Next Article in Journal
Identification of Therapeutic Targets for Medulloblastoma by Tissue-Specific Genome-Scale Metabolic Model
Previous Article in Journal
Development of Quantitative Chemical Ionization Using Gas Chromatography/Mass Spectrometry and Gas Chromatography/Tandem Mass Spectrometry for Ambient Nitro- and Oxy-PAHs and Its Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Low Pt Loading Membrane Electrode Assembly for Proton Exchange Membrane Fuel Cells

1
National Laboratory of Solid State Microstructures (NLSSM), Collaborative Innovation Center of Advanced Microstructures, College of Engineering and Applied Sciences, Nanjing University, Nanjing 210093, China
2
School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210093, China
3
Heraeus Precious Metal Technology (China) Co., Ltd., Nanjing 211511, China
4
Institute of Energy Power Innovation, North China Electric Power University, Beijing 102206, China
5
State Key Laboratory of Mechanics and Control of Mechanical Structures, Key Laboratory for Intelligent Nano Materials and Devices of the Ministry of Education, College of Material Science and Engineering, Nanjing University of Aeronautics and Astronautics, Nanjing 210016, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(2), 773; https://doi.org/10.3390/molecules28020773
Submission received: 29 November 2022 / Revised: 3 January 2023 / Accepted: 6 January 2023 / Published: 12 January 2023

Abstract

:
Hydrogen has the potential to be one of the solutions that can address environmental pollution and greenhouse emissions from traditional fossil fuels. However, high costs hinder its large-scale commercialization, particularly for enabling devices such as proton exchange membrane fuel cells (PEMFCs). The precious metal Pt is indispensable in boosting the oxygen reduction reaction (ORR) in cathode electrocatalysts from the most crucial component, i.e., the membrane electrode assembly (MEA). MEAs account for a considerable amount of the entire cost of PEMFCs. To address these bottlenecks, researchers either increase Pt utilization efficiency or produce MEAs with enhanced performance but less Pt. Only a few reviews that explain the approaches are available. This review summarizes advances in designing nanocatalysts and optimizing the catalyst layer structure to achieve low-Pt loading MEAs. Different strategies and their corresponding effectiveness, e.g., performance in half-cells or MEA, are summarized and compared. Finally, future directions are discussed and proposed, aiming at affordable, highly active, and durable PEMFCs.

1. Introduction

Energy demand is constantly increasing, and the accompanying environmental pollution and greenhouse effect are intensifying. Developing renewable energy technologies and promoting their commercialization are the keys to meeting energy demands while protecting the environment [1,2,3,4]. Hydrogen is one of the most promising forms of renewable energy because it has various forms of utilization and is an environmentally friendly product. Additionally, its advantages in safety, energy density, and storage simplicity are noteworthy [5]. However, its large-scale application is currently hindered by the high costs of devices that convert hydrogen [6,7,8]. Proton exchange membrane fuel cells (PEMFCs) are crucial for efficiently converting the chemical energy stored in hydrogen into clean electricity [9,10,11,12]. Hence, increasing the output power density and decreasing the cost of PEMFCs have become the goals of researchers in the field. These measures are shown in Scheme 1.
A single-cell PEMFC is shown in Figure 1a [13]. Its components include a flow field plate, a sealing layer, a gas diffusion layer (GDL), a catalyst layer (CL) on both sides of the anode and cathode, and a proton-exchange membrane (PEM) in the middle. GDL, CL, and PEM constitute the membrane electrode assembly (MEA), which is regarded as the most critical part of a PEMFC. The key electrochemical process of unleashing the chemical energy occurs in MEA, and the reaction speed determines the output power density in practical application. The so-called three-phase interface, where the gas, ionomer, and catalyst converge and where electrochemical reactions occur, is found in the CL. During PEMFC operation, electrochemical reactions occur simultaneously at both sides of the electrode interface. As shown in Figure 1b, H2 from the anode side first enters the flow field plate. It then passes through the GDL and goes inside the CL. A hydrogen oxidation reaction occurs at the anode electrocatalyst, producing protons and electrons. Protons are then transported through the PEM to the cathode side, and electrons travel through the external circuit. Similarly, O2 reaches the CL on the cathode side through a similar diffusion path and undergoes an oxygen reduction reaction (ORR) at the three-phase interface at the cathode electrocatalysts. The fuel cell continuously produces electricity with a continuous supply of hydrogen and air. Current research suggests that PEMFC performance is mainly determined by the activity of the cathodic catalyst and CL; a sluggish ORR is a rate-determining step.
Hence, many electrocatalysts (in most cases, Pt-based catalysts) are required to boost electrochemical reactions, increasing PEMFC manufacturing costs. Moreover, electrocatalyst stability and MEAs that can survive long-term service are critical for practical application. Therefore, optimizing the electrocatalyst and the CL structure (determining interfaces) to achieve minimum precious metal usage, high activity, and long-term stability is critical. Thus, PEMFCs require MEAs with low-Pt loading and with ideal output power density and durability. Researchers have proposed different strategies. Introducing additional elements [14,15,16], modulating the morphologies of electrocatalysts [17,18,19], and adjusting engineering parameters for CL [20,21,22,23] have been proven effective, and considerable progress has been achieved.
This review presents a summary of recent publications about low-Pt loading in MEAs for PEMFCs. Two perspectives on different scales are discussed: ORR electrocatalysts and the CL structure in MEA. The progress in ORR electrocatalysts is divided into three parts: nanocatalyst alloying, morphology design, and catalyst–support development. For CL structures, preparation methods, catalyst ink modulation, and unique CL structure designs are described. Furthermore, a comprehensive review is provided, ranging from enhancing microscopic electrochemical reactions to lowering mesoscopic mass transfer resistance and managing the multiphase flow of gas/water reactant/product. We also provide an outlook on future developments in durable electrocatalysts and CLs. This review is a valuable introduction for readers new to this field and a source of new insights and perspectives for experienced researchers.

2. Electrocatalysts for ORR in MEA

2.1. Alloying with Transition Metals

The widely adopted electrocatalyst for ORR is Pt nanoparticles loaded onto carbon supports. However, its catalytic activity is limited. Moreover, its stability is limited under harsh operating conditions due to the Ostwald ripening and particle migration processes [24]. Alloying can optimize the electronic structure of Pt nanoparticles and regulate their adsorption of oxygenated species, enhancing catalyst performance while reducing Pt loading [25,26]. Therefore, alloying Pt with other transition metals has been an important research direction in recent years [27]. Markovic et al. conducted a comprehensive study on the fundamental relationship between the surface electronic structure and ORR performance of two different Pt3M (M represents other transition metals) catalysts. They found that different compositions affect the electronic structure and alter the catalytic activity of the catalysts [28]. Pt3Co was found to exhibit the best specific activity based on the volcano-shaped relationship between the ORR catalytic activity and the corresponding descriptor of electronic structure, the d-band center, in Pt3M. Therefore, the study has proven that appropriate composition adjustment is a way to improve the ORR activity of PtM alloys.
In previous studies, PtCo and PtNi alloy catalysts have shown great potential as substitutes for Pt in PEMFCs. Huang et al. used wet-chemical methods to prepare Pt3Co nanowires with abundant high-index surfaces enriched with Pt atoms and ordered structures [29]. Because of this unique structure, the catalyst is highly active in boosting ORR, with a mass activity (MA) and specific activity (SA) 33.7 and 39.6 times superior to commercial Pt/C, respectively. Furthermore, the catalyst composition and structure remain almost unchanged after 20,000 dynamic voltage cycles in the accelerated degradation test (ADT). The MA retains 91.9% of the initial value. Xia et al. prepared one-dimensional PtNi bunched nanospheres with different Pt and Ni ratios (Figure 2a) [30]. The final product, one-dimensional PtNi bunched nanocages, was synthesized using selective removal of nickel species under acidic conditions. The MA and SA of the best-performing catalyst are 3.52 A and 5.16 mA cmPt−2, far exceeding those of commercial catalysts (0.21 and 0.36 mA cmPt−2 for MA and SA, respectively). The catalyst activity did not decay considerably after 50 k dynamic voltage cycles in ADT, reflecting its excellent durability. The outstanding ORR performance of PtNi bunched nanocages should be attributed to the unique structure, the suitable Pt/Ni ratio, and the synergistic effects of stress and coordination. This finding provides new insights into the structural modulation of alloy catalysts.
Structurally ordered intermetallic compounds (IMCs), members of PtM catalysts, have been reported to be the most promising ORR electrocatalysts due to their unique properties [33]. By carefully controlling the structure of the catalysts to a highly ordered form, the unique IMC electronic structure remarkably improves the ORR catalytic performance [34]. By using structural and electrochemical characterizations, Mayrhofer et al. found that ordered, structured alloy catalysts exhibit superior activity and stability than disordered, structured ones [35]. However, their preparation usually involves a high-temperature annealing treatment. This treatment inevitably causes the IMC particles to grow in size, yielding low catalyst surface exposure. The result is limited performance on MEA in single-cell tests, despite promising activity observed in rotating disk electrodes (RDE) in half-cell tests [36]. To address this issue, Liang et al. employed a novel sulfur anchoring method (Figure 2b) [31]. This method takes advantage of the strong interaction between sulfur and platinum atoms to anchor the Pt-based nanoparticles on the carbon support like “solid glue.” The interaction prevents nanoparticle aggregation and enhances its catalytic durability. As a result, they were able to synthesize small-sized IMC catalysts and construct a material library consisting of 46 Pt-based binary and multiple IMC catalysts, which exhibit excellent MEA performance. In a later work, they further developed a synthetic strategy using heteroatom-containing molecules as additives in the impregnation process. A universal synthesis of 18 small-sized binary Pt IMC catalysts through the dual action of Pt-S chemical bonding and physical domain limiting could be achieved. The prepared PtCo catalysts exhibit outstanding MEA performance at low-Pt loading [37].
Doping is another way of improving the activity and durability of Pt alloy nanocatalysts. Li et al. synthesized ultrafine L10-PtCo nanoparticles doped with a small number of metals [38]. The catalyst performance is remarkably better than that of commercial Pt/C catalysts. Moreover, excellent MEA activity and stability have been observed after 50 k dynamic voltage cycles at 80 °C. Extended X-ray absorption fine structure analysis and density functional theory (DFT) simulations have also indicated that W doping stabilizes the intermetallic structure and adjusts the Pt-Pt bond distance, thus weakening the adsorption of oxygen-containing species at the Pt surface. In another recently published work, Duan et al. prepared sub-2 nm Co-doped Pt nanoclusters using micro-wave assisted polyol reduction (Figure 2c) [32]. Unprecedented activity and stability were achieved from the as-prepared nanoclusters, mainly due to the extensive exposure of unique low-coordination sites on the surface of these nanoclusters. With a small amount of cobalt doping, the electrochemically active surface area (ECSA) of Pt nanoclusters is remarkably higher than that of undoped and commercial catalysts. Structural and compositional characterizations and DFT simulations show that the total energy of the nanoclusters decreases after doping. This decrease reflects nanocluster stabilization. Doping further reduces the d-band center and optimizes the binding energy of oxygen-containing intermediates, reducing the energy-determining step barrier and enhancing nanocluster activity. In real H2-air fuel cell evaluations, the catalysts easily outperform commercial Pt/C. More importantly, the commonly observed considerable performance loss in the high-current density region due to decreased Pt loading is substantially alleviated. In addition, doping also increases vacancies and improves nanocluster stability. As a result, the Co-doped nanoclusters have been proven to be more stable than a commercial catalyst in an ADT test of 30 k dynamic voltage cycles.
Currently, the effects of introducing additional transition metals are recognized as ligand and strain effects. The ligand effect refers to the direct electron transfer between Pt and the metal atoms, while the strain effect means lattice compression induced by nearby atoms with different atomic radii. These two effects shift the Pt d-band center downward and reduce the Pt binding energy and reaction intermediates, thus optimizing the ORR pathways [39]. However, the most serious problem faced by current Pt alloys is poor durability. Under high potential and strong acidic electrolytes during PEMFC operation, the introduced transition metals are susceptible to oxidation and dissolution. Additionally, the dissolved metal ions may cause ionomer and PEM degradation, resulting in poor MEA performance [40]. Fine-tuning the morphology of Pt alloy nanoparticles may be a possible solution. As a surface reaction, the geometry and structure of the catalyst influence the catalytic activity and stability. Therefore, besides modulating the Pt-based nanocrystal components, optimizing the morphology is also an effective strategy for increasing Pt-based catalyst efficiency [41].

2.2. Optimization of Morphology

In recent years, researchers have designed and fabricated diverse Pt-based alloy geometry structures, including nanowires [42], nanoplates [43], nanoframes [44], and other special structures [45,46]. In general, increased active site exposure and accelerated electronic conduction afford excellent performance. Moreover, unique morphologies and enhanced crystallinity improve durability, making PEMFCs suitable for practical applications [47]. Huang et al. reported a novel catalyst based on PtGa ultrafine nanowires (Pt4.31Ga NWs/C), which exhibit strong p-d orbital hybridization interactions by doping with p-region metal elements (Figure 3a). This strategy inhibits component dissolution and precipitation of the alloy catalyst, thereby achieving high durability [48]. Notably, as a result, the Pt4.31Ga NWs/C catalyst shows only a 15.8% loss in MA after 30 k dynamic voltage cycles. In contrast, 79.6% are lost in commercial Pt/C under the same condition. DFT simulations show that the p-d orbital hybridization interaction between Pt and Ga atoms optimizes the surface electronic structure and improves the Pt oxidation resistance. Additionally, Ga dissolution is also inhibited, causing a remarkable improvement in ORR performance.
Particle size also affects electrocatalytic activity because ORR is a surface reaction. Small nanoparticles have high ECSA and expose many active sites under the same Pt loading, yielding a high MA [51]. However, catalyst particle size reduction brings a disadvantage in surface energy. They are more likely to agglomerate, causing a decrease in durability. However, this issue could also be mitigated using a delicate design of catalyst structures [52]. Huang et al. designed a specially structured catalyst in which a graphene shell encapsulates PtCo nanoparticles (Figure 3b). The nanocatalyst exhibits excellent electrochemical activity. Most impressively, it shows outstanding durability at an ultra-low-Pt loading of 0.07 mgPt cm−2 due to the noncontact shell of the graphene nanopocket [49]. With this design, catalyst particle aggregation and shedding are limited without affecting mass transport during ORR.
The surface properties of nanoparticles are closely related to their structure. Specifically, surface catalysis in transition metals depends mainly on the crystal facets. Furthermore, different nanocrystal morphologies afford varying degrees of exposure from different crystal facets, causing variation in efficiency and selectivity during catalytic processes [53]. Previous studies have reported that ORR kinetics at the Pt surface in 0.1 M HClO4 varies with the Miller index. The trend follows the order Pt (100) < Pt (111) < Pt (110) [54]. Therefore, modulating the catalyst structure to expose specific preferred facets is also a promising and important way to enhance ORR activity. For example, PtNi octahedral nanocrystals have a higher proportion of (111) facets than cubic nanocrystals; hence, their theoretical activity is higher than that of cubic nanocrystals. Xia et al. synthesized Pt2.5Ni octahedra to promote (111) facet formation and found that the activity of PtNi octahedra is remarkably enhanced due to the clean and well-preserved surfaces of the catalyst, with 17 and 51 times more MA and SA than commercial Pt/C, respectively [55]. Similarly, Huang et al. prepared highly dispersed PtNi octahedra on carbon supports, which improved surface exposure and catalytic activity. Good stability is observed after 6 k cycles of potential cycling [56]. Furthermore, the design of the core–shell structure can improve the utilization of Pt atoms and further decrease the amounts of precious metals. Because of this, ideal core–shell interactions can be obtained to enhance the performance of catalysts. Zeng et al. synthesized Pd@Pt1.8Ni octahedra with an ultra-thin alloy shell. The structure improves catalytic performance mainly due to the PtNi alloy and the interaction between the shell layer and the nucleus, which changes the electronic structure and atomic geometry [57]. Previous reports have proven that the high-index Pt crystal planes have better catalytic activity than other low-index crystal surfaces. Moreover, the core–shell structure with Pt skin reduces noble metal requirements while providing considerable catalytic performance. If the two structural features above can be combined into one morphology, the combined advantages might further promote electrocatalytic reactions [58,59]. Correspondingly, Guo et al. designed a novel synthetic route to obtain one-dimensional Pt3Fe nanowires with high-index facets and formed Pt-skin structures approximately two atomic layers thick on their surface [60]. The MA and SA of Pt3Fe nanowires are 17 and 13 times higher than those of commercial Pt/C, respectively. Moreover, excellent electrochemical stability is obtained. DFT simulations have revealed that the coordination and stress effects exhibited by this unique structure optimize the adsorption energy of oxygen-containing species, thus greatly increasing the catalytic performance.
The support material does not directly affect the overall performance; however, investigating it is still vital. Durability is a decisive factor in determining whether an electrocatalyst can be employed. Additionally, conductivity, which affects output power density, is closely related to the support type. Hence, designing support materials to meet the challenges of high potential is also an important topic to address.

2.3. Modulation of Catalyst Support

In addition to optimizing catalyst structure and composition, modulating support materials is also important for improving catalytic performance. Currently, the most commonly adopted catalyst–support combination in MEA is Pt-based nanoparticles loaded on a carbon support. The weak interaction between the metal particles and the carbon material is not strong enough for stabilization; therefore, metal particle migration and sintering are likely to occur, causing a remarkable decrease in long-term catalytic activity [33]. Moreover, carbon supports are vulnerable to oxidation and corrosion under harsh conditions, which can lead to metal particle detachment and agglomeration [61]. Introducing transition metal oxides as catalyst supports is a promising way. It can alleviate support corrosion and anchor the metal particles with strong metal–support interactions (SMSI) [62]. Liao et al. reported a novel high-performance catalyst (TiNiN@Pt), which shows only a slight loss of performance after 10 k dynamic potential cycles. Its structure also remained intact [63]. The excellent durability of the TiNiN@Pt catalyst is attributed to the good stability of the nitride and the synergy effect between Pt atoms and the stable TiN supports. In addition, Qiao et al. produced hybrid nanocatalysts by combining Pt nanoparticles with CoO nanorods on conductive carbon fiber paper substrates [64]. The electronic interaction between Pt nanoparticles and CoO nanorods and the direct growth of CoO nanorods on the carbon fiber paper substrate prevents the aggregation and separation of Pt nanoparticles, providing remarkably better performance than commercial Pt/C. Despite the excellent stability of metal oxides and the SMSI of metal particles, these materials have a drawback. Their limited electronic conductivity is insufficient to meet the targets for large-scale PEMFC applications.
Another way to enhance stability is by using highly graphitized nanocarbons as supports, such as carbon nanotubes, carbon nanocages, and graphene. Their oxidation and corrosion are less severe than those of traditional supports [65]. Li et al. prepared Pt/SH-CNT (carbon nanotube) catalysts with –SH groups on the surface of carbon nanotubes [66]. The strong interaction between Pt and S atoms causes a uniform distribution of Pt on the carbon nanotube surface while inhibiting Pt dissolution and agglomeration. Excellent electrochemical properties of the catalysts have been observed during performance evaluation. Despite the excellent electronic conductivity and stability of graphene as catalyst support, its pristine surface is not conducive to dispersing Pt nanoparticles uniformly, which implies the need for graphene surface functionalization to promote good Pt nanoparticles dispersion [65]. Ding et al. demonstrated the excellent stability of Pt nanoparticles on the surface of N-doped graphene by studying the structural and electronic properties of Pt nanoparticles on different kinds of graphene [67].
In order to address the disadvantages of carbon and noncarbon supports, researchers combined the two to form composite support. This approach may have great potential to improve catalyst durability [68]. Xiao et al. designed a TiO2-C composite support with dispersed TiO2 nanocrystals on the carbon surface and then prepared Pt/TiO2-C catalysts by photo-deposition method [69]. The TiO2 nanocrystals remarkably enhance the stability of carbon support under harsh conditions. Moreover, the SMSI between TiO2 nanocrystals and Pt nanoparticles accelerated the electron transfer process in the catalytic reaction. The Pt nanoparticles are firmly anchored, preventing migration and aggregation on the support, providing a remarkable improvement in the stability of Pt/TiO2-C catalysts. Similarly, Ye et al. used atomic layer deposition to deposit Ta2O5 nanoparticles onto carbon (Figure 3c). Pt nanoparticles were then deposited on this composite support. The support prevents Pt nanoparticles from detachment, migration, and agglomeration during PEMFC operations [50].
From the research works introduced above, it is fair to say that many highly active and durable catalysts for ORR have been developed. We summarized these catalyst properties (Table 1). Note that most of these results are based on half-cell tests on RDE. For practical applications, engineering and optimization of the CL structure are essential to obtaining high-performance MEA [70].

2.4. Progresses in Theoretical Simulation

Despite experimental explorations, theoretical simulations via DFT could help understand the catalytic mechanism and screen potential candidates. Norskov et al. first proposed to use free-energy changes in the adsorbed intermediate such as oxygen or hydroxyl, to describe the origin of ORR overpotential [71]. Additionally, the superiority of Pt in the volcano-shaped curve for ORR activity was explained by its moderate d-band center. Furthermore, they investigated the binary alloy system and explained the experimental observation of the excellent performance of Pt-Co/Ni/Fe from the DFT perspective [26]. Since then, researchers have extensively applied theoretical simulation as a powerful tool for explaining the origin of the ORR activity of electrocatalysts. It should be noted that, more importantly, in recent years, researchers have applied machine learning (ML) and artificial intelligence (AI) models to boost the high throughput screening for potential ideal material components and systems [72,73,74,75,76,77,78,79]. As surrogate models, as-trained ML models could significantly reduce the massive demand for computational resources by DFT calculations. For instance, Kang et al. trained a high-dimensional neural network potential based on a dataset of 13,877 DFT results for PtNi, PtCu, CuNi, and PtCuNi nanoparticles as a function of size and composition [80]. By combining Monte Carlo and molecular dynamics, the authors found a stable and active ternary nanoalloy electrocatalyst for ORR in acids: a 2.6 nm icosahedron containing 60% Pt and 40% Ni/Cu. Similarly, Rück et al. focused on Pt alloy nanoparticles with a core–shell structure [81]. They trained ML models as an alternative to the costly DFT calculations to predict the corresponding strains directly from geometrical structural features. The optimal size of Pt@Cu and Pt@Ni for ORR was modeled as 1.9 nm, while 2.8 nm was suitable for Pt@Au and Pt@Ag. More importantly, the authors demonstrated in the paper that the advantages of the ML model in terms of computational efficiency become superior as the size of the particles studied increases. Additionally, for the high entropy alloy system, which is considered a hot topic and more complex than binary/ternary alloys, such advantages have been proven by a series of works conducted by Rossmeisl and coworkers. They extensively built a workflow based on ML to reveal the ideal component in Ag-Ir-Pd-Pt-Ru and Ir-Pd-Pt-Rh-Ru systems finally. Such results could be impossible via traditional DFT simulation with astronomical computational costs. Therefore, ML might be a powerful novel tool in theoretically predicting potentially efficient catalysts that deserve attention.

3. CL Structure in MEA

3.1. Preparation of CL

The core component of PEMFC is the MEA, where the cathode CL is where ORR occurs. The catalyst ink consists of a catalyst, an ionomer, and a solvent, which are then coated onto the GDL or PEM to form the CL. The ideal CL requires a high surface area catalyst/ionomer interface that boosts the reaction, a uniform ionomer distribution, and a suitable pore structure to facilitate proton, O2, and H2O transport. The structure of the PEM/CL interface has an important influence on the charge transfer and mass transfer processes in the MEA [82]. CL preparation can be divided into two types (Figure 4): gas diffusion electrode (GDE) and catalyst-coated membrane (CCM), which are distinguished by their different substrates. For the former, the catalyst ink is sprayed onto the GDL, and the surfaces are bonded together by a hot-pressing process. When the interfacial bonding disconnects, especially at the PEM/CL interface, local charge transfer stops, causing insufficient proton supply to assist the chemical reaction and thereby reducing catalyst utilization efficiency. Therefore, CCM preparation using PEM as the CL substrate is becoming more common to reduce adhesion failure at the PEM/CL interface [83]. Here, the catalyst ink is sprayed onto both sides of the PEM to improve bonding, making it less likely to detach and thus improving catalyst utilization [84].
Squeegee coating, screen printing, spray coating, and sputtering are generally used for CL formation. Ensuring thin-layer coating uniformity or large-scale production is challenging for these methods. In contrast, the ultrasonic spraying method (Figure 5a) is a novel next-generation approach for MEA preparation. Thin and uniform coatings are achieved using ultrasonic spraying, improving catalyst utilization and reducing MEA costs [85]. Pollet et al. believe that ultrasonic spraying technology can create special spraying conditions that form highly dispersed Pt nanoparticles, reducing agglomeration and providing high catalytic performance for ORR [86]. In addition, Eroglu [87] and Su [88] also find that MEAs prepared using the ultrasonic spraying technique exhibit excellent electrochemical properties.

3.2. Regulation of Ink Composition

The CL structure is closely related to the catalyst ink, especially the catalyst and ionomer dispersions in the solvent. These determine the three-phase interface, the distribution of ionomers, and the pore structure in the CL. Therefore, building the ideal CL structure by adjusting the ink composition and recipe is necessary [91]. The ratio of ionomer to carbon carrier (so-called IC ratio) is the first thing that can be associated with the modulation of catalyst inks. The ionomer content in the catalyst ink can have a remarkable impact on gas permeability and proton conductivity. When the ionomer content is too low, proton conduction is blocked, degrading electrochemical performance. However, when the ionomer content is too high, the ionomer blocks pore structures and hinders gas and water transport, also degrading electrochemical performance [92]. Lee et al. demonstrated, through a series of electrochemical performance evaluations, that the best activity and durability can be obtained at a moderate IC ratio of 1, whether the ink is water- or organic solvent-based [93]. In contrast, Yu et al. found that, for situations where Pt loading on MEA is relatively low, the appropriate IC ratio should be 0.3. This value changes with the support material [94]. In a recent study, Alink et al. employed two catalysts with different Pt loadings to test the gas and proton conductivity in the MEA under different conditions. They found that the optimal ionomer content correlates with both the Pt weight ratio in the catalyst and the Pt mass loading in the CL [95]. Therefore, the optimal IC ratio in the CL still needs to be investigated more deeply as an engineering issue in different cases. Advanced characterization methods are required to demonstrate its effect on the formed CL structure. Finally, water, proton, and gas conduction need to be evaluated.
The solvent also influences MEA performance. The solvent affects the size, viscosity, curing rate, and other properties of the ionomer particles in the catalyst ink. These then affect the newly formed CL structure [96,97]. Most studies choose water or ethanol as solvents; however, these media may not disperse ionomers well [98,99,100,101]. Hence, some researchers used mixed solvents for dispersion. Kim et al. noted that ionomers (e.g., Nafion) tend to form large particle network structures when dispersed in aqueous/alcoholic mixed solvents but form cylindrical rods or coil-like small particles when dispersed in polar solvents (Figure 5b) [89]. The chemical properties of the solvent molecules are the main reason for the varying structures. Uchida found that ionomers are dispersed uniformly in solvents with dielectric constants above 10 but form colloids between 3 and 10. Ionomers precipitate in solvents with dielectric constants below three [102]. However, the prevailing view is that the solvent influences the state of the ionomer; the relationship is still not thoroughly understood due to a lack of reliable approaches for characterizing the microstructure of catalyst inks [103].
Finally, ionomer properties also influence MEA performance. Park et al. analyzed CLs prepared from various ionomers containing different solvents [104]. They measured the thickness and contact angles of the ionomer films, tested the polarization curves, and investigated the high-frequency resistance of different samples. They found that the propylene glycol solvent in the ionomer yields small particles, allowing well-interconnected ionomers. Proton conduction is enhanced, resulting in excellent MEA performance. In another recent study, Zhang et al. incorporated ionic covalent organic framework nanosheets into Nafion to modify the interaction between the ionomer and the catalyst (Figure 5c). The approach yields a unique three-phase interfacial structure suitable for mass transfer in the CL. The MEA prepared using the composite ionomer exhibits good proton and gas transport and weakens –SO3H group adsorption on the Pt surface. These enhancements improve the Pt utilization, and the MA and peak power density in the MEA sample reach 1.6 times that of conventional MEA [90].

3.3. Fine-Tuning of CL Structure

In addition to modulating the catalyst ink composition, designing the CL structure can also be investigated. The CL can be fine-tuned to enhance electron, proton, water, and gas transport, thus achieving improved MEA performance. As mentioned previously, the two common ways of preparing CL are GDE and CCM. The latter has now become more commonly adopted because CLs with much lower interfacial impedance than those prepared using GDE can be obtained [105]. However, the ordered electrode (Figure 6a) developed in recent years is regarded as the future direction. In the ordered electrode, the transfer direction for electrons, protons, oxygen, and water is perpendicular to the PEM, enabling the maximum catalyst exposure. Mass transfer resistance is also reduced [106,107,108,109]. Vliet et al. developed nanostructured thin-film electrocatalysts based on this approach. The approach provides a large ECSA while reducing the contact resistance between the catalyst and the support, providing high catalyst utilization efficiency and low mass transfer resistance [110]. In addition, ordered CL structures based on carbon nanotube [111], carbon nanofiber [112], and polyaniline [113] arrays have also been investigated. These well-defined structures have proven effective in decreasing the dependence on precious metals while enhancing MEA performance.
CL design also involves adding some pore-forming materials and constructing a multi-stage CL. Zhao et al. reported uniformly dispersed catalysts with large surface pores and low gas diffusion resistance on the MEA surface by adding NH4HCO3 [116]. Chen et al. prepared CL-containing electrocatalysts with different Pt mass fractions and different ionomer contents to improve MEA catalyst utilization [117]. After electrochemical characterization, they found that intentionally placing more catalysts and ionomers near the PEM provides smaller ohmic and charge transfer resistances, causing excellent performance. Similarly, a bilayer CL structure with a gradient distribution of Pt content and pore structure was proposed by Ye et al. [118]. By changing the Pt and C contents, fabricating the CL near the PEM side, which has a high Pt mass loading, to promote the ORR reaction is possible. The CL near the GDL side should have a large pore size that can facilitate gas and water transport. This “allocation on demand” measure is able to decrease the amount of Pt effectively. Such an idea could be called the design of gradient structures. Investigations aimed at ionomer and catalyst particles are also conducted. Holdcroft et al. developed and evaluated GDEs with ionomer (Nafion) gradient distribution [119]. The results show that cathode performance is improved by placing more ionomers near the CL/PEM interface at a medium- to high-current density area. On the contrary, when the ionomer content gradient was reversed, the performance of the cell decreased remarkably. Unhindered proton transport near the CL/PEM interface must be ensured and high ionic conductivity in this region is required. Zhang et al. designed a highly durable MEA by combining particle size gradients in the CL structure with Pt loading (Figure 6b) [114]. By using mathematical models, they evaluated several gradient CL structures of Pt/C catalysts based on ECSA evolution and Pt mass during the cycling process. They found that Pt utilization and MEA performance could be improved by optimizing the particle size and the Pt loading gradient distribution. Therefore, a reasonable gradient design of the CL structure improves MEA catalytic activity. In addition, ionomer distribution in the CL can be improved by using functionalized carbon supports. Ott et al. obtained N-doped carbon supports by pre-oxidation and high-temperature ammonolysis processes (Figure 6c), which allowed uniform distribution of ionomers on their surfaces and reduced mass transfer resistance [115]. The pore structure of the porous carbon support was also optimized so that Pt nanoparticles were uniformly dispersed on the outer surface of the support. Excellent MEA performance was obtained. We summarized the performance of MEAs from related publications in Table 2 to reflect the influence of CL design on MEA performance.

4. Summary and Outlook

Recently, remarkable progress has been made in reducing PEMFC cost by improving the intrinsic activity of catalysts and optimizing the structure within the CL. In this review, we summarized and discussed the developments in catalysts and CL structures in recent years. For catalysts, alloying is the most common strategy to increase activity. However, catalyst stability is crucial; thus, studies focus on optimizing catalyst morphology. In addition to these, corrosion of the carbon support can occur under harsh conditions, leading to agglomeration and the shedding of catalyst nanoparticles. These changes then lead to activity loss and device degradation. Thus, further work on catalyst supports is essential to break through this bottleneck. Balancing catalytic activity, conductivity, and durability requires novel electrocatalyst designs. Furthermore, large-scale CL studies are also needed for PEMFC to be employed in practical applications. CL is not only the place where an electrochemical reaction takes place but also involves the transport of oxygen, electrons, protons, and water. Therefore, its influence on PEMFC performance is wide-ranging. Researchers proposed adopting modulation of the catalyst ink composition or designing the CL structure to obtain a reasonable three-phase interface that enhances Pt utilization and reduces Pt loading without sacrificing the performance of MEA.
Despite the current progress, there is still a long way to go in the commercialization process of PEMFCs. Novel catalysts and CLs designed for scientific research often encounter problems in practical applications. Thus, current catalysts and MEA structures used in most commercial power reactors are still conventional. For this reason, we propose several directions that might be able to bring a real change in the current situation.
  • In many cases, high activity or Pt utilization efficiency for the catalyst on RDE does not translate to good catalytic performance in MEAs [59,70,120]. This difference is due to different working conditions. Numerous studies continue to focus on RDE electrocatalyst performance due to their low cost. We suggest that electrocatalysts be evaluated and optimized in MEA and single cells. The findings would then be more convincing and valuable for commercial development. Low-Pt loading in MEA needs to be conducted for both catalyst and CL structures. The most pressing issue is MEA durability.
  • In developing low-Pt loading MEAs, a large ECSA is needed to address mass transfer issues in current catalysts. Therefore, catalyst particle size needs to be reduced [32,49,52]. However, this reduction causes migration and ripening, leading to reduced durability, and durability is sometimes more important than activity [121,122,123]. Currently, many studies still focus on obtaining enhanced catalytic activity. Stability is rarely the goal. ADTs are often performed after the best sample is chosen at the end. We suggest that researchers consider stability as equally important as activity in their research and development.
  • Nonprecious metal electrocatalysts have recently been proven as an alternative to traditional Pt-based electrocatalysts, even in MEA [124,125,126,127]. However, the stability of nonprecious metal electrocatalysts could not meet the demands of practical applications. Additionally, activity and stability are still significant challenges, especially in larger-scale application circumstances, such as in actual stacks [128,129]. Like the Pt-based material system that has been discussed, most studies of nonprecious metal-based electrocatalysts are still on RDE [72]. However, recent reports have indicated that combining nonprecious and precious metal catalysts may benefit from a synergistic effect [130,131,132]. This integrated field might be promising for enhancing ORR catalytic performance while keeping costs down.
  • The problem of mass transport severely affects the performance of low-Pt loading MEA [133,134,135,136]. Therefore, for CL, a good structure suitable for mass transfer needs to be developed. However, a cheap and effective solution has not yet emerged. Hence, in-depth studies focusing on suitable pore structures and homogeneous three-phase interfaces are needed.
  • It should also be noted that the reduction in noble metal loadings is not only important. Considering the limited crustal reserves, it is very economical to recycle precious metals from obsolete and defunct MEAs [137,138]. Specific manufacturing routes of the MEAs may have used less Pt but would increase the difficulties in recycling. Most researchers have not yet realized that the cost should eventually be “life-long”. Hence, it is suggested that such prospects be concerned in the future.

Author Contributions

F.C. and R.D. contributed equally. Conceptualization, F.C. and R.D.; methodology, Z.R.; software, R.D.; validation, B.Z., W.D. and Z.M.; investigation, J.L. (Jia Li); data curation, J.L. (Jianguo Liu); writing—original draft preparation, F.C. and R.D.; writing—review and editing, X.W., Z.M. and X.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (52272039, 51972168, 51672124).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cullen, D.A.; Neyerlin, K.C.; Ahluwalia, R.K.; Mukundan, R.; More, K.L.; Borup, R.L.; Weber, A.Z.; Myers, D.J.; Kusoglu, A. New roads and challenges for fuel cells in heavy-duty transportation. Nat. Energy 2021, 6, 462–474. [Google Scholar] [CrossRef]
  2. Shi, F.; Chen, C.; Xu, Z.-L. Recent Advances on Electrospun Nanofiber Materials for Post-lithium Ion Batteries. Adv. Fiber Mater. 2021, 3, 275–301. [Google Scholar] [CrossRef]
  3. Li, D.; Long, X.; Wu, Y.; Hou, H.; Wang, X.; Ren, J.; Zhang, L.; Yang, D.; Xia, Y. Hierarchically Porous and Defective Carbon Fiber Cathode for Efficient Zn-Air Batteries and Microbial Fuel Cells. Adv. Fiber Mater. 2022, 4, 795–806. [Google Scholar] [CrossRef]
  4. Wang, Q.; Fang, Z.; Zhang, W.; Zhang, D. High-Efficiency g-C3N4 Based Photocatalysts for CO2 Reduction: Modification Methods. Adv. Fiber Mater. 2022, 4, 342–360. [Google Scholar] [CrossRef]
  5. Wang, X.X.; Swihart, M.T.; Wu, G. Achievements, challenges and perspectives on cathode catalysts in proton exchange membrane fuel cells for transportation. Nat. Catal. 2019, 2, 578–589. [Google Scholar] [CrossRef]
  6. Zaman, S.; Huang, L.; Douka, A.I.; Yang, H.; You, B.; Xia, B.Y. Oxygen Reduction Electrocatalysts toward Practical Fuel Cells: Progress and Perspectives. Angew. Chem. Int. Ed. 2021, 60, 17832–17852. [Google Scholar] [CrossRef]
  7. Suter, T.A.M.; Smith, K.; Hack, J.; Rasha, L.; Rana, Z.; Angel, G.M.A.; Shearing, P.R.; Miller, T.S.; Brett, D.J.L. Engineering Catalyst Layers for Next-Generation Polymer Electrolyte Fuel Cells: A Review of Design, Materials, and Methods. Adv. Energy Mater. 2021, 11, 2101025. [Google Scholar] [CrossRef]
  8. Kodama, K.; Nagai, T.; Kuwaki, A.; Jinnouchi, R.; Morimoto, Y. Challenges in applying highly active Pt-based nanostructured catalysts for oxygen reduction reactions to fuel cell vehicles. Nat. Nanotechnol. 2021, 16, 140–147. [Google Scholar] [CrossRef]
  9. Xiao, F.; Wang, Y.-C.; Wu, Z.-P.; Chen, G.; Yang, F.; Zhu, S.; Siddharth, K.; Kong, Z.; Lu, A.; Li, J.-C.; et al. Recent Advances in Electrocatalysts for Proton Exchange Membrane Fuel Cells and Alkaline Membrane Fuel Cells. Adv. Mater. 2021, 33, 2006292. [Google Scholar] [CrossRef]
  10. Ahn, C.-Y.; Park, J.E.; Kim, S.; Kim, O.-H.; Hwang, W.; Her, M.; Kang, S.Y.; Park, S.; Kwon, O.J.; Park, H.S.; et al. Differences in the Electrochemical Performance of Pt-Based Catalysts Used for Polymer Electrolyte Membrane Fuel Cells in Liquid Half- and Full-Cells. Chem. Rev. 2021, 121, 15075–15140. [Google Scholar] [CrossRef]
  11. Tang, M.; Zhang, S.; Chen, S. Pt Utilization in Proton Exchange Membrane Fuel Cells: Structure Impacting Factors and Mechanistic Insights. Chem. Soc. Rev. 2022, 51, 1529–1546. [Google Scholar] [CrossRef]
  12. Zhao, L.; Zhu, J.; Zheng, Y.; Xiao, M.; Gao, R.; Zhang, Z.; Wen, G.; Dou, H.; Deng, Y.P.; Yu, A.; et al. Materials Engineering toward Durable Electrocatalysts for Proton Exchange Membrane Fuel Cells. Adv. Energy Mater. 2021, 12, 2102665. [Google Scholar] [CrossRef]
  13. Debe, M.K. Electrocatalyst approaches and challenges for automotive fuel cells. Nature 2012, 486, 43–51. [Google Scholar] [CrossRef]
  14. Wang, F.; Zhang, Q.; Rui, Z.; Li, J.; Liu, J. High-Loading Pt–Co/C Catalyst with Enhanced Durability toward the Oxygen Reduction Reaction through Surface Au Modification. ACS Appl. Mater. Interfaces 2020, 12, 30381–30389. [Google Scholar] [CrossRef]
  15. Hua, K.; Rui, Z.; Zhang, T.; Cao, Y.; Shen, X.; Wang, S.; He, P.; Li, J.; Liu, J. Efficient synthesis of Pt3Co/NC alloy catalysts with enhanced durability and activity for the oxygen reduction reaction. Int. J. Hydrogen Energy 2022, 47, 13022–13029. [Google Scholar] [CrossRef]
  16. Lin, Z.; Sheng, Y.; Li, J.; Rui, Z.; Liu, Y.; Liu, J.; Zou, Z. Ternary heterogeneous Pt–Ni–Au nanowires with enhanced activity and stability for PEMFCs. Chem. Commun. 2020, 56, 4276–4279. [Google Scholar] [CrossRef] [PubMed]
  17. Cheng, Q.; Yang, S.; Fu, C.; Zou, L.; Zou, Z.; Jiang, Z.; Zhang, J.; Yang, H. High-loaded sub-6 nm Pt1Co1 intermetallic compounds with highly efficient performance expression in PEMFCs. Energy Environ. Sci. 2022, 15, 278–286. [Google Scholar] [CrossRef]
  18. Zhang, H.; Liu, J.; Li, X.; Duan, X.; Yuan, M.; Cao, F.; Sun, K.; Zhang, Y.; Wang, Y.; Gu, Z.; et al. A TiN@C core–shell support for improving Pt catalyst corrosion resistance. RSC Adv. 2022, 12, 25035–25040. [Google Scholar] [CrossRef]
  19. Li, J.; Song, Y.; Zhang, G.; Liu, H.; Wang, Y.; Sun, S.; Guo, X. Pyrolysis of Self-Assembled Iron Porphyrin on Carbon Black as Core/Shell Structured Electrocatalysts for Highly Efficient Oxygen Reduction in Both Alkaline and Acidic Medium. Adv. Funct. Mater. 2017, 27, 1604356. [Google Scholar] [CrossRef]
  20. Ehelebe, K.; Knöppel, J.; Bierling, M.; Mayerhöfer, B.; Böhm, T.; Kulyk, N.; Thiele, S.; Mayrhofer, K.J.J.; Cherevko, S. Platinum Dissolution in Realistic Fuel Cell Catalyst Layers. Angew. Chem. Int. Ed. 2021, 60, 8882–8888. [Google Scholar] [CrossRef]
  21. Zhu, F.; Luo, L.; Wu, A.; Wang, C.; Cheng, X.; Shen, S.; Ke, C.; Yang, H.; Zhang, J. Improving the High-Current-Density Performance of PEMFC through Much Enhanced Utilization of Platinum Electrocatalysts on Carbon. ACS Appl. Mater. Interfaces 2020, 12, 26076–26083. [Google Scholar] [CrossRef]
  22. Weber, A.Z.; Kusoglu, A. Unexplained transport resistances for low-loaded fuel-cell catalyst layers. J. Mater. Chem. A 2014, 2, 17207–17211. [Google Scholar] [CrossRef]
  23. Lopez-Haro, M.; Guétaz, L.; Printemps, T.; Morin, A.; Escribano, S.; Jouneau, P.H.; Bayle-Guillemaud, P.; Chandezon, F.; Gebel, G. Three-dimensional analysis of Nafion layers in fuel cell electrodes. Nat. Commun. 2014, 5, 5229. [Google Scholar]
  24. Ferreira, P.J.; la O’, G.J.; Shao-Horn, Y.; Morgan, D.; Makharia, R.; Kocha, S.; Gasteiger, H.A. Instability of Pt/C Electrocatalysts in Proton Exchange Membrane Fuel Cells. J. Electrochem. Soc. 2005, 152, A2256. [Google Scholar]
  25. Bing, Y.; Liu, H.; Zhang, L.; Ghosh, D.; Zhang, J. Nanostructured Pt-alloy electrocatalysts for PEM fuel cell oxygen reduction reaction. Chem. Soc. Rev. 2010, 39, 2184–2202. [Google Scholar] [CrossRef] [Green Version]
  26. Greeley, J.; Stephens, I.E.L.; Bondarenko, A.S.; Johansson, T.P.; Hansen, H.A.; Jaramillo, T.F.; Rossmeisl, J.; Chorkendorff, I.; Nørskov, J.K. Alloys of platinum and early transition metals as oxygen reduction electrocatalysts. Nat. Chem. 2009, 1, 552–556. [Google Scholar] [CrossRef] [PubMed]
  27. Abdel-Hady, E.E.; Shaban, M.; Abdel-Hamed, M.O.; Gamal, A.; Yehia, H.; Ahmed, A.M. Synthesis and Characterization of NiCoPt/CNFs Nanoparticles as an Effective Electrocatalyst for Energy Applications. Nanomaterials 2022, 12, 492. [Google Scholar]
  28. Stamenkovic, V.R.; Mun, B.S.; Arenz, M.; Mayrhofer, K.J.J.; Lucas, C.A.; Wang, G.; Ross, P.N.; Markovic, N.M. Trends in electrocatalysis on extended and nanoscale Pt-bimetallic alloy surfaces. Nat. Mater. 2007, 6, 241–247. [Google Scholar] [CrossRef] [PubMed]
  29. Bu, L.; Guo, S.; Zhang, X.; Shen, X.; Su, D.; Lu, G.; Zhu, X.; Yao, J.; Guo, J.; Huang, X. Surface engineering of hierarchical platinum-cobalt nanowires for efficient electrocatalysis. Nat. Commun. 2016, 7, 11850. [Google Scholar] [CrossRef] [Green Version]
  30. Tian, X.; Zhao, X.; Su, Y.-Q.; Wang, L.; Wang, H.; Dang, D.; Chi, B.; Liu, H.; Hensen, E.J.M.; Lou, X.W.; et al. Engineering bunched Pt-Ni alloy nanocages for efficient oxygen reduction in practical fuel cells. Science 2019, 366, 850–856. [Google Scholar] [CrossRef]
  31. Wang, X.X.; Sokolowski, J.; Liu, H.; Wu, G. Pt alloy oxygen-reduction electrocatalysts: Synthesis, structure, and property. Chin. J. Catal. 2020, 41, 739–755. [Google Scholar] [CrossRef]
  32. Gamler, J.T.L.; Ashberry, H.M.; Skrabalak, S.E.; Koczkur, K.M. Random Alloyed versus Intermetallic Nanoparticles: A Comparison of Electrocatalytic Performance. Adv. Mater. 2018, 30, 1801563. [Google Scholar] [CrossRef] [PubMed]
  33. Hodnik, N.; Jeyabharathi, C.; Meier, J.C.; Kostka, A.; Phani, K.L.; Rečnik, A.; Bele, M.; Hočevar, S.; Gaberšček, M.; Mayrhofer, K.J.J. Effect of ordering of PtCu3 nanoparticle structure on the activity and stability for the oxygen reduction reaction. Phys. Chem. Chem. Phys. 2014, 16, 13610–13615. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, D.; Xin, H.L.; Hovden, R.; Wang, H.; Yu, Y.; Muller, D.A.; DiSalvo, F.J.; Abruña, H.D. Structurally ordered intermetallic platinum–cobalt core–shell nanoparticles with enhanced activity and stability as oxygen reduction electrocatalysts. Nat. Mater. 2013, 12, 81–87. [Google Scholar] [CrossRef]
  35. Yang, C.-L.; Wang, L.-N.; Yin, P.; Liu, J.; Chen, M.-X.; Yan, Q.-Q.; Wang, Z.-S.; Xu, S.-L.; Chu, S.-Q.; Cui, C.; et al. Sulfur-anchoring synthesis of platinum intermetallic nanoparticle catalysts for fuel cells. Science 2021, 374, 459–464. [Google Scholar] [CrossRef]
  36. Song, T.-W.; Xu, C.; Sheng, Z.-T.; Yan, H.-K.; Tong, L.; Liu, J.; Zeng, W.-J.; Zuo, L.-J.; Yin, P.; Zuo, M.; et al. Small molecule-assisted synthesis of carbon supported platinum intermetallic fuel cell catalysts. Nat. Commun. 2022, 13, 6521. [Google Scholar] [CrossRef]
  37. Liang, J.; Li, N.; Zhao, Z.; Ma, L.; Wang, X.; Li, S.; Liu, X.; Wang, T.; Du, Y.; Lu, G.; et al. Tungsten-Doped L10-PtCo Ultrasmall Nanoparticles as a High-Performance Fuel Cell Cathode. Angew. Chem. Int. Ed. 2019, 58, 15471–15477. [Google Scholar] [CrossRef]
  38. Duan, X.; Cao, F.; Ding, R.; Li, X.; Li, Q.; Aisha, R.; Zhang, S.; Hua, K.; Rui, Z.; Wu, Y.; et al. Cobalt-doping stabilized active and durable sub-2 nm Pt nanoclusters for low-Pt-loading PEMFC cathode. Adv. Energy Mater. 2022, 12, 2103144. [Google Scholar] [CrossRef]
  39. Xia, Z.; Guo, S. Strain engineering of metal-based nanomaterials for energy electrocatalysis. Chem. Soc. Rev. 2019, 48, 3265–3278. [Google Scholar] [CrossRef]
  40. Zhu, E.; Wu, M.; Xu, H.; Peng, B.; Liu, Z.; Huang, Y.; Li, Y. Stability of Platinum-Group-Metal-Based Electrocatalysts in Proton Exchange Membrane Fuel Cells. Adv. Funct. Mater. 2022, 32, 2203883. [Google Scholar] [CrossRef]
  41. Chen, S.; Li, M.; Gao, M.; Jin, J.; van Spronsen, M.A.; Salmeron, M.B.; Yang, P. High-Performance Pt-Co Nanoframes for Fuel-Cell Electrocatalysis. Nano Lett. 2020, 20, 1974–1979. [Google Scholar] [CrossRef]
  42. Chen, Y.; Cheng, T.; Goddard Iii, W.A. Atomistic Explanation of the Dramatically Improved Oxygen Reduction Reaction of Jagged Platinum Nanowires, 50 Times Better than Pt. J. Am. Chem. Soc. 2020, 142, 8625–8632. [Google Scholar] [CrossRef]
  43. Luo, S.; Chen, W.; Cheng, Y.; Song, X.; Wu, Q.; Li, L.; Wu, X.; Wu, T.; Li, M.; Yang, Q.; et al. Trimetallic Synergy in Intermetallic PtSnBi Nanoplates Boosts Formic Acid Oxidation. Adv. Mater. 2019, 31, 1903683. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, C.; Kang, Y.; Huo, Z.; Zhu, Z.; Huang, W.; Xin, H.L.; Snyder, J.D.; Li, D.; Herron, J.A.; Mavrikakis, M.; et al. Highly Crystalline Multimetallic Nanoframes with Three-Dimensional Electrocatalytic Surfaces. Science 2014, 343, 1339–1343. [Google Scholar] [CrossRef] [PubMed]
  45. Gan, L.; Heggen, M.; O’Malley, R.; Theobald, B.; Strasser, P. Understanding and Controlling Nanoporosity Formation for Improving the Stability of Bimetallic Fuel Cell Catalysts. Nano Lett. 2013, 13, 1131–1138. [Google Scholar] [CrossRef]
  46. Cui, C.; Gan, L.; Heggen, M.; Rudi, S.; Strasser, P. Compositional segregation in shaped Pt alloy nanoparticles and their structural behaviour during electrocatalysis. Nat. Mater. 2013, 12, 765–771. [Google Scholar] [CrossRef]
  47. Huang, L.; Zaman, S.; Tian, X.; Wang, Z.; Fang, W.; Xia, B.Y. Advanced Platinum-Based Oxygen Reduction Electrocatalysts for Fuel Cells. Acc. Chem. Res. 2021, 54, 311–322. [Google Scholar] [CrossRef]
  48. Gao, L.; Li, X.; Yao, Z.; Bai, H.; Lu, Y.; Ma, C.; Lu, S.; Peng, Z.; Yang, J.; Pan, A.; et al. Unconventional p–d Hybridization Interaction in PtGa Ultrathin Nanowires Boosts Oxygen Reduction Electrocatalysis. J. Am. Chem. Soc. 2019, 141, 18083–18090. [Google Scholar] [CrossRef] [PubMed]
  49. Zhu, S.; Wang, X.; Luo, E.; Yang, L.; Chu, Y.; Gao, L.; Jin, Z.; Liu, C.; Ge, J.; Xing, W. Stabilized Pt Cluster-Based Catalysts Used as Low-Loading Cathode in Proton-Exchange Membrane Fuel Cells. ACS Energy Lett. 2020, 5, 3021–3028. [Google Scholar]
  50. Cheng, H.; Cao, Z.; Chen, Z.; Zhao, M.; Xie, M.; Lyu, Z.; Zhu, Z.; Chi, M.; Xia, Y. Catalytic system based on sub-2 nm Pt particles and its extraordinary activity and durability for oxygen reduction. Nano Lett. 2019, 19, 4997–5002. [Google Scholar] [CrossRef]
  51. Zhao, Z.; Liu, Z.; Zhang, A.; Yan, X.; Xue, W.; Peng, B.; Xin, H.L.; Pan, X.; Duan, X.; Huang, Y. Graphene-nanopocket-encaged PtCo nanocatalysts for highly durable fuel cell operation under demanding ultralow-Pt-loading conditions. Nat. Nanotechnol. 2022, 17, 968–975. [Google Scholar] [PubMed]
  52. Sánchez-Sánchez, C.M.; Solla-Gullón, J.; Vidal-Iglesias, F.J.; Aldaz, A.; Montiel, V.; Herrero, E. Imaging Structure Sensitive Catalysis on Different Shape-Controlled Platinum Nanoparticles. J. Am. Chem. Soc. 2010, 132, 5622–5624. [Google Scholar] [CrossRef] [PubMed]
  53. Marković, N.M.; Ross, P.N. Surface science studies of model fuel cell electrocatalysts. Surf. Sci. Rep. 2002, 45, 117–229. [Google Scholar] [CrossRef]
  54. Choi, S.-I.; Xie, S.; Shao, M.; Odell, J.H.; Lu, N.; Peng, H.-C.; Protsailo, L.; Guerrero, S.; Park, J.; Xia, X.; et al. Synthesis and Characterization of 9 nm Pt–Ni Octahedra with a Record High Activity of 3.3 A/mgPt for the Oxygen Reduction Reaction. Nano Lett. 2013, 13, 3420–3425. [Google Scholar] [CrossRef]
  55. Huang, X.; Zhao, Z.; Chen, Y.; Zhu, E.; Li, M.; Duan, X.; Huang, Y. A rational design of carbon-supported dispersive Pt-based octahedra as efficient oxygen reduction reaction catalysts. Energy Environ. Sci. 2014, 7, 2957–2962. [Google Scholar] [CrossRef]
  56. Zhao, X.; Chen, S.; Fang, Z.; Ding, J.; Sang, W.; Wang, Y.; Zhao, J.; Peng, Z.; Zeng, J. Octahedral Pd@Pt1.8Ni Core–Shell Nanocrystals with Ultrathin PtNi Alloy Shells as Active Catalysts for Oxygen Reduction Reaction. J. Am. Chem. Soc. 2015, 137, 2804–2807. [Google Scholar] [CrossRef]
  57. Stamenkovic, V.R.; Fowler, B.; Mun, B.S.; Wang, G.; Ross, P.N.; Lucas, C.A.; Marković, N.M. Improved Oxygen Reduction Activity on Pt3Ni(111) via Increased Surface Site Availability. Science 2007, 315, 493–497. [Google Scholar] [CrossRef] [Green Version]
  58. Shao, M.; Chang, Q.; Dodelet, J.-P.; Chenitz, R. Recent Advances in Electrocatalysts for Oxygen Reduction Reaction. Chem. Rev. 2016, 116, 3594–3657. [Google Scholar] [CrossRef] [Green Version]
  59. Luo, M.; Sun, Y.; Zhang, X.; Qin, Y.; Li, M.; Li, Y.; Li, C.; Yang, Y.; Wang, L.; Gao, P.; et al. Stable High-Index Faceted Pt Skin on Zigzag-Like PtFe Nanowires Enhances Oxygen Reduction Catalysis. Adv. Mater. 2018, 30, 1705515. [Google Scholar] [CrossRef]
  60. Chattot, R.; Le Bacq, O.; Beermann, V.; Kühl, S.; Herranz, J.; Henning, S.; Kühn, L.; Asset, T.; Guétaz, L.; Renou, G.; et al. Surface distortion as a unifying concept and descriptor in oxygen reduction reaction electrocatalysis. Nat. Mater. 2018, 17, 827–833. [Google Scholar] [CrossRef]
  61. Cao, F.; Zhang, H.; Duan, X.; Li, X.; Ding, R.; Hua, K.; Rui, Z.; Wu, Y.; Yuan, M.; Wang, J.; et al. Coating Porous TiO2 Films on Carbon Nanotubes to Enhance the Durability of Ultrafine PtCo/CNT Nanocatalysts for the Oxygen Reduction Reaction. ACS Appl. Mater. Interfaces 2022, 14, 51975–51982. [Google Scholar] [CrossRef] [PubMed]
  62. Tian, X.; Luo, J.; Nan, H.; Zou, H.; Chen, R.; Shu, T.; Li, X.; Li, Y.; Song, H.; Liao, S.; et al. Transition Metal Nitride Coated with Atomic Layers of Pt as a Low-Cost, Highly Stable Electrocatalyst for the Oxygen Reduction Reaction. J. Am. Chem. Soc. 2016, 138, 1575–1583. [Google Scholar] [CrossRef] [PubMed]
  63. Meng, C.; Ling, T.; Ma, T.-Y.; Wang, H.; Hu, Z.; Zhou, Y.; Mao, J.; Du, X.-W.; Jaroniec, M.; Qiao, S.-Z. Atomically and Electronically Coupled Pt and CoO Hybrid Nanocatalysts for Enhanced Electrocatalytic Performance. Adv. Mater. 2017, 29, 1604607. [Google Scholar] [CrossRef] [PubMed]
  64. Singh, K.; Tetteh, E.B.; Lee, H.-Y.; Kang, T.-H.; Yu, J.-S. Tailor-Made Pt Catalysts with Improved Oxygen Reduction Reaction Stability/Durability. ACS Catal. 2019, 9, 8622–8645. [Google Scholar] [CrossRef]
  65. Chen, S.; Wei, Z.; Guo, L.; Ding, W.; Dong, L.; Shen, P.; Qi, X.; Li, L. Enhanced dispersion and durability of Pt nanoparticles on a thiolated CNT support. Chem. Commun. 2011, 47, 10984–10986. [Google Scholar] [CrossRef] [PubMed]
  66. Tian, Y.; Liu, Y.-j.; Zhao, J.-x.; Ding, Y.-h. High stability and superior catalytic reactivity of nitrogen-doped graphene supporting Pt nanoparticles as a catalyst for the oxygen reduction reaction: A density functional theory study. RSC Adv. 2015, 5, 34070–34077. [Google Scholar] [CrossRef]
  67. Zhang, J.; Yuan, Y.; Gao, L.; Zeng, G.; Li, M.; Huang, H. Stabilizing Pt-Based Electrocatalysts for Oxygen Reduction Reaction: Fundamental Understanding and Design Strategies. Adv. Mater. 2021, 33, 2006494. [Google Scholar] [CrossRef]
  68. Wang, J.; Xu, M.; Zhao, J.; Fang, H.; Huang, Q.; Xiao, W.; Li, T.; Wang, D. Anchoring Ultrafine Pt Electrocatalysts on TiO2-C via Photochemical Strategy to Enhance the Stability and Efficiency for Oxygen Reduction Reaction. Appl. Catal. B 2018, 237, 228–236. [Google Scholar] [CrossRef]
  69. Song, Z.; Banis, M.N.; Zhang, L.; Wang, B.; Yang, L.; Banham, D.; Zhao, Y.; Liang, J.; Zheng, M.; Li, R.; et al. Origin of achieving the enhanced activity and stability of Pt electrocatalysts with strong metal-support interactions via atomic layer deposition. Nano Energy 2018, 53, 716–725. [Google Scholar] [CrossRef]
  70. Fan, J.; Chen, M.; Zhao, Z.; Zhang, Z.; Ye, S.; Xu, S.; Wang, H.; Li, H. Bridging the gap between highly active oxygen reduction reaction catalysts and effective catalyst layers for proton exchange membrane fuel cells. Nat. Energy 2021, 6, 475–486. [Google Scholar]
  71. Nørskov, J.K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J.R.; Bligaard, T.; Jónsson, H. Origin of the Overpotential for Oxygen Reduction at a Fuel-Cell Cathode. J. Phys. Chem. B 2004, 108, 17886–17892. [Google Scholar] [CrossRef]
  72. Ding, R.; Wang, R.; Ding, Y.; Yin, W.; Liu, Y.; Li, J.; Liu, J. Designing AI-Aided Analysis and Prediction Models for Nonprecious Metal Electrocatalyst-Based Proton-Exchange Membrane Fuel Cells. Angew. Chem. Int. Ed. 2020, 59, 19175–19183. [Google Scholar] [CrossRef]
  73. Ding, R.; Ding, Y.; Zhang, H.; Wang, R.; Xu, Z.; Liu, Y.; Yin, W.; Wang, J.; Li, J.; Liu, J. Applying machine learning to boost the development of high-performance membrane electrode assembly for proton exchange membrane fuel cells. J. Mater. Chem. A 2021, 9, 6841–6850. [Google Scholar] [CrossRef]
  74. Ding, R.; Yin, W.; Cheng, G.; Chen, Y.; Wang, J.; Wang, R.; Rui, Z.; Li, J.; Liu, J. Boosting the optimization of membrane electrode assembly in proton exchange membrane fuel cells guided by explainable artificial intelligence. Energy AI 2021, 5, 100098. [Google Scholar] [CrossRef]
  75. Ding, R.; Chen, Y.; Chen, P.; Wang, R.; Wang, J.; Ding, Y.; Yin, W.; Liu, Y.; Li, J.; Liu, J. Machine Learning-Guided Discovery of Underlying Decisive Factors and New Mechanisms for the Design of Nonprecious Metal Electrocatalysts. ACS Catal. 2021, 11, 9798–9808. [Google Scholar] [CrossRef]
  76. Wang, J.; Ding, R.; Cao, F.; Li, J.; Dong, H.; Shi, T.; Xing, L.; Liu, J. Comparison of state-of-the-art machine learning algorithms and data-driven optimization methods for mitigating nitrogen crossover in PEM fuel cells. Chem. Eng. J. 2022, 442, 136064. [Google Scholar] [CrossRef]
  77. Ding, R.; Zhang, S.; Chen, Y.; Rui, Z.; Hua, K.; Wu, Y.; Li, X.; Duan, X.; Wang, X.; Li, J.; et al. Application of Machine Learning in Optimizing Proton Exchange Membrane Fuel Cells: A Review. Energy AI 2022, 9, 100170. [Google Scholar] [CrossRef]
  78. Ding, R.; Chen, Y.; Rui, Z.; Hua, K.; Wu, Y.; Li, X.; Duan, X.; Wang, X.; Li, J.; Liu, J. Guiding the Optimization of Membrane Electrode Assembly in a Proton Exchange Membrane Water Electrolyzer by Machine Learning Modeling and Black-Box Interpretation. ACS Sustain. Chem. Eng. 2022, 10, 4561–4578. [Google Scholar] [CrossRef]
  79. Ding, R.; Ma, M.; Chen, Y.; Wang, X.; Li, J.; Wang, G.; Liu, J. Inspecting design rules of metal-nitrogen-carbon catalysts for electrochemical CO2 reduction reaction: From a data science perspective. Nano Res. 2022. [Google Scholar] [CrossRef]
  80. Kang, J.; Noh, S.H.; Hwang, J.; Chun, H.; Kim, H.; Han, B. First-principles database driven computational neural network approach to the discovery of active ternary nanocatalysts for oxygen reduction reaction. Phys. Chem. Chem. Phys. 2018, 20, 24539–24544. [Google Scholar] [CrossRef]
  81. Rück, M.; Garlyyev, B.; Mayr, F.; Bandarenka, A.S.; Gagliardi, A. Oxygen Reduction Activities of Strained Platinum Core–Shell Electrocatalysts Predicted by Machine Learning. J. Phys. Chem. Lett. 2020, 11, 1773–1780. [Google Scholar] [CrossRef] [PubMed]
  82. Shangguan, Z.; Li, B.; Ming, P.; Zhang, C. Understanding the functions and modifications of interfaces in membrane electrode assemblies of proton exchange membrane fuel cells. J. Mater. Chem. A 2021, 9, 15111–15139. [Google Scholar] [CrossRef]
  83. Gong, X.; Ruan, M.; Song, P.; Li, H.; Cao, J.; Yao, P.; Xu, W. A cost-effective and highly efficient dissymmetric membrane electrode assembly designed for fuel cells. J. Power Sources 2021, 489, 229485. [Google Scholar] [CrossRef]
  84. Gomes Bezerra, C.A.; Deiner, L.J.; Tremiliosi-Filho, G. Unexpected Performance of Inkjet-Printed Membrane Electrode Assemblies for Proton Exchange Membrane Fuel Cells. Adv. Eng. Mater. 2019, 21, 1900703. [Google Scholar] [CrossRef]
  85. Pollet, B.G. The use of ultrasound for the fabrication of fuel cell materials. Int. J. Hydrogen Energy 2010, 35, 11986–12004. [Google Scholar] [CrossRef]
  86. Millington, B.; Whipple, V.; Pollet, B.G. A novel method for preparing proton exchange membrane fuel cell electrodes by the ultrasonic-spray technique. J. Power Sources 2011, 196, 8500–8508. [Google Scholar] [CrossRef]
  87. Devrim, Y.; Erkan, S.; Baç, N.; Eroglu, I. Improvement of PEMFC performance with Nafion/inorganic nanocomposite membrane electrode assembly prepared by ultrasonic coating technique. Int. J. Hydrogen Energy 2012, 37, 16748–16758. [Google Scholar] [CrossRef]
  88. Huang, T.-H.; Shen, H.-L.; Jao, T.-C.; Weng, F.-B.; Su, A. Ultra-low Pt loading for proton exchange membrane fuel cells by catalyst coating technique with ultrasonic spray coating machine. Int. J. Hydrogen Energy 2012, 37, 13872–13879. [Google Scholar] [CrossRef]
  89. Gong, Q.; Li, C.; Liu, Y.; Ilavsky, J.; Guo, F.; Cheng, X.; Xie, J. Effects of Ink Formulation on Construction of Catalyst Layers for High-Performance Polymer Electrolyte Membrane Fuel Cells. ACS Appl. Mater. Interfaces 2021, 13, 37004–37013. [Google Scholar] [CrossRef]
  90. Suzuki, T.; Tsushima, S.; Hirai, S. Effects of Nafion® ionomer and carbon particles on structure formation in a proton-exchange membrane fuel cell catalyst layer fabricated by the decal-transfer method. Int. J. Hydrogen Energy 2011, 36, 12361–12369. [Google Scholar] [CrossRef]
  91. Lee, M.R.; Lee, H.Y.; Yim, S.D.; Kim, C.S.; Shul, Y.G.; Kucernak, A.; Shin, D. Effects of Ionomer Carbon Ratio and Ionomer Dispersity on the Performance and Durability of MEAs. Fuel Cells 2018, 18, 129–136. [Google Scholar] [CrossRef]
  92. Yu, H.; Roller, J.M.; Mustain, W.E.; Maric, R. Influence of the ionomer/carbon ratio for low-Pt loading catalyst layer prepared by reactive spray deposition technology. J. Power Sources 2015, 283, 84–94. [Google Scholar] [CrossRef]
  93. Alink, R.; Singh, R.; Schneider, P.; Christmann, K.; Schall, J.; Keding, R.; Zamel, N. Full Parametric Study of the Influence of Ionomer Content, Catalyst Loading and Catalyst Type on Oxygen and Ion Transport in PEM Fuel Cell Catalyst Layers. Molecules 2020, 25, 1523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Lee, S.-J.; Yu, T.L.; Lin, H.-L.; Liu, W.-H.; Lai, C.-L. Solution properties of nafion in methanol/water mixture solvent. Polymer 2004, 45, 2853–2862. [Google Scholar] [CrossRef]
  95. Kim, Y.S.; Welch, C.F.; Mack, N.H.; Hjelm, R.P.; Orler, E.B.; Hawley, M.E.; Lee, K.S.; Yim, S.D.; Johnston, C.M. Highly durable fuel cell electrodes based on ionomers dispersed in glycerol. Phys. Chem. Chem. Phys. 2014, 16, 5927–5932. [Google Scholar] [CrossRef]
  96. Ding, R.; Yin, W.; Cheng, G.; Chen, Y.; Wang, J.; Wang, X.; Han, M.; Zhang, T.; Cao, Y.; Zhao, H.; et al. Effectively Increasing Pt Utilization Efficiency of the Membrane Electrode Assembly in Proton Exchange Membrane Fuel Cells through Multiparameter Optimization Guided by Machine Learning. ACS Appl. Mater. Interfaces 2022, 14, 8010–8024. [Google Scholar] [CrossRef]
  97. Orfanidi, A.; Rheinländer, P.J.; Schulte, N.; Gasteiger, H.A. Ink Solvent Dependence of the Ionomer Distribution in the Catalyst Layer of a PEMFC. J. Electrochem. Soc. 2018, 165, F1254. [Google Scholar] [CrossRef] [Green Version]
  98. Yang, F.; Xin, L.; Uzunoglu, A.; Qiu, Y.; Stanciu, L.; Ilavsky, J.; Li, W.; Xie, J. Investigation of the Interaction between Nafion Ionomer and Surface Functionalized Carbon Black Using Both Ultrasmall Angle X-ray Scattering and Cryo-TEM. ACS Appl. Mater. Interfaces 2017, 9, 6530–6538. [Google Scholar] [CrossRef] [Green Version]
  99. Ngo, T.T.; Yu, T.L.; Lin, H.-L. Influence of the composition of isopropyl alcohol/water mixture solvents in catalyst ink solutions on proton exchange membrane fuel cell performance. J. Power Sources 2013, 225, 293–303. [Google Scholar] [CrossRef]
  100. Welch, C.; Labouriau, A.; Hjelm, R.; Orler, B.; Johnston, C.; Kim, Y.S. Nafion in Dilute Solvent Systems: Dispersion or Solution? ACS Macro Lett. 2012, 1, 1403–1407. [Google Scholar] [CrossRef]
  101. Uchida, M.; Aoyama, Y.; Eda, N.; Ohta, A. New Preparation Method for Polymer-Electrolyte Fuel Cells. J. Electrochem. Soc. 1995, 142, 463. [Google Scholar] [CrossRef]
  102. Xu, F.; Zhang, H.; Ilavsky, J.; Stanciu, L.; Ho, D.; Justice, M.J.; Petrache, H.I.; Xie, J. Investigation of a Catalyst Ink Dispersion Using Both Ultra-Small-Angle X-ray Scattering and Cryogenic TEM. Langmuir 2010, 26, 19199–19208. [Google Scholar] [CrossRef] [PubMed]
  103. Park, J.-H.; Kim, B.-S.; Park, J.-S. Effect of ionomer dispersions on the performance of catalyst layers in proton exchange membrane fuel cells. Electrochim. Acta 2022, 424, 140680. [Google Scholar] [CrossRef]
  104. Zhang, Q.; Dong, S.; Shao, P.; Zhu, Y.; Mu, Z.; Sheng, D.; Zhang, T.; Jiang, X.; Shao, R.; Ren, Z.; et al. Covalent organic framework-based porous ionomers for high-performance fuel cells. Science 2022, 378, 181–186. [Google Scholar] [CrossRef] [PubMed]
  105. Sassin, M.B.; Garsany, Y.; Gould, B.D.; Swider-Lyons, K.E. Fabrication Method for Laboratory-Scale High-Performance Membrane Electrode Assemblies for Fuel Cells. Anal. Chem. 2017, 89, 511–518. [Google Scholar] [CrossRef] [PubMed]
  106. Middelman, E. Improved PEM fuel cell electrodes by controlled self-assembly. Fuel Cells Bull. 2002, 2002, 9–12. [Google Scholar] [CrossRef]
  107. Zeng, Y.; Zhang, H.; Wang, Z.; Jia, J.; Miao, S.; Song, W.; Xiao, Y.; Yu, H.; Shao, Z.; Yi, B. Nano-engineering of a 3D-ordered membrane electrode assembly with ultrathin Pt skin on open-walled PdCo nanotube arrays for fuel cells. J. Mater. Chem. A 2018, 6, 6521–6533. [Google Scholar] [CrossRef]
  108. Pan, S.; Qin, J.; Ning, F.; Bai, C.; Wen, Q.; Shen, M.; Li, Y.; Song, Y.; Chen, J.; Huang, Y.; et al. Well-Dispersed Nafion Array Prepared by the Freeze-Drying Method to Effectively Improve the Performance of Proton Exchange Membrane Fuel Cells. ACS Sustainable Chem. Eng. 2021, 9, 16770–16777. [Google Scholar] [CrossRef]
  109. Liu, H.; Qin, J.; Rockward, T.; Wu, J.; Li, J.; Li, G.; Mao, Q.; Lv, Y.; Wang, X.; Zhang, S.; et al. Photo-driven growth of a monolayer of platinum spherical-nanocrowns uniformly coated on a membrane toward fuel cell applications. J. Mater. Chem. A 2020, 8, 23284–23292. [Google Scholar] [CrossRef]
  110. van der Vliet, D.F.; Wang, C.; Tripkovic, D.; Strmcnik, D.; Zhang, X.F.; Debe, M.K.; Atanasoski, R.T.; Markovic, N.M.; Stamenkovic, V.R. Mesostructured thin films as electrocatalysts with tunable composition and surface morphology. Nat. Mater. 2012, 11, 1051–1058. [Google Scholar] [CrossRef]
  111. Cheon, J.Y.; Ahn, C.; You, D.J.; Pak, C.; Hur, S.H.; Kim, J.; Joo, S.H. Ordered mesoporous carbon–carbon nanotube nanocomposites as highly conductive and durable cathode catalyst supports for polymer electrolyte fuel cells. J. Mater. Chem. A 2013, 1, 1270–1283. [Google Scholar] [CrossRef] [Green Version]
  112. Li, M.; Xiong, Y.; Liu, X.; Han, C.; Zhang, Y.; Bo, X.; Guo, L. Iron and nitrogen co-doped carbon nanotube@hollow carbon fibers derived from plant biomass as efficient catalysts for the oxygen reduction reaction. J. Mater. Chem. A 2015, 3, 9658–9667. [Google Scholar] [CrossRef]
  113. Sun, R.; Xia, Z.; Shang, L.; Fu, X.; Li, H.; Wang, S.; Sun, G. Hierarchically ordered arrays with platinum coated PANI nanowires for highly efficient fuel cell electrodes. J. Mater. Chem. A 2017, 5, 15260–15265. [Google Scholar] [CrossRef]
  114. Zhao, J.; He, X.; Wang, L.; Tian, J.; Wan, C.; Jiang, C. Addition of NH4HCO3 as pore-former in membrane electrode assembly for PEMFC. Int. J. Hydrogen Energy 2007, 32, 380–384. [Google Scholar] [CrossRef]
  115. Chen, G.-Y.; Wang, C.; Lei, Y.-J.; Zhang, J.; Mao, Z.; Mao, Z.-Q.; Guo, J.-W.; Li, J.; Ouyang, M. Gradient design of Pt/C ratio and Nafion content in cathode catalyst layer of PEMFCs. Int. J. Hydrogen Energy 2017, 42, 29960–29965. [Google Scholar] [CrossRef]
  116. Ye, L.; Gao, Y.; Zhu, S.; Zheng, J.; Li, P.; Zheng, J.P. A Pt content and pore structure gradient distributed catalyst layer to improve the PEMFC performance. Int. J. Hydrogen Energy 2017, 42, 7241–7245. [Google Scholar] [CrossRef]
  117. Xie, Z.; Navessin, T.; Shi, K.; Chow, R.; Wang, Q.; Song, D.; Andreaus, B.; Eikerling, M.; Liu, Z.; Holdcroft, S. Functionally Graded Cathode Catalyst Layers for Polymer Electrolyte Fuel Cells: II. Experimental Study of the Effect of Nafion Distribution. J. Electrochem. Soc. 2005, 152, A1171. [Google Scholar] [CrossRef]
  118. Zheng, Z.; Yang, F.; Lin, C.; Zhu, F.; Shen, S.; Wei, G.; Zhang, J. Design of gradient cathode catalyst layer (CCL) structure for mitigating Pt degradation in proton exchange membrane fuel cells (PEMFCs) using mathematical method. J. Power Sources 2020, 451, 227729. [Google Scholar] [CrossRef]
  119. Ott, S.; Orfanidi, A.; Schmies, H.; Anke, B.; Nong, H.N.; Hübner, J.; Gernert, U.; Gliech, M.; Lerch, M.; Strasser, P. Ionomer distribution control in porous carbon-supported catalyst layers for high-power and low Pt-loaded proton exchange membrane fuel cells. Nat. Mater. 2020, 19, 77–85. [Google Scholar] [CrossRef]
  120. Sun, Y.; Polani, S.; Luo, F.; Ott, S.; Strasser, P.; Dionigi, F. Advancements in cathode catalyst and cathode layer design for proton exchange membrane fuel cells. Nat. Commun. 2021, 12, 5984. [Google Scholar] [CrossRef]
  121. Borup, R.L.; Kusoglu, A.; Neyerlin, K.C.; Mukundan, R.; Ahluwalia, R.K.; Cullen, D.A.; More, K.L.; Weber, A.Z.; Myers, D.J. Recent developments in catalyst-related PEM fuel cell durability. Curr. Opin. Electrochem. 2020, 21, 192–200. [Google Scholar] [CrossRef]
  122. Sandbeck, D.J.S.; Inaba, M.; Quinson, J.; Bucher, J.; Zana, A.; Arenz, M.; Cherevko, S. Particle Size Effect on Platinum Dissolution: Practical Considerations for Fuel Cells. ACS Appl. Mater. Interfaces 2020, 12, 25718–25727. [Google Scholar] [CrossRef] [PubMed]
  123. Watanabe, M.; Yano, H.; Uchida, H.; Tryk, D.A. Achievement of distinctively high durability at nanosized Pt catalysts supported on carbon black for fuel cell cathodes. J. Electroanal. Chem. 2018, 819, 359–364. [Google Scholar] [CrossRef]
  124. Ding, R.; Liu, Y.; Rui, Z.; Li, J.; Liu, J.; Zou, Z. Facile grafting strategy synthesis of single-atom electrocatalyst with enhanced ORR performance. Nano Res. 2020, 13, 1519–1526. [Google Scholar] [CrossRef]
  125. Li, J.; Zhu, X.; Wang, J.; Rui, Z.; Zhang, S.; Li, Y.; Ding, R.; He, W.; Liu, J.; Zou, Z. Iron-Containing Porphyrins Self-Assembled on ZnO Nanoparticles as Electrocatalytic Materials for Oxygen Reduction. ACS Appl. Nano Mater. 2020, 3, 742–751. [Google Scholar] [CrossRef]
  126. Wu, Y.; Yuan, M.; Li, X.; Ding, R.; Duan, X.; Li, J.; Wang, Y.; Li, X.; Zhang, Y.; Liu, J. Assistance of rearrangement of active sites in Fe/N/C catalyst for harvesting ultra-high power density PEMFCs. Appl. Catal., B 2022, 312, 121365. [Google Scholar] [CrossRef]
  127. Wei, H.; Su, X.; Liu, J.; Tian, J.; Wang, Z.; Sun, K.; Rui, Z.; Yang, W.; Zou, Z. A CeO2 modified phenylenediamine-based Fe/N/C with enhanced durability/stability as non-precious metal catalyst for oxygen reduction reaction. Electrochem. Commun. 2018, 88, 19–23. [Google Scholar] [CrossRef]
  128. Osmieri, L.; Meyer, Q. Recent advances in integrating platinum group metal-free catalysts in proton exchange membrane fuel cells. Curr. Opin. Electrochem. 2022, 31, 100847. [Google Scholar] [CrossRef]
  129. Kiani, M.; Tian, X.Q.; Zhang, W. Non-precious metal electrocatalysts design for oxygen reduction reaction in polymer electrolyte membrane fuel cells: Recent advances, challenges and future perspectives. Coord. Chem. Rev. 2021, 441, 213954. [Google Scholar] [CrossRef]
  130. Chong, L.; Wen, J.; Kubal, J.; Sen, F.G.; Zou, J.; Greeley, J.; Chan, M.; Barkholtz, H.; Ding, W.; Liu, D.-J. Ultralow-loading platinum-cobalt fuel cell catalysts derived from imidazolate frameworks. Science 2018, 362, 1276–1281. [Google Scholar] [CrossRef]
  131. Wang, X.X.; Hwang, S.; Pan, Y.-T.; Chen, K.; He, Y.; Karakalos, S.; Zhang, H.; Spendelow, J.S.; Su, D.; Wu, G. Ordered Pt3Co Intermetallic Nanoparticles Derived from Metal–Organic Frameworks for Oxygen Reduction. Nano Lett. 2018, 18, 4163–4171. [Google Scholar] [CrossRef] [PubMed]
  132. Huang, L.; Wei, M.; Qi, R.; Dong, C.-L.; Dang, D.; Yang, C.-C.; Xia, C.; Chen, C.; Zaman, S.; Li, F.-M.; et al. An integrated platinum-nanocarbon electrocatalyst for efficient oxygen reduction. Nat. Commun. 2022, 13, 6703. [Google Scholar] [CrossRef]
  133. Kongkanand, A.; Mathias, M.F. The Priority and Challenge of High-Power Performance of Low-Platinum Proton-Exchange Membrane Fuel Cells. J. Phys. Chem. Lett. 2016, 7, 1127–1137. [Google Scholar] [CrossRef] [PubMed]
  134. Baker, D.R.; Caulk, D.A.; Neyerlin, K.C.; Murphy, M.W. Measurement of Oxygen Transport Resistance in PEM Fuel Cells by Limiting Current Methods. J. Electrochem. Soc. 2009, 156, B991. [Google Scholar] [CrossRef]
  135. Cheng, X.; Shen, S.; Wei, G.; Wang, C.; Luo, L.; Zhang, J. Perspectives on Challenges and Achievements in Local Oxygen Transport of Low Pt Proton Exchange Membrane Fuel Cells. Adv. Mater. Technol. 2022, 7, 2200228. [Google Scholar] [CrossRef]
  136. Greszler, T.A.; Caulk, D.; Sinha, P. The Impact of Platinum Loading on Oxygen Transport Resistance. J. Electrochem. Soc. 2012, 159, F831–F840. [Google Scholar] [CrossRef]
  137. Duclos, L.; Lupsea, M.; Mandil, G.; Svecova, L.; Thivel, P.-X.; Laforest, V. Environmental assessment of proton exchange membrane fuel cell platinum catalyst recycling. J. Cleaner Prod. 2017, 142, 2618–2628. [Google Scholar] [CrossRef]
  138. Duclos, L.; Svecova, L.; Laforest, V.; Mandil, G.; Thivel, P.X. Process development and optimization for platinum recovery from PEM fuel cell catalyst. Hydrometallurgy 2016, 160, 79–89. [Google Scholar] [CrossRef]
Scheme 1. An overview of strategies for PEMFC cost reduction and performance improvement.
Scheme 1. An overview of strategies for PEMFC cost reduction and performance improvement.
Molecules 28 00773 sch001
Figure 1. (a) Composition structure and (b) operational principle of PEMFC. Reprinted with permission from Ref. [13]. Copyright 2012, Springer Nature.
Figure 1. (a) Composition structure and (b) operational principle of PEMFC. Reprinted with permission from Ref. [13]. Copyright 2012, Springer Nature.
Molecules 28 00773 g001
Figure 2. Catalyst performance enhancement through alloying, ordering, and doping: (a) Synthesis route and structure of PtNi alloy. Reprinted with permission from Ref. [30]. Copyright 2019, AAAS. (b) Diagram of the sulfur-anchored ordered alloy method. Reprinted with permission from Ref. [31]. Copyright 2021, AAAS. (c) Schematic of transition metal doping method. Reprinted with permission from Ref. [32]. Copyright 2022, John Wiley & Sons.
Figure 2. Catalyst performance enhancement through alloying, ordering, and doping: (a) Synthesis route and structure of PtNi alloy. Reprinted with permission from Ref. [30]. Copyright 2019, AAAS. (b) Diagram of the sulfur-anchored ordered alloy method. Reprinted with permission from Ref. [31]. Copyright 2021, AAAS. (c) Schematic of transition metal doping method. Reprinted with permission from Ref. [32]. Copyright 2022, John Wiley & Sons.
Molecules 28 00773 g002
Figure 3. Catalyst performance enhancement through structural optimization: (a) Diagram of the design of PtGa NWs. Reprinted with permission from Ref. [48]. Copyright 2019, ACS. (b) Diagram showing the protective effect of special support on ultrafine catalysts. Reprinted with permission from Ref. [49]. Copyright 2022, Springer Nature. (c) Schematic of the SMSI effect between Pt and support. Reprinted with permission from Ref. [50]. Copyright 2018, Elsevier.
Figure 3. Catalyst performance enhancement through structural optimization: (a) Diagram of the design of PtGa NWs. Reprinted with permission from Ref. [48]. Copyright 2019, ACS. (b) Diagram showing the protective effect of special support on ultrafine catalysts. Reprinted with permission from Ref. [49]. Copyright 2022, Springer Nature. (c) Schematic of the SMSI effect between Pt and support. Reprinted with permission from Ref. [50]. Copyright 2018, Elsevier.
Molecules 28 00773 g003
Figure 4. Preparation process of MEA: (a) CCM and (b) GDE methods. Reprinted with permission from Ref. [83]. Copyright 2021, Elsevier.
Figure 4. Preparation process of MEA: (a) CCM and (b) GDE methods. Reprinted with permission from Ref. [83]. Copyright 2021, Elsevier.
Molecules 28 00773 g004
Figure 5. Modulation of spraying materials and methods to improve MEA performance: (a) Diagram of the ultrasonic spraying method. Reprinted with permission from Ref. [85]. Copyright 2010, Elsevier. (b) Dispersion of Nafion in different solvents. Reprinted with permission from Ref. [89]. Copyright 2012, ACS. (c) Modulation of the ionomer to optimize the structure of the three-phase interface. Reprinted with permission from Ref. [90]. Copyright 2022, AAAS.
Figure 5. Modulation of spraying materials and methods to improve MEA performance: (a) Diagram of the ultrasonic spraying method. Reprinted with permission from Ref. [85]. Copyright 2010, Elsevier. (b) Dispersion of Nafion in different solvents. Reprinted with permission from Ref. [89]. Copyright 2012, ACS. (c) Modulation of the ionomer to optimize the structure of the three-phase interface. Reprinted with permission from Ref. [90]. Copyright 2022, AAAS.
Molecules 28 00773 g005
Figure 6. Optimization of CL structure to improve MEA performance: (a) Diagram of conventional MEA and ordered MEA. Reprinted with permission from Ref. [107]. Copyright 2018, RSC. (b) Design of Gradient CL Structure. Reprinted with permission from Ref. [114]. Copyright 2020, Elsevier. (c) Modified supports to regulate the distribution of ionomers. Reprinted with permission from Ref. [115]. Copyright 2020, Springer Nature.
Figure 6. Optimization of CL structure to improve MEA performance: (a) Diagram of conventional MEA and ordered MEA. Reprinted with permission from Ref. [107]. Copyright 2018, RSC. (b) Design of Gradient CL Structure. Reprinted with permission from Ref. [114]. Copyright 2020, Elsevier. (c) Modified supports to regulate the distribution of ionomers. Reprinted with permission from Ref. [115]. Copyright 2020, Springer Nature.
Molecules 28 00773 g006
Table 1. ORR performance of catalysts in the relevant literature.
Table 1. ORR performance of catalysts in the relevant literature.
CatalystSpecific Activity (mA cm−2)Mass Activity (A mgPt−1)Electrochemically Active Surface Area (m2 gPt−1)Accelerated Durability TestMass Activity RetentionElectrolyteReference
Au–
Pt–Co/C-0.015
0.5350.38672.20.6–1.0 V
30 k cycles
90.6%0.1 M HClO4[14]
Pt3Co/NC1.2360.38231.00.6–1.0 V
30 k cycles
104.5%0.1 M HClO4[15]
Pt–Ni–Au NWs2.590.65125.10.6–1.0 V
30 k cycles
82.3%0.1 M HClO4[16]
700-
Pt1Co1-IMC@Pt/C-2.5
1.110.5343.50.6–1.1 V
30 k cycles
76.6%0.1 M HClO4[17]
Pt/TiN@C 0.15569.21.2 V constant
for 400 h
95.9%0.1 M HClO4[18]
Ni4Co2Pt/CNFs
Ni6Pt/CNFs
NA35.7 A/g @ 799 mV for ethanol
37.4 A/g @ 799 mV for ethanol
NA0.8 V constant
chronoamperometry
57.4% (1000 s for methanol)
70.8% (1000 s for urea)
1 M
KOH
[27]
Pt3Co NWs/C7.123.7152.10.6–1.1 V
20 k cycles
91.9%0.1 M HClO4[29]
PtNi
BNCs/C
5.163.5268.20.6–1.1 V
50 k cycles
98.7%0.1 M HClO4[30]
PtCu33.84.18 0.1 M HClO4[31]
Co-doped
Pt
0.4980.579116.20.6–1.0 V
30 k cycles
91.0%0.1 M HClO4[32]
PtCo3.212.2570.1 0.1 M HClO4[37]
L10-W-PtCo/C3.602.2161.40.6–1.0 V
10 k cycles
86.4%0.1 M HClO4[38]
PtGa NWs/
C
3.281.8953.40.6–1.1 V
30 k cycles
84.2%0.1 M HClO4[48]
PtCo@Gnp1.621.1968.7 0.1 M HClO4[49]
Pt/N-ALDTa2O5/C 0.2870.30.6–1.0 V
10 k cycles
90%0.1 M HClO4[50]
Pt2.5Ni/C7.33.3450.65 V (5 s)
0.95 V (5 s)
5 k cycles
60%0.1 M HClO4[55]
PtNiCo/C3.882.3361.60.6–1.1 V
6 k cycles
43.3%0.1 M HClO4[56]
Pd@Pt1.8Ni 0.79178.010.6–1.05 V
6 k cycles
NA (MA retention)
90% (ECSA retention)
0.1 M HClO4[57]
Pt-skin
Pt3Fe z-NWs/C
4.342.1134.00.6–1.1 V
50 k cycles
75.4%0.1 M HClO4[60]
PtCo/TiO2/CNT0.6280.47675.80.6–1.0 V
30 k cycles
88.8%0.1 M HClO4[62]
TiNiN@Pt0.490.8355.40.6–1.05 V
10 k cycles
91.6%0.1 M HClO4[63]
Pt/TiO2-C 0.2181.70.6–1.0 V
10 k cycles
99.1%0.1 M HClO4[69]
Table 2. Properties of MEAs with different structures and components.
Table 2. Properties of MEAs with different structures and components.
CatalystCathode Pt Loading (mgPt cm−2)Test ConditionsPeak Power Density (W cm−2)Area (cm2)Mass Activity (A mgPt−1)Accelerated Durability TestMass Activity RetentionReference
Pt–Ni–Au NWs0.1H2/O2
200/200 sccm
100 kPaabs
0.714 A cm−2
@ 0.6 V
1 0.6 V (3 s)
0.95 V (3 s)
10 k cycles
0.520 A cm−2
@ 0.6 V
[16]
Pt1Co1-
IMC@Pt/C
0.2H2/O2
1000/400 sccm
100 kPaabs
1.45 A cm−2
@ 0.65 V
6.250.180.6 V (3 s)
0.95 V (3 s)
30 k cycles
75.2%[17]
PtNi
BNCs/C
0.15H2/Air
150/300 sccm
30 psi
1.5 A cm−2
@ 0.6 V
5 [30]
PtCo0.02H2/O2
200/200 sccm
150 kPaabs
51.520.6–1.0 V
30 k cycles
77%[31]
PtCo0.125H2/Air
250 kPaabs
1.17 1.080.6 V (3 s)
0.95 V (3 s)
30 k cycles
75%[37]
L10-W-PtCo/C0.11H2/O2
200/500 sccm
1.5 kPaabs
0.570.6–1.0 V
30 k cycles
82.5%[38]
PtCo@Gnp0.07H2/O2
835/2000 sccm
150 kPaabs
1.0151.210.6 V (3 s)
0.95 V (3 s)
30 k cycles
73%[49]
Sigracet
SGL 10BC GDL
0.15H2/O2 (stoichiometry 1.3/2.2) 200 kPaabs0.5316 [86]
Pt/C@SDT-Nafion0.07H2/Air
150 kPaabs
1.0850.350.6 V (3 s)
0.95 V (3 s)
30 k cycles
62%[90]
RA1 MEA H2/Air (stoichiometry 1.5/3) 100 kPaabs1.3 A cm−2
@ 0.6 V
0.6–1.0 V
30 k cycles
54.3% (ECSA retention)[93]
Pt/Vulcan (Cabot)0.07H2/Air
500/1000 sccm
0.91250.23 [94]
Ptskin@PdCo NTAs-4003.5 μgPt cm−2H2/O2
100/200 sccm
200 kPaabs
0.782.56 0.6–1.0 V
5 k cycles
63.5% (power density retention) [107]
Pt/N-KB0.11H2/Air
1000/2000 sccm
230 kPaabs
1.391.40.2020.6–1.0 V
1.5 k cycles
98.5%[115]
Gradient design MEA0.2H2/Air
stoichiometry 1.5/2.5
100 kPaabs
0.6925 [117]
Gradient design MEA0.09H2/Air
100 kPaabs
0.486.25 [118]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cao, F.; Ding, R.; Rui, Z.; Wang, X.; Meng, Z.; Zhang, B.; Dong, W.; Li, J.; Liu, J.; Jiang, X. Advances in Low Pt Loading Membrane Electrode Assembly for Proton Exchange Membrane Fuel Cells. Molecules 2023, 28, 773. https://doi.org/10.3390/molecules28020773

AMA Style

Cao F, Ding R, Rui Z, Wang X, Meng Z, Zhang B, Dong W, Li J, Liu J, Jiang X. Advances in Low Pt Loading Membrane Electrode Assembly for Proton Exchange Membrane Fuel Cells. Molecules. 2023; 28(2):773. https://doi.org/10.3390/molecules28020773

Chicago/Turabian Style

Cao, Feng, Rui Ding, Zhiyan Rui, Xuebin Wang, Zhen Meng, Bin Zhang, Weiwen Dong, Jia Li, Jianguo Liu, and Xiangfen Jiang. 2023. "Advances in Low Pt Loading Membrane Electrode Assembly for Proton Exchange Membrane Fuel Cells" Molecules 28, no. 2: 773. https://doi.org/10.3390/molecules28020773

Article Metrics

Back to TopTop