number of electronic devices and environmental awareness of the awakening, catalysts for oxygen reduction reaction (ORR) which are the most critical parts for high energy capacity and cleaning power sources including fuel cells and metal-air batteries have attracted highly attention and intense researches recently. Despite massive efforts, developing catalysts of high efficiency, long-term durability and excellent methanol tolerance with low cost for ORR is still a big challenge. Pt and Pt–based alloys are still the best choices for catalyzing oxygen reduction process by now. However, their high cost, low abundance, inferior stability and declining activity hinder their large-scale uses in industry. Therefore, selecting alternative materials to replace Pt and improve its performance is of great concern.

Chitosan, as the second most abundant biopolymer in nature, is rich in the outer shell of some arthropod (prawns, crabs, insects) and the inner shell and cartilage of mollusks (squid, cuttlefish) with some attractive properties, such as non-toxic, biocompatibility and safety, especially, its natural nitrogen rich, strong chelation effects, sustainable and environmental friendly and low cost depend that it can be heavy used to synthesize metal and nitrogen co-doped carbon (MNC) nanomaterials , which can provide an enhanced catalytic activities.

In the present work, taking advantage of strong chelation effects between hydroxyl group and metal ions and abundant natural nitrogen of chitosan, we developed a novel Co and N co-doped carbon (Co-N-C) nanocomposites as high performance electrocatalysts for ORR by using a simple and facile hydrothermal method, followed by high-temperature calcination at inert gas atmosphere. This material shows a delectable catalytic activity compared to the commercial 20% Pt/C electrode in the alkaline solution for ORR with a relative positive onset potential of -0.045V (vs. Hg/HgO), a much improved by 16.7% current density than Pt/C (loading ~ 0.4mg/cm2) as well as excellent durability performance and superior methanol tolerance. In addition, the method we used in this research is also suitable for preparing other MNC, which demonstrates that we found a new, straightforward and low-cost way to synthesis a series of catalysts for ORR which are comparable to commercial Pt/C.

Cyclic voltammetry (CV) was employed to test the ORR catalytic activity of Co-N-C-800 and the commercial 20% Pt/C catalyst in saturated O2 alkaline solution (0.1 M KOH) and shown in Figure 1a, an obvious cathodic peak can be found at -0.10 V vs. Hg/HgO which implies excellent ORR performance, and according to figure 1b, we can find that our Co-N-C-800 has a very close onset potential (-0.045V) and much higher current density (16.7% improved) in comparison to that of the commercial Pt/C, respectively, superior to the performances of pure chitosan without Co (C-800), which can confirm that Co-doping is crucial to the outstanding activity performance of Co-N-C-800. Owning to small metal nanocrystals of cobalt can improve the conductivity of materials and charge transfer efficiency between carbon and cobalt and give more active sites1. RDE measurements at various rotating speeds at a scan rate of 10 mV s-1 in O2-saturated solution were carried out for Co-N-C-800 and shown in Figure 1c. The current density shows a typical increase with rotation rate due to the shortened diffusion layer and the linearity of K-L plots for Co-N-C-800 indicates a good four-electron reaction path selectivity with the electron transfer numbers (n) is ≈ 3.9, which is close to that of Pt/C catalyst and further demonstrates an outstanding catalytic performance of this catalyst.

Acknowledgments

This work was financially supported by the Shenzhen Peacock Plan (KQCX20140522150815065), the Natural Science Foundation of Shenzhen (JCYJ20150331101823677) and the Science and Technology Innovation Foundation for the Undergraduates of SUSTC (2015x19 and 2015x12).

Reference

1. Xie, S.; Huang, S.; Wei, W.; Yang, X.; Liu, Y.; Lu, X.; Tong, Y., Chitosan Waste-Derived Co and N Co-doped Carbon Electrocatalyst for Efficient Oxygen Reduction Reaction. ChemElectroChem 2015,2 (11), 1806-1812.

Figure 1

2481

, and

The development of energy storage and conversion devices provides a beneficial approach for the renewable energy application, which can help relieve the severe reliance on fossil fuels and also address problems related to global climate change. Currently, the efficiencies of energy conversion devices such as the metal–air batteries and fuel cells are mainly limited by the sluggish kinetics of the oxygen reduction reaction (ORR) and oxygen evolution reaction (OER) processes. Developing catalytically active and cost effective catalyst for the ORR and OER is of prime importance. So far, noble metals and their alloys, such as Pt and Pt-Pd, have been exclusively used as electrochemical bifunctional catalyst in ORR and OER due to their superior catalytic activity. However, the high cost, limited availability, poor durability and sluggish electron transfer kinetics of noble metal based bifunctional catalysts have impeded the practical application of metal-air batteries. Therefore, the discovery of cost effective, catalytically active alternative bifunctional catalyst such as non-precious metals, carbonaceous materials, and transition metal oxides is highly desirable.

Perovskite oxide (ABO3) possesses a unique electronic structure and chemical defect properties, and has been demonstrated to be a promising non-precious metal catalyst among the transition metal oxides in the application of metal air batteries and alkaline fuel cells. It is known that the intrinsic electrochemical catalytic activity is mainly determined by the B site cation in perovskite oxide. Extensive research conducted by Yang et al. demonstrated that the ORR and OER catalytic activities of perovskite oxides follow a volcanic relationship with the filling of electrons in antibonding orbitals [1]. Among the typical LaMO3 (M= Mn, Co, Fe, Ni, Cr), LaMnO3 and LaCoO3 have exhibited highest ORR and OER catalytic activities, respectively. In addition, due to the desirable electronic properties of the perovskite oxides, strategies such as cation partial substitution and oxygen non-stoichiometry formation could, therefore, be utilized as the effective approaches to fine-tune the catalytic properties and to achieve a better bifunctional activity [2,3].

To enhance the bifunctional catalytic performance, improved specific surface area as well as enhanced oxygen channel are desired. Researchers have found that electrochemical catalysts with porous nanofiber structures could favor the ORR and OER pathways by providing uniform O2 electrolyte distribution, and beneficial oxygen diffusion channels. Zhao et al. have synthesized mesoporous La0.5Sr0.5CoO2.91 nanowires through the multistep microemulsion method, showing significant enhanced catalytic performance [4]. Xu et al. have demonstrated that the porous La0.75Sr0.25MnO3 nanotube catalyst fabricated through facile electrospinning technologies provides favorable advantages to the availability of the catalytic active sites in the organic solvent electrolyte in Li-O2 batteries [5].

Herein, we developed a bifunctional perovskite catalyst towards both ORR and OER in alkaline solution. Cobalt cations were doped into Pr0.5Ba0.5MnO3-δ perovskite to achieve the higher intrinsic ORR and OER activities by engineering the structure symmetry, electronic and the oxygen vacancy defects of the material. The bifunctional catalyst with a unique porous nanofiber structure was fabricated by electrospinning technology (Figure.1). A single phase perovskite oxide Co doped Pr0.5Ba0.5MnO3-δ was obtained after sintering the electrospun precursor at high temperature. The ORR and OER activities of the composite catalysts consisting 50 wt% of the as-synthesized perovskite oxides and 50 wt% carbon black were investigated in 0.1 KOH with rotating disk electrode (RDE). Significant enhancement in ORR and OER performance was achieved via using the composite catalyst with Co doped Pr0.5Ba0.5MnO3-δ nanofiber/carbon black with respect to the reduced overpotential and improved current density. In particular, the Co doped Pr0.5Ba0.5MnO3-δ nanofiber/carbon black composite demonstrated enhanced electron transfer number in the ORR process, indicating preferable dominance of four electron transfer pathway. Moreover, the Co doped Pr0.5Ba0.5MnO3-δ composite exhibited high stability during the cycling tests, indicating its promising applications in metal-air batteries.

References:

(1) Suntivich, Jin, et al. Nature chemistry 3.7 (2011): 546-550.

(2) Zhu Yinlong, et al. Chemistry of Materials, 28.6 (2016):1691−1697

(3) Jung, Jae-ll, et al. Angewandte Chemie 126.18 (2014): 4670-4674.

(4) Zhao, Yunlong, et al. Proceedings of the National Academy of Sciences 109.48 (2012): 19569-19574.

(5) Xu, Jing Ji, et al. Angewandte Chemie International Edition, 52.14 (2013): 3887-3890.

Figure 1

2482

and

Manganese oxide/poly(3,4-ethylenedioxythiophene) (MnOx/PEDOT) nanostructured hybrid thin films are readily prepared using a simple anodic electrodeposition process from aqueous solution. Using this approach, MnOx/PEDOT films with thicknesses and mass loadings up to 100 nm and 40 μg cm-2 were prepared, then tested for oxygen reduction reaction (ORR) activity in alkaline electrolyte using rotating disk electrode and rotating ring disk electrode methods. MnOx/PEDOT provided improvements over MnOx-only and PEDOT-only control films, with > 0.2 V decrease in onset and half-wave overpotentials and > 1.5 times increase in current density. The MnOx/PEDOT film exhibited only a slightly lower reaction order (n = 3.86-3.92) than the 20% Pt/C benchmark electrocatalyst (n = 3.98) across all potentials. MnOx/PEDOT also displayed a more positive half-wave potential and superior electrocatalytic selectivity for the ORR upon methanol exposure than 20% Pt/C. The high activity and synergism of MnOx/PEDOT towards the ORR is attributed to effective intermixing/dispersion of the two materials, intimate substrate contact and the improved charge transfer processes attained by co-electrodepositing MnOx with PEDOT. PEDOT has previously been demonstrated in fuel cells to limit methanol crossover. PEDOT has also been used as a catalyst support in fuel cells. These facts, combined with the substantially lower cost of Mn versus Pt (Mn is ~ 17,000 times cheaper), augurs for further development and testing of MnOx/PEDOT hybrid materials as ORR electrocatalysts with application in alkaline fuel cells.

This work was supported by the Laboratory Directed Research and Development program at Sandia National Laboratories, a multi-program laboratory managed and operated by Sandia Corporation, a wholly owned subsidiary of Lockheed Martin Corporation, for the U.S. Department of Energy's National Nuclear Security Administration under contract DE-AC04-94AL85000.

D-12 Core-Shell Cathode Catalysts & Catalyst Layers 1 - Oct 4 2016 2:00PM

2483

, , and

The slow kinetics of the oxygen reduction reaction (ORR) at the cathode and high cost of electrocatalysts (commonly Pt and Pt-based alloys) for this reaction are major hurdles that hinder the large scale production of proton exchange membrane fuel cells (PEMFCs).[1] A core-shell structure consisting of an atomically thin layer (up to several atoms thick) of Pt deposited on Pd has been recognized as a promising ORR electrocatalysts to replace bulk Pt-based materials.[2] Ideally, all of the Pt atoms will be fully utilized since they are all on the surface in the case of an atomically thin layer of Pt on a core material. The large scale synthesis of Pd@Pt core-shell catalysts, however, has not been fully realized. This study reported two synthesis methods to prepare gram scale batch of Pd@Pt/C electrocatalysts with the assistance of citric acid.

The first is Cu-mediated-Pt-displacement method involving the displacement of an underpotentially deposited (UPD) Cu monolayer by Pt.[3] In this method, the Cu monolayer deposited on Pd then is displaced by Pt via a surface limited redox replacement (SLRR) reaction: Pd@Cu + PtCl42- → Pd@Pt + Cu2+ + 4Cl-. Recent studies have shown that the SLRR reaction is rather complicated and poorly-controlled resulting in formation of 3D Pt islands instead of a uniform overlayer, especially in the large batch size synthesis. By adding citric acid in the displacement reaction, the activity of Pd@Pt/C can be enhanced by 2 times compared with that synthesized without citric acid.

The second method is direct chemcial reduction of PtCl42- with the assistance of citric acid without preformation of a Cu UPD layer. Comparing with other reducing agents, the reducing power of citric acid is neither too strong nor too week and allows the reduction reaction to occur at the room temperature, which significantly simplifies the reaction process. In addition, the strong interaction of citric acid with metals may reduce the Pd dissolution and introduce charge transfer from adsorbed citric acid molecules to Pd surface, which can also be used to reduce Pt cations. 

Both methods resulted good quality of core-shell catalyts in terms of activity and durability. The Pt activities of Pd@Pt/C preapred by the Cu-mediated-Pt-displacement and chemcial reduction methods were 0.95 and 0.78 A mg-1 at 0.9 V, respectively. Fuel cell performance will be also reported and the role of citric acid will be discussed in the presentation. 

References

[1] in: M. Shao (Ed.) Electrocatalysis in Fuel Cells: A Non- and Low- Platinum Approach, Springer, London, 2013.

[2] R.R. Adzic, Electrocatalysis, 3 (2012) 163-169.

[3] R.R. Adzic, J. Zhang, K. Sasaki, M.B. Vukmirovic, M. Shao, J.X. Wang, A.U. Nilekar, M. Mavrikakis, J.A. Valerio, F. Uribe, Top. Catal., 46 (2007) 249-262.

2484

, , , , , , , and

Introduction

Carbon supported Pd core-Pt shell catalyst (Pt/Pd/C) and Pt-Pd alloy catalyst (PtPd/C) are attractive alternatives to a carbon supported Pt catalyst (Pt/C) because their ORR specific activities are enhanced with an accelerated durability test (ADT) performed at 80°C [1, 2]. However, ECSA of the catalysts drastically decreased with the ADT, resulting in slight increase of ORR mass activities. We explored the ADT protocol and developed a high activation protocol (HAP) to mitigate the ECSA decay and enhance the ORR mass activities. Furthermore, we developed a Cu-O2 treatment which mimics the HAP carried out on GC electrode and is suitable for mass-production of highly activated catalysts. We also applied SiO2 coating technique [3, 4] to prevent agglomeration of the catalyst particles with the ADT and improve durability of the catalysts.

 

Experimental

Pt/Pd/C was synthesized with a modified Cu-UPD/Pt replacement process [1]. A carbon supported Pd core (Pd/C, Pd size: 4.6 nm, Pd loading: 30 wt.%, Ishifuku Metal Industry) was dispersed in 50 mM H2SO4 containing 10 mM CuSO4 and the solution was stirred with co-existence of a metallic Cu sheet at 5°C under Ar atmosphere. After stirring for 5 h, the Cu sheet was removed and K2PtCl4 was added to replace under potentially deposited Cu shell on the Pd core surface with Pt shell. PtPd/C alloy catalyst was synthesized with an impregnation method. Pt(acac)2 and Pd(acac)2 were impregnated onto a carbon support using t-butylamine as a solvent (Pt20Pd80 in at.%), followed by reduction under 15% H2/Ar atmosphere at 600°C for 4 h. The catalysts were characterized by TG, XRF, XRD, TEM, CV and XAFS techniques. ORR activity of the catalysts was evaluated with RDE technique in O2 saturated 0.1 M HClO4 at 25°C. ADT was conducted using rectangular wave potential cycling of 0.6 V (3 s)-1.0 V (3 s) vs. RHE in Ar saturated 0.1 M HClO4 at 80°C for 10,000 cycles.

 

Results and Discussion

Changes in ECSA and ORR mass activity of the Pt/Pd/C and PtPd/C catalysts with ADT, HAP and Cu-O2 treatment are summarized in Fig. 1 (dotted black lines show Pt/C catalyst; Pt size: 2.8 nm, Pt loading: 46 wt.%, TEC10E50E, TKK). The ECSA of the catalysts drastically decreased with the ADT, by which the ORR mass activity showed slight increases. On the contrary, the ECSA decay was mitigated with the HAP (rectangular wave potential cycling of 0.4V (300 s)-1.0 V (300 s) vs. RHE in Ar saturated 0.1 M HClO4 at 80°C for 30-40 cycles) and the ORR mass activity was highly enhanced, showing ca. 3-fold values of the Pt/C catalyst.

Since the HAP is performed on the GC electrode, treated amount of the catalysts is extremely small (ca. 30 µg), we developed a Cu-O2 treatment (Fig. 2) to scale-up the HAP. In the Cu-O2 treatment, the catalyst powders are stirred for 300 s in 2 M H2SO4 containing 0.1 M CuSO4 at 80°C under an inert atmosphere (N2) with co-existence of a metallic Cu sheet, where the equilibrium potential of Cu2+/Cu (ca. 0.3 V) is applied to the catalysts when they contact with the Cu sheet. Next, the Cu sheet is removed and O2 gas is introduced in the solution, where the equilibrium potential of the ORR (ca. 1.0 V) is applied to the catalysts. As shown in Fig. 1, the ECSA decay by the Cu-O2 treatment is equivalent to that by the HAP and the ORR mass activity of the catalysts were equivalently enhanced, indicating that the Cu-O2 treatment mimics the HAP on the GC electrode and is suitable for mass-production of highly activated Pt/Pd/C and PtPd/C catalysts.

However, the highly activated Pd@Pt NPs were severely agglomerated with the ADT and the ORR mass activity was decayed. In order to prevent the NPs agglomeration with the ADT, we applied SiO2 coating onto the Pd@Pt NPs [3, 4]. The SiO2 coating suppressed the NPs agglomeration and the ECSA decay was mitigated (Fig. 3), which highly enhanced the ORR mass activity of the Pt/Pd/C catalyst even after the ADT (Fig. 4).

 

Acknowledgement

This work was supported by NEDO, Japan.

 

References

[1] M. Inaba and H. Daimon, J. Jpn. Petrol. Inst., 58(2), 55 (2015).

[2] K. Okuno et al., The 228th Electrochemical Society Meeting, #1405, Phoenix, USA, October 2015.

[3] S. Takenaka et al., J. Phys. Chem. C, 111, 15133 (2007).

[4] S. Takenaka et al., Appl. Catal. A Gen., 409- 410, 248 (2011).

Figure 1

2485

, , , , and

Introduction

Pd@Pt core-shell catalyst is one of the promising cathode electrode materials for polymer electrolyte fuel cell (PEFC) from the view-points of a great potential to reduce Pt usage and to improve the oxygen reduction reaction (ORR) activity of Pt [1]. Recently, some researchers have reported that ORR activities of the core-shell catalysts increase during applying potential cycles, as a result of structural changes at the near surface regions [2]. Their results suggest that comprehensive understandings of dynamic behavior of the surface atomic structures during potential cycles are crucial for developing highly-active and highly-durable Pd@Pt core-shell catalysts. In this study, we investigate the relation between surface structures and electrochemical stabilities for the Pt/Pd(111) model electrocatalysts prepared by molecular beam epitaxy (MBE).

Experimental

Sample fabrication processes of the Pt/Pd(111) were conducted in UHV. Pd(111) single crystal substrate was cleaned by repeating cycles of Arsputtering and subsequent annealing at 800 °C. 4ML-thick Pt was deposited onto the cleaned Pd(111) (4ML-Pt/Pd(111)) by an electron-beam evaporation method at the substrate temperatures of 300 °C. The resulting surface structures were verified with reflection high-energy electron diffraction (RHEED), scanning tunneling microscope in UHV (UHV-STM), and low-energy ion scattering (LEIS). Then, the prepared samples were transferred without being exposed to air to an electrochemical system set in a N2-purged glove box. Cyclic voltammogram (CV) of the samples were recorded in N2-purged 0.1 M HClO4, and, then, linear sweep voltammetry (LSV) was conducted by using a rotating electrode (RDE) method after saturating the solution with O2. The ORR activities were estimated by kinetic-controlled current density (jk) at 0.9V vs. RHE by using Koutecky-Levich plots. The electrochemical stabilities were evaluated by applying potential cycles (PCs) between 0.6 V and xV (x = 0.8, 0.85, 0.9 and 1.0V) vs. RHE for each 3 seconds in O2-saturated 0.1 M HClO4 at 80 °C. The potential cycle conditions are hereafter denoted as "xV-PCs".

Results and Discussion

Surface structures for the clean Pd(111) and 4ML-Pt/Pd(111) are summarized in Fig.1 (a) and (b). The UHV-STM image indicates that the clean Pd(111) surface has atomically flat terraces. On the other hand, UHV-STM image of the 4ML-Pt/Pd(111) shows islands-like structures having hexagonal-shaped terraces with ca. 20 nm width. However, the corresponding RHEED pattern indicates sharp streaks, suggesting that deposited Pt atoms have epitaxially grown (111) lattice on the substrate.

Initial ORR activities for the 4ML-Pt/Pd(111), clean Pt(111) surfaces are shown in Fig.1 (c). The results indicates that the 4ML-Pt/Pd(111) exhibits ca. 4.5 times higher ORR activity than the clean Pt(111).Fig.1 (d) summarizes changes in ORR activities of the 4ML-Pt/Pd(111) during applying PCs at 80 °C. The activity of 1.0V-PCs applied sample drastically decreased with increasing PC numbers and dropped to less than that of the clean Pt(111) at 1000 PCs. However, it can be seen that decrease in upper limit potential of PCs suppress activity deterioration of the 4ML-Pt/Pd(111). In particular, the activities of 0.8V- and 0.85V-PCs applied samples after 5000 PCs were 120 % and 90 % of the initials, respectively. To investigate effects of surface structural changes in the ORR activity, the surface structures and compositions were evaluated by UHV-STM and LEIS after 5000 PCs. LEIS spectra (Fig.1 (e)) obtained after the PCs indicate that Pd surface concentration of 0.9V-PCs applied sample is higher by ca. 30 % than 0.8V-PCs applied sample, suggesting that activity deterioration during PCs is derived from increase in surface composition of Pd which show much lower ORR activity than Pt. Furthermore, judging from the corresponding UHV-STM images, although height roughness of the 0.8V-PCs applied surface is suppressed to lower than 0.7 nm, the 0.9V-PCs applied surface comprises agglomerated particle-like structures with 20 nm width and 3 nm heights. These results suggest that the structural and electrochemical stabilities for Pd@Pt core-shell catalysts strongly depend on operating conditions of PEFC.

Acknowledgement

This study was supported by the new energy and industrial technology development organization (NEDO) of Japan.

References

[1] R. Adzic, et al., Top. Catal., 46, 249 (2007).

[2] a) H. Daimon et al., 226th ECS Meeting, Abstract #1063, Cancun, Mexico (2014); b) K.A. Kuttiyiel, et al., Electrochim. Acta, 110, 267 (2013); c) W. Liu, et al., Angew. Chem. Int. Ed., 52, 9849 (2013).

Figure 1

2486

Developing a novel catalyst for oxygen reduction reaction (ORR) for real world applications should be focused on not only optimizing its activity and stability, but also synthesizing it by a simple process and thus being capable for mass production cost effectively. By taking a reliable recipe and facile one-pot synthesis method, we fabricated a wide range of unique nanoframe structures with high complexity, including the octopod, polyhedral, and hierarchical nanoframe architectures. Specially, the octopod nanoframe architectures possess three-dimensional catalytic surfaces, favorable platinum atomic arrangement and beneficial high-index facets (HIFs), thus can serve as models of advanced catalysts. The PtCu octopod nanoframe catalyst achieved 20-fold enhancement in mass activity for ORR compared to the commercial platinum-carbon catalyst. More importantly, this nanoframe catalyst maintained its structure over storing for months, after rigorous electrochemical reactions, and heat-treatment process. We also demonstrated that the one-pot synthesis of high-performance octopod nanoframe catalysts can be easily scaled up in high quality.

Fig. 1.Structural and compositional analysis of typical Pt-based nanoframe architectures.

Figure 1

2487

, and

Fuel cells are a promising source of renewable, clean energy. Widespread commercialization of fuel cell technology, however, has been hampered by technological and economic hurdles due to inadequate device efficiencies and expensive Pt electrocatalysts. To address these challenges, considerable research has focused on enhancing the kinetics of the oxygen reduction reaction (ORR) at the cathode which is the dominant source of voltage loss in state-of-the-art fuel cells.1 Recent work has sought to increase intrinsic catalyst activity and reduce Pt loading via core-shell nanoparticles where a Pt shell surrounds the core of a different metal or alloy. These nanostructures provide the ability to increase Pt mass activity through a larger Pt surface area to volume ratio as well as enhance the Pt intrinsic activity by altering its electronic structure though stain and ligand effects. These electronic changes dictate the binding energies of reaction intermediates to the catalyst surface and thus reactivity for the ORR. To identify a core material capable of improving ORR activity, we utilized a systematic approach for catalyst design developed in a previous study. This study found that activity can be improved by selecting a core material that is theoretically predicted to over-weaken the oxygen binding energy when covered with a Pt monolayer.2 The oxygen binding energy can then be strengthened towards its optimum by increasing the Pt shell thickness and through nanoscale effects where undercoordinated sites demonstrate stronger binding to adsorbates. Through this method, iridium was identified as a promising core material. Therefore, this work focuses on the synthesis, characterization, activity, and stability of Ir-Pt core-shell nanoparticles. Ir-Pt nanoparticles with three different compositions were synthesized using a highly scalable, inexpensive polyol method. TEM, STEM-EDS, and XPS were used to characterize particle size and composition. ORR activities of the Ir-Pt catalysts were compared to synthesized Ir-only and Pt-only nanoparticles as well as to the state-of-the-art commercial standard Pt catalyst, TKK. Electrochemical analysis demonstrates the high activity of the Ir-Pt catalyst with a specific activity of approximately 2 times that of TKK. This catalyst also demonstrates high stability after 10,000 stability cycles showing improvements in both mass and specific activity upon cycling with the latter increasing to 2.6 times that of TKK.

1. Stephens, I.E.L., et al., Energy & Environ. Sci. 5, 6744-6762, (2012).

2. A. Jackson et al., ChemElectroChem 1, 67-71 (2014).

2488

, , , , , and

In order to simultaneously meet the automotive PEMFC targets of cost, performance and durability, Pt-alloys electrocatalysts are being intensively investigated.1,2,3 Pt-alloys show promise of higher ORR mass activities that will permit the use of lower Pt loadings (~0.1 mgPt/cm2) in the cathode catalyst layers. In this work we have studied several Pt-Co/C catalysts that exhibit mass activities that exceed the DOE target of 440 mA/mgPt at 0.90V, 80oC, 100% RH for ORR activity. The performance at rated power density of 1W/cm2 at 0.60V is more difficult to achieve at low loadings due to unanticipated transport resistances. In addition to carrying out ORR activity studies both in TF-RDE set-ups4,5 as well as in MEAs of subscale cells, we have also applied complementary diagnostics of wet and dry H2-Air curves, microscopy of catalyst powders and catalyst layer films, EIS and dilute oxygen limiting current studies6 on these materials. Rigorous base-lining was carried out to obtain the ORR activity of a commercial Pt/C catalyst in order to benchmark the Pt-Co/C catalysts under study using identical protocols.7,8 Although we easily achieve the mass activity targets at 0.90V, we only approach the rated power targets indicating that further understanding of the transport resistances at high current densities and low Pt loadings is necessary to identify and mitigate the losses.

Acknowledgements: This work was funded through the DOE FC-PAD Consortium.

References Mathias MF, Makharia R, Gasteiger HA, Conley JJ, Fuller TJ, Gittleman CJ, Kocha SS, Miller DP, Mittelsteadt CK, Xie T, Van SG. Two fuel cell cars in every garage. Electrochem. Soc. Interface. 2005;14(3):24-35.

  • Gasteiger HA, Kocha SS, Sompalli B, Wagner FT. Activity benchmarks and requirements for Pt, Pt-alloy, and non-Pt oxygen reduction catalysts for PEMFCs. Applied Catalysis B: Environmental. 2005 Mar 10;56(1):9-35.

  • Kongkanand A, Mathias MF. The Priority and Challenge of High-Power Performance of Low-Platinum Proton-Exchange Membrane Fuel Cells. The journal of physical chemistry letters. 2016 Mar 11;7:1127-37.

  • Kocha SS, Zack JW, Alia SM, Neyerlin KC, Pivovar BS. Influence of ink composition on the electrochemical properties of Pt/C electrocatalysts. ECS Transactions. 2013 Mar 15;50(2):1475-85.

  • Shinozaki K, Zack JW, Pylypenko S, Pivovar BS, Kocha SS. Oxygen Reduction Reaction Measurements on Platinum Electrocatalysts Utilizing Rotating Disk Electrode Technique II.

  • Baker DR, Caulk DA, Neyerlin KC, Murphy MW. Measurement of oxygen transport resistance in PEM fuel cells by limiting current methods. Journal of The Electrochemical Society. 2009 Sep 1;156(9):B991-1003.

  • Kocha SS. Principles of MEA Preparation in Handbook of Fuel Cells: Fundamentals, Technology, and Applications; Vol. 3, edited by W. Vielstich, A. Lamm, and HA Gasteiger. Ch.;43:538.

  • Neyerlin KC, Gu W, Jorne J, Clark A, Gasteiger HA. Cathode catalyst utilization for the ORR in a PEMFC analytical model and experimental validation. Journal of The Electrochemical Society. 2007 Feb 1;154(2):B279-87.

2489

, , , and

Polymer electrolyte fuel cells (PEFCs) are regarded as a promising alternative power source for automobiles (fuel cell electric vehicle; FCEV). Cost reduction and increasing power density of PEFCs are the key challenges for commercialization of FCEVs, and reduction of Platinum amount in the catalyst layer (CL) is essential to achieve cost reduction of FCEVs. In order to accomplish the reduction of Platinum loading without sacrificing PEFC performance under high current density operation, it is important to make more effective use of Pt used in the CL of the membrane electrode assembly (MEA), and the mass transport of reactants within the CL should be enhanced significantly.

From this background, fabrication process of the CL should be investigated in more detail. The fabrication process of the CL includes an ink preparation process and a coating and dry process. In the former process, the materials composing the CL are mixed and dispersed in the solvent. A blend of water and alcohol is typically used as a dispersion medium for Pt-supported carbon and ionomer. The composition of the dispersion medium, the choice of alcohol, solid contents, and mixing conditions are major controlling parameters. The catalyst ink is deposited on a substrate or directly onto a PEM. The screen printing, die coating, or spray-coating is a major choice for the deposition method. The CL structure is finally formed after drying the deposited CL ink under elevated temperature. Extensive number of studies have been conducted to understand the impact of each parameters on PEFC performance, but the mechanism of structural formation of the CL has not been fully understood.

In this study, Impact of stirring condition of the catalyst ink on the CL structure, cell performance and oxygen transport phenomena was investigated by using cyclic voltammetry (CV) measurement, limiting current density measurement, and soft X-ray imaging. To investigate the impact of stirring condition, we prepared two catalyst ink with the same composition but the different stirring condition; stirred with Zr ball (ink A) and without Zr ball (ink B).

Figure 1 shows the result of CV measurement, and it showed that the CL prepared from the ink A have larger electrochemical surface area. Because the amount of Pt-supported carbon contained within the MEA should be the same, this result clearly suggested that the stirring condition of the catalyst ink affect the CL structure and its Pt utilization ratio. In addition, the i-V characteristics of these CLs (Figure 2) clearly showed that the CL prepared from the ink A have better performance than the CL prepared from the ink B, and the performance drop of the CL prepared from ink B was serious under the cell temperature of 30 deg.C. To understand this performance drop in more detail, the soft X-ray imaging technique and the oxygen transport resistance measurement were carried out. Results of the soft X-ray imaging (Figure 3) showed that more liquid water was accumulated within the MEA prepared from ink B. In addition, the oxygen transport resistance suggested that the oxygen transport resistance within the MEA prepared by using the ink A is smaller than the MEA prepared by using the ink B.

These results clearly showed that the stirring condition of the catalyst ink strongly affect the structure of the CL and the transport phenomena within the MEA. The stirring condition might affect not only the aggregation of the Pt-supported carbon but also the ionomer coverage on it, and their impact on the PEFC performance is still not clear. It is highly important to understand the CL fabrication process more scientifically, and it helps to develop high-performance CL.

Acknowledgement:

This work was supported by the New Energy and Industrial Technology Development Organization (NEDO), Japan.

Figure 1

2490

, , , , , , and

The development of stratified catalyst layers promises an increase in catalyst layer performance compared to conventional flat catalyst layers.1,2 Irregular catalyst layer thickness and porosity can lead to enhanced gas and water transport in and out of the catalyst layer, respectively. The stratified structure is expected to have the same performance in the kinetic region where the performance is controlled by the overall Pt loading. However at high current densities, the thinner sections of the stratified catalyst layers should allow for better mass transport properties.

Electrode fabrication is done in house with a custom designed spray coating procedure and catalyst ink recipe. Various approaches are being explored to achieve the desired electrode structure. One approach is to densify the catalyst layer in localized regions, and is based on Ion Power proprietary manufacturing techniques. This can involve the use of glass epoxy masks during the spray coating process to directly create a patterned electrode on the membrane that has thicker and thinner regions. Figure 1 illustrates the effect of stratification on the fuel cell performance of a MEA.

The second approach is to pattern thicker and thinner sections of an electrode using masks. During spray coating, one mask exposes the lands and the other mask exposes the channels. This way the two extreme cases can be investigated and the resulting performance compared. The concept of having more catalyst material in the channels may be beneficial due to rapid reaction times during influx of reactant gases through the channels. However, the presence of a thicker catalyst layer under the channel may later become an issue for product water removal and oxygen diffusion. On the other hand, spray coating more of the electrode in the land areas may support better product water removal and oxygen diffusion by leaving a thinner catalyst layer under the channels. The best performance is expected from a combination of ordered thin and thick regions in the catalyst layer. Results from various MEAs tested in a single cell on a fully automated test station to evaluate catalyst performance will be presented.

The evaluation of new catalyst ink recipes with reduced ionomer to carbon ratios is included in this study to further increase mass transport by reducing ionomer swelling in the catalyst layer at high relative humidity.

Acknowledgments

This research is supported by DOE Fuel Cell Technologies Office, through the Fuel Cell Performance and Durability (FC-PAD) Consortium; Fuel Cells program manager: Dimitrios Papageorgopoulos.

References

1. T. E. Springer, M. S. Wilson, and S. Gottesfeld,  J. Electrochem. Soc., 140 (12), 3513–3526 (1993).

2. R. Borup and T. Rockward, US Dep. Energy Annu. Merrit Rev., Project ID: FC052(2015). https://www.hydrogen.energy.gov/pdfs/review15/fc052_rockward_2015_p.pdf

Figure 1. Performance of textured and baseline MEA at 275kPa in H2/Air @ 80oC and 100%RH.                 

Figure 1

2491

, , and

For polymer electrolyte fuel cells (PEFCs) to achieve broad commercialization several technological hurdles have to be resolved. These include the high cost associated with the use of Pt as electrocatalyst in the electrodes, and durability problems due to carbon corrosion. PEFCs' conventional carbon-supported platinum (Pt/C) electrodes use 0.25 mg/cm2 of Pt loading, whereas the set target for 2020 by the U.S. Department of Energy (DOE) is 0.125 mg/cm2. Non-conventional thin-film electrodes prepared by atomic layer deposition (ALD) are a promising alternative to conventional designs1-2.

In the ALD process, monolayers of Pt are deposited onto the substrate by dissociative chemisorption in a self-assembling reaction. In this work, we propose an ALD thermal exposure mode to deposit Pt nanostructures onto a sacrificial anodized aluminum oxide (AAO) substrate using the trimethylcyclopentadienylmethylplatinum (IV) and oxygen gas as precursors. Thermal exposure mode provides additional time for the reactions in the ALD chamber and allows higher penetration of the precursor into the substrate. After the deposition the ALD electrodes were hot-pressed onto Nafion XL membrane and sacrificial AAO template was etched. We have demonstrated the feasibility of long free-standing structures (6-8 μm), as shown by Figure 1a.

The ALD cathode electrodes were electrochemically characterized within a fuel cell hardware where the cell had custom-made anode gas diffusion electrodes (GDE) with a loading of 0.2 mg/cm2. Polarization curves, cyclic voltammetry and electrochemical impedance spectroscopy (EIS) were performed at dry and wet conditions to assess activity and water management of these electrodes. A cyclic voltammogram is shown in Fig 1c, measured at a sweep rate of 40mV/s, 100% RH at 35° C and an inset to the figure shows a Nyquist plot of EIS under 60° C, 100%RH, H2/N2 at 0.8V, with a 5 mV perturbation measured between a range of frequencies from 200 MHz to 100 mHz.

In this presentation we will compare various ALD electrode fabrication conditions, ionomer presence and provide detailed electrochemical analysis (electrochemical surface area, polarization curves, cyclic voltammetry) of these nanoelectrode arrays.

Acknowledgements

This work was performed in part at the Center for Nanoscale Systems (CNS), a member of the National Nanotechnology Coordinated Infrustructure Network (NNCI), which is supported by the National Science Foundation under NSF award no. 1541959. CNS is part of Harvard University. We thank Mr. Jake Berliner for helping in ALD electrodes characterization and Mr. Berney Peng for SEM sample preparation. Fabrication of the structures was carried out in part at the Tufts Micro and Nano Fabrication Facility.

References

1. Galbiati, S.; Morin, A.; Pauc, N., Supportless Platinum Nanotubes Array by Atomic Layer Deposition as Pem Fuel Cell Electrode. Electrochimica Acta 2014, 125, 107-116.

2. Galbiati, S.; Morin, A.; Pauc, N., Nanotubes Array Electrodes by Pt Evaporation: Half–Cell Characterization and Pem Fuel Cell Demonstration. Applied Catalysis B: Environmental 2015, 165, 149-157.

Figure 1

2492

, , , , , and

Significant advances in the development of electrocatalysts that exceed the DOE target of 440 mA/mgPt (0.90V, 80oC, 100% RH, ptotal=150 kPa) for ORR activity have placed fuel cells on a promising path towards achieving the DOE target of a 0.1 mgPt/cm2elec cathode loading by 2020. However, unanticipated voltage losses that manifest at high current density and low Pt loading have prevented the attainment of the 0.125 gPGM/kWrated 2020 target.1 While some of the observed voltage loss for low Pt electrodes (< 0.1 mgPt/cm2elec) has been attributed to oxide dependent kinetics that manifest at the lower iR-free potentials (<0.75V),2 it has been observed that a significant portion of this unanticipated loss stems from a transport resistance local to or associated with electrochemically available Pt surface area.1 While the exact cause of this phenomenon remains unknown, studies have demonstrated that this loss both; 1) scales with total Pt surface area1 and 2) can be associated with the incorporation of ionomer into the cathode electrode.3Several diagnostics have been developed and applied to PEMFC materials in an attempt to isolate the exact cause and/or quantify the magnitude of this so termed "local Pt resistance".4-6 However insightful the results of these studies may be, the idealized conditions under which the diagnostics have been performed leave questions as to the relevance of the extracted resistance values. In this work we will expand the parameter space used previously for oxygen limiting current measurements4as a means to elucidate trends and gain further insight into the cause and magnitude of the local Pt resistance at relevant PEMFC operating conditions (e.g. cell temperature, oxygen partial pressure and relative humidity).

Experiments were performed using a 5cm2 differential cell with MEAs consisting of Pt/V and Pt/HSC at loadings of 0.2 – 0.05 mg Pt/cm2. An example measurement is shown in Figure 1 where clear increases in local Pt resistance (RO2,local) can be observed commensurate with; 1) an increase in oxygen partial pressure (pO2) and 2) a decrease in relative humidity (RH).

Figure 1. Local Pt resistance (RO2,local) as a function of oxygen partial pressure (pO2) and relative humidity (RH). Data was obtained on 50 wt% Pt/HSC electrodes of 0.05, 0.10 and 0.25 mgPt/cm2nominal loading, which were diluted with carbon to maintain a uniform thickness.

Acknowledgements: This work was funded in part under the DOE FC-PAD consortium and also by the U.S. Department of Energy under CRADA #CRD-14-539.

References

1. A. Kongkanand and M. F. Mathias, The Journal of Physical Chemistry Letters, 7, 1127 (2016).

2. N. Subramanian, T. Greszler, J. Zhang, W. Gu and R. Makharia, Journal of The Electrochemical Society, 159, B531 (2012)

3. A. Kongkanand, J.E. Owejan, S. Moose, M. Dioguardi, M. Biradar, R. Makharia, Journal of TheElectrochemical Society, 159, F676 (2012).

4. T. A. Greszler, D. Caulk, P. Sinha, Journal of The Electrochemical Society 159, F831 (2012).

5. H. Liu, W.K. Epting, S. Lister, Langmuir, 93, 8361 (1990).

6. H. Iden, S. Takaichi, Y. Furuya, T. Mashio, Y. Ono, A. Ohma, Journal of Electroanalytical Chemistry 694, 37-44 (2013).

Figure 1

A-12 Imaging - Oct 4 2016 2:00PM

2493

, , , and

Commercially viable polymer electrolyte membrane fuel cell (PEMFC) systems for transportation and stationary power applications require optimal combination of performance and durability at a competitive cost. Optimized water management across a fuel cell is key to improving performance and maintaining high conductivity of the ionomer in PEM and catalyst layers (CLs), while preventing excessive flooding in CLs, gas diffusion layers (GDLs), channels, and manifolds. Durability is also greatly affected by water transport as certain degradation mechanisms are accelerated depending on the humidity levels. As well, catalyst layers that suffered carbon corrosion due to long operation and/or high number of starts/stops are more susceptible to flooding.

Imaging in situ and ex situ enables crucial insight into the microstructure of the fuel cell components and materials, associated water transport, and their effect on cell performance, dynamic response, and degradation mechanisms. Several case studies are presented to highlight the use of optical imaging, neutron imaging, and micro X-ray computed tomography (microXCT) to investigate fuel cell components across a range of length scales. Laboratory-scale microXCT instruments are used ex situ to resolve the three-dimensional structure of the GDL and electrodes, before and after various accelerated stress tests.

Optical imaging in operating fuel cells offers a cost-effective combination of high temporal resolution (tens of frames per second) and high spatial resolution (several µm). The imaging requires special cell design with visual access to the imaged area (GDL surface and channels [1] or catalyst layer surface [2]). We employed optical visualization in a variety of studies, such as to visualize liquid water dynamics in the cathode and anode flow fields, or at the interface between the CL and the GDL. Besides water transport studies, optical imaging can quantify in-plane gas distribution in the flow fields during startups and shutdowns [3].

Neutron imaging is a powerful tool to visualize and quantify water transport in operating fuel cells. It has a high sensitivity to small amounts of water inside a cell, while having high transmission through the common fuel cell hardware. At a lower spatial resolution of 250 µm, neutron imaging is used to measure in-plane water distribution with high temporal resolution and large field of view of up to 20 cm by 20 cm. We studied the performance with novel NSTF (nano structured thin film) electrodes and GDL materials [4], and also visualized water and ice formation when cell was subjected to repeated starts at sub-freezing temperatures [5]. Since neutron imaging provides water content integrated along the neutron beam, we combined neutron imaging with the optical visualization to study water transport in different flow field geometries [6] as well as in an operating electrolyzer [7]. Such approach with simultaneous imaging provides additional information about the water location, and in certain situations allows distinguishing between channel and GDL water, or between anode and cathode channels and manifolds.

High-resolution neutron imaging is able to measure water distribution across the cell thickness at spatial resolution as high as 10 µm. Image processing procedure is developed to measure the water uptake and observe Schroeder's paradox in situ in PEMs [8,9]. Further, properties of the microporous layer (MPL) of the GDL were manipulated to prevent excessive flooding in cathode catalyst layers.

Anode flooding is evidenced by optical visualization, simultaneous imaging, and high-resolution neutron imaging. Increased water transport across the membrane, from cathode to anode, can be detrimental as liquid water on the anode side may cause localized fuel starvation and cause irrecoverable degradation, similar to accelerated carbon corrosion during unassisted startups and shutdowns. However, for certain novel cathodes, which are prone to flooding, e.g. nano-structured thin film (NSTF) electrodes and non-precious group metal (non-PGM) catalysts, water removal through the anode has proven to be a viable option to improve the performance by reducing cathode water content.

The authors acknowledge support from US Department of Energy EERE FCTO, Los Alamos National Laboratory LDRD (Laboratory Directed Research and Development) program, Federal Transit Administration, US Department of Commerce, and NIST Center for Neutron Research.

References:

  • D. Spernjak et al., J. Power Sources 170 pp334 (2007)

  • F.-Y. Zhang et al.J. Electrochem. Soc. 154 (11) B1152 (2007)

  • Y. Ishigami et al., J. Power Sources 269 pp556 (2014)

  • A. J. Steinbach et al., ECS Trans33 (1) 1179 (2010)

  • N.Macauley et al, J. Electrochem. Soc. submitted (2016)

  • D. Spernjak et al., J. Power Sources 195 pp3553 (2010)

  • O.F. Selamet et al., Int. J. Hydrogen Energy 38 pp5823 (2013)

  • D. Hussey et al., J. Appl. Phys. 112 (10), 104906 (2012)

  • D. Spernjak et al., ECS Trans33 1451 (2010)

2494

and

Polymer electrolyte fuel cells (PEFC) require an effective water management to operate efficiently, especially at high current densities which are needed to reach system cost targets [1]. The description of the complicated two-phase water transport remains a challenge in PEFC models and requires experimental validation on various length scales. In this work, operando X-ray tomographic microscopy (XTM) with scan times of 10 s was used to depict the liquid water at defined differential conditions at the technically relevant cell temperature of 80 °C at the TOMCAT beamline of the Swiss Light Source.

Cells with Toray TGP-H060 GDL with MPL were operated at high feed gas humidity (0.75 A/cm2, 108 % rH, H2/Air) or high current density (3.0 A/cm2, 78 % rH, H2/O2) conditions. The liquid water in the cathode GDL was analysed in terms of the local and average saturation, water cluster sizes, connectivity of the water clusters and the permeability of the water filled GDL pores. While fully through-plane connected water clusters contribute to the transport of water in the liquid phase, only partly connected water clusters do not (see Figure 1a). The majority of liquid water volume (> 80 %) is found in a limited number of fully through-plane connected water clusters (< 20 %) (see Figure 1b). Even though the water phase occupies in average the larger pores of the GDL, as reported earlier [2], still some large pores and percolation paths remain liquid free. Nevertheless, the available fully through-plane liquid paths provide liquid permeabilities that result in no noteworthy pressure losses to drain the product water of today's and upcoming current densities.

As research in industry and academia strive for the separation of the transport paths of the liquid and the vapour phase [3], the effect of the presence or absence of the partly connected water clusters on the gas phase transport parameters (permeability & effective diffusivity) of the GDL is discussed as well.

References

[1] O. Gröger et al., J. Electrochem. Soc. 162 (2015), A2605-A2622.

[2] T. Rosén et al., J. Electrochem. Soc. 159 (2012), F536-F544.

[3] A. Forner et al., Advanced Materials 27 (2015), 6317-6322.

Figure 1: a) Sketch of the different cluster types that can be found in the GDL domain; blue: full connected water clusters (FC); green: top connected (TC); red: bottom connected (BC); purple: non connected (NC). b) Volume and cluster number fractions of the different cluster types for the high humidity condition.

Figure 1

2495

, , and

Neutron imaging is an ideal method to study water transport phenomena in proton exchange membrane fuel cells (PEMFCs). Neutrons readily penetrate aluminum, carbon and platinum, yet have a high sensitivity to hydrogenous materials, including water. In addition, unlike x-rays, neutrons are a non-destructive probe and do not alter the performance of PEMFCs due to irradiation. As such, there have been numerous studies that revealed the heat and mass transport phenomena in the flow fields and gas diffusion layers [1], freeze operation [2], impacts of carbon corrosion [3], and the impact of porosity in thick non-precious metal catalyst layers [4].

The primary limitation of neutron imaging is the achievable spatial and temporal resolutions. Current state of the art spatial resolution is about 20 µm, with a 20 minute image acquisition time. The spatial resolution is primarily limited by the range of the charged particles that are the result of the neutron capture reaction and are used to detect the neutron. This range is of order 5 µm for many common detector schemes. The temporal resolution is limited by the brightness of neutron sources, which are in general many orders of magnitude less bright than synchrotron x-ray sources; thus to have reasonable time resolution conventional neutron imaging requires the use of beam defining apertures that are of order 1 mm for high resolution imaging of PEMFCs, so that one cannot make use of geometric image magnification.

We are pursuing three efforts to improve the achievable image spatial resolution and time resolution. The simplest approach to improving the spatial resolution is to place a narrow slit in front of the test section and scan it across the active area. The challenge of this method is to create a narrow, well-defined, completely opaque slit with opening area ~1 µm. KAERI and Pusan National University have devised a Gadox powder filling method that enables creating such slits, enabling imaging with 1 µm resolution [5]. By amplifying the scintillation light and using a high speed camera, it is possible to record individual neutron scintillation events and using a centroid algorithm determine the position of the neutron with high precision (less than 10 µm). The measured through-plane water content of a PEMFC is shown in figure 1 for the present state of both methods. The final effort is the development of a neutron microscope objective based on reflective neutron optics [6]. With a lens, one no longer requires the use of small beam defining apertures and thus the image temporal resolution can increase by at least a factor of 100. We will report on the progress of the first optics to reach 20 µm resolution and discuss the possibility of achieving 1 µm spatial resolution through neutron image magnification.

References

  • T.A Trabold et al, "Use of neutron imaging for proton exchange membrane fuel cell (PEMFC) performance analysis and design", Handbook of fuel cells v. 5, 2009.

  •  R Mukundan et al, "Performance of PEM fuel cells at sub-freezing temperatures", ECS Transactions 11 (1), 543-552, 2007.

  • JD Fairweather et al, "Effects of cathode corrosion on through-plane water transport in proton exchange membrane fuel cells", JECS 160 (9), F980-F993, 2013.

  • D Spernjak et al, "Water Management in PEM Fuel Cells with Non-Precious Metal Catalyst Electrodes", 228th ECS Meeting abstracts, p. 1540, 2015.

  • J. Kim et al, "Fabrication and characterization of the source grating for visibility improvement of neutron phase imaging with gratings", RSI 84(6):063705, 2013.

  • D. Liu et al "Demonstration of achromatic cold-neutron microscope utilizing axisymmetric focusing mirrors", Applied Physics Letters 102 (18), 183508, 2013.

Figure 1 Comparison of the image quality of the current state of the art (a), the slit imaging method with 4 µm resolution (b), and the reconstructed image from centroiding (c).

Figure 1

2496

, , , , , , , , and

Automakers around the world are investing in polymer electrolyte membrane (PEM) fuel cells for the next generation drivetrains for their low emission automobiles. For PEM fuel cell vehicles to become commercially competitive, fuel cell efficiency and durability must increase. However, in order to advance the performance of PEM fuel cells, the dynamic behaviour of liquid water transport must be taken into consideration, particularly because fuel cell cars will be driven with dynamic load (power draw) conditions1. The relationship between PEM fuel cell performance and dynamic load must be well understood so that next generation fuel cells can be tailored for optimal dynamic operation.

Effective water management is key to obtaining the optimal performance of fuel cells2,3. It has also been shown that the transient response of two-phase flows of water is significantly longer compared to the electrochemical response1. Therefore, the changes in two-phase flows in the fuel cell can have a larger impact on the dynamic performance of fuel cells4,5. Liquid water evolves in the porous layers of the fuel cell until it reaches a steady state for the current density of operation6. However, there is a scarcity of experimental observations focused on GDL liquid water saturation as a function of time with changes in operational current density.

In this work, the transient response of the cell potential and the dynamic change in water saturation of the gas diffusion layer (GDL) are investigated. The current density of the cell was increased from 0 A/cm2 to prescribed values at various rates of increasing current density (ramp rates), while the response of the cell potential was concurrently measured. Figure 1 shows the transient response of the cell potential due to a change in the current density from 0 to 1.2 A/cmat various rates of increase. With a step change in the current density, the voltage fluctuated for a few minutes before the potential of the cell dropped, leading to performance failure that was indicative of flooding. However, when a ramp was applied to gradually reach the same current density as the step change, the cell was able to maintain a steady potential. This shows that the ramp rate of increasing current density has a direct impact on the cell performance and its transient response.

In addition, X-ray radiographic evidence of water in the porous layers of the PEM fuel cell during transient operation will be presented. The dynamic water evolution at the microporous layer (MPL)|catalyst layer interface and the MPL|GDL interface are highly influenced by the rate of increasing current density. The cell performance was correlated to the time-dependent water evolution patterns in order to identify the interfacial liquid water accumulation as a function of changing current density.

References:

1. R. Banerjee and S. G. Kandlikar, Int. J. Hydrog. Energy, 40, 3990–4010 (2015).

2. S. G. Kandlikar, Heat Transf. Eng., 29, 575–587 (2008).

3. J. P. Owejan, J. J. Gagliardo, J. M. Sergi, S. G. Kandlikar, and T. A. Trabold, Int. J. Hydrog. Energy, 34, 3436–3444 (2009).

4. R. Banerjee and S. G. Kandlikar, Int. J. Hydrog. Energy, 39, 19079–19086 (2014).

5. R. Banerjee and S. G. Kandlikar, J. Power Sources, 268, 194–203 (2014).

6. Y. Wang and C.-Y. Wang, J. Electrochem. Soc., 154, B636 (2007).

Figure 1

2497

, , and

Phosphoric acid used as the electrolyte in high temperature polymer electrolyte fuel cells (HT-PEFC) is not covalently bound to the membrane polymer. The interaction of phosphoric acid with the nitrogen of the pyridine group of the polybenzimidazole (PBI) polymer backbone [1] is limited to two acid molecules per PBI repeating unit. Consequently, the bulk of phosphoric acid is mobile and the acid concentration and distribution in/and between membrane, catalyst and gas diffusion layers (GDL) is varying, depending on the operating conditions of the fuel cell.

X-ray imaging technology, initially developed for imaging liquid water in the porous structures of LT-PEFC, has been adapted [2] for the imaging of phosphoric acid. Operando tomographic microscopy imaging has shown that acid migration to the anode due to the non-zero transference number of the hydrogen phosphate anion is an important process in acid redistribution [3]. The acid can invade the anode GDL, similar to the injection of water into the cathode GDL of LT-PEFC.

So far, not much has been known about the characteristics of the interaction of phosphoric acid with the porous structure of the gas diffusion layer. Ex-situ acid injection and withdrawal experiments at temperatures up to 160 °C are performed with concurrent tomographic imaging to understand the characteristics, as well as the similarities and differences to the wetting of GDLs with water.

[1] Q. Li, J. O. Jensen, R. F. Savinell, and N. J. Bjerrum, Prog. Polym. Sci., 34, 449–477 (2009)

[2] S.H. Eberhardt, F. Marone, M. Stampanoni, F.N. Büchi, and T.J. Schmidt, J. Synchrotron Radiat., 21, 1319–1326 (2014)

[3] S.H. Eberhardt, M. Toulec, F. Marone, M. Stampanoni, F.N. Büchi, and T.J. Schmidt, J. Electrochem. Soc., 162, F310-F316 (2015)

2498

, , and

Neutrons and X-rays provide rich complementary information for investigations of three-phase flows. A hydrogenous liquid phase is well resolved from the solid and gas phases using neutrons due to the large neutron attenuation of hydrogen while the solid phase can be separated from the liquid and gas phases using X-rays. By combining these two imaging techniques it is possible to achieve a full understanding of the three-phase flow with high sensitivity to all phases. A critical application for this combined technique is in proton exchange membrane fuel cells where understanding of the liquid water buildup in porous materials is paramount to achieving high performance and reliability. To achieve this goal, NIST has brought online a unique system for simultaneous neutron and X-ray imaging. A 90 keV microfocus X-ray source is oriented 90° to the neutron beam and the sample is positioned at the intersection of the two beams allowing for the simultaneous imaging with both modalities. The system will facilitate both serial radiography using the X-rays to enhance the current state-of-the-art neutron radiography and simultaneous tomography for 3D resolved water content and solid material structure. This presentation will describe the current work to develop this new capability for the fuel cell research community, fuel cell hardware development for simultaneous tomography, and initial results obtained from this system.

2499

, , , , , , , , and

Synchrotron X-ray radiography has been proven to be a powerful tool for visualizing and quantifying the in operando liquid water distributions in polymer electrolyte membrane (PEM) fuel cells. Water thickness, the accumulated water content at an image pixel, was calculated by normalizing the images collected to isolate the presence of liquid water. Images containing liquid water are called wet images, and are evaluated with respect to a reference image in the absence of liquid water, called the dry image. This normalization process is based on the Beer-Lambert law, and the attenuation coefficient of liquid water (at the applied photon energy level) is vital for this calculation. The attenuation coefficient for monoenergetic X-ray photons, provided by the National Institute of Standards and Technology (NIST), describes the probability that a photon will interact with the particles in a material as it travels a unit distance in the direction of the beam. Hence, the liquid water thickness calculation is highly dependent on the correct use of the attenuation coefficient.

The incorrect use of a monoenergetic attenuation coefficient in synchrotron X-ray image analysis may lead to significant inaccuracies attributed to the neglect of scattering effects and higher harmonics contamination, which may be present at a synchrotron beamline. In our previous work (1), a calibration experiment was developed to experimentally obtain the water attenuation coefficient based from six known thicknesses of liquid water. The calibrated attenuation coefficient was 18.9% lower than the value for the monoenergetic photons at 20 keV. This deficit implied that at least one of the two prescribed phenomena (scattering and higher harmonics) contributed to the measured intensity in the obtained images. The purpose of this study is to determine the ratios of produced X-ray intensities that arise due to the scattering effect and higher harmonics, respectively. From this work, a correction method for calculating the water thickness can be proposed based on the scattering ratio. This correction method is vital for increasing the accuracy of water quantification.

The calibration experiments were performed with a range of photon energy levels and filters at the Canadian Light Source.  It was observed that the scattered signal increased as a function of water thickness, and this trend was more significant for lower experimental energy levels. The results suggested that the calibration experiment should be conducted with customized devices that contain liquid water with comparable quantities to that which one would find in the PEM fuel cell, in order to accurately simulate the in operando scattering effect.

The existence of the higher harmonics was verified at the experimental energy level of 20 keV, and the water attenuation coefficient was experimentally measured to be 0.258 cm-1, which is in agreement with the value reported by NIST at 40 keV (0.268 cm-1). From the results obtained at 20 keV with the high-transmission filter, the ratio of higher harmonics was determined to be 0.9%–1.4% of the incident beam intensity. From this work, the authors urge that caution should be exercised when using selecting energy levels that could contain higher harmonic photons.

References

1. N. Ge, S. Chevalier, J. Hinebaugh, R. Yip, J. Lee, P. Antonacci, T. Kotaka, Y. Tabuchi and A. Bazylak, Journal of Synchrotron Radiation., 23, 2 (2016).

2500

, , , , and

The catalyst layer (CL) of polymer electrolyte fuels cells (PEFCs) plays a critical role in their operation and also the production process of CL is of significant importance in terms of the cell performance as well as the system cost. The decal method has been used in the commercial production of CL in membrane electrode assemblies, in which a catalyst ink is prepared by mixing catalyst particles and ionomers in the solvent, coated on a decal substrate, dried to form a particulate layer with a specified thickness, and transferred to a membrane via hot pressing [1,2]. Since the three-phase interface (catalyst-void-ionomer) is required for the transport phenomena taking place in the electrochemical reactions, it is desirable to control the microstructure of CL during the decal process [3]. Therefore, the catalyst ink and its decal process should be studied in detail.

Herein, we propose a novel approach to investigate the characteristics of catalyst inks by using magnetic resonance imaging (MRI) combined with nuclear magnetic resonance (NMR) spectroscopy. Although traditional NMR studies of ionomers have been reported [4,5], the detailed discussion regarding the spacial inhomogeneity of particles in catalyst inks and their dynamics during the drying processs has been lacking. MRI enables us to visualize the spacial variations in the sample characteristics which cannot be accessed through other spectroscopy techniques [6,7]. Because the dispersion stability of catalyst particles and the morphology of ionomer depend on the solvent condition, we first examined the solvent dynamics by determining the diffusion coefficients. We developed a technique to acquire NMR spectra from a selected volume of a liquid sample and determine the local diffusion coefficients of component molecules. The proposed technique was verified using the reference values of diffusion coefficients of water and alcohols (D = 2.13×10-9 and 0.57×10-9 m2/s for water and n-propyl alcohol, respectively) [8-10] and then applied to our samples. The diffusion coefficients of water and n-propyl alcohol (NPA or 1-propanol) typically used in the solvent of PEFC catalyst ink were obtained at various mixture compositions. The diffusion coefficients of both water and NPA remarkably changed depending on the mixuture composition, which indicates the change in molecular microstructure [10]. Then, we mixed ionomers with the solvent and examined the effect of ionomer addition on the solvent dynamics. We observed the decrese in the diffusion coefficients of water and NPA due to the existence of mixed ionomers. In addition, we obtained MRI images of catalyst inks and the precipitated bed of carbon particles was visualized. We determined the local diffusion coefficient of water in the particle bed region with a thickness of 2 mm, and the dependence of diffusion coefficient on the diffusion time was investigated. As we increased the diffusion time from 15 to 50 ms, the diffusion coefficient of water existing in the carbon bed monotonically decreased. The result indicated that the molecular diffusion of solvent molecules were distracted by the closely-packed carbon particles. This approach might be useful to investigate the packing condition of CL obtained through the decal method.

We demonstrated that our approach provided a unique capability to investigate the catalyst ink and its decal process. It will be proved that this technique is useful in investigating many samples prepared at various conditions and better understand the decal process.

Acknowledgment

This paper is based on results obtained from a project commissioned by the New Energy and Industrial Technology Development Organization (NEDO).

References

[1] H.J. Cho et al., Int. J. Hydrogen Energy 36 (2011) 12465.

[2] X. Liang et al., Fuel 139 (2015) 393.

[3] Zamel, J. Power Sources 309 (2016) 141.

[4] S. Ma et al., Solid State Ionics 178 (2007) 1568.

[5] C. Welch et al., ACS Macro Lett. 1 (2012) 1403.

[6] M.H. Levitt, "Spin dynamics" 2nd edition, Wiley (2008).

[7] W.S. Price, "NMR studies of translational motion", Cambridge Univ. Press (2009).

[8] M. Holz et al., Phys. Chem. Chem. Phys 2 (2000) 4740.

[9] P.S. Tofts et al., Magn. Reson. Med. 43 (2000) 368.

[10] R. Li et al., J. Phys. Chem. B 118 (2014) 10156.

Figure 1

2501

and

Widespread commercialization of polymer electrolyte fuel cells (PEFCs) requires improvement in the performance and durability of its key components. Catalyst layer degradation is a major factor contributing to the performance loss [1]. The relatively slower oxygen reduction reaction (ORR) kinetics and mass transport in the cathode contribute to a significant overpotential and thus requires noble metal catalysts to be used. The cathode catalyst layer (CCL) must maintain the heterogeneous catalyst structure for the species transport to/from the reaction sites which facilitates durable operation and low Pt loading. During operation, the heterogeneous CCL structures may experience localized non-uniform potential loads, fuel starvation, development of hot spots, and potential cycling that eventually leads to carbon corrosion, Pt agglomeration, ionomer redistribution, Pt oxidation, and Pt isolation in nano holes [2]. The affected structure may result in reduced catalytic activity, reduced Pt utilization and fuel utilization, inaccessible active sites, inefficient water management, increased electronic resistance, and increased mass transport resistance [3]. Hence, spatially resolved imaging of the catalyst layer and its constituents is required to unravel the effects of elemental and structural degradation. The present study aims to investigate the effects of high voltage excursions, which are typically encountered during start up/shut down conditions, on the heterogeneous nanoscale structure of the cathode catalyst layer.

Membrane electrode assemblies (MEAs) with CCL comprising of a commercial 50:50 wt.% Pt/C catalyst and Nafion ionomer were subjected to a voltage cycling accelerated stress test (AST) with upper and lower potential limits of 1.3 and 0.6 V, respectively. The AST was operated until 4,700 cycles. Further details on the MEAs and AST can be found elsewhere [4]. The potential cycling is intended to induce CCL degradation, similar to that observed during field operation of fuel cells. The low potential operation leads to a higher current to be drawn from the fuel cell and causes temperature rise and flooding of the CCL. High voltage operation on the other hand triggers Pt oxidation and dissolution, carbon corrosion, and ionomer dehydration. Beginning-of-life (BOL) and end-of-test (EOT) MEAs extracted from in-situ AST were analyzed using transmission electron microscopy (TEM). The samples were embedded in epoxy resin and sliced to thin films (~70-90 nm) using ultramicrotome to reveal the cross-section and collected on copper grids for imaging. The nanoscale distribution of solid/void phases and the elemental maps were recorded at high resolution using a Tecnai Osiris TEM instrument from FEI. Elemental mapping of the CCLs was carried out by selectively capturing the generated X-rays with characteristic energy due to electron interaction with the sample. The X-ray energy distinguishes different elements present in the sample. The energy dispersive X-ray (EDX) analysis offers the sophistication of imaging the ionomer phase through the fluorine elemental map and carbon and platinum through their respective elemental maps. The BOL catalyst layer exhibiting uniformly distributed fluorine is presented in Figure 1(A). In contrast, the fluorine map of the EOT CCL (Figure 1(B)) reveals regions of increased concentration and closely compacted ionomer. The carbon and platinum elemental maps (not shown) at EOT reveal vacant carbon and aggregated platinum regions as a result of voltage cycling. The solid/void morphology of CCLs seen from the dark field images also revealed an increased solid content at EOT. Images indicate that the catalyst layer experienced compaction which will affect the porosity and triple-phase regions promoting ORR. The obtained results are in good qualitative agreement with the lower resolution structure visualized by X-ray computed-tomography and the present images confirm the trends observed in the 2-D virtual slices of ionomer dominated and Pt/C dominated catalyst layer regions [5]. Overall, the observed effects of compositional, structural, and morphological degradation of the CCL during voltage cycling contributes to the fundamental understanding of its complex degradation process and can aid development of catalyst layers with improved performance and extended durability.

ACKNOWLEDGMENTS

Research funding provided by Automotive Partnership Canada (APC), Natural Sciences and Engineering Research Council of Canada (NSERC) and Ballard Power Systems is gratefully acknowledged. We also thank Ballard Power Systems for experimental support. This work made use of the 4D LABS shared facilities supported by the Canada Foundation for Innovation, British Columbia Knowledge Development Fund, Western Economic Diversification Canada, and Simon Fraser University.

REFERENCES

[1] M. K. Debe, Nature, 486 43 (2012).

[2] T. Saida et al., Angew. Chem., 124 10457 (2012).

[3] R. Borup et al., Chem. Rev.,107 3904 (2007).

[4] A. P. Hitchcock et al., J. Power Sources, 266 66 (2014)

[5] A. Pokhrel et al., Meeting Abstracts. No. 37. 1357, The Electrochemical Society (2015).

Figure 1

2502

, , , , , , and

Extended structure Pt catalysts are viewed as a promising alternative to state-of-the art catalysts based on carbon-supported Pt nanoparticles due to improved specific activities and durability. However, these types of catalysts posses lower surface areas and demonstrate lower mass activities than traditional catalysts. We have recently developed high surface area PtNi nanowires that meet the requirements for high mass activity, specific activity, and durability. These extended structure PtNi nanowires were synthesized via galvanic displacement producing extended Pt surfaces with high surface areas (>90 m2/g Pt).1Various post-processing treatments have been applied to alter the chemistry and structure of the nanowires in order to optimize their performance. Annealing the nanowires in hydrogen to alloy the Pt to the Ni, acid leaching to remove unalloyed Ni, and annealing in oxygen was found to significantly enhance performance and durability in a rotating disk electrode (RDE) half-cell. The initial mass activity was found to be 7 times higher than Pt/HSC and after durability testing 97% of performance was retained.

Our current efforts focus on optimization of the performance of these extended surface electrode structures in membrane electrode assemblies (MEAs), which requires detailed information about electrode composition and structure. Transmission X-ray microscopy (TXM) allows for non-destructive analysis of the extended surface catalysts electrode structure with 3-D visualization.2,3Another major advantage of the TXM is the ability to selectively image Ni or Pt by tuning the incident X-ray energy to the element's specific absorption edge and thus analyze their respective distribution throughout the entire electrode structure.

A series of MEAs was analyzed to explore the effects of ink formulations (i.e. amount of ionomer, amount and type of carbon, addition of poly(acrylic acid) (PAA)) on electrode structure. MEAs were also analyzed after post-treatments involving acid leaching to remove undesired nickel species. The results showed clear differences in the PtNi nanowire structure with the various ink compositions and post-treatments. Addition of graphitized carbon nanofibers (GCNFs) resulted in significantly less densely packed, but more homogeneous nanowire distribution, as seen in figure 1. Rotation of the MEA confirms the observed structure persists throughout the MEA. Leaching the MEA in acid resulted in a decrease in Ni weight percent and changes in the nickel distribution throughout the MEA. TXM was correlated with analysis from other microscopy techniques, including scanning electron microscopy (SEM), transmission electron microscopy (TEM), and scanning TEM (STEM) combined with energy dispersive elemental mapping. The results from these studies provide guidance for further optimization of extended surface catalysts and electrodes.

Figure 1: (a) PtNi nanowires with PAA and nafion (b) PtNi nanowires with PAA, nafion, and GCNFs

1. S.M. Alia, B.A. Larsen, S. Pylypenko, D.A. Cullen, D.R. Diercks, D. R.; K.C. Neyerlin, S.S. Kocha, B.S. Pivovar, Acs Catal4(2014) 1114-1119.

2. Y. Singh, O. Luo, F. Orfino, M. Dutta, E. Kjeang, Meeting Abstracts 2015,MA2015-02(37), 1355-1355.

3. J. N. Weker, X. Huang, M. F. Toney, Current Opinion in Chemical Engineering 2016, 12, 14-21.

Figure 1

F-22 Oxygen Evolution Reaction - Oct 4 2016 2:00PM

2503

and

Hydrogen production through the electrolysis of water continues to be a key next-generation energy conversion and storage strategy with advantages including minimal greenhouse gas emissions, the lack of fossil fuels, and the potential as an alternative fuel source for fuel cells and reactant for industries such as ammonia production. Recent work by the Boettcher research group and others has demonstrated that iron incorporation into nickel hydroxide based catalysts can significantly enhance the activity of the catalyst for the oxygen evolution reaction (OER) [1-3], which is the half reaction of water electrolysis known to limit the overall process with high overpotential and slow kinetics. With these recent developments in understanding the role of iron incorporation into nickel hydroxide and the importance of iron as the catalytic site for OER within these bimetallic electrocatalysts [1], there is now an opportunity to further develop nanostructured bimetallic iron-nickel hydroxide catalysts for improved OER performance in an alkaline electrochemical environment.

In this presentation, our results on the synthesis, characterization, and electrochemical testing of an iron-nickel core-shell hydroxide nanoparticle catalyst will be presented. Bimetallic iron-nickel nanoparticles were synthesized using a multi-step procedure in water under ambient conditions. When compared to monometallic iron and nickel nanoparticles, the Fe-Ni nanoparticles show enhanced catalytic activity for OER under alkaline conditions (1 M NaOH). The bimetallic nanoparticles demonstrated an improvement in OER overpotential as well as a significant increase in maximum measured current density, as compared to the monometallic iron and nickel nanoparticles. At 1 mA/cm2, the overpotential for the monometallic iron and nickel nanoparticles was 421 mV and 476 mV, respectively, while the bimetallic Fe-Ni nanoparticles had a greatly reduced overpotential of only 256 mV. At 10 mA/cm2, bimetallic Fe-Ni nanoparticles had an overpotential of 311 mV. Electron microscopy and elemental analysis results (Figure 1) will be presented with a detailed discussion of unique aspects of the FeNi nanoparticle catalyst. Results suggest that while the nanoparticles are nominally in a core-shell morphology, there is significant migration of iron into the nickel shell as well as incorporation of some phosphorus into the nanoparticle shell, likely originating from the phosphonate-based stabilizer used during nanoparticle synthesis. X-ray photoelectron spectroscopy suggests that the primary phase of nickel is nickel hydroxide, and x-ray absorption spectroscopy characterization suggests that the primary phase of nickel is the more disordered alpha phase of nickel hydroxide. The presence of a small amount of nickel (oxy)hydroxide is also likely present, based on characterization results.

References

[1] M.S. Burke, L.J. Enman, A.S. Batchellor, S.H. Zou, S.W. Boettcher, Oxygen evolution reaction electrocatalysis on transition metal oxides and (oxy)hydroxides: Activity trends and design principles, Chemistry of Materials, 27 (2015) 7549-7558.

[2] M.S. Burke, S.H. Zou, L.J. Enman, J.E. Kellon, C.A. Gabor, E. Pledger, S.W. Boettcher, Revised oxygen evolution reaction activity trends for first-row transition-metal (oxy)hydroxides in alkaline media, Journal of Physical Chemistry Letters, 6 (2015) 3737-3742.

[3] L. Trotochaud, S.L. Young, J.K. Ranney, S.W. Boettcher, Nickel-iron oxyhydroxide oxygen-evolution electrocatalysts: The role of intentional and incidental iron incorporation, J. Am. Chem. Soc., 136 (2014) 6744-6753.

Figure 1

2504

, , and

Increasing demand for clean energy have triggered researches on alternative energy sources and devices to reduce use of fossil fuel. Hydrogen has been considered as one of the most promising energy source for future due to its high energy density and no air pollutant emission. Splitting water into hydrogen and oxygen is an environmentally friendly method for producing hydrogen gas. This technology can store excess electric energy in the form of chemical bonds of hydrogen, which can resolve an issue about surplus electric power of present renewable energy systems caused by irregular energy source such as airflow and sunlight.

Water electrolysis reaction is divided into two half reactions; hydrogen evolution reaction (HER) and oxygen evolution reaction (OER). High overpotential of both HER and OER is the most significant problems to hamper reaction rate and overall efficiency of water electrolysis, especially OER has much higher overpotential than HER. Therefore, recently major efforts have been devoted to exploring active catalysts for the OER in water electrolysis cell.

Among many kinds of candidate materials for OER catalyst, cobalt (Co) and various Co based materials, including nanostructured Co3O4, CoSe2, Co based perovskites, CoP, CoB and Co/N-doped carbon, have drawn much attention for use in the alkaline water electrolysis system. These Co based catalysts have high OER activity in alkaline media comparable with precious metal based catalysts, such as IrO2 and RuO2. However, previous studies has focused mainly on the exploring desirable composites for low OER overpotential without careful mechanistic study. OER mechanism on Co based catalysts and descriptors for designing more efficient catalysts have been unclear yet.

Herein, we tried to modify the electronic structure of Co through the strain effect. The strain effect modifies the adsorption energy of gaseous or ionic species for OER on Co surface that can be an effective way to design an active electrocatalyst. For making a strain in the Co lattice, we alloyed Co with other transition metals, such as iron (Fe), manganese (Mn), nickel (Ni), and combined with carbide and carbon through a facile solution based process. Graphitic carbon-nitride (g-C3N4) was used as a carbon source. Synthesized alloyed catalyst exhibited higher activity and durability for OER compared with Co and ruthenium dioxide (RuO2) catalysts in alkaline media (0.1 and 1 M KOH). And, the catalysts with tensile strain have much enhanced OER activity than compressive strained one.

2505

, , , and

Polymer exchange membrane (PEM) water electrolysers are based on scarce and expensive platinum nanoparticles materials, seriously affecting the manufacturing cost of commercial electrolyzers. In addition to the cost aspect, studies regarding the long term stability of PEM electrolysers have shown a coarsening of the carbon supported platinum particles leading to an increase of the cell voltage over time (degradation). Unfortunately, the mechanisms responsible for the cathode degradation were not identified in these studies.[1,2] The literature also indicates that the degradation of Pt/C starts at a potential > 0.4 vs. RHE (reversible hydrogen electrode).[3,4,5] However, the working potential of a PEM electrolyser cathode is lower than 0 V vs.RHE, and a degradation of the platinum catalyst on the cathode still occurs. Thus, conversely to fuel cells, it appears that different mechanisms are responsible for the degradation of the platinum active material under cathodic PEM electrolysis conditions.

Here, we employed the Identical Location TEM method to unveil the mechanisms like particle coarsening, detachment, and migration of platinum under cathodic PEM electrolysis conditions. For this purpose we simulated the steady operation mode of an electrolyser in a three electrode setup with a catalyst-coated TEM-Grid as a working electrode under different overpotentials. Identical locations on the TEM-Grid were examined before and after the electrochemical experiments. In addition, the same experiments were conducted with catalyst coated gas diffusion electrodes in order to determine the change of the oxidation states of the catalytic material, and the amount of oxygen containing functional groups via X-ray photoelectron spectroscopy (XPS). In the fuel cell literature it is assumed that this functional groups work as an anchor holding the platinum particles in place. A decrease of these functional groups should induce a migration of platinum particles.[6]

Our study reveals a dependence between the migration of platinum particles and the applied over potential. A more negative potential induces an increase of platinum particle migration leading to the formation of platinum aggregates. The XPS analysis shows no correlation between the applied potential and the amount of functional groups. Thus, the reduction of the functional groups is not responsible for the particle migration. Finally, we postulate that due to a more negative overpotential at the cathode, the hydrogen coverage rate of the platinum particles is increased leading to a bigger gap between the platinum particles and the carbon support. Consequently, the attractive Van-der-Waals forces are weakened, increasing the mobility of the platinum particles.

It is clear from the literature that a reduction of the cathode platinum loading leads to a more negative cathode overpotential.[7] Our study adds a new contribution to this theme, suggesting that the same Pt loading reduction increases the degradation of the cathode.

1) P. Millet, R. Ngameni, S. A. Grigoriev, N. Mbemba, F. Brisset, A. Ranjbari, C. Etievant, Int. J. Hydrogen Energ., 2010, 35(10),5043 - 5042.

2) G. Wei, Yuxin Wang, C. Huang, Q. Gao, Z. Wang, L. Xu, Int. J. Hydrogen Energ., 2010, 35(9), 3951 - 3957.

3) Q. Xu, E. Kreidler, D. O. Wipf, and T. He, J. Electrochem Soc., 2008, 155(3), B228 – B231.

4) Q. Xu, D. O. Wipf, and T. He, Langmuir, 2007, 23 (17), 9098 – 9103.

5) J. Willsau, J. Heitbaum, J. electroanal. chem. interfacial electrochem.,1984, 161(1), 93 – 101.

6) L. Dubau, L. Castanheira, G. Berthome, F. Maillard, Electrochim. Acta, 2013, 110, 273 - 281.

7) H. Gasteiger, J. Power Sources, 2004,127, 162 – 171.

2506

, , , , , and

The development of high-performance catalysts for the oxygen evolution reaction (OER) is critical for cost-effective conversion of renewable electricity to fuels and chemicals. Here we report the significant improvement in the OER activity of electrodeposited NiOx films resulting from the combined effects of using gold as a metal support and cerium as a dopant. This NiCeOx-Au catalyst delivers high OER activity in alkaline media, and is among the most active OER electrocatalysts reported to date. Based on theoretical modelling coupled with experimental observations, we ascribe the activity to a combination of electronic, geometric and support effects, where highly active under-coordinated sites at the oxide support interface are modified by the local chemical binding environment and by doping the host Ni oxide with Ce. The NiCeOx-Au catalyst is further demonstrated in a device environment by pairing it with a nickel-molybdenum hydrogen evolution catalyst in a water electrolyser, which delivers 50 mA consistently at 1.5 V over 24 hours of continuous operation. This work deepens the understanding of complex multi-component catalytic systems by showing how electronic, geometric and support effects can be combined to enhance catalytic activity.

2507

and

Hydrogen is an attractive energy carrier and electrolysis of water is an important route to hydrogen generation. Alkaline water electrolysis is preferred over electrolysis in acidic medium due to the possibility of lowering stack costs and enhancing he library of stable electrocatalyst materials available for the electrochemical reactions. The high anode overpotential arising from the sluggish oxygen evolution reaction (OER) has led to significant interest in developing stable and active OER electrocatalysts. IrO2 (state of the art catalyst), RuO2 and PGM-based pyrochlores are suitable catalyst materials that exist today, but there is benefit in finding cost-effective alternatives.

In this study, perovksite oxides of the form La[Ni(1-x-y)CoxFey]O3 (where 0≤x≤1 and 0≤y≤1) were synthesized by the co-precipitation method. Many of these perovskites exhibited electron conductivities greater than 0.1S/cm, eliminating the need to add carbon for OER studies and implying the likelihood of making conducting electrodes without additives with these materials. The electrocatalytic activity for the OER was studied using a rotating disk electrode (RDE) in 0.1M KOH. A decrease in OER activity was observed with increasing cobalt content in perovskites with y=0. A similar trend was observed for perovskites containing iron with x=0. All the perovskite electrocatalysts exhibited higher onset overpotentials than state-of-the-art-IrO2 but many exhibited significantly higher specific activities than IrO2 at 1.6V vs RHE. Specific activities of 0.21 A/m2, 7.08 A/m2 and 1.01 A/m2 at 1.6V vs RHE and 0.05 A/m2, 0.1 A/m2 and 0.07 A/m2 at 1.5V vs RHE were obtained for IrO2, LaNi0.6Co0.4O3 and LaNi0.6Fe0.4O3 respectively. 

2508

, , , , and

Hydrogen production by proton exchange membrane water electrolysis (PEMWE) is an attractive way to store renewable energy. At the current stage of development, however, this technology suffers from a high price and a relatively low efficiency, in particular due to sluggish oxygen evolution reaction (OER) electrocatalysis. Only iridium oxide electrocatalysts are proven to provide the required longevity of operation with relatively low overpotential of the OER at the same time.1 In this connection iridium based mixed-oxide systems, where scarce iridium is diluted in oxide matrix of less expensive element, are considered as promising ones. Such an approach allows one to decrease the loading of low abundant iridium while improving the overall performance of an anode. For example, utilization of valve metal oxides can improve stability of iridium2 while addition of ruthenium may increase electrocatalytic activity of electrode towards the OER.3 During the last decades different material combinations were suggested in the relevant literature, and many more can be considered feasible.4 In this situation, an effective alleviation strategy in finding novel, more efficient catalysts is using high-throughput approaches for fast material library screening both, for activity and stability.

In the present work we utilize a scanning flow cell (SFC) connected to an inductively coupled plasma mass spectrometer (ICP-MS) as a tool for the simultaneous activity and stability screening of the most promising mixed oxide libraries for the acidic OER. Material gradient libraries covering a broad composition range are prepared by magnetron sputtering. As a working example, performance of iridium and ruthenium mixed oxide libraries with iridium content changing from 0 to 100% is studied in detail. It is shown that the stabilization of ruthenium versus dissolution is a result of the stability of iridium at the surface, but not due to the formation of a common band due to mixing, as reported in literature. The change of composition during the OER is further analyzed by ex-situ XPS analysis. It is found that the OH/O ratio in the oxides is highly affected by ruthenium leaching, a behavior similar to that reported for the iridium nickel system.5 Finally, the obtained information is used to fill the gap in understanding of the effect of composition on stability-activity relationship for iridium and ruthenium based mixed oxide OER catalysts.

References

[1] Dmitri Bessarabov, Haijiang Wang, Hui Li, N. Zaho, PEM Electrolysis for Hydrogen Production, CRC Press, Boca Ranton, 2015.

[2] V. V. Gorodetskii, V. A. Neburchilov, V. I. Alyab'eva, Russian Journal of Electrochemistry, 41, 1111 (2005).

[3] R. Kötz, S. Stucki, Electrochimica Acta, 31, 1311 (1986).

[4] I. Katsounaros, S. Cherevko, A. R. Zeradjanin, K. J. J. Mayrhofer, Angew. Chem. Int. Ed.53, 102 (2014).

[5] T. Reier, Z. Pawolek, S. Cherevko, M. Bruns, T. Jones, D. Teschner, S. Selve, A. Bergmann, H. N. Nong, R. Schlögl, K. J. J. Mayrhofer, P. Strasser, Journal of the American Chemical Society, 137, 13031 (2015).

2509

, , , , and

INTRODUCTION

PEMWE can produce hydrogen gas at high pressussure, directly. The PEMWE at high pressure operation is expected to be a hydrogen gas charger to FCV at car dealer and/or isolated district in prevailing period of hydrogen station. It is expected to be a hydrogen production device to highly utilize sustainable energy. However, hydrogen gas crossover is a problem to address. High pressure difference, such as several tens MPa, imposed through PEM causes a significant amount of crossover, in which hydrogen gas produced at cathode permeates to anode, resulting in decrease of current efficiency. Also, some amount of the hydrogen gas crossed accumulates in anode and anode cylinders, and the accumulation will become safety issue in case of utilizing oxygen gas, such for regenerative fuel cell. In this study, we try to reduce the crossover by controlling the wettability of cathode current collector. The current collector controlled in wettability is expected to enhance the hydrogen bubble separation from there, leading to a less crossover and higher current efficiency. Controlling the wettability changes the contact angle between water, hydrogen bubble and the matrix of current collector, and also changes bubble dynamics, pressure and occupation in cathode current collector, leading to decreasing the crossover and increasing current efficiency. Three current correctors with different wettability are fabricated, and embedded into a PEMWE. Operating the cell under 2 MPa condition and visualizing hydrogen bubble in cathode examines the wettability effect.

EXPERIMENTAL APPATUS

Table 1 shows the cathode current collectors prepared. Every collector is fabricated based on carbon paper. Among of them, two indicate hydrophobicity, and one does hydrophilicity. They are named in hydrophobic, high-hydrophobic and hydrophilic current collector. Because the wettability changes in elapsed time, contact angle before and after the operation is shown in the table. The PEM used is Nafion324, on which IrO2 and Pt/C layer is formed as for anode and cathode electrode. The electrode area is 15.4 cm2. Titanium sintered compact with coating platinum is used for anode current collector.

PEMWE comprising the above components is mounted on a cell evaluation system. Water at 80 and 150 cc/min is supplied into anode and cathode channel, respectively. Produced gas drains through gas-liquid separator. Hydrogen gas produced boosts up to 2 MPaG with back-pressure valve. Applied current is 1 A/cm2, corresponding to 0.065 A/cm2. This density, which is lower than that in practical cases, is expected to highlight the wettability effect. The current efficiency is obtained through the hydrogen flow rate measured with MFM located at lower reach of the back pressure valve and comparison with theoretical flow rate. Hydrogen bubble in cathode channel is visualized by a high speed camera through windows mounted on cathode separator.

RESULTS AND DISCUSSION

Figure 1 shows the current efficiency measured. Higher operation pressure indicates higher pressure difference of hydrogen gas between cathode and anode, resulting in increasing the crossover and decreasing current efficiency. This figure also indicates the wettability effect, in which current efficiency in the hydrophilic case is highest. Comparing the hydrophobic and high hydrophobic case suggests that later case is higher in current efficiency. The wettability effect confirmed is discussed with visualization simultaneously done. Fig. 2 is frequency and diameter when hydrogen bubble separated from the surface of cathode current collector. Highest current efficiency appeared in the hydrophilic case can be explained by the highest frequency in this case. For example, under 0.3 MPaG condition, the frequency in the hydrophilic and hydrophobic case is 30 and 17 Hz, respectively. High frequency appeared in the hydrophilic case implies that hydrogen bubble formed in catalyst layer immediately moves to channel through the collector and that accumulation of hydrogen bubble at the interface between catalyst layer and current collector is suppressed, leading to a less chance of hydrogen crossover. This mechanism explains that hydrophilic current collector functions to increase current efficiency.

The reason why the high hydrophobic case indicates higher current efficiency comparing with the hydrophobic case is possibly explained by the diameter, shown in Fig. 2 (b). The large diameter measured in the high hydrophobic case suggests that hydrogen gas occupies collector pores over a wide region, and that many gas paths form in through-plane direction. The paths result in a smaller pressure difference of hydrogen gas through the collector and smaller local pressure at catalyst layer, where hydrogen gas crossover starts. Thus, high-hydrophobicity contributes to high current efficiency.

Figure 1

2510

and

The increasing demand for solar fuels and hydrogen as most promising energy carrier requires new materials for electro-catalytic splitting of water or the back-transformation of hydrogen into electricity. Especially, the oxygen electrode caused by sluggish kinetics is hampering progress for cheap hydrogen made by electrolysis. Moreover the low abundance and the high costs of current highly active noble metal catalysts require new developments for novel cheap and abundant materials. Therefore we have developed a new two-component catalyst system consisting of Nickel-Iron layered double hydroxide (NiFe-LDH) and Fe-N doped carbon material (Fe-N-C) as non noble bifunctional catalyst for the oxygen electrode in alkaline unitized reversible fuel cells/electrolyzers (URFC). The combination of NiFe-LDH known as most active non-noble catalyst for the oxygen evolution reaction (OER) in alkaline media and Fe-N-C known as highly active catalyst for the oxygen reduction reaction (ORR) resulted in a combined overpotential with their respective advantages. Highly crystalline NiFe-LDH was prepared by a simple and time-saving microwave assisted solvothermal synthesis route. The crystallinity was confirmed by X-ray diffraction (XRD) spectroscopy and Transmission electron microscopy (TEM) verified morphological characteristics. Rotating Ring disk (RRDE) measurement revealed high selectivity for the multi-component catalyst. We further investigated the behaviour in a membrane based reversible electrolyzer using an anion exchange membrane (AEM, Tokuyama A201). Compared to noble catalyst systems such as Iridium and Platinum the membrane electrode assembly (MEA) measurements proved the ability to compete and supported the initial high activities.

2511

, , , , and

Introduction

In order to reduce carbon dioxide emissions, a significant number of renewable energies that are uneven distribution with fluctuation must be introduced. Therefore, to increase renewable energies, energy carrier technology is needed for storage and transportation. Toluene-methylcyclohexane organic chemical hydride system is one of promising technologies as hydrogen storage and transportation. Electrohydrogenation of toluene with water splitting has higher theoretical energy conversion efficiency compare to a series process of water electrolysis and hydrogenation. Cathode side is a cathode membrane assembly with PtRu/C, which is applied PEFC technology. Anode is a dimensionally stable electrode for oxygen evolution reaction in acidic electrolyte using industrial electrolysis technology. In our previous study, we demonstrated good performance of the electrolyzer with hydrophilized membrane; however, hydrogen generated with low concentration of toluene feed, which should be improved (1).

In this study, the effect of the design of toluene feed flow field on the cell voltage and the current efficiency has been investigated to increase conversion ratio from toluene to methylcyclohexane without hydrogen generation.

 

Experimental

A single cell electrolyzer made of titanium with 100 cm2of projected electrode area was used to determine the performance. Figure 1 shows the schematic drawing of flow field for the cathode. A parallel flow, which is usual for industrial electrolysis and liquid electrolyte fuel cells, a serpentine flow, which is conventional for polymer electrolyte fuel cells, and an interdigitated flow have been investigated to improve the performance of the electrolyzer. The anode flow field was parallel.

A cathode was a carbon paper (35BC, SGL) coated 0.5 mgcm-2 of PtRu (TEC61E54, TKK) with Nafion dispersion. The cathode was pressed on a perfluoroethylene sulfuric acid (PFSA) membrane (Nafion® 117, DuPont) for a cathode membrane assembly. The membrane of the cathode side was mechanically hydrophilized. A DSE® anode with IrO2 based electrocatalyst is used for oxygen evolution. Backing of the anode was titanium felt. The anode was uniformlypressed on the membrane by elastic force of the titanium felt.

10 cm3 min-1 of toluene or 50% toluene-methylcyclohexane mixture and 1M (=moldm-3) of H2SO4were supplied to the cathode and anode for hydrogenation of toluene, respectively.

Cell voltage was determined with 4 mVs-1 of voltage sweep from 1 V for toluene hydrogenation up to 0.5 A cm-2of the current density. Current efficiency was determined with constant cell voltage electrolysis with the volume measurement of gas evolution from cathode during the electrolysis.Internal resistance (iR) was determined with higher frequency intercept of AC impedance method.

 

Results and discussion

Figure 1 shows the cell voltage and the current efficiency as a function of the current density for electrohydrogenation of toluene at 60oC with various flow fields for the cathode. 100% of toluene or 50% toluene – methylcyclohexane mixture was fed to the cathode chamber. The internal resistances by AC impedance method were 0.25 Ω cm2 for all electrolyzers, which is almost same as the membrane area resistivity (2). The cell voltages for 50 and 100% toluene feed as a function of current density were almost the same for each flow design of the cathode. The cell voltage of parallel and interdigitated flow was 2.0 V at 0.4 Acm-2, and the cell voltage of the serpentine was a little larger than the other flow patterns. The current efficiency of hydrogenation decreased with the increase of current density, and the order from high current efficiency was the serpentine, interdigitated, and parallel flows. The difference among flow patterns was significantly for 100 % toluene feed, but it was very small for 50 % toluene feed. Turbulence flow would be better for mixing than laminar flow to feed toluene to the catalyst layer, and the order seems to be same to the flow velocity. Flow field design is important to increase toluene conversion. In addition, improvement of catalyst layer should also be important, because current efficiency of 50 % toluene feed seems to be controlled by mass transfer in the catalyst layer. 

Acknowledgment

This work was supported by Council for Science, Technology and Innovation (CSTI), Cross-ministerial Strategic Innovation Promotion Program (SIP), "energy carrier" (Funding agency: JST). The Institute of Advanced Sciences (IAS) in YNU is supported by the MEXT Program for Promoting Reform of National Universities. We appreciate the person concerned them.

 

References

1) S. Mitsushima, Y. Takakuwa, K. Nagasawa, Y. Sawaguchi, Y. Kohno, K. Matsuzawa, Z. Awaludin, A. Kato, Y. Nishiki, Electrocatalysis, 2016, 7, 238.

2) S. Slade, S. Campbell, T. Ralph, F. Walsh, J. Electrochem. Soc., 2002, 149, A1556.

Figure 1

2512

, , and

The design of cost-effective and highly active, durable electrocatalysts for the sluggish oxygen evolution reaction (OER) is critical for promoting various energy conversion process such as water splitting and rechargeable metal-air batteries. Due to a wider selection of earth-abundance and low cost catalyst candidates (e.g., mixed 3d-transition metal oxides), OER in alkaline environment has attracted considerable interests during past years. Besides the recently developed Ni-Fe layered double hydroxides (LDH) catalysts, spinel phase transition metal oxides have also drawn particular interests owing to their well-defined rigid structure and thus potentially high stability during electrocatalysis. While improved OER activities have been demonstrated, currently most spinel transition metal oxides (mainly Co-based spinel oxides) still underperformed the state-of-the-art benchmark OER catalyst IrO2and the Ni-Fe LDH catalysts. Moreover, atomic insights into the surface structure that governs their OER reactivities and stabilities still remain very limited.

We report here the exploration of iron-based spinel oxide nanoparticles (MxFe3-xO4, M= Mn, Fe, Co, Ni, Cu) as a new class of alkaline OER electrocatalysts with potentially both high electrocatalytic activity and high stability. Emphasis will be placed on atomistic understanding on the surface cation chemistry and surface atomic structures of the spinel oxide nanoparticles and their correlations with OER activities and stabilities. This will be achieved by comparative microscopic and spectroscopic studies of the nanoparticle surfaces before and after OER electrocatalysis using high-resolution (scanning) transmission electron microscopy, electron energy loss spectroscopy and X-ray photon spectroscopy, which will be further complemented by density functional theory calculations. The results will provide important insights into the structure-activity-stability relations of spinel phase transition metal oxides as well as other mixed metal oxides for OER electrocatalysts.

2513

, , and

The electrocatalytic oxygen evolution reaction (OER) is a critical anode reaction that is often coupled with an electronic/photoelectronic water splitting reaction or used in rechargeable metal-air batteries for renewable energy conversion and storage. [1] However, the sluggish OER reaction kinetics require the use of efficient electro-catalysts, such as metal oxides (e.g. IrO2, RuO2, MnO2, Co3O4, NiFeOx, etc.) or metal-free materials (e.g. N- and P- doped graphene, MWCNT, graphitic carbon nitrite, etc.). To enhance the electro-catalytic activity of OER, great efforts have been devoted to the design of hybrid materials with functionalities tuned by tailoring their micro/nano structure. In addition, pore structure and surface chemistry of supporting materials also play an important role in determining the efficiency of OER catalysis. Traditionally OER catalysts are deposited on a plenary glassy carbon electrode or indium titanium oxide (ITO) glasses, of which the low surface area greatly limits the catalyst loading. Additionally, owning to the lack of surface functionality, these substrates interact weakly with the OER catalyst, adversely affecting OER electro-catalytic activities. More importantly, the bulky and rigid electrodes make them incompatible for broad applications in energy conversion and storage systems.

Herein, we present a flexible, free standing oxygen electrode featuring 3-D architectures of high activity and durability for OERs. Conductive microfibers and paper sheets were fabricated using layer-by-layer (LbL) nano-assembly techniques, in which the cationic poly(ethyleneimine) (PEI) has been used in alternate deposition with anionic conductive PEDOT–PSS and solubilized CNT–PSS on lignocellulose wood microfibers [2]. By creating alternating layers of oppositely charged components on the surface of wood microfibers, we have produced a nanocoating of 20–150 nm thickness that enables the microfibers to exhibit electrical conductivity. Moreover, the surface charge of resultant paper can be well tuned by the LbL polyelectrolyte coating. P-, N- dual doped 3-D graphenes were then prepared with an imposing opposite charge to the composite paper. When casting functional porous graphenes onto the composite paper, it is observed that they interconnected, driven by the strong electro-static interaction. The resulting graphene coated paper, which shows high porosity and functionality, was then employed as the host framework, in which a various types of metal oxide particles of high OER activity were decorated in-situ. The obtained oxygen electrode of this hierarchy structure has exhibited improved mass and ion transport and OER efficiency was significantly enhanced. In summary, the developed smart paper consisting of orderly assembled P-, N- doped graphene layers incorporated with active OER catalyst particles performs as a flexible oxygen electrode that shows great promise in being directly applied in flexible electronic devices for efficient energy storage and conversion.

References:

[1] Y. Jiao, Y. Zheng, M. Jaroniec, S.Z. Qiao, Design of electrocatalysts for oxygen- and hydrogen-involving energy conversion reactions, Chemical Society Reviews, 44 (2015) 2060-2086.

[2] M. Agarwal, Q. Xing, B.S. Shim, N. Kotov, K. Varahramyan, Y. Lvov, Conductive paper from lignocellulose wood microfibers coated with a nanocomposite of carbon nanotubes and conductive polymers, Nanotechnology, 20 (2009).

A-21 Degradation - Oct 5 2016 8:00AM

2514

, , and

The estimation and increase of the lifetime of PEMFC fuel cell under dynamic conditions is one of a major challenge about PEMFC. Increasing the durability of the fuel cell must be treated by both the development of new material and design but also by optimal strategies and control of the operating conditions of the fuel cell. To validate some new strategies for lifetime and also the durability of new material, experimental tests must be coupled with modeling and numerical simulations. Indeed, the different irreversible degradation mechanisms in the MEA (Membrane Electrode Assembly) (as the platinum dissolution, Ostwald ripening, carbon support corrosion and chemical membrane degradation) and reversible degradation mechanisms (platinum oxidation, liquid water management) are strongly coupled to the local conditions into the MEA. Moreover, the local conditions into the MEA are depending of the material properties and the operating conditions of the fuel cell stack. The strong coupling of the different phenomena and the heterogeneities along the surface of the cell and through the thickness of the MEA can be solved only by numerical simulations.

In this talk, to predict the degradation mechanisms along the surface of the cell and function of the dynamic operating conditions, three multi-physics models are linked together (see Figure). The three models are described below.

i/ EDMOND model is a 0D double layer model to calculate the local surface potential at the surface of the catalyst as well as the coverage of the various reaction intermediates, based on a dynamic coupling between the local operating conditions and the kinetics of the various reaction steps. Both the surface potential and the coverage are involved in the mechanistic models, which makes the EDMOND framework required for such a modeling approach.

ii/ The MEA model is a 2D CFD model cross-section of the MEA. The model is able to compute the differences in operation under the rib and channel of the bipolar plate. It takes into account the gas diffusion, thermal, ionic and electrical transport and electrochemical response based on Butler-Volmer approach. Anisotropy and compression of the materials are also taken account. The catalyst layer is meshed through the thickness.

iii/ The fuel cell model (called PS++ code) is a 2D+1D fuel cell model, based on bond graph approach. PS++ is a dynamic multi-physic model taking into account two phases flow and heat transport equations. The model is used to calculate the local conditions along the surface of the cell function of dynamic operating conditions. Degradation mechanisms are added (by bottom-up or top-down approach) to estimate the fuel cell lifetime.

The irreversible and reversible degradation mechanism are modeled in the local model and up-scaled into the cell model PS++. The local double layer model (EDMOND) is used to calculate the irreversible degradations of the loss of catalyst (Ostwald ripening). The Ostwald ripening model integrated in EDMOND relies on a multiscale mechanistic approach where parameters come from DFT calculations. It is able to calculate the dependency of both the size and the distribution of the particles, the voltage and the hydration in the degradation rate. The MEA model is used to calculate the reversible degradations. The catalyst oxidations are introduced and are in competition with the electrochemical reactions (ORR and HOR), allowing to compute dynamically the evolution of the active catalyst surface.

The upscaling of the degradations and the activity of the catalyst are introduced into the cell model to simulate the loss of performance of the cell under dynamic solicitations and to calculate the effects of local degradations on the repartition of the current density along the surface of the cell.

The experimental validation of the approach is realized by two 2000 hours experimental tests in a 30 cells stack. The experimental tests have been carried out in a fuel cell stack durability tests with a permanent current density for the first test and a dynamic solicitation for the second one. Current density measurement card and periodic electrochemical characterizations are realized during the tests and post-mortem characterizations at the end of the tests. The results show a good qualitative agreement with the model simulations, in particular to the current density distribution function of ageing and the reversible and irreversible mechanisms.

In conclusion, the approach of lifetime prediction by a full multi-scale approach is demonstrated and open new perspectives about the coupling of modeling and experimental tests to increase the durability of the fuel cell.

Figure 1

2515

, , , , and

A major problem with performance for transportation applications is associated with degradation. The performance degradation losses can be separated into two categories: irreversible and recoverable losses. Significantly more work has examined the irreversible degradation losses compared with the recoverable losses because the recoverable losses often can be simply recovered by performing polarization curves[1]. Delineating the component contributions of recoverable performance losses and mitigation of these losses appears to be a growing concern for fuel cell developers. While these degradation mechanisms are recoverable, recovery processes are not necessarily trivial to accomplish in operando in a vehicle.

MEA performance degradation is sensitive to operating conditions and both irreversible and reversible materials changes. Some known recoverable losses for PEMFCs include the loss of cathode activity due to surface oxide (hydroxide) formation at high potentials,[2] catalyst poisoning by membrane degradation products [3] and recoverable transport losses due to water transport [1,4].

Platinum oxidation at the cathode results in decreased ORR (Oxygen Reduction Reaction) and with Pt often considered to have two different Tafel slopes; one for a metallic Pt surface, and one for an oxidized Pt surface. In other cases, indications are that membrane degradation products, such as (bi)sulfate are readily adsorbed onto the catalyst and are a large source of recoverable degradation. These require removal from the catalyst surface and then from the electrode layer for the performance recovery.[5]

To examine the relative contributions to recoverable degradation, we have conducted in situ and ex situ experiments to separate the effects of Pt oxidation, (bi)sulphate adsorption, water management plus other operational effects including spatial degradation mapping, potential cycling and different levels of RH. One method to identify the effect of membrane degradation products on the performance decay is to compare the effect of non-chemically-stabilized membranes versus the effect of chemically-stabilized membranes. This is shown in Figure 1, where the OCV is shown during two-24-hr periods (with recovery in-between the 24-hr periods) for MEAs with (a) a non-chemically stabilized membrane and (b) a chemically stabilized membrane. Past results have shown significantly more degradation products of fluoride and sulphate anions without the chemical stabilization. Both MEAs show decreasing OCVs, however the decay is significantly more for MEA using the non-chemically stabilized membrane. Segmented cell measurements, made in a 10x10 co-flow cell, do not indicate a substantial spatial difference in the rate of decay during these types of tests. Various methods of recovering the losses were examined to separate the Pt oxidation from the other effects, e.g. decreasing the potential to 0.6 V to reduce the Pt oxide films compared with much lower recovery voltages to desorb adsorbed anions. Recovery protocols in fuel cell mode (H2/air) are also compared without current generation (H2/N2) at potentials of 0.6 V, 0.4 V, 0.3V and 0.2 V including with and without liquid water injection. Liquid water injection was found to be detrimental to the recovery of the fuel cell performance and a voltage of 0.3 V required to good recovery with little difference below that potential.

Acknowledgments

This research is supported by DOE Fuel Cell Technologies Office, through the Fuel Cell Performance and Durability (FC-PAD) Consortium; Fuel Cells program manager: Dimitrios Papageorgopoulos.

References:

[1] Cleghorn, S. J. C.; Mayfield, D. K.; Moore, D. A.; Moore, J. C.; Rusch, G.; Sherman, T. W.; Sisofo, N. T.; Beuscher, U., A polymer electrolyte fuel cell life test: 3 years of continuous operation, Journal of Power Sources (2006), 158 (1), 446-454.

[2] Yu Seung Kim and Piotr Zelenay in Polymer Electrolyte Fuel Cell Durability, eds. Felix N. Büchi, Minoru Inaba, Thomas J. Schmidt, Springer New York (2009) pp 223-240

[3] Yu Seung Kim, Melinda Einsla, James E. McGrath, and Bryan S. Pivovar, The Membrane–Electrode Interface in PEFCs: II. Impact on Fuel Cell Durability Journal of The Electrochemical Society, 157 11 B1602-B1607 2010.

[4] Qing Li, Dusan Spernjak, Piotr Zelenay, Yu Seung Kim, Micro-crack formation in direct methanol fuel cell electrodes, Journal of Power Sources 271 (2014) 561e569

[5] Jingxin Zhang, Brian A. Litteer, Frank D. Coms, and Rohit Makharia, Recoverable Performance Loss Due to Membrane Chemical Degradation in PEM Fuel Cells, Journal of The Electrochemical Society, 159 (7) F287-F293 (2012)

Figure 1

2516

, , , , , , and

The cost and durability of Polymer Electrolyte Membrane Fuel Cells (PEMFCs) are the two major barriers to the commercialization of these systems for stationary and transportation power applications.1 Given the stringent performance requirements that PEMFCs have to meet for automotive applications, Pt-alloy catalysts are being considered as cathode catalysts. While PtCox alloy catalysts exhibit better mass activity than Pt/C catalysts and can meet the beginning of life performance requirements, their durability is not proven. Several studies have indicated that the base metal leaches out of these catalysts over time and any benefits due to alloying is lost over the lifetime of the Membrane Electrode Assembly (MEA).2

The U.S. DOE Fuel Cell Tech Team has recommended various ASTs for PEMFC components.3,4 Los Alamos National Laboratory (LANL) has previously reported on the durability of several Pt/C catalysts during both simulated drive cycle testing and accelerated stress tests (ASTs).5 A catalyst AST that involved applying a square wave with 3 sec at 0.6V and 3 sec at 0.95V was determined to be an appropriate AST for evaluating the durability of Pt electrocatalysts.6In this study the performance of PtCo/C cathode catalyst based MEAs subjected to this AST were evaluated.

The durability of PtCo/C catalysts was found to be comparable or better than that of the Pt/C based catalysts. Moreover the activity of the PtCo/C catalyst based MEA was also enhanced with respect to the Pt/C MEA. Figure 1 illustrates the evolution of the electrochemical surface area (ECSA) over 30,000 cycles of the square wave AST. The PtCo/C based MEA exhibited little change in ECSA over the course of the test. The ECSA initially increased due to insufficient conditioning and then decreasing due to catalyst particle size increase. In contrast, the Pt/C catalyst showed significant increase in ECSA over the course of the test. Finally the starting ECSA of the PtCo/C catalyst based MEA was significantly lower than that of the Pt/C MEA, while the end of test (EOT) ECSA was almost identical. These results are consistent with the small (≈ 2nm) starting particle size of the Pt and relatively large (≈ 5nm) starting particle size of the PtCo catalyst.

The performance of the two MEAs before and after the durability AST is presented in Figure 2. It should be noted that the Pt/C based MEA had a loading of 0.15mg.Pt/cm2 whereas the PtCo/C MEA had an initial loading of 0. 21mg.Pt/cm2. The beginning of test (BOT) performance of the PtCo/C based MEA was significantly better in the kinetic region due to both the alloying effect and higher Pt loading. However, the BOT performance of the Pt/C MEA was better in the mass transport region due to extra mass transport limitations observed in alloy catalysts. The End of test (EOT) performance of the PtCo/C based MEA is only slightly higher than that of the EOT performance of the Pt/C based MEA, consistent with the fact that most of the Co has leached out of the PtCo catalyst.

In this presentation the durability implications of using PtCo alloy catalysts will be discussed in detail with results presented from TEM, SEM, XRF, XRD and electrochemical testing.

References

1. R. Borup, et al., Chemical Reviews, V. 107, No. 10, 3904-3951 (2007).

2. S.C. Ball, S.L. Hudson, B.R.C. Theobald and D. Thompsett, "PtCo, a Durable Catalyst for Automotive Proton Electrolyte Membrane Fuel Cells?" ECS Trans., 11 (1), 1267 (2007).

3. DOE Cell Component AST and polarization curve Protocols for PEM Fuel Cells (Electrocatalysts, Supports, Membranes and MEAs), Revised December 16, 2010.

4. N. L. Garland, T.G. Benjamin, J. P. Kopasz, ECS Trans., V. 11 No. 1, 923 (2007).

5. R. Mukundan, G. James, J. Davey, D. Langlois, D. Torraco, W. Yoon, A. Z. Weber, and R. Borup, "Accelerated Testing Validation" ECS Trans., 41 (1), 613 (2011)

6. R. Borup et al. DOE Hydrogen and Fuel Cell Program, FY15 Annual Progress Report. https://www.hydrogen.energy.gov/pdfs/progress15/v_e_1_borup_2015.pdf

Acknowledgements

The authors wish to acknowledge the financial support of the Fuel Cell Technologies Program and Technology Development Managers: Dimitrios Papageorgopoulos and Nancy Garland. The authors also wish to acknowledge SGL Carbon for the GDLs used in this study.

Figure 1

2517

, , , , , , , , , et al

The polymer electrolyte membrane (PEM) fuel cell is a promising alternative power source for electric vehicles, forklifts, and stationary backup power due to its high power density, low operating temperature, and clean emissions. The gas diffusion layer (GDL) facilitates the transport of heat, reactant gases, water vapor, and liquid water in the PEM fuel cell. The GDL is a highly porous material composed of dense arrays of carbon fibers, resins, and polytetrafluoroethylene (PTFE). Additionally, it is typically coated with a micro porous layer (MPL), which results in the reduction of contact resistance at the catalyst layer interface and the overall improvement in water management. Effective water management must be achieved within the GDL over the entire operating lifetime in order to improve the commercial success of PEM fuel cells. Investigations into the effects of ageing on water management within the GDLs can be useful tools to guide this success.

Water is required to maintain polymer electrolyte membrane hydration and ionic conductivity; however, excess water tends to accumulate in the catalyst layer, GDL, and flow channels. This accumulation limits mass transport and leads to performance losses. Therefore, effective water management in the GDLs is essential for improving cell performance, durability, and stability. Most GDL research has been focused on the impact of GDL materials and microstructure design on overall performance, rather than durability. Decreased GDL hydrophobicity has been found after only a few hundred hours of operation (1, 2). Therefore, a clear understanding of the influence of GDL degradation on the liquid water behavior is essential.

As a tool to visualize this behaviour, synchrotron X-ray radiography is a powerful technique for visualizing liquid water transport behavior within an operating PEM fuel cell. High intensity X-rays generated at the BMIT-BM beamline of the Canadian Light Source were used for identifying liquid water inside the GDLs and capturing the multiphase flow behavior in an operating PEM fuel cell. Figure 1 is an example synchrotron X-ray radiograph of the fuel cell in operando, illustrating the spatial distribution of liquid water. The observed pixel brightness was correlated to the presence of liquid water, and the saturation profile across the GDL thickness was obtained.

In this study, as-received fresh GDLs were immersed in a 35wt % solution of H2O2 at 90o C for 12 hours in order to perform an accelerated artificial ageing process. As-received fresh and artificially aged GDLs were then used to assemble specialized fuel cells for synchrotron imaging. Both the electrochemical performance and liquid water transport behavior were observed for a range of current densities. Additionally, the effect of changes in relative humidity on performance was compared to investigate the effect of artificial ageing. Notable differences in cell performance and liquid water transport behavior were observed at steady state operating conditions. The cell with aged GDLs showed lower cell voltage and higher quantities of liquid water than those of as-received fresh GDLs. The impacts of aging on the liquid water distributions within the GDL and on cell performance will be presented in this work. The objective of current work is to highlight the potential degrading effect on liquid water management in long term fuel cell operation. 

References: 

1. S. Kandlikar, M. Garofalo and Z. Lu, Fuel Cells., 11, 6 (2011).

2. P. K. Das, A. Grippin, A. Kwong and A. Z. Weber, J.Electrochem.Soc., 159, 5 (2012).

Figure 1

2518

, , , , , and

The use of a three-electrode cell, in which the potential of a working electrode can be accurately measured vs. a reference electrode (RE), is a standard practice in liquid-phase electrochemical experiments. However, the geometric constraints of a working PEMFC make use of a RE challenging. Incorporation of a RE into an operating PEMFC typically requires experimental compromises that prevent precise and accurate measurements from being obtained, and do not allow accurate spatial resolution of potential variation along the electrodes. As a result, PEMFC experiments are often performed using the PEMFC anode, which remains within a few mV of 0 V vs RHE during typical operation, as a pseudo RE.

Recently, the National Physical Laboratory (NPL) developed an effective technique for accurate in situ monitoring of local electrode potentials in a PEMFC. By inserting REs from the back side of the MEA, the NPL configuration eliminates the edge effects, lack of anode vs. cathode specificity, Ohmic losses, and current density distribution disruptions that have compromised previous fuel cell RE experiments.

We report here on the use of this RE technique to monitor local electrode potentials in an operating PEMFC during exposure to CO gas (Figure 1). CO is a common impurity in H2 gas and is known to poison Pt-based anodes. Levels of CO present in H2 fuel are expected to be in the ppb range for cells operating on pure H2 and the ppm range or higher for cells operating on reformate. Use of an RE can reveal the distribution of CO losses, and can also enable losses on the anode to be distinguished from those on the cathode. The results of this work demonstrate a new diagnostic capability that enables improved understanding of local conditions during PEMFC operation during transient or steady state operation.

Figure 1. Cell voltage (orange) and anode potential (blue) of a MEA operating at 1 A/cm2 with anode exposed to 20 ppm CO. CO was removed from the anode feed at t = 47 min. Gore MEA, with anode/cathode loading of 0.4/0.1 mg/cm2, 80°C, 100% RH, 100 kPaabs.

Figure 1

2519

, , , and

Polymer electrolyte fuel cells (PEFCs) have been growing in popularity as an alternative energy source for a multitude of applications, including the automotive industry. The use of fuel cells in these applications requires long-term durability with minimal degradation to be cost competitive against conventional technology sources. Current targets for automotive applications are >5,000 hours, under realistic operating conditions. One primary degradation pathway associated with these operating conditions is that of cathode catalyst support corrosion, which occurs due to the oxidation of carbon support for platinum nanoparticles, leaving the platinum unsupported and inactive. The pathway for this degradation mechanism is at elevated cathode potentials greater than 1.2 VRHE, where significant carbon corrosion, in the presence of water, occurs at rates high enough to cause significant structural degradation effects. These elevated potentials can occur during fuel starvation or gas switching during start-up and shutdown procedures [1].

A significant amount of effort has been devoted toward research regarding cathode degradation rates and mitigation, primarily using methods such as electrochemical impedance spectroscopy (EIS), cyclic voltammetry (CV), and scanning electron microscopy (SEM). These methods provide information on a global, two-dimensional perspective. That is, SEM typically provides information on catalyst layer thinning by 2-D cross-sectional images, and EIS and CV provide overall electrochemical changes affected by changes in surface area from carbon corrosion. The usefulness of the proposed X-ray computed tomography (XCT) technique is in its non-invasive nature, excellent spatial resolution, and three-dimensional imaging abilities. These advantages allow for observations to be made on a local and global level providing further insight into the cathode catalyst layer degradation process. Typically research regarding XCT is performed at a synchrotron beamline which is significantly limiting in that it is expensive, impractical and available in only short time intervals. This means that investigating in-situ degradation effects is extremely difficult. With advances in commercial X-ray sources, optics and detectors for laboratory use, many researchers are now being able to take advantage of the power of XCT scans with much lower cost and increased availability.

In this work, we present an investigation toward understanding cathode catalyst layer degradation mechanisms through in-situ visualization by commercial XCT using a fully functional dual-channel, small-scale, fuel cell fixture [2]. This small-scale fixture allows imaging of the membrane electrode assembly (MEA) at multiple stages of its lifecycle during an accelerated stress test, targeting the cathode catalyst layer, in this case causing carbon corrosion. Differences under land and channel are investigated as well as water distribution, which is shown to have a significant effect on the degradation rate using a sample containing catalyst layer cracks. Image processing techniques used to obtain quantitative results are discussed which include histogram deconvolution and thresholding. Figure 1 shows the thresholding procedure results, indicating differences found under land and channel. Continued research using this tool hopes to further our understanding of the interconnectivity within a fuel cell.

Acknowledgements

Funding for this research was provided by the Natural Sciences and Engineering Research Council of Canada, Canada Foundation for Innovation, British Columbia Knowledge Development Fund, and Ballard Power Systems through an Automotive Partnership Canada grant.

References

[1] A. Young, V. Colbow, D. Harvey, E. Rogers, and S. Wessel. J. Electrochem. Soc. 160 (4) F381-F388 (2013)

[2] R. White, M. El Hannach, O. Luo, F. Orfino, M. Dutta and E. Kjeang. ECS abstract 58394, presented at 228thECS meeting, Phoenix, AZ. Oct. 11-16, 2015

Figure 1: a) 3D visualization of a full MEA highlighting local thresholding ability b) 2D segmented area of cathode catalyst at BOL and EOL c) plot showing area fraction of solid to crack change during accelerated stress test d) cross-section of regions used in calculations for the plot above. The image shows a single slice from the End of Life (EOL) sample 3D reconstruction

Figure 1

2520

, , , and

Air pollution is a challenge for the commercialization of proton exchange membrane fuel cells (PEMFCs) due to the detrimental impact of pollutants on cell performance and durability (1-5). Some of these pollutants adsorb onto the Pt surface and compete with the crucial oxygen reduction reaction (ORR). Contaminants such as chloride anions also undergo an irreversible complexing reaction with Pt (5).

We have previously reported how chlorobenzene, a halocarbon and commodity production intermediate inhibits the ORR and causes rapid and significant performance loss when introduced into the cathode of a PEMFC (6). Chlorobenzene adsorption, reactions and molecular orientation on the Pt surface depend on the electrode potential. Cl- is created and remains in the membrane/electrode assembly (MEA). Cl-binds to the Pt surface much more strongly than chlorobenzene but is slowly flushed out by liquid water.

We have recently carried out PEMFC-cathode poisoning studies with bromomethane, another halocarbon, is a solvent and a chemical manufacture precursor (7). Chlorobenzene and bromomethane have an electron withdrawing halogen moiety. The Br- adsorbate on Pt is more strongly bound than Cl- in HClO4(8). These observations suggest that bromomethane and chlorobenzene may share similar features during PEMFC contamination, although methyl and aromatic group adsorption differ. A comparison between these 2 species supports the development of the bromomethane contamination mechanism.

Bromomethane contamination was investigated with a 50 cm2 active area single cell and a constant current of 1 A cm-2. Electrochemical impedance spectroscopy, cyclic voltammetry, linear sweep voltammetry and polarization measurements were used to characterize the temporary effect and the permanent performance loss. X-ray absorption spectroscopy was carried out in a spectro-electrochemical cell with similar procedures as for the chlorobenzene contamination study, to identify chemical changes to Pt and bromomethane and adsorbates on Pt (6,7). A contamination mechanism and a performance-recovery method were derived from these results.

While chlorobenzene poisoned the PEMFC within an hour and performance could be largely recovered with voltammetric cycling, poisoning with bromomethane took days and the losses to the PEMFC were less reversible. We surmised that unlike chlorobenzene, bromomethane is hydrolyzed to bromide and methanol before reaching the Pt surface, as shown in Figure 1a. Bromomethane also permeates through the ionomer film and is subsequently hydrolyzed on the Pt electrode. The methanol product is readily oxidized on the Pt surface, thus favoring the hydrolysis. Bromide anions outside the ionomer film cannot access the Pt surface due to Donnan exclusion. Bromide anions created at the Pt interface are difficult to remove for the same reason.

X-ray absorption near edge structure analysis revealed identical mechanisms for bromomethane and chlorobenzene adsorption on Pt above 0.3 V vs SHE. However, figure 1b illustrates that during bromomethane and bromide adsorption, the Br atom is in contact with the Pt surface within the accessible cathode potential range (0–1 V vs SHE). In contrast, for chlorobenzene, the aromatic ring is lying on the Pt surface below 0.3 V vs SHE. At cathode potentials near or lower than the point of zero charge (PZC), bromomethane, bromide and chloride desorb from the Pt surface which facilitate their removal by dissolution into liquid water in the catalyst layer. These considerations were synthesized into an effective method to recover the remaining cell performance loss after contamination injection was interrupted.

We conclude that contamination research is still an important research field as such two seemingly similar poisoning compounds behave so differently in a practical PEMFC.

Acknowledgments

The authors are grateful to the United States Department of Energy (award DE-EE0000467) and the Office of Naval Research (award N00014-13-1-0463) for financial support of this project. The authors are also grateful to the Hawaiian Electric Company for their ongoing support to the operations of the Hawaii Sustainable Energy Research Facility.

References

[1] R. Borup, J. Meyers, B. Pivovar, et al., Chem. Rev. 107, 3904 (2007).

[2] M. Debe, Nature 486, 43 (2012).

[3] J. Moore, P. Adcock, J. Lakeman, G. Mepsted, J. Power Sources 85, 254 (2000).

[4] Y. Garsany, O. Baturina, K. Swider-Lyons, J. Electrochem. Soc. 154, B670 (2007).

[5] O. Baturina, A. Epshteyn, P. Northrup, K. Swider-Lyons, J. Electrochem. Soc. 158, B1198 (2011).

[6] Y. Zhai, O. Baturina, D. Ramaker, et al., J. Phys. Chem. C 119, 20328 (2015).

[7] Y. Zhai, O. Baturina, D. Ramaker, et al., Electrochim. Acta, submitted.

[8] N. Marković, P. Ross, Surf. Sci. Rep. 45, 117 (2002).

Fig. 1. a) Proposed bromomethane and Br- transport paths and bromomethane hydrolysis reaction, b) adsorption configurations of bromomethane and chlorobenzene on the cathode catalyst of a PEMFC under different potentials.

Figure 1

2521

and

We have heard it over and over again – internal combustion engines (ICE) operated with fossil fuels can cause severe damage to our health and environment.1 Nevertheless, they are still very widespread and will not disappear until a compelling alternative is found. One possible alternative is the low temperature polymer electrolyte fuel cell (LT-PEFC). They are commonly operated with hydrogen and air and only emit water. However, before LT-PEFCs can replace ICEs on a large scale mainly two barriers need to be overcome. Those barriers are cost and lifetime. Hence, reducing the costs of a fuel cell and increasing its lifetime are the two most relevant research fields regarding fuel cell commercialization. Whereas the costs are mainly dictated by the raw material prices (e.g. platinum) and the manufacturing process the reasons for limited lifetime are numerous.

PEFCs in general are affected by various degradation effects at different locations of the fuel cell.2,3 The membrane suffers for example from pinhole formation and membrane thinning. The hydrophobicity of the gas diffusion layer decreases over operation time which decreases its water transport ability. The bipolar plates are prone to surface corrosion which reduces their conductivity. The catalyst and catalyst layer also represent a major degradation area. Some specific degradation mechanisms are carbon corrosion, catalyst detachment and catalyst particle growth. All degradation effects are complex functions of the specific operational parameters, making it challenging to identify them individually. This in turn is necessary to gain insights into limitations in fuel cell lifetime and is the key to successful development of mitigation strategies to reach a desired lifetime of >5,000 hours for automotive drive cycle operation.

There are different techniques available to identify and characterize the individual degradation mechanisms. Some common ones are polarization curve or electrochemical active surface area measurements which are based on electrochemistry. Furthermore, imaging techniques like secondary electron microscopy (SEM) are frequently used for degradation analysis. Another established imaging technique is X-ray tomography. The main advantages of X-ray tomography over SEM imaging are its ability to capture the fuel cell area of interest as a whole and that the tomography scan can be recorded nondestructively and in-operando.4

For some time, fuel cell X-ray tomography was mostly carried out at large scale synchrotron beamline facilities. Unfortunately, they tend to have only very limited access and therefore only few experiments can be conducted in a short time period. In recent development though, lab sized X-ray microscopes have gained interest in the community.5 Such microscopes have excellent availability and therefore enable the possibility to carry out long term experiments.

The present contribution will utilize this particular advantage. Our unique approach is to retrace fuel cell degradation with X-ray tomography imaging (cf. Figure 1) over time and correlate it with individual failure mechanisms. In particular, electrode degradation will be discussed. MEAs, with e.g. different carbon/platinum ratios, were exposed to accelerated stress tests and imaged at beginning and end of life. With the aid of X-ray tomography, we are able to segment the MEA and subsequently locate and analyze the degradation effects at the electrode. In combination with electrochemical measurements, exclusive in depth insights of fuel cell degradation are provided. In addition, a novel approach for post X-ray tomography processing is discussed. In summary, the combined findings lead to an improved understanding of LT-PEFC degradation which is a necessary first step in order to investigate new mitigation strategies to increase the fuel cell lifetime.

Figure 1. A membrane electrode assembly visualized by X-ray computed tomography recorded at beginning of life.

1. S. M. Platt et al., 'Two-stroke scooters are a dominant source of air pollution in many cities', Nat. Commun., 5, p. 3749, 2014.

2. J. Wu et al., 'A review of PEM fuel cell durability: Degradation mechanisms and mitigation strategies', J. Power Sources, 184, pp. 104–119, 2008.

3. T. Engl, L. Gubler, and T. J. Schmidt, 'Think Different! Carbon Corrosion Mitigation Strategy in High Temperature PEFC: A Rapid Aging Study', J. Electrochem. Soc., 162, pp. F291–F297, 2015.

4. J. Eller et al., 'Progress in In Situ X-Ray Tomographic Microscopy of Liquid Water in Gas Diffusion Layers of PEFC, J. Electrochem. Soc., 158, B963 (2011).

5. M. Andisheh-Tadbir, F. P. Orfino, and E. Kjeang, 'Three-dimensional phase segregation of micro-porous layers for fuel cells by nano-scale X-ray computed tomography', J. Power Sources, 310, pp. 61–69 (2016)

Funding for this research was provided by the Natural Sciences and Engineering Research Council of Canada, Canada Foundation for Innovation, British Columbia Knowledge Development Fund, and Ballard Power Systems through an Automotive Partnership Canada grant.

Figure 1

2522

, , , , , , and

Despite the advantages of high temperature polymer electrolyte fuel cells (HT-PEM FCs) over low temperature PEM-FC, like higher tolerances against CO, degradation rates are higher in comparison to LT-PEM fuel cells. The best published degradation rates, achieved with HT-PEM FCs so far, constitute 4.9 µV/h under pure hydrogen and air supply and an operating temperature of 160 °C [1-2]. Low temperature PEM fuel cells under optimal conditions and a working temperature of 75 °C show very low degradation rates of 1-2 µV/h [3].

Therefore, one of the most important tasks within the European project CISTEM (Construction of Improved HT-PEM MEAs and Stacks for Long Term Stable Modular CHP Units, GA-No. 325262) is, next to the development of a new HT-PEM fuel cell based Combined Heat and Power (CHP) technology, to realize degradation rates with values less than 4 µV/h. The test conditions have been defined by the FCH JU:

  • Single cell test with 25 cm²-MEAs

  • Constant current density: 0.3 A/cm²

  • Hydrogen and air supply (λ=1.5/2)

  • Temperature: 160 °C

  • Min. test duration: 2,000 hours

The tests have been performed with Dapozol®-G55 MEAs with thermally cured polybenzimidazole (PBI) membrane and Pt/C based electrodes (BoA1-MEAs), developed and provided by Danish Power Systems and delivered to the facilities of different partners. The distribution of new MEAs and testing at three different lab facilities under agreed identical operating conditions support and verify achieved degradation results perfectly. Electrochemical investigation with polarization curves have been performed at Begin of Life (BoL), every 1,000 hours of operation and at the End of Test (EoT). The characterization has been completed with ante- and post-mortem micro-computed tomography (µ-CT).

Several tests have been performed, three representatives are shown in Figure 1 and the corresponding degradation rates are listed in

Table 1. The average degradation rate of the long term tests presented in this work results in 1.7 µV/h and this value correlates with the average degradation rate of LT-PEM fuel cells [3].

To verify the influence of different test conditions on the degradation rates, these experiments have been compared with constant load (0.3 A/cm²) long term tests under different fuel and oxidant compositions, with different flow field designs (serpentine and grid flow fields) and performed with complete in-situ electrochemical characterization procedures (polarization curves, electrochemical impedance spectroscopy, cyclic voltammetry and linear sweep voltammetry). These investigations enable the determination of the best possible operating conditions for HT-PEM fuel cells to achieve ultralow degradation rates.

Figure 1:Voltage changes as function of time under hydrogen and air supply (λ=1.5/2.0) and degradation rates, 160°C, 0.3 A/cm², Dapozol®-G55-MEAs. Test at UCTP still in progress.

References:

1. S. Yu, L. Xiao, B.C. Benicewicz, Fuel Cells, 08, 3-4, 165-174 (2008).

2. T.J. Schmidt, J. Baurmeister, ECS Transactions 3, 1, 861-869 (2006).

3. F.A. de Bruijn, V.A.T. Dam, G.J.M. Janssen, Fuel Cells, 08, 1, 3-22 (2008).

Figure 1

2523

, and

Polymer electrolyte fuel cells (PEFCs) are very promising as clean energy sources for future automotive applications. However, the performance and durability of PEFCs can be significantly decreased by the presence of contaminants. Such species may adsorb on the Pt catalyst, penetrate into the ionomer and membrane, or compete with the main reactions, resulting in a decrease of the active surface area, ionic conductivity of the catalyst layer and membrane, and cell performance. Also, H2O2 is a side product of the oxygen reduction reaction (ORR) and its associated yield often increases in the presence of organic contaminants, such as acetonitrile, propene, naphthalene, and methyl methacrylate [1-3]. A larger H2O2 generation rate exerts an undetermined influence on the performance and durability of PEFCs. The formation of H2O2 has been linked to the radical attack decomposition of the Nafion® ionomer, which is accelerated by the presence of Fe2+ and Cu2+ ions [4,5]. The Nafion® ionomer is not only used as an electrolyte membrane, but is also added to the catalyst layer to exploit its binding property and improve catalyst utilization. Consequently, the decomposition of the Nafion® ionomer is a significant risk to PEFC performance and durability. Therefore, it is necessary to investigate the ORR, hydrogen oxidation reaction (HOR), and H2O2 yield in the presence of contaminants to understand their effects on PEFCs.

Ethylene glycol (EG) is of interest because it is widely used as a coolant, antifreeze agent, and de-icing solution. De-icing of airport runways and airplanes is the primary source of EG in the environment. EG is also dispersed in the environment by the disposal of products that contain it. Caprolactam is another potential contaminant released by PEFC system materials either as a result of degradation or leaching. It has been detected in many leachates including from polyphthalamide materials that are being considered for use as balance-of-plant structural materials [6,7]. Both EG and caprolactam can poison ORR Pt/C catalysts [8-10], but the H2O2 yield has not been measured. In contrast to the slow ORR kinetics, the HOR kinetics on Pt/C catalysts is so fast that the cell voltage losses are negligible even for very low Pt loadings. However, attention has not been given to the poisoning effects of EG and caprolactam on the HOR. These considerations are important for prevention because tolerance levels are currently missing. The rotating ring/disk electrode (RRDE) is the equipment of choice to reveal the H2O2 yield and reaction pathway. In this presentation, the contaminant impacts of EG and caprolactam on the kinetics of ORR and HOR in acid media using RRDE are reported. Results will emphasize the changes in the electrochemical surface area, ORR and HOR kinetic currents, Tafel slope, H2O2 yield,and reaction pathway as a function of EG and caprolactam concentration.

Acknowledgments

Authors are grateful to the Office of Naval Research (award N00014-13-1-0463) and the Hawaiian Electric Company for their ongoing support to the operations of the Hawaii Sustainable Energy Research Facility.

References

[1] J. Ge, J. St-Pierre, and Y. Zhai, Electrochim. Acta, 134, 272 (2014).

[2] J. Ge, J. St-Pierre, and Y. Zhai, Electrochim. Acta, 138, 437 (2014).

[3] J. Ge, J. St-Pierre, and Y. Zhai, Int. J. Hydrogen Energy, 39, 18351 (2014).

[4] T. Kinumoto, M. Inaba, Y. Nakayama, K. Ogata, R. Umebayashi, A. Tasaka, Y. Iriyama, T. Abe, and Z. Ogumi, J. Power Sources, 158, 1222 (2006).

[5] J. Qiao, M. Saito, K. Hayamizu, and T. Okada, J. Electrochem. Soc., 153, A967 (2006).

[6] C. Macomber, H. Wang, K. O'Neill, S. Coombs, G. Bender, B. Pivovar, and H.N. Dinh, ECS Trans., 33 (1), 1637 (2010).

[7] C.S. Macomber, J. Christ, H. Wang, B.S. Pivovar, and H.N. Dinh, ECS Trans., 50 (2), 603 (2013).

[8] N. Travitsky, L. Burstein, Y. Rosenberg, and E. Peled, J. Power Sources, 194, 161 (2009).

[9] D. Morales-Acosta, L.G. Arriaga, L. Alvarez-Contreras, S.F. Luna, and F.J.R. Varela, Electrochem. Commun., 11, 1414 (2009).

[10] H.-S. Cho, M. Das, H. Wang, H.N. Dinh, and J.W. Van Zee, J. Electrochem. Soc., 162, F427 (2015).

E-22 Membranes for DFC and AMFC 2 - Oct 5 2016 2:00PM

2524

, , , , , , and

The potential of anion exchange membrane (AEM) fuel cells to provide inexpensive compact power from a wider variety of fuels than is possible with a proton exchange membrane (PEM) fuel cell, has continued to drive the research interest in this area. Alkaline catalysis in fuel cells has been demonstrated with non-precious metal catalysts, and with a variety of fuels beyond H2 and methanol. Alkaline fuel cells (AFCs), based on aqueous solutions of KOH, have serious drawbacks associated with system complexity and carbonate formation. Anion exchange membrane (AEMs) fuel cells have a number of advantages over both PEM fuel cells and traditional AFCs; however, although anionic conductivity in AEMs can be comparable to PEMs the chemical stability of membrane attached cations in hydroxide is still not always sufficient for practical applications. Recently, it has been recognized that a number of advanced cations, may give AEMs the needed chemical stability. In some circumstances simple trimethyl benzyl ammonium cations are stable up to 60oC allowing us to being to study hydroxide and water transport in these systems. We use in-plane conductivity and multi-nuclear PFGSE to measure self-diffusion coefficients of the water and where possible, the ion, i.e. F-, carbonate and bicarbonate. Together with temperature and RH dependent SAXS we couple this information with the morphological changes in the materails.

By their nature these organic cations form dipoles, which have a tendency to interact, see Figure below for a trimethylbenzyl cation with bromide for illustrative purposes. For maximum ion transport the cation should be distributed along the polymer chain, however, cation clustering is a common phenomena. It occurs in phase separated diblock polymers when an attempt is made to raise the IEC, no net increase is observed in ionic conductivity above a certain IEC. In many other less structured polymers it occurs in the fully humidified state when the distributed disordered cations order in agglomerates at a disorder/order Tδ. This Tδ is often observed in the operating temperature range expected for a device, in both DMA, DSC, and broadband electric spectroscopy. Interestingly, even though the SAXS clearly shows an agglomeration on the nano-scale, little difference is seen in the bulk conductivity with temperature. However, steric hindernce effects are expected to discourage aggregation, and may allow the cation to tune all properties of the AEM. As these phenomena are so prevalent in AEM materials the implications and science behind them are being investigated by us and will be discussed in this presentation. The implications of more complex cations proposed for stability in AEMs will be described.

Figure 1

2571

and

The design and optimization of stable and highly conductive anion exchange membranes (AEMs) is essential to the development of energy and conversion technologies.1, 2 There has been an increased interest and research to address the low stability and ionic conductivity of polymeric AEMs. However, a fundamental understanding of stability and transport in AEMs is still in its infancy3 and the effects of the choice of the cationic functional group and polymer architecture are still not known. Recently developed elastomeric AEMs4 exhibiting satisfactory chemical stability are investigated in this study. These AEMs are based on the elastomeric triblock copolymer, polystyrene-b-poly(ethylene-co-butylene)-b-polystyrene (SEBS), that are functionalized with various cationic groups.

Dissipative particle dynamics (DPD) is a meso-scale simulation technique5, 6 which enables the simulation of large time- and length-scales with reasonable computational expense has been selected to investigate the morphology and dynamical behavior of such systems. The system of hydrated AEM is coarse-grained to a degree that the consideration of both chemical distinction and fine structural variants are satisfied. The structures of the DPD beads were optimized using first principles electronic structure calculations and the interaction parameters were determined using a recently developed methodology.7

The effects of several microstructural parameters including the length of the tether chains, type of the cationic group (trimethylammonium [(CH3)3NH]+; dimethylimidazolium [DMIm+]; and triphenylphosphonium [(C6H5)3PH]+ ), and the polymer architecture on the morphology and transport of anions in the aforementioned elastomeric AEMs were investigated. The effect of water content in the membrane and the effect of ion exchange capacity (IEC) were also studied. The results are compared with the widely studied proton exchange membrane Nafion.

References:

1. G. Couture, A. Alaaeddine, F. Boschet and B. Ameduri, Prog Polym Sci, 2011, 36, 1521-1557.

2. J. R. Varcoe, P. Atanassov, D. R. Dekel, A. M. Herring, M. A. Hickner, P. A. Kohl, A. R. Kucernak, W. E. Mustain, K. Nijmeijer, K. Scott, T. W. Xu and L. Zhuang, Energ Environ Sci, 2014, 7, 3135-3191.

3. M. G. Marino, J. P. Melchior, A. Wohlfarth and K. D. Kreuer, J Membrane Sci, 2014, 464, 61-71.

4. A. D. Mohanty, C. Y. Ryu, Y. S. Kim and C. Bae, Macromolecules, 2015, 48, 7085-7095.

5. P. Espanol and P. B. Warren, Europhys Lett, 1995, 30, 191-196.

6. R. D. Groot and P. B. Warren, J Chem Phys, 1997, 107, 4423-4435.

7. F. Sepehr and S. J. Paddison, Chem Phys Lett, 2016, 645, 20-26.

2572

, and

Anion exchange polymer electrolyte is a key element for alkaline membrane fuel cell (AMFC) applications. In fuel cells, the polymer electrolytes require alkaline stable covalently tethered cationic group to conduct hydroxide ion. Benzyltrimethyl ammonium cationic group has been mostly often employed in anion exchange polymer electrolytes because it is reasonably basic, stable and easily synthesized. However, the demands on more stable and less catalyst poisoning cationic groups are growing.

 One promising stable cationic groups that potentially can be tethered in the polymer backbone is guanidinium. However, benzyl guanidinium tethered polymers prepared via chloromethylation and subsequent cationization of the halomethyl group showed a limited stability in high pH conditions due to the nucleophilic degradation of the benzylic carbon. In 2011, we had developed a new synthetic method to synthesize guanidinium tethered polymer via activated fluorophenyl-amine reaction.1 This new synthetic method allows to prepare phenyl guanidinium (instead of benzyl guanidinium via conventional method) tethered polymer. The direct connection between phenyl group and guanidinium can stabilize the cationic group by whole resonance structure. However, continuing study indicates that the poly(arylene ether) polymer that we employed undergoes ether-cleavage degradation under high pH conditions.2

In this study, we report the phenyl guanidinium tethered ether-free poly(phenylene)s. Activated fluorophenyl-amine reaction was used to synthesize the poly(phenylene)s. The alkaline stability of the polymer is exceptional: no chemical structural degradation was found during 1000 h in 0.5 M NaOH at 80°C. The tailored poly(phenylene) anion exchange ionomers is well suited for the ionomeric binders, particularly for hydrogen oxidation reaction in alkaline membrane fuel cells.

References:

1 D. S. Kim, A. Labouriau, M. D. Guiver, and Y.S. Kim, Chem. Mater. 2011, 23, 3795-3797.

2 C. Fujimoto, D. S. Kim, M. Hibbs, D. Wrobleski, Y.S. Kim, J. Memb. Sci. 2012, 423, 438-449.

2573

, , and

Anion exchange membranes are an integral part of a low temperature polymer electrolyte fuel cell. They serve the purpose of transporting hydroxide ions from cathode to anode while acting as a separator for fuel and air.

In this work we are focused on synthesis and device testing of some novel poly(2,6-dimethyl-1,4-phenylene oxide) based AEMs. The idea is to start with a polymer that is readily available in the market and synthesize large quantities of functionalized materials. Our work centers around attaching conventional quaternary ammonium cations as well as long spacer chain quaternary ammonium cations on the polymer backbone. We solvent cast and melt press the synthesized polymer to fabricate a membrane. Melt pressing crosslinks the membrane and adds the required mechanical and dimensional stability. We are able to achieve high degree of chlorination (~80%) of polyphenyle oxide and still maintain the swelling properties of the polymer because of crosslinking.

Our initial experiments show that the membrane has promising chloride conductivity, mechanical and dimensional stability thus making it suitable for fuel cell testing. We are also testing the device performance of these membranes using H2/O2fuel cell with Pt/C catalyst at various temperatures and different humidification levels to find the optimum operating conditions without causing significant degradation of the membrane.

Overall we are reporting a novel AEM material with simple chemistry which can be synthesized in large batches with ease and can serve as s good fuel cell membrane in the future.

2574

, , , , , , and

Alkaline anion exchange membrane fuel cells are widely investigated recently as promising alternative to well-established proton exchange membrane fuel cells. Anion exchange membranes, using as solid electrolytes, are required to be thermally, chemically and mechanically stable in base at the operation temperature and humidity, as well as have an efficient OH- transport along the membrane across two electrodes.

This study aims at studying AEMs functionalized with three different bulky cations. They are 1,4,5-trimethyl-2-(2,4,6-trimethoxyphenyl) imidazolium functionalized polyphenylene oxide (Figure 1(a)) [1], tris(2,4,6-trimethoxyphenyl) phosphonium functionalized PPO (Figure 1(b)) [2], and phenylene cobaltocenium functionalized norbornene (Figure 1(c)).

Degradation study on 1,4,5-trimethyl-2-(2,4,6-trimethoxyphenyl) imidazolium functionalize PPO and unsubstituted imidazolium PPO by soaking both two membranes into 1 M sodium hydroxide at 80°C indicates attachment of 2,4,6-trimethoxyphenyl group efficiently improve OH- stability due to the enhanced volumetric steric and basicity of the central cation. Super hydrophobic and bulky phosphonium cation attached by three 2,4,6-trimethoxyphenyl groups has the chemical stability surpassing benzyl trimethyl ammonium by 30% after degrading in 1 M KOH at 80°C for 20 days. The cobaltocenium cation, which has 18 valance electrons further improve the chemical stability by 60% higher than trimethyl ammonium after treating in 1 M NaOH/CH3OH at 80°C for 20 days (Figure 2).

Besides enhanced chemical properties from these bulky cations, their influence on the other polymer properties include thermal stability, water uptake, conductivity as well as transport property are also discussed.

Figure 1. Structure of 1,4,5-trimethyl-2-(2,4,6-trimethoxyphenyl) imidazolium functionalized PPO (a), tris(2,4,6-trimethoxyphenyl) phosphonium functionalized PPO (b) and phenylene cobaltocenium functionalized norbornene (c).

Figure 2. Degradation studies performed at 80°C on 1,4,5-trimethyl-2-(2,4,6-trimethoxyphenyl) imidazolium (solid diamond, 1 M KOH), tris(2,4,6-trimethoxyphenyl) phosphonium (solid square, 1 M KOH), phenylene cobaltocenium (solid triangle, 1 M NaOH/CH3OH) and benzyl trimethyl ammonium (open square, 1 M KOH; open triangle, 1 M NaOH/CH3OH).

Reference

[1] Y. Liu, J. Wang, Y. Yang, T.M. Brenner, S. Seifert, Y. Yan, M.W. Liberatore, A.M. Herring, J. Phys. Chem. C 118 (2014) 15136-15145.

[2] Y. Liu, B. Zhang, C.L. Kinsinger, Y. Yang, S. Seifert, Y. Yan, C.M. Maupin, M.W. Liberatore, A.M. Herring, J. Membr. Sci. 506 (2016) 50-59.

Figure 1

2575

, and

To circumvent the shortcomings of traditional synthetic routes relying on chloromethylation or bromination of benzylic methyl groups, a novel approach, chloromethylbenzoylation, has been proposed to synthesize commercially available poly(aryl ether)s (PAEs) with side-chain-type benzylic cationic groups for anion exchange membranes (AEMs). Previous works have demonstrated that adding alkyl side chain to PAE backbones can improve the alkaline stability of the AEMs. In this work, side-chain-type benzylchloride groups have been directly introduced onto poly resorcinol ketone (PRK) via an acylation reaction between 4-(chloromethyl)benzoic acid and PRK in P2O5/CH3SO3H at 60 oC. 1H NMR and GPC confirm that 100% substitution occurred, without side reactions such as crosslinking or degradation. Robust AEMs were then obtained through amination upon introduction of trimethylammonium (TMA+) groups. The chloride ionic conductivity of the AEM was 8.8 mS/cm at 30 °C, which is competitive with that of a commercial TokuyamaÒ A201 membrane at similar IEC values. In summary, this approach represents a simple and straightforward route capable of quantitative introduction of side-chain-type benzylchloride groups without using expensive or toxic reagents, with potential applications of the resultant AEMs in energy conversion and storage and other fields.

2576

, and

There is a growing interest in using anion exchange membranes (AEMs) as separators in alkaline membrane fuel cells (AMFCs) and in other energy conversion and storage systems such as redox flow batteries (RFBs), alkaline water electrolyzers (AWEs) and reverse electrodialysis (RED) cells. The most commonly used cation in AEMs is the benzyl trimethylammonium cation. It is easy to synthesize due to the basicity of the trimethylamine, results in relatively large ion exchange capacities (IECs), and the resulting AEMs have good ionic conductivities. AEMs based on the benzyl trimethylammonium cation are good candidates for separators in redox flow batteries. However, it had been shown that the quaternary-ammonium-based AEMs are sensitive towards Hofmann elimination and direct nucleophilic elimination reactions under alkaline conditions, which impedes their use in AMFCs and AWEs. To solve the alkaline stability problem encountered by quaternary ammonium cations, researchers have investigated cations without alpha hydrogens to avoid degradation through ylide formation. Gu and Yan (1) first synthesized a polysulfone based ionomer containing the benzyl tris-(2,4,6-trimethoxyphenyl) phosphonium (TTMP+) cation and they found it had an outstanding stability under alkaline conditions. In our lab, we attempted to attach TTMP to polyphenylene oxide (PPO) and other polymer backbones to make alkaline stable AEMs, but the resultant AEMs were extremely brittle. Therefore, we have devoted attention to studying more robust backbones.

Polystyrene-block-poly-(ethylene-ran-butylene)-block-polystyrene (SEBS) is a chemically stable and elastomeric triblock copolymer, which deforms elastically to accommodate bulky cations (TTMP+) without becoming too brittle. There have been several attempts to make SEBS-based AEM by chloromethylation, however, the degree of functionalization (DF) obtained were relatively low due to gelation during the reaction (2, 3). In our work, we have optimized the chloromethylation conditions by selecting the appropriate solvent, temperature and reaction time. The resulting polymers have DFs up to 0.3 (mol of chloromethyl groups per polymer repeat unit). We have been able to synthesize highly robust SEBS-based AEMs containing benzyl trimethyl ammonium (TMA+) and benzyl tris- (2,4,6-trimethoxyphenyl) phosphonium (TTMP+) cations. 1H-NMR spectroscopy, ion exchange capacity and tensile tests were employed to characterize the AEMs. We will report the results obtained.

References

1. S. Gu, R. Cai, T. Luo, Z. Chen, M. Sun, Y. Liu, G. He and Y. Yan, Angewandte Chemie International Edition, 48, 6499 (2009).

2. R. Vinodh, A. Ilakkiya, S. Elamathi and D. Sangeetha, Materials Science and Engineering: B, 167, 43 (2010).

3. L. Sun, J. Guo, J. Zhou, Q. Xu, D. Chu and R. Chen, Journal of Power Sources, 202, 70 (2012).

2577

, , and

Alkaline membrane fuel cells (AMFCs), which use anion exchange membrane (AEM) as a solid electrolyte, continue to receive increased attention because of their advantages of fast oxygen reduction reaction in alkaline media, efficient water management, and the ability to use non-precious metal catalysts (e.g., nickel, cobalt, silver, iron, and etc.). Hydrocarbon-based aromatic polymers have been widely used as AEMs due to their good thermal and mechanical stabilities and easiness of synthesis and functionalization. However, poor chemical stabilities under strong alkaline environment and low ionic conductivity of AEMs should be improved for commercialization of AMFCs. Hydrophilic and hydrophobic phase separation is helpful for improvement in chemical stability and ionic conductivity. Well-developed hydrophilic domains form effective ion transport pathways resulting in enhanced ionic conductivity, while hydrophobic domains maintain mechanical and chemical stabilities because most nucleophilic sources, which can attack polymer backbone and cationic groups, are in hydrophilic domains. It is well known that block copolymers are the most promising chemical architecture for developing hydrophilic and hydrophobic phase separation.

Based on these backgrounds, we synthesized poly(arylene ether sulfone) based block copolymers to investigate the effect of chemical structure on AEM properties. These block copolymers were functionalized with cations after selective chloromethylation of hydrophilic blocks. The degree of chloromethylation was controlled to optimize ion exchange capacity values of block copolymers. The molecular weight of each oligomer was controlled by changing the feed ratio of monomers and confirmed by nuclear magnetic resonance (NMR) and gel permeation chromatography (GPC). All block copolymers showed well-developed morphology and outstanding AMFC performances. More detailed results will be discussed.

2578

and

In this study, synthesis and characterization of Anion Exchange Blend Membranes (AEBMs) are introduced. AEBMs were prepared by mixing polymer solutions of a halomethylated polymer (Br-PPO, brominated poly(2,6-dimethyl-1,4-phenylene oxide)), polybenzimidazole (PBI, F6PBI or PBI-OO) and a sulfonated polymer (S-polymer, SAC 098 (a sulfonated poly(phenylethersulfone), nonfluorinated) or SFS 001 (a sulfonated aromatic polyether, partially fluorinated)).

 A halomethylated polymer was used as the anion exchange ionomer precursor. Polybenzimidazole was used as a matrix polymer in order to enhance the mechanical strength of the AEBN. A minor amount of sulfonated polymer was added, forming ionic cross-links, and covalent bonds were formed in the blend membrane by reaction of a small amount of the CH2Br groups of Br-PPO with imidazole N-H groups of the PBI, respectively, leading to an increase of the chemical and dimensional stability of the novel AEBMs [1].

 Membranes were fabricated by following procedure 1) mixing of the polymer solutions to homogeneity, 2) solvent evaporation in oven, 3) quaternization of the blend membrane by soaking the membrane in amine solution, 4) Ion exchange into chloride form in sodium chloride solution.

 It is well known that many types of anion exchange membranes are chemically unstable under alkaline condition. S. Holdcroft published a novel polybenzimidazolium AEM with a mesitylene building block shielding the dimethylimidazolium moiety of the polybenzimidazolium from OH- attack, leading to excellent hydroxide stability of this AEM: after immersion of this AEM in 2M KOH for 10 days at 60°C, no significant change of the NMR spectrum was observed [2]. Therefore it can be concluded that introduction of bulky steric hindrance groups in the vicinity of the cationic groups of AEMs can prevent the quaternized amine in the polymer structure from nucleophilic attack of hydroxide ions. Consequently in this study several types of sterically hindered tertiary amines were introduced in the polymer blend system for quaternization with Br-PPO, among them 1,2,2,6,6-pentamethylpiperidine (pempidine), quinuclidine, and 1,2,4,5-tetramethyl-1H-imidazole (abbreviated TMIm). Among these AEBMs, the TMIm-based membranes showed the best properties in terms of conductivity and stability.

We investigated different combinations of polymers for AEBMs. The membrane containing PBI-OO and SAC098 showed good weight maintenance than other material combinations in DMAc extraction test at 90oC. This membrane was prepared and post-treated with TMIm as mentioned above. The membrane exhibited a chloride conductivity of 6.5 mS/cm at 30oC and 90% RH which is higher than that of commercial membrane (Tokuyama A202) under the same condition. The change of ion exchange capacities (IECs) was also investigated in order to examine the hydroxide stability. The membrane was exposed to 1M KOH solution at 90oC for 10days, and ion-exchange capacities (IECs) were compared to each other before and after KOH treatment. The IECs before and after KOH treatment were 2.63 and 2.59 mmol/g respectively, indicating excellent hydroxide stability. Apart from Br-PPO, other halomethylated polymers such as polyvinylbenzylchloride and polyepichlorohydrine were also investigated as precursors for AEBMs in this study, yielding good Cl- conductivities (in excess of 30 mScm-1) and alkaline stabilities with the novel 1,2,4,5-tetramethyl-1H-imidazolium head group. Addition of a hydrophilic polymer phase to the novel AEBMs resulted in even higher Cl- conductivities (e. g. 131 mScm-1@90°C@90% r.h.) after 10 d of KOH treatment (1M, 90°C).

 

References

[1] C. G. Morandi, R. Peach, H. M. Krieg, J. Kerres, J. Mater. Chem. A, 2015, 3, 1110

[2] O. D. Thomas, K. J. W. Y. Soo, T. J. Peckham, M. P. Kulkarni, S. Holdcroft, J. Am. Chem. Soc., 2012, 134, 10753

2579

and

Development of anion-exchange membrane fuel cells (AEMFCs) is of great interest among scientific community because of their potential of commercialization at low cost for their use in automotive transportation, portable batteries and power generators [1,2]. However, they currently show serious losses in efficiency associated principally with the low ionic conductivity of the polymeric membranes used as electrolytes [1,3,4].

In order to improve ionic conductivity, different theoretical approaches have been proposed to represent the transport phenomena involved in the mobility of hydroxide ions through hydrated anion-exchange membranes [1,4,5]. However, the characteristics of the mechanisms responsible of that transport are not well known. Particularly, the mechanism responsible of the major contribution to hydroxide mobility in aqueous media, called structural diffusion or Grotthuss mechanism is not completely understood in different systems [1,4].

Although theoretical studies have described how the Grotthuss mechanism takes place in aqueous media and which are its characteristics [6], currently it has not been made extensions of those studies to describe the Grotthuss mechanism in hydrated anion-exchange membranes, taking into account structural aspects of the polymer such as its hydration degree and type of cationic functional group [1].

The mentioned conceptual gaps hinder the improvement of the anion-exchange membranes for engineer and technological applications, considering this from a theoretical point of view. For that reason, the objective of this research is to determine and describe the characteristics of the Grotthuss mechanism for hydroxide ions in the hydrated quaternized polystyrene-block-poly(ethylene-ran-butylene)-block-polystyrene (QSEBS) membrane for different humidification values. This will be carry out by means of Car-Parinello Ab Initiomolecular dynamics (AIMD), which is considered the most suitable simulation technique in terms of prediction and accuracy given the characteristic length and time scales of the Grotthuss mechanism, and also taking into account that to date there are no formal studies in this topic.

From the simulations carried out in this study, qualitative descriptions are derived for the Grotthuss mechanism for hydroxide ions through hydrated QSEBS membrane. Also, radial distribution functions, mechanism characteristic times and diffusion coefficients associated with it are also estimated and analyzed. It is expected that the new understandings derived from this research could be used to propose the structural characteristics of a certain polymeric material which maximize the ion mobility due to Grotthuss mechanism and therefore its ionic conductivity [1], so making a strong contribution to the improvement of the efficiency of AEMFC.

References:

[1] G. Merle, M. Wessling, K. Nijmeijer, J. Memb. Sci. 377 (2011) 1.

[2] Pivovar, B. Proceedings from the Alkaline Membrane Fuel Cell Workshop, Arlington, Virginia, May 8-9, 2011.

[3] E. Antolini, E.R. Gonzalez, J. Power Sources 195 (2010) 3431.

[4] K.N. Grew, W.K.S. Chiu, J. Electrochem. Soc. 157 (2010) B327.

[5] A.Z. Weber, R.L. Borup, R.M. Darling, P.K. Das, T.J. Dursch, W. Gu, D. Harvey, A. Kusoglu, S. Litster, M.M. Mench, R. Mukundan, J.P. Owejan, J.G. Pharoah, M. Secanell, I. V Zenyuk, J. Electrochem. Soc. 161 (2014) 1254.

[6] D. Marx, A. Chandra, M.E. Tuckerman, Chem. Rev. 110 (2010) 2174.

E-21 Membranes for DFC and AMFC 1 - Oct 5 2016 8:20AM

2525

, , , , and

There has been a growing interest in noble metal-free alkaline fuel cells, in which anion conducting polymer electrolyte membranes (AEMs) required highly alkaline durability because of high temperature and highly basic fuel cell operating condition in the fuel cells. In recent years, we have developed N-vinylimidazolium-type AEMs by radiation-induced grafting for direct hydrazine hydrate fuel cells (DHFCs).1,2 The prepared poly(N-vinylimidazolium)-grafted AEMs (NVIm-AEM) showed moderate alkaline stability, however, these AEMs degraded via the initial β-elimination at the initial stage and hydrolytic ring opening reaction of the imidazolium unit at the later stage.3 In this work, we synthesized the AEM based on a homo- and co-polymer grafts of 2-methyl-N-methyl-4-vinylimidazolium with styrene (2Me-NMe-4VIm-AEM and 2Me-NMe-4VIm/St-AEM) to prevent the β-elimination and ring opening reaction (Figure 1).

According to the literature, 4-vinylimidazole (4VIm) was transformed to N-methyl-4-vinilimidazole (NMe-4VIm) by the N-methylation using sodium hydrate and methyl iodide in a 64% yield.4 NMe-4VIm was then reacted with n-BuLi and methyl iodide to give a 2-methyl-N-methyl-4-vinylimidazole monomer (2Me-NMe-4VIm) in a 56% yield (Scheme 1).

The poly(ethylene-co-tetrafluoroethylene) (ETFE) films were irradiated with a 60Co γ-ray source (QST Takasaki, Gunma, Japan) at room temperature in argon atmosphere. The pre-irradiated ETFE films were immediately immersed into the argon-purged monomer solution consisting of 5.3 mol/L 2Me-NMe-4VIm or a mixture of 2Me-NMe-4VIm and St (80 : 20 molar ratio) in 1,4-dioxane. The graft reaction proceeds smoothly and gave 2Me-NMe-4VIm and 2Me-NMe-4VIm/St graft ETFE films with the grafting degree of 33% and 54%, respectively. The grafted films were then N-alkylated and ion exchanged to obtain the hydroxide from of 2Me-NMe-4VIm-AEM and 2Me-NMe-4VIm/St-AEM with IECs of 1.38 and 1.74 mmol/g (Scheme 2). The N-alkylation reaction proceeded quantitatively while the un-methylated analogue, a 4-VIm graft ETFE film, showed less reactivity and saturated at around 66% level. 2Me-NMe-4VIm-AEM and 2Me-NMe-4VIm/St-AEM showed conductivities of 175 and 157 mS/cm with water uptakes of 53 and 86%. These AEMs show significantly higher conductivities with lower water uptakes, which are desirable properties for fuel cell applications.

The alkaline stability of AEMs was evaluated by the change in conductivity of AEMs in 1M KOH at 80°C. As shown in Figure 2, the conductivity of NVIm-AEM (N-vinyl type AEM) decreased to 0 mS/cm after 100 h, while 2Me-NMe-4VIm-AEM showed higher conductivities than 20 mS/cm after 100 h. The decreasing profile of the 2Me-NMe-4VIm-AEM conductivity indicates that the initial conductivity drop resulting from β-elimination was clearly suppressed by the shift of vinyl substituent from 1- to 4-positon of the imidazolium ring. Also, the decreases of conductivity at the later period, arising from ring opening degradation become much slower for 2Me-NMe-4VIm-AEM owing to the methyl substitution effect at 2-position of the imidazolium ring. The copolymer type 2Me-NMe-4VIm/St-AEM possesses excellent initial and long term alkaline stabilities. Thus, sharp drops of conductivity at the initial stage probably caused by the graft chain detachment, and the introduction of copolymerized styrene units should suppress the detachment phenomenon by hydrophobicity.5 In conclusion, for vinylimidazolium-containing graft-type AEMs, the control of vinyl position, the methyl-protection at 2-positon, and copolymerization with styrene should be promising strategy to improve the alkaline stability, which is the essential property for alkaline fuel cells.

Acknowledgement

This work was supported by the Advanced Low Carbon Technology Research and Development Program (ALCA) from the Japan Science and Technology Agency (JST).

 

References

[1] K. Asazawa, K. Yamada, H. Tanaka, A. Oka, M. Taniguchi and T. Kobayashi, Angew. Chem., Int. Ed., 46, 8024 (2007).

[2] K. Yoshimura, H. Koshikawa, T. Yamaki, H. Shishitani, K. Yamamoto, S. Yamaguchi, H. Tanaka and Y. Maekawa, J. Electrochem. Soc., 161(9), F889 (2014).

[3] Y. Ye and Y. A. Elabd. Macromolecules, 44, 8494 (2011).

[4] C. G. Overberger and Y. Kawakami, J. Polym. Sci. Part A Polym. Chem, 16, 1237 (1978).

[5] K. Enomoto, S. Takahashi, T, Iwase, T. Yamashita and Y. Maekawa, J. Mater. Chem., 21, 9343 (2011).

Figure 1

2526

, and

There is a growing interest in using anion exchange membranes (AEMs) as separators in alkaline membrane fuel cell (AMFCs) and in other energy conversion and storage systems such as redox flow batteries (RFBs), alkaline water electrolyzers (AWEs) and reverse electrodialysis (RED) cells. The most commonly used cation group in AEMs is the benzyl trimethylammonium cation. However, it had been shown that quaternary ammonium-based AEMs are sensitive towards Hofmann elimination (1) and direct nucleophilic elimination reactions (2) that result in loss of ion exchange capacity (IEC) and ionic conductivity. To solve the alkaline stability issue inherent to quaternary-ammonium-group-containing AEMs, several alternative cations, including imidazolium, benzimidazolium, guanidinium, phosphonium and metal cations have been proposed and investigated.

Imidazolium-based AEMs have drawn researchers' interests mainly because of their high hydroxide ion conductivities. In addition to 1-methylimidazole and 1,2-dimethylimidazole, a series of modified imidazole bases, namely 1-heptyl-2-methyl-1H-imidazole and 1-dodecyl-2-methyl-1H-imidazole, were synthesized, wherein these modified bases contained a long alkyl chain at the N-3 position . Anion exchange membranes (AEMs) were prepared with derived imidazolium cations either grafted onto the benzyl position of poly(phenylene oxide)(PPO) or affixed using a hexyl (6-carbon) spacer chain. AEMs with cations affixed to the benzyl position were synthesized by bromination of PPO followed by reaction with the imidazole bases. AEMs with the hexyl spacer were synthesized by Friedel-Crafts acylation of 6-bromo-1-hexanoyl chloride on PPO, followed by reduction of the ketone group and reaction with the imidazole bases. The alkaline stability of the various imidazolium-cation-containing AEMs was evaluated by: 1) measuring the change in IEC after immersion in 1M KOH at 60°C for up to 14 hours; and 2) using 1-D and 2-D NMR spectroscopy to investigate any changes in chemical structure. Both NMR and IEC experiments showed that PPO functionalized with modified imidazolium cations (with long alkyl chain at the N-3 position) did not exhibit better alkaline stability than benchmark AEMs made with 1-methylimidazole. However, the use of a six-carbon spacer did improve AEM alkaline stability when the AEM was prepared directly using 1-methylimidazole and 1,2-dimethylimidazole, but not when the AEM was prepared with the modified imidazolium cations (with a long alkyl chain at the N-3 position).

References

1. C. G. Arges and V. Ramani, Journal of The Electrochemical Society, 160, F1006 (2013).

2. S. Chempath, B. R. Einsla, L. R. Pratt, C. S. Macomber, J. M. Boncella, J. A. Rau and B. S. Pivovar, The Journal of Physical Chemistry C, 112, 3179 (2008).

2527

, , , and

H2 fuel cells and water electrolysers containing solid alkaline electrolytes represents a promising approach for the development of a cost-effective H2-based infrastructure. The alkaline media promises the use of cheaper non-precious metal catalysts [1]. Therefore, there is a push to fulfil the current performance requirements of such electrochemical devices. The employment of alkaline anion-exchange membrane (AAEM) in fuel cells and water electrolysers allows the reduction of the inter-electrode distance with better energy efficiencies without an increase in undesirable gas crossover and a reduction of the purity of the product H2 [2]. However, the stability of the central AAEM at operation temperatures (up to 90°C) remains the biggest challenge of this technology.

Numerous researchers around the world are seeking the most stable chemistries of AAEMs in strongly basic media. Among the cationic head group options, the use of quaternary ammoniums (QA) is common, due to the accessibility of the starting materials and their demonstrated ion-conducting properties. Numerous studies of hydroxide-derived degradation products have enabled the identification of the main degradation pathways with such tetraalkylammonium functionalities [3].

Herein, we have employed radiation-induced grafting for the preparation of conductive AAEMs (> 100 mS cm-2 at 80°C, fully hydrated) with various benzylic QA head-groups (cyclic and non-cyclic). The radiation induced grafting of commodity polymer films affords a reproducible method for the cost competitive, large lab-scale preparation of ion-exchange materials that is scalable for further applications [4]. The spectroscopic and composition analysis of the AAEMs after aqueous hydroxide (1 mol dm-3) treatment at 80°C revealed various degradation mechanisms are occurring. According to this, we have developed chemistries to introduce QA head groups that afforded AAEMs with improved alkali stabilities. Furthermore, fuel cell testing has demonstrated that some of these new chemistries significantly enhance the in situ electrochemical performance of the AAEMs compared to the radiation-grafted benzyltrimethylammonium benchmark [5] (> 800 mW cm-2 H2/O2 peak power densities at 60°C, no back-pressurization, with 50 μm thick AAEMs and Pt- benchmark catalysts).

 [1] J. R. Varcoe, P. Atanassov, D. R. Dekel, A. M. Herring, M. A. Hickner, P. A. Kohl, A. R. Kucernak, W. E. Mustain, K. Nijmeijer, K. Scott, T. W. Xu and L. Zhuang, Energy Environ. Sci., 2014, 7 3135.

[2] S. Vengatesan, S. Santhi, S. Jeevanantham and G. Sozhan, J. Power Sources, 2015, 284, 361–368.

[3] A. D. Mohanty and C. Bae, J. Mater. Chem. A, 2014, 2, 17314–17320; M. G. Marino and K. D. Kreuer, ChemSusChem, 2015, 8, 513–23; M. R. Sturgeon, C. S. Macomber, C. Engtrakul, H. Long and B. S. Pivovar, J. Electrochem. Soc., 2015, 162, F366–F372.

[4] M. Nasef, Chem.Rev, 2012, 114, 12278.

[5] J. R. Varcoe, R. C. T. Slade, E. Lam How Yee, S. D. Poynton, D. J. Driscoll and D. C. Apperley, Chem. Mater., 2007, 19, 2686–2693.

Figure 1

2528

, , and

Bipolar membranes fuel cells utilizing both anion and cation conductive materials have several advantages compared to their purely acidic or alkaline counterparts. The potential advantages include improved water management through self-hydration and facile electrode kinetics. Material transport properties play an important role in determining viability of membrane and ionomer materials. In addition, the cation-anion junction itself is a critical element in determining device performance because it must be sufficiently conductive to ionic species and mechanically stable to withstand the internal pressure from water formation. A series of electrochemical devices using different bipolar membranes have been fabricated for use in direct methanol and hydrogen fuel cells. This study examines both single- and two-membrane systems containing cation-anion junctions. These fuel cells were characterized by performance metrics and electrochemical impedance spectroscopy to determine specific areas for improvement in the bipolar devices. Operation under varying humidity was studied in order to understand water management necessary for bipolar fuel cells. These results will be used in the future optimization of bipolar devices.

Financial support from US Office of the Deputy Assistant Secretary of the Army for Defense Exports and Cooperation (DASA-DE&C) is gratefully acknowledged.

2529

, and

The use of anion exchange membranes (AEMs) as separators in alkaline polymer electrolyte fuel cells (AFCs) has received considerable interest over the past decade. AFCs offer several advantages over more traditional PEM fuel cells, in particular operating at high pH enables the use of non-precious metal catalysts, drastically reducing the cost of the cells. However, widespread adoption of AEMs has been hampered due to poor stability, in part due to base-induced decomposition of the polymer backbone and cation group as well as conversion to carbonate form, all of which contribute to reduction of ion exchange capacity (IEC), conductivity, and mechanical strength. Furthermore, there is an insufficient understanding on how the polymer nanostructure and microstructure affects stability, transport, and cell performance.

Nuclear magnetic resonance (NMR) spectroscopy is well suited to investigate these issues. We will demonstrate how 1D and 2D NMR spectroscopy may be utilized to characterize the degradation of AEMs using both model cations and ex situ membranes. Measurements of relaxation, e.g. T1, for a series of AEMs will be correlated with ionic conductivity. We also have used pulsed-field gradient NMR experiments to measure water self-diffusion to assist in deconvoluting how tortuosity, connectivity, and the mobility of water in hydrophobic and hydrophilic domains are governed by the nature of the cation and polymer backbone. The ultimate goal of this work is to develop a structure-activity relationship to assist in the design of novel materials with desirable properties.

2530

, , , , , and

Single ion conducting polymer electrolyte membranes (PEMs) are found at the heart of many electrochemical devices (e.g., fuel cells, flow batteries, electrolyzers, electrodialysis, etc.).(1) The key requirements for such materials include high ionic conductivity, electron insulation, mechanical resilience, selectivity (i.e., high transference #), and chemical, thermal, and mechanical stability.(2, 3) Ionic conductivity is particularly important because it strongly influences the ohmic overpotential in the said electrochemical devices affecting device efficiency.(1, 4)

Block copolymer electrolytes represent an attractive set of PEM materials because their micro-phase architecture yields greater ion conduction over their random copolymer counterparts.(5-7) Furthermore, the block copolymer charateristic gives rise to a variety of morphological architectures, while engineering of the block copolymer's self-assembly influences domain alignment.(8-12) However, there is lack of systematic studies correlating molecular level structural design to bulk material properties like ion transport.

In this work, a model lamellae-forming diblock copolymer electrolyte system was manipulated to examine the extent of ion domain connectivity on ion conduction.(13, 14) Careful control of the model system's volume fraction gave diblock copolymer structures with slightly rich electrolyte domains or slightly rich hydrophobic domains. The slightly rich electrolyte domains had greater contiguous area fraction while simultaneously demonstrating fewer terminal defects. Having a contiguous area fraction from 0.95 to 1.0 resulted in a 2x improvement in ionic conductivity over a non-micro-phase separated block copolymer electrolyte or a micro-phase separated block copolymer electrolyte with poor connectivity. Incremental adjustment in the extent of connectivity revealed an exponential growth curve for ion conductivity as a function of contiguous area fraction. Furthermore, the benefits that a micro-phase separated block copolymer electrolyte affords in terms of ion conductivity were not realized when the block copolymer electrolyte had poor connectivity. The results of this work have far reaching implication into the rationale design of PEM materials based upon block copolymer designs. This talk will emphasize the importance of maximizing ion domain connectivity while taking great strides to minimize terminal defects to boost ion transport in PEM materials. This talk will close with future directions on how molecular level engineering of block copolymer electrolytes offer the potential to reveal how other structural features (e.g., tortuosity(15), grain boundaries(5), and counterion condensation) alter ion transport in PEM materials(16, 17).

1. H. Strathmann et al., Ion-Exchange Membranes in the Chemical Process Industry. Industrial & Engineering Chemistry Research52, 10364-10379 (2013).

2. H. Strathmann, Ion-Exchange Membrane Separation Processes, Volume 9. Membrane Science and Technology (Elsevier Science, Amsterdam, The Netherlands, 2004), vol. 9.

3. T. Sata, Ion Exchange Membranes: Preparation, Characterization, Modification and Application. (Royal Society of Chemistry, Cambridge, UK, 2004).

4. A. Z. Weber, J. Newman, Modeling Transport in Polymer-Electrolyte Fuel Cells. Chemical Reviews104, 4679-4726 (2004).

5. Y. A. Elabd, M. A. Hickner, Block Copolymers for Fuel Cells. Macromolecules 44, 1-11 (2011).

6. N. Li, M. D. Guiver, Ion Transport by Nanochannels in Ion-Containing Aromatic Copolymers. Macromolecules 47, 2175-2198 (2014).

7. Y. Schneider et al., Ionic Conduction in Nanostructured Membranes Based on Polymerized Protic Ionic Liquids. Macromolecules46, 1543-1548 (2013).

8. H. Hu, M. Gopinadhan, C. O. Osuji, Directed self-assembly of block copolymers: a tutorial review of strategies for enabling nanotechnology with soft matter. Soft Matter10, 3867-3889 (2014).

9. M. Luo, T. H. Epps, III, Directed Block Copolymer Thin Film Self-Assembly: Emerging Trends in Nanopattern Fabrication. Macromolecules46, 7567-7579 (2013).

10. S. Ji, L. Wan, C.-C. Liu, P. F. Nealey, Directed self-assembly of block copolymers on chemical patterns: A platform for nanofabrication. Progress in Polymer Science, 54-55, 76-127 (2016).

11. S.-J. Jeong et al., Directed self-assembly of block copolymers for next generation nanolithography. Materials Today 16, 468-476 (2013).

12. M. P. Stoykovich, P. F. Nealey, Block copolymers and conventional lithography. Materials Today9, 20-29 (2006).

13. I. P. Campbell, G. J. Lau, J. L. Feaver, M. P. Stoykovich, Network Connectivity and Long-Range Continuity of Lamellar Morphologies in Block Copolymer Thin Films. Macromolecules  45, 1587-1594 (2012).

14. C. G. Arges, Y. Kambe, H. S. Suh, L. E. Ocola, P. F. Nealey, Perpendicularly Aligned, Anion Conducting Nanochannels in Block Copolymer Electrolyte Films. Chemistry of Materials28, 1377-1389 (2016).

15. X. Feng et al., Scalable Fabrication of Polymer Membranes with Vertically Aligned 1 nm Pores by Magnetic Field Directed Self-Assembly. ACS Nano8, 11977-11986 (2014).

16. K. M. Beers, D. T. Hallinan, X. Wang, J. A. Pople, N. P. Balsara, Counterion Condensation in Nafion. Macromolecules44, 8866-8870 (2011).

17. K. M. Beers, N. P. Balsara, Design of Cluster-free Polymer Electrolyte Membranes and Implications on Proton Conductivity. ACS Macro Letters1, 1155-1160 (2012).

Figure 1

2531

It is required to form nanophase-segregated structures in anion exchange polymer electrolyte membrane to develop and enhance the ion transport. In this study, we employ multi-scale modeling method to investigate the correlation of the ion transport with the nanophase-segregation in anion exchange polymer electrolyte membrane. The atomistic force field is developed using density functional theory (DFT) method, and used for molecular dynamics (MD) simulations. The annealing procedure is applied to model the nanophase-segregation. The equilibration is achieved through NPT MD simulation, and the extent of nanophase-segregation is evaluated by the structure factor analysis. The ion transport is evaluated using mean-square displacement. The nanophase-segregated structures and the transport properties are compared to the proton exchange membrane consisting of the same polymer backbone except for the acidic functional group. The coarse-grained model is also constructed based on atomistic model. Using dissipative particle dynamics (DPD) simulation, the meso-scale structures are simulated to elucidate large-scale correlation in structure.

2532

, , , , , , and

Anion-exchange membrane fuel cells (AEMFCs) provide significant advantages over their proton-exchange membrane counterparts. In the alkaline environment, the oxygen reduction reaction (ORR) is more facile, there is diminished fuel crossover, and a greater flexibility regarding fuel and catalyst choice. The membrane at the heart of AEMFCs not only facilitates the ion exchange but also separates the fuel feedstocks and acts as a support for the membrane-electrode assembly (MEA). However, to date there are still no membrane materials that satisfy all the needs (long-term stability in alkaline environment, high ionic conductivity, low swelling and good structural integrity) for use in AEMFCs and this remains one of the larger obstacles for further AEMFC development.

The amination and subsequent quarternisation of polyketone leads to a new family of ionomers containing N-substituted pyrrole moieties. The degree of amination can be controlled by manipulating reaction conditions, allowing the composition and resulting structural properties of the polymer to be tuned. Membrane fabrication results in thermally stable (TD > 250 oC), structurally robust polymer electrolytes that exhibit ionic conductivity ( > 10-3 S cm-1). These new solid-state ion conducting materials have the potential to be used in a variety of applications including AEMFCs.

Here we present an in-depth study focusing on the structure-property relationships of this new polypyrrole/polyketone polymer. A variety of analytical techniques are used to probe the thermal and structural properties of the polymers, these include high-resolution thermogravimetric analysis, modulated differential scanning calorimetry, dynamic mechanical analysis, vibrational, NMR and UV-Vis spectroscopies. In addition, broadband electrical spectroscopy is used to gauge the interplay between the structural properties and electrical response.

Figure 1

2533

, , and

Anion Exchange Membrane (AEM) fuel cells are a possible route to overcoming fundamental issues with acid-based fuel cells. The issues include the high cost of platinum catalysts, complex water management, and sluggish electrochemical reactions. However, the performance of AEM fuel cells is not as good as that of Proton Exchange Membrane (PEM) fuel cells partially because of the limitations of current anion exchange membranes, such as low ionic conductivity, poor stability at high pH, and high water uptake.

A series of partially fluorinated multiblock copoly(arylene ether)s with long head-group tether were synthesized for use in AEM fuel cells and electrolyzers. The multiblock copolymers were synthesized via polycondensation of hydroxyl-terminated oligomers and fluoro-terminated oligomers. The resulting multiblock structure has one hydrophilic block and one hydrophobic block. It was designed so that nanophase-separation occurs and efficient conductive channels are formed with low water uptake. Multiblock copolymers with different block lengths and ion exchange capacity (IEC) were synthesized to maximize ion conductivity and explore the relationship between chemical structures and membrane properties. Table 1 shows the membrane properties. A non-linear relationship was found between the number of head-groups on a hydrophilic block and the conductivity. Doubling the number of head-groups more than doubled the hydroxide conductivity. Hydroxide conductivity as high as 119 mS/cm at 80°C have been observed with a specific size block size: X5.4Y7, where X is the hydrophobic block and Y is the hydrophilic block. The hydrophobicity of the backbone has allowed synthesis of polymers with minimal water uptake. The ratio of conductivity-to-water uptake shows that less water is absorbed compared to conventional materials. The number of waters per ion within the polymer is as low as 4. This is reflected in the measurement of the amount of free-water and bound-water. No conductivity loss was observed after soaking the membrane in 1M NaOH solution at 60°C for over 700 hours.

Financial support from US Office of the Deputy Assistant Secretary of the Army for Defense Exports and Cooperation (DASA-DE&C) is gratefully acknowledged.

Figure 1

2534

, , , , and

Due to the growing concerns on the depletion of petroleum based energy resources and the climate change, polymer electrolyte membrane fuel cells (FCs) technologies have received much attention in recent years owing to their high efficiencies and low emissions. Among them, alkaline anion exchange membrane fuel cell (AEM-FC) is a good choice, because it avoids the consumption of acid-resistant precious metal catalysts and costs up.

However, the biggest challenge in developing AEM-FCs is to fabricate AEM with high ion conductivity and mechanical stability without chemical deterioration at elevated pH and temperatures. So far, most strategies were focused on synthesizing new thermally and chemically durable fluorinated and aromatic polymers. For the first time, our group tried to develop the new type of 2Me-AEMs by radiation-induced grafting of 2-methyl-1-vinylimidazole and styrene into poly(ethylene-co-tetrafluoroethylene) (ETFE) films and a subsequent N-alkylation with methyliodide. The resultant AEMs exhibit high ion conductivity (> 100 mS/cm) and longer alkaline durability, owing to the fact that the methyl protecting group at 2-imidazole position prevented the ring-opening degradation. 2Me-AEM with an IEC(ion exchange capacity) of 1.82 mmol/g shows the best well-balanced properties required for fuel cell applications. All these findings on one hand, are the result of sample preparation procedure of the radiation grafting method and the introduction of alkylimidazolium cations as an anion conducting group, and on the other hand, are believed to be controlled by the microphase separated structures of the membranes in the hydrated state. In this work, we aim to elucidate the morphology of these 2Me-AEMs and understand the structure related unique properties such as the mechanical property and the anion conductivity.

We investigated the morphology and swelling behavior of these new graft-type AEMs by using contrast variation small angle neutron scattering (SANS) technique, performed mainly on KWS-2 SANS diffractometer operated by Juelich Centre for Neutron Science at the neutron source Heinz Maier-Leibnitz (FRM II reactor) in Garching, Germany. Our results showed that the scattering intensity and the shape of the profiles vary significantly upon grafting, but change little upon alkylation; while the incorporation of water in the hydrophilic domains upon swelling, not only increases the intensity and domain size, but also leads to the excess scattering at high-q region with q being the scattering vector. From these results, we concluded that the crystalline lamellar and crystallite structures originating from the pristine ETFE films were more or less conserved in these AEMs, but the lamellar d-spacing in both dry and wet membranes were enlarged, indicating an expansion of the amorphous lamellae due to the graft chains introduced in the grafting process and the water incorporated in the swelling process. We further quantitatively studied the swelling behavior of the AEMs in various water mixtures of water and deuterated water with different volume ratios (contrast variation method), and the morphology of these membranes was elucidated by three phases: phase 1) crystalline ETFE domains, which offer good mechanical properties; phase 2) hydrophobic amorphous domains, which are made up of amorphous ETFE chains and offer a matrix to create conducting regions; phase 3) interconnected hydrated domains, which are composed of the entire graft chains and water and play a key role to promote the conductivity.

2570

, , , and

Anionic exchange membrane fuel cell is a promising new technology which potentially meets the challenges faced by other low-temperature fuel cells like Proton-exchange membranes. Anionic exchange membrane can be operated with non-noble catalyst. To design a robust material which is hydroxide conductive, chemically stable and mechanically strong is the biggest challenge to make this technology commercial. This research focuses on the characterization of a novel anionic exchange membrane with PTFE backbone. The experimental ion-exchange capacity of this polymer is 0.9 mmol/gm. The conductivity of this membrane was measured in its chloride form. We observed the highest value of 0.039 S/cm conductivity at 80 0C and 95%RH. The polymer also shows water uptake of 11 % and λ value of 6.8 at 60 0C and 95% RH. The small-angle x-ray scattering analysis of this membrane shows the polymer's ionomer peak and we understand that the swelling of the polymer is higher at higher temperature. From this analysis, this material can be expected to exhibit higher hydroxide conductivities and good mechanical stability and hence this polymer is a promising candidate for anionic exchange membranes.

Keywords: Anionic exchange membrane, PTFE backbone

Figure 1

D-21 Catalyst Layers 2 & Carbon-based Supports 1 - Oct 5 2016 8:00AM

2535

, , , and

A simple but effective electrochemical route to functionalize graphene is demonstrated. Cyclic voltammetric sweeps are performed in 0.5 M H2SO4 aqueous solution on electrodes containing carbon cloth, graphene, and Nafion ionomer. With supply of ambient oxygen, the formation of hydroxyl radicals from the oxygen reduction reaction during CV cycles initiates the decomposition of Nafion ionomer that engenders oxygenated functional groups on the graphene surface. Raman analysis suggests a minor change for the graphene structure. Exploring various amounts of Nafion ionomer, we determine the optimized conditions for graphene functionalization. Contact angle is used to evaluate the effect of Nafion ionomer decomposition. X-ray photoelectron spectroscopy is employed to study relevant functional groups. Afterwards, nanoparticles of Pd9Ru are synthesized and impregnated on those functionalized graphene (FGN) via a wet chemical reflux process. X-ray diffraction patterns of the as-synthesized samples suggest successful formation of alloy without presence of individual Pd and Ru nanoparticles. Images from transmission electron microscope confirm the average size of 3-4 nm. Subsequently, the Pd9Ru/FGN undergoes Cu under potential deposition, followed by a galvanic displacement reaction to deposit a Pt monolayer on the Pd9Ru surface (Pd9Ru@Pt). Intensive experimental work is ongoing to investigate the electrochemical active surface area (ECSA) and electrocatalytic activities of Pd9Ru@Pt/FGN.

2536

, and

Support carbon corrosion has been identified as a major degradation source in a polymer electrolyte fuel cell. Start-up/Shutdown (SUSD) [1], fuel starvation [1], and voltage reversal [2] are widely-known conditions that can trigger non-negligible carbon corrosion in the anode or cathode. SUSD carbon corrosion is associated with current reversal while there is no external load. As a comparison, fuel starvation due to nitrogen blanketing and/or water droplets in an operating cell also can result in similar current reversal. The region with current reversal, i.e., a proton flux from the cathode to the anode, features extremely severe hydrogen depletion in the anode and occurrence of carbon corrosion in the opposite cathode. Differently, the voltage reversal occurs in one or a few cells (not all) in a stack. Those cells show negative cell voltages under extended reversal, but the overall current is flowing as usual thanks to the other normal cells in the stack. The voltage reversal leads to carbon corrosion in the anode, as compared to carbon corrosion in the cathode for current reversal. The local potential for carbon corrosion could be much more severe in an extended voltage reversal than that in SUSD.

This paper/presentation will elucidate the fundamental differences of these three carbon corrosion scenarios with model simulation and analysis. The models [3, 4] will also be used to perform comprehensive parametric studies, aiming to understand the impacts from key design and operating parameters including pseudo-capacitance, selective oxygen evolution reaction (OER) catalysis, reactant gas flow rate, temperature, membrane hydration state, and current load drawn from the stack. It is anticipated to shed light on the otherwise confusing carbon corrosion scenarios in a polymer electrolyte fuel cell and the fundamental ideas for alleviation strategies. For example, in the fuel starvation induced carbon corrosion, the simulation indicates that the decreasing cell voltage due to diluted hydrogen concentration appears to limit the local potential in the cathode and associated carbon corrosion. In the SUSD carbon corrosion, pseudo-capacitance from Pt oxidation may raise another concern in parallel to carbon corrosion. In general, a reduced temperature or enhanced OER activity is helpful in reducing carbon corrosion.

References:

[1] W. Gu, P. T. Yu, R. N. Carter, R. Makharia, and H. A. Gasteiger, "Modeling and Diagnostics of Polymer Electrolyte Fuel Cells— Local H2Starvation and Start–Stop Induced Carbon-Support Corrosion", Modern Aspects of Electrochemistry 49, Springer Science + Business Media, LLC, New York (2010).

[2] S. Knights, D. Wilkinson, S. Campbell, J. Taylor, J. Gascoyne, and T. Ralph, "Solid Polymer Fuel Cell with Improved Voltage Reversal Tolerance", U.S. Patent, 6,936,370 B1 (2005).

[3] J. Chen, J. Siegel, T. Matsuura, and A. Stefanopoulou, "Carbon Corrosion in PEM Fuel Cell Dead-Ended Anode Operations", Journal of the Electrochemical Society, 158, B1164-B1174 (2011).

[4] J. Chen, J. Hu, and J. Waldecker, "A Comprehensive Model for Carbon Corrosion during Fuel Cell Start-Up", Journal of the Electrochemical Society, 162, F878-F889 (2015).

Figure 1. The schematic of carbon corrosion mechanisms from current reversal during start-up (upper subfigure) and voltage reversal (lower subfigure) in a polymer electrolyte fuel cell.

Figure 1

2537

, , , and

Lifetime of polymer electrolyte fuel cells (PEFCs) is limited by degradation mechanisms, often caused by undesirable operating conditions. Depending on the application, the likeliness of these operation modes varies; e.g. mobile applications inevitably undergo more start-up/shut-down cycles and dynamic load and humidity changes than stationary systems [1]. On the other hand, in stationary applications with steady load conditions, other critical events, such as cathode starvation, become lifetime defining issues. In case of oxygen undersupply, protons are reduced on the cathode catalyst instead of oxygen, leading to hydrogen evolution. However, with oxygen still present near the cathode inlet, a strong current and temperature gradient is induced. The temperature rises near the inlet, whereas it remains low in the areas of proton reduction reaction. This is accompanied by voltage oscillation due to the low inlet flux and thus insufficient water removal [2,3].

In order to study this particular situation, accelerated stress tests (ASTs) are used. However, so far, most ASTs were not suitable for portraying real cathode supply failure, as they do not alter the cathode gas flow under stable load conditions. Therefore, in this work, the cathode gas flow rate was varied in order to minimise the effects of other operating modes, contributing to the degradation of fuel cells.

Multiple ASTs with short intervals of reduced cathode gas flow were conducted analogously to the AST displayed in Figure 1. All tests were performed in a single hydrogen fuel cell with a segmented S++ current scan shunt device on the cathode for a spatial detection of current and temperature in 10x10 and 5x5 segments, respectively (Figure 2). The fuel cell was electrochemically characterised in-situ via polarisation, electrochemical impedance, linear sweep and cyclic voltammetry measurements [4]. Additionally, the fluoride emission rate (FER) was determined by effluent water analysis.

While a mild cathode starvation did not lead to a fundamental voltage drop, the cell voltage decreased to values below 0 V for a more detrimental cathode undersupply. A very distinct gradient in current and temperature was detected for the latter tests. Both temperature and current exhibited peak values near the cathode inlet. The current did, however, not drop to values below zero at the outlet, indicating that a second cathode reaction is taking place in areas of oxygen starvation. Thus, protons were inevitably reduced instead of oxygen at the cathode outlet in order to maintain the set current. The change in local behaviour of current and temperature before and after 100 hours of AST at a set current density of 0.6 A cm-² is shown in Figure 3.

Acknowledgement

The research leading to these results has received funding from the European Union's Seventh Framework Programme (FP7/2007-2013) for the Fuel Cells and Hydrogen Joint Technology Initiative under grant agreement n° 621216.

[1] Brightman E, Hinds G. In situ mapping of potential transients during start-up and shut-down of a polymer electrolyte membrane fuel cell. J Power Sources 2014;267:160–70. doi:10.1016/j.jpowsour.2014.05.040.

[2] Gerard M, Poirot-Crouvezier J-P, Hissel D, Pera M-C. Oxygen starvation analysis during air feeding faults in PEMFC. Int J Hydrogen Energy 2010;35:12295–307. doi:10.1016/j.ijhydene.2010.08.028.

[3] Wang YX, Xuan DJ, Kim YB. Design and experimental implementation of time delay control for air supply in a polymer electrolyte membrane fuel cell system. Int J Hydrogen Energy 2013;38:13381–92. doi:10.1016/j.ijhydene.2013.06.040.

[4] Bodner M, Hochenauer C, Hacker V. Effect of pinhole location on degradation in polymer electrolyte fuel cells. J Power Sources 2015;295:336–48. doi:10.1016/j.jpowsour.2015.07.021.

Figure 1

2538

Pt based fuel cell catalysts are well known since many decades. The main principles of tuning activities and stabilities of especially Pt based Carbon Black fuel cell catalysts seems to be well understood. Therefore only little interest has been shown in new research and developments activities in Pt fuel cell catalysts based on commercial Carbon Blacks. Form pure scientific research view there seems to be only little or nothing new requiring a deeper understanding of the main catalytic principles and properties regarding pure Pt based Carbon Black fuel cell catalysts.

On the other hand Pt based Carbon Black fuel cell catalysts are the most important commercial fuel cell catalysts. The starting and growing mass production of these type of fuel cell catalysts for all different applications requires a well-controlled synthesis process and quality check procedure including activity testing in MEA´s. Controlling the synthesis process on industrial scale on one hand requires a deep understanding of all properties of the final Pt based carbon Black catalyst. On the other hand the property control of the catalyst material is done by controlling all process steps including the understanding of correlations in-between different process parameters.

From industrial view the challenge of producing such simple and for decades well known Pt based Carbon Black catalysts starts with the reproducibility of the catalytic important properties from batch to batch and in all different batch sizes. In addition to this challenge of manufacturing reproducible amounts of fuel cell catalysts the question arises how to measure the main catalytic important properties, such as the mass activity, air performance and ECSA, in a reproducible way using dependably equipment and test procedures.

In this talk all aspects regarding the question of reproducible synthesis processes on different manufacturing scales as well as determining the catalytic properties in a reproducible way will be discussed using the example of such a simple catalyst type such as the Pt based on Carbon Black fuel cell catalyst.

2539

, , , , , , , , and

Development and application of the proton exchange membrane fuel cell (PEMFC) based transport systems is one of the urgent problems for mega-cities and densely populated areas. However, low redox stability of the carbon supports in strongly oxidising/reducing fuel cells working conditions is the major unsolved problem in materials technology, hindering the wide scale application of PEMFC. Therefore the novel hierarchically microporous-mesoporous carbon supports for deposition of various catalytically active particles have been tested in order to establish the best materials with high cyclability. The carbon materials developed and used as PEMFC catalyst supports in this work have been tested for many years in electrical double layer supercapacitors under very high current density pulsation conditions.

In the last years the synthesis methods for carbide derived carbon (C(VC), C(Mo2C), C(TiC), C(SiC) [1, 2] and carbon nanospheres [3, 4] have been developed and corresponding Pt, Ir, Ru-Ir, as well as d-metals nanoclusters activated catalysts have been synthesised [5-8].

All carbon as well as catalyst materials synthesised have hierarchically microporous-mesoporous structure verified by SEM-EDX, FIB-TOF-SIMS, high-resolution TEM and Brunauer-Emmett-Teller gas adsorption methods. The carbon materials have been tested by small angle neutron scattering methods and variable shape of pores has been verified [9].

The Ru, Pt and Ir nanoparticles were deposited onto a carbon support, using the well established liquid-phase reduction methods [5-8]. Thermogravimetric and thermodynamic analysis, X-ray diffraction and Raman spectroscopy methods were used in order to characterize the structure of the materials synthesised. The catalysts prepared were tested as electrodes in the three-electrode as well as in the single cell conditions. The polarization, rotating disk electrode, as well as chronoamperometric measurements were carried out in order to evaluate the activity and stability of the catalyst materials synthesised. The rotating disk electrode data show that the degradation of catalytic activity during 30 cycles was moderate.

The electrochemically active surface area has been calculated from the hydrogen adsorption data. The electrochemical impedance method has been used to calculate the electrolyte resistance, polasization resistance and activation energy values.

The assembled single cells demonstrated excellent cyclability (3000 cycles) and very high catalytic activity, thus noticeably better performance than those based on the traditional Vulcan electrode materials.

Acknowledgements

This work was supported by Estonian Target research project No. IUT20-13, Estonian Centre of Excellence projects (Nos. 3.20101.11-0030 and 2014-2020.4.01.15-0011), and Personal Research Grant PUT55.

References

  • A. Jänes, L. Permann, M. Arulepp, E. Lust, Electrochem. Commun. 6 (2004) 313−318.

  • E. Tee, I. Tallo, H. Kurig, T. Thomberg, A. Jänes, E. Lust, Electrochim. Acta 161 (2015) 364−370.

  • T. Thomberg, T. Tooming, T. Romann, R. Palm, A. Jänes, E. Lust, J. Electrochem. Soc. 160 (2013) A1834−A1841.

  • I. Tallo, T. Thomberg, H. Kurig, K. Kontturi, A. Jänes, E. Lust, Carbon 67 (2014) 607−616.

  • E. Lust, K. Vaarmets, J. Nerut, I. Tallo, P. Valk, S. Sepp, E. Härk, Electrochim. Acta 140 (2014) 294-303.

  • S. Sepp, K. Vaarmets, J. Nerut, I. Tallo, E. Tee, H. Kurig, J. Aruväli, R. Kanarbik, E. Lust, Electrochim. Acta 2016, doi:10.1016/j.electacta.2016.03.158.

  • S. Sepp, E. Härk, P. Valk, K. Vaarmets, J. Nerut, R. Jäger, E. Lust, J. Solid State Electrochem. 18 (2014) 1223-1229.

  • R. Jäger, P. Ereth Kasatkin, E, Härk, E. Lust, Electrochem. Commun. 35 (2013) 97-99.

  • H. Kurig, M. Russina, I. Tallo, : Siebenbürger, T. Romann, E. Lust, Carbon 100 (2016) 617-624.

2540

, , , , and

Catalyst poisoning by specific adsorption of the sulfonic acid anion of ionomers has been known as one of the main performance-limiting factors of PEFCs. In a recent study, it was reported that the catalyst poisoning can be suppressed by using a mesoporous carbon (mesoporous carbon nano-dendrites) as a cathode catalyst support [1]. Here, we present that a similar effect can be obtained by using a MgO-templated mesoporous carbon [2] having meso-pores the sizes of which are more-systematically controllable by changing the size of the MgO templates than for conventional carbon supports.

An MgO-templated carbon denoted as CNovel® was purchased from Toyo Tanso Inc. The mean diameter of its mesopores is ca. 5 nm (the template MgO size was 5 nm). The sample was graphitized by a heat treatment at 2,100 °C for improving the chemical stability under electrochemical conditions and was crushed by a bead mill to form secondary particles smaller than hundreds nm in size. The obtained powder was immersed in 1 M HNO3 at 80 °C for 1 h to introduce hydrophilic surface groups for facilitating Pt deposition. After thoroughly washing it with pure water, platinum nanoparticles were deposited in aqueous media by using hexahydroxy-platinate as a platinum source. The ORR activity of the prepared catalyst was evaluated by the thin-film RDE method (without ionomers) in aqueous media (0.1 M HClO4) and in MEAs prepared with a Nafion® ionomer (I/C=1.0). The results were compared those of Pt/Vulcan.

Figures 1a and 1b show that TEM images of the catalysts. Platinum particles of ca. 2 nm are observed on both the carbons. While Pt/Vulcan shows platinum particles only on the outer edge of the carbon particle, Pt/CNovel shows platinum particles not only on the outer edge but also on the mesopores inside the carbon particle. Figure 1c shows the I-V curves of MEAs using Pt/CNovel and Pt/Vulcan. Pt/CNovel exhibits higher cell-voltage than Pt/Vulcan in the low current density region. This result indicates that the former has a higher ORR activity than the latter as specifically indicated in the specific activities at 0.9 V shown in Figure 1d. The trend is, however, opposite in the RDE measurements while the specific activities for both samples by RDE are higher than those measured by the MEAs. These results indicate that the Pt/CNovel suffers less catalyst poisoning by the sulfonic acid anion than Pt/Vulcan in MEAs probably because the anion linked to the ionomer cannot reach the surfaces of the Pt nanoparticles inside the mesopores of CNovel while all Pt nanoparticles on Vulcan® can be reached by the ionomer.

References

[1] M. Hori, et al., 225th ECS Meeting Abstract, #1500 (2013)

[2] T. Morishita, et al., Carbon, 48, 2690 (2010)

Figure 1

2541

, , , , and

In the development of PEMFC fuel cells much attention and development is devoted to the "active" portion of the catalysts system, the precious metal nano-particles. While this approach has led to great improvement in specific activity, existing catalysts are still not meeting the combination of performance and cost necessary for the commercialization of PEMFC.

A key limitation to catalysts today arises from reliance on carbon black(CB) supports that serve to disperse and stabilize the precious metal catalyst nanoparticles, but which corrode under certain operating conditions. Carbon support modifications that rely on "top-down" graphitization of CB's have so far improved resistance to carbon corrosion, but at the cost of reduced stabilization of precious metal dissolution.

Pajarito Powder developed a "bottom-up" approach where Engineered Carbon Support are made from organic precursors that yield a highly tailorable carbon support with great corrosion resistance. This method which is based on technologies licensed from the University of New Mexico, Northeastern University, and Los Alamos National Laboratory allows manufacture of highly graphic carbonaceous powders with excellent control of pore sizes and particle roughness.

Pt/C catalysts made using these Engineered Carbon Supports maintain performance under carbon corrosion test protocols and thus demonstrate their utility in achieving the performance/price balance needed for commercializing PEMFCs. The catalyst characteristics and performance, as well as economics of this approach will be presented and discussed.

Figure 1

2542

and

Pt has been considered the best element for ORR in acidic condition in the point of activity, stability and power density. Pt catalysts for ORR have been often prepared using Pt compounds as starting material in aqueous solutions. The melting temperature, Tamman temperature and Huettig temperature of Pt are much higher than those of Pt compounds such as chlorides and oxides unlike Ni, Cu (1). It might be effective to deposit Pt metal directly on support to get better dispersed catalyst particles compared with the conventional methods. A liquid phase process and gas phase processes were found to be able to afford higher activity than conventional methods for various supporting materials. It was found that specific CV curves and extremely high mass activity in the case of CNT with low D/G as a catalyst support. as shown in Fig.1. The measurement condition was executed under the condition deciced based on the discussion with Y. Takasu and W. Sugimoto (2). It is very interesting that high ORR mass activity and the strong shift of HUPD to lower potential can be observed at the same time(3,4) in the electronic structure of Pt surface. We have started to discuss this behavior related with the surfacee electronic structure of Pt, the bonding between Pt and C, stability of the catalyst.  

References

1. J.A. Moulijn, B. Silberova, M. Makkee and F. Kapteijn, "Catalyst Deactivation and Process Design", 10th Internatinal Symposium on Catalyst Deactivation, February 5-8, 200 in Berlin/Germany.

2. Y. Takasu, W. Sugimoto, M. Yoshitake, "Examination of the Evaluation Method with the Common Catalysts for PEFC Cathode-3(Summary)", State-of-the-art Fuel Cells and Hydrogen Technology in Japan, - Collected from the 20th FCDIC Fuel Cell Symposium and the 1st FCV Forum, Fuel Cell Development Information Center, (2014).

3. M. Yoshitake et al, "Electrode material and method for producing same", WO2012053561 A1.

4. M. Yoshitake, "Development of PEFC at Asahi Glass Co., Ltd.", IUMRS-ICEM2010, Symposium H. Materials and Devices for FUel Cells, August 26,2010, Korea.

Figure 1

2543

, , , , , and

Carbon blacks supporting Pt, currently widely used as electrocatalysts in Polymer Electrolyte Membrane Fuel Cells (PEMFC) are thermochemically unstable in PEMFC operating conditions. This is especially true at the cathode side where, on top of relatively elevated temperature (80°C) and acidic conditions, both the potential and the relative humidity may be high. The resulting carbon oxidation is partially responsible for the PEMFC performance decrease observed over time. Hence, long term durability still needs to be improved in order to consider PEMFC as credible alternatives to conventional power sources for automotive, stationary or portable applications.

Much effort have been directed to identify and synthesize alternative carbon materials as catalyst supports for PEMFCs. One strategy to decrease carbon support corrosion is to use carbon with high extent of graphitization, which is supposed to decrease defect sites on the carbon structure, where carbon oxidation starts [1], [2]. However high graphitic content of carbon can be a brake for particle nucleation and dispersion. Among the different forms of carbon, graphene and carbon nanotubes have attracted tremendous interest materials for various energy applications [3], [4]. We will show how the combination of their high surface area, high conductivity and high chemical stability makes these 1D and 2D materials promising candidates for cathode catalyst support in PEMFCs.

In this work, platinum and bimetallic (PtM, M=Co or Ni) nanoparticles catalysts have been prepared on N-doped Multi-Walled Carbon Nanotubes (MWCNTs) (home synthetized and commercial) and Few Graphene Layers (FGLs) using three synthetic routes: polyol route [5], impregnation and reduction under H2 [6] and SBH reduction [7]. The Platinum loading, the dispersion and size of metallic nanoparticles were analyzed by UV, XRD, SEM. Our materials showed good dispersion on nanostructured support and size of metallic particles reached 2 nm (figure 1).

We investigated the properties of 2D FGLs/1D MWCNTs Pt and PtM catalysts for electrocatalysis of oxygen reduction. By comparing the electrochemical properties of these hybrid materials with a commercial Pt/C catalyst using carbon blacks as carbon support, it is found that this hybrid material demonstrated an enhancement of electro-catalyst performances in RDE tests. Moreover accelerated stress tests in half cell and fuel-cell setup demonstrated that the use of this hybrid FGLs/MWCNTs support can be promising in effectively reducing the carbon corrosion and then increase lifetime of the cell.

This work is funded by the FP7 NANOCAT European program (SP1-JTI-FCH.2012.1.5) and French region Midi-Pyrénées.

REFERENCES

[1] Xin Wang , Wenzhen Li , Zhongwei Chen , Mahesh Waje , Yushan Yan, "Durability investigation of carbon nanotube as catalyst support for proton exchange membrane fuel cell", Journal of Power Sources, 2006, 158, pp 154–159

[2] D. Bom, R. Andrews, D. Jacques, J. Anthony, B. Chen, M. S. Meier, John P. Selegue , "Thermogravimetric Analysis of the Oxidation of Multiwalled Carbon Nanotubes: Evidence for the Role of Defect Sites in Carbon Nanotube Chemistry", Nano Letters, 2002, 2, pp 615-619

[3] Yingwen Cheng, Songtao Lu, Hongbo Zhang, Chakrapani V. Varanasi⊥, and Jie Liu, "Synergistic Effects from Graphene and Carbon Nanotubes Enable Flexible and Robust Electrodes for High-Performance Supercapacitors", Nano Letters, 2012, 12 (8), pp 4206–4211

[4] Zhou, X., Qiao, J., Yang, L. and Zhang, J "A Review of Graphene-Based Nanostructural Materials for Both Catalyst Supports and Metal-Free Catalysts in PEM Fuel Cell Oxygen Reduction Reactions", Adv. Energy Mater., Vol. 4, (2014) p.1301523

[5] Yongjie Li, Wei Gao, Lijie Ci, Chunming Wang, Pulickel M. Ajayan "Catalytic performance of Pt nanoparticles on reduced graphene oxide for methanol electro-oxidation", Carbon (2010), 48, 1124-1130.

[6] S. M. Choi, M. H. Seo, H. J. Kim, W. B. Kim, "Synthesis and characterization of graphene-supported metal nanoparticles by impregnation method with heat treatment in H2 atmosphere", Synthetic Metals 161 (2011) 21–22, 2405–2411.

[7] D. Wang et al., Nature Materials, "Synthesis and characterization of Cocore–Ptshell electrocatalyst prepared by spontaneous replacement reaction for oxygen reduction reaction" 2013, 12, 81-87.

Figure 1

2544

, , , , and

Polymer electrolyte membrane fuel cells (PEMFCs) have been widely accepted as a clean energy device because of its capability of generating high energy density and only water as a product via direct converting of the chemical energy into the electrical energy. However, durability problems originating from the carbon corrosion and Pt agglomeration have been one of the major issues for practical applications. The high potentials formed in a cell during the fuel cell operations lead to serious carbon corrosion in catalyst layers, which causes the decrease of cell performance. To resolve the carbon corrosion, highly durable carbon materials such as carbon nanotubes (CNTs) have been intensively studied. Although the corrosion-resistant characteristics of CNTs were efficient to utilize as an electrocatalyst support for PEMFCs, the difficulty in processing the CNTs mainly due to poor dispersibility and easy bundle formation in a catalyst ink limited their practical applications. In this respect, vertically aligned carbon nanotubes (VACNTs) have been tried and showed promising results in a PEMFC MEA. However, VACNTs need complex process to make MEAs, remaining the issues of mass production.

In this work, we report CNTs grown on Al2O3 (Al2O3-CNT) as an efficient support material for PEMFC MEAs. The catalyst support was designed to utilize durable CNT properties while resolving the CNT bundling problems. The catalyst was characterized by transmission electron microscope (TEM), field emission scanning spectroscopy (FESEM), and X-ray photoelectron spectroscopy (XPS) in order to investigate structural and electronic properties. The kinetics of O2 reduction is examined with the rotating disk electrode (RDE) technique. A potential cycling test was applied to evaluate durability of the catalyst. The electrochemical surface area (ECSA) of Pt/Al2O3-CNT slightly decreased (17 %), while Pt/C loses nearly 46 % of its ECSA of its initial value over repeated cycling test under carbon corrosion potential range, representing more durable properties of CNTs. The single cell test was also performed with Pt/Al2O3-CNT and Pt/C after fabricating the catalysts to MEAs. The performance of the Pt/Al2O3-CNT was comparable to that of Pt/C and showed better performance at high current density regions. Furthermore, after the accelerated stress test (AST), the Pt/Al2O3-CNT also showed approximately 2 times higher stability than Pt/C in a single cell test.

2545

, , , and

A new approach for the preparation of PEMFC cathode electrocatalysts based on thin layers of platinum deposited on a support instead of conventional Pt nanoparticles supported on carbon black could be achieved adopting various Pt deposition techniques including atomic layer deposition, electrochemical atomic layer epitaxy, self-terminated Pt electrodeposition, thermal dealloying, and galvanostatic displacement of Ni and Cu ions with Pt1,2. This morphology allows for a more effective utilisation of the highly priced noble metal, while increases the durability of the electrodes over time.

The morphology and nature of the support material also plays a crucial role. Nanostructured 1D materials have attracted significant research attention due to the influence of their nanostructure and porosity on performance and durability of the PEMFC3. In this regard, electrospinning is a reliable and scalable technique for the preparation of nanofibers with controlled and uniform diameters and structures. Its versatility allows the production of organic, hybrid and inorganic nanofibres, as well as the possibility of creating different geometries, assemblies and architectures4.

We developed novel nanofibrous electrodes (NFE) based on 3D nanostructured networks of carbon nanofibers covered by Pt nanoislands, merging the advantage of Pt thin layers catalyst and a highly porous robust support. This was achieved by combining electrospinning and high overpotential pulsed Pt electrodeposition.

These catalysts are based on self-standing electrospun carbon nanofibers covered by platinum thin platinum nanoislands. NFEs are highly durable when compared to commercial standards and have a higher degree of platinum exploitation. Electrochemical active specific surface areas as high as 140 m2g-1Pt with retention in the order of 70 % after accelerated stress testing are measured, making them a suitable alternative to currently available commercial catalyst.

1. S. M. Alia, Y. Yan, and B. Pivovar, Catal. Sci. Technol., 4, 3589–3600 (2014)

2. G. Ercolano, S. Cavaliere, D. Jones, and J. Rozière, ECS Trans., 69, 1237–1242 (2015).

3. S. Cavaliere et al., ChemElectroChem, 2, 1966–1973 (2015).

4. S. Subianto et al., in Electrospinning for Advanced Energy and Environmental Applications, vol. 1, p. 29–60 (2015).

C-21 Membrane Fabrication and Testing I - Oct 5 2016 8:00AM

2546

High temperature polymer electrolyte membranes containing basic units that are able to interact with strong protic acids providing ionically conductive composite membrane have shown respectable performance and long operational stability. Further increase of the operation temperature and even an increase on the obtained conductivity values would facilitate the construction of efficient systems providing both electricity and heat, increasing thus the total efficiency.

Initial attempts to stabilize the HTPEM electrolyte membranes through covalent crosslinking have proven that the operation temperature can be shifted up to 220oC 1,2. Different crosslinking methodologies have been tested to stabilize either PBI or aromatic polyethers bearing pyridine units3,4. In the present work we focused on the optimization of the crosslinking methodology using side double bonds and we also develop new crosslinking methodologies leading to wholly aromatic crosslinked structures. Different type of crosslinkers have been used, controlling thus the acid doping ability of the final membranes while the use of dopable crosslinkers that enable the high acid uptake and retention. Moreover, depending on the chemical structure of the crosslinked membranes, a significant increase of the conductivity values of the final acid doped composite membranes was obtained. High temperature operation as the one achieved here enables the use of liquid feed like methanol in a compact methanol reformer-high temperature PEM fuel Cell setup that is under construction and testing5.

References: 1. K.D. Papadimitriou, F. Paloukis, S.G. Neophytides, J.K. Kallitsis, "Cross-Linking of Side Chain Unsaturated Aromatic Polyethers for High Temperature Polymer Electrolyte Membrane Fuel Cell Applications" Macromolecules 44, 4942–4951 (2011)

2. C.I Morfopoulou, A.K. Andreopoulou, M.K. Daletou, S.G. Neophytides J.K. Kallitsis "Cross-linked high temperature polymer electrolytes through oxadiazole bond formation and their applications in HT PEM fuel cells" Journal of Materials Chemistry A: Materials for Energy and Sustainability1 (5) 1613-1622, (2013)

3. "Pyridine Containing Aromatic Polyether Membranes" J.K Kallitsis, A.K Andreopoulou, M. Daletou, S. Neophytides; in "Part I: Approaches to HTPEM Fuel Cells" of "High Temperature Polymer Electrolyte Fuel Cells - Approaches, Status and Perspectives" Springer International Publishing AG, Cham; Edited by J.O. Jensen, D. Aili, H.A. Hjuler, Q. Li, 2016.

4. K.D. Papadimitriou, M. Geormezi, S.G. Neophytides, J.K. Kallitsis "Covalent cross-linking in phosphoric acid of pyridine based aromatic polyethers bearing side double bonds for use in high temperature polymer electrolyte membrane fuel cells' Journal of Membrane Science 433, 1–9 (2013)

5. G. Avgouropoulos, S. Schlicker, K.-P. Schelhaas, J. Papavasiliou , K.D. Papadimitriou, E. Theodorakopoulou, N. Gourdoupi, A. Machocki, T. Ioannides, J.K. Kallitsis, G. Kolb, S. Neophytides "Performance evaluation of a proof-of-concept 70W internal reformingmethanol fuel cell system. Journal of Power Sources, 307 875-882 (2016)

Acknowledgment: Financial support from the Fuel Cell and Hydrogen Joint Undertaking (FCH JU), program "Development of a Portable Internal Reforming Methanol High Temperature PEM Fuel Cell System" IRMFC-FCH-JU-325358 is acknowledged.

2547

, , and

Perfluorosulfonic acid (PFSA) ionomers have long been used to fabricate membranes for proton exchange membrane fuel cells (PEMFC). When used in combination with mechanical reinforcements and peroxide scavenging additives, this class of ionomers has demonstrated excellent durability and performance in both accelerated lab testing and in end-use applications.

Automotive traction applications, among others, have very demanding performance targets at relatively high temperature and low humidity, driven by the need to reduce overall system cost. In order to meet these targets, ionomers with increased proton conductivity are required. Under dry conditions, conductivities of commercial ionomers are insufficient to achieve several of the targets, as they are limited to equivalent weights (EWs) of about 700 g/mol. Not only are low EW PFSA ionomers difficult to polymerize, they also tend to become water soluble as the amount of the tetrafluoroethylene (TFE) co-monomer is reduced.

One strategy for achieving lower equivalent weight, while maintaining suitable TFE content, is a multi-acid side chain (MASC) approach. We have developed a synthetic process to use the 3M precursor polymer to prepare a new ionomer containing both bis(sulfonyl)imide and sulfonic acid groups in each sidechain, to form the perfluoro-imide acid (PFIA) ionomer. Both functional groups are strong acids, and their combination results in EWs in the range of 620-650 g/mol, with increased conductivity under all conditions, including low humidity. This approach has been extended to include multiple bis(sulfonyl)imide groups per side chain to form materials classified as perfluoro-ionene chain extended (PFICE) ionomers, with EWs as low as 440 g/mol.

Characterization of these new ionomers for conductivity, solubility, and performance will be presented, along with a discussion of the practical limits of the MASC approach. Like PFSA ionomers, the PFIA and PFICE ionomers require both mechanical reinforcement and stabilizing additives in order to fabricate a viable PEMFC membrane. Development of membrane design considerations will be presented.

2548

, , , , , , and

Inorganic-organic nanostructured hybrid membranes of polyethersuflone (PES)-polyvinylpyrrolidone (PVP) were prepared with mesoporous silica materials. Hollow mesoporous silica is synthesized via cationic surface-assistant etching method, while amino-functionalized hollow mesoporous silica and amino-functionalized mesoporous silica spheres were synthesized via post-grafting strategy. After the addition of mesoporous silica and amino-functionalized mesoporous silica into the matrix of PA doped PES-PVP composite membranes, all composite membranes show similar PA uptake. However, the proton conductivity of the composite membranes increases significantly with the substantial decrease in the activation energy for proton diffusion, especially for the amino-functionalized hollow mesoporous silica material. That reason is most likely due to the facilitated proton transportation in the ordered mesoporous channels via the hydrogen bond between the –NH2 groups and H3PO4. Cell performance also confirms the superiority of the addition of inorganic fillers in PES-PVP membrane, and the highest peak power density at 180 oC was 480 mW cm-2 for NH2-HMS based composite membrane, which is 92.7 % higher than that of PA doped PES-PVP composite membrane at the identical condition. The results show promising application of NH2-HMS based PES-PVP composite membrane for elevated temperature proton exchange membrane fuel cells.

Figure 1

2549

, and

Intermediate Temperature PEFC (IT-PEFC) operating between 100 and 120 °C benefit from faster chemical reactions and more facile water and heat management as compared to conventional PEFC operating below 100 °C. As the operating temperature is increased, the temperature difference between external environment and the fuel cell also increases and heat rejection and removal by heat exchangers is enhanced. Smaller cooling systems can therefore be used reducing the weight and economic costs of PEFC systems. Moreover, the specific operating conditions of IT-PEFC enable water to remain purely in the vapour phase. Thus, the water balance is easier to maintain unlike in ordinary PEFC, which contain water in dual phase. As a consequence, water will not condensate in the flow channels, minimising flooding and interference with the gas flow. Despite these advantages, the main issue in IT-PEFCs is the durability and performance of the polymer electrolyte membrane (PEM), which lies in the heart of the fuel cell. Commercial Nafion membranes and equivalent perfluorinated sulphonic acid (PFSA) membranes are not able to hold enough water at 120 °C and low humidification conditions will eventually cause them to dry out, resulting in poor IT-PEFC performance. This is due to the dependence of the proton transport mechanisms (hopping and vehicular) and hence proton conductivity on the water content of the membrane.

Due to various oxygen groups on the basal and edge planes, graphene oxide (GO) is highly hydrophilic and a good electronic insulator. These properties led GO to be used in PEMs. In the last ten years, GO and functionalised GO have been used as additives in composite membranes to improve the membrane properties. These composite membranes have been fabricated with varied polymers such as Nafion, PBI and PSU. In recent work, GO has also been used in multilayer membranes. In composite membranes water retention due to GO can be extremely high leading to irregular expansion and affecting the membrane dimensionality. In multilayer membrane (MM) systems, however, the external layers can limit the expansion. Besides, the external layers of Nafion can avoid water retained in the inner layer of GO being transported towards the electrodes. GO is able to hold water at higher temperatures trapping the molecules in the structure even at intermediate temperature conditions, thus making it ideal for use in IT-PEFC.

This work investigates a MM with three layers composed of Nafion in the external layers and GO in the inner layer for application in an IT-PEFC. GO multilayer membranes were prepared by two different routes, 1) hot pressing, and 2) casting. Hot pressing is a faster process and minimises solvent issues. However, the interlayer-interaction in this case is only mechanical and hence relatively weak. This may result in delamination. Casting is a more involved route, but the interaction between the layers is chemical in nature and stronger. In order to compare the influence of the preparation method on the interfacial properties these two membrane types were evaluated for water retention and ion exchange capability. Proton conductivity and single cell IT-PEFC testing were carried out to evaluate the overall performance of the IT-PEFC. Preliminary results revealed that the cast N/GO/N MM presented better results, especially when comparing the proton conductivity. In both cases, the adherence between the inner layer of inorganic GO and the external layers of organic Nafion was not ideal. As such, a small percentage of Nafion was mixed into the inner GO layer. For both methods, hot pressing and casting, optimisation of the Nafion and GO content in the inner layer was performed in order to enhance interlayer interaction.

2550

, , , , , and

Graphene is increasingly being studied for fuel cell catalysis because of its high conductivity, large surface, and chemical stability.[1,2] Graphene oxide (GO) is a monolayer carbon material with carboxyl, hydroxyl and epoxy oxygen functional groups on the surface and edges. These make GO hydrophilic, and the sp2/sp3-hybridization is believed to make it insulating to electrons.[3] Fuel cell membranes have several important requirements. They must be insulating to electrons, have significant ionomer conductivity, provide sufficient barrier to prevent hydrogen/methanol crossover, and have the mechanical strength to be able to support the cell. GO could potentially fulfil these requirements. [4,5] Here, we investigate the suitability of GO for fuel cell membranes by measuring the gas barrier properties, proton conductivity, anion conductivity, electronic conductivity, tensile strength, and fuel cell performance.

Initially we reported the characterization of a graphene oxide membrane fuel cell (GOMFC).[6] Free-standing flexible GO membranes were prepared from GO dispersion in water by vacuum-filtration. GO was found to have higher tensile strength and water uptake compared with Nafion. The power density of a fuel cell with a 30 μm thick GO membrane was 35 mW/cm2 at 30°C, despite the fact that the proton conductivity is several orders of magnitude lower than Nafion. The device had a very high open circuit voltage >1 V indicating low crossover. This indicates that thinner membranes compared with Nafion can be utilized, reducing the overall cell resistance. Indeed, by changing the fabrication method to further reduce the thickness of the GO membranes to several microns, we recently achieved improved power density of up to 80 mW/cm2.

We also have performed a detailed study on the through-plane conductivity and permittivity of GO membranes over a wide temperature and humidity range.[7] It was found that the proton conductivity is strongly dependant on humidity, and that under dry warm conditions significant electronic conductivity is observed. This opens the opportunity for tuneable mixed ionic-electronic conductivity and may be extremely useful in sensing application. In addition, we performed in-situ scanning electron microscopy (SEM) and electron energy loss spectroscopy (EELS) on humidified and dried GO membranes, directly observing expansion and contraction ("breathing") of the membranes.[8] Even after drying and under high vacuum conditions, the EELS signal for crystalline ice was observed after freezing, suggesting that water is extremely strongly bound to GO, which may have implications for proton conductivity.

Finally, a novel class of alkaline anion exchange membrane (AAEM) is presented, in the form of KOH-modified multilayer graphene oxide paper (GOKOH).[9] SEM investigations showed that the morphology of GO changes after KOH-treatment, whilst X-ray photoelectron spectroscopy (XPS) measurements and X-ray diffraction (XRD) analysis confirmed successful chemical modification. The hydrogen gas permeability was several orders of magnitude lower than conventional polymer-based ionomer membranes. The maximum anion conductivity was 6.1 mS/cm at 70 °C, and the dominant charge carrier was confirmed to be OH− by utilization of anion and proton-conducting blocking layers. The ion exchange capacity was 6.1 mmol/g, measured by titration. A water-mediated reverse Grotthuss-like mechanism is proposed as the main diffusion mode of OH− ions. A prototype AAEM fuel cell was fabricated using a GOKOH membrane, confirming the applicability to real systems.

References:

1. J. Liu, T. Daio, K. Sasaki, S. M. Lyth, Journal of the Electrochemical Society, 161, F834 (2014).

2. J. Liu, T. Daio, D. Orejon, K. Sasaki, S. M. Lyth, Journal of The Electrochemical Society 161, F544-F550 (2014)

3. D. R. Dreyer, S. Park, C. W. Bielawski, and R. S. Ruoff, Chem. Soc. Rev., 39, 228–240 (2010).

4. R. R. Nair, H. A. Wu, P. N. Jayaram, I. V. Grigorieva, and A. K. Geim, Science, 335, 442–444 (2012).

5. K. Hatakeyama et al., Angew. Chemie, 53, 6997–7000 (2014).

6. T. Bayer, S. R. Bishop, M. Nishihara, K. Sasaki, and S. M. Lyth, Journal of Power Sources, 272, 239–247 (2014).

7. T. Bayer, S. R. Bishop, N. Perry, K. Sasaki, S. M. Lyth, ACS Applied Materials and Interfaces, In press (2016)

8. T. Daio, T. Bayer, T. Ikuta, T. Nishiyama, K. Takahashi, Y. Takata, K. Sasaki, S. M. Lyth, Scientific Reports, 5, 11807 (2015)

9. T. Bayer, B. V. Cunning, R. Selyanchyn, T. Daio, M. Nishihara, S. Fujikawa, K. Sasaki, S. M. Lyth, Journal of Membrane Science, 508, 51 (2016)

2551

, , , , , and

Historically, the general strategy for developing new ion-exchange membranes focused almost exclusively on the synthesis of new polymers with a high IEC, where the ion-exchange groups are strongly dissociable. Unfortunately, the use of polymer chemistry alone has not been successful in achieving the performance goals of next-generation proton conducting membranes for fuel cells and redox flow batteries. For example, simply increasing the ion-exchange capacity of a polymer to improve conductivity has lead to polymers that are brittle when dry and swell excessively in water, with a loss in mechanical strength and counterion/coion selectivity. To overcome the problems with highly charged single polymer membrane systems, researchers have: (i) investigated block copolymers which self- assemble into separate phases, where one block provides selective transport pathways and the second block imparts mechanical strength, (ii) combined the desirable properties of two different polymers in a single membrane through blending, (iii) impregnated a functional polymer into a microporous inert support, where the support provides the requisite mechanical properties that the functional polymer lacks, and (iv) crosslinked the ionomer.

The present paper will focus on blended ion exchange membranes, composed of perfluorosulfonic acid (PFSA) and polyvinylidene fluoride (PVDF), where nanofiber electrospinning is used to control/alter the nano/micro distribution of PFSA and PVDF. Three membrane fabrication strategies will be presented, using 1100 and 825 EW PFSA and Kynar PVDF: (1) dual fiber electrospinning, where the final membrane morphology is an ionomer matrix with an embedded network of uncharged PVDF reinforcing nanofibers, (2) electrospinning of submicron fibers composed of a PFSA/PVDF polymer blend, followed by fiber mat processing to create a dense and defect-free film, where the morphology is that of randomly oriented bundles of PFSA and PVDF nanofibrils, and (3) core-shell single fiber electrospinning, where the core is PVDF and the shell is a mixture of PFSA and PVDF, followed by compaction and densification to create a dense membrane, where the final membrane structure is similar to that of (1). The methods for fiber electrospinning and follow-on mat processing will be described. Physical property data will be presented for the different fabrication schemes and for different PFSA/PVDF membrane compositions. The use of selected membranes in hydrogen/air and DMFC fuel cells and H2/Br2 redox flow cells will be discussed.

2552

and

In PEMFC, nanocomposites are drawing attention as the alternative membrane because of selected various acid moiety functionalized additives, increased water retention and improved mechanical and thermal properties. But in modification of nanocomposites, when more than 100 nm-sized particles are introduced in ion channel, ion channel development is affected and hydrogen ion movement is reduced.

So this study explores the concept of improving fuel cell membrane performance without compromising proton conductivity by using a 1 ~ 3 nm sized polyhedral oligomeric silsesquioxane (POSS). Polyhedral oligomeric silsesquioxane (POSS) is a molecule that has a spatially defined, rigid and hydrophobic cube-octameric siloxane skeleton (about 1 ~ 3 nm in size) with eight organic vertex groups, one or more of which are reactive or polymerizable. [1 ~ 3]. These particular structural features render POSS be a versatile additive for acquiring enhanced thermomechanical properties, better thermal stability, atom oxygen resistance, abrasion resistance and low water uptake [4 ~ 6].

Only one POSS-based nanocomposite for PEMFC h as been reported [7]. An open-cage POSS carrying three glycidyl epoxy groups was reacted with 4-hydroxybenzenesulfonic acid, the resulting sulfonated POSS was blended with polyvinylalcohol and the blend was crosslinked using ethylenediaminetetracetic dianhydride (EDTAD). This system differs considerably from the system described in this study (open-cage versus closed-cage POSS, crosslinked versus noncross-linked structure and aliphatic versus aromatic composition) and also has several disadvantages as follows: it requires a multistep fabrication process and it contains chemically unstable methylene groups. Additionally, the mechanical and chemical stability was not reported.

In this study, POSS nanoadditives carrying phosphonic acid moieties were synthesized, characterized and formulated into Nafion®. As mention in the ref. [8 ~ 12], phosphonic acid groups are good candidate for high temperature proton carriers with the relatively high proton conductivity without water due to its self-dissociation natures. Also we developed Nafion®/3-aminopropyl triethoxysilane (APTES) hybrid membranes with phosphonic acid groups in the previous work in order to improved high temperature performances [11]. These hybrid membranes show high proton conductivities (0.028 S/cm) at 120 oC and 40% RH, so we can confirm the strong influence of phosphonic acid groups in low humid conditions.

Also, in contrast to common additives, it can be possible to progressive improvement of proton conductivity because of eight vertex groups which substitute by phosphonic acid groups. We expect to innovatively improve proton conductivities by molecular-closed particle size and many vertex groups of POSS.

The phosphonic acid grafted POSS prepared were characterized for particle size and chemical structure. Also Nafion®/POSS-PA nanocomposites were characterized for morphologies, mechanical strength, proton conductivities at different temperatures and humid conditions, and the cell performances at different temperatures.

2553

, , , , , and

In the "Direct Membrane Deposition" (DMD) approach for polymer electrolyte fuel cells the conventional catalyst coated membrane (CCM) is replaced by two gas diffusion electrodes (GDE) coated with ionomer via inkjet printing. Assembling the ionomer-coated GDEs creates a fuel cell with a very thin membrane (12 µm) and improved ionic contact of membrane and electrodes. Fuel cells fabricated with DMD therefore showed peak power densities of 4 W/cm² at 70°C, 300 kPa and with oxygen as fuel exceeding the peak power density of the Nafion HP reference fuel cell by a factor of 2 [1]. Despite the thin membrane DMD fuel cells showed no increased hydrogen crossover (< 2 mA/cm²). Furthermore, DMD fuel cells reached a power density of more than 1 W/cm² even under very dry conditions (zero gas humidification) and with air at the cathode.

In a second work we demonstrated a record Pt-utilization efficiency of 88 kW/gPt of a DMD fuel cell with low Pt-loaded electrodes (anode/cathode 0.029 mgPt/cm2) at 80°C, 300 kPa and with oxygen as fuel [2]. The DMD approach also proved its suitability for medium temperature fuel cells: by incorporating TiO2nanoparticles into the directly deposited membrane the fuel cell showed stable operation at 120°C with a power density of 2 W/cm² (300 kPa and oxygen at the cathode) [3]. Extensive electrochemical characterization showed that fuel cells fabricated with DMD have an ionic resistance and a mass transport resistance half that of reference fuel cells with CCMs at high current densities. Impedance spectroscopy revealed that the reduction of mass transport losses is responsible for the major part of the improvement in power density. Besides the increased power density, DMD bears the potential to simplify the fabrication process of fuel cells by successively spray-coating all layers including the membrane onto a gas-diffusion-layer [4].

This talk provides an overview of the DMD activities, its future potential and gives detailed insight into the underlying reasons for the increased power density of DMD fuel cells.

Fig. 1Conventional catalyst coated membranes (CCM) with gas diffusion layers are replaced by gas diffusion electrodes with direct membrane deposition (DMD) assembled with a subgasket. Taken from [1] - Published by The Royal Society of Chemistry.

References

1. Klingele, M., Breitwieser, M., Zengerle, R., Thiele, S.: Direct deposition of proton exchange membranes enabling high performance hydrogen fuel cells. J. Mater. Chem. A 3(21), 11239–11245 (2015). doi: 10.1039/c5ta01341k

2. Breitwieser, M., Klingele, M., Britton, B., Holdcroft, S., Zengerle, R., Thiele, S.: Improved Pt-utilization efficiency of low Pt-loading PEM fuel cell electrodes using direct membrane deposition. Electrochem. Commun. 60, 168–171 (2015). doi: 10.1016/j.elecom.2015.09.006

3. Wehkamp, N., Breitwieser, M., Büchler, A., Klingele, M., Zengerle, R., Thiele, S.: Directly deposited Nafion/TiO 2 composite membranes for high power medium temperature fuel cells. RSC Adv 6(29), 24261–24266 (2016). doi: 10.1039/c5ra27462a

4. M. Breitwieser, M. Klingele, B. Britton, S. Holdcroft, R. Zengerle and S. Thiele: High power fuel cells with direct membrane deposition via ionomer spray-coating. (Poster), Bad Zwischenahn (Germany) (2015)

Figure 1

2554

In this study the synthesis of low-T and high-T proton-conducting blend membranes and of anion-exchange blend membranes utilizing the acid-base blend concept by usage of the following blend components is presented: (i) a cation-exchange polymer, (ii) an anion-exchange polymer, and (iii) a stable basic polymer such as a polybenzimidazole (PBI, F6PBI or PBIOO) as the stabilising matrix.

For the preparation of low-T proton-conductive blend membranes, DMAc solutions of a sulfonated cation-exchange polymer in molar excess were mixed with a DMAc-dissolved halomethylated polymer, a tertiary amine or alkylated imidazole, and a DMAc-dissolved polybenzimidazole. During membrane formation by solvent evaporation quaternization of the halomethylated polymer took place. A blend membrane composed of a partially fluorinated aromatic sulfonated polymer, F6PBI, and a partially fluorinated aromatic anion-exchange polymer with pendent 1-ethyl-2-methylimidazolium anion-exchange groups was investigated in a SO2-depolarized electrolyser and yielded an excellent performance (operation temperature 95°C, 0.5 Acm-2@0.85V) which was markedly better than that of a Nafion®115 membrane (operation temperature 95°C, 0.4 Acm-2@0.92V), even after voltage stepping [1].

High-T proton-conducting blend membranes were prepared by mixing DMAc solutions of an excess PBI amount with a halomethylated polymer, a sulfonated polymer and a tertiary N base. During membrane formation, following reactions took place: quaternization of the halomethylated polymer with the tertiary N base to formation of an anion-exchange polymer, ionic cross-linking of the anion-exchange polymer´s cationic groups with the anionic groups of the sulfonated polymer, and a minor extent of covalent cross-linking by reaction of residual halomethyl groups of the halomethylated polymer with N-H groups of the PBI blend component, yielding base-excess blend membranes which were subsequently doped with phosphoric acid to achieve H+-conductivity. As an example, DMAc solutions of F6PBI (70 wt%), a partially fluorinated aromatic anion-exchange polymer with pendent 1-ethyl-2-methylimidazolium anion-exchange groups (15 wt%) and a sulfonated PPSU polymer (15 wt%) were mixed together. After solvent evaporation the blend membrane was doped with PA (H3PO4, PA doping degree 167%) and operated in a high-T fuel cell at T=140°C. The blend membrane showed excellent performance (0.5 A/cm-2@0.5 V) and good longevity (fuel cell experiment stopped after 400 hrs without membrane failure) [2].

Anion-exchange blend membranes were prepared mixing DMAc solutions of a halomethylated polymer (40-60 wt%), a PBI (30-50 wt%), a tertiary N base, and a sulfonated polymer (5-15 wt%). During solvent evaporation, formation of the anion-exchange polymer and of ionical and covalent cross-links took place. As an example, DMAc solutions of polyvinylbenzylchloride, PBIOO and sulfonated PPSU were mixed with tetramethylimidazole. After solvent evaporation, an anion-exchange blend membrane was obtained which was composed of 66.35% anion-exchange polymer, 26.15% PBIOO and 7.5% sulfonated PPSU. After KOH treatment (90°C, 1M, 10 days) the membrane exhibited a Cl- conductivity of 65 mScm-1@90°C@90% r.h. (IEC=2.9 meq OH-/g) which is an excellent value for anion-exchange membranes.

 

References

 

[1] R. Peach, H. M. Krieg, A. J. Krϋger, J. JC Rossouw, D. Bessarabov, J. Kerres, Int. J. Hydrogen Energy 2016, accepted

[2] J. Kerres, V. Atanasov, Int. J. Hydrogen Energy 2015, 40, 14723-14735

2555

, , and

Heteropoly acids (HPA)s are a large class of inorganic acids that are known to have extremely high proton conductivities.[1] It has been demonstrated that doping perfluorinated sulfonic acid and hydrocarbon-based membranes with HPA is able to improve protonic conductivity, particularly under reduced humidity and elevated temperatures, but this method suffers from a lack of stability to water.[2,3] Even when the cesium salt (water-insoluble) is used, it is likely that agglomeration will occur overtime, as the HPA is not truly immobilized. Silicotungstic acid (SiW12O40-4) is of particular interest for use in a fuel cell because it has been shown to be stable under relevant conditions.[4]

In an effort to mitigate agglomeration and water instabilities that plague most HPA containing films, we have covalently attached silicotungstic to 3M FC-2145 (polyvinylidene-co-hexafluoropropylene), forming a highly crosslinked polymer/hybrid network. First, phenol phosphonic acid (PPA) side chains are attached to FC-2145, and then two PPA groups react with a single, mono-lucunary silicotungstic acid (SiW11O39-8), thus forming covalent crosslinks. This synthesis produces materials that have loadings of HPA, stable to boiling, in excess of 60wt%. The resulting material has high proton conductivity (>200 mS/cm at 90°C/95%RH) under humidified conditions as well as relatively high proton conductivity (>30mS/cm at 110°C/50%RH) under lower humidity conditions. This talk will include discussion on synthesis, transport properties, and preliminary chemical stability.

References:

[1] Nakamura, O.; Kodama, T.; Ogino, I.; Miyake, Y., High-Conductivity Solid Proton Conductors - Dodecamolybdophosphoric Acid and Dodecatungstophosphoric Acid Crystals. Chemistry Letters 1979, 17-18.

[2] Meng, F.; Aieta, N. V.; Dec, S. F.; Horan, J. L.; Williamson, D.; Frey, M. H.; Pham, P.; Turner, J. A.; Yandrasits, M. A.; Hamrock, S. J.; Herring, A. M., Structural and transport effects of doping perfluorosulfonic acid polymers with the heteropoly acids, H3PW12O40 or H4SiW12O40. Electrochimica Acta 2007, 53, 1372-1378.

[3] Wang, Z.; Ni, H.; Zhao, C.; Li, X.; Fu, T.; Na, H., Investigation of sulfonated poly(ether ether ketone sulfone)/heteropolyacid composite membranes for high temperature fuel cell applications. Journal of Polymer Science Part B: Polymer Physics 2006, 44, 1967-1978.

[4] Sweikart, M. A.; Herring, A. M.; Turner, J. A.; Williamson, D. L.; McCloskey, B. D.; Boonrueng, S. R.; Sanchez, M., 12-Tungstophosphoric Acid Composites with Sulfonated or Unsulfonated Epoxies for High-Temperature PEMFCs. Journal of The Electrochemical Society 2005, 152, A98.

2556

Proton exchange membranes (PEMs) are central to determine the fuel cell (FC) performance and durability. The PEMs employed in existing applications are perfluorinated sulfonic acid polymers (PFSAs), such as Nafion by DuPont, Flemion by Asahi Glass, Aquivion by Solvay. For example, it is known that Gore select series membranes (with reinforcement) are adopted to Toyota Mirai on sale since December 2014. However, the PFSAs are not the perfect PEMs when it comes to mass production phase of FC vehicles (FCVs), in light of material cost and its environmental friendliness. FCVs need to be designed as cheaply as possible with a maintained high performance, including durability and safety, of the FC stacks and systems. Many types of PEMs including PFSAs and non-PFSAs have been investigated in academy and industry, however often leading to a complexity of manufacturing themselves (synthesis, processing, raw material price, etc.) and their applicability to existing FC systems. PEMs should be, no matter what type of materials are chosen, as simple as possible towards industrial applicability. For example, multiblock copolymers were heavily discussed for the last decades as an alternative to PFSAs in terms of proton conductivity as a funciton of relative humidity (RH). It is true that the property of such block copolymers would get close to that of PFSAs, however casting doubt on reproducibility/reliability for production phase. Therefore, in this study, a simple method was selected, i.e., blending of simple structural polymers for both acidic and basic functions, with an aim at comparable PEM properties to the state-of-the-art PFSAs represented by Nafion.

Acid-base blend membranes based on non-fluorinated hydrocarbon typed chemistry were prepared, by involving syntheses of novel base monomers and polymers. Two kinds of monomers were successfully synthesized via lithiation chemistry using electrophiles including pyridyl groups. The versatility for the monomers towards polycondensations for preparation of the high molecular weights poly(arylene ether sulfone)s was confirmed, after optimizing the reaction conditions against transetherification. Molecular weights (Mn) of the new basic polymers found to be from 20 to 50 kDa, in relation to ca. 35 kDa of one UDEL polymer commercialized by Solvay, as confirmed by gell permeation chromatography. Each of the basic polymers prepared was blended with a disulfonated alternating poly(arylene sulfone) (acid/base = 80/20 wt./wt.). Neutralization treatment prior to the processing was no need in order to form their uniform blend membranes, unlike the case of blending acidic polymers with basic polybenzimidazole. All the acid-base blend membranes showed high proton conductivities and good dimensional stabilities in addition to reasonable mechanical properties. For instance, the proton conductivity at 20%RH of the blend membranes showed superior to that of Nafion, whose IECs ranged till ca. 2.0 meq./g. The water uptake of the membranes under immersed conditions remained below 30 of the lambda (water molecule per sulfonic acid) at water temperatures ranging from room temperature to 90 degrees C. In conclusion, the acid-base blend membranes containing the new basic polymers would be a promising candidate to future PEMs with impact on cost reduction of the FC stack.

B-31 Optimizing Stack Performance - Oct 5 2016 8:10AM

2557

and

By adding moisture preserve materials in the anode catalyst layer, we prepared a series of low humidity membrane electrode assembly (MEAs), and all the MEAs exhibited excellent performances at low humidity for both anode fuel and cathode air. The effects of the structure, used amount of the materials on the low humidity performances of the MEAs were investigated, and we also explored the mechanism for the excellent low-humidity performances of the MEAs with moisture preserve material added at anode catalyst layer.

2558

, , , , and

Overview

 Performances of cell stacks are simulated, focusing on driving state of anode system. This simulation technique takes overall transport phenomena in PEFC fuel cell systems, such as transport of gas species and liquid water in both anode and cathode channels and across MEA, thermal balance, electrochemical reactions and current distributions. Injector and exhaust systems are modeled by boundary conditions, mimicking actual driving conditions. Relation between cell performances and driving state of anode system, for example, consumption of hydrogen, build-up of liquid water and cross-over nitrogen and purging by control of injector and/or ejector can be discussed based on simulated results. Over-all performance of anode system is one of crucial issue for durability and cost-reduction of fuel cell systems. These results show good applicability of this kind of simulations on design of PEFC stack system.

Numerical Modeling and Results

 Simulation is performed by our own simulation software for PEFCs, which is capable of simulating full-stack fuel cell system for fuel cell vehicles (up to 400 cells) under transient operating conditions. An important feature of this software is that macroscopic models are applied for microscopic phenomena in the MEA such as electrochemical reactions and proton/water transport. Heat transfer and fluid of liquid-gas two phase fluid with phase changing are also taken into account. These models are coupled with multi-dimensional heat and mass transport equations, including effective parameters for gas diffusivity and permeability in the GDL and equivalent hydraulic diameter in the flow channel. This software can also simulate cathodic degradation by carbon corrosion reactions.

 Performances of up to 400 cells stacks are simulated. A numerical model of anode system is constructed by adding transient boundary conditions for an injector and an exhaust system with an ejector. These transient boundary conditions represent open / close timings of injector and ejector under actual driving conditions. Distributions and consumptions of hydrogen, build-up of cross-leak nitrogen and condensate liquid water can be shown under constant (up to 30A) loading.

 Simulated results show difference of cell potential between of stacked cells, and relation between performance distributions and balances of transport phenomena in stack systems. Effects of configuration of stack systems are also shown. Based on these results, design of stack systems and control schemes of anode system can be discussed. Those discussion can provide useful information for improving durability of stacked cells by avoiding depletion of hydrogen, and / or examination of control schemes which can efficiently utilize hydrogen fuel.

2559

, and

It is possible to design and operate a fuel cell with non-humidified reactant gases, by relying on water produced by electrochemical reaction inside the fuel cell. However, in order to do so there must be adequate water and heat management applied. For example, temperature distribution along the cathode channel must be such that relative humidity is maintained close to 100%. As water is produced on the cathode side there is a water concentration gradient across the mebrane, and in order to humidify hydrogen on the other side there must be back-diffusion flux greater than or at least equal to the electroosmotic drag. Membrane thickness, thus, plays an important role in realization of this concept.

A segmented fuel cell is designed such that each segment can be maintained at different pre-determined temperature. The temperatures of the segments along the cathode channel are selected by the help of Molliere h-x diagram such that relative humidity is maintained close to 100%. Operating conditions are selected such that the product water is sufficient to saturate the cathode exhaust without appearance of liquid water. Humidity measurements are taken between each segment on both anode and cathode sides. The anode and cathode exhaust streams are condensed so that water balance may be established.

The entire setup is modeled with FLUENT CFD package, and the experiments are conducted with two different mebrane thickness. Good agreement is achieved between the modeling and experimental results.

The results indicate that it is possible to establish the required temperature profile along the cathode channel and keep the cathode stream relatively well humidified without external humidification. Performance of such variable temperature fuel cell is better than the uniform temperature fuel cell. The thinner membrane results in better humidification of the hydrogen side, which is important especially at higher current densities, i.e., at higher electro-osmotic drags.

Figure 1

2560

Toyota Motor Corporation launched the Prius, the world's first mass-production hybrid vehicle (HV), in 1997. The global popularization of the Prius and other HVs has helped to counteract the negative effects of vehicles on the environment. As emerging markets continue to develop economically and the world's population grows, vehicles are likely to have an even greater impact on the environment. Therefore, if vehicles are to remain of benefit and interest to society in the next one-hundred years, it will be even more important to find answers to issues related to energy and the ideal form of mobility in the future. Through the development of hybrid, plug-in hybrid, and battery electric vehicles, Toyota is already acting proactively to encourage the use of electricity. In addition, Toyota has also taken the step of developing fuel cell vehicles (FCVs) to help promote the use of hydrogen for contribution to environment.

A fuel cell (FC) generates power by electrochemical reactions through an electrocatalyst using hydrogen and oxygen in the air. Since these reactions generate only water as a by-product, FCs are regarded as an extremely clean and efficient means of generating power, with the potential to help resolve energy and environmental issues in the future. Greatly enhancing the performance and reducing the size and cost of the FC stack is an important part of measures aimed to facilitate the widespread adoption of FCVs.

A FC stack consists of a wide range of materials, including various types of mechanical elements and electronic parts. In particular, cells inside the FC use many materials that are not found in conventional vehicle parts. In contrast to the research phase, which prioritized efforts to enhance the functional performance of the FC stack, the new mass-production MIRAI FCV was developed as a normal industrial product with an emphasis on both performance and the applicability of manufacturing methods. This article describes several examples of manufacturing methods adopted to ensure high quality and productivity while reducing cost.

 

2561

, and

In this study a validated model for a PEM fuel cell system is developed to actively control the supplied reactants and to manage water within the cell structure and correctly provide system cooling. Due to the substantial difference in response times of each of the system volumes, a model of each components system is developed for analysis and controller design to achieve thermal tracking while adequately rejecting system disturbances.

Control-oriented transient models have been developed to account for the formation of liquid water within the gas channels or within both the channels and the GDL, however the advantages of having a validated model will help achieving adequate humidity regulation. Due to the notoriously slow response of humidity transducers, especially near saturated conditions, this is an advances the field of fuel cell reactant pre-treatment. In developing this strategy, a critical step will be accomplished by properly selecting the controller references used for temperature feedback. Although this may at first seem like a simple step, the selected temperature references have a profound impact on the resulting thermal and humidity regulation. If not properly considered, the system response could be unnecessarily slow or produce an undesirable excursion in humidity. The establishment of this relationship allows for accurate dynamic voltage estimation under a range of operating conditions, and is a necessary step towards further understanding physical phenomena.

2562

, and

One way of improving the durability and reliability of a fuel cell stack or system is to isolate faults within the fuel cell and perform targeted condition-based maintenance on the affected components, to remove or limit the conditions that caused the fault. This can be done by on-line monitoring of the fuel cell system's state of health to locate and isolate faults and prevent sudden failures before they occur and correcting faulty conditions to prolong the lifetime of the fuel cell. This paper presents electrochemical impedance spectroscopy (EIS) results obtained on PEMFC stacks, with the aim of investigating the fuel cell stacks' response to different faults, for a possible use of EIS as an advanced monitoring and diagnostics tool. The stacks used in this study are from Dantherm Power A/S, intended for stationary microCHP application. Four different faults were considered; change in fuel composition, air starvation, fuel starvation and water management, both for flooding and drying. Each fault simulation was performed in a separate fuel cell stack to avoid the mixing of the impedance signatures of the different faults. The same continuous, steady state operations under nominal condition were recorded prior to fault simulations as healthy-state references. Then the different faults were induced at varying current densities and varying fault levels.

2563

, , , , , , , and

Proton Exchange Membrane Fuel Cell (PEMFC) technologies are considered as a promising and clean energy supply for both transportation and stationary applications. Whereas the overall electrical performances and power densities have now reached most of the targeted specifications for integration, the overall durability and cost are still major hurdles to the mass market. The mechanisms of performances loss under PEMFC operation on individual components (catalyst, membrane, gas diffusion layer) are well-known in the literature at the material or small electrode scales. However in real stack configuration, the degradations may be strongly accelerated and localized since all phenomena are highly inter-correlated and very dependent of both the stack/cell designs and the operating conditions.

The decrease in performances (under continuous or cycling modes) is composed of both irreversible losses (essentially due to materials degradations) and reversible losses which are attributed to the evolution of cell water management and electrode pollution (CO-like contamination, Pt-oxide formation) over time. Long stop phases make it possible to rejuvenate these reversible degradations but may not always be convenient or possible according to the applications or user needs. Developing reliable operation strategies either to recover or avoid reversible performance loss are necessary to adjust system dimensioning and prevent from using oversized stacks.

When considering particularly the reversibility losses hypothetically caused by PtOx formation at high cathode potential, a modification during a short period of local conditions enabling to remove the oxides can be proposed to stabilize the performances [1]. Air-starvation have already been studied at single cell level to regenerate the reversible losses in Direct Methanol Fuel Cell (DMFC) [2] but the impact on the long-term durability remains uncertain. Our work is thus on PEMFC stack operation under both automotive Fuel Cell Dynamic Load Cycles (FC-DLC) and stationary conditions at fixed load to evaluate and discuss the interest and reliability of transient and repeated air-starvation periods.

Start-up and shut-down protocols must also be taken into account since they can cause harsh degradation phenomena in the fuel cell caused by the so-called reverse current mechanism arising from the H2|O2 (or air) front at the anode side. Since this configuration is hardly avoidable after long stops in real systems, mitigation strategies must also be developed to enhance the lifetime of the PEMFC stack despite the embedded system integration constraints. In this perspective, we determined in-situ current density distribution mapping using an S++® equipment inside our stack in order to localize the degradation under repetitive fuel starvation conditions. Based on the degradation mapping, we propose to discuss the possibility to tune and adapt the 2D/3D structure and composition of the Membrane Electrode Assembly to modify internal current distribution and mitigate corrosion phenomena within the cells.

References

[1] Y. Huang et al. , J. Electrochem. Soc., 2013, vol. 161, no1, p. F10

[2] C. Eickes et al. , J. Electrochem. Soc., 2006, vol. 153, no 1, p. A171

2564

, , and

To further improve the cell performance of a proton exchange membrane fuel cell (PEMFC), thus meeting the high requirement for its automotive and stationary application, great efforts have been made on the optimization of the flow field pattern[1,2]. In this work, two novel flow field patterns are proposed and the influences on the cell performance are investigated. Both two novel flow field designs lead to superior cell performance than the conventional flow field patterns do, and this can be attributed to their superiority in removing water at high current densities while increasing the flow velocity near the outlet. The velocity magnitude in flow fields is examined and compared through the computational fluid dynamics modeling.

Acknowledgements

This work was supported in part by National Natural Science Foundation of China (Grant No. 21373135 and 21533005) and Science Foundation of Ministry of Education of China ( Grant No. 413064).

References

[1] Soler, J., E. Hontanon, and L. Daza, Electrode permeability and flow-field configuration: influence on the performance of a PEMFC. Journal of Power Sources, 2003. 118(1): p. 172-178.

[2] Lee, B., K. Park, and H.-M. Kim, Numerical Optimization of Flow Field Pattern by Mass Transfer and Electrochemical Reaction Characteristics in Proton Exchange Membrane Fuel Cells. Int. J. Electrochem. Sci, 2013. 8: p. 219-234.

B-32 Gas Diffusion Media and Bipolar Plates - Oct 5 2016 2:00PM

2565

, , and

Laboratory-made membrane electrode assemblies (MEAs) for proton exchange membrane fuel cells typically perform poorly compared to commercial ones. A key issue is the difficulty with making a high performance catalyst coated membranes (CCMs) at the laboratory scale by unautomated processes that often leads to irreproducible results due to lack of control of important parameters in the catalyst layer (CL). We have developed a method to fabricate reproducible CCMs with high activity using direct ultrasonic spray deposition of electrocatalysts onto a heated commercial perfluorosulfonic membrane. The CCMs are optimized by varying the number of catalyst layers sprayed, the height of spay nozzle from the membrane, and the pressure used for hot pressing on CLs after spraying. Nitrogen adsorption and scanning electron microscopy (SEM) are used to investigate the effects of ionomer/carbon ratio (I/C) on the surface area, pore structure and morphology of the CLs. How the CCMs are assembled into MEAs is another key issue in the fuel cell performance, particularly the compression, which governs the porosity of the CL and gas diffusion media and resistance. The resulting MEAs are analyzed by I-V polarization, cyclic voltammetry and electrochemical impedance spectroscopy. A comparative study is also made with CCMs prepared with low equivalent weight Nafion ionomer in the CLs. Operating the cell at 80°C and 50 % RH, a cell voltage of 0.6 V a current density of 971 ± 41mA cm-2 is achieved in Air/H2 at a stoichiometric ratio of 2/2 at ambient pressure and increased to 1432 ± 131 mA cm-2 at a back pressure of 150 KPa. These tools can be used to guide the laboratory development and understanding of new catalyst layer structures and catalyst compositions.

2566

, and

A deep understanding of the behavior of microstructural parameters in proton exchange fuel cells (PEFCs) will help to reduce the material cost and to predict the performance of the device at cell scale. Changes in morphological configuration, i.e., fiber diameter and fiber orientation, of the gas diffusion layers (GDLs) result in variations of fluid behavior throughout the layer, and therefore; the microstructural parameters are affected. The aim of this study is to analyze, for three selected fiber diameters and different percentages presence of inclined fibers, the behavior of the different microstructural parameters of the GDLs.

This study is carried out over digitally created two-dimensional GDL models in which the fluid behavior is obtained by means of the Lattice Boltzmann method (LBM). Once the fluid behavior is determined, the microstructural parameters, i.e., the porosity, gas-phase tortuosity, obstruction factor, through-plane permeability, and inertial coefficient, are computed. Several relationships are found to predict the behavior of such parameters as function of the fiber diameter, presence of inclined rods or porosity. The results presented in this work are compared and validated by previous theoretical and experimental studies found in the literature.

2567

, , , and

The proton exchange membrane fuel cell (PEMFC) offers the potential for an efficient, and reliable power source for vehicles. The bipolar plates are a key component in the stack providing mechanical support, supplying reactant gases to both anode and cathode, removing the reaction products, and providing electrical and thermal conductivity. [1] Bipolar plate materials are broadly divided into metallic and carbon-based materials. Carbon-based plate materials do well mechanically; still fabrication is difficult and costly. As a result, metallic bipolar plates are becoming increasingly popular.[2] Metal bipolar plates have all the necessary properties (including high thermal and electrical conductivity, low gas permeability, high manufacturability, and relatively low cost materials); however, chemical instability due to the highly corrosive environment of the PEMFC leads to the formation of a passive surface oxide layer.

Altering the surface by coating with a corrosion resistant material can be cost prohibitive when considering the amount of coating necessary to overcome the harsh conditions in a PEMFC, fabrication of the coating, and the number of plates in a stack. Less costly metals for coating, such as titanium, typically form an oxidative layer while non-corrosive metals can form pits where the underlying metal becomes exposed. Our multilayer technique combines both metal types for a corrosion resistant economical solution. [3] This design prevents oxide formation from percolating throughout the layered structure via thin films of non-corrosive layers alternated with thicker Ti layers. The specialized structure of multilayers of thin films (Au/Ti) allows the noncorrosive layer, the most costly, to be <1 nm while maintaining the viability and conductivity of the coating and the metal bipolar plate.

[1] H. Tsuchiya and O. Kobayashi, "Mass production cost of PEM fuel cell by learning curve," International Journal of Hydrogen Energy, 29 (2004) 985–990

[2] K. RoBerg and V. Trapp, "Graphite-based bipolar plates," in Handbook of Fuel Cells Fundamentals, Technology and Applications, W. Vielstich, A. Lamm, and H. A. Gasteiger, Eds., JohnWiley & Sons (2003) 308–314

[3] S Wang, J. Peng, W. Lui, and J. Zhang, Performance of the Gold-plated Titanium Bipolar Plates for the Light Weight PEM fuel cells," Journal of Power Sources 162 (2006) 486–491  SHAPE  \* MERGEFORMAT

Figure 1. Multilayer design model.

Figure 1

2568

, , , and

Operating a proton exchange membrane fuel cell (PEMFC) in a planar, open cathode configuration can enable significant reductions in overall device weight and system complexity compared to a conventional fuel cell stack. This decrease in device weight makes an array of open-cathode, planar fuel cells a compelling candidate power source for unmanned aerial vehicles. The flow of ambient air over vehicle-mounted cells can improve oxygen, waste heat, and water transport between the catalyst layer and bulk air. We previously studied the effect of environmental conditions on cell polarization and found that ambient temperature, relative humidity, altitude and air speed all influence cell temperature, which affects water balance and oxygen transport to the catalyst layer. In this work, we relate gas diffusion media (GDM) properties to cell polarization behavior in a broad range of anticipated flight conditions. We study the effect of gas diffusion layer thickness and porosity on cell polarization and discuss its influence on cell temperature, oxygen transport and water management. Similarly, we investigate the polytetrafluoroethylene (PTFE) content in the GDM and the influence of hydrophobicity on cell water management and performance in a broad range of anticipated flight conditions. We use cyclic voltammetry and AC impedance spectroscopy to characterize performance losses associated with GDM properties to guide GDM selection.

2569

, and

As fuel cells gain increased volumetric power density, there is a need for thinner GDLs that give increased thermal and electrical conductivity, and mechanical stiffness. This has been demonstrated by the introduction of advanced GDL/bipolar plate designs that have been implemented in commercial fuel cells such as that seen in the Toyota Mirai (1).

The use of metal GDLs has been previously examined in ethanol and methanol fuel cells (2,3) but have been largely ignored in PEMFCs due to the corrosive conditions present in PEMFCs which considerably reduce the lifetime of cells with non-passive metals present. Spun carbon fibre GDLs, such as Toray and Avcarb, have previously been the main choice for PEMFC GDLs due to the passivity and gas and water transport properties of these materials. With increased power density in fuel cells however these materials can incur losses due to a lack of thermal and electrical conductivity as well as being relatively thick and brittle.

In this study we have examined a number of metal meshes used in place of carbon paper GDLs in fuel cells. Meshes with different aperture sizes were examined with regard to fuel cell performance as well as flooding of the electrodes and performance under high mass transport conditions. The aperture sizes of these meshes played a significant role in performance in the cell, both in terms of electrical conductivity and water management.

Surface passivation of the meshes by nitriding and using carbon and other passive coatings was also examined in order to increase their durability in the cell. The surface energy properties of these coatings was also examined ex situ via droplet shape analysis in order to achieve the best properties for use in a fuel cell. Adjustment of the hydrophilicity and hydrophobicity of these coatings was also used to improve water management.

The performance of some of these meshes and foams was seen to exceed some of the commercially available carbon papers even without significant optimisation. These materials could soon replace carbon fibre GDLs in PEMFCs leading to much higher volumetric power densities and improved performance.

References

  • T. Yoshida, K. Kojima, Electrochemical Society Interface, 24 (2015), pp. 45–49.

  • S. Arisetty, A. K. Prasad, S. G. Advani, Journal of Power Sources, 165, 1, (2007).

  • W. Yuan, Y. Tang, X. Yang, Z. Wan, Applied Energy, 94 (2012).

Figure 1 (a) Water droplet on metal mesh surface before surface treatment, and (b) after surface treatment. (c) PEM fuel cell performance of a metal mesh GDl compared with two common commercial carbon papers (Toray HGP-H-080 and SGL 24 BC).

Figure 1

D-22 Carbon-based Supports 2 - Oct 5 2016 2:00PM

2580

, and

Production of energy through renewable and sustainable processes are important goals for the future in the context of depleting fossil fuels and environmental pollution. In this regard fuel cell technology offers an attractive combination of highly efficient fuel utilization and environmentally friendly operation. For successful commercialization of low temperature proton exchange membrane (PEM) fuel cells, they should have a long life and a high performance.

Recently, high performance fuel cell electrodes were obtained by electrospinning a solution containing proton conducting Nafion and commercial platinum catalyst powder along with PAA as electrospinning polymer. Electrospinning of the catalyst together with Nafion ionomer resulted in enhanced triple phase boundaries which resulted in high fuel cell performance. Moreover, it was also found that the stability of the electrodes is significantly higher than the electrodes prepared from the same catalyst by conventional methods. However, carbon corrosion still ensued.

In this work, we demonstrate the stability of polyacrylic acid (PAA) - Nafion composite as a stable support due to which the Pt after the carbon corrosion does not come out of the system. On the contrary, it stays entrapped in the membrane electrode assembly giving significance performance even after the corrosion protocols/measurements. Here, we first produce Pt nanoparticles that are produced by a photochemical reaction of the precursor induced by UV light. To this, PAA and Nafion (1:2 by weight) were added and stirred overnight. The resulting solution was electrospun. (Small amount of carbon was added in some cases to get the desired conductivity) to get a Pt/PAA-Nafion catalyst as shown in transmission electron micrograph. The electrodes are first tested for electrochemical stability by potential cycling and then in the fuel cell test bench using the FCCJ recommended cell evaluation protocol.

Figure 1

2581

, , , , , and

Proton exchange membrane fuel cell (PEMFC) is one of the most promising candidates as an environmentally clean power source in various applications. The efficiency of a catalyst depends strongly on the selection of an appropriate support material. The instability of carbon supports in fuel cell working conditions is the major obstacle hindering the commercialization of PEMFCs. Therefore, novel carbon supports are being sought for fuel cell application [1, 2]. Hierarchical microporous-mesoporous carbon supports were prepared from molybdenum carbide using the high temperature chlorination method at different synthesis temperatures within the range from 600 °C to 1000 °C. Prepared materials exhibited high specific surface area, but they differed from one another in their structural properties, such as pore size distribution and the ratio of micro- and mesopores volumes [1, 3-4].

The platinum nanoparticles were deposited onto the carbon support using the sodium borohydride method. Thermogravimetric analysis, X-ray diffraction, low temperature nitrogen sorption and high resolution transmission electron microscopy methods were used to characterize the structure of the synthesized materials.

Prepared catalysts were used in the single cell measurements. The polarization curves and chronopotentiometric measurements were conducted to evaluate the activity and stability of the catalyst materials synthesized. The electrochemically active surface area values were calculated and used to estimate the contact surface of platinum nanoparticles and Nafion electrolyte. The resistance of electrolyte, polarization resistance and activation energy values have been established from electrochemical impedance spectroscopy data. Prepared materials with well-defined properties are excellent objects to study the impact of the carbon catalyst support properties on the performance and stability of a PEMFC single cell.

The main aim of this work was to study the time stability (Fig. 1) of Pt nanoclusters activated microporous-mesoporous carbon supports in fuel cell application and to compare the properties of these materials with commercial carbon Vulcan XC72 used as catalyst support. It was found that the novel synthesized carbon supports are very stable and thus suitable for PEMFC application.

Acknowledgements.

This work was supported by Estonian target research project IUT20-13, Personal Research Grant PUT55, Estonian Centre of Excellence Projects No. 2014-2020.4.01.15-0011 and 3.20101.11-0030.

 

References

  • S. Sepp, K. Vaarmets, J. Nerut, I. Tallo, E. Tee, H. Kurig, J. Aruväli, R. Kanarbik, E. Lust, Electrochim. Acta 2016, doi:10.1016/j.electacta.2016.03.158

  • M. Uchida, Y. Fukuoka, Y. Sugawara, N. Eda, A. Ohta J. Electrochem. Soc. 1996, 143, 2245–2252.

  • S. Sepp, E. Härk, P. Valk, K. Vaarmets, J. Nerut, R. Jäger, E. Lust, J. Solid State Electrochem. 2014, 18(5), 1223 – 1229.

  • E. Lust, K. Vaarmets, J. Nerut, I. Tallo, P. Valk, S. Sepp, E. Härk, Electrochim. Acta 2014, 140, 294-303.

Figure 1. Galvanostatic life-time measurements for single cells with synthesized and commercial carbon used as catalyst support.

Figure 1

2582

Understanding the relationship between structures and properties in PEFC materials and/or components are essential in designing better PEFCs and their materials. Recent progress in analysis methods point toward this direction: combination of basic and advanced structure analyses and imaging methods using x-rays and electron probes, as well as computational simulations correlate structural information and electrode properties.

Quick overview of x-ray and electron beam analysis methods using various physical phenomena, such as, diffraction, scattering, x-ray absorption, x-ray photo-emission, and latest challenges of their application to the catalyst and catalyst-layer developments for both platinum-based and non-platinum catalysts will be presented.

2583

, , , , , , , and

One of the main bottlenecks in the operation of proton-exchange membrane fuel cells (PEMFCs) is the sluggish kinetics of the oxygen reduction reaction (ORR). The latter gives rise to large overpotentials, which degrade significantly the energy conversion efficiency of the device [1]. Our work addresses this issue by the development of innovative ORR electrocatalysts (ECs) characterized by an improved turnover frequency in comparison with state-of-the-art materials [2]. The proposed ECs comprise a hierarchical support "core" based on graphene flakes, which is covered by a carbon nitride (CN) "shell" embedding the ORR active sites. The hierarchical support "core", which typically comprises both graphene flakes and a carbon black spacer, is devised to facilitate the charge and mass transport phenomena by: (i) reaping the benefits of graphene (e.g., a very high electron mobility, up to 200000 cm2·V-1·sec-1, and specific surface area, up to ca. 2630 m2·g-1); and (ii) fine-tuning the morphology of the final ECs [3-6].

The proposed ECs are obtained by the optimization of the preparation protocol devised in our research group [7]. In particular, the physicochemical properties and the morphology of the final materials are modulated by a post-synthesis activation step carried out by electrochemical de-alloying. The final ECs bear bimetallic ORR active sites comprising Pt as the "active metal" and Ni as the "co-catalyst" [8]. Preliminary results clearly evidence that this approach is capable to yield hierarchical ECs exhibiting a very promising performance in the ORR despite a very low loading of Pt. In detail, the best EC exhibits an ORR onset potential ca. 30 mV higher with respect to that of the Pt/C reference (see Figure).

The chemical composition of the ECs is determined by inductively-coupled plasma atomic emission spectroscopy (ICP-AES) and microanalysis. The structure is elucidated by wide-angle X-ray diffraction (WAXD) and vibrational spectroscopies (e.g., confocal micro-Raman); the morphology is probed by, both conventional and high-resolution, scanning electron microscopy (SEM) and transmission electron microscopy (TEM); the "ex-situ" electrochemical performance and ORR reaction mechanism are gauged by cyclic voltammetry with the rotating ring-disk electrode method (CV-TF-RRDE). Finally, the ECs are implemented at the cathode of single fuel cell prototypes which are tested in operating conditions.

REFERENCES

[1] I. Katsounaros, S. Cherevko, A. R. Zeradjanin, K. J. J. Mayrhofer, Angew. Chem. Int. Ed., 53, 102 (2014).

[2] J. Zhang, Front. Energy, 5, 137 (2011).

[3] S. Sharma, B. G. Pollet, J. Power Sources, 208, 96 (2012).

[4] M. Liu, R. Zhang, W. Chen, Chem. Rev., 114, 5117 (2014).

[5] A. C. Ferrari, F. Bonaccorso, V. Fal'ko et al., Nanoscale, 7, 4587 (2015).

[6] J. H. Chen, C. Jang, S. Xiao, M. Ishigami, M. S. Fuhrer, Nature Nanotech., 3, 206 (2008).

[7] V. Di Noto, E. Negro, K. Vezzù, F. Bertasi, G. Nawn, L. Toncelli, S. Zeggio, F. Bassetto, Patent application 102015000055603 filed on 28 September 2015. Applicants: Università degli Studi di Padova and Breton S.p.A. (2015).

[8] V. Di Noto, E. Negro, K. Vezzù, F. Bertasi, G. Nawn, The Electrochemical Society Interface, Summer 2015, 59 (2015).

Figure 1

2584

, , , , , , , , and

The cost and durability of the electro-catalysts used in the proton exchange membrane fuel cells (PEMFCs) are two critical factors affecting the early market penetration for both transportation and stationary applications. Although progress has been made in the rational designs of highly active ORR electro-catalysts with excellent performance characterized by rotating disc electrode (RDE), [1, 2] there still remains a huge challenge to implement these RDE findings into an actual membrane electrode assembly (MEA) for PEMFC systems. Because the ORR performance is largely influenced by the ionomer/catalyst interface within the catalyst layers of a MEA, the "ideal" interface should contain 100% ionomer coverage for maximizing catalyst utilization. In addition, the thickness of ionomer film over the catalyst nanoparticles should be optimal to facilitate gas diffusion and water balance without sacrificing its protonic conductivity. Additionally, the appropriate pore structure in the catalyst layer is also necessary for providing transport paths for both reactants (O2 and H2) to reach reaction sites as well as allow water mobility throughout the catalyst layer. In some of the more conventional MEA fabrication processes, the ionomer coverage of the catalyst particles can be partial or non-uniform. This is partly due to the lack of control of ionomer deposition onto the catalyst surface. As a consequence, some of the catalyst becomes electrochemically inactive. Furthermore, insufficient ionomer coverage suppresses the amount of H+ transport, thus impacting the oxygen reduction reaction (ORR). On the other hand, the aggregation of ionomer with increasing ionomer content (to increase catalyst coverage) in a catalyst layer leads to a thicker ionomer film, which results in an increased gas and water diffusional barrier.

We will present a novel approach to construct an innovative ionomer/catalyst interface with high ionomer coverage and a thin ionomer layer. This is done by an electrostatic charge attraction of a negatively charged "-SO3-" on the surface of ionomer particles and a positively charged "-NH3+" on the surface of catalyst particles (catalyst surface charge is realized by chemically grafting different functional groups via diazonium reaction). In our approach, the improved ionomer/catalyst interface is formed during the ink preparation. Using a unique method of combined ultra-small angle x-ray scattering (USAXS) and cryo-TEM, we observed a significant increase on carbon aggregate size after the NH3+- functionalized carbon black (CB) particles were mixed with Nafion ionomer, whereas only a negligible size change was observed when SO3- functionalized CB was used. These results suggest that the coulombic attraction facilitates the mixing of catalyst and ionomer. The ionomer/catalyst interface was characterized by Scanning Transmission Electron Microscopy (STEM) combined with Energy-dispersive X-ray Spectroscopy (EDS) mapping. Figure 1(b) shows the EDS mapping of F distribution for both NH3+-CB and SO3--CB. We will report on the performance of MEAs using this approach.

Reference

[1] C. Chen, Y. Kang, Z. Huo, Z. Zhu, W. Huang, H.L. Xin, J.D. Snyder, D. Li, J.A. Herron, M. Mavrikakis, M. Chi, K.L. More, Y. Li, N.M. Markovic, G.A. Somorjai, P. Yang, V.R. Stamenkovic, Highly Crystalline Multimetallic Nanoframes with Three-Dimensional Electrocatalytic Surfaces, Science, 343 (2014) 1339-1343.

[2] X. Huang, Z. Zhao, L. Cao, Y. Chen, E. Zhu, Z. Lin, M. Li, A. Yan, A. Zettl, Y.M. Wang, X. Duan, T. Mueller, Y. Huang, High-performance transition metal–doped Pt3Ni octahedra for oxygen reduction reaction, Science, 348 (2015) 1230-1234.

Figure 1

2585

, , , , , and

Graphene synthesized from the exfoliation of graphite has a great potential in polymer electrolyte membrane fuel cell (PEMFC) applications.[1] The highly graphitic structure composed of the conjugated carbons endows graphene with extremely high electric conductivity and chemical/electrochemical stability. Improved stability is required for PEMFC catalyst support materials as the stability plays a critical role in determining the overall durability of PEMFC systems. While graphene is an attractive material, there are three major barriers when using graphene-supported Pt-based catalysts in PEMFCs: 1) the lack of bonding sites for catalyst landing on the basal plane of graphene causes the migration/aggregations of Pt nanoparticles when subjected to harsh accelerated stress tests (ASTs), 2) the highly hydrophobic surface of graphene is difficult to wet when mixed with the Nafion ionomer particles, leading to a poor catalyst/ionomer interface, 3) 2D graphene sheets tends to restack back to graphite structure through π-π interactions, which can severely block the O2 diffusion and retard the catalytic reactions, in particular when PEMFCs are operated at high current density (>1.5 A/cm2),and 4) the exfoliation of natural graphite using wet chemistry method usually yield graphene that contains some defects and oxygenated functional groups, which can de-stabilizes its conjugated electronic structure which negatively affects electronic conductivity and stability. Furthermore, when subjected to harsh potential cycling (e.g. 1.0-1.5 V; potentials which can be observed during start-up/shut-down or local fuel starvation), these defect sites are vulnerable to carbon corrosion.

To overcome these challenges, we use a novel approach to make the 2D graphene sheets into 3D composite materials with many channels and pores to facilitate the facile mass transport by developing the highly stable hierarchical polybenzimidazole (PBI) -grafted graphene hybrids supported Pt catalysts for PEMFCs. PBI- functionalization is found to homogenize the chemical environment of graphene surface, resulting in a uniform dispersion of Pt nanoparticles that exposes more active sites for electro-catalytic reactions. In order to construct appropriate pore structures in the catalyst layers, spacers were introduced using graphitized carbon materials or metal oxides. The surface charge was imposed on the surfaces of these spacer-particles as opposed to PBI-graphene, so that the columbic attraction forces drive them to anchor onto graphene sheets. As a result, the secondary pore volume (characterized by mercury porosimetry) significantly increases, leading to the improved mass transport of reactants (H2 and O2) and water, and consequently an improvement in the PEMFCs performance at high current density (i.e. 2.0 A/cm2). Additionally, we prepared graphene nanoplatelet with smaller dimensions as the catalyst supports, which leads to extra voltage gains at the high current density. Covalent grafting polymers onto graphene sheets containing semi-flexible backbones on graphene defects can help to re-build the conjugated carbons and de-localize the electron, as well seal the dangling bond resulting from graphene synthesis. The PEMFC durability studies carried out by our research group has shown significant improvement on support stability when they were cycled between 1.0 V to 1.5 V for 5000 cycles compared to traditional carbon support materials,[RB1]during which the mass activity is only decreased by 47% for Pt/PBI-graphene and 25% for Pt/PBI-nanographene. The reported results are very close to DOE 2020 targets on catalyst support stability [2].

In addition, these materials show improved Pt stability. Strengthened interaction between functional groups and Pt particles, as characterized by X-ray photoelectron spectroscopy (XPS) and X-ray absorption spectroscopy (XAS) greatly improves the catalyst stability. Our developed PBI-graphene supported Pt demonstrated a 27% decrease in mass activity and 15% loss of ECSA after 10,000 cycles between 0.6 V and 1.0 V. This exceeds the DOE 2020 targets for automotive PEM fuel cell electro-catalysts.[2] The unique advantages of the hierarchical PBI-functionalized graphene/nanographene hybrids demonstrates the potential for these materials to be the next generation of catalyst supports for PEMFCs.

[1]. Geim, A. K.; Novoselov, K. S. Nat. Mater. 2007, 6, 183

[2]. U.S. DRIVE FuelCell Tech Team Cell Component Accelerated Stress Test Protocols for PEM Fuel Cells (2013)

2586

and

Cathode durability, cost, and performance are some of the key issues preventing the commercialization of polymer electrolyte membrane fuel cells (PEMFCs) for automotive applications. The support degradation during start-up/ shut-down and the catalyst degradation during driving potential cycling limit overall cathode catalyst lifetime. The ultra-low Pt cathode catalyst development project at the University of South Carolina (USC) has developed highly-stable activated carbon composite supports (ACCS) as well as highly-active and stable compressive Pt lattice cathode catalysts, Pt*/ACCS (Pt* = Compressive Pt lattice catalyst) that retain their activity after accelerated stress tests (AST) to simulate start-up/shut-down and driving potential cycling [1-7].

The novel USC technology is based on a two-step patented process. In the first step, the following major constraint was addressed: the support should be chemically and electrochemically stable at high potentials, low pH, and high temperature. To accomplish these requirements, ACCS was synthesized with optimization of: (i) BET surface area, porosity, pore-size and distribution, (ii) hydrophilic/ hydrophobic ratio, (iii) structural properties (amorphous/ crystalline ratio), and (iv) Pt-support interaction through inclusion of active surface functional groups.

In the second step, compressive Pt lattice catalyst (Pt*) was synthesized through a USC-developed annealing procedure that controls the particle size during annealing and forms several monolayers of Pt* by diffusing Co atoms into Pt which is deposited on ACCS support. The Pt*/ACCS shows high rated power density (0.174 gPt kW−1), excellent support stability (8 mV loss at 1.5 A cm−2), and enhanced catalyst durability (24 mV loss at 0.8 A cm−2) under AST conditions.

References

[1] T. Kim, B. N. Popov, Int. J. Hydrogen Energ., 41 (2016) 1828-1836.

[2] T. Kim, T. Xie, W. Jung, F. Gadala-Maria, P. Ganesan, B.N. Popov, J. Power Sources, 273 (2015) 761-774.

[3] W. Jung, T. Xie, T. Kim, P. Ganesan, B.N. Popov, Electrochim. Acta, 167 (2015) 1-12.

[4] T. Xie, W. Jung, T. Kim, P. Ganesan, B.N. Popov, J. Electrochem. Soc., 161 (2014) F1489-F1501.

[5] Á. Kriston, T. Xie, B. N. Popov, Electrochim. Acta, 121 (2014) 116-127.

[6] Á. Kriston, T. Xie, D. Gamliel, P. Ganesan, B. N. Popov, J. Power Sources, 243 (2013) 958-963.

[7] Á. Kriston, T. Xie, P. Ganesan, B. N. Popov, J. Electrochem. Soc., 160 (2013) F406-F412.

Figure 1

2587

, , , , , , , , , et al

Proton exchange membrane fuel cells (PEMFC) are energy converters that can power nomad, automotive or stationary systems without emission of pollutants. This technology is already used for niche markets, but some challenges must be overcome to enable large-scale deployment. In particular, the electrochemical corrosion of the carbon support at the cathode remains a major concern especially under start-stop conditions. This corrosion causes massive detachment of the supported Pt-based nanocrystallites, negatively affects water management and ultimately results into a collapse of the electrode structure.

In this study, we bridged the textural and the structural properties of a set of carbons supports to their resistance to corrosion. The stability of carbon nanotubes (CNT, graphitized but with a low specific surface area), carbon blacks (featuring either large specific surface area but poorly-structured, CBe, or low specific surface area and high degree of graphitization, CBs) and carbon aerogel (controlled texture but low degree of organization, CA) was investigated via accelerated stress tests derived from the FCCJ organization. A trade-off was found in terms of properties: graphitic carbons are more resistant to oxidation, whereas carbon supports featuring high specific surface areas are more favorable to the dispersion of a large amount of catalyst nanoparticles per unit volume.

Controlled fluorination of these model carbon supports was performed so as to improve their resistance to electrochemical corrosion [1]. The fluorination was carried out by an easily scalable solid gas process using pure molecular fluorine [2]. The parameters were set in order to limit the fluorination below 0.2 fluorine atom per carbon atom and to avoid any carbon decomposition into gaseous fluorinated compounds such as CF4 and C2F6 [3]. Platinum nanoparticles were also deposited (40 wt.%) by a colloid (polyol) method on the bare and fluorinated carbons. The samples were physically and chemically characterized by XRD, TEM (Figure 1), nitrogen sorption, FTIR and TGA. The catalytic activity of these electrocatalysts towards the oxygen reduction reaction was determined by linear sweep voltammetry (rotating disk electrode technique). Accelerated stress tests, load cycle (0.6-1.0 V) and start-up/shutdown (1.0-1.5 V) protocols, conducted at T= 80°C in a four-electrode cell [4], were performed to investigate the robustness of the bare and fluorinated Pt electrocatalysts. The results and the impact of the fluorination are discussed and compared to those of a commercial 40 wt.% electrocatalyst.

[1] S. Berthon-Fabry et al. First insight into fluorinated Pt/carbon aerogels as more corrosion-resistant electrocatalysts for Proton Exchange Membrane Fuel Cell cathodes, Electrocatalysis, 6, 6 (2015) 521-533

[2] N. Batisse, P. Bonnet, M. Dubois, K. Guérin, Fluorination of 0D, 1D and 2D nanocarbons" In: Carbon Nanomaterials Sourcebook (Eds. Taylor & Francis Publisher), 2016[3] Y Ahmad et al., The synthesis of multilayer graphene materials by the fluorination of carbon nanodiscs/nanocones. Carbon, 50, 10 (2012) 3897-3908

[4] L. Dubau, F. Maillard, "Unveiling the crucial role of temperature on the stability of oxygen reduction reaction electrocatalysts", Electrochem. Commun., 63 (2016) 65-69.

This work is funded by the French National Research Agency programme, (ANR-14-CE05-0047 project CORECAT and is supported by Capenergies and Tenerrdis

Figure 1: TEM images of bare CBe (left), Pt/CBe (middle) and Pt/F-CB (right) samples

Figure 1

2588

, , and

The polymer electrolyte membrane fuel cells (PEMFCs) are regarded as a promising energy conversion device because of high efficiency and eco-friendly energy generation. However, the prohibitive cost of PEMFCs still remains big problem for its a commercialization. The platinum which is considered promising metal for oxygen reduction reaction (ORR) accounts for the biggest part of the cost. Therefore, most researches have concentrated on reducing the amount of platinum or developing non-precious metal catalyst.

In this study, we employed nitrogen doped carbon to improve performance of Pt catalyst. A lot of researches have shown that Pt supported on the nitrogen doped carbon is prospective catalyst for PEMFC.[1] And also, we applied the nitrogen doped carbon to non-precious metal catalyst. The nitrogen doped carbon was promising candidate to substitute Pt based catalyst, because the nitrogen doped carbon have catalytic activity without metal such as platinum.[2]

The synthesis method to make nitrogen doped carbon is very diverse. In this study, we adopted a method that utilized carbonized polyaniline (PANI). The PANI have been used as a source of nitrogen doped carbon. Gavrilov et al. reported carbonized PANI supported Pt catalyst for ORR.[3] Their catalysts had good catalytic activity in acidic and alkaline media. Wu et al. reported that PANI-M-C (M = Fe and/or Co) which was comparable to the performance of conventional Pt/C catalyst was successfully synthesized with PANI.[4]

We suggested differentiated method from other researches to synthesized PANI based catalyst. Our experiment process to make PANI based catalysts are like follows. First of all, metal precursor (Pt or Fe), aniline monomer and carbon black was fully mixed. The mixture was irradiated by ultrasound and the irradiated mixture was filtered. After filtering, we obtained M-PANI-C (M = Pt or Fe, C = carbon black). Finally, the M-PANI-C was heat-treated in nitrogen atmosphere to carbonize PANI. In our experiment, the role of metal precursor was not only metal source but also oxidant for polymerization of aniline. And the ultrasound made the metal particle well dispersed in PANI matrix. Consequently, this method suggested more simple process than other methods and made it possible to polymerize aniline without any oxidant.

Fig 1. is the linear sweep voltammetry(LSV) result of Pt-PANI-C. The LSV curves shows that the synthesized Pt-PANI-C catalyst has better performance for ORR than conventional Pt/C catalyst. The Pt contents of Pt-PANI-C was 10 wt%, which is only half the Pt content of Pt/C catalyst. (Pt contents could be obtained by ICP-MS analysis) This result indicates that the PANI was successfully applied to ORR catalyst and the PANI could enhance performance of Pt based catalyst. Fig 2. is the LSV result of Fe-PANI-C catalyst in alkaline electrolyte. The Fe-PANI-C catalyst was acid leached for removing impurity and iron oxide before electrochemical analysis. The ORR performance of Fe-PANI-C catalyst is comparable to Pt/C catalyst. The both of catalysts were analyzed with various tools to characterize catalysts and determine the reason why performance was improved. And furthermore, the catalysts were applied to single cell test and we evaluated the single cell performance of catalysts. All the results investigated in the research will be presented with in-depth discussion.

References

1. Wang, C.-H., et al., High methanol oxidation activity of electrocatalysts supported by directly grown nitrogen-containing carbon nanotubes on carbon cloth. Electrochimica Acta, 2006. 52(4): p. 1612-1617.

2. Shao, Y., et al., Nitrogen-doped carbon nanostructures and their composites as catalytic materials for proton exchange membrane fuel cell. Applied Catalysis B: Environmental, 2008. 79(1): p. 89-99.

3. Gavrilov, N., et al., Carbonized polyaniline nanotubes/nanosheets-supported Pt nanoparticles: Synthesis, characterization and electrocatalysis. Materials Letters, 2011. 65(6): p. 962-965.

4. Wu, G., et al., High-performance electrocatalysts for oxygen reduction derived from polyaniline, iron, and cobalt. Science, 2011. 332(6028): p. 443-7.

Figure 1

2589

, , , and

As the fuel cell market has been growing, the price competitiveness of polymer electrolyte fuel cell (PEFC) became the most important factor for its commercialization. In particular, the production of membrane-electrode assembly (MEA) consisting of Pt-based electrocatalyst is a critical hurdle for the cost reduction of PEFCs. Therefore, many researchers have veen focusing on the development of cost effective electrocatalysts such as non-precious metal based catalysts (NPMCs) or low content platinum based catalysts to minimize Pt loading by morphological and structural tunings with inexpensive transition metals. Recently, we reported the M-N doped ordered mesoporous porphyrinic carbon (M-OMPC, M=Fe, Co) which showed a comparable ORR activity to that of Pt/C in an acid medium. However, an inferior MEA performance of M-OMPC is a practical barrier because of high mass transport resistance and electrical resistance, which mainly comes from the thicker electrode of M-OMPC. In this study, we proposed a new approach for enhanced ORR activity as well as reduction of noble metal loading content via adding a trace of platinum nanoclusters on M-OMPC. The Pt/M-OMPC (Pt~5 wt%) hybrid catalyst showed an increased Pt based mass activity and specific activity by a factor of 7, compared to that of Pt/C (5 wt%) at 0.85 V RHE.

A-22 Catalyst Layer 1 - Oct 5 2016 2:00PM

2590

, and

The catalyst layer (CLs) are the central part of Polymer Electrolyte Fuel Cells. Consisting of catalyst particles, ionomer, carbon black plus numerous gases and liquid water during operation, the CL contains not less than five different phases with a plethora of mutual physical interactions. Further, all these phases have characteristic length scales in the nanometer range. Due to its nanoscale, multi-phase nature the CL bears complex physical phenomena that are very challenging to describe and understand.

The precise morphology of all phases within a CL plays a key role for reactant transport and thus  performance: Platinum catalyst particles with poor connection to ion conducting, electron conducting or gas transporting phases cannot contribute to the fuel cell reaction. Also, the precise shape of the pathways of the reaction species determines their effective transport properties. Wetting or non-wetting pore wall areas influence water precipitation and therefore the generation of liquid water networks. Such networks both improve ionic transport and adulterate gas transport. 

Facing the tremendous influence morphology has on PEMFC performance, cost and durability, tools to image morphology are an obvious necessity. While there is a variety of imaging methods that allow three-dimensional reconstructions, only few are able to resolve the nanostructure of CLs: x-ray tomography (Xt), focused ion beam / scanning electron microscopy tomography (FIB-SEMt) and transmission electron microscopy tomography (TEMt) [1].

Both Xt and FIB-SEMt have been proven to enable differentiating porous and solid phase within PEMFC CLs [2,3]. TEMt allows imaging with resolutions below 1 nm and can be directly used to image Platinum particles [4]. To image all important phases multiple tomographic techniques must be used. In this talk we give an overview on the latest tomographic analysis techniques for PEMFC CL reconstruction with a particular emphasis on FIB-SEMt, multi-scale imaging approaches and validity of the existing reconstruction methods.

References

[1]  G. Möbus, B.J. Inkson, Materials Today 10 (2007) 18–25.

[2]  C. Ziegler, S. Thiele, R. Zengerle, J. Power Sources 196 (2011) 2094–2097.

[3]  W.K. Epting, J. Gelb, S. Litster, Advanced Functional Materials (2012).

[4]  H. Uchida, J.M. Song, S. Suzuki, E. Nakazawa, N. Baba, M. Watanabe, The Journal of Physical Chemistry B 110 (2006) 13319–13321.

[5]          S. Thiele, T. Fürstenhaupt, D. Banham, T. Hutzenlaub, V. Birss, C. Ziegler, R. Zengerle, J. Power Sources 228 (2013) 185–192.

2591

, , , and

In polymer electrolyte fuel cells, the freezing of product water deteriorates fuel cell performance at sub-zero temperatures. When ice forms in the membrane electrode assembly (MEA), catalyst sites are blocked from oxygen resulting in large portions of the MEA to become inactive during Freeze Start- up (FSU).

Improving the ice tolerance of the CCM is important to increase the robustness of the MEA to Freeze-Start up failure mechanisms. It is believed that the cathode catalyst layer is one of the key sub-components for ice tolerance, which is the ability of the MEA to accumulate ice before the MEA becomes inoperable. Isothermal Constant Current (ICC) measurements in units of (C/cm2) are commonly used to indicate this capability.

ICC data shows a clear correlation of ice tolerance with thickness of the catalyst layer. However, many of the control factors which can be used to improve ice tolerance in MEAs by increasing the thickness of the catalyst layer are limited by design constraints related to material cost and performance trade-offs. It would be difficult to achieve catalyst layer thickness >20um without seriously impacting performance under hot conditions or mass transport at high current densities.

Using a carbon/ionomer layer adjacent to the cathode catalyst layer allows to tune-in cathode attributes related to water management and ice distribution under freezing conditions, such as: thickness, porosity, hydrophilicity, interfaces and electric resistance. By changing attributes of the carbon bilayers separately from the cathode catalyst layer, the impact to the ORR reaction under normal or hot/dry operating conditions and mass transport region can be minimized.

2592

, , , and

Proton exchange membrane fuel cells (PEMFCs) are promising energy conversion devices due to their high efficiency, high energy density and low operating temperatures [1]. One of the major obstacles for high performance PEMFCs is that the sluggish oxygen reduction reaction (ORR) and the ORR activity in a membrane electrode assembly (MEA) are dependent of the structure of the ionomer/catalyst interface in catalyst layers. Coverage of the ionomer over the surface of catalyst particles and the thin ionomer film over the catalyst particles, as well as the porosimetry of the catalyst layer are critical factors to be considered in the fabrication of MEAs. Clearly, the determination of ionomer coverage over the surface of the catalyst particles, the morphology of ionomer film over the surface of catalyst particles and porosimetry of catalyst layer of MEAs are of paramount significance for developing high performance and low cost MEAs. Transmission electron microscopy (TEM) is a powerful tool to provide direct observations of the structures.

In this work, we present the TEM analysis of MEAs for (1) ionomer coverage over carbon particles, (2) the thickness of ionomer film over carbon particles and (3) the porosimetry of catalyst layer using different TEM techniques. MEAs with different functional groups (which were chemically grafted on the surface of carbon support particles via diazonium reactions) were fabricated using (1) NH3+- functionalized carbon black (CB) and (2) SO3- functionalized CB. These two types of MEAs were analyzed in terms of ionomer/catalyst interface and porosimetry. In order to prepare electron transparent TEM samples, different methods were applied, including normal/partial embedding ultra-microtomy and cryo-microtomy. Figure 1 show scanning transmission electron microscopy (STEM) images of catalyst layers (CLs) of two MEAs, as well as energy-dispersive X-ray spectroscopy (EDS) mapping of the same area. Combining these results, it is evident that the fluorine in the NH3+- functionalized carbon CL is uniformly distributed over the entire region. Meanwhile, the SO3- functionalized carbon CL (as circled in Figure 1c) shows non-uniform distribution; the F signal is much weaker in some areas than others or even disappeared. These maps indicate that the NH3+- functionalized carbon CL has better ionomer coverage than that of the SO3- functionalized carbon CL. On the other hand, the X-ray spectra of both samples are shown in Figure 2. The C/F ratio in NH3+- functionalized carbon CL is 216 while it is only 16 in SO3- functionalized carbon CL. Such a large discrepancy in the C/F ratio between two different CLs are mainly due to the CL's intrinsic property - the ionomer coverage difference - as these two TEM samples were prepared using the same method. As these two MEAs are prepared with the same catalyst support/ionomer ratio, higher C/F signal ratio in NH3+- functionalized carbon CL indicates more complete coverage and thinner ionomer layer. An overview of catalyst layers in both CLs is shown by the TEM bright-field images in Figure 3. The thickness of catalyst layer made of NH3+- functionalized CB is approximately 6.3μm, while it is only approximately 1.4 μm of that made of SO3- functionalized CB. It suggests that the carbon support aggregates are much smaller in the catalyst layer made of NH3+- functionalized CB, based on the fact that two MEAs are synthesized with same amounts of carbon supports and ionomer. Furthermore, 3D tomography TEM and reconstruction of both MEAs based on focused ion beam (FIB) supports this conclusion. Highly porous catalyst layer of NH3+- functionalized CB facilitates mass transport for both reactants (H2 and O2) to reach reactive sites and water dissipated out.

In summary, TEM characterization shows that NH3+- functionalized CB catalyst layer has better ionomer coverage, thinner ionomer layer, as well as better porosimetry than SO3- functionalized CB catalyst layer. In consistency with the TEM characterizations, the NH3+- functionalized CB MEA provide better electrochemical performance.

Reference

[1] Kocha, S. S. Principles of MEA preparation. In Handbook of Fuel Cells − Fundamentals, Technology and Applications, ed. 1; Vielstich, W., Lamm, A., Gasteiger, H. A., Eds.; Wiley: Chichester, UK, 2003; Vol. 3, 538

Figure 1

2593

, , , , , , and

Proton exchange membrane fuel cells for automotive applications (PEMFC) frequently experience transient conditions such as start-up/shut-down, freeze start-up and local fuel starvation at the anode catalyst layer (ACL) that are responsible for the loss of performance and durability as a result of irreversible carbon corrosion [1].

 

To mitigate such degradation, oxygen evolution catalysts have been incorporated into ACL and studied for effects on membrane electrode assembly (MEA) performance and reversal tolerance. Ru and Ir oxides are among the most studied catalysts for oxygen evolution reaction (OER) in both PEMFCs and water electrolyzers [2-5]. Although comparative measurements have been performed to investigate the OER activity trends, systematic studies focusing on effects of particle/agglomerate size and distribution within ACL on PEMFC MEA reversal tolerance have not yet been reported. 

 

In order to be able to design anode catalyst layer structures that meet the performance and reversal tolerance requirements, it is necessary to obtain a better understanding of structure versus performance relationships. This requires the capability to fabricate different anode catalyst layer structures, as well as efforts in development of methodologies to characterize the spatial distribution of all components within the catalyst layers. The ultimate goal is to use the experimentally measured structural parameters as inputs for the development of models for understanding of the relationship between anode structures and reversal tolerance. The objective of this work is to present the methodology for structural characterization of anode catalyst layers containing OER materials. Structural parameters correlating with the extended reversal tolerance will be discussed.

 

References

  • S. Enz, T.A. Dao, M. Messerschmidt and J. Scholta, Investigation of degradation effects in polymer electrolyte fuel cells under automotive-related operating conditions, J. Power Sources, 274 (2015) 521-535

  • E. Antolini, Ir as catalyst and cocatalyst for oxygen evolution/reduction in acidic polymer electrolyte membrane electrolyzers and fuel cells, ACS Catal. 4 (2014) 1426-1440

  • T. Reier, M. Oezaslan and P. Strasser, Electrocatalytic oxygen evolution reaction (OER) on Ru, Ir and Pt catalysts: a comparative study of nanoparticles and bulk materials, ACS Catal. 2 (2012) 1765-1772

  • S. Cherevko, T. Reier, A.R. Zeradjanin, Z. Pavolek, P. Strasser and K.J.J. Mayrhofer, Stability of nanostructure iridium oxide electrocatalysts during oxygen evolution reaction in acidic environment, Electrochem. Commun. 48 (2014) 81-85

  • M. Bernicke, E. Ortel, T. Reier, A. Bergmann, J.F. de Araujo, P. Strasser and R. Kraerhnet, Iridium oxide coatings with template porosity as highly active oxygen evolution catalysts: structure-activity relationship, ChemSusChem, 8 (2015) 1908-1915

2594

, , , , , , , and

Fuel cell electric vehicles (FCEVs) are an effective solution to reduce the emission of carbon dioxide (CO2) and the consumption of petroleum fuel. The biggest issues to popularize FCEVs are cost reduction of polymer electrolyte fuel cell (PEFC) power systems, and the development of hydrogen infrastructure. On the development of PEFC stack for automotive application, the cost of catalyst and application accounts for 49% of the total stack cost at 500,000 units per year1). This estimation indicates that reducing platinum loading is important for the cost reduction of PEFC stack. However, low platinum loading increases the activity and mass transport losses and lowers the performance of PEMFC stack. Increasing oxygen reduction reaction (ORR) activity of catalyst materials such as platinum alloys is an effective way to decrease the above losses. Optimizing the structure of catalyst layer is also required to reduce the mass transport loss. This means the preparation process of catalyst layer is the important factor to reduce the performance losses on low platinum loading.

In general, wet coating process is used for forming PEFC catalyst layer. Catalyst ink, dispersed platinum loaded carbon (Pt/C) and ionomer in solvent, are coated on the substrate or polymer electrolyte membrane. The catalyst ink coated substrate is dried to obtain the catalyst layer. Therefore, understanding the structure of catalyst ink is needed to control the structure formation of catalyst layer. Due to the fluidity of catalyst ink, visualization by cryogenic electron microscopy is an effective way to analyze the structure of catalyst ink. We observed the dispersion of Pt/C in catalyst ink by cryogenic scanning electron microscopy (Cryo-SEM)2). The result suggests that the dispersion of Pt/C aggregates correlates to the structure and performance of catalyst layer. However, the resolution of Cryo-SEM was insufficient to detect ionomer molecules in catalyst ink because of their molecular size.

For above reason, we tried to visualize the dispersion of ionomer molecules by Cryo-TEM. Sample preparation for the observation of Cryo-TEM was based on the rapid freezing technique used in structural analysis of biological specimens3). The frozen thin film of catalyst ink were prepared by immersion into liquid Ethane. Observation samples were cooled by liquid helium in order to reduce the damage for the ionomer by electron irradiation.

In this study, we focused on the effect of two factors, which were the solvent and Pt/C, on the dispersion of ionomer molecules. Cryo-TEM image of catalyst ink is shown in Fig.1. The net-like structure formed by ionomer strings were observed in case of catalyst ink. The average width of the ionomer string formed the net-like structure is less than 10nm approximately, which are the same dimension as the result analyzed by small angle X-ray scattering in previous study4). These results suggest the possibility that observed ionomer strings are bundled ionomer molecules together. The end part of the ionomer net-like structure seems to connect to the surface of the Pt/C aggregate. In this presentation, we will report the considerable mechanism that ionomer molecules form the net-like structure.

 

References:

1) D. Papageorgopoulos, "Fuel Cells Program", 2014 Annual Merit Review and Peer Evaluation Meeting, June 16-20 (2014)

2) S. Takahashi, T. Mashio, N. Horibe, K. Akizuki, A. Ohma, ChemElectroChem., 2, 1560 (2015).

3) A. Miyazawa, Y. Fujiyoshi, M. Stoewll, N. Unwin, J. Mol. Biol., 288, 765 (1999)

4) M. Yamaguchi, T. Matsunaga, K. Amemiya, A. Ohira, N. Hasegawa, K. Shinohara, M. Ando, T. Yoshida, J. Phys. Chem. B, 118, 14922 (2014)

Figure 1

2595

, , and

Introduction

Commercializing polymer electrolyte membrane fuel cells (PEMFC) is ultimately a matter of achieving the necessary cost for broad market penetration. Cost is a function of materials, manufacturability, performance, and durability. Performance and durability targets have been set by the DOE [1] to provide guidance and focus to the PEMFC industry for specific applications such as the automotive sector.

To cascade these challenging targets down to material and transport requirements we employ a fully integrated performance model to conduct parameter optimization studies in tandem with experimental validation. Design curves are utilized as inputs to the model to provide the current state-of-the-art capability regarding catalyst activity and ionomer proton conductivity.

High performing commercial membrane and GDL components were selected with a standard catalyst (Figure 1) to provide a baseline for this evaluation. The gap between modelled MEA performance and DOE targets will be highlighted and a recommended path forward will be provided.

Results & Discussion

Aside from catalyst activity and available surface area, the cathode catalyst layer performance is dictated by the mass transport of protons (proton conductivity), oxygen (gas diffusivity), and water (gas & liquid permeability) [2-4]. Several material sets have been evaluated experimentally in-situ to provide a range in catalyst activity, proton conductivity, and effective layer diffusivity as shown in figures 1 and 2 (diffusivity not shown).

During operation the distribution of current through the porous three dimensional catalyst layer structure [6] is dictated by the catalyst activity and layer transport properties. We will describe in this work the relationship between proton conductivity and voltage performance. Based on these results we can set relevant conductivity targets, which can be used for both ionomer development and material down selection.

These parameters will be utilized in the Ballard/DOE funded FC-Apollo performance model to provide the current status toward achieving the DOE's 2020 automotive targets listed in Table 1. Further, technology gaps will be identified in conjunction with a strategy toward meeting these objectives.

Acknowledgement

The authors would like to acknowledge Joey Jickain for his testing support and National Resources Canada (NRCan), National Research Council of Canada (NRC-IRAP), and the US Department of Energy (DOE) for funding various aspects of this work.

Reference

1. http://energy.gov/eere/fuelcells/doe-technical-targets-polymer-electrolyte-membrane-fuel-cell-components

2. Egushi, M., Baba, K., Onuma, T., Yoshida, K., Iwasawa, K., Kobayashi, Y., Uno, K., Komatsu, K., Kobori, M., Nishitani-Gamo, M., Ando, T., Polymers, 4, p. 1645 (2012)

3. Xie, J., Xu, F., Wood, D.L., More, K.L., Zawodzinski, T.A., Smith, W.H., Electrochimica Acta 55 p. 7404 (2010)

4. Gode, P., Jaouen, F., Lindbergh, G., Lundblad, A., Sundholm, G., Electrochimica Acta 48 p. 4175 (2003)

5. Young, AP., Gyenge, E., Stumper, J., J. Electrochem. Soc., 156, B913 (2009)

6. Young, AP., Knights, S., Gyenge, E., Stumper, J., J. Electrochem. Soc., 157, B425 (2010)

Figure 1

2596

and

Low-temperature proton exchange membrane fuel cells (PEMFC) may become essential energy conversion devices in the automotive sector and contribute to a future hydrogen economy. Corrosion of the cathode catalyst support is known to degrade performance but many specific details of the changes in performance are still poorly understood. A focused ion beam with scanning electron microscope (FIBSEM) method of nanotomography was developed for postmortem direct observation of corroded cathode microstructures. Unique to corroded cathodes, additional charge-reduction measures were taken to ensure high-fidelity imaging and low-error rate image processing. From three-dimensional reconstructed volumes, computational microstructural characterization methods were developed to extract pertinent transport parameters (e.g. porosity and tortuosity distributions) and characterize changes between pristine and corroded samples. FIBSEM and computational methods were coupled with traditional electrochemical techniques and carbon mass loss detection. Taken together, this work helps clarify and connect the roles of microstructure, corrosion, performance, and durability in PEMFCs.

Figure 1

2597

, , , , and

Among the different Fuel Cell technologies, one is based on the implementation of Polymer Electrolyte Membrane fuel cells (PEMFC). Significant research effort has been directed towards replacement of the platinum by materials mainly consisting of metal–nitrogen–carbon (MNC) network. The knowledge derived from tens of years of optimization of Pt-based PEMFC can not be directly translated to MNC catalytic systems due to different approaches in making catalyst layers.

The structure of the active site/sites of the PGM-free ORR electrocatalysts remains contentious even after 50 years of research. The structure of the active site within the Membrane Electrode Assembly (MEA) has not been studied at all. It is necessary to understand the effect of the interaction of ionomer and catalyst on the structure of the active site, morphology of catalyst layer and durability.

There is a missing link between durability and activity parameters as well as chemical and morphological changes that occur inside the catalyst layer during the oxygen reduction reaction in a fuel cell.

In this study, we are investigating a series of MNC electrocatalysts synthesized by the same sacrificial support route[1] and their performance in MEA. Chemistry of the electrocatalysts and catalyst layers is studied by XPS. The types of nitrogen and iron-nitrogen functionalities that are present in these materials are in-plane defects such as graphitic N and N-coordinated to three or four nitrogens and a multitude of possible edge sites such as pyridinic, pyrrolic, quaternary and Fe-N2/Fe-N sites. [2, 3] In the catalyst layers, the makeup of the active site may be affected by the interactions with negatively charged sulfonate group of nafion. Due to this interaction between groups on the catalyst surface and sulfonate groups of nafion, the rearrangement in high binding energy range has been observed. The higher relative amount of peaks between 401-403 eV in the catalyst layers is due to the shift in the position of other peaks due to interaction with an ionomer.

DFT calculations were used to evaluate the strength of the interaction between different types of nitrogen containing defects and sulfonate groups and to calculate binding energy shifts of N 1s spectra upon ionomer binding. [4] Figure 1 shows DFT calculated adsorption energies of sulfonate groups on nitrogen defects. High adsorption energies particularly for protonated nitrogens which may be either pyrrolic or protonated pyridine nitrogen is observed. Moreover, the higher amount of protonated nitrogens in catalyst layer results in worse MEA performance.

The pore structure is critical to the transport of oxygen to active sites and removal of water. The change in pore structure induced by the chemical changes introduced during fuel cell operation has to be understood in order to design PGM-free electrocatalyst with highest possible lifetime. The morphology of catalyst layers will be analyzed by focused ion beam/scanning electron microscopy (FIB-SEM) sectioning. This will allow to obtain a 3D visual representation of morphology of catalysts layers and to estimate in detail the evolution of structural parameters such as: specific surface area, total porosity, connectivity of pores and others as a result of degradation studies.

1. Serov, A., et al., Nano-structured non-platinum catalysts for automotive fuel cell application. Nano Energy, 2015. 16: p. 293-300.

2. Jia, Q., et al., Spectroscopic Insights into the Nature of Active Sites in Iron-Nitrogen-Carbon Electrocatalysts for Oxygen Reduction in Acid and the Redox Mechanisms. Nano Energy, 2016.

3. Artyushkova, K., et al., Chemistry of Multitudinous Active Sites for Oxygen Reduction Reaction in Transition Metal-Nitrogen-Carbon Electrocatalysts. Journal of Physical Chemistry C, 2015. 119(46): p. 25917-25928.

4. Kabir, S., et al., Binding energy shifts for nitrogen-containing graphene-based electrocatalysts – experiments and DFT calculations. Surface and Interface Analysis, 2016.

Figure 1

2598

, , , , , , , , and

In recent years, numerous Pt-based nanoparticle catalysts with very high oxygen reduction reaction activity have been developed, most of these based on transition metal alloys of Pt with or without post-synthesis de-alloying. As intrinsic electrocatalytic activities have been improved, the performance limitation of the cathode catalyst layers has shifted from activity to facile transport of reactants to catalytic sites. Electrode structure has thus become increasingly important in defining electrode performance. Effective transport of reactants through the thickness of the electrode (~10 µm) is highly dependent on the agglomerate structure and spatial distribution of the conducting phases and the pores. Ultra-small angle scattering and TEM analyses have shown that the agglomerate structure of electrodes based on representative high activity Pt alloy catalysts is different than that of Pt catalyst-based electrodes. In this presentation, we will compare the structure of Pt alloy-based electrodes and Pt electrodes, characterized by a variety of techniques, including ultra-small-angle X-ray scattering, TEM, and X-ray tomography. The information obtained from these techniques has been used as input to an electrode structural model. The model has been probed to analyze the transport processes of protons, electrons, oxygen, and water. The results from alloy and Pt-based electrodes fabricated from inks with different ionomer to carbon ratio, solvent, and ionomer equivalent weight have been examined to establish the empirical relationship between the microstructure and ink composition and the ultimate performance of the electrode layer.

2599

, , and

The high cost of platinum (Pt) is a challenging barrier to broad commercialization of proton exchange membrane fuel cells (PEMFCs). Over recent decades efforts have been made to decrease Pt loading by increasing its mass and specific activities through nanostructuring and alloying. One such approach is the nanostructured thin film (NSTF) electrode1, which has Pt or Pt alloy catalyst deposited onto high surface area organic whiskers and lacks the addition of ionomer binder. NSTF cathodes typically have a lower Pt loading than conventional carbon supported Pt (Pt/C) catalyst cathodes, but tend to lose performance in low relative humidity and cold, flooded conditions. Despite the lack of ionomer, the ion conductivity for NSTF and ionomer-free Pt black electrodes has been shown to be good under high relative humidity conditions.2,3 However, the apparent proton transport mechanism is not well understood. The high mobility of adsorbed proton or hydroxyl species and the diffuse charge in the electric double layer could account for high ion conductivity on the metal surface away from the membrane.4 Such mechanisms are complicated by their sensitivities to water activity and interfacial electric potentials.

The present work investigates the role of water on the ion conductivity and oxygen reduction reaction (ORR) activity of metal/water interfaces by measuring the dependence of accessible metal surface area on humidity and potential for two ionomer/binder-free metal surfaces: Pt and gold (Au). Pt is studied because of its relevance to PEMFCs and Au is selected as a model surface because of its close ideal behavior and oxide-free surface within the envelope of PEMFC cathode potentials. However, Au catalyst provides low ORR activity relative to Pt. A membrane electrode assembly (MEA) was fabricated by painting a layer of Pt black or Au micro powder (0.5 m2/g) ink on a gas diffusion layer (GDL). A commercial Pt gas diffusion electrode (GDE) was used as the anode. Cyclic voltammetry (CV) was conducted for different relative humidity conditions as shown for Au in Figure 1a. As the inlet gas relative humidity increases, the double layer and Au-OH redox currents increase, indicating increased electrochemical active surface area on the wetted Au. As Figure 1b shows, the increased active area also yields increased current density when operated the cell as an air/H2 fuel cell and polarization curves are measured. This indicates that water films on the Au surface play an essential role in ion conduction and that the active area made accessible by the water also facilitates the ORR.

References

(1) An, S. J., and Litster, S. (2013) In Situ, Ionic Conductivity Measurement of Ionomer/Binder-Free Pt Catalyst under Fuel Cell Operating Condition. ECS Trans.58, 831–839.

(2) McBreen, J. (1985) Voltammetric Studies of Electrodes in Contact with Ionomeric Membranes. J. Electrochem. Soc.132, 1112.

(3) Sinha, P. K., Gu, W., Kongkanand, A., and Thompson, E. (2011) Performance of Nano Structured Thin Film (NSTF) Electrodes under Partially-Humidified Conditions. J. Electrochem. Soc.158, B831.

(4) Zenyuk, I., and Litster, S. (2014) Modeling ion conduction and electrochemical reactions in water films on thin-film metal electrodes with application to low temperature fuel cells. Electrochim. Acta.

Figure 1

C-22 Membrane Fabrication and Testing II - Oct 5 2016 2:00PM

2600

This presentation describes the membrane development and transport study at PNNL for the redox flow batteries. Several topics will be covered in the talk including the development of the PTFE/SiO2 porous separator and the research on the correlation between the Nafion membrane microstructure and the vanadium redox flow battery performance. The impact of the equivalent weight on the microstructure of the membrane and thus the performance of VRBs is systematically studied. By tailoring the equivalent weight and membrane thickness, the new membrane contains optimal pore geometry with extremely low vanadium-ion permeability was developed. In addition, its area resistance is comparable to, and its cost is significantly less than, the widely used Nafion 115 membrane. Excellent VRB single-cell performance (89.3% energy efficiency at 50mAbold dotcm-2) was achieved along with a stable cyclical capacity over prolonged cycling (>200 cycles). The chemical composition (Figure 1), pore geometry, transport properties and flow cell performance will be reported.

In addition, a detailed study on the ions transport behavior across the membrane and its impact on the redox flow battery performance will be presented, using vanadium redox flow battery as an example. In operando study on the vanadium ion crossover unveiled the mechanisms contributing to the battery capacity fading, as well as the state-of-charge changes along cycling. The studies were carried out on various membranes, such as cation exchange membrane and porous separators.

Figure 1.19F MAS-NMR spectra measured at an 11.7 T magnetic field with a spinning speed of 13.5 kHz for NDM142, NDM220, and NDM221 membranes.  The circles and red line represents the experimental spectra and the fitted curve with deconvolution analysis, respectively (see text for details).  The asterisk (*) indicates MAS spinning side bands.

Figure 1

2601

, , and

Polymer electrolyte membrane (PEM) as one of the most important components of fuel cells should satisfy the requirements such as high chemical and thermal stability, high proton conductivity, impermeability against the fuel, and good mechanical properties. In order to achieve the requirements, charged-transfer (CT) complex method was developed and applied to improve the properties of sulfonated polyimide (SPI) for high temperature fuel cell (HT-PEFC) application. The developed aromatic SPI CT films showed high proton conductivity and charged-transfer could control several properties of the films. The obtained aromatic SPI CT films, however, had low flexibility to be applied in a real MEA test. In order to solve this technical problem, an aliphatic monomer consisting of six carbons was added to SPI main chain to improve the flexibility and mechanical strength of the membrane. Aliphatic SPIs were synthesized by polycondensation of 4,4-diamino-2,2-byphenyldisulfanoic acid (DAPS), 1,4,5,8-napthalene tetracarboxilicdianhydride (NTCD), and 1,6-hexanediamine as an aliphatic monomer, at 180 ˚C. The composition and incorporation of aliphatic monomer to all monomer units was 20% and it was confirmed by 1H NMR and FTIR. The obtained aliphatic SPI contained electron-accepting unit (NTCD) in the main chain. To create CT complex in the membranes the electron-donating molecule, 2,6-dihydroylnaphtalene (DHN), was introduced to the aliphatic SPI DMSO solution in some molar ratios.  The mixture DMSO solutions were cast on the glass dishes, and DMSO was evaporated at 60 ˚C in vacuo. The obtained films showed darker color than the intact aliphatic SPI film indicating that CT complex was formed between electron-accepting unit (NTCD) in the aliphatic SPI and the electron-donating molecule, 2,6-dihydroylnaphtalene (DHN). Moreover, the further confirmation of CT complex formation in the film was carried out by visible spectroscopy, and the CT absorption peak at 520~530 nm was observed. This result proved that the electron-donating DHN and the electron-accepting NTCD could form CT complex in the films.

     The introduction of electron donor DHN to the aliphatic SPI resulted in reduced water uptake with increasing donor molecule molar ratio. It proved that CT complex suppress water uptake of the aliphatic SPI CT films. In addition, the aliphatic SPI CT films with different DHN molar ratio exhibited fairly small variation in proton conductivity, which spanned a range of 85 to 140 mS cm-1, measured at 120 ˚C and 100 % relative humidity. The obtained proton conductivities at elevated temperature (120 ˚C) and high relative humidity (100%) were up to 1.5 times higher than that of Nafion 212 . Nevertheless, the obtained results of the proton conductivity indicated a significant dependence on relative humidity. On the other hand, the thermal and mechanical properties of the aliphatic SPI CT films were investigated by TGA, DSC, and tensile measurements. The aliphatic SPI CT films showed high glass transition temperatures (Tg),  ~160 °C with higher tensile modulus, and elongation at break than the aromatic SPI CT films. XRD of the aliphatic SPI films shows a broad peak centered around 22.5 ~ 23.1˚ 2θ. And the addition of donor molecule did not affect to the distance between polymer main chains significantly. The aliphatic SPI CT membrane had been demonstrated to be promising candidates for high temperature PEM in fuel cell applications.

2602

and

Polymer electrolyte fuel cells (PEMFCs) have been spotlighted as one of the promising eco-friendly energy technologies for stationary and automotive applications owing to zero CO2 emission, high energy density and moderate operation conditions. In this technology sector, polymer electrolyte membrane, one of the key components of PEMFC, has been intensely studied for several decades. Conventionally,perfluorinated sulfonic acid (PFSA) membranes like Nafion are used due to their high proton conductivity and mechanical stability. However, their high cost has been pointed out as a significant drawback interrupting mass commercialization of fuel cell electric vehicle. In this regard, hydrocarbon (HC) membranes, as cheaper alternatives, have been intensively studied in replacing PFSA membrane.

However, until now, the challenge in adopting cost-effective HC membrane for PEMFCs has been the poor interfacial adhesion between catalyst layers (CLs) and HC membrane, which causes the membrane to delaminate easily, losing efficiency with use. Here, we present scalable mechanical nano-faster featured by three-dimensional interlocked interfacial structure between HC membrane and PFSA-based CL as a novel strategy to tackle the interfacial issue. It is realized by forming nano-porous skins on the both side of HC membrane and successively fiiling the pores with PFSA ionomer with scalable wet coating methods. The interlocking interface tightly binds the HC membrane and CL owing to its highly-interlocked ball and socket joint structure. The interfacial adhesion is dramatically enhanced by 37-fold with the nano-fastener. The membrane electrode assembly (MEA) with the three-dimensional interlocking interface exhibits 17 times higher durability than that with flat interface, paving a way to realize highly robust and cost-effective HC membrane-based PEMFCs for automotive use.

2603

, , , , , and

Fuel cells are the key technology towards establishing a hydrogen energy based society, as they convert the chemical energy of hydrogen into electricity with high efficiency, and without emitting harmful exhaust gases. Polymer electrolyte membrane fuel cells (PEMFCs) are already successfully commercialized and are being used in e.g. fuel cell vehicles such as the Toyota MIRAI and the Honda Clarity Fuel Cell. However, the aim for next-generation PEMFCs is to operate at temperatures above the boiling point of water, therefore new membrane materials are needed. Stable operation at high temperature, sufficient high proton conductivity, high gas barrier as well as low cost are essential requirements.

Cellulose is the most abundant biopolymer on earth, and has been utilized in various forms by mankind for thousands of years. Cellulose polymer units form nanofibrils during biosynthesis, with diameters of several tens of nanometers and lengths of up to a few microns. These group together to form micro- and macrofibrils. In 1977 nanofibrils were successfully observed for the first time by Turbak et al., kick-starting the field of nanocellulose science. Paper made from nanocellulose has intriguing properties such as high tensile strength (214 MPa), high gas barrier and thermal stability up to 150°C.1–3 During the fabrication process of nanocellulose, acid group functionalities (i.e. carboxyl and sulfonic acid groups) are introduced,4,5which increase hydrophilicity, and additionally act as proton donors and acceptors, thus enabling proton conduction.

We investigated nanocellulose for its suitability as fuel cell membrane in PEMFCs for the first time. Two different nanocellulose materials were characterized for their morphology and chemical composition, as well as mechanical properties and gas barrier. Nanocellulose membranes showed four times higher tensile strength compared to the industry standard membrane Nafion and a roughly three orders of magnitude lower hydrogen permeability. A maximum proton conductivity of 4.7 mS/cm at 120°C and 100% RH was observed. For the first time, nanocellulose membranes were assembled into membrane electrode assemblies and operated as hydrogen fuel cell. We will report latest results of our research.

References:

1. Henriksson, M., Berglund, L. A., Isaksson, P., Lindstro, T. & Nishino, T. Cellulose Nanopaper Structures of High Toughness. Biomacromolecules9, 1579–1585 (2008).

2. Nair, S. S., Zhu, J., Deng, Y. & Ragauskas, A. J. High performance green barriers based on nanocellulose. Sustain. Chem. Process.2, 1–7 (2014).

3. Nogi, M. et al. High thermal stability of optical transparency in cellulose nanofiber paper. Appl. Phys. Lett.102, 181911 (2013).

4. Lin, N. & Dufresne, A. Nanocellulose in biomedicine: Current status and future prospect. Eur. Polym. J.59, 302–325 (2014).

5. Barbosa, L. et al. A Rapid Method for Quantification of Carboxyl Groups in Cellulose Pulp. BioResources8, 1043–1054 (2013).

Figure 1

2604

, , and

Perfluorosulfonic acid (PFSA) polymers are the most promising state-of-the-art materials as proton exchange membranes (PEM) for fuel cells, however, there are some drawbacks such as high production cost and hydrothermal stability. Aromatic membranes based on hydrocarbon polymers have been studied as alternatives. Among them, multi-block copolymers seem to be promising to compete with PFSA resulting high proton conductivity even under low RH conditions.[1, 2]

In this report, novel sulfonated poly(arylene ether sulfone)s multi-block copolymer membranes containing highly sulfonated hydrophilic blocks were synthesized.[3] Different local concentration of sulfonic acid in their hydrophilic blocks affected chemical and physical properties of the SPAES. To investigate the effects of chemical composition on their membrane properties, different hydrophilic oligomers sharing same hydrophobic blocks gave us exact comparison of effect of hydrophilic blocks. The higher concentration of sulfonic acid groups resulted in higher proton conductivity under certain relative humidity conditions than that of the state-of-the-art perfluorinated sulfonic acid membrane and showed that the well-developed phase separation of SPAES. Moreover, physical properties of theses SPAES including water behavior, humidity dependence of proton conductivity were investigated along with morphology characterizations by transmission electron microscopy (TEM) for PEMFC application. Two types of oligomers, F-terminated and OH-terminated telechelic oligomers, were synthesized by controlling the feed ratio of dihydroxyl- and difluoro-monomers. Their number of repeating unit (X and Y) was analyzed by GPC and 1H NMR. Copolymerization with F-terminated and OH-terminated telechelic oligomers via nucleophilic aromatic substitution, gave high-molecular-weight multi-block PESs. Each block length was controlled to have different values with X5Y10, X10Y10, X20Y10 and X20Y20.

The SPAES X10Y10 membrane showed highest proton conductivity than that of our previous random and block copolymers at wide range of humidity. The best balanced membrane was SPAES X10Y5 membrane with high proton conductivity and lower water uptake. Consequently, the SPAES X10Y5 membrane showed high cell performance under various conditions (510 mA/cm2, 290 mA/cm2 @ 0.6V under 80% and 50 % RH conditions at 80℃, respectively). Systematic approaches for developing alternative block SPES membrane for low RH condition will be discussed.

 References

[1] H.-S. Lee, A. Roy, O. Lane, J.E. McGrath, Polymer, 49 (2008) 5387-5396.

[2] B. Bae, T. Yoda, K. Miyatake, H. Uchida, M. Watanabe, Angew. Chem. Int. Ed., 49 (2010) 317-320.

[3] S. Lee, J. Ann, H. Lee, J.-H. Kim, C.-S. Kim, T.-H. Yang, B. Bae, J. Mater. Chem. A, 3 (2015) 1833-1836.

2605

, , , , , and

The 1,2,3-triazole functionality demonstrates high chemical stability towards severe hydrolytic, oxidizing, and reducing conditions, even at elevated temperatures. Furthermore together with aromatic character and ability to form hydrogen bonds, it is reported to possess anti-oxidant properties1. All these features make polymers functionalized with 1,2,3-triazole of interest in proton exchange membranes as reinforcement or matrix for application in low and high temperature fuel cells.

We will describe a fast and high-yielding synthesis of 1,2,3-triazole functionalized polysulfone through microwave-assisted copper-catalyzed azide-alkyne cycloaddition (CuAAC) click chemistry. Polymers with various degrees of substitution of the side-chain functionality of 1,4-substituted 1,2,3-triazole with alkyl and aryl pendant structures were prepared and characterized to assess the completeness of the reaction. The thermal and thermo-chemical properties of the modified polymers were also characterized.

In one approach, cast membranes of 1,2,3-triazole functionalized polysulfone were doped with phosphoric acid for application in HT-PEMFC and their properties compared with conventional polybenzimidazole based systems. These novel membranes presented promising performance with high proton conductivity in anhydrous conditions and satisfactorily high elastic modulus2.

In another strategy with targeted low temperature PEMFC applications, nanocomposite membranes were prepared based on functionalized polysulfone and a perfluorosulfonic acid ionomer. As already evidenced for PBI-PFSA composites3, the specific interaction between the basic triazole functionalized polymer and the acidic ionomer lead to electrolytes with improved mechanical properties and without significant loss in proton conductivity.

1. J. Totobenazara and A. J. Burke, Tetrahedron Lett., 56, 2853–2859 (2015).

2. R. Sood, A. Donnadio, S. Giancola, A. Kreisz, D.J. Jones, S. Cavaliere, ACS Appl. Mater. Interf., in press.

2606

, , and

Sterically-encumbered, sulfonated poly(phenylene)s are an interesting group of hydrocarbon materials for use in low Pt-content catalyst layers because they potentially offer high conductivity, high thermo-oxidative stability, and appear less-prone to strong adsorption on Pt – a process that typically reduces the electroactive surface area. Frequently, polymers like poly(phenylene)s are randomly post-functionalized in order to introduce proton-conducting properties.(1-3) However, as randomly-functionalized materials are not ideal as proton conductors, a new synthetic route has been developed in order to obtain well-defined pre-functionalized monomers for the synthesis of sulfonated poly(phenylene)s. Using a variety of NMR pulse sequences, the isomers resulting from different couplings (e.g., meta-meta) of arylene moieties has been studied through an examination of several small molecule, model compounds. Furthermore, a series of random copolymers has been made in which the ionic content was varied by alteration of the monomer feed ratios. Finally, the properties (e.g., proton conductivity, chemical stability, fuel cell performance as membrane and catalyst ionomer) have been extensively studied.(4)

(1) Maier, G.; Meier-Haack, J. Advances in Polymer Science2008, 216, 1

(2) Otsuki, T.; Kanaoka, N.; Iguchi, M.; Mitsuta, N.; Soma, H.; (Honda Motor Co., Ltd., Japan; JSR Corporation). Application: US, 2004, 13

(3) Fujimoto, C. H.; Hickner, M. A.; Cornelius, C. J.; Loy, D. A. Macromolecules2005, 38, 5010

(4) T. J. G. Skalski, B. Britton, T. J. Peckham and S. Holdcroft, Journal of the American Chemical Society, 2015, 137, 12223

Figure 1

2607

, and

While perfluorinated sulfonic acid (PFSA) membrane and ionomer materials such as Nafion® form the standard for high-performance proton-exchange membrane fuel cells (PEMFCs), the limited and difficult chemistry of perfluorinated materials hampers further material development to extended fuel cell performances and lifetimes, while the high cost of perfluorinated materials contributes to the cost barrier limiting the ubiquity adoption of fuel cells for energy generation. One method of reducing cost is the US Department of Energy target of 0.0625-0.125 mg PGM/cm2 as a catalyst loading. However, at low to ultra-low catalyst loadings, PFSAs exhibit a substantial increase in resistivity attributed to oxygen mass transport losses, particularly in the far Ohmic region of polarization data, causing disproportionate reductions to achievable power densities.

Hydrocarbon proton-exchange materials offer the potential for solutions to material issues inherent to PFSAs, with established, varied chemistry allowing for versatile material innovation. Properties such as lower fuel crossover and improved operation in desirable operational conditions such as reduced relative humidity (RH) or high temperature has been well established. Together with the lower cost of basic materials and the ability to recycle catalyst layers, these properties make the demonstration of high performance, fully hydrocarbon PEMFC operation significantly of interest to the field.

Here, hydrocarbon proton-exchange materials are shown that exhibit significant radical stability and high conductivities. Furthermore, these materials form thin membranes and are soluble in low-boiling, polar solvents necessary for incorporation as ionomer into catalyst inks that create high-quality catalyst layers. MEAs and catalyst layers were formed, incorporated into fuel cells, and characterized in situ by IV polarization, CV, CA, LSV, and EIS, and ex situ by mercury porosimetry and SEM. As ionomers, these exhibit reduced mass transport losses compared to PFSA ionomers, a result of substantially smaller increases to oxygen mass transport losses than PFSAs exhibit as catalyst layer loadings are decreased. As fully hydrocarbon fuel cells, improved interfaces lead to improved water transport and lower total resistivity attributable to membrane and ionomer congruency, thereby achieving higher power densities than directly comparable PFSA references using rigorously optimized conditions and catalyst loadings. These high-performance hydrocarbon proton-exchange materials may thereby represent a fully hydrocarbon alternative to PFSA materials for hydrogen fuel cells.

2608

, and

Toray had been developing highly-reliable hydrocarbon-based (HC) membrane for PEFC in NEDO FY2013-2014 Next Generation Project. Newly-developed HC membrane using newly-developed organic peroxide decomposition catalysts showed higher cell performance under high temperature and low humidity condition, and more than 5.2 times higher chemical durability, in comparison with the reinforced PFSA membrane NafionHPas a reference membrane. Consequently, newly-developed HCmembrane accomplished the FY2014 target of NEDO Project on both cell performance and chemical/mechanical durability.

Acknowledgement: This work was partially supported by the New Energy and Industrial Technology Development Organization (NEDO) of Japan.

2609

, and

One of the most promising technologies for power generation are Proton Exchange Membrane Fuel Cells (PEMFCs). PEMFCs are electrochemical devices which convert directly the chemical energy of oxidation-reduction reactions into electrical energy. These electrochemical devices have high electric efficiency and low environmental impact [1].

Several authors are working in the upcoming commercialization of this technology. However, there are technical aspects that must be overcoming before their implementation. The proton exchange membrane, which is the key component of a PEMFC, needs a considerable improvement regarding its performance without external humidification [2]. This is of great importance because it allows working at temperatures above 80 ºC and eliminate the management of water.

Concerning this aspect, many authors are working incorporating ionic liquids as proton exchange components in PEMFCs. Ionic liquids, which are in liquid state at room temperature, have attractive properties for this application, such as thermal and electrochemical stability and high anhydrous conductivity.

There are several strategies in order to incorporate ionic liquids inside the proton exchange membrane. One of them deals with the polymerization of an ionic liquid monomer [3]. This option avoids the leak of the ionic component during the fuel cell operation.

In this work, membranes based on polymeric ionic liquids have been designed for their use as electrolytes without external humidification. For this purpose, the ionic liquids 1-(4-sulphobutyl)-3-vinylimidazolium trifluoromethanesulphonate [HSO3-BVIM][OTF], 1-sulfobutyl-3-metylimidazolium 2-sulfoethylmethacrylate [SBMIm][SEM] and 1-sulfobutyl-3-vinylimidazolium 2-sulfoethylmethacrylate [SBVIm][SEM] were polymerized under ultraviolet light. The ionic conductivity and fuel cell performance were tested without external humidification.

References

[1] Z. Wojnarowska, J. Knapik, M. Díaz, A. Ortiz, I. Ortiz, M. Paluch, Macromolecules, 2014, 14, 4056-4065.

[2] M. Díaz, A. Ortiz, I. Ortiz, J. membra. Sci., 2014, 469, 379-396

[3] D. Mecerreyes, Prog. Polym. Sci., 2011, 36, 1629-1648.

 

Acknowledgements

This research was supported by the Ministry of Education and Science under the project CTQ2012-31639 (MINECO, SPAIN-FEDER 2001-2013). Mariana Díaz also thanks MINECO for the FPU fellowship (AP2012-3721).

2610

, , , and

Current fuel cells for transportation and residential applications use polymer electrolyte membrane fuel cells (PEMFCs) which require platinum catalysts and operate below 80 °C due to water management issues. However, PEMFCs operating at 100 °C or higher temperatures in anhydrous conditions do not need a humidifier, and do not suffer from water management issues created by the presence of liquid water. Excessive amount of liquid water in the electrode can flood the backing and catalysts layers at lower temperatures leading to poorer fuel cell performance. Furthermore, such intermediate temperature PEMFCs can solve the problem of carbon monoxide (CO) poisoning of the anode electrocatalyst leading to relaxation of the hydrogen fuel quality standards, thus lowering infrastructure costs. Finally, higher operating temperatures allow for simpler cooling systems in transportation applications, which results in smaller and lighter radiators. These advantages mean that a paradigm shift in fuel cells for transportation and residential applications can be achieved with an intermediate temperature proton conducting solid-electrolyte that can operate above 150 °C.

Unfortunately, electrolyte membranes and ionomers developed so far do not have satisfactory conductivity and performance in the above temperature ranges, at low-humidity, or in non-humidified conditions. The further advancement of higher temperature PEMFCs relies on the development of membranes that possess the desired electrochemical and mechanical properties in dry (< 30% RH) and hot (> 100 oC) conditions.

In this study, we have developed an organic/inorganic composite membrane using a SnP2O7 (TPP) inorganic conductor and evaluated the fuel cell performance under various conditions. Membrane electrode assemblies (MEA) were prepared from these membranes and gas diffusion electrodes (GDE) were optimized using either LANL ionomer dispersion or various commercial gas diffusion layer (GDL). A maximum power density (> 600mW/cm2 ) has been obtained for fuel cell operation at ≈ 220 oC under less than 2 % humidification condition.

I01-C Poster Session - Oct 5 2016 6:00PM

2611

, and

Degradation of the polymer membrane is known as one of the important factors to determine the lifetime of polymer electrolyte fuel cells (PEFCs). The membrane electrode assembly of a PEFC is made from various components, e.g. the catalyst layers, the gas diffusion layers, the micro-porous layers and the polymer electrolyte membrane. Generally, the lifetime of a PEFC is specified by the degradation of a polymer electrolyte membrane. Degradation of the membrane can be classified into three categories: thermal, mechanical, and chemical. For Nafion membranes which are often used in PEFCs, the chemical degradation is more significant because Nafion membranes have excellent mechanical and thermal stability. Since the chemical degradation, which is mainly due to the attack of radicals formed during the operation, is inevitable, the knowledge to estimate the performance of the degraded membrane is required. Proton conductivity is one of the typical indicators for chemical degradation of the polymer electrolyte membranes. The proton conductivity in a degraded Nafion membrane becomes lower than the one in the pristine Nafion membrane. The transport properties in the membranes strongly influenced by the molecular morphology of the multicomponent system which includes polymers, water/hydronium molecules, and several fragments generated by the degradation. Although the degraded Nafion membranes had been investigated experimentally and numerically, evaluation of the relationship between transport properties and degradation levels, taking into account several fragments generated by the degradation, had not been considered. In addition, it is difficult to control the degradation experimentally because the degradation occurs in molecular scale.

In the present study, we have carried out molecular dynamics simulations of a degraded Nafion membrane system. The effect of a degradation mechanism on transport properties, which is identified by what kind of radical react on, is important because the combination of fragments is considered to determine the water structure. However the degradation mechanism is still under discussion. Here, a proposed mechanism by Ghassemzadeh et al.1 have been adopted to create a degraded Nafion and generated fragments. In their mechanism, the attack of the hydroxyl radical, which is one of the most reactive species in a degraded Nafion membrane, causes unzipping of side chain. Consequently, HF, CO2, CF3, and HOC2F4SO3 molecules are generated in the reaction. A modified DREIDING force field2-4 have been employed for degraded Nafion molecules and fragments. Water/ hydronium molecule interacts with anharmonic potential5 to describe Grothus mechanism with aTS-EVB model6. For side chain molecules of a degraded Nafion monomer and fragment molecules, the equilibrium bond lengths, the equilibrium angles and the partial charges have been determined by DFT calculations. In the simulation systems, there were four ten-unit degraded Nafion molecules, whose degradation level is defined by changing the number of broken side chains, equimolar generated fragments, and water/hydronium molecules with varying water contents. We have conducted simulations in various conditions, such as different water contents and degradation levels. Diffusion coefficient of each molecule have been used as their transport properties. To evaluate molecular structures, radial distribution functions have been calculated in various degradation levels. We have evaluated the relationship between diffusion coefficient and degradation levels.

References

1. L. Ghassemzadeh et al., J. Am. Chem. Soc., 135, 15923 (2013).

2. S. L. Mayo et al., J. Phys. Chem., 94, 8897 (1990).

3. S. S. Jang et al., Macromolecules, 36, 5331 (2003).

4. S. S. Jang et al., J. Phys Chem. B, 108, 3149 (2004).

5. K. Park et al., J. Phys Chem. B, 116, 343 (2012).

6. T. Mabuchi et al., J. Chem. Phys., 143, 014501 (2015).

2612

, and

Significant advances in the performance of polymer-electrolyte fuel cells (PEFCs) can be attributed to perfluorinated sulfonic-acid (PFSA) ionomers, like Nafion and 3M-ionomer. They serve both as electrolyte membranes and as conductive binders in porous catalyst-layer structures. In PEFC catalyst structures, ionomers are thought to be major contributors to unexplained large mass-transfer overpotentials.[1-3] Thin (< 1 µm) and ultra-thin (< 0.025 µm) ionomer films in the catalyst layer serve as ionic charge carriers to catalytic sites, where low resistance to gas transport is required for optimal performance, yet it is believed that such resistance is increased compared to bulk ionomer. This resistance is further exasperated with decrease in platinum loading, which is an essential consideration for commercial viability and cost reduction of PEFCs.[4] Much has been done to shed light on the aberration of ionomer thin-film from bulk in other properties such as proton conductivity, morphology and water uptake. This has allowed many to attribute the rise in local mass-transfer resistance at the catalyst layer to the decrease in permeability of thin-film ionomers as compared to that of bulk membrane. However, direct measurement of permeability of thin-film ionomer has yet to be accomplished.

A significant amount of work has been done to study gas transport and permeation in the bulk ionomer membrane.[5] However, much remains to be explored in thin and ultra-thin ionomer film thickness regime. Effects such as confinement, processing conditions, and surface and interfacial interactions have been pointed to as the prime cause of deviation of thin-film polymer behavior from that of the bulk. Linking the extent of impact of these effects to ionomer properties can elucidate source of transport resistance in the catalyst layer. In this work, we will present data for gas permeation in thin and ultra-thin PFSA films including free-standing thin ionomer films (> 400nm) as well as thin and ultra-thin ionomer films supported on well studied, highly permeable rubbery poly(dimethylsiloxane) (PDMS). O2, N2 and H2 permeability of the ionomer film is measured using a constant-volume, variable-pressure permeation system. Experimental results demonstrate dependence of gas permeation on thickness, and influence of physical aging on gas transport.

 

Acknowledgement

We would like to thank Norman Su for providing assistance in permeation system assembly and operation. This work made use of facilities at the Biomolecular Nanotechnology Center at University of California, Berkeley and the Molecular Foundry at Lawrence Berkeley National Laboratory. This work was funded in part by the University of California Chancellor's Graduate Fellowship, National Science Foundation Graduate Fellowship and the Fuel Cell Technologies Office, Office of Energy Efficiency and Renewable Energy, U. S. Department of Energy under contract number DE-AC02-05CH11231.

References

[1]Nonoyama, N. et al (2011). Journal of Electrochemical Society. 158(4)

[2]A. Kongkanand, M. F. Mathias, The Journal of Physical Chemistry Letters, 7, 1127−1137(2016)

[3]Gresler, T. et al (2012). ). Journal of Electrochemical Society. 159(12)

[4]Weber, A. et al (2012). Journal of Materials Chemistry A. 2,17207

[5] Kocha, S. et al (2006).AIChE Journal. 52(5)

2613

, , , and

For decades, in polymer electrolyte membrane fuel cell (PEMFC) field, hydrocarbon (HC) membranes have been constantly studied to replace perfluorinated sulfonic acid membranes owing to their low cost and high fuel efficiency. However, adopting HC membranes to practical PEMFCs has been not successful due to their poor mechanical stability which causes a mechanical failure with generating pin-hole under repeated volume expansion/shrinkage during cell operation. Conventionally, the problem has been addressed by inserting a porous mechanical supporter in HC membrane which is denoted as 'internal reinforcement'. However, the introduction of the inert support decreases proton conduction and increases membrane cost.

Here we present an external reinforcement of HC membrane as a new strategy to enhance mechanical durability of HC membrane. It features the incorporation of mechanically tough porous fibrous network into catalyst layers and the strong connection of HC membrane and the toughened catalyst layers with introducing an interlocking interface. The mechanically toughened catalyst layers and interfaces can effectively mitigate the volume change of HC membrane, lowing the membrane failure. Under an accelerated humidity cycling test, the externally reinforced HC membrane exhibits an enhanced durability compared to un-reinforced counterpart. Furthermore, contrary to the internal reinforcement, the external reinforcement strategy does not cause any loss of proton conductivity of HC membrane. Therefore, the external reinforcement coupled with toughed catalyst layers and interfaces can provide an effective way to enhance durability with preserving the proton conductivity of pristine HC membrane.

2614

, , and

Polymer electrolyte membranes (PEMs) are important components for PEFC and are required the several functions such as high proton conductivity, film durability, and so on. Current strategy of the molecular design for PEMs is mainly the formation of phase-separated structure like Nafion, and many hydrocarbon PEMs which were composed of block copolymers have been developed1,2. These block copolymer PEMs showed high proton conductivities and enough film durability and are the promising alternative PEMs to Nafion. On the other hand, PEMs with no phase separation were not applied practically. One of the main reasons is that highly proton-conductive homogeneous PEMs are easy to dissolve in water because large amount of sulfonic acids as a proton conductor are required. If, however, we can develop stable PEMs without any phase separation, all surfaces of PEMs can be used for proton conductivity. In order to develop stable PEMs without any phase separation, we have developed charge-transfer (CT) complex hybrid films which are composed of an electron-accepting sulfonated polyimide (SPI) and electron-donating additives3,4. Functionality of CT films such as mechanical strength and optical property can be expanded widely by using functional donor molecules. Therefore, we applied the CT films as stable PEMs without phase separation. In order to reinforce the CT films for the more practical PEMs, we developed new semi-interpenetrating polymer network (IPN) films by cross-linking reaction of the cross-linkable donor molecules. Various properties such as proton conductivity of the obtained films are reported.

IPN films were prepared from electron-withdrawing SPI and electron-donating 2,6-dihydroxynaphthalene (2,6-DHN) with cross-linkers for 2,6-DHN. As cross-linkers, isocyanate, carboxylic acid chloride which can react with hydroxyl group of 2,6-DHN and 2,6-DHN with vinyl group were used. The IPN films were prepared by casting method. The DMSO solution of SPI and DHN was concentrated in vacuo, and the cross-linkers were added to the concentrated mixture and mixed in the glove box. DMSO was removed at 60 ˚C under the vacuum condition.

Some of the obtained films showed dark brown color, indicating that CT complex between SPI and 2,6-DHN was formed in the films. Especially, adipoyl chloride (AC) as a typical carboxylic acid chloride, showed the proper film shape and mechanical stability. Therefore, IPN films consisting of AC were used for the further characterization. To confirm the CT formation in the films, visible spectroscopy of the obtained films was carried out. The maximum wavelength of the CT films with and without AC were 552 nm and 529 nm, respectively. The intact SPI film did not show any peaks in this region. This result indicated that CT film with AC would have different molecular structure from CT films without AC. To determine the molecular weight change after the cross-linking reaction, GPC measurement was carried out. The molecular weight of the IPN films were larger than that of the intact SPI, indicating that some of the sulfonic acid reacted with AC and the molecular weight after cross-linking would increase. Mechanical strength of the IPN films was also evaluated. The obtained results indicated that Young's modulus tended to increase with the increase of AC, and stress and strain at break tends to decrease with the increase of AC. This result indicated that IPN introduced the film strength compared with the CT films without AC. The proton conductivity measurement of the obtained IPN films was also carried out. The proton conductivity of the obtained IPN film (SPI : DHN : AC = 100 : 25 : 25) was 80 mS/cm at 70˚C, 80%RH. The IPN films showed the slightly higher proton conductivity of the CT films without AC, indicating that enough sulfonic acid in SPI remained after cross-linking reaction and the carboxylic acid generated from AC would also work as a proton conductor.

Reference

  • K. Yamazaki, H. Kawakami, J. Membr. Sci., 43, 7185 (2010).

  • B. Bae, K. Miyatake, M. Watanabe, Macromolecules, 43, 2684 (2010).

  • R. Watari, M. Nishihara, H. Tajiri, H. Otsuka, A. Takahara, Polymer J., 45, 839 (2013).

  • L. Christiani, S. Hilaire, K. Sasaki, M. Nishihara, J. Polym. Sci. A, Polym. Chem. 52, 2991 (2014). 

Figure 1

2615

, and

Cost reduction and performance improvements are still major issues for polymer electrolyte fuel cells (PEFCs). To solve these problems, the relationship between internal transport phenomena and performance in PEFCs are an essential factor.

 Particularly, the mass transport phenomena in the catalyst layer comprehends several unclear parts, which is desired to clarify. Evaluation of membrane electrode assemblies (MEAs) with different kind of ionomer materials under operating conditions are effective to clarify the impact of ionomer proton conductivity and gas diffusivity in the performance. However, these experiments are difficult due to less variety of ionomer materials.

We have already introduced poly (p-phenylene) based diblock hydrocarbon polymers has comparable properties to perfluorinated polymer electrolytes. Moreover, these materials can be used as model materials because their compositions and structures can easily be controlled. [1, 2] In this study, we synthesized several hydrocarbon polymer electrolytes for cathode ionomer, and the effect on the fuel cell performance was investigated.

The catalyst coated membranes (CCMs) were prepared for this study. Fig.1 shows the structure of hydrocarbon polymer electrolyte which was utilized for the cathode ionomer. It was synthesized by catalyst transfer polycondensation of the hydrophobic and hydrophilic monomers. The synthesized polymer electrolyte was dissolved in tetrahydrofuran (THF). Platinum supported on carbon black (Pt/C) catalyst (TEC10E50E, Tanaka Kikinzoku) and pure water were added to the solution, and the solution was mixed by a bead mill (BSG02, AIMEX). Nafion was utilized for the anode ionomer. Pt/C, isopropyl alcohol, and water were added to the Nafion solution (D2020, DuPont), and mixed by a bead mill. These catalyst inks were applied to polymer electrolyte membranes (NRE211, DuPont) by a pulse spray system (V5R02, Nordson).

A prepared CCM was installed in a single cell with two gas diffusion layers (SGL25BC, SGL). The single cell consists of separators with flow channels (1mm depth, 1mm width and 1mm rib), current correctors, and end-plates.

Electrochemical impedance spectroscopy (EIS) was performed by a potentio/galvanostat (HZ-7000, Hokutodenko). The cell temperature was remained at 80ºC, and 0.5 L/min hydrogen and 1.0L/min nitrogen were fed into the single cell. Both gases were humidified at 100%RH.

The cathode proton conductivity of the CCM with the hydrocarbon ionomer was close value of the CCM with Nafion ionomer. This result shows the hydrocarbon ionomer work well under practical operating conditions.

Details of their electrochemical activities and performance evaluation results are presented at the symposium.

Reference

[1] Umezawa, K., Oshima, T., Yoshizawa-Fujita, M., Takeoka, Y., & Rikukawa, M. (2012). ACS Macro Letters1 (8), 969-972.

[2] Takeoka, Y., Umezawa, K., Oshima, T., Yoshida, M., Yoshizawa-Fujita, M., & Rikukawa, M. (2014). Polymer Chemistry5 (13), 4132-4140.

Figure 1

2616

and

In this paper, proton exchange membranes are presented on the basis of multiblock co-ionomers. The multiblock-co-ionomers are prepared from hydrophobic, partially fluorinated polymer [1] and hydrophilic, sulfonated ionomer blocks [2]. The synthesis of the individual polymer blocks is done by nucleophilic aromatic step-growth polycondensation. The multiblock-co-ionomer membranes are cast and dried according to standard procedure. The interest in multiblock-co-ionomer membranes for use in PEMFCs is that they form microphase- or nanophase-separated regions [3]. The areas with hydrophobic character ensure the stability of the membrane and the areas with hydrophilic character of the proton conduction. In case of longer hydrophilic blocks it is necessary to blend the multiblock-co-ionomers with basic polybenzimidazoles like PBI-OO or F6-PBI for mechanical stability.

In this contribution the first results in terms of characterization by means of NMR spectroscopy, molecular weight distribution, thermal stability and proton conductivity will be presented.

Literature:

1: Kerres. J. A.; Xing, D.; Schöneberger, F. J.Polym. Sci. PART B: Polym. Phys. 2006, 44, 2311-2326

2: Kerres. J. A.; Schöneberger, F. J.Polym. Sci. PART A: Polym. Chem. 2007, 45, 5237-5255

3: Kerres, J. A. Polymer Reviews2015, 55, 273-306

2617

, and

Polymer electrolyte fuel cells (PEFCs) have attracted attention because of their high efficiency, low environmental load and small application size. Development of novel proton exchange membranes (PEMs) with high proton conductivity under low humidity is strongly desired to enhance a performance of PEFCs. In this context, precedence research based on quantum chemical calculations shows that effective proton transfer through the counter anion (e.g. -SO3 in the case of sulphonic acid-based PEMs) occurs by increasing their density1. From this viewpoint, we focused on heterocyclic ring systems such as benzothiaziazole (BT) and thiazolothiazole (TT) units. These units promote a hydrophobic structure owing to its planar conformation and strong intermolecular interactions (S-N or S-S interactions)2,3. Therefore, we first designed novel aromatic polymers based on BT units with the expectation that the BT units would enhance to suppress membrane swelling and increase the density of the sulphonic acid groups (Figure 1-a, b).

 The designed polymers which composed of sulphonated poly (ether sulphone) (SPES) and BT units were synthesized by typical polycondensation. The obtained polymers were characterized by 1H-NMR, 13C-NMR and GPC. AFM analysis was used to observe the morphology of membrane surface, which suggested that the BT-based membrane exhibit a regular structure (Figure 2). Hence, the membrane structure is greatly influenced by the introduction of BT units. Given the different structuring described above, the density of sulphonic acid group for the BT-based membrane was investigated by FT-IR measurements. Consequently, we observed an interesting tendency that the absorption peaks arised from the sulphonic acid group are gradually shifted to the higher wavenumbers, which suggested the progression of high-density growth favorable for the efficient proton transfer. Then, the water contents of humidity dependency for these membranes were calculated from their weights. The results indicate that the introduction of BT unit is effective to prompt the swelling resistance of PEMs.

 Being motivated by the above results, we evaluated the proton conductivity of the BT-based membrane as a function of RH at 80°C (Figure 2). The BT-based membrane exhibits high proton conductivity over a wide range of RH. In particular, the conductivity of the BT-based membrane is 4 times higher than that of SPES at 30% RH. Furthermore, the activation energy from the Arrhenius plot at 40% RH is 19.3 kJ/mol, which is lower than that of SPES (24.8 kJ/mol), indicating that effective proton transition occurs in the BT-based membrane4.

 In conclusion, we found that the BT-based membrane afford high proton conductivities, particularly in low RH conditions along with a unique structure by the introduction of just a small amount of the BT unit. We believe that this design concept based on BT units is greatly effective for constructing a proton conduction favorable structure when developing novel PEMs.

Reference

(1) T. Ogawa, K. Kamiguchi, T. Tamaki, H. Imai, and T. Yamaguchi, Anal. Chem., 86, 9362 (2014).

(2) M. Akhtaruzzaman, N. Kamata, J. Nishida, S. Ando, H. Tada, M. Tomura, and Y. Yamashita, Chem. Commun., 25, 3183 (2005).

(3) S. Ando, J. Nishida, H. Tada, Y. Inoue, S. Tokito, and Y. Yamashita, J. Am. Chem. Soc., 127, 5336 (2005).

(4) N. Hara, H. Ohashi, T. Ito, and T. Yamaguchi, J. Phys. Chem. B., 113, 4656 (2009).

Figure 1

2618

, , and

Polymer electrolyte membrane fuel cells (PEFCs) as one of fuel cell systems, offer a board range of benefits, including: (1) high efficiency, (2) clean process (no CO2 emission), (3) compact design1, and (4) quiet operation. As one of solutions to current technical problems, PEFC operation at high temperature (low humidity) has been considered as a promising system to reduce cost of PEFCs. High temperature PEFCs have some additional merits such as fast reaction kinetic, low activation energy for power generation and high CO catalyst poisoning tolerance2. These advantages make high temperature PEFCs very attractive. However, there are also some technical barriers to develop this attractive system. Especially, harsh environment such as high temperature and low humid condition are very severe for polymer electrolyte membranes (PEMs) as components of PEFCs.

Sulfonated poly(ether sulfone)s (SPESs) have been commonly developed as hydrocarbon-type PEMs3,4for high temperature PEFCs, because of the high durability at high temperature condition. Several kinds of SPESs with different structures were investigated previously. However, most of them used only aromatic chain as backbone, which lead the membranes to be tough. In this research, we have investigated a series of copolymerized sulfated poly (ether sulfone) (SPES) with aliphatic chain for high temperature PEFCs. Sulfuric acid was attached to the hydroxyl group connected to the aliphatic main chain. The attached sulfuric acid would be able to move easily and work as a proton conductor because aliphatic unit is more flexible than aromatic unit. We evaluated the basic property of the aliphatic SPES as high temperature PEMs although chemical stability of the aliphatic unit is not higher than that of the aromatic unit.

SPES shown in Figure 1 was prepared by sulfation of the hydroxyl group of aliphatic PES. Purification and protonation of the obtained SPES were carried out thoroughly by dialysis from SPES HCl aq. The structure of SPES was confirmed by 1H NMR. SPES films were prepared by cast method. SPES DMSO solution was cast on a petri dish and dried at 60 ˚C in vacuo overnight. The membrane showed hydrophobicity. This property indicated that it showed strong durability against water, which is necessary in fuel cell system. Thermal stability of PEMs is significantly required for high temperature PEFC operation. The obtained SPES was investigated by TGA/DSC experiment as shown in Figure 2. From the result of TGA, SPES had three stages of weight loss. The weight loss was assigned to the removing of water from 66 ˚C to 100 ˚C, thermal decomposition of sulfuric acid2from 176 ˚C to 343 ˚C and the thermal decomposition of the main chain from 343 ˚C to 600 ˚C. And DSC result indicated that no glass transition temperature was observed from SPES at operation temperature region for high temperature PEFC (<120 ˚C). In addition, the proton conductivity of SPES film was observed. The obtained proton conductivity was 0.3 mS/cm at 120 ˚C, 100%RH, which may be caused the high hydrophobic and long distance between sulfonic acid group. We introduce more sulfuric acid group in SPES, and the effect of the sulfuric acid amount and introduction position of sulfonic acid on the proton conductivity is discussed and an optimum condition for preparing MEA for high temperature PEFC is investigated.

Reference

  • T. Husaboe, J. A. Rittenhouse, M. D. Polanka, P. J. Litke, J. Hoke, AIAA Paper. DOI, 10, (2013).

  • S. Bose, T. Kuila, T.X. Nguyen, N. H. Kim, K. T. Lau, J. H. Lee, Prog. Polym. Sci., 36, 813 (2011).

  • Y. S. Ye, J. Rick, B. J. Hwang, Polymers, 4, 913 (2012).

  • M. Ulbricht, Polymer, 47, 2217 (2006).

Figure 1

2619

, , and

Proton exchange membrane fuel cells (PEMFCs) have received much attention in recent years because of their high power density, efficiency and zero-environmental pollution. As one of the key components in fuel cells, the proton exchange membrane is expected to have high proton conductivity and good electrochemical stability. In the attempt to promote PEMCFs commercialization, high cost of fuel cell systems and short lifecycle are the two main issues that need to be addressed, thus large research effort has been devoted in developing new polymer electrolytes that can replace the usually employed proton conductors, e.g. Nafion®, with other membranes of comparable performances but lower cost.As a low-cost and eco-friendly biopolymer, Chitosan (CS)-based membrane electrolyte is currently studied as alternative candidate for PEMFC application to possibly produce economically viable fuel cells. Several works have shown that Heteropolyacids (HPAs) can be used to prepare Chitosan polyelectrolytes (PECs) to be employed as proton exchange membrane in low temperature fuel cell. HPAs, such as phosphotungstic acid (PTA) or phosphomolybdic acid (PMA), are strong Bronsted acid as well as solid electrolytes, thus being very effective in the fabrication of high proton conductive organic–inorganic nanocomposite membranes for fuel cell. A survey of the already published works on the use of CS/HPA composite membranes show that they have been tested as proton conductors in direct methanol, in borohydride and hydrogen fed fuel cells [1-3 and refs therein]. In previous works [1-3] we have shown that CS/PTA membranes, prepared using an alumina porous medium for the slow release of H3PW12O40, show good performances when employed as electrolyte in H2 fed fuel cell with proton conductivity of ~ 14 mS cm-1. Notably, in ref. [4] CS/PMA membranes are reported to exhibit proton conductivity higher than that measured with CS/PTA, which is mainly because the proton conductivity of PMA is higher than that of PTA. However, a careful inspection of published papers revealed that CS/PMA membranes haven't been yet tested in low temperature fuel cell. Starting from these findings, this paper in focused on the characterization of CS/PMA and mixed CS/PMA-PTA membranes prepared according to the above described procedure with the aim to evaluate their performance as proton conductors in hydrogen fed fuel cell. X-ray diffraction and Fourier Transform Infrared Spectroscopy analyses are performed to study the structure and composition of the membranes, while Scanning Electron Microscopy is used to get information on the membranes morphology and thickness as a function of the preparation conditions. The membranes are tested in a H2/O2 fuel cell working at low temperature (25°C) and fixed Pt loading (1 mg cm-2). Impedance Spectroscopy is used to get information of the conductivity of the membrane as well as to model the overall electrical behaviour of the cell. The role of humidification in influencing the proton conductivity of the membranes as a function of the employed HPA is also studied.

References

1 -M. Santamaria, C.M. Pecoraro, F. Di Quarto, P. Bocchetta, J. Power Sources 276 (2015) 189-194.

2 - C.M. Pecoraro, M. Santamaria, P. Bocchetta, F. Di Quarto, Int. J. Hydrogen Energy, 40 (2015) 14616-14626.

3 - M. Santamaria, C.M. Pecoraro, F. Di Franco , F. Di Quarto, I. Gatto, A. Saccà, Int. J. Hydrogen Energy (2016), http://dx.doi.org/10.1016/j.ijhydene.2016.01.133.

4 - Z. Cui, W. Xing, C. Liu, J. Liao, H. Zhang, J. Power Sources 188 (2009) 24-29.

2620

, and

Introduction

Polymer electrolyte membrane fuel Cells (PEMFC) have received much attention, because PEMFC are promising power generator for many applications such as automobiles and portable devices. One of the key componenet for PEMFC is polymer electrolyte membrane (PEM) that can transfer protons from cathode to anode, and the performance of PEM highly depends on their proton conductivity. Most-frequently studied PEM is Nafion due to their high proton conductivity even at room temperature. However, Nafion has serious drawbacks such as high cost and high humidity dependency, and also the operation temperature is limited up to 100 oC. Since higher temperature can give better reaction kinetics and reduce poisoning of Pt catalyst, materials that can show high proton conductivity at lower humidity and higher temperature have been required.

Polybenzimidazole (PBI) doped by Phosphoric acid (PA) has been developed to provide high proton conductivity at high temperature under low humidity. However, leakage of PA was observed after long-term operation1, and this hinders further applications of PA doped PBI.

To solve this problem, Kawahara et al2 synthesized propanesulfonic acid-grafted PBI (PBI-PS) by substitution of benzimidazoles of PBI. PBI-PS showed improved solubility and 10-3 S/cm as its proton conductivity over 100 oC. However, no fuel cell test was conducted, and higher proton conductivity is required to be commercialized.

In this research, poly(2,5-benzimidazole) (ABPBI) was chosen instead of PBI due to the shorten distance between benzimidazole groups, which enables to shorten the distance between grafted propanesulfonic acids. In addition, ABPBI has advantages in the cost and the ease of synthesis compared to PBI since ABPBI can be easily obtained from the self-condensation of one monomer. ABPBI grafted by propanesulfonic acid (ABPBI-PS) was synthesized, and their properties were compared with PBI-PS.

Experiment

Synthesis of PBI-PS

In 30 mL of dimethylacetamide solution of PBI (32.4 mM), NaH (0.051 g, 2.13 mmol) was added, and the temperature was kept at 90 oC for 3 h. Then, 3-bromopropanesulfonic acid sodium salt (0.877 g, 3.44 mmol) was added to the solution, and the temperature was kept at 30 oC for 2 h and 80 oC for 22 h. The solution was reprecipitated in acetone-water mixture (9:1) and then, the precipitate was filtrated and washed with acetone. After drying under vacuum at 60 oC for 24 h, yellow powder (0.262 g, 45.2 %) was obtained.

Synthesis of ABPBI-PS

In 30 mL of dimethyl sulfoxide solution of ABPBI (86.1 mM), NaH (0.186 g, 7.75 mmol) was added, and the temperature was kept at 120 oC for 24 h. Then, 3-bromopropanesulfonic acid sodium salt (1.74 g, 7.73 mmol) was added to the solution, and the temperature was kept at 30 oC for 24 h. The solution was reprecipitated in acetone-water mixture (9:1) and then, the precipitate was filtrated and washed with acetone. After drying under vacuum at 60 oC for 24 h, brown powder (0.668 g, 99.4) was obtained.

Membrane fabrication

DMSO solutions of ABPBI-PS (10 mL, 20 mg/mL) were poured into petri dishes and heated at 60 oC for 24 h. The obtained film (45.4 μm) was treated with HCl (1.0 M). PBI-PS membrane was prepared in similar manner.

 

Result and Discussion

PBI-PS and ABPBI-PS were synthesized based on previous reports2,3. From FT-IR measurements, both compounds showed new peaks at 1190 and 1050 cm-1 assignable to asymmetric and symmetric stretching of SO3. Based on the 1H NMR proton integration, the grafting ratio of PBI-PS and ABPBI-PS are 94.5 % and 65.0 %, respectively.

The obtained membranes were yellow and self-standable. We found that the thickness of the membrane could be controlled by changing the concentration of the solutions. Proton conductivity measurements were conducted from 30 oC to 120 oC at 100 % relative humidity. Similar to the result from Kawahara, PBI-PS showed 1.5 mS/cm-1 as its maximum conductivity at 110 oC. Interestingly, the maximum conductivity of ABPBI-PS was 16 mS/cm-1 at 120 oC, which was much higher than PBI-PS. This gap was probably brought by the shorter distance between grafted propanesulfonic acids of ABPBI-PS.

References

(1) Oono, Y.; Sounai, A.; Hori, M. J. Power Sources2013, 241, 87–93.

(2) Kawahara, M.; Rikukawa, M.; Sanui, K.; Ogata, N. Solid State Ionics2000, 136, 1193–1196.

(3) Namazi, H.; Ahmadi, H. J. Memb. Sci.2011, 383, 280–288.

Figure 1

2621

, , , and

Although the big progress in developing heteroatom doped carbon catalysts for oxygen-reduction reaction (ORR), the current noble metal-free catalysts are still facing the poor ORR activity which restricts its application in fuel cell. Developing new fuel cell catalysts with excellent ORR performance may be a potential way to solve this drawback. Herein, we developed a facile and green method to synthesize nitrogen and sulfur co-doped three-dimensional reduced graphene oxide networks-supported cobalt nanoparticles (Co/N-S-3DrGO) using hydrothermal method. The doping of N and S not only provide the graphene networks with many defects, but also play significant roles in engineering the covalent coupling between cobalt nanoparticles and the 3D graphene networks. The Co/N-S-3DrGO displays electrocatalytic activity toward ORR, which is comparable to the commercially available Pt/C (20%) catalyst, in addition to their long-term stability and high methanol tolerance. Such superior performances may be attributed to the synergistic effect, which includes the high catalytic sites of N-S heteroatom, the distribution activity of cobalt nanoparticles and the high electron transfer rate of the 3D graphene networks. Our results indicate a simple strategy to design highly efficient graphene-based non-noble catalysts for fuel cell application.

2622

, and

The catalyst layer, one of the components of membrane electrode assembly in polymer electrolyte fuel cells, was modeled by molecular dynamics methods, and proton transport in ionomer thin films was simulated to elucidate the effect of surface wetting interaction on the ionomer thin film morphology and proton transport property. Ionomer thin film properties are considered to be affected by the surface property of supported carbon due to their nano-scale thickness (4-10 nm thick). One of the surface properties is surface wetting interaction basically attributed to the physical or chemical property of surfaces.

In this study, modeled Nafion thin films with 4 nm and 10 nm thick were generated on the two ideal LJ smooth surfaces which have different strength of interactions to focus on the effect of the surface wetting interaction on the ionomer thin film morphology and proton transport. Protons exist as hydronium ions in ionomer thin films for chemical stability and are transported by the two transport mechanisms which are the vehicle mechanism (just move as hydronium ions) and the Grotthuss mechanism called proton hopping. It is common that proton transport model is restricted to only the vehicle mechanism when using classical molecular dynamics method, however, our proton transport model includes the Grotthuss mechanism by using aTS-EVB method (anharmonic two-state empirical valence bond method) in order to make proton transport model more reproducible.

From the analysis in the ionomer thin films with 4 nm thick, surface wetting interaction was found to affect the orientation of sulfonic group near the surface. The orientation is well-ordered on the higher hydrophobic surface, and the sulfonic groups point to the gas phase direction, while it is randomly distributed on the lower hydrophobic surface, which make differences in the water cluster connectivity and size in ionomer thin films between the two surfaces. Proton transport property was evaluated by the self-diffusion coefficient. The proton self-diffusion coefficient in ionomer thin films was indicated to be improved on higher hydrophobic surface probably because higher hydrophobic surface makes ideal water cluster structures for proton transport in ionomer thin films.

As compared to the ionomers with the different film thickness, the proton self-diffusion coefficient in ionomer thin films on higher hydrophobic surface was indicated to decrease in the increase of the film thickness, while that on lower hydrophobic surface was indicated to increase slightly. This result is expected to come from the confinement and interface effects which can be remarkable when the thickness decrease. The ionomer morphology on higher hydrophobic surface has advantages for proton transport probably thanks to the effects as observed in the ionomer with 4 nm thick, while that on lower hydrophobic surfaces has disadvantage. Since the effects can be weaker with the increase in the film thickness, the proton self-diffusion coefficient in ionomer thin films on higher hydrophobic surface was found to decrease with the increase in the film thickness, while that on lower hydrophobic surface was found to increase slightly.

2623

, and

The performance of membrane electrode assemblies (MEA) is considerably affected by water balance or content in ionomer solution. The three-dimensional structure of the ionomer in electrode is very important for getting higher current density in MEA. In this study, some representative water-based and alcohol-based commercial ionomers were compared for getting higher MEA performance under different humidity level. The contact angle of the catalyst layer was measured to determine the variation in wetting properties with different ionomer solution. Various MEAs with different electrode combination were tested in single cell station. And some electrochemical techniques were applied to this MEA such as polarization curve, electrochemical impedance spectroscopy (EIS) and cyclic voltammetry. The results show that a MEA of water-based ionomer as anode and alcohol-based ionomer as cathode exhibits higher performance at most humidity levels.

2624

, , , , and

Operating polymer electrolyte membrane fuel cell(PEMFC) at high temperatures (HT, 100oC~200oC) is beneficial owing to fast kinetics for electrochemical reactions and high tolerance to carbon monoxide(CO) poisoning. Also, the HT operation allows a simpler system design by eliminating water management and PROX. However, current HT-PEMFCs still suffer from insufficient power performance partly originating from large membrane resistance. Typically, HT-PEMFC employs the phosphoric acid-doped PBI membrane and its proton conductivity depends on the acid doping level. Therefore, increasing the doping level is prerequisite for high power performance of HT-PEMFC. However, phosphoric acid doping level is practically limited by a decreased mechanical stability of the membrane with acid doping. To mitigate the problems caused by the loss of low mechanical strength with the acid doping, a thick PBI membrane (>100um) is conventionally employed although its large membrane resistance and consequent negative influence on power performance. Against this backdrop, in this study, we present a novel approach for addressing the trade-off between proton conduction and mechanical robustness of PBI based HT membrane.

For fabricating an reinforced PBI membrane coupled with a porous PTFE membrane, PBI filling in a porous PTFE membrane is highly challenging due to low compatibility between hydrophobic PTFE and hydrophilic PBI. In this work, to make the PTFE surface hydrophilic, a porous PTFE membrane is treated with dopamine which is as known nature-inspired adhesive and self-polymerized into polydopamine via a pH-induced oxidation. A polydopamine layer can be formed on the PTFE surface by simply immersing the porous PTFE membrane in the aqueous buffered dopamine solution at room temperature. Compared to the pristine PTFE membrane, the polydopamine-treated PTFE membrane is more densely filled with PBI after the impregnation of PBI solution, indicating an effective intrusion of the PBI solution into the modified hydrophilic pores. The reinforced membrane shows a higher mechanical strength after phosphoric acid doping process even at a thin membrane thickness(~50um) in comparison with a 100 micron thick un-reinforced PBI membrane.

The second feature of this approach is the introduction of plasticizer in the PBI matrix for increasing the acid doping level for the reinforced PBI/PTFE membrane. In general, mechanical supports suppress swelling of the PBI matrix with phosphoric acid solution, resulting in a lower acid doping level compared to that of pristine PBI membrane. However, poly(ethylene glycol) (PEG), which can be exchanged with phosphoric acid, provides an additional space for accommodating phosphoric acid in the PBI matrix when incorporated in the PBI matrix. PEG and PBI forms a miscible blend and PEG can be extracted out during the acid doping process, being exchanged with phosphoric acid as confirmed by DSC and FT-IR analysis. The PEG/PBI blend exhibits a higher mechanical stability owing to its lower expansion but a higher proton conductivity owing to its high acid doping level after the acid doping, clearly demonstrating the benefit of the approach. When the PEG-plasticized PBI matrix and polydopamine-treated PTFE support are combined, the mechanical stability can be significantly improved compared with the PEG-plasticized PBI membrane and the proton conductivity be enhanced compared with the PTFE-reinforced PBI membrane. The power performance of an in-house membrane electrode assembly (MEA) employing the novel reinforced membrane is comparable to that of a commercial HT-MEA. Therefore, the combined strategy can overcome the trade-off between mechanical stability and proton conductivity in the design of HT-PBI membranes.

2625

, , , and

The widespread utilization of polymer electrolyte fuel cells for transportation applications such as fuel cell vehicles requires high power density operation. However, as the current density increases, the diffusion polarization, which is attributed to the insufficient mass transport, degrades the cell performance. In particular, the oxygen transport resistance in catalyst layers (CLs) and micro porous layers (MPLs) at the cathode side is one of the major factors of the performance degradation. Therefore, the reduction in the oxygen transport resistance is important for the high power density operation.

In the previous research, Kinefuchi et al. have investigated the oxygen diffusion resistance in MPLs and CLs using the direct simulation Monte Carlo (DSMC) method. They reported that the simulation well reproduces the experimental result in MPLs. However, the DSMC results underestimate the oxygen diffusion resistance in CLs. This discrepancy is caused by the drawback to a scattering model of gas molecules on surfaces, which plays a crucial role in DSMC analysis. In these calculations, the same scattering model is used for the scattering of oxygen molecules on surfaces regardless of the presence or absence of ionomer on micro porous carbons. An advanced scattering model that can reproduce the scattering and surface diffusion phenomena on ionomer surface, is required for more accurate analysis of the oxygen transport in CLs. The construction of the model needs the detailed understanding of the gas‒surface interaction.

In this study, we have investigated the mechanics of oxygen scattering and surface diffusion on ionomer surface using molecular dynamics simulation. We modeled the solid surface in CLs as an ionomer film covering the carbon layers. The simulation domain was 51.1 × 44.2 × 100 Å3 in size. The ionomer film was composed of water molecules hydronium ions, and perfluorosulfonic acid molecules as typified by Nafion, water molecules, and hydronium ions. Oxygen molecules were directed to the ionomer surface and the trajectory calculations were performed for 20 ps. The initial translational energy and incident angle were the same for all the incident molecules, but the initial position and orientation were given randomly in order to obtain the sufficient statistics. If an oxygen molecule desorbed from the surface and reached the initial height, it was defined to be scattered.

Firstly, we have evaluated the energy transfer between oxygen molecules and the ionomer surface by calculating between translational energy before and after the scattering. Oxygen molecules with low incident energy tend to receive energy when they leave the surface after the collisions.

Next, we have evaluated the scattering phenomena of oxygen molecules on ionomer surface by examining the energy and angular probability distributions of scattered molecules. The energy distribution depends on the incident energy, and deviates from the distribution given by diffusive scattering model, that the incident molecules are assumed to accommodate completely with the surface. On the other hand, the scattering angle distribution is independent of the incident energy, and well-reproduced by the diffusive scattering model. These results suggest that oxygen molecules do not accommodate completely with the ionomer surface when they are reflected, however they are reflected following the diffusive scattering model because of the corrugation of the ionomer surface.

2626

, and

In cold regions, ensuring startup ability and durability of polymer electrolyte fuel cells (PEFCs) at subfreezing conditions is one of critical issues for the practical use of PEFC. At cold startup of PEFCs near 0 °C like –10 °C, the produced water remains as supercooled state before it freezes, and then the shutdown occurs with ice layer formation between the cathode catalyst layer (CL) and micro-porous layer (MPL) (1). It was estimated from our previous study (2) that water transport in the CL occurs by concentration gradient in ionomer. This study investigated the water transport in ionomer and the ice formation behavior by measurement of cold start characteristic and observation of membrane electrode assembly (MEA) by a cryo-SEM. To evaluate the water transport, we measured change in polymer membrane resistance corresponding to water content during the supercooling release. Further, to improve the cold start characteristic, we introduced a hydrophilic MPL and confirmed that the hydrophilic MPL makes the startup period longer and improves ice distribution.

A single cell with an active area of 25 cm2(5 cm × 5 cm) was used in this study. After a preconditioning process to enhance the performance of a MEA, the initial conditions of the water in the cell were carefully controlled by the procedure described in Ref. 1. Then, the cell and the chamber were cooled to –10 °C, and cell operation was started and maintained at a constant current density. The anode and the cathode supplied gases were dry hydrogen and dry air. For the cryo-SEM observation, the sample of the MEA as removed from the cell after the stop of operation was immediately immersed in liquefied nitrogen, and the sample was set to a sample holder. The sample was cut by a cold knife in a vacuumed chamber, and the cut section area was coated with gold-palladium alloy (Au-Pd), and observed at –150 °C.

To estimate the changes in the water content of a polymer electrolyte membrane from supercooled water to ice, we measured the change in the in-plane high frequency resistance of the polymer membrane. It is expected that when supercooled water changes to ice, the chemical potential of water decreases and the water content of the polymer membrane contacting with ice also decreases. Fig.1 shows change in resistance and temperature of the polymer electrolyte membrane. Supercooling release occurs after around 70 min at –7 °C, and the temperature slightly rises due to heat of solidification. Here, resistance increases even after the temperature becomes again –7 °C. This resistance increase suggests the decrease in the water content of the polymer electrolyte membrane due to ice formation. ΔR is the difference between the resistance contacting with ice and with supercooled water. Fig.2 shows relationship of ΔR and supercooling degree (ΔT). ΔR increases as ΔT increases, and this supports the above estimation that ice freezing from the supercooled state takes up water from the ionomer.

Secondly, we introduced a hydrophilic MPL made by the gas diffusion electrode (GDE) method (3) to prevent water from freezing at the interface between the CL and MPL and to improve the cold start characteristics. In the previous study (1)(2), we observed the ice layer at the interface with a hydrophobic MPL. Fig.3 shows change in cell voltage and resistance at 0.02 A/cm2, –10 °C. This result suggests that hydrophilic MPL increases the startup period than conventional hydrophobic MPL. Fig.4 is observation results of cross-section of the cathode MEA with hydrophilic MPL after the 0.02 A/cm2shutdown. Figs.4(a) and (b) are the cryo-SEM images with low magnification and with high magnification. Fig.4(a) shows that there is no ice layer at the interface and water freezes in the MPL. Fig.4(b) shows that there is little ice in CL, and this suggests that the hydrophilic MPL enhances the migration of water through ionomer in the CL. These results also suggest that the hydrophilic MPL is effective to improve the characteristics at cold startup.

Reference

(1) Y. Tabe, et al. , Journal of Power Sources208 (2012), 366.

(2) Y. Ishima, et al. , National Symposium on Power and Energy Systems(2015), 101.

(3) Y. Aoyama, et al. , ECS Trans. 69 (17) (2015), 743.

Figure 1

2627

, , , , and

Polymer electrolyte membrane fuel cell (PEMFC) technology faces considerable demands of cost competitiveness for mass commercialization of fuel cell electric vehicle. A significant challenge to the cost reduction is the reduction of Pt loading for membrane electrode assembly (MEA). Besides a lowered catalytic activity, low Pt-loaded catalyst layer (CL) suffers from large mass transports resistance, resulting in low power performance at high current densities. Such larger mass transport resistance at a lowered Pt loading can be attributed to a larger oxygen transport resistance from the ionomer film covering catalysts and more significant water flooding due to a lower amount of the pore volume in cathode CL. For conventional CL structure, gaseous oxygen and liquid water share the same network of meso-pores (< 50 nm) inside a flat CL for their transport, therefore, the water condensation in the pores inevitably causes a blocking of oxygen transport. In order to address this issue, an ionomer fiber-induced macro-porous CL is presented. It is fabricated via electrospinning of ionomer fibers onto the membrane, followed by spray-coating catalyst ink on the ionomer fiber-decorated membrane. The ionomer fiber deposition on the membrane induce a roughening of the membrane surface, which allows the formation of a CL, morphology of which dictates that of the ionomer fibers, and of micron-scale pores between the CL and gas diffusion layer. The new CL structure dramatically improves power performances at high current densities owing to an effective oxygen and water transport through the micron-scale mass transport pathway. From polarization curve and oxygen transport resistance analysis, and stability under constant current operation for various feeds, the efficacy of the unique CL morphology in enhancing power performances is demonstrated and understood. Also, the relationship between the macro-pore in CL and mass transport resistance is investigated with systematically varying the pore size and porosity.

2843

, , , and

The degradation of the polymer electrolyte is one of the dominant factors, which limits the lifetime of the polymer electrolyte fuel cells (PEFCs). Therefore, the understanding and mitigation of all contributing degradation phenomena is fundamental in order to establish the PEFC as competitive product.

Membrane degradation is often quantified by analysing the fluoride emission rate by determining the fluoride content in the effluent water. However, when oligomers (i.e. larger polymer fragments) are released, the ion conductivity of the ionomer declines without extensive fluoride release, as the carbon backbone fluorine bonds are still intact. Thus, it is necessary to conduct effluent water treatment in order to determine the total fluorine emission rate (tFER) instead of the fluoride emission rate (FER) only [1].

In the presented work, an accelerated stress test (AST) was conducted in a single cell with a segmented S++ current scan shunt device with a sensor plate on the cathode for spatial current and temperature measurements in 10x10 and 5x5 segments, respectively. A membrane electrode assembly was characterised in situ [2]. Throughout 100 hours of stress testing, the effluent water was collected and used to determine the FER, using a fluoride selective electrode (Figure 1) [3]. Furthermore, effluent water samples were treated with KOH following different protocols. The alkaline treatment decomposes residual ionomer fragments for the detection of the total fluorine emission rate. The tFER allows more profound assumptions regarding the extent of the membrane degradation and further, gives insight into the underlying degradation mechanism.

In preliminary results it is shown that by conducting alkaline effluent water treatment at a temperature of 500 °C the amount of detected fluoride was increased significantly (Figure 2). Thus it can be concluded that not all emitted fluorine is detectable without previous treatment, indicating that the electrolyte membrane decomposes into oligomeric fragments.

Acknowledgement

The research leading to these results has received funding from the European Union's Seventh Framework Programme (FP7/2007-2013) for the Fuel Cells and Hydrogen Joint Technology Initiative under grant agreement n° 621216.

[1] Aarhaug TA, Svensson AM. Degradation Rates of PEM Fuel Cells Running at Open Circuit Voltage. ECS Trans., vol. 3, ECS; 2006, p. 775–80. doi:10.1149/1.2356197.

[2] Bodner M, Cermenek B, Rami M, Hacker V. The Effect of Platinum Electrocatalyst on Membrane Degradation in Polymer Electrolyte Fuel Cells. Membranes (Basel) 2015;5:888–902. doi:10.3390/membranes5040888.

[3] Bodner M, Hochenauer C, Hacker V. Effect of pinhole location on degradation in polymer electrolyte fuel cells. J Power Sources 2015;295:336–48. doi:10.1016/j.jpowsour.2015.07.021.

Figure 1

I01-D Poster Session - Oct 5 2016 6:00PM

2628

, and

We suggest a simple, inexpensive approach for enhancing the durability of automotive proton exchange membrane fuel cells by selective promotion of the hydrogen oxidation reaction (HOR) and suppression of the oxygen reduction reaction (ORR) at the anode in startup/shutdown events. Dodecanethiol forms a self-assembled monolayer (SAM) on the surface of Pt particles, thus decreasing the number of Pt ensemble sites. Interestingly, by controlling the dodecanethiol concentration during SAM formation, we can precisely tune the number of ensemble sites to an optimum such that it is sufficient for the HOR but insufficient for the ORR. Thus, a Pt surface with an SAM of dodecanethiol clearly shows HOR-selective electrocatalysis. We demonstrate clear HOR selectivity in unit cell tests with the actual membrane electrode assembly as well as in an electrochemical three-electrode setup with a thin film rotating disk electrode configuration.

2629

, , , , , and

Abstract

Cost and durability are two major factors that delay large-scale production and commercialization of PEMFC's. One of the technologically ready application of the proton exchange membrane fuel cells (PEMFC) is in the combined heat and power systems (µCHP), which are used in the individual households or buildings. The performance of the µCHP systems greatly depends on the purity of the hydrogen stream, which is produced via methane reforming process. To overcome low CO tolerance of the commercially used Pt electrocatalyst and to lower the catalyst content we have prepared non-stoichiometric tungsten oxide as a Pt based catalyst support. We have prepared several catalysts designated as 10% Pt/WO3-C, 20% Pt/WO3-C, 40% Pt/WO3-C. The structure and morphology characteristics of the prepared catalysts were investigated using XRD, TEM and SEM/EDX techniques. Investigations concerning electroactivity of these catalysts towards the hydrogen oxidation reaction (HOR) were performed using cyclic voltammetry, linear sweep voltammetry, forming an ultra thin catalyst layer onto RDE. Mechanism and the kinetics of the prepared catalysts towards HOR were evaluated and if was found that increased mass activity of the 10% Pt/WO3-C could be attributed to the interactive naure of the WO3 catalyst support. Obtained results clearly show increased CO tolerance of Pt/WO3-C catalyst compared to commercial Pt/C, which was confirmed by lowering the stripping potential of the CO, adsorbed on the surface of the 10% Pt/WO3-C. catalyst is more facile than that on commercial 40% Pt/C. These catalysts were employed as anode catalyst in the MEA, and the performance of single cell PEMFC were compared to commercial catalyst.

2630

, , , , , and

Abstract

Cost and durability are two major factors that delay large-scale production and commercialization of PEMFC's. One of the technologically ready application of the proton exchange membrane fuel cells (PEMFC) is in the combined heat and power systems (µCHP), which are used in the individual households or buildings. The performance of the µCHP systems greatly depends on the purity of the hydrogen stream, which is produced via methane reforming process. To overcome low CO tolerance of the commercially used Pt electrocatalyst and to lower the catalyst content we have prepared non-stoichiometric tungsten oxide as a Pt based catalyst support. We have prepared several catalysts designated as 10% Pt/WO3-C, 20% Pt/WO3-C, 40% Pt/WO3-C. The structure and morphology characteristics of the prepared catalysts were investigated using XRD, TEM and SEM/EDX techniques. Investigations concerning electroactivity of these catalysts towards the hydrogen oxidation reaction (HOR) were performed using cyclic voltammetry, linear sweep voltammetry, forming an ultra thin catalyst layer onto RDE. Mechanism and the kinetics of the prepared catalysts towards HOR were evaluated and if was found that increased mass activity of the 10% Pt/WO3-C could be attributed to the interactive naure of the WO3 catalyst support. Obtained results clearly show increased CO tolerance of Pt/WO3-C catalyst compared to commercial Pt/C, which was confirmed by lowering the stripping potential of the CO, adsorbed on the surface of the 10% Pt/WO3-C. catalyst is more facile than that on commercial 40% Pt/C. These catalysts were employed as anode catalyst in the MEA, and the performance of single cell PEMFC were compared to commercial catalyst.

2631

, , , and

Carbon supported Pt-Ru systems such as PtRu/C and Pt2Ru3/C have been investigated as anode catalysts of polymer electrolyte membrane fuel cells. However, hydrogen oxidation reaction (HOR) activity for Pt-Ru/C is significantly degraded due to adsorption of carbon monoxide (CO) on the active Pt sites and dissolution of Ru. The addition of metal oxide nanosheets enables to enhance the performance of catalysts. We have reported that RuO2 nanosheets [1] as an additive to Pt/C (RuO2ns-Pt/C) and PtRu/C (RuO2ns-PtRu/C) enhances CO tolerance at room temperature [2]. The HOR current of RuO2ns-Pt/C was improved by 25% compared to Pt/C. RuO2ns-PtRu/C had higher HOR activity and enhanced CO tolerance. Thus, CO poisoning is suppressed by the addition of RuO2 nanosheets. However, it is necessary to measure the performances at higher temperature near actual operating of PEFC. In this study, we report the addition effect of RuO2 nanosheets to PtRu/C and Pt2Ru3/C at 25◦C as well as 70◦C.

Chronoamperograms (CA) of commercial and RuO2 nanosheests modified catalysts at 25 and 70◦C are shown in Figures A and B. At 25◦C (Fig. A), the addition of RuO2 nanosheets increased the initial HOR activity and the activity after accelerated durability tests (ADT) in 300 ppm CO containing H2 (CO/H2) as well as pure H2. The current was enhanced in the entire range of CA. The addition of RuO2 nanosheets slightly enhanced CO tolerance and durability which were represented by ratio of density current j (in H2) divided by j (initial in CO/H2) and j (initial in CO/H2) divided by j (after ADT in CO/H2), respectively. This behavior is similar to the previous result of comparing RuO2ns-PtRu/C and PtRu/C [2].

The HOR current at 70◦C showed the different behavior compared to 25◦C (Fig. B). At 70◦C, there was a difference of HOR current between composite and commercial catalysts in CO/H2 while HOR activity of the catalysts was almost the same in pure H2. Thus, the decline in HOR current of composite catalysts was slower than non-modified catalysts. Therefore, the composite catalysts suppresses CO adsorption. In particular, the durability of RuO2ns-Pt2Ru3/C at 70◦C shows higher (93%) than the other catalysts (78~86%).

This work was supported in part by the "Polymer Electrolyte Fuel Cell Program" from the New Energy and Industrial Technology Development Organization (NEDO), Japan

[1] W. Sugimoto, H. Iwata, Y. Yasunaga, Y. Murakami, and Y. Takasu, Angew.Chem. Int. Ed., 42, 4092 (2003).

[2] D. Takimoto, T. Ohnishi, and W. Sugimoto, ECS Electrochem. Lett., 4(5), F35-F37 (2015).

Figure 1

2632

, , , and

For a power generation of a polymer electrolyte fuel cell (PEFC), reformed hydrogen gas is fueled to the anode. The reformed hydrogen gas contains a small amount of carbon monoxide and 10-30% carbon dioxide [1]. The carbon dioxide is reported to be reduced at Pt electrode [2, 3]. In this study, CO2 reduction is performed at various potential of Pt/C-based membrane electrode assembly(MEA). The product of CO2reduction was analyzed by a mass spectrometer.

Electrocatalyst of Pt/C (Pt: 45.7 wt%) was sprayed on a gas diffusion layer. The electrolyte membrane was sandwiched by two pieces of the gas diffusion layers which were coated electrocatalyst layers (metal amount: 1.0 mg cm-2), and then, it was hot-pressed to obtain on MEA. Geometric electrode area of the MEA was 3×3 cm2. The prepared MEA was installed in a single cell having a dynamic hydrogen electrode (DHE) as a reference electrode. The temperature of the single cell was controlled at 80°C. Humidified carbon dioxide gas (purity: 99.995%) and humidified hydrogen gas (purity: 99.999%) were supplied to the working electrode and counter electrode, respectively, at a flow rate of 50 cm3 min-1. Cyclic voltammograms were measured at a potential scan rate of 10 mV s-1and the potential scan range was between 0.00-1.38 V vs. DHE. The exhaust gas was introduced directly into the mass spectrometer (JEOL, JMS-Q1050GC Ultra Quad GC/MS ).

Figure 1 shows a result of differential electrochemical mass spectroscopy measured at 0-1.38 V vs. DHE. In Fig. 1 (a), cyclic voltammograms at the Pt/C-based MEA were obtained under CO2 and Ar atmospheres. Under the CO2 atmosphere, a typical anodic current at around 0.5 V vs. DHE is observed. This is due to the re-oxidation of the CO2 reductant which was generated at < 0.1 V vs. DHE by the CO2 reduction. Mass spectra of m/z = 2, 16, 45 versus electrode potential under CO2 atmosphere shown in bottom of Figure 1 (b)-(d). The m/z = 2, 16, 45 were originaled from hydrogen, methane and formic acid. Generation of m/z = 16 is seen at < 0.2 V vs. DHE, and m/z = 2, m/z = 45 are seen at < 0.1 V vs. DHE. It was found that there is a potential dependency for CO2reduction at Pt/C-based MEA. Consequently, adsorbed CO2reductant on the Pt remains in the potential range of 0-1.38 V vs. DHE which influences the hydrogen oxidation reaction.

[1] K. Kortsdottir, C. D. Fernandez, R. W. Lindstrom, ECS Electrochem. Lett.,2, F41-F43 (2013).

[2] N. Hoshi, S Kawatani, M. Kudo, Y. Hori, J. Electroanal. Chem., 467, 67–73 (1999).

[3] S. Shironita, K. Karasuda, K. Sato, M. Umeda, J. PowerSources, 240, 404-410 (2013).

Figure 1

2633

, , and

The research and development of highly active cathode catalysts for the oxygen reduction reaction (ORR) is one of the most important subjects to achieve high efficiency in polymer electrolyte fuel cells (PEFCs). In order to understand the ORR mechanism and find clues to the design of high-performance cathode catalysts, it is essential to analyze surface the oxidation states of catalysts and the adsorbed oxygen species in the ORR. Recently, we have clarified the surface oxidation states of Pt polycrystalline and low-index single-crystal electrodes as a function of the electrode potential by using X-ray photoelectron spectroscopy combined with an electrochemical cell (EC–XPS).1 In practical PEFCs, however, Pt nanoparticles supported on carbon black have been employed as the anode and cathode catalysts. In the present research, we have prepared a model electrode of Pt nanoparticles supported on a carbon substrate and have examined the electronic structure of the Pt nanoparticles and the oxygen species adsorbed on the surface by using EC–XPS.

Pt nanoparticles were prepared by a colloidal method,2 and dispersed on polished glassy carbon (GC). It was clarified by scanning electron microscopy (SEM) that Pt nanoparticles of ca. 3 nm diameter were well dispersed on the carbon substrate. We employed angle-resolved, grazing incidence X-ray photoelectron spectroscopy (AR-GIXPS), which is more surface-sensitive than conventional XPS, in order to analyze the small amount of Pt nanoparticles supported on the flat GC substrate.

Figure 1 shows Pt 4f XP spectra of the Pt/GC model electrode after emersion from the 0.1 M HF solution at various electrode potentials. It was found that the peaks assigned to Pt 4f7/2 and 4f5/2 shifted to higher binding energies with increasing electrode potential. This peak shift can be ascribed to the surface core level shift induced by the surface oxidation of the Pt nanoparticles. At E > 1.1 V, peaks for bulk PtO, with binding energies of 73 and 77 eV, commenced to increase due to further oxidation of the Pt nanoparticles.

Figure 2a shows O 1s XP spectra at the Pt/GC model electrode and a GC electrode (without any Pt) at 0.9 V, indicated by red and black lines, respectively. The O1s spectrum of the Pt/GC model electrode was found to include photoelectron signals attributed to oxygen species formed on the GC substrate, such as quinone and carboxyl groups. Then, we extracted the photoelectron signals originating from oxygen species adsorbed on the Pt nanoparticles by subtracting the O1s spectrum of the GC from that of the Pt/GC model electrode. Figure 2b shows the O1s difference spectrum for oxygen species adsorbed on the Pt nanoparticles. Thus, we succeeded in deconvoluting the difference spectrum into three components, H2Oad, OHad and Oad formed on the Pt nanoparticles. On the basis of the deconvoluted photoelectron intensities, it was found that H2Oad decreased with increasing electrode potential, while OHad increased. The Oad species was found to appear at E > 0.8 V. The onset potential of the OHad formation at the Pt/GC electrode was less positive than that for polycrystalline Pt, suggesting that the surface oxidation proceeded more easily at the Pt nanoparticle surfaces than at the polycrystalline surface.

This work was supported by funds for the SPer-FC Project of NEDO, Japan

 

References

1 M. Wakisaka, H. Suzuki, S. Mitsui, H. Uchida, and M. Watanabe, Langmuir, 25, 1897 (2009).

2 M. Watanabe, M. Uchida, and S. Motoo, J. Electroanal. Chem., 199, 311 (1986).

Figure 1

2634

, , , and

  • Introduction

Polymer electrolyte fuel cells (PEFCs) are expected as one of highly efficient power generation systems. Moreover, the product is only water, using hydrogen gaseous as fuel. However, many problems remain to be solved for its commercialization. H2O2 formation on anode and cathode catalyst surface in PEFCs is one of the most important issues for the durability of the MEA in PEFCs.  We reported that lager amount of H2O2 formed as an intermediate in oxygen reduction reaction (ORR) on Pt/C catalysts with lower catalyst loadings, and Fe2+ cation accelerated a deterioration of the electrolyte membrane1, 2). In addition, similar phenomenon was reported in the case of Pt/C catalysts3).  In present study, we evaluated the ORR activity and H2O2 formation rate on prepared model electrodes with different Pt/C loadings by RRDE technique.  

 

  • Experimental

The RRDE consisted of a glassy carbon (GC) disk and Pt ring sealed in a polytetrafluoroethylene (PTFE) holder used in ORR and hydrogen peroxide reduction reaction (HPRR) measurements. Commercially available 45.6 wt% Pt/C (TKK Inc, TEC-10E50E) catalyst was used for RRDE measurement. The ultrasonicated Pt/C catalyst suspension was carefully dropped on the mirror polished GC disk electrode surface with a microsyringe at each Pt/C loading densities. A thermostated three-compartment electrochemical cell was used for all electrochemical measurements. Counter electrodes were Pt and Au wire for ORR and HPRR measurements, respectively. A reversible hydrogen electrode (RHE) was used as a reference electrode. Linear sweep voltammetry (LSV) was performed from 0.05 to 1.0 V at 10 mV s-1 in oxygen saturated 0.1 M HClO4 for the ORR measurement on prepared electrodes. And then, the potential of the ring electrode held at 1.2V where it achieved to diffusion limit, and the rotating rate of the RRDE were between 400 and 3000 rpm. HPRR measurement was carried out at the same disk potential range and scan rate, in Ar saturated 0.1 M HClO4 with H2O2 additive, while ring electrode potential held at 0.1V.  

 

  • Results and Discusssion

Cyclic voltammograms (CV) per mass of Pt on prepared electrodes with different catalyst loadings shows similar behavior. A characteristic CV curves such as Hupd adsorption/desorption and Pt oxide formation, and its reduction were clearly observed. The each CV current per mass of Pt was almost equal to each catalyst loading. Therefore, the prepared electrodes were successfully highly dispersed and modified on GC disk electrodes. Hydrodynamic voltammograms for ORR at the different Pt/C catalyst loadings was obtained and achieved to diffusion limit above the catalyst loading of 2.82 µgPt cm-2. The kinetic current (IK) for ORR can estimated by the Koutechy-Levich equation, and a Mass activity (MA) and a specific activity (SA) was obtained. The MA and the SA decreased under the catalyst loading of 7.05 µgPt cm-2. This decrease indicated that the amount of the Pt/C catalyst is not enough to remove the O2 diffusion effect under the catalyst loading of 7.05 µgPt cm-2. H2O2 formation rate were calculated from LSV in the ORR.  H2O2 formed below 0.8V, and the amount was larger with lowering disk potential and decreasing catalyst loadings.  Fig. 1 shows LSVs in HPRR. The disk currents (ID), which is a reduction current of H2O2, decreased with decreasing of catalyst loadings, similar to the ORR.  On the other hand, IR , which is reduction currents of H2O2 and O2  formed on the disk electrode, increased with decreasing of catalyst loadings, while it is generally known that O2 produced from the H2O2 oxidation reaction above 0.8V.  The increase of the IR caused by decreasing catalyst loadings indicated that chemically O2 formation such as disproportionation reaction occurred on the Pt/C catalyst surface.

 

4. References

1) M. Inaba, H. Yamada, J. Tokunaga and A. Tasaka, Electrochemical and Solid-State Letters, 7 (12), A474-A476 (2004).

2) M. Inaba, H. Yamada, R. Umebayashi, M. Sugishita and A. Tasaka, Electrochemistry, 75 (2), 207-212 (2007).

3) H. Itaya, S. Shironita, A. Nakazawa, M. Inoue and M. Umeda, International Journal of Hydrogen Energy, 41(1), 534-542, 2015.

Figure 1

2635

, , , and

The carbon black-supported Pt nanoparticle catalyst (Pt/C) is a representative cathode catalyst of polymer electrolyte fuel cell (PEFC). Unfortunately, Pt/C deteriorates with loss of the electrochemical surface area (ECSA) under potential cycling operations and start-stop cycles, and is one of the major reasons for the deterioration of cell performance. Several degradation mechanisms have been considered to be responsible for the loss of ECSA[1] and then, we have been investigating the degradation mechanisms of individual Pt particles in Pt/C by "identical location field emission scanning electron microscopy (IL-FE-SEM)", which is our original technique to visualize as well as monitor the change of Pt nanoparticles at nanometer scale, in order to quantitative analysis the degradation process as well as design durable catalyst[2]. In this presentation, the potential pulse acceleration degradation test (ADT) of Pt/C was carried out in HClO4 solution, and simultaneously the degradation of individual Pt particles was monitored by using IL-FE-SEM. Furthermore, Pt dissolution amount in the solution was also measured by using inductively coupled plasma-mass spectrometry (ICP-MS) to discuss the material balance owing to the deterioration of Pt/C.

Pt/C (TEC10E50E, Tanaka Kikinzoku Kogyo, Pt weight = 46.7 wt%) was dispersed on a glassy carbon disk, and then Nafion thin film was formed on the catalyst (calculated thickness = 80 nm), which was employed as the test electrode. FE-SEM observation was carried out and the observation point was defined for IL-FE-SEM. After that, the test electrode was transferred to a glass half-electrochemical cell filled with 0.1 mol dm-3 HClO4 aqueous solution. Potential pulse ADT (0.6 V - 1.0 V, pulse time = 3 s, recommended by Fuel Cell Commercialization Conference of Japan; FCCJ) was carried out under N2 atmosphere at 60oC. IL-FE-SEM observation and cyclic voltammetry measurements were carried out after an arbitrary test period. We randomly selected 750 pieces of Pt nanoparticles to investigate the degradation mechanism. ICP-MS analysis was also carried out by sampling a tiny amount of the electrolyte solution during the ADT.

Figure 1 shows typical IL-FE-SEM images of Pt/C catalyst taken during ADT. Individual (specific) Pt particles in Pt/C were clearly monitored. The typical morphological changes were observed for Pt particles such as shrinkage and growth of the particles, disappearance, migration, coalescence, precipitation on carbon support. The typical degradation processes were observed for Pt particles such as shrinkage and growth of the particles, disappearance, migration, coalescence, precipitation on carbon support.

The ECSA decreased to 65% after 10,000th pulse. The total number of Pt particles was degreased from 750 to 697, showing the number of Pt particles is surely decreased during the ADT. Among the degradation processes, disappearance and shrinkage were recognized as main process, suggesting that dissolution of Pt should be severely taken place under the ADT. It was found from ICP-MS results that 2% of Pt was dissolved out from Pt/C catalyst after 10,000th pulse. In this presentation, the material balance of Pt during the ADT will be also discussed based on IL-FE-SEM and ICP-MS.

  • Y. Shao-Horn, et al., Top. Catal., 46, 285 (2007).

  • T. Kinumoto, et al., Electrochemistry, 83, 12 (2015).

Figure 1

2636

, and

1. Purpose

 Fuel cell is high-efficiency and clean energy system. However, high overpotential of oxygen reduction reaction (ORR) and large Pt loading in electrocatalyst are grave issues for the general use.

 Miyake et al. reported that platinum nanoparticles modified with octylamine (OA) and alkylamine with pyrene group (PA) give higher activity and durability at the molar ratio OA/PA=7/3 compared with bare Pt nanoparticles [1]. However, the platinum nanoparticles are not shape controlled or well-defined. We have investigated surface structures which enhance the ORR activity and durability using Pt single crystal electrodes modified with OA/PA. The surfaces examined are Pt(100), Pt(110), Pt(111), Pt(331)=3(111)-(111) that has the highest ORR activity [2] and Pt(775)=7(111)-(111). In the notation of high index planes such as n(hkl)-(mno), the value of n, Miller indices (hkl) and (mno) show the number of terrace atomic rows, the structures of terrace and step, respectively.

 

2. Experimental

 Linear sweep voltammograms of the ORR were measured with rotating disk electrode (RDE). Potential was scanned from 0.05 V (RHE) to 1.0 V (RHE) at scanning rate 0.010 V s-1 and rotation rate 2000 rpm. The activity for the ORR was estimated using the specific activity at 0.90 V (RHE) jk according to the Koutecky-Levich equation.

 Durability test was conducted using rectangular wave from 0.60 V step to 1.0 V.

 All the potentials were referred to RHE.

 

3. Result and Discussion

 The ORR activities of Pt(111) modified with OA/PA = 9/1, 8/2, 7/3 (molar ratio) were twice or three times higher than that of bare Pt(111). The ORR activity increases with the increase of the ratio of PA. The activity of Pt(775) = 7(111)-(111) modified with OA/PA=9/1 was also higher than that of bare Pt(775). Pt(100) and Pt(110) modified with OA/PA=9/1 gave lower ORR activity than bare electrodes. These results show that PA on wide (111) terrace enhances the ORR activity.

 The ORR activity of Pt(111) was enhanced after the durability test because active (110) type defects are formed on the surface. The ORR activity of Pt(111) modified with OA/PA was 1.8 mA/cm2 after 2000 cycles, higher than that of bare Pt(111) 1.0 mA/cm2. Adsorbed amines on wide (111) terrace preserved the high activity after the durability test. The activity after 2000 cycles decreased with the increase of the ratio of PA. This result differs from the previous study [1]. The ORR activity of Pt(775) modified with OA/PA was as high as that of bare Pt(775) after the durability test.  Durability of Pt(775) was not improved by OA/PA modification.

 

4. Acknowledgement

 This work was supported by New Energy Development Organization.

 

5. References

[1] K. Miyabayashi, H. Nishihara, M. Miyake, Langmuir, 30, 2936 (2014)

[2] N. Hoshi, M. Nakamura, A. Hitotsuyanagi, Electrochim. Acta, 112, 899 (2013)

2637

, and

1. Purpose

 Polymer electrolyte fuel cells (PEFCs) attract attention as a clean and highly efficient energy system. It is necessary to activate the oxygen reduction reaction (ORR) for the wide spread of fuel cells because of the high cost and the limitation of natural resources of Pt.

 One of the strategies for the enhancement of the ORR activity is modification of the Pt electrode surfaces. Kodama et al. reported that the ORR activity of Pt(322) = 5(111)-(100) is enhanced by Au-deposition at the terrace edge [1]. Feng et al. revealed that metal-porphyrin and its metal-organic frame (MOF) have high ORR activity [2]. Higher density of d-band vacancy also enhances the ORR activity [3]. Cobalt tetraphenyl porphyrin (CoTPP) has electron attracting property to increase the density of d-band vacancy; improvement of ORR activity is expected.

 In this paper we have studied the ORR activity on Pt high index planes modified with cobalt tetraphenyl porphyrin (CoTPP) and Au.

2. Experimental

 Pt(322) = 5(111)-(100) and Pt(553) = 5(111)-(111) electrodes, were prepared by Clavilier's method. Both electrodes have 5 atomic rows of (111) terrace, but the step structure of Pt(322) is different. from that of Pt(533) (Pt(322):(100)-step, Pt(553):(111)-step).

 Au was deposited on Pt surface according to the method of Kodama et al. [1], CoTPP modification was done according to Itaya et al. [4]. Linear sweep voltammograms of the ORR were measured using rotating disk electrode (RDE) in 0.1 M HClO4 saturated with O2. Potential was scanned from 0.05 V (RHE) up to 1.0 V (RHE) at scanning rate 0.010 V s−1 and rotation rate 1600 rpm. The activity for the ORR is estimated using the specific activity jkat 0.90 V (RHE).

3. Results and discussion

 Voltammograms of bare and Au-modified Pt(322) and Pt(553) are shown in Fig. Redox peaks due to the adsorption/desorption of hydrogen at the terrace edge appear at 0.27 and 0.12 V (RHE) on Pt(322) and Pt(553), respectively. Other broad peaks are due to the hydrogen desorption/adsorption at terrace. The peaks due to the step disappear after Au deposition on Pt(322), whereas the peaks due to the terrace are intact. These results suggest that Au atoms were deposited selectively at terrace edge. On Pt(553), however, the peaks due to the step and terrace shrink after the Au deposition, suggesting that both step and terrace are modified by Au atoms. Selectivity of Au modification depends on the step structure of Pt substrate.

 The ORR activity of Pt(322) increases from 2.8 mA cm-2 to 3.1 mA cm-2 after the Au deposition, however that of Pt(553) decreases from 2.4 mA cm-2 to 0.9 mA cm-2. Au deposition on the terrace prevents the ORR on Pt(533). This fact indicates that Au deposition at only terrace edge is necessary for the enhancement of ORR activity.

 Au deposited Pt(322) was further modified by CoTPP. The ORR activity decreases from 3.1 mA cm-2 to 2.3 mA cm-2after the modification of CoTPP. Five atomic rows of (111) terrace is as large as the size of CoTPP; CoTPP may block the active sites for the ORR at the terrace. It is necessary to investigate the ORR on surfaces with wider terrace.

4. Acknowledgement

This work was supported by New Energy Development Organization (NEDO).

5. References

[1] K. Kodama, R. Jinnouchi, N. Takahashi, H. Murata, Y. Morimoto, J. Am. Chem. Soc. 10.1021, jacs.6b00359 (2016).

[2] P. Feng, Q.Lin, X.Bu, A.Kong, C.Mao, X.Zhao, F.Bu, J. Am. Chem. Soc. 137, 2235 (2015).

[3] S. Mukerjee, S. Srinivasan, M. P. Soriaga, J. Electrochem. Soc. 142, 1409 (1995).

[4] S. Yoshimoto, A. Tada, K. Suto, S. Yau, K. Itaya, Langmuir. 20, 3159 (2004).

Figure 1

2638

, , , and

Pt-based alloys such as Pt-Fe, Pt-Co and Pt-Ni, have attracted much attention as cathode catalysts with enhanced activities for the oxygen reduction reaction (ORR) in polymer electrolyte fuel cells (PEFCs).1 Recently, we have found by use of an electrochemical quartz crystal microbalance (EQCM) and a rotating disk electrode (RDE) that the specific adsorption of perchlorate anion on a Pt-skin/Pt-Co alloy polycrystalline electrode was stronger than that on bulk-Pt, and that the kinetically-controlled current densities jk for the ORR at the Pt-skin/Pt-Co decreased with increasing HClO4 concentration more steeply than that in the case of bulk-Pt, due to the blocking of the active sites by the specifically adsorbed perchlorate anion. 2

In practical PEFCs, Pt-based nanoparticles dispersed on carbon supports are employed in the catalyst layer, and the nanocatalyst surfaces usually consist of low-index crystal facets such as (111) and (100). Recently, we have established a facile method for preparation of Pt-Co alloy single crystals with desired Co contents,3 and have revealed the strong dependence of the ORR activity jk on the Co content at Pt-skin/Pt100-xCox(111) electrodes prepared by annealing under pure H2 atmosphere.4 In the present research, we have evaluated jk at well-defined Pt-skin/Pt-Co alloy single crystal electrodes as a function of HClO4 concertation [HClO4] by using the rotating disk electrode (RDE) method in order to clarify the effect of specific adsorption of perchlorate on the ORR activity.

The preparation of Pt-Co alloy single crystal electrodes has been described elsewhere.3 Prior to the electrochemical measurements, well-defined (1 × 1) surfaces of polished crystal electrodes were freshly prepared by heating to 1170 K in H2, and subsequent cooling in H2 for the (111) and (100) planes or 1% CO/He streams for the (110) plane.3 This heat treatment resulted in the formation of a Pt-skin layer on the Pt-Co alloy single crystal surface with the same atomic arrangement as the underlying crystal, confirmed by low energy ion scattering (LEIS) and low energy electron diffraction (LEED).4

Figure 1 shows cyclic voltammograms (CVs) at a Pt-skin/Pt81Co19(111) electrode acquired in various concentrations of HClO4 purged with N2. The ohmic loss was corrected based on impedance measurements. The CVs exhibited characteristic, highly reversible features, i.e., the hydrogen underpotential deposition (HUPD) wave and the so-called butterfly wave, with a broad peak around 0.85 V, in the surface-oxidation region. At 0.05 M ≤ [HClO4] ≤ 0.2 M, no significant changes were observed in the HUPD region. In contrast, with increasing [HClO4], it was found that the double layer currents increased and the surface oxidation wave shifted to higher potentials due to specific adsorption of perchlorate anion, similar to the polycrystalline Pt-Co alloy.2

Figure 2 shows jk values of Pt73Co27(111)-RDEs at 0.9 V vs. RHE in air-saturated HClO4 solutions as a function of [HClO4]. As [HClO4] increased from 0.05 to 0.2 M, the jk values decreased, similar to the case of the polycrystalline Pt-Co alloy.2 The decrease in jk can be ascribed to the blocking of the ORR active sites on the Pt-skin layer by perchlorate anions. However, the jk value in the 0.01 M HClO4 solution was significantly lower than those for other concentrations. The lower jk value in 0.01 M HClO4was presumably due to the low activity of protons, which are consumed in the ORR on the surface.

This work was supported by the funds for the "SPer-FC Project'' of NEDO (Japan) and a Grant-in-Aid No. 25410007 for Scientific Research MEXT of Japan.

References

1. T. Toda, H. Igarashi, H. Uchida, and M. Watanabe, J. Electrochem. Soc., 146, 3750 (1999).

2. J. Omura, H. Yano, D. A. Tryk, M. Watanabe, and H. Uchida, Langmuir, 30, 432(2014).

3. M. Wakisaka, Y. Hyuga, K. Abe, H. Uchida, M. Watanabe, Electrochem. Commun, 13, 317 (2011).

4. M. Wakisaka, S. Kobayashi, S. Morishima, Y. Hyuga, D. A. Tryk, M. Watanabe, A. Iiyama, and H. Uchida, Electrochem. Commun, 67, 47 (2016).

Figure 1

2639

, , and

Introduction

Decrease of Pt usage in the PEFCs is essential for their worldwide commercialization. Pt-M alloy catalyst (M: 3d transition metals such as Fe, Co, Ni and Cu) is an attractive candidate for the decrease due to its high ORR activity [1, 2]. However, since the 3d transition metals M have lower redox potentials than the Pt, the metals M dissolve-out under the PEFC's cathode condition, which decays the ORR activity of the Pt-M alloy catalysts. It has been reported that the decay of the ORR activity evaluated at room temperature (e.g., 25oC) can be suppressed with a crystallographic ordering of the catalyst [3]. Thus, in this study, we synthesized disordered and L10 ordered Pt-Co alloy catalysts and investigated their durability at PEFCs' working temperature of 80oC using an accelerated durability test.

Experimental

Carbon supported PtCo alloy catalysts (PtCo/C) were synthesized by the following procedures. 0.5 mmol of Pt(acac)2 and 0.5 mmol of Co(acac)2 were dissolved in 50 ml of oleylamine (OAm) at room temperature and the solution was heated to 300oC under N2 atmosphere and kept for 1 h with stirring [4]. After the solution was cooled to room temperature, PtCo NPs were precipitated by adding 40 ml of ethanol and separated by centrifugation, followed by re-dispersion in n-hexane. These processes were repeated for 3 times to wash the PtCo NPs. Carbon support, Ketjen black EC-300J, was added to the n-hexane solution and stirred overnight to load the PtCo NPs, followed by filtration and drying. As-synthesized PtCo/C catalyst was heated at 400oC for 4 h to remove the OAm, followed by heating at 700oC for 1 h and 900oC for 1 h to promote crystallographic ordering under 15%-H2/Ar atmosphere. PtCo/C catalysts were characterized with TG-DTA, XRF, XRD, TEM, TEM-EDX and CV. ORR activity of the catalysts was evaluated by RDE technique in O2 saturated 0.1 M HClO4 at 25oC. Accelerated durability test (ADT) was performed at 25oC and 80oC using a rectangular wave potential cycling of 0.6 V (3 s)-1.0 V (3 s) vs. RHE in Ar saturated 0.1 M HClO4 for 10,000 cycles.

Results and Discussion

XRF analysis revealed that bulk composition of the PtCo/C catalysts was Pt46Co54 (at.%) and TG analysis showed that metal loading was ca. 30 wt.%. XRD patterns of the PtCo/C catalysts are depicted in Fig. 1. The PtCo/C catalyst heated at 400oC showed a typical face-centered cubic (fcc) structure with disordered phase (hereafter, called as 400oC-Disorder). The broad peak at around 25o is attributed to the carbon support. The other four diffraction peaks are consistent with the fcc structure, corresponding to (111), (200), (220) and (311) planes. After heating at 700oC and 900oC, the XRD patterns transformed to an ordered intermetallic L10 Pt-Co phase (black triangles) (hereafter, called as 700oC-Order and 900oC-Order, respectively). TEM images of the PtCo/C catalysts are shown in Fig. 2. Mean diameters of 400oC-Disorder, 700oC-Order and 900oC-Order PtCo NPs were 5.4 nm, 5.8 nm and 6.0 nm, respectively, indicating that the mean diameter increased with increase of the heating temperature.

It is well known that high ORR activity of the Pt-M alloy catalysts arises from ligand effect with the 3d transition metals M existing underneath of the topmost Pt shell in the catalysts [1, 2]. Therefore, durability of the Pt-M alloy catalysts strongly depends on dissolution of the M atoms under PEFC's cathode condition. The Co dissolution from the PtCo/C catalysts with ADT performed at 25oC and 80oC is demonstrated in Fig. 3. The 400oC-Disorder PtCo/C catalyst lost all Co with the ADT. On the contrary, the 900oC-Order PtCo/C catalyst retained ca. 30% Co after the ADT performed at 80oC, indicating that the durability of the PtCo/C catalyst was enhanced with the crystallographic ordering from the disordered fcc phase to the ordered L10 phase. After the ADT, the 900oC-Order PtCo/C catalyst showed higher ORR mass activities than those of the 400oC-Disorder PtCo/C catalyst, also showing higher durability of the 900oC-Order PtCo/C catalyst (Fig. 4, black dotted line indicates a reference Pt/C catalyst; Pt mean diameter: 5.1 nm, Pt loading: 51 wt.%, TEC10E50E-HT, TKK). At the meeting, size effect of the L10 ordered PtCo/C catalyst on the durability will be also presented.

Reference

[1] T. Toda et al., J. Electrochem. Soc., 146, 3750 (1999).

[2] V. R. Stamenkovic et al., Nat. Mater., 6, 241 (2007).

[3] D. Wang et al., Nat. Mater., 12, 81 (2013).

[4] Y. Yu et al., Nano Lett., 14, 2778 (2014).

Figure 1

2640

, , , and

Development of cathode catalysts with high oxygen reduction reaction (ORR) activity and high durability is one of the most important issues for the large-scale commercialization of polymer electrolyte fuel cells (PEFCs). To obtain high mass activity (MA), Pt alloy nanoparticles dispersed on high-surface-area carbon support have been prepared. However, the durability of conventional PtM alloys (M = Fe, Co, Ni) have been insufficient due to an appreciable dealloying of M component at high temperatures (> 60 °C).1, 2 Recently, to increase both the MA and durability, we have succeeded in preparing new catalyst with two atomic layer of Pt skin (stabilized Pt skin layer) formed on PtCo-core alloys nanoparticles supported on graphitized carbon black (GCB; 150 m2 g-1) 2, 3 and high surface area carbon black (C; 800 m2 g-1).4

The preparation is briefly described below.3, 4 The PtCo alloy (1:1 atomic ratio) nanoparticles with uniform size and composition were highly dispersed on GCB or C was performed by the nanocapsule method.5 The PtCo/GCB or PtCo/C was heat-treated in H2 atmosphere in order to form nearly one atomic layer (Pt1AL) of Pt skin on the surface of the PtCo nanoparticles (denoted as Pt1AL–PtCo) by the segregation of Pt atoms from interior of the alloy core. Then, one more atomic layer of Pt-skin was formed on Pt1AL–PtCo (denoted as Pt2AL–PtCo) via controlled Pt deposition from a Pt-complex-containing aqueous solution with H2 bubbling. Finally, the Pt2AL–PtCo/GCB or Pt2AL–PtCo/C catalysts thus obtained were heat-treated in H2 at 200 oC.

The kinetically-controlled mass activity (MAk) and specific activity (jk) for the ORR at Nafion-coated Pt2AL–PtCo/GCB and Pt2AL–PtCo/C electrodes were examined in O2-saturated 0.1 M HClO4 solution at 65 oC by multi-channel flow double electrode cell.1 The accelerated durability test (ADT) was examined by a standard potential step protocol (E = 0.6 V ↔ 1.0 V vs. RHE, holding 3 s at each E) simulated the load-change for the fuel cell vehicle (FCV)6 in N2-purged 0.1 M HClO4 solution at 65 oC.

Figure 1 shows the plots of MAk and jk values at Nafion-coated Pt2AL–PtCo/GCB and Pt2AL–PtCo/C electrodes as a function of the potential step cycles (N). The initial values of jk at Pt2AL–PtCo/GCB (dTEM = 2.9 ± 0.2 nm) and Pt2AL–PtCo/C (dTEM = 3.3 ± 0.5 nm) were identical with that of PtCo/GCB, which was 2.5 times larger than that of a commercial standard catalyst c-Pt/C. The value of MAk at c-Pt/C decreased severely to 10% of the initial value after N = 30,000, whereas the jk was nearly unchanged. The MAk and jk values at PtCo/GCB were drastically decreased with increase N and were equal to that for c-Pt/CB after N = 30,000 due to the dealloying of Co. In contrast, the value of jk at the Pt2AL–PtCo/GCB electrode was maintained constant value during the ADT, and the retention of MAk at N = 30,000 was as high as 68%. A similar trend was seen for the Pt2AL–PtCo/C, at least, up to N= 10,000. Thus, the dealloying of Co was almost completely suppressed by the stabilized Pt skin layer on the surface regardless of GCB and C support.

We have also succeeded in preparing the Pt2AL–PtM alloy (M = Fe and Ni) dispersed on C support by the same method as described above. The average particle sizes dTEM of the Pt2AL–PtFe/C and Pt2AL–PtNi/C were 2.9 ± 0.4 nm and 3.2 ± 0.4 nm, respectively. The Pt2AL–PtFe/C and Pt2AL–PtNi/C also exhibited high MAk. We will present the effect of alloy components on the ORR activity and durability.

This work was supported by the funds for the "SPer-FC" project from the NEDO of Japan.

References

1. H. Yano, J. M. Song, H. Uchida, M. Watanabe, J. Phys. Chem. C,112, 8372 (2008).

2. K. Okaya, H. Yano, K. Kakinuma, M. Watanabe, and H. Uchida, ACS Appl. Mater. Interfaces, 4, 6982 (2012).

3. M. Watanabe, H. Yano, D. A. Tryk, and H. Uchida, J. Electrochem. Soc., 163, F455 (2016).

4. M. Chiwata, H. Yano, S. Ogawa, M. Watanabe, A. Iiyama, H. Uchida, Electrochemistry, 84, 133 (2016).

5. H. Yano, M. Kataoka, H. Yamashita, H. Uchida, and M. Watanabe, Langmuir, 23, 6438 (2007).

6. http://fccj.jp/pdf/23_01_kt.pdf

Figure 1

2641

, , , and

Polymer electrolyte fuel cells (PEFCs) have been attracting much attention as a clean and efficient power source. However, the degradation of polymer electrolyte membranes (PEMs), such as perfluorosulfonic acid or sulfonated hydrocarbon, made the cell performance decrease via not only a decrease in the proton conductivity but also a decrease in the cathode performance by the decomposition products. In our previous work, we have investigated the effect of the specific adsorption of sulfuric acid (a typical decomposition product from PEMs) in the electrolyte solution on the ORR activity at Nafion-coated commercial Pt/C (c-Pt/C) electrode at 30 to 80 oC by the channel flow double electrode (CFDE) method.1 Very recently, we have developed a new cathode catalyst, by forming two uniform atomic layers of stabilized Pt skin on PtCo nanoparticles supported on graphitized carbon black or high-surface-area carbon black (Pt2AL–PtCo/GCB or Pt2AL–PtCo/C), with high ORR activity and high durability at high temperature.2, 3 In the present research, we have investigated the effect of sulfate anion on the ORR activity at the Pt2AL–PtCo/C electrode in 0.1 M HClO4 supporting electrolyte solution in the practical temperature range from 30 to 80 °C.

The Pt2AL–PtCo/C catalyst (30 wt%-metal loading, particle size d = 3 nm) was prepared in the same manner as described previously.2, 3 The kinetically-controlled ORR activities at Nafion-coated Pt2AL–PtCo/C working electrode was evaluated from the hydrodynamic voltammograms in O2-saturated 0.1 M HClO4 + X mM H2SO4 (X = 10-3, 1, 5, and 50) solution by using the CFDE cell.1 At the collecting electrode located downstream of the working electrode, the H2O2 yield, P(H2O2), was quantified.

 Figure 1 shows the values of P(H2O2) at 0.76 V vs. RHE on the Nafion-coated Pt2AL–PtCo/C and c-Pt/C as a function of log [H2SO4]. The P(H2O2) values on both Pt2AL–PtCo/C and c-Pt/C increased with increasing [H2SO4]. This indicates that the specific adsorption of sulfate anion was the major factor for increasing H2O2 formation on c-Pt/C or Pt2AL–PtCo/C catalysts. However, the values of P(H2O2) at the Pt2AL–PtCo/C were markedly lower than that of c-Pt/C at all temperature. Especially, the P(H2O2) at 80 oC and [H2SO4] ≤ 50 mM on the Pt2AL–PtCo/C was nearly identical value to that in sulfate-free solution. Thus, the use of Pt2AL–PtCo/C in place of c-Pt/C provides a great advantage of suppressing the H2O2generation appreciably, resulting in a mitigation of the decomposition of the PEMs.

 Figure 2 shows plots of the apparent rate constant kapp per active surface area of Nafion-coated Pt2AL–PtCo/C and c-Pt/C electrodes for the ORR at 0.80 V vs RHE as a function of log [H2SO4]. In sulfate-free 0.1 M HClO4 solution, the kapp values at Pt2AL–PtCo/C electrode were ca. 2-3 times higher than those of c-Pt/C electrode at all temperatures examined. The kapp values at [H2SO4] = 1 μM on both electrodes approximately accord with those obtained in sulfate-free solution, followed by the decrease linearly with log [H2SO4] in the temperature range between 50 and 80 °C. Such a behavior can be explained well with the Frumkin–Temkin adsorption isotherm for sulfate anions.1, 4 In the high-temperature range (≥ 70 °C), the activity loss on the Pt2AL–PtCo/C was greatly suppressed, i.e. the loss of kapp value at 80 °C and [H2SO4] = 50 mM was only 7 % in contrast with the case of 50°C (50 %). This is consistently explained by a weakening of the adsorption of sulfate at high temperature on the Pt2AL–PtCo alloy catalysts, probably due to a modified electronic structure.

 Hence, we have proposed with certainty that a short-term high current density operation is effective to recover the performance loss suffered from the specific adsorption of sulfate anions, because both large amounts of water and heat are produced.

 This work was supported by the funds for "SPer-FC" projects from the New Energy and Industrial Technology Development Organization (NEDO) of Japan.

References

  • H. Yano, T. Uematsu, J. Omura, M. Watanabe, and H. Uchida, J. Electroanal. Chem., 747, 91 (2015).

  • M. Watanabe, H. Yano, D. A. Tryk, and H. Uchida, J. Electrochem. Soc., 163, F455 (2016).

  • M. Chiwata, H. Yano, S. Ogawa, M. Watanabe, A. Iiyama, and H. Uchida, Electrochemistry, 84, 133 (2016).

  • J. Omura, H. Yano, D. A. Tryk, M. Watanabe, and H. Uchida, Langmuir, 30, 432 (2014).

Figure 1

2642

, , , , and

Enhancing the activity and stability of cathode catalysts in proton exchange membrane fuel cells (PEMFCs) is of great importance for their widespread commercialization. While nanostructured Pt-based nanoparticles supported on carbon have been the most active cathode catalysts, their long-term durability and their implementation n single cells represent challenges. In this work, we report the self-supported, mesostructured Pt-based bimetallic (Meso-PtM; M=Ni, Fe, Co, Cu) nanospheres containing intermetallic phase, which can combine the beneficial effects from transition metal (M), intermetallic phase, interconnected framework, and porous structure. All Meso-PtM nanospheres showed enhanced ORR activity, compared to Pt black and Pt/C catalysts. Particularly, Meso-PtNi with intermetallic phase exhibited ultrahigh stability, showing enhanced ORR activity even after 50,000 potential cycling, whereas Pt/C underwent a dramatic degradation. Importantly, the superior performance of Meso-PtNi with intermetallic phase was also demonstrated in a PEMFC single cell. The initial mass activity of Meso-PtNi far exceeded that of Pt/C and the US DOE target value, and represents one of the best activities among Pt nanocatalyst-based PEMFCs.

2643

, and

Introduction

Although Proton exchange membrane fuel cells (PEMFCs) are clean and efficient energy converters, the amount of platinum used as the catalyst for the oxygen reduction reaction (ORR) at the cathodic electrode is still large and has to be reduced to drive down cell manufacturing costs. Core-shell structured Pt alloys with relatively abundant transition metals, such as Cu, Co, and Ni, have demonstrated significant improvements for the ORR activities, suggesting one solution to tackle this problem1). However, it has been reported that Pt-M alloys catalysts degrade more easily than pure Pt because of selective dissolution of M, leading to severely reduced performance in long run operation2). Taylor et al. have shown that Pt/C-catalyzed GDL fabrication using the electrodeposition method enables the high Pt utilization ratio in the catalyst layer because it ensures that Pt is deposited only on the electrochemically active area3). Woo et al. have reported considerable single cell performance of Pt-Co catalysts prepared on unloaded GDL using the electrodeposition method4). However, more investigations are needed to clarify their corrosion and degradation behaviors.

In this work, we present the preparation of Pt-Cu catalyst particles using the electrodeposition method, and the evaluation of durability and ORR activity using electrochemical measurements such as immersion test and ORR measurement.

 

Experimental

Glassy Carbon was used as a substrate on which Pt-Cu nanoparticles were deposited. Direct current electrodeposition of Pt-Cu nanoparticles was conducted in a conventional three-electrode cell. A double junction KCl-saturated Ag / Ag-Cl electrode was used as a reference electrode, and a carbon rod was used as the counter electrode. Electrodeposition was carried out in a plating bath containing a solution of 6 mM K2PtCl4 and various concentrations(1-40 mM) of CuSO4・5H2O dissolved in 0.5M Na2SO4. The parameters for direct current electrodeposition were an electrodeposition potential of -0.05 V vs. SHE and a total charge density of 0.11 C cm−2.

Chemical compositions and particle diameters were determined by SEM-EDX. The duration of the immersion test was 3 h in 0.5 M H2SO4, followed by the investigation of the dissolved amount of Cu using ICP-MS. ORR activities were determined by measuring the potentiostatic polarization curve in 0.5 M H2SO4 from OCP to 0.6 V vs. SHE at a scan rate of 0.2 mV s-1. All electrodeposition processes and electrochemical measurements were conducted open to the atmosphere and at 25 oC.

 

Results and Discussion

SEM-EDX observations confirmed that various compositions of Pt-Cu were prepared and homogeneous nanoparticles were obtained in each sample. Chemical compositions were controlled from 13 to 85 at. % Cu by changing the concentration of precursor in a plating bath.

The amounts of dissolved Cu ions during the immersion test drastically increased with atomic ratio of Cu as shown in Fig. 1(a). The corrosion potential monitored during the immersion test, all Pt-Cu catalysts formed the Pt enriched layer on the surface by selective dissolution of Cu. However, Cu-rich alloys did not suppress the dissolution of Cu. In addition, SEM observations after the immersion test displayed that the surface morphology of Cu-rich Pt-Cu nanoparticles completely changed.

Fig. 1(b) shows specific activities at 0.9 V initial against Cu composition. Pt58Cu42 indicated the highest ORR activity, 2.6 times higher than pure Pt. Fe group metal content enhances Oxygen adsorption and weakens O-O bond5) and thus ORR activity is enhanced by the effect of alloying when Cu content increases. From our results, however, the Pt enriched layer which weakens the effect of alloying becomes thicker as Cu content increases because the amount of dissolved Cu ions also increases. As a result, optimum composition is determined by the balance between the effect of alloying and the thickness of the Pt enriched layer.

Pulse current electrodeposition of Pt-Cu nanoparticles on GDL will be further discussed in the session.

References

1) Mehtap Oezaslana, Peter Strasser, Journal of Power Sources196 (2011) 5240–5249.

2) S. Mukerjee, S. Srinivasan, Handbook of Fuel Cells, vol2 (2003) 502.

3) E.J. Talyor, E.B. Anderson, N.R.K. Vilambi, J. Electrochem. Soc.139 (1992) L45.

4) Seunghee Woo, In Kim, Jae Kwang Lee, Sungyool Bong, Jaeyoung Lee, Hasuck Kim, Electrochim. Acta56 (2011) 3036–3041.

5) T. Toda, H. Igarashi, H. Uchida, and M. Watanabe, J. Electrochem. Soc.146 (1999) 3750.

Figure 1

2644

, , , and

Finding a highly active and durable nanocatalyst under routinely used proton exchange membrane fuel cell operating conditions (60 - 80 oC) is urgent for the commercialization of fuel cells. Many nanocatalysts have been reported to have much improved oxygen reduction reaction (ORR) activity compared to the commercially used Pt/C catalyst. However, few of them can meet the durability requirement. In this report, monodisperse face-centered-tetragonal (fct) (L10) FePt nanoparticles with size of ~ 9 nm were synthesized from dumbbell-like face-centered-cubic (fcc)-FePt-Fe3O4 nanoparticles. The durability of C-fct-FePt nanoparticles as ORR catalysts at 60 oC was studied and it was found that these nanoparticles served as a highly active and durable ORR catalyst. Its mass activity was found to be about 5.4 times that of commercial Pt/C. Also, no obvious loss in ORR activity was observed after 10,000 cycles. A perchloric acid-treatment and subsequent annealing treatment were found to improve the durability of fct-FePt nanoparticles at 60 oC. The highly ordered L10 structure, with alternating layers of Fe and Pt atoms, as well as the Pt skin formed after the acid-treatment, were hypthesized to be the main reason for the durability enhancement. Our work demonstrates a reliable approach to structurally ordered FePt nanoparticles as a promising candidate for practical use in fuel cells.

2645

and

Carbon-supported PdFe nanoparticles (NPs) were prepared with composition-controlling via a modified chemical synthesis and a heat-treatment at high temperature under a reductive atmosphere. The synthesis combined the polyol reduction method and hydride reduction method, which was used to obtain monodispersed PdFe NPs. In addition, to induce structural modifications, the as-prepared PdFe NPs received heat-treatment under a reductive atmosphere. Results of structural characterizations including highresolution powder diffraction (HRPD), X-ray photoelectron spectroscopy (XPS), and X-ray absorption spectroscopy (XAS) analysis, indicated that heat-treated PdFe NPs exhibited a higher degree of alloying and surface Pd atomic composition compared with as-prepared ones. Furthermore, new crystalline phases were detected after heat-treatment. Thanks to the structural alterations, heat-treated PdFe NPs showed ~3 and ~18 times higher mass- and area-normalized oxygen reduction reaction (ORR) activities, respectively than commercial Pt/C. Single cell testing with heat-treated PdFe catalysts exhibited a ~2.5 times higher mass-normalized maximum power density than the reference cell. Surface structure analyses, including cyclic voltammetry (CV), COad oxidation, and XPS, revealed that, after heat-treatment, a downshift of the Pd d-band center occurred, which led to a decrease in the affinity of Pd for oxygen species, resulting in more favorable ORR kinetics.

2646

, , , , and

Introduction

Carbon supported Pd core-Pt shell catalyst (Pt/Pd/C) is a promising alternative to the conventional Pt/C catalyst because of high Pt utilization and enhancement of ORR activity [1]. Recently, we found that ORR specific activity of the Pt/Pd/C catalyst was drastically enhanced with an accelerated durability test (ADT) conducted at 80°C [2]. During the ADT, the Pt shell rearranged associated with the Pd core dissolution and a compressive strain was induced in the Pt shell, which is considered to enhance the ORR specific activity [2]. However, ECSA of the catalyst decreased after the ADT. Thus, we developed a high activation protocol (HAP) using GC electrode to mitigate the ECSA decay and enhance ORR mass activity [3]. In this study, we explored influence of potential range in the HAP on ECSA decay and ORR activity enhancement of the Pt/Pd/C catalyst. Furthermore, we developed a Cu-O2 treatment to scale-up the HAP on the GC electrode for mass-production of highly activated Pt/Pd/C catalyst.

 

Experimental

Pt/Pd/C catalyst was synthesized with a modified Cu-UPD/Pt replacement process [2]. A carbon supported Pd core (Pd/C, Pd size: 4.6 nm, Pd loading: 30 wt.%, Ishifuku Metal Industry) was stirred in 50 mM H2SO4 containing 10 mM CuSO4 with co-existence of a metallic Cu sheet at 5°C under Ar atmosphere. After stirring for 5 h, the Cu sheet was removed and K2PtCl4 was added to replace under potentially deposited Cu shell on the Pd core surface with the Pt shell. ADT was carried out using rectangular wave potential cycling of 0.6 (3 s)-1.0 V (3 s) vs. RHE in Ar saturated 0.1 M HClO4 at 80℃ for 10,000 cycles. The Pt/Pd/C catalyst was characterized by TG, XRF, XRD, TEM and CV. ORR activity of the catalyst was evaluated with RDE technique in O2 saturated 0.1 M HClO4 at 25°C.  

Results and Discussion

We explored influence of potential range in the HAP on electrochemical properties of the Pt/Pd/C catalyst using rectangular potential cycling of 0.05~0.8 V (300 s for low potential) to 1.0 V (300 s for high potential) vs. RHE performed in Ar saturated 0.1 M HClO4 at 80℃ for 30 cycles. Changes in ECSA and ORR activity of the catalyst are demonstrated in Fig. 1. In comparison to ADT, ECSA decay of the Pt/Pd/C catalyst was mitigated with the HAP. Interestingly, ORR specific activity of the catalyst was largely enhanced when the low potential range was 0.2~0.6 V, which largely enhanced ORR mass activity of the catalyst. At the low potential range (0.2~0.6 V), it is considered that Pt shell rearrangement associated with Pd core dissolution was advanced due to sufficient oxidation/reduction of Pt and Pd. At low potential of 0.05 V, since hydrogen adsorbs on the Pt, it is considered that hydrogen adsorption hindered rearrangement of the Pt shell [4] and the ORR activity was not largely enhanced. We further developed a Cu-O2 treatment to scale-up the HAP performed on GC electrode. In the Cu-O2 treatment, the Pt/Pd/C catalyst powder (200 mg) is stirred at 80℃ for 300 s in 2 M H2SO4 containing 0.1 M CuSO4 with co-existence of a metallic Cu sheet under N2 atmosphere, where equilibrium potential of Cu2+/Cu (ca. 0.3 V) is applied to the catalyst powder when it contacts with the Cu sheet. Next, the Cu sheet is removed and O2 gas is introduced for 300 s, where equilibrium potential of ORR (ca. 1.0 V) is applied to the catalyst (Fig. 2). Figure 3 summarizes changes in ECSA and ORR mass activity of the Pt/Pd/C catalyst with the HAP and the Cu-O2 treatment (30 cycles). Compared with the HAP, the Cu-O2 treatment equivalently mitigated ECSA decay and enhanced ORR mass activity of the catalyst by ca. 3 times of reference carbon supported Pt catalyst (Pt/C, Pt size: 2.8 nm, Pt loading: 46 wt.%, TEC10E50E, TKK), indicating that the Cu-O2 treatment mimics the HAP on the GC electrode and is suitable for mass-production of highly activated Pt/Pd/C catalyst. 

Acknowledgement

This work was supported by NEDO, Japan.

 

References

[1] A. U. Nilekar et al., Top Catal., 46, 276 (2007).

[2] M. Inaba and H. Daimon, J. Jpn. Petrol. Inst., 58(2), 55 (2015).

[3] H. Daimon et al., The 228th Electrochemical Society Meeting, #1375, Phoenix, USA, October 2015.

[4] Q. Xu et al., J. Electrochem. Soc., 155(3), B228 (2008).

Figure 1

2647

, , , , and

High cost, sluggish kinetics, and poor durability of Pt catalysts predominantly hinder the wide-spread commercialization of fuel cells. Pt-based core-shell nanosheet catalyst may resolve these problems listed above because of the following 3 advantages. 1) The Pt content can be greatly reduced via core−shell nanostructures consisting of a Pt shell on appropriate monometallic or alloy cores; 2) Nanosheets have high surface-to-volume ratio and terrace sites. The active sites for the oxygen reduction reaction (ORR) have been suggested to be located on the terrace sites of the nanocrystals.[1] 3) Nanosheets should dramatically reduce Pt dissolution due to few edges and corners with their low coordinate sites. Terraced facets have been reported to be more stable than edges and corners.[2]

Here, we present a novel method to synthesize palladium (Pd) nanosheets by a wet-chemical preparation at room temperature (Fig.A). The Pd precursor (Pd(acac)2) is reduced by CO in the presence of decylamine (DA). The protection agent (DA) was easily removed by washing with acetic acid. Using this nanostructure as a core, Pd@Pt core-shell nanosheets were prepared by surface limited redox replacement (SLRR). AFM indicates Pd nanosheets are synthesized with a thickness around 0.9 nm to 1.6 nm and lateral size from tens to several hundred nanometers (Fig.B). After dispersing the Pd nanosheet on a carbon support, Pd@Pt core-shell nanosheets with different Pt shell thickness were prepared, i.e. Pd6.0ML@Pt2.7ML, Pd6.0ML@Pt4.4ML, and Pd6.0ML@Pt5.9ML. The electrochemically active surface area (ECSA) of Pd@Pt nanosheets are 56, 48, 60 m2/g-PGM (90, 66, 77 m2/g-Pt), respectively (Fig. C). The linear sweep voltammetry is applied by using a rotating disc electrode (RDE) in 0.1 M HClO4 at room temperature. The ORR activity is evaluated from jk values at 0.9 V (vs RHE) estimated from Koutecky-Levich Plots. The Pd@Pt nanosheets demonstrated mass activity of 364, 685, 598 A/g-Pt (227, 501, and 492 A/g-PGM), which is much higher than the benchmark Pt/C catalyst (177 A/g-Pt) for the oxygen reduction (Fig. D).

This research was supported in part by the "Polymer Electrolyte Fuel Cell Program" from the New Energy and Industrial Technology Development Organization (NEDO) of Japan.

Fig. A) TEM of Pd@DA nanosheets; B) AFM image of Pd nanosheets; C) CV profiles of the Pd@Pt nanosheets recorded in N2-saturated 0.1 M HClO4 solution at a sweep rate of 50 mV/s; and D) The mass activity of Pd@Pt nanosheets and commercial Pt/C.

Reference

  • F. J. Perez-Alonso, D. N. McCarthy, A. Nierhoff, P. Hernandez-Fernandez, C. Strebel, I. E. L. Stephens, J. H. Nielsen, I. Chorkendorff, Angew. Chem. Int. Ed.2012, 51, 4641–4643.

  • F. N. Büchi, M. Inaba, T. J. Schmidt, Polymer Electrolyte Fuel Cell Durability, Springer Science+ Business Media, New York, 2009.

Figure 1

2648

, , , and

Introduction

It is of great importance to decrease Pt usage in the polymer electrolyte fuel cells (PEFCs) for their cost reduction. Pd core-Pt shell catalyst (Pt/Pd/C) is a promising candidate for the decrease due to its high Pt utilization efficiency and enhancement of oxygen reduction reaction (ORR) activity [1, 2]. Recently, we have reported that ORR specific activity of the Pt/Pd/C catalyst is drastically enhanced with an accelerated durability test (ADT) [3]. It is considered that the enhancement arises from rearrangement of the Pt shell associated with the Pd core dissolution, in which number of low-coordinated surface Pt atoms decreased and a compressive strain was induced in the Pt shell [3].In our previous study, the Pt/Pd/C catalysts were successfully synthesized with a modified Cu-UPD/Pt replacement process in a large-scale (50-100 g/batch) [4]. However, the degree of the ORR activity enhancement after the ADT had poor reproducibility although their initial physical and electrochemical properties were almost equivalent. In this study, structural change of the Pt/Pd/C catalysts with the ADT was analyzed using STEM-EDX and XAFS techniques to understand the poor reproducibility in ORR performance after the ADT. 

Experiment

80 g of carbon supported Pd core (Pd/C; particle size: 6.0 nm, Pd loading: 30 wt.%) was ultrasonically dispersed in 50 mM H2SO4 containing 20 mM CuSO4 and stirred at 5oC with coexistence of a metallic Cu sheet under Ar atmosphere. After pre-determined stirring time, the Cu sheet was removed and K2PtCl4 was added to replace under-potentially deposited Cu monolayer on the Pd core surface with Pt monolayer, forming Pt/Pd/C core-shell catalyst. The Pt/Pd/C catalyst was characterized with ICP, XRD and CO adsorption. ORR activity of the Pt/Pd/C catalyst was evaluated with RDE technique in O2 saturated 0.1 M HClO4 at 25oC. Accelerated durability test (ADT) was performed using a rectangular wave potential cycling of 0.6 V (3 s)-1.0 V (3 s) vs. RHE at 60oC in Ar saturated 0.1 M HClO4for 10,000 cycles. Structure of the Pt/Pd/C catalyst was analyzed by STEM, STEM-EDX and XAFS (BL14B2 beam line at SPring-8).

Results and Discussion

Physical and electrochemical properties of two Pt/Pd/C catalysts, (a) and (b), which were synthesized in different batches (100 g/batch), are summarized in Table 1. Although initial properties are almost equivalent, ORR specific activity of (a) Pt/Pd/C catalyst is much enhanced by the ADT (434→966 µA/cm2, 2.2-fold) compared with that of (b) Pt/Pd/C catalyst (437→731 µA/cm2, 1.7-fold).HAADF-STEM images and STEM-EDX line analyses of the catalysts after the ADT are demonstrated in Fig. 1, revealing that there are two different structures in the catalysts. One retained Pd core-Pt shell structure with thickened Pt shell (Fig. 1 (A)) and the other transformed to a Pt enriched nanoparticle with large Pd dissolution (Fig. 1 (B)). The STEM-EDX analysis showed that the core-shell structure was much retained in the (a) Pt/Pd/C catalyst. Furthermore, XAFS analysis indicated that Pt-Pt bond distance of the Pt shell was more shortened in the (a) Pt/Pd/C catalyst after the ADT. Therefore, it can be concluded that the higher ORR specific activity of the (a) Pt/Pd/C catalyst after the ADT arises from retention of the core-shell structure and the shortened Pt-Pt bond distance (i.e., higher compressive strain in the Pt shell).Since it is considered that the inferior ORR specific activity of the (b) Pt/Pd/C catalyst after the ADT is mainly due to size distribution of the Pd core nanoparticles, synthesis of the Pd nanoparticles with narrower size distribution should be developed. At the meeting, cell performance using the Pt/Pd/C catalysts will be presented.

Acknowledgement

This work was supported by NEDO, Japan. 

References

[1] J. Zhang et al., J. Phys. Chem. B108, 10955 (2004).

[2] J. Zhang et al., Angew. Chem. Int. Ed., 44, 2132 (2005).

[3] M. Inaba and H. Daimon, J. Jpn. Petrol. Inst., 58(2), 55 (2015).

[4] N. Aoki et al., The 224th ECS Meeting, Abstract #1522, San Francisco, USA, October 2013.

Figure 1

2649

, , and

Introduction

Pt–M (M = Fe, Co, Ni, Cu, Pd, etc.) alloy and Pt/M core–shell nanoparticles have been intensively studied as oxygen reduction reaction (ORR) catalysts for polymer electrolyte fuel cells (PEFCs). As for the core–shell catalysts, the core elements easily dissolve in the electrolyte under the PEFCs operating conditions, leading to degradation of the catalytic activity. To control dissolution behavior of the core elements, Kuttiyiel et al. proposed the structure of Pt–shell/nitride–core nanostructures and demonstrated that nitride core is effective not only for enhancing ORR activity but also for improving durability [1]. For developing the practical Pt–M catalysts, it is significant to clarify a role of the nitrogen atoms located at the core–shell interface. In this study, we fabricate model catalyst structures that composed of Pt(111) epitaxial layers on low–energy N2–beam irradiated Pt25Ni75(111) substrate surfaces by using molecular beam epitaxy (MBE) and discuss the ORR activity and the durability.

Experimental

Model catalyst sample surfaces are fabricated as follows. First, Pt25Ni75(111) single crystal substrate was cleaned by Ar+ sputtering and annealing at 1173 K in ultra–high vacuum (UHV). Next, neutralized N2–beam (≤ 100eV) was irradiated onto the cleaned Pt25Ni75(111) at room temperature by using low–energy neutralized nitrogen ion beam gun. Then, 3ML–thick–Pt was deposited on the N2–beam irradiated Pt25Ni75(111) substrate by an electron–beam evaporation method at room temperature, followed by annealing at 673 K for 10 minutes in UHV to flatten the topmost surfaces. The resulting surface structure was verified by scanning tunneling microscopy (STM). The sample thus fabricated was transferred to electrochemical evaluation systems set in a N2–purged glove box without air exposure [2]. CV and LSV measurements were conducted in N2–purged and O2–saturated 0.1 M HClO4, respectively, and ORR activity was evaluated from jk values at 0.9 V vs. RHE by using Koutecky–Levich equation. The electrochemical stability of the sample was investigated by applying potential cycles between 0.6 V (3s) and 1.0 V (3s) in O2–saturated 0.1 M HClO4.

Results and Discussion

UHV–STM images of the as–prepared 3ML–Pt/Pt25Ni75(111) and 3ML–Pt/100eV–N2–beam irradiated Pt25Ni75(111) (3ML–Pt/N2–beam–Pt25Ni75(111)) surfaces are shown in Fig. 1(a). The line profiles of white arrows are depicted below the corresponding STM images. The STM images clearly show that atomically flat surfaces can be obtained through the UHV–process, irrespective of the 100eV–N2–beam irradiation. Furthermore, the both surfaces exhibited moiré–like height modulations (line profiles) that derived from lattice mismatches between the topmost Pt–shell layers and corresponding substrates. The results suggest that compressive surface strain should be induced for both the shell layers. Corresponding CV curves of the 3ML–Pt/Pt25Ni75(111) (blue), 3ML–Pt/N2–beam–Pt25Ni75(111) (red) and Pt(111) (black–dashed) are summarized in Fig. 1(b). Compared with the clean Pt(111), the total Hupd charges for the 3ML–thick–Pt surfaces considerably decreased and the onset of Oads positively shifted. Such CV features are common for highly–active Pt skin–type surfaces [2, 3]. Furthermore, the oxidation of the Pt shell layers for the latter N2–beam irradiated sample is suppressed in comparison to the former. ORR activities evaluated before and after 1000 potential cycles loading are summarized in Fig. 1(c). As for the initial activity enhancement factor, the 3ML–Pt/Pt25Ni75(111) andthe N2–beam irradiated one reveals ×15 and ×8 vs. Pt(111), respectively. The results suggest that the nitrogen atoms located at the Pt–shell/Pt25Ni75(111) interface should release the surface strain of the Pt–shell layers, leading to decrease in pristine ORR activity. In contrast, N2–beam irradiated sample is relatively stable against the potential cycle loading, indicating that the interface nitrogen atoms dominate not only activity but also durability of the Pt– shell/nitride–core type catalysts. At the PRiME 2016, I will discuss ORR activity and durability for various–thick Pt shells prepared on N2–beam–Pt25Ni75(111).

Acknowledgement

This study was supported by new energy and industrial technology development organization (NEDO) of Japan.

References

[1] K. A. Kuttiyiel et al., Nano Energy, 13, 442 (2015).

[2] T. Wadayama et al., Electrochem. Commun., 12, 1112 (2010).

[3] V. R. Stamenkovic et al., Science, 315, 493 (2007).

Figure 1

2650

, , , , , , , and

Improvement in the activity and durability of electrocatalysts for the oxygen reduction reaction (ORR) is an essential issue for wide-spread commercialization of polymer electrolyte fuel cells. Nanoparticle catalysts with alloy, skin, and/or core-shell structures have been developed to improve the activity and durability. However, these nanoparticle catalysts often suffer from degradation under harsh conditions. On the other hand, extended surfaces including thin-films possess high durability; however, the surface area is lower than that of nanoparticles due to low surface area.

Nanosheets with atomic thickness have large specific surface area and two dimensionally extended surface. We previously reported the synthesis of metallic ruthenium nanosheets with monoatomic thickness via thermal reduction of RuO2 nanosheets.1,2In this study, we demonstrate the synthesis of Ru-core@Pt-shell nanosheets and the catalytic performance towards the oxygen reduction reaction.

Metallic Ru nanosheets supported on carbon composite was prepared via thermal reduction.3 Pt shell was successively formed on metallic Ru nanosheets supported on carbon via galvanic displacement reaction between Cu and Pt2+. The electrochemically active Pt surface area of Ru-core@Pt-shell nanosheets with 3.5 monolayer Pt-shell (Ru@Pt-3.5ML(ns)/C) was 148 m2 (g-Pt)‒1and showed 4.5 times higher activity than benchmark commercial Pt/C catalyst for the oxygen reduction reaction. In addition, the activity retention of Ru@Pt-3.5ML(ns)/C after durability test (0.6-1.0 V and 1.0-1.5 V for 5000 cycles) was higher than that of Pt/C. The high activity could be attributed to the core-shell structure, and the durability originates from the extended surface structure.

This research was supported in part by the "Polymer Electrolyte Fuel Cell Program" from the New Energy and Industrial Technology Development Organization (NEDO) of Japan.

1) W. Sugimoto, H. Iwata, Y. Yasunaga, Y. Murakami, and Y. Takasu, Angew. Chem. Int. Ed., 42, 4092 (2003).

2) K. Fukuda and K. Kumagai, e-Journal Surf. Sci. Nanotechnol., 12, 97 (2014).

3) D. Takimoto, T. Ohnishi, Y. Ayato, D. Mochizuki, and W. Sugimoto, J. Electrochem. Soc., 163, F367 (2016).

Figure 1

2651

, , and

Introduction

The core-shell type nanostructure, that is composed of ultra-thin Pt coating on other transition metal nanoparticles have attracted much attention from the viewpoints of cathode electrode catalysts in polymer electrolyte fuel cell (PEFC). Compressive strain of the Pt-shells induced by lattice mismatch and direct electronic interaction between the Pt shell and core elements influences electronic properties of the Pt shell layers and, thereby, contributing to enhanced oxygen reduction reaction (ORR) activity [1]. However, because of operating conditions of PEFC, i.e., low pH, high positive potentials, ORR activity of the core-shell catalysts decline through dissolution of core elements and rearrangements of surface Pt atoms. Considering standard electrode potentials, iridium is stable among the transition metals and Pt-shell/Ir-core nanoparticles might be stable under the PEFC operating conditions. However, at present, relationship between Pt-Ir bimetallic surface structures and ORR properties has been less reported [2]. In this study, we prepared well-defined Pt/Ir(111) bimetallic surfaces by molecular beam epitaxy (MBE) in ultra-high vacuum (UHV) and studied the ORR activity and durability.

Experimental

Ir(111) single crystal substrate was cleaned by Ar+ sputtering and annealing at 1273 K in UHV. Various-monolayers(ML)-thick Pt was deposited onto the clean Ir(111) substrate by an electron beam evaporation method at the substrate temperature of 673K. The resulting surface structures were verified by reflection high energy electron diffraction (RHEED) and scanning tunneling microscopy (STM) in UHV. The UHV-fabricated samples were transferred to electrochemical evaluation systems set in a N2-purged glove box without air exposure. CV and LSV measurements were conducted in N2-purged and O2-saturated 0.1 M HClO4, respectively. The ORR activity was evaluated from jk values at 0.9 V vs. RHE by using the Koutecky-Levich equation. EC degradation of the Pt/Ir(111) surface was evaluated by applying potential cycles between 0.6 V (3s) and 1.0 V (3s) vs. RHE in the O2-saturated solution at room temperature.

Results and Discussion

Fig. 1 (a) summarizes CV curves of the MBE-prepared nML-thick Pt/Ir(111) (n=1-4: PtnML/Ir(111)) and clean Pt(111). All CV curves of the PtnML/Ir(111) show symmetrical peaks similar to the clean Pt(111) (so-called butterfly peak) in the region of OH adsorption and desorption (0.6-1.0 V), accompanied with remarkable decrease in hydrogen adsorption and desorption charges (0.05-0.35 V). The CV features are typical for highly-ORR-active Pt-based bimetallic surfaces reported to date, suggesting that the topmost surfaces comprise pure Pt(111) lattice.

Fig. 1 (b) presents changes in jk during potential cycles loading between 0.6 and 1.0 V. Initial ORR activity enhancement factors for the 2ML, 3ML and 4ML-thick Pt shell layers can be estimated to be ×24, ×12 and ×8, respectively compared with the clean Pt(111). Because Ir lattice constant is smaller than Pt, the Ir(111) substrate should induce compressive surface strain on the Pt-shell. As a result, the ORR activity is enhanced. Furthermore, because compressive strain of the Pt shell layers should be relaxed with increasing the Pt shell layer thickness, the initial activity enhancement factors decreased with increasing the thickness. After the 5000 potential cycles loading, the enhancement factors for the Pt2ML/Ir(111) decreased to ×2 vs. clean Pt(111) indicating that compressive surface strain is easily released by the potential cycle loading. In contrast, particularly, the initial activity of the 4ML-Pt-shell remains nearly unchanged even after the 5000 cycles loading: the activity is the highest among the other well-defined Pt-M(111) bimetallic systems, e.g., Pt/Pd(111) [3], Pt-skin surfaces of Pt-Ni(111) [4] and Pt-Co(111) [5], reported in our laboratory. The results can be explained by 3D migration of the surface Pt atoms during the potential cycle loading and by release in compressive strain in Pt shell. Therefore, the results suggest that Pt-Ir bimetallic system can be expected to be applied to ORR catalysts for practical use.

 

Acknowledgement

This study was supported by new energy and industrial technology development organization (NEDO) of Japan and Grant- in Aid for young scientists (B) from the Japan society for the promotion of science (N. T.).

 

References

[1] P. Mani et al., J. Power Sources,196, 666 (2011).

[2] J. Zhang et al., Angew. Chem. Int. Ed., 44, 2132 (2005).

[3] Y. Band et al., J. Electroanal.Chem.,162, 463 (2015).

[4] N. Todoroki et al., Phys. Chem. Chem. Phys., 15, 17771 (2013).

[5] Y. Yamada et al., Surface Science,607, 54 (2013).

Figure 1

2652

, , , and

Introduction

The relatively low activity and durability of Pt nanoparticles (NPs) supported carbon (Pt/C) catalyst limit wide commercialization of polymer electrolyte fuel cell (PEFC). To address this problem, Pt-shell/transition-metal-core type NPs have been investigated. Toyoda and co-workers have been demonstrated superior ORR activity of the Pt layers formed on TaB2 single crystal substrate [1]. They explain the activity enhancement to be caused by the electronic effect of Ta on Pt. Because nitride of 4-6th group metals, e.g., Hf, Ta, and W are thermodynamically stable, Ta-nitride can be expected to be highly durable core-material under PEFC operating condition. However, no study has been conducted for Pt-Ta-N ternary alloy NPs for ORR catalysts. In this study, Pt-shell/Ta nitride-core NPs (Pt/TaNx) model catalyst were fabricated on a highly-oriented pyrolytic graphite substrate (HOPG) through Arc-plasma depositions of TaNx NPs followed by electro-beam deposition of Pt in ultra-high vacuum: ORR activity and durability of the Pt/TaNxNPs are discussed.

Experimental

HOPG substrate was scraped by Scotch tape in air and cleaned by annealing for 30 min in ultra-high vacuum (UHV). Ta and TaNx core are synthesized by arc plasma deposition on the HOPG substrate at 873K or 1173K in UHV (Ta NPs) and N2 partial pressure of 0.1 Pa (TaNx NPs). Subsequently, Pt was deposited on the Ta or TaNx NPs at 873K by using electron-beam deposition method in UHV. The deposition amounts of Ta and Pt were estimated to be 0.7 and 1.9 μg/cm2, respectively. UHV-fabricated Pt/Ta and Pt/TaNx NPs samples were transferred from UHV to electrochemical (EC) setup without exposure to air. ORR properties of the samples are discussed by cyclic voltammetry (CV) and linear sweep voltammetry (LSV) in 0.1 M HClO4. Durability of the samples discussed on the bases of mass activity that estimated based on ikvalues at 0.9V during applying square potential cycles (PCs) between 0.6 (3 s) and 1.0 V (3 s) in oxygen-saturated 0.1 M HClO4 at room temperature. Surface morphologies of the sample alloy NPs were observed by scanning tunneling microscopy (STM).

Results & Discussions

STM images of the samples were shown in Fig. 1(a). Monodispersed NPs with average diameter of ca. 5nm are formed on the substrate irrespective of the core Ta and TaNx preparation conditions. Fig.1(b) shows CV curves for the corresponding samples. Oxygen species adsorption charge of the Pt/TaNx-1173K is larger than that of Pt/Ta-873K and Pt/TaNx-873K, suggesting that high-temperature nitridationaffect the electronic effect of the core Ta on Pt. Fig. 1(c) shows changes in mass activity of the samples under applying the PCs. The Pt/TaNx-1173K exhibits a 2.4-fold higher initial mass activity compared with that of the commercial Pt/C [2] and is the highest among the samples tested in this study. Furthermore, the Pt/TaNx-1173K shows high ORR durability: the values estimated after 10k PCs were about ca. 80% of the initial one. The results obtained in this study clearly reveal that nitridization of transition-metal-cores is effective for improving ORR activity and durability of Pt-shell/transition-metal-core type NPs catalysts.

Acknowledgement

This study was supported by the new energy and industrial technology development organization (NEDO) of Japan and Grant-in-Aid for challenging exploratory research from the Japan society for the promotion of science (T. W.).

References

[1] E. Toyoda, R. Jinnouchi, T. Ohsuna, T. Hatanaka, T. Aizawa, S. Otani, et al., Angew. Chem. Int. Ed. Engl., 52 (2013) 4137–4140.

[2] M.K. Debe, Nature, 486 (2012) 43–51.

Figure 1

2653

, , , , , , , , and

The role played by the surface properties of carbon blacks used as electrocatalyst supports in fuel cell electrodes is crucial in determining the performance and durability characteristics of the final membrane electrode assembly. These key surface properties include the polar character arising from the presence of heteroatoms, and the proportion of the surface area that corresponds to graphitic carbon sites.

We have used flow calorimetry to characterise the surface properties of a family of carbon blacks having the same BET surface area (400±10 m2/g), but different pore size characteristics and different properties in terms of polar and graphitic surface sites. The preferential heat of adsorption of n-dotriacontane and n-butanol for basal plane sites and polar sites respectively has been determined for the carbon sites immersed in n-heptane1. The overall hydrophobic character of the carbon surface was evaluated by measuring the heat of adsorption of n-butanol from its aqueous solution2.

Platinum nanoparticles prepared by a microwave-assisted polyol method2 were deposited on each of the carbons, and the electrochemical surface area, the mass and specific activity of the Pt/C samples were determined. Membrane-electrode assemblies were prepared using the electrocatalysts at the cathode and they were tested in fuel cell over a range of temperatures and relative humidities, when the observed performance could be related to the relative hydrophobic/hydrophilic properties of the carbons. Further, the stability of the carbons to electrochemical corrosion was assessed by submitting them to high voltage (1.4 V/RHE) and determining the corrosion current and the corresponding mass loss, and the results related to the graphitic character of the carbons determined from flow microcalorimetry and Raman spectroscopy.

This understanding will be beneficial in the tuned design of PEMFC electrocatalyst supports for the targeted operation conditions. This work demonstrates the importance of calorimetric methods in characterising carbons for electrochemical applications.

References

1. J. Zajac, A. J. Groszek, Carbon 35 (1997) 1053-1060.

2. L. Kaluža et al., Electrocatalysis, (2016) http://dx.doi.org/10.1007/s12678-016-0312-3.

2654

, , , and

Introduction

Catalyst layers in PEFCs are the most important components affecting their performance. Such catalyst layers are mostly composed by carbon materials. Therefore, optimization of the carbon structure in the catalyst layers is essential. In our study, we have successfully improved PEFC durability by suppressing the aggregation of Pt particles through the encapsulation in nano-channels of mesoporous carbon (MC).1 However, the improvement on the mass transfer, such as diffusion of the fuel gases and the removal of the produced water, still required. Deficient mass transfer is mainly caused by the insufficient structure of carbon materials in the macro scale. Therefore, the purpose of this study is to improve the performance and further durability of PEFCs by controlling carbon structures over the nanometer to micrometer scales.

Experimental methods

For the MEA preparation, the anode and the cathode were spray-printed using a standard catalyst made by TKK Co., Ltd. (TEC10E50E 46%Pt/KB) and our original MC made in the laboratory, respectively. IV characteristics of MEAs were evaluated while changing the cathode conditions. In order to control the structure over the nanometer to micrometer scales, two separate approaches were tried. As approach 1 to control the structure in the micrometer scale, cathode layers were made by original MC with the addition of carbon nanofibers (CNF). Then, as approach 2 to control the nanometer scale, new MC whose mesopores were expended, was in used. IV performance under the conditions of 80 oC-RH100% was evaluated by supplying 100 cc/min of hydrogen to the anode and 100 cc/min of the air to the cathode. Furthermore, overvoltages were carefully separated into ohmic, activation, and concentration overvoltages. Regarding to the cathode layer structures, the porosity was quantitatively analyzed by measuring the thickness of the cathode layer form cross-section SEM images.

Results and discussion

As a result form approach 1, the addition of CNF to original MC lead to improvement on IV performance owing to reduced concentration overvoltage. In order to see the structure change in the cathode layer by the addition of CNF, SEM observation was done, and images are shown in Fig. 1. If two images are compared, MC +CNF (Fig. 1(b)) clearly shows the more suitable structure for the mass transfer even from two-dimensional images. Furthermore, based on quantitative analyses of the porosity, the porosity was found to be improved by about 4%. Additionally, activation overvoltage resulted in lower most likely owing to improved conductive path in cathode layer.

In the case of approach 2, mesopores of new MC resulted in larger (20~25 nm). IV performance of MEA with this new MC was also improved as a result of reduced concentration overvoltage. Based on the observation of the cathode structure, two factors, enlargement of mesopores and refinement of carbon particles, have most likely influence on PEFC performance.

(1) Y. Minamida et al., ECS Trans., vol.64(3), pp.137-144, 2014.

Figure 1

2655

and

One of the main issues restricting wide-spread commercialization of Polymer Electrolyte Membrane Fuel Cells (PEMFCs) is the gradual decline in their performance during operation, primarily due to degradation of cathode catalyst layer by sintering, dissolution and re-deposition of Pt and corrosion of carbon support. Development of durable carbon-based supports for PEMFC electrodes can be beneficial for improving catalytic activity and stability of the electrocatalysts. Accordingly, various carbons (carbon blacks, carbon nanotubes, carbon nanofibers, and graphenes) as well as metal oxides are extensively explored as support materials. Among them, mesoporous carbon (MC) is highly attractive from the viewpoint of pore structure and pore-size distribution desired for mass-transport and enhanced Pt-utilization, which enables encountering problems inherent with microporous carbon black support (i.e., Vulcan XC-72R). On top of that, improvement of catalytic performance can be obtained by modifications in both surface chemistry and pore structure of MC. Therefore, in situ polymerization of 3,4-ethylenedioxythiophene(EDOT) with mesoporous carbon is prepared as catalyst support for platinum nanoparticles by sol-gel method. This conducting polymer was chosen owing to its characteristics such as high chemical, thermal stability, and high electrical conductivity of Poly(3,4-ethylenedioxythiophene)(PEDOT). Pt nanoparticles are then impregnated onto the MC-PEDOT composite using conventional formaldehyde reduction method. Resulting Pt supported MC-PEDOT composite exhibits promising electrocatalytic activity toward oxygen reduction reaction (ORR), which make it attractive for use as electrode material for fuel cell application. The morphology and nanosturcture of Pt supported MC-PEDOT was confirmed by several analysis equipment including SEM, TEM, BET and FT-IR. Electrochemical techniques such as cyclic voltammetry (CV) and impedance measurements are also used to evaluate the extent of degradation in the catalyst layer.

2656

, , , , , , , and

Proton exchange membrane fuel cells (PEMFCs) that use hydrogen as fuel to produce electricity have attracted extensive interests over the past decades.[1] The high efficiency and zero carbon- emission make PEMFCs promising candidates for viable propulsion systems in electric vehicels.[2] One of major challenges that hinder the commercialization of PEMFCs vehicles is the poor stability of Pt nanoparticles on carbon supports, which is mainly due to (1) the dissolution and re-deposition of Pt and (2) migration of Pt nanoparticles over carbon supports.[3],[4] Both of them will cause the agglomeration of Pt nanoparticles, resulting in loss of Pt electrochemical surface area (ECSA) and the increasing of activation over-potential. In this regard, the enhancement of Pt nanoparticle anchoring strength and dispersion on carbon supports is highly desirable in PEMFCs as well as in other catalysis processes.

Presented here is a comprehensive study of the interaction between catalyst nanoparticles and carbon supports in terms of the electronic structure change and its effects on the electrocatalytic performance of supported catalysts. Graphene was chosen as an ideal model system for catalyst support because the unique 2-D structure allows the direct investigation of the interaction with supported metal nanoparticles at their interface. We developed a facile strategy to covalently graft p-phenyl SO3H or p-phenyl NH2 groups onto the graphene surface. The functional groups were found to not only facilitate the homogeneous distribution of Pt nanoparticles over the surface of graphene supports and reduce the Pt average particle size but also strengthen the interaction of the Pt atoms with the functional groups and, consequently, minimize the migration/coalescence of the Pt nanoparticles in the course of accelerated durability tests. The experimental results from both X-ray photoelectron spectroscopy (XPS) and X-ray absorption spectroscopy (XAS) demonstrate the electron density shift from Pt to graphene supports with the strength of the Pt−graphene interaction following the trend of Pt/p-phenyl NH2-graphene > Pt/p-phenyl SO3H-graphene > Pt/graphene. This study will shed light on strategies to improve not only the durability but also the activity of the metal nanoparticles via the functionalization of the catalyst supports in the catalysis field

(1) Gasteiger, H. A.; Markovic, N. M. Science 2009, 324, 48.

(2) Debe, M. K. Nature 2012, 486, 43.

(3) Xie, J.; Wood, D. L.; More, K. L.; Atanassov, P.; Borup, R. L. Journal of The Electrochemical Society 2005, 152, A1011.

2657

, and

Introduction

PEFCs are paid much more attention since the commercialization of FCVs and are deeply related to carbon materials, which are components of PEFCs themselves and also related systems. In our group, we have been studying mesoporous carbon (MC), which has meso-channel structures with about 10 nm diameter, as a catalyst support and have successfully improved durability of PEFCs, so far. In order to further improve performance and durability of PEFCs and related systems, it is important to understand the fundamentals of adsorption properties of fuels involving hydrogen gas, oxygen gas, and water vapor. In this research, a correlation between structures of mesoporous carbon and gas adsorption characteristics is investigated.

Experimental

General MC was synthesized by mixing Pluronic® F127, resorcinol, formaldehyde, and trimethyl orthacetate and heated under nitrogen atmosphere. Other MC materials1 were also made by slightly different conditions as shown in Table 1. Then, gas adsorption properties of each sample were evaluated after the ball milling and heat treating at 900 oC.

Results & discussion

Hydrogen adsorption dependence on the pore size of MC is discussed here. If MC#1 and MC#2 in Table 1are compared, MC#2 has larger pores based on the pore distribution obtained from nitrogen adsorption measurements. The values of BET specific surface area of the MC#1 and the MC#2 results in not very much different, 615 m2/g and 650 m2/g, respectively. The maximum hydrogen absorption content (wt%) at 7 MPa has resulted in 0.6 % for MC#1 and 0.35 % for MC#2. Considering Chahine rule2 for carbon materials, revealing that linear relationship between the hydrogen absorption content and the specific surface area, our results are against that rule. Therefore, hydrogen adsorption properties can be explained by other than the specific surface area in our materials. The possible mechanism of hydrogen adsorption is further discussed through the analyses of several MC materials.

(1) Y. Sonoda, A. Hayashi, Y. Minamida, E. Akiba, Chemistry Letters, 44, 503-505 (2015).

(2) R. Chahine, T.K. Bose, 11th WHEC, 1996; Pergamon Press: Oxford, U.K., 1259 (1996).

Figure 1

2658

, , and

Polymer electrolyte membrane fuel cells (PEMFCs) have been regarded as a key element for the development of sustainable and renewable energy conversion devices, owing to their high efficiency and environmental benignity. In PEMFCs, carbon supported Pt-based catalysts (Pt/C) have shown the best activity, yet they are easily deteriorated via several degradation processes such as Pt particle detachment, Pt dissolution, and/or Ostwald ripening. Therefore, understanding and enhancing the durability of Pt catalysts is one of the major challenges for long-term operation of PEMFC. To improve the stability of Pt/C catalysts, recently, heteroatom-doped carbons have been exploited as support materials. In this work, we investigated the relationship between metal-support interaction and durability of Pt catalysts for the oxygen reduction reaction (ORR). For this purpose, three different types of heteroatom-doped ordered mesoporous carbon (OMC) supports, namely O-doped OMC (O-OMC), N,O-doped OMC (N,O-OMC), and S,O-doped OMC (S,O-OMC), were prepared and Pt nanoparticles were supported on these carbon supports. The interaction between Pt and doped OMC supports was investigated by using atomic force microscopy (AFM). It was revealed that the S,O-OMC exhibited the strongest adhesion with Pt, whereas the N,O-OMC showed the weakest adhesion. X-ray diffraction patterns and X-ray photoelectron spectroscopy corroborated the adhesion results. In accelerated degradation tests for the ORR, Pt/S,O-OMC exhibited the highest durability, followed by Pt/O-OMC and Pt/N,O-OMC. These results suggest that sulfur has the strong adhesive strength with Pt, thereby enhancing the durability, but nitrogen exert less adhesive force with Pt, undermining the durability.

2659

, , and

With the development of various energy technologies, much interest is focused on the feasibility and efficiency study of energy devices such as fuel cells, batteries, supercapacitors and water electrolysis systems. Among them, polymer electrolyte fuel cell (PEFC), which convert hydrogen into electric energy with zero emission of pollutants, is one of the most promising environmentally friendly technologies. Despite innumerable studies for more than half a century, degradation of the key component, membrane electrode assembly (MEA), is still a big obstacle for the commercialization of PEFC system.

For the high performance and long-term stability of Pt/C electrode catalyst, the agglomeration of Pt particles and dissolution or detachment of Pt particles from carbon support have to be improved. For this purpose, we adopted a new shape of nano-carbon material for the surface modification of Pt catalyst.

Graphene has been an attractive two-dimensional carbon allotrope having large surface area and electronic conductivity. Graphene also shows high flexibility and mechanical strength so that it can be used for large number of applications such as flexible display or printable electronics, etc [1-2]. In most cases, people need large-area graphene films with little or no defect for high thermal and electronic conductivity. Also, people need to transfer the as-prepared graphene film for each application. All of these processes are not easy or simple, and they are also labor-intensive processes. However, in the case of catalysis, porous graphene films can be synthesized via very simple one-step process and can be used as effective protective layers for Pt catalysts.

In this study, we developed porous graphene films in order to improve the long-term stability of Pt catalysts maintaining the high performance of them. The graphene films were synthesized by single-step vaporization process, where the number of graphene layers and the defects in their structure are manipulated by temperature and composition of the precursors. In this process, the amounts of structural defects, pyridine was simultaneously introduced to the vaporization process, which is much easier and cost-effective compared to the conventional NH3-treatment at high temp [3].

Consequently, our Pt/C catalysts coated with porous graphene films showed similar initial activity compared with the commercial catalysts (Pt 40wt%, Johnson Matthey) showing more than 150% higher long-term stability [4].

 [1] A.K. Geim and K. S. Novoselov, Nature Mater. 6 (2007) 183.

[2] X. L. Li, G. Y. Zhang, X. D. Bai, X. M. Sun, X. R. Wang, E. Wang and H. J. Dai, Nature Nanotech. 3 (2008) 538.

[3] Y. Wang, Y. Shao, D. W. Matson, J. Li and Y. Lin, ACS nano 4(4) (2010) 1790.

[4] H. Kim, A. Robertson, S. O. Kim, J. M. Kim and J. H. Warner, ACS nano 9(6) (2015) 5947.

 

 

2660

, and

High cost and low durability are unresolved issues that impede the commercialization of proton exchange membrane fuel cells (PEMFCs). To overcome these limitations, Pt/TiO2 is reported as an alternative electrocatalyst for enhancing the oxygen reduction reaction (ORR) activity and/or durability of the system. However, the low electrical conductivity of TiO2 is a drawback that may be addressed by doping. To date, most reports related to Pt/doped-TiO2 focus on changes in the catalyst activity caused by the Pt-TiO2 interaction (metal-support interaction), instead of the effect of doping itself; doping is merely considered to enhance the electrical conductivity of TiO2. In this study, we discuss the variation in the electronic fine structure of Pt caused by the dopant, and its correlation with the ORR activity. More extensive contraction of the Pt lattice in Pt/M-TiO2 (M = V, Cr, and Nb) relative to Pt/TiO2 and Pt/C leads to outstanding ORR specific activity of Pt/M-TiO2. Notably, a fourfold increase of the specific activity is achieved with Pt/V-TiO2 relative to Pt/C. Furthermore, an accelerated durability test (ADT) of Pt/V-TiO2 demonstrates that this system is three times more durable than conventional Pt/C due to the metal-support interaction.

2661

, , and

One of the technical problems impeding commercial use of polymer electrolyte fuel cells (PEFCs) is the degradation of the carbon support under startup/shutdown (SU/SD) conditions, which leads to severe and eventually drastic reduction of the performance due to agglomeration, dissolution, or detachment of Pt nanoparticles. Thus, nanoceramic supports (NCSs) with both high electrical conductivity and durability under SU/SD operating conditions have been actively developed to replace the commercial carbon support.[1-6] Titanium dioxide (TiO2) doped with aliovalent cations is also a candidate as a high durability cathode NCS. We synthesized Pt catalysts supported on tantalum and tungsten-doped TiO2with unique microstructure, involving a fused-aggregated network, and evaluated the oxygen reduction reaction (ORR) activity and durability under SU/SD operation conditions.

We prepared the NCSs by a flame spray synthesis method. The morphology of the NCSs, which was investigated by use of transmission electron microscopy (TEM), had nanoparticles with a partially fused aggregated network structure, similar to that of carbon black (Fig. 1(a)(b)). The crystal structures of the NCSs included rutile and anatase. The surface areas, which were estimated using the Brunauer-Emmett-Teller (BET) technique, were greater than 30 m2g-1. The electrical conductivities, which were measured by use of the two-probe method, were as high as 10-4 S/cm at room temperature under a pressure of 19 MPa (Fig. 2). The NCSs were confirmed to be extremely stable in 1.0 M H2SO4 at 80 oC under air atmosphere. Pt nanoparticles, which were loaded on the NCSs by a colloidal method, were uniformly dispersed on the surface (Fig. 1(c)(d)). The Pt loadings were 19 wt% (Pt/Ta-TiO2) and 19.4 wt% (Pt/W-TiO2), determined by ICP-MS. Pt particle sizes were 4 nm (Pt/Ta-TiO2) and 10 nm (Pt/W-TiO2), determined by TEM. The ORR activity increased with increasing heat treatment temperature and reached that of a commercial carbon-supported Pt catalyst (c-Pt/CB). The durability of Pt/Ta-TiO2 under SU/SD conditions was superior to that of the c- Pt/CB.

Figure 1

2662

, , , , , , , , , et al

Fuel cell is one of the most promising electrochemical energy conversion devices that has many advantages such as high efficiency, high power density, CO2 free, noise free, rapid start up, etc. Especially, polymer electrolyte membrane fuel cell (PEMFC) has been shown a lot of attention due to its high power density and relatively good portability. Generally, Pt supported on carbon black is used as an electrocatalyst in PEMFC. However, carbon corrosion is one of the main problem for electrode structure collapse and catalyst activity loss, which can be further accelerated at high temperature. Therefore, we need to find an alternative catalyst support material that is electrochemically stable and interacts strongly with the catalyst active sites.

Recently, metal oxide supports have been reported as promising materials due to their excellent mechanical resistance and inherently higher stability. Among many candidates, titanium dioxide (TiO2) support has an extraordinary stability under severe acidic atmosphere, which provides the possibility to directly use as a support material. However, the low electrical conductivity, catalytic activity and surface area of TiO2hinder the direct apply to the PEMFC electrode.

In order to solve these problems, we prepared TiO2 nanofibers by electrospinning method and platinum nanoparticles are deposited by microwave-assisted polyol method to synthesize an effective electrode catalyst. Then, CNT was winded around the catalyst surface to boost up the electrical conductivity. Furthermore, we found a modified Pt electronic structure that takes advantage of the strong synergetic interactions of TiO2 nanofibers, Pt nanoparticles and winded-CNT. This structure influences on a decrease of the d-band vacancy of Pt due to electron transfer from the support, resulting in an improved oxygen reduction reaction. Therefore, the cathode with the CNT modified Pt/TiO2 nanofiber composite shows higher catalytic activity due to the elimination of the drawback associated with conductivity and enhanced electronic active structure. As from various PEMFC tests, our CNT-Pt/TiO2 catalyst shows superior performance and durability, which is definitely distinguishable at high temperature condition of 120 oC and RH 40% compare to the commercial Pt/C.

2663

, , , , and

Introduction 

Electrocatalyst layer of polymer electrolytefuel cells (PEFCs) has complex 3-dimensional structure, and therefore control and optimization of microstructure is necessary for higher I-V performance. Carbon black supported cathode catalyst is generally used, however carbon supports can be degraded through oxidation on the cathode side. Thus, SnO2 support which is conductive and stable under the strongly-acidic cathode condition could be alternative support materials [1-3]. The objective of this study is thus to develop MEAs with higher performance and durability using SnO2 supports on conductive fillers for optimizing microstructural design of electrocatalysts.

Experimental

 Sn0.98Nb0.02O2 was synthesized via the ammonia co-precipitation method or the homogeneous precipitation method. Pt nanoparticles were decorated on such a metal oxide support by using platinum acetylacetonate complex as a precursor. Microstructure of MEAs was modified by changing Nafion-to-electrocatalyst ratio (hereafter, Nafion ratio) systematically. The performance of MEAs was evaluated by measuring I-V characteristics and separating each overvoltage. Two series of MEAs were prepared and evaluated: MEAs with different conductive fillers (VGCF and CNT) and MEAs with different SnO2loadings on the conductive fillers.

Results and Discussion

I-V performance with different Nafion ratios is shown in Fig. 1. MEA with Pt/SnO2(Nb)/VGCF + 23wt. % Nafion exhibited the highest I-V performance in this study. In case Nafion ratio decreased from 23wt. %, I-V performance was degraded. However, the differences of each I-V performance were small. When Nafion ratio increased from 23wt. %, I-V performance decreased notably mainly due to higher concentration overvoltage. This is because open pores as gas transport pathway were filled with Nafion ionomer. I-V performance with different conductive fillers and different SnO2 loadings are shown in Fig. 2. I-V performance of MEA with CNT as the conductive filler was comparable to that of MEA with VGCF as the conductive filler. In case SnO2 loading increased, I-V performance was improved. This is attributed to better dispersion of Pt nano-particles. In case CNT was used as the conductive filler, the distribution of SnO2 particles on CNTs was often inhomogeneous so that Pt nano-particles were aggregated. Accordingly, surface area of Pt was reduced and activation overvoltage became higher. When SnO2 loading was increased, surface area of SnO2 was increased too and Pt nano-particles were highly dispersed, leading to lower activation overvoltage. In this study, cell voltage of the MEA with Pt/SnO2(Nb)/VGCF (SnO2 loading: 50wt. %) and MEA with Pt/SnO2(Nb)/CNT (SnO2 loading: 70wt. %) was 95% of that of MEA with 46.4wt. % Pt/KB at 0.2 A/cm2. High I-V performance of MEAs with SnO2support on conductive fillers is thus confirmed.

Reference

[1]K. Kanda, Z. Noda, Y. Nagamatsu et al., ECS Electrochem. Lett., 3(4) F15-F18 (2014)

[2]T. Tsukatsune, Y. Takabatake, Z. Noda et al., J.Electrochem.Soc., 161(12) F1208-F1213 (2014)

[3]F. Takasaki, S. Matsuie, Y. Takabatake et al., J.Electrochem.Soc., 158(10) B1270-B1275 (2011)

Figure 1

2664

, , , , and

Abstract

Pt/C is widely used as polymer electrolyte fuel cell (PEFC) electrocatalysts. However, under the high potential at the cathode, Pt detachment and aggregation due to carbon corrosion can occur. Therefore, Pt/C has a difficulty in durability. Electrocatalysts with tin oxide (SnO2) support are thus developed where vapor grown carbon fiber (VGCF-H) act as a conducting backbone, achieving both high activity and high durability [1-3]. However, their electrocatalytic activity should be further improved. For this reason, we are focusing on further activation by Pt-Co alloying (PtCox/Nb-SnO2/VGCF-H). Co is one of the relatively stable metals under the PEFC electrocatalyst environment and PtCox alloy catalysts are known to be capable of exhibiting higher activity over Pt catalysts [4]. The aim of this study is to develop Pt-Co alloy electrocatalysts on oxide support.

Sn0.98Nb0.02O2 was prepared on the VGCF-H (Sn0.98Nb0.02O2/VGCF-H) by the ammonia coprecipitation method or the homogeneous precipitation method. Pt(acac)2 and Co(acac)2 were applied to impregnate Pt and Co nanoparticles to obtain the electrocatalysts (acac method). In the acac method, two different procedures were applied: 1-step (simultaneously impregnating Pt and Co) and 2-step (after impregnating Pt, Co was impregnated). The Pt : Co ratio was adjusted to prepare Pt3Co alloy to obtain Pt3Co/Sn0.98Nb0.02O2/VGCF-H electrocatalyst.

In order to evaluate Pt and Co loadings, ICP analysis was performed. We also investigated Pt3Co alloying by XRD measurement for each different heat treatment condition. We then made half-cell tests to evaluate electrochemical activities of these electrocatalysts. Electrochemical surface area (ECSA) was measured by cyclic voltammetry (CV), and oxidation reduction reaction (ORR) activity was derived from kinetically controlled current density (ik) in the rotating disk electrode (RDE) measurement.

From the ICP measurement results, we confirmed that we could impregnate both Pt and Co close to the initial ratio in the acac method. In XRD measurement, after 600℃ heat treatment in a short time, SnO2 was reduced and Pt-Sn alloy was formed. On the other hand, when 300℃ heat treatment was applied (2-step), PtCo alloying proceeded. Figure 1 shows that, in the 2-step process, the XRD peak of the catalyst is shifted to a higher angle when compared with the XRD peak of pure Pt. Figure 2 shows EDS analysis by high-resolution STEM of Pt3Co/Sn0.98Nb0.02O2/VGCF-H prepared by the acac method (2-step, 300℃, 3h). These micrographs clearly show that Pt and Co are distributed in the same position to cover the SnO2 surface. We confirmed that Pt : Co was 7.23 : 2.73 (atomic ratio) by quantitative analysis of the catalyst particles. It is therefore concluded that PtCo alloy catalysts impregnated on SnO2-based support can be prepared.

References

[1] A. Masao, S. Noda, F. Takasaki, K. Ito, and K. Sasaki, Electrochemical and Solid-State Letters, 12 (9) B119-B122 (2009)

[2] F. Takasaki, S. Matsuie, Y. Takabatake, Z. Noda, A. Hayashi, Y. Shiratori, K. Ito, and K. Sasaki, Journal of The Electrochemical Society, 158 (10) B1270-B1275 (2011)

[3] K. Kanda, S. Hayashi, F. Takasaki, Z. Noda, S. Taniguchi, Y. Shiratori, A. Hayashi, and K. Sasaki, ECS Transactions, 41 (1) 2325-2331 (2011)

[4] T. Tada, Y. Yamamoto, K. Matsutani, K. Hayakawa, and T. Namai, ECS Transactions, 16 (2) 215-224 (2008)

Figure 1

2665

, , and

Pt and Pt alloy nanoparticles supported on carbon were applied as cathode catalysts for polymer electrolyte fuel cells (PEFCs). The carbon support is appropriate for the support due to its high electrical conductivity and high surface area, although it corrodes under high potential conditions. Carbon corrosion leads to aggregation and detachment of the catalyst and reduces the performance of the PEFC. In our group, a highly durable Pt/Ta-SnO2 catalyst was developed for the cathode of PEFCs1). In this study, we synthesized Pt-Co alloy particles supported on Ta-SnO2 (PtCo/Ta-SnO2) in order to improve the activity while maintaining high durability. The Ta-SnO2 support was synthesized by the flame spray synthesis method. Pt67Co33, Pt75Co25 and Pt80Co20 catalysts were loaded on the Ta-SnO2 by the colloidal method. After heat treatment, the crystal phases, Pt/Co content, and microstructure were evaluated by X-ray diffractometry (XRD), inductively coupled plasma mass spectrometry (ICP-MS), and scanning transmission electron microscope (STEM), respectively. The electrochemical measurements were carried out by the half-cell method using the rotating disk electrode (RDE). The electrochemically active surface area (ECA) of each catalyst was estimated by cyclic voltammetry (CV). The kinetically controlled current density (jk) and mass activity (MAk) at 0.85 V were also estimated from linear sweep voltammetry (LSV) with Koutecky-Levich plots.

The Pt:Co composition ratios (mol%) of the catalysts were 67.1:32.9(Pt67Co33), 74.6:25.4(Pt75Co25), and 80.9:19.1(Pt80Co20), which were well controlled to the desired ratios. The metal loading amounts (wt%) of the catalysts were 17.9, 17.4 and 15.4, respectively. The particle diameters (nm) were 2.7 ± 0.9, 2.8 ± 0.9, and 3.1 ± 0.8, respectively. The catalysts were highly dispersed on the support with hemispherical shape (Fig. 1 (a)). Based on the HAADF-STEM image, it was confirmed that the lattice planes of the catalyst particles were oriented parallel to those of the support. This is expected to suppress the agglomeration of the catalyst particles during startup/shutdown. The Ta-SnO2 support particles formed a fused aggregated network structure, which is able to function as a pathway for both gas diffusion and electrical conduction2). The ECA values of the catalysts, estimated from the CV, were 56.2, 70.6, and 71.3 m2g-1. The LSV values of these catalysts (1750 rpm, Fig. 1 (a)) showed the same limiting current density of that of commercial Pt/carbon black (Pt/CB, TEC10E50E). The jk values at 0.85 V (Fig. 1 (b)) showed that the Pt75Co25/Ta-SnO2 catalyst had the highest oxygen reduction reaction (ORR) activity, which reached 2.5 times larger than that for commercial Pt/CB.

Acknowledgement

This work was supported by funds for the SPer-FC project from the New Energy and Industrial Technology Development Organization (NEDO) of Japan.

References

1) Y. Senoo, K. Taniguchi, K. Kakinuma, M. Uchida, H. Uchida, S. Deki, M. Watanabe, Electrochem. Commun. 51, 37 (2015).

2) K. Kakinuma, Y. Chino, Y. Senoo, M. Uchida, T. Kamino, H. Uchida, S. Deki, M. Watanabe, Electrochemica Acta110, 316 (2013).

Figure 1

2666

, , , and

Nanofibre antimony and niobium doped tin oxides (Sb-SnO2, Nb-SnO2) were prepared by electrospinning and calcination, and were functionalised by depositing platinum nanoparticles prepared by a microwave-assisted polyol method. Catalysts inks of Pt/Sb-SnO2 and Pt/Nb-SnO2 were used to prepare cathodes that were assembled with a commercial anode and membrane into MEAs. These MEAs were submitted to voltage cycling between 0.9 and 1.5 V, and to stop/start cycling. I-V characteristics recorded at the beginning and end of the accelerated stress test protocols. The relative stability of Pt/Sb-SnO2, Pt/Nb-SnO2 under these conditions, and the results of probe spectroscopic methods used to investigate ageing mechanisms in doped tin oxides, will be described.

2667

, , , , , and

Proton exchange membrane fuel cells (PEMFCs) are electrochemical energy conversion devices used for portable, residential and vehicular applications because of their low emissions, high efficiency and quick start-up characteristics. However, PEMFCs use Pt-based electrocatalysts as electrode catalysts. Due to the high cost and limited availability of platinum, research and development of non-precious catalysts is of paramount importance. A promising alternative is nitrogen-doped carbons. Doping graphene with nitrogen alters the chemical structure and modulates the electronic properties, allowing a degree of control over the catalytic properties1-4. Here we present the synthesis, characterization and electrochemistry of nitrogen-doped carbon foams.

A novel bottom-up chemical synthesis method was developed in our lab 5. Triethanolamine (or any other nitrogen-containing alcohol) is mixed with ethanol in various ratios and reacted with sodium to form a nitrogen-containing alkoxide. In this case, this precursor is pyrolysed under nitrogen at 600˚C to form nitrogen-doped carbon foam mixed with sodium oxide by-product. This by-product is removed by washing with water. The resulting nitrogen-doped carbon foams are then pyrolysed at various temperatures in order to improve the conductivity and vary the nitrogen content.

Electron microscopy reveals that the carbon foams have relatively large micron-scale pores separated by monolayer / few-layer graphene-like walls. The samples large surface area varying from ~1000 to 2500 m2/g. The nitrogen content can be varied over a wide range (e.g. <0.5 to 15 at%) by changing the precursor ratios and/or the pyrolysis temperature. In addition, the ratio of e.g. pyridinic to tertiary nitrogen bonding can be tailored by changing the pyrolysis temperature; the relative proportion of tertiary bonded nitrogen increases with increasing temperature.

Electrochemical characterization was performed by rotating ring-disk electrode (RRDE) voltammetry. The oxygen reduction reaction (ORR) activity increases with increasing temperature. Detailed analysis of the effect of nitrogen content and type will be presented. In particular, the samples prepared by this method are truly metal-free electrocatalysts and these results give insight into the role of nitrogen in the ORR in non-precious catalysts.

References

1. Lyth, S. M., Nabae, Y., Moriya, S., Kuroki, S., Kakimoto, M., Ozaki, J., Miyata, S. Carbon nitride as a nonprecious catalyst for electrochemical oxygen reduction. J. Phys. Chem. C113,20148–20151 (2009).

2. Lyth, S. M., Nabae, Y., Islam, M., Kuroki, S., Kakimoto, M., Miyata, S. Electrochemical Oxygen Reduction Activity of Carbon Nitride Supported on Carbon Black. J. Electrochem. Soc.158,B194–B201 (2011).

3. Wang, H., Maiyalagan, T. & Wang, X. Review on recent progress in nitrogen-doped graphene: Synthesis, characterization, and its potential applications. ACS Catal.2,781–794 (2012).

4. Shao, M., Chang, Q., Dodelet, J.-P. & Chenitz, R. Recent Advances in Electrocatalysts for Oxygen Reduction Reaction. Chem. Rev.acs.chemrev.5b00462 (2016). doi:10.1021/acs.chemrev.5b00462

5. Liu, J., Takeshi, D., Orejon, D., Sasaki, K. & Lyth, S. M. Defective Nitrogen-Doped Graphene Foam: A Metal-Free, Non-Precious Electrocatalyst for the Oxygen Reduction Reaction in Acid. J. Electrochem. Soc.161,F544–F550 (2014).

2668

, , , , , and

New carbon-based particles prepared using methane by microwave-assisted catalytic decomposition has been developed. The prepared carbon-based particles contained a nanographene-covered nickel, carbon nanotube, and others. In non-aqueous solvent, an oxygen reduction reaction (ORR) at the prepared carbon-based particles treated with acidic solution in advance was evaluated by using a porous microelectrode and milimeter-size diskelectrode based on electrochemical methods. The ORR was compared to commercially available glassy carbon, carbon fiber, Au, and Pt electrodes. As a result, the onset potentials for ORR at the prepared carbon-based particle were slightly better than other electrodes.

2669

, , and

One of the barriers hindering the commercialization of polymer electrolyte membrane fuel cell (PEMFC) is the high cost, in large part due to the expensive Pt-based catalyst for both the cathodic oxygen reduction reaction (ORR) and anodic hydrogen oxidation. The ORR requires more Pt catalyst than anodic hydrogen oxidation because of the sluggish kinetics. Therefore, ternary Fe/N/C material, as a non-Pt catalyst for ORR, is being studied intensively as one of the most promising alternatives for expensive Pt-based catalysts. However, the catalytic activity of Fe/N/C materials is considerably inferior to that of Pt-based catalyst in acidic medium. What's more, the stability of Fe/N/C is poor and the decay mechanism remains unclear.

Here, we found that iron salts can also greatly influence the ORR activity of Fe/N/C, and the use of Fe(SCN)3 instead of FeCl3 can nearly double the ORR activity. The improvement may be correlated with S doping and high surface area. The Pmax of PEMFC using this improved catalyst reached 0.94 and 1.03 Wcm-2 at 1 and 2 bar back-pressure, respectively. Moreover, we find that oxidized environment in PEMFC, mainly caused by oxygen, high voltage and hydrogen peroxide, can deteriorate Fe/N/C catalyst.

Figure 1

2670

, and

A new facile template-free method is presented to synthesize Fe-treated N-doped carbon (Fe/N-C) catalysts for oxygen reduction reaction (ORR) by employing a synthesis protocol of pyrolysis-leaching-stabilization (PLS) sequence of polypyrrole in the presence of ferric source, which serves dual purposes of an oxidant for pyrrole polymerization and an iron source. Each step in the PLS sequence is assessed in detail in terms of related structural properties of resulting carbon catalysts, and their effects on ORR activities are elaborated to confirm the validity of the current synthesis protocol. It is found that as-prepared carbon catalyst exhibits outstanding high catalytic activity in both alkaline and acidic conditions. The carbon catalyst prepared at pyrolysis temperature of 900 oC (FePPyC-900) shows remarkably high ORR activity with onset potential of 0.96 V (vs. RHE), which is similar to that of Pt/C, whereas the half-wave potential (E1/2) of FePPyC-900 is 0.877 V, which is more positive than that of Pt/C at the same catalyst loading amount in alkaline condition. Furthermore, the FePPyC-900 catalyst also illustrates exceptionally high activity in acidic condition with onset and half-wave potentials, which are almost comparable to those of the state-of-the-art Pt/C catalyst, which is rarely observed for non-Pt based carbon catalyst. In addition, the FePPyC-900 catalyst displays much better stability and methanol tolerance than the Pt/C and exhibits four electron transfer pathway in both alkaline and acidic conditions. Such extraordinary high ORR activity and stability the FePPyC samples can be attributed to the implementation of extra stabilization step in addition to conventional sample preparation steps of pyrolysis and subsequent leaching in current PLS synthesis protocol as well as to the use of highly conducting PPy as a single precursor of carbon and nitrogen in the presence of Fe.

2671

, , , , and

It is necessary to develop renewable and environmentally benign energy systems to offset the impact of human development on the environment. Polymer electrolyte membrane fuel cells (PEMFCs) can be a transformative technology but they are costly, with the Pt-based catalyst contributing most to the price of a PEMFC.1,2 Developing earth-abundant non-precious metal catalysts (NPMCs) for the oxygen reduction reaction (ORR) can significantly reduce the cost of PEMFCs. Advances in the performance of NPMCs have been made, however the connection between chemical and morphological properties and electrochemical performance are still not well understood.3,4 Difficulties in elucidating the exact chemical environment and morphology of highly active catalysts is largely due to their heterogeneous nature.1,2

A particularly promising NPMC catalyst was synthesized from iron nitrate and nicarbazin precursors using a sacrificial support method.5 By varying several parameters in this synthesis process, a set of iron containing, nicarbazin derived (Fe-NCB) catalyst materials were prepared with slight variations in chemical and structural properties. In order to deconvolute the contribution of different properties to electrochemical performance, measured in rotating disk electrode (RRDE) and membrane electrode assemblies (MEAs) a study spanning multiple sophisticated characterization techniques was employed.

Transmission electron microscopy (TEM) and energy dispersive x-ray spectroscopy (EDS) mapping were used to investigate the morphology of Fe-NCB catalysts and the relative distribution of nitrogen and iron species within the porous carbon network. Figure 1 shows a high-angle annular dark field (HAADF) image and variation in N/Fe ratio as determined by area-averaged EDS for one of the investigated Fe-NCB catalysts. This data was correlated to surface elemental composition and chemical speciation information from X-ray photoelectron spectroscopy (XPS), and by combining these techniques with electrochemical performance, the contributions of various active sites were investigated. The effect of different levels of porosity on the distribution of Nafion® ionomer and its effects on Fe-NCB catalyst performance were also studied. While TEM/EDS analysis conducted in ultra-high vacuum yields important information about a material's properties, it is limited as compared to the impact of characterization conducted in a realistic operating environment. To gain better understanding of composition-morphology-performance correlations, a series of Fe-NCB catalyst materials was investigated using an in-situ/in-operando TEM cell for electrochemical studies. In-situ/in-opreando observations of material changes under relevant conditions (in the presence of electrolyte, applied potential) add another layer of understanding to the material properties and provide direction for further improvements in the catalyst composition and structure.

References

1. Wood, K.; O'Hayre, R.; Pylypenko, S., Recent progress on nitrogen/carbon structures designed for use in energy and sustainability applications. Energy & Environmental Science 2014,7 (4), 1212-1249.

2. Bashyam, R.; Zelenay, P., A class of non-precious metal composite catalysts for fuel cells. Nature 2006,443 (7107), 63-66.

3. Jaouen, F.; Proietti, E.; Lefevre, M.; Chenitz, R.; Dodelet, J.; Wu, G.; Chung, H.; Johnston, C.; Zelenay, P., Recent advances in non-precious metal catalysis for oxygen-reduction reaction in polymer electrolyte fuel cells. Energy & Environmental Science 2011,4 (1), 114-130.

4. Proietti, E.; Jaouen, F.; Lefevre, M.; Larouche, N.; Tian, J.; Herranz, J.; Dodelet, J., Iron-based cathode catalyst with enhanced power density in polymer electrolyte membrane fuel cells. Nature Communications 2011,2; Wu, G.; More, K.; Johnston, C.; Zelenay, P., High-Performance Electrocatalysts for Oxygen Reduction Derived from Polyaniline, Iron, and Cobalt. Science 2011,332 (6028), 443-447.

5. Serov, A., Artyushkova, K., Niangar, E., Wang, C., Dale, N., Jaouen, F., ... Atanassov, P. (2015). Nano-structured non-platinum catalysts for automotive fuel cell application. Nano Energy, 16, 293–300. http://doi.org/10.1016/j.nanoen.2015.07.002

Figure 1

2672

, , , and

To date, Pt-based catalysts have remained the sole, though expensive, solution in electrocatalysis of the sluggish oxygen reduction reaction (ORR) at the hydrogen fuel cell cathode.1 Recently, platinum group metal-free (PGM-free) materials have drawn attention as promising low cost replacement for Pt-based ORR catalysts.2 A prevailing approach in the synthesis of highly active PGM-free catalysts is based on a two-stage heat treatment of precursors containing C, N and transition metals.2,3,4 Studies suggest that Fe-N moieties formed during the heat treatment may play critical role as active sites.5 However, this state-of-the-art synthesis approach involves the acid-leaching of excess non-active iron-rich phases, which inevitably damages some of the active sites. A prolonged second heat treatment at high temperature (700°C-1000°C) is essential to recover the lost active sites. This procedure not only increases the chemical demands and cost to produce ORR PGM-free catalysts, but, more importantly, may not lead to full recovery of all the active sites initially formed. More economical and efficient methods are in demand to purify the PGM-free catalysts.

In this presentation, we demonstrate an approach to magnetically purify active PGM-free catalysts while preserving the initial density of active sites. During the process of magnetic purification, the pyrolysis products are subjected to shear mixing in inert solvent to break down the agglomerations. The aggregates containing ferromagnetic phases, such as iron and iron carbides, are extracted efficiently in the magnetic field of a permanent magnet. The recovered active carbon-rich phase is collected and further characterized using X-ray diffraction, Raman spectroscopy, scanning electron microscopy, and scanning transmission electron microscopy to understand the chemical and physical properties of the active phase. Our results show that the ferromagnetic metal-rich phases are efficiently removed, leaving behind carbon-based phases. Electrochemical characterization and fuel cell testing show that the magnetically purified catalyst outperforms catalysts prepared with the conventional two-stage heat treatment. The nature of active sites in the purified carbon phases are also studied with X-ray photoelectron spectroscopy and discussed. This innovative approach enables the green and economical purification of active non-precious metal ORR catalysts, without compromising their catalytic performance.

Acknowledgement

Financial support for this research by DOE-EERE through Fuel Cell Technologies Office is gratefully acknowledged.

References

  • Spendelow, J.; Marcinkoski, J., DOE Fuel Cell Technologies Office, Fuel Cell System Cost-2014, (2014).

  • Wu, G.; Zelenay, P., Nanostructured Nonprecious Metal Catalysts for Oxygen Reduction Reaction. Acc. Chem. Res.46, 1878-1889 (2013).

  • Wu, G.; More, K. L.; Johnston, C. M.; Zelenay, P., High-Performance Electrocatalysts for Oxygen Reduction Derived from Polyaniline, Iron, and Cobalt. Science332, 443-447 (2011).

  • Zhao, D.; Shui, J.-L.; Grabstanowicz, L. R.; Chen, C.; Commet, S. M.; Xu, T.; Lu, J.; Liu, D.-J., Highly Efficient Non-Precious Metal Electrocatalysts Prepared from One-Pot Synthesized Zeolitic Imidazolate Frameworks. Adv. Mater.26, 1093-1097 (2014).

  • Holby, E. F.; Wu, G.; Zelenay, P.; Taylor, C. D., Structure of Fe–Nx–C Defects in Oxygen Reduction Reaction Catalysts from First-Principles Modeling. J. Phys. Chem. C118, 14388-14393 (2014).

2673

, , , , and

With their efficiency, high power density, and low/zero emissions, polymer electrolyte fuel cells (PEFCs) are recognized as promising energy converting system. However, high cost and insufficient durability currently hinder their commercialization. Especially, one of the major barriers is the high cost and low stability of platinum (Pt)-based catalysts 1). Thus, developing alternative catalysts with low cost and high stability has been strongly required for the commercialization of PEFCs.

In the effort to reduce the cost of PEFCs, we have been developing nitrogen doped titanium oxide (N- TiO2) as a non-noble metal electrocatalyst in NEDO (New Energy and Industrial Technology Development Organization, Japan) project. In this paper, we introduce a new and convenient synthesis method of N-doped TiO2electrocatalysts and their catalytic activities for oxygen reduction reaction (ORR).

In this study, we used phthalocyanine (Pc) as nitrogen source in order to dope nitrogen into TiO2. For the high surface area of electrocatalyst, TiO2 nanoparticles were deposited on high surface area carbon supports (C). The carbon supported TiO2 (TiO2/C) and phthalocyanine (Pc) were mixed and heat-treated at 900oC in inert gas atmosphere containing a small amount of hydrogen (H2) and oxygen (O2) in order to finally obtain N-TiO2/C electrocatalysts.

TiO2 nanoparticles of ca. 20 nm size deposited on the carbon supports were observed by transmission electron microscope (TEM), indicating that the synthesized N-TiO2/C may have considerably high surface areas. In X-ray photoelectron spectroscopy (XPS) and X-ray diffraction (XRD) studies, it was found that the XPS core-level peaks of Ti was very similar to that of Ti-N binding energy and the XRD peak of TiO2 shifted to higher angles, respectively. These results well explain that the nitrogen was successfully doped into TiO2 nanoparticles by our synthesis process. Fig. 1 compares the ORR catalytic activities of the synthesized N-TiO2/C and TiO2/C. The synthesized N-TiO2/C electrocatalyst shows significantly higher ORR catalytic activity than that of TiO2/C. It is believed that the enhanced catalytic activity of N-TiO2/C is due to the doping effect of nitrogen into TiO2.

 

Acknowledgements

The authors wish to thank NEDO (New Energy and Industrial Technology Development Organization) for financial support.

 

References

  • Ken-ichiro Ota, Yoshiro Ohgi, Kyung-Don Nam, Koichi Matsuzawa, Shigenori Mitsushima, Akimitsu Ishihara, J. Power. Sources., 196, 5256, ( 2011).

Figure 1

2674

, , , , , , and

Introduction

The development of polymer electrolyte fuel cells (PEFCs) has attracted tremendous interest, because they have many advantages, including high power density, high energy conversion efficiency, and low operating temperatures. Recently, PEFCs has been already commercialized as power supply of fuel cell vehicles and residential use. Generally, platinum-based alloys supported carbon black are used as catalysts in both the anode and cathode of PEFCs. However, Pt is high cost and limited resources. In addition, carbon support is fundamentally unstable under cathode conditions to oxidize at high potential region. The development of alternative catalysts with high stability for oxygen reduction reaction (ORR) is required for wide spread of PEFCs. This catalyst needs two components, precious-metal-free active sites for ORR and carbon-free electro-conductive supports. We have focused on group 4 and 5 metal oxide-based compounds as both precious-metal-free materials with active sites and electro-conductive materials. As an alternative support of carbon , we focused on titanium Magnéli sub-oxides (TinO2n-1; 4≦n≦10), particularly Ti4O7, because of its high chemical stability in acid electrolyte and high conductivity. However, the formation of Ti4O7 needs to reduce at high temperature, 1050 oC, under 4% H2 for 20-60 h1). Such a high temperature caused the drastic particle growth and morphology destruction. In this study, we applied solid-phase reduction at low temperature by using NaBH42) to investigate the appropriate condition for the formation of reduced titanium oxides as support.

 

Experimental

Rutile type TiO2 powder (ca. 30 nm) was mixed with NaBH4. The mixture was heat-treated under Ar atmosphere at 300 - 400oC for 24 h. After heat-treatment, we washed products with deionized water and 0.1 M hydrochloric acid to obtain the reduced titanium oxide powders. X–ray diffraction spectroscopy (XRD, Rigaku Ultima IV) was performed to determine the crystalline structure of the powder.

Results and discussion

Figure 1 shows the XRD patterns of the reduced powders heat-treated at various temperatures. The peaks near 27.4o identical to first main peak of TiO2 rutile were gradually decreased from 300 to 360 oC, and the peaks near 23.8 o and 33.0 o identical to Ti2O3 appeared above 350 oC, indicating that the formation of bulk Ti2O3 gradually proceeded above 350 oC. On the other hand, the temperature over 375 oC caused transition to amorphous phase. These results suggested that it was difficult to form the Ti4O7 phase by solid-phase reduction at low temperature. We need to find an appropriate condition to form the Ti4O7 phase. On the other hand, it was reported that the bulk Ti2O3 has much larger conductivity than TiO23). However, the all reduced powders showed large resistance measured by two-terminal resistance measurement. The large resistance was caused by the contact resistance due to the passive layer formed on the nano-particles. In order to reduce the contact resistance, we think that the formation of the electro-conductive network structure is effective by control the morphology of the reduced oxides.

Acknowledgments

The authors thank New Energy and Industrial Technology Development Organization (NEDO) for financial support. The Institute of Advanced Sciences (IAS) in Yokohama National University was supported by the MEXT Program for Promoting Reform of National Universities.

References

1) M. Hamazaki, A. Ishihara, Y. Kohno, K. Matsuzawa, S. Mitsusima, and K. Ota, Electrochemistry, 83, 817 (2015).

2) S. Tominaka, Chem. Commun., 48, 7949 (2012).

3) S. Tominaka, Inorg. Chem, 51, 10136 (2012)

Figure 1

2815

, , , , and

The limitations of low-temperature polymer electrolyte fuel cells (LT-PEFCs) can be addressed by elevating the operating temperature above 100 °C. High-temperature (HT-) PEFC systems benefit from a simplified heat and water management, a higher tolerance toward fuel impurities such as carbon monoxide and faster oxygen reduction reaction (ORR) kinetics [1,2]. Operating temperatures higher than 100 °C involve the usage of other electrolytes than conventional perfluorosulfonic acid membranes (e.g. Nafion) due to the absence of liquid water in the system and the resulting loss of proton conductivity in the membrane. For PEFC operation in the range of 150-200 °C, humidification-independent phosphoric acid doped polybenzimidazole (PBI) is typically used as membrane material [1,2]. However, the usage of phosphoric acid determines also the activity of the catalysts due to the strong adsorption of phosphate species on the platinum surface and the corresponding reduced number of available active sites. In order to overcome the significant deactivation of Pt, the employed catalyst system needs to be modified. It was shown that combining Pt with Au via electroplating results in increased ORR activity also in the presence of phosphoric acid [3].

We present a facile and straight-forward synthesis of Pt-Au nanoparticles supported on high surface area carbon with varying metal stoichiometry. The as-prepared Pt-Au/C catalysts are further benchmarked to a standard Pt/C catalyst for usage as cathode catalysts in HT-PEFC systems. The ORR activity and stability of the catalyst systems are evaluated ex situ by means of cyclic voltammetry, Levich-Analysis and accelerated stress tests using a rotating disk electrode (RDE) setup. Furthermore, the catalyst systems are characterized in situ by means of polarization curves, continuous operation, accelerated stress tests and electrochemical impedance spectroscopy measurements at single cell and stack level.

The as-prepared Pt-Au/C catalysts show increased tolerance toward phosphoric acid in comparison to the standard Pt/C (see Figure 1 and Figure 2).

Acknowledgment

Financial support was provided by The Climate and Energy Fund of the Austrian Federal Government and The Austrian Research Promotion Agency (FFG) through the program Energieforschung (e!Mission).

[1] T. Ossiander, M. Perchthaler, C. Heinzl, F. Schönberger, P. Völk, M. Welsch, A. Chromik, V. Hacker, C. Scheu, Influence of membrane type and molecular weight distribution on the degradation of PBI-based HTPEM fuel cells, J. Memb. Sci. 509 (2016) 27–35.

[2] A. Schenk, C. Grimmer, M. Perchthaler, S. Weinberger, B. Pichler, C. Heinzl, C. Scheu, F.-A. Mautner, B. Bitschnau, V. Hacker, Platinum–cobalt catalysts for the oxygen reduction reaction in high temperature proton exchange membrane fuel cells – Long term behavior under ex-situ and in-situ conditions, J. Power Sources. 266 (2014) 313–322.

[3] J.E. Lim, U.J. Lee, S.H. Ahn, E. Cho, H.J. Kim, J.H. Jang, H. Son, S.K. Kim, Oxygen reduction reaction on electrodeposited PtAu alloy catalysts in the presence of phosphoric acid, Appl. Catal. B Environ. 165 (2015) 495–502.

Figure 1

I01-E Poster Session - Oct 5 2016 6:00PM

2675

and

Platinum is the most widely used catalyst for fuel cells. The morphology of platinum has great influences on its catalytic activity and selectivity. Thus, the morphology controlled synthesis of platinum nanocrystals is attracting more and more attention in recent years [1-3]. Polyvinylpyrrolidone (PVP) is one of the most commonly used surface active agent for the morphology control of platinum nanocrystal. However, since PVP could strongly adsorb on the platinum surface and was difficult to remove thoroughly, the platinum surface couldn't be utilized effectively and thus its catalytic activity will be reduced to a certain extent.

This work focused on the decontamination of PVP on platinum by electrochemical treatment. Platinum nanocrystals was synthesized by colloidal method using chloroplatinic acid as the metal precursor and PVP K30 as a stabilizer. Spherical, cubic, octahedral and truncated octahedral platinum nanocrystals were obtained by adding different concentrations of silver nitrate to the above colloidal solution. The morphology of platinum nanocrystals was characterized by transmission electron microscope (TEM). It was found that the as-preparation platinum nanocrystals were well dispersed with a narrow size distribution. Electrochemical decontamination was applied to the platinum nanocrystals by short pulse potential step before electrocatalytic performance measurements. Repeating potential steps were set up between high voltage and low voltage. As the potential step proceeding, more and more surface sites were exposed, which could be confirmed from the more and more clear hydrogen and oxygen adsorption/desorption peak of the cyclic voltammograms (CV) (as shown in Fig. 1A). The CV curves stabilized eventually. The electrocatalytic activity of the electrochemical cleaned platinum nanocrystals was characterized by CV (as shown in Fig. 1B) and chronoamperometry measurements. It was found that truncated octahedral platinum nanocrystals showed the best catalytic activity for metha nol electrooxidation in 0.5 mol·L-1 H2SO4 + 0.5 mol·L-1 CH3OH solution. The methanol electrooxidation peak current was much higher than the current on platinum nanocrystals with other morphology. This study put forward a simple and effective electrochemical decontamination method of the platinum surfaces capped by PVP.

Figure 1

2676

, , , , , and

  • Introduction

    One of the challenges to be addressed in facilitating widespread dissemination of polymer electrolyte fuel cell (PEFC) is to reduce the amount of Pt catalysts used in PEFC. The rate of oxygen reduction reaction (ORR) (O2 + 2H+ + 4e- → 2H2O) is relatively lower at Pt surface and quite a large amount of Pt catalysts is required to increase O2 reduction rate, which is a serious problem for cathode electrodes. For instance, 100g of Pt is needed for a fuel cell vehicle. Pt usage of this sort is impractical in terms of resources as well as economics. In order to ensure widespread commercialization of PEFC, the development of non-Pt catalysts is essential. The other challenge is to improve the stability of catalysts. The activity of Pt catalysts is known to be decreased by surface poisoning during long-term use. The objective of this study is to realize non-Pt catalyst with ORR activity and higher stability by using amorphous carbon (a-C) incorporating nitrogen atoms. It has been reported that carbon alloy with quinolizinium structure at graphene edges exhibited the higher ORR activity [1]. Our research group has reported that a-C including amorphous phase and nano-sized sp2 cluster could be a conductive material by incorporating nitrogen atoms. It works as an ideal polarizable electrode with higher overpotential toward water discharge and higher stability to electrochemically-induced corrosion [2]. In the study, a-C based catalysts with higher ORR activity and higher stability was tried to be realized by introducing quinolizinium structure at the sp2cluster surface in a-C.

  • Experimental

    a-C catalysts with quinolizinium structure were synthesized by plasma-enhanced CVD method. Vaporized acetonitrile + pyridine (6:4 at a molar ratio) mixture was used as a source material. a-C catalysts were synthesized under various conditions of RF power and substrate temperature. Chemical compositions and structures of the resulting a-C were examined by XPS and Raman spectroscopy. The reactivity of ORR was examined by linear sweep voltammograms using rotating disk electrode (RDE).

  • Results and Discussion

    The size of sp2 clusters of the resulting a-C catalysts was estimated from Raman spectra. The ratios of sp2/sp3-hybridized carbons and the density of N atoms in quinolizinium structure were estimated from XPS results. Hydrodynamic voltammograms using RDE (Figure 1) were carried out to estimate the number of electrons for O2 reduction in O2 saturated 1M KOH solution. The peak correspond to N atoms in quinolizinium structure (graphitec N) was included in N1s peak in XPS spectra of all a-C catalysts and observed at approximately 402 eV. These results indicate that functional groups active for O2 reduction are introduced in graphite regions in a-C. With increasing RF power of CVD synthesis, the ratio of sp2/sp3 from XPS exhibited a tendency to increase, the size of sp2 cluster to decrease, and the density of N atoms in quinolizinium structure to increase. It indicates that the active site for O2 reduction can be controlled by deposition condition. The number of electrons for O2 reduction was enhanced by these structural changes and reached 2.82 at a maximum (Figure 1). The value over 2 suggests that a-C catalysts with the activity of 4-electron ORR were successfully realized. It can be expected that the ORR activity is further enhanced by optimizing sp2/sp3, the size of sp2cluster, and the density of N atoms in quinolizinium structure.

  • References

    [1] Hainbo Wang et. al.ACS Ctal., 2, 781-794, (2012).

    [2] Yoriko Tanaka et. al., Electrochimica Acta, 56 (3), 1171-1182 (2011).

2677

, and

Fuel cell is environment-friendly power source which has high efficiency in generating electricity. Polymer electrolyte fuel cell uses hydrogen as a fuel that only produces water as byproduct which does not cause any pollution. Currently, commercialized Nafion-based polymer electrolyte membrane fuel cell (PEMFC) applied Pt/C as both cathode and anode catalyst which cause cost inefficiency in manufacturing fuel cell stack.

Anion exchange membrane fuel cell (AEMFC) is being developed to combine advantages of both PEMFC and traditional alkaline fuel cell (AFC). By adopting solid membrane as electrolyte, leakage problem and insoluble salt precipitation could be solved. Furthermore, alkaline condition enhances oxygen reduction reaction (ORR) kinetics which obstructs cell power in PEMFC. However, stability of membrane is still in issue, and hydrogen oxidation reaction (HOR) kinetics are significantly diminished [1], which leads to higher Pt loading. This research is focusing at slow HOR kinetics and reducing catalyst cost by replacing Pt into non-precious metals such as transition metals.

One of the well-known HOR catalyst for alkaline fuel cell is nickel. Several studies show nickel has catalytic activity of hydrogen oxidation in alkaline media [2,3], but it's not enough to overcome precious metals such as platinum. Therefore, diverse researches tried to enhance activity of nickel by expanding surface area or doping other elements [4]. However, these catalytic activities were only tested in 6 M KOH, 60~80 ℃ condition, which is identical to real cell operating condition.

NiMo catalyst was synthesized by electrodeposition method which is cost-effective way to produce. Nickel chloride is used for nickel source, and sodium molybdate worked as molybdenum source. Galvanostatic electrodeposition was applied to synthesize NiMo catalyst. Rotating disk electrode (RDE) was used to remove evolved hydrogen efficiently.

HOR activity was measured in hydrogen saturated 0.1 M KOH solution. RDE was used to control diffusion with 1600 rpm speed. Experiment was held at room temperature. To remove capacitive current, steady state polarization method was applied for measuring HOR activity.

Figure 1 shows HOR current of commercial Pt/C, electrodepositied NiMo, Ni. Adding small amount of Mo in Ni increases HOR current significantly. At overpotential of 20 mVRHE, Pt/C shows current density of 1.12 mA/cm2, where NiMo has 0.86 mA/cm2which is about 77% of Pt/C HOR activity.

Reference

[1] W. Sheng, et al., J. Electrochem. Soc. 157 (11), B1529 (2010)

[2] T. Tomida, et al., J. Electrochem. Soc. 136 (11), 3296 (1989)

[3] H. K. Lee, et al., Mat. Chem. Phys., 55, 89 (1998)

[4] J. A. Linnekoski, J. Fuel Cell Sci. Tech., 4, 45 (2007)

Figure 1

2678

, and

Immobilized benzimidazolium cations as functional groups in anion exchange polymers can be used in alkaline anion exchange membrane fuel cells (AAEM-FCs), electrolyzers, or water purification systems, but are prone to hydroxide attack. Steric protection by proximal methyl groups has been shown to drastically increase hydroxide stability (A. Wright, S. Holdcroft ACS Macro Lett20143, 444-447.). To further improve stability, model compounds, representing the ion exchange sites of AAEMs, were investigated for their hydroxide stability. By means of density functional theory (DFT), we studied degradation mechanisms, such as de-methylating SN2 reaction of methylated benzimidazolium cations with hydroxide ions and the attack of hydroxide on the C2 position of the benzimidazolium. Some of these results have also been compared to experimental stability tests of model compounds and polymers (A. G. Wright, T. Weissbach, S. Holdcroft Angew. Chem. Int. Ed.2016, 55, 4818-4821.). The findings of this study enable the design of new materials for AAEM-FCs.

2679

, , , and

Polymer electrolyte fuel cells (PEFCs) directly fueled liquid fuels are suitable power source for portable energy applications. Liquid alcohol fuels such as methanol and ethanol are attractive energy carriers with good energy density and low cost. However, the problems of their low flash points (below room temperature) and/or toxicity should be overcome for the safe application of alcohol-fueled PEFCs. The liquid oligomers of poly oxymethylene dimethyl ether (POMMn: CH3-O-(CH2-O)n-CH3, n = 3~8) have been proposed as alternative fuels, due to their advantages, including their higher energy density, lower toxicity and higher flash point1-3). In this study, we synthesized Pt and PtRu catalysts supported on Ta-SnO2 (Pt/Ta-SnO2, PtRu/Ta-SnO2) for the direct oxidation of POMM and evaluated their performance in a PEFC.

The Ta-SnO2 support4) was synthesized by the flame spray synthesis method. The oxide supports obtained were nanometer-sized particles with a carbon-like fused-aggregate structure. Pt and PtRu catalysts were loaded on the Ta-SnO2 by the colloidal method. The Pt/Ta-SnO2 (Pt loading amount, 14.0 wt%; particle size, 6.3 nm) and PtRu/Ta-SnO2 (Pt/Ru loading amount, 6.3/3.5 wt%; particle size, 2.2 nm) were obtained after heat treatment. The Pt and PtRu particles were highly dispersed on the Ta-SnO2 support. As shown in the high-angle annular dark-field STEM (HAADF-STEM) image and STEM-EDX line profile, Sn metal diffused into the Pt catalysts with the sintering process. X-ray photoemission spectroscopic (XPS) Sn 3d5/2 spectra for each catalyst indicated that part of the Sn was converted from Sn4+ to Sn0. We concluded that the catalysts formed binary Pt-Sn and ternary Pt-Ru-Sn alloys during the sintering procedure (Fig. 1). The onset potential for POMM2 oxidation on PtRu/Ta-SnO2, which was measured by a half-cell with a rotating disk electrode (RDE), was 0.3 V less positive than that of a commercial Pt2Ru3/CB catalyst (Fig. 2). The mass activity for POMM2 oxidation of PtRu/Ta-SnO2 at 0.5 V was also 3.5 times higher than that for a commercial Pt2Ru3/CB catalyst. The anodic activity of Pt/Ta-SnO2 was also higher than that of the commercial catalyst, by a factor of 2. The electrochemical performances of these anode catalysts were also evaluated by the use of membrane-electrode assemblies (MEAs, NRE117 membrane, commercial Pt/CB(TEC10E50E) 0.5 ± 0.025 mgPt cm2 cathode, Pt/Ta-SnO2 or PtRu/Ta-SnO2 0.25 or 0.27 mgPt cm2 anode, Fig. 3). The current densities of the MEAs using the Pt/Ta-SnO2 and PtRu/Ta-SnO2 catalysts as the anodes and Pt/CB as the cathode (POMM2 simulated fuel, O2 oxidant, 80oC operation temperature) were more than 10 times higher than that using the commercial Pt2Ru3/CB anode. The current density at 0.6 V for formaldehyde fuel supplied to an MEA applying the Pt/Ta-SnO2anode was 15 times higher than that for the commercial anode, even though the performance for methanol remained at the same level. We consider that the Sn on the surface of the Pt catalyst may suppress CO poisoning by weakening the adsorption strength of CO formed during the fuel oxidation processes, even at such low potentials as 0.2 to 0.4 V, and directly accelerate the oxidation of carbon monoxide at more positive potentials by a bifunctional mechanism, similar to that for Pt-Ru catalysts.

Acknowledgements

This work was partially supported by funds for the A-STEP project from the Japan Science and Technology Agency (JST), JSPS KAKENHI (B24350093) from the Ministry of Education, Culture, Sports, Science and Technology (MEXT), and the SPer-FC project from the New Energy and Industrial Technology Development Organization (NEDO) of Japan.

References

1) D. Devaux, H. Yano, H. Uchida, J.L. Dubois, M. Watanabe, Electrochim. Acta 56 (2011) 1460.

2) S. Baranton, H. Uchida, D.A. Tryk, J.L. Dubois, M. Watanabe, Electrochim. Acta 108 (2013) 350.

3) K. Kakinuma, I.T. Kim, Y. Senoo, H. Yano, M. Watanabe, and M. Uchida, ACS Appl. Mater. Interfaces, 6 (2014) 22138.

4) Y. Senoo, K. Taniguchi, K. Kakinuma, M. Uchida, H. Uchida, S. Deki, and M. Watanabe, Electrochem. Commun., 51 (2015) 37.

Figure 1

2680

Anion-exchange membrane direct ethanol fuel cells are attractive owing to the fact that: i) the alkaline media speeds up the electro-kinetics of both the ethanol oxidation reaction (EOR) and oxygen reduction reaction (ORR), ii) low-cost Pd catalyst presents better catalytic activity for the EOR than Pt in alkaline environment, and iii) the ethanol possesses the advantages of high-specific energy and producing from cellulosic biomass.

Carbon foam promises to be a better monolithic catalyst support due to its three-dimensional honeycomb-type open cell structure, resulting in the high porosity and the large specific surface area [1]. In this work, the carbon foam-supported PdNi nanocatalysts (PdNi/CF) were synthesized by the layer-reduction method using NaBH4 as a reductant [2], and characterized by different physicochemical methods. The experimental results indicated that the PdNi nanocatalysts were evenly dispersed on the surface of the skeleton of the carbon foam, facilitating the mass and charge transfer. The electrocatalytic activity of PdNi/CF was optimized by testing different Pd/Ni atomic ratios. It has been demonstrated that the electrocatalytic activity of PdNi/CF was higher than that of Pd/CF, suggesting promising application in alkaline direct ethanol fuel cells.

References

[1] Y.S. Li, J.H. Lv, Y.L. He, A Monolithic Carbon Foam-Supported Pd-Based Catalyst towards Ethanol Electro-Oxidation in Alkaline Media, J. Electrochem. Soc., 2016, 163 (5): F424-F427.

[2] Y.S. Li, Y.L. He, Layer Reduction Method for Fabricating Pd-coated Ni Foams as High-performance Ethanol Electrode for Anion-exchange Membrane Fuel Cells, Rsc Adv., 2014, 4 (32): 16879-16884.

2681

, and

Introduction

Anion exchange membrane fuel cells have recently attracted considerable attention as promising energy conversion devices because of possible use of non-precious metals as electrocatalysts such as Ni and Co. In order to achieve high fuel cell performance, it is essential to develop electrocatalysts that have high oxygen reduction reaction (ORR) activity and durability in alkaline media. Previously, our group reported that PtCo alloy nanoparticles with controlled particle size, alloy composition, and metal loading could be prepared by nanocapsule method.1 The PtCo/C thus obtained showed high ORR activity and reasonable durability in acidic media compared to those of the commercial Pt/C and Pt/C catalysts.2 In contrast, ORR properties of Pt alloy catalysts in alkaline media have not been investigated as much. In this study, we report ORR activity, and durability of our Pt/Co catalysts prepared by nanocapsule method in alkaline media, in particular, the effect of alloy composition on these properties. The results will be compared with those of the commercial Pt/C and PtCo/C catalysts in details.

Experimental

Platinum/cobalt alloy nanoparticle catalysts were prepared by nanocapsule method,1 and supported on acetylene black (AB) that had 800 m2 g-1 specific surface area. The prepared catalysts (PtCo/AB) were characterized by X-ray diffraction (XRD), transmission electron microscopy (TEM), and X-ray fluorescence (XRF). The loaded amounts of the metals on carbon support were quantified by thermogravimetric analyses (TGA) in air from room temperature to 600 °C. Electrochemical measurements were performed with a typical rotating ring disk electrode (RRDE) equipment with a gas-tight Pyrex glass cell in 0.1 M KOH. A ring-shaped platinum wire and a reversible hydrogen electrode (RHE) were used as the counter electrode and the reference electrode, respectively. All electrode potentials are stated relative to the RHE. The ORR catalytic activity was measured at 25 °C under air saturated conditions. The durability test of PtCo/AB was performed in 0.1 M KOH saturated with nitrogen at 40 °C. The potential was stepped between 0.6 V and 1.0 V, with a holding period of 3 sec at each potential (6 sec for one cycle). After a given number of potential step cycles, cyclic voltammetry and RRDE measurement were performed as mentioned above to examine the changes in ECA and the ORR activities.

Results & discussion

XRD patterns of PtCo/AB showed characteristic peaks assignable to platinum/cobalt alloys and carbon support. TEM images revealed that the obtained catalysts contained platinum/cobalt alloy nanoparticles well-distributed on AB support. The average particle sizes were ca. 3 nm. TG measurement suggested that metal loading amounts were ca. 20 wt% nearly comparable to the target values. Five catalysts with different compositions (Pt100-xCox: x = 31, 39, 42, 51, 73 atom%) were obtained. Cyclic voltammograms of PtCo/AB catalysts indicate characteristic peaks of oxidation/reduction of platinum and cobalt and hydrogen adsorption/desorption. The onset potential for ORR of PtCo/AB catalysts were higher than that of the commercial Pt/C catalyst. It was found that the PtCo/AB (x = 34 and 42) was the most active for ORR. The highest mass activity (MA) at 0.85 V was ca. 767 A g-1Ptat the optimum composition. The optimum composition of Co was higher than that (ca. 25 atom%) in acidic media. The yield of hydrogen peroxide via two electron transfer mechanism was similar to that of the commercial Pt/C catalyst. In the durability test, however, MA of PtCo/AB catalysts decreased with the potential step cycle. After 30000 cycles, the MA of PtCo/AB catalysts was comparable to that of the commercial Pt/C catalyst, implying the dealloying during the durability test. Details of the differences between in alkaline media and in acidic media will be discussed.

Acknowledgement

This work was partly supported by Japanese Science Technology Agency (JST), CREST.

References

1. H. Yano, M. Kataoka, H. Yamashita, H. Uchida, and M. Watanabe, Langmuir, 23, 6438 (2007).

2. H. Yano, J. M. Song, H. Uchida, and M. Watanabe, J. Phys. Chem., C, 112, 8372-8380 (2008).

Figure 1

2682

, and

Polymer electrolyte fuel cells (PEFCs) are considered to be one of the prospective power sources for automobile and domestic use due to their outstanding energy conversion efficiency and compactness. Recently, anion-exchange membrane fuel cells (AEMFCs) are promising due to the practicablity of eliminating the need for platinum-group metal catalysts, and employing non-precious metal electrocatalysts such as nickel, cobalt, silver are feasible, conducing a potentially low-cost technology. Also, an alkaline environment can inherently enhance the electrochemical kinetics of reaction such as the oxygen reduction reaction (ORR) on cathode, enabling higher power density and energy conversion efficiency. Hence, AMEFCs are attracting strong interest as the substitutes of proton-exchange membrane fuel cells requiring platinum-group metal catalysts[i]. However, the AEMFCs are still suffer from low performance and low OH- conductivities. In order to enhance the performance and durability, it's required to enhance the ionic conductivity of anion-exchange electrolyte for both membrane and electrocatalyst. Typically, the fabrication of anion-exchange electrolyte is required for both polymer main chains and side chains which include cationic functional groups and charged anions[ii]. In order to enhance the anion conductivity, increasing the cationic density of anion-exchange electrolyte was considered to be one of the promising method for designing anion-exchange electrolyte with superior performance[iii].However, the critical drawback of anion-exchange electrolyte with high cationic density of side chains is their water solubility, due to their hydrophilicity of ionic moieties, which may cause the loss of anion-exchange electrolyte by leaching in water generated in the humidified AMEFCs cell.

To overcome this issue, approach of hydrophobic part was introduced by condensation copolymerization to inhibit their water solubility and leaching, while, the reduction of ionic conductivity has been found. Also, another approach of crosslinking by directly casting on the electrode to insolubilize the electrolyte was reported[iv], while the uniform modification of the surface of the electrocatalyst is probably difficult.

In our previous studies, we suggested an anchoring of the anion-exchange electrolyte on a polymer backbone, which can prevent the leaching of the anion-exchange electrolyte with high cationic density, uniformly and individually wrap on the surface of carbon supporting materials. We chose multi-walled carbon nanotubes (MWNT) as the carbon supporting material since they provide an outstanding electrical conductivity and lower impurities[v]. We used polybenzimidazole (PBI) as polymer backbone of anion-exchange electrolyte since we found that PBI can uniformly and individually wrap MWNT based on π-π interactions. Since high yield chemical modification of PBI on MWNT has already being reported, this strategy is expected to homogeneously decorate supporting material by anion-exchange electrolyte, which can probably offering a high ionic conductivity. Furthermore, PBI was chosen because Pt nanoparticles can be loaded onto the PBI wrapped MWNT (MWNT/PBI) with a homogeneous distribution in high yield.

Based on this strategy, we studied on grafting approach by anchoring 1,4-diazabicyclo[2,2,2]octane (DABCO) polymer moieties as anion-exchange electrolyte onto MWNT/PBI (MWNT/PBI-g-DABCO), and loaded Pt nanoparticles onto MWNT/PBI-g-DABCO to fabricated MWNT/PBI-g-DABCO/Pt as an electrocatalyst for AEMFCs. This approach enabled to developed anion-exchange electrolyte for AEMFCs with high content of cationic moieties and water insolubility[vi].

In this study, instead of grafting approach, we designed a new approach to introduce cationic group by methylation of PBI, due to the advantages of convenient synthesis, high potential of ionic conductivity[vii], and excellent durability under alkaline environment[viii]. Hence, we fabricated methylated polybenzimidazole (mPBI) as electrolyte for AEMFCs electrocatalyst as well as the membrane (Fig. 1). Single cell test of membrane electrode assembly (MEA) using mPBI as electrocatalyst wrapped on MWNT was performed using H2 and conditioned air as the fuel for anode and cathode, respectively, and power density of 24.6 mWcm-2was obtained.

[i] Z. Chen et al., Int. J. Hydrogen energy2014, 39, 18405.

[ii] K. Nijmeijer et al., J. Membrane Sci. 2011, 377, 1.

[iii] M. A. Hickner et al., Macromolecules2013, 46, 9270.

[iv] Park et al., Int. J. Hydrogen Energy2014, 39, 16556.

[v] N. Nakashima et al., Small. 2009, 5, No. 6, 735.

[vi] N. Nakashima. Submitted.

[vii] Sun et al., Int. J. Hydrogen Energy 2011, 36, 11955.

[viii] S. Holdcroft et al. J. Am. Chem. Soc2012, 134, 10753.

Figure 1

2683

, , , and

Non-precious-metal catalysts (NPMCs) for oxygen reduction reaction (ORR) have been of tremendous interests in energy conversion and storage devices. Among several classes of NPMCs, iron and nitrogen supported on carbon (Fe-N/C) have shown the most promising ORR activity. It has been widely suggested that an active site structure for Fe-N/C catalysts contains Fe-Nx coordination. However, the preparation of Fe-N/C catalysts mostly involves a high-temperature pyrolysis step, which generates not only Fe-Nx sites, but also a significant portion of less active large iron-based particles. This poses a great challenge to rational design of Fe-N/C catalysts with abundant Fe-Nx species. We developed "silica-protective-layer-assisted" synthetic approach that can preferentially generate the catalytically active Fe-Nx sites in Fe-N/C catalysts while suppressing the formation of large Fe-based particles. The catalyst preparation consisted of an adsorption of Fe porphyrinic precursor on carbon nanotubes (CNTs), silica layer overcoating, high-temperature pyrolysis, and silica layer etching, which yielded CNTs coated with thin layer of porphyrinic carbon (CNT/PC) catalysts. In situ X-ray absorption spectroscopy during the preparation of CNT/PC catalyst revealed that the interaction between the silica layer and Fe-N4 in a porphyrin precursor appears to protect the Fe-N4 site and to prevent the formation of large Fe-based particles. The CNT/PC catalyst showed very high ORR activity and remarkable stability in alkaline media. Importantly, an alkaline anion exchange membrane fuel cell (AEMFC) with a CNT/PC-based cathode exhibited record high current and power densities among NPMC-based AEMFCs. In addition, a CNT/PC-based cathode exhibited a high volumetric current density of 320 A cm-3 in acidic proton exchange membrane fuel cell, comparable with 2020 DOE target (300 A cm-3). We further demonstrated the generality of this synthetic strategy to other carbon supports including reduced graphene oxides and carbon blacks.

2684

, , , and

Introduction

Direct ethanol fuel cells (DEFCs) have attracted attention for their practical high energy density and low environmental burden. To generate electricity with high efficiency, ethanol needs to be completely oxidized to CO2. To date, Xia et al. have reported that the ethanol oxidation reaction (EOR) is sensitive to the topmost surface atomic arrangements of Pt [1]. Furthermore, Pt-based alloy catalysts have been studied extensively for EOR with low overpotentials. Therefore, it is important to clarify relations between EOR activity and topmost surface structures of Pt-based alloys to develop practical Pt-based alloy catalysts for DEFCs. In this study, we focused on Sn/Pt catalysts which have been studied well as alloy catalysts for EOR. On-line electrochemical mass spectrometry (OLEMS) should be effective for in-situ analysis of gaseous reaction products. Therefore, the relations of EOR onset potentials, potential-dependent gaseous products, and catalyst surface structures are investigated for well-defined Sn/Pt(hkl) surfaces prepared by vacuum depositions of Sn on Pt(hkl) single crystal substrates.

Experimental

Sample fabrication processes of the well-defined Sn/Pt(hkl) were conducted in ultra-high vacuum (UHV). Pt single crystal substrates (Pt(hkl): Pt(111), (110) and (100)) were cleaned by Ar+ sputtering and subsequent annealing at 1273 K in UHV. 0.1 nm-thick Sn was deposited onto the cleaned Pt(hkl) by an electron-beam evaporation method at room temperature, followed by annealing at 1000 K to flatten the topmost surfaces. The UHV-prepared Pt(hkl) and Sn/Pt(hkl) were transferred from UHV chambers to the electrochemical system set in a N2-purged glove box without being exposed to air. Cyclic voltammetry (CV) of the surfaces was conducted in N2-purged 0.1 M HClO4. Subsequently, ethanol was added to the solution to prepare a 2 M solution of ethanol in 0.1 M HClO4, and then CV and OLEMS measurements were performed.

Results and Discussion

The anodic currents of EOR for the Pt(hkl) and Sn/Pt(hkl) surfaces recorded in ethanol added HClO4 solutions are presented in Fig.1 (a). EOR onset potentials of the Sn/Pt(hkl) surfaces (solid curves) shift negatively in comparison to the corresponding Pt(hkl) surfaces (dashed curves). The negative potential shifts suggest that the Sn/Pt(hkl) surfaces exhibit higher EOR activity than the corresponding Pt(hkl) surfaces, probably because of the ensemble effect of the surface Pt and Sn atoms [2]. Fig.1 (b) shows the OLEMS results of the Sn/Pt(hkl) surfaces. The onset potentials of CO2 evolution are summarized in Table.1. Fig.1 (b) and Table.1 suggest that potential sweeps in the anodic direction (0.2 V to 0.9 V) cause CO2 evolution, and that the EOR onset potentials depend on the substrate atomic arrangements of Pt. In particular, the Sn/Pt(110) shows the lowest EOR onset potentials, as judged by EOR currents (Fig.1 (a)) and CO2 evolution (Fig.1 (b)). The results can be attributed to the effectiveness of Pt(110) steps against ethanol C-C bond cleavage [3] and the ensemble effect of the surface Pt and Sn atoms against CO poisoning of the Pt atoms [2].

Acknowledgement

This work was partly supported by a Grant-in-Aid for scientific research (A) from the Japan society for the promotion of science (T. W.).

References

[1] X.H. Xia, H.D. Liess and T. Iwasita, J. Electroanal. Chem., 437, 233 (1997).

[2] A.A. El-Shafei and M. Eiswirth, Surf. Sci., 604, 862 (2010).

[3] S.C.S. Lai and M.T.M Koper, J. Phys. Chem. Lett., 1, 1122 (2010).

Figure 1

2685

, , and

Introduction

Recently, biodiesel fuel (BDF) attracts attention due to the progress of global warming. Glycerol as a byproduct is produced with BDF, and practically used for pharmaceuticals, cosmetics, food additives and so on. With the mass production of BDF, new applications of the glycerol byproduct need to be found. The application of glycerol as anode for direct alcohol fuel cells (DAFCs) can be an important solution because it has low toxicity and high specific energy (5.0 kWh kg-1). However, glycerol has two C-C bonds and three OH groups, so the mechanism for glycerol oxidation reaction (GOR) is complex,1 and the rate for GOR is so slow that the complete oxidation to CO2 is very difficult.

Palladium is known to be one of active materials for GOR. We have found the alloy nanoparticles of Pd with Ag or Au improved the GOR activity and tolerance to the poisoning species on Pd nanoparticle in alkaline medium due to the electronic and bi-functional effects.2 Meanwhile, the Pt electrode had higher GOR activity than the Pd electrode in terms of the onset potential of GOR current, and Ag would improve the GOR activity of Pt. In this study, the effect of the modification with Ag on the GOR activity of a Pt substrate was evaluated by cyclic voltammetry (CV). Moreover, the mechanism for GOR on the Ag-modified Pt (Ag/Pt) electrode was discussed based on in situ infrared reflection absorption spectra at different potentials.

 

Experimental

The Ag/Pt electrode was prepared by underpotential deposition of Cu (Cu-upd) on a Pt substrate and the following galvanic replacement with Ag. The coverage of Ag (θAg) on the Pt substrate were calculated from the charge for H desorption before and after the Ag modification. The Ag/Pt electrodes with θAg = 0.8 and 0.5 were used in this study, which are denoted as Ag(0.8)/Pt and Ag(0.5)/Pt, respectively. The electrolytes were 1 M KOH or (1 M KOH + 0.5 M glycerol) solutions. Pt plate (1 cm × 1 cm) and a Hg/HgO electrode were used as counter and reference electrodes, respectively. In situ infrared reflection-absorption spectroscopy (IRAS) was used to qualitatively analyze the products of GOR on the Pt, Ag(0.8)/Pt and Ag(0.5)/Pt electrodes in alkaline solution. All electrochemical measurements were performed at room temperature.

 

Results and Discussion

The GOR activity for the Pt, Ag(0.8)/Pt and Ag(0.5)/Pt electrodes in alkaline medium was evaluated by CV (Fig. 1). Both Ag(0.8)/Pt and Ag(0.5)/Pt electrodes had twice higher current density and about 150 mV lower onset potential than the Pt substrate. This means that the enhancement of the GOR activity is caused by the modification of Ag atomic layers. Moreover, as θAg increased, a GOR wave was separated into two.

The durability for GOR was evaluated by potentiostatic electrolysis for 60 min at -0.1 V and -0.3 V (vs. Hg/HgO). At -0.1 V, the durability for GOR was decreased in the order of Ag(0.5)/Pt > Pt > Ag(0.8)/Pt. On the other hand, at -0.3 V, the Ag(0.5)/Pt electrode still showed the best tolerance to the poisoning species, but the durability for GOR of Ag(0.5)/Pt was poorer than that of Pt. These results indicated that the modification of Ag on Pt improved the GOR activity and durability.

IRAS gives us the information on the GOR mechanism. In the case of the Pt electrode, the mainly absorption band are observed at 1310 cm-1, 1335 cm-1, 1350 cm-1, 1385 cm-1 and 1575 cm-1 which can be attributed to glyceraldehyde or glycerate, dihydroxyacetone, hydroxypyruvate, symmetric and symmetric O-C-O stretching of carboxylate, respectively. At higher potentials, the absorption band at 1310 cm-1 was obviously observed. On the other hand, in the case of both Ag(0.8)/Pt and Ag(0.5)/Pt electrodes, the absorption band at 1335 cm-1 due to dihydroxyacetone was more obviously observed at -0.4 to -0.2 V, whereas at -0.2 to 0.1 V, the absorption band at 1310 and 1350 cm-1 was strengthen. This results indicate that the oxidation of primary and secondary OH groups depends on potential, and the secondary OH (-0.4 to -0.2 V) and the primary OH ( -0.1 V) are preferentially oxidized on the Ag modified Pt electrodes.

 

Acknowledgement

This work was partially supported by JSPS KAKENHI Grant Number 15H04162.

 

References

1) M. Simoes et al., Appl. Catal. B: Environ., 93, 354 (2010).

2) B. T. X. Lam et al., J. Power Sources, 297, 149 (2015).

Figure 1

2686

, , and

Anion exchange membrane fuel cells (AEMFCs) have attracted a considerable attention because of the potential use of non-precious metal catalysts and the lower overpotentials for the oxygen reduction reaction. However, a low hydroxide ion conductivity of the anion exchange membranes (AEMs) is a drawback, which eventually lowers the power density of AEMFCs. To achieve higher performances of AEMFCs, understanding the nano-scale phase separations inside the membranes is essential. The ion conduction at the catalyst layer/membrane interfaces also needs to be understood. Small angle X-ray scattering (SAXS) is known as a technique to reveal the structures inside the bulk membranes on the nanometer scale, whereas current-sensing atomic force microscopy (CS-AFM) is capable of observing the morphology and the ion conductive region simultaneously on the membrane surfaces on the nanometer scale. At present, most of the structural studies have been carried out on proton exchange membranes, and few results were reported on AEMs.1,2) In this study, the phase separations and the distributions of anion conductive areas of AEMs with different but similar chemical structures were systematically investigated by SAXS for bulks and by CS-AFM for surfaces.

Fig. 1 shows the chemical structures of two membranes, QPE-bl-11a(C1) and QPE-bl-11b(C3), synthesized in our laboratory.3) The membranes have the same hydrophobic structure but slightly different hydrophilic structures. SAXS analyses were carried out in an environmental chamber. The X-ray wavelength was 0.154 nm. The range of the scatter vector, q, was from 0.10 to 2.0 nm-1 (ca. 3.1 – 63 nm). The temperature was set at 40 ºC, while the humidity was controlled from 30% to 90% RH under the nitrogen atmosphere. The membrane samples were equilibrated at least for 2 h at each humidity. For the CS-AFM measurements, the membrane was pressed on a gas-diffusion electrode (GDE) with a catalyst layer composed of Pt/C and a commercial binder (AS-5, Tokuyama Corp.). The membrane on the GDE was installed in a homemade environment-control chamber.2,4) The CS-AFM measurements were carried out in an ultrapure air at 40 ºC and 70% RH using a Pt-coated cantilever at the contact mode with the contact force of 10 nN. The sample voltage between the AFM tip and the GDE was set at -2.0 V.

In the SAXS profile of QPE-bl-11a(C1), a well-defined but broad peak was observed at d = ca. 20 nm. The peak intensity increased with increasing humidity, which might indicate the development of a periodic structure with increasing humidity. For QPE-bl-11b(C3), a well-defined peak was observed at d = ca. 14 nm, smaller than that for QPE-bl-11a(C1). Interestingly, a small difference in the length of the pendant aliphatic groups made a considerable difference in the phase separation. Fig. 2 shows the topographies and the current images obtained on the two AEMs. The surface of QPE-bl-11a(C1) was very flat with the maximum height difference of only 10 nm in the scanned area of 1 μm x 1 μm, but small, nanometer corrugations were also observed over the surface. The surface of QPE-bl-11b(C3) was flatter than that on QPE-bl-11a(C1). Fig. 2c shows the current image on QPE-bl-11a(C1). Anion conductive spots were observed over the surface, and the non-conductive regions (current < 0.5 pA of the background) were only 1% of the entire surface area. On the other hand, it should be also noted that the current differed from 0 to 30 pA at different locations on the surface. In the current image on QPE-bl-11b(C3) (Fig. 2d), the anion conductive regions were also observed over the surface. Interestingly, each conductive spot was not evident, showing a larger homogeneousness. The current difference was smaller, 2 – 20 pA, than that on QPE-bl-11a(C1). For QPE-bl-11a(C1) and QPE-bl-11b(C3), the pseudo current density,4) or the averaged current measured in a unit area, was 2.29 and 2.65 pA nm-2, respectively, whereas the OH- conductivities at 40 ºC in pure water were 23.6 and 34.7 mS cm-1, respectively. The difference in the pseudo current density was 14%, whereas that in the OH- conductivity was 32%. The homogeneousness of the surface conductivity, the pseudo current density, and the OH- conductivity of the membrane bulk must be together considered for the cell performance, 2) which will be further discussed using the membranes.

This work was supported by a CREST program of Japanese Science and Technology Agency (JST).

1) Q. He and X. Ren, J. Power Sources, 220, 373, (2012).

2) M. Hara et al., ACS Appl. Mater. Interfaces, submitted.

3) R. Akiyama et al., Macromolecules, submitted

4) M. Hara et al., J. Phys. Chem. B, 117, 3892 (2013).

Figure 1

2687

, , and

Direct ethanol fuel cells raise the possibility for clean and high efficient energy utilization. Noble Pt-group metals are still essential for electrocatalysts related to such fuel cells. Yet, the high cost and limited reserve of Pt-group metals (especially Pt) as well as the sluggish oxidation kinetics of ethanol molecule call for facile synthesis of low-Pt-content efficient catalysts. Both experimental and simulated investigations suggested alcohol oxidation activity could be improved by coating a monolayer (ML) of Pt on Au [1]. To address the major challenge of scalable synthesis of practical carbon-supported catalysts, here we report a successive and convenient synthesis of Au@Pt-qML/C (qML: quasi-monolayer) using CO as the reducing and capping agents for Pt-qML deposition on Au/C [2]. The Au core with an average diameter of ca. 5 nm was first obtained by reducing HAuCl4 with NaBH4, and Pt-qML was later coated on Au surfaces in one pot through introducing CO and controlled amount of Pt(II). The Au core- Pt shell structure was examined through XRD, EDX, IR and electrochemical measurements. A remarkable enhancement towards ethanol oxidation reaction was obtained on the as-formed Au@Pt-ML/C as compared to that on the state-of-the art commercial Pt/C.

AcknowledgementsFinancial supports from NSFC (grant No. 21273046 and 21473039)and the 973 Program (No. 2015CB932303) of MOST are highly appreciated

References:

[1] M. Li, P. Liu, R.R. Adzic, J. Phys. Chem. Lett., 3, 3480 (2012).

[2] H. Wang, K. Jiang, Q.L. Chen, Z.X. Xie, W.B. Cai. Chem. Commun., 52, 374 (2016).

Figure 1

2688

Active layers of the electrocatalysts for oxidation of methanol and ethanol were prepared by ion beam assisted deposition (IBAD) of platinum and activating rare earths metals (Ce, Yb) onto carbon (AVCarb® Carbon Fiber Paper Р50 and Toray Carbon Fiber Paper TGP-H-060 Т) supports. The deposition method is characterized by the use of deposited-metal ions as assisting ions. Metal deposition and mixing between the precipitable layer and surface of the substrate by accelerated ions of the same metal were carried out on the experimental unit from a neutral vapor fraction and the vacuum-arc discharge plasma of a pulsed electric arc ion source, respectively. Ion accelerating voltage is 10 kV; vacuum – 102 Pa.

Investigation of the composition and microstructure of layers was carried out by RBS, SEM, EPMA and XRF methods. It has been established that the obtained catalytic layers contain atoms of the deposited metals and substrate material, as well as impurity oxygen atoms; their thickness reaches ~30–100 nm. Content of platinum atoms in the layers is ~2×1016 cm–2, concentration of deposited metals atoms in the maximum of distribution equals about a few at.%.

According to investigations with use of cyclic voltammetry the electrocatalysts with prepared layers exhibited high catalytic activity in the reactions of electrochemical oxidation of methanol and ethanol, which form the basis for the principle of operation of low temperature fuel cells (DMFC and DEFC) (Figure 1).

Figure 1. Cyclic voltammograms in the ethanol comprising solution of electrocatalyst prepared using ion beam deposition of ytterbium and platinum onto Toray Carbon Fiber Paper TGP-H-060 Т

In comparison with the traditional multistage chemical methods of preparation of the deposited catalysts, the proposed IBAD method appears to be promising and often more preferable. It allows of the introduction of micro amounts of a doping impurity in the near-surface of a substrate under non equilibrium conditions and of the formation of cohesive catalytic layers at ultra-low platinum consumption.

Figure 1

2689

and

The structure of novel polymers for anion exchange membranes (AEMs) is investigated using small and wide angle X-ray scattering (SAXS & WAXS,) combined with molecular dynamics (MD) simulations using the CHARMM36 force field. The polymers studied are poly(benzimidazole) (PBI) derivatives including mesitylene-poly(benzimidazole) (mes-PBI), mesitylene poly(dimethylbenzimidazole) (mes-PDMBI), and hexamethyl-p-terphenylene poly(methylbenzimidazole). WAXS reveals an amorphous structure with two main length scales; SAXS shows no microphase separation. By comparing simulation results to scattering data, we attribute features observed in the scattering data to stacking between polymer chains, `kinked' monomer lengths, and chain-ion spacings. Overall, we are able to validate the interpretation of scattering data by combining MD simulations and scattering experiments and propose an interesting ion conduction mechanism.

2690

, and

Hydrogen is the most popular fuel source for fuel cell applications due to fast electrocatalytic reaction and very high energy to weight ratio. But its storage and distribution are still major issues. To solve these problems, it is necessary to use alternative other materials as hydrogen source.

Ammonia has attracted attention as one of the candidate hydrogen carrier materials1-3). Ammonia has following features.

-CO2-free energy carrier

-Simple storage as liquid

-High storage density at low pressure

-Essentially non-flammable, non-explosive

-Easy and efficient to crack to hydrogen

-Well-established transport and storage infrastructure already in place

Generally, a few ppm ammonia remains in ammonia cracking to hydrogen3). In case of this ammonia cracked hydrogen for proton exchange membrane fuel cells (PEMFC), significant performance loss will happen because of the decreased conductivity caused by the conducting ions' convert to NH4+ from H+ 4). On the other hand, it has been expected that this fuel can be used for anion exchange membrane fuel cells (AEMFC), because its conducting species is OH- and will not be substituted by NH4+. However, there is little study dealing the effect of ammonia on AEMFC until today.

 In this study, we operated AEMFC with hydrogen with different amount of NH3. As the results, we could confirm that their performance would not be affected by the existence of NH3 in hydrogen up to around 20%. We succeeded in demonstrating that AEMFC has the good adaptability of NH3-H2fuel compared to PEMFC.

This work was supported by Council for Science, Technology and Innovation (CSTI), Cross-ministerial Strategic Innovation Promotion Program (SIP), "energy carrier" (Funding agency: JST).

References

1. K. Kordesch, V. Hacker, J. Gsellmann, M. Cifrain, G. Faleschini, P. Enzinger, R. Franlhauser, M. Ortner, M. Muhr, R. Aronson, J. Power Sources, 86, 162 (2000)

2. W. Wang, J. M. Herreros, A. Tsolakis, A. P.E. York, Int. J. Hydrogen Enegy, 38, 9907 (2013)

3. A. Klerke, C.H. Christensen, J.K. Nrslov, T. Vegge, J. Mater. Chem. 18, 2304 (2008)

4. F.A. Uribe, S. Gottesfeld, T.A. Zawodzinski, J. Electrochem. Soc. 149, A293 (2002)

2691

, , , , , and

Abstract

The kinetics of the oxygen reduction reaction (ORR) greatly influences the performance and the costs of electrodes in fuel cells. Commercial platinum based electrocatalysts exhibit the highest performance, but also increase the cost of the fuel cells. On the other hand, carbon based platinum catalyst support is subject to oxidation, which causes agglomeration of the Pt nanoparticles and decrease in performance of the fuel cells. In order to overcome these issues, we have developed and prepared novel catalyst support based on cobalt and molybdenum carbides. Synthesis of CoMoC was done by chemical method, followed by the high temperature treatment. Since the high temperature preparation of the carbides usually produces low surface area materials, we have developed and prepared a specific carbon based matrix, where Co and Mo based precursors were added. High temperature treatment was done at several temperatures, from 750°C to 1200°C. Structural and morphological characterizations were done using XDR, SEM/EDX and BET analysis, and the obtained results show the formation of the mesoporous non-stoichiometric CoMoC, having the highest value of surface area of 97 m2g. It was shown that the increase in the preparation temperature leads to increase in the mixed carbide stoichiometry, which was ascribed to the high temperatures needed to transform molybdenum oxide into stable carbide. This catalyst support was used to deposit platinum nanoparticles via boron hydride reduction method to obtain several electrocatalyst having 1-10% Pt on the catalyst support. The electrochemical measurements with the prepared CoMoC and different Pt/CoMoC were done using cyclic voltammetry and linear sweep voltammetry on the RDE in alkaline solution. The stability of the prepared catalyst support and catalysts was evaluated using accelerated test procedure by cycling the potential between 0.65 and 1.0 V in oxygen rich alkaline solution. The ORR kinetics parameters were determined and we found comparable performance of the 10% Pt/CoMoC with the commercial Pt/C (40%), which was ascribed to the non-stoichiometric nature of the catalyst support.

2692

, , , , , and

The reduction of CO2 emissions from cars has recently become an important issue, and the development of fuel cell vehicle (FCV) is now in progress. Two problems to be solved for its wide commercial use of fuel cells are the efficient use of platinum (Pt) and ruthenium (Ru) catalysts and the improvement of power density. Carbon black, nanometer-size carbon particles, is commercially used as a catalyst support in fuel cells owing to its high surface area, porosity, electric conductivity, low density, and low cost. Carbon nanomaterials have unique characteristics. In the previous work, we supported PtRu catalysts on various carbon nanomaterials with different geometry and evaluated the catalytic activity of the supported catalysts for direct methanol fuel cell (DMFC) [1]. In this study, we used carbon nanoballoon (CNB) as a catalyst support and measured the catalytic activity of CNB-supported PtRu catalysts.

Arc black (AcB) was produced by an arc discharge of graphite in N2 atmosphere as the precursor of CNB. The twin-torch arc discharge apparatus was used for AcB synthesis. AcB is mainly composed of cocoon-shaped carbon nanoparticles with a lot of amorphous ingredients. CNB was prepared by a heat treatment of AcB in Ar gas at 2600°C for 2 h [2]. AcB and CNB comprised spherical particles of 50 nm in diameter. The particle shape of CNB is hollow. CNB is graphitic and is expected to have high conductivity. Carbon nanocoil (CNC) was synthesized using an automatic chemical vapor deposition system with consecutive substrate transfer mechanism [1]. The fiber diameters of the CNCs was ~300 nm, and the coil diameters of the CNCs was ~1000 nm. We prepared PtRu catalysts for the DMFC anode. The PtRu catalysts were loaded onto CNB by the reduction method using sodium boron hydrate (NaBH4), and counterparts employing the commercial Vulcan-supported PtRu catalyst and CNC-supported PtRu catalyst were also prepared. Hydrogen hexachloroplatinate (IV) hexahydrate (H2PtCl6·6H2O) and ruthenium trichloride (RuCl3) were used as the Pt and Ru precursors, respectively. The molar ratio of Pt and Ru was set at 1:1. Each of the carbon nanomaterials (200 mg) was dispersed in 500 mL of deionized water by sonication for 20 min. H2PtCl6·6H2O and RuCl3 were stirred in 50 mL of deionized water at 60 rpm for 10 min. The solutions were mixed and stirred at 600 rpm for 10 min. Next, a 30-fold molar excess of NaBH4 with respect to the metal precursors was added to 400 mL deionized water. This NaBH4 solution was added to the metal precursor and carbon nanomaterial mixture and stirred. The solution was then filtered, washed and dried to obtain the supported catalyst. The carbon nanomaterials were characterized using scanning electron microscopy (SEM), a laser Raman spectroscopy, Brunauer–Emmet–Teller (BET) measurements, and compressive resistivity measurements. The morphological characteristics and crystalline structure of the prepared catalysts were analyzed using transmission electron microscopy (TEM) and X-ray diffraction (XRD). The amounts of catalyst loaded on the carbon nanomaterials were analyzed by thermo-gravimetric analysis (TGA).

We measured the electrochemical property of the PtRu catalysts supported on the carbon nanomaterials with different shapes, size, and electrical properties. The same amounts of catalyst were loaded on each carbon nanomaterial. The catalyst amount was measured to be 30 wt.% by TGA. Their catalytic activities were measured by cyclic voltammetry (CV) in H2SO4 and CH3OH/H2SO4 electrolytes. The highest methanol oxidation reaction (MOR) current densities were observed for the PtRu catalysts supported on CNB. The catalyst activity of the CNB-supported catalysts was higher than that of the Vulcan-supported catalysts. This is mainly due to the higher electrical conductivity benefiting from the structure of CNB.

References: [1] Yoshiyuki Suda, Yoshiaki Shimizu, Masahiro Ozaki, Hideto Tanoue, Hirofumi Takikawa, Hitoshi Ue, Kazuki Shimizu, Yoshito Umeda, "Electrochemical properties of fuel cell catalysts loaded on carbon nanomaterials with different geometries", Materials Today Communications., Vol. 3, pp. 96-103 (2015); [2] Takashi Ikeda, Shota Kaida, Toshiyuki Satou, Yoshiyuki Suda, Hirofumi Takikawa, Hideto Tanoue, Shinichiro Oke, Hitoshi Ue, Takashi Okawa, Nobuhiro Aoyagi, Kazuki Shimizu, "Preparation of arc black and carbon nanoballoon by arc discharge and their application to a fuel cell", Japanese Journal of Applied Physics, Vol. 50, 01AF13 (2011).

2693

, , and

Hydrazine hydrate, which has high energy density, is one of desirable fuels for direct fuel cells. However, toxicity and volatility of this compound might prevent the wide spread of direct hydrazine hydrate fuel cells. In this context, hydrazine derivatives such as N, N-diaminourea (DAU) and methyl carbazate would be a possible remedy for this problem. These hydrazine derivatives are solid, and evaporation of them would be suppressed. Unfortunately, conventional metal catalysts such as Ni, Pt, and Co hardly oxidize these hydrazine derivatives.

In this work, we report a metallocomplex-based electrocatalyst for the oxidation of hydrazine derivatives. These catalysts completely differ from conventional metal-based electrocatalysts in that single atoms form active sites, and hence new oxidation activity would be expected.

We examined metalloporphyrin-based electrocatalysts for the oxidation of DAU. These complexes can oxidize DAU to give considerable current, while conventional Pt catalysts cannot oxidize DAU. The catalytic activities depend on the kinds of central metals; Co-, Rh-, Fe-, Ru-porphyrins give high current. The catalytic activity increased with the change in the ligand from porphyrins to phthalocyanines. Especially, Fe phthalocyanine gave high catalytic activity for the oxidation of DAU; the current exceeded 150 mA/cm2 at 0.6 V vs. a reversible hydrogen electrode (RHE).

There are two possible pathways of the electro-oxidation of DAU by Fe-PC: (1) DAU is directly oxidized by the electrode (direct pathway) and (2) DAU undergoes hydrolysis to produce hydrazine, and the generated hydrazine is oxidized by the electrode (indirect pathway). HPLC analysis and the dependence of the activity on the concentrations of DAU and hydrazine revealed that the oxidation of DAU proceeds mainly via a direct oxidation pathway.

Finally, we made a membrane electrode assembly (MEA) using Fe-PC as an anode catalyst, and examined the performance of the MEA (cathode catalyst is Pt/C and anion-exchange membrane is A201 (Tokuyama)). The cell gave a considerable power; open circuit potential is close to 0.6 V, and the short-circuit current is 180 mA/cm2 at 80 °C. To the best of our knowledge, this is the first report of a polymer electrolyte fuel cell that uses Fe-PC as an anode catalyst, while the electro-oxidation activity of Fe-PC has been studied for several decades.

The authors are grateful to Tokuyama Corporation for donating an ionomer (AS-4) and A201 membrane.

2694

, , , , and

As an effort to enhance oxygen reduction reaction (ORR) activity of transition metal –nitrogen–graphene (M-N-C) catalysts, we propose a facile and effective strategy through forming edge site and doping S in M-N-C by facile ball milling method. The ORR performance of edge activated S doped Fe-N-graphene (EA-SFeNG) was dramatically enhanced than that of Fe-N-graphene by doping S and forming edge site. Its onset potential and half wave potential were positively shifted from 0.91, 0.77 VRHE to 1.02, 0.848 VRHE respectively, which was comparable to commercial 20 wt.% Pt/C (Vonset: 1.05V, V1/2: 0.865V) and it also exhibit better stability than Pt/C in alkaline media. The reasons for the high ORR activity can be attributed to the increase in defect density and the formation of SOx from the XPS and Raman analysis. Furthermore, we also experimentally confirm the correlation between ORR activity and the promoting factors (doping S, forming edge site) through measurement of work function by kelvin probe force microscopy. The KPFM results showed that doping S and edge activation treatment in FeNG reduce its work function from 4.10 eV to 4.01 eV, which facilitated ORR kinetics of EA-SFeNG.

2695

, and

Anion exchange membrane fuel cells (AEMFCs) are one of the promising energy conversion devices because of their high efficiency and possible use of non-precious metal catalysts. However, the existing anion exchange membranes, main component of AEMFCs, do not show sufficient ion conductivity and stability. In this study, we report a novel series of AEMs composed of perfluoroalyl chains and ammonium-functionalized fluorenyl groups (Fig. 1). The perfluoroalkyl groups are expected to provide chemical stability as well as flexible membrane forming capability, while the ammonium-functionalized fluorenyl groups contain high ionic density for high ion conductivity. We have investigated in details synthetic procedure, structure, and properties of the title AEMs.

The hydrophobic monomer containing perfluroaklyl groups was synthesized by Cu-catalyzed Ullmann coupling reaction, and fluorenyl-containing monomers were synthesized by Sandmeyer reaction. Precursor polymers were synthesized by Ni-mediated Ullman coupling reaction. Then, chloromethyl groups were introduced onto the precursor polymers by Friedel-Crafts reaction. The chloromethyl groups were quarternized with trimethyl amine by Menshutkin reaction.

It is known that the chloromethylation reaction often accompanies with unfavorable side reactions such as cross-linking. By carefully optimizing the Friedel-Crafts reaction conditions, we have successfully controlled the degree of chloromethylation. In fact, the number and the position of chloromethyl groups substituted on fluorenyl groups were controllable.

The chloromethylated precursors were soluble in organic solvent and the solutions were cast to thin films. The ion exchange capacities obtained by titration were lower than calculate from DC. The difference of IEC value was increased with increase of DC. It is suggest that when ionic group density becomes higher, there is an ammonium group which is not functioning as ionic groups by their steric hindrance. The quaternized polymer membranes exhibited high hydroxide ion conductivity and the maximum conductivity was 29 mS/cm in water at 80 °C for the membrane with IEC = 1.6 meq/g. It was found that meta-linked polymers in the hydrophilic component showed slightly higher ion conductivity than that of para-linked polymers. Other properties such as morphologies and stabilities will also be reported.

Acknowledgement

This work was partly supported by Japanese Science and Technology Agency (JST), CREST.

 

 

References

1) Manai Shimada, Shigefumi Shimada, Junpei Miyake, Makoto Uchida, Kenji Miyatake, J. Polym. Sci., A: Polym. Chem.,2015, 7,935-944.

2) Hideaki Ono, Junpei Miyake, Shigehumi Shimada, Makoto Uchida and Kenji Miyatake, J. Mater. Chem. A, 2015, 3, 21779–21788.

Figure 1

2696

, , , , , and

The development of electrocatalysts with low or zero noble metal content is a key factor in the large-scale commercialization of fuel cells. Platinum and platinum alloys are the most efficient catalysts for the oxygen reduction reaction (ORR), which occurs in the cathode of the fuel cell. Nanocarbon materials doped with nitrogen (ex. carbon nanotubes and graphene) have been recognized as a promising alternative for the development of platinum-free electrocatalysts for the ORR. However, the synthesis of these carbon materials often involves the use of harmful organic reagents, synthesis´s extreme conditions, high costs and often are restricted to small scale production. Recently, the use of biomass as source of carbon and nitrogen has been explored to obtain metal-free electrocatalysts. The results reported in the literature indicate that biomass is a very promising alternative for obtaining of electrocatalysts of low-cost and highly efficient in the ORR.

This work describes a two-step method for producing metal-free electrocatalysts from biomass. Two sources of biomass with high nitrogen content in their molecular structure were selected: sargassum and leather. First, each raw material was subjected to pyrolysis treatment at 700 C for 90 minutes under nitrogen, then an activation treatment was given with KOH in nitrogen at 750 C for 90 minutes. The materials were labeled as PAS and PAL for sargassum and leather respectively.

The structural properties of the materials were analyzed by Raman spectroscopy and X-ray diffraction. To determine the morphology scanning electron microscope was used. The electrochemical performance was evaluated by rotating disk electrode technique (RDE).

Raman spectra showed the D and G bands characteristic of carbon materials. This results indicate that after pyrolysis and activation treatments carbonaceous materials with high disorder in their molecular structure were obtained. BET analysis confirmed high surface areas (up to 2000 m2 g-1) and SEM showed the formation of larger porosity in the surface of the electrocatalysts from biomass.

Finally, electrochemical characterization showed high activity for ORR in alkaline media for both electrocatalysts (PAL and PAS). The sample PAL has major current density than platinum electrocatalyst and very close on-set potential (shift 1145 mV). The results obtaining in this work indicate that electrocatalyst from biomass as sargassum and leather are a promising alternative for cathodic reaction in fuel cells.

2697

, and

Recently, Alkaline fuel cells (AFCs) have regained attention because, compared to the acid environment, an alkaline media provides a less corrosive environment to the catalysts and electrodes. One of the main advantages of AFCs is the possibility to replace Pt-based electrocatalysts with non-Pt electrocatalysts.[1] In recent studies, catalytic activity was improved by the influence of interaction between metal and metal oxide. Also researcher have studied to find the ORR mechanisms.[2,3] In this study, the influence of interaction between Ag and perovskite oxide, also electrocatalytic performances in alkaline medium were investigated.

I01-A Poster Session - Oct 5 2016 6:00PM

2698

, , , , , and

The performance of membrane electrode assembly (MEA) greatly depends on the structure of catalyst layers. Carbon particles with Pt nanoparticles sitting on its surface and Nafion ionomer network which bind the catalyst Pt/C particles together and provide the proton conduction path (proton transport to reaction sites) are two major components in a catalyst layer (1-3). A good catalyst/ionomer interface is important because such interface provides 3-phase reaction zone for oxygen reduction reaction (ORR). The ionomer coverage over carbon particles directly affects the catalyst utilization, consequently, the mass activity and the thickness of the ionomer layer over a carbon particle determines the O2 diffusion barrier. Hence, achieving the balanced catalyst/ionomer interface with appropriate H+conductivity and thin layer (low diffusion barrier) is critical step for a high performance MEA. The fabrication of MEA starts from the catalyst ink in which catalyst powder and Nafion ionomer are mixed in a solvent. It will be desired if the catalyst/ionomer interface can be formed in a catalyst ink. To study the formation of such an interface, the dispersion of catalyst powder and ionomer particles in an ink system needs to be studied to see if the particle size of carbon changes after Nafion ionomer addition—increased carbon particle size generally indicates the formation of the interface.

Ultra-small angle X-ray scattering (USAXS) characterization had been proved to be a great method to investigate the particle size of aggregate system (4). We have developed a unique method (5) to characterize the catalyst dispersion in inks using the combined USAXS and Cryo-TEM by which USAXS can provide the geometry and size distribution of different particles in a ink while cryo-TEM can confirm the USAXS results with direct observation of geometry and size distribution of these particles (Cryo-TEM is usually used to characterize biology samples by fast freezing the liquid samples to form a very thin ice to lock all particles within the liquid to keep their original structure and size of these particle without interruption). In this work, we studied the interaction between Nafion ionomer and two carbon blacks which are NH2 functionalized carbon black (XC72-NH2) and SO3H functionalized carbon black (XC72-SO3H). By comparing the particle size change of carbon black particles before and after addition of Nafion ionomer in two ink systems, we can study the interaction between Nafion and carbon black aggregates.

The USAXS results and corresponding Cryo-TEM images of four ink systems are shown in figure 1 and 2. The particle size from USAXS fitting is shown in Table 1. It can be seen in figure 1 and table 1 that the size of the aggregates of carbon black in NH2-CB ink system which is in the 2nd level of USAXS fitting increases a lot (87.1%) after adding Nafion ionomer. However, according to the observation from figure 2 and table 1, the size of the carbon black aggregates which is in the 2nd level of the USAXS fitting only decreased a little (13.5%). Nafion ionomer particles with negative charges (-SO3-) on the surface attract the positively charged (-NH3+) carbon particles via electrostatic forces to form larger aggregates. Opposite results can be found in XC72-SO3H ink system because XC72-SO3H particles with negative charges (-SO3-) was repelled by the negatively charged (-SO3-) Nafion ionomer particles.

From the results above, it is clearly that the charged groups on carbon can effectively affect the formation of catalyst/ionomer interface which can be used as an effective tool to build the interface. Understanding the interaction between Nafion ionomer and catalyst support will lead to the rational design a high performance MEA.

Reference

1. Z.-F. Li, L. Xin, F. Yang, Y. Liu, Y. Liu, H. Zhang, L. Stanciu and J. Xie, Nano Energy, 16, 281 (2015).

2. M. S. Wilson and S. Gottesfeld, Journal of Applied Electrochemistry, 22, 1 (1992).

3. L. Xin, F. Yang, S. Rasouli, Y. Qiu, Z.-F. Li, A. Uzunoglu, C.-J. Sun, Y. Liu, P. Ferreira, W. Li, Y. Ren, L. A. Stanciu and J. Xie, ACS Catalysis, 6, 2642 (2016).

4. H. K. Kammler, G. Beaucage, D. J. Kohls, N. Agashe and J. Ilavsky, Journal of Applied Physics, 97, 054309 (2005).

5. F. Xu, H. Zhang, J. Ilavsky, L. Stanciu, D. Ho, M. J. Justice, H. I. Petrache and J. Xie, Langmuir, 26, 19199 (2010).

Figure 1

2699

and

In polymer electrolyte fuel cells (PEFC), cathode catalyst layers may experience high overpotentials during cell start up/shut down and even during normal operation. High overpotentials undesirably facilitate corrosion kinetics of the carbon catalyst support which results in a severe increase in mass transport resistance in the cathode. Focused ion beam and scanning electron microscope tomography (FIBSEM) is an emerging technique for nanometer-resolution direct observation of three-dimensional microstructure and shows promise for improving understanding of microstructural changes to PEFC cathodes in response to corrosion. However, the high porosity of PEFC cathodes leads to challenges in downstream image processing. Filling the pore space prior to tomography may allow circumvention of image processing challenges but has not yet been definitively shown to provide good results. The present paper assesses the viability of an organoplatinum electron beam induced deposition (PtEBID) step in order to fill the pore space prior to FIBSEM. PtEBID methods were investigated and validated with complementary AFM and STEM to quantify precursor diffusion into porous cathodes, kinetics of adsorption and cracking of organoplatinum on the carbon support surface, and damage induced by the focused ion beam used during FIBSEM.

2700

, and

As one of the renewable energy, polymer electrolyte membrane fuel cell (PEMFC) is a promising candidate because of its good performance efficiency and environment-friendly technique. However, high cost and poor durability of Pt catalys are main obstacles to a commercialization of PEMFC. As a solution for the problems, some researchers have focused on thin film catalyst layer instead of Pt/C catalyst for reducing the amount of platinum use and for enhancing performance at high current density. For example, NSTF (nanostructured thin film) catalyst layer and inverse opal structured catalyst layer were studied for improving mass-transfer and enhancing durability and stability of catalyst.[1-2]

In this study, we used ultrasound to form a thin catalyst layer. By using ultrasound, we synthesized Pt nanoparticle and coated it on MPL (microporous layer), simultaneously. It has being revealed that ultrasound is an effective way to create metal nanoparticle because of ultrasound's cavitation effect.[3] Bang et al explained about cavitation effect. Bubble explosion provides huge pressure and energy for a chemical reaction if the bubble size reached the critical point under ultrasound irradiation.[4] In addition, ultrasound has a micro jet effect, which has ability to clean a substrate and coat nanoparticles on it.[5] In our study, ultrasound applied to form a catalyst layer and the we adopted it for fabricating MEA (Membrane Electrode Assembly) of single cell. The MEAs were compared with MEA fabricated by a conventional method with Pt/C catalyst.

To verify a formation of nanoparticle via a cavitation effect of ultrasound, we used UV-visible spectroscopy and observed spectrum change nearby 450 nm which is known for Pt4+ after ultrasound's irradiation.[6] To coat Pt nanoparticles on the carbon paper, we prepared carbon papers fixed at an acrylic frame in platinum precursor solution with reducing agent. Ultrasound was irradiated at various conditions. For example, time and power of irradiation, concentrations of the precursor and the reducing agent were changed to find an optimum condition for a sonochemical coating. As shown in Figure 1, images of the carbon papers acquired by FE-SEM (Field Emission Scanning Electron Microscope) and EDS (Energy Dispersive Spectroscopy) were used for checking platinum distribution. Cathode's catalyst layer of MEAs was prepared by sonochemical method and anode's catalyst layer was formed by spraying Pt/C on carbon paper. The MEAs were fabricated using hot-pressing without ionomer. Figure 2 shows I-V curve of the MEAs made at the 12 W of ultrasound power for 120 min and 90 min. To figure out experiment condition effect, we are still carrying out single tests and perform a characterization of MEA to elucidate the difference between MEAs. All the results investigated in the study will be presented with in-depth discussion.

Figure 1

2701

, and

Air is the most practical and economic oxidant for fuel cell operation, so any impurities in the air will be a significant concern for performance and durability of proton exchange membrane fuel cells (PEMFCs). Exposure of the cathode to airborne contaminants was found to cause serious performance loss and degradation in many cases (1). Seven compounds were chosen for a detailed investigation from the inventory of 260 possible air pollutants suggested by Environmental Protection Agency (2, 3). CH3Br was selected for further studies as a potential air pollutant due to its natural and anthropogenic origin. The application of CH3Br as an agricultural pesticide has been reduced from 2005 since it is an ozone-depleting compound, however, CH3Br has a major natural emission source from oceans (4). The work focuses on comprehensive analysis of localized long-term PEMFC performance exposed to 5 ppm CH3Br for improving adaptability and durability and understanding the poisoning mechanism.

A segmented cell and data acquisition system were used (5) with a commercially available 100 cm2 membrane/electrode assembly (MEA). Each electrode contained a Pt/C catalyst with a loading of 0.4 mgPt cm-2. A segmented SGL 25BC gas diffusion layer (GDL, 10 segments of 7.6 cm2) and a Teflon gasket were employed at the cathode whereas a single GDL piece was applied at the anode. The MEA was operated under galvanostatic control of the whole cell current. Other operating conditions were: 80°C, 48.3 kPagback pressure, 100/50% relative humidity and 2/2 stoichiometry for the anode and cathode respectively. The dry contaminant was injected into the humidified cathode air stream. The poisoning proceeded until the cell voltage reached a steady value. MEAs were analyzed by SEM, TEM, XPS and electrochemical methods.

Fig. 1 a) shows the voltage response and normalized current density for each segment and 1.0 A cm-2. For the first 18 hours, the cell was operated with air resulting in a cell voltage of 0.650 V. The injection of 5 ppm CH3Br decreased the voltage over a long transition period (~50 h) and eventually resulted in a steady stateof 0.335 V. The voltage decrease was accompanied by a redistribution of local current densities. Operation with pure air for 70 h after the poisoning phase did not recover the original cell voltage. XPS analysis of the MEA exposed to CH3Br for 147 h showed the presence of Br- that suggests bromomethane hydrolysis yielding CH3OH, Br- and H+ (6). CH3OH appears to be oxidized at the conditions of the cathode, whereas Pt can adsorb Br-. Chemisorption of Br- results in a decrease of the electrochemical area (ECA), suppression of O2 adsorption and shift of the oxygen reduction from a 4-electron to a 2-electron mechanism with H2O2intermediate formation, which negatively impact PEMFC performance (7).

Cyclic voltammetry (CV) measurements demonstrated that the ECA of both anode and cathode after CH3Br exposure for 147 h decreased by 50%. The cathode ECA loss was only 30% for an MEA aged under the same conditions without CH3Br. The observed ECA drop is explained by Pt particles size growth to 6-8 nm under contamination (Fig. 1 b) and 4-5 nm without CH3Br (Fig. 1 c) compared to the initial particle size of 2-2.5 nm. The absence of a self-recovery indicates that Br- can be strongly adsorbed on Pt which reduces performance. However, polarization curves measured after CV scans revealed a cell recovery and performance loss was only 25-50 mV mainly due to increased activation overpotential. Thus, CV accompanied by operation with fully humidified gases could remove Br-. A detailed discussion of results, CH3Br poisoning mechanism and possible mitigation procedures will be presented.

ACKNOWLEDGMENTS

We gratefully acknowledge ONR (N00014-13-1-0463), DOE EERE (DE-EE0000467) and Hawaiian Electric Company.

REFERENCES

  • O.A. Baturina, Y. Garsany, B.D. Gould, K.E. Swider-Lyons, in: H. Wang, H. Li, X.-Z. Yuan (Eds.), PEM fuel cell failure mode analysis, CRC Press, 2011, p. 199.

  • T.V. Reshetenko, J. St-Pierre, J. Power Sources, 293, 929 (2015).

  • T.V. Reshetenko, J. St-Pierre, J. Power Sources, 287, 401 (2015).

  • S.A. Yvon-Lewis, E.S. Saltzman, S.A. Montzka, Atmos. Chem. Phys., 9, 5963 (2009).

  • T.V. Reshetenko, G. Bender, K. Bethune, R. Rocheleau, Electrochim. Acta,56, 8700 (2011).

  • W. Mabey, T. Mill, J. Phys. Chem. Ref. Data, 7, 383 (1978).

  • N.M. Marković, H.A. Gasteiger, B.N. Grgur, P.N. Ross, J. Electroanal. Chem., 467, 157 (1999).

Figure 1

2702

, and

Abstract:

One of the current barriers listed in the Safety Codes and Standards section of the Fuel Cell Technologies Office Multiyear Research, Development, and Demonstration Plan is: 'insufficient technical data to revise standards' [1]. More specifically, SAE J2719 and ISO-14687-2 are two recently developed standards for gaseous hydrogen fuel specifications whose revisions are being planned [2,3]. Both standards are applicable for fuel cell vehicles and each governs contaminants and their maximum allowable concentration in H2. Of the contaminants listed in these fuel specifications, we focus our efforts on carbon monoxide (CO), whose allowable level is 200 parts per billion (ppb). In our Polymer Electrolyte Fuel Cell (PEFC) parametric study, we utilized 25cm2 Membrane Electrode Assemblies (MEAs) with 2015 DOE target loadings for platinum (0.15mg Pt/cm2). While it is virtually impossible to test the impact of CO using all of the possible testing conditions, we have systematically formed a set of experiments (Table 1) that encompasses a broad spectrum of Fuel Cell (FC) operating parameters. For each experiment, we fixed the cell temperature at 80oC and operated the FC at 1 A/cm2 with H2 and air stoichiometric flows equivalent to 83% and 50% utilization, respectively. The table reflects the parameters we vowed such as: relative humidity (32, 50, and 100% RH), back pressure (ambient, 150kPa, and 275kPa) and CO concentration (0.2, 0.5, 1.0ppm). The CO dosage remained constant throughout the experiments conducted in this study. In this presentation we will provide data to modelers for enhancing their predictive capabilities in order to help the FC community deduce the CO impact at other operating condition.

References:

  • http://energy.gov/sites/prod/files/2015/06/f23/fcto_myrdd_safety_codes.pdf

  • SAE J2719: Hydrogen Fuel Quality for Fuel Cell Vehicles, www.sae.org

  • ISO 14687-2, Hydrogen Fuel – Product Specification, Part 2: PEM fuel cell applications for road vehicles, http://www.iso.org/iso/catalogue_detail.htm?csnumber=55083

 

Acknowledgements:

The authors gratefully acknowledge the financial support of the DOE Fuel Cell Technologies Office and the support of Technology Development Manager, Charles (Will) James, Jr

Figure 1

2703

, , , , , , , , and

In a polymer electrolyte membrane (PEM) fuel cell, the effective removal of product liquid water is necessary for achieving high power output. Liquid water produced at the cathode catalyst layer (CL) accumulates in the pores of the gas diffusion layer (GDL) and impedes oxygen diffusion from gas channels to the catalytic reaction sites. To reduce liquid water build up in the GDL, dual-layer GDLs comprised of a carbon fiber macro-porous substrate and a micro-porous layer (MPL) are typically used. The MPL covers the large surface pores of the substrate yielding a smooth interfacial contact region between the CL and the GDL. This smooth interface results in superior thermal and electrical conductivities as well as reduced water accumulation in the region between the CL and GDL.

The MPL is typically a mixture of carbon black particles and hydrophobic agents. The structural properties of the MPL, such as surface crack size and porosity distribution, are dependent on its composition. An MPL containing 20-40 wt. % of hydrophobic agent was shown to facilitate product water removal at high current density operation1. Recently, it was found that the addition of carbon nanotubes (CNT) led to the strong adhesion between carbon black particles and the reduction of ohmic and mass transport losses 2-4.

In this work, the physical properties of the SGL 25 series GDLs (SGL Group) were characterized by various imaging methods. Three types of commercially available GDLs were studied: SGL 25BC, 25BI and 25BN. SGL 25BC is the standard MPL with 23 wt.% PTFE, while SGL 25BI contains a reduced PTFE content of 10 wt.%. In SGL 25BN, CNTs are added to the standard MPL (i.e. 23 wt.% PTFE with CNT). Scanning electron microscopy (SEM) images of each GDL are shown in Figure 1. High intensity X-rays generated at the BMIT-BM beamline of the Canadian Light Source were utilized to image the fuel cell in operando, and the liquid water distribution in the resulting radiographs was identified with in-house post-processing algorithms. Images were obtained with a pixel resolution of 6.5 µm at a frame rate of 0.33 frames per second. In this work, the performance of these materials will be discussed, in particular, within the context of their varying microstructure.

Reference

1. Qi Z, Kaufman A. Improvement of water management by a microporous sublayer for PEM fuel cells. J Power Sources. 2002;109(1):38-46.

2. Lin S, Chang M. Effect of microporous layer composed of carbon nanotube and acetylene black on polymer electrolyte membrane fuel cell performance. Int J Hydrogen Energy. 2015;40(24):7879-7885.

3. Fan C, Chang M. Improving proton exchange membrane fuel cell performance with carbon nanotubes as the material of cathode microporous layer. Int J Energy Res. 2016;40(2):181-188.

4. Schweiss R, Steeb M, Wilde PM, Schubert T. Enhancement of proton exchange membrane fuel cell performance by doping microporous layers of gas diffusion layers with multiwall carbon nanotubes. J Power Sources. 2012;220:79-83.

Figure 1

2704

, , , and

Introduction

 Elevation of the PEFC operation temperature over 100 oC is expected for the next generation FCVs in order to further increase the efficiency. Even so, only a few reports studying MEA performance and durability over 100 oC are available. In our laboratory, we have been examining MEAs at the high temperature and low humidity condition, and particularly paying attention to the durability of electrocatalysts. During our study, carbon corrosion was found to be promoted at the high temperature and low humidity condition when the accelerating degradation test simulating the start/stop cycles of FCVs was performed. This carbon deterioration was observed at the interface between the cathode layer and the Nafion electrolyte membrane in particular. We believe that this phenomenon is derived from the decreased proton conductivity of Nafion under the low humidity and carbon corrosion homogeneously occurs over the cathode layer if the proton conductivity is high enough through the cathode layer. Therefore, the aim of this study is clarifying the phenomenon of cathode layer deterioration using the electrolyte with higher proton conductivity

Experimental

 Aquivion membrane was used in this study. Here, proton conductivity of Aquivion is higher than that of Nafion, and is assumed to be high enough even over 100 oC. Then, MEAs were made through spray-printing catalyst layers containing 46.2%Pt/KB electrocatalyst (TEC10E50E) and Aquivion ionomer. IV performance was evaluated at 105 oC-RH57 % by supplying 100 cc/min of hydrogen and 100 cc/min of air to the anode and the cathode, respectively. Then, the durability test was carried out by fluctuating the potential between 1.0 and 1.5 V based on the protocol1simulating the start/stop cycles of FCVs. Besides IV characteristics, ECSA and impedance measurements were also performed. Finally, the change in the cathode layer structure was evaluated three dimensionally using the FIB-SEM technique.

Results and discussion

 Durability tests were performed up to 10,000 cycles at 105 oC.-RH57 %. Then, the change in IV characteristics, ECSA, and cathode structures was compared between Aquivion MEA and Nafion MEA. Even after 2000 cycles, with Aquivion MEA, large increase in activation overvoltage and also in ECSA was observed. Based on this observation, the degradation phenomenon is assumed to be different between the two MEAs as shown in Fig. 1. For Nafion MEA, carbon oxidation reactions only occur at the interface between Nafion membrane and the cathode layer because protons cannot easily transfer into the cathode layer owing to decreased proton conductivity of Nafion (Fig. 1a). On the other hand, in Aquivion MEA, protons can more easily move into the cathode layer, and then deterioration of carbon is promoted (Fig. 1b). In order to confirm our assumption, the structure change in the cathode layer was analyzed by the FIB-SEM techique. After the three dimensional reconstruction based on 100 SEM images, relatively large pores produced by carbon corrosion were more often observed near Nafion membrane in Nafion MEA. On the other hand, in Aquivion MEA, the pores were homogeneously distributed. In other words, deflection of cathode deterioration observed in Nafion MEA does not take place even at the high temperature if the proton conductivity is high enough.

(1)A. Ohma et al., ECS Trans, 41 (2011) 755.

Figure 1

2705

, and

In polymer electrolyte fuel cells (PEFCs), the catalyst layer (CL) of the cathode needs a large amount of Pt because of the slow oxygen reduction reaction. Since electron, proton, and oxygen are necessary for the cathode reaction, achieving the optimum structure of electrode CL and the efficient transport of the reactants is significantly effective to reduce the usage of Pt catalyst. In this study, the effects of the cathode CL structure on the IV characteristics were experimentally evaluated using CLs fabricated under several conditions. Furthermore, the oxygen transport resistance in the CL was evaluated. Pore size distribution (PSD) of the CL was measured by nitrogen physisorption method, and the CL structural parameters were estimated using the estimation model developed by the authors (1). Using the model analysis with the estimated parameters, we evaluate various oxygen transport resistances in the CL and discuss the contributions of these resistances to the cell performance.

Figure 1 shows the IV characteristics for different ratios of ionomer to carbon (I/C ratios) in the CL. To investigate the effects of ionomer thickness surrounding the carbon agglomerate, the weights of carbon with Pt were set to be similar. Then the Pt loadings of the CLs were 0.16, 0.20, and 0.20 mg/cm2for the I/C ratios, 0.8, 1.0, and 1.2. The oxygen transport resistances distinguished using limiting current density method (2)(3) are summarized in Table 1. It can be considered that the pressure dependent resistance is the oxygen transport resistance in the GDL and the channel, and the pressure independent resistance is the oxygen transport resistance in the CL. It is confirmed that the pressure dependent resistances are similar in all of the I/C ratios because of using the same GDL and channel, and the pressure independent resistance in the I/C=1.2 is the largest. This result shows that the oxygen transport resistance in the CL depends on the structure of the CL.

Figure 2 shows the PSDs in the CLs measured by nitrogen physisorption method for the different I/C ratios, 0.8, 1.0, and 1.2. The mode pore diameter becomes smaller as I/C ratio becomes higher, and the porosities were 0.62, 0.55, and 0.42 for the I/C ratios, 0.8, 1.0, and 1.2. From these results, the structural parameters of the CL were estimated using the estimation model developed by the authors (1). The estimated parameters are summarized in Table 2. It is confirmed that the ionomer thickness with the lower I/C ratio is thinner than with the higher I/C ratio, and the carbon agglomerate radiuses are around 15nm in all of the cases.

The oxygen transport resistances in the CL were distinguished in further detail using the estimated parameters and the simplified evaluation formula (4), where the effect of the oxygen dissolution resistance was introduced additionally. The oxygen transport resistance in the CL is assumed to consist of three resistances: oxygen dissolution resistance into ionomer, oxygen diffusion resistances in ionomer, and that in pore. We estimated combinations of the oxygen dissolution rate kdiss, the oxygen diffusion coefficient in polymer DpO2, and the effective oxygen diffusion coefficient in pore DCL,effO2 by fitting them to simulate the values of the total oxygen transport resistance in the CL estimated from experiments. Here, we show an example using values, kdiss = 2.35×10-3 m/s and DpO2 = 1.36×10-9 m2/s. The results are summarized in Table 3. It is confirmed that the oxygen dissolution resistance is the largest with I/C=0.8 because of the smallest surface area of carbon agglomerate with polymer (as shown in Table 2). On the other hand, the oxygen diffusion resistance in pore with I/C=0.8 is the smallest. This is because the resistance in pore depends on the porosity of the CL and decreases with the higher porosity.

Figure 3 shows the I-V characteristics calculated by the model analysis using the estimated parameters in Table 2. The result can simulate the tendency of the measured IV characteristics and the limiting current densities for the different I/C ratios well. In this paper, we evaluated the relationships between the oxygen transport resistances and the structure in the CL for only three I/C ratios. We plan to evaluate these relationships using various structures of the CL, and to elucidate the effective structure to achieve small oxygen transport resistance in the CL.

Reference

(1) S. Akabori, et al., ECS Trans., 64(3), 305 (2014).

(2) D.R. Baker, et al., J. Electrochem. Soc., 156, B991 (2009).

(3) J.P. Owejan, et al., J. Electrochem. Soc., 160, F824 (2013)

(4) Y. Tabe, et al., J. Electrochem. Soc., 158, B1246 (2011).

Figure 1

2706

, , and

Much progress in reducing the platinum loading in proton exchange membrane (PEM) fuel cell cathodes to a target of 0.1 mg-Pt/cm2 has been made. However, large performance losses are observed at high current densities at low cathode platinum loadings. Recent studies have shown that the large losses are caused due to resistances at or near the catalyst-ionomer interface.[1] This resistance can be countered through the use of ionomers designed to interact with Pt in a way that does not limit the oxygen reduction reaction. However, the development of tools to quantify and characterize this resistance is needed to enhance our understanding of this resistance and correlate it with performance loss.

In the early 1990's Feliu et al. developed a method for measuring adsorbed ion charges on Pt single crystals in an aqueous electrolyte.[1] We have validated this CO displacement method on polycrystalline Pt and carbon-supported Pt nanoparticles in an aqueous electrolyte and present here results from implementing this methodology in a fuel cell. Figure 1 illustrates the current transient measured during chronoamperometry as CO gas is introduced to the cathode of a fuel cell. The measured displacement charge at potentials less than 0.2 V represent a displacement of adsorbed protons, corresponding to a positive displacement current. At potentials greater than 0.3 V, a negative current is measured corresponding to the displacement of negatively charged adsorbed species. In the case of a fuel cell cathode, these negative species are the sulfonate group present in the ionomer. The adsorption charge can then be integrated to determine the amount of species adsorbed on the surface at each potential. Using this tool, ionomer adsorption for different ionomer types at different operating conditions can be examined.

[1] Kongkanand A, Mathias MF, J. Phys. Chem. Let. 7 (2016) 1127.

[2] Feliu JM, Orts JM, Gomez R, Aldaz A, Clavilier J. J. Electroanal. Chem. 372 (1994) 265.

Figure 1. Chronoamperometry at different potentials during CO Displacement. Introduction of CO to the cathode occurs at t=120s.

Figure 1

2707

, , , , and

1. Introduction

In the durability evaluation of the PEFC (Polymer electrolyte fuel cell), the degradation characteristics are different due to the difference of the potential cycle test method. It is inferred that the influence of the PEFC by the potential cycle test is different every the constructional element. In order to improve the cycle characteristics, it is considered that to clarify the correlation between the degradation acceleration part and the test method is one of the important factors.

In this study, for the MEA (Membrane and electrode assembly) made by the commercial catalyst and electrolyte materials, two potential cycle tests were carried out. The MEAs was analyzed by EPMA (Electron probe micro analyzer), LC/MS (Liquid chromatography–mass spectrometry), et al., and the degradation part was evaluated.

 

2. Experimental

The electrolyte membrane and the ionomer were composed of Nafion®, and the Pt / C was used as the catalyst powder in both the anode electrode and the cathode electrode. The catalyst ink was applied by the spray method, and the CCMs (Catalyst coated membrane) were created. The durability tests in the two conditions were carried out after the conditioning operation. One of the operating conditions was the Start-Stop cycle test. In that condition, the MEA was operated with the sawtooth-shaped potential cycle of between 1.0 V and 1.5 V (2 s/cycle). With the Load cycle test which is another condition, the MEA was operated with the square-shaped potential cycle of between 0.6V and 1.0V (6 s/cycle). About the CCMs that the durability tests were carried out, the elemental distribution of the cross section was analyzed by EPMA. In addition, the information of the degradation products about the ionomer was obtained by LC/MS.

 

3. Result and discussion

Figure 1 shows the SEM images (upper side) and the line profiles (lower side). From the SEM image of Initial, each thickness of the anode layer, the electrolyte membrane and the cathode layer is about 7 mm, 60 mm and 7 mm, respectively. Each thickness of the anode layer and the electrolyte membrane in Start-Stop cycle is also the same value as Initial. However, the thickness of the cathode layer is thinner than Initial. It is inferred that the change of the layer thickness is due to the influence of the durability test. On the other hand, there is no significant different about each region in Load cycle. In the line profile of the elements by EPMA, the profile shape of Load cycle is almost the same as Initial. The concentration of Pt in the cathode layer of Start-Stop cycle is higher than the other samples. In the cathode layer of Load cycle, because the layer thickness became thin, it is inferred that Pt concentrated.

As the result of the other analysis, it was revealed by LC/MS that the degradation product quantity for the ionomer of Load cycle was the largest and the degradation product quantity for the electrolyte of Initial was the smallest.

From those results, it is inferred that the influence to the catalyst layer is large on the Start-Stop cycle test and the influence to the ionomer is large on the Load cycle test.

Figure 1

2708

, , , and

High power density along with low gravimetric and volumetric characteristics for Proton Exchange Membrane Fuel Cell stacks are desired in order to be successfully implemented in the stationary applications market.

Liquid cooling is currently the most widely used cooling strategy used in PEMFC stacks construction, especially for those who incorporate a large active area and high number of cells, since the heat transfer coefficients with liquid flow are much higher than airflow for the same pumping power.

This paper presents the manufacturing steps for an in house production of graphite type PEMFC bipolar plates with internal liquid cooling circuit and investigation results for a short stack construction based on 92cm2 active area.

2709

and

Efficient direct electro-oxidation of liquid fuels, such as methanol and formic acid for low-temperature portable fuel cell applications, remains a challenge limiting commercialization. The major performance detractors are: (1) the formation of strongly adsorbed reaction intermediates, (2) fuel crossover through the membrane depolarizing the cathode, and (3) mass transport for fast fuel delivery into the anode catalyst layer and gaseous product removal.

Direct formic acid fuel cells (DFAFCs) have the advantage over direct methanol fuel cells in that it is possible for formic acid to be electro-oxidized via a direct non-strongly adsorbed reaction intermediate pathway.[1] In addition to being less susceptible to fuel crossover due to electrostatic repulsion of the polarized formic acid dipole by the negatively charged sulfonic groups in the proton exchange membrane (PEM).[2] However, the catalyst selection is pivotal for enhanced performance. Previous research from our group has shown highly active carbon-supported platinum (Pt/C) catalyst layers spontaneously decorated with bismuth exhibiting low overpotential for formic acid electro-oxidation, shown in Fig 1. This enhancement of catalytic ability has been attributed to promotion of the direct oxidation pathway by the 'third-body' as well as an electronic effect to promote CH-down adsorption of formic acid on the unoccupied Pt/C surface.[3] However, adsorbed Bi is unstable under oxidizable potentials, as demonstrated in our previous studies, Figure 2.[4] Pore-former was incorporated into the catalyst layer to increase formic acid mass transport. During open potential acid wash removal of the pore-former the Bi was lost from the Pt surface and performance was also lost.

Research by Antaño-López et al. has shown that the spontaneous adsorption of citrate, a common nanoparticle capping agent, suppresses the formation of platinum oxide.[5] This implies that citrate is a stable surface adsorbate in the oxide region. It is predicted that citrate will exhibit a similar 'third-body' effect as Bi with an optimal surface coverage fraction yet to be determined due to the chelation effect of citrate that Bi does not exhibit.

References:

  • Parsons, R. and T. VanderNoot, 'The oxidation of small organic molecules: A survey of recent fuel cell related research', Journal of Electroanalytical Chemistry and Interfacial Chemistry, 1988, 257, 9.

  • Rhee, Y.-W., S. Y. Ha, R. I. Masel, and A. Wieckowski, 'Crossover of formic acid through Nafion membranes', Journal of Power Sources, 2003, 117, 35.

  • Bauskar, A. S. and C. A. Rice, 'Spontaneously Bi decorated carbon supported Pt nanoparticles for formic acid electro-oxidation', Electrochimica Acta, 2013, 93, 152.

  • Pistono, A.O., C.S. Burke, J.W. Cisco, C. Wilson, B.G. Adams, and C.A. Rice, 'Inhibition of Bismuth Dissolution during Anode Catalyst Layer Pore Former Removal in a Direct Formic Acid Fuel Cell', ECS Electrochemistry Letters, 2014, 3, F65.

  • González-Peña O. I., T. W. Chapman, Y.M.Vong, and R. Antaño-López, 'Study of adsorption of citrate on Pt by CV and EQCM', Electrochimica Acta, 2015, 53, 5549.

Figure 1

2710

, , , and

Platinum (Pt) is the state-of-the-art catalyst for both anodic and cathodic reactions that occur in polymer electrolyte fuel cells (PEFCs). However, the cost of platinum has severely hindered the commercialization and distribution of PEFCs. One of the challenges for the next generation of fuel cells is to significantly reduce the Pt content of both electrodes while maintaining excellent activity and high durability. By anchoring platinum to carbon substrates, the amount of platinum can be substantially reduced without compromising its activity. While this approach may address the cost concerns; it may also induce durability issues depending on the operating conditions. In this work, we have investigated the effects of aging on different carbon supported, low Pt loading catalysts, using accelerated stress tests (AST). These tests consist of series of voltammetric cycles conducted in solution to accelerate the rate of degradation of the carbon support. Structural, chemical, and morphological information will be presented via X-ray techniques (XRD and XRF) along with electron microscopy (SEM,TEM). Electrochemically active surface area (EASA) will be measured via H2 adsorption and desorption along with CO stripping techniques. The electrochemical activity of the catalysts versus the number of cycles will be shown using RRDE as well as some complementary fuel cell data. Preliminary results show a decrease the EASA with time; however, the rate and extent of decay varied depending on the type of carbon support.

2711

, , and

Porosity and pore size of cathode catalyst layer was controlled by changing catalyst layer coating methods. Two-layered catalyst layer of dense and porous layers was also compared in order to get optimum porosity of catalyst layer in the view point of both performance and durability. Performance and durability of MEA's with different porosity were compared and impedance and ECSA of them were analyzed. Porosity and pore size were measured and observed by mercury intrusion method and SEM images. Porous catalyst layer was generally favorable. However porosity of catalyst layer could be further advanced with two-layered structure of different pore structures.

2712

and

The water transport in a gas diffusion layer (GDL) has been investigated using two-dimensional lattice Boltzmann (LB) simulation. The LB model is developed to simulate the dynamic behavior of liquid water and enables to visualize the water-invasion process through micro-pores in GDL. To investigate the effect of rib structure on water invasion process in GDL, two different cases (i.e., with and without rib structure) are compared. The numerical model is verified by the comparison of the flow permeabilities in GDL. The validation result indicates that the LB model can properly predict the permeability of GDL and enables to simulate the water transport behavior in the GDL. The reconstruction of GDL is established by randomly placing the particles in GDL and ignoring the GDL deformation due to clamping force. The results of LB simulation confirm that the liquid water transport inside GDL is strongly governed by capillary force and the rib structure greatly impacts on the water transport behavior. The rib structure influences on the location of water breakthrough by comparing the simulation results of two different cases. This is due to the higher resistance force underneath the rib, resulting in the change of flow path which preferentially selects the lower resistance force. The water saturation level under the channel is higher than that under the rib caused by the suppression of growth of water cluster. After water breakthrough, the liquid water distribution under the channel has little change, whereas that under the rib keeps stretching for a while. The result indicates that a careful control of rib structure would enhance the water removal from the GDL. Therefore further studies for the optimum design of rib structure are needed. In an operating PEMFCs, the mechanism of water transport and wetting characteristics play an important role on flooding behavior. Therefore the results of the present study would contribute to the novel design for better water removal and flooding alleviation from the GDL.

2713

, and

Polymer electrolyte fuel cell (PEFC) has been developed as a power source for fuel cell vehicle (FCV). To have FCV used more commonly, it is required to reduce PEFC system cost[1]. Increasing output density is effective measures for cost reduction because it is necessary to reduce total cost of PEFC in addition to quantity of platinum catalyst. In recent years, high performance cathode catalyst layers (CLs) are proposed, which made contrivances to its catalyst structure[2-3]. However, the method for developing catalyst is trial and error and there is no clear design guideline for further improvement. The purpose of this study is to propose the optimum design of cathode catalyst (Pt/CB) by simulating the electrochemical reaction and PEFC transport phenomena. In this study, based on the our previous research[4], diffusion phenomenon in the catalyst is newly reflected on the 1D model to compare catalyst structures.

Fig.1 shows the image of transport phenomena in the catalyst pore. As shown in the structure (A), if Pt is covered with ionomer, oxygen reduction reaction (ORR) rate per unit Pt surface area is reduced by sulfonic group in ionomer. Thus it is assumed that if Pt exists in the pore as shown in the structure (B), ORR rate increases compared with structure (A) because there is no sulfonic group in the pore. However, oxygen concentration on the surface of Pt decreases compared with structure (A) because oxygen diffusion resistance increases. In this research, reflecting this trade-off between ORR rate and diffusion resistance, we constructed the model to compare catalyst structures.

Based on the porous electrode theory[4], ORR and mass transport at the fuel cell evaluation test were simulated. In this research, generated water was considered as vapor, carbon corrosion reaction and suppress elution of platinum were neglected. The temperature profile in the cell was presumed to be uniform (80 ℃). Fig.2 shows some catalyst structures for simulation. Spherical diameter of carbon black (CB) and Pt catalyst carrying rate of CB are same in all structures. Assuming that pores of the porous CB is filled with the generated water. In addition, diffusion coefficient of GDL and MPL are determined by the experimental correlation equation for porous structures of PEFC. Each structure was simulated under various exchange current density. And oxygen distribution and relative reaction ratio were investigated in the catalyst.

Fig. 3 shows the current-voltage characteristic curves obtained from the simulation (Exchange current density in the pore is five times as large as covered with ionomer.) Str.2-6 has larger output density at low current density region. In this case, this result shows porous carbon black has larger output density than Str.1 for PEFC because practical voltage region is 0.6 V - 0.8 V. Fig.4 shows the radial relative reaction rate in the pore.Reaction rate decreasing for the center direction, we were able to reflect oxygen diffusion resistance and compare catalyst structures. Furthermore, we attempt to separate factors of performance degradation of fuel cell and evaluate each factors quantitatively to confirm the appropriateness of the model.

References

[1] Department of Energy, "2012 Annual Merit Review and Peer Evaluation Meeting", http://www.hydrogen.energy.gov/pdfs/review12/fc000_papageorgopoulos_2012_o.pdf

[2] Y. C. Park et al., 228th ECS Meeting, 1499(2015)

[3] Y. Chino et al., J. Electrochem. Soc., 163(2015), F97-105

[4] G. Inoue et al., ECS Trans., 50(2013), 461-468

Figure 1

2714

, , , , and

Direct methanol fuel cell (DMFC) is promising energy source for portable and automotive applications, mainly due to their low operating temperature, direct use of liquid fuel, and simple structure without the stringent need for a reformer [1-3]. Carbon black, nanometer-size carbon particles, is commercially used as the catalyst support in fuel cell owing to its high surface area, porosity, electric conductivity, low density, and low cost. In the previous work, we have used various carbon nanomaterials as a catalyst support for DMFC [4]. In this study, we measured the powder conductivity of carbon nanomaterials including carbon nanocoil (CNC), carbon nanoballoon (CNB), Vulcan XC-72R (Vulcan), and vapor-grown carbon fiber (VGCF-H). Under compression of these materials, it is shown that the electrical conductivity of carbon nanomaterials did not only depend on its intrinsic morphological properties, which determine the degree of packing of the material and hence the change in density, but also on such extrinsic factors as the applied pressure and the ambient humidity. In addition, this study investigates the effects of the conductivity and structure of the carbon nanomaterials used in the anode catalyst layer (CL) on the performance of DMFC using transmission and scanning electron microscopies, polarization technique, and electrochemical impedance spectroscopy (EIS).

 CNCs were synthesized using an automatic chemical vapor deposition system with a consecutive substrate transfer mechanism. The fiber diameter of the CNCs is ~300 nm, the coil diameter is ~1000 nm, and the coil length is ~10 μm. Arc black (AcB) was synthesized using the twin-torch arc discharge apparatus developed in our laboratory. CNB was obtained by heating AcB in a Tammann oven in Ar gas at 2600 ºC for 2 h. Commercially available Vulcan (Cabot Corp., Boston, MA, USA) and VGCF-H (SHOWA DENKO K. K., Tokyo, Japan) were used as the Vulcan and VGCF-H samples, respectively. CNB and Vulcan were composed of spherical with a particle diameter of ~50 nm. The fiber diameters of the VGCF-H were ~15 nm, and their length was ~3 μm. Powder conductivity of each sample was measured by a source meter, with an applied voltage of 0.1 V at room temperature. 300 mg of the sample was set in acrylic pipe. Subsequently the sample was compressed between the brass pistons. The compressive force was varied from 0.01 to 1.0 MPa.

 Nafion®115 membrane (Dupont) was used as electrolyte membrane. The anode and cathode catalysts used were 30-wt.% PtRu/CNC, /CNB, /Vulcan, and /VGCF-H and 50-wt.% Pt/C (Tanaka Kikinzoku International K.K), respectively. The membrane electrode assembly (MEA) was mounted into the DMFC cell (Japan Automobile Research Institute). In this performance testing, 0.5 M (M = moldm-3) methanol solution were supplied to the anode at a flow rate of 0.1 mLs-1, and dry air was supplied to the cathode at a flow rate of 5 mLs-1. The DMFC was operated 60 ºC and its polarization characteristics and EIS were measured using a fuel cell impedance meter (Kikusui Electronics Corp., KFM2030). The carbon nanomaterials and the surface morphologies of the anode catalyst layer were examined by TEM (JEM-2100F, JEOL, Tokyo, Japan) and SEM (S-4500 II and SU8000, Hitachi, Tokyo, Japan), respectively.

 The density of the carbon nanomaterials by compressive forces depends on the rearrangement and fragmentation of agglomerates [5]. In addition, the powder conductivity during compaction is mainly governed by the increase of particle contact area. From the measurement results of the powder compression electrical conductivity, CNB showed comparable powder conductivity to Vulcan. Moreover, VGCF-H showed the highest conductivity.

 The figure shows the cell polarization and power density of the DMFCs with different carbon nanomaterials in the anode CL. The DMFC performance exhibited the highest power density (15.3 mW cm-2) when CNC was used as the catalyst support in the anode CL, while that using CNB showed the lowest power density (8.1 mW cm-2). Therefore, the DMFC performance was not correlated with the results obtained by the measurement of the powder compression electrical conductivity. From the results of the EIS measurement and SEM images, it is suggested that the interface state of the catalyst supports was also an important factor in the DMFC performance.

References: [1] J. S. Yi and T. V. Nguyen, J. Electrochem. Soc., 145, 1149 (1998); [2] S. H. Ge and B. L. Yi, J. Power Sources., 124, 1 (2003); [3] C. K. Dyer, J. Power Sources., 106, 31 (2002); [4] Y. Suda, Y'. Shimizu, et al, Materials Today Communications., 3, 96 (2015); [5] B. Marinho, J. Powder Technology., 221, 351 (2012).

Figure 1

2715

We studied the degradation and durability of polymer electrolyte membrane fuel cell (PEMFC) at membrane-electrode-assembly (MEA) level by injection of octamethylcyclotetrasiloxane (D4) as a representative siloxane, which has been found in many industrial and personal products. Specifically, i) GC/MS analysis demonstrated that the ring-opening polymerization of D4 could result in the formation of various linear and cyclic siloxanes in both electrodes of MEA; ii) post-test analysis revealed that the transformed siloxanes were transported from the anode to the cathode via free-volumes in the polymer membrane; iii) RDE measurement and DFT calculation revealed that D4 was not directly responsible for the electrocatalytic activity of Pt; iv) electrochemical analysis demonstrated that the residual methyl groups of siloxane and various siloxanes did not hinder the proton transport in the polymer membrane; and v) siloxanes accumulated in the primary and secondary pores with the exception of an external surface of carbon, causing an increase in the oxygen reactant's resistance and resulting in a decrease of the cell performance. In addition, we confirmed that injection of D4 did not affect the carbon corrosion adversely because the siloxane had little influence on water sorption in the catalyst layer.

2716

and

The successful commercial deployment of proton exchange membrane fuel cells (PEMFCs) depends on the achievement of stringent performance and durability requirements. Currently, air is the most convenient oxidant for fuel cell applications, and its quality is an important consideration for operation because airborne contaminants can negatively affect fuel cell performance, cause premature degradation and decrease durability (1). Aromatic compounds are hazardous pollutants produced or used in many industrial processes. Benzene and naphthalene are the main representatives of aromatic hydrocarbons and are widely used as precursors for chemical syntheses. More than half of the entire benzene production is processed to styrene to manufacture polymers and plastics. Benzene is originated in the air from emissions of chemical plants, burning coal and oil, gasoline stations and vehicle exhausts. Naphthalene is used in the production of phthalic anhydride and as a pest control agent. Determination of the impact of C6H6 and C10H8 on PEMFC performance is critical to establish environmental requirements for fuel cell usage, define specifications for air filtration systems and support understanding of fundamental aspects of PEMFC operation and maintenance.

A segmented cell and data acquisition system were used (2) with a commercially available 100 cm2 membrane/electrode assembly (MEA). Each electrode contained a Pt/C catalyst with a loading of 0.4 mgPt cm-2. A segmented SGL 25BC gas diffusion layer (GDL, 10 segments of 7.6 cm2) and a Teflon gasket were employed at the cathode whereas a single GDL piece was applied at the anode. The MEA was operated under galvanostatic control of the whole cell current. Other operating conditions were: 80°C, 48.3 kPagback pressure, 100/50% relative humidity and 2/2 stoichiometry for the anode and cathode respectively. The dry contaminant was injected into the humidified cathode air stream. The poisoning proceeded until the cell voltage reached a steady value. Subsequently, the contaminant injection was stopped to evaluate the cell self-recovery.

Fig. 1 a) shows the voltage response and normalized current density for each segment at 1.0 A cm-2 under benzene contamination. For the first 18 hours, the cell was operated with pure air resulting in a cell voltage of 0.670 V. The injection of 2 ppm C6H6 decreased the voltage to a steady state of 0.560 V and caused a redistribution of local current densities. Operation with pure air fully recovered the initial cell performance. Effects of naphthalene are shown in Fig. 1 b). The introduction of 2.3 ppm C10H8to air stream led to a significant voltage drop within 10 h and a different current redistribution pattern. Voltage oscillations from 0.100 to 0.170 V were observed as soon as the cell reached 0.12 V. Recovery took 2 h and was accompanied by further redistribution of localized currents.

Benzene and naphthalene have similar electrochemical properties (3, 4). Cathodic desorption of the adsorbed species on Pt occurs at hydrogen adsorption potentials (< 0.1 V) and is accompanied by partial hydrogenation, while electrooxidation takes place at 1.35 V with formation of CO2 as the main product. The strong adsorption of C6H6 and C10H8 occurs at 0.1-0.6 V without electrochemical reactions, and results in an in-plane adsorbate configuration due to the interaction between the aromatic ring and the Pt surface (5). Contaminant adsorption results in a decrease of the electrochemical area, suppression of O2 adsorption and shift the oxygen reduction from a 4-electron to a 2-electron mechanism, which negatively impact PEMFC performance. The data demonstrated that C10H8 has a severer effect on PEMFC than C6H6 which is most likely due to a higher adsorption energy and abilty to form multilayer adsorption (6, 7). A detailed discussion of the results and a poisoning mechanism will be presented.

ACKNOWLEDGMENTS

We gratefully acknowledge ONR (N00014-13-1-0463), DOE EERE (DE-EE0000467) and Hawaiian Electric Company.

REFERENCES

  • O.A. Baturina, Y. Garsany, B.D. Gould, K.E. Swider-Lyons, in: H. Wang, H. Li, X.-Z. Yuan (Eds.), PEM fuel cell failure mode analysis, CRC Press, 2011, p. 199.

  • T.V. Reshetenko, G. Bender, K. Bethune, R. Rocheleau, Electrochim. Acta, 56, 8700 (2011).

  • F. Montilla, F. Huerta, E. Morallon, J.L. Vazquez, Electrochim. Acta, 45, 4271 (2000).

  • T. Löffler, E. Drbalkova, P. Janderka, P. Königshoven, H. Baltruschat, J. Electroanal. Chem., 550-551, 81 (2003).

  • M.P. Soriaga, A.T. Hubbard, J. Am. Chem. Soc., 104, 2735 (1982).

  • C. Morin, D. Simon, P. Sautet, J. Phys. Chem. B, 108, 12084 (2004).

  • J.M. Gottfried, E.K. Vestergaard, P. Bera, C.T. Campbell, J. Phys. Chem. B, 110, 17539 (2006).

Figure 1

2717

, , and

A novel nitrogen-rich support material (CNNF-G) consisting of graphitic carbon nitride (g-C3N4) nanoflakelets (CNNF) and reduced graphene oxide (rGO) is designed and fabricated for loading Pd nanoparticles. Structural characterizations indicates that the CNNF is formed via splitting decomposition of the g-C3N4 polymer on rGO at higher temperatures and the resulting CNNF is intimately coupled to the rGO sheets. The CNNF can provide more exposed edge sites and active nitrogen species for the high dispersion of Pd NPs. It is found that the Pd NPs with an average diameter of 3.92 nm are uniformly dispersed on CNNF-G sheets. DFT computations reveal that CNNF can trap Pd adatom and thus act as a Pd nucleation site at which Pd atoms tend to accumulate to form Pd clusters. The Pd-CNNF-G nanocatalyst exhibits excellent electrocatalytic activity for both formic acid and methanol oxidation reactions, including large electrochemically active surface area (ECSA) values, significantly high forward peak current densities, and reliable stability and durability, far outperforming the Pd-graphene, commercial activated carbon-supported Pd catalyst or Pd-carbon nanotubes. Such a stable Pd/CNNF-G nanocatalyst may bring new design opportunities for high-performance direct formic acid fuel cell (DFAFC) and direct methanol fuel cell (DMFC) in the future.

Figure 1

2718

, , , and

In electrode catalysts used for solid polymer fuel batteries, reducing the usage quantity of Pt has become a critical issue. For that reason, catalysts with greater oxygen reduction activity are developed daily. Among those catalysts, Pt alloy catalysts using transition metals are known for their property of high oxygen reduction activity. On the other hand, such Pt alloy catalysts, which use transition metals, have a problem in that their performances degrade depending on the usage conditions or environment.Conventionally, the degradation behavior of catalysts has been analyzed on the basis of the changes in oxygen reduction performance or the structure observation and composition reduction after their measurements. Though those methods can verify the performances before and after the experiment, it is difficult for them to analyze what behaviors the gradations occur with. In the study here, dissolution behaviors of Pt model electrodes were analyzed by the EQCM method, which can perform in situ measurement on the mass change caused by electrochemical reactions.

2719

, , and

In order to familiarize fuel cell vehicles (FCV), its system cost reduction is a critical issue. High current density operation is an effective approach, but it is necessary to reduce activation overvoltage, IR loss and concentration overvoltage. In particular, the concentration overvoltage must be decreased to increase oxygen flux at high current density. In order to reduce concentration overvoltage, oxygen gas diffusion characteristics in gas diffusion layer (GDL) and microporous layer (MPL), oxygen dissolution and diffusion characteristics in ionomer, reduction of oxygen local flux to platinum and influence analyses of water retention in MEA to PEFC performance have been researched by various approaches. In recent years, not only water retention in GDL / MPL but the effect of water retention in catalyst layer (CL) on cell performance has been reported by advancement of MEA inside visualization technique. However, relationship between the catalyst layer structure parameters and liquid water retention characteristics has not been found out yet. In this study, to find out this relationship and influence of water retention in CL to oxygen transport resistance and PEFC performance, porosity/tortuosity (ε/τ) measuring method has been developed under operating condition of PEFC as first step. The ε/τ of CL did not change with temperature (25-80 deg.C) under dry environment, which is reasonable due to low thermal expansion coefficient of each CL material. Currently, we are planning this technique to humidity control and the results will be reported.

 Acknowledgments

This research was financially supported by New Energy and Industrial Technology Development Organization (NEDO), Japan.

I01-B Poster Session - Oct 5 2016 6:00PM

2720

and

With the concern of global warming, air pollution and energy security, the use of polymer electrolyte fuel cells (PEFCs) has achieved substantial momentum for future sustainable and renewable energy conversion systems. The fuel cell (FC) is not a new invention and its principle dates back to 1838. FC science and technology cut across multiple disciplines, including (i) materials science, (ii) chemistry, (iii) (iv) mechanical/chemical engineering and (v) catalysis.

The PEFC has emerged as the leading fuel cell type for automotive and some portable applications, and also as back-up power due to its operation at low temperature, comparative simplicity in construction, high power density, and ease of operation. In spite of tremendous scientific advance as well as engineering progress over the last decades, the commercialization of PEFCs has been delayed, due primarily to the following aspects: (1) technical problems mainly concerning water management, (2) economic viability associated with high prices for materials and components, (3) membrane fragileness and (4) membrane hydration. The difficulty in understanding water management lies mostly in the two phase multi-component flow involving phase-change in porous media, coupled heat and mass transfer, interactions between the porous layers and the gas channel (GC), and the complex relationship between water content and cell performance. In PEFCs the electrochemical reactions are strongly coupled to the transport of gas phase and liquid phase species, momentum, charge and heat. The transport, phase-change and reaction processes within PEFCs occur at different length and time scales simultaneously. Due to the low-temperature of the operation, water generated by the electrochemical reactions often condenses into liquid phase, potentially flooding the catalyst layer, gas diffusion layer (GDL), micro porous layer (MPL) and GC (see Fig. 1). Insight into the fundamental processes of liquid water evolution and transport is still lacking, preventing further enhanced fuel cell development.

Diffusion media characterization and development still rely heavily on in-situ testing because of well-established correlations between in-situ performance results and ex-situ characterization data are not yet available. This limitation has resulted in the development of detailed computational fluid dynamics (CFD) based models where the ability to predict local and global characteristics such as voltage, current density, temperature and concentrations have been demonstrated. However, research combining experimental and modeling activities in a systematic iterative way is still missing. Some models assumes the GDL (and MPL) water transport as singe phase only. CFD models make it possible to reduce the number of experiments needed for cell design and development, and a decreased amount of tests are then required to validate the accuracy of the models. Modeling can also be used to confirm experimental results and conclusions. Various assumptions are made in CFD models, e.g., the gas phase, liquid phase, ionic and electronic tortuosities as well as the contact angles are normally treated as fitting parameters, used in the respective governing equations, thus unrealistic values may be assumed, or the property values might not be representative for the corresponding microstructures.

2721

Xin Gao[1], Torsten Berning, Søren Knudsen Kær

Department of Energy Technology

Aalborg University

Pontoppidanstraede 111, 9220 Aalborg, Denmark

Proton exchange membrane fuel cells (PEMFC's) are becoming increasingly popular for uninterrupted power supply especially in remote areas. In the case of telecom back-up operations, PEMFC systems are often placed in areas of extreme climates, e.g. in Norway or Canada where the temperatures drop below -20 °C in the winter which make liquid-cooled fuel cells impossible. In such cases, air-cooled fuel cell systems are deployed where the air that is fed to the fuel cell serves both as reactant supplier and coolant to remove the waste heat that is generated during fuel cell operation. In some cases the warmer exhaust air is used to pre-heat and also humidify the incoming colder and dryer air stream using an enthalpy wheel.

It is important to thermodynamically understand such a fuel cell system, and in this work the enthalpy streams and the humidity stream are followed throughout the fuel cell system in order to optimize the operating conditions and the performance of such a system. The adjustable parameters include the fan speed that determines the amount of air that is brought into the system, and the size and rotating speed of the rotating enthalpy wheel. In addition, computational fluid dynamics simulations have been carried out to better understand the distribution of the reactant air over the fuel cell stack and the resulting temperature distribution across the stack. These results suggest that the humidifying function of the current enthalpy wheel is negligible and a smaller enthalpy wheel or an ordinary heat exchanger can fulfill the heat recovery demand. Despite the fact that the air enters the stack at a cold temperature, even the forefront of the stack is at a much elevated and desired stack temperature with the help of supplying an acceptable amount of power to an electric stack heater. So a preheating of the air stream by the enthalpy wheel might be unnecessary during steady-state operation.

[1] Corresponding Author: xga@et.aau.dk

2722

, , , , and

Steam reforming (SR) of dimethyl ether (DME) is one of the promising ways to produce hydrogen for fuel cells (FCs) [1]. DME is harmless with a high H/C ratio, handling DME is easy because it is liquefied under ca. 6 atm and conventional facilities providing LPG can be used due to the similarity of DME to LPG.Coppper based materials were considered as good catalyst for DME SR [2]. However the durability of copper based catalysts is not good because of the copper sintering as reaction running. There are two ways to improve the durability of copper based catalysts, one is to enhance the dispersion of copper or to improve the thermal stability of copper by forming a spinel oxide or alloy [3].

γ-Al2O3 was used as support. The Cu/ZnO/γ-Al2O3, Cu/ZnO/Cr2O3/γ-Al2O3 and Cu/ZnO/Fe2O3/γ-Al2O3 catalysts were prepared by impregnation method. The nitrates were impregnated into support for 6 h, then dried at 323 K for 12 h, calcined at 823 K for 4h. The DME SR was carried out with a fixed bed flow reactor under atmospheric pressure using 1 g of catalyst at 773 K. After the catalyst was reduced by 12 v% H2/Ar at 503 K for 3 h, the mixed gas of DME and N2 (20 v%) were fed into the reactor with GHSV=500 ml/(g·h), the water was pumped into the reactor though a 473 K heater. The H/C ratio was 3. The catalysts were characterized by means of XRD, TPR, SEM, and BET.

Steam reforming of DME was carried out at 773 K over Cu/ZnO/γ-Al2O3, Cu/ZnO/Cr2O3/γ-Al2O3 and Cu/ZnO/Fe2O3/γ-Al2O3 catalysts. The conversion of DME and the hydrogen yield reached 100% and 87% respectively for Cu/ZnO/γ-Al2O3 at the initial stage. The durability of this catalyst exhibted degradation. After running for 97 h, the hydrogen yield decreased to less than 60%. The reason for this degradation was supposed to be the sintering of copper, therefore, a metal oxide with better thermal stability was added into the catalyst to prevent the sintering of copper. The time-on-stream results indicate that the addtion of Cr2O3 or Fe2O3 can improve the durability of the catalyst for DME SR significantly. After running for 100 h, the hydrogen yield was kept at 84% and 85%, respectively. Figure 1 is the time-on-stream result of DME conversion over Cu/ZnO/ Cr2O3. Figure 1 is the time-on-stream result of DME conversion over Cu/ZnO/ Fe2O3. The mechanism of the improvement of the durability of catalysts were further investigated by XRD, SEM, BET and TPR.

Figure 1

2723

, , and

Recently, in order to fulfill the requirements for decreasing air pollution caused by the use of fossil fuels and achieving low or zero emission, some renewable sources such as solar, hydrogen fuel cells, wind, geothermal, biomass, and tidal energy have been taken center stage under the international spotlight. Among them, the fuel cell has been broadly considered as a renewable and environmental energy-conversion solution for settling aforementioned problem. Specifically, the polymer electrolyte membrane fuel cell (PEMFC) has numerous advantages including low-operating temperature, fast start-up ability, high current density, and high potential for low cost when compared to other sources. Then, in order to utilize this PEMFC efficiently, it is surely necessary to have an optimal fuel cell management system. For this purpose, an equivalent electrical-circuit model (ECM) should be elaborately designed to describe the dynamic characteristic of the fuel cell terminal voltage in a PEMFC. Therefore, the most important thing in the ECM-based fuel cell management is to achieve an accurate experimental PEMFC voltage signal (EPVS). Unfortunately, there may be at risk for instantaneous and unwanted sensing of noisy. This absolutely results in erroneous fuel cell management system caused by noise-including EPVS loss. As a result, there should be an imperative to investigate a sophisticated approach to eliminate the noise from this inevitable loss. So far, no definitive solution has been provided to settle this problem.

This approach newly gives insight to the design and implementation of the wavelet transform-based multi resolution analysis (WTBMRA) for noise elimination and ultimately an efficient PEMFC operation accomplishment. As is well known, the WTBMRA is more remarkable for processing of signal that having non-stationary and transient phenomena characteristics. Fortunately, because of various fuel cell voltage forms by different time intervals and magnitudes, the EPVS can sufficiently considered as an original signal in the WTBMRA. The WTBMRA has the decomposition and reconstruction abilities with a vigorous function of both time and frequency localization of the EPVS. Therefore, it is capable of obtaining low- and high-frequency components An and Dn. Like other studies dealing with noise elimination, there are three steps for noise elimination in this approach. The noticeable difference between previous approaches and the proposed approach is to compare the analytic results of noise elimination with various decomposition levels based on the WTBMRA. According to this analysis, it can be expected to select an optimal decomposition level that clearly shows the high performance on noise elimination under the same condition of mother wavelet. Basically, this approach finally considered the Daubechies as a mother wavelet which belongs to the family of orthogonal wavelet filtering the WTBMRA. For reference, the Daubechies wavelet (db) is extensively used in solving a board rage of problems due to its orthogonal property. In this approach, the decomposition and reconstruction scale in the WTBMRA is selected as 3. With regard to the signal-to-noise (SNR) ration, all comparative analyses are elaborately evaluated for determining the optimal decomposition and reconstruction levels. Two techniques on noise elimination such as hard-thresholding and soft-thresholding with threshold value by VisuShrink calculation are implemented. Consequently, our analytic results sufficiently the suggest the clear comparison by showing the SNR difference using two techniques. From this approach, there are two conclusions. The first conclusion is that the performance on noise elimination using soft-thresholding is always inferior to that of hard-thresholding. The second conclusion is that there are the highest SNR values at each specified decomposition and reconstruction levels. Our experimental apparatus basically designed for obtaining the noise-including EPVS from the PEMFC by the 'Materials and Electro-Chemistry Laboratory' in Inha University.

Figure 1

2724

, , , , and

This paper describes a Portable Polymer Electrolyte Membrane Fuel Cell (PEMFC) System developed at HySA Systems, UWC. The system has a rated maximum output power of 130 W at 240 VAC and 5 VDC USB output port. Hydrogen is supplied to the PEMFC using Metal Hydride (MH) on the basis of a multi-component AB2 – type hydrogen storage alloy (A=Ti,Zr; B=Mn,Fe,Cr,Ni; Ti:Zr=0.55:0.45). The stainless steel MH canister with external aluminium fins was developed in-house. The heat transfer performance was optimized by utilizing external fins and compacting thermal expanded graphite (TEG) with the MH powder to form pellets. The MH canister has a maximum hydrogen storage capacity of 90 NL H2.

It is well-known that the endothermic desorption of H2 from MH results in the significant cooling, which decrease H2 supply flow rate. To overcome this problem and improve the dynamic performance, the MH canister was placed in front of the fuel cell exhaust to utilise the waste heat generated by PEMFC. By utilisation of the PEMFC waste heat in combination with external fins and MH / TEG compacts, the MH canister allowed for >40 minutes-long stable operation at the stack power of 130 W (225 W/kg(MH)) with the utilisation of >90% of the stored H2. Figure 1 illustrates the MH storage unit integrated with 200 W PEMFC stack, Figure 2 shows the rated power of the PEMFC and temperature of the MH canister at different loads connected to the system.

Figure 1

2725

, , and

Proton Exchange Membrane Fuel Cell (PEMFC) is an electrochemical device that convert the chemical energy of reactants directly to electrical energy. The water-based, acidic polymer membrane usually used as its electrolyte. Therefore, the bipolar plate, as one of the most important components in PEMFC system, is required to be high corrosion resistance in the acidic condition. Currently, researchers have tried to develop metallic bipolar plate using a stainless steel. In case of using stainless steel, a passivation or protective coatings layers should be applied to improve its stability in the harsh environment. However, the general intrinsic defects (columnar structures, pinholes, pores, discontinuities) in the protective films, which can greatly affect its high performance. Therefore, a passive thin film with low defects and stable property was needed to be developed.

In this work, TiN thin films were deposited on 316L stainless steel as the protective layer by using plasma enhanced atomic layer deposition (PEALD). For comparison, tetrakis-dimethylamido-titanium (TDMAT) and titanium tetrachloride (TiCl4) were used as different Ti sources, respectively. The microstructure, chemical composition and the corrosion behavior of TiN protective layer deposited by different sources were investigated by 4-point probe, XRD, FESEM, AES and potentiodynamic polarization equipment. The SEM results showed that the uniform TiN protective layers were successfully deposited on the 316L stainless steel substrate by both two sources. and the FCC crystal structure was confirmed by the XRD analysis. In addition, the TDMAT-deposited TiN has lower the corrosion current density (Icorr) and higher corrosion potential (Ecorr) than those of TiCl4-deposited TiN, which could be inferred that the TDMAT-deposited TiN will have a higher corrosion resistance and protective efficiency during the fuel cell application.

2726

, , and

One of the key issues to ensure the safety of a ship in the case of a gas leakage is the condition of air flow inside a fuel cell stack package. Unstable air flow increases the explosion risk by allowing the accumulation of leaked gas in one place. Therefore, a place for installing a ventilating fan and its capacity should be properly selected in order to make the air flow smoothly inside the package. Furthermore, an alarm unit and gas detector should be installed at the proper place for detecting leaked gas.

This paper studied the characteristics of air flow inside the stack package using CFD techniques (FLUENT flow analysis) in order to ensure safety in the case of a gas leak for the 300kW MCFC stack package model for ships. The analysis of a leaked gas accumulation is to decide for the installing sites of the alarm unit, gas detector, and ventilating fan, as well as its capacity. As a subsequent, a model for flow field including a fuel cell stack module, BOP and various pipes was prepared.

 With the developed stack package model, a temperature flow analyzation was performed for the heating conditions of each component or pipe to study where the temperature was evenly distributed inside the package and the temperature rise at the outlet of the model. Using the analytical model for the gas leakage that may occur in the valve, analysis of the gas distribution inside the package was also conducted.

2727

, and

Methane steam reforming reaction is the most important chemical process to produce hydrogen or synthesis gas. Hydrogen is heavily consumed for ammonia production, the cryogenics industry and methanol production. Recently, the hydrogen demand is expected to increase as fuel cells become more widely accepted and are used more in the near future. For effective production of hydrogen or synthesis gas, the role of the reforming catalyst becomes more significant. Especially, highly active and stable catalyst is necessary for an on-site reformer for fuel cell systems.

In conventional technology, the methane steam reforming reaction is conducted on supported noble metals- (Pt, Pd, Rh, Ru, and Ir) or nickel-based catalysts at temperature up to 700~800°C and steam to methane rations between 2 and 4. However, these catalysts suffer from the deactivation by agglomeration and carbon deposition. Noble metal-based catalysts are less sensitive to carbon deposition than nickel-based catalyst. However, high cost and limited availability are major concern.

In this study, nickel-based nanocomposite catalysts were fabricated by exolution process. The exsolution means the process to precipitate particles from solid solution by means of the heat treatment in a specially-controlled atmosphere. This process is distinguished from the infiltration in which particles are precipitated from solutions by evaporation. First, Mg1-xNixO solid solution powders were synthesized from aqueous magnesium and nickel nitrate solution by precipitation technique and then the powder was heat-treated in reducing atmosphere at 600 to 900°C. SEM and TEM images revealed that the nano-sized nickel particles were homogeneously dispersed in the Mg1-xNixO solid solution matrix and the size and morphology of nano nickel particles can be controlled by the heat-treatment condition. Catalytic activity for the methane steam reforming or methane particle oxidation reaction and durability of the Ni/Mg1-xNixO was investigated in terms of nickel contents and sizes.

2728

, and

In generally, Bipolar plate is vital component of PEM fuel cells, which supplies fuel and oxidant to reactive sites, removes reaction products, collects produced current and provides mechanical support for the cells in the stack. At the present, the bipolar plate for fuel cell vehicle is metallic separators formed of metallic materials such as stainless steel. The existing bipolar plate includes a land(contacting) portion directly bonded to the gas diffusion layer(GDL) and a channel portion that serves as a supply passage of reaction gases and an exhaust passage of water between land portions. In the typical separator, the land portion and the channel portion are disposed in a longitudinal direction and a flow field area in which the channel portion is formed both have a longitudinal structure. Also, the land portion and the channel portion are distinctly separated from each other.However, in the case of a typical separator, this non-uniformity causes a concentration difference between the area that the channel portion contacts and the area that the land portion contacts among the whole area of the MEA where the electrochemical reaction occurs and the performance of fuel cell is decreased.In this paper, a porous separator with various structures for a fuel cell, which causes the high diffusion of reactant gases was studied for increasing the performance of fuel cell. Also, we conducted an analysis of the properties(contact resistance, corrosion etc.), machinability(uniformity of thickness, change of hole structure etc.) and performance according to hole and pitch structure.In conclusions, the porous separator with optimum structure shows the performance above 200mA/cm2 at 0.6V in comparision with the typical separator. It is considered to be minimize concentration differences between area in a gas diffusion layer and achieve uniform electrochemical reaction and electricity generation over the whole reaction are, by improving the structure of a flow field in which reactant gases flow.

2729

, and

Methanol is a liquid fuel which can be produced from renewable energy sources and has high energy density. It has a low boiling point, a high H:C ratio of 4:1 without any C-C bond, therefore can be reformated at a relatively low temperature. In this work, we investigated the application of CGO(Gd:CeO2)-based catalyst for hydrogen production via methanol steam reforming. CGO-based powder material for a catalyst of methanol steam reforming was synthesized by glycine-nitrate process(GNP). GNP is one of the solution combustion synthesis methods, Metal nitrate and glycine were mixed with deionized water and solution heated by hot plate. After evaporation of water, solution converted to a viscous gel. The gel ignited with gases and powders. The synthesized powders were calcined for 4 hours at 800°C to remove moisture and impurity. Their properties such as micro structure, surface morphology, surface area and composition were investigated with XRD(X-ray diffraction), SEM(Scanning electron microscopy), BET (N2 adsorption-desorption), EDS(Energy dispersive spectroscopy) and ICP-MS(Inductively coupled plasma mass spectrometry). The performance of catalysts was examined with reactor at 200 - 400°C range. The reactants and reaction products were analyzed by GC(Gas chromatography). The catalytic performance for hydrogen production was analyzed in terms of methanol conversion and selectivity of products.

2730

and

It is important to uniformly supply the fuel gas into the reaction activity area in polymer electrolyte membrane fuel cell (PEMFC). Recent studies have shown that the cell performance can be significantly improved by employing metal foam gas distributor as compared with the conventional bipolar plate types. The metal foam gas distributor has been reported to be more efficient to fuel transport. In this study, three-dimensional computational fluid dynamics (CFD) simulations have been performed to examine the effects of metal foam flow field design on the fuel supply to the reaction site. Darcy's law is used for the flow in the porous media. By solving additional advection equation for fluid particle trajectory, the gas transport has been visualized and examined for various geometrical configuration of metal foam gas distributor.

†This work was supported by the International Collaborative Energy Technology R&D Program of the Korea Institute of Energy Technology Evaluation and Planning (KETEP) (No. 20148520120160)

2731

and

For optimizing the performance of proton exchange membrane fuel cells (PEMFCs), proper water management in the cell is one of the most important issues because liquid water in PEMFCs have a significant impact on the cell performance. If excessive water accumulate in the cell, including that produced by chemical reactions and the humidification of supplied gases, power generation performance will decrease since the transport of the fuel gas will be blocked by the liquid water that has accumulated in the void of a gas diffusion layer (GDL) or a gas flow channel. On the other hand, the drying of the polymer electrolyte membrane (PEM) causes not only a performance decrement due to an increased proton transport resistance, but also a degradation of membrane. Therefore it is very important to understand the water transport in PEMFCs. Some researchers have investigated the water transport phenomena and mass transfer characteristics in the GDL. Utaka and Koresawa have been proposed a novel GDL (Hybrid type GDL, Fig. 1) which has two different wettability to improve the oxygen diffusivity by controlling the water distributions and movement in the GDL, and have examined the effects of the wettability distribution and PTFE content in the hydrophobic region of the GDL on efficient oxygen diffusivity [1]. Additionally, they also have been proposed a novel gas channel with micro-grooves [2]. The cell performance of PEMFC that combined the hybrid GDL and the gas channel with micro-grooves were experimentally investigated in an attempt to improve the power generation of PEMFC. Figure 2 shows the effects of hybrid GDL and novel gas channel on cell performance, and from this figure, the PEMFC combined the hybrid GDL and gas channel with micro-grooves could enhanced the performance under high current density.

A three-dimensional numerical model, which includes the gas channels with micro-grooves, GDLs, catalyst layers in both cathode and anode side and PEM, has been developed to investigate the water transport and power generating characteristics. In addition, we investigate the effectiveness of the hybrid GDL to mitigate liquid water saturation under flooding conditions by using our comprehensive model, and compared with experimental results.

Numerical result of liquid water distribution at two different wettability region in GDL shown in Fig. 3. Liquid water distribution biased to hydrophilic regions due to the drawing the water from the hydrophobic region into the hydrophilic region. Furthermore, the performance of PEMFC improved by combined the hybrid GDL and novel gas channel with micro-grooves, and discussed quantitatively.

Reference

[1] R. Koreaswa and Y. Utaka, J. Power Sources, 271, 16 (2014).

[2] Y. Utaka, A. Okabe and Y. Omori, J. Power Sources, 279, 533 (2015).

Figure 1

2732

, and

In generally, degradation of membrane & electrode assembly(MEA) according to driving conditions of fuel cell electric vehicle focused on analyzing the state of catalyst, membrane, gas diffusion layer before and after durability evaluation . At present, the exact mechanism for degradation of membrane electrode assembly(MEA) according driving conditions of fuel cell vehicle does not hold and the mechanism were approximately the level that analogize the whole out of a part.

In case of the degradation of MEA by operating modes, it was shown on three sources; (1) the loss of apparent catalytic activity according to the time(ηa), (2) conductivity loss by decreasing ion conductivity per the time, (3) the loss of rate of mass transfer according to the time. In other words, the equation about degradation of MEA can be seen that the basic form can be created by analyzing the polarization according to time.In this study, the 10 driving modes were devised by cell voltage behavior according to driving conditions of fuel cell vehicle. The operating modes were basically designed by high current, middle current and low current. Also, these modes were considered about the most commonly used driving power and time variation in accordance with the current change. The result of analyzing the MEA degradation by current/voltage variation was as below.

The activation overpotential shows a tendency to be rapidly decreased in early state and decrease or increase gradually without reference to operating conditions. The ohmic overpotential tends to decrease in proportion to the time. OCV(Open Circuit Voltage) shows the result of decreasing rapidly in high current mode than low current mode. In this paper, we conducted an analysis of the equation for degradation of MEA according to the time, current/voltage variation.

In conclusions, the system of fuel cell vehicle must be operated in high current range in order to prevent degradation of the MEA in similar load change condition. It is considered to be occurred in oxidation of electrode catalyst under high voltage condition.

2733

, and

Partial pressure of oxygen is a significant factor for determining performance of fuel cell. As is well known, the oxygen concentration in air is generally 21%, but it is about 20% at urban area which has numerous vehicles and is reduced by 18% at underground parking lot with stuffy ventilation facilities. Also, the oxygen concentration in air is 21% at a high altitude, but efficiency of fuel cell vehicle can be decreased by low oxygen partial pressure.

In this study, change of fuel cell performance was analyzed by applying empirical equation about I-V curves using experimental results according to oxygen concentration.

The performance of fuel cell in oxygen concentration of 18%~23% and altitude of 0~4km was prospected by a deducted empirical equation. In case of stoichiometric ratio(SR) 2.0, performance of fuel cell in oxygen concentration of 18% was decreased over 10% in comparison with concentration of 21%, the results was revealed as reduction in performance of 22%. Also, the performance of fuel cell at 4km altitude appeared decrease of 25% owing to low partial pressure.

2734

, , and

PBI(polybenzimidazole) based HT-PEFC(high temperature polymer electrolyte fuel cell) has been received a lot of attention as a great alternative of the LT-PEFC(low temperature polymer electrolyte fuel cell). Currently, life time of HT-PEFC was shorter than LT-PEFC, so life extension of HT-PEFC has been studied. But existing research works on the life time of HT-PEFC focus on finding out the causes of its performance degradation in the single cell or small size of sub-stack levels. Researches on the life time of HT-PEFC in a full stack level have nearly been conducted. In this paper, effective operational algorithm for HT-PEFC such as minimizing OCV conditions, selecting the proper operation temperature and purging the water inside of the HT-PEFC stack were established to avoid a performance degradation of HT-PEFC in a full stack level. 5kW HT-PEFC stack was manufactured by ourselves and conducted long-term operating test using the proposed operating algorithm. As a result, HT-PEFC stack was successfully operated with a less degradation rate during the target life time.

I01-F Poster Session - Oct 5 2016 6:00PM

2735

, and

 Hydrogen, produced by electrolysis, is one of the most promising means to store electricity coming from renewables and to address its intermittency. Due to the increase of global electricity share from renewable energy sources, large-scale storage systems are required. Alkaline electrolysis can be a suitable solution for the large-scale energy storage system. However, high overpotential and slow kinetics, particularly of the oxygen evolution reaction (OER), still limit the efficiency of the system. To enhance the efficiency of alkaline electrolysis cells, catalysts with high electrocatalytic activity for OER should be developed.

 Platinum-like behavior of tungsten carbide in surface catalysis was first discovered by Levy and Boudart in 1973[1]. In addition, tungsten carbide was reported to have a Pt-like electronic structure by Colton in 1975[2]. Since then, tungsten carbide has been studied widely and expected to replace noble metal catalysts which are used in many electrochemical reactions. In this study, we researched tungsten carbide as a support for nickel, and formation of bi-metallic surfaces for developing a high activity OER catalyst.

  WC was synthesized by two-steps. First, WC-Co was synthesized by heat treatment. Carbon nitride (g-C3N4) and tungsten chloride were used as a carbon and tungsten precursor. Cobalt precursor was also used to help carburization of tungsten. After synthesizing WC-Co, the acidic leaching process was followed by using sulfuric acid (H2SO4) to eliminate cobalt. Through additional mixing with 3d metal precursors and heat treatment, Ni/WC and NiFe/WC were synthesized. Furthermore, Ni/C was also prepared using same precursors as a counter group. Figure 1 shows the SEM images of the synthesized WC and Ni/WC. Cyclic voltammetry was performed to evaluate OER activity in 1 M KOH in potential range from 1 to 1.8 V.

  Figure 2 shows the OER catalytic activities of synthesized catalysts and Ir/C. Pure WC and Ni/C had 430 mV and 394 mV overpotential at 10 mAbold dotcm-2 respectively. In 3d Metal/WC cases, Ni/WC and NiFe/WC showed only 335 mV and 314 mV overpotential at 10 mAbold dotcm-2. Both catalysts had higher electrocatalytic activity for OER than Ir/C in alkaline media.  

References

[1] R. B. Levy, M. Boudart, Science, New series, Vol. 181, No. 4099 (1973) 547-579

[2] Richard J. Colton, Jan-Tsyu j. Huang, J.Wayne Rabalais, Chemical Physics Letters, Vol. 34, No. 2 (1975) 337-339

Figure 1

2736

, , , and

Two-dimensional layered transition metal dichalcogenides (2D TMDs) have been of tremendous recent interests. Understanding their growth behavior on specific substrate or support can provide critical insights for designing the TMDs with desirable structure and functionality for targeted applications. Particularly for the electrocatalytic hydrogen evolution reaction (HER), where porous carbons are predominantly used as conductive supports, revealing the growth behavior of TMDs on porous carbon materials can suggest rational design concept for the TMD-based HER catalysts. With an aim to investigate the growth orientation on porous carbon support, we synthesized MS2 (M = W or Mo) nanoplates embedded on porous carbon nanorod arrays by limiting their growth space at the nanoscale. We found that the horizontal growth is preferred in WS2 giving rise to monolayer nanoplates, which is contrast to the growth of MoS2 that favors vertical stacking to generate multilayer structure. Density functional theory calculations of adhesion energy of TMDs on porous carbons and their stacking energies revealed that WS2 favor the basal plane bonding with carbon support and horizontal growth, wheareas the edge bonding and vertical growth is favored for MoS2, supporting experimental results. The orientation-controlled WS2 monolayer NPs embedded in mesoporous carbon exhibited highly efficient electrocatalytic activity for the HER with a low overpotential of 179 mV (vs RHE) at ‒10 mA cm-2 and low Tafel slope of 63 mV sec-1.

2737

, , , , and

Polymer electrolyte membrane water electrolysis (PEMWE) is a promising technology which enables a production of high-purity hydrogen with high energy conversion efficiency. However, conventional PEMWE cells are very expensive because noble metal catalysts (a few mg cm–2) are used. The aim of our research is the reduction of the amount of noble metal catalysts to 1/10 maintaining high-efficiency ≥ 90%. We reported that IrOx nanoparticles dispersed on M-doped SnO2 (M=Nb, Ta and Sb) supports with fused aggregated structures exhibited much higher mass activities for the oxygen evolution reaction (OER) than a conventional catalyst (mixture of commercial IrO2 and Pt black, 50:50 in weight ratio) in 0.1 M HClO4 electrolyte at 80 oC.1 In the present research, we examined performances of membrane-electrode assemblies (MEAs) with the IrOx/SnO2 based anode catalysts.

Uniform size IrOx nanoparticles (ca. 2 nm) highly dispersed on M-doped SnO2 (M=Nb, Ta and Sb) supports were prepared in the same manner as reported previously.1 The amounts of Ir loaded were quantified to be 11.3, 10.4 and 11.0 wt% for IrOx/M-SnO2 (M=Nb, Ta and Sb), respectively. The IrOx/M-SnO2, Nafion® binder (the volume ratio of Nafion binder to oxide support = 0.67), ethanol and pure water were mixed by using a planetary ball mill. The anode catalyst ink thus obtained was sprayed onto the Nafion® membrane with the thickness of 50 μm, followed by hot-pressing. The Ir loading at the anode was ca. 0.1 mgIr cm−2. Commercial Pt/graphitized carbon black (GCB) catalyst was used as the cathode with the Pt loadings of ca. 0.35 mgPt cm−2. The Japan Automobile Research Institute (JARI) standard cell with a geometric area of 25 cm2 was used for PEMWE measurements. Before the measurements, pure water at 80 oC was supplied to the cell, at least, for 12 h, and the ohmic resistance of the cell was then recorded by AC milli-ohmmeter. The I-V curves of the cells were measured at 80 oC.

Figure 1 shows I-V curves of the cells with the IrOx/M-SnO2 anode catalysts. The data of catalyst loading, ohmic resistance and cell potential at 1 A cm−2 are summarized in Table 1. The cell with IrOx/Ta-SnO2 catalyst showed a lower cell potential at 1 A cm-2 and ohmic resistance than that with IrOx/Nb-SnO2 catalyst, probably due to an increase in the apparent electrical conductivities σapp of the catalyst. However, the overpotential of the cell with IrOx/Ta-SnO2 catalyst was fairly larger than the conventional cell with Pt black + IrO2. We have already reported that Pt nanoparticle contributed to improving electric conductivities for SnO2 supports, which may be related to a decrease in the electronic depletion region of SnO2.4 Then, we synthesized Pt-loaded IrOx/Ta-SnO2 catalyst by the colloidal method. The use of Pt-IrOx/Ta-SnO2 catalyst significantly reduced both ohmic resistance and overpotential, achieving to the cell potential of 1.63 V at 1 A cm−2. The improved cell performance can be mainly ascribed to increases in the σapp (as shown in Fig. 2) and utilization efficiency of the catalyst. In addition, it was found that IrOx/Sb-SnO2 catalyst prepared very recently exhibited ca. 7 times higher σapp than Pt-IrOx/Ta-SnO2 catalyst. The evaluation of cell performance with IrOx/Sb-SnO2catalyst is now in progress.

Acknowledgement

This work was supported by the New Energy and Industrial Technology Development Organization (NEDO) of Japan.

References

1) H. Ohno, S. Nohara, K. Kakinuma, A. Miyake, S. Deki, M. Watanabe, and H. Uchida, 229th ECS Meeting,abstract 1422 (San Diego, May 30, 2016).

2) K. Kakinuma, M. Uchida, T. Kamino, H. Uchida, and M. Watanabe, Electrochim. Acta, 56, 2881 (2011).

3) M. Watanabe, M. Uchida, and S. Motoo, J. Electroanal. Chem., 229, 395 (1987).

4) Y. Senoo, K. Kakinuma, M. Uchida, H. Uchida, S. Deki, and M. Watanabe, RSC Adv., 4, 32180 (2014).

Figure 1

2738

, and

PEM water electrolysis is a key component for closing the loop of the renewable energy eco-system. These high response water electrolysers are particular suitable for fluctuating power sources. Conventional PEM water electrolysers are typically operated at a current density of around 1 A/cm2 and are fairly expensive. One means of increasing the hydrogen yield to cost ratio of such systems, is to increase the operating current density. However, at high current densities, management of heat transfer and fluid flow in the anode GDL and channel becomes crucial. This entails that further understanding of the gas-liquid flow in both the porous media and the channel is necessary for insuring proper oxygen, water and heat management of the electrolysis cell.

In this work, the vertical upward gas-liquid flow pattern in a 0.5×1×94 mm micro-channel is both numerically and experimentally analysed. A sheet of titanium felt is used as a permeable wall for permeation of air to the column of water similar to the phenomenon encountered in Oxygen Evolution Reaction (OER).

The transparent setup is operated in situ and the gas-liquid flow regimes are identified using a high-speed camera. The picture shows how the transparent cell is made which consists of a channel for the inlet air and a channel for the water-bubble flow. The transparent material is Plexiglas that is sealed with a sheet of silicon.

The conventional co-current gas-liquid two-phase flow patterns, such as bubbly flow, slug flow and annular flow, are observed in the present micro-channel. The phenomenon is also analysed numerically in 3D, unsteady, euler/euler multiphase method using the commercial software ANSYS FLUENT 17 and the results show good agreement with the experimental data.

Influence of each multiphase flow regime is described in the study as well as the recommendation for improving the performance. A well management of the multiphase flow regime along the whole micro-channel length can assure a proper distribution of water inside the titanium felt.

Figure 1

2739

, , , , , and

Introduction

Hydrogen energy has been paid more attention since the commercialization of FCVs. In order to further distribute FCVs, efficient production of hydrogen fuel becomes important. Water electrolysis is one possible method to produce hydrogen without emitting any carbon dioxide. Moreover, water electrolysis is the best technology to successfully use excess renewable energy, which is often hard to use due to the strong energy fluctuation. With this method, renewable energy can be converted into hydrogen, which can be stored for a long time and used anytime in any places.

Among various water electrolysis systems, the system based the polymer electrolyte has many advantages. Easy and rapid starting/stopping of the system is possible. Also, there exist many similar points to the PEFC system, and the technologies developed for PEFC can be shared. On the other hand, anode catalysts must be newly developed for the electrolysis cell since the carbon support dispersing Pt nanoparticles in PEFC cannot be used under high potential like nearly 2.0 V. Therefore, in this study, novel iridium-based SnO2/VGCF electrocatalysts were developed. Since SnO2, mostly working as a support material for iridium-based catalysts, is highly durable even under such high potential, iridium-based SnO2/VGCF is most likely a good candidate for anode catalysts. Therefore, the aim of this study is development and electrochemical evaluation of iridium-based SnO2/VGCF electrocatalysts.

Experimental

Sn0.98Nb0.02O2/VGCF was synthesized through an ammonia co-precipitation method, using SnCl2・2H2O and NbCl5. Then, IrO2(Ir) nanoparticles were decorated on Sn0.98Nb0.02O2/VGCF via the method that the aqueous solution mixture containing Sn0.98Nb0.02O2/VGCF and H2IrCl6・nH2O was completely evaporated while stirring, followed by the heat treatment at 440 oC in air. Actual amounts of Sn0.98Nb0.02O2 and IrO2were determined based on weight loss by VGCF obtained from TG-DTA. Electrocatalysts were further analyzed by STEM and XPS

Results and discussion

The composition of IrO2/Sn0.98Nb0.02O2/VGCF was determined to be 23 wt.% of IrO2, 58 wt.% of Sn0.98Nb0.02O2, and 19 wt.% of VGCF. From STEM observation, IrO2 (or Ir) particles were successfully observed on the surface of SnO2. Furthermore, in order to clarify either IrO2 or Ir, Ir 4f binding energy was analyzed by XPS and is shown in Figs. 2 and 3. When the electrocatalyst was synthesized in air, iridium-based catalysts most likely exist as IrO2 according to the previous report[1] as seen in Fig. 2. On the other hand, when it was made under 5% H2-N2, the catalysts mostly stay as Ir metal particles as shown in Fig. 3. Electocatalytic activity on water electrolysis is further discussed with these newly prepared electrocatalysts.

References

[1] D.Xu, P.Diao, T.Jin, Q.Wu, X.Liu, X.Guo, H.Gong, F.Li, M.Xiang, Y.Ronghai, ASC Apple. Mater. Interfaces 7, 16746 (2015).

Figure 1

2740

and

The interest in harnessing energy from renewable sources to achieve environmental cleanliness in the energy industry keeps gaining momentum. The need to be able to convert and store all that renewable energy has rekindled interest in hydrogen as a clean and environmentally benign energy carrier. Likewise, the rising star of hydrogen enabled mobility (motive fuel cells) ads further impetus to addressing needs for economical means to generate hydrogen - the ultimate key to success for hydrogen as fuel in societal everyday transportation solutions.

Polymer Electrolyte Membrane (PEM) water electrolysis emerged recently as one of the better choice today to address the need for hydrogen fuel by virtue of being compatible with highly variable and unpredictable nature of the renewable energy generation. Even though PEM has in fact been used for quite a few years now without undergoing substantial improvements over those years now, however, with the new focus on hydrogen as the energy carrier there is much more interest in low cost/high efficiency H2 production. There are two main ways to lower the cost of hydrogen production via PEM water electrolysis: to lower the capital expenses (CAPEX) and/or to lower the operating expenses (OPEX). We at 3M have recently demonstrated a way to address reducing the high CAPEX by widening the range of current densities where electrolyzers can operate from a maximum of about 2.0 A/cm2, as used today in commercial electrolyzers, to as much as 20 A/cm2 by employing a novel 3M's proprietary Nano Structured Thin Film (NSTF) catalyst and more conductive 3M PFSA based electrolytes in the electrolyzer MEA1-3. What the cited work does not touch upon is the need to further address high electrolyzer CAPEX by lowering costs of individual components of electrolyzer cells. This could be done independently of the operational cost savings and in addition to them. One such area where improvements can still be made is in the use of more economical Gas Diffusion Layers. These differ in function from fuel cell GDLs by the fact that reactants move through them in different directions, the requirements for performance, durability and mechanical robustness are also much more stringent. The standard materials in use as GDLs today are based on sintered titanium plates and although they work well and perform satisfactory such GDLs are neither inexpensive nor are they compatible with high speed/low cost roll-to-roll manufacturing – a necessary requirement for substantial cost reductons.

In this work, we intend to present results of our attempts to investigate materials available in the fuel cell realm for compatibility and performance in electrolyzer cathodes as possible replacement for Ti sinters. We will also present data evaluating alternatives to rigid Ti sinters for electrolyzer anodes. All materials are selected such that compatibility of these candidate GDL materials with high speed/low cost roll-to-roll manufacturing process is not negatively affected by their properties and/or modifications.

 

  • Krzysztof A. Lewinski, Sean M. Luopa, (invited) "High Power Water Electrolysis as a New Paradigm for Operation of PEM Electrolyzer" (abstract 1948), Spring ECS Meeting, Chicago, IL, May 2015.

  • Krzysztof A. Lewinski, Dennis van der Vliet, and Sean M. Luopa, "NSTF Advances for PEM Electrolysis - the Effect of Alloying on Activity of NSTF Electrolyzer Catalysts and Performance of NSTF Based PEM Electrolyzers" (Abstract 1457), Fall ECS Meeting, Phoenix, AZ, Oct 2015.

  • Krzysztof A. Lewinski, Dennis van der Vliet, and Sean M. Luopa, "NSTF Advances for PEM Electrolysis - the Effect of Alloying on Activity of NSTF Electrolyzer Catalysts and Performance of NSTF Based PEM Electrolyzers", (10.1149/06917.0893ecst), ECS Transactions, 69 (17) , p. 893-917 (2015).

2741

, , , , , , and

 In recent, huge efforts have been conducted to relieve the global warming problem by the electrochemical conversion of CO2 into useful chemicals. Particularly, it has been reported that the Ag and Au are highly active materials to produce the CO with high selectivity. However, the cost problem originated from use of noble metal catalysts limits their practical applications.

 In this study, the Ag and Au foam structures were electrochemically fabricated as a catalyst for CO2 reduction to produce CO. As a first step, the Cu foam was prepared on Ti foil substrate by galvanostatic pulse deposition which consisted of low current for continuous Cu film and following high current for porous Cu. The pore size (22 ~ 56 μm) and wall thickness (14 ~ 42 μm) were controlled by concentration of Cu precursor and applied deposition currents. Then, the Cu foams were immersed in the electrolyte which contained noble metal precursors and additives, in order to minimize the noble metal usage as well as to increase the electrochemical surface area. The coverage and morphology of noble metals on the surface of Cu form were affected by the precursor and additives concentrations in displacement step. The catalytic activity of fabricated noble metal foams were examined for electrochemical CO2 reduction. As a result, it was revealed that the CO Faradaic efficiency was significantly related to the morphology and composition of noble metal foams.

A-31 GDL - Oct 6 2016 7:40AM

2742

, , , , , and

Two components of the PEM fuel cells play a critical role in obtaining a good performance, namely the gas diffusion layer (GDL) and the micro porous layer (MPL). The target of the reported investigation was the influence of the required gas diffusion layer compression, during stack assembly, onto PEM fuel cell performance. For developing a fuel cell it is necessary to assembly all the components together in order to ensure a good connection and to reduce the leakage problem, and this is done under a compression load. The compression applied can deform the GDL and modify its thickness, porosity, hydrophobicity and electrical resistance, and these can result in damaging or decreasing the performance of the PEM fuel cell. Also, the GDL may intrude in the flow field channels and can lead in significant disturbance of the reactants flow in the channels and consequently to a drop in the performance.

Numerical simulations based on ANSYS software were carried out in order to investigate the effect of the clamping force on the fuel cell performance. Therefore, 4 geometries were considered in our study, by taking into account the uncompressed and compressed GDL and the presence or absence of MPL. Different levels of GDL compression and intrusion were considered and simulated. The 3D multiphase model revealed that the GDL compression and intrusion influence the in-plane gradient in liquid saturation, oxygen concentration, membrane water content, and especially current density profiles. Also, our study showed that inserting a micro porous layer between catalyst layer and GDL can lead to an improvement of the cell performance because it helps ensuring a good water management, reducing liquid saturation.

2743

, , , and

In proton exchange membrane fuel cells (PEMFCs), appropriate water management is critical to achieve high power density operation with increased robustness. Proton exchange membranes (PEMs) require sufficient hydration to fulfill its function as proton conductor, while flooding at the cathode side can hamper the transport of reactant, resulting in deterioration of cell operation. Liquid water accumulation and its transportation is one of major issue for PEMFCs.

For diagnosis the water and mass transport ability the following experimental methods can be used: polarization, oxygen gain, limiting current density, calculate oxygen transportation resistance, electrochemical impedance spectrum (EIS), and visualization technique etc. However, there are not so many literatures point out where the water accumulation can cause the mass transport resistance arc. In this study, the cell operated temperature and cathode humidify has been selected as main parameters for investigating the mass transfer phenomenon. The EIS was using for indicating mass transport occurs condition and the soft X-ray radiography was using for verify the location of the water accumulation. The soft X-ray radiography [1-13] had been proved it is a powerful tool to investigate liquid water within membrane electrode assemblies (MEAs).

Fig. 1 (a) shows EIS and water extraction image with different humidification conditions. The liquid water accumulation within gas diffusion substrates/channel is barely effects nothing on the mass transport resistance art. The water extraction image implies the water accumulates within/nearby catalyst layer is the main reason to cause the mass transport resistance art. Fig. 1(b) shows EIS and water extraction image with different operating conditions. It is barely no mass transport resistance arc at high temperate condition, due to generation water is mainly as vapor form. This result indicates the mass transport resistance arc is important indicator which points out there are liquid water accumulate within/nearby catalyst layer.

Acknowledgement

This work has been supported by the New Energy and Industrial Technology Development Organization (NEDO) of Japan.

References

[1] T. Sasabe, S. Tsushima, S. Hirai, International Journal of Hydrogen Energy, 35 (2010) 11119-11128.

[2] P. Deevanhxay, T. Sasabe, S. Tsushima, S. Hirai, International Journal of Hydrogen Energy, 36 (2011) 10901-10907.

[3] P. Deevanhxay, T. Sasabe, S. Tsushima, S. Hirai, in: 11th Polymer Electrolyte Fuel Cell Symposium, PEFC 11 - 220th ECS Meeting, October 9, 2011 - October 14, 2011, Electrochemical Society Inc., Boston, MA, United states, 2011, pp. 403-408.

[4] T. Sasabe, P. Deevanhxay, S. Tsushima, S. Hirai, Journal of Power Sources, 196 (2011) 8197-8206.

[5] T. Sasabe, P. Deevanhxay, S. Tsushima, S. Hirai, Electrochemistry Communications, 13 (2011) 638-641.

[6] P. Deevanhxay, T. Sasabe, S. Tsushima, S. Hirai, Electrochemistry Communications, 22 (2012) 33-36.

[7] P. Deevanhxay, T. Sasabe, S. Tsushima, S. Hirai, in: 12th Polymer Electrolyte Fuel Cell Symposium, PEFC 2012 - 222nd ECS Meeting, October 7, 2012 - October 12, 2012, Electrochemical Society Inc., Honolulu, HI, United states, 2012, pp. 335-341.

[8] T. Sasabe, G. Inoue, S. Tsushima, S. Hirai, T. Tokumasu, U. Pasaogullari, in: 12th Polymer Electrolyte Fuel Cell Symposium, PEFC 2012 - 222nd ECS Meeting, October 7, 2012 - October 12, 2012, Electrochemical Society Inc., Honolulu, HI, United states, 2012, pp. 735-744.

[9] S. Tsushima, P. Deevanhxay, T. Sasabe, S. Hirai, in: 12th Polymer Electrolyte Fuel Cell Symposium, PEFC 2012 - 222nd ECS Meeting, October 7, 2012 - October 12, 2012, Electrochemical Society Inc., Honolulu, HI, United states, 2012, pp. 327-333.

[10] P. Deevanhxay, T. Sasabe, S. Tsushima, S. Hirai, Electrochemistry Communications, 34 (2013) 239-241.

[11] P. Deevanhxay, T. Sasabe, S. Tsushima, S. Hirai, Journal of Power Sources, 230 (2013) 38-43.

[12] T. Sasabe, S. Tsushima, S. Hirai, K. Minami, K. Yada, in: 9th Proton Exchange Membrane Fuel Cell Symposium (PEMFC 9) - 216th Meeting of the Electrochemical Society, October 4, 2009 - October 9, 2009, Electrochemical Society Inc., Vienna, Austria, 2009, pp. 513-521.

[13] T. Sasabe, P. Deevanhxay, S. Tsushima, S. Hirai, in: ASME 2011 9th International Conference on Fuel Cell Science, Engineering and Technology. Collocated with ASME 2011 5th International Conference on Energy Sustainability, FUELCELL 2011, August 7, 2011 - August 10, 2011, American Society of Mechanical Engineers, Washington, DC, United states, 2011, pp. 163-169.

Figure 1

2744

, , and

Unified fuel cell reaction modeling was proposed to predict the effects of local deformation of porous-streaming paths (i.e. gas diffusion media (GDM)) on transport phenomena of polymer electrolyte fuel cells (PEFCs) by structural and electrochemical approaches. First, structural analysis is conducted to figure out non-uniform behavior of GDM with bipolar plate channels under the stack clamping pressure. The deformation results enables the exact determination of flow properties of porous media (i.e. permeability and porosity). After recalculation process of flow paths, computational fuel cell analysis in the deformed regions is performed to visualize flow patterns and estimate the electrochemical performance of PEFCs. The results show that water transport in deformed GDM restricts the smooth transport of reactant gas leading to significant mass transport losses. Therefore, combined structural-fluidic analyses with electrochemical reaction provide insight into high-current density PEFC system.

2745

, , , , , and

Through-plane thermal conductivity and thickness variation under different compaction pressures were measured for a composite region of commercial gas diffusion layer (GDL), namely Freudenberg H1410 GDL, as received as well as saturated with a custom-made carbon ink/microporous layer (MPL). 3D X-Ray pictures were taken of the materials.

Thermal conductivity for the composite region is higher than for the GDL alone. The compressibility of the composite region is in the same order of magnitude as pure GDL material. The found results motivate further investigation of this composite region that has gotten very little attention in the literature.

A PEMFC consists of several components, i.e. the membrane electrolyte assembly (MEA) sandwiched between a thin MPL and a somewhat thicker GDL on each side. The MEA consists of a membrane (PEM) coated with catalyst layers (CL) on each side. MPL and GDL are often treated as separate layers in the literature. However, there exists a considerable interfacial region where the two different materials are intertwined. The fine material of the MPL can intrude considerably into the fiber structure of the GDL material. It is this region we are interested in.

The temperature difference across the PEMFC can reach several °C despite being less than a millimeter thick between the gas flow plates. Temperature differences arise mainly across the GDL. Several research efforts have led to a good understanding of the thermal conductivity of the GDL and how it changes with compression, temperature, PTFE content, different fabrics, and water content. This work presents an effort to determine the thermal conductivity as well as the compressibility of the aforementioned composite region. [2]

At a compaction pressure of 9.2 bar the thermal conductivity of untreated Freudenberg H1410 (114 μm), was found to be 0.111±0.009 W K-1 m-1 and for the custom-MPL-drenched Freudenberg H1410 material it was 0.124±0.005 W K-1 m-1. The untreated Freudenberg H1410 material was compacted to 87% of its original thickness at 9.2 bar compaction pressure. The MPL-treated H1410 material was compacted to 77% of its original thickness at 9.2 bar compaction pressure. The MPL-treated Freudenberg H1410 has a larger compression from atmospheric pressure to first compaction pressure, but a less steep compression gradient with rising compaction pressure than the original H1410 material, see figure.

The thermal conductivity at 9.2 bar compaction pressure of another untreated Freudenberg GDL, namely H2315 (182 μm), was reported earlier to be 0.15±0.02 W K-1 m-1, roughly 50% higher than for the H1410 we measured. This complies with a trend seen in Toray paper, where the thermal conductivity at 9.3 bar compaction increases from 0.53±0.03 W K-1 m-1 for Toray TGP-H-060 (165 μm) to 0.65±0.02 W K-1 m-1 for Toray TGP-H-090 (265 μm) and to 0.81±0.03 W K-1 m-1for Toray TGP-H-120 (333 μm). [1]

SIGRACET GDL 10 AA, also an untreated GDL, was measured to have a thermal conductivity of 0.38±0.03 W K-1 m-1 at 9.2 bar compaction pressure, also much higher than the Freudenberg GDL. A SolviCore PTL had a very similar thermal conductivity of 0.36±0.08 W K-1 m-1 at 9.2 bar compaction pressure. [1]

The way the GDL is produced, the Teflon content, and the binder content all have influence on the final value of the thermal conductivity for a GDL material. Toray paper has a higher thermal conductivity when the thickness increases. An increase in PTFE content will lower the thermal conductivity. In general, Freudenberg GDLs have a lower thermal conductivity than other commercially available GDLs.

By depositing MPL material into the GDL, thermal contact between GDL fibers is enhanced, hence the overall thermal conductivity of the material increases. This could be a means to improve thermal conductivity of GDL materials with low thermal conductivity to achieve better temperature management in the fuel cell.

Acknowledgment

The Advanced Light Source is supported by the Director, Office of Science, Office of Basic Energy Sciences, of the U.S. Department of Energy under Contract No. DE-AC02-05CH11231. The authors would like to acknowledge assistance of Dula Parkinson with the experimental set-up at the beamline.

References

[1] O. Burheim, J. Pharoah, H. Lampert, P. Vie, S. Kjelstrup, "Through-Plane Thermal Conductivity of PEMFC Porous Transport Layers", J. FC Sci. Techn 8 (2011) 021013 1-13.

[2] O. Burheim, G. Crymble, R. Bock, N. Hussain, S. Pasupathi, A. du Plessis, S. le Roux, F. Seland, H. Su, B. Pollet, "Thermal conductivity in the three layered regions of micro porous layer coated porous transport layers for the PEM fuel cell", Int J. Hydr Energy. 40 (2015) 16775–16785.

Figure 1

2746

, , and

 Proper water management in proton exchange membrane fuel cell (PEMFC) is critical to achieve a high power density and durability of construction materials. The gas transport is blocked by liquid water remained in GDL, and it causes performance decrement with increase of concentration overpotential. The location of water accumulation is influenced by the temperature distribution inside GDL. At the same time, the temperature distribution is affected by the liquid water distribution, because the liquid water inside GDL decrease the effective thermal resistance. Then, the objective of this study was set to clear the mutual interaction between the liquid water distributions and temperature fields. For this purpose, temperature at the CL surface was measured using thin film thermocouple fabricated by MEMS based technology, and liquid water inside the GDL was visualized by X-ray CT.

The experimental conditions were shown in the table. The GDL thermal property, especially the thickness, was changed intentionally to get symmetrical temperature distributions, and the influence of the GDL thermal property on the liquid water distributions and temperature have been examined.

The thickness of the sensor was only about 9µm and the width is about 100µm. The electromotive force of the MEMS based sensor was confirmed to be proportional to the temperature difference. The MEMS sensor was inserted at the boundary between MPL and CL. The temperature rise from OCV was plotted in the bottom left figure. The temperature increased quadratically except for the high current density region of condition 4.

The visualized liquid water distributions were shown in right hand side figures. The accumulated water was observed in cathode side GDLs in all conditions, however, the highest saturation of water was observed near the rib surface of condition 4. The accumulated water at the rib surface was supposed to reduce the contact thermal resistance and decrease the temperature at the CL surface.

Figure 1

2747

, , and

In polymer electrolyte fuels (PEFC) carbon fiber based gas diffusion layers (GDL) are used for fine distribution of reactant gases from the flow field channels to the catalyst layer (CL). It is a key component for PEFC as it serves also for water removal from the CL as well as heat and electricity transport between CL and the flow field plates. GDL materials can be classified as woven, felts, papers, where the later type requires a binder to mechanically stabilize the fiber arrangement.

While most GDL binders are solid, the binder of SGL GDLs appears porous and it's influence on gas phase transport is unclear [1]. In SEM surface images of SGL, GDL binder domains of appear as flake like fine-structure with particles of about 100 to 300 nm (see Fig. 1a). Here, we apply multi-dimensional X-ray imaging techniques and multi-scale numerical simulations to clarify possible contributions of the binder porosity to overall GDL gas transport.

Absorption contrast X-ray tomographic microscopy measurements at the TOMCAT beamline of the Swiss Light Source (SLS) with voxel sizes of minimum 0.16 µm revealed a high porosity (59 %) of SGL 24 BA binder domains of (Fig. 1b). Computational transport simulations on the segmented binder structures showed, that the binder domains provide isotropic effective diffusivity of 0.30. In order to quantify the contribution of the porous binder to the overall fluid transport in the GDL, the obtained binder diffusivity values were used in multi-scale transport simulations of ternary segmented XTM data of representative GDL samples with 2.2 µm voxel size. If the binder was treated as solid, through-plane effective diffusivity of the GDL of 0.1 was obtained. If we applied a multi-scale simulation approach, that treats the binder domain as porous and uses the transport parameters of the binder domain as input values for the coarse scale simulations of the whole GDL domain, the through-plane effective diffusivity increased to 0.4 (+300 %). These numbers will be verified by X-ray ptychograptic computed tomography measurements of the binder fine structure at the cSAXS beamline of the SLS that proved resolutions of up to 16 nm [2].

References

[1]Rashapov, A; Gostick, J.T., In-Plane Effective Diffusivity in PEMFC Gas Diffusion Layers, Transport in Porous Media, 2016, In press.

[2]Holler, M.; Diaz, A.; Guizar-Sicairos, M.; Karvinen, P.; Färm, E.; Härkönen, E.; Ritala,M.; Menzel, A.; Raabe, J. & Bunk, O., X-ray ptychographic computed tomography at 16 nm isotropic 3D-resolution, Sci. Rep., 2014, 4, 3857.

Figure 1: SGL 24 binder fine structure a) SEM image and b) X-ray tomographic microscopy with 0.16 µm voxel size.

Figure 1

2748

, and

Polymer electrolyte fuel cells (PEFCs) generally have external humidifiers to supply humidified fuel and oxidant gases, preventing dehydration of the membrane electrode assembly (MEA). However, if a PEFC could be operated without humidification, these external humidifiers could be removed, resulting in a simplified PEFC system with increased overall efficiency and reduced cost. One of the most important goals related to advancing the commercial viability of PEFCs has been the development of a high-performance PEFC that can operate without humidification. The humidification requirements of the anode and cathode are different. At the anode, humidified pure hydrogen gas that is not used for the electrochemical reaction can be recirculated, and it is therefore possible to remove an external humidifier. At the cathode, air is exhausted without recirculation and so fresh humidified air is supplied using an external humidifier. It is therefore important to find a method to enhance the performance of a PEFC that does not require humidification at the cathode.

The present study was carried out to clarify the effect of anode gas recirculation on PEFC performance without humidification. The cell temperature was set at 75°C. Air utilization was set to 60% and the hydrogen flow rate supplied from the hydrogen tank was set so as to maintain a hydrogen utilization of 100%. The anode gas recirculation flow rate was varied between 0 and 100 cm3 min-1. The relative humidity (RH) of the cathode inlet gas was set at 0%. The RH was maintained at a very low value of 30% at the anode. The gas diffusion layer (GDL) used at the anode was a commercial SGL24BA GDL without a microporous layer (MPL). The MPL-coated GDLs were used at the cathode. The hydrophobic MPL consisted of an SGL24BA GDL coated with an MPL composed of 80 mass% carbon black and 20 mass% PTFE. The MPL containing hydrophilic carbon nanotubes (CNTs) was made from an SGL24BA GDL coated with an MPL composed of 4 mass% CNTs, 76 mass% carbon black and 20 mass% PTFE. The CNT surfaces were modified via an oxidation treatment. As a result, the contact angle of a CNT sheet was 30°, demonstrating that the CNT surfaces exhibited high hydrophilicity. Increasing the anode gas recirculation flow rate from 0 to 100 cm3 min-1 reduced the IR (ohmic) overpotential, which enhanced PEFC performance without humidification. Increasing the anode gas recirculation flow rate to more than 100 cm3 min-1 raised the gas velocity in the flow channel to over 2 m s-1. This is effective at promoting water transport from the anode gas to the MEA, thereby enhancing PEFC performance without humidification. In the case of the MPL without CNTs, decreasing the MPL mean flow pore diameter from 10 to 2 μm lowered the gas permeability. This effectively enhanced the ability of the MPL to prevent membrane dehydration, which improved PEFC performance without humidification. The PEFC performance obtained when employing the MPL with CNTs was higher than that for the MPL without CNTs. Because the pore diameter was constant at 2 μm both with and without CNTs, the difference in the gas permeability between these MPL-coated GDLs was negligible. Since the hydrophobicity of the MPL containing hydrophilic CNTs was reduced, its ability to retain humidity of the MEA was improved. This resulted in much higher PEFC performance compared with that obtained from the MPL without CNTs.

Even when a PEFC is operated without humidification, it is also essential to prevent flooding due to water produced in the cell under high current density conditions. Therefore, oxygen transport resistances were also assessed based on the limiting current density values of polarization curves to evaluate the ability of the MPL-coated GDL to reduce flooding under high humidity conditions. In the case of an MPL without CNTs, decreasing the MPL pore diameter to 2 μm increased the water breakthrough pressure. This degraded the ability of the MPL to prevent flooding under high humidity conditions. The appropriate MPL pore diameter for enhanced performance was different under no and high humidity conditions. In the case of an MPL with CNTs, it was possible to decrease the MPL pore diameter to 2 μm without lowering PEFC performance under high humidity conditions. An MPL with CNTs allows much higher performance under both no and high humidity conditions compared to that obtained with a hydrophobic MPL-coated GDL.

The authors would like to thank NITTA Corporation for supplying the hydrophilic CNTs used in this study.

2749

, , , and

Enhancement of the fuel cell performance at higher current densities is important to improve the power density and reduce the cost of proton exchange membrane fuel cell (PEMFC) system. Mass transport overpotential is the major barrier to achieving high performance at high current density. The overpotential at the cathode is significantly large and the oxygen partial pressure in the oxygen reduction reaction (ORR) area is vital. Condensed water in the flowfield and the gas diffusion layer (GDL) reduces oxygen transport to the ORR area. Direct investigation of oxygen transport has been limited by an inability to resolve water saturation dependent properties. A novel diagnostic method to analyze the boundary between flowfield and GDL surface is required. A measurement of the oxygen partial pressure in the flowfield while the fuel cell is operating was introduced [1-3]. This method uses oxygen sensitive fluorophore materials whose fluorescent luminescence is a function of oxygen quenching as described by the Stern-Volmer equation. In the previous work, the oxygen quenching rate was modified in order to enable in-situ oxygen partial pressure measurement on the surface of the cathode GDL with typical operating conditions of an automotive PEMFC [3]. In this work, a computational fluid dynamic (CFD) based two-phase fuel cell model was newly developed to be validated with measured oxygen fractions. Figure 1 shows the geometric configuration of this tested single cell. Figure 2 shows that simulation of oxygen partial pressure distribution shows similarity to the previous experimental data. Figure 3 shows measured oxygen partial pressure on the cathode GDL surface along with fraction of flowfield channel length. This oxygen fraction is deviated from the average stoichiometric line. The spatial variability of current density, which is proportion to the oxygen consumption rate, is considered to be a large contributor. Combined empirical and modeling approaches enable to gain the mechanistic understanding of two-phase fluid flow in the flowfield and the surface of GDL, including water condensation and implication with the fuel cell operating conditions. Further analysis will be discussed.

 

References:

  • J. Inukai, et. al, Angewandte Chemie International Edition, 47(15):2792-2795

  • S. Hirano, et. al, ECS Trans. 58 (1) 1791-1797 (2013)

  • S. Hirano, et. al, ECS Trans. 2015 69(17): 1331-1339 (2015)

Figure 1

2750

, , and

Water transport within the gas diffusion layer (GDL) depends on the material's wettability. In a PEM fuel cell, hydrophobic substrates facilitate water transport at the cathodes under high current density operation. We report a new approach to quantify the relative, ex situ GDL wettability by measuring the solid- liquid (S-L) interfacial area. We applied electrochemical techniques to measure the double layer capacitance, and exploited the proportionality between this quantity and the S-L interfacial area. We combined this electrochemical approach with traditional techniques (e.g., contact angle, and capillary pressure) to characterize commercial GDL materials. The measurements demonstrate that the fibre structure and morphology affect the wetting behaviour of the GDL materials. We describe the experimental approach and the underlying physical mechanisms, as well as the method's applicability to characterize water intrusion processes for GDL materials.

2751

, and

To improve the performance of a polymer electrolyte fuel cell (PEFC), it is essential to control liquid water behavior in the cell. Gas diffusion layer (GDL) plays an important role for the water transport, and optimization of the structure is one of the major research topics. As the structure of GDL is quite complex and small with the scale order of 200 mm thickness, measurement of liquid water behavior is difficult, and only a few results are reported with X ray measurement. In this study a scale model experiment is proposed to observe the behavior in a model GDL made with a 3D printer in the scale of 300 times of the actual one.

The parameters relating to the liquid water motion in a complex fiber structure are Capillary number, Weber number and Bond number. As the pore scale in the GDL is quite small and the speed of the water is very slow, it was assumed that capillary force is the dominant factor for the motion and other effects such as viscosity, inertial and gravity can be ignored. This gives that the behavior in the GDL is organized by Capillary number, Ca = /σ, where, u is velocity, μ is viscosity and σ is interface tension. It represents the ratio of viscous to capillary forces.

By using silicone oil as water and water as air, the scale model can be magnified to 300 times of actual size. Gravitational effect can be also removed, as the densities of the two liquids are similar. The scale model was made by 3D printer from the structural data obtained by X ray CT images of actual GDL.

Figure 1 shows the GDL made by 3D printer. The length is 25mm ´ 25 mm ´ 45 mm. The bottom of the model faces to a flat plate with a hole assuming a crack of micro porous layer (MPL). Silicone oil representing water is dyed in red. Viscosity ratio of silicone oil to liquid water is 55, which is same as the ratio of air to liquid water. Inter face tension is σ = 3.1´10-2 N/m, equivalent order of water and air. The contact angle of the material and the silicone oil was about 130°, which is the same order with the actual GDL condition. The GDL structure is set in the box made of acrylic and the GDL is immersed in liquid water representing air.

Figure 2 shows comparison of liquid water behavior in the scale model and they are compared to the results simulated by Lattice Boltzmann Method (LBM). Length of the simulation area is 80 μm ´ 80 μm ´ 160μm. The interfacial (surface) tension is σ = 7.29´10-2 N/m. Viscosity ratio of liquid phase to gas phase is 55 and the contact angle is set 130° in the simulation. In each case Capillary number, Ca, is adjusted to ~10-2 by inlet velocity. The developments of liquid water in the fibers are quite similar between the scale model and the simulation. Particularly the selection of pores in different scales in the water motion is very similar. Variety of observations in different Ca conditions showed good similarity between the model and the simulation. These results indicate the possibility of the scale modeling.

Investigation was made for the water behavior for the wide range of different Capillary numbers. Figure 3 shows pictures with different capillary numbers. In Ca ~10-1 liquid phase immerses small pores in bottom region, whereas the small pores are not filled by the water in Ca less than 10-2. This is due to the fact that viscous and inertial forces become large relative to the capillary force when Capillary number is large, and the capillary force becomes dominant when Capillary number is small. In less than Ca ~10-2, the phenomena were almost same and independent to the Capillary numbers. In this study the characteristics of liquid water behavior was investigated intensively for wide range of Capillary numbers.

Using the scale model experiment, optimization of the fiber structures was attempted by comparing the results with LBM simulation. Water motion in the vicinity of libs and channels facing to the GDL was also investigated. Some of the results will be presented to show clues for the optimization of the GDL structure.

Figure 1

2752

, , and

Water flooding at cathode catalyst layer (CL) is a critical issue for high power density and long-term operation of polymer electrolyte fuel cells (PEFCs). Especially, at high current densities, excessive water produced by oxygen reduction reaction (ORR) on cathode side is rapidly condensed and accumulated inside porous electrode. When open pores in CL and gas diffusion layer (GDL) are filled with liquid water, oxygen cannot be sufficiently fed to reaction sites. To alleviate this issue, it is necessary to design the optimum channel/GDL structure for accelerating through-plane water removal. Turhan et al. [1] investigated the through-plane liquid storage, transport and flooding mechanism as a function of channel wall wettability with the use of high-resolution neutron imaging. Results revealed that hydrophilization of cathode channel forms liquid film layers around channel walls, which is difficult to purge. However, hydrophilic channel effectively enhances liquid water suction from under-land locations into gas channels. Nishida et al. [2] also demonstrated that the through-plane water transport from GDL to channel is encouraged by channel hydrophilization, and the voltage drop due to flooding is reduced. On the other hand, several researchers presented a concept of perforated GDL structure to secure sufficient water passages through porous electrode. Gerteisen et al. [3] developed the customized GDL which is structured with water transport pathways by laser perforation, and revealed that this modified GDL improves the limiting current density of 8-22%. The perforated GDL structure beneficially enhances in-plane water discharge from porous media toward large penetration holes. However, the laser perforations may act as water pooling locations under high-current and high-humidity conditions because of the removal of PTFE coating, leading to performance degradation [4].

This study proposes the novel channel/GDL combinational structure of channel hydrophilization and GDL perforation as shown in Fig. 1, in order to promote water removal through cathode electrode.Liquid water accumulated at the CL is discharged into the large penetration hole. Subsequently, the water droplets gathered in the hole gradually grow up and reach the hydrophilic channel. When these droplets are attached to the channel sidewall, they are immediately spread out on the hydrophilic surface and liquid films are moved upward along the sidewall. This water suction through the penetration hole effectively alleviates water flooding near the cathode CL. In this experiment, the liquid droplet behavior inside the cathode channel of an operating PEFC is firstly observed based on a cross-sectional visualization technique, and the effect of hydrophilic treatment of cathode channel on the enhancement of through-plane water transport and the performance improvement is investigated. Secondly, the impact of combinational structure of channel hydrophilization and GDL perforation on the reduction of water flooding is demonstrated under high-current and high-humidity conditions. The GDL perforation is carried by two different methods of electric discharge machining (EDM) and manual micro-drilling technique. The EDM method has the disadvantage of removing the PTFE coating of GDL, resulting in local water flooding. In contrast, the micro-drilling technique does not damage the hydrophobic treatment of GDL.

Fig. 2 shows the changes of cell voltage during startup for two different channel/GDL structures. The red line denotes the result for the untreated channel/GDL structure; the blue line denotes the combinational structure with hydrophilized channel and perforated GDL. The effective electrode area of the cell is 2.88 cm2. The width, depth and total length of the gas channel are 1.0, 1.0 and 88.5 mm. The contact angle of water of the hydrophilized channel is 17 deg. In the modified cell, 35 penetration holes (Hole diameter: 0.3 mm) are installed into the cathode electrode by the micro-drilling method. The fuel cell operation is performed at 0.73 A/cm2 for 1000 s. The inlet gas temperature and humidity are 45 deg. C and 80% RH, respectively. As shown in the figure, in the case of the untreated structure, the cell voltage drastically drops after starting due to water flooding. On the other hand, the modified channel/GDL structure gradually recovers the voltage after t=150 s, because liquid water is removed from the cathode CL.

References:

[1] A. Turhan, et al., Electrochim. Acta, 55, 2734 (2010).

[2] K. Nishida, et al., ECS Trans., 69(17), 1121 (2015).

[3] D. Gerteisen, et al., J. Power Sources, 177, 348 (2008).

[4] M.P. Manahan, et al., J. Power Sources, 196, 5573 (2011).

Figure 1

2753

, , , , , , , , and

The accumulation of liquid water formed by the electrochemical reaction of hydrogen and oxygen in a polymer electrolyte membrane (PEM) fuel cell causes a gas diffusion layer (GDL) to become effectively less porous and more tortuous. To achieve designs directed for water and gas transport in PEMFC diffusion media, a nuanced understanding of the nature and extent of GDL water as an impedance to oxygen flux is required.

In this work, a limiting current approach was used to measure the mass transport resistance in a custom PEM fuel cell. Using a small active area at high stoichiometric ratios, it was assumed that the concentration of gases was constant along the length of the reactant flow channels, creating a uniform concentration gradient across the entire diffusion medium. If the fuel cells are then operated at the limiting current density, where cell potential approaches zero, it can also be assumed that all the reactants are being consumed and the concentration of reactants at the reaction site is effectively zero. Consequently, oxygen transport resistance in these conditions can be inferred directly from the limiting current density using the following equation, where RT is the transport resistance in s/cm, c0 is the oxygen concentration at the reactant channel and iL is the limiting current: 

RT = 4Fc0/iL

Previously, through a thorough set of limiting current experiments at a range of operating conditions, Baker et al.1 showed that the carbon-fiber diffusion medium is the dominant contributor to oxygen transport resistance. Owejan et al.2 combined limiting current methods with in situ neutron radiography to demonstrate a strong qualitative correlation between GDL saturation and oxygen transport resistance.

Synchrotron radiography has been established as a powerful tool for visualizing water distribution in PEMFC diffusion media3. In this work, synchrotron X-ray radiographs were obtained with a pixel size of 6.5 µm with a spatial resolution of 10 μm.

GDL materials were imaged in operando at limiting current over a range of relative humidities (set in the inlet cathode gas stream). Figure 1 (a) is an example radiograph of a cell operating at limiting current density processed according to the Beer-Lambert law5,6. A distinct in-plane variation in liquid water distributions is visible between regions over ribs and regions over channels (Figure 1 (b)). This work exploits the high spatial resolution for capturing liquid water accumulation trends, information which is essential for building towards a direct, quantitative empirical relation between oxygen transport resistance and GDL saturation. 

References:

1. Baker DR, Caulk DA, Neyerlin KC, Murphy MW. Measurement of oxygen transport resistance in PEM fuel cells by limiting current methods. J Electrochem Soc. 2009;156(9):B991-B1003.

2. Owejan JP, Trabold TA, Mench MM. Oxygen transport resistance correlated to liquid water saturation in the gas diffusion layer of PEM fuel cells. Int J Heat Mass Transfer. 2014;71(0):585-592.

3. Lee J, Hinebaugh J, Bazylak A. Synchrotron X-ray radiographic investigations of liquid water transport behavior in a PEMFC with MPL-coated GDLs. J Power Sources. 2013;227:123-130.

4. Lee J, Yip R, Antonacci P, Ge N, Kotaka T, Tabuchi Y. Synchrotron investigation of microporous layter thickness on liquid water distribution in a PEM fuel cell. J Electrochem Soc. 2015;162(7):F669--676.

5. Manke I, Hartnig C, Gruenerbel M, et al. Investigation of water evolution and transport in fuel cells with high resolution synchrotron x-ray radiography. Appl Phys Lett. 2007;90(17):174105.

6. Hartnig C, Manke I, Kardjilo N, et al. Combined neutron radiography and locally resolved current density measurements of operating PEM fuel cells. J Power Sources. 2008;176(2):452-459.

Figure 1

C-31 Membrane Structure and Properties - Oct 6 2016 7:40AM

2754

The properties of ion conducting polymers (ionomers) used as proton conducting membranes are strongly dictated by their nanostructure. Due to the strong difference in polarity between the acidic sulfonic groups and the apolar main polymer chain, these materials exhibit strong phase separation. X-ray and neutron scattering are among the best experimental techniques to investigate the ionomer nanostructure. In particular, synchrotron based small and wide angle X-ray scattering (SAXS and WAXS) are particularly suited to perform in-situ investigation on ionomers.

In this contribution, I will show our most recent results obtained by using synchrotron SAXS and WAXS on the investigation of the structure-property relationships in different ionomers. The structural knowledge acquired on the benchmark Nafion membranes is compared to that of its short-side chain counterpart (Aquivion) and to other non-perfluorosulfonated ionomers like sulfonated poly(ether ether ketone)s and sulfonated polysulfones. The nano morphology of Nafion and Aquivion can be qualitatively and (almost) quantitatively explained by a locally flat morphology, describing the morphology as an alternated assembly of water/ions and polymeric 2D domains. The same morphological model can be applied successfully to other fluorine-free ionomers, suggesting that a local flat morphology could be common to many proton conducting ionomers.

In the second part of my talk, I will show some recent in-situ SAXS experiments on microstructural changes occurring in different ionomer membranes under controlled temperature and humidity conditions. The data suggest local re-distribution of water inside the membranes that can lead either to membrane swelling (low temperature and high humidity) or to local ionic structure collapse (high temperature and low humidity). These experiments mimic very closely what may happen in a real fuel cell and the understanding of the results is crucial to improve the membrane performances.

2755

and

The proton-exchange membrane (PEM) is one of the principal components for polymer electrolyte fuel cells. In the nanoscopic structure of the membrane, proton transport is one of the dominant factors that governs power generation efficiency, which are largely attributed to the nanoscopic structure of PEMs and water aggregations. Therefore, it is critical to understand the proton transport mechanisms through PEM and clarify an important link between the membrane nanostructure and the proton transport properties. As the phase separation is considered one of the primary factors affecting its performance, the morphological properties of PEM have been studied experimentally and been characterized by proposing the cluster models (e.g., the cylinder model and the lamellar model) which reasonably fit the experimental scattering spectra. However, a detailed relationship between the morphological features and proton transport properties is still under debate. Therefore, in this study, reactive molecular dynamics simulations have been performed to study the effects of water cluster structure on proton transport properties by constructing the cluster models in the simulations. The anharmonic two-state empirical valence bond (aTS-EVB) model has been used to incorporate excess proton transport efficiently through the Grotthuss hopping mechanism, providing an accurate estimation of proton transport properties in multiproton environments within the simplicity of the theoretical framework. The proton transport properties have been estimated in terms of diffusion coefficient and proton distributions in the two hydrophilic cluster structures (i.e., the cylinder model and the lamellar model) that are the most typical proposed morphological models in PEMs. The cylindrical systems with radii from 0.5 nm to 1.7 nm and the lamellar systems with the thicknesses from 0.6 nm to 1.6 nm were constructed for the purpose of comparison. The water contents for all of the models were kept at λ = 7, where the parameter λ indicates the ratio of the number of water molecules to that of SO3-, which is equivalent to ~10 wt %, typical operating conditions in real PEM fuel cells. To construct the desired water cluster models in the PEM system, water domains with the desired geometric shapes were obtained from equilibrated simulations of bulk water. The systems were built by first placing the water domains into an empty box and then growing polymers around the water. The sulfonate groups were arranged so that they were in close contact with the surfaces of the water domains. The simulation cell for each model was then equilibrated while the positions of water molecules and hydronium ions were constrained. In the production run, a carbon atom in the backbone of each monomer was fixed to maintain the proper geometric shape of the water domain, while still allowing reasonable fluctuations of the sulfonate groups. All other atoms in the system were allowed to relax. The diffusion coefficients in each dimension are calculated and correlated with the cluster size and the type of cluster models. It is found that the proton diffusion shows the peaks in the 0.8 nm radius of the cylinder model and in the 0.9 nm thickness of the lamellar model, and decreases with increasing the cluster size. Nevertheless, the proton diffusions at any cluster size calculated in this study show the higher value than that in the random PEM model using the aTS-EVB method (~0.23×10-5 cm2/s), suggesting that the proton can be transferred effectively in the non-random cluster models. In the cylinder model with 0.8 nm radius, where the proton shows the higher diffusion, a comparatively large number of protons that are farther than the radius of the solvation shell are found in the center of cylinder, and thus the protons are freer to explore the free water region. Similarly, the more protons distribute in the surface of the lamellar model at larger thickness because the surface density of the sulfonate groups increases at larger thickness of the lamellar model. In the lamellar model with 0.9 nm thickness protons show the higher distributions in the center of the lamellar model, suggesting that protons are able to cross from interface to interface and are less hindered by the sulfonate groups, resulting in the faster proton diffusion. In the larger size of clusters, the protons are trapped in the interface of the water domains because the surface density of the sulfonate groups becomes higher at larger cluster sizes, resulting in the slower proton diffusion. Our simulation results provide insight into quantitative information about the water cluster structure dependence of the proton transport properties at an atomic level.

2756

, , , and

Molecular-level understanding of dynamic behaviors of complex mass transports, especially proton and water- in polymer electrolytes is essential to design better membrane materials used for polymer electrolyte fuel cells (PEFCs). NMR is one of the suitable experimental techniques for analyzing such molecular dynamics, since both complex static structures and dynamical motions of molecules can be tracked by high-resolution NMR spectra and a pulsed field gradient NMR (PFG-NMR). This technique is widely used to understand proton dynamics however, it cannot clearly distinguish vehicle and grotthuss mechanisms.

Recently, we have developed a combination method of 2H, 17O MAS NMR and 1H PFG NMR measurement to overcome the drawback. We applied the technique to an organic-inorganic composite membrane, Zr-SPP-SPES, which forms controlled configuration of proton functional groups [1,2]. In the paper we proposed enhanced grottuss mechanism in a water-freezing temperature region called "packed acid mechanism". [6]

In the present study we prepared perfluorinated sulfonic acid membranes with different equivalent weight (EW) 500, 600, 700 and 900, to understand relationship between value of ion exchange capacity (IEC = 1000/EW) and proton conduction mechanism in water freezing region. 17O labeled water was used in order to distinguish signals from non-exchangeable protons in the membranes, thus natural abundance 17O (0.04%) NMR signals of water in membrane cannot be detected.

1H and 17O MAS NMR measurements were carried out on Agilent NMR systems 400WB at 9.4 T. A 4 mm diameter air tight rotor was used for sample spin at 12 kHz. 1H and 17O 1D-NMR spectra were obtained by using a single pulse sequence. Sample temperature was calibrated by measuring the 79Br signal of KBr powder under the same measurement condisions as described elsewhere. [3] 1H PFG NMR measurements were carried out on JNM LA-400 at 9.4 T. NMR spectrometer equipped with maximum gradient strength of 12 T/m. The diffusion coefficient (D) was measured by using a pulsed-field-gradient stimulated-echo pulse sequence. [4, 5]

 17O NMR spectrum in all the membrane show a sharp peak around 0ppm, from temperature 6 to 42℃. With lowering temperature from -20 to -45 ℃, the width of the spectra of all the membranes is drastically increased then finally the spectrum is vanished at T = -45 ℃. The behavior of 17O NMR in all samples were almost similar. Since the width of the NMR spectrum is inversely proportional to the movement of the observed nucleus, it indicates that absorbed water molecules are completely freezing.

In contrast to the 17O NMR, solid 1H NMR spectra show a sharp peak even at the low temperature range, from -25 to -45 ℃. This result indicates that proton conductivity is still remaining, even while absorbed water molecules are freezing.

1H PFG-NMR also revealed that proton conduction is present at low temperature as well as the results of 1H NMR. The sample with highest IEC value ( IEC ~ 2 ) shows highest proton conductivity D = 7 x 10-7 (m2/s) at T = -25 ℃ and D = 0.5 x 10-7 (m2/s) at T = -60 ℃.

These three results indicate that protons in the membranes with high IEC move independently like a "packed acid mechanism" in freezing temperature region. This conduction mechanism would happen efficiently because sulfonic acid is tightly-packed, which is a noteworthy property of the high IEC membranes.

 

References:

[1] G. M. Anilkumar, S. Nakazawa, T. Okubo, T. Yamaguchi, Electrochem. Commun., 8 (2006) 133.

[2] J. M. Lee, H. Ohashi, T. Ito, T. Yamaguchi, J.Chem. Eng. Japan,42(2009) 918.

[3] K. R. Thurber, R. Tycko, J. Magn. Reson., 196 (2009) 84.

[4] O. E. Stejskal, E. J. Tanner, J. Chem. Phys., 42 (1965)288.

[5] J. E. Tanner, J. Chem. Phys., 52 (1970) 2523.

[6] T. Ogawa, K. Kamiguchi, T. Tamaki, H. Imai and T. Yamaguchi, Anal. Chem., 86 (2014) 9362.

 

2757

, , , , , , , , , et al

Nafion has been well known as the standard perfluorosulfonated material in the field of Polymer Electrolyte Membrane Fuel Cell and has been irreplaceable materials as a membrane and an ionomer of catalyst layer. Many researchers have investigated the proton conduction, water-uptake, mechanical property and nano-structure1,2).As for nano-structure researches, small angle X-ray scattering and small angle neutron scattering revealed the morphology characterization of Nafion solution3). X-ray irradiates many molecules in Nafion solution.The characterization was attributed to average information.TEM is very useful apparatus to observe nano-structure of materials. It is applicable to samples which withstand a vacuum condition.Nafion solution is not able to be observed by TEM because of high vapor pressure of solution. Freeze method is an idea to measure Nafion solution. In case of freeze method using liquid nitrogen, vaporized nitrogen severely damaged the sample. In case of another freeze liquid ethane method Figure1, liquid ethane was cooled by liquid nitrogen to avoid vaporization of ethane.Figure 2 shows TEM photograph of Nafion solution. Objects of rod structure were observed, some rods seemed to be linked each other and dark points also observed.

Rod structure with radius 1.2 to 3.6 nm and length 16nm was predicted when entangle molecular weight 54904), EW=1000 and Nafion molecular weight between 0.1million and 1million were applied.Rod structure of TEM photograph is interpreted as the prediction.Dark points of TEM Photograph might be nano-crystals of PTFE part of Nafion.SAXS measurement of Nafion solution also showed nanocrystal with interval of 18nm. This interval value fairly fits to distances between dark points.

This presentation is based on results obtained from a project commissioned by the New Energy and Industrial Technology Development Organization (NEDO).

References

1)Kenneth A. Mauritz and Robert B. Moore, "State of Understanding of Nafion", Chem. Rev. 2004, 104, 4535-4585

2)Adam Z. Weber and John Newman," Modeling Transport in Polymer-Electrolyte Fuel Cells", Chem. Rev. 2004, 104, 4679-4726

3)Makoto Yamaguchi, Takuro Matsunaga, Kazuki Amemiya, Akihiro Ohira, Naoki Hasegawa, Kazuhiko Shinohara, Masaki Ando and Toshihiko Yoshida, "Dispersion of Rod-like Particles of Nafion in Salt-Free Water/1-Propanol and Water/Ethanol Solutions", J. Phys. Chem. B 2014, 118, 14922-14928

4)Souheng Wu; "Characterization of Polymer Molecular Weight Distribution by Transient Viscoelasticity: Polytetrafluoroethylenes", Polymer Engineering & Science, 1988 - Wiley Online Library

Figure 1

2758

, , , and

In the construction of polymer electrolyte membrane fuel cells (PEMFCs) and polymer electrolyte membrane electrolyzers (PEMELs), Nafion® (sulfonated tetrafluorethylene) is used as a membrane material and as proton conductor in the porous catalysts layers of membrane electrode assemblies (MEAs). Efficiency can be increased and costs reduced by optimizing and controlling the porous catalyst layer in MEAs. It is important to understand the polymer properties, which can be achieved by using different solutions and varying concentrations of one mixture for manufacturing the dispersion. Furthermore, the findings are of particular importance for the coating process. The efficiency of the PEMFCs and PEMELs depends on the number of active centers in the porous catalyst layer. The formation of the triple phase boundary is of great importance. This can be ensured by controlling the structural properties of the dispersion. Thereby, the static and dynamic surface tensions are analyzed. Surface tension and interfacial tension are necessary for understanding the coating process and adhesion to a transfer layer. The complex system was simplified and research on the surface properties performed with an aqueous Nafion® solution in an alcohol concentration series. Hence, the only aspect that changed was the alcohol, while the dynamic surface tension was measured with a bubble pressuretensiometer (BP100, KRÜSS GmbH). The static surface tension and interfacial surface tension were carried out with a force tensiometer (K100, KRÜSS GmbH) after Wilhelmy plate respectively the duNouy Ring method. In addition the solution was analyzed by means of IR- and 19F-NMR- spectroscopy.

It will be demonstrated in this contribution that the static and dynamic surface tensions differ significantly at lower alcohol concentrations in solution series with and without Nafion®. While at higher alcohol concentrations the static surface tension between both series is similar, the dynamic surface tension shows a different trend with varied bubble age. This can be explained by the interaction between alcohol and Nafion®. In accordance with the previous interface outcomes, the 19F-NMR- and IR- spectra are analyzed. According to the first results, it is also apparent that there are interactions between Nafion® and alcohol. The application of these methods provides an understanding of the interaction between the molecules in the catalyst dispersions. This information is essential for understanding the interactions between the components of catalyst dispersion during electrode manufacturing.

2759

, , , , , , , and

Perfluorosulfonate ionomers such as Nafion consist of a polytetrafluoroethylene backbone having pendant side chains terminating with sulfonic acid groups. These materials have attracted much attention as proton-conducting materials in polymer electrolyte membrane fuel cells (PEMFCs), due to their excellent chemical stability and high proton conductivity. The catalyst layer is fabricated by deposition of catalyst ink containing catalyst, ionomer, and solvents. In the catalyst layers, ionomer provided an ultra-thin film, coating the catalyst and the catalyst support surfaces. These ultra-thin films have a significant influence on the electrochemical activity and transport phenomena that determine the whole cell performances. Recently, precise structural analyses of ionomer in the solutions and solid-state thin films have been reported. To the best of our knowledge, however, the dynamics of ionomer thin-film formation process remain unclear.

Herein, time-resolved measurements of grazing-incidence small- and wide-angle X-ray scattering (GISAXS/GIWAXS) utilizing synchrotron radiation at BL45XU/SPring-8 were carried out to observe the nanostructural evolution of thin film during solidification processes (i.e., spin-coating) from the Nafion solution. GISAXS and GIWAXS are synchrotron-based X-ray techniques that have been used extensively to provide nanostructural insight of polymeric thin films. In the present study, effects of solvent evaporation and centrifugal force on the structural formation behaviors of Nafion were observed for water/1-propranol (NPA) solution of Nafion (3.5 wt% Nafion in water/NPA (1/1 in volume ratio)). The wavelength of the X-ray beam was 0.1 nm, and camera length was 270 mm and 2503 mm for GIWAXS and GISAXS measurements, respectively. The scattering intensity at a certain scattering vector q was measured as a function of time after deposition of Nafion solution on the rotated Si substrate (2000 rpm). The 2D-scattering image was acquired using a photon counting detector (PILATUS3X 2M, DECTRIS Ltd.). In the first stage from t = 0 to 20 s, the 2D GIWAXS profile showed strong solvent peak and crystalline peak arising from the packing of the Nafion backbone (qz ~ 11.7 nm-1, d-spacing is defined as 2π/qz, d ~ 0.5 nm) quickly appeared. After 180 s, the solvent peak disappeared, and only the crystalline peak was observed. This supports that fine crystallites were contained in Nafion solution. The 2D GISAXS profile, on the other hand, significantly changed with time. Selected GISAXS patterns during thin-film formation process are shown in Fig. 1. In the first stage (from t = 0 to 110 s), an interaction peak was observed at q > 0.1 nm-1, the position of which is related to the main first neighbor distance D between the scatter structures in Nafion solution, and the peak position shifted towards larger q values with time. This indicates that solvent evaporation decreases the distance between the scatter structures in Nafion solution. At the second stage (from t = 110 to 135 s), the new peak of qz ~ 1.9 nm-1 (d ~ 3.3 nm) corresponding to so-called ionomer peak appeared. At the third stage (form t = 135 to 180 s), the interaction peak disappeared, and only the peak of qz ~ 1.9 nm-1 was observed, indicating that the ordered structure were fixed in solid-state thin film after the solvent evaporation. This work clearly indicates that second-scale nanostructural evolution of Nafion during the film forming process can be traced by utilizing synchrotron GISAXS/GIWAXS measurements. Further investigations on the effects of process parameters during film forming with a film applicator on nanostructural formation behaviors of Nafion are now in progress.

This presentation is based on results obtained from the PEMFC Research and Development Program for "Highly‐Coupled Analysis of Phenomena in MEA and its Constituents and Evaluation of Cell Performance" commissioned by the New Energy and Industrial Technology Development Organization (NEDO). The synchrotron radiation experiments were performed at BL45XU in SPring-8 with the approval of JASRI (Proposal No.2015B1105).

Figure 1

2760

, , , , and

Proton exchange membrane (PEM)-based fuel cells are being aggressively targeted as an alternative energy source for transportation and stationary power due to their high charge density and low operating temperatures. While the structure, transport and mechanical properties of bulk PEMs have been extensively studied, there is increasing interest on the behavior of these materials at interfaces and under confinement, since there exist many heterogeneous interfaces within an actual fuel cell assembly. One could argue that the performance and ultimate failure of the membrane electrode assembly is greatly impacted by the structure and transport of ionomers confined to extremely small length-scales.

At the heart of the PEM fuel cell is the polymer membrane that must transport protons but not electrons, while also acting as a barrier to the reactant gases (H2 and O2). The most widely studied material for this purpose is perfluorosulfonic acid (PFSA) ionomers such as Nafion. Over its service life, the polymer membrane undergoes hundreds if not thousands of cycles between wet and dry conditions, which leads to swelling and shrinking of the membrane material. This can ultimately lead to mechanical failure in the membrane. This mechanical fatigue is of particular importance in the catalyst layer where the PFSA ionomers, often used as an ionically conductive binder, are confined in domains on the order of tens of nanometers in thickness. A further understanding of the swelling-induced stresses experienced in both the catalyst layer as well as the PEM is necessary.

In this talk, we will describe our efforts to develop and implement a cantilever bending technique1 in order to investigate the swelling induced stresses in Nafion thin films. By monitoring the deflection of a cantilever beam coated with a thin Nafion film as it is exposed to varying humidity environments, the swelling induced stresses of the polymer film can be measured. Using this technique, we have measured the mechanical response to humidity cycling of thin Nafion films as a function of film thickness and thermal annealing, as well as hygrothermal aging.

  • K. A. Page, J. W. Shin, S. A. Eastman, B. W. Rowe, S. Kim, A. Kusoglu, K. G. Yager, and G. R. Stafford. In Situ Method for Measuring the Mechanical Properties of Nafion Thin Films during Hydration Cycles, ACS Applied Materials and Interfaces, 7, 17874-17883 (2015).

2761

, , , , and

We have previously shown1,2 that there is a suppression in both swelling and water diffusivity when Nafion is confined to thin and ultrathin films. We have also shown that the modulus of ultrathin Nafion films increases upon confinement3,4, as well as upon thermal annealing. Interestingly, we also observe a decrease in swelling and an increase in modulus for ultrathin films aged in high humidity environments. Similar behavior has been shown to occur in bulk Nafion films.5,6 Some researchers suggest that the root cause of such a decrease in swelling and increase in modulus is due to the formation of sulfonic anhydrides, while others suggest that it is due to internal hydrogen bonding of dehydrated sulfonic acids.

In this talk, we will present our measurements results on thin and ultrathin Nafion films. We apply polarization-modulation infrared reflection-absorption spectroscopy to study the kinetics of dehydration and formation of sulfonate species in thin Nafion films. We also employ cantilever bending during humidity-induced swelling to probe the mechanical properties of these films. Grazing-incidence small-angle x-ray scattering is used to elucidate changes in morphology upon annealing and/or aging. By combing the results from these techniques, we can paint a more holistic picture of the origins of suppressed swelling and increased mechanical properties in thin and ultrathin Nafion films.

  • S. A. Eastman, S. Kim, K. A. Page, B. W. Rowe, S. Kang, C. L. Soles, and K. G. Yager. Effect of Confinement on Structure, Water Solubility, and Water Transport in Nafion Thin Films, Macromolecules, 45, 7920-7930 (2016).

  • E. M. Davis, C. M. Stafford, and K. A. Page. Elucidating Water Transport Mechanisms in Nafion Thin Films, ACS Macro Letters, 3, 1029-1035 (2014).

  • K. A. Page, J. W. Shin, S. A. Eastman, B. W. Rowe, S. Kim, A. Kusoglu, K. G. Yager, and G. R. Stafford. In Situ Method for Measuring the Mechanical Properties of Nafion Thin Films during Hydration Cycles, ACS Applied Materials and Interfaces, 7, 17874-17883 (2015).

  • K. A. Page, A. Kusoglu, C. M. Stafford, S. Kim, R. J. Kline, and A. Z. Weber. Confinement-Driven Increase in Ionomer Thin-Film Modulus, Nano Letters, 14, 2299-2304 (2014).

  • F. M. Collette, C. Lorentz, G. Gebel, F. Thominette. Hygrothermal aging of Nafion, Journal of Membrane Science, 330, 21-29 (2009).

  • S. Shi, T. J. Dursch, C. Blake, R. Mukundan, R. L. Borup, A. Z. Weber, and A. Kusoglu. Impact of Hygrothermal Aging on Structure/Function Relationship of Perfluorosulfonic-Acid Membranes, Journal of Polymer Science B: Polymer Physics, 54, 570-581 (2016).

2762

, , , and

The structure and properties of the polymer electrolyte membrane has a major impact on the MEA performance and durability. Polymers with different chemical structures can propose different mechanisms for water and proton transport which lead to differences in performance and durability. For membranes with similar chemical structures, factors such as thickness and equivalent weight (EW) are crucial in transport and therefore could cause differences in the performance. The focus of this work was a closer look at the impact of these properties, and their changes due to degradation, on the performance and durability of the MEA and its components.

The impact of membrane thickness and equivalent weight was studied by comparing the water content of different PFSA membranes as a function of relative humidity (RH). The data was then used to further discuss the conductivity and hydrogen permeability of the membranes in different regimes of temperature and RH. The impact of membrane chemical degradation on transport property was studied by comparing the membrane conductivity, thickness, hydrogen permeability and performance before and after membrane degradation. To study the impact of membrane properties on cathode catalyst performance and degradation, the cathode degradation test was applied on MEAs with and without membrane degradation, and the results were compared.

The water content (Lambda) of all membranes showed a type II isotherm increase by increasing the relative humidity. Lambda was found to be independent of membrane thickness and equivalent weight, while, as expected, the absolute water content (g water/g membrane) showed a decrease at higher equivalent weights. Increasing the relative humidity and operational temperature showed an increase in membrane conductivity. Higher thickness or higher equivalent weight both resulted in higher membrane resistance; supporting a more general trend which is the linear increase of membrane conductivity with increase in the absolute water content.

Cathode degradation testing showed an impact on the ECSA loss and performance. The effect was more intense for thinner or lower EW membranes suggesting that membrane water content was the contributing factor. After the membrane degradation test, losses in membrane thickness, membrane conductivity, performance and ECSA were observed. In addition, the hydrogen crossover increased due to the membrane thinning. The membranes with pre-existing degradation showed less ECSA loss but increased performance losses. This trend was explained by the combination of higher catalyst layer ionomer resistance and platinum depletion which lead to higher catalyst layer ionic losses related to a reaction distribution shift further into the catalyst layer.

Acknowledgements: Funding was provided by Department of Energy EERE Hydrogen and Fuel Cell Technology Program (Project DE-EE0006375).

2763

, , and

Perfluorosulfonic-acid (PFSA) ionomers play key roles in polymer-electrolyte fuel-cell (PEFCs), primarily as a proton-exchange membrane (PEM), and also as an electrolyte film in porous catalyst structures binding the catalytic agglomerates.1 Despite their wide adoption in PEFCs, most studies on PFSAs have thus far focused on Nafion, and studies on other PFSAs, in particular with varying side-chain chemistries and equivalent weights (EWs), have been relatively scarce. Therefore, it is of interest to elucidate how such changes in chemistry and EW influence the PFSA behavior, not only as a bulk membrane (i.e., PEM), but also as a catalyst-ionomer film (i.e., thin-film). In particular, the latter phenomenon could have important implications on explaining the mass-transport limitations in catalyst layers, which have been recently shown to be strongly related to the transport resistances at the ionomer thin-film.1,2

In this talk, the impact of EW and side-chain chemistry on structure/property relationship of various PFSAs will be explored as bulk membranes (> 20 μm) and dispersion-cast thin films (< 100 nm). PFSAs owe their combination of good transport properties and mechanical stability to their phase-separated morphology of conductive hydrophilic ionic domains and the hydrophobic backbone, which are connected via side-chains. It will be shown how this PFSA morphology is controlled by the fraction of backbone (via EW) and the side-chain length (via chemistry), which together govern the interrelation between transport and stability. Moreover, our investigations demonstrate the existence of universal correlations between the water-uptake and conductivity (chemical energies), and backbone crystallinity (mechanical energies), with a key role of the side-chain length in altering this balance between the chemical-mechanical energies. We will report the results of a systematic investigation on hydration, conductivity, mechanical properties and crystallinity of various types and EWs of PFSA ionomers to develop universal structure/function relationships. Lastly, it will be explored how these EW and side-chain effects influence the structure and properties of PFSA thin films when confined to nanometer thicknesses, and control their deviation from bulk membrane behavior. Our results provide new insights into the understanding and optimizing the PFSA ionomers' functionalities in fuel-cells, both as bulk membranes and as thin-films.

References

1. A. Z. Weber and A. Kusoglu, Journal of Materials Chemistry A, 2, 17207 (2014).

2. A. Kongkanand and M. F. Mathias, The Journal of Physical Chemistry Letters, 1127 (2016).

Acknowledgement

This work was funded under the Fuel Cell Performance and Durability Consortium (FC-PAD), by the Fuel Cell Technologies Office (FCTO), of the office of the Energy Efficiency and Renewable Energy (EERE), of the U. S. Department of Energy under contract number DE-AC02-05CH11231 and Program Development Manager Dimitrios Papageorgopoulos.

2764

, , and

Performance of polymer-electrolyte-fuel-cells (PEFCs) links strongly to thermodynamic and mass-transport properties of the polymer membrane. The quintessential PEFC-membrane material is perfluorosulfonic-acid (PFSA) polymer, which phase separates into nanoscale hydrophilic water-filled domains through which ions and water transport and into hydrophobic polymer-matrix domains that provide structural integrity and durability. Although considerable progress has been made to reduce reactant mass-transport limitations in PFSA membranes, mass-transport is still a critical area of concern. Aqueous electrolyte transport across a phase-separated membrane is inherently a multiscale problem with aqueous-dissolved species moving through nanoscale domains that are connected to form a transport network at the intermediate mesoscale. To guide the optimization of PEFC material design and operation, we decouple the impact of micro and macro length-scale on mass transport in PEFC membranes.

Specifically, molecular-scale interactions between aqueous-electrolyte and polymer in hydrophilic domains of PFSA membranes are modeled with mean-field local-density theory aligned with an experimentally consistent 3D domain geometry. Our molecular-scale description of membrane conductivity accounts for solvation, electrostatics, solvent dielectric saturation, finite size, and confinement effects. The microscale framework is validated against atomistic simulations and, subsequently, up-scaled to predict macroscale conductivity and solvent absorption by accounting for the interconnectedness of the hydrophilic domains. Excellent agreement is found with experimental data, as shown in Figure 1. Our proposed multiscale model for membrane conductivity and solvent sorption provides a new tool to explore potential avenues for improving PFSA membrane performance.

Acknowledgements

This work was funded by the Assistant Secretary for Energy Efficiency and Renewable Energy, Fuel Cell Technologies Office, of the U. S. Department of Energy under contract number DE-AC02-05CH11231 and by the National Science Foundation Graduate Research Fellowship under Grant No. DGE 1106400

Figure 1.3M PFSA membrane conductivity of proton-form membranes as a function of water content as determined from the model (open) and experiments (filled, with lines to guide the eyes) for equivalent weights of 1100 (circles), 1000 (squares), and 825 g/mol (SO3-) (diamonds)

Figure 1

2765

, , and

Perfluorinated sulfonic-acid (PFSA) polymers are a class of ion-conducting polymers (ionomers) heavily utilized in fuel cells and similar electrochemical technologies. The hydrophobic backbone gives excellent mechanical properties, while the side-chain terminated with a hydrophilic acid group bestows superior proton conductivity. The crystallinity of the backbone is believed to be inversely proportional to conductivity and proportional to the mechanical modulus in bulk membranes. Demonstrated previously, the transport and mechanical properties of PFSA ionomers deviate from bulk membranes as the ionomers are confined into thin-film geometries.[1]Under such confinement, conductivity and water uptake decrease, while the modulus increases. Studying this deviation from bulk behavior is crucial as ionomer is present as thin films in the catalyst layer of these electrochemical devices.

In this work, we correlate the relative crystallinity of PFSA thin-films, calculated from Grazing-incidence Wide-angle X-ray Scattering (GIWAXS), to their mechanical properties measured using a cantilever bending system.[2]Using a laser array, thin-film stress can be measured as a function of humidity from which mechanical properties are calculated. This system allows the study of thin films on hard impenetrable surfaces that induce confinement effects, and probe the effect of interfacial interactions on different substrates. Understanding crystallinity and mechanical properties as part of the structure-property relationship will allow optimization of ionomer performance within the catalyst layer.

Acknowledgement

This work was sponsored by the Army Research Office under MIPR0010754069. The work utilizes beamline 7.3.3 at the Advanced Light Source, which is funded by DOE Basic Energy Science under contract number DE-AC02-05CH11231.

References

1. Page, K.A., et al., Confinement-driven increase in ionomer thin-film modulus. Nano Lett, 2014. 14(5): p. 2299-304.

2. Page, K.A., et al., In Situ Method for Measuring the Mechanical Properties of Nafion Thin Films during Hydration Cycles. ACS Applied Materials & Interfaces, 2015. 7(32): p. 17874-17883.

Figure 1. The cantilever bending system utilizes a laser array to measure thin film stresses. The setup has been modified to also measure film swelling simultaneously. Stoney's equation shows the relationship between the radius of curvature of the substrate and the force exerted by the film. Es is the Young's modulus, hs is the thickness and vsis Poisson's Ratio, all of the substrate

Figure 1

2766

, and

Sub-micron Nafion films are relevant to several electrochemical devices including polymer electrolyte fuel cells, sensors and actuators. Such thin films are invariably fabricated from dispersion of the ionomer in suitable solvent, e.g. by drop-casting, spin-coating or self-assembly. Often a thermal treatment or annealing step is applied to improve the mechanical properties of the film and relax the polymers. Thus, understanding the response of the ionomer thin film subjected to thermal changes is important. Furthermore, thermal characteristics of the ionomer such as glass transition temperature and thermal expansion coefficient can provide fundamental insight into structure and the nature of the interaction with the substrate.

We have initiated an in-situ thermal ellipsometry and thermal FTIR study to probe the response of 20-650 nm thick Nafion thin films supported on SiO2 substrate. Ellipsometry experiments yield the thickness change as a function of temperature. This set of data yields thermal expansion coefficient of the Nafion films. The change in slope of the thickness change with temperature provides information on the nature of phase transition – glassy to rubbery or disorder-to-order. The thermal FTIR is being used as a complimentary technique to probe whether the phase change phenomenon manifests as a dramatic change in response in one of the characteristic peaks.

The results obtained so far indicate that the thermal expansion coefficient of Nafion films 60 nm and lower thickness is changes with thickness but those of the films above 60 nm thickness converges to that of the bulk material. No distinct transition in slope of thickness change is observed. That is, no glass transition like behavior is observed. No thickness-dependent response in FTIR peak is observed. A sharp change in the 638-626 cm-1 doublet region is observed at temperature corresponding to the so-called a transition for Nafion.

To the best of our knowledge, there have been no prior studies on in-situ thermal ellipsometry or in-situ thermal FTIR of Nafion thin films.

D-31a Non-PGM Cathode Catalysts 1 - Oct 6 2016 8:00AM

2767

, and

Recently, non-noble metal catalysts has been intensively studied for proton exchange membrane fuel cell to achieve low cost and high performance. Among them, carbon nanotubes have attracted much attention because of their lower cost and recyclable property. It is suggested that defective structures on the carbon surface with nitrogen doping and small amount of transient metal to form Fe-C-N moiety are important to have high activity.[1][2] However, the role of carbon defects for the oxygen reduction reaction (ORR) is still not clarified. We have been interested in the ORR activity of the defective pure carbon nanotube and found that with defective structures even without nitrogen doping or annealing, the onset potential can also reach to 0.73V vs RHE. [3]

So far high ORR activity always achieved by high temperature annealing process in Argon or ammonia atmosphere. In this study, aiming to further investigate the characteristics of the defective carbon nanotube, optimization of the chemical drilling processes conduct to control the defect structure on MWCNT. By this method high ORR activity MWCNT material could be obtained without annealing or heating treatment.

The MWCNT were provided by Showa Denko KK Japan, acid and heat treatment has been done before experiment. The procedure of making defect on MWCNT is prepared by cobalt deposition and oxidation. Here various conditions have been applied which will lead to different defect structure. These differences cause essential influences on the ORR property of MWCNT. To evaluate the performance of defected MWCNT, both cyclic voltammetry and power generation of cathode catalyst in PEMFC have been measured.

Linear sweep voltammetry test shows that high defect density and shallow defect structure could bring benefits to the ORR property, and 0.65V vs. RHE onset potential has already been achieved by pure MWCNTs in Fig 1, without N doping, metal impurities or annealing process. Temperature program desorption results also proved that different defect making methods would lead to different characteristics of functional group. The details of different defect structures and principles will be furtherly studied by XPS, Raman, ICP analysis and Power Generation.

Acknowledge:

This work partially was supported by COI STREAM from the Ministry of Education, Culture, Sports, Science and Technology (MEXT) and Shin-Etsu Chemical Co., Ltd. Japan. Their contribution is greatly appreciated.

Reference:

[1] Michel Lefèvre, Eric Proietti, Frédéric Jaouen, Jean-Pol Dodelet. "Iron-Based Catalysts with Improved Oxygen Reduction Activity in Polymer Electrolyte Fuel Cells" Science, 324, 71-74, 2009.

[2] Kuanping Gong, Feng Du, Zhenhai Xia, Michael Durstock, Liming Dai. "Nitrogen-Doped Carbon Nanotube Arrays with High Electrocatalytic Activity for Oxygen Reduction" Science, 323, 760-764, 2009.

[3] K. Waki, R. A. Wong, H. S. Oktaviano, T. Fujio, T. Nagai, K. Kimoto and K. Yamada. "Non-nitrogen doped and non-metal oxygen reduction electrocatalysts based on carbon nanotubes: mechanism and origin of ORR activity" Energy Environ. Sci., 7, 1950-1958, 2014.

Figure 1

2768

, and

Nitrogen-doped non-precious metal catalysts show high oxygen reduction reaction (ORR) activity, which are considered as possible alternatives for fuel cell catalysts. It has been reported that the transition metals and nitrogen doped carbon complex shows outstanding performance of ORR activity, suggesting that metal impurities in carbon nanotube play a very important role to the ORR activity. [1] On the other hand, recently the ORR active sites in N—doped carbon materials are carbon atoms with Lewis basicity next to pyridinic N was proposed. [2] However, the mechanism and the ORR active site are still in controversy. In our previous work, we reported that annealing multi-walled carbon nanotubes (MWCNT) with nano-drilled defective structure in Argon atmosphere could reach high ORR activity. [3] We suggest that the edge of defects is essential for the formation of active site. In order to clarify the active site and further enhance the ORR activity, here, by making defective edges on the MWCNT structure accompanied with nitrogen doping, we showed that the ORR activity has been improved with onset potential reached up to 1.1V vs RHE in 0.1M KOH and 0.88V vs RHE in 0.1M HClO4.

The MWCNT used were provided by Showa Denko KK Japan (VGFX-XA, diameter: 15nm, approximate length: 1μm; containing Fe impurities <1 wt%). As a precursor, functionalized and purified MWCNT was obtained by heat and acid treatment in the mixture of H2SO4/HNO3. [3] Defective MWCNTs were then prepared following our previous report by nano-drilling purified MWCNT using CoOx as oxidation catalyst. [3,4] Then defective MWCNTs were nitrogen doped in 10% NH3/ Argon at 900℃.

The ORR activity of nitrogen doped defective MWCNT in acid was shown in Fig.1 and in alkaline was shown in Fig.2. We found that nitrogen doped defective MWCNTs show very high ORR activity and efficiently doping nitrogen to the defective edge is essential for the enhancement of ORR activity. The optimization of nitrogen doping process and the nitrogen contain along with the role of defect structure on the ORR activity will be discussed further through detailed characterizations with X-ray photoelectron spectroscopy, Temperature Programmed Desorption and Raman Spectroscopy.

Acknowledgement

This work partially was supported by COI STREAM from the Ministry of Education, Culture, Sports, Science and Technology (MEXT) and Shin-Etsu Chemical Co. ,Ltd. Japan. Their contribution is greatly appreciated.

Reference

[1] Li, Yanguang, et al. "An oxygen reduction electrocatalyst based on carbon nanotube-graphene complexes." Nature nanotechnology 7.6 (2012): 394-400.

[2] Guo, Donghui, et al. "Active sites of nitrogen-doped carbon materials for oxygen reduction reaction clarified using model catalysts." Science 351.6271 (2016): 361-365.

[3] K. Waki, R. A. Wong, H. S. Oktaviano, T. Fujio, T. Nagai, K. Kimoto and K. Yamada, Energy Environ. Sci., 2014, 7, 1950-1958.

[4] Oktaviano, Haryo S., Koichi Yamada, and Keiko Waki. "Nano-drilled multiwalled carbon nanotubes: characterizations and application for LIB anode materials." Journal of Materials Chemistry 22.48 (2012): 25167-25173.

Figure 1

2769

, , , , , and

The oxygen reduction reaction (ORR) is an essential reaction for energy applications such as fuel cells and metal-air batteries. There are several drawbacks to the current ORR involving the kinetically sluggish process and expensive catalysts which impede the mass production of noble-metal materials for application in clean and efficient energy conversion devices. Herein, a type of ultrafine carbon nanoparticle with high nitrogen doping concentration can be simply prepared from a conveniently available precursor through a green and cost-effective hydrothermal process with sterculia scaphigera as the precursor, which is abundantly available and a constantly renewed natural source. With the aid of water, not only is the product's specific surface area enlarged, but the pore structure is also enriched, and there is improvement in the degree of graphitization and nitrogen-doped content. 1 Meanwhile, the size of the carbon particle can be readily tuned to a nanoscale by changing the duration of the hydrothermal reaction.2 Interestingly, electrocatalytic activity is dramatically enhanced with the transformation of the carbon size and nitrogen structure. Further considering the effect of temperature on the structural and functional properties, the carbon material treated at 900 oC serves as an optimal ORR electrocatalyst with high selectivity, remarkable stability and excellent methanol tolerance in comparison to the commercial 20% Pt/C electrode in alkaline and acidic media. This kind of carbon material may be a promising alternative for costly Pt-based electrocatalysts in fuel cells. The characterizations reveal that we have achieved a highly satisfactory electrocatalytic activity for ORR, a relatively positive onset potential of -0.02 V (vs. Hg/HgO) in alkaline media and 0.55 V (vs. Ag/AgCl) in acidic media, a higher diffusion-limiting current density than that of Pt/C, as well as outstanding stability and superior tolerance durability to methanol. The preparation method validated in this study responds to the requirement to produce high quality and outstanding performance carbon-based electrocatalysts. Both the simplicity of the operation and the employed biomass precursor exactly meet the criteria for significant cost-saving, easy scale-up and eco-friendly demands for energy storage.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (21001117), the Starting-Up Funds of South University of Science and Technology of China (SUSTC) through the talent plan of the Shenzhen Government and the Shenzhen Peacock Plan (KQCX20140522150815065). S.Y. thanks the support from the Guangdong and Shenzhen Innovative Research Team Program (No. 2011D052, KYPT20121228160843692),Special Funds for the Cultivation of Guangdong Collage Student's Scientific and Technological Innovation ("Climbing Program" Special Funds)

Reference

1. Chenghang, Y.; Xiaoyuan, Z.; Xiaochang, Q.; Fangfang, L.; Ting, S.; Li, D.; Jianhuang, Z.; Shijun, L., Fog-like fluffy structured N-doped carbon with a superior oxygen reduction reaction performance to a commercial Pt/C catalyst. Nanoscale 2015,7(8), 3780-3785.

2. Guo, D.; Shibuya, R.; Akiba, C.; Saji, S.; Kondo, T.; Nakamura, J., Active sites of nitrogen-doped carbon materials for oxygen reduction reaction clarified using model catalysts. Science 2016,351 (6271), 361-365.

Figure 1

2770

, , , , , and

Carbon foams are ideal materials for electrochemical applications due to their large surface area, high electrical conductivity, and stability over a wide electrochemical potential window. However, many of these materials are expensive; use complex multi-step synthesis protocols; and require the use of nanostructured template materials which must be removed with strong acid or alkaline. It is often difficult to control the porosity (and thereby optimize mass diffusion); and it is difficult to dope effectively with heteroatoms. We synthesize carbon foams at low cost and gram-scale, with large micron-scale pores, very thin walls, and very large surface area (e.g. 2500 m2/g), by thermal decomposition of sodium ethoxide.[1,2]

Crucially, by synthesizing and subsequently decomposing nitrogen-containing metal alkoxides, nitrogen-doped carbon foams can be made.[3] The nitrogen content can be varied over a wide range (e.g. <0.5 to 15 at%) by simply changing e.g. the precursor ratios. In addition, the ratio of e.g. pyridinic to tertiary nitrogen bonding can be tailored by changing the pyrolysis temperature. These materials do not contain transition metal contamination (confirmed by ICP-AES), unlike carbon nanotubes or carbon black. Therefore they can be used to probe the fundamental oxygen reduction reaction (ORR) activity of nitrogen-doped carbon in acid – a relatively under-represented topic, plagued by contamination issues.[4,5] We observe surprisingly high mass activity and onset potential for the ORR, with high electron transfer number. This work shows that 4-electron ORR is possible even in the absence of transition metals, most likely associated with tertiary nitrogen sites.[6] In alkaline medium, these catalysts are comparable to Pt/C, and undergo negligible degradation even over 60,000 load potential cycles.[7]

Finally, these materials are investigated for their electrochemical CO2 conversion activity to carbon monoxide, formic acid, methane, and ethane.

References

[1] S. M. Lyth, H. Shao, J. Liu, K. Sasaki, Int. J. Hydrogen Energy, 2014, 39, 376

[2] J. Liu, D. Takeshi, K. Sasaki, S. M. Lyth, J. Electrochem. Soc. 2014, 161, F838

[3] S. M. Lyth, Y. Nabae, N. M. Islam, et al., eJournal of Surface Science and Technology, 2012, 10, 29-32.

[4] S. M. Lyth, Y. Nabae, N. M. Islam, S. Kuroki, M. Kakimoto, S. Miyata, J. Nanosci. Nanotechnol. 2012, 12, 4887

[5] S. M. Lyth, Y. Nabae, N. M. Islam, S. Kuroki, M. Kakimoto, S. Miyata, J. Electrochem. Soc. 2011, 158, B194

[6] J. Liu, D. Takeshi, D. Orejon, K. Sasaki, S. M. Lyth, J. Electrochem. Soc. 2014, 161, F544

[7] J. Liu, K. Sasaki, S. M. Lyth, ECS Trans. 2014, 64, 1161

2771

, , , and

Highly active and durable electrocatalysts for the oxygen reduction reaction (ORR) are undoubtedly essential for the commercialization of polymer exchange membrane fuel cells. Recently, two types of iron-based catalysts have been extensively studied and reported to be active and stable towards ORR in acidic electrolyte. One is the well-recognized FeNx/C catalyst, and the other is the catalyst with structures of graphitic layer encapsulated metal particles. Generally, catalysts synthesized by pyrolysis of different iron- and nitrogen-containing precursors always contain both structures, making it impossible to distinguish the role of each structure during the ORR. Using a high temperature autoclave approach, we synthesized two model catalysts with almost exclusively one desired structure in each catalyst. Based on detailed characterizations in terms of catalyst structure and composition as well as electrochemical tests, the active site structures of these two catalysts and their correlation with the ORR performance are explored.

Figure 1. Schematic illustration of the two possible active sites.

Figure 1

2772

and

The widespread use of fuel cells is currently limited by the lack of efficient and cost-effective catalysts for the oxygen reduction reaction (ORR). Fe-based non-precious metal (NPM) catalysts exhibit promising activity and stability as an alternative to state-of-the-art Pt catalysts. However, the identity of the active species in NPM catalysts remains elusive, impeding the development of new catalysts. For the first time, we report the identification the catalytic species in an active NPM catalyst using Cl2 and H2 treatments to decrease heterogeneity. Additionally, we demonstrate the reversible deactivation and reactivation of the catalyst using the Cl2 and H2 treatments. Mössbauer and X-ray absorption spectra reveal that carbon-encapsulated Fe nanoparticles present in the as-prepared and H2-treated catalyst, but absent in the Cl2-treated catalyst are responsible for the observed ORR activity and stability of the catalyst. These findings suggest a new direction for the design and synthesis of enhanced NPM ORR catalysts.

2773

and

1. Introduction

After the major discovery of the platinum based electro-catalysts for the cathodic reduction of the oxygen in the polymer electrolyte membrane fuel cells, commercialization of PEM fuel cells begins for automobile applications. However, it is necessary to replace the platinum-based electro-catalysts due to the high cost and scarcity by using other cheap and abundant non-precious metal catalysts. Fe-containing nitrogen-doped carbon catalysts (Fe-N-C) have been demonstrated as the potential alternative for the Pt-based catalysts for the last few years. However the poor understanding of the active sites of Fe-N-C catalysts prevents the development of highly active durable catalysts. Our recent studies[1] of the ORR on the Fe-N-C and N-C catalyst demonstrates that the ORR predominantly proceeds via the step wise reduction. Though the conventional Damjanovic model[2] (as shown in Scheme 1) concludes the direct 4-electron ORR, it actually overestimates the k1 and underestimates the k2 and k3. Since the metal free catalysts are very active towards the reduction of O2 to H2O2, the role of the Fe on the reduction of H2O2 to H2O is studied and the estimated rate constants for the reduction of H2O2 to H2O will be used to estimate the real current for direct 4-electron ORR (I1), 2-electron ORR to H2O2 (I2) and 2-electron H2O2 reduction to H2O (I3).

2. Methodology

The precursor synthesis has been already discussed in the previous papers[3,4]. Briefly, the stoichiometric amount of pyromellitic acid dianhydride and 4,4'-oxydianiline was polymerized in the presence of tris(acetylacetonato) iron (III) (Fe(acac)3) yielding polyimides/Fe(acac)3 (the size of the polymer is controlled by maintaining low temperature during polymerization). The precursor was subjected to the multi-step pyrolysis reported earlier and the resulting catalyst is designated as Fe/PI. The Fe/PI-modified GC electrode with the loading density of 100 µg cm-2, carbon cloth and reversible hydrogen electrode was used as working, counter and reference electrodes, respectively. Both oxygen reduction reaction (ORR) and H2O2 reduction reaction was measured using rotating disk and rotating ring (Pt)-disk electrode voltammetry in 0.5 M H2SO4 solution. The H2O2 reduction rate constants (k3) with potential were estimated using Koutecký‒Levich plot. From the ORR voltammograms, the value of I1, I2and I3 were determined by the following equations;

I2=0.5*{Ir/N+ [(Ir/N)2(1+4k3/(Zω0.5)]0.5} (1)

 I3=0.5*{-Ir/N+ [(Ir/N)2(1+4k3/(Zω0.5)]0.5} (2)

 I1=I- [(Ir/N)2(1+4k3/(Zω0.5)]0.5 (3)

where Id and Ir represents disk and ring current of the oxygen reduction reaction, respectively, k3 represents the rate constant of the H2O2 reduction reaction and ω refers the rotational speed of the ORR voltammogram. The value of the Z can be estimated from the kinematic viscosity (v) and diffusion coefficient (D) of H2O2 in 0.5 M H2SO4 (Z=0.62D2/3v-1/6)

3. Results and Discussion

The H2O2-reduction steady-state voltammograms indicates that the Fe/PI catalyst is more active for both H2O2 oxidation and reduction indicated by the sharp intersection on the potential axis around 0.8 V vs. RHE. This simple method will be helpful separate the currents of many ORR voltammograms to describe the kinetics because this does not requires any parameters to be plotted against the rotation speed. The I2 and I3 values estimated using this method is significantly improved than the conventional Damjanovic model currents estimated form the rate constants (Figure 1).

4. Conclusion

  The newly developed methodology successfully separated I1, I2 and I3 for ORR over Fe/N/C cathode catalysts.

5. Acknowledgements

This work was financially supported by the New Energy and Industrial Technology Development Organization (NEDO), Japan.

6. References

[1] A. Muthukrishnan, Y. Nabae, T. Okajima, T. Ohsaka, ACS Catalysis, 5 (2015) 5194-5202.

[2] A. Damjanovic, M.A. Genshaw, J.O.M. Bockris, J. Chem. Phys., 45 (1966) 4057-4059.

[3] Y. Nabae, Y. Kuang, M. Chokai, T. Ichihara, A. Isoda, T. Hayakawa, T. Aoki, J. Mater. Chem. A, 2 (2014) 11561-11564.

[4] Y. Nabae et al., Sci. Rep., 6 (2016) 23276.

Figure 1

2774

, , , and

The limited natural abundance, high cost and carbon monoxide deactivation of Pt and other noble metals possess a major barrier in its applications for hydrogen or methanol fuel cells and therefore; development of alternative electrocatalytic materials based on non-precious metal-N4 is a key challenge in the current growing demand for the clean-energy fuel cell research. In this regard, in our group, a new cobalt-corrole molecule functionalized with ferrocene has been successfully synthesized and its structural characterization was done by UV-Vis, NMR and HR-Mass spectroscopic studies. The molecular structure of the complex was also confirmed by single crystal X-ray diffraction and explained with structural orientation of the atoms and attached moieties. Furthermore, this molecule mixed with activated carbon was pyrolyzed at different temperatures and among the resulting electrocatalysts, particularly at 500 °C was shown to be performed as promising ORR active catalyst for the PEM Fuel-Cell application. The unique Co-corrole integrated with mono substituted peripheral Fe complex with its pyrolyzed form of low symmetric structure provides a bimetallic (Co and Fe) active centre and facilitates the oxygen reduction reaction (ORR) via a 4-electron pathway. The prepared carbon black supported catalyst at 500 °C pyrolyzed temperature exhibits higher electron transfer number 3.96 for ORR in acidic medium with better stability, which is superior to the previously reported pyrolyzed Co-corroles. The enhancement of the ORR activity of the well characterized Co-corrole with peripheral Fe-metallocene bimetallic N4 macrocyclic complex provides a new prospect for the next-generation of non-precious metal-N4 electrocatalysts for fuel cell application. In this presentation, I will talk more in details of the electrochemical experimental process of the catalyst and its characterization using different analytical techniques.

Key words: N4-macrocyclic, corrole, oxygen reduction reaction, electrocatalysts, fuel cell

2775

The development of highly active, durable, and low-cost oxygen reduction reaction (ORR) catalysts is central to making polymer electrolyte fuel cells (PEFCs) commercially viable. In this direction, tremendous recent efforts have been devoted to replacing expensive, scarce Pt-based electrocatalysts with non-precious metal catalysts (NPMCs) for the ORR. Among a wide range of NPMCs, the Fe-N/C catalysts have emerged as the most promising ORR catalysts due to their high ORR activities. A growing body of literature based on spectroscopic studies suggests that the active sites of these catalysts involve Fe-Nx coordination. We have endeavored to develop general synthetic strategies towards high-performance Fe-N/C catalysts that can host a high density of catalytically active Fe-Nx sites. We have prepared transition metal-doped ordered mesoporous porphyrinic carbons (M-OMPCs) by nanocasting mesoporous silica templates. The M-OMPC catalysts contain predominantly molecularly dispersed Fe-Nx sites and have high surface areas and tunable pore structures. Among the M-OMPC catalysts, the FeCo-OMPC catalyst exhibited an excellent ORR activity in an acidic medium, higher than other non-precious metal catalysts. It showed higher kinetic current at 0.9 V and superior long-term durability and MeOH-tolerance than Pt/C catalyst. Second strategy is based on a generalized "silica-protective-layer-assisted" method that can preferentially produce catalytically active Fe-Nx sites towards highly efficient Fe-N/C electrocatalysts. This method is applicable to any type of carbon supports including carbon nanotubes, reduced graphene oxide, and carbon black. One of resulting catalysts, consisting of CNT wrapped with thin porphyrinic carbon layer (CNT/PC), contained relatively high density of Fe-Nx sites, and showed very high ORR activity and remarkable stability in alkaline media. Importantly, the CNT/PC-based cathode demonstrated excellent performances in both alkaline anion exchange membrane fuel cell as well as acidic proton exchange membrane fuel cell.

2776

, , , , , , , and

Research into alternative catalysts for the oxygen reduction reaction in PEMFCs continues to advance due to ongoing efforts to understand structure-to-property relationships in these materials. Though much work has been done on platinum group metal free (PGM-free) catalysts for the ORR, there is still a need to develop tangible relationships between synthesis methods, morphology, chemistry, and performance of these materials. One catalyst that has shown promise for both performance and durability is synthesized from precursors of iron salts and nicarbazin (Fe-NCB) using the sacrificial support method (SSM).1 Though synthesis procedures for this catalyst have been well optimized, there has not been a systematic study of structure-to-property relationships and how they relate to synthesis parameters.

In addition to fundamental properties of the catalyst itself, knowledge of how the material integrates into the electrode in an operational fuel cell is critical. To address this issue, both electrochemical testing and data from an operational fuel cell are required. Further, to be able to understand interactions between individual catalyst particles, and their interactions with ionomer in the catalyst layer, analysis of the catalyst must include measurement of chemistry and morphology both for the bulk and the surface.

To elucidate relationships between morphology, chemistry, and performance, a multivariate study on synthesis parameters and outcomes will be presented. TEM-EDS maps will be acquired and analyzed to examine nanoscale elemental distributions. Concentrations of atomically dispersed Fe with respect to particle and carbon plane edges, as well as the presence of Fe nanoparticles will be addressed. Correlations between location specific relative concentrations of Fe and N will be explored for correlations with performance, morphology, peroxide generation, and potentials for the Fe+2-Fe+3 transition. To address bulk carbon crystallite structure including stacking order, layer strain, and lateral crystallite size, a curve fitting algorithm will be applied to XRD spectra of these materials.2 Graphitic content and strain analyses will be related to both TEM observations and XPS chemical analysis. Morphology characterization will be completed by pore size analysis with nitrogen isotherms, catalyst surface analysis using the discrete wavelet transform, and particle size distributions. Relationships between catalyst structure from sub-nanometer to micron scale, elemental distributions within individual catalyst particles, electrochemical performance, and fuel cell performance will be elucidated and presented.

1. Serov, A.; Artyushkova, K.; Niangar, E.; Wang, C.; Dale, N.; Jaouen, F.; Sougrati, M.-T.; Jia, Q.; Mukerjee, S.; Atanassov, P. Nano-structured non-platinum catalysts for automotive fuel cell application. Nano Energy 2015, 16, 293-300.

2. Shi, H.; Reimers, J.; Dahn, J. Structure-refinement program for disordered carbons. Journal of applied crystallography 1993, 26, 827-836.

Figure 1

2777

, , and

Two of the largest hurdles facing platinum group metal-free (PGM-free) oxygen reduction reaction (ORR) electrocatalysts for polymer electrolyte fuel cell (PEFC) cathodes are the need for improved activity and increased durability. Quantum chemical modeling (in particular, density functional theory, DFT, coupled with thermochemical models) can be utilized to better understand both activity and durability of these materials at the atomic scale. In this work, we apply quantum chemical modeling approaches to study the activity of N-coordinated Mn, Fe, Co, and Ni containing zig-zag edge sites. Both Metal-N4 as well as Metal2-N5edge structures are considered. Spontaneous *OH ligand modification in which the *OH intermediate is persistently bound to metal centers above a relatively low potential is found to be applicable to both structures but only for certain metal species. Experimentally determined trends in activity as a function of metal species fit well with theoretical predictions based on the spontaneous *OH ligand modification hypothesis. In addition, strain-modification of structures and resultant shifts in activity are explored. It is found that, due to anisotropic changes in binding energies with strain, this approach can lead to minor improvements of the computationally obtained ORR activity descriptor.

With regards to the durability hurdle of PGM-free electrocatalysts, a structurally sensitive and computationally accessible descriptor is explored. First-principles molecular dynamics derived knock-on displacement threshold energy may serve as a high-throughput, corrosion-mechanism-agnostic descriptor of active site structure durability that does take into account the kinetics of local bond breaking. Further comparisons between computational predictions and experimental findings are required to ascertain the applicability of this simplified durability descriptor for non-PGM active site structures. Current progress and future directions will be discussed.

E-31 Alkaline & DFC Electrocatalysis 3 - Oct 6 2016 8:00AM

2778

, , , , and

Fe/N/C is a promising electrocatalyst for oxygen reduction reaction (ORR). This type catalyst can perform in both acidic and alkaline media. In acidic media, the active sites of Fe/N/C can be corroded; however, in alkaline media, Fe/N/C can exhibit high stability. Therefore, it is promising to explore the applications of Fe/N/C catalyst in alkaline fuel cell. Herein, we will talk about our recent advances in preparation of Fe/N/C catalysts and the application of them in alkaline fuel cells: (1) By using 2-aminothiazole, a molecule containing both N and S atoms, we prepared S co-doped Fe/N/C catalyst with graphene nanosheets. In alkaline solution, this catalyst exhibited high ORR activity with half-wave potential of 0.926 V and mass activity of 0.56 A g-1 @ 1.0 V. Furthermore, the catalyst displayed excellent durability, and only lost 9% of initial activity after 100 h of durability test at 0.80 V. In alkaline anion exchange membrane fuel cell (AEMFC) test, the peak power density could reach 164 mW cm−2. (2) We used the unzipped carbon nanotubes as carbon support and melamine as nitrogen source, and prepared a Fe/N/C catalyst with high ORR activity in alkaline media. The edge sites of unzipped carbon nanotubes play a key role for high catalyst activity. (3) we explored a series of nitrogen source including binary polymer of melamine-terethalaldehyde, ternary polymer of cynuric acid, 2, 4-diamino-6-phenyl-1, 3, 5-triazine, and melamine. N-doped carbon nanotubes with encapsulated Fe nanoparticles were prepared. The catalyst exhibited considerably high ORR activity in alkaline solution. We further tested its performance in alkaline membrane direct ethanol fuel cell, and the peak power density was about 64 mW cm-2. These results demonstrated that Fe/N/C is a promising candidate for ORR electrocatalyst in alkaline fuel cell.

References

[1] C. Chen et al. Chem. Commun., 2015, 51, 17092-17095.

2779

, , , , and

Platinum shows an exceptionally high activity for the hydrogen oxidation reaction (HOR) in acidic environment,[1] enabling ultra-low Pt-loadings on the anode side of proton exchange membrane fuel cells (PEMFCs).[2] Unlike in acid, this reaction is about two orders of magnitude slower on platinum and other platinum metals in alkaline electrolyte,[3]; [4] which has recently been suggested to be due to entropic effects.[5] Amongst other issues, the low HOR activity of Pt-based electrocatalysts in alkaline environment hinders the development of anion exchange membrane fuel cells (AEMFCs).[4]

Bimetallic Pt-Ru catalysts were demonstrated to exhibit significantly higher HOR activities in acid[6]; [7] and base[8]–[10] compared to Pt catalysts. One hypothesis for the high HOR activity of Pt-Ru alloys in base is a bi-functional mechanism,[9]; [10] with hydrogen being adsorbed on Pt and hydroxide being supplied by more oxophilic Ru, whereby the presence of hydroxide in the vicinity of Pt-Hads would accelerate the rate of the hydrogen oxidation. The second hypothesis considers a modification of the electronic structure of platinum by ruthenium,[8] leading to a lower Pt-Hads binding energy and ultimately to a higher activity of platinum towards the oxidation of hydrogen. While the exposure of ruthenium on the surface of the catalyst is absolutely mandatory for a bi-functional mechanism, this is not the case for an activity enhancement due to a modification of the electronic structure of platinum. Based on this fundamental difference, we have attempted to identify the actual HOR mechanism on bimetallic Pt-Ru catalysts by evaluating Ru@Pt core-shell nanoparticles with various Pt-coverages and shell thicknesses. Pt-coverage and shell thickness were determined electrochemically via the Ru-oxide reduction peak and COads stripping voltammetry, a method established by El Sawy et al.[11] Considering a pure, bi-functional mechanism where hydroxide on Ru reacts with Hads on Pt, ruthenium is required to be on the surface of the catalytically active particle. Hence, we expect the highest activity at submonolayer Pt-coverage.

For the bifunctional mechanism, covering the whole Ru-surface with Pt would result in a performance equal or similar to that of pure Pt (s. red line in Fig. 1). In contrast to that, for a purely electronic effect, Ru surface atoms would not be required for effective HOR catalysis and one would expect the best performing Ru@Pt catalysts to have a fully Pt-covered Ru surface or even multilayers of Pt; the optimum Pt-shell thickness would mainly depend on the range of the electronic effect (s. black line in Fig. 1). Based on this concept, we will present data to disentangle the mechanism of the hydrogen oxidation on bimetallic Pt-Ru catalysts. The activity of the developed catalysts will be compared to Pt catalysts prepared by the same method with a similar particle size as that of the Ru@Pt core-shell nanoparticles.

Acknowledgment

The authors of this work thank the microanalytical laboratory at TUM for quick and reliable elemental analysis.

References

1 J. Durst; C. Simon; F. Hasche and H. A. Gasteiger, J. Electrochem. Soc., 1, F190-F203 (2015).

2 H. A. Gasteiger; J. E. Panels and S. G. Yan, J. Power Sources, 1-2, 162 (2004).

3 P. J. Rheinlander; J. Herranz; J. Durst and H. A. Gasteiger, J. Electrochem. Soc., 14, F1448–F1457 (2014).

4 W. Sheng; H. A. Gasteiger and Y. Shao-Horn, J. Electrochem. Soc., 11, B1529 (2010).

5 J. Rossmeisl; K. Chan; E. Skúlason; M. E. Björketun and V. Tripkovic, Catalysis Today, 36 (2016).

6 J. X. Wang; Y. Zhang; C. B. Capuano and K. E. Ayers, Scientific reports, 12220 (2015).

7 J. X. Wang; P. He; Y. Zhang and S. Ye, ECS Transactions, 3, 121 (2014).

8 K. Elbert; J. Hu; Z. Ma; Y. Zhang; G. Chen; W. An; P. Liu; H. S. Isaacs; R. R. Adzic and J. X. Wang, ACS Catal., 11, 6764 (2015).

9 S. St. John; R. W. Atkinson; R. R. Unocic; T. A. Zawodzinski and A. B. Papandrew, J. Phys. Chem. C, 24, 13481 (2015).

10 Y. Wang; G. Wang; G. Li; B. Huang; J. Pan; Q. Liu; J. Han; L. Xiao; J. Lu and L. Zhuang, Energy Environ. Sci., 1, 177 (2015).

11 El Sawy, Ehab N.; H. A. El-Sayed and V. I. Birss, Chem. Commun., 78, 11558 (2014).

Figure 1

2780

, and

The electrooxidation of C3 alcohols (1-propanol, 2-propanol, 1,2-propanediol, 1,3 propanediol and glycerol) has been studied in alkaline medium on Pt/C and Pt9Bi1/C catalysts by cyclic voltammetry and in situ FTIR spectroscopy. The modification of Pt by 10 at% of Bi decreases the oxidation onset potentials of C3 alcohols down to ca. 200 mV. In situ FTIR spectroscopy measurements indicated clearly that the presence of Bi also led to avoid the C-C bond cleavage during alcohol electrooxidation reactions. Systematic evaluation of the positions and intensity changes of absorption bands as a function of the electrode potential allowed determining reaction pathways of electrooxidation of C3 alcohols. It has been shown that the secondary alcohol groups are more reactive than the primary ones, but also that in the case of glycerol steric limitations due to the presence of two primary alcohol groups could be responsible of the higher oxidation onset potential. At potentials above 0.6 V, the linear1-propanol bearing a single primary alcohol group leads to the highest activity due to lower steric hindrance of the surface compared with the other alcohols studied.

Figure 1

2781

, , and

In the present study, we demonstrate a one-pot capping agent free synthesis of metal nanoparticles deposited on low dimensional carbon supports. A laser ablation mediated synthesis (LAMS) of gold-silver on graphene and/or carbon nanotube composite is reported here. We aim to remove the use of surfactants, which are known to mask the performance of the catalyst. The composite materials have been well characterized morphologically as well as, functionally using microscopic and spectroscopic techniques. It has been observed that the introduction of Ag into the bimetallic catalyst has made the electro-oxidation of ethylene glycol and glycerol more facile. The synergistic influence of the two metal nanoparticles, with the added advantage of its pristine surface, pose an enhanced catalytic effect for alcohol electrooxidation. In addition, the onset potential for the reaction has been found to be more negative that translates to its plausible application in Direct Alcohol Fuel Cells.

Figure 1

2782

, , , , and

The electrochemical oxidation of methanol on electrocatalysts is intensively investigated since direct methanol fuel cells (DMFCs) were proposed as a promising alternative as power source for mobile and back-up power applications. In DMFCs, methanol is oxidized at the anode and oxygen is reduced at the cathode to convert the chemical energy of the fuels to electric power. In recent years, the electrochemical oxidation of methanol in alkaline environment is rising in interest, because other electrocatalysts than the expensive platinum are feasible as anode and cathode catalysts in alkaline DMFCs. While multiple catalysts are feasible on the cathodic site, Palladium was proven to be a suitable substitute for Pt as anode catalyst for the electrochemical oxidation of alcohols in alkaline media [1], but needs further improvement to compete with Pt especially regarding methanol oxidation reaction. The catalytic activity of Pd catalysts was improved by alloying or mixing other metals (or metal oxides) to Pd [2,3] or by controlling the size and shape of Pd particles [4]. Besides high catalytic activity, the complete oxidation of methanol to CO2is of enormous interest, because the complete oxidation results in the highest possible faradaic efficiency with six electrons transferred per methanol molecule. In contrast, two or four electrons are transferred to the anode per methanol molecule, if methanol is not completely oxidized. Differential electrochemical mass spectrometry (DEMS) is used by different research groups to quantify a catalyst's ability to completely oxidize methanol. While the methanol electrooxidation on Pt-based catalysts was investigated intensively via DEMS in acidic media, DEMS studies on Pd-based catalysts for methanol electrooxidation in alkaline media are rare.

In former studies, we reported on PdXNi/C and PdXRu/C catalysts for methanol electrooxidation in alkaline media synthesized by wet-chemical reduction with sodium borohydride [2,3]. It was possible to show that these oxophilic metals have a positive effect on the catalytic activity towards methanol oxidation of Pd-based catalysts. Taking these studies into account, we herein investigated the influence of these metals on Pd-based electrocatalysts regarding methanol oxidation reaction via DEMS. Besides analyzing the product distribution from methanol bulk oxidation on PdXNi/C or PdXRu/C, the main interest of this work is to investigate how the addition of the metals to Pd influences the methanol oxidation mechanism and catalyst poisoning. Hence, standard electrochemical tests like cyclic voltammetry, chronoamperometry, COadsstripping or methanol adsorbate stripping were conducted in the DEMS flow cell.

Figure 1 shows cyclic voltammograms and corresponding MS signals for produced CO2 (m/z = 44) and methyl formate (m/z = 60) for Pd/C, Pd5Ni/C and Pd3Ru/C during methanol electrooxidation. While both modified catalysts show higher current densities and lower onset potentials than Pd/C for the electrooxidation of methanol, the product distribution differs strongly depending on the added metal. For Ni-modified catalyst the formation of CO2 is significantly lower than for Pd/C. While the addition of Ni to Pd leads to a negative effect on the ability to completely oxidize methanol, Pd3Ru/C is able to oxidize methanol more efficiently to CO2 than Pd/C. For Pd3Ru/C, the CO2 formation is shifted to lower potentials while methyl formate is produced at higher potentials. This hints to a more selective production of CO2 at low potentials on Pd3Ru/C than on Pd/C or Pd5Ni/C.

Besides DEMS experiments, measurements were carried out in dependence of methanol concentration or oxidation temperature using a 3-electrode setup with stationary electrolyte to determine the rate determining step and the activation energy of methanol electrooxidation on Pd/C and the modified catalysts.

Fig. 1 – Methanol bulk oxidation on Pd/C, Pd5Ni/C and Pd3Ru/C investigated via cyclic voltammetry in a DEMS flow cell at 20 mV s-1 from 0.1 to 1.2 V (vs. RHE) with 0.1 M CH3OH + 0.5 M KOH solution as electrolyte. Corresponding mass spectrometry signals for CO2 (m/z 44) and HCOOCH3(m/z 60) are displayed to evaluate the product distribution during methanol oxidation reaction.

References

[1] C. Xu, L. Cheng, P. Shen, Y. Liu, Electrochemistry Communications 9 (2007) 997-1001.

[2] T. Jurzinsky, C. Cremers, F. Jung, K. Pinkwart, J. Tübke, International Journal of Hydrogen Energy 40 (2015) 11569-11576.

[3] T. Jurzinsky, P. Kammerer, C. Cremers, K. Pinkwart, J. Tübke, Journal of Power Sources 303 (2016) 182-193.

[4] A. Serov, T. Asset, M. Padilla, I. Matanovic, U. Martinez, A. Roy, K. Artyushkova, M. Chatenet, F. Maillard, D. Bayer, C. Cremers, P. Atanassov, Applied Catalysis B: Environmental 191 (2016) 76–85.

Figure 1

2783

, , and

Hybrid structure of Co3O4 nanocrystals supported on carbon nanotubes (CNTs) was previously synthesized in our lab, and it exhibits excellent bifunctional oxygen reduction reaction (ORR)/oxygen evolution reaction (OER) activities in alkaline media. More importantly, the hybrid nanomaterials showed great durability in a broad cycling potential range, from 0.0 V to 1.9 V (vs. RHE), which is a harsh environment that no previous hybrid structure reported has demonstrated stability in.

In this study, we further explored other types of structure using similar synthesizing procedure, but different metal precursors (forming single, binary or ternary oxide solid solution with cobalt acetate, nickel acetate, manganese acetate, iron acetate or copper acetate) and solvents to pursue better performance. Other parameters such as CNT size and post-treatment conditions were also altered to optimize the synthesis procedure.

The CNT-based hybrid electro-catalysts were made into membrane electrode assemblies (MEAs) to be tested in reversible alkaline fuel cells. The MEA fabrication was optimized in aspects of: catalyst ink preparation (by tuning the ionomer type, ionomer/catalyst ratio, and sonication length), membrane selection (alkaline membrane types, pre- and post- activation), catalyst-coated membrane fabrication (painting, spraying, hot pressing condition) and cell assembly. The MEAs were then tested in fuel cell mode and electrolyzer mode intermittently. MEA structural characterizations were carried out before and after these tests to study catalyst degradation mechanisms.

This work may provide perspectives for bifunctional catalyst synthesis and MEA design for reversible alkaline fuel cells.

Acknowledgement:The project is financially supported by the Department of Energy's Fuel Cell Technology Office under the Grant DE-EE0006960.

2784

, and

Recently, carbon materials functionalized with transition metals and nitrogen (labelled as Me-N-C, where Me = Fe, Co, etc.) have been reported as a competitive class of non-precious metal catalysts (NPMCs) for the oxygen reduction reaction (ORR).1,2 However, the intrinsic nature of the active site is poorly understood and still remained controversial.3,4 Various kinds of non-crystalline MeNxCy moieties5 or the crystalline metal-based nanoparticles encased in surrounding graphitic carbon layers6 have been proposed as the main active sites. Me-N-C composite pyrolyzed at high temperaturegenerally contained metal-based nanoparticles encapsulated in carbon layers that are hard to remove completely by acid leaching. As a result, the co-existence of metal-based nanoparticles and non-crystalline MeNxCy moieties on the carbon surface makes its further theoretical analysis and experimental characterizations very difficult.

Here we develop a novel template-free strategy (see Fig.1) to synthesize a series of iron and nitrogen co-doped carbon materials (Fe-N-C) with controllable structure for ORR. These catalysts with well-defined structures not only possess remarkable ORR activity but also provide unique models for probing active sites of Fe-N-C composites. The most active catalyst shows a half wave potential at 0.86 V, which is much higher than that of Pt/C with the same mass loading. The activities of various Fe-N-C catalysts synthesized by different methods will be compared and the correlation of the activity with the structure will be presented. Possible active sites will be discussed.

Fig. 1 Synthesis of ORR catalysts with four different kinds of typical morphologies and the corresponding LSV curves.

 

References

(1) Wu, G; More, K. L.; Johnston, C. M.; Zelenay, P. Science, 2011, 332, 443-447.

(2) Masa, J.; Xia,W.; Muhler, M.; Schuhmann,W. Angewandte Chemie International Edition, 2015, 54, 10102-10120.

(3) Hu,Y.; Jensen, J. O.; Zhang, W.; Cleemann, L. N.; Xing, W. N.; Bjerrum, J.; Li, Q. Angewandte Chemie International Edition, 2014, 53, 3675-3679.

(4) Zhu, Y.; Zhang, B.; Liu, X.; Wang, D. W.; Su, D. S. Angewandte Chemie International Edition, 2014, 53, 10673-10677.

(5) Zitolo, A.; Goellner, V.; Armel, V.; Sougrati, M. T.; Mineva, T.; Stievano, L.; Fonda, E.; Jaouen, F. Nature Materials, 2015, 14, 937-942.

(6) Deng, D.; Yu, L.; Chen, X.; Wang, G.; Jin, L.; Pan, X.; Deng, J.; Sun, G.; Bao, X. Angewandte Chemie International Edition, 2013, 52, 371-375.

2785

, , , , , and

Direct formic acid fuel cells, DFAFCs, have received considerable attention due to their higher power density in the direct liquid fuel cells. Recently, 550 mW/cm2 of power density at 30 oC was reported by Chan et al. using Pd–Ni2 P/C catalyst for DFAFC anode [1]. In order to obtain the higher power density, the operation temperature of DFAFC should be elevated since anode and cathode reaction kinetics increases with the increasing operation temperature, however, it has been known that the power density of DFAFC above 50 oC showed plateau with the increasing temperature. As a reason of the plateau in the power desnity above 50 oC, the increase of the mass transport loss was considered [2]. Therefore, controlling the mass transport in the catalyst layer is one of the key factor to obtain high power density. In this context, the silica embedded carbon nanofiber, SECNF, support for DFAFC anode have been suggested in this study. By using SECNF as a catalyst support, both of the high porosity catalyst layer which can enhance the mass transport rate in the catalyst layer and the high catalystic activity due to the support metal effect of the silica have been expected. It was found that the formic acid oxidation reaction activity significantly increased by using SECNF as a support of Pd comparing to conventional Pd/CNF and Pd/C due to the support effect of the SiO2. Moreover, the high porosity anode electrode could be successfully fabricated by using Pd/SECNF. On the other hand, the power density of DFAFC was not enhanced due to thick catalyst layer. In the presentation, the effect of the porosity and thickness of catalyst layer on the anode mass transport loss and power generation characteristics will be discussed in detail.

Reference

[1] J. Chang, L. Feng, C. Liu, W. Xing, and X. Hu, Angew. Chemie Int. Ed., 53, 122–126 (2014).

[2] T. Tsujiguchi, F. Matsuoka, Y. Hokari, Y. Osaka, and A. Kodama, Electrochim. Acta, 197, 32–38 (2016)

2786

and

Application of metal oxides as active matrices in electrocatalysis is particular importance. The hydrous behavior, which favors proton mobility and affects overall reactivity, reflects not only the oxide's chemical properties but its texture and morphology as well. What is of importance to electrochemical science and technology, certain nonstoichiometric mixed-valence oxides could exhibit pseudo-metallic conductivity and possess appreciable catalytic activity.

Recently, mixed oxide systems stabilized through Zr-O-W bonds have been demonstrated to be very attractive acid catalysts exhibiting high catalytic activities and good stabilities in many demanding industrial reactions. For electrocatalytic applications, mixed-valent tungsten(VI,V) oxide and zirconium(IV) oxide have been sequentially deposited and integrated through voltammetric potential cycling to form sub-microstructured films on glassy carbon electrode. The mixed WO3/ZrO2 systems are characterized by fast charge (electron, proton) propagation during the system's redox transitions. By dispersing metallic Pt and bimetallic platinum-ruthenium (PtRu) electrocatalytic nanoparticles over such active WO3/ZrO2 supports, the electrocatalytic activities of the respective systems toward the oxidation of hydrogen and small organic molecules (formic acid, methanol, ethanol or dimethyl ether) have been enhanced even at decreased loadings of noble metal nanostructures (for hydrogen oxidation) as well as in terms of both increasing the electrocatalytic currents and lowering the onset potentials of organic fuels' reactions. The enhancement effects should be attributed to features of the mixed metal oxide support such as porosity and high population of hydroxyl groups (due to presence of ZrO2), high Broensted acidity of sites formed on mixed WO3/ZrO2, fast electron transfers coupled to unimpeded proton displacements (e.g. in HxWO3), as well as strong metal-support interactions between nanosized noble metals (Pt, Pd or PtRu) and metal oxo species. The fact that WO3/ZrO2 nanostructures are in immediate contact with the metallic catalytic sites leads to the competitive interactions (via the surface hydroxyl groups) with undesirable reaction intermediates (including CO adsorbates). Thus their desorption ("third body effect") or even oxidative removal (e.g. of CO to CO2) are feasible.

 Furthermore, during oxidation of ethanol, when rhodium nanoparticles have been dispersed in between WO3 and ZrO2 layers (toward formation of the nanoreactor system), significant electrocatalytic current enhancements are observed. The result can be rationalized in terms of the formation of sub-microstructured nano-reactors in which Rh induces splitting of C-C bonds in C2H5OH molecules toward CHx/CH4 and CO adsorbate species, or even methanol-type intermediates, that could be further electrooxidized at PtRu catalytic centers. Further mechanistic studies along this line are planned.

2787

, , , and

Anion exchange membrane (AEM) fuel cells and electrolyzers have advanced over recent years to become competitive with proton exchange membrane (PEM) devices. The primary benefit of AEM systems is the potential ability to use less expensive, non-platinum (Pt) group metal (PGM) catalysts, and the non-PGM stability that the alkaline environment allows. While these systems have had polymer limitations, recent improvements in the chemical stability of alkaline membranes suggest that catalysts will become the limiting factor in the near future.[1, 2]

Catalyst development in AEM systems typically focuses on oxygen reduction and evolution, since the reactions are kinetically orders of magnitude slower than hydrogen oxidation and evolution. Alternatives, however, exist to PGMs in these reactions: silver in oxygen reduction; and nickel and cobalt in oxygen evolution.[3, 4] While hydrogen oxidation and evolution are kinetically faster reactions, they are roughly two orders of magnitude slower on Pt in base compared to acidic environments. Non-PGM catalyst options are also less clear, and generally struggle to justify the AEM cost benefit, producing activities orders of magnitude lower than PGMs at higher overpotentials.[5]

Recently, advanced Pt electrocatalysts have been developed in an effort to thrift the amount of PGMs in AEM fuel cells and electrolyzers. Pt-nickel (Ni) nanowires, previously developed for acidic oxygen reduction, were studied for their activity in hydrogen oxidation and evolution.[6] These materials were formed by spontaneous galvanic displacement, a process that occurs when a metal template contacts a nobler metal cation. At low levels of displacement, small amounts of Pt were deposited to produce high electrochemical surface areas. Subsequent post-synthesis processing was used to integrate the Pt-rich and Ni-rich zones, compressing the Pt lattice and improving its activity for hydrogen oxidation and evolution. Compared to carbon-supported Pt nanoparticles (Pt/HSC), Pt-Ni nanowires produced hydrogen oxidation/evolution exchange current densities 9 times greater.

Figure 1. Comparison of the mass (red) and site-specific (blue) exchange current densities of Pt-Ni nanowires, Pt-copper nanowires, and Pt/HSC in the hydrogen oxidation and evolution reactions in a 0.1 m sodium hydroxide electrolyte.[7]

[1] B. Pivovar, Alkaline Membrane Fuel Cell Workshop Final Report, in: U.S. Department of Energy (Ed.), http://www1.eere.energy.gov/hydrogenandfuelcells/pdfs/amfc_may2011workshop_report.pdf, 2011.

[2] K.J.T. Noonan, K.M. Hugar, H.A. Kostalik, E.B. Lobkovsky, H.D. Abruña, G.W. Coates, Journal of the American Chemical Society, 134 (2012) 18161-18164.

[3] R. Subbaraman, D. Tripkovic, K.-C. Chang, D. Strmcnik, A.P. Paulikas, P. Hirunsit, M. Chan, J. Greeley, V. Stamenkovic, N.M. Markovic, Nat Mater, 11 (2012) 550-557.

[4] J.S. Spendelow, A. Wieckowski, Physical Chemistry Chemical Physics, 9 (2007) 2654-2675.

[5] J.K. Nørskov, T. Bligaard, A. Logadottir, J.R. Kitchin, J.G. Chen, S. Pandelov, U. Stimming, Journal of The Electrochemical Society, 152 (2005) J23-J26.

[6] S.M. Alia, B.A. Larsen, S. Pylypenko, D.A. Cullen, D.R. Diercks, K.C. Neyerlin, S.S. Kocha, B.S. Pivovar, ACS Catalysis, 4 (2014) 1114-1119.

[7] S.M. Alia, B.S. Pivovar, Y. Yan, Journal of the American Chemical Society, 135 (2013) 13473-13478.

Figure 1

2788

, , , and

In this study the CeO2/graphene and Nb2O5/graphene supported Pt and PtCo catalysts denoted as PtCeO2/GR, PtNb2O5/GR, PtCoCeO2/GR and PtCoNb2O5/GR were prepared by polyol method using the simple and rapid microwave heating method. To obtain the catalysts, required amounts of CeO2/graphene or Nb2O5/graphene, H2PtCl6 and CoCl2 were heated in ethylene glycol at 170 oC for 30 min in the microwave reactor. It was found that Pt nanoparticles were successively deposited onto the surfaces of CeO2/graphene and Nb2O5/graphene. For comparison, the PtCo/graphene (PtCo/GR) and Pt/carbon (Pt/C) catalysts were prepared in the same manner. The composition and morphology of catalysts were detected by Inductively Coupled Plasma Optical Emission Spectroscopy (ICP-OES), Field Emission Scanning Electron Microscopy (FESEM), and Transmission Electron Microscopy (TEM). The synthesized PtCeO2/GR, PtNb2O5/GR, PtCoCeO2/GR, PtCoNb2O5/GR, PtCo/GR and Pt/C catalysts were examined as electrocatalysts towards the electro-oxidation of methanol and ethanol, and the reduction of oxygen.

It has been found that the CeO2/GR or Nb2O5/GR supported Pt and PtCo catalysts show an enhanced electrocatalytic activity towards the oxidation of ethanol and methanol in an alkaline medium as compared with that of the Pt/C catalyst. Depending on catalysts composition, ethanol and methanol oxidation current densities were found to be ca. 3 – 12 times higher at the CeO2/GR and Nb2O5/GR catalysts, respectively, in comparison with those at the Pt/C catalyst.

In the case of oxygen reduction, the PtCoCeO2/GR and PtCoNb2O5/GR catalysts outperform the PtCeO2/GR, PtNb2O5/GR, PtCo/GR and Pt/C catalysts towards oxygen reduction and shows higher onset potential, as well as higher current density towards the oxygen reduction reaction as compared with those at the aforementioned catalysts.

D-31b New Support Materials & Anode Catalysts - Oct 6 2016 8:00AM

2789

, , , , , , , , , et al

Graphitic carbon nitride is a new class of semiconducting graphene-like polymeric material with visible light absorption and photocatalytic properties. In addition to high nitrogen content and tunable structure, it was shown that graphitic carbon nitride based on polytrazine imide (PTI) sheets exhibit excellent anti-corrosion ability in ex-situ fuel cell environment.1 However, in bulk form, their low surface area and poor conductivity limits their applications in fuel cells.

In this work, an exfoliation route was established to produce PTI ink containing single to few-layer nanosheets. The ink was then processed to produce 3D networks of carbon nitride nanosheets/reduced graphene oxide (rGO-PTI) hybrid nanostructure with large interconnecting pores for fast mass transport of reactants (Figure 1a) and high surface area. The material was decorated with low loadings of platinum which also acts as spacers to prevent the restacking of the nanosheets, and then investigated for its electrochemical properties and applications as a durable catalyst support for PEM fuel cells. Initial results show that the cathode catalytic activity of Pt/rGO-PTI hybrid is significantly improved in comparison to Pt/PTI or Pt/rGO (Figure 1b), and long-term potential cycling shows that it is more durable than commercial Pt/C. The work is now being extended to develop high-performance and stable membrane electrode assemblies (MEA) for PEM fuel cells.

  • N. Mansor, A. Belen Jorge, F. Corà, C. Gibbs, R. Jervis, P. F. McMillan, X. Wang, D. J. L. Brett, J. Phys. Chem. C 2014, 118 (13), 6831-6838

Figure 1

2790

, , and

Pt-nanoparticles activated porous carbon composite catalysts are widely used in PEMFC applications. The instability of these catalysts in harsh fuel cell working conditions is the main problem hindering the large-scale application of the fuel cell technology in commercial and energy recuperation systems [1].

Hierarchical microporous-mesoporous carbon supports C(Mo2C) were prepared from molybdenum carbide using high temperature chlorination method at different synthesis temperatures from 600 °C to 1000 °C. The Pt-nanoparticles were deposited onto the carbon support using the NaHB4method [2,3]. The X-ray diffraction, low temperature nitrogen sorption, thermogravimetric analysis (TGA) and high-resolution scanning electron microscopy (HRSEM) methods were used to characterize the catalysts materials. The medium size of the Pt-nanoparticles deposited was 4.5±0.6 nm.

Corrosion of the Pt-C(Mo2C) catalysts were evaluated using an accelerated durability test (ADT) method (in oxygen (99.999%, AGA) saturated 0.5 M H2SO4 aqueous solution, from 0.58 V to 0.98 V vs SHE, at potential scan rate 50 mV s−1) [4]. The oxygen electroreduction reaction kinetics and the value of electrochemically active surface area (ECA) of the catalysts were determined after 10, 100, 1000, 3000, 10 000, 20 000 and 30 000 cycles.

Electrochemical impedance spectra were obtained at 0.9 V vs SHE within ac frequencies from 0.01 to 10 000 Hz to determine the changes of the electrolyte resistance and capacitance values during the ADT. The results show that only 20% decrease in electrode capacitance and 15% increases in series resistance takes place during 30,000 potential cycles.

The values of ECA were estimated in the solution saturated with nitrogen (99.999%, AGA) using the cyclic voltammetry at different potential scan rates from 30 to 1000 mV s−1. The rotating disc electrode method was used to evaluate the oxygen electroreduction kinetics in solution saturated with oxygen at rotation rate 1600 rev min−1 and at potential scan rate 20 mV s−1within the region of potentials from 1.03 V to 0.03 V vs SHE. Based on the cyclic voltammetry and rotating disk electrode data the main decrease in ECA took place during the first 1000 cycles and thereafter the ECA value stabilized. The half wave potential values depend weakly on the number of cycles (up to 30,000) measured.

The content of Pt in the catalyst was determined before and after the ADT using the electrochemical dissolution of platinum from the electrode in 6 M HCl solution. The average weight percentage of Pt in different catalyst materials was 71.8±1.7. The content of Pt in the catalysts before the ADT was in a good agreement with the TGA data.

References

[1] B.C.H. Steele, A. Heinzel, Nature 414 (2001) 345–352.

[2] K. Vaarmets, J. Nerut, E. Härk, E. Lust, Electrochimica Acta 104 (2013) 216–227.

[3] E. Lust, K. Vaarmets, J. Nerut, I. Tallo, P. Valk, S. Sepp, E. Härk, Electrochimica Acta 140 (2014) 294–303.

[4] U.S Department of Energy, Office of Energy Efficiency and Renewable Energy, Rotating Disk-Electrode Aqueous Electrolyte Accelerated Stress Tests for PGM Electrocatalyst/Support Durability Evaluation (2011).

Acknowledgements

This work was supported by Estonian target research project IUT20-13, Personal Research Grant PUT55, Estonian Centre of Excellence Projects No. 2014-2020.4.01.15-0011 and 3.20101.11-0030.

2791

, , and

Long recognised as key to the reduction of precious metal loading in PEMFC is the use of small catalyst nanoparticles (c.a. 3nm) stabilised by supporting them on high surface area carbon. However, PEMFCs using this technology achieve only 50% of the lifetime required by the US DOE's target [1] with the oxidation of carbon supports a major lifetime limitation [2]. We have previously reported that catalysts on composite boron carbide, graphite supports show an increased durability to potential cycling below 1V and remarkably high ORR specific activities, 85% higher than commercial Pt/C [3]. Despite their excellent performance at low potentials, the presence of carbon in these catalysts still limits their durability during cycling to high potentials.

We will present our latest work on catalyst supports in the Si-B-C system with the focus on eliminating carbon. Hybrid DFT/HF calculations were used to highlight boron doped boron carbide and boron doped silicon carbide as ceramics with intrinsic conductivity [Figure 1a]. To synthesise these doped ceramics with high purity, low temperature synthetic routes based on the carbo-thermal reduction of polymers were developed. These routes gave carbides with low levels of carbon impurities and provided control over the morphology of the support. This allowed the creation of high surface area supports on which platinum nanoparticles have been deposited [Figure 1b]. Electrochemical characterisation (ORR activity, accelerated degradation, etc) of the produced catalysts will also be presented.

[1] Fuel Cell Technical Team Roadmap, 2013, US DOE

[2] E. Antolini, and E. R. Gonzalez. Solid State Ionics, 2009, 180.9, 746

[3] C. Jackson et al. 18th Topical ISE Meeting 2016. [Manuscript in preparation]

Figure 1

2792

, , and

Carbon black is an essential material for PEFC. Due to its specific surface area, conductivity, pore-making structure and stability for acidic environment, carbon black is so far the best choice as the catalyst support for cathode and anode. However, carbon black is being one of the critical elements which limit the lifetime of PEFC. Especially in cathode, oxidative degradation of carbon black support is resulting in the depression of catalyst activity.

In order to prevent carbon black support from its oxidative degradation, surface coating with silicon oxide (silica) on acetylene black was tried using sol-gel method. Nanometer-thick amorphous silica layer was formed on almost the whole surface on the acetylene black. Its effect on resistivity against oxidation, Pt supporting function and total behavior of the catalyst is to be discussed.

2793

, , , , and

Catalyst supports for Polymer Electrolyte Fuel Cells (PEFC) are currently made up of carbon blacks. This material is however not thermodynamically stable in fuel cell operating conditions and loss of performance is observed with time, especially at the cathode side. To improve PEFC durability and make this technology a credible alternative to conventional power sources, carbon free cathodes were prepared. With a remarkable morphology, aerogels have already proven their ability to efficiently support catalysts for PEFC application [1, 2]. In this study doped tin dioxide aerogels are proposed as alternative support presumably stable in PEFC operating conditions.

Antimony and niobium doped tin dioxide aerogels were synthetized using sol-gel route in acidic media from alkoxide precursors. These materials have shown particularly adapted physico-chemical properties [3]. Platinum catalyst supported on doped SnO2 aerogels was prepared by two methods. Method A was based on the impregnation of a platinum salt followed by a reduction under UV and a heat treatment in oxidative or reducing atmosphere. Method EG is a conventional polyol method using ethylene glycol. Electrocatalysts structures and morphologies were investigated by X-ray diffraction and transmission electron spectroscopy. Active Electrochemical Surface Areas (ECSA) and catalytic activities for oxygen reduction reaction (ORR) were measured on Rotating Disk Electrode (RDE). Method A leads to the formation of particularly well dispersed Pt nanoparticles on aerogel surface (Figure 1), whereas filament form was observed after the use of Method EG. Heat treatments have shown direct influence on Pt structure and crystallinity. Highest ECSA was recorded after method A (45 m². mgPt-1) while highest ORR mass activity was measured after method EG (40 mA. mgPt-1). This value is even higher than that of the chosen carbon based electrocatalyst reference, TEC10E40E, measured in the same conditions (23.4 mA. mgPt-1). Half-cell and MEA singlecell test measurements will also be presented to complete the characterization panel.

The work presented here is funded by the European Union's Seventh Framework Program for the Fuel Cells and Hydrogen Joint Technology Initiative under grant agreement n325239 (FCH-JU project Nano-CAT) and the French National Research Agency PROGELEC programme, (ANR-12-PRGE-007 project SURICAT). It was supported by Capenergies and Tenerrdis.

Figure 1. TEM image of a Pt/Sb doped (10 at%) SnO2 aerogel from method A

REFERENCES

1. M. Ouattara-Brigaudet, , S. Berthon-Fabry, C. Beauger, M. Chatenet, N. Job, M. Sennour, Influence of the carbon texture of platinum/carbon aerogel electrocatalysts on their behavior in a proton exchange membrane fuel cell cathode. International Journal of Hydrogen Energy37, 9742-9757 (2012).

2. M. Ouattara-Brigaudet, C. Beauger, S. Berthon-Fabry, P. Achard, Carbon Aerogels as Catalyst Supports and First Insights on Their Durability in Proton Exchange Membrane Fuel Cells. Fuel Cells11, 726-734 (2011).

3. G. Ozouf, C. beauger, Niobium- and antimony-doped tin dioxide aerogels as new catalyst supports for PEM fuel cells, Journal of Materials Science 51, 11, 5305-5320, DOI 10.1007/s10853-016-9833-7 (2016)

Figure 1

2794

, , , , and

Pt nanoparticles supported on carbon black (Pt/C) are widely used as polymer electrolyte fuel cell (PEFC) electrocatalysts. However, under the cathode high potential, carbon corrosion can occur leading to cathode electrocatalyst degradation. In addition, when air flows into the anode at the fuel starvation, carbon corrosion also occurs at the anode as well as at the cathode [1]. Therefore, it is desired to develop alternative electrocatalyst support materials. Metal oxide is one of the promising candidates. Especially, titanium oxide (TiO2) possesses excellent stability in PEFCs environment in both anode and cathode. Pt/TiO2 electrocatalyst has high durability, but the initial catalytic activity is much lower than that of Pt/C because the electrical conductivity of TiO2 is too low around room temperature [2],[3]. In this study, we use SrTi(Nb)O3 with higher electrical conductivity compared to TiO2. SrTi(Nb)O3 is a perovskite-type oxide with strontium ions on the A-site and titanium ions on the B-site. As strontium ions can easily dissolve in acid, the surface of SrTi(Nb)O3 will become titanium-rich by applying an acid treatment. In this way, we can prepare "Core-Shell support" consisting of TiO2-based "shell" with high stability in strongly-acidic PEFC environment and SrTi(Nb)O3 "core" with a high conductivity. The aim of this study is to develop such electrocatalyst with both high durability and activity using TiO2–based support.

We made the acid treatment using HClO4 to SrTi(Nb)O3 prepared by sol-gel method to remove strontium ions from the surface of SrTi(Nb)O3. To impregnate Pt particles on this core-shell support (SrTi(Nb)O3 core - Ti(Nb)O2 shell), Pt(acac)2 method were applied. We then obtained Pt/SrTi(Nb)O3-Ti(Nb)O2.

SrTi(Nb)O3 before and after the acid treatment was characterized by XPS to analyze the titanium-rich surface. We made half-cell tests to evaluate electrochemical activities of these electrocatalysts. Electrochemical surface area (ECSA) was evaluated by cyclic voltammetry (CV), and oxidation reduction reaction (ORR) activity was derived from kinetically controlled current (ik) in rotating disk electrode (RDE) measurements.

Figure 1 shows the XPS spectra of SrTi(Nb)O3. It is confirmed that the surface of SrTi(Nb)O3 become Ti-rich as the Ti(2p) spectrum after the acid treatment was larger than that before the acid treatment. Figure 2 shows FE-SEM micrograph of the Pt/SrTi(Nb)O3 core - Ti(Nb)O2 shell electrocatalyst. This micrograph shows the highly dispersed impregnation of several-nanometer Pt particles on this support. We will report electrochemical activities in our presentation.

References

[1] A. Taniguchi, T. Akita, K. Yasuda, and Y. Miyazaki, J. Power Sources, 130, 42-49 (2004).

[2] Y. Takabatake, Z. Noda, S.M. Lyth, A. Hayashi, and K. Sasaki, Int. J. Hydrogen Energy, 39, 5074-5082 (2014).

[3] T. Kuroki, K. Sasaki, H. Kusaba, Y. Teraoka, Extended Abstract#1527, 206th Meeting of Electrochem. Soc., Hawaii, 2004. (http://www.electrochem.org/dl/ma/206/pdfs/1527.pdf).

Figure 1

2795

, and

 

Introduction

Carbon monoxide (CO) in hydrogen fuel is known to degrade a polymer electrolyte fuel cell (PEFC) performance [1-5]. The maximum allowable concentration of CO is 0.2 ppm, according to the quality standards for hydrogen fuel (ISO14687-2) in fuel cell vehicles (FCVs) [6]. Normally, a single-cell evaluation system is a hydrogen one-way pass system, which allows fuel unused by the cell to be released into the atmosphere. However, FCVs usually have a hydrogen circulation system, which returns the fuel from the anode outlet to the inlet. A comparison between the hydrogen one-way pass system and hydrogen circulation system has not been reported with respect to the effect of CO on hydrogen fuel. Our previous study showed that CO is not accumulated in hydrogen fuel when using a hydrogen circulation system [2]. However, the effect of CO on the actual FCV condition has not yet been fully understood because the experimental conditions in our previous study (CO concentration was 4.8 ppm and anode platinum loading was 0.4 mg cm−2) were far from the actual FCV conditions. In this study, using a hydrogen circulation system, the effect of CO on the PEFC's performance degradation is investigated and is compared to that using a hydrogen one-way pass system under the conditions of low platinum loading at the anode and a low CO concentration.

 

Experimental

A JARI standard single cell (25 cm2 electrode area), whose temperature was controlled by a heating medium, and commercially available membrane electrode assemblies (MEA) were used for the single-cell tests. The platinum loading on the anode and the cathode were 0.1 and 0.4 mg cm−2, respectively. The electrolyte membrane thickness was 15 μm. The cell temperature was 60 ºC, the anode was non-humidified, and the cathode dew point was 40 ºC. The single-cell operation tests were conducted using both a conventional hydrogen one-way pass system and the hydrogen circulation system (Fig. 1). The cell was operated at a constant current density of 1000 mA cm−2. The anode gas was hydrogen mixed with CO (up to 1.0 ppm), and the cathode gas was a mixture of N2 (79%) and O2 (21%). The stoichiometry of the fuel and the oxidant was 2 and 2.5, respectively. The CO and CO2 in the purge gas in the hydrogen circulation system were sampled at a 1% purge rate and measured by gas chromatography coupled with pulsed discharge helium ionization detector.

 

Results and Discussion

Figure 2 shows, for comparison, the voltage change by adding CO at a steady state in the hydrogen one-way pass system and in the hydrogen circulation system and the CO concentration in the hydrogen circulation system. The voltage change in the hydrogen circulation system is smaller than that in the hydrogen one-way pass system. At 0.2 ppm CO concentration in the hydrogen circulation system, the voltage change induced by adding CO is about 1/10 of that incurred when the hydrogen one-way pass system is used. The CO concentration in the purge gas in the hydrogen circulation system is lower than that in the supplied CO/H2 mixture. Most of the CO in the hydrogen fuel is oxidized to CO2 and gets accumulated in the hydrogen circulation system when a low concentration of CO is supplied to the cell. When the hydrogen circulation system is used, the exhaust gas is sent to the anode inlet again. Consequently, the CO concentration at the anode inlet decreases because the gas in the cylinder, which contains CO, is diluted by the gas from the hydrogen circulation system. In addition, our previous study of the hydrogen one-way pass system revealed that a large excess of O2, compared with CO, existed in the anode [3]. O2 in the hydrogen circulation system is probably circulated and consumed by CO oxidation in the anode electrocatalyst via a non-electrochemical reaction [4]. Consequently, the voltage change in the hydrogen circulation system is smaller than that in the hydrogen one-way pass system.

 

Acknowledgements

This work was supported by the New Energy and Industrial Technology Development Organization (NEDO).

 

References

[1] T. E. Springer, T. Rockward, T. A. Zawodzinski, S. Gottesfeld, J. Electrochem. Soc., 148 (1) A11–A23 (2001).

[2] Y. Matsuda, Y. Hashimasa, D. Imamura, M. Akai, S. Watanabe, Rev. Automot. Eng.30, 167–172 (2009).

[3] Y. Matsuda, T. Shimizu, S. Mitsushima, J. Power Sources, 318 (30) 1–8 (2016).

[4] T. V. Reshetenko, K. Bethune, R. Rocheleau, J. Power Sources, 218, 412–423 (2012).

[5] Y. Hashimasa, Y. Matsuda, M. Akai, ECS Trans.,26(1), 131–142 (2010).

[6] ISO 14687-2 (2012).

Figure 1

2796

, , , , , , , and

Development of Pt-based catalysts with higher activity and durability is essential for implementation of cost-competitive polymer electrolyte membrane fuel cells. Nanoparticle catalysts with core-shell, skeleton, or skin-type structures have shown to outperform Pt nanoparticles. However, complications arise when such catalysts are exposed to harsh operation conditions of the fuel cell where Pt and the alloying metal undergoes consecutive oxidation/reduction and dissolution, de-activating the catalyst. Compared to nanoparticles, extended surfaces (bulk electrodes and nanostructured thin films) are more stable, but suffers from low surface area.

In this study, we describe the concept of "nanosheet catalysts" with surface area, activity, and durability that are higher than conventional core-shell nanoparticles for the oxygen reduction reaction as well as the hydrogen oxidation of reformate fuel. Our approach utilizes metallic Ru nanosheets [1] with atomic-scale thickness prepared via thermal reduction of exfoliated RuO2nanosheets [2,3]. The metallic Ru nanosheet is used as a core for the synthesis of two-dimensional Ru-core@Pt-shell catalysts.

Sub to a few monolayer Pt shell was successively formed on metallic Ru nanosheets with monoatomic thickness via galvanic displacement reaction between Cu and Pt2+. The electrochemically active Pt surface area of Ru-core@Pt-shell nanosheets with 1.5-4.5 monolayer Pt-shell was 112-151 m2 (g-Pt)1. This corresponds to Pt nanoparticles with 1-1.5 nm of diameter. A carbon supported core-shell nanosheet catalyst with a 3.5 monolayer Pt-shell (Ru@Pt-3.5ML(ns)/C) showed 4.5 times higher activity than the benchmark Pt/C catalyst for the oxygen reduction reaction with a slower degradation rate even at high potentials. For the anode reactions, Ru@Pt-1.5ML(ns)/C had 2 times higher hydrogen oxidation activity in pure H2 as well as 300 ppm CO containing H2, and better stability against potential cycling. The developed nanosheet catalysts should provide the high utilization of Pt towards catalytic reactions and concurrently the stability of extended surfaces, offering a practical solution to the trade-off issue.

This work was supported in part by the "Polymer Electrolyte Fuel Cell Program" from the New Energy and Industrial Technology Development Organization (NEDO), Japan.

[1] K. Fukuda and K. Kumagai, e-Journal Surf. Sci. Nanotechnol., 12, 97 (2014).

[2] W. Sugimoto, H. Iwata, Y. Yasunaga, Y. Murakami and Y. Takasu, Angew. Chem. Int. Ed., 42, 4092 (2003).

[3] K. Fukuda, H. Kato, J. Sato, W. Sugimoto and Y. Takasu, J. Solid State Chem., 182, 2997 (2009).

Figure 1

2797

, , , and

For residential co-generation polymer electrolyte fuel cell (PEFC) systems operated with reformates, the mass activity for the hydrogen oxidation reaction (HOR) at the state-of-the-art commercial Pt-Ru/C anode catalyst is insufficient, especially when operated under high CO concentration (> 500 ppm) in a fuel processing system simplified for cost reduction. Recently, we have developed a novel CO-tolerant catalyst by forming two uniform atomic layers of stabilized Pt skin on Pt-Co alloy nanoparticles (Pt2AL–PtCo/C) and found its superlative CO-tolerant HOR activity together with high robustness with respect to air exposure.1

In the present research, to establish a clear strategy for designing CO-tolerant catalysts, we have examined the effect of the nonprecious metal species M (M = Fe, Co, Ni) in Pt2AL–Pt-M/C on the CO-tolerance and the robustness.

The Pt2AL–PtFe/C, Pt2AL–PtCo/C, and Pt2AL–PtNi/C were prepared by the use of a high-surface-area carbon black support (specific surface area = 780 m2 g−1) in the same manner as that described previously.2, 3 Two commercial catalysts, c-Pt2Ru3/C and c-Pt/C, were used for comparison. The experimental procedure, in which the channel flow electrode cell (CFE) was used, was the same as that described in ref. 1. Each catalyst was uniformly dispersed on an Au substrate as the working electrode with a constant loading of carbon support of 11 µg cm−2, which corresponds to approximately two monolayers in height of the carbon black particles. Nafion film was coated on the catalyst layer with an average thickness of 0.075 µm. All electrode potentials were referred to the reversible hydrogen electrode (RHE).

Figure 1 shows the apparent mass activities, MAapp, at 20 mV vs. RHE for the HOR at various catalysts in H2-purged 0.1 M HClO4 solution at 70°C and 90°C without (tad = 0, CO-free) and with adsorbed CO (tad = 90 min at E = 50 mV in 1000 ppm CO/H2 saturated solution). For the HOR activity on the CO-free surface of the catalysts (tad = 0), the Pt2AL–Pt-M/C was found to exhibit higher MAapp than those of c-Pt/C and c-Pt2Ru3/C. The largest MAapp was seen at Pt2AL–PtFe/C, 330 A gmetal−1 at 70°C, which was about 1.8 times higher than that of c-Pt/C and 2.5 times higher than that of c-Pt2Ru3/C. All of the catalysts showed a decrease in MAapp at 90°C compared to that at 70°C, which is ascribed mainly to the decreased H2 concentration in the solution at high temperature, but the trend of the MAappvalues was unchanged.

After 90-min CO poisoning (tad = 90 min) at 70°C, as shown by the black bar for the c-Pt/C electrode, MAapp decreased greatly, due to a decrease in the hydrogen adsorption sites blocked by strongly adsorbed CO. In contrast, the Pt2AL–Pt-M/C and c-Pt2Ru3/C catalysts exhibited excellent CO-tolerance, suggesting that the HOR active sites were not so rigidly blocked by CO, due to its enhanced mobility. Specifically, the CO tolerance increased in the order Pt2AL–PtNi/C < c-Pt2Ru3/C < Pt2AL–PtCo/C < Pt2AL–PtFe/C.

At 90°C, the CO tolerance for all the catalysts was improved. The Pt2AL–Pt-M/C still showed higher MAapp than those of c-Pt/C and c-Pt2Ru3/C, e.g., the MAapp at Pt2AL–PtFe/C was 3.4 times higher than that of c-Pt/C and 2.8 time higher than that of c-Pt2Ru3/C. Thus, in the practical temperature range, the highest CO-tolerant HOR activity was demonstrated for Pt2AL–PtFe/C. The effect of particle size and the optimization of preparation methods are in progress in our laboratory.

This work was supported by funds for the "SPer-FC" project from NEDO of Japan.

References

1. G. Y. Shi, H. Yano, D. A. Tryk, M. Watanabe, A. Iiyama and H. Uchida, Nanoscale, 2016, doi: 10.1039/C6NR00778C.

2. M. Chiwata, H. Yano, S. Ogawa, M. Watanabe, A. Iiyama, and H. Uchida, Electrochemistry, 84, 133 (2016).

3. M. Watanabe, H. Yano, D. A. Tryk, and H. Uchida, J. Electrochem. Soc., 163, F455 (2016).

Figure 1

2798

, , , and

Electrochemical CO oxidation (CO + H2O → CO2 + 2H+ + 2e) is important to counteract CO poisoning of Pt-based anode catalysts, since the reaction can reduce CO concentration on an anode of PEFC. However, the CO oxidation at anode potentials hardly occurs on conventional Pt-Ru electrocatalysts.

From these backgrounds, we have developed Rh-porphyrin-based electrocatalysts for CO oxidation [1, 2]. In the catalysts, molecules of Rh porphyrin are dispersed on a carbon black support. These catalysts can oxidize CO at much lower overpotentials than the Pt-Ru electrocatalysts [1-3]. A membrane-electrode assembly (MEA) that uses a Rh porphyrin as an anode catalyst gave significant power (>40 mW/cm2) when CO was supplied as a fuel [4]. A combined catalyst of a Rh porphyrin and Pt-Ru catalyst functions as a CO-tolerant anode catalyst [5].

However, the reason why Rh porphyrins can catalyze CO electro-oxidation at low overpotentials remains unclear. Mechanistic analyses on carbon-supported Rh porphyrin are needed to clarify the reason. The analysis would also contribute to further improvement of this catalyst. In this work, we report a possible reaction mechanism of CO oxidation by the catalyst. First, we examined the reactivity of Rh porphyrins dissolved in solution. This analysis enables us to discuss the CO activation mechanism by Rh porphyrins without the effect of carbon black support. A Rh porphyrin was treated with CO gas in dichloromethane to generate CO adduct of the Rh porphyrin. We characterize this CO-adduct, a possible intermediate of CO oxidation, by IR, NMR, and X-ray crystallography. The results indicate that CO coordinated on Rh atom is feasible to nucleophilic attack by water. Actually, the CO-adduct reduces an soluble electron acceptor in the presence of water. A possible reaction mechanism is shown in Fig. 1. An electrode works as an electron acceptor in the electrocatalytic CO oxidation.

In actual electroctalysts using Rh porphyrins, molecules are adsorbed on a carbon black support. Configuration of molecules on a carbon support would be also important for the catalytic activity. Then, we examined possible effects of the configuration by observing molecules on a highly oriented pyrolytic graphite with an atomic force microscope. The results show that certain ligand structure facilitates the configuration desirable for the high catalytic activity.

These analyses would be an answer for the question; why does Rh porphyrin can oxidize CO at lower potentials than conventional Pt-based catalysts.

This work was supported by Grant-in-Aids for Scientific Research (B) (No. 15H03853).

References

[1] S. Yamazaki et al., Phys. Chem. Chem. Phys. (2010) 12,8968.

[2] S. Yamazaki et al., Electrochem. Solid-State Lett. (2011) 14,B23.

[3] J. F. van. Baar et al., Electrochim. Acta (1982) 27,57.

[4] S. Yamazaki et al., Angew. Chem. Int. Ed. (2006) 45,3120.

[5] S. Yamazaki et al., J. Phys. Chem. C (2010) 114, 21856.

Figure 1

2799

, , , and

Design of highly durable, electro-active and cost-effective catalyst to replace prevalent Pt has been a major issue to researchers in a polymer electrolyte membrane fuel cell (PEMFC) community. For past decades, the catalyst degradation, during the transient conditions, such as start-up/shut­down and cell reversal, in automotive fuel cells has gained large attention due to its irreversible consequences in the membrane electrode assembly (MEA). 

It was reported earlier that binary aIloy of Ir and Ru supported on carbon can be used for hydrogen oxidation reaction (HOR) catalysts for PEMFCs. Among various Ir:Ru atomic compositions, Ir:Ru=1:4 showed the best HOR activity, confirmed by rotating disk electrode (RDE) half-cell test. Furthermore, Ir and Ru are known to possess water oxidation properties, in other words, each can catalyze the oxygen evolution reaction (OER) from the water. Due to this OER-enabling feature, the Ir and Ru were often used as additives to promote the water oxidation over carbon oxidation reaction in order to protect the catalyst layer from the collapse when the catalyst layer was abruptly exposed to a high potential for various reasons. The Ir and Ru exhibit oxygen reduction reaction (ORR) activity to certain extent, however they typically show inferior ORR characteristics compared to the platinum so they are scarcely applied to the cathode catalysts in MEAs.

In this presentation we propose to utilize multi­functional IrRu4 nanoparticles supported on carbon, IrRu4/C, as an anode catalyst for MEAs, particularly to be used in the fuel cell electric vehicles (FCEVs). IrRu­4's material costs only as high as 40% of widely used platinum. Therefore during the normal fuel cell operation, IrRu4/C can be used as anode electro-catalyst at a reduced cost while showing the Pt-similar performance. And during the transient conditions of FCEV operations, the MEA durability can be retained with IrRu4's other interesting properties.

IrRu4/C was synthesized by a simple impregnation method using metal salts and carbon support and a successive reduction in a hydrogen atmosphere at higher temperature. The synthesized catalysts were characterized with XRD, ICP, TGA, and TEM for physicochemical properties. And it underwent the RDE half-cell tests for various electrochemical activity measurements, such as HOR, OER, and ORR. IrRu4/C and Pt/C were each applied to the anode in the MEA, and the single-cell IV performance and anode polarization tests were carried out at various operating conditions (cell temperature, relative humidity, and back pressure). The cell reversal tolerance of IrRu4/C and commercial Pt/C anode MEAs was also measured by subjecting each MEA to the anode's fuel starvation condition. The IrRu4/C anode catalyst showed Pt-anode comparable MEA performance and Pt-similar hydrogen oxidation catalytic activity. Moreover, IrRu4/C anode exhibited superior durability (~120 times better) over Pt/C anode under cell reversal condition. This is because during the cell reversal condition, IrRu4 promoted water oxidation reaction so that carbon oxidation reaction was avoided. For Pt anode MEA, since Pt has low OER activity, the catalyst layer oxidation took place to severely impact the MEA during the fuel starvation.

The prospective benefits earned from this work, that is replacing Pt/C anode to IrRu4/C anode for a PEMFC in a FCEV are the following: firstly, the cost down of the MEA while preserving the performance, secondly the anode catalyst layer protection during the cell reversal, thirdly, although not studied in detail in this work, the cathode catalyst layer protection during the start-up/shut-down.

A-32 Catalyst Layer 2 - Oct 6 2016 1:40PM

2800

, , , and

In this study, a set of experimental techniques have been adopted to characterize the fundamental properties of machine prepared catalyst layers (CLs) formed by different Pt:C ratios, highly graphitized carbon and 3M ionomer (825EW). Single cell performance with these catalyst layers was evaluated in order to analyze the relationship with the structural properties.

Scanning electron microscopy (SEM), transmission electron microscopy (TEM) and Brunauer-Emmett-Teller (BET) nitrogen adsorption were conducted in a series of CLs with Pt:C ratio variations at a fixed ionomer:carbon (IC) ratio to investigate the microstructure including materials dispersion and porosity. This analysis was performed on both the CL in a fresh (as-prepared) condition and after subjected to a standard PEMFC protocol.

TEM images along with Pt size distributions of the 50:50 Pt:C fresh and post-operation samples are shown in Fig. 1. A significant widening of the Pt size distribution and a possible tendency towards a multimodal distribution is observed after FC operation. This behavior can be explained by a combination of Pt particle coalescence and dissolution/re-precipitation processes within the solidionomer, as has been suggested in previous studies.1

In order to assess the ionomer behavior in the machine prepared CLs, proton conductivity measurements and estimation of the proton conduction tortuosity was performed as a function of relative humidity (RH) in the single cell. Water uptake isotherms of the as-prepared CLs have also been acquired under well controlled RH levels. In contrast to typical Nafion®-based fuel cells, the 3M fuel cell reached its optimum performance at 60%-70% RH and suffered dramatic mass transfer losses at saturated humidity levels. Therefore, the 3M fuel cells are able to function fully under drier conditions than the Nafion units. However, the high sensitivity to the cells' water content requires efficient water management during operation, especially at higher current densities.

References

  • More, K.L., Borup, R., and Reeves K.S. ECS Transactions, 3 (1) 717-733 (2006)

Figure 1

2801

, , and

Polymer electrolyte membrane (PEM) fuel cells have been demonstrated as a promising technology in meeting some of the challenges faced with shifting towards a clean and renewable energy economy. High Temperature PEM (HTPEM) fuel cells provide an attractive alternative to low temperature PEM fuel cells with respect to fuel flexibility, simplified cooling system, balance of plant and higher value waste heat without reactant humidification (1). Furthermore, the higher operating temperature range of 140-180 °C dramatically increases the tolerance of the noble metal catalysts to impurities in the reactants at both the anode and cathode. To date, most of the success in the field of HT-PEM fuel cell development has been realized through the implementation of a phosphoric acid-doped polybenzimidazole (PBI) membrane electrolyte [1]. The use of such a material allows for operating temperatures in the range of 140-180 °C.

We have previously presented work regarding the effect of m-PBI molecular weight on membrane oxidative stability and MEA performance [2,3]. In the present work, we seek to provide an update regarding challenges related to the optimization of the properties of the catalyst layers, especially the properties of the cathode as this have a significant impact on the performance and durability of HTPEM fuel cells. The performance and Pt utilization in the cathode is highly depending on parameters like porosity, pore size distribution, hydrophilic/hydrophobic properties, catalyst and acid content. In this communication we will focus on the performance of two different catalysts by varying the hydrophobic/hydrophilic properties and the Pt loading..

The hydrophobic/hydrophilic properties are modified by addition of PTFE (increased hydrophobicity), PBI (increased hydrophilicity) or "binderless" (neutral) to the cathode. The loading is varied by adding more catalyst to the catalyst layer and thus making it thicker. The two different catalysts are Pt/C and PtCo/C. Figure 1 shows examples of the performance during the first ~100 hours of operation for a HTPEM MEAs using hydrophilic cathodes.

Figure 1: Performance of HTPEM MEAs using cathodes with increased hydrophilicity and Pt/C catalyst. Pt loading on anode and cathode: 0.9 mg/cm2 on anode. λH2 and λair  was 1.5 and 2.5, respectively. Temperature: 160 °C.

The performance during the first ~100 hours of operation of HTPEM MEAs using these hydrophilic, hydrophobic and binderless cathodes is evaluated under the hydrogen/air mode at 160 °C and the results will be presented. The correlation of the MEA performance with the catalyst layer structures is explored and further optimization is outlined in order to achieve high performance while reducing the Pt loading.

References

[1] Li, Q., Jensen, J. O., Savinell, R. F. & Bjerrum, N. J. High temperature proton exchange membranes based on polybenzimidazoles for fuel cells. Prog. Polym. Sci., 2009, 34, 449–477.

[2] Valle, F., Zuliani, N., Marmiroli, B., Amenitsch, H. & Taccani, R. SAXS Analysis of Catalyst Degradation in High Temperature PEM Fuel Cells Subjected to Accelerated Ageing Tests. Fuel Cells, 2014. 14, 938–944.

[3] Lifetime and degradation of high temperature PEM membrane electrode assemblies, R. Kerr, H.R. García, M. Rastedt, P. Wagner, S.M. Alfaro, M.T. Romero, C. Terkelsen, T. Steenberg and H.A. Hjuler, International Journal of Hydrogen Energy, 2015, DOI: 10.1016/j.ijhydene.2015.07.152, in press.

Figure 1

2802

, , , , and

The electrocatalyst layer of the polymer electrolyte fuel cells (PEFCs) has a complicated porous microstructure. Since the structure allows gas, proton, and electron transport, and water management, cell performance depends strongly on their 3-dimensional microstructure. Understanding on the correlations between the electrocatalyst microstructure and the cell performance is therefore essential in improving the PEFC electrochemical performance. The purpose of this study is thus to clarify these correlations using focused-ion-beam coupled scanning electron microscopy (FIB-SEM).

We used the standard electrocatalyst (TEC10E50E) and varied the Nafion-to-electrocatalyst ratio (hereafter the Nafion ratio) for the electrocatalyst layer because the Nafion ratio affects the electrocatalyst microstructure and the cell performance[1],[2]. After model cell preparations, current-voltage (IV) performance was measured and thus each electrode overvoltage contribution was separated. We then considered FIB-SEM working condition and method of image processing to observe electrocatalyst microstructure by the FIB-SEM technique. In particular, because it was difficult to distinguish between solid and pore in SEM images due to the complicated porous microstructure, we solved this problem by separating solid and pore objectively by auto-image thresholding technique[3]. Then, we examined the correlations between the 3D microstructure and the cell performance.

According to the results of IV characteristics of several cells with different Nafion contents, in case Nafion ratio was 28 wt.%, IV performance was the highest. In case Nafion ratio was less than 28 wt.%, whilst IV performance was lower than that of 28 wt.%, the cells can still generate high current densities. On the other hand, in case Nafion ratio was higher than 28 wt.%, the cells can exhibit a lower cell voltage due to a much higher concentration overvoltage. In order to examine the correlations between the 3D microstructure and the cell performance, we observed electrocatalyst microstructure by the FIB-SEM. Figure 1 describes the observation process. According to the results of 3D analysis of samples with different Nafion contents, we could confirm that gas transport pathway in the electrocatalyst layer was filled with Nafion ionomer with increasing Nafion ratio. Details on the FIB-SEM observation conditions and the microstructure-performance correlations are presented at the symposium.

(1) S. Jeon, J. Lee, G.M. Rios, H.-J. Kim, S.-Y. Lee, E Cho, Intl. J. Hydrogen Energy 35 (2010) 9678-86.

(2) T. Suzuki, S. Tsushima, S. Hirai, Intl. J. Hydrogen Energy 36 (2011) 12361-12369.

(3) J. Bernsen, 8th ICPR (1986) 1251-1255.

Figure 1

2803

, , , , , and

The membrane electrode assembly (MEA) is the core component of polymer electrolyte membrane fuel cells (PEMFCs) and directly determines the performance and cost of PEMFCs. Recent research has been focused on developing high-performance, low-cost, and durable MEAs. The dispersion of the Pt/C aggregates and the ionomer particles in a solvent is the first and very critical step to achieve a fine structure of the catalyst layer (1-6). In order to prepare an appropriate ink, the Pt/C catalyst powder must be dispersed well in a solvent. Therefore, to verify whether a formulated catalyst ink disperses the Pt/C aggregates and the ionomer particles well is of great importance. In a catalyst ink, the solvent plays a critical role to disperse both Nafion ionomer and Pt/C catalyst particles. The polarity of the solvent is indicated by its dielectric constant. For instance, when the Nafion ionomer is mixed with a variety of organic solvents, the geometry and the micro-structure of both the Nafion ionomer and Pt/C particles tends to change (7). These changes are strongly associated with the surface energy of solid/liquid, to great extent, due to the dielectric constant (ε) of an organic solvent. Therefore, it is of great interest to study the effects of different organic solvents on the dispersion of both Nafion ionomer and Pt/C catalyst particles. Instead of a "trial-and-error" approach—measuring the MEA fuel cell performance of different ink formulation, designing an "ideal" ink guided by the knowledge of ionomer and catalyst particles in different solvents is a rational approach.

In 2010, our group developed a unique method of combined the ultra-small angle x-ray scattering (USAXS) and cryo-TEM to study the size and geometry of both the Pt/C aggregates and the ionomer particles in catalyst inks and the effects on dispersion in catalyst ink (8). USAXS can be used to measure the size and geometry of Nafion ionomer particles and Pt/C aggregates in liquid media without the issues of incident beam absorption by inks from using light scattering. With direct observation of particle dispersion from cryo-TEM imaging, by which the solid particles of Nafion ionomer and carbon aggregates in a liquid ink are frozen to lock their geometry and particle size distribution for TEM imaging, the USAXS fitting data can be well validated.

In this work, three different solvents with different dielectric constants, (1) ethanol (ETH), (2) glycerol (Glyc) and (3) mixture of isopropanol and water (1:4) were used to study the effect of different solvents on the dispersion of carbon particles in the ink systems. The USAXS results of carbon particles in different solvents are shown in figure 1. It can be seen that the size of carbon aggregate in solvent with medium dielectric constant (glycerol) decreases after adding Nafion. However, the carbon black aggregates sizes in solvents with higher and lower dielectric constant decrease after adding Nafion ionomer. This significant difference indicates that solvents have an important role in ink formulation.

From the results above, it is very clear that solvent plays an important role in ink formulation which affects the geometry and size of both Nafion and carbon particles, consequently, the structure of catalyst layer, which ultimately determines the performance of MEA. This also demonstrate that the unique method of USAXs combined with cryo-TEM is an effective means to study the ink formulation.

Reference

1. Z.-F. Li, L. Xin, F. Yang, Y. Liu, Y. Liu, H. Zhang, L. Stanciu and J. Xie, Nano Energy, 16, 281 (2015).

2. M. S. Wilson and S. Gottesfeld, Journal of Applied Electrochemistry, 22, 1 (1992).

3. J. Xie, F. Garzon, T. Zawodzinski and W. Smith, Journal of The Electrochemical Society, 151, A1084 (2004).

4. J. Xie, K. L. More, T. A. Zawodzinski and W. H. Smith, Journal of The Electrochemical Society, 151, A1841 (2004).

5. J. Xie, F. Xu, D. L. Wood Iii, K. L. More, T. A. Zawodzinski and W. H. Smith, Electrochimica Acta, 55, 7404 (2010).

6. L. Xin, F. Yang, S. Rasouli, Y. Qiu, Z.-F. Li, A. Uzunoglu, C.-J. Sun, Y. Liu, P. Ferreira, W. Li, Y. Ren, L. A. Stanciu and J. Xie, ACS Catalysis, 6, 2642 (2016).

7. M. Uchida, Y. Aoyama, N. Eda and A. Ohta, Journal of The Electrochemical Society, 142, 463 (1995).

8. F. Xu, H. Zhang, J. Ilavsky, L. Stanciu, D. Ho, M. J. Justice, H. I. Petrache and J. Xie, Langmuir, 26, 19199 (2010).

Figure 1

2804

, , , and

Tiny optical fiber sensor was applied to realize non-destructive measurement of oxygen concentration on PEFC catalyst layer.

Insufficient supply of oxygen to the cathode catalyst layer caused by transport resistance is critical issue. Fundamental study in order to achieve efficient oxygen transport is required, however; non-destructive and in-situ measurement of oxygen concentration on the catalyst layer was not achieved. In this study, oxygen concentration on the catalyst layer under operating condition was measured by using a tiny optical fiber and oxygen indicator.

Figure 1 shows experimental setup. Platinum tetrakis pentrafluoropheny porphine (PtTFPP) was employed as oxygen indicator[1]. When the PtTFPP is exposed to excitation light (l = 405 nm), phosphorescence emission (l = 650 nm) is produced. Phosphorescence intensity is determined by oxygen partial pressure. Thus, quantitative oxygen concentration can be obtained by measuring phosphorescence intensity and using calibration data. In this study, PtTFPP was painted at the edge of fused silica optical fiber of which diameter was 110 mm. In order to measure the phosphorescence intensity, bifurcated optical fibers were connected to excitation light source and spectrometer, respectively.

The cell has an active area of 4 cm2 (2 × 2 cm) with straight channels. The channel width and depth were 1.0 and 1.0 mm, respectively, and the rib-to-channel ratio was 1. The carbon paper GDL without MPL (SIGRACET®24BA, SGL Group, USA) was used for both the anode and the cathode side. In order to measure the oxygen concentration on the catalyst layer with reducing insertion effect, optical fiber was inserted from the intrinsic large pore of GDL without boring a hole, and the edge of the fiber was touched on the catalyst layer surface. Flow rates of hydrogen and air were 50NmL/min and 125 NmL/min, respectively. Cell was operated under 33 °C because phosphorescence intensity becomes low under high temperature. Bubbler temperature was 22 °C (relative humidity was 50%)

In the experiment, cell voltage was varied from 0.9 V to 0.2 V, and change of current density and oxygen concentration were measured. Figure 2 shows current-voltage (IV) characteristics. In the higher current density condition, variation of plots suggests emergence of flooding in the cathode. Since the flooding decreased gas diffusivity in the cathode catalyst layer, IV characteristics fluctuate at high current density and low cell voltage condition. As shown in Figure 3, change of oxygen concentration on the catalyst layer was measured successfully. Oxygen concentration decreases monotonically with decreasing cell voltage. Minimum value of the concentration was below 2 % at 0.2 V cell voltage.

Figure 2 indicated that total cell performance was declined by flooding. On the other hand, local oxygen concentration of the catalyst layer was able to be measured by using a tiny optical fiber. Figure 3 shows large amount of oxygen was still consumed under high current density and it means that power generation concentrated on local surface of the catalyst layer where surface not covered with the liquid water.

As discussed above, oxygen concentration on cathode catalyst layer was measured non-destructively. This technique holds further potential to measure oxygen transport phenomena on the catalyst layer under 60-80 °C cell operation by improving optical oxygen sensor sensitivity.

 

Acknowledgement

This study is based on results obtained from a project commissioned by the New Energy and Industrial Technology Development Organization (NEDO). Experimental setup in present study was assisted by Yoshihiko Aoki (Tokyo Institute of Technology).

 

Reference

[1] CS Chu., CA Lin, Sensors and Actuators B: Chemica, 2014, 195, 259-265.

Figure 1

2805

and

Automotive Proton Exchange Fuel Cells (PEFCs) are approaching technology readiness levels where personal transportation manufacturers are predicting eminent preliminary adaptation into the commercial market within the next decade. To further decrease PEFCs cost, reductions in platinum (Pt) catalyst loadings on the cathode are being sought. As the structure and the volume of the cathode catalyst layer (CL) are altered, a tangential impact will be made to sub-zero cold-start performance and durability will be realized.1 Automotive cold-starts below 0°C are challenging due to the inherent presence and production of water. The hydrophilic domains of the proton conducting ionomer in the CLs and the membrane must maintain a minimum hydration level of freezing point depressed water for operational proton conduction even at subzero temperatures. A sub-zero cold-start actively hydrates the ionomer due to production and mobility of freezing point depressed water. Saturation of the ionomer results in product water migration to the cathode open pore space which then likely exists as metastable supercooled water or ice. A successful cold-start is characterized by elevating the cell temperature above freezing for product water removal before the porous space within the cathode CL becomes blocked restricting mass transport of oxygen.

Sub-zero cold-start performance in PEFCs is impacted by preconditioning hydration levels, CLs/membrane materials, and sub-zero operational conditions. The amount, state, and connectivity of water is controlled by the purge process after operational shutdown prior to freeze. Multiple research groups have established that the product water fill capacity of the CLs/membrane is extended for lower relative humidity (RH) preconditioning levels, which is at odds with the minimum parasitic losses mandated by the US Department of Energy.2

The hydration state of the CLs/membrane below zero can be examined via electrochemical impedance spectroscopy (EIS) using symmetric 4% H2 feeds.3 The impedance spectra acquired at -20°C in Figure 1 compares the preconditioning RH for two distinct shutdown methods prior to freeze: (grey) capped immediately after operational 80°C polarization curve and (black) equilibrium purged with RH inert gasses. The capped tests retained additional water that remind non-frozen at -20°C as seen by the low resistance shift in the high frequency intercept and a smaller high frequency arc related to improved charge transfer resistance.

The subzero isothermal water storage capacity (WSC) is material processing dependent. Figure 2 compares the WSC at 10 mA/cm2 (-20°C) for two distinct catalyst layer (CL) fabrication methods with the same initial water content. The cathode CL prepared by decal transfer had a lower initial voltage, while the one directly sprayed onto the membrane suffered from limited WSC. Differences in initial voltage suggest significant CL ohmic contributions to overpotential. Differences in water storage capacity suggest CL structure differences. This may suggests a more ohmically connected, yet less porous electrode is produced using the direct spray.

Finally, cold-start testing (Figure 3) is being evaluated using a quasi-adiabatic single cell fixture developed in-house, expanding upon the original proto-type designed at United Technologies Corporation4,5. Heating pads were integrated into the fixture and powered at levels equivalent to 1x and 2x adjacent cells to compensate for non-adiabatic losses. Higher water content favorable impact initial performance, however at long time scales seems to insignificantly impact cold start probability. Cold start probability is significantly impacted by the heating pad condition indicating significance of bipolar plate and coolant thermal mass impact on ability for CL to self-heat to above 0°C.

References:

1. Nandy, A., F. Jiang, S. Ge, C.-Y. Wang, and K.S. Chen, 'Effect of Cathode Pore Volume on PEM Fuel Cell Cold Start', J. Electrochem. Soc., 2010, 157, B726.

2. US Department of Energy, Hydrogen and Fuel Cells Program. 2014; Available from: http://www.hydrogen.energy.gov/.

3. Pistono, A., C.A. Rice-York, and V. Boovaragavan, 'Electrochemical Impedance Spectroscopy Detection of Saturation Level in a Frozen Polymer Electrolyte Membrane Fuel Cell', Journal of The Electrochemical Society, 2011, 158, B233.

4. Patterson, T., J. O'Neill, M. Perry, P. Hagans, C. York, and R. Zaffou, 'PEM Fuel Cell Freeze Durability and Cold Start Project', Final Technical Report, 2006.

5. Rice-York, C., J. Needham, N. Gupta, and P.L. Hagans, 'Platform for Rapid Prototyping of PEM Fuel Cell Designs with Enhanced Cold-Start Performance and Durability', ECS Transactions, 2006, 1, 383.

Figure 1

2806

, , , and

Polymer electrolyte membrane fuel cells (PEMFCs) are energy conversion devices that offer high power densities at low operating temperatures making PEMFCs the most promising technology for many applications such as automobiles, back-up power generating units, and portable devices. While design and material considerations for PEMFCs have a large impact on cost, it is also necessary to consider a transition to high volume production of fuel cell systems, including MEA components, to enable economies of scale and reduce per-unit cost. The fuel cell industry has identified quality control as a critical barrier for continuous production of MEA components, i.e. membranes, electrodes, and GDLs. One of the critical manufacturing tasks is developing and deploying techniques to provide in-process measurement of fuel cell components for quality control. This work focuses on a necessary subsidiary task: The study of the effect of manufacturing irregularities on performance with the objective to establish validated manufacturing tolerances for fuel cell electrodes.

Membrane electrode assemblies with nominal active areas of 50 cm2 were prepared by spraying a catalyst ink formulation directly onto NRE212 polymer membrane material held at 80°C. The spray system was an ExactaCoat from Sono-Tek with an ultrasonic spray head. Catalyst loading variations were created by masking off 0.0625 – 1 cm2 areas. Within the defect area the nominal catalyst loading was reduced by various degrees up to 100% of the nominal value. The sample pool includes MEAs with different locations, shapes, and severity of coating irregularities, different nominal loading, as well as CCM vs. GDE structures. For spatial interrogation NREL's high resolution segmented cell system was employed. The system consists of 121 segments of 0.41 cm2 area each arranged along the path of a quadruple serpentine flow-field. The segmented cell system was operated with a state-of-the-art single cell fuel cell test station.

Figure 1 shows one sample set used to understand the performance effect of reducing total catalyst loading by 1% of the nominal loading. Rather than reducing catalyst loading in a single location only, variations of the 1% reduction were introduced to investigate if (i) location, (ii) shape, and (iii) intensity of the 1% coating variation matter. Differential spatial data was computed by subtracting the current distribution measured with a sample MEA from that of a pristine MEA. Figure 2 shows two such differential data sets. Areas with reduced performance are red, those with increased performance are blue, and those with unchanged performance are black. The data shows results for a sample with 2x 0.25 cm2 coating variations each having a 100% loading reduction, one near the inlet and one near the outlet (left) and for a second sample with a 1 cm2 50% reduction coating variation in the center. In addition, Figure 3 shows the total cell performance of the sample set. The total cell performance alone, as shown in Figure 3, does not indicate the presence of the defect. Instead, the detection of the impact of the defect requires a high resolution spatial diagnostic tool. As shown in Figure 2, the effect of the defect fades when reducing defect size. The areas that are 0.25 cm2 and have no catalyst loading show a smaller performance impact than the 1 cm2 area that only has a loading reduction of 50%. Obviously the impact of the coating variation has been distributed to more than one location. In any case, with respect to the total cell performance impact, the coating variation may not be classified as a defect, since no impact has been detected. A second criteria to classify the studied coating variation as defects is the effect on lifetime. These studies are currently underway and will be addressed in a second presentation.

Figure 1

2807

, , and

For the large-scale commercialization of polymer electrolyte fuel cells (PEFCs), it is very important to reduce the amount of Pt by developing highly efficient cathode catalyst layers (CLs) due to the high cost and limited availability of Pt resources. In previous work, we demonstrated the importance of two factors for the design of high performance cathode CLs. The first is higher effective Pt surface area (S(e)Pt), which practically contributes to the oxygen reduction reaction (ORR) [1]; the n-Pt/AB250 catalyst, for which all of the Pt particles exist only on the AB250 exterior surface, is the most attractive in order to generate the large current densities required by actual fuel cell operation [2]. The second is the optimized distribution of ionomer on the surfaces both of Pt particles and carbon particles; short-side-chain (SSC) perfluorosulfonic acid ionomers with high ion-exchange capacity (IEC) showed better continuity and uniformity on the Pt and graphitized carbon black (GCB) particles than the conventional ionomer, which might have led to the improvement of both the mass transport and the proton-conducting network in CLs [3].

In this study, we investigate the effects of high oxygen permeability ionomers on the cathode performance of PEFCs. The high oxygen permeability ionomers are expected to increase the flux of oxygen near the Pt surface of the three-phase boundary, in the case of extremely low Pt loadings [4]. The low Pt loading cathode CLs (ca. 0.05 ± 0.003 mg-Pt cm-2) were prepared from the AB250-supported Pt catalyst, with a higher effective Pt surface area (n-Pt/AB250: S(e)Pt = 107 m2 gPt-1, specific surface area of AB250: 219 m2 g-1, Denka Co., Ltd.) and ionomers developed by Asahi Glass Co., Ltd. (AGC); AGC PFSA (IEC = 1.13 meq g-1), AGC sample A (IEC = 1.17 meq g-1, O2 permeability 1.5 times higher than AGC PFSA) and AGC sample B (IEC = 1.50 meq g-1, O2 permeability 1.5 times higher than AGC PFSA).

Figure 1 (a) shows the resistance-corrected (IR-free) polarization curves of the high oxygen permeability ionomers for the ORR at 80 oC fed with ambient pressure air humidified at 80% RH. The mass activity (MA) at 0.85 V and the mass power values, calculated from the maximum outputs of the IR-free polarization curves, are plotted as a function of RH in Figure 1 (b) and (c), respectively. The cathode cell performance was significantly increased by using the high oxygen permeability ionomers, i.e., AGC sample A and AGC sample B. However, the ionomers showed different performance behavior at low and high current densities. In the high potential regions (low current densities), in which there is a smaller effect of mass transport, the AGC sample B with the higher IEC of 1.50 exhibited higher cell performance than the other ionomers, as can be confirmed in the MA results in Figure 1 (b). In contrast, the AGC sample A with the lower IEC of 1.17 led to a significant improvement of the cathode performance in the high current density region, which led to the achievement of very high performance, greater than 22 W mgPt-1at 80-100% RH, as presented in the mass power results in Figure 1 (c). These results show that the improved oxygen permeability of the ionomers could result in highly efficient cathode CLs for PEFCs.

Acknowledgment

This work was supported by funds for the "Superlative, Stable, and Scalable Performance Fuel Cell" (SPer-FC) project from the New Energy and Industrial Technology Development Organization (NEDO) of Japan. The authors are grateful to Asahi Glass Co., LTD. for kindly providing the experimental ionomers. Also, the authors are grateful to Denka Co., Ltd. for kindly providing the experimental AB supports.

References

1. M. Uchida, Y.C. Park, K. Kakinuma, H. Yano, D.A. Tryk, T. Kamino, H. Uchida, M. Watanabe, Phys. Chem. Chem. Phys., 15, 11236 (2013).

2. Y.C. Park, H. Tokiwa, K. Kakinuma, M. Watanabe, M. Uchida, J. Power Sources, 315, 179 (2016).

3. Y.C. Park, K. Kakinuma, H. Uchida, M. Watanabe, M. Uchida, J. Power Sources, 275, 384 (2015).

4. K. Yamada, S. Hommura, T. Shimohira, ECS Trans, 50, 1495 (2013).

Figure 1

2808

, and

The structure of a catalyst layer (CL) in a proton exchange membrane fuel cell (PEMFC) varies according with the deposition technique and the chemical composition that build it up. Then, the electrical, ionic and catalytic properties depend on the final structure. These characteristics have also an impact on the overall response of the PEMFC. In this work, stochastic reconstruction of the microstructure of a PEMFC CL and a scaling method are used to determine the effective transport coefficients and a statistical electrochemical surface area of platinum associated with the active sites. The case study contains three stages: 1) the charge transport continuity equation is used to determine conduction efficiency. 2) The platinum surface is determined by a statistical algorithm that considers the percolation and the three phase union condition. 3) Finally, an analytical model for electrochemical behavior is applied. These numerical and analytical models are applied to hypothetical microstructures of the carbon support, ionomer electrode load and platinum to carbon weight ratio.

2809

, , , , , and

Electrodes based on platinum group metal-free (PGM-free) catalysts are under development as an alternative to costly Pt-based cathodes for automotive fuel cell systems.[1,2] Novel PGM-free catalysts with much improved activity and durability must be developed in order to become competitive with their PGM nanoparticle counterparts for transportation applications. Analytical characterization by transmission electron microscopy plays an important role in identifying electrochemically active sites at the atomic scale as well as to exploring electrode morphology and porosity at the microscale.

In this work, a hybrid PGM-free catalyst developed at Los Alamos National Laboratory was studied by multiscale and multidimensional electron microscopy methods. This promising catalyst, branded CM-PANI-Fe-C, has been synthesized by heat-treating two nitrogen precursors, polyaniline (PANI) and cyanamide (CM). The catalyst is comprised of fibrous carbonaceous agglomerates intermixed with few-layer graphene sheets. Low voltage aberration-corrected scanning transmission electron microscopy (STEM) and atomic resolution electron energy loss spectroscopy (EELS) were utilized to identify single Fe atoms coordinated with nitrogen within the few-layer graphene sheets (Fig. 1a,b). These atomic features have been identified by modeling as a possible active site which exhibits high four-electron selectivity.

The thick electrode structures required to accommodate high non-PGM loadings also play a critical role in fuel cell performance. Hierarchal pore structures (spanning the micro to macro) are necessary for maintaining high mass transport through the thick electrodes. Complementary imaging and compositional mapping by energy dispersive X-ray spectroscopy (EDS) has been utilized to study the bulk electrode structures, as shown in Fig. 1c. These analytical studies will be extended to understand the role of different process parameters on catalyst and electrode structures at multiple length scales.

References

  • G. Wu, K. L. More, C. M. Johnston, P. Zelenay, Science 332 (2011) 443.

  • Z. Chen, D. Higgins, A. Yu, L. Zhang, J. Zhang, Energy Environ. Sci. 4 (2011) 3167.

Acknowledgements

Research sponsored by the Fuel Cell Technologies Office, Office of Energy Efficiency and Renewable Energy, U.S. Department of Energy (DOE), and through a user project supported by ORNL's Center for Nanophase Materials Sciences (CNMS), which is a DOE Office of Science User Facility.

Figure 1

2810

, , , and

The establishment of a fully renewable energy system is predicated on the use of hydrogen as a fuel. Polymer electrolyte membrane (PEM-) water electrolyzers for its generation and (PEM-) fuel cells for its reconversion will be essential components of such a system. In order for this to be economically feasible, however, their costs have to be reduced. The quantity of expensive materials utilized for the construction of cells, including platinum, iridium or titanium must be reduced and overall efficiency increased. The efficiency of these devices is a function of the electrochemical processes that take place in the electrodes and is inseparably tied to their structure. The structure can be affected by controlling various steps in the manufacturing progress. Aside from the chemical composition of the wet coat (consisting of the supported catalyst, Nafion®, solvents and additives) the drying plays a significant role during the self-organization progress in the catalyst layer. Therefore, the drying aspect is one of the main focuses of interest.

Nowadays the most commonly used catalyst-containing dispersions in the manufacturing of membrane electrode assemblies (MEAs) are based on a mixture of organic solvents. In order to clarify the drying process a test rig was constructed. By using a gas phase infrared spectrometer the drying can be monitored by analyzing the waste gas. The set up enables a wide range of conditions in terms of the temperature and the type of drying gas and its preloading. Furthermore, the contact time between the drying gases with the coated layer is adjustable.

As various substances differ in their boiling points a different drying rate is observed for each organic component inside the electrode layer. Consequently, the physical and chemical parameters of the solvent mixture which evaporates during the drying step change dramatically over time. The test rig gives time-resolved information about the selectivity of drying, and the drying rates of the single mixture components in general. The measurement system provides accurate and consistent results. The low measurement uncertainty (< 2% rel.) offers an ideal tool for observing slight modifications during the drying progress.

2811

, , , and

Polymer electrolyte membrane fuel cell (PEMFC) has been extensively studied because of its many desirable qualities for the automotive application, including high power density, low operating temperature (~80 oC) as well as quick start-up and match-up to the power demand. Although appealing, a challenge caused by the high Pt loading in the cathode should be addressed before the mass production of fuel cell vehicles. In this regard, the research topic is now urgently focused on to use ultra-low Pt loading (≦0.1 gPt/kW) in the cathode of PEMFCs. It is noted that both the electrode kinetics of oxygen reduction reaction (ORR) [1] and mass transport for conventional cathodes with low Pt loading do not apply to the cathodes with ultra-low Pt loading, especially for the local mass transport surrounding Pt particles in the catalyst layer [2]. Herein, the challenges in the ultra-low Pt cathode in terms of ORR kinetics and local transport resistance will be clarified and the corresponding solution in terms of the electrocatalyst design will be provided. How the ORR kinetics on the cathode changes with the Pt loading decreasing will be examined in detail and the difference in the transport resistance in each component (Figure 1), especially the O2local transport between the ultra-low Pt cathode and high Pt cathode will be elucidated. The effects of catalyst type, specific activity, Pt loading as well as the cathode structure on the cell performance will be discussed, thus providing a solid scientific guideline for developing cathodes with ultra-low Pt loading and accelerating the commercialization of PEMFCs using cathodes with ultra-low Pt loading.

Acknowledgements

This work was supported in part by National Natural Science Foundation of China (Grant No. 21373135 and 21533005) and Science Foundation of Ministry of Education of China ( Grant No. 413064).

References

[1] Y. Huang, J.L. Zhang, A. Kongkanand, F.T. Wagner, J.C.M. Li, J. Jorné, Transient platinum oxide formation and oxygen reduction on carbon-supported platinum and platinum-cobalt alloy electrocatalysts, J. Electrochem. Soc., 2014, 161 (1): F10-F15.

[2] Y. Ono, T. Mashio, S. Takaichi, A. Ohma, H. Kanesaka, K. Shinohara, The analysis of performance loss with low platinum loaded cathode catalyst layers, ECS Transactions, 2010, 28 (27) 69-78.

Figure 1

D-32b Anode and Cathode for HT-PEMFC - Oct 6 2016 2:00PM

2812

, , , and

Jet fuels like the Jet-A and Jet-A1 fuel for commercial aviation or the military JP-8 fuel are important fuels for many types of application. Recently the replacement of these fuels by synthetic drop-in fuels is being considered. As fuel cell based auxiliary power units are of interest in many of the applications using jet fuels for the main propulsion, the question arises if avoiding some of the impurities which need to be expected in reformed fossil jet fuels can enhance the operation of the fuel cells.

To further investigate this question tests have been performed on catalyst and single cell level. Catalyst tests were performed with commercial 20 wt.% Pt/C catalyst from Johnson & Matthey using circular gas diffusion electrodes with about 1 cm diameter which were mounted in a special test cell allowing for inline mass spectrometry during the electrochemical characterisation (1, 2). Parts of the measurements were verified by single cell measurements in a commercial fuel cell test bed (Evaluator C050, Fuelcon AG, Germany). The tested MEAs were commercially procured from either Danish Power Systems (Denmark) or Advent (Greece). For the tests an accelerated stress tests based on start-stop cycling described elsewhere (3) was employed. Three types of impurities were considered each represented by one or two model substances. The first class of impurities are alkenes which can result from the reforming of hydrocarbon fuels of any composition. The effect of alkenes was studied using ethene as model substance. The second type of impurities is aromatic compounds which are contained in fossil jet fuels but not in all replacement fuels. The effect of aromatic compounds was tested with toluene. The last type of impurities is sulphurous compounds which are not prersent in most replacement fuels. Here two types of model substances were investigated. Beside hydrogen sulphide thiophene was tested in combination with toluene as it is both an aromatic compound and a sulphur organic compound. Gaseous impurities were tested using pre-mixed gases purchased from Linde-Gase Germany. Toluene and thiophene were added by bubbling the gas feed through a flask with the toluene solution containing thiophene at ambient temperature so that the saturation partial pressure off 2.9 kPa should have been established.

The addition of small amounts of ethene two the hydrogen feedstock caused slightly increased currents in the catalyst testing. The MS data also showed a slight increase of the CO2 production at high potentials (cf. fig 1). Tests with Advent MEAs and hydrogen with 10 ppm of ethene confirmed that no effect became apparent during the test phase of 130 start-stop cycles (cf. fig 2).

At low potentials toluene has no effect of the hydrogen oxidation at Pt/C catalyst. At high potentials the activity is however slightly reduced (cf. fig 3). This reduction could be attributed to the oxidation of the thiophene which was present in the toluene. This can be concluded from the observed reduction of the thiophene peak at m/z 83 in the mass spectrum at the relevant potentials shown in fig 4. The effect of H2S has already been studied by Schmidt and Baurmeister (4) as well as by our group (2, 5). On single cell level H2S concentration of 10 – 20 ppm can be tolerated for Pt/C anodes.

In total sulphur remains to be the most critical impurity if it occurs at to high concentrations, smaller amounts even of sulphur-aromatic compounds can however be tolerated. The use of fuel cells should thus profit from the introduction of sulphur reduced drop in fuels. In the contribution further results of ongoing single cell tests will be reported.

Acknowledgements

Financial support by the German Federal Ministry of Defence, Bundeswehr Research Institute for Materials, Fuels and Lubricants (WIWeB) under contract E/E210/AF020/CF062 is gratefully acknowledged.

The authors would like to thank Ms. Florian Jung for her support in electrode preparation.

References:

1. C. Niether, M. S. Rau, C. Cremers, D. J. Jones, K. Pinkwart and J. Tübke, Journal of Electroanalytical Chemistry, 747, 97 (2015).

2. M. Rau, C. Cremers and J. Tübke, International Journal of Hydrogen Energy, 40, 5439 (2015).

3. M. S. Rau, A. Niedergesäß, C. Cremers, S. Alfaro, T. Steenberg and H. A. Hjuler, Fuel Cells (submitted).

4. T. J. Schmidt and J. Baurmeister, ECS Transactions, 3, 861 (2006).

5. M. Rau, C. Cremers, K. Pinkwart and J. Tübke, ECS Transactions, 64, 983 (2014).

Figure 1

2813

, , and

The effect of hydrogen sulphide (H2S) poisoning on the anode of H3PO4-polybenzimidazole based fuel cells has been investigated. High temperature PEM fuel cells (HT-PEMFC) offer a number of advantages that allow for a simple electricity generating system. Of the main importance is the ability to operate on reformate hydrogen without extensive purification [1]. Most of the research is concentrated on CO poisoning mitigation, but as the feedstock often contains sulphur contamination or sulphur-containing odorant additives, the poisoning effect of sulphur on Pt electrodes at elevated temperatures should not be overlooked. Little literature is available on the topic [2, 3]. Present work shows the effect of anode H2S poisoning on 25 cm² HT-PEMFCs operated at 160 °C, evaluated by polarisation curves and electrochemical impedance spectroscopy (EIS).

Cells exposed to 0.2 – 50 ppm H2S in the fuel supply show no significant signs of voltage degradation when operated at 160 °C and 200 mA/cm². The main effect is seen at higher current densities with 10 – 50 ppm levels of contamination, where concentration overpotentials cause the voltage to drop off. The effect is partly reversible, as the voltage at high current densities recovers upon switch back to pure hydrogen operation. This is contrary to the low temperature PEM fuel cells, where the voltage degradation with only trace amounts of H2S is rapid even at low currents and irreversible [4, 5].

Prolonged exposure to 1 – 50 ppm H2S in the fuel supply show a time-dependant nature of the sulphur poisoning with the major voltage degradation occurring within the first 24 hours. Again, the effect is only visible at intermediate and high currents and is partly reversible when the fuel is changed to pure hydrogen after several days of operation with H2S contaminated fuel.

EIS measurements at 200 mA/cm² showed no change in the series resistance. Occurrence of a semicircle at low frequencies, growing throughout the prolonged exposure measurements could indicate increased levels of adsorption of sulphur on the Pt sites of the electrode, thus explaining the concentration overpotentials recorded at high current densities.

[1]        Q. Li et al., Progress in Polymer Science, 2009, 34 (5), p. 449-477.

[2]        T.J. Schmidt and J. Baurmeister, ECS Transactions, 2006, 3 (1), p. 861-869.

[3]        G. Qian and B.C. Benicewicz, Ecs Transactions, 2011, 41 (1), p. 1441-1448.

[4]        V.A. Sethuraman and J.W. Weidner, Electrochimica Acta, 2010, 55 (20), p. 5683-5694.

[5]        I. Urdampilleta et al., ECS Transactions, 2007, 11 (1), p. 831-842.

2814

, , and

Metal phosphides are world-widely investigated as highly-active and selective catalysts for hydrodesulfurization, hydrodenitrogenation, and hydrodeoxygenation. These reactions require catalytic ability for dissociation of H-H bond as well as C-S, C-N, or C-O bond. In addition to the catalytic activity, metal phosphides are known to be thermally and chemically stable and to possess electric conductivity comparable to metals and hardness similar to ceramics [1]. Because of these features, metal phosphides are expected to function as electrodes in electrochemical cells. Recently, metal phosphide based electrodes have been investigated as an electrode for hydrogen evolution reaction in H2O electrolysis. McEnaney et al. carried out cyclic voltammetry of phosphide anodes in sulfuric acid and obtained stable current-voltage curves [2]. This indicates that metal phosphides have potentials to be used as anodes of fuel cells employing acidic electrolytes. So far, application of metal phosphides to a fuel cell anode is quite limited. Chang et al. developed anodes of Pd supported on Ni2P/C for direct formic acid fuel cells [3].The Pd anode with nickel phosphide exhibited improved power generation characteristics when they were compared with the cell without the nickel phosphide. The role of the nickel phosphide seems, however, still unclear in the anode, since the power generation temperature was ca. 30ºC, which is too low for the phosphide-based catalysts to exert catalytic activity.

In this study, metal phosphides such as Ni2P, CoP, FeP, WP, and MoP have been investigated as anode catalysts for intermediate temperature fuel cells. Composites composed of CsH2PO4 and SiP2O7, the metal phosphates, and a commercial Pt/C electrode were used as the electrolyte, anode, and cathode, respectively, and power generation characteristics were evaluated at 220ºC as H2-O2 fuel cells. Cesium dihydrogen phosphate, CsH2PO4, was prepared by dissolving stoichiometric quantities of Cs2CO3 and H3PO4 in distilled water and drying overnight at 120 °C. SiP2O7 was prepared from mesoporous SiO2 and H3PO4 as reported previously [4]. As a cathode catalyst Pt/C-loaded carbon paper (Pt loading 1 mg cm-2, ElectroChem Inc.) was used. As an anode catalyst, the phosphide-loaded carbon paper was prepared by filtration of the phosphide dispersed in an ethanol solution. The membrane electrode assembly (MEA) was prepared by uniaxial pressing at 250 MPa for 10 min to form the MEA. In power generation experiments, humidified H2 and O2 were supplied to anode and cathode, respectively, at 50 cm3 min-1. A water vapor concentration of 30 vol% was obtained by bubbling the gas flow through water at 70 °C. Current-voltage characteristics were measured at 220 °C using a potentiostat (Solartron 1287), and AC impedance measurements were carried out at open circuit condition with a frequency-response analyzer (Solartron 1260).

Figure 1 summarizes current-voltage characteristics of H2-O2 fuel cells with phosphide anodes at 220°C. The phosphide electrocatalysts were found to function as anodes of intermediate temperature fuel cells. The power generation characteristics are in the following order: MoP > WP > FeP > CoP > Ni2P. Impedance analysis showed that ohmic resistance of the cells with the phosphides except FeP was small and comparable to that of the Pt/C anode. This result indicates that these phosphide anodes possess good conductivity applicable to fuel cell anodes. On the other hand, the non-ohmic resistance of FeP, CoP, and Ni2P was considerably large, which corresponds to the order of the power generation performance.

[1] S.T. Oyama, T. Gott, H. Zhao, Y.-K. Lee, Catal. Today 143, 94-107 (2009).

[2] J.M. McEnaney, J.C. Crompton, J.F. Callejas, E.J. Popczun, A.J. Biacchi, N.S. Lewis, R.E. Schaak, Chem. Mater. 26, 4826-4831(2014).

[3] J. Chang, L. Feng, C.P. Liu, W. Xing, X. Huet, Angew. Chem. Int. Ed. 53, 122-126 (2014).

[4] T. Matsui, T. Kukino, R. Kikuchi, K. Eguchi, Electrochem. Solid-State Lett. 8, A256-A258 (2005).

Figure 1

2816

, , and

Carbon nanotubes are recognized as a promising alternative support material for fuel cell electrocatalysts due to their higher electrical conductivity and higher durability, when compared to carbon black. In fact, using Pt nanoparticles supported on carbon nanotubes (CNTs) wrapped in poly (vinylphosphonic acid)-doped polybenzimidazole (PVPA-PBI) makes a remarkably high durable fuel cell, thus opening the door for the next generation of these devices [1].

In this work, the degradation mechanism of 3.4 nm Pt nanoparticle catalysts supported on PVPA-PBI wrapped CNTs was investigated by aberration corrected TEM, before and after voltage cycling. In order to carry this experiment, Pt/PVPA-PBI/CNT powder was deposited on a gold grid attached to a gold plate, which was used as a working electrode in a three electrode electrochemical cell. To simulate the effect of fuel cell cycling, the TEM grid was cycled between 1 and 1.5 V RHE in N2 saturated 0.1 HClO4liquid electrolyte for 1000, 1500, and 2000 cycles. In this fashion, pre-defined locations of the electrocatalyst on the TEM grid were observed before and after cycling, by an aberration-corrected JEOL ARM 200F.

In the first 1000 cycles, the main mechanism for the loss of electrochemical active surface area was found to be particle motion, followed by coalescence. Severe structural deformation of the carbon nanotubes during voltage cycling is another source of degradation. The wave-like structure of the carbon formed after voltage cycling is the result of the appearance of the defect sites on the carbon nanotubes which convert the flat hexagon structure to curved heptagon and pentagon carbon rings. In order to understand if there is any correlation between carbon degradation and particle movement on the carbon support, carbon degradation was accelerated under the electron beam, while the behavior of the nanoparticles was observed. It is shown that the particles start to move as soon as carbon atoms at the interface of the carbon/particle interface are removed. During voltage cycling, carbon corrosion occurs through a series of reactions at the carbon/ionomer, ionomer/particle, and carbon/particle interfaces. These reactions result in the formation of carbon dioxide and removal of carbon atoms at the particle/carbon interface. When carbon atoms at the carbon/particle interface are removed from the surface of carbon, leaving a carbon defect site, the particle tends to make new bonds with the next carbon atoms to possibly decrease its free surface energy.

In the next 1000 cycles, single atoms and atomic clusters appears on the ionomer phase at the surface of carbon nanotubes. These single atoms and atomic clusters move subsequently toward large particles and deposit on their surfaces. The atomic clusters will either dissolve again to single atoms or ions and redeposit on the large particles or move toward particles and redeposit between them with consequent bridging.

Reference:

[1] M. Berber, T. Fujigaya, N. Nakashima, Chemcatchem, 6 (2) 567-572 (2014)

2817

, , , , and

High temperature PEM fuel cell based on phosphoric acid doped polybenzimidazole have reached a quite mature state of development. However, durability is still a critical issue and the key to improving it is a better understanding of the degradation mechanisms. It can be expected that the cells have most, if not all, degradation modes of the low temperature PEM fuel cell (corrosion, particle growth, membrane thinning etc.), but additionally there are issues with the phosphoric acid that the membrane is filled with to provide ionic conductivity at high temperature. In terms of degradation, one could argue that the cell combines the challenges of the low temperature PEM fuel cell and the phosphoric acid fuel cell. Nevertheless, lifetimes on the order of 10.000 hours are demonstrated by several groups [1]

In the present study, a large number of cells manufactured by Danish Power Systems were tested with hydrogen in three in-house built multichannel test rigs over several years. The working temperature, the oxidant and fuel flow rates and the current load were varied between 160 and 200 °C, stoichiometry up to about 10 and current densities between 200 and 800 mA cm-2, respectively. The cells were characterized repeatedly by polarization curves and/or electrochemical impedance spectroscopy. After test, selected cells were subjected to post mortem analyses by means of cross section examination by microscopy, X-ray diffraction for platinum particle sized and acid titration.

All the major degradation mechanisms were seen (catalyst particle growth, membrane thinning and acid loss). The natural question to answer is which mechanisms dominate under which conditions. It was a general trend that degradation rates increased significantly with temperature and with current density. The temperature of the cell is normally measured outside the gas diffusion layers and even outside the channel plates. Since the produced heat is liberated within less than a hundred micrometers (membrane and catalyst layers) and since the transport away from that area is driven by the temperature gradient, one must expect a significantly higher temperature in the center of the cell than the temperature measured in the channel plates or end plates of the test cell. At higher current densities, the heat evolution rate is higher and thus a higher central temperature is expected. The temperature was measured inside the membrane, but the temperatures could not alone explain the increased degradation rate at higher current densities.

A stronger effect seems to origin from the reactant flow rates. Some test series were carried out at high flow rates to mimic an air-cooled stack and in those cases, degradation was significant even at moderate current densities. A combination of acid loss and acid dehydration is suggested as decisive, although dehydration should at the same time limit acid evaporation. The presentation will review the findings of the temperature/load/flow matrix studies and suggest conclusions on the main factors limiting durability of high temperature PEM fuel cells.

 

[1] M. T. Dalsgaard Jakobsen, J. O. Jensen, L. N. Cleemann and Q. Li. Chapter 22. Durability Issues and Status of PBI Based Fuel Cells. In Q. Li, D. Aili, H. A. Hjuler and J. O. Jensen (eds). High Temperature Polymer Electrolyte Membrane Fuel Cells- Approaches, Status and Perspectives. Springer-Verlag 2016

Figure 1

E-32 Alkaline & DFC Electrocatalysis 4 - Oct 6 2016 2:00PM

2818

and

The electrolysis of water for hydrogen generation and the electrochemical oxidation of alternative fuels such as methanol for alkaline fuel cells currently relies on the use of precious metal catalysts. However, the scarcity and high cost of these precious metals lead to the search for earth-abundant alternatives with similar electrocatalytic properties. Key challenges in the development of catalysts for the oxygen evolution half reaction (OER) of water electrolysis remain; these challenges include reducing the overpotential of the reaction and developing non-precious metal catalysts that are stable and active with minimal mass transport limitations. For direct electrooxidation of alternative fuels such as methanol, non-precious metal alternatives to precious metals such as platinum and palladium are often much less active. Among the non-precious metals, multimetallic nickel-based catalysts show particular promise for their high stability and high activity for the oxygen evolution reaction (OER) half-reaction of water electrolysis. In particular, it has recently been demonstrated by Trotochaud, Burke, Boettcher, and co-authors that iron incorporation into nickel hydroxide and cobalt hydroxide catalysts causes a significant increase in the OER activity of the catalyst [1-3]. It is now understood that the iron is in fact the active site of such bimetallic catalysts [2]. Additionally, nickel has exhibited high activity for methanol oxidation, particularly when in the form of a bimetallic catalyst such as NiPd [4]. However, promising active catalysts that are comprised entirely of non-precious metals such as nickel and iron have not been often reported for alternative fuel electrooxidation. Furthermore, there is a need for the development of nanostructured catalysts that can address mass transport limitations inherent in bulk or film-based catalyst materials [2]. As such, there continues to be an opportunity in alkaline electrocatalyst development to discover and understand nanostructured catalyst materials for both OER and alternative fuel electrooxidation.

In this presentation, our results on the development and optimization of an FeNi hydroxide core-shell nanoparticle catalyst (Figure 1a) will be discussed for both OER and methanol electrooxidation. Through control of critical synthesis parameters, such as the addition rate of reducing agent, ratio of reducing agent to iron, and reaction time for nanoparticle synthesis, we have demonstrated the ability to tune nanoparticle catalyst activity for either OER or methanol electrooxidation. In Figure 1b, results for maximum OER current as a function of the rate of addition of reducing agent sodium borohydride illustrate the sensitivity of catalyst performance to changes in synthesis parameters. Characterization of nanoparticle structure and phase will be discussed and correlated to electrochemical performance. Results suggest that tuning of the crystalline structure and order are critical to electrocatalyst performance optimization, as well as the presence of oxide and hydroxide.

References

[1] L. Trotochaud, S.L. Young, J.K. Ranney, S.W. Boettcher, Nickel-iron oxyhydroxide oxygen-evolution electrocatalysts: The role of intentional and incidental iron incorporation, J. Am. Chem. Soc., 136 (2014) 6744-6753.

[2] M.S. Burke, L.J. Enman, A.S. Batchellor, S.H. Zou, S.W. Boettcher, Oxygen evolution reaction electrocatalysis on transition metal oxides and (oxy)hydroxides: Activity trends and design principles, Chemistry of Materials, 27 (2015) 7549-7558.

[3] M.S. Burke, M.G. Kast, L. Trotochaud, A.M. Smith, S.W. Boettcher, Cobalt-iron (oxy)hydroxide oxygen evolution electrocatalysts: The role of structure and composition on activity, stability, and mechanism, J. Am. Chem. Soc., 137 (2015) 3638-3648.

[4] A. Dutta, J. Datta, Energy efficient role of Ni/NiO in PdNi nano catalyst used in alkaline DEFC, J. Mater. Chem. A, 2 (2014) 3237-3250.

Figure 1

2819

, , and

Electrooxidation of methanol has been studied intensively in the scientific community, and various catalysts have been developed for direct methanol fuel cells.

In this study, platinum-cerium oxide nanoparticles have been synthesised, exhibiting high activity towards methanol electrooxidation. The materials were electrochemically characterised using rotating disk electrode, cyclic voltammetry, chronoamperometry and electrochemical impedance spectroscopy methods. Scanning electron microscopy, X-Ray diffraction, inductively coupled plasma mass spectrometry and BET surface area mesurements were performed to characterise the physical properties of the synthesised materials.

The methodology of the synthesis of the materials was straightforward and resource efficient. The synthesised nanocatalysts have a very uniform dispersion on the catalyst support material, and a narrow size distribution.

Even at low platinum loadings, the materials have a lower overpotential towards methanol oxidation than commercial Pt-Vulcan catalysts. Based on cyclic voltammetry measurement data, the peak potential for the synthesised materials was 40 mV more negative than for the commercial material in the anodic potential scan, and 60 mV more negative in the cathodic potential scan.

In addition to the high currents for methanol oxidation, extremely high currents were achieved in the cathodic potential sweep, in the region attributable to the oxidation of carbon monoxide. Methanol and carbon monoxide poisoning of the nanocatalysts was unnoticeable during the experiments.

The results of the study lay a promising foundation for the investigation of nanocatalysts containing other rare earth metal oxides, as well as to the investigation of the electrooxidation of ethanol or other organic fuels on composite catalysts.

Acknowlegements

The authors wish to thank P. Paiste for ICP-MS measurements and J. Aruväli for XRD measurements

This work was supported by the EU through the European Regional Development Fund TK141, the institutional research funding project IUT20-13 of the Estonian Ministry of Education and Research, the Estonian Energy Technology Program Project No. SLOKT10209T, Estonian Materials Technology project 3.2.1101.12-0019, Estonian Center of Excellence in Science Project TK117T, grant ETF9352 and PUT55.

2820

, and

Ru@Pt core-shell catalytic particles consist of a thin Pt shell covering a Ru core, and may be supported on carbon to provide catalysts with properties intermediate between those of Pt and RuPt alloy particles of similar size [1]. Whereas carbon-supported Pt typically oxidizes adsorbed CO in stripping experiments with a peak potential around 0.8 V vs. a reversible hydrogen reference electrode (RHE) in 0.5 mol dm-3 HClO4(aq) the current for the core-shell catalysts peaks around 0.6 V. The value for the alloy is around 0.5 V. However, although the catalytic properties for for example electrooxidation of CO may not be quite as good as those of the alloy, the core-shell particles display superior stability. These core-shell catalysts should therefore be of interest in applications such as direct methanol fuel cells [2] or as CO-tolerant catalysts in PEM fuel cells.

The potential of total zero charge (pztc), the potential corresponding to zero electrode charge (including both free charges and bonded charged species), may be assessed through the CO-displacement technique by Climent et al. [3] and shown by Mayrhofer et al. [4] to correlate with catalytic acticity. The method is based on CO displacing more weakly adsorbed cations and anions at the electrode or catalyst surface.

The pztc has been found to correlate with the work function of metal electrodes, as has also the enthalpy of adsorption of CO and O2 [5]. This suggests that it should be possible to infer the enthalpy of adsorption for these species directly from the pztc. Moreover, scaling relations established by density-functional theory (DFT) calculations [6] imply that it is possible to estimate adsorption enthalpies for a number of other adsorbates relevant for for example the methanol oxidation reaction (MOR) or the oxygen reduction reaction (ORR), if the adsorption enthalpies of CO and oxygen are known. With a knowledge of adsorption enthalpies for a given reaction mechanism involving reactions of the type A + 1/2 H2 → AH the free energy of the steps may be inferred as explained by Rossmeisl and co-workers [7,8]. Through this sequence of correlations we have been able to infer the rate-determining step and rank the catalytic activity from the pztc. The pztc measurements thus predict that the onset potentials for the ORR should increase in the order RuPt < Ru@Pt < Pt, whereas that of CO-stripping and the MOR should decrease in the same order, RuPt > Ru@Pt > Pt, in close agreement with experimental results, as shown in the the figure; the figure shows the dependence of the specific activity of the ORR and the apparent rate constant, the exponent of -FEp/RT [9] where Epis of the peak potential and the other symbols take their usual meaning, for CO-stripping to the potential of zero total charge vs. RHE. From left to right are PtRu, Ru@Pt, and Pt, respectively. Moreover, this combination of CO-displacement measurements and DFT results appears to predict the right order of magnitude differences in the onset potential between the catalysts as well.

Akcnowledgement

This work was funded by NTNU.

References

1. P. Ochal, J. L. G. de la Fuente, M. Tsypkin, F. Seland, S. Sunde, N. Muthuswamy, M. Rønning, D. Chen, S. Garcia, S. Alayoglu, and B. Eichhorn, J. Electroanal. Chem. 655 (2011) 140

2. D. Bokach, J. Gomez de la Fuente, M. Tsypkin, P. Ochal, I. C. Endsjø, R. Tunold, S. Sunde, and F. Seland, Fuel Cells, 11 (2011)735

3. V. Climent, R. Gomez, J. M. Orts, A. Rodes, A. Aldaz, and J. M. Feliu, Interfacial Electrochemistry, Marcel Dekker: New York (1999) p. 463

4. K. J. J. Mayrhofer and B. B. Blizanac and M. Arenz and V. R. Stamenkovic and P. N. Ross and N. M. Markovic, J. Phys. Chem. B, 109 (2005) 14433

5. S. Trasatti, The Work function in Electrochemistry', in H. Gerischer and C. W. Tobias (Eds.), Adv. Electrochem. Electrochem. Eng., Vol. \textbf{10}, John Wiley & Sons, New York (1977) pp. 213 - 321

6. G. A. Tritsaris and J. Rossmeisl, J. Phys. Chem. C, 116 (2012) 11980

7. J. Rossmeisl, Z.-W. Qu, H. Zhu, G.-J. Kroes, J.K. Nørskov, J. Electroanal. Chem. 607 (2007) 83

8. J. K. Nørskov, J. Rossmeisl, A. Logadottir, L. Lindqvist, J. R. Kitchin, T. Bligaard, H. Jonsson, J. Phys. Chem. B 108 (2004) 17886

9. M. Tsypkin, J. Luis~Gomez de la Fuente, S. Garcia Rodriguez, Y. Yu, P. Ochal, F. Seland, O. Safonova, N. Muthuswamy, M. Rønning, D. Chen, and S. Sunde, J. Electroanal. Chem., 704 (2013) 57

Figure 1

2821

Design and Fabrication of Pt-Au Nanocatalyst with High Performance

Fanpeng Kong[a], Chunyu Du[a]*, Jinyu Ye[b], Lei Du[a], Guangyu Chen[a], Geping Yin[a]*

[a] MIIT Key Laboratory of Critical Materials Technology for New Energy Conversion and Storage, Harbin Institute of Technology. Email: cydu@hit.edu.cn, yingphit@hit.edu.cn.

[b] State Key Lab of Physical Chemistry of Solid Surfaces, Xiamen University.

Abstract: Direct formic acid fuel cell (DFAFC), as a promising power source, can greatly improve energy conversion efficiency of small organic molecules and dramatically reduce the environment pollution.[1-2] Pt is a key catalytic component at both anode and cathode in DFAFC. However, the scarcity and high cost of Pt require the design and fabrication of Pt-based nanostructure with high performance and reduced usage of Pt.

Here, we found that the formic acid oxidation (FAO) activity of Pt and Au nanoparticles supported on carbon (Pt1-Au1/C) increases more than 40 fold relative to that of Pt/C. Generally, ensemble and electronic effects are proposed to explain the enhanced FAO performance.[3-4] However, we have to exclude the ensemble effect since there is no Au adatoms on Pt. From X-ray photoelectron spectroscopy (XPS) patterns, the binding energy of Pt4f and Au4f peak in Pt1-Au1/C shifts positively and negatively compared to that in Pt/C and Au/C, respectively, indicating that Au modifies the electronic structure of adjacent Pt. This electronic modification is further confirmed by the higher onset potential of CO oxidation on Pt1-Au1/C than that on Pt/C in the CO stripping voltammetry, demonstrating stronger binding of CO on Pt1-Au1/C.

Unfortunately, the pathway of formic acid oxidation on Pt-Au system is ambiguous because Pt and Au are closely coupled together, so that the in-situ information on the respective role of Au and Pt is difficult to obtain. To this end, a novel selective surface engineering method is proposed to probe the main active sites. We separately employed CO adsorption and selective under potential deposition to engineer respectively Pt and Au surfaces. However, the activity of Pt1-Au1/C with engineered surfaces is obviously lower than that of Pt1-Au1/C. Therefore, the coexistence of Pt and Au, rather than electronic effect, is indispensable for the high FAO activity. In-situ FTIR measurement of Pt/C, Au/C and Pt1-Au1/C is carried out to probe more detailed molecular information. Formate on Pt1-Au1/C is more than that on Pt/C and Au/C, which should result from the coverage of CO on Au, instead of Pt, at low potentials. Since Pt has higher affinity to H than Au, the increased formate is attributed to the accelerated rupture of O-H bond in formic acid on Au by Pt. However, formate is not an active intermediate but a spectator species, so that it cannot account for the increased activity. Recently, it is proved that –COOH is the active species during FAO. In view of strong affinity between Pt and H, we can rationally assume that Pt ruptures the C-H bond of formic acid on Au to produce –COOH, which decomposes to CO2 quickly, so that Pt1-Au1/C possesses a remarkable FAO activity.

Reference:

  • J. V. Perales-Rondon, A. Ferre-Vilaplana, J. M. Feliu, E. Herrero, J. Am. Chem. Soc. 2014, 136, 13110-13113.

  • M. E. Scofield, C. Koenigsmann, L. Wang, H. Liu, S. S. Wong, Energy Environ. Sci. 2015, 8, 350-363.

  • Q. S. Chen, Z. Y. Zhou, F. J. Vidal-Iglesias, J. Solla-Gullon, J. M. Feliu, S. G. Sun, J. Am. Chem. Soc. 2011, 133, 12930-12933.

  • G. R. Zhang, D. Zhao, Y. Y. Feng, B. S. Zhang, D. S. Su, G. Liu, B. Q. Xu, ACS Nano 2012, 6, 2226-2236.

2822

and

One of the major obstacles in fuel cell commercialization is the cost of the precious metal catalysts used for oxygen reduction reaction (ORR) and hydrogen oxidation reaction(1). Recently, non-precious metal catalysts are mainly obtained by the pyrolysis of the transition metal, carbon and nitrogen sources together and these catalysts are viewed as potential alternative to the expensive platinum based catalysts(2). Pyrolysis step leads to formation of several nitrogen and transition metal based functionalities covalently embedded on the carbon surface(3). Hence, identifying the exact chemical structure of the ORR active site in the pyrolyzed catalyst remains elusive. Hence, designing a synthesis process (which includes pyrolysis) with an objective of maximizing the number of active sites is a work in progress. To avoid difficulties associated with such process, we propose a metalorganic complex capable of catalyzing ORR. As it is a molecular material, its characterization by single crystal XRD, 1H and 13C NMR unambiguously predicted the structure. In this study we report ORR on bis(2,2'-bipyridine-N,N')carbonatocobalt(III)nitrate pentahydrate. The above said complex was immobilized on the Ketjenblack carbon and its ORR activity was measured in O2 saturated 0.1 N KOH solution (Figure (a)). This complex crystalizes in monoclinic (C 2/c) crystal structure and its crystal structure is shown in Figure (b).

Figure Shows (a) the cyclic voltamograms recorded at different rotation rate in O2-saturated (solid lines) and in N2-saturated (dash line) 0.1 N KOH electrolyte and (b) crystal structure of the metalorganic complex employed in this work.

References:

1. Y. Wang, K. S. Chen, J. Mishler, S. C. Cho and X. C. Adroher, Appl. Energy, 88, 981 (2011).

2. X. Li, H.-J. Zhang, H. Li, B. Zhao and J. Yang, J. Electrochem. Soc., 161, F925 (2014).

3. K. Artyushkova, A. Serov, S. Rojas-Carbonell and P. Atanassov, J. Phys. Chem. C, 119, 25917 (2015).

Figure 1

2823

In this study, we report about the activity of different catalysts based on nanoplatinum and/or platinum alloys and nanocarbons for alcolhol oxidation.

Firstly, we introduce about the synthesis of nanocatalysts by reduction in aqueous solution using NaBH4 and Ethylene Glycol as reductant. Then we present the synthesis of thin films of Platinum and bimetallic platinum-ruthenium on carbon paper and fluorine-doped tin oxide coated glass (FTO) by electrostatic spray deposition (ESD). The alcolhol electro-oxidation in acidic and alkaline media were investigated by cyclic voltammetry (CV) and chronoamperometry (CA). Oxygen reduction reaction was investigeted using a rotating disk electrode and the three-electrode cell operating MetrOhm AutoLab-302N. Compared to Pt, the bimetallic catalysts displayed an enhanced activity toward alcolhol electro-oxidation reaction and some for oxygen reduction reaction. The beneficially effect of combination Ru with Pt was also found to be influenced by the nature of metallic precursors and substrates used in the ESD technique.

D-32a Non-PGM Cathode Catalysts 2 - Oct 6 2016 2:00PM

2824

, , , , , , , and

In recent years, the inherent activity of M/N/C (M=Fe, Co, or Mn) non precious metal catalyst (NPMCs) has advanced to a stage where these catalysts can be considered as potential alternatives to Pt/Pt-alloy catalysts for certain applications(1-4), provided stability/durability can be improved(5). Due to these achievements, NPMCs have advanced beyond purely rotating disc electrode (RDE) studies, and are now ready to be thoroughly evaluated at the membrane electrode assembly (MEA) level. While previous work has looked at in-situ MEA evaluation of NPMCs, few of these studies have investigated maximizing performance through rational design of the cathode catalyst layer (CCL). However, recent work by Serov et al. (3) has highlighted the importance of CCL design parameters in maximizing the performance of NPMC-based CCLs.

The significant difference in CCL layer thickness of NPMC-based CCLs vs. conventional Pt/C CCLs means any optimization used for conventional Pt/C CCLs will likely not apply to these novel NPMC CCLs. This provides the research community with significant opportunity in this largely unexplored area.

In this work, we will present some of our recent findings on designing and optimizing NPMC-based CCLs for portable power and backup power applications. The NPMC used in this work is developed by Nisshinbo Holdings, and has demonstrated world-leading performance in laboratory scale (50 cm2) MEA testing. After a brief discussion on the synthesis of this NPMC, the results of optimizing the CCL design (e.g. ionomer type/content) will be shown. Performance data recently obtained under conditions relevant for portable power/backup power will be presented and compared against product requirements.

Acknowledgements

The authors thank Alan Young for many helpful discussions,

References

1. M. Lefèvre, E. Proietti, F. Jaouen and J.-P. Dodelet, Science, 324, 71 (2009).

2. F. J. Eric Proietti, Michel Lefèvre, Nicholas Larouche, Juan Tian, Juan Herranz, Jean-Pol Dodelet, Nature Communications, 2 (2011).

3. A. Serov, K. Artyushkova, E. Niangar, C. Wang, N. Dale, F. Jaouen, M.-T. Sougrati, Q. Jia, S. Mukerjee and P. Atanassov, Nano Energy, 16, 293 (2015).

4. A. Serov, M. H. Robson, B. Halevi, K. Artyushkova and P. Atanassov, Electrochemistry Communications, 22, 53 (2012).

5. D. Banham, S. Ye, K. Pei, J.-i. Ozaki, T. Kishimoto and Y. Imashiro, Journal of Power Sources, 285, 334 (2015).

2825

, , , , , , , and

Platinum group metal-free (PGM-free) oxygen reduction reaction (ORR) electrocatalysts have demonstrated substantial improvements in electrochemical cell and fuel cell performance in the last decade. However, the durability of these catalysts is still insufficient for the implementation of these catalysts in polymer electrolyte fuel cells (PEFCs), particularly in automotive power systems. Recently, carbon oxidation has been reported as a main cause for the destruction of ORR active sites in heat-treated metal-nitrogen-carbon (M-N-C) PGM-free catalysts (1). This result suggests that in order to improve their durability, M-N-C catalysts should be synthesized at as high a temperature as possible to maximize the graphitic content thereby making the catalyst more oxidation resistant. An optimum heat-treatment temperature for the best ORR activity of M-N-C catalysts depends on many factors, including the synthesis approach, type of precursors used, and the precursor ratio. One more factor to account for, based on the reasoning above, is the need to maximize the heat-treatment temperature without sacrificing catalyst activity.

In this work we have found that for the high-temperature synthesis of active M-N-C catalysts, the most important factors are the selection of the nitrogen precursor, adjustment of the precursor ratio, and the choice of the synthesis route. Without fine-tuning of the synthesis conditions, the ORR activity of the PGM-free catalysts usually decreases when heat-treated above 900° C. By selecting an appropriate synthesis route and catalyst precursors we have been able to perform synthesis at 1100° C and obtain catalysts that show high RDE activity (E½ of 0.81 V vs. RHE, Figure 1). In this presentation, we will summarize the electrochemical and fuel cell performance of these catalysts, with particular emphasis on their durability.

Acknowledgement

Financial support of this research by DOE-EERE through Fuel Cell Technologies Office is gratefully acknowledged. Microscopy research was supported through a user project at ORNL's Center for Nanophase Materials Sciences, which is an Office of Science User Facility.

Reference

  • Choi Chang Hyuck, Claudio Baldizzone, Jan-Philipp Grote, Anna K. Schuppert, Frédéric Jaouen, and Karl J. J. Mayrhofer, Stability of Fe-N-C Catalysts in Acidic Medium Studied by Operando Spectroscopy. Angew. Chem. Int. Ed.54, 1 (2015).

Figure 1

2826

, , , and

The kinetically sluggish oxygen reduction reaction (ORR) at the cathode of polymer electrolyte membrane fuel cells (PEFCs) requires higher content of platinum at the cathode than the facile hydrogen oxidation at the anode. Due to the scarcity and high cost of platinum, the development of inexpensive platinum group metal-free (PGM-free) ORR catalyst is one of the most urgent needs in the PEFC field. Transition metal-nitrogen-carbon (M-N-C) catalysts obtained by heat-treating transition metals, nitrogen, and carbon precursors have been viewed as the promising PGM-free catalysts for PEFC cathodes though not yet meeting all the performance characteristics required. High ORR activity, excellent mass-transport properties and practical long-term durability must be achieved simultaneously for PGM-free catalyst to replace platinum-based catalysts in the PEFC system.

Specific catalyst precursors and heat-treatment conditions are critical to obtaining PGM-free catalysts with the desired structure. Iron- and cobalt-based zeolitic imidazolate frameworks (ZIFs), with uniformly distributed transition metals coordinated by N-containing ligands, have been viewed as highly suitable precursors for ORR catalysts. However, catalysts derived from these ZIFs often have low surface area and porosity, which are not conducive to efficient mass transport. On the contrary, ZIF-8, a zinc based ZIF, yields carbon with high surface area and porosity. Dodelet et al. used ZIF-8 as a microporous host for Fe and N precursors [1]. Following a heat-treatment at a relatively high temperature of 1050 ºC, they obtained an ORR catalyst with enhanced fuel cell performance. Unlike Fe and Co in Fe- and Co-based ZIFs, zinc in ZIF-8 is volatile during high-temperature treatment, likely acting as pore-forming agent. Base on that assumption, we used in this work a zinc salt instead of zinc ZIF as a precursor responsible for the formation of micropores in the catalyst. By using high heat-treatment temperature we also achieved more corrosion-resistant catalysts, while increasing the porosity and specific surface area from 315 m2 g-1 to 910 m2 g-1due to Zn evaporation.

Fe-CM-PANI(Zn) catalyst was synthesized using the two nitrogen-precursor (cyanamide and PANI) approach, developed previously by LANL[2]. ZnCl2 was mixed with nitrogen precursors and heat-treated at 1000 ºC to ensure complete removal of Zn. The resulting CM-PANI-Fe(Zn) catalyst had much higher surface area than the catalyst obtained without Zn under the same conditions. The half-wave potential (E½) for CM-PANI-Fe(Zn) in RDE testing in acidic electrolyte, 0.79 V vs. RHE, was similar to that obtained with Zn-free CM-PANI-Fe heat-treated at 900 ºC (optimal heat-treatment temperature for the Zn-free system), but by 0.16 V higher than E½ measured with the Zn-free catalyst heat-treated at 1000 ºC (Figure 1). While no immediate increase in the activity can be demonstrated with CM-PANI-Fe(Zn), this catalyst promises to have much improved stability thanks to being treated at a much higher temperature. Electrochemical and fuel cell testing of the Zn-derived catalyst, focusing in particular on the durability, is underway, as is the catalyst structure optimization. The results will be present at the meeting.

References

[1] E. Proietti, F. Jaouen1, M. Lefèvre, N. Larouche1, J. Tian, J. Herranz J.-P. Dodelet. Iron-based cathode catalyst with enhanced power density in polymer electrolyte membrane fuel cells. Nat. Commun. 2011, 2, 416.

[2] P. Zelenay. Non-Precious Metal Fuel Cell Cathodes: Catalyst Development and Electrode Structure Design. 2014 Hydrogen and Fuel Cells Program Annual Merit Review and Peer Evaluation Meeting (2014) available online at: http://www.hydrogen.energy.gov/pdfs/review14/fc107_zelenay_2014_o.pdf

Figure 1

2827

, , and

Platinum group metal (PGM)-free catalysts with high volumetric activity and low hydrogen peroxide yields are needed as replacements for expensive and scarce platinum-based catalysts for the oxygen reduction reaction (ORR). State-of-the-art PGM-free ORR catalysts are currently synthesized from a variety of carbon-nitrogen-iron (C-N-Fe) precursors. While these C-N-Fe catalysts have shown promising ORR activity and selectivity, a concern exists on the possibility of harmful Fenton (Fe2+/H2O2) and Fenton-like (Fe3+/H2O2) side reactions generating hydroxyl radicals risking the durability of polymer electrolyte fuel cells due to membrane degradation. Therefore, in an effort to mitigate that risk, development of Fe-free catalysts with similar ORR activity to the C-N-Fe system but with alternative transition metal precursors are being pursued.

Our previous results have shown improved ORR activity from (CM+PANI)-Co-C catalysts with a half-wave potential of 0.77 V nearing that of Fe-based catalysts [1]. While this improvement is significant, further development is necessary to enhance both activity and selectivity of Fe-free catalysts. In this work, we will exploit the synergism between theoretical and experimental approaches to develop binary Fe-free transition metal ORR catalysts. Earlier results from quantum-chemical models linking atomic scale structure to ORR activity based on density functional theory (DFT) suggest increased ORR activity for mixed-metal-species active site structures in the Fe-Co system [2]. The same effect was observed from the Mn-Co system. A Sabatier principle relationship between ORR activity and single-transition-metal active sites was observed in both theoretical and experimental studies. A similar experimental-theoretical relationship will be developed from our findings of binary Fe-free transition metal sites.

Acknowledgement

Financial support for this research by DOE-EERE through Fuel Cell Technologies Office is gratefully acknowledged.

[1] Martinez, U., Holby, E.F., Dumont, J.H., Chung, H.T., Zelenay, P. "Non-PGM ORR Catalysts Based on Transition Metals Alternative to Iron." ECS Meeting Abstracts, 2016.

[2] Holby, E.F., Taylor, C.D. "Activity of N-coordinated multi-metal-atom active site structures for Pt-free oxygen reduction reaction catalysis: Role of *OH ligands," Scientific Reports, 2015, 5, 9286.

2828

, , , , , , and

Due to the high cost of Pt-based catalysts for the anode and the cathode of polymer electrolyte fuel cells (PEFCs), their commercialization and distribution have been limited. Non-precious metal catalysts (NPMCs) with high activity and low peroxide yields [1, 2] represent a viable option for replacing the Pt-based catalysts for oxygen reduction reaction in the PEFC cathode. Graphene and graphene oxide-based catalysts have shown promising properties, such as good chemical stability and excellent conductivity, and they can be modified in a controllable manner by inducing functionalities [3]. However, the ORR activity of nitrogen-doped graphitic systems is lower than that of conventional nitrogen-doped mesoporous carbons. This is primarily due to fewer active sites on the basal plane than graphene edges and, secondly, due the re-stacking of n-doped graphitic sheets that forms an impermeable film, which limits the diffusion of reactants to the active sites. These key challenges have precluded the use of N-doped graphene-oxide catalysts for practical applications. In this presentation, we will show how the treatment of graphene oxide with solvents chosen based on Hansen's solubility parameters, followed by nitrogen doping, results in the formation of active ORR catalysts [4]. Furthermore, we will use electrochemical treatment to enhance the four-electron selectivity of catalysts in oxygen reduction. Structural, chemical, and morphological information by X-ray techniques (XRD and XPS) will be presented, along with electron microscopy (SEM, TEM). The ORR activity graphene of oxide-based catalysts in both alkaline and acidic media, determined using a rotating ring-disk electrode (RRDE), will be presented. The results of this study are expected to aide in the synthesis of well performing graphitic ORR electrocatalysts with tunable activity and could lead in the future to better understanding of the mechanism ORR active site formation.

Acknowledgement

Financial support from the Los Alamos National Laboratory, Laboratory-Directed Research and Development (LDRD) is gratefully acknowledged.

References

[1] Wu, G., More, K. L., Johnston, C. M., Zelenay, P., Science, 332, 443-447 (2011).

[2] Bashyam, R., Zelenay, P., Nature, 443, 63-66 (2006).

[3] Qu, L., Liu, Y., Baek, J.-B., Dai, L., ACS Nano, 4, 1321-1326 (2010)

[4] Martinez, U., Dumont, J. H., Holby, E. F., Artyushkova, K., Purdy, G. M., Singh, A., Mack, N. H., Atanassov, P., Cullen, D. A., More, K. L., Chhowalla, M., Zelenay, P., Dattelbaum, A. M., Mohite, A. D.,Gupta, G., Science Advances, 2, 3

2829

The sluggish kinetics of oxygen reduction reaction (ORR) at the cathode of proton exchange membrane fuel cells (PEMFC) demand a high loading of precious platinum (Pt) or Pt alloy electrocatalysts [1]. The extensive implementation of fuel cell technologies for stationary and vehicular applications, however, demands the replacement of high cost and less abundant Pt with low cost and natural abundant electrocatalysts. Recent studies show the use of various non-precious metals, metal alloys, metal oxides and nitrogen doped carbon as alternative catalysts for ORR [2, 3]. Even though, these non-precious metal catalysts experience severe dissolution and large agglomeration under vigorous PEMFC operational conditions and subsequent catalyst performance degradation. In this work, we developed a facile method for making non-precious ORR catalyst using nitrogen (3.7 at. %), sulfur (2.4 at. %) and iron (1.3 at. %) co-doped porous graphene (Fe-NSG). The ORR activity and durability of catalyst in acidic media was investigated by half (Fig. 1) and full cell electrochemical measurements. RDE measurements of the Fe-NSG catalyst confirm a four electron transfer ORR process with high current density. PEMFC full cell measurements of the catalyst give a maximum power density of 225 mW cm-2 at 80 °C with good stability [4]. The high ORR activity and good stability of Fe-NSG electrocatalyst can be ascribed due to, (a) the comparatively higher electronegativity of doped N and S atoms (N: 3.04, S: 2.58) with respect to carbon (C: 2.55), which induces more charged sites within the graphene lattice favorable for oxygen adsorption and reduction, (b) presence of the large amount of pyridinic nitrogen, which can coordinate with iron cations and increase the density of ORR active Fe-Nx centers, (c) an optimum amount of sulfur within the non-precious catalyst suppress the iron carbide formation, and increases the formation of Fe-NORR active species, (d) the unique confined morphology of Fe nanoparticles within the graphene layers suppress the agglomeration/dissolution of metal particles and increase their interfacial contact and durability.

 

 References 

  • F. Jaouen, E. Proietti, M. Lefevre, R. Chenitz, J.-P. Dodelet, G. Wu, H. T. Chung, C. M.

  • Johnston and P. Zelenay, Energy & Environmental Science, 2011, 4, 114-130.

  • Z. Chen, D. Higgins, A. Yu, L. Zhang and J. Zhang, Energy & Environmental Science, 2011, 4, 3167-3192.

  • M. Lefèvre, E. Proietti, F. Jaouen and J.-P. Dodelet, Science, 2009, 324, 71-74.

  • B. P. Vinayan, T. Diemant, R. J. Behm and S. Ramaprabhu, RSC Advances, 2015, 5, 66494-66501.

Figure 1

2830

, , , and

The most active non-precious metal catalysts (NPMCs) for oxygen reduction (ORR) to date are the pyrolyzed catalysts, inspired from Heme-like complexes. These are usually composed of iron and/or cobalt coordinated by nitrogens, claimed to resemble the structure of porphyrins and phthalocynines, on the surface of a carbon support. Unfortunately, the exact structure of the catalytic sites remains a mystery and the solution for this conundrum seems be almost impossible. The advantages of using such catalysts are: (1) their high activity and (2) low price, which is derived from the cost of their precursors. The disadvantages are: (1) low durability compared to precious metal catalysts, and (2) their unknown structure, which limits their further improvement in order to obtain both the activity and durability benchmarks needed to become good alternatives for precious metals.

One of the most promising options to resolve these issues is to find non-pyrolyzed molecular non-precious metal catalysts for ORR that could be tuned to match the necessary requirements. Recently, Metallo-corroles, a relatively new family of molecular catalysts was reported to have very good potential as non-precious metal catalysts for ORR. Inspired from the extensive work on the electropolymerization of metalloporphyrins, and its merits, we electropolymerized of metallocorroles on carbon electrodes to (1) increase their site density in order to overcome their relatively low electrocatalytic turnover frequency (when compared to platinum) and (2) use the proximity of active sites to move from 2- to 4-electroreduction of oxygen.  Metallocorroles substituted with anilines were studied and electropolymerized. They show a significant enhancement in the overpotential required for the execution of ORR and tendency for 4-electron reduction to water. The electropolymerization, its characterization and the effect on the electrocatalytic activity of the electropolymerized metallo-corroles vs. the monomers will be presented.

2831

, , , , and

Polymer electrolyte fuel cells are expected for the residential and transportable applications, due to their high power density and low operating temperature. The ENEFARMs (micro-co-generation system using 700 W PEFC system) are operating more than 150,000 units in Japan. Fuel cell vehicles are also commercially available in Japan.

However, the estimated amount of Pt reserve is too small to supply for the huge number of fuel cell devices. Pt is not stable in acid solution although Pt is relatively stable among the platinum group metals. The dissolution of Pt cathode might be the final problem to be solved related to the stability in the present PEFC system. In order to commercialize PEFCs systems more widely, the development of non-precious metal cathode is strongly required.  Additionally, the instability of carbon support is also a big problem especially for fuel cell vehicles. Carbon including graphite is thermochemically unstable at room temperature in air or oxygen containing atmosphere. A stable non-precious metal oxide cathode with stable metal oxide support without carbon might be the final goal for the cathode of PEFC for fuel cell vehicles.

We have reported that partially oxidized group 4 and 5 metal carbonitrides and organometallic complexes are stable in an acid solution and have definite catalytic activity for the oxygen reduction reaction (ORR) (1-4).  The stability of metal oxide were evaluated by the solubility of a metal oxide in the acid solution. In oxygen containing atmosphere most of metals except Au form metal oxide thermochemically. In this paper we will report our recent advancement of the group 4 and 5 metal oxide catalyst with metal oxide support without carbon.

Recently, we published the results of precious-metal-free and carbon-free cathodes based on oxides and demonstrated the superior durability of oxide-based cathodes by preparing titanium-niobium oxides mixed with Ti4O7 (TixNbyOz + Ti4O7) (5). The ORR activity of the TixNbyOz + Ti4O7 is higher than that of the Ti4O7, indicating that the TixNbyOz might have active sites for the ORR. The highest onset potential of the TixNbyOz +Ti4O7 was over 1.1 V with respect to reversible hydrogen electrode. No degradation of the ORR performance of TixNbyOz + Ti4O7 was observed during both start-stop and load cycle tests. Therefore, we successfully demonstrated that the precious-metal and carbon-free oxide-based cathodes had superior durability under the cathode conditions of a polymer electrolyte fuel cell.

In order to qualify the role of Nb oxide for ORR, we used TiO2-Nb (Nb; 0.5 or 5atm%) rods (TOSHIMA Manufacturing Co., Ltd) as working electrodes.  In addition, TiO2-Nb (0.5atm%) rod was heat-treated at 800 oC in 4%H2/Ar to examine the effect of heat-treatment under reductive atmosphere (0.5atm%, 800oC reduction).  All electrochemical measurements were performed in 0.1 mol dm-3 H2SO4 at 30 oC with a 3-electrode cell.  Chronoamperometry (CA) was performed from 0.2 to 1.2 V vs. RHE under O2 atmosphere to obtain ORR current. The ORR current density was normalized by the electric charge of the double layer capacitance under N2 atmosphere.

From the results of the potential-ORR current curves from 0.2 to 1.0 V of TiO2-Nb(0.5 or 5atm%) and (0.5atm%, 800oC reduction) electrodes, the ORR activity of the TiO2-Nb(0.5atm%) electrode was higher than that of the TiO2-Nb(5atm%) electrode.  Although Nb doping is necessary to have some electric conductivity, small amount of Nb doping might be enough to get high ORR activity. In addition, the ORR activity of the TiO2-Nb(0.5atm%_800oC reduction) was higher than that of the TiO2-Nb(0.5atm%). Therefore, we found that the heat-treatment under reductive atmosphere enhanced the ORR activity.

 

The authors wish to thank to the New Energy and Industrial Technology Development Organization (NEDO) for their financial support.

REFERENCES

1) A. Ishihara, Y. Shibata, S. Mitsushima, K. Ota, Journal of Electrochemical Society, 155, 2008, B400-B406 (2008).

2) A. Ishihara, M. Tamura, Y. Ohgi, M. Matsumoto, K. Matsuzawa, S. Mitsushima, H. Imai, K. Ota,

Journal of Physical Chemistry, ser. C, 117, 18837-18844 (2013).

3) A. Ishihara, M. Chisaka, Y. Ohgi, K. Matsuzawa, S. Mitsushima, K. Ota, Physical Chemistry Chemical Physics, 17, 7643-7647 (2015).

4) N Uehara, A. Ishihara, M Matsumoto, H. Imai, Y. Kohno, K. Matsuzawa, S. Mitsushima, K. Ota, Electrochimica Acta, doi:10.1016/j.electacta.2015.03.125.

5) A. Ishihara, M. Hamazaki, M. Arao, M. Matsumoto, H. Imai, Y. Kohno, K. Matsuzawa, S. Mitsushima, and K. Ota, Journal of The Electrochemical Society, 163 (7) F603-F609 (2016)

2832

, , , , , , , , , et al

Development of non-platinum electrocatalysts for oxygen reduction reaction (ORR) is required for wide commercialization of polymer electrolyte fuel cells. We focused on titanium oxide-based electrocatalysts, because of their high chemical stability and much less expensive than platinum. We found that Ti-Nb complex oxide mixed with magneli phase titanium oxide (Ti4O7) had high ORR onset potential and superior durability [1]. However, the ORR current was small because of their large particle size. In this study, we have tried to prepare titanium oxide-based nano particles using multi-walled carbon nano-tubes (CNTs) as support via hydrolysis method. We examined the effect of heat-treatment condition such as temperature and time on the ORR activity. We prepared Nb doped titanium oxide [mass ratio of TiO2:Nb2O5= 8:2] supported CNTs with total oxide mass ratio of 20wt%. These catalysts were heat treated at 600-900 oC for 0-60 min under 4%H2/Ar atmosphere. The catalyst without heat-treatment had poor ORR activity. The heat-treatment under reductive atmosphere enhanced the ORR activity of the catalysts. Figure 1 shows the dependence of the ORR current at 0.6 and 0.7 V of the catalysts prepared at several temperatures for 10 min under 4%H2/Ar in 0.1M H2SO4 at 30 oC. The catalyst prepared at 800 oC showed highest ORR activity. The SEM observation indicated that the sizes of the oxide-based particles were around 20 nm for all catalysts with and without heat treatment. Therefore, the surface area of the oxide-base particles was almost the same. The dependence of the ORR activity was not responsible for the change of the surface area. The XRD patterns revealed that the catalyst without heat treatment was composed of anatase phase of TiO2, and the peaks due to the rutile TiO2 apparently appeared at around 700-800oC. In particular, the peaks due to Rutile TiO2 shifted at lower angular, suggesting that the niobium doped into the TiO2 phase. Thus, the Nb doping into rutile TiO2 could affect the ORR activity. Figure 2 shows the Ti 2p XPS spectra of these catalysts. As shown in Fig.2, the peak at 457.8 eV corresponding to Ti3+ (Ti2O3) as well as the peak at 459.2 eV corresponding to Ti4+ (TiO2) were observed. The peak due to Ti3+ increased from 600 to 800 oC, and decreased at 900oC. This behavior was well corresponded to the ORR activity. Therefore, we concluded that Ti3+ could act as active sites for the ORR. According to the XRD patterns, the Nb doping into rutile TiO2 proceeded with increasing the temperature from 700 to 800oC. The niobium doping generated the Ti3+ via electro-neutral principle. The formation of the Ti3+ in TiO2 rutile phase was considered to be essential for emergence of the ORR. [1] A. Ishihara et al., J. Electrochem. Soc, 163(7), F603 (2016).

Figure 1

2833

, and

Conductive carbon black powders have been used to support platinum-group-metal (PGM) catalysts in polymer electrolyte fuel cells (PEFCs). The PEFC-powered vehicles have been commercialized recently, yet their prices remain high due to the scarce and expensive PGMs in the catalysts, as well as the need to protect carbon-black supports from corrosion.1,2 In PEFC, the PGM loading required for oxygen reduction reaction (ORR) at the cathode is an order of magnitude higher than that for hydrogen oxidation reaction at the anode.3 At potentials above 0.207 V versus the standard hydrogen electrode, the carbon support corrodes. The corrosion rate is especially high during the startup/shutdown of the cell, due to the so-called reverse-current decay mechanism which increases the cathode potential up to ~1.5 V.4 Therefore, non-PGM cathode catalysts and/or PGM catalysts on carbon-free supports have been extensively developed to reduce the cost of PEFC. Most of the non-PGM catalysts developed to date utilize graphitic carbon materials such as carbon black,5 carbon nanotube,6,7 graphene,7 carbonized polymers,5 carbonized heterocyclic compounds8 or carbonized Metal Organic Frameworks.9 Development of non-PGM cathode catalysts free from carbon supports remains challenging and only a few papers have been published, in which the reported geometrical current densities were moderate, the order of micro ampere per square centimeters in a practical potential range.10

In this work, support-free titanium oxynitride and zirconium oxynitride catalysts were synthesized using a recently developed solution phase combustion route11 with the modification of removing carbon-support. In 0.1 mol dm–3 H2SO4 solution, titanium oxynitride showed three orders of magnitude larger current densities than those studies.10 Compared with the previously synthesized carbon-supported titanium oxynitride,11 the activity was enhanced by 0.05 V of half-wave potential but the selectivity was similar, indicating that ORR proceeded on titanium oxynitride and carbon-support was not necessary. The support-free zirconium oxynitride showed no activity owing to the insulating nature, indicating that activity of the titanium oxynitride is not from the carbon traces from precursors. The higher conductivity of titanium oxynitrides as compared with zirconium oxynitrides is due to amenable nature to the incorporation of oxygen defects. The effect of the source of titanium, dispersant of precursors on the activity will be shown at the meeting.

 

Acknowledgments

The authors gratefully acknowledge Mr. Yusei Tsushima for his help with acquisition of transmission electron microscopy images. This work was partially supported by a Grant-in-Aid for Scientific Research (C) (26420132) from the Ministry of Education, Culture, Sports, Science, and Technology (MEXT) of Japan; a research grant from the Naoji Iwatani Foundation of Japan; a grant for chemical research from the Foundation for Japanese Chemical Research and a research grant from Nippon Sheet Glass Foundation for Materials Science and Engineering in Japan. The X-ray photoelectron spectra were acquired with the support by Nanotechnology Platform, 12024046 of the MEXT of Japan.

 

References

(1) T. Yoshida and K. Kojima, ECS Interface24, 45–49 (2015).

(2) U. Eberle, B. Müller, and R. von Helmolt, Energy Environ. Sci.5, 8780–8798 (2012).

(3) H. A. Gasteiger, J. E. Panels and S. G. Yan, J. Power Sources127, 162–171 (2004).

(4) C. A. Reiser, L. Bregoli, T. W. Patterson, J. S. Yi, J. D. Yang, M. L. Perry, and T. D. Jarvi, Electrochem. Solid-State Lett.8, A273–A276 (2005).

(5) Y. C. Wang, Y. J. Lai, L. Song, Z. Y. Zhou, J. G. Liu, Q. Wang, X. D. Yang, C. Chen, W. Shi, Y. P. Zheng, M. Rauf and S. G. Sun, Angew. Chem. Int. Ed.54, 9907–9910 (2015).

(6) M. Chisaka, A. Ishihara, N. Uehara, M. Matsumoto and K. Ota, J. Mater Chem. A, 3, 16414–16418 (2015).

(7) Y. Li, W. Zhou, H. Wang, L. Xie, Y. Liang, F. Wei, J. C. Idrobo, S. J. Pennycook and H. Dai, Nature Nanotechnol.7, 394–400 (2012).

(8) A. Serov, K. Artyushkova and P. Atanassov, Adv. Energy Mater., 4, 1301735-1– 1301735-7 (2014).

(9) E. Proietti, F. Jaouen, M. Lefevre, N. Larouche, J. Tian, J. Herranz and J. P. Dodelet, Nature Commun.2, 416-1–416-9 (2011).

(10) C. Gebauer, J. Fischer, M. Wassner, T. Diemant, J. Bansmann, N. Hüsing, and R. J. Behm, Electrochim. Acta 146, 335–345 (2014).

(11) M. Chisaka, Y. Ando and N. Itagaki, J. Mater Chem. A4, 2501–2508 (2016).

2834

and

Fuel cells offer potentially efficient, silent and low-pollution electrical power generation. Technical and economic problems have however, restricted their widespread development. One of the principal factors hampering the extensive commercialization of fuel cells is their dependence on platinum as electrocatalysts [1]. In recent years, many attempts have been made to reduce the platinum usage by introducing some metal oxide-based electrocatalysts. As a search of new metal oxide-based electrocatalysts, titanium oxide catalysts were studied for their catalytic activity during the oxygen reduction reaction (ORR) and stability in acid media, especially as a support material. It was reported that titanium oxide supported the oxygen reaction in acid [2] and alkaline [3] media. Early studies on titanium oxide catalysts showed that the catalytic activity for the ORR was low [2], however, Kim et al. [4] later showed that it depends on oxidation state, crystalline structure and work function of the titanium oxide catalysts. Moreover, although previous studies showed that the particle size of nanoparticles electrocatalysts can significantly affect their activity for the ORR [5], no work has been done to investigate the effect of particle size on catalytic performance of titanium oxide nanoparticles electrocatalysts. Therefore, due to some ambiguities and lack of knowledge in this research field, titanium oxide is worth being studied as an independent non-platinum electrocatalyst.

In this study, titanium dioxide (TiO2) nanoparticle powders with four different average particle size were purchased and characterized by X-ray diffraction analysis (XRD), small-angle X-ray scattering (SAXS) and scanning electron microscopy (SEM). The SAXS profiles showed that the particle size maximum probability for TiO2 powders with average particle size of 5 nm, 18 nm, 50 nm and 100 nm were 4 nm, 12 nm, 43 nm and 96 nm, respectively. The SEM images of TiO2 nanoparticle powders showed agglomerated particles with spherical structure that this agglomeration increased with decreasing the size of particles. The XRD patterns indicated anatase and rutile phases for TiO2 nanoparticles. The diffraction intensity of crystal planes for both anatase and rutile phases increased with increasing the particle size which can impact the catalytic activity of TiO2 catalysts for the ORR. The catalytic activity of the catalysts for the ORR was evaluated using cyclic voltammograms (CVs) and linear sweep voltammograms (LSVs) in 0.1 M HClO4 acid solution at 25 ºC. The results shed light on the effect of TiO2 particle size for the ORR.

References 

[1] C.D.A. Brady, E.J. Rees, G.T. Burstein, Electrocatalysis by Nanocrystalline Tungsten Carbides and the Effects of Codeposited Silver, J. Power Sources 179 (2008) 17-26.

[2] V.B. Baez, E. Graves, D. Pletcher, The Reduction of Oxygen on Titanium Oxide Electrodes, J. Electroanal. Chem. 340 (1992) 273-286.

[3] A. Fujishima, K. Honda, Electrochemical Photolysis of Water at a Semiconductor Electrode, Nature 238 (1972) 37-38.

[4] J-H Kim, A. Ishihara, S. Mitsushima, N. Kamiya, K-I Ota, Catalytic activity of titanium oxide for oxygen reduction reaction as a non-platinum catalyst for PEFC, Electrochim. Acta 52 (2007) 2492-2497.

[5] K. Kinoshita, Particle size Effects for Oxygen Reduction on Highly Dispersed Platinum in Acid Electrolytes, J. Electrochem. Soc. 137 (1990) 845-848.

C-32 Membrane Durability and Degradation - Oct 6 2016 2:00PM

2835

and

Electrolytic transport models available in the literature primarily deal with liquid media; models for solid electrolytes in polymer-electrolyte-membrane (PEM) fuel cells [1] largely derive from theories originally formulated to describe fluids. Since a solid can sustain a deformation at equilibrium under an applied stress or strain, there is a fundamental difference in the transport physics. Accurately modelling water-induced swelling in ionomer membranes requires that solid mechanics be reconciled with the transport and kinetic behaviour of electrochemical systems [2]. Previous efforts made to consider membrane swelling usually operate on a global basis, accounting only for an overall volume change; it is not clear whether models that include local swelling effects do so in thermodynamically consistent ways. This talk will deal specifically with swollen ionomers, although the principles introduced apply equally well to other solid ion conductors. In the case of Nafion, as Fig. 1 shows, water and ionic transport can produce, and dynamically interact with, non-uniform local volume changes and associated distributions of stress, making electrochemical/mechanical coupling significant.

Our strategy is to return to first principles, identifying and undoing the assumptions made specifically for liquid mixtures, and finding ways to ensure that the simultaneous use of ostensibly independent theories of kinematics, thermodynamics, and transport does not induce internal inconsistencies in the model. The incorporation of additional state variables such as pressure, stress, and strain requires the inclusion of new balance equations and constitutive laws: one must carry along a local balance of momentum, which relates the local pressure within the electrolyte to the distributions of stress and inertia; there must be a stress constitutive law that characterises the mechanical response (viscous, elastic, viscoelastic etc.); and strain must be defined in such a way that it ties deformed coordinates to a reference state while simultaneously abiding by the thermodynamic volume-explicit equation of state. Modifying the system of equations in the manner discussed above and tackling questions of model closure have helped us to produce a more robust and comprehensive set of equations to model Nafion.

The underlying aim is to combine developments in continuum poroelasticity theory with advanced electrochemical transport models. Ultimately, such a model will help us rationalise the physics associated with swollen vapour-equilibrated Nafion-like membranes in fuel cells under isothermal, non-isobaric, and constrained or free-swelling conditions. Such a model could inform better water management schemes and efforts to improve membrane durability. It will also be useful to have a general framework for modelling solid electrolytes that can be applied to materials beyond Nafion where electrochemical/mechanical coupling is important.

[1] Weber, A.Z. and J. Newman, Transport in polymer-electrolyte membranes - II. Mathematical model. Journal of the Electrochemical Society, 2004. 151(2): p. A311-A325.

[2] Kusoglu, A. and A.Z. Weber, Electrochemical/Mechanical Coupling in Ion-Conducting Soft Matter. The Journal of Physical Chemistry Letters, 2015. 6(22): p. 4547-4552.

Figure 1

2836

, , and

Membrane stability is an important consideration for the overall durability and lifetime of polymer electrolyte fuel cells. During operating conditions, the ionomer membrane is subjected to chemical degradation due to the formation of radicals and their subsequent attack on the chemical bonds of the polymer [1,2]. The membrane is also subjected to mechanical degradation due to the hygrothermal cycles during operation, which causes the membrane to expand and contract repeatedly leading to the development of cyclic stresses [3,4]. Recent studies show that the mechanical and chemical degradation processes are connected and, once combined, cause accelerated degradation and reduced integrity of the membrane [3–6].

On one hand, the chemical degradation causes a change in the ionomer membrane mechanical properties, transforming it from a ductile material to a brittle material [3]. This transformation makes the membrane less resistant to crack formation and propagation and increases the mechanical damage that can be caused by typical hydrothermal cycles. On the other hand, it was observed that applying compression to the membrane increases the rate of chemical decomposition [7]. Understanding the interaction between the chemical and mechanical degradation processes is therefore an important step in developing more durable membranes and improving the performance and lifetime of the fuel cell.

We propose a statistical model to establish a correlation between the chemical degradation at the level of the polymer backbone chains and the initiation of microcracks due to mechanical stress. The representation of the morphology of the ionomer is based on the fibrillary structure described in [6,8]. In this representation a bundle of backbone chains is considered as a basic building block of the structure. We generate a network of bundles representative of a microcrack initiation site.

A crack initiation in the membrane structure is the tipping point that leads to an accelerated degradation in terms of hydrogen leaks and ultimate failure of the fuel cell under regular operating conditions. Once a crack is initiated, it increases the amount of reactant crossover leading to an increasing chemical decomposition [4]. The crack also constitutes a stress concentration site where the cyclic mechanical load participates further in its propagation. The combination of these processes accelerates the overall rate degradation leading to a complete failure of the ionomer.

Clearly, the crack initiation process is a critical root cause for the accelerated cycle of degradation of the ionomer. Our model allows for the estimation of the time of the crack initiation under relevant mechanical stresses and follows a realistic chemical degradation pattern. The statistical analysis provides a comprehensive understanding of the interaction between the local amounts and distribution of chemical decomposition sites and the mechanical stressors causing the initiation of the physical damage in the ionomer structure.

[1] K.H. Wong, E. Kjeang, J. Electrochem. Soc. 161 (2014) F823.

[2] K.H. Wong, E. Kjeang, ChemSusChem 8 (2015) 1072.

[3] A. Sadeghi Alavijeh, M.A. Goulet, R.M.H. Khorasany, J. Ghataurah, C. Lim, M. Lauritzen, E. Kjeang, G.G. Wang, R.K.N.D. Rajapakse, Fuel Cells 15 (2015) 204.

[4] A. Kusoglu, A.Z. Weber, J. Phys. Chem. Lett. 6 (2015) 4547.

[5] R.M.H. Khorasany, A. Sadeghi Alavijeh, E. Kjeang, G.G. Wang, R.K.N.D. Rajapakse, J. Power Sources 274 (2015) 1208.

[6] P.-É.A. Melchy, M.H. Eikerling, J. Phys. Condens. Matter 27 (2015) 325103.

[7] A. Kusoglu, M. Calabrese, A.Z. Weber, ECS Electrochem. Lett. 3 (2014) F33.

[8] L. Rubatat, G. Gebel, O. Diat, Macromolecules 37 (2004) 7772.

2837

and

Cationic contamination in polymer electrolyte membrane fuel cells (PEFCs) can lead to accelerated degradation and loss of performance. There are several mechanisms to introduce these cationic impurities; including transport with the inlet fuel/air streams, or from degradation of the fuel cell and balance of plant components. Many common cations tend to have a higher affinity to the sulfonic acid side chains of the polymer electrolyte membrane (PEM) than protons, resulting in preferential uptake of these contaminants into the PEM. Once occupying these sites, the impurity cations contribute to the loss of performance by lowering the water content of the membrane and reducing the ionic conductivity.[1] To compliment experimental results, numerical studies have been performed[2,3] to understand the coupled impact of cation occupancy in the membrane with the water transport in the cell and the overall cell performance. Early generations of these models had assumed a constant cation occupancy in the cathode that was used as a boundary condition to solve the cation transport via a Stefan-Maxwell diffusion process. By coupling the contaminant occupancy to the water content of the membrane and electro-kinetics of the catalyst layers, cell performance has been shown to decrease down to as little as one third its non-contaminated levels.[3] This study builds off of previous models to develop a steady state, one dimensional, cation contamination model of a PEFC to examine the impact of cation transport to the membrane electrode assembly (MEA) via a water bridge across the gas diffusion layer (GDL) as shown in Figure 1. The multiphase mixture (M2) model, proposed by Wang and Cheng[4], is used to derive species transport equations for the reactant gases and water that permits two phase water transport in the GDL and catalyst layers. This resulting water distribution acts as to promote cation transport from the gas flow channels across the GDL, where the dissolved cations can evolve out of the liquid phase water and into the MEA, occupying the sulfonic acid sites of the membrane. A baseline non-contaminated case is presented to study the cell performance and species distribution normal operating conditions. Then, the model is employed to show how the saturation of the cathode effects cation distribution across the GDL and how the presence of dissolved cations in the catalyst layer impacts the contaminant occupancy and overall cell performance.

References

 1. T. Okada, in Handbook of Fuel Cells, Cells– Fundamentals, Technology and Applications., W. Vielstich, H. A. Gasteiger, A. Lamm and H. Yokokawa. Editors, John Wiley & Sons, Ltd (2010).

 2. M. F. Serincan, U. Pasaogullari and T. Molter, Int. J. Jydrogen Energy, 35, 11 (2010).

 3. M. A. Uddin and U. Pasaogullari, J. Electrochem. Soc., 161, 10 (2014).

 4. C. Y. Wang and P. Cheng, Int. J. Heat Mass Transfer, 39, 17 (1996).

Figure 1

2838

, , and

Introduction

Polymer electrolyte fuel cells (PEFCs) for automotive applications must operate over a wide range of operation conditions. Humidity cycle tests that simulate the operation of fuel cell under dry and wet conditions are widely used to test the mechanical durability of polymer electrolyte membrane of PEFCs in the form of membrane electrode assembly (MEA). The degradation phenomenon of membrane under conditions of cycling humidity has been studied by many researchers[1-5], but the mechanism is not clear enough. In this study, correlation between the humidity cycle durability of MEAs equipped with different type of membranes and their mechanical properties obtained by dynamic mechanical analysis (DMA) and thermal mechanical analysis (TMA) was investigated.

 

Experimental

MEAs of different catalyst layer thickness were prepared using NR-211(25μm) membranes or sulfonated polyethersulfone (SPES) membranes (32μm). Humidity cycle tests using a rectangular waveform of 0-150% RH (4min per cycle) were carried out on a JARI standard single cell (25cm2 electrode area, one serpentine flow channel). The cell temperature was set at 80°C. Linear sweep voltammetry was used to measure the hydrogen crossover rate through membrane as a diagnostic of the degradation status of the membrane in the humidity cycle tests. The cycle tests were finished when the hydrogen crossover rate through the membrane reached 10 times the initial value. To consider the reason of difference of the cycle life of MEAs, swelling tests in water was conducted at 80°C.Moreover, dynamic mechanical analysis (DMA), thermal mechanical analysis (TMA) of MEAs were carried out by changing the humidity of atmospheric gas.

 

Result and Discussion

The cycle life, which was defined as the terminated cycle, by humidity cycle tests were shown in Fig.1 The cycle life of SPES MEA was greatly shorter than that of NR-211 MEA at the same catalyst layer thickness (10μm). To consider the difference of humidity cycle durability of both membranes, DMA of membrane in dry nitrogen gas was carried out. Fig.2 shows the relationship between temperature and storage elastic modulus of membranes measured under the condition of 0%RH of relative humidity. Storage elastic modulus of SPES membrane was higher than that of NR-211 membrane. This result shows that the membrane stiffness of SPES was higher than that of NR-211 membrane at 0%RH. The difference of humidity cycle durability of both membranes was not explained by the result of DMA.

Swelling tests of membranes were conducted to consider the reason of difference of cycle life of both MEAs. Swelling ratio of NR-211 and SPES were shown in Fig.1. Swelling ratio of SPES were greatly larger than that of NR-211. Our previous research revealed that the mechanical durability in humidity cycle test was influenced by the catalyst layer thickness of MEA[6]. In the case of this study, the thickness of both MEAs were same(10μm). The difference of cycle life of both MEAs was thought to be influenced by the difference of swelling ratio of both MEAs.These results indicate that it is important to prevent membrane swelling in order to improve the mechanical durability under humidity cycle .

 

ACKNOWLEDGMENTS

This work was supported by the New Energy and Industrial Technology Organization (NEDO).

 

REFERENCES

 [1] A. Kusoglu, A. M. Karlsson, M. H. Santare, S. Cleghorn, and W. B. Johnson, J. Electrochem. Soc. 157, B705-B713 (2010).

[2] Y.-H. Lai, Y. Li, and J. A. Rock, J. Power Sources, 195, 3215-3223 (2010).

[3] F. E. Hizir, S. O. Ural, E. C. Kumbur, and M. M. Mench, J. Power Sources, 195, 3463-3471 (2010).

[4] M. N. Silberstein and M. C. Boyce, J. Power Sources, 196, 3452-3460 (2011).

[5] T. T. Aindow and J. O'Neill, J. Power Sources, 196, 3851-3854 (2011).

[6] Y. Hashimasa, T. Numata, N. Yoshimura, J. Power Sources, 265, 30-35 (2014).

Figure 1

2839

, and

Polymer electrolyte fuel cell (PEFC) is strongly expected as a next generation power source because of its high efficiency, high power density, and purity of exhaust gas. However, there are some problems left such as infrastructure of hydrogen gas, durability of cell, and production cost. Performance of PEFC depends strongly on proton transport property in polymer electrolyte membrane (PEM). Proton transport property of PEM has been evaluated by experiments and numerical simulations by many researchers. The parts of separator used for holding PEM and gas flow channel are typically made of graphite materials because they require mechanical strength, corrosion resistance, and electric conductivity. However, using graphite material causes increasing the cost. Alternatively, using metal materials, such as stainless alloy, for the separator can reduce the cost although this causes the metal ion liquation, which leads to decrease in power density of PEFC [1]. To deal with this problem, more information about the behavior of metal ion and its influence on proton transport in PEM is required. However, it is difficult to analyze the effect of metal ion directly in experiment because the proton transport property depends on the nano-scale structure of PEM and water molecules.

In our previous work [2], characteristic changes in diffusion coefficients of proton were observed with ferrous ion contamination by using molecular dynamics simulations. Ferrous ion contamination has two totally different effects on proton transport. One is the connecting water clusters and enhancing proton transport at ER = 50%. The other is disintegrating water clusters and inhibiting proton transport at ER >= 70%. However, the accuracy of those results was insufficient because the model of ferrous ion was treated as the static charged particle with coulomb interaction. In regard to this problem, the polarizable force field model is effective to describe ferrous ion complex [3]. In this study, we analyze the structural property around the ferrous ion with more accurate force field model of ferrous ion considering with induced dipole interaction. Polarizability is considered on the ferrous ion atom and the oxygen atom in water molecule. The model of water molecule is developed based on aSPC/Fw model [4]. We investigated hydrated Nafion membrane in the case of water content λ=3, 6, 9, 12. We evaluated the radial distribution function between ferrous ion and water molecule, hydronium ion, or sulfonate groups. The cluster analysis of water region will be carried out. Finally we discussed the correlation between the concentration of ferrous ion and nanoscale structural changes of Nafion membrane, and estimated the change of proton transport property.

References

[1] A. Pozio, R. F. Silva, M. De Francesco, L. Giorgi, Electrochim Acta48, 1543 (2003).

[2] K. Kawai, T. Mabuchi, and T. Tokumasu, ECS trans., 69, 17, 579 (2015).

[3] D. Semrouni, W. C. Isley III, C. Clavaguéra, J. P. Dognon, C. J. Cramer, and L. Gagriardi, J. Chem. Theory Comput., 9, 3062 (2013).

[4] K. Park, W. L. F. Paesani, J. Phis. Chem. B, 116, 343 (2012).

2840

, , , and

The durability of polymer electrolyte fuel cells (PEFCs) is one of the main challenges facing the commercialization of automotive fuel cells. The lifetime of the membrane, a critical structural component of PEFCs, is one of the principal obstacles in achieving the fuel cell industry durability targets [1]. Under dynamic automotive operating conditions and duty cycles, the membrane is subjected to chemical and mechanical degradation, which could cause hydrogen leaks and ultimate cell failure. Chemical degradation is linked to the polymer molecular decomposition caused by radical species formed during the fuel cell operation as by-products of electrochemical reactions [2]. On the other hand, mechanical degradation is attributed to the fracture caused by the induced mechanical and hygrothermal stresses in a constrained cell [2]. The US Department of Energy (DOE) introduced standardized in-situ accelerated stress test (AST) protocols [3]. Following the DOE mechanical AST protocol, the membrane mechanical durability under pure humidity cycling was investigated [4]. In this manner, two customized in-situ mechanical AST protocols were utilized by our group to evaluate the mechanical durability of PEFCs, indicating significant decay in mechanical properties, formation of microstructural cracks, and initiation of failure [5]. Despite the valuable outputs of the in-situ studies, the current protocols are time consuming and costly. As hygrothermal fatigue is expected to dominate the membrane mechanical lifetime, an ex-situ mechanical fatigue-creep based AST was recently developed by our group as a more convenient alternative [6]. The proposed ex-situ tensile fatigue-creep test demonstrated in this work is intended to evaluate the mechanical durability of catalyst coated membranes (CCMs) in a fraction of the time required for the conventional in-situ tests [7].

The proposed ex-situ tensile fatigue-creep accelerated stress test (TFC-AST) was conducted on dog bone shaped [8] CCM samples using a dynamic mechanical analyzer (DMA) equipped with environmental chamber. After equilibration at 80°C and 50% RH, as illustrated in Figure 1, a high frequency sinusoidal cyclic tensile load with stress ratio (R) of 0.2 and 6.1 MPa mean stress was applied on the CCM samples to certain fractions of fatigue lifetime (~140,000 cycles) [7]. When compared to the well-defined standard in-situ mechanical ASTs which last about 8 weeks [5], the proposed TFC-AST results in mechanical failure ~400 times faster than the in-situ tests due to higher frequency and magnitude of fatigue and creep loading. Depending on the total lifetime, TFC-ASTs were interrupted at different fractions of the CCM lifetime, i.e. 20%, 40%, 60%, and 80%, and partially fatigued samples were extracted for analysis.

The obtained partially degraded CCMs were further studied through mechanical and microstructural techniques. The mechanical properties of the extracted CCMs were investigated via tensile and hygrothermal expansion experiments in the same manner as reported in [9] using DMA. Tensile tests revealed remarkable increase in the tensile strength of the partially degraded samples indicating the alignment of the polymer molecules along the TFC stress direction. CCM thermal and hygral expansions were evaluated by stepwise increase in temperature and relative humidity, respectively. Interestingly, the CCM was found to contract, which is contradictory to the typical hygrothermal expansion behaviour of these materials [5,9]. This behaviour can be attributed to the exclusively tensile loading of the TFC-AST. In addition to the mechanical testing, morphological evolution of the TFC-AST degraded CCMs was also examined and compared with the analogous in-situ mechanical AST degraded CCMs using transmission electron microscopy (TEM). The TEM micrographs provided supplementary evidence regarding the reorientation of membrane molecules along the TFC stress direction. However, the mechanical failure of the specimens was found to be dominated by fatigue, similarly to the corresponding in-situ tests.

Figure 1. Schematic of the proposed tensile fatigue-creep accelerated stress test (TFC-AST) protocol for rapid mechanical durability testing of fuel cells.

Acknowledgements:

This research was supported by Ballard Power Systems and the Natural Sciences and Engineering Research Council of Canada through an Automotive Partnership Canada grant.

References:

[1] Fuel Cell Technical Team Roadmap Hydrogen Storage Technologies Roadmap, 2013.

[2] C.S. Gittleman, et al. (Eds.), Polymer Electrolyte Fuel Cell Degradation, Elsevier Inc., 2012, pp. 15–88.

[3] U.S. Department of Energy, DOE Cell Component Accelerated Stress Test, 2010.

[4] Y.H. Lai, et al., J Fuel Cell Sci Tech, 6, 021002, 2009.

[5] A. Sadeghi Alavijeh, et al., J ElectrochemSoc. 162, F1461, 2015.

[6] R. Khorasany et al. J Power Sources, 274, 1208, 2015.

[7] A. Sadeghi Alavijeh, et al., J Power Sources, 312, 123, 2016.

[8] R. Khorasany et al., Int J Hydrogen Energy, in press.

[9] A. Sadeghi Alavijeh, et al., Fuel Cells, 15, 204, 2015.

Figure 1

2841

, , , and

Perfluorosulfonic acid (PFSA) ionomer membranes are subjected to simultaneous chemical and mechanical degradation under fuel cell operation. Open circuit voltage (OCV) operation is typically employed as an in-situ accelerated stress test (AST) to expedite the chemical degradation which plays a key role in membrane thinning. The in-situ mechanical degradation is typically produced using wet/dry humidity cycles leading to micro crack initiation/propagation in the membrane [1]. Characterization of morphological changes in the membrane electrode assembly (MEA) is helpful in understanding the degradation mechanisms at play during fuel cell operation. 2D visualization techniques, such as optical and electron microscopy [2], have typically been used to determine the structural features of the MEA. However, these techniques are destructive and 2D in nature, and demand elaborate sample preparation. The X-ray computed tomography (XCT) technique overcomes these limitations by integrating several 2D images acquired at diverse incident angles into a virtual 3D image. Furthermore, modern XCT systems achieve sub µm resolution which is adequate for imaging membrane cracks and other damage features. In this work, the XCT technique is employed to investigate the structural evolution of membrane degradation over time in the presence of combined chemical and mechanical stressors. Additionally, the observed trends are correlated with various material and diagnostic properties to develop a better understanding of the degradation patterns.

Partially degraded MEAs extracted after different numbers of operating cycles when subjected to cyclic open circuit voltage (COCV) AST protocol are visualized using the XCT technique and the evolution of membrane degradation is studied by mapping the 3D structural/morphological changes over time. Crack distribution and morphology are examined from various perspectives by studying the 2D planar and cross-sectional views of the 3D reconstructed images. No through-thickness membrane crack is detected up to 60% of the MEA lifetime which suggests that sizeable crack development occurs mainly during the 60-85% period of the MEA lifetime. A detailed survey exhibits eight cracks at end of life (EOL) and five cracks at 85% degradation across a survey area of 0.44 mm2. All identified membrane cracks are observed to have a single distinct fragment of X or I-shape on the surface of the membrane. The X-cracks in their initial stage of development observed after 85% degradation are likely to grow along their width, while I-cracks tend to be slender, as shown in Figure 1. In addition, about 31% of cracks after 85% degradation and 50% of cracks at EOL are exclusive membrane cracks without any connectivity with the catalyst layers. The membrane crack width is almost uniform along its length after 85% degradation but has greater variation along the length at EOL resulting in an increased maximum crack width to crack length ratio likely caused by in-plane stresses [3]. The rapid decrease in the fracture strain of the membrane during the AST suggests increasing brittleness, which represents the most pronounced change in the mechanical properties [4]. This illustrates that the membrane encounters reduced mechanical strength and leads to more branching of cracks with increasing AST cycles. The combined effect of mechanical and chemical degradation is likely to have created favorable conditions for crack initiation and propagation inside the mechanically weakened membrane.

The work summarized here is a unique attempt to study the evolution of membrane degradation with a 3D perspective. These new findings demonstrate that adoption of XCT technology can provide a distinct advantage in understanding the pattern of membrane degradation, thereby enabling the capture of critical failure modes that may be invoked at different stages of fuel cell operation.

Acknowledgement

This research was funded by the Natural Sciences and Engineering Research Council of Canada, Canada Foundation for Innovation, British Columbia Knowledge Development Fund, and Ballard Power Systems through an Automotive Partnership Canada (APC) grant.

References

[1]R.M.H. Khorasany, A.S Alavijeh, E. Kjeang, G.G. Wang, R.K.N.D. Rajapakse. J. Power Sources 274 (2015) 1208-1216

[2]K.S Alcantara, D.C. Martınez-Casillas, K.B. Zheng, Q. Zhu, M. Abdellah, D. Haase, T. Pullerits, O. Solorza-Feria, S.E. Canton. Int. J. Hydrogen Energy, 39 (10) 5358–5370.

[3]Y. Singh, O. Luo, F.P. Orfino, M. Dutta, E. Kjeang, 3D Analysis of PEM Fuel Cell Membrane Cracks using X-Ray Computed Tomography, in: 228th Meet. Electrochem. Soc., Phoenix, 2015.

[4] C. Lim, L. Ghassemzadeh, F. Van Hove, M. Lauritzen, J. Kolodziej, G.G. Wang, S. Holdcroft, E. Kjeang. J. Power Sources. 257 (2014) 102.

Figure 1

2842

and

Ceria supported membranes have been proposed to mitigate chemical membrane degradation in polymer electrolyte fuel cells. It was confirmed that ceria as a radical scavenger protects the membrane under open circuit voltage (OCV) or ex-situ Fenton's durability conditions [1-7]. However, its effectiveness has not been examined for cell voltages below OCV, which are necessary conditions for field operation of fuel cells. On the other hand, ceria membrane additive can be considered as a cation contamination since it is dissolved in the membrane [8]. Performance tradeoffs have been observed experimentally with the use of ceria membrane additive, and the tradeoffs are more significant at lower cell voltages [9]. Therefore, a comprehensive investigation on chemical membrane mitigation and performance tradeoffs in ceria-supported fuel cells is required.

In the present work, a transient in-situ chemical degradation model for simulating the transport and reaction of cerium redox couples in the membrane electrode assembly (MEA) is developed and integrated with the chemical membrane degradation models in which the effects of iron redox couples on membrane degradation and detailed ionomer degradation processes are included [10-11].The developed model is then applied to investigate the mitigation effectiveness of cerium redox couples and the fundamental mechanisms for the performance tradeoffs under different cell voltage conditions.

The modeling results reveal that abundant Ce(III) ions are available in the membrane to quench hydroxyl radicals at high cell voltages. Since the hydroxyl radical is the dominant reactive species to attack the membrane ionomer, the modeling results demonstrate the primary mechanism for the significant mitigation observed in ceria supported MEAs at OCV conditions. However, this type of mitigation is found to be suppressed at low cell voltages, where electromigration drives Ce(III) ions into the cathode catalyst layer and reduces the available Ce(III) ion in the membrane. Ce(III) ion is the dominant species for hydroxyl radical quenching, and inadequate amount of Ce(III) ion in the membrane leads to a ten-fold reduction in the mitigation effectiveness at cell voltages below 0.7 V.

The simulated Ce(III) ion migration to the cathode catalyst layer is found to be responsible for the performance losses observed in ceria supported MEAs. The modeling results reveal that proton starvation can occur in the cathode catalyst layer due to the local Ce(III) ion accumulation. Without the adequate supply of protons in the ionomer of the cathode catalyst layer, proton conductivity and oxygen reduction kinetics are reduced. Significant performance tradeoffs in the form of combined ohmic and kinetic voltage losses are therefore evident and shown to increase with current density.

Overall membrane durability and fuel cell performance management is shown to be possible at high cell voltages with the use of ceria membrane additive, where cerium migration and the associated performance loss are insignificant. Unfortunately, at low cell voltages additional steps must be taken to address proton starvation in the cathode catalyst layer and inadequate amount of Ce(III) ion in the membrane in order to achieve a durable membrane without compromising fuel cell performance.

Acknowledgement:

This research was supported by Ballard Power Systems and the Natural Sciences and Engineering Research Council of Canada through an Automotive Partnership Canada (APC) grant. The authors wish to thank their colleagues at SFU FCReL and Ballard for providing valuable comments and advices.

References:

[1] Coms et al., ECS Trans. 16 (2) (2008) 1735–1747

[2] Xiao et al., J. Power Sources 195 (16) (2010) 5305 – 5311

[3] Wang et al., J. Membrane Sci. 421422 (0) (2012) 201 – 210

[4] Pearman et al., J. Power Sources 225 (0) (2013) 75 – 83

[5] Pearman et al., Polym. Degard. Stabil. 98 (9) (2013) 1766 – 1772

[6] Wang et al., Electrochimica Acta 109 (0) (2013) 775 – 780

[7] Lim et al., ECS Electrochem. Lett. 4 (4) (2015) F29–F31

[8] Hayes et al., J. Electrochem. Soc. 149(12) (2002) C623–C630

[9] Cheng et al., J. Electrochem. Soc., 160(1) (2013) F27–F33

[10] Wong et al., J. Electrochem. Soc., 161(9) (2014) F823–F832

[11] Wong et al., ChemSusChem, 8(6) (2015) 1072–1082

2844

A fuel cell vehicle (FCV) is a vehicle that utilizes fuel cells as its power source. Development of fuel cells for FCV is always aimed at a more compactsize by means of higher power density and longer driving range through enhanced energy efficiency. Fuel cells also require endurance reliability comparable to internal combustion engines, and cost reductions will be required to encourage widespread consumer use.

The power generating part of a fuel cell is called the membrane electrode assembly (MEA), which consists of a polymer electrolyte membrane (PEM) and electrodes with a catalyst. The performance of the fuel cell is mainly determined by the activity of the catalyst on each electrode and by the proton conductivity of the PEM. Because the performance, energy efficiency, durability, and cost of fuel cells are all related and involve trade-offs, optimization of the PEM design cannot be readily achieved. A particular issue that has faced optimization of the PEM design is the amount of time it takes to estimate PEM durability. The only method available has been to test it over a long period of time, and this is why FCV has had such an extended development period.

In order to optimize the PEM design in less time, a new method capable of estimating durability without durability testing must be developed. To this aim, we first made it possible to quantitatively evaluate the rate of chemical degradation of PEM to estimate durability and predict its lifetime. Next, we developed a mathematical formula for the rate of chemical degradation of the PEM to predict its service life.

Two possible reaction mechanisms have been proposed as the chemical degradation of the perfluorosulfonic acid (PFSA) membrane. One is the scission of the main chains of the polymer and the other is an unzipping reaction in which the end groups of the main chains progressively degrade. In order to quantitatively estimate the rate of chemical degradation of the PFSA membrane, we numerically modelized these two reaction mechanisms. When a scission of the main chain occurs, the location of the split becomes two new end groups, which suggests that the number of points of origin for an unzipping reaction increases exponentially. The fluoride release behavior was calculated based on this model. As a result, we confirmed that the experimental fluoride release data can be explained by these two reaction mechanisms of chemical degradation. In order to describe the fluoride release quantity F as a function of duration time t, exponential function Eq. (1) was formulated to closely match the experimental results.

F(t) = a [exp(bt) - 1] (1)

One of the factors that influence the chemical degradation rate of the PEM is contamination by metal ions. Adding a radical quencher is a well-known method to assure durability of the PEM by mitigating chemical degradation. In order to estimate the acceleration of PEM degradation caused by metal ion contamination or the mitigation effects offered by a radical quencher, durability testing was carried out using different levels of metal ion contamination or radical quencher additives. The results of these tests show that fluoride release behavior can be represented by Eq. (1) by fixing coefficient b and varying only coefficient a. Coefficient ais therefore defined as the accelerating factor of chemical degradation. The accelerating factor can be obtained by making Eq. (1) fit the experimental fluoride release data, and it can be used as an indicator of the severity of chemical degradation. Introducing the concept of accelerating factor to represent the fluoride release rate made it possible to deal quantitatively with the effect of factors that accelerate or mitigate chemical degradation. Moreover, the size of acceleration or mitigation effect caused by multiple factors can be indicated with a single variable.

We confirmed changes in accelerating factor during FCV operation with bench testing, which demonstrated the driving modes of an actual vehicle using a fuel cell stack with the same specifications as stacks installed in actual vehicles. The fluoride release rate was calculated from the accelerating factor envisioning actual FCV operation. Fluoride release behavior can be converted into membrane thickness loss. In this way, decrease in membrane thickness during FCV operation and the lifetime of the PEM can be predicted using calculations.

2845

, , , , and

Cerium enhances polymer electrolyte membrane (PEM) fuel cell durability by scavenging damaging radical species which are generated during operation. However, during cell fabrication, conditioning, and discharge, cerium migrates between the PEM and catalyst layers (CLs), which can reduce cell performance [1]. In this work, cerium migration within the cell and washout from it were quantified in membrane electrode assemblies (MEAs) which were fabricated using PEMs containing cerium concentrations of ~6.0µg/cm2. MEAs were subjected to accelerated stress tests (ASTs) at open circuit voltage (OCV) with cell temperatures of 90°C in hydrogen/air. ASTs were performed at three different humidity conditions: 100% RH, 30% RH, and wet/dry cycling ASTs, where the cell was cyclically exposed to humidified and dry reactant gasses for 30/45s, respectively [2].

After ASTs, through-thickness and in-plane cerium migration profiles were characterized in MEAs using X-ray fluorescence. In-plane cerium migration from the gasketed, inactive PEM border region into the active area was observed and correlated to the active area water content during the AST. Cerium concentration in the active area increased from ~6.0 to 15.3µg/cm2 after 2,000h of 100% RH operation. After 456h of 30% RH operation, active area water content was less than at 100% RH, and cerium concentration increased to 9.84µg/cm2. After 823h of humidity cycling, cerium concentration remained unchanged.

Through-thickness cerium migration from the PEM into the CLs is enhanced by low humidity operation. 30% RH operation resulted in uniform migration of cerium into both the anode and cathode CLs. PEM cerium was reduced from ~6.0 to 3.7µg/cm2, while anode and cathode CL concentrations were uniformly increased from 0.0 to 2.3 and 3.5µg/cm2, respectively. Conversely, after 100% RH operation, more cerium remained in the active area of the PEM. In-plane concentrations were maintained at ~6.0µg/cm2 near the inlet, however, concentration increased linearly to ~15µg/cm2 near the outlet. These concentration gradients were attributed to the condensation and subsequent flow of liquid water from cell inlet to outlet. Subjecting the MEA to humidity cycling resulted in both significant through-thickness cerium migration out of the PEM, as well as concentration gradients from PEM inlet to outlet. Under these conditions, average PEM concentration was reduced to 0.98µg/cm2. PEM cerium was depleted from the inlet and its concentration was only 1.5µg/cm2 near the outlet. Average anode and cathode CL cerium concentrations increased to 1.9 and 3.1µg/cm2, respectively. Cerium depletion which results from such conditions could reduce the amount of available radical scavengers, leaving the PEM more susceptible to chemical attacks.

Effluent cell water samples were collected from the anode and cathode during ASTs and analyzed to measure fluoride and cerium concentrations using ion chromatography and inductively coupled plasma mass spectrometry. Fluoride and cerium emissions are correlated, which suggests that ionomer degradation products serve as counter-ions for emission from the cell. The most aggressive test conditions, however, only reduced the total cerium inventory in the MEA by < 0.5%.

Fluoride emission rates (FERs) are also correlated to the average PEM cerium content after ASTs, which indicates a relationship between membrane degradation and cerium migration (Figure 1). We propose that during ASTs, cerium migration from the PEM into the CLs is driven by both concentration gradients between the PEM and CL, which arise due to cell humidification, and from degradation, itself, which is proportional to AST aggressiveness (wet/dry cycling > 30% RH >> 100% RH). We also postulate that cerium interacts with carbon catalyst support particles in the CLs which prevents its ionic equilibration with PEM ionomer. Ex situ experiments demonstrate the ability of carbon black to stabilize quantities of cerium identical to those measured in the CLs after ASTs.

In addition to water content and degradation, potential gradients and proton flux strongly influence cerium migration from the PEM into the CLs. Window cell experiments will be discussed, which determine the relative influence of these mechanisms and their effects in operating cells. Further understanding of these mechanisms will enable cerium stabilization within the active area of the PEM, in order to mitigate performance losses and further enhance cell durability.

Acknowledgements

This research is supported by DOE Fuel Cell Technologies Office, through the Fuel Cell Performance and Durability (FC-PAD) Consortium; Fuel Cells program manager: Dimitrios Papageorgopoulos.

References

[1] D. Banham, S. Ye, S. Knights, S. M. Stewart, M. Wilson and F. Garzon, J. Power Sources, 281, 238-242, 2015.

[2] R. Mukundan, D. Langlois, D. Torraco, R. Lujan, K. Rau, D. Spernjak, A. M. Baker and R. L. Borup, ECS Meeting Abstracts, 37, 1483-1483, 2015.

Figure 1

E-33 Alkaline & DF Cells and Catalysts - Oct 6 2016 4:20PM

2846

, , , , , and

Recent advances in highly conductive and base stable anion exchange membrane chemistries have enabled the widespread fabrication of alkaline membrane fuel cells (AMFC).(1) A significant limitation to AMFC commercialization, however, is the sluggish hydrogen oxidation reaction (HOR) kinetics of current anode catalysts. Much work has been performed to understand the fundamental limitations of HOR in an alkaline environment(2, 3) and to study the HOR in an alkaline membrane electrode assembly (MEA)(4). It is of significant interest to the AMFC community to continue studying the HOR reaction to understand the in-situsources of voltage losses, and to develop a diagnostic tool to test the kinetics of new/alternative catalysts in an alkaline MEA.

In this study, a series of MEAs were fabricated where the membrane (Tokuyama A201) and binder polymer (Tokuyama AS-4) are held constant, but where the catalyst content and composition are varied. A hydrogen pump technique was utilized where an external current source drives HOR on one electrode of the MEA and hydrogen evolution (HER) on the other. It should be noted that there is no traditional reference electrode in this system, as each electrode reaction contributes a significant overpotential. The polarization curves acquired from the hydrogen pump are analyzed for sources of voltage losses including mass transport, ohmic, and kinetic overpotentials. The causes of such losses will be discussed and presented with insight gained from modeling and experimental work. Exchange current densities of HOR and HER for both platinum on carbon and platinum ruthenium on carbon will also be presented, where all data was taken in an in-situalkaline MEA. The potential for hydrogen pump as a diagnostic method for future anode catalyst testing will also be included.

Figure 1. HOR and HER overpotentials for Pt/C electrodes between 0.119 and 0.8 mg/cm2 loading as a function of current density per Pt site, with the corresponding Butler-Volmer model fit

1. J. R. Varcoe, P. Atanassov, D. R. Dekel, A. M. Herring, M. A. Hickner, P. A. Kohl, A. R. Kucernak, W. E. Mustain, K. Nijmeijer, K. Scott, T. Xu and L. Zhuang, Energy & Environmental Science(2014).

2. J. Durst, A. Siebel, C. Simon, F. Hasche, J. Herranz and H. A. Gasteiger, Energy & Environmental Science, 7, 2255 (2014).

3. Y. Wang, G. Wang, G. Li, B. Huang, J. Pan, Q. Liu, J. Han, L. Xiao, J. Lu and L. Zhuang, Energy & Environmental Science, 8, 177 (2015).

4. M. D. Woodroof, J. A. Wittkopf, S. Gu and Y. S. Yan, Electrochemistry Communications, 61, 57 (2015).

Figure 1

2847

Active and durable electrocatalysts for methanol oxidation reaction are of critical importance to the commercial viability of direct methanol fuel cell technology. Unfortunately, current methanol oxidation electrocatalysts fall far short of expectations and suffer from rapid activity degradation. Here we report platinum–nickel hydroxide–graphene ternary hybrids as a possible solution to this long-standing issue. The incorporation of highly defective nickel hydroxide nanostructures is believed to play the decisive role in promoting the dissociative adsorption of water molecules and subsequent oxidative removal of carbonaceous poison on neighbouring platinum sites. As a result, the ternary hybrids exhibit exceptional activity and durability towards efficient methanol oxidation reaction. Under periodic reactivations, the hybrids can endure at least 500,000 s with negligible activity loss, which is, to the best of our knowledge, two to three orders of magnitude longer than all available electrocatalysts.

2848

, and

With the recent developments of high-performance anion-exchange membranes (AEM), there has been an increased interest towards studying various electrochemical reactions such as ethanol electrooxidation, hydrogen oxidation reactions and oxygen reduction reactions in AEM fuel cells (AEMFC). This increased attention has been attributed to the faster electro oxidation kinetics, increased stability and lower poisoning of the catalysts in alkaline media. And since AEMFCs are not limited to platinum - unlike in proton exchange membrane fuel cells (PEMFCs) with highly acidic environments - intensive efforts towards developing alternatives such as palladium (Pd) and Pd-based alloys are currently underway.1, 2

Various studies have shown the enhanced stability and higher activities of Pd-based catalysts compared to Pt in alkaline medium using potentiodynamic and potentiostatic techniques. However, there are only a limited number of studies that have investigated the performance of Pd-based electrocatalysts in AEMFCs. Also, majority of these studies have demonstrated relatively low performances in AEMFCs, mainly due the complications associated with forming the triple-phase boundary (TPB) structure in the membrane layer, where reactions are taking place in the electrolyte, gaseous fuel and electrode interface. The density of the TPB, along with the intrinsic activities of the catalysts can play an important role in determining the overall performance of AEMFCs. Moreover, Pd-based are usually synthesized using surfactants, organic stabilizers or reducing agents, that can get adsorbed onto the surface of Pd and inhibit ionomer-catalyst-fuel interactions in the TPB.

To mitigate the limitations associated with TPB structure and subpar catalyst activities, macroporous three-dimensional Graphene nanosheet (3D-GNS) supports with controlled morphologies and porosities were fabricated by utilizing silica sacrificial templates.3 The sacrificial templates were then etched to leave a network of porous channels within its matrix, and utilized as supports for Pd nanoparticles. The porous structure of these highly graphitized 3D-GNS supports can facilitate mass-transport kinetics by enabling the ionomer and polymer electrolyte getting the reactants to get catalyzed by Pd nanoparticles. Also, to enhance ionomer-nanoparticle interactions, the Pd nanoparticles of an average size of 3-5 nm were synthesized using a previously established surfactant-free soft alcohol reduction method.4

In this study, anion exchange catalyst-coated membranes (CCMs) were prepared using Pd nanoparticles supported on both commercial carbon blacks (Vulcan) and 3D-Graphene supports as both cathode and anode catalyst in H2/O2 fed AEMFCs. The Pd catalyst inks were prepared by mixing the dispersed Pd-catalyst in an optimized solution of isopropyl alcohol and quaternary ammonium-functionalized AS4 ionomer. The inks were then applied onto an anion exchange membrane (A201, Tokuyama) with an active area of 5 cm2. The CCMs was then hydrated in 0.5 M KOH for 24 hours, followed by rinsing in de-ionized water for 24 hours. It was then sandwiched between gas diffusion layers and annealed with gaskets by pressing under 500 psi for 5 minutes. The cell was then assembled and activated by operating at 0.3V at 60ºC for 10 minutes under humidified O2 and H2, followed by the measuring the polarization curves.

Preliminary results show that Pd nanoparticles synthesized using surfactant-free method SARM and deposited on commercial carbon supports have and OCV of 0.95 V and a peak power output of 200 mWcm-2. These results already show the promising potential of using surfactant-free non-alloyed monometallic Pd nanoparticles. Further tests will also be performed to show the effect of porous 3D-Graphene supports towards modifying the TPB interphase and its effects on MEA performance in both Ethanol/O2 and H2/Ofed AEMFCs. The results from this study will not only contribute towards the development of Pt-free oxidative and reductive electrocatalysts, but also lead to further advancements in AEMFC technology that will enable it to operate with other fuels such as methanol and hydrazine.

References

1. M. Alesker, M. Page, M. Shviro, Y. Paska, G. Gershinsky, D. R. Dekel and D. Zitoun, Journal of Power Sources, 2016, 304, 332-339.

2. S. Y. Shen, T. S. Zhao and Q. X. Wu, International Journal of Hydrogen Energy, 2012, 37, 575-582.

3. S. Kabir, A. Serov, K. Artyushkova and P. Atanassov, Electrochimica Acta, 2016, 203, 144-153.

4. A. Serov, N. I. Andersen, S. A. Kabir, A. Roy, T. Asset, M. Chatenet, F. Maillard and P. Atanassov, Journal of The Electrochemical Society, 2015, 162, F1305-F1309.

Figure 1

2849

and

Anion exchange membrane (AEM) stability is a long-standing challenge that has limited the widespread development and adoption of AEM fuel cells. It is essential to understand the mechanism of AEM degradation during fuel cell operation. There are multiple modes of AEM degradation, broadly classified as chemical, mechanical and thermal degradation. Chemical degradation is among the most destructive modes, and can be further sub-divided into nucleophilic degradation (induced by the hydroxide ion), and oxidative degradation (induced by reactive oxygen species). While the former has been extensively studied, there is minimal work on oxidative AEM degradation.

The reactive oxygen chemical species produced during the operation of an AEM fuel cell have hitherto not been detected during operation. Given the high pH, it is postulated that superoxide anion radicals (O2·-), as opposed to hydroxyl radicals, are primarily involved in the degradation progress. The objective of this study was to confirm the O2·- formation during AEM fuel cell operation and to monitor in real-time the rate of O2·- generation in an operating fuel cell using in-situ fluorescence spectroscopy.

1,3-diphenlisobenzofuran (DPBF) was chosen as the fluorescence probe, the sensitivity of which towards O2·- was evaluated by performing ex-situ experiments in a semi-batch reactor. We demonstrate that the fluorescence intensity of this dye selectively decreased upon exposure O2·-. DPBF was then incorporated into an AEM (membrane was solution cast after mixing the dye with the casting solution), which was assembled into a fuel cell. O2·- generation in an operating AEM fuel cell was then monitored via in-situ fluorescence spectroscopy using a bifurcated optical probe, when the cell was operated in H2/O2 mode. To confirm the impact of O2·-  on AEM degradation, independent experiments (without dye) were performed under identical conditions, under both H2/O2 and N2/N2 modes, and the ionic conductivity and ion exchange capacity were monitored to estimate degradation extent. From our in-situ fluorescence studies, we were able to estimate the rate constants and activation energy for oxidative AEM degradation in an operating AEM fuel cell.

2850

, , , , , and

Anion-exchange ionomers are used in wide range of applications such as water treatment (desalination, purification, decontamination), chromatography and fuel cells. They consist of polymer chains functionalized with cationic groups to carry anionic molecules and can therefore be used as membrane and catalyst binder in anion-exchange membrane fuel cells (AEMFCs). In the last years there were many improvements related to alkaline stability and ion conductivity of anion-exchange membranes making these fuel cells more feasible.

Anion-exchange membrane direct methanol fuel cells (AEM-DMFCs) are a subcategory of AEMFCs and convert the chemical energy of methanol into electric current by oxidizing methanol and reducing oxygen on the anode and cathode, respectively. Advantages of the alkaline over the acidic environment are enhanced methanol oxidation kinetics, lower overpotentials for oxygen reduction reaction, usage of cheaper catalysts than platinum, no need of highly alloyed steels due to less corrosive medium and possible usage of alcohol tolerant oxygen reduction reaction catalysts. It was recently demonstrated that AEM-DMFCs free from platinum reach power densities of up to 0.1 W cm-2 [1,2] making them an interesting alternative as power source for mobile and back-up power applications.

In AEM-DMFCs, an ionomeric binder is needed to fix the carbon supported catalyst onto the liquid/gas diffusion layer. The catalyst layer consisting of the ionomer and the catalyst is supposed to build a three-phase boundary (TPB). The TPB facilitates the simultaneous transport of hydroxide ions, electrons and reactants/products. Research groups working on AEM-DMFCs currently rely on highly stable polytetra-fluoroethylene (PTFE) as binder as it is forming a microporously structured catalyst layer. On the downside, PTFE is non-anion conductive polymer. Hence, AEM-DMFCs with PTFE as binder need KOH added to the anodic fuel as electrolytic supplement to reach high current densities. As most anion-exchange membranes are not stable in highly alkaline media at elevated temperatures the addition of KOH to the fuel is considered to be problematic leading to the failure of the cell. Besides this, the addition of KOH is not consumer friendly and could therefore lead to problems in commercialization of the AEM-DMFCs. Thus, it is favorable to eliminate KOH as a fuel supplement. To reach this, ionomeric binders have to be designed for the needs of the AEM-DMFC especially on the anode compartment as the liquid methanolic fuel and the methanol oxidation reaction demand special properties from the ionomeric binder: Anionic conductivity, low swelling in water/methanol, high stability in alkaline environment and formation of a microporous catalyst layer structure.

We herein present the synthesis, characterization and implementation in single cells of a comb-shaped anion-conductive ionomer with poly(2,6-dimethyl-1,4-phenylene oxide) (PPO) as backbone polymer for the use as anodic binder in AEM-DMFCs. To synthesize the ionomer, PPO was brominated to further functionalize it. The comb-shaped ionomer resulted from a subsequent reaction of linking (2-methylimidazole, 2MIm), side chain (octafluoro-1,4-diiodobutane, DIFB) and quaternization (1,2,4,5-tetramethylimidazole, TMIm) agents. Successful synthesis was confirmed by 1H-NMR-spectroscopy. In Figure 1 the alkaline stability of comb-like PPO-2MIm-DIFB-TMIm is compared with a non-comb-shaped ionomer (PPO-TMIm). It is observed that the alkaline stability is greatly enhanced especially at high temperatures when the cationic group is not directly linked to the PPO backbone. Besides this, the comb-shaped ionomer shows several other advantageous properties over PPO-TMIm like lower swelling ratios while having similar water/methanol uptakes, higher thermal stability and higher ionic conductivity. Further investigations were done by implementing the ionomer as anodic binder in membrane electrode assemblies and conducting CO stripping and performance experiments on single cell level. Resulting UI and power curves (Fig. 2) showed a significant performance gain for the single cells with the comb-shaped ionomer used as binder.

Fig. 1 (A) Loss of ionic conductivity of PPO-TMIm (black) and PPO-2MIm-DIFB-TMIm membranes (red) immersed in 3M KOH solution for 10 days at room temperature (straight lines), 60 °C (dashed lines) and 80 °C (dotted lines). (B) UI and power density curves of single cells fed with 4M CH3OH and O2 as anodic and cathodic fuels in dependency of the ionomer content in the anodic catalyst layers. Anodes consisted of various amounts of comb-like ionomer or PTFE and 2 mgPt cm-2 Pt/C. Cathodes consisted of 2 mgPt cm-2 Pt/C and 12.5wt% PTFE binder. Membrane: Tokuyama A201.

References

[1]        T. Jurzinsky, R. Bär, C. Cremers, J. Tübke, P. Elsner, Electrochimica Acta 176 (2015) 1191-1201.

[2]        T. Jurzinsky, P. Kammerer, C. Cremers, K. Pinkwart, J. Tübke, Journal of Power Sources 303 (2016) 182-193.

Figure 1

2851

, , , and

Understanding oxidation kinetics and reaction mechanism at elevated temperatures is paramount to making efficient fuel cells that can run at temperatures higher than 100oC. We describe a method that allows for studies of aqueous electrochemistry at temperatures above the normal boiling point of water, and this system is used to study potential fuels for direct fuel fed fuel cells. Two alcohols with very different partial pressures at temperatures above 100oC are used in this study: methanol and glycerol. A self-pressurized glass autoclave heated in an oil bath is used for this purpose, and a model of the autoclave is shown in Fig. 1a. Slow ramping of the temperature allows for efficient data acquisition by using cyclic voltammetry up to 140oC, exemplified by Fig. 1b, and kinetic parameters such as onset potentials, Tafel slopes and activation energies are found and discussed for the relevant fuels. For methanol oxidation, dissociative water adsorption plays a key limiting role in the total reaction of methanol to CO2 at all temperatures. For glycerol oxidation, some partial oxidation products do not require an oxygen donor in the form of adsorbed water for their formation. This is deemed to play a fundamental role giving a large drop in overpotential at temperatures above 110oC, and making glycerol a potential fuel cell candidate at these temperatures. Therefore, a partial oxidation of glycerol in a fuel cell offers a potential method for a combined fuel cell and value-adding process for abundant glycerol.

Figure 1