Next Article in Journal
Sensitive Mechanism and Instability Modeling Methods of Flexible Sensing Films Based on 2D Materials
Next Article in Special Issue
Enhancing Perovskite Solar Cell Stability by TCO Layer Presence Beneath MACl-Doped Perovskites
Previous Article in Journal
Luminescence Investigation of BaMgF4 Ceramics Under VUV Synchrotron Excitation
Previous Article in Special Issue
Artificial Intelligence and Li Ion Batteries: Basics and Breakthroughs in Electrolyte Materials Discovery
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advanced Strategies for Mitigating Catalyst Poisoning in Low and High Temperature Proton Exchange Membrane Fuel Cells: Recent Progress and Perspectives

1
Department of Environment and Energy Engineering, Sungshin Women’s University, Seoul 01133, Republic of Korea
2
Department of Chemical Engineering, Chungbuk National University, Choengju 28644, Republic of Korea
*
Authors to whom correspondence should be addressed.
Crystals 2025, 15(2), 129; https://doi.org/10.3390/cryst15020129
Submission received: 26 December 2024 / Revised: 19 January 2025 / Accepted: 21 January 2025 / Published: 24 January 2025

Abstract

:
Catalyst poisoning remains a persistent barrier to the efficiency and longevity of electrocatalytic energy conversion devices, namely fuel cells. To address this challenge, this review provides a systematic investigation of recent advancements in mitigation strategies, with particular emphasis on surface engineering, alloying, and combined approaches. Notable developments include the rational design of Pt-alloy catalysts with enhanced CO, H2S, and H3PO4 tolerance as well as the implementation of anti-poisoning molecular architectures and carbon-based protective layers. These methods collectively show considerable promise for improving catalytic activity by fine-tuning electronic structures and minimizing interactions with undesired adsorbates. In addition to presenting a comprehensive overview of the current progress, this review identifies promising future directions, guiding the design and realization of robust, poison-tolerant catalysts crucial for sustainable energy technologies.

1. Introduction

With their exceptional efficiency, high power density, rapid start-up, and lightweight design, hydrogen fuel cells are poised to transform the automotive and portable device sectors, offering a major step forward toward cleaner and more efficient energy solutions. Despite the significant promise of large-scale pure hydrogen production, it remains cost-prohibitive. As a result, approximately 95% of industrial hydrogen continues to be generated through economically viable methods such as natural gas steam reforming and coal gasification. In the former, natural gas reacts with steam to produce hydrogen, carbon monoxide, and trace amounts of carbon dioxide, whereas in the latter, coal is converted into syngas (a mixture of hydrogen and carbon monoxide), which is subsequently processed to extract hydrogen. These processes predominate largely because of their cost-effectiveness and well-established infrastructure.
However, hydrogen streams derived from these methods inevitably contain impurities including CO, H2S, and NH3, which can enter a fuel cell’s anode, degrade key materials, and curtail system performance. This contamination constrains the durability and commercial appeal of fuel cell electric vehicles. To further advance fuel cell technology, it is therefore critical to develop cells that tolerate these common impurities. Among them, CO and H2S pose the greatest challenges. Strong CO adsorption onto transition metals severely compromises catalytic processes in many applications. Specifically, in fuel cell anodes, CO competes with hydrogen for active sites on platinum, hindering the dissociation of H2 into atomic hydrogen. Similarly, hydrogen sulfide (H2S) exhibits a strong affinity for precious metals. When H2S interacts with platinum in an anode catalyst, sulfide formation ensues, diminishing the catalyst’s activity and negatively impacting the overall fuel cell performance.
Low temperature proton exchange membrane fuel cells (LT-PEMFCs), regarded as a valuable environmentally friendly technology for the next generation, have been facing the challenge of the aforementioned poisoning [1]. In LT-PEMFCs, the hydrogen oxidation reaction (HOR) occurs at the anode, while the oxygen reduction reaction (ORR) occurs at the cathode [2,3,4]. In general, the hydrogen from reformed natural gas serves as the fuel source for HOR, and even a small amount of carbon monoxide (CO) and hydrogen sulfide (H2S) in the parts per million (ppm) range within this gas can lead to catalyst poisoning and a subsequent reduction in fuel cell performance [5,6,7,8,9,10,11]. Considering the present technical constraints in producing pure hydrogen, which is derived from water electrolysis [12], as a fuel source, the utilization of reformed hydrogen is necessity. As a result, the development of catalysts that are tolerant to poisoning has emerged as a top priority [13,14]. Thus, numerous studies are currently underway to address this critical challenge [15,16,17,18,19].
Phosphoric acid fuel cells (PAFCs) or high temperature PEMFCs (HT-PEMFCs) are a class of fuel cells in which a phosphoric acid-impregnated membrane serves as the proton-conducting medium [20]. A notable advantage of HT-PEMFCs is their ability to operate at elevated temperatures exceeding 200 °C, which minimizes the susceptibility to poisoning by impurities in the fuel—a common limitation observed in LT-PEMFCs. However, HT-PEMFCs face a significant challenge due to the adsorption of bulky anionic phosphate species on catalyst surfaces, such as platinum, which are critical for oxygen adsorption. This adsorption can result in catalyst poisoning and performance degradation. Phosphate ions, characterized by their large size and strong binding affinity, have a pronounced detrimental effect on catalytic activity, even when adsorbed in small quantities. This phenomenon represents a critical and persistent limitation in the operation of HT-PEMFCs. In response, considerable research efforts are being directed toward developing strategies to mitigate phosphate-induced poisoning including the modification of catalyst surfaces and the incorporation of additional physical barriers. Such advancements are essential to enhance the durability and efficiency of HT-PEMFCs, thereby addressing one of the key challenges in their practical application.
This issue is not exclusive to FCs; it extends to various devices and chemical processes employing catalysts [21,22,23]. In conclusion, our ultimate goal is to develop catalysts with high catalytic activity that remain resistant to poisoning. Therefore, this paper aims to provide an extensive review of past research and propose a path for future exploration. The method for designing the anti-poisoning catalysts to be discussed was categorized into three main approaches: surface engineering, alloy development, and combination strategies.
While there have been previous reviews covering specific aspects of catalyst poisoning or targeting certain types of fuel cells, few have offered a comprehensive comparative analysis across LT-PEMFC and HT-PEMFC system. In particular, most existing works have focused on either surface engineering or alloy-based strategies alone, leaving a gap in discussions of combined, synergistic approaches. Therefore, the present review is both timely and urgent, as it consolidates diverse anti-poisoning methodologies—from molecular coatings to core–shell alloys—and examines how these can be adapted or optimized to address real-world operational challenges in a broad range of fuel cell technologies. We anticipate that our critical examination of recent advances, along with our proposed future directions, will guide researchers in designing more robust, multi-tolerant electrocatalysts. Ultimately, we hope that this review will serve as a catalyst itself, inspiring further innovations aimed at sustaining high-performance fuel cells in increasingly demanding applications.

2. Overview of State-of-the-Art Poison-Tolerant Electrocatalysts

2.1. Surface Engineering

Surface engineering aims to tailor the catalytic surface to improve the overall performance and tolerance toward common poisoning species [24,25]. Broadly, this strategy involves modifying a catalyst’s outermost layers through doping with heteroatoms, coating with protective shells, or constructing core–shell architectures. By adjusting the surface composition and atomic arrangement, researchers can influence key parameters such as the adsorption energy, charge transfer, and local electronic structure. Common materials used in surface engineering include noble metals (e.g., Pt, Au), transition-metal carbides/nitrides (e.g., TiWC, TiWN), and various organic/inorganic coatings (e.g., carbon shells, molecular layers). These engineered surfaces find particular utility in electrocatalysts for fuel cells, where even trace amounts of contaminants like CO or H2S can severely degrade activity. Applications span low-temperature-PEMFCs and HT-PEMFCs, all of which demand robust catalysts that remain active over extended operation. In this section, we explore how distinct surface-engineering methods—ranging from nitridation processes to molecular canopy-like structures—can mitigate poisoning effects and enhance catalyst longevity under realistic electrochemical conditions.
Aaron Garg et al. [26] presented the synthesis of early transition-metal nitride nanoparticles (TMN NPs) with a coating of thin noble-metal (NM) layers [27,28,29,30,31]. The production of Pt/TiWN core–shell NPs involves the nitridation process of Pt/TiWC NPs supported on carbon black (CB) under NH3 flow at 800 °C. A novel high-temperature self-assembly technique was devised to generate noble metal (NM)/transition-metal carbide (TMC) nanoparticles. This method entails carburizing silica-encapsulated transition metal oxide nanoparticles, which are coated with NM salt precursors.
The powder X-ray diffraction (PXRD) pattern showed that the Pt/TiWC NPs initially aligned closely with that of the face-centered cubic (fcc) WC, featuring a calculated lattice parameter of 0.422 nm (Figure 1a). The scarcity of reflections associated with fcc Pt implies a close interaction between Pt and the TiWC surface, preventing the formation of separate particles. To confirm the preservation of the core–shell structure in Pt/TiWN, scanning transmission electron microscopy combined with energy dispersive X-ray spectroscopy (STEM-EDX) was employed. Due to a weak Pt signal, STEM-EDX mapping could only distinctly discern Pt shells that were a few monolayers thick (Figure 1b).
The X-ray photoelectron spectroscopy (XPS) spectra of the core–shell nitrides and carbides, with Pt coverages ranging, exhibited distinct patterns in both the W 4f and Pt 4f binding energies (refer to Figure 2a,b) For all cases, the W 4f energies exhibited higher energy shifts for nitrides compared with carbides with equivalent Pt loading, while the W 4f and Pt 4f peak binding energies decreased with increasing Pt coverage. These alterations, termed core level shifts (CLSs), indicate bonding interactions between the Pt and W atoms, leading to a charge redistribution impacting the core levels. Although the cause of CLSs remains uncertain, they suggest significant changes in Pt’s electronic and chemical properties due to the presence of the transition-metal carbide (TMC) or nitride (TMN) core. CLSs offer insights into potential modifications in Pt reactivity, with larger CLSs in Pt/TiWN indicating weaker adsorbate binding compared with Pt/TiWC. Overall, these findings manifest the influence of core–shell composition on the electronic structure and reactivity of Pt.
The HOR activity results demonstrate the CO tolerance of core–shell catalysts (Figure 2c). In the absence of CO, all materials demonstrated exceptional HOR catalytic activity, requiring overpotentials below 50 mV for mass transfer limitation. Upon introducing 1000 ppm of CO, Ptcomm displayed suppressed HOR activity until the potentials surpassed 0.5 V. In contrast, both core–shell materials readily recovered their activity at potentials as low as 0.1 V, showcasing noteworthy CO tolerance. The observed CO tolerance can be ascribed to a combination of reduced CO binding energy and OH adsorption onto accessible W sites at low potentials. Significantly, consistent with the XPS CLSs data, Pt/TiWN exhibited superior CO tolerance compared with Pt/TiWC, displaying higher initial activity over Ptcomm and Pt/TiWC. This indicates weakened CO binding, facilitating CO oxidation even at 0.025 V, and the catalyst surface remained active without complete deactivation under these conditions.
A comparison of the pre-peak maxima of CO stripping voltammograms for Pt/TiWN and Pt/TiWC (Figure 2d) at 0.274 and 0.309 V, respectively, revealed that Pt/TiWN exhibited weaker binding to CO than Pt/TiWC. Additionally, the disparity in pre-peak area implies that Pt/TiWN has a higher proportion of weakly bound CO compared with Pt/TiWC (44% vs. 23%). These findings provide direct evidence of the distinctive catalytic properties arising from the utilization of a nitride core to modulate the electronic structure of Pt.
To address the issue of surface poisoning in HT-PEMFCs MEA, Kara et al. [32] introduced an electrocatalyst that eliminated noble metals at the cathode electrode interface. They employed non-platinum group metal (non-PGM) compounds with metal-organic framework (MOF) precursors for oxygen reduction [33,34,35]. Preliminary studies demonstrated the improved resistance to anion poisoning in various electrolytes, showcasing the oxygen reduction reaction (ORR) activity of Fe-based non-PGMs. The same group investigated the use of a MOF-supported Fe-based non-PGM (FePhen@MOF-ArNH3) as a potential electrocatalyst for ORR in HT-PEM fuel cells.
In contrast to FePhen@MOF-ArNH3 (Figure 3a), the introduction of only 10 mM H3PO4 into the electrolyte for Pt/C led to a slight decrease in the limiting current magnitude. This phenomenon suggests a hindrance to O2 reaching the electrode surface. Furthermore, compared with the profile without H3PO4, the half-wave potential (E1/2), a pseudo-kinetic parameter reflecting ORR activity, shifted cathodically by approximately 150 mV (0.87 vs. 0.72 V).
Tafel plots from the polarization data (Figure 3b) revealed that Pt/C experienced a significant increase in overvoltage in the presence of 10 mM H3PO4, indicating susceptibility to phosphate anion adsorption. In contrast, FePhen@MOF-ArNH3 showed negligible changes in Tafel performance, suggesting resistance to phosphoric acid poisoning. This clearly demonstrates that the active sites of FePhen@MOF-ArNH3 remain unaffected in the presence of phosphate anions.
To investigate adsorbate coverage on the electrocatalyst surface, the study employed the equation below and monitored the adsorbate coverage on FePhen@MOF-ArNH3.
Δ μ = μ   ( V ,   A r 100 m M   H   P O ) μ   ( 0.3 V ,   A r )
A comparison of the experimental Δ μ spectra (Figure 3c) with the theoretical models (Figure 3e) representative of phosphate anion adsorption on an Fe6 cluster revealed negligible Δ μ amplitude (<0.005 | Δ μ |) in the presence of H3PO4, indicating an absence of phosphate anion adsorption. This observation suggests that Fe particles are subsurface to graphitic layers, preventing direct contact with the electrolyte and hindering phosphate anion adsorption. These results corroborate the earlier claim of the absence of direct Fe−N coordination on the FePhen@MOF-Ar-NH3 catalyst. While the catalyst exhibited immunity to phosphate adsorption at room temperature, the authors emphasize the necessity for additional investigation under operational conditions. This is particularly crucial due to potential alterations in the electrolyte structure at elevated phosphate concentrations during fuel cell operation.
Tao Wang et al. [36] proposed a new strategy to enhance the resilience of local Pt active sites during H2 oxidation, providing high tolerance to CO and H2S [37,38,39,40,41]. Researchers introduced an organic molecular architecture using 2,6-diacetylpridine (DAcPy), a pyridine derivative with two carbonyl groups. DAcPy was reduced under the reaction condition and strongly bound to the Pt surface through tridentate coordination, forming a canopy-like structure. This architecture allowed Pt atoms to remain accessible for small-sized H2 while effectively blocking relatively larger CO and H2S molecules.
Comparing the two HOR polarization curves (Figure 4a,b), Figure 4a shows that the bare Pt/C is highly susceptible to H2S poisoning. As the Na2S concentration increased, the bare Pt/C was completely poisoned, resulting in the absence of HOR current. In contrast, the addition of DAcPy to the solution could enhance the H2S tolerance by spontaneous adsorption of DAcPy onto the Pt surface (Figure 4b). The XPS results demonstrate successful blocking of sulfur poisoning by DAcPy (Figure 4c). After the addition of DAcPy, no observable signal of the S element was detected on the Pt surface modified by DAcPy [42,43].
Density functional theory (DFT) calculations were conducted to confirm the adsorption of DAcPy on Pt (Figure 4d,e). In its pristine state, DAcPy tends to adsorb in parallel to the surface through van der Waals forces due to steric hindrance stemming from its acetyl groups, with an adsorption energy of 1.40 eV. It is lower than the adsorption energy of Py (1.61 eV) due to the N atom adsorption. However, under reductive adsorption, the energy dramatically increases to 3.50 eV, which is dramatically higher than that of CO (1.91 eV) and H2S (1.12 eV). In this reductive state, DAcPy accepts two proton–electron pairs, facilitating the reduction of the carbonyl groups to hydroxyl groups, thus forming Pt–C–OH bonds with Pt. This process is thermodynamically favorable, with a reaction energy of up to 2.10 eV. In this configuration, the reduced DAcPy adsorbs through three coordination sites (two Pt–C bonds and one Pt–N bond) with the pyridine ring in a tilted orientation (Figure 4d), allowing it to strongly compete with H2S and CO adsorption. Calculated adsorption energies for AcPy and its reduced form were 1.22 and 2.51 eV, respectively, which corresponded to the superior sulfur tolerance of DAcPy over AcPy.
When DAcPy adsorbed onto the Pt surface in a tilted orientation with a pyridine ring (Figure 4e), it formed a wedge-shaped space beneath it, creating a barrier against CO and H2S during the hydrogen oxidation reaction (HOR). This configuration established a height-limited space with a 2.41 Å gap between Pt and the pyridine ring. Within this space, small-sized H2 (2.30 Å) could approach isolated Pt atoms beneath the pyridine ring (depicted as light green spheres), while larger CO (2.65 Å) and H2S (4.47 Å) were impeded by the pyridine ring. This novel design marks the pioneering instance of creating molecular architectures for Pt-based catalysts to resist CO and H2S by strategically managing steric hindrance at local active sites. Notably, this approach avoids dependence on highly ordered and compact 2D molecular patterns, hinting at its potential versatility in mitigating various poisoning species beyond CO and H2S in the context of HOR.
Mengmeng Sun et al. [44] employed an ammonia borane (NH3BH3, AB) dehydration method to modify PtRu@h-BN/C core–shell electrocatalysts [45]. By utilizing the PtRu-catalyzed polymerization of ammonia borane and subsequent pyrolysis into graphitic h-BN in the presence of NH3, the researchers successfully enhanced the CO tolerance of PtRu/C nanocatalysts through encapsulation with few-layer h-BN shells.
Based on the TEM and HR-TEM images (Figure 5a,b), well-dispersed nanoparticles ranging in size from 3 to 6 nm were observed. Additionally, a significant portion of these NPs was surrounded by few-layer graphitic shells. Despite treatment at 600 °C in flowing NH3, the particle sizes of the PtRu@h-BN/C catalysts remained similar to those of the untreated PtRu/C catalyst.
HS-LCIS (Figure 5c) analysis showed that a significant portion of PtRu NPs were fully encapsulated within h-BN covers, forming PtRu@h-BN core–shell structures. This encapsulation was further confirmed through the benzothiazole adsorption test (Figure 5d,e). As expected, benzothiazole molecules exhibited strong adsorption on the surface of the PtRu/C catalyst (Figure 5d). In contrast, due to the large size of the molecules, benzothiazole diffusion through the h-BN shells was hindered, allowing only selective adsorption on the bare PtRu surface (Figure 5e).
In a CO tolerance durability experiment, the PtRu@h-BN/C catalyst demonstrated strong CO-tolerance in the fuel cell applications. As shown in Figure 6a,b, the single cell employing the PtRu/C anode catalyst manifested noticeable CO poisoning in contrast to the cell utilizing the PtRu@h-BN/C anode.
The current densities (Figure 6c,d) showed that PtRu@h-BN/C catalyst exhibited a minimal decrease in current density at 0.9 V, far less than the PtRu/C catalyst. In the voltage range of 0.4 to 0.8 V, the PtRu/C catalyst showed a 60% decrease, while the PtRu@h-BN/C catalyst achieved a reduction below 20%. These findings highlight the superior catalytic performance and enhanced CO tolerance of the PtRu@h-BN/C anode catalyst.
In this section, we explored various surface-engineering strategies for modifying the outer layers of catalysts to control the adsorption characteristics of reactants and poisonous species. By implementing techniques such as core–shell architectures, nitride/carbon coatings, and molecular canopy-like layers [26,32,36,44], researchers can effectively mitigate the detrimental effects of CO, H2S, and phosphate ions, thus enhancing both the durability and stability of fuel cells.
Nevertheless, surface engineering alone may not be sufficient for all operating conditions. In many cases, it is coupled with alloying (Section 2.2) in order to exploit synergistic effects. The following section will illustrate how combining different metal elements fine-tunes the electronic structure to achieve superior catalytic performance and enhanced tolerance to common poisoning species.

2.2. Alloying

Alloying represents a pivotal technique in heterogeneous catalysis, wherein combining different metals can yield synergistic effects unattainable by single-metal catalysts alone [46,47,48,49,50,51]. By incorporating second or third metal elements (e.g., Fe, Ru, Ni, Co, Ag), scientists can tune parameters such as the d-band center and the adsorption energies of reaction intermediates. This approach often improves catalytic activity, enhances selectivity, and bolsters resistance to poisoning species like CO or phosphate anions. Practical synthesis routes include one-pot solvothermal reactions, galvanic exchange, and high-temperature annealing, each tailored to control alloy composition and nanoparticle morphology. These alloyed catalysts have proven invaluable in fuel cell applications—particularly at the anode for hydrogen oxidation and at the cathode for oxygen reduction—where high durability and poisoning resilience are paramount. In this section, we introduce fundamental alloying concepts and illustrate their importance through recent examples of ternary and core–shell electrocatalysts, highlighting improvements in CO tolerance, long-term stability, and overall electrocatalytic efficiency.
Zhao et al. [52] modified a ternary electrocatalyst with surface Au-decorated PtFe nanocrystals (NCs) [53]. The resulting PtFeAu catalyst exhibited enhanced resistance to CO poisoning compared with bare Pt and binary PtFe. Surface engineering likely creates its distinctive electronic structure.
A ternary electrocatalyst comprising surface Au-decorated PtFe nanocrystals (NCs) (Figure 7a) was synthesized using a one-pot solvothermal method to form binary PtFe NCs, followed by a cation redox reaction in the presence of Au3+. The structural analysis was evaluated by HR-TEM and XRD (Figure 7b–d), confirming successful Au alloying. The TEM and HR-TEM images (Figure 7b,c) demonstrated the mono-dispersity of the binary PtFe NCs with an average particle size of around 3.2 nm. Subsequent galvanic reaction between Fe and Au3+ preserved the size and morphology of the resulting PtFeAu NCs, where the discontinuity in the PtFeAu lattice indicated the introduction of Au atoms at the surface.
In the XPS and CO stripping experiments, a significant negative shift was observed. The XPS (Figure 7e) findings confirmed the transfer of electron density from Pt to Au/Fe, resulting in a reduction in the Fermi level and the repositioning of the d-band center. Consequently, the binding strength of CO on the Pt site was weakened. CO stripping experiments (Figure 7f) revealed that CO stripping on the PtFeAu surface was the most facile reaction, which noted the weakened binding of CO on the Pt centers.
Chronoamperometric (CA) measurements and accelerated durability tests (Figure 7g,h) showed that ternary PtFeAu electrocatalysts outperformed PtFe NCs and Pt/C catalysts in current density and ECSA.
Lee et al. [54] reported research on fabricating a PtRuNi/C ternary electrocatalyst with a composition gradient shell using a combination of high-temperature heat treatment and the polydopamine (PDA) protective coating method [55,56,57]. Their work highlighted the effectiveness of protective coating layers in creating core–shell structures with a Pt-rich shell for optimal ORR catalysis. The protective coating method was found to prevent particle sintering during high-temperature treatment (600–800 °C).
The MEA performance showed (Figure 8a) that the development of PtRuNi/C ternary electrocatalysts is crucial for achieving high CO tolerance. The addition of Ni to the PtRu/C binary electrocatalyst resulted in enhanced CO tolerance in the ternary catalysts. The higher heat treatment temperature correlated with an increase in CO tolerance levels; however, the catalyst’s CO tolerance began to diminish at 700 °C, where Ru segregation was initiated. Notably, the catalyst treated at 800 °C exhibited the lowest CO resistance. Thus, it is imperative to subject the particles to heat treatment at a temperature that promotes a high alloying degree without inducing Ru segregation.
Table 1 summarizes the CO tolerance of various electrocatalysts from the literature, comparing them with the Pt1Ru1Ni0.75/C-PDA-650 catalyst in a fuel cell. The Pt1Ru1Ni0.75/C-PDA-650 catalyst demonstrated significant CO resistance, showing only an 11% performance reduction at 0.6 V in the presence of CO, outperforming other catalysts. This improved CO tolerance was attributed to the enhanced ligand effect.
The XPS spectra of Pt 4f in Pt1Ru1Ni0.75/C-PDA-650 and Pt1Ru1/C (Figure 8b) exhibited a negative shift in binding energies. This shift implies an electron transfer from Ni and Ru to Pt, leading to a reduction in the Pt-CO binding energy. As a result, the Pt1Ru1Ni0.75/C-PDA-650 catalyst showed enhanced CO tolerance.
Jingyu et al. [58] presented an innovative surface-engineering technique involving the application of PtBi alloy shells to coat Pt/Pt-alloys [59]. Despite the PtBi alloys’ downshifted d-band center, the introduction of Bi induced interlayer tensile strain, adversely affecting ORR electrocatalysis. However, the incorporation of large Bi atoms in the Pt overlayers resulted in in-plane shearing, causing surface compression and shortened Pt@Pt bonds (Figure 9a). Through annealing, solid-solution alloys were transformed into ordered intermetallic Pt-alloy NPs, and both materials were coated with PtBi alloy shells using a leach-embed-rearrangement process. The results inferred that the addition of PtBi led to improved activity as well as stability for all materials. Particularly notable was the exceptional ORR performance of FePt@PtBi, demonstrating a high resistance to methanol and CO poisoning.
The XRD patterns (Figure 9b) of FePt@PtBi and FePt revealed characteristic peaks for tetragonal FePt, intermetallic phases, and ordered L10-FePt, with similar (110)/(111) intensity ratios indicating highly ordered crystal structures. High angle annular dark field imaging-STEM (HAADF-STEM) images (Figure 9c) of FePt@PtBi depicted a 2–3 layer-thick PtBi shell covering the ordered intermetallic core with (110) planes of L10-FePt, and the FFT patterns confirmed the core’s face-centered tetragonal (fct) structure.
The same research group demonstrated the resistance of FePt@PtBi to poisoning in the presence of CH3OH or CO. When exposed to CH3OH (Figure 9d), FePt@PtBi maintained a consistent current response, in contrast with the commercial Pt/C electrocatalyst, which suffered a 70% reduction in current. Similarly, the introduction of CO (Figure 9e) had minimal impact on FePt@PtBi, while the commercial Pt/C electrocatalyst experienced an approximate 50% decline in the initial current.
Jang et al. [60] synthesized carbon-supported PtNi alloying nanoparticles (NPs) with varying thicknesses of carbon shells by controlling the concentration of oleylamine [61,62,63,64,65]. The oleylamine carbonization process was revealed through TEM images (Figure 10a), employing a high-temperature heat treatment in a H2-reductive atmosphere. This surface coating induced a geometric barrier effect in the catalyst. As the concentration of oleylamine increased, the thickness of the carbon shell also increased (Figure 10b). The higher standard deviation in thicker carbon shells indicates an uneven surface state in the metal nanoparticles (NPs), negatively impacting the electrochemical reaction by causing the loss of certain active sites.
Half-cell test results confirmed that the carbon shells acted as geometric barriers, preventing phosphate anion adsorption (Figure 11a). For PtNi@Cx, the carbon shells acted as spacers, restraining the adhesion of phosphate ions to the Pt surface due to the distance between the two materials. The oxygen reduction reaction (ORR) test results illustrated this phenomenon (Figure 11b,c), presenting scenarios with and without a carbon shell, respectively. Catalysts exhibited a similar reduction in ORR activity when exposed to oleylamine, dependent on the presence of phosphoric acid. In contrast, PtNi nanoparticles with carbon shells consistently recovered their ORR activities, irrespective of the thickness of the carbon shells. This implies the existence of a geometric barrier effect caused by the carbon shells. PtNi@C2’s superior catalytic performance in high-temperature proton exchange membrane fuel cells resulted from the synergistic effects of its geometry and electronics, enabled by a simple synthesis method.
In high-temperature PEMFCs, sustaining the molecular sieve effect is challenging due to the detachment of organic molecules from the metal nanoparticles. Sourabh et al. [66] proposed a solution involving the creation of a durable carbon molecular sieve layer (MSL) on Pt-based nanoparticles, thus maintaining the particle size and preventing metal dissolution [67]. This MSL enhances the ORR activity in phosphoric acid by reducing phosphate anion adsorption and facilitating selective O2 permeation, addressing challenges in long-term HT-PEMFC operation.
The HR-TEM images (Figure 12) illustrate the formation of ultrathin carbon shells on the Pt nanoparticles in the Pt@MSL catalysts through annealing in various gas atmospheres. Notably, annealing with 20% H2/N2 gas induced carbon etching, removing most carbon shells. The precisely controlled carbon MSLs are anticipated to evoke distinct phosphate poisoning effects in electrochemical tests.
A comparison of the ORR performance between the commercial Pt/C catalyst and Pt@MSL7 (Figure 13a–d) was conducted to observe the molecular sieve effect by carbon MSLs. Due to decreased active sites from phosphate anion absorption, the commercial Pt/C catalyst portrayed reduced ORR activity in the H3PO4-mixed HClO4 solution. Conversely, the Pt@MSL catalysts demonstrated interesting changes in their electrochemical properties. The ORR activity in both solutions enhanced with increasing carbon shell porosity from Pt@MSL7-0 to Pt@MSL7-20. The difference in ORR performance between electrolyte solutions increased with larger carbon shell pores, suggesting a pronounced molecular sieve effect. Smaller differences in ORR activity correlated with higher molecular sieving effects.
The Pt@MSL7-0 catalyst, with minimal mass activity for ORR in the H3PO4-mixed HClO4 solution, exhibited a notable molecular sieving effect (~90%). This suggests the effective prevention of catalyst poisoning by anions through the high-density carbon shell encapsulating the Pt nanoparticles. In contrast, the Pt@MSL7-10 catalyst displayed heightened mass activity and an appropriate molecular sieving effect value (44%) in the H3PO4-mixed solution, indicating selective O2 gas permeation through small carbon pores (defects) with the effective prevention of phosphate anion adsorption. Considered optimized among the prepared Pt@MSL catalysts, Pt@MSL7-10 demonstrated improved ORR activity, slightly lower than Pt@MSL7-20.
Jeong-Hoon et al. [68] investigated Pt-based phosphide catalysts for high-temperature PEMFCs (HT-PEMFCs) [69,70]. The catalysts were prepared using trioctylphosphine (TOP) as the phosphorus source, and the synthesis involved Ketjen black (KB) carbon. Unlike conventional methods using phosphine gas or tristrimethylphosphine, this study employed TOP as both a phosphorus source and a coordinating solvent. The PtP2 phase synthesis followed a unique conversion-based mechanism, setting it apart from previous strategies.
The synthesis process involved the initial conversion of Pt salt into Pt atoms, followed by their reaction with dissociated P atoms to yield the PtP2 phase (Figure 14a). To obtain a pure PtP2 phase, the P concentration was adjusted while keeping the Pt concentration constant. The procedure included vacuum treatment, heating at 390 °C for 10 h, washing with an ethanol and toluene (50:50) solution to remove unreacted TOP, and drying for enhanced crystallization, resulting in the catalyst PtP2/C-ASP.
TEM revealed well-dispersed PtP2 particles with a 2–3 nm diameter on the carbon surface of PtP2/C-ASP (Figure 14b), lacking long-range ordering as observed in the HR-TEM analysis. No diffraction spots were observed in the fast Fourier transform (FFT) pattern (Figure 14c,d). The amorphous PtP2 particle in PtP2/C-ASP, revealed by HAADF and energy-dispersive X-ray (EDX) imaging (Figure 14e), contained both Pt and P atoms. Upon heat treatment, a slight increase in the PtP2 particle size to approximately 3.9 nm was observed in PtP2/C-800 (Figure 14f), accompanied by the development of crystalline ordering (Figure 14g,h). The inverse FFT patterns (Figure 14i,j) indicated polycrystalline ordering in PtP2/C-800, which was supported by EDX images confirming the presence of the PtP2 crystalline phase (Figure 14k).
The authors employed DFT calculations to elucidate the resistance of PtP2 catalysts to phosphate poisoning and their exceptional activity. The results demonstrated stronger phosphate adsorption on PtP2 compared with Pt (Figure 15a), with P atoms attracting HPO4 anions, preserving Pt sites. Contrary to common assumptions, materials with stronger phosphate binding, such as PtP2, proved advantageous. The Pourbaix diagram of the surface (Figure 15b) revealed the formation of a protective layer of phosphoric oxide-based oligomers on PtP2 at oxidative potentials, ensuring the maintenance of active Pt sites during HPO4 adsorption.
Here, we discuss how alloying different metal components—such as Fe, Ru, Ni, Co, and others—can improve the catalytic properties beyond those of single-metal systems. By tuning the d-band center and modulating the adsorption energies of key intermediates, alloy catalysts exhibit increased activity, stability, and resistance to poisoning agents like CO and phosphates [52,54,58,60,66,68].
However, surface engineering and alloying each addresses different facets of catalyst design. While alloying manipulates the core and near-surface electronic structure, surface-engineering techniques provide targeted modifications at the atomic or molecular level. In the next section, we highlight how these two approaches can be combined to achieve even greater enhancements in catalyst performance under demanding fuel cell conditions.

2.3. Combined Approach

A growing trend in catalyst design involves merging both surface engineering and alloying approaches to optimize performance [71,72,73,74]. In these dual strategies, alloying adjusts the bulk and near-surface electronic structure, while targeted surface modifications—such as shell growth, protective coatings, or molecular-level design—fine-tune the active site accessibility and selectivity. By simultaneously exploiting the “bulk” electronic effects of alloys and the “local” control from surface coatings, researchers have achieved heightened activity and robust poison resistance, even under severe operating conditions found in high-temperature or direct methanol fuel cells. This combined methodology also facilitates the development of catalysts tailored to specific operational needs, such as selective blocking of large contaminant molecules, while still permitting smaller reactants to reach active sites. In this section, we review recent advances illustrating how the interplay between alloy composition and surface functionalization can produce cutting-edge electrocatalysts with outstanding CO, H2S, and phosphate-tolerance, ultimately underscoring the significance of integrated design principles in next-generation fuel cell systems.
A remarkable investigation was reported by Daisuke et al. [75], who proposed replacing RuO2.1 nanosheets with Ru nanosheets to enhance the CO tolerance of PtRu/C [76]. Through metallization, the Ru nanosheets established close contact with the surface of the PtRu nanoparticles while preserving the high surface area of Pt, aiming to boost the catalyst’s CO tolerance.
Figure 16a–c shows the mean size of the PtRu nanoparticles in Ru(ns)-PtRu/C (4.3 ± 0.4 nm) was similar to that of the original PtRu/C (4.2 ± 0.3 nm). This indicates that the metallization process does not lead to an increase in the size of the PtRu nanoparticles. Moreover, Ru(ns)-PtRu/C showed an identical fcc(220) reflex and width (d-spacing of 0.1369 nm) compared with PtRu/C and RuO2.1ns-PtRu/C (Figure 16d). This similarity suggests that the alloying state of PtRu nanoparticles was consistent among all catalysts.
Comparing the hydrodynamic voltammograms of the HOR contingent on the CO (Figure 17a,b), it is evident that the behavior of HOR in pure H2 was comparable among PtRu/C, RuO2.1ns-PtRu/C, and Ru(ns)-PtRu/C. This observation underscores that the introduction of metallic Ru nanosheets does not impede the activity of HOR. Conversely, when HOR was conducted in the presence of CO, the HOR current for all catalysts was lower than that observed in pure H2 due to CO poisoning. Upon closer examination, RuO2.1ns-PtRu/C exhibited a higher HOR current than PtRu/C, and Ru(ns)-PtRu/C surpassed RuO2.1ns-PtRu/C in HOR current. These findings suggest that integrating metallic Ru nanosheets effectively enhanced the catalyst’s CO tolerance.
Similarly, the chronoamperogram results in pure H2 (Figure 17c) indicate that RuO2.1ns PtRu/C and Ru(ns)-PtRu/C displayed comparable or slightly higher HOR activity than PtRu/C. As anticipated, a sharp initial decrease in HOR activity for all catalysts was observed in 300 ppm CO/H2 due to CO accumulation (Figure 17d), followed by a quasi-steady state. At 5 h, Ru(ns)-PtRu/C exhibited 1.3 times higher HOR activity than PtRu/C. The normalized current decay ratio (Figure 17e) illustrates that Ru(ns)-PtRu/C mitigated CO poisoning in the initial 60 min. The steady-state current at 3 h for Ru(ns)-PtRu/C was in contrast to PtRu/C’s continuous decrease even after 5 h, suggesting that the improved CO tolerance due to metallic Ru nanosheet modification was most likely from the suppression of CO adsorption on the Pt surface.
Furthermore, the HOR activity after ADT (Table 2) demonstrated superior retention in HOR activity for Ru(ns)-PtRu/C after both 1000 and 3000 cycles. In conclusion, the superior activity and stability of HOR in the presence of CO are due to the Ru nanosheets effectively suppressing the adsorption of CO on the catalyst surface.
Yezhou Hu et al. [77] introduced a one-step pyrolysis method for synthesizing Pt-Fe intermetallic compounds with in situ N-doped carbon encapsulation [78,79]. The resulting O-Pt-Fe@NC/C demonstrated improved ORR activity and long-term stability in 0.1 M HClO4, along with enhanced tolerance to CO, SOx, and POx compared with Pt/C, attributed to the protective carbon layer. O-Pt-Fe@NC/C, as a cathode catalyst in HT-PEMFCs, exhibited significantly enhanced maximum power density.
O-Pt-Fe@NC/C was synthesized via one-step pyrolysis (Figure 18), involving the reduction of metal salts and the transformation of dicyandiamide into graphitic carbon nitride (g-C3N4) at temperatures below 550 °C. As the temperature increased to 800 °C, the decomposition of g-C3N4 occurred, forming an ultrathin N-doped carbon layer catalyzed by Pt-Fe nanoparticles and coating their surface, while simultaneously promoting the formation of the ordered Pt-Fe intermetallic phase.
Comparing the XRD patterns (Figure 19a) of Pt/C, the disordered Pt-Fe alloy (D-Pt-Fe/C), and O-Pt-Fe@NC/C, both O-Pt-Fe@NC/C and D-Pt-Fe/C exhibited reflex shifts to higher angles, signifying lattice contraction from the introduced Fe. The TEM images (Figure 19b) revealed well-dispersed nanoparticles (average size ~3.8 nm) on the carbon support, demonstrating effective size control by the carbon layer, which resisted Ostwald ripening or coalescence during high-temperature treatment. The HR-TEM images (Figure 19c) depicted Pt-Fe nanoparticles encapsulated in a ~0.7 nm thick carbon shell, suggesting a strong coupling effect. Pores in the carbon layer (Figure 19d) resulted from N doping and CNx gases released during the C3N4 heat treatment. The EDS elemental mapping (Figure 19e–j) confirmed the presence of an N-doped carbon shell adjacent to nanoparticles. XPS validated the existence of N in the carbon shell, indicating an interface confinement effect between the N-doped carbon and Pt-Fe ordered intermetallic nanoparticles.
In the CO stripping tests (Figure 20a,b), O-Pt-Fe@NC/C exhibited superior anti-poisoning capability with the lowest CO oxidation peak potential and minimal half-wave decay in the presence of SO32− compared with the others. The authors attributed this effect to nitrogen introduction. Nitrogen’s higher electronegativity changes the local charge on carbon, which improves the charge transfer and reduces the adsorption of poisonous molecules.
Additionally, the carbon layer acts as a physical protective barrier, effectively isolating nanoparticles from contact with poisonous molecules. Furthermore, O-Pt-Fe@NC/C was systematically compared with previously reported findings (Figure 20c). The results highlighted O-Pt-Fe@NC/C as a promising material for HT-PEMFCs operating at elevated temperatures such as 160 °C.
Douglas et al. [80] improved CO tolerance in low-temperature PEM hydrogen oxidation by utilizing a Pt/C and WO3 powder mixture [81,82,83]. The authors elucidated the CO tolerance enhancement mechanism in the Pt/WO3 system, focusing on the role of WV/WVI in the bifunctional mechanism and WO3’s influence on CO coverage and binding modes to Pt. This investigation provides insights into potential electronic effects.
The TEM images illustrated the sizes of the synthesized WO3 nano-platelets, Pt particles, and mixed catalyst electrodes (Figure 21a–c). In the mixed state, there was an interdispersion of the Pt/C and WO3 components, with a stacking of WO3 nanoplatelets. The XPS analysis (Figure 21d) revealed a ~0.2 eV shift in the Pt4f electron peak position, indicating a potential two-phase boundary or equilibration of the electronic energy levels in the Pt/C/WO3 catalyst mixture. The result addressed the possibility of W soluble species depositing on the platinum surface during the measurements, considering the stability of WO3 electrodes toward dissolution under operation conditions.
The authors proposed a mechanism of catalytic cycle for CO oxidation by WO3 (Figure 21e), which involves WVI as the active redox species oxidizing COads,Pt in the presence of water (reaction a) to form CO2 and tungsten bronze. The rapid electrochemical oxidation of tungsten bronze (reaction b) occurs at potentials exceeding ~0.3 VRHE, completing the catalytic cycle. Despite the thermodynamic feasibility of the electrochemical oxidation of WV above ~0.2 VRHE, the slow electron transfer kinetics between 0.2 and 0.3 VRHE make reaction “b” the rate-determining step in this potential range.
To investigate the oxidation state of tungsten oxide and its potential influence on enhanced CO tolerance in Pt/C/WO3 catalysts, in situ Raman spectroscopy was conducted (Figure 22a,b). The Pt/C/WO3 electrode exhibited peaks at 271, 714, and 807 cm−1 at open-circuit potential, corresponding to δ(O−W−O) bending and W−O stretching modes of WO3. Reducing potential led to peak disappearance, indicating the formation of hydrogenated tungsten bronze. At 0.1–0.25 VRHE, a peak at 820 cm−1 suggested partially reduced tungsten centers, while potentials above 0.3 VRHE revealed fully oxidized WO3 (WVI). The upward shift in peak position at lower potentials signified a stronger W-O bond in the hydrogenated tungsten bronze than in WO3, attributed to the increased covalency of the WV-O compared with WVI-O bonds (Figure 22b). This effect may enhance CO oxidation at positive potentials. The reduced peak area in more hydrogenated tungsten bronze was attributed to increased coloration, reducing scattered light as well as lowering the optical skin depth in more conductive tungsten bronze compared with semiconducting WO3. Although the comparison of peak areas is qualitative, it strongly suggests the presence of distinct phases at different potentials. Additionally, Pt is known to induce oxidation state changes in WO3 through H2 spillover at room temperature, a phenomenon directly applicable to the Pt/C/WO3 electrochemical system in this study.
In situ ATR-IR spectroscopy was employed to investigate the CO oxidation behavior of Pt/C catalysts with and without WO3. Spectra (Figure 22c) revealed characteristic Pt-CO features with bands at 2000–2100 cm−1 and 1820–1920 cm−1, corresponding to linear (COL) and bridge-bonded (COB) CO configurations on Pt. At an initial potential of 0.04 VRHE, the broad and asymmetric COL band suggested multiple CO adsorption environments, similar to the carbon-supported Pt catalysts. Incremental potential steps from 0.04 to 1.09 VRHE resulted in shifts and intensity reductions in both the COL and COB bands, indicating CO removal from the surface, with a surface structure sensitivity observed in the Pt-CO oxidation process. While the Pt/C/WO3 catalyst displayed a similar behavior to Pt/C, differences were observed in the shape and position of the linear-bonded CO band (Figure 22d). Pt/C/WO3 showed a primary peak at a lower wavenumber (2009 cm−1) with shoulders at 2027 and 2053 cm−1, indicating a unique Pt−CO bonding environment induced by WO3. The lower wavenumbers suggest an electronic/polarization effect or reduced dipole−dipole coupling, implying lower coverage or higher mobility. The intensity decreased with increasing potential, and preferential removal of the lower wavenumber feature mirrored the Pt/C behavior.
Therefore, this study emphasizes WO3’s effectiveness in improving CO tolerance in Pt/C catalysts, particularly in the potential range 0.3 ≥ E ≥ 0.4 VRHE. This improvement makes them suitable for applications with an anodic overpotential, such as in direct methanol fuel cells, by revealing WO3‘s electronic effect on CO binding to Pt.
In this section, we presented examples of merging surface engineering and alloying methods to create advanced electrocatalysts—such as alloy core–shell structures with protective coatings or molecular modifications [75,77,80]. These combined strategies allow for the precise tuning of both bulk electronic effects and local surface chemistry, resulting in outstanding stability and poison resistance in severe operating environments.
As we move forward to the conclusion, we will integrate the key insights from Section 2.1, Section 2.2 and Section 2.3, discussing how the interplay between surface modifications and alloy design drives next-generation catalysts with excellent durability, activity, and tolerance in diverse fuel cell systems.

3. Converging Anti-Poisoning Strategies

Table 3 highlights that, despite the diversity in catalyst compositions, synthesis strategies, and target poisoning species, all of the studies featured shared the same overarching goal: mitigating catalyst deactivation due to adsorbed contaminants. By comparing parameters such as catalyst materials, surface engineering methods, alloying compositions, and resulting performance metrics, we can see how seemingly unrelated research has ultimately converged on controlling the surface adsorption energies and modulating local electronic structures. This unified perspective underscores that every approach—ranging from molecular coatings on Pt to carbide or nitride core–shell architectures—seeks to preserve active sites from poisoning by CO, H2S, phosphate anions, or other foulants. In doing so, each study contributes a complementary piece to the larger puzzle of designing robust, long-lived catalysts for fuel cell technologies and related electrochemical applications.

4. Conclusions and Perspectives

Catalyst poisoning represents a critical challenge in electrochemical reactions, significantly impeding catalytic activity and accelerating the development of stable electrocatalysts for PEMFCs. Comprehensive investigations in this review identified three primary strategies for mitigating catalyst poisoning:
  • Surface engineering: This approach involves direct surface modifications and the implementation of protective barriers to manipulate catalyst geometric properties and enhance poisoning resistance.
  • Alloying: A strategic method that modifies the electronic structure of catalytic materials to improve their tolerance to poisoning agents.
  • Combined approaches: Generation of synergistic effects by combining multiple mitigation strategies to optimize catalyst performance.
Among these approaches, integrated methodologies have demonstrated the most promising results in developing high-performance, poison-resistant catalysts. The effectiveness of combined strategies highlights the potential for advanced catalyst design that can overcome chronic limitations in electrochemical systems. The current research progress necessitates a critical transition from laboratory-scale investigations to various evaluations under realistic operational conditions. Critical considerations for future development include:
  • Rigorous and standardized characterization protocols of catalyst performance under industrially relevant environments;
  • Development of advanced in situ monitoring techniques;
  • Comprehensive understanding of long-term catalyst stability and degradation mechanisms.
Sustained research efforts and interdisciplinary collaborations are essential to address the complex challenges of catalyst poisoning. Future investigations should focus on the development of robust methodologies that can reliably mitigate the poisoning effects and achieve the severe requirements of large-scale commercial applications in PEMFCs. These advancements will pave the way for the reliable and efficient large-scale commercialization of PEMFCs in automotive and stationary power applications.
Although the present review primarily focused on LT-PEMFCs and HT-PEMFCs, the anti-poisoning strategies outlined here including surface engineering, alloying, and combined approaches can be extended to a broad range of electrochemical energy conversion and storage devices. For instance, catalytic materials used in CO2 reduction systems, electrolyzers, and metal–air batteries can face analogous poisoning challenges arising from contaminants, by-products, or strongly adsorbing intermediates [12,84,85]. The design principles highlighted in this review, particularly those involving the manipulation of d-band centers, the selective blocking of large adsorbates, and the integration of protective or functional coatings, are readily transferrable to other platforms that require sustained catalytic activity over prolonged operation.
Advanced in situ monitoring techniques such as operando X-ray absorption spectroscopy and high-resolution electron microscopy can be leveraged not only to examine degradation pathways in fuel cells, but also to elucidate poisoning mechanisms in diverse electrochemical systems [80]. By doing so, researchers can expedite the transition of promising anti-poisoning methods from laboratory proof-of-concept to large-scale deployment. Moreover, interdisciplinary collaborations encompassing materials science, chemical engineering, and computational modeling will be instrumental in realizing catalyst designs that exhibit high selectivity and durability across multiple applications [36,68]. This synergy offers significant potential for accelerating innovation beyond the confines of fuel cell technologies, paving the way for robust catalysts that can withstand both conventional and emerging contaminants in numerous electrochemical processes.
In summary, while this review has underscored the urgency and feasibility of combating catalyst poisoning in LT-PEMFCs and HT-PEMFCs, it also serves as a roadmap for developing broadly applicable solutions that address the multifaceted demands of modern electrochemical energy conversion and storage. The lessons learned here, such as tuning surface geometry, employing alloying for electronic effects, and combining multiple strategies for synergistic gains, can inform the next generation of catalysts designed to sustain activity and minimize deactivation in any high-performance electrochemical environment.

Author Contributions

Writing—original draft preparation, S.C.; Writing—review and editing, I.J. and S.L.; Supervision I.J. and S.L.; Funding acquisition, S.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Sungshin Women’s University Research Grant of 2022, grant number H20220087.

Data Availability Statement

No new data were created or analyzed in this study.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Molochas, C.; Tsiakaras, P. Carbon Monoxide Tolerant Pt-Based Electrocatalysts for H2-PEMFC Applications: Current Progress and Challenges. Catalysts 2021, 11, 1127. [Google Scholar] [CrossRef]
  2. Park, J.-H.; Saito, N.; Kawasumi, M. Improving the Durability and Anti-Poisoning Properties of Platinum Catalysts by Carbon Shell Encapsulation: A Promising Approach for Both Cathode and Anode Catalysts for Polymer Electrolyte Membrane Fuel Cells. Int. J. Hydrog. Energy 2024, 72, 642–651. [Google Scholar] [CrossRef]
  3. Zhao, W.; Xu, G.; Dong, W.; Zhang, Y.; Zhao, Z.; Qiu, L.; Dong, J. Progress and Perspective for in Situ Studies of Oxygen Reduction Reaction in Proton Exchange Membrane Fuel Cells. Adv. Sci. 2023, 10, 2300550. [Google Scholar] [CrossRef]
  4. Xu, X.; Pan, Y.; Zhong, Y.; Ran, R.; Shao, Z. Ruddlesden–Popper Perovskites in Electrocatalysis. Mater. Horiz. 2020, 7, 2519–2565. [Google Scholar] [CrossRef]
  5. Wang, M.; Pei, P.; Xu, Y.; Fan, T.; Ren, P.; Zhu, Z.; Chen, D.; Fu, X.; Song, X.; Wang, H. CO-Tolerance Behaviors of Proton Exchange Membrane Fuel Cell Stacks with Impure Hydrogen Fuel. Appl. Energy 2024, 366, 123326. [Google Scholar] [CrossRef]
  6. Wei, X.; Wang, R.-Z.; Zhao, W.; Chen, G.; Chai, M.-R.; Zhang, L.; Zhang, J. Recent Research Progress in PEM Fuel Cell Electrocatalyst Degradation and Mitigation Strategies. EnergyChem 2021, 3, 100061. [Google Scholar] [CrossRef]
  7. Liu, Z.; Cai, S.; Tu, Z.; Chan, S.H. Recent Development in Degradation Mechanisms of Proton Exchange Membrane Fuel Cells for Vehicle Applications: Problems, Progress, and Perspectives. Energy Storage Sav. 2024, 3, 106–152. [Google Scholar] [CrossRef]
  8. Patil, V.; Reshmi, P.; Prajna, S.; Haleshappa, D.; Jayarama, A.; Pinto, R. Degradation Mechanisms in PEM Fuel Cells: A Brief Review. Mater. Today Proc. 2023, in press. [CrossRef]
  9. Okonkwo, P.C.; Ige, O.O.; Uzoma, P.C.; Emori, W.; Benamor, A.; Abdullah, A.M. Platinum Degradation Mechanisms in Proton Exchange Membrane Fuel Cell (PEMFC) System: A Review. Int. J. Hydrog. Energy 2021, 46, 15850–15865. [Google Scholar] [CrossRef]
  10. Haıdar, F.; Arora, D.; Soloy, A.; Bartoli, T. Study of Proton-Exchange Membrane Fuel Cell Degradation and Its Counter Strategies: Flooding/Drying, Cold Start and Carbon Monoxide Poisoning. Int. J. Automot. Sci. Technol. 2024, 8, 96–109. [Google Scholar] [CrossRef]
  11. Labata, M.F.; Li, G.; Ocon, J.; Chuang, P.-Y.A. Insights on Platinum-Carbon Catalyst Degradation Mechanism for Oxygen Reduction Reaction in Acidic and Alkaline Media. J. Power Sources 2021, 487, 229356. [Google Scholar] [CrossRef]
  12. Sun, H.; Xu, X.; Kim, H.; Jung, W.; Zhou, W.; Shao, Z. Electrochemical Water Splitting: Bridging the Gaps between Fundamental Research and Industrial Applications. Energy Environ. Mater. 2023, 6, e12441. [Google Scholar] [CrossRef]
  13. Ramli, Z.A.C.; Pasupuleti, J.; Zaiman, N.F.H.N.; Saharuddin, T.S.T.; Samidin, S.; Isahak, W.N.R.W.; Sofiah, A.; Kamarudin, S.K.; Tiong, S.K. Evaluating Electrocatalytic Activities of Pt, Pd, Au and Ag-Based Catalyst on PEMFC Performance: A Review. Int. J. Hydrog. Energy 2024, in press.
  14. Tian, H.; Cui, X.; Shi, J. Emerging Electrocatalysts for PEMFCs Applications: Tungsten Oxide as an Example. Chem. Eng. J. 2021, 421, 129430. [Google Scholar] [CrossRef]
  15. Luo, W.; Jiang, Y.; Wang, M.; Lu, D.; Sun, X.; Zhang, H. Design Strategies of Pt-Based Electrocatalysts and Tolerance Strategies in Fuel Cells: A Review. RSC Adv. 2023, 13, 4803–4822. [Google Scholar] [CrossRef]
  16. Valdés-López, V.F.; Mason, T.; Shearing, P.R.; Brett, D.J. Carbon Monoxide Poisoning and Mitigation Strategies for Polymer Electrolyte Membrane Fuel Cells–A Review. Prog. Energy Combust. Sci. 2020, 79, 100842. [Google Scholar] [CrossRef]
  17. Zhang, J.; Zhang, C.; Li, J.; Deng, B.; Fan, M.; Ni, M.; Mao, Z.; Yuan, H. Multi-Perspective Analysis of CO Poisoning in High-Temperature Proton Exchange Membrane Fuel Cell Stack via Numerical Investigation. Renew. Energy 2021, 180, 313–328. [Google Scholar] [CrossRef]
  18. Liu, L.; Zeng, R.; Hao, L. Advances in CO-Tolerant Anode Catalysts for Proton Exchange Membrane Fuel Cells. Prog. Nat. Sci. Mater. Int. 2024, 34, 654–675. [Google Scholar] [CrossRef]
  19. Bhuvanendran, N.; Ravichandran, S.; Lee, S.; Sanij, F.D.; Kandasamy, S.; Pandey, P.; Su, H.; Lee, S.Y. Recent Progress in Pt-Based Electrocatalysts: A Comprehensive Review of Supported and Support-Free Systems for Oxygen Reduction. Coord. Chem. Rev. 2024, 521, 216191. [Google Scholar] [CrossRef]
  20. Guo, Z.; Perez-Page, M.; Chen, J.; Ji, Z.; Holmes, S.M. Recent Advances in Phosphoric Acid–Based Membranes for High–Temperature Proton Exchange Membrane Fuel Cells. J. Energy Chem. 2021, 63, 393–429. [Google Scholar] [CrossRef]
  21. 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] [PubMed]
  22. Zhang, W.; Liu, C.; Kucernak, A.; Liu, H.; Wu, J.; Li, S. Synergizing Single-Atom and Carbon-Encapsulated Nanoparticles of Fe for Efficient Oxygen Reduction and Durable Zn–Air Batteries. ACS Appl. Energy Mater. 2024, 7, 5398–5407. [Google Scholar] [CrossRef]
  23. Wang, X.; Lu, Y.; Xie, C.; Wang, Y.; Li, H.; Ding, X. Hetero-Structured Composite Cathodes by Electrospinning with High CO2-Poisoning Tolerance for Solid Oxide Fuel Cells. Int. J. Hydrog. Energy 2024, 50, 1137–1146. [Google Scholar] [CrossRef]
  24. Yue, X.; Zhang, X.; Zhang, M.; Du, W.; Xia, H. The Enhancement in the Performance of Ultra-Small Core–Shell Au@ AuPt Nanoparticles toward HER and ORR by Surface Engineering. Nanoscale 2023, 15, 4378–4387. [Google Scholar] [CrossRef]
  25. Long, D.; Ping, X.; Ni, J.; Chen, F.; Chen, S.; Wei, Z.; Guo, L.; Zheng, J. Strengthening Pt/WO x Interfacial Interactions to Increase the CO Tolerance of Pt for Hydrogen Oxidation Reaction. Chem. Commun. 2023, 59, 13583–13586. [Google Scholar] [CrossRef]
  26. Garg, A.; Milina, M.; Ball, M.; Zanchet, D.; Hunt, S.T.; Dumesic, J.A.; Román-Leshkov, Y. Transition-metal Nitride Core@ Noble-metal Shell Nanoparticles as Highly CO Tolerant Catalysts. Angew. Chem. Int. Ed. 2017, 56, 8828–8833. [Google Scholar] [CrossRef]
  27. Chen, L.; Zhang, P.; Jin, Y.-Q.; Yang, H.; Sheng, T.; Yan, Y.; Wang, T.; Chen, Z.; Tian, N.; Li, X. Enhancing CO Tolerance in PEMFC Anodes via Thermal Oxidation Induced RuO2 Blocking Shell on a PtRu/C Catalyst. Nano Lett. 2024, 24, 10642–10649. [Google Scholar] [CrossRef]
  28. Pan, H.-R.; Tang, T.; Jiang, Z.; Ding, L.; Xu, C.; Hu, J.-S. CO-Tolerant Hydrogen Oxidation Electrocatalysts for Low-Temperature Hydrogen Fuel Cells. J. Phys. Chem. Lett. 2024, 15, 3011–3022. [Google Scholar] [CrossRef]
  29. Wang, R.; Du, Y.; Yan, Y.; Yan, S.; Zou, Z. Dopamine-Carbonized Coating PtCo Catalyst with Enhanced Durability toward the Oxygen Reduction Reaction. J. Phys. Chem. Lett. 2024, 15, 8459–8466. [Google Scholar] [CrossRef]
  30. Park, H.; Kim, D.-K.; Kim, H.; Oh, S.; Jung, W.S.; Kim, S.-K. Binder-Coated Electrodeposited PtNiCu Catalysts for the Oxygen Reduction Reaction in High-Temperature Polymer Electrolyte Membrane Fuel Cells. Appl. Surf. Sci. 2020, 510, 145444. [Google Scholar] [CrossRef]
  31. Wen, Z.; Wu, D.; Banham, D.; Chen, M.; Sun, F.; Zhao, Z.; Jin, Y.; Fan, L.; Xu, S.; Gu, M. Micromodification of the Catalyst Layer by CO to Increase Pt Utilization for Proton-Exchange Membrane Fuel Cells. ACS Appl. Mater. Interfaces 2022, 15, 903–913. [Google Scholar] [CrossRef] [PubMed]
  32. Strickland, K.; Pavlicek, R.; Miner, E.; Jia, Q.; Zoller, I.; Ghoshal, S.; Liang, W.; Mukerjee, S. Anion Resistant Oxygen Reduction Electrocatalyst in Phosphoric Acid Fuel Cell. Acs Catal. 2018, 8, 3833–3843. [Google Scholar] [CrossRef]
  33. Zhang, L.; Li, L.; Chen, H.; Wei, Z. Recent Progress in Precious Metal-free Carbon-based Materials towards the Oxygen Reduction Reaction: Activity, Stability, and Anti-poisoning. Chem.—Eur. J. 2020, 26, 3973–3990. [Google Scholar] [CrossRef]
  34. Orellana, W.; Zuniga, C.; Gatica, A.; Ureta-Zanartu, M.-S.; Zagal, J.H.; Tasca, F. Effect of Electrolyte Media on the Catalysis of Fe Phthalocyanine toward the Oxygen Reduction Reaction: Ab Initio Molecular Dynamics Simulations and Experimental Analyses. ACS Catal. 2022, 12, 12786–12799. [Google Scholar] [CrossRef]
  35. Tang, M.; Yan, H.; Zhang, X.; Zheng, Z.; Chen, S. Materials Strategies Tackling Interfacial Issues in Catalyst Layers of Proton Exchange Membrane Fuel Cells. Adv. Mater. 2023, 2306387. [Google Scholar] [CrossRef]
  36. Wang, T.; Chen, Z.-X.; Yu, S.; Sheng, T.; Ma, H.-B.; Chen, L.-N.; Rauf, M.; Xia, H.-P.; Zhou, Z.-Y.; Sun, S.-G. Constructing Canopy-Shaped Molecular Architectures to Create Local Pt Surface Sites with High Tolerance to H 2 S and CO for Hydrogen Electrooxidation. Energy Environ. Sci. 2018, 11, 166–171. [Google Scholar] [CrossRef]
  37. Zhang, D.; Liu, W.; Ye, K.; Li, X. High CO and Sulfur Tolerant Proton Exchange Membrane Fuel Cell Anodes Enabled by “Work along Both Lines” Mechanism of 2, 6-Dihydroxymethyl Pyridine Molecule Blocking Layer. J. Colloid Interface Sci. 2024, 653, 413–422. [Google Scholar] [CrossRef]
  38. Cai, B.; Chen, X.; Wang, L.; Fu, H. Advanced Progress for Promoting Anodic Hydrogen Oxidation Activity and Anti-CO Poisoning in Fuel Cells. ACS Catal. 2024, 14, 13602–13629. [Google Scholar] [CrossRef]
  39. Tang, M.; Yang, T.; Yang, X.; Li, Y.; Shi, Z.; Wang, X.; Liu, C.; Xing, W.; Ge, J. Single-Atom Catalysts for Proton Exchange Membrane Fuel Cell: Anode Anti-Poisoning & Characterization Technology. Electrochim. Acta 2023, 446, 142120. [Google Scholar]
  40. Lee, D.W.; Choi, D.; Lee, M.J.; Jin, H.; Lee, S.; Jang, I.; Park, H.-Y.; Jang, J.H.; Kim, H.-J.; Lee, K.-Y. Tailoring of Pt Island RuO2/C Catalysts by Galvanic Replacement to Achieve Superior Hydrogen Oxidation Reaction and CO Poisoning Resistance. ACS Appl. Energy Mater. 2021, 4, 8098–8107. [Google Scholar] [CrossRef]
  41. Sun, C.; Qiu, L.; Ke, S.; Zhang, J.; Xu, G.; Dou, M. Improved Chloride Tolerance of PtCo/C with a Pt-Skin Structure toward the Oxygen Reduction Reaction Due to a Weakened Pt–Cl Interaction. ACS Appl. Energy Mater. 2022, 5, 7503–7514. [Google Scholar] [CrossRef]
  42. Li, X.; Xi, J.; Tang, L.; Wang, H.; Liu, Y.; Min, C.; Gao, S.; Liu, F.; Yang, X. A High-Performance and Anti-Sulfur Poisoning Oxygen Reduction Reaction Catalysts Derive from Molybdenum-Doping on PtCo Alloy Nanoparticles Surface. Electrochim. Acta 2023, 462, 142658. [Google Scholar] [CrossRef]
  43. Wang, M.; Cao, L.; Liu, Z.; Yang, W.; Sun, S.; Hou, M.; Shao, Z. Sulfur Dioxide-Resistant Platinum-Based Intermetallic Nanocatalysts Encaged by Porous Nitrogen-Doped Carbon for Oxygen Reduction Reaction. Chem. Eng. J. 2024, 492, 152162. [Google Scholar] [CrossRef]
  44. Sun, M.; Lv, Y.; Song, Y.; Wu, H.; Wang, G.; Zhang, H.; Chen, M.; Fu, Q.; Bao, X. CO-Tolerant PtRu@ h-BN/C Core–Shell Electrocatalysts for Proton Exchange Membrane Fuel Cells. Appl. Surf. Sci. 2018, 450, 244–250. [Google Scholar] [CrossRef]
  45. Byeon, J.; Kim, S.; Lee, S.; Jang, J.H.; Kim, S.-K.; Lee, J. Co-Tolerant Electrocatalysts for Hydrogen Fuel Cells: Fundamental Study-Based Design and Real-Life Applications. Chem. Eng. J. 2024, 493, 152626. [Google Scholar] [CrossRef]
  46. Li, X.; Duan, X.; Zhang, S.; Wang, C.; Hua, K.; Wang, Z.; Wu, Y.; Li, J.; Liu, J. Strategies for Achieving Ultra-Long ORR Durability—Rh Activates Interatomic Interactions in Alloys. Angew. Chem. Int. Ed. 2024, 63, e202400549. [Google Scholar] [CrossRef]
  47. Ehteshami, S.M.M.; Jia, Q.; Halder, A.; Chan, S.; Mukerjee, S. The Role of Electronic Properties of Pt and Pt Alloys for Enhanced Reformate Electro-Oxidation in Polymer Electrolyte Membrane Fuel Cells. Electrochim. Acta 2013, 107, 155–163. [Google Scholar] [CrossRef]
  48. Sun, C.; Wen, R.; Qin, Y.; Wang, L.; Wang, Y.; Dou, M.; Wang, F. Origin of Pt Site Poisoning by Impurities for Oxygen Reduction Reaction Catalysis: Tailored Intrinsic Activity of Pt Sites. ACS Appl. Energy Mater. 2023, 6, 5700–5709. [Google Scholar] [CrossRef]
  49. Shroti, N.; Daletou, M.K. The Pt–Co Alloying Effect on the Performance and Stability of High Temperature PEMFC Cathodes. Int. J. Hydrog. Energy 2022, 47, 16235–16248. [Google Scholar] [CrossRef]
  50. Prokop, M.; Domin, V.; Burianek, J.D.; Bystron, T.; Gatalo, M.; Pavko, L.; Hodnik, N.; Paidar, M.; Bouzek, K. Impact of the Catalyst Type and Dopant Composition on the Performance of High-Temperature PEM Fuel Cell; The Electrochemical Society, Inc.: Pennington, NJ, USA, 2023; p. 1759. [Google Scholar]
  51. Ugartemendia, A.; Peeters, K.; Ferrari, P.; de Cózar, A.; Mercero, J.M.; Janssens, E.; Jimenez-Izal, E. Doping Platinum with Germanium: An Effective Way to Mitigate the CO Poisoning. ChemPhysChem 2021, 22, 1603–1610. [Google Scholar] [CrossRef]
  52. Cai, Z.; Lu, Z.; Bi, Y.; Li, Y.; Kuang, Y.; Sun, X. Superior Anti-CO Poisoning Capability: Au-Decorated PtFe Nanocatalysts for High-Performance Methanol Oxidation. Chem. Commun. 2016, 52, 3903–3906. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, C.; Zhang, Q.; Hu, Y.; Hu, H.; Yang, J.; Yang, C.; Zhu, Y.; Tu, Z.; Wang, D. N-Doped Carbon Confined Ternary Pt2NiCo Intermetallics for Efficient Oxygen Reduction Reaction. Chin. Chem. Lett. 2024, 36, 110429. [Google Scholar] [CrossRef]
  54. Lee, H.; Park, S.; Kim, H. Preparation of CO-Tolerant PtRuNi/C Ternary Electrocatalyst Having a Composition Gradient Shell. Chem. Eng. J. 2021, 414, 128792. [Google Scholar] [CrossRef]
  55. Davletbaev, K.; Chougule, S.S.; Min, J.; Ko, K.; Kim, Y.; Choi, H.; Choi, Y.; Chavan, A.A.; Pak, B.; Rakhmonov, I.U. Effect of Heat Treatment on Structure of Carbon Shell-Encapsulated Pt Nanoparticles for Fuel Cells. Nanomaterials 2024, 14, 924. [Google Scholar] [CrossRef]
  56. Meyer, Q.; Yang, C.; Cheng, Y.; Zhao, C. Overcoming the Electrode Challenges of High-Temperature Proton Exchange Membrane Fuel Cells. Electrochem. Energy Rev. 2023, 6, 16. [Google Scholar] [CrossRef]
  57. Rivas, N.A.R.; Manjón, A.G.; Vega-Paredes, M.; Kim, S.-H.; Gault, B.; Jun, H.; Jung, C.; Berova, V.; Hengge, K.; Jurzinsky, T. Chemistry and Microstructure of C-Supported Ru Catalyst Nanoparticles: A Correlative Study. Ultramicroscopy 2023, 254, 113831. [Google Scholar] [CrossRef]
  58. Guan, J.; Yang, S.; Liu, T.; Yu, Y.; Niu, J.; Zhang, Z.; Wang, F. Intermetallic FePt@ PtBi Core–Shell Nanoparticles for Oxygen Reduction Electrocatalysis. Angew. Chem. 2021, 133, 22070–22075. [Google Scholar] [CrossRef]
  59. Tian, H.; Wang, X.; Ge, J.; Wan, H.; Ma, W.; Xie, G.; Ge, J. Pt-based Intermetallic Compound Catalysts for the Oxygen Reduction Reaction: From Problems to Recent Developments. J. Energy Chem. 2024, 9, 302–324. [Google Scholar] [CrossRef]
  60. Jang, I.; Ahn, M.; Lee, S.; Yoo, S.J. Surfactant Assisted Geometric Barriers on PtNi@ C Electrocatalyst for Phosphoric Acid Fuel Cells. J. Ind. Eng. Chem. 2022, 110, 198–205. [Google Scholar] [CrossRef]
  61. Min, J.; Chougule, S.S.; Sravani, B.; Ko, K.; Kim, Y.; Jung, N. A Bottom-up Approach to Solving Technical Challenges in Fuel Cell Systems through Innovative Catalyst Design. Curr. Opin. Electrochem. 2023, 39, 101257. [Google Scholar] [CrossRef]
  62. An, Z.; Li, H.; Zhang, X.; Xia, Z.; Zhang, H.; Chu, W.; Yu, S.; Wang, S.; Sun, G. Ultrastable and Phosphoric Acid-Resistant PtRhCu@ Pt Oxygen Reduction Electrocatalyst for High-Temperature Polymer Electrolyte Fuel Cells. ACS Catal. 2024, 14, 2572–2581. [Google Scholar] [CrossRef]
  63. Xu, M.; Xiao, Y.; Liu, T.; Wang, H.; Fan, B.; Dong, Y.; Wang, C.-A. Pt3Co/C Catalyst on the Nitrogen-Doped Carbon Support with High Catalytic Activity and Stability. ACS Appl. Energy Mater. 2024, 7, 2451–2459. [Google Scholar] [CrossRef]
  64. Dong, A.-Q.; Li, H.; Wu, H.-M.; Li, K.-X.; Shao, Y.-K.; Li, Z.-G.; Sun, S.-H.; Wang, W.-C.; Hu, W.-B. Weakening CO Poisoning over Size-and Support-Dependent Pt n/X-Graphene Catalyst (X= C, B, N, N= 1–6, 13). Rare Met. 2023, 42, 1138–1145. [Google Scholar] [CrossRef]
  65. Berova, V.; Manjón, A.G.; Paredes, M.V.; Schwarz, T.; Rivas, N.A.; Hengge, K.; Jurzinsky, T.; Scheu, C. Influence of the Shell Thickness on the Degradation of Ru@ Pt Core-Shell Catalysts in PEM Fuel Cells. J. Power Sources 2023, 554, 232327. [Google Scholar] [CrossRef]
  66. Chougule, S.S.; Jeffery, A.A.; Roy Chowdhury, S.; Min, J.; Kim, Y.; Ko, K.; Sravani, B.; Jung, N. Antipoisoning Catalysts for the Selective Oxygen Reduction Reaction at the Interface between Metal Nanoparticles and the Electrolyte. Carbon Energy 2023, 5, e293. [Google Scholar] [CrossRef]
  67. Huang, C.; Liu, H.; Tang, Y.; Lu, Q.; Chu, S.; Liu, X.; Shan, B.; Chen, R. Constructing Uniform Sub-3 Nm PtZn Intermetallic Nanocrystals via Atomic Layer Deposition for Fuel Cell Oxygen Reduction. Appl. Catal. B Environ. 2023, 320, 121986. [Google Scholar] [CrossRef]
  68. Yu, J.-H.; Singh, K.P.; Kim, S.-J.; Kang, T.-H.; Lee, K.-S.; Kim, H.; Ringe, S.; Yu, J.-S. Active and Stable PtP 2-Based Electrocatalysts Solve the Phosphate Poisoning Issue of High Temperature Fuel Cells. J. Mater. Chem. A 2023, 11, 6413–6427. [Google Scholar] [CrossRef]
  69. Zhang, W.; Xu, S.; Yang, Y.; Gao, Q.; Zhu, S.; Ding, K.; Li, Q.; Shi, P.; Min, Y. Rhombohedral Platinum-Copper Intermetallic Compound: A High Phosphate Tolerance Electrocatalyst for HT-PEMFC. Chem. Eng. J. 2024, 487, 150348. [Google Scholar] [CrossRef]
  70. Seselj, N.; Alfaro, S.M.; Bompolaki, E.; Cleemann, L.N.; Torres, T.; Azizi, K. Catalyst Development for High-Temperature Polymer Electrolyte Membrane Fuel Cell (HT-PEMFC) Applications. Adv. Mater. 2023, 35, 2302207. [Google Scholar] [CrossRef]
  71. Long, D.; Xie, Z.; Wang, M.; Chen, S.; Wei, Z. A Phosphate Tolerant Pt-Based Oxygen Reduction Catalyst Enabled by Synergistic Modulation of Alloying and Surface Modification. Chem. Commun. 2023, 59, 14277–14280. [Google Scholar] [CrossRef]
  72. Yang, T.; Yang, X.; Gao, N.; Tang, M.; Wang, X.; Liu, C.; Xing, W.; Ge, J. High Density Iridium Synergistic Sites Boosting CO-Tolerate Performance for PEMFC Anode. eScience 2024, 4, 100230. [Google Scholar] [CrossRef]
  73. Ekinci, A.; Büyükkanber, K.; Akdag, A.; Şahin, Ö. Improved Catalytic Activity in PdCo Nanocatalysts Synthesized via Ultrasonic Spray Method for PEMFC Applications. Int. J. Hydrog. Energy 2024, 92, 810–820. [Google Scholar] [CrossRef]
  74. Fu, H.; Chen, Y.; Lu, S.; Zhang, Z.; Zhu, T.; Li, H.; Lai, F.; Zhang, N.; Liu, T. Anti-Poisoned Oxygen Reduction Reaction by Rice-like Pd–Sb Nanoparticles. J. Mater. Chem. A 2024, 12, 9031–9037. [Google Scholar] [CrossRef]
  75. Takimoto, D.; Ohnishi, T.; Ayato, Y.; Mochizuki, D.; Sugimoto, W. Suppression of CO Adsorption on PtRu/C Catalysts Modified with Metallic Ruthenium Nanosheets. J. Electrochem. Soc. 2016, 163, F367. [Google Scholar] [CrossRef]
  76. Kobayshi, T.; Chida, Y.; Todoroki, N.; Wadayama, T. Enhanced Oxygen Reduction Reaction Activity on the Melamine-Modified Pt-High-Entropy Alloy Single-Crystal Lattice Stacking Surface. ACS Catal. 2024, 14, 11512–11521. [Google Scholar] [CrossRef]
  77. Hu, Y.; Shen, T.; Zhao, X.; Zhang, J.; Lu, Y.; Shen, J.; Lu, S.; Tu, Z.; Xin, H.L.; Wang, D. Combining Structurally Ordered Intermetallics with N-Doped Carbon Confinement for Efficient and Anti-Poisoning Electrocatalysis. Appl. Catal. B Environ. 2020, 279, 119370. [Google Scholar] [CrossRef]
  78. Liu, D.; Zhang, J.; Liu, D.; Li, T.; Yan, Y.; Wei, X.; Yang, Y.; Yan, S.; Zou, Z. N-Doped Graphene-Coated Commercial Pt/C Catalysts toward High-Stability and Antipoisoning in Oxygen Reduction Reaction. J. Phys. Chem. Lett. 2022, 13, 2019–2026. [Google Scholar] [CrossRef]
  79. Zhou, F.; Yan, Y.; Guan, S.; Guo, W.; Sun, M.; Pan, M. Solving Nafion Poisoning of ORR Catalysts with an Accessible Layer: Designing a Nanostructured Core-shell Pt/C Catalyst via a One-step Self-assembly for PEMFC. Int. J. Energy Res. 2020, 44, 10155–10167. [Google Scholar] [CrossRef]
  80. Stewart, D.W.; Scott, K.; Wain, A.J.; Rosser, T.E.; Brightman, E.; Macphee, D.; Mamlouk, M. The Role of Tungsten Oxide in Enhancing the Carbon Monoxide Tolerance of Platinum-Based Hydrogen Oxidation Catalysts. ACS Appl. Mater. Interfaces 2020, 12, 37079–37091. [Google Scholar] [CrossRef]
  81. Borbáth, I.; Salmanzade, K.; Pászti, Z.; Kuncser, A.; Radu, D.; Neaţu, Ş.; Tálas, E.; Sajó, I.E.; Olasz, D.; Sáfrán, G. Strategies to Improve CO Tolerance and Corrosion Resistance of Pt Electrocatalysts for Polymer Electrolyte Membrane Fuel Cells: Sn-Doping of the Mixed Oxide–Carbon Composite Support. Catal. Today 2024, 438, 114788. [Google Scholar] [CrossRef]
  82. Hu, Y.; Zhang, J.; Shen, T.; Lu, Y.; Chen, K.; Tu, Z.; Lu, S.; Wang, D. A Low-temperature Carbon Encapsulation Strategy for Stable and Poisoning-tolerant Electrocatalysts. Small Methods 2021, 5, 2100937. [Google Scholar] [CrossRef] [PubMed]
  83. Al-Tememy, M.G.H.; Devrim, Y. Development of Effective Bimetallic Catalyst for High-temperature PEM Fuel Cell to Improve CO Tolerance. Int. J. Energy Res. 2021, 45, 3343–3357. [Google Scholar] [CrossRef]
  84. Chatterjee, S.; Griego, C.; Hart, J.L.; Li, Y.; Taheri, M.L.; Keith, J.; Snyder, J.D. Free Standing Nanoporous Palladium Alloys as CO Poisoning Tolerant Electrocatalysts for the Electrochemical Reduction of CO2 to Formate. Acs Catal. 2019, 9, 5290–5301. [Google Scholar] [CrossRef]
  85. Álvarez-Gómez, J.M.; Varela, A.S. Review on Long-Term Stability of Electrochemical CO2 Reduction. Energy Fuels 2023, 37, 15283–15308. [Google Scholar] [CrossRef]
Figure 1. (a) PXRD analysis comparing 1 ML Pt/TiWN and Ptcomm before and after nitridation, (b) TEM image, particle size distribution, and STEM-EDX map of Pt/TiWN. Reproduced with permission. Copyright 2017, Wiley-VCH [26].
Figure 1. (a) PXRD analysis comparing 1 ML Pt/TiWN and Ptcomm before and after nitridation, (b) TEM image, particle size distribution, and STEM-EDX map of Pt/TiWN. Reproduced with permission. Copyright 2017, Wiley-VCH [26].
Crystals 15 00129 g001
Figure 2. XPS results for the core level shifts for W 4f and Pt 4f binding energies in (a) Pt/TiWN and (b) Pt/TiWC at different Pt coverages. (c) HOR polarization curves at 1600 rpm in 0.1 M HClO4 saturated with H2, with/without 1000 ppm CO. (d) CO stripping voltammograms at 1600 rpm in Ar-saturated 0.1 M HClO4 after saturation with CO at 0.025 V versus reversible hydrogen electrode (RHE). Reproduced with permission. Copyright 2017, Wiley-VCH [26].
Figure 2. XPS results for the core level shifts for W 4f and Pt 4f binding energies in (a) Pt/TiWN and (b) Pt/TiWC at different Pt coverages. (c) HOR polarization curves at 1600 rpm in 0.1 M HClO4 saturated with H2, with/without 1000 ppm CO. (d) CO stripping voltammograms at 1600 rpm in Ar-saturated 0.1 M HClO4 after saturation with CO at 0.025 V versus reversible hydrogen electrode (RHE). Reproduced with permission. Copyright 2017, Wiley-VCH [26].
Crystals 15 00129 g002
Figure 3. RRDE electrochemical performance of FePhen@MOF-ArNH3 and Tanaka Pt/C. (a) RDE ORR polarization plots with/without 100 mM H3PO4. (b) mass-transport corrected Tafel plots with/without 10 mM H3PO4 in O2 saturated electrolyte (0.1 M HClO4) at 20 mV/s and 1600 rpm at room temperature (22 °C). Anion adsorption investigation of FePhen@MOF-ArNH3 at 0.3 V vs. RHE in N2 saturated 0.1 M HClO4 with and without 100 mM H3PO4. (c) Fe K-edge XANES with Δ μ = μ   0.9   V μ   ( 0.3 V )  (inset). (d) Fourier transform EXAFS, theoretical Δ μ signatures calculated by FEFF 8 (e) of the illustrated PO4 adsorption on the Fe6 cluster, (f) and atop and fcc-inverted PO4 adsorption on the Pt6 cluster. Reprinted with permission from [32]. Copyright 2018 American Chemical Society.
Figure 3. RRDE electrochemical performance of FePhen@MOF-ArNH3 and Tanaka Pt/C. (a) RDE ORR polarization plots with/without 100 mM H3PO4. (b) mass-transport corrected Tafel plots with/without 10 mM H3PO4 in O2 saturated electrolyte (0.1 M HClO4) at 20 mV/s and 1600 rpm at room temperature (22 °C). Anion adsorption investigation of FePhen@MOF-ArNH3 at 0.3 V vs. RHE in N2 saturated 0.1 M HClO4 with and without 100 mM H3PO4. (c) Fe K-edge XANES with Δ μ = μ   0.9   V μ   ( 0.3 V )  (inset). (d) Fourier transform EXAFS, theoretical Δ μ signatures calculated by FEFF 8 (e) of the illustrated PO4 adsorption on the Fe6 cluster, (f) and atop and fcc-inverted PO4 adsorption on the Pt6 cluster. Reprinted with permission from [32]. Copyright 2018 American Chemical Society.
Crystals 15 00129 g003
Figure 4. HOR polarization curves of commercial Pt/C (a) in 0.5 M H2SO4, (b) and 0.5 M H2SO4 + 10 mM DAcPy. (c) XPS of Pt foils poisoned by 10 μM Na2S with/with DAcPy adaptation. (d) DFT calculation of adsorption structure of DAcY on Pt(111). (e) The proposed model of the reduced form of DAcPy adsorbed on Pt(111) from the side. Reproduced from Ref. [36] with permission from the Royal Society of Chemistry.
Figure 4. HOR polarization curves of commercial Pt/C (a) in 0.5 M H2SO4, (b) and 0.5 M H2SO4 + 10 mM DAcPy. (c) XPS of Pt foils poisoned by 10 μM Na2S with/with DAcPy adaptation. (d) DFT calculation of adsorption structure of DAcY on Pt(111). (e) The proposed model of the reduced form of DAcPy adsorbed on Pt(111) from the side. Reproduced from Ref. [36] with permission from the Royal Society of Chemistry.
Crystals 15 00129 g004
Figure 5. (a) TEM (b) and HRTEM image of PtRu@h-BN/C comparing PtRu/C and PtRu@h-BN/C with (c) HS-LEIS spectra and (d,e) UV–Vis absorption spectra of benzothiazole in solution. Reproduced with permission. Copyright 2018, Elsevier [44].
Figure 5. (a) TEM (b) and HRTEM image of PtRu@h-BN/C comparing PtRu/C and PtRu@h-BN/C with (c) HS-LEIS spectra and (d,e) UV–Vis absorption spectra of benzothiazole in solution. Reproduced with permission. Copyright 2018, Elsevier [44].
Crystals 15 00129 g005
Figure 6. H2-O2 fuel cell test in H2 with/without introduction of 30 ppm CO and 25% CO2 in the single cell. Polarization curves (a) with the PtRu/C anode and PtRu@h-BN/C anode (b) and galvanostatic curves at 0.2268 A cm−2. (c) Current densities at different cell voltages (d) and current density decrement. (e) Effect of CO concentration at 0.2268 A cm−2. Reproduced with permission. Copyright 2018, Elsevier [44].
Figure 6. H2-O2 fuel cell test in H2 with/without introduction of 30 ppm CO and 25% CO2 in the single cell. Polarization curves (a) with the PtRu/C anode and PtRu@h-BN/C anode (b) and galvanostatic curves at 0.2268 A cm−2. (c) Current densities at different cell voltages (d) and current density decrement. (e) Effect of CO concentration at 0.2268 A cm−2. Reproduced with permission. Copyright 2018, Elsevier [44].
Crystals 15 00129 g006
Figure 7. (a) Synthesis process of surface Au-decorated PtFe nanocrystals, (b,c) TEM (scale bar: 40 nm) and HRTEM (scale bar: 2 nm) images of the PtFe and PtFeAu nanocrystals. (d) XRD, (e) XPS curves, (f) CO stripping curves of Pt/C, PtFe and PtFeAu nanocrystals, (g) chronoamperometry curves, and (h) cyclic voltammetry after ADT. Reproduced from Ref. [52] with permission from the Royal Society of Chemistry.
Figure 7. (a) Synthesis process of surface Au-decorated PtFe nanocrystals, (b,c) TEM (scale bar: 40 nm) and HRTEM (scale bar: 2 nm) images of the PtFe and PtFeAu nanocrystals. (d) XRD, (e) XPS curves, (f) CO stripping curves of Pt/C, PtFe and PtFeAu nanocrystals, (g) chronoamperometry curves, and (h) cyclic voltammetry after ADT. Reproduced from Ref. [52] with permission from the Royal Society of Chemistry.
Crystals 15 00129 g007
Figure 8. (a) MEA performance of the Pt1Ru1Ni0.75/C-PDA at different heating temperatures. (b) XPS of Pt 4f. Reproduced with permission. Copyright 2021, Elsevier [54].
Figure 8. (a) MEA performance of the Pt1Ru1Ni0.75/C-PDA at different heating temperatures. (b) XPS of Pt 4f. Reproduced with permission. Copyright 2021, Elsevier [54].
Crystals 15 00129 g008
Figure 9. (a) Synthesis procedure of FePt@PtBi. (b) XRD of FePt and FePt@PtBi. (c) HAADF-STEM image of L10-FePt (inset: fast Fourier transform of interior). Chronoamperometric responses of commercial Pt/C and FePt@PtBi adjusted electrodes (d) by adding 0.5 M CH3OH and (e) CO into an O2 saturated 0.1 M HClO4 solution. Reproduced with permission. Copyright 2021, Wiley-VCH [58].
Figure 9. (a) Synthesis procedure of FePt@PtBi. (b) XRD of FePt and FePt@PtBi. (c) HAADF-STEM image of L10-FePt (inset: fast Fourier transform of interior). Chronoamperometric responses of commercial Pt/C and FePt@PtBi adjusted electrodes (d) by adding 0.5 M CH3OH and (e) CO into an O2 saturated 0.1 M HClO4 solution. Reproduced with permission. Copyright 2021, Wiley-VCH [58].
Crystals 15 00129 g009
Figure 10. (a) Synthesis procedure. (b) PtNi@Cx NPs carbon shell thickness according to different C content. Reproduced with permission. Copyright 2022, Elsevier [60].
Figure 10. (a) Synthesis procedure. (b) PtNi@Cx NPs carbon shell thickness according to different C content. Reproduced with permission. Copyright 2022, Elsevier [60].
Crystals 15 00129 g010
Figure 11. ORR tests with phosphoric acid. (a) Representation of the geometric barrier-induced generation of ORR. ORR LSV curves for the (b) PtNi@OAmx series and (c) PtNi@Cx series in 0.1 M H3PO4 + 0.1 M HClO4. (d) Geometric effect of PtNi@C2 and the effect of PO43− poisoning. (e) Half-wave potential enhancement of PtNi@C2 in only PA. Reproduced with permission. Copyright 2022, Elsevier [60].
Figure 11. ORR tests with phosphoric acid. (a) Representation of the geometric barrier-induced generation of ORR. ORR LSV curves for the (b) PtNi@OAmx series and (c) PtNi@Cx series in 0.1 M H3PO4 + 0.1 M HClO4. (d) Geometric effect of PtNi@C2 and the effect of PO43− poisoning. (e) Half-wave potential enhancement of PtNi@C2 in only PA. Reproduced with permission. Copyright 2022, Elsevier [60].
Crystals 15 00129 g011
Figure 12. Schematic for the strategy to control the carbon shell of Pt-based NPs for the molecular sieve layer effect. Reproduced with permission. Copyright 2023, Wiley-VCH [66].
Figure 12. Schematic for the strategy to control the carbon shell of Pt-based NPs for the molecular sieve layer effect. Reproduced with permission. Copyright 2023, Wiley-VCH [66].
Crystals 15 00129 g012
Figure 13. Electrochemical properties of (a) Pt@MSL7-0, (b) Pt@MSL7-5, (c) Pt@MSL7-10, and (d) Pt@MSL7-20. (e) E1/2 of the catalysts in a 0.1 M HClO4 solution with/without 0.1 M H3PO4. (f) Mass activities and molecular sieving effect values in a 0.1 M HClO4 solution with 0.1 M H3PO4. Reproduced with permission. Copyright 2023, Wiley-VCH [66].
Figure 13. Electrochemical properties of (a) Pt@MSL7-0, (b) Pt@MSL7-5, (c) Pt@MSL7-10, and (d) Pt@MSL7-20. (e) E1/2 of the catalysts in a 0.1 M HClO4 solution with/without 0.1 M H3PO4. (f) Mass activities and molecular sieving effect values in a 0.1 M HClO4 solution with 0.1 M H3PO4. Reproduced with permission. Copyright 2023, Wiley-VCH [66].
Crystals 15 00129 g013
Figure 14. (a) Synthesis procedure of PtP2/C-ASP and PtP2/C-800. PtP2/C-ASP: (b) TEM, (c) HR-TEM, and (d) FFT patterns of the red square area in (c). (e) HAADF and corresponding EDX of (c). PtP2/C-800: (f) TEM, (g) HR-TEM, and (h) FFT patterns of the red square area in (g). (i,j) Inverse FFT images of the red square in (g), where (i) shows the distribution of 220 planes and (j) shows the presence of 220 planes. (k) HAADF and corresponding EDX of (g). Reproduced from Ref. [68] with permission from the Royal Society of Chemistry.
Figure 14. (a) Synthesis procedure of PtP2/C-ASP and PtP2/C-800. PtP2/C-ASP: (b) TEM, (c) HR-TEM, and (d) FFT patterns of the red square area in (c). (e) HAADF and corresponding EDX of (c). PtP2/C-800: (f) TEM, (g) HR-TEM, and (h) FFT patterns of the red square area in (g). (i,j) Inverse FFT images of the red square in (g), where (i) shows the distribution of 220 planes and (j) shows the presence of 220 planes. (k) HAADF and corresponding EDX of (g). Reproduced from Ref. [68] with permission from the Royal Society of Chemistry.
Crystals 15 00129 g014
Figure 15. Computational simulations. (a) Adsorption free energy of H2PO4− anions on the (111) facets of Pt and PtP2 as a function of excess surface charge density. (b) The Pourbaix diagram of the surface. The (*) mark indicates the site where each chemical species is adsorbed on the catalyst surface. Reproduced from Ref. [68] with permission from the Royal Society of Chemistry.
Figure 15. Computational simulations. (a) Adsorption free energy of H2PO4− anions on the (111) facets of Pt and PtP2 as a function of excess surface charge density. (b) The Pourbaix diagram of the surface. The (*) mark indicates the site where each chemical species is adsorbed on the catalyst surface. Reproduced from Ref. [68] with permission from the Royal Society of Chemistry.
Crystals 15 00129 g015
Figure 16. Ru(ns)-PtRu/C: (a) HR-SEM, (b) TEM, (c) and corresponding PtRu particle distribution obtained from (b), and (d) XRD patterns: PtRu/C, RuO2.1 ns-PtRu/C, and Ru(ns)-PtRu/C. Reproduced with permission. Copyright 2016, IOP [75].
Figure 16. Ru(ns)-PtRu/C: (a) HR-SEM, (b) TEM, (c) and corresponding PtRu particle distribution obtained from (b), and (d) XRD patterns: PtRu/C, RuO2.1 ns-PtRu/C, and Ru(ns)-PtRu/C. Reproduced with permission. Copyright 2016, IOP [75].
Crystals 15 00129 g016
Figure 17. Hydrodynamic voltammograms saturated with (a) pure H2, (b) 300 ppm CO/H2 for PtRu/C (black), RuO2.1 ns-PtRu/C (blue), and Ru(ns)-PtRu/C (red) with 400 rpm in 0.1 HCIO. Chronoamperograms saturated with (c) pure H2, (d) 300 ppm He/CO for PtRu/C (black), RuOz,ns-PtRu/C (blue), and Ru(ns)-PtRu/C (red) with 400 rpm in 0.1 M HCIO at 20 mV vs. RHE., and (e) normalized current of (b). Reproduced with permission. Copyright 2016, IOP [75].
Figure 17. Hydrodynamic voltammograms saturated with (a) pure H2, (b) 300 ppm CO/H2 for PtRu/C (black), RuO2.1 ns-PtRu/C (blue), and Ru(ns)-PtRu/C (red) with 400 rpm in 0.1 HCIO. Chronoamperograms saturated with (c) pure H2, (d) 300 ppm He/CO for PtRu/C (black), RuOz,ns-PtRu/C (blue), and Ru(ns)-PtRu/C (red) with 400 rpm in 0.1 M HCIO at 20 mV vs. RHE., and (e) normalized current of (b). Reproduced with permission. Copyright 2016, IOP [75].
Crystals 15 00129 g017
Figure 18. Scheme of O-Pt-Fe@NC/C synthesis. Reproduced with permission. Copyright 2020, Elsevier [77].
Figure 18. Scheme of O-Pt-Fe@NC/C synthesis. Reproduced with permission. Copyright 2020, Elsevier [77].
Crystals 15 00129 g018
Figure 19. (a) XRD patterns of O-Pt-Fe@NC/C, D-Pt-Fe/C, and Pt/C with ordered intermetallic Pt-Fe (blue vertical line, standard PDF card # 0106346) and Pt (black, #00-004-0802). (b) TEM image of O-Pt-Fe@NC/C. (c,d) HR-TEM images of Fe@NC/C and the corresponding FFT pattern. (ej) HAADF-STEM and EDS elemental mapping of Pt, Fe, C, N, and the composite of C and N. Reproduced with permission. Copyright 2020, Elsevier [77].
Figure 19. (a) XRD patterns of O-Pt-Fe@NC/C, D-Pt-Fe/C, and Pt/C with ordered intermetallic Pt-Fe (blue vertical line, standard PDF card # 0106346) and Pt (black, #00-004-0802). (b) TEM image of O-Pt-Fe@NC/C. (c,d) HR-TEM images of Fe@NC/C and the corresponding FFT pattern. (ej) HAADF-STEM and EDS elemental mapping of Pt, Fe, C, N, and the composite of C and N. Reproduced with permission. Copyright 2020, Elsevier [77].
Crystals 15 00129 g019
Figure 20. (a) CO stripping result with 50 mM NaHSO3. (b) LSV curves of 0-Pt-Fe@NC/C and Pt/C in 0.1 M HCIO4, with and without the addition of 0.2 M H3PO4. (c) High-temperature MEA test of previous catalysts. Reproduced with permission. Copyright 2020, Elsevier [77].
Figure 20. (a) CO stripping result with 50 mM NaHSO3. (b) LSV curves of 0-Pt-Fe@NC/C and Pt/C in 0.1 M HCIO4, with and without the addition of 0.2 M H3PO4. (c) High-temperature MEA test of previous catalysts. Reproduced with permission. Copyright 2020, Elsevier [77].
Crystals 15 00129 g020
Figure 21. Structural analysis: (a) TEM image of WO3 nanoparticles, (b) TEM image of the Pt/C catalyst (Tanaka), (c) HR-SEM images of Pt/C/WO3, (d) XPS spectra for Pt/C and Pt/C/WO3 electrodes, and (e) proposed catalytic reaction pathway. Reprinted with permission from [80]. Copyright 2020 American Chemical Society.
Figure 21. Structural analysis: (a) TEM image of WO3 nanoparticles, (b) TEM image of the Pt/C catalyst (Tanaka), (c) HR-SEM images of Pt/C/WO3, (d) XPS spectra for Pt/C and Pt/C/WO3 electrodes, and (e) proposed catalytic reaction pathway. Reprinted with permission from [80]. Copyright 2020 American Chemical Society.
Crystals 15 00129 g021
Figure 22. In situ Raman spectroscopy: (a) Pt/C/WO3 electrodes (0.1M H2SO4 solution) and (b) peak positions and integrated areas (at ~810 cm−1) from the in situ spectra in (a). ATR-IR spectra: (c) Pt/C catalyst and (d) Pt/C/WO3 catalyst after the adsorption of CO in Ar-purged 0.1 M H2SO4 at a range of applied potentials. Reprinted with permission from [80]. Copyright 2020 American Chemical Society.
Figure 22. In situ Raman spectroscopy: (a) Pt/C/WO3 electrodes (0.1M H2SO4 solution) and (b) peak positions and integrated areas (at ~810 cm−1) from the in situ spectra in (a). ATR-IR spectra: (c) Pt/C catalyst and (d) Pt/C/WO3 catalyst after the adsorption of CO in Ar-purged 0.1 M H2SO4 at a range of applied potentials. Reprinted with permission from [80]. Copyright 2020 American Chemical Society.
Crystals 15 00129 g022
Table 1. Comparison of the CO tolerance of different electrocatalysts in a fuel cell. Reproduced with permission. Copyright 2021, Elsevier [54].
Table 1. Comparison of the CO tolerance of different electrocatalysts in a fuel cell. Reproduced with permission. Copyright 2021, Elsevier [54].
Anode CatalystMetal Loading (mg cm−2)MembraneOperating Temperature (°C)CO Gas (ppm)Decreasing Ratio at 0.6 V (%)
PtSn/C0.3N112801030
PtPd/C0.3N212801050
PtMo/C0.5N212702527
PtRu/C0.2N117801027
PtRu/C0.4N212751030
PtNi/C1.0N212753057
PtRuNi/C0.4N212751011
Table 2. HOR activity measured before and after the durability test of catalysts from chronoamperometry (at 20 mV vs. RHE). Reproduced with permission. Copyright 2016, IOP [75].
Table 2. HOR activity measured before and after the durability test of catalysts from chronoamperometry (at 20 mV vs. RHE). Reproduced with permission. Copyright 2016, IOP [75].
HOR Activity/A (g-PtRu)−1
CatalystInitialAfter 1000 CyclesAfter 3000 Cycles
PtRu/C896455
RuO2.1ns-PtRu/C1118675
Ru(ns)-PtRu/C12411297
Table 3. A comparison of previously proposed mitigation strategies categorized by composition, method, poisoning molecules, mechanism, and performance.
Table 3. A comparison of previously proposed mitigation strategies categorized by composition, method, poisoning molecules, mechanism, and performance.
#Catalyst CompositionSynthesis/Key MethodTarget PoisonKey Mechanism for Poison Mitigation1. Performance Gains
/2. Highlights
Reference
1Pt/TiWNNitridation of Pt/TiWC (800 °C in NH3)COCore–shell synergy (Pt shell on transition-metal nitride), lowered CO binding energy1. Rapid HOR recovery at ~0.1 V in presence of CO
2. Higher CO tolerance than commercial Pt (Ptcomm)
Aaron Garg et al. [26]
2FePhen@MOF-ArNH3 (Fe-based N-PGM)MOF-based pyrolysis (Fe-Phen@MOF) with ammonia treatmentPhosphate (H3PO4)Subsurface Fe encapsulation; minimal phosphate adsorption on Fe sites1. Negligible Tafel slope changes under phosphate
2. Enhanced ORR stability and reduced anion poisoning
Kara et al. [32]
3Pt/C + DAcPyIn situ adsorption/reduction of 2,6-diacetylpyridine (DAcPy)CO, H2S“Molecular canopy” (tridentate binding); size-selective blocking of CO/H2S while allowing H2 access1. Nearly complete prevention of sulfur poisoning
2. No detectable S on Pt surface by XPS
Tao Wang et al. [36]
4PtRu@h-BN/CAmmonia borane (AB) polymerization + pyrolysis; formation of few-layer h-BN shellCOh-BN encapsulation partially blocks CO from PtRu surface1. Significantly improved CO tolerance in single-cell tests
2. Only ~20% drop vs. ~60% drop with PtRu/C
Mengmeng Sun et al. [44]
5PtFeAuOne-pot solvothermal synthesis + partial galvanic replacementCOAu decoration shifts electron density away from Pt: => weaker CO binding1. Largest negative shift in CO stripping
2. High ECSA and stable chronoamperometry
Zhao et al. [52]
6Pt1Ru1Ni0.75/C-PDA-650High-temp (600–800 °C) heat treatment + polydopamine coating to form gradient shellsCOStrong ligand effect (Ni and Ru → Pt): downshift of Pt-CO binding energy1. Only 11% performance loss at 0.6 V under CO
2. Superior CO tolerance compared with PtRu/C
Lee et al. [54]
7FePt@PtBiLeach-embed-rearrangement to create PtBi shells on ordered FePt intermetallic coreCO, CH3OHCompressive strain + larger Bi atoms: d-band shift, weakened adsorbate binding1. ~70% higher current retention vs. commercial Pt/C
2. High methanol and CO tolerance
Jingyu et al. [58]
8PtNi@CxOleylamine carbonization + H2 reduction forming carbon shells of variable thicknessPhosphate (H3PO4)Geometric barrier: carbon shell prevents phosphate adsorption on PtNi surface1. ORR activity quickly recovers even in phosphoric acid
2. Maintains active sites for prolonged operation
Jang et al. [60]
9Pt@MSL (Pt with carbon molecular sieve layer)Annealing in different gas atmospheres to form controllable carbon shellsPhosphate (H3PO4)Selective O2 permeation via dense carbon shell; blocks anion diffusion1. Increased ORR activity with high molecular-sieving effect
2. Up to ~90% poison-blocking efficiency
Sourabh et al. [66]
10PtP2Trioctylphosphine (TOP) route; Pt salt converted to PtP2 phase (amorphous or crystalline)Phosphate (H3PO4)Strong P−HPO4 interactions preserve Pt sites: stable at higher potentials1. Maintains ORR activity under phosphate poisoning
2. Protective phosphoric oxide oligomers on PtP2 surface
Jeong-Hoon et al. [68]
11Ru(ns)-PtRu/CMetallization of RuO2.1 nanosheets to form metallic Ru nanosheets on PtRu/CCOMetallic Ru nanosheets in close contact with Pt: suppressed CO adsorption1. 1.3× higher HOR current after 5 h vs. PtRu/C
2. Superior CO tolerance and durability
Daisuke et al. [75]
12O-Pt-Fe@NC/COne-step pyrolysis: metal salts + dicyandiamide => N-doped carbon shell + ordered PtFe alloyCO, SOx, POxN doping + thin carbon encapsulation: electron modulation + physical blocking1. Excellent multi-poison tolerance (CO, SO32−, PO43−)
2. Enhanced performance as HT-PEMFCs cathode
Yezhou Hu et al. [77]
13Pt/C + WO3Simple physical mixing of Pt/C with WO3 nanoplateletsCOW5+/W6+ redox cycle in WO3: catalytic CO oxidation (tungsten bronze)1. Improved CO removal at ~0.3 VRHE
2. Potential for direct methanol fuel cell applications
Douglas et al. [80]
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

Choi, S.; Jang, I.; Lee, S. Advanced Strategies for Mitigating Catalyst Poisoning in Low and High Temperature Proton Exchange Membrane Fuel Cells: Recent Progress and Perspectives. Crystals 2025, 15, 129. https://doi.org/10.3390/cryst15020129

AMA Style

Choi S, Jang I, Lee S. Advanced Strategies for Mitigating Catalyst Poisoning in Low and High Temperature Proton Exchange Membrane Fuel Cells: Recent Progress and Perspectives. Crystals. 2025; 15(2):129. https://doi.org/10.3390/cryst15020129

Chicago/Turabian Style

Choi, Suyeon, Injoon Jang, and Sehyun Lee. 2025. "Advanced Strategies for Mitigating Catalyst Poisoning in Low and High Temperature Proton Exchange Membrane Fuel Cells: Recent Progress and Perspectives" Crystals 15, no. 2: 129. https://doi.org/10.3390/cryst15020129

APA Style

Choi, S., Jang, I., & Lee, S. (2025). Advanced Strategies for Mitigating Catalyst Poisoning in Low and High Temperature Proton Exchange Membrane Fuel Cells: Recent Progress and Perspectives. Crystals, 15(2), 129. https://doi.org/10.3390/cryst15020129

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop