University of Birmingham

Introduction to Materials for PEMFC Mardle, Peter; Du, Shangfeng

License: None: All rights reserved

Document Version Peer reviewed version Citation for published version (Harvard): Mardle, P & Du, S 2020, Introduction to Materials for PEMFC Electrodes. in A-G Olabi (ed.), Encyclopedia of Smart Materials. Elsevier.

Link to publication on Research at Birmingham portal

General rights Unless a licence is specified above, all rights (including copyright and moral rights) in this document are retained by the authors and/or the copyright holders. The express permission of the copyright holder must be obtained for any use of this material other than for purposes permitted by law.

•Users may freely distribute the URL that is used to identify this publication. •Users may download and/or print one copy of the publication from the University of Birmingham research portal for the purpose of private study or non-commercial research. •User may use extracts from the document in line with the concept of ‘fair dealing’ under the Copyright, Designs and Patents Act 1988 (?) •Users may not further distribute the material nor use it for the purposes of commercial gain.

Where a licence is displayed above, please note the terms and conditions of the licence govern your use of this document.

When citing, please reference the published version. Take down policy While the University of Birmingham exercises care and attention in making items available there are rare occasions when an item has been uploaded in error or has been deemed to be commercially or otherwise sensitive.

If you believe that this is the case for this document, please contact [email protected] providing details and we will remove access to the work immediately and investigate.

Download date: 27. Sep. 2021 Introduction to Materials for PEMFC Electrodes Peter Mardle and Shangfeng Du* School of Chemical Engineering, University of Birmingham, UK *Corresponding email: [email protected] Keywords: PEMFC, Polymer membrane, , , Electrocatalyst, Catalyst support, Platinum, Nanoparticle, Nanostructure, Oxygen Reduction Reaction, Alloy.

Abstract

• As part of a sustainable energy future, much attention has been drawn to the use of hydrogen fuel as it can be formed renewably through the electrolysis of water. As part of this proposed hydrogen economy, polymer electrolyte membrane fuel cells (PEMFCs) play a key role as clean power generators in vehicles, building and portable devices. However, for extensive commercialization, high performance and durable devices need to be produced. This chapter introduces the material used for electrodes and how advanced development can help to address the challenge faced by PEMFCs.

1. Introduction Serious concerns about the sustainability and the environmental consequence of using fossil fuels as our primary energy source have led to several ingenious technologies, such as energy generation from hydrogen through the use of fuel cells. Such a technology provides a clean and sustainable way for producing electricity on-site such as portable power generators or in the transport sector. By avoiding the thermodynamic limitations of the Carnot cycle, fuel cells also have the additional benefits of providing a high energy efficiency while releasing extremely low levels of pollutants (Ramachandran & Stimming, 2015). Among various types of fuel cells, polymer electrolyte membrane fuel cells (PEMFCs) are favored for automotive and portable applications due to their low operating temperature, flexible scale, and quick start up/shut down. In recent decades, PEMFCs have therefore received the greatest efforts towards their development. However, despite significant progress has been achieved for this technology, PEMFCs still face some big challenges including poor reliability and robustness, and also high cost. In order to meet these challenges and move this clean technology for extensive commercialization, further understanding of the materials and structure of PEMFC components is essential.

1.1. PEMFC History

1

In 1839, William R. Grove constructed what can be considered to be the first hydrogen fuel cell, and published the results as a mere postscript to his investigations of acidic salt solutions (Grove, 1839). The first PEMFC, however, was developed 120 years later by Thomas Grubb for General Electric (GE). After refinements by Leonard Niedrach in the mid of the 1960s, the PEMFC then saw active use by NASA as a power source in Project Gemini. The ‘Grubbs-Niedrach fuel cell’ employed for these missions was used in place of the batteries used in project Mercury as it could be used for longer durations. In over 50 years since the first PEMFC, the basic design and the materials used have barely changed. Initially phenolic membranes were used, but a series of developments led to a breakthrough in 1966 with the development of the NFI Nafion® membrane by DuPont (Zaidi & Matsuura, 2010). The use of the Nafion® membrane effectively tripled the durability of the electrolyte, and as such is still the conventionally used membrane today. It also so happens that the platinum (Pt) used in the Grove cell is inherently the most active single metal catalyst for use in PEMFCs (Nørskov et al., 2004). Interest dropped in PEMFCs until the 1990s where advances in nanotechnology and the understanding of electrode processes led to the common practice of dispersing Pt nanoparticles on high surface area carbon to act as the catalyst. Since then, heavy research and development has been carried out on PEMFCs, not least on the electrode materials leading to the release of commercial PEMFC powered vehicles in the form of the Hyundai iX35 (2013) and Toyota Mirai (2014) fuel cell electric vehicles (FCEVs) and more recently; the Honda Clarity (2017) and Hyundai Nexo (2018). However, for the mass adoption of such clean technology, the development of cheaper and more powerful PEMFC systems is required. In recent years, PEMFC development has been guided by the US Department of Energy (DoE) ultimate target of developing the PEMFC system at a 65% electricity efficiency with a durability of 8,000 hours, all at a cost of $30 kW-1 (Papageorgopoulos, 2019).

1.2. PEMFC Principles The fuel cell is a device which can directly convert the chemical energy in fuels to electricity through electrochemical reactions. The scheme of a PEMFC is shown in Figure 1. At the , molecular hydrogen is split into two protons and two electrons by the hydrogen oxidation reaction (HOR), as shown by Equation 1. The protons then transfer though the polymer electrolyte membrane (PEM) to the . As the PEM does not conduct electrons, they are forced to transfer to the cathode through the external load, allowing an electrical current to be drawn from the fuel cell. At the cathode, molecular oxygen then reacts with the transferred protons and electrons to form water (Equation 2). This process is called the oxygen reduction reaction (ORR). The overall chemical reaction (Equation 3) is simply therefore the recombination of oxygen and hydrogen to produce water.

Figure 1: Scheme illustration of a typical PEMFC with electrodes and overall reactions.

2

Under standard conditions the Gibbs free energy of the forwards reaction for the production of liquid water (Equation 3) is –237 kJ mol-1 (Dicks & Larminie, 2003), corresponding to a potential difference of 1.23 V between both sides of the fuel cell. The PEMFC can therefore be thought of as a method of splitting this simple redox reaction into its two half-cell processes. In doing so the change in Gibbs free energy is harnessed in the form of electrical energy. Figure 2 depicts the main components of a PEMFC, where in the center resides the membrane electrode assembly (MEA). In a conventional PEMFC the MEA consists of the cathode (GDE) and the anode GDE hot pressed to the PEM. The GDE comprises of a polytetrafluoroethylene (PTFE) treated carbon paper/cloth-based gas diffusion layer (GDL) with the catalyst layer (CL) deposited on top. Gases are distributed to the GDL through the flow field plate (FFP). In this chapter, we only discuss the electrodes, while leaving the PEM, GDLs and FFPs to next chapter.

Figure 2: The components of a single PEMFC. The power performance of PEMFCs is usually presented by a polarization curve, or voltage-current density curve, as shown in Figure 3a. The polarization curve is usually determined by holding the cell potential at a series of set values for 15-30 minutes prior to recording the current (Gasteiger, Kocha, Sompalli, & Wagner, 2005; Multi-Year Research, Development, and Demonstration Plan: 3.4 Fuel Cells, 2017). It can also be replotted as power-current density curve, representing the power output from the PEMFC to the external load. The durability of the PEMFC is usually tested either by monitoring the power performance over a long-term operation at a specified condition, e.g. holding the cell potential at 0.6 V for a few thousand hours, or by running dynamic load cycles which can accelerate the fuel cell degradation rate, thus reducing the time required for the test (Tsotridis, Pilenga, Marco, & Malkow, 2015). A typical dynamic load cycle for fuel cells in automotive applications is shown in Figure 3b.

Figure 3: (a) A typical polarization curve of a single PEMFC, and (b) dynamic load cycle for testing fuel cells in automotive applications. Reproduced (adapted) with permission from (Tsotridis et al., 2015).

1.3. Electrode requirements and fabrication From Figure 1, it can be seen that in order to get a larger power output from a PEMFC, highly efficient fuel cell reactions (Equations 1-3) are necessary. To achieve this, both reactants and products need to be quickly supplied to and removed from the reaction sites, respectively. Therefore, the electrodes need to be highly electrically conductive to allow for rapid transfer of electrons between the reaction sites and external circuit. While the PEM has a high proton conductivity requirement to facilitate the transport

3 of protons from the reaction sites of the anode to the cathode, proton conducting ionomer (typically a dispersion of the electrolyte material) is required throughout the CL to provide the final proton transfer pathways to the catalytic sites and increase overall catalyst utilization (Figure 4). Additionally, the electrodes should possess a large degree of porosity for both hydrogen and oxygen to be delivered to the reaction sites in time. The electrode structure and composition must also allow for the efficient removal of the produced water from the reaction sites. Without efficient water transport, the water will cover the reaction sites or even block the pores, resulting in water flooding, hindering the transfer of gaseous fuels to the reaction sites.

Figure 4: An illustration of mass transport in the CL. Reproduced (adapted) with permission from (Inoue & Kawase, 2016). The PEMFC electrodes can be fabricated by either applying catalysts onto the GDL surface to form GDEs, or by coating onto the membrane surface to get catalyst coated membranes (CCMs). Focusing initially on the GDE fabrication, a simple application technique is the colloidal method (Litster & McLean, 2004). In this method, an ink is generally formed by dispersing the catalysts in a solvent such as iso-propyl alcohol (IPA) and a defined amount of proton conducting ionomer. To assist with the dispersion, sonication by means of a sonic bath or probe can be implemented. The ink can then be applied to the GDL by means of brushing, spraying (i.e. sputter coating) or screen-printing to get a GDE. The GDE is subsequently hot pressed with the PEM to fabricate the complete MEA. Although the construction of a GDE is still a popular MEA fabrication method today, a relatively poor contact is commonly obtained between the GDE and the membrane. This is a result of the fact that the relative soft GDE structure can be deformed in the hot press process meaning that the force applied cannot be fully transferred to the surface. An improved approach is to use CCM technology by coating catalysts directly onto the membrane (W. Wang, Chen, Li, & Wang, 2014), for which a better power performance is usually achieved through an enhanced catalyst to membrane interface. Alternatively, a CCM can be formed by the indirect method of decal transfer. This is where the catalyst is coated onto an inert substrate, i.e. PTFE film, which is in turn hot pressed with the membrane. The inert substrate can then be peeled off, leaving behind the CCM (Kim et al., 2010; Mehmood & Ha, 2012). In the fabricated electrodes, the network formed by the catalysts and electrolyte ionomer provides the transfer path for electrons and protons, respectively. The pores formed by both materials then provide the required channels for efficient gas transport. Figure 5 shows a typical well-fabricated catalyst layer represented by the uniform ionomer distribution through fluorine element mapping technique (More, Cullen, Sneed, & Reeves, 2016).
Figure 5: A 3D ‘F map’ obtained by STEM showing ionomer distribution in a catalyst layer. Reproduced (adapted) with permission from (More et al., 2016).

4

2. Catalyst Materials When comparing the kinetic rates of HOR and ORR in the PEMFC, ORR at the cathode is comparatively a much more sluggish reaction, which is about 6 orders slower than HOR (Shao, Chang, Dodelet, & Chenitz, 2016). Therefore, in PEMFC development, most efforts have been attracted to developing advanced cathodic catalysts towards a fast ORR. The amount of effective reaction sites on the catalyst surface, with respect to the total amount of catalyst material, is defined by the electrochemical effective surface area (ECSA) given in m2/mg. For Pt catalysts, this can be determined electrochemically by measuring the number of hydrogen molecular adsorbed on the effective reaction sites on the catalyst surface. A large ECSA indicates a high degree of catalyst utilization whereby there are more surface sites available to catalyze the reaction. The catalytic activity of the catalysts is defined as the current obtained at a specific voltage, i.e. 0.9 V divided by the mass of catalyst material present. This is often referred to as the mass activity, the most commonly used figure to compare different catalysts. The ratio of the mass activity to the ECSA then provides the specific area activity, indicating how active the catalyst surface is towards a fuel cell reaction. Electrochemical testing of the fuel cell catalysts can be conducted both ex-situ half- cell measurement (outside of the fuel cell), or in-situ (testing in the MEA). The most popular ex-situ test is that of the rotating disk electrode (RDE) in aqueous (e.g. H2SO4 or HClO4) where a three electrode setup allows the intrinsic catalytic activity of the catalysts to be quickly measured (Garsany, Baturina, Swider-Lyons, & Kocha, 2010; Gasteiger et al., 2005). Although it is a useful screening technique, the recorded activities often do not correlate well with the performance in fuel cells for novel non-Pt/C catalysts and is a key challenge in effective use of novel catalyst materials in operating PEMFCs (Stephens, Rossmeisl, & Chorkendorff, 2016). Asides from measuring catalytic activity, accelerated degradation testing (ADT) is frequently used to determine the stability of the catalysts. ADT is usually conducted by cycling the fuel cell operation for a fixed time, e.g. 30,000 cycles between low current densities (low load) and high current densities in set conditions (Multi-Year Research, Development, and Demonstration Plan: 3.4 Fuel Cells, 2017). Such cycling accelerates the dissolution and aggregation of the catalyst particles and so the susceptibility of the catalyst to processes such as Ostwald ripening can be examined. The dynamic load cycle test is one type of ADT and better simulates the voltage changes in an operating PEMFC in practical driving of vehicles.

2.1. Platinum Nanoparticles For both HOR and ORR, Pt is recognized as the most efficient catalyst where Pt nanoparticles supported on carbon nanospheres (Pt/C) are still the commonly used catalysts today. As such, little has changed with regards to the materials used in PEMFC electrodes since 1839 and the reasons were cleared by the work conducted by Nørskov et al. (Nørskov et al., 2004) and Trasatti et al (Trasatti, 1972). Utilizing density functional theory (DFT) it was found that the ORR activity is directly related to the oxygen and hydroxyl binding energies on the catalyst surface. If the binding energy is too strong, then the rate of ORR would be limited by desorption of the

5 reaction products from the surface. If the binding energy is too weak, then the adsorption of oxygen becomes less favorable and so the reaction becomes limited by the proton and electron transfer to dissociate molecular oxygen. Because of this trade- off, the volcano plot in Figure 6 forms (Nørskov et al., 2004). Closest to the peak is Pt, demonstrating that it is inherently the most effective single metal catalyst towards ORR. A similar plot was constructed by Wendt et al (Wendt, Spinacé, Oliveira Neto, & Linardi, 2005) who made the analogous argument for HOR. With lower bond strengths there would be a reduced activation of the hydrogen-hydrogen bond. Conversely, if the metal-hydrogen bond strength is too high, then the rate of the heterogeneous reaction would be limiting.

Figure 6: Volcano plot for the oxygen and hydroxyl binding energies on different metal surface. Reprinted (adapted) from (Nørskov et al., 2004) with permission from American Chemical Society. Copyright (2004). For catalyst applications, small nanoparticles are usually preferred due to their high specific surface area, and in the same access ratio a large ECSA and mass activity are expected. The fuel cell reactions are structure-sensitive reactions which are more active at specific crystal facets, e.g. (111) has a much higher specific area activity than (100) or amorphous surfaces towards the ORR, but less active at the edge and corner sites due to the extremely high oxygen binding energy. Larger particles have a higher proportion of crystal facets in comparison to edge and corner sites, leading to higher specific area activities. However, the low ECSA results in a smaller mass activity that can be obtained with smaller particles. Based on this particle size effect, in 1990 Kinoshita (Kinoshita, 1990) calculated that the peak mass activity is obtained for particles of an average radius of ca. 3.5 nm. Therefore, the small nanoparticle with a controlled morphology possessing highly active crystal facets is preferable for fuel cell catalyst application. Besides activity, stability must be also considered in developing catalysts. Ostwald ripening is the main cause of performance loss of the catalysts in PEMFC electrodes. Due to a difference in surface free energy, there is a thermodynamic tendency for smaller particles to agglomerate to form larger ones (Parthasarathy & Virkar, 2013). The larger particle size results in a smaller ECSA and thus a drop of mass activity and catalyst performance. This effect can also be described in terms of surface free energy where a larger value results in a stronger driving force for agglomeration. In order to obtain consistent small particles, the favored chemical synthesis method for Pt nanoparticles is to deposit them onto carbon nanosphere supports. In one case, this can efficiently reduce the Ostwald ripening of Pt nanoparticles in operation due to the hindering effect of the support. On the other hand, the excellent electrical conductivity of the carbon support can reduce the impact of the conductivity loss of nanoparticles that results from the size effect, thereby maintaining effective electron transfer. Conventionally, the deposition of Pt on carbon can be done by a one-step reduction of chloroplatinic acid or potassium hexachloroplatinate (IV) using sodium borohydride (E. Antolini, Salgado, dos Santos, & Gonzalez, 2005; Brown & Brown, 1962; Islam, Bhuiya, & Islam, 2014; K. W. Park et al., 2002; Salgado, Antolini, & Gonzalez, 2004). can also be used as the reducing agent in place of sodium borohydride (Jiang & Kucernak, 2003). A second common wet chemical route is the

6

‘polyol process.’ This is a process where a polyol such as ethylene-glycol is used as both the solvent and reducing agent for the metallic salt (Fievet, Lagier, & Figlarz, 1989). Microwave assistance can be employed to increase the reaction rate to mere minutes while producing consistently small particle sizes (Figure 7) (Sharma et al. 2019; H. Li, Zhang, Yan, Lin, & Ren, 2013). Modifications of this process such as the use of capping agents, e.g. oleic acid have also been used to achieve size control of the precipitated product (Y. Chen, Liang, Yang, Liu, & Chen, 2011; Santiago, Varanda, & Villullas, 2007).

Figure 7: TEM images of Pt/C catalysts. Reprinted (adapted) from (Sharma et al., 2019) with permission from American Chemical Society. Copyright (2019). To give an example of size and morphology control, Wang et al (C. Wang, Daimon, Onodera, Koda, & Sun, 2008) synthesized unsupported platinum nanoparticles via the reduction of platinum (II) acetylacetonate by a mixture of octadecene, oleic acid and oleylamine. By then injecting the reaction mixture with iron pentacarbonyl at various temperatures, the size and morphology of the nanoparticles could be better controlled. For example, at an elevated temperature, 7 nm cubic nanoparticles could be formed instead of 3 nm polyhedral nanoparticles at a low temperature. Such work indicates that if the ideal catalyst structure can be deduced, it can almost certainly be made.

2.2. Platinum alloys and hybrids Alloying platinum with transition metals has been proven to vastly improve the sluggish ORR kinetics of the cathode catalyst because of geometric and electronic effects. By extending the aforementioned work, Nørskov et al (Nørskov et al., 2004), Stamenkovic et al , Greeley et el (Greeley et al., 2009) and Chorkendorff et al (Stephens et al., 2012) extended the volcano plot to include platinum-transition metal alloys (Figure 8). By reducing the d-band center for the platinum-oxygen bond, the bond strength is weakened and the activities for the ORR therefore increased. The ORR activity of these alloys is also enhanced through lattice strain effects. As a result of the strained lattice structure, surface Pt-Pt bonds contract, weakening the Pt-OH bond strength through an increased downwards d-band shift (Mukerjee, Srinivasan, Soriaga, & Mcbreen, 1995; Xiaoming Wang et al., 2013). By simply adding a second precursor alongside the chloroplatinic acid in the salt reduction process, other metal can be easily incorporated into the reaction mixture to form alloyed nanoparticles, e.g. Pt-Cr/C (E. Antolini et al., 2006), Pt-Ni/C (E. Antolini et al., 2005; K. W. Park et al., 2002), Pt-Co/C (Salgado et al., 2004), Pt-Cu/C (Zhou & Zhang, 2015) alloyed nanoparticles have all been successfully prepared for fuel cell catalyst application.

Figure 8: Volcano plot for different catalysts with Pt-overlayers: experimental ORR activity enhancement as a function of hydroxyl binding energy, ΔGHO*, both relative to pure Pt. All data are at U = 0.9 V, with respect to a reversible hydrogen electrode (RHE). Reprinted (adapted) from (Stephens et al., 2012) with permission from the Royal Society of Chemistry. Copyright (2012).

7

Although vastly improved ORR activities can be obtained by this alloying approach, popular alloy materials such as Co, Fe and Ni are prone to leaching out of the catalyst nanoparticles in the highly acidic environment in PEMFC operation. Such leaching not only leads to a drop of the catalytic activity of the catalysts, but also results in the degradation of both ionomer in CL and membrane due to the contamination of the leaching metal ions to the ionomer molecular chain (Yu, Pemberton, & Plasse, 2005). As a result of these durability concerns, core-shell structures have attracted much interest in recent years as the shell of Pt can provide a degree of protection for the transition metals in the core. The vast majority of alloy structures investigated have been those with a Pt3M lattice structure due to superior performances over its other stoichiometric counterparts (Fernandes, Paganin, & Ticianelli, 2010; Ma & Balbuena, 2008; Mukerjee et al., 1995). It is also well known that for the majority of alloys, Pt readily segregates to the surface (Ma & Balbuena, 2008) to form a Pt-rich shell layer after annealing. The second layer for Pt3Fe, Pt3Co and Pt3Ni however contains the alloyed M element. In theory then the electrochemically active sites consist of surface Pt, with the ORR activity enhancement provided by the alloyed metal directly underneath the surface. In addition to the simple annealing, a prior de-alloying process is usually employed to assist the formation of the Pt-rich skin (Snyder, McCue, Livi, & Erlebacher, 2012; Strasser et al., 2010). The favored method is of electrochemical de-alloying where potential cycling of the catalyst in acidic electrolyte dissolves the exposed non-noble metal particles. De-alloying is highly effective for smaller particles with a large Pt:M ratio, but for larger particles (> 15 nm diameter) the de-alloying process results in a nanoporous structure (Snyder, Livi, & Erlebacher, 2013; Snyder et al., 2012) rather than prompting the surface re-arrangement to core-shell nanoparticles (Figure 9). The porosity of such structures can also be advantageous for heterogeneous as higher attempt frequencies of the reactants can be achieved, although the leaching is still a problem.

De-alloying processes have also been used to form intricate Pt3M frameworks. By the subsequent chemical de-alloying of PtNi3 polyhedra, Pt3Ni nanoframes were developed by Chen et al (C. Chen et al., 2014). Thermal annealing of the obtained nanoframes formed a highly active Pt skin on the surface of the frame. Remarkably this catalyst exhibited a 22-fold increase in specific activity with respect to commercial Pt/C. This enhancement was in part attributed to the electronic and geometric effects of the alloy, but also due to the accessibility provided by the open structure for the efficient transport of reactants and products to/from the catalytic sites.

Figure 9 (a): TEM and (b): High resolution TEM images of nano-porous NiPt nanoparticles. Reprinted (adapted) from (Snyder et al., 2012) with permission from American Chemical Society. Copyright (2012). Core-shell structures can also be prepared from a direct approach. As an example, Chen et al (Y. Chen et al., 2011) prepared core-shell Ni-Pt/C catalysts by firstly using a modified polyol process to synthesize Ni/C particles. Only after the formation of the core nanoparticles was chloroplatinic acid slowly added with more polyol. By physically stagnating the reduction of the metal salts, a core-shell structure was obtained. In this investigation it was found that a monolayer of platinum provides the

8 best catalytic performance and durability for the catalyst. With the sequential reduction method, the ratio of Pt:M is commonly controlled simply by the quantity of the precursors used in the synthesis (Y. Chen et al., 2011; G. Wang, Wu, Wexler, Liu, & Savadogo, 2010). Furthermore, individual monolayers of platinum can also be added by using a sacrificial metal such as Cu. Zhang et al (J. Zhang et al., 2005) synthesized a variety of alloy nanoparticles by reducing an “almost dry slurry of carbon and metal salts” using hydrogen gas at high temperatures. The temperatures involved help to induce surface segregation of Pt to the surface of the alloy particle. A copper monolayer was then adsorbed on the surface of the particle, to be replaced by platinum using galvanic displacement (J. Zhang et al., 2004).

2.3. 1D Platinum Nanostructures Following on from the nanoframe structures and with mass transport limitations in mind, 1D structures have gained significant interest in recent years. 1D Pt nanostructures can provide durability enhancements for PEMFC catalysts as due to their inherent structures, the effects of Ostwald ripening can be significantly reduced (Sun, Jaouen, & Dodelet, 2008; Sun et al., 2011). In addition, 1D nano-structures usually possess a much higher electrical conductivity which does not suffer from the size effect as much as their 0D nanoparticle analogues. Both advantages mean that no carbon support is required as that in Pt/C. Thus, the carbon corrosion that has been considered as one of the largest challenges in PEMFC development can be fully avoided. The preparation methods of 1D nanostructures for PEMFC applications was recently reviewed by Du (Du, 2012). One of the simple ways is by using template methods. For example, silica, polymer and anodic aluminum oxide (AAO) templates can be used to form highly regular arrays of nanowires (S. M. Choi, Kim, Jung, Yoon, & Kim, 2008; W. C. Choi & Woo, 2003; Zhao, Xu, Guo, Li, & Li, 2006). By a further kinetic control of the chemical reduction of the metal precursor salts, single crystal nanowires can be achieved with preferred exposure of highly active crystal facets so that catalytic activity can be maximized. For the polyol process this can be achieved by adding a trace amount of Fe2+ or Fe3+ ions (J. Chen, Herricks, Geissler, & Xia, 2004; Xia et al., 2003) to reduce the saturation of Pt atoms. Alternatively, in the wet chemical reduction method in aqueous solution using formic acid, temperature control can be used to control the growth of nanowires (Du, 2010; Du, Lin, et al., 2014; Lu, Du, & Steinberger-Wilckens, 2015; Sun et al., 2011) (Figure 10). At room temperature, crystal growth favors the lower energy crystallographic planes and so single crystal nanowires can form. However, even though the 1D nanostructures provide better stability and very high specific area activities, such performance enhancements must overcome the reduced ECSA caused by the large bulky volume, i.e. to synthesize ultrathin catalyst nanowires.

Figure 10: SEM image of Pt nanowires grown directly on the carbon paper gas diffusion surface. Reprinted (adapted) from (Du, 2010) with permission from Elsevier. Copyright (2010).

9

A class of ultrathin Pt and Pt alloys nanowires have therefore been developed in -1 recent years resulting in the highest recorded ORR mass activity of 13.6 A mgPt (31- -1 fold the DoE 2020 target of 0.44 A mgPt ). A variety of Pt alloy and jagged Pt nanowires have been synthesized by refining oleylamine based methodologies whereby highly strained crystalline surfaces are complemented by extremely high ECSAs (Bu et al., 2015; M. Li et al., 2016). Improving scalability and more in-situ tests are required for these materials to find real use in PEMFC applications.

2.4. Non-platinum catalysts The most efficient way to reduce the amount of expensive Pt is to simply use none of it. Pt group metal (PGM) free catalysts have shown much promise in recent years. Carbon based materials that are commonly used as catalyst support materials (see Section 3) have demonstrated catalytic activity towards the ORR. Wang et al (Xiqing Wang et al., 2010) showed that nitrogen doped mesoporous carbon can be used as ORR catalysts where a computational study by Ikeda et al (Ikeda et al., 2008) suggested that a graphitized nitrogen atom could promote catalytic activity on a nearby carbon site. Hetero-atom dopants have also shown to greatly increase the activity of graphene (Molina-garcía & Rees, 2018) and carbon nanotube (CNT) catalysts (Shui, Wang, Du, & Dai, 2015) toward the ORR, however, performance is not comparable to Pt in acidic media. Metal oxides, nitrides, carbide and chalcogenides are just a few of the other materials that have been investigated as PGM-free catalysts (Shao et al., 2016; Wang et al., 2019). However, a most promising class of materials are M-N-C based where the metal is typically a transition metal such as Fe. Development of these materials has intensified where Ballard in conjunction with Nisshinbo have developed an MEA which has a peak power density of 570 mW cm-2 under practical (air) operating conditions (Banham et al., 2018). Current research challenges are in overcoming the poor stability of these catalysts as well as circumventing high mass transport limitations imposed by the relatively thick CL and high catalyst loadings required (Banham, Choi, Kishimoto, & Ye, 2019).

3. Catalyst Support Materials To achieve a uniform size distribution in the synthesis and reduce Oswald ripening in PEMFC operation, catalyst nanoparticles are usually supported on high surface area materials, in particular carbon black. The excellent electrical conductivity of the support materials can also facilitate the electron transfer through the active sites of the poorly conductive catalyst nanoparticles.

3.1. Carbonaceous catalyst supports To meet the requirements of the high electrical conductivity, large specific surface area (SSA) and high resistance to corrosion, carbon blacks are commonly used at the minimal cost. Carbon blacks are composed primarily of graphite agglomerates/nanospheres with size of up to 50 nm (Sharma & Pollet, 2012). The

10 agglomerates of carbon black provide a porous network, which allows the maximization of surface area without sacrificing electrical conductivity. Perhaps most importantly is the fact that carbon blacks can be formed easily by the incomplete thermal decomposition of hydrocarbons. The catalyst is then deposited onto the support by means of sputter deposition (Zacharia, Rather, Hwang, & Nahm, 2007), electrochemical deposition (Day, Unwin, & Macpherson, 2007) or simple impregnation processes (Matsumoto et al., 2004). Despite these benefits, the corrosion of carbon black in the acidic PEMFC, in particular the high potential during start-up is still a big challenge prompting the investigation of other carbonaceous support materials. More stable carbonaceous materials including CNTs and graphene nanosheets have been studied for this application as both of them possess a larger SSA than carbon black (Sharma & Pollet, 2012). Also with the development of materials synthesis technologies, even though the cost of CNTs and graphene is still high (Dai, 2002), when compared with the expensive Pt catalyst this is not so much of an obstacle for PEMFC application. A more significant drawback of these supports is the difficulty for the direct deposition of Pt nanoparticles on the pristine inert surface, and prior functionalization is usually required (Wildgoose, Banks, & Compton, 2006). The biggest challenge with both materials is caused by their anisotropic particle morphology compared to the carbon black, i.e. 1D CNTs and 2D graphene nano- sheets. The PEMFC catalyst layer made from CNT-based catalysts shows a very loose structure and it is not easy for the electrolyte ionomer coated on surface to form a continuous network through the whole catalyst layer to membrane, leading to low robustness and performance in comparison to other supports (Du, Lu, Malladi, Xu, & Steinberger-Wilckens, 2014; Mardle et al., 2020). Graphene nano-sheets also tend to stack on top each other due to their 2D morphology forming large aggregates (Si & Samulski, 2008). This dramatically reduces the catalyst utilization where reaction sites become embedded between the stacked nanosheets, consequently leading to a poor electrode performance (Ermete Antolini, 2012; Mardle, Fernihough, Du, 2018). Fully erect graphene nanosheets for the whole catalyst layer or a scaffolding architecture (Du, Lu, & Steinberger-Wilckens, 2014; S. Park et al., 2011; Si & Samulski, 2008) (Du, Lu and Steinberger-Wilckens, 2014; Park et al, 2011; Si and Samulski, 2008) can be potential solutions (Figure 11(a)-(b), but their real performance in PEMFCs is still unclear.

Figure 11 (a): Schematic of the use of spacers between graphene sheets. Reproduced (adapted) from (S. Park et al., 2011) with permission from Elsevier. Copyright (2011) (b): Reduced graphene oxide nanosheets with erect PtPd nanowires. Reproduced (adapted) from (Du, Lu, & Steinberger-Wilckens, 2014) with permission from Elsevier. Copyright (2014).

11

3.2. Non-Carbonaceous Catalyst Supports The very nature of carbonaceous materials means it’s impossible to completely mitigate the issue of carbon corrosion in PEMFC operation. Metal oxides have thus been studied for this application because they can avoid this corrosion problem, albeit at the price of reduced electrical conductivities (E. Antolini & Gonzalez, 2009). Therefore, a large degree of effort has focused on the doping of metal oxides in order to improve their electrical conductivities. For example, the conductivity of titanium dioxide is known to increase by introducing dopants, e.g. niobium (K.-W. Park & Seol, 2007) or by using sub-stoichiometric technology. In fact it has been reported that Ti4O7 exhibits a higher electrical conductivity compared to carbon (Ioroi, Siroma, Fujiwara, Yamazaki, & Yasuda, 2005). However, this did not entirely translate into an improvement in catalytic activity due to the comparatively low surface area. Other categories of non-carbonaceous supports include tin dioxide (Okanishi, Matsui, Takeguchi, Kikuchi, & Eguchi, 2006) which can also promote the electro-oxidation of poisonous carbon monoxide on Pt catalyst, and tungsten trioxide (E. Antolini & Gonzalez, 2009) which also adds a benefit in that the support itself can demonstrate some degree of proton transfer (Nakajima & Honma, 2002). Although a wide variety of catalyst support materials have been investigated, significant drawbacks are present with each of them. Although imperfect, the simplicity of carbon black has made this catalyst support very difficult to be replaced on an industrial scale.

4. Proton Conducting Ionomers in Catalyst Layers One of the main goals of PEMFC development is to reduce the PGM content and so -2 lower the overall cost. The biggest reduction in loading from around 4 mgPt cm to -2 0.4 mgPt cm was made in the 1990’s where alongside use of the carbon support, the incorporation of proton conducting ionomer into the catalyst layer much improved catalyst utilization (Wilson & Gottesfeld, 1992). Despite huge advancements in catalyst material in recent three decades, the common PGM loadings in the catalyst layer have not much decreased since then. Recent studies have shown that one major factor is that at low Pt loadings, the diffusion resistance of oxygen through the ionomer thin film coated on the catalyst surface becomes a significant contributor to the overall mass transport resistance (Jinnouchi, Kudo, Kitano, & Morimoto, 2016). Paradoxically to the PEM, high gas permeability is therefore required of the ionomer material. For example, highly oxygen soluble ionic liquids have been shown to improve the ORR activity of Pt based catalysts Zhang, Munoz, & Etzold, 2015, Wang et al., 2019, Zhang et al., 2019). Due to environmental concerns with the production of Nafion®, efforts are made to replace current PEM materials with non-fluorinated alternatives, typically consisting of a poly-aromatic backbone. These alternative ionomer materials exhibit reduced gas permeability, a desirable feature for PEM materials in order to reduce crossover and hence degradation, but undesirable for an ionomer. Differences in ionic exchange capacity (IEC), water uptake and catalyst utilization are also found with these materials and are to be much considered when incorporating in a PEMFC (Peron, Shi, & Holdcroft, 2011). This is also the case with short side chain ionomer materials such

12 as AquivionTM (Hyflon® in Figure 12) which through the removal of the ether group demonstrate increased chemical stability (J. Li, Pan, & Tang, 2014).

Figure 12: PFSA ionomer structures. Reproduced (adapted) with permission from (J. Li et al., 2014). Anion adsorption on Pt surfaces also effects the rate of ORR and is a consideration for ionomer materials. The groups in Nafion® are crucial for proton conductivity, but they bind to Pt surfaces and produce (bi)sulfate which also binds strongly, inhibiting ORR kinetics (Liu & Huang, 2018; Shinozaki, Morimoto, Pivovar, & Kocha, 2016). Phosphate similarly binds to Pt, inhibiting ORR in high temperature PEMFCs (He et al., 2010). One technique to reduce the effects of adsorption is through using carbon supports with specific pore size that do not allow ionomer penetration, but are still proton accessible (Yarlagadda et al., 2018). This additionally shows that when considering electrode materials, the synergy between all components needs to be well considered.

5. Extended surface area catalysts Extended surface area (ESA) catalysts have gained a lot of interest in recent years. They comprise large area surfaces extended in two dimensions such as thin films, possessing larger radii of curvature and having greater resistance to surface area loss (Debe, 2012). The specific area activities towards ORR are usually higher than Pt/C, owing to electronic structure properties versus nanoparticles. In ESA catalysts, thin film catalyst layer technology on high-aspect-ratio nanoparticles shows great potential for PEMFC applications. The most famous ESA catalyst is nanostructured thin-film (NSTF) catalysts developed by 3M. This advanced technique was deliberate from 1995 and has been reviewed recently (Debe, 2013). The thin catalyst layer is firstly introduced by using a monolayer array of perylene whiskers (1 µm tall, 30 nm×55 nm in cross-section) with surface coated 20 nm polycrystalline PtCoMn film. This extremely small thickness of the catalyst layer enables a much higher catalyst utilization ratio thus significantly enhancing the power performance. However, the approach is intrinsically limited by challenges facing with water management issues due to the hydrophilic Pt-transitional 2 metal alloy catalyst surface and the very low ECSA (only 10-15 m /gPt). Recently, based on this concept, Du’s group developed an approach by directly growing ultrathin Pt nanowire arrays on GDL surface in aqueous solution using a very simple method (Du, 2010). Pt nanowires have a diameter about 3-4 nm a length between 10- 2 200 nm. This Pt nanowire electrode shows an enhanced ECSA of 34.4 m /gPt in MEA, together with the advantage of high catalyst utilization ratio of the thin film catalyst nanostructure, a better power performance is achieved in PEFCs than the conventional one from Pt/C nanoparticles, especially at a high catalyst loading (Lu et al., 2015). This work was reviewed in detail by the group very recently (Lu, Du, & Steinberger-Wilckens, 2016). However, there is a challenge with this approach to directly grow Pt alloy nanowires to further improve the electrode performance.

13

6. Conclusions and Perspectives Even though the core materials in PEMFC technology haven’t changed much over two centuries, vast improvements in electrode performance are clearly obtainable by creating novel nanoscale materials. In this chapter it has been shown that for the catalyst in particular, remarkable size, morphology, structure and compositional control of precious group and transition metals can be achieved. Pt-based transitional metal alloy nanoparticles deposited on carbon are still the most preferred commercialized catalysts in the conventional PEMFC electrodes, and electrodes from extended surface area catalysts also show promising potential, including both NSTF and Pt-nanowire array electrodes. These progresses have led to much improved understanding of the first principles underlying PEMFC technology. With the expansion of such knowledge, focused design can begin to supersede the development of new preparation methods. It is of importance that the research focus shifts to considering electrodes as a single entity rather than just its materials. The interaction and behavior of catalysts, support and electrolyte ionomer with the whole electrode, their impact on the catalytic performance, water and gas distribution, electrical and proton conductivity, and their contact with polymer electrolyte membrane must all be cleared. It is only through considering all of the variables that we will successfully pass the demonstration stage today and the extensive commercialization will become possible, allowing the potential role of PEMFCs in a sustainable energy future to be fully realized.

7. References

Antolini, E., & Gonzalez, E. R. (2009). Ceramic materials as supports for low- temperature fuel cell catalysts. Solid State Ionics, 180(9–10), 746–763. https://doi.org/10.1016/j.ssi.2009.03.007 Antolini, E., Salgado, J. R. C., dos Santos, a. M., & Gonzalez, E. R. (2005). Carbon- Supported Pt-Ni Alloys Prepared by the Borohydride Method as Electrocatalysts for DMFCs. Electrochemical and Solid-State Letters, 8(4), A226–A230. https://doi.org/10.1149/1.1870632 Antolini, E., Salgado, J. R. C., Santos, L. G. R. A., Garcia, G., Ticianelli, E. A., Pastor, E., & Gonzalez, E. R. (2006). Carbon supported Pt-Cr alloys as oxygen-reduction catalysts for direct methanol fuel cells. Journal of Applied Electrochemistry, 36(3), 355–362. https://doi.org/10.1007/s10800-005-9072-0 Antolini, Ermete. (2012). Applied Catalysis B : Environmental Graphene as a new carbon support for low-temperature fuel cell catalysts. “Applied Catalysis B, Environmental,” 123–124, 52–68. https://doi.org/10.1016/j.apcatb.2012.04.022 Banham, D., Choi, J. Y., Kishimoto, T., & Ye, S. (2019). Integrating PGM-Free Catalysts into Catalyst Layers and Proton Exchange Membrane Fuel Cell Devices. Advanced Materials, 31(31), 1–6. https://doi.org/10.1002/adma.201804846 Banham, D., Kishimoto, T., Zhou, Y., Sato, T., Bai, K., Ozaki, J., … Ye, S. (2018).

14

Critical advancements in achieving high power and stable nonprecious metal catalyst – based MEAs for real-world proton exchange membrane fuel cell applications. Science Advances, 4(March), 1–8. https://doi.org/10.1126/sciadv.aar7180 Brown, H. C., & Brown, C. A. (1962). Reaction of Sodium Borohydride with Platinum Metal Salts in the Presence of Decolorizing Carbon-A Supported Platinum Catalyst of Markedly Enhanced Activity for Hydrogenations. Journal of the American Chemical Society, 84(14), 2827. https://doi.org/10.1021/ja00873a037 Bu, L., Ding, J., Guo, S., Zhang, X., Su, D., Zhu, X., … Huang, X. (2015). A General Method for Multimetallic Platinum Alloy Nanowires as Highly Active and Stable Oxygen Reduction Catalysts. Advanced Materials, 27(44), 7204–7212. https://doi.org/10.1002/adma.201502725 Chen, C., Kang, Y., Huo, Z., Zhu, Z., Huang, W., Xin, H. L., … Stamenkovic, V. R. (2014). Highly Crystalline Multimetallic Nanoframes with Three- Dimensional Electrocatalytic Surfaces. Science, 343(6177), 1339–1343. https://doi.org/10.1126/science.1249061 Chen, J., Herricks, T., Geissler, M., & Xia, Y. (2004). Single-Crystal Nanowires of Platinum can be Synthesized by Controlling the Reaction Rate of a Polyol Process. Journal of the American Chemical Society, 126(35), 10854–10855. https://doi.org/10.1021/ja0468224 Chen, Y., Liang, Z., Yang, F., Liu, Y., & Chen, S. (2011). Ni-Pt core-shell nanoparticles as oxygen reduction electrocatalysts: Effect of Pt shell coverage. Journal of Physical Chemistry C, 115(49), 24073–24079. https://doi.org/10.1021/jp207828n Choi, S. M., Kim, J. H., Jung, J. Y., Yoon, E. Y., & Kim, W. B. (2008). Pt nanowires prepared via a polymer template method: Its promise toward high Pt- loaded electrocatalysts for methanol oxidation. Electrochimica Acta, 53(19), 5804–5811. https://doi.org/10.1016/j.electacta.2008.03.041 Choi, W. C., & Woo, S. I. (2003). Bimetallic Pt-Ru nanowire network for anode material in a direct-methanol fuel cell. Journal of Power Sources, 124(2), 420–425. https://doi.org/10.1016/S0378-7753(03)00812-7 Dai, H. (2002). Carbon Nanotubes : Synthesis , Integration , and Properties. 1035– 1044. Day, T. M., Unwin, P. R., & Macpherson, J. V. (2007). Factors controlling the electrodeposition of metal nanoparticles on pristine single walled carbon nanotubes. Nano Letters, 7(1), 51–57. https://doi.org/10.1021/nl061974d Debe, M. K. (2012). Electrocatalyst approaches and challenges for automotive fuel cells. Nature, 486(7401), 43–51. https://doi.org/10.1038/nature11115 Debe, M. K. (2013). Tutorial on the Fundamental Characteristics and Practical Properties of Nanostructured Thin Film (NSTF) Catalysts. Journal of The

15

Electrochemical Society, 160(6), 522–534. https://doi.org/10.1149/2.049306jes Dicks, A., & Larminie, J. (2003). Fuel Cell Systems Explained. Wiley. Du, S. (2010). A Facile Route for Polymer Electrolyte Membrane Fuel Cell Electrodes with in situ Grown Pt Nanowires. Journal of Power Sources, 195(1), 289–292. https://doi.org/10.1016/j.jpowsour.2009.06.091 Du, S. (2012). Pt-based nanowires as electrocatalysts in proton exchange fuel cells. International Journal of Low-Carbon Technologies, 7(1), 44–54. https://doi.org/10.1093/ijlct/ctr027 Du, S., Lin, K., Malladi, S. K., Lu, Y., Sun, S., Xu, Q., … Dong, H. (2014). Plasma nitriding induced growth of Pt-nanowire arrays as high performance electrocatalysts for fuel cells. Scientific Reports, 4: 6439, 1–6. https://doi.org/10.1038/srep06439 Du, S., Lu, Y., Malladi, S. K., Xu, Q., & Steinberger-Wilckens, R. (2014). A simple approach for PtNi–MWCNT hybrid nanostructures as high performance electrocatalysts for the oxygen reduction reaction. Journal of Materials Chemistry A, 2(3), 692–698. https://doi.org/10.1039/c3ta13608f Du, S., Lu, Y., & Steinberger-Wilckens, R. (2014). PtPd nanowire arrays supported on reduced graphene oxide as advanced electrocatalysts for methanol oxidation. Carbon, 79(1), 346–355. https://doi.org/10.1016/j.carbon.2014.07.076 Fernandes, A. C., Paganin, V. A., & Ticianelli, E. A. (2010). Degradation study of Pt- based alloy catalysts for the oxygen reduction reaction in proton exchange membrane fuel cells. Journal of Electroanalytical Chemistry, 648(2), 156–162. https://doi.org/10.1016/j.jelechem.2010.07.013 Fievet, F., Lagier, J. P., & Figlarz, M. (1989). Preparing Monodisperse Métal Powders in Micrometer and Submicrometer Sizes by the Polyol Process. MRS Bulletin, 29–34. https://doi.org/10.1557/S0883769400060930 Garsany, Y., Baturina, O. a, Swider-Lyons, K. E., & Kocha, S. S. (2010). Experimental Methods for Quantifying the Activity of Platinum Electrocatalysts for the Oxygen Reduction Reaction. Analytical Chemistry, 82(15), 6321–6328. https://doi.org/10.1021/ac100306c Gasteiger, H. A., Kocha, S. S., Sompalli, B., & Wagner, F. T. (2005). Activity benchmarks and requirements for Pt, Pt-alloy, and non-Pt oxygen reduction catalysts for PEMFCs. Applied Catalysis B: Environmental, 56(1-2 SPEC. ISS.), 9–35. https://doi.org/10.1016/j.apcatb.2004.06.021 Greeley, J., Stephens, I. E. L., Bondarenko, a S., Johansson, T. P., Hansen, H. a, Jaramillo, T. F., … Nørskov, J. K. (2009). Alloys of platinum and early transition metals as oxygen reduction electrocatalysts. Nature Chemistry, 1(7), 552–556. https://doi.org/10.1038/nchem.367 Grove, W. R. (1839). On Voltaic series and the Combination of Gases by Platinum. 16

Philosophical Magazine Series 3, 14, 127–130. He, Q., Yang, X., Chen, W., Mukerjee, S., Koel, B., & Chen, S. (2010). Influence of phosphate anion adsorption on the kinetics of oxygen electroreduction on low index Pt(hkl) single crystals. Physical Chemistry Chemical Physics, 12(39), 12544–12555. https://doi.org/10.1039/c0cp00433b Ikeda, T., Boero, M., Huang, S. F., Terakura, K., Oshima, M., & Ozaki, J. (2008). Carbon alloy catalysts: Active sites for oxygen reduction reaction. Journal of Physical Chemistry C, 112(38), 14706–14709. https://doi.org/10.1021/jp806084d Inoue, G., & Kawase, M. (2016). Understanding formation mechanism of heterogeneous porous structure of catalyst layer in polymer electrolyte fuel cell. International Journal of Hydrogen Energy, 41(46), 21352–21365. https://doi.org/10.1016/j.ijhydene.2016.08.029 Ioroi, T., Siroma, Z., Fujiwara, N., Yamazaki, S. I., & Yasuda, K. (2005). Sub- stoichiometric titanium oxide-supported platinum electrocatalyst for polymer electrolyte fuel cells. Electrochemistry Communications, 7(2), 183– 188. https://doi.org/10.1016/j.elecom.2004.12.007 Islam, A., Bhuiya, M. A. K., & Islam, M. S. (2014). A Review on Chemical Synthesis Process of Platinum Nanoparticles. Asia Pacific Journal of Energy and Environment, 1(2), 107–121. https://doi.org/10.15590/apjee/2014/v1i2/53749 Jiang, J., & Kucernak, A. (2003). Electrooxidation of small organic molecules on mesoporous precious metal catalysts II: CO and methanol on platinum- ruthenium alloy. Journal of Electroanalytical Chemistry, 543(2), 187–199. https://doi.org/10.1016/S0022-0728(03)00046-9 Jinnouchi, R., Kudo, K., Kitano, N., & Morimoto, Y. (2016). Molecular Dynamics Simulations on O2 Permeation through Nafion Ionomer on Platinum Surface. Electrochimica Acta, 188, 767–776. https://doi.org/10.1016/j.electacta.2015.12.031 Kim, K. H., Lee, K. Y., Kim, H. J., Cho, E. A., Lee, S. Y., Lim, T. H., … Jang, J. H. (2010). The effects of Nafion?? ionomer content in PEMFC MEAs prepared by a catalyst-coated membrane (CCM) spraying method. International Journal of Hydrogen Energy, 35(5), 2119–2126. https://doi.org/10.1016/j.ijhydene.2009.11.058 Kinoshita, K. (1990). Particle Size Effects for Oxygen Reduction on Highly Dispersed Platinum in Acid Electrolytes. Journal of The Electrochemical Society, 137(3), 845–848. https://doi.org/10.1149/1.2086566 Li, H., Zhang, S., Yan, S., Lin, Y., & Ren, Y. (2013). Pd/C catalysts synthesized by microwave assisted polyol method for methanol electro-oxidation. International Journal of Electrochemical Science, 8(2), 2996–3011. Li, J., Pan, M., & Tang, H. (2014). Understanding short-side-chain perfluorinated

17

sulfonic acid and its application for high temperature polymer electrolyte membrane fuel cells. RSC Advances, 4(8), 3944–3965. https://doi.org/10.1039/c3ra43735c Li, M., Li, M., Zhao, Z., Cheng, T., Fortunelli, A., Chen, C., … Duan, X. (2016). Ultrafine jagged platinum nanowires enable ultrahigh mass activity for the oxygen reduction reaction. Science, 9050(November), 1414–1419. https://doi.org/10.1126/science.aaf9050 Litster, S., & McLean, G. (2004). PEM fuel cell electrodes. Journal of Power Sources, 130(1–2), 61–76. https://doi.org/10.1016/j.jpowsour.2003.12.055 Liu, J., & Huang, Y. (2018). Oxygen Reduction Reaction on PtCo Nanocatalyst: (Bi)sulfate Anion Poisoning. Nanoscale Research Letters, 13. https://doi.org/10.1186/s11671-018-2574-6 Lu, Y., Du, S., & Steinberger-Wilckens, R. (2015). Temperature-controlled growth of single-crystal Pt nanowire arrays for high performance catalyst electrodes in polymer electrolyte fuel cells. Applied Catalysis B: Environmental, 164, 389–395. https://doi.org/10.1016/j.apcatb.2014.09.040 Lu, Y., Du, S., & Steinberger-Wilckens, R. (2016). One-dimensional nanostructured electrocatalysts for polymer electrolyte membrane fuel cells—A review. Applied Catalysis B: Environmental, 199, 292–314. https://doi.org/10.1016/j.apcatb.2016.06.022 Ma, Y., & Balbuena, P. B. (2008). Pt surface segregation in bimetallic Pt3M alloys: A density functional theory study. Surface Science, 602(1), 107–113. https://doi.org/10.1016/j.susc.2007.09.052 Mardle, P., Fernihough, O., Du, S.F. (2018) Evaluation of the scaffolding effect of Pt nanowires supported on reduced graphene oxide in PEMFC electrodes. Coatings 8 (2), 48, https://doi.org/10.3390/coatings8020048 Mardle, P., Ji, X., Wj, J., Guan, S.L., Dong, H.S., & Du, S.F. (2020) Thin film electrodes from Pt nanorods supported on aligned N-CNTs for proton exchange membrane fuel cells. Applied Catalysis B: Environmental, 260, 118031. https://doi.org/10.1016/j.apcatb.2019.118031 Matsumoto, T., Komatsu, T., Arai, K., Yamazaki, T., Kijima, M., Shimizu, H., … Nakamura, J. (2004). Reduction of Pt usage in fuel cell electrocatalysts with carbon nanotube electrodes. Chemical Communications, (7), 840–841. https://doi.org/10.1039/B400607K Mehmood, A., & Ha, H. Y. (2012). An efficient decal transfer method using a roll- press to fabricate membrane electrode assemblies for direct methanol fuel cells. International Journal of Hydrogen Energy, 37(23), 18463–18470. https://doi.org/10.1016/j.ijhydene.2012.09.045 Molina-garcía, M. A., & Rees, N. V. (2018). “Metal-free ” electrocatalysis : Quaternary-doped graphene and the alkaline oxygen reduction reaction.

18

Applied Catalysis A, General, 553(August 2017), 107–116. https://doi.org/10.1016/j.apcata.2017.12.014 More, K., Cullen, D. A., Sneed, B. T., & Reeves, K. S. (2016). New Fuel Cell Materials: Characterization and Method Development. DOE Hydrogen and Fuel Cells Annual Merit Review Meeting. Retrieved from https://www.hydrogen.energy.gov/pdfs/review16/fc020_more_2016_o.pdf Mukerjee, S., Srinivasan, S., Soriaga, M. P., & Mcbreen, J. (1995). Role of Structural and Electronic Properties of Pt and Pt Alloys on Electrocatalysis of Oxygen Reduction. Journal of The Electrochemical Society, 142(5), 1409–1421. https://doi.org/10.1149/1.2048590 Multi-Year Research, Development, and Demonstration Plan: 3.4 Fuel Cells. (2017). US Department of Energy (DOE). Nakajima, H., & Honma, I. (2002). Proton-conducting hybrid solid electrolytes for intermediate temperature fuel cells. Solid State Ionics, 148(3–4), 607–610. https://doi.org/10.1016/S0167-2738(02)00127-3 Nørskov, J. K., Ressmeisl, J., Logadottir, A., Lindqvist, L., Kitchin, J. R., Bligaard, T., & Jonsson, H. (2004). Origin of the overpotential for oxygen reduction at a fuel cell cathode. J. Phys. Chem. B, 108, 17886–17892. https://doi.org/10.1021/jp047349j Okanishi, T., Matsui, T., Takeguchi, T., Kikuchi, R., & Eguchi, K. (2006). Chemical interaction between Pt and SnO2 and influence on adsorptive properties of carbon monoxide. Applied Catalysis A: General, 298(1–2), 181–187. https://doi.org/10.1016/j.apcata.2005.09.035 Papageorgopoulos, D. (2019). Fuel Cell R&D Overview. 2019 Annual Merit Review and Peer Evaluation Meeting. US DoE - Fuel Cells Technologies Office. Park, K.-W., & Seol, K.-S. (2007). Nb-TiO2 supported Pt cathode catalyst for polymer electrolyte membrane fuel cells. Electrochemistry Communications, 9(9), 2256–2260. https://doi.org/10.1016/j.elecom.2007.06.027 Park, K. W., Choi, J. H., Kwon, B. K., Lee, S. A., Sung, Y. E., Ha, H. Y., … Wieckowski, A. (2002). Chemical and electronic effects of Ni in Pt/Ni and Pt/Ru/Ni alloy nanoparticles in methanol electrooxidation. Journal of Physical Chemistry B, 106(8), 1869–1877. https://doi.org/10.1021/jp013168v Park, S., Shao, Y., Wan, H., Rieke, P. C., Viswanathan, V. V., Towne, S. A., … Wang, Y. (2011). Design of graphene sheets-supported Pt catalyst layer in PEM fuel cells. Electrochemistry Communications, 13(3), 258–261. https://doi.org/10.1016/j.elecom.2010.12.028 Parthasarathy, P., & Virkar, A. V. (2013). Electrochemical Ostwald ripening of Pt and Ag catalysts supported on carbon. Journal of Power Sources, 234, 82–90. https://doi.org/10.1016/j.jpowsour.2013.01.115 Peron, J., Shi, Z., & Holdcroft, S. (2011). Hydrocarbon proton conducting polymers for fuel cell catalyst layers. Energy and Environmental Science, 4(5), 19

1575–1591. https://doi.org/10.1039/c0ee00638f Ramachandran, S., & Stimming, U. (2015). Well to wheel analysis of low carbon alternatives for road traffic. The Royal Society of Chemistry, 8(8), 3313–3324. https://doi.org/10.1039/c5ee01512j Salgado, J. R. C., Antolini, E., & Gonzalez, E. R. (2004). Pt-Co/C electrocatalysts for oxygen reduction in H-2/O-2 PEMFCs synthesized by borohydride method. Journal of the Electrochemical Society, 151, A2143–A2149. https://doi.org/10.1149/1.1819832 Santiago, E. I., Varanda, L. C., & Villullas, H. M. (2007). Carbon-supported Pt-Co catalysts prepared by a modifled polyol process as for PEM fuel cells. Journal of Physical Chemistry C, 111(7), 3146–3151. https://doi.org/10.1021/jp0670081 Shao, M., Chang, Q., Dodelet, J.-P., & Chenitz, R. (2016). Recent Advances in Electrocatalysts for Oxygen Reduction Reaction. Chemical Reviews, 116(6), 3594–3657. https://doi.org/10.1021/acs.chemrev.5b00462 Sharma, R., Wang, Y., Li, F., Chamier, J., & Andersen S.M. Particle Size-Controlled Growth of Carbon-Supported Platinum Nanoparticles (Pt/C) through Water- Assisted Polyol Synthesis. ACS Omega, 2019 Sep 24; 4(13): 15711–15720. https://doi.org/10.1021/acsomega.9b02351 Sharma, S., & Pollet, B. G. (2012). Support materials for PEMFC and DMFC electrocatalysts - A review. Journal of Power Sources, 208, 96–119. https://doi.org/10.1016/j.jpowsour.2012.02.011 Shinozaki, K., Morimoto, Y., Pivovar, B. S., & Kocha, S. S. (2016). Suppression of oxygen reduction reaction activity on Pt-based electrocatalysts from ionomer incorporation. Journal of Power Sources, 325, 745–751. https://doi.org/10.1016/j.jpowsour.2016.06.062 Shui, J., Wang, M., Du, F., & Dai, L. (2015). N-doped carbon nanomaterials are durable catalysts for oxygen reduction reaction in acidic fuel cells. Science Advances, 1(1), 1–8. https://doi.org/10.1126/sciadv.1400129 Si, Y., & Samulski, E. T. (2008). Exfoliated graphene separated by platinum nanoparticles. Chemistry of Materials, 20(21), 6792–6797. https://doi.org/10.1021/cm801356a Snyder, J., McCue, I., Livi, K., & Erlebacher, J. (2012). Structure/Processing/Properties Relationships in Nanoporous Nanoparticles As Applied to Catalysis of the Cathodic Oxygen Reduction Reaction. Journal of American Chemical Society, 134, 8633–8645. https://doi.org/10.1021/ja3019498 Snyder, J., Livi, K., & Erlebacher, J. (2013). Oxygen reduction reaction performance of [MTBD][beti]-encapsulated nanoporous NiPt alloy nanoparticles. Advanced Functional Materials, 23(44), 5494–5501. https://doi.org/10.1002/adfm.201301144

20

Stamenkovic, V., Mun, B. S., Mayrhofer, K. J. J., Ross, P. N., Markovic, N. M., Rossmeisl, J., … Nørskov, J. K. (2006). Changing the Activity of Electrocatalysts for Oxygen Reduction by Tuning the Surface Electronic Structure. Angewandte Chemie, 118(18), 2963–2967. https://doi.org/10.1002/ange.200504386 Stephens, I. E. L., Bondarenko, A. S., Rossmeisl, J., Grønbjerg, U., Rossmeislc, J., & Chorkendorff, I. (2012). Understanding the electrocatalysis of oxygen reduction on platinum and its alloys. Energy & Environmental Science, 5, 6744-6762. https://doi.org/10.1039/C2EE03590A Stephens, I. E. L., Rossmeisl, J., & Chorkendorff, I. (2016). Toward sustainable fuel cells. Science, 354(6318), 1378–1380. https://doi.org/10.1126/science.aal3303 Strasser, P., Koh, S., Anniyev, T., Greeley, J., More, K., Yu, C., … Nilsson, A. (2010). Lattice-strain control of the activity in dealloyed core–shell fuel cell catalysts. Nature Chemistry, 2(6), 454–460. https://doi.org/10.1038/nchem.623 Sun, S., Jaouen, F., & Dodelet, J. P. (2008). Controlled growth of Pt nanowires on carbon nanospheres and their enhanced performance as electrocatalysts in PEM fuel cells. Advanced Materials, 20(20), 3900–3904. https://doi.org/10.1002/adma.200800491 Sun, S., Zhang, G., Geng, D., Chen, Y., Li, R., Cai, M., & Sun, X. (2011). A highly durable platinum nanocatalyst for proton exchange membrane fuel cells: Multiarmed starlike nanowire single crystal. Angewandte Chemie - International Edition, 50(2), 422–426. https://doi.org/10.1002/anie.201004631 Trasatti, S. (1972). WORK FUNCTION , ELECTRONEGATIVITY , AND ELECTROCHEMICAL BEHAVIOUR OF METALS III . ELECTROLYTIC HYDROGEN EVOLUTION IN ACID SOLUTIONS Although the dependence of hydrogen overvoltage ( qH ) on electrode material was experimentally recognized at the beginning. 39, 163–184. Tsotridis, G., Pilenga, A., Marco, G. De, & Malkow, T. (2015). EU Harmonised Test Protocols for PEMFC MEA Testing in Single Cell Configuration for Automotive Applications; JRC Science for Policy report. https://doi.org/10.2790/54653 Wang, C., Daimon, H., Onodera, T., Koda, T., & Sun, S. (2008). A general approach to the size- and shape-controlled synthesis of platinum nanoparticles and their catalytic reduction of oxygen. Angewandte Chemie - International Edition, 47(19), 3588–3591. https://doi.org/10.1002/anie.200800073 Wang, M., Zhang, H.X., Thirunavukkarasu, G., Salam, I., Varcoe, J.R., Mardle, P., Li, X.Y., Mu, S.C., & Du, S.F. (2019) Ionic Liquid-Modified Microporous ZnCoNC- Based Electrocatalysts for Polymer Electrolyte Fuel Cells. ACS Energy Letters, 4 (9), 2104-2110. https://doi.org/10.1021/acsenergylett.9b01407 Wang, G., Wu, H., Wexler, D., Liu, H., & Savadogo, O. (2010). Ni@Pt core-shell

21

nanoparticles with enhanced catalytic activity for oxygen reduction reaction. Journal of Alloys and Compounds, 503(1), 2–5. https://doi.org/10.1016/j.jallcom.2010.04.236 Wang, W., Chen, S., Li, J., & Wang, W. (2014). Fabrication of catalyst coated membrane with screen printing method in a proton exchange membrane fuel cell. International Journal of Hydrogen Energy, (August 2016). https://doi.org/10.1016/j.ijhydene.2015.02.027 Wang, Xiaoming, Orikasa, Y., Takesue, Y., Inoue, H., Nakamura, M., Minato, T., … Uchimoto, Y. (2013). Quantitating the lattice strain dependence of monolayer Pt shell activity toward oxygen reduction. Journal of the American Chemical Society, 135(16), 5938–5941. https://doi.org/10.1021/ja312382h Wang, Xiqing, Lee, J. S., Zhu, Q., Liu, J., Wang, Y., & Dai, S. (2010). Ammonia- treated ordered mesoporous carbons as catalytic materials for oxygen reduction reaction. Chemistry of Materials, 22(7), 2178–2180. https://doi.org/10.1021/cm100139d Wendt, H., Spinacé, E. V., Oliveira Neto, A., & Linardi, M. (2005). Electrocatalysis and electrocatalysts for low temperature fuel cells: fundamentals, state of the art, research and development. Química Nova, 28(6), 1066–1075. https://doi.org/10.1590/S0100-40422005000600023 Wildgoose, G. G., Banks, C. E., & Compton, R. G. (2006). Metal nanoparticles and related materials supported on Carbon nanotubes: Methods and applications. Small, 2(2), 182–193. https://doi.org/10.1002/smll.200500324 Wilson, M. S., & Gottesfeld, S. (1992). Thin-film catalyst layers for polymer electrolyte fuel cell electrodes. Journal of Applied Electrochemistry, 22(1), 1– 7. https://doi.org/10.1007/BF01093004 Xia, B. Y., Yang, P., Sun, Y., Wu, Y., Mayers, B., Gates, B., … Yan, H. (2003). One- Dimensional Nanostructures : Synthesis , Characterization , and Applications **. (5), 353–389. https://doi.org/10.1002/adma.200390087 Yarlagadda, V., Carpenter, M. K., Moylan, T. E., Kukreja, R. S., Koestner, R., Gu, W., … Kongkanand, A. (2018). Boosting Fuel Cell Performance with Accessible Carbon Mesopores. ACS Energy Letters, 3, 618–621. https://doi.org/10.1021/acsenergylett.8b00186 Yu, P., Pemberton, M., & Plasse, P. (2005). PtCo/C cathode catalyst for improved durability in PEMFCs. Journal of Power Sources, 144(1), 11–20. https://doi.org/10.1016/j.jpowsour.2004.11.067 Zacharia, R., Rather, S. ullah, Hwang, S. W., & Nahm, K. S. (2007). Spillover of physisorbed hydrogen from sputter-deposited arrays of platinum nanoparticles to multi-walled carbon nanotubes. Chemical Physics Letters, 434(4–6), 286–291. https://doi.org/10.1016/j.cplett.2006.12.022

22

Zaidi, J., & Matsuura, T. (2010). Polymer Membranes for Fuel Cells. Springer. Zhang, G. R., Munoz, M., & Etzold, B. J. M. (2015). Boosting performance of low temperature fuel cell catalysts by subtle ionic liquid modification. ACS Applied Materials and Interfaces, 7(6), 3562–3570. https://doi.org/10.1021/am5074003 Zhang, H.X., Liang, J.Y., Xia, B.W., Li, Y., & Du, S.F. (2019) Ionic liquid modified Pt/C electrocatalysts for cathode application in proton exchange membrane fuel cells. Frontiers of Chemical Science and Engineering, 13 (4), 695-701. https://doi.org/10.1007/s11705-019-1838-8 Zhang, J., Lima, F. H. B., Shao, M. H., Sasaki, K., Wang, J. X., Hanson, J., & Adzic, R. R. (2005). Platinum monolayer on non-noble metal-noble metal core-shell nanoparticles electrocatalysts for O2 reduction. The Journal of Physical Chemistry B, 109(48), 11973. https://doi.org/10.1021/jp055634c Zhang, J., Mo, Y., Vukmirovic, M. B., Klie, R., Sasaki, K., & Adzic, R. R. (2004). Platinum monolayer electrocatalysts for O 2 reduction: Pt monolayer on Pd(111) and on carbon-supported Pd nanoparticles. Journal of Physical Chemistry B, 108(30), 10955–10964. https://doi.org/10.1021/jp0379953 Zhao, G. Y., Xu, C. L., Guo, D. J., Li, H., & Li, H. L. (2006). Template preparation of Pt-Ru and Pt nanowire array electrodes on a Ti/Si substrate for methanol electro-oxidation. Journal of Power Sources, 162(1), 492–496. https://doi.org/10.1016/j.jpowsour.2006.06.082 Zhou, Y., & Zhang, D. (2015). Nano PtCu binary and PtCuAg ternary alloy catalysts for oxygen reduction reaction in proton exchange membrane fuel cells. Journal of Power Sources, 278, 396–403. https://doi.org/10.1016/j.jpowsour.2014.12.088

23