Next Article in Journal
Fast Kinetic Response and Efficient Removal of Methyl Blue and Methyl Green Dyes by Functionalized Multiwall Carbon Nanotubes Powered with Iron Oxide Nanoparticles and Citrus reticulata Peel Extract
Previous Article in Journal
Bipolar Switching Properties and Reaction Decay Effect of BST Ferroelectric Thin Films for Applications in Resistance Random Access Memory Devices
Previous Article in Special Issue
Mesoporous Nitrogen-Doped Carbon Support from ZIF-8 for Pt Catalysts in Oxygen Reduction Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Rare Earth Ce/CeO2 Electrocatalysts: Role of High Electronic Spin State of Ce and Ce3+/Ce4+ Redox Couple on Oxygen Reduction Reaction

Natural Science Research Institute, College of Natural Sciences, Keimyung University, 1095 Dalgubeol-daero, Daegu 42601, Republic of Korea
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(8), 600; https://doi.org/10.3390/nano15080600
Submission received: 17 March 2025 / Revised: 9 April 2025 / Accepted: 11 April 2025 / Published: 14 April 2025
(This article belongs to the Collection Micro/Nanoscale Open Framework Materials (OFMs))

Abstract

With unique 4f electronic shells, rare earth metal-based catalysts have been attracting tremendous attention in electrocatalysis, including oxygen reduction reaction (ORR). In particular, atomically dispersed Ce/CeO2-based catalysts have been explored extensively due to several unique features. This review article provides a comprehensive understanding of (i) the significance of the effect of Ce high-spin state on ORR activity enhancement on the Pt and non-pt electrocatalysts, (ii) the spatially confining and stabilizing effect of ceria on the generation of atomically dispersed transition metal-based catalysts, (iii) experimental and theoretical evidence of the effect of Ce3+ ↔ Ce4+ redox pain on radical scavenging, (iv) the effect of the Ce 4f electrons on the d-band center and electron transfer between Ce to the N-doped carbon and transition metal catalysts for enhanced ORR activity, and (v) the effect of Pt/CeO2/carbon heterojunctions on the stability of the Pt/CeO2/carbon electrocatalyst for ORR. Among several strategies of synthesizing Ce/CeO2 electrocatalysts, the metal–organic framework (MOF)-derived catalysts are being perused extensively due to the tendency of Ce to readily coordinate with O- and N-containing ligands, which upon undergoing pyrolysis, results in the formation of high surface area, porous carbon networks with atomically dispersed metallic/clusters/nanoparticles of Ce active sites. This review paper provides an overview of recent advancements regarding Ce/CeO2-based catalysts derived from the MOF precursor for ORR in fuel cells and metal–air battery applications and we conclude with insights into key issues and future development directions.

Graphical Abstract

1. Introduction

Increasing global warming due to the extensive use of carbon-based fossil fuels escalating environmental concerns has led to a demand for alternative technologies to meet the growing demand of energy [1,2]. The electrochemical oxygen reduction reaction (ORR) is a central, cathodic electrochemical reaction which plays a critical role in various energy conversion devices like fuel cells and metal–air batteries like Zn–air and Al–air batteries [3,4,5]. In these devices, the ORR occurs on the cathodes, where the electrocatalysts reduce the atmospheric oxygen into H2O/OH. Due to the process being inherently sluggish and requiring multiple steps, kinetically controlled ORR often requires efficient electrocatalysts to enhance effective electron transfer and the reduction of O2 [6,7]. In general, the best state-of-the-art electrocatalysts for ORR are Pt nanoparticles supported by high surface-area carbon support (Pt/C) [8]. Despite the best ORR activity of Pt/C catalysts, successful commercial application of the energy conversion devices is severely hindered due to the high cost and scarcity of Pt, together with the poor stability of the Pt/C catalysts [9]. It is well known that a commercial Pt/C catalyst undergoes a variety of degradation pathways, such as electrochemical carbon corrosion, Pt nanoparticle dissolution and agglomeration, catalyst poisoning in the presence of gasses such as CO and SOx, etc. Pt nanoparticles detach from the carbon support due to the loss of supported solid carbon due to carbon corrosion [10]. Consequently, research is being conducted to develop effective and stable electrocatalysts, which incorporate stable supports to enhance catalyst stability, reducing the Pt loading, as well as cost-effective catalysts utilizing transition and inner-transition metals, including lanthanides, to diminish dependence on costly platinum-based materials [9,11]. To substantially decrease the quantity of Pt loading, many solutions have been implemented, such as alloying Pt with other metals. The noble Pt metal have been alloyed with various transition metals (such as Fe, Co, Ni, Mn, Cu, etc.) and inner-transition metals of lanthanides (such as, Ce, La, Y, Gd, Dy, etc.) and it was found that Pt-alloy catalysts exhibit excellent ORR activity over traditional Pt/C catalysts [12,13,14]. However, even the Pt-alloys catalysts also suffer from issues like dissolution, Ostwald ripening, and nanoparticle sintering. In addition, the Pt-alloy catalysts show poor stability, suggesting the need for active and stable electrocatalysts for ORR both in acidic and alkaline electrolytes [15].
Though basic and advanced research on Pt, Pt-alloy, and non-precious metal catalysts with active sites composed of Mx-Ny-C (M=TM, such as Fe, Co, Mn, Cu, Ni, etc.) is so popular, with tens to thousands of papers being published each month [16,17], it is very recently that research exploring the possibility of using inner-transition metals of lanthanides from the f-block series has been attracting tremendous attention [18]. Rare earth elements are composed of 17 elements: lanthanum (La), cerium (Ce), praseodymium (Pr), neodymium (Nd), promethium (Pm), samarium (Sm), europium (Eu), gadolinium (Gd), terbium (Tb), dysprosium (Dy), holmium (Ho), erbium (Er), thulium (Tm), ytterbium (Yb), lutetium (Lu), scandium (Sc), and yttrium (Y). While rare earth elements are not great electrocatalysts on their own, even in trace amounts they can change the characteristics of active sites composed of other noble and transition metals [19]. Rare earth metals find several applications in day-to-day life, for example, several rare earth metal oxides have been used as polishing powders in the making of glass products with various rare earth metal oxides. The quality of gasoline is also improved by using catalysts that contain rare earth elements [20]. Cerium oxide is used in the three-way catalytic converter of waste gas treatment systems to change toxic waste gas into a harmless gas that people can breathe in. Due to their use and applications in various industrial products, today, rare earth elements are termed “Industrial Gold” or “Industrial Vitamins” [21]. Rare-earth (RE) elements exhibit various oxidation states such as RE2+, RE3+, and RE4+, of which RE3+ is the most common. The presence of these various oxidation states suggests that RE with empty, half-filled, and fully filled electron energy levels tends to be more stable, as suggested by Hende’s rule [21]. Rare earth elements are specially characterized by the presence of unique f-orbitals. In general, 4f electrons are inert, meaning they do not participate in chemical bonding with other elements and are part of the ionic core of the element. However, the positioning and incomplete filling of 4f electrons gives rare earth elements unique physical and chemical properties that are suitable for electrocatalytic applications [22,23]. In addition, the rare earth metallic elements exhibit strong coordination ability with reactants’ molecular orbitals and as rare earth metal oxides have the features of oxygen affinity and oxygen vacancies, in various catalytic reactions, rare earth oxides act as co-catalysts by transferring the stored oxygen species and hence promote higher catalytic activity of the main catalysts [24,25].
Among the various RE elements, cerium (Ce)-based catalysts, either in an atomically dispersed state or in the form of Ce-oxide (CeO2), has attracted considerable attention for various electrochemical reactions such as CO oxidation, NO reduction, water–gas-shift reaction, and for ORR [26]. The rich oxygen vacancies and affinity and facile interchange of the oxidation state of Ce3+ and Ce4+ facilitates the ORR. In view of the fuel cell reaction, the ability to perform the interexchange of Ce3+/Ce4+ in the oxidation state is so important; for this, it is essential that Ce inactivates the hydrogen peroxide, formed from the 2-electron O2 reduction reaction. CeO2 has high stability under acidic conditions and CeO2 also improves the corrosion-resistant ability of carbon in acidic electrolytes [27]. CeO2 also exhibits strong interaction with noble metals such as Pt and non-precious metals, thus enhancing the metal/metal oxide interactions in promoting the ORR by facilitating optimal adsorption of O2 and ORR intermediates [28]. Furthermore, the 4f orbit of the Ce atom allows for both unoccupied and one-electron occupied states, and the 4f shell of the Ce atom is suitable for electron sharing and bonding [29]. Furthermore, ceria plays a key function in promoting the dispersion of noble metals; ceria can significantly enhance O2 adsorption and ORR kinetics. Ceria-modified carbon compounds have been shown to serve as very efficient catalysts for a direct 4-electron reduction in ORR [30,31]. These distinct features of ceria motivated us to search for the possibility of Ce/CeO2 materials being employed as ORR catalysts. This review paper provides an overview of the research that had been carried out recently on the applications of Ce and CeO2 for ORR for fuel cells and metal–air batteries. The first part of this review paper describes the characteristics and properties of Ce/CeO2 and the second part of the review paper describes the ORR kinetics of the electrocatalysts of Ce/CeO2 alone or in combination with Pt and non-Pt metallic active sites, derived from metal–organic framework (MOF) precursors. MOFs are a class of porous materials comprising metal ions/organic linkers; after undergoing pyrolysis, they result in the formation of a porous carbon network with atomically dispersed metallic/clusters/nanoparticles of active sites [32].
MOF-derived catalysts have recently gained tremendous interests regarding their application in various fields, including gas storage and separation, luminescence, magnetism, biology and medicine, biomass conversion, photovoltaic cells, water purification, and catalysis; this includes ORR [33]. Due to the tendency of RE metals to readily coordinate with O- and N-containing ligands, particularly -COOH ligands, together with their high coordination numbers and varied coordination modes, the synthesis of novel structural metal–organic frameworks (MOFs) is feasible. Nevertheless, owing to steric hindrance, rare earth elements are unable to fully coordinate with ligands, resulting in one or more coordination sites being available to interact with solvent molecules, thereby forming end groups. Upon heat treatment of the synthesized rare earth MOFs, solvent molecules are expelled from the rare earth MOF, revealing coordinatively unsaturated sites within the structure that can serve as active Lewis’s acid/base centers, that can interact with reactant species like O2 to promote the ORR [34]. Rare earth ions possess unsaturated 4f orbitals and readily coordinate with the lone electron pairs of organic groups. The ligands most commonly used for the synthesis of RE MOFs include terephthalic acid; 1,4-benzene dicarboxylic acid (BDC); trimesic acid (TMA), also known as 1,3,5-benzenetricarboxylic acid (H3BTC); pyridine-2,4-dicarboxylic acid (2,4-pydc); 2-aminoterephthalic acid (BDC-NH2); zeolite imidazole framework (ZIF); Zn-triazolates; Prussian blue; and various other organic ligands [35].
To briefly conclude, this review article provides a comprehensive understanding of (i) the significance of Ce high-spin state on the ORR activity enhancement on the Pt and non-pt electrocatalysts, (ii) the spatially confining and stabilizing effect of ceria on the generation of the atomically dispersed transition metal-based catalysts, (iii) experimental and theoretical evidence of the effect of Ce3+ ↔ Ce4+ redox pain on radical scavenging, (iv) the effect of the Ce 4f electrons on the d-band center and electron transfer between Ce to the N-doped carbon and transition metal catalysts for enhanced ORR activity, and (v) the effect of Pt/CeO2/carbon heterojunctions on the stability of the Pt/CeO2/carbon electrocatalyst for ORR (Figure 1).

2. Properties of Cerium Dioxide/Atomically Dispersed Cerium for ORR

The element cerium (Ce), which has the atomic number 58, is by far the most abundant element from the group of rare earth metals, with almost the same abundance as Ni, Cu, and Zn in the Earth’s crust, making up around 0.0046 wt% of the crust of the Earth [36]. The exploration of RE elements in combination with transition metals goes back to the 1970s, with the first synthesis of LaNi5, used for water electrolysis, and the demonstration of the electrocatalytic hydrogen evolution activity [37] from the RE-TM and noble metals including Pt, Au, and Pd, gaining considerable interest in 21st century. Among several RE metals, Ce is by far the most explored element for electrocatalytic applications due to its superior redox capability, i.e., interconversion of Ce4+ ↔ Ce3+. Cerium, both in the form of nanoparticles and in an atomically dispersed state, is explored as an ORR electrocatalyst. The fluorite crystal structure of the CeO2 nanocrystal is with the space group Fm3m and the cell parameters are 0.5411 nm (a=b=c). In the CeO2 unit cell, each Ce4+ is coordinated with eight adjacent O2− to form an octahedral interstitial, and each O2− is coordinated with four adjacent Ce4+ to form a tetrahedral unit [38]. Because of its high electrical energy levels and unoccupied 4f orbital, CeO2 shows significant promise as a catalytic agent. The remarkable physical and chemical properties of Ce are mostly derived from its unusual electronic configuration, which is [Xe]4f15d16s2. Electrocatalytic activity can be enhanced by using Ce, a naturally occurring occupier of 4f orbital electrons, as an electronic modulator to build charge transfer highways close to the Fermi level [39]. Furthermore, Ce’s electronic properties and large atomic radius cause it to frequently display a highly coordinated structure with a minimum of four ligands, and its electronic configuration allows for a flexible variation in the coordination number [40]. This results in Ce-based materials having a highly tunable local coordination/geometry. Ce is a dopant that can be used to regulate the adsorption/desorption of chemical intermediates for electrocatalytic reactions, thanks to its high atomic radius [41]. In addition, the multivalence feature of CeO2 provides the potential to create robust interactions synergistically with other catalysts, hence improving the efficacy of electrocatalysis. The distinctive crystal structure and reversible valence properties of CeO2 facilitate the creation of oxygen vacancies, resulting in a defect-rich architecture that underpins its superior catalytic activity.
The oxygen vacancy, resulting from the reduction of Ce4+ to Ce3+ or the migration of lattice oxygen, significantly affects the electrical and chemical characteristics of CeO2. In catalysts, oxygen vacancies not only stabilize active component nanoparticles or clusters but also modulate the electrical structure of the catalysts and provide a faster charge transfer rate and excellent reaction kinetics [42,43]. During ORR electrocatalysis, the 4f band of Ce can hybridize with a 2p band of O2 intermediates because the oxygen-defective CeO2 lowers the 4f energy level of Ce below the Fermi level [44]. As a result, CeO2 shows promising results as an electrocatalytic promoter for improving the electrocatalytic activity of adjacent TM-based active sites [45,46]. Despite the several advantages of CeO2, poor electrical conductivity hinders its direct use as an electrocatalyst; therefore, the CeO2 nanoparticles, in general, require conducting carbon support to improve the interfacial electron transfer and to provide a platform to host the CeO2 nanoparticles. The combination of the high conductivity of rGO and the ample active sites on CeO2 renders the nanocomposite an effective electrocatalyst for ORR [47]. In addition, in several studies, it is found that the noble metal nanoparticles/CeO2/carbon support heterojunctions create a unique electrocatalyst in which CeO2 found to improve the catalyst activity and stability [48]. Not only are CeO2 nanoparticles known to improve electrocatalyst stability, but they are also known to improve the nafion membrane stability by averting membrane degradation. It is well known that during ORR, the electrocatalysts produce a considerable amount of H2O2 and OOH, which is produced by a 2+2 O2 reduction pathway, detrimental to the nafion membrane and even the catalyst layer. CeO2 is the best known OH· radical scavenging agent for fuel cells that include both membrane and catalysts. Therefore, plenty of researchers have introduced CeO2 nanoparticles incorporated into the membranes, aiming to improve stability [49,50].
Platinum group metals supported on a carbon matrix are the most popular catalysts for heterogeneous catalysts. As stated earlier, RE metals in the form of nanoparticles of metallic and metal oxides are often used as co-catalysts/promoters in combination with transition and noble metals [51]. However, when either catalysts, or co-catalysts with heterogenous catalysis, are used, only the surface atoms participate in the reaction, with most reactants never reach atoms in the bulk, therefore rendering them surplus [52]. Metal nanoparticles, for example, often have more than 40 atoms and have a diameter of about 1 nm. Despite environmental and chemical reaction-induced changes to the geometric configurations of exposed surface atoms (facets, corners, edges, metal–support interfaces, etc.), their geometric structures remain relatively un-stable and are thus less vulnerable [53]. In addition, the specific electronic structure of the metal–oxide interface in supported catalysts is crucial to the catalytic process since it can influence the selectivity of the products, the absorption and activation of reactants/intermediates, and the catalysts themselves [54]. Hence, a lot of work had gone into finding ways to make active metals more efficiently and to make supporting metal catalysts work better by making the active metals smaller [55]. The most efficient use of precious metals would be possible with nanoparticles that were smaller; this would be especially true for metals that were disseminated in the form of single atoms [56]. Since then, catalysis at single-atom sites has been expanding rapidly. Single-atom catalysts (SACs) illustrate the smallest dimension of metal catalysts, which not only enhances the atomic efficiency of supported metals but also boosts the quantity of interfacial sites; in contrast, the unexposed inner atoms of conventional nano-sized catalysts cannot interact with the support at interfacial sites [57]. Accordingly, several studies have been conducted in which SACs of RE metals have been proposed for ORR and other energy conversion and storage electrochemical reactions [51].
The rare earth metal Ce is special among the several rare earth metals with regard to electrocatalysis, especially for ORR, as shown in Figure 2. The special features of Ce include (i) the redox couple of Ce3+/Ce4+ which accompanies the oxygen vacancy and OH radical scavenging, whereas most other RE elements largely exists in stable RE+3 oxidation states and lack the dynamic redox coupling ability; (ii) the unique fluorite structure of CeO2 which can host O2 vacancies without any lattice distortions which act as active sites for O2 adsorption, while most other rare earth oxides, for instance, La2O3 and Y2O3 with their rigid hexagonal structures, offer less oxygen vacancies and O2 storage capacity; (iii) the radical scavenging capacity of Ce, whereas other rare earths are known to lack radical scavenging capacity; (iv) the structure stability of CeO2, owing to its partially filled 4f orbitals; and (v) its ability to change the high-spin state in the atomically dispersed state, a unique property of Ce. These factors stand out as important aspects of Ce for ORR (Figure 2)

3. Pt/CeO2 Catalysts for ORR

It is well accepted that various metal oxides, including RE oxides, can induce a positive effect on Pt NPs, thereby enhancing the ORR process due to restructuring of Pt surface defects induced by the RE oxides [58]. In addition, the presence of Ce3+ was also found to enhance the metallic ability of Pt nanoparticles and effectively inhibit the oxidation of Pt nanoparticles [59]. In this regard, CeO2 as a metal oxide component presence in Pt-based catalysts could bring several advantages such as Pt/CeO2/C heterojunctions; the Pt nanoparticles’ binding strength at the CeO2 phase was found to enhance the stability of the catalyst in strong metal support interactions (SMSI). CeO2 can donate electrons to Pt and influence its electronic structure through electronic effects and oxygen absorption energy. Pt NPs can also be stabilized by CeO2 anchoring. CeO2 acts as a co-catalyst in Pt-CeOx/C catalysts, helping the main Pt catalytic center electronically and preventing Pt oxide production at high potentials through the substitutional Ce3+ to Ce4+ redox reaction and oxygen buffering. Pt-CeOx/C triple-phase hetero-interface structures improve ORR activity and fuel cell performance.
In this regard, in several studies, CeOx-interfaced Pt catalysts have been proposed as durable ORR catalysts. One such notable work is carried out by Luo et al. [60], who synthesized an advanced Pt/CeOx/C nanocomposite with porous carbon with multiwalled carbon nanotubes (MWCNTs) by utilizing Ce-incorporated MOFs synthesized with an ATPT = 2-aminoteraphthalate organic ligand. After a high temperature treatment, the MOF(Ce) transformed into Ce-Ox, resulting in the formation of an oxide/carbon composite. The TEM images show that Pt nanoparticles of nearly spherical shapes, together with some amorphous carbon mixed oxide, are found to be present in an intimate, interconnecting way with Pt, Ce, O and C, forming a nanocomposite system. TEM images also clearly show the formation of fragments of Ce oxides in close connection with Pt nanoparticles (Figure 3a,b). The ORR analysis shows proportionately increased half-wave potential and improved Tafel slopes indicate that the formation of hetero-junctions of Pt NPs and CeOx/C helps in enhancing the ORR activity in acidic medium (Figure 3c,d). The mass and specific activities of the Pt/CeOx/C catalyst are shown to be 10 times greater than the commercial Pt/C catalyst. Metal oxides such as TiO2, SnO2, and WO3, including CeO2, have been used as potential carbon composites structures aiming for improved metal–metal oxide support interactions and improved corrosion resistance catalysts in fuel cells and metal–air batteries [61]. The triple-phase interface structures formed from the Pt, CeOx, and C, are known to be responsible for the strong metal–support interactions (SMSI). Several studies have concluded that the presence of SMSI is responsible for optimized ORR activity and stability of the catalysts. Among several metal oxides, CeO2 is the one most investigated as an active component of triple-phase interface structures, due to its defective crystal structure and rapid interchangeable oxidation states, from Ce3+ to Ce4+ and vice versa. In addition to this, CeO2 has excellent oxygen storage capacity, meaning that it has the ability to release O2 in environments with low O2 levels by absorbing O2 atoms from NO, H2O, and O2. CeO2 also performs excellently in scavenging hydroxyl (HO⋅) and hydroperoxyl (HOO⋅) radicals formed during the ORR process and mitigates the chemical degradation of the membrane and catalyst. Such triple-phase hetero-structures are constructed by packing nanosized CeO2 with ZIF-8-derived NC as a novel composite, supported in a study by Zhao et al. [62]. An abundant triple-phase hetero-junction of a Pt/CeO2-NC catalyst is created after Pt nanoparticles are deposited (Figure 3e). The Pt NPs are the ORR’s primary catalytic centers, and NC makes it possible for the Pt NPs to be evenly distributed and guarantees enough electrical conductivity. By donating electrons to Pt and influencing its electronic structure through electronic effects and oxygen absorption energy, CeO2 can play the role of an electron donor. Additionally, the anchoring effect between Pt and CeO2 can stabilize the Pt NPs. The catalyst made of 20% Pt/CeO2-NC exhibits an E1/2 of 0.922 V, and no loss after 10,000 cycles of ADTs in acidic media, according to RRDE tests, which is better than the commercial Pt/C. This is because of the interactions between Pt NPs, CeO2, and NC. Furthermore, when comparing 20% commercial Pt/C to 20% Pt/CeO2-NC in PEMFC single-cell measurements, the latter exhibits superior electrochemical activity and durability, even at lower cathode Pt loadings. To construct the effective triple-phase hetero-junctions, first, Ce@ZIF-8 structures are synthesized from 2-MIM, Ce3+, and Zn2+ structures in methanol solution which are then pyrolyzed under N2 atmosphere, which results in the formation of CeO2 nanoclusters on to which Pt nanoparticles are deposited using a microwave polyol method.
TEM images clearly reveal the formation of abundant three-phase (i.e., Pt NPs, CeO2, and NC) interfacial structures in Pt/CeO2-NC catalyst. It is seen that the CeO2 (111) phase is in close proximity to the Pt (111) NPs, suggesting an efficient catalyst with abundant triple-phase interfacial structures. The strong anchoring effect between Pt NPs and CeO2-NC is responsible for the more uniform distribution of Pt NPs on CeO2-NC supports compared to NC. The strong SMSI effect is further revealed by the XPS analysis, in which there is a shift in the Pt 4f peaks which is ascribed to the strong interactions between CeO2 and Pt. The ORR activity of the Pt/CeO2-NC catalyst showed excellent kinetics, with obtained half-wave potentials of 0.922 V, much higher than those of Pt/C, Pt-NC, and Pt/CeO2; also, the highest ECSA of Pt/CeO2-NC is 86.8 m2⋅g Pt−1, which is the largest one among these catalysts. In addition to the ORR activity, it was found that Pt/CeO2-NC exhibited excellent stability, with a loss of just 10 mV in its half-wave potential after 10,000 ADT cycles, in contrast to 35 mV loss in half-wave potential for Pt/C, with a proportional loss of ECSA of 26%, whereas the loss in ECSA is just 9% for Pt/CeO2-NC catalyst. Because of this, Pt/CeO2-NC outlasts commercial 20% Pt/C in terms of durability. This is because CeO2-NC has an anchoring effect on Pt NPs, which forbids their growth and aggregation. Interestingly, when applied as a cathode catalyst in PEMFC fuel cell configuration, the Pt/CeO2-NC catalyst delivered a power density of 1.08 W cm−2, higher than the Pt/NC and Pt/C catalysts (Figure 3f–j). The enhanced ORR activity, stability, and fuel cell performance of the Pt/CeO2-NC catalyst was attributed to several causes, among which “an effective three phase interface structure” was the main one [63]. By influencing the electronic structure of Pt and the energy of oxygen absorbed on Pt sites, CeO2 can increase the activity of the ORR. Pt nanoparticles are uniformly anchored by the CeO2-NC supports because of the anchoring effect [64,65]. Because the anchoring effect prevents Pt NPs from growing into aggregates, they enhance the stability of the catalysts [66,67]. The free radical scavenger properties of CeO2 nanoparticles in PEMFCs are well-known. Ce3+ and Ce4+ are present in CeO2 due to the presence of oxygen vacancies in the material. The antioxidant properties of CeO2 prevent the breakdown of the nafion membrane and catalyst support, which prolongs the life of Pt/CeO2-NC in PEMFCs [68].
Alloying Pt with other metals and metal oxides is known to down-shift the d-band center of Pt; as a result, the ORR intermediates absorb optimally on the active sites [69]. However, the alloyed element or metal oxide leaches out gradually, resulting in long-term instability of the electrocatalyst [70]. The mitigation of the effects of the alloying element and metal oxide component on the electrocatalyst is highly desired to improve the electrocatalytic performance of Pt for ORR. It is well known that CeO2 can enhance carbon corrosion resistance in acid due to its strong electron interaction with Pt and its high stability in acidic environments [71,72]. Nevertheless, the use of Pt-CeO2 binary catalysts in ORR has been greatly hindered by CeO2’s poor electronic conductivity. An effective strategy to improve electronic conductivity while maintaining the strong interaction between Pt and CeO2 is to construct a Pt-CeO2/C ternary nanostructure by introducing carbon into the hybrid catalyst. Du et al. [73] fabricated a Pt/CeO2-C ternary mutual interacted nanostructure from a crystal feature of Ce-MOF, resulting from the thermal treatment of Ce-MOFs, resulting in the formation of abundant tiny CeO2 nanoclusters (~2 nm) on to which Pt nanoparticles were deposited (Figure 4a). The electronic conductivity, ORR catalytic activity, and long-term durability of ORR are all greatly enhanced by the closely packed structure, which can form enough interfaces for electronic interaction and electron transfer. HRTEM images suggest that CeO2 nanocrystals are well-structured, with a cluster size of 2 nm, and CeO2 NCs produced by Ce-MOF avoided aggregation, which may be explained by the unique backbone properties of Ce [74] that cause the nanorod morphology of CeO2/C (Figure 4b). XPS analysis reveals that the Ce3+ molar ratio in CeO2/C and Pt/CeO2/C catalysts were found to be 14.3% and 20.5%, respectively. It is possible that the partial reduction of Ce4+ during Pt deposition is responsible for the increase in the Ce3+ component, and that Ce3+ acts as a reducing agent to create small Pt NCs embedded on the surface of CeO2 in an alkaline environment. A prior study using in situ electrochemical X-ray absorption fine structure demonstrated that the small amount of Ce3+ in CeO2 hinders the production of Pt oxide at high potentials through the substitutional oxidation of Ce3+ to Ce4+. This has positive implications for enhancing the catalytic performance of Pt in ORR [59,75]. The ORR studies suggest that the Pt/CeO2/C catalyst showed excellent ORR activity, with its kinetics equaling commercial Pt/C catalysts, which is attributed to the enhanced charge associated with H2 adsorption/desorption, with nearly 4 electron transfer of O2 reduction (Figure 4c–e). The enhanced Pt stability may be due to the confinement of Pt NCs in the CeO2/C hybrid and the synergistic effect between the NCs and the CeO2 support, specifically the electron deficiency of small Pt NCs caused by the transfer of electrons from Pt to CeO2. The decrease in the Pt-Pt bond distance because of the Pt-O interaction may be responsible for the enhancement of ORR activity [76].
In addition to their dual role as co-catalysts and supports, metal oxides like CeO2 can modify the electronic structure of Pt NPs and thereby produce a synergistic effect with Pt. Among the oxygen catalytic reactivity compounds, CeO2 stands out due to its superior chemical stability and noticeably reduced cost. On the other hand, a carbon layer may be necessary to enhance the anchoring effect of Pt on CeO2 due to its reduced electrical conductivity. Wang et al. [77] developed a “Pt-oxide”-based composite electrocatalyst of “CeO2 overlapped with nitrogen-doped carbon layer anchoring Pt nanoparticles” (Pt-CeO2@CN). The TEM measurements make it very clear that the catalyst contains Pt NPs, CeO2, and N-doped carbon. Specifically, it shows that the CeO2 layer overlaps with the N-doped carbon layer, which makes Pt NPs well-anchored. Furthermore, the HR-TEM analysis also clearly reveals that the distinct crystal plane distances are 0.223 nm for the Pt (1 1 1) plane and 0.311 nm for the CeO2 (1 1 1) plane, respectively. XPS analysis reveals a high abundance of graphitic-N and pyridinic-N species. Pt 4f spectra further confirm the anchoring effect of Pt on the CeO2 interfaced by N-doped carbon, which enhances the electron transport from the support to the Pt catalytic active sites. The ORR activity of the Pt-CeO2@CN is found to be higher than the commercial Pt/C catalyst. The onset potential and half-wave potential of the Pt-CeO2@CN catalyst is superior by 16 and 29 mV, respectively, when compared to Pt/C catalyst, suggesting that CeO2 and N-doped carbon helps to enhance the catalytic activity.
In a similar trend, Chen et al. [78] developed an atomic layer deposition-produced Pt/CeO2/CNT triple-junction interface, showing improved ORR activity and durability. SMSI have been proven to not only to enhance the strength of supported metal nanoparticles but also to help in the dispersion of Pt nanoparticles. In addition, the SMSI helped to prevent Pt detachment and further aggregation during long-term potential cycling, therefore helping to enhance the catalyst stability. Kinetically, the SMSI also promote the absorption of oxygen and cleavage of the O=O bond on the Pt surface [79]. Although Pt-MOx-C triple junctions are typically produced by wet chemical synthesis, it is challenging to precisely control the catalyst’s nucleation and growth during the reaction, leading to potentially larger size and uneven distribution. On the other hand, atomic layer deposition is a great way to make nano catalysts, thin films, and metal and metallic compound catalysts with clusters, single atoms, and highly dispersed nanoparticles. In this work, clusters of CeO2-nanoparticles were deposited on a CNT which is then used as a substrate on to which Pt nanoparticles are deposited using the ALD method. It was found that the Pt/CeO2/CNT-A triple-junction catalyst showed enhanced ORR activity compared to the Pt/CeO2/CNT-W that is synthesized by a wet chemical synthesis method. The XRD analysis reveals that Pt nanoparticles deposited by ALD show smaller crystallite size, with an average size of 8.30 nm, whereas, for wet chemically synthesized catalysts, it is 10 nm. TEM images clearly show that the Pt NPs were deposited on the of CeO2 and no visible CNTs were seen with Pt NPs, suggesting that most of the Pt NPs were deposited on the CeO2/CNT, confirming that effective Pt/CeO2/CNT-A triple-junction structures were generated. Furthermore, the XPS analysis revealed that the Pt/CeO2/CNT-A catalyst contains abundant oxygen vacancies (Vo), which are essential for the interaction between Pt and CeO2. The XPS spectra also reveal that Pt/CeO2/CNT-A contains high amount of Ce3+, resulting from the electron transfer from Pt to the CeO2 due to the reduction in Ce4+ ions [80]. Pt 4f spectra of Pt/CeO2/CNT-A and Pt/CeO2/CNT-W also suggest a clear differentiation between the Pt nanoparticle deposition method-induced ORR activity. A negative shift of 0.14 eV is observed for Pt/CeO2/CNT-A compared with Pt/CeO2/CNT-W. According to the d-band theory, negative shift in the binding energy indicates stronger electronic interaction, which is consistent with the observed increase in Ce3+ content. The negative shift in binding energy also means the lowering the d-band center of Pt NPS, which further helps in enhancing the electrocatalytic activity [81]. The ORR studies indicate that both Pt/CeO2/CNT-A and Pt/CeO2/CNT-W catalysts show enhanced ORR activity, with 25 and 10 mV-higher half-wave potential representing the paramount interfacial structures of CeO2 and CNT in enhancing the ORR activity. Furthermore, electron transfer from the Ce3+ to Ce4+ oxidation state further strengthens the nucleation and Pt-CeO2 interface interaction. In addition, the high Vo and oxygen storage capacity of CeO2 promotes the optimal adsorption of oxygen and cleavage of the O=O bond in the Pt surface [82]. In addition to the excellent ORR activity, Pt/CeO2/CNT-A also showed excellent stability.
To briefly conclude, it is found that Pt-CeOx/C triple-phase hetero-interface structures with Pt enhanced the binding strength of Pt nanoparticles with the Pt-CeOx/C junctions, which results in SMSI which is responsible for the enhanced ORR activity and stability of the catalysts. By donating electrons to Pt and influencing its electronic structure through electronic effects and oxygen absorption energy, CeO2 can play the role of an electron donor. Additionally, the anchoring effect between Pt and CeO2 can stabilize the Pt NPs. In Pt-CeOx/C catalysts, CeO2 plays the role of co-catalyst, assisting in the main Pt catalytic center electronically, and hinders the production of Pt oxide at high potentials through the substitutional Ce3+ to Ce4+ redox reaction and CeO2 oxygen buffering capacity. The presence of Pt-CeOx/C triple-phase hetero-interface structures not only enhances the ORR activity but also influences the fuel cell performance.

4. Fe/CeO2/C Catalysts for ORR

CeO2 is an important type of metal oxide, useful for various electrochemical reactions including CO oxidation, NO reduction, and for ORR, due to its unique oxygen storage capacity and facile interconversion of Ce3+ to Ce4+, in addition to its superior property of corrosion resistance and optimal absorption of reaction species on to its surface [26]. Ce is found to regulate Fe’s electronic structure and spatially confines and stabilizes Fe atoms. O2 interaction on Fe active sites is enhanced by 4f1 (Ce3+) localized electron transfer to d-orbitals due to differences in electronegativity values between Ce and Fe. Transition metal atoms like Fe interact well with Ce 4f1 of (Ce3+). Hybridizing 4f-3d orbitals lowers the band gap and improves conduction band dispersion, boosting ORR. Thus, the catalyst must have a high Ce3+ ratio. Hybridizing d-f orbitals reduces the Fe atom’s d-band center, making electron transfer to adsorbed OH intermediates easier and speeding up the rate-determining step and ORR. Thus, in Fe-Ce dual-atom-based catalysts, Fe is the main active center and rare earth Ce 4f electrons improve charge transfer, reduce the d-band center, and improve ORR kinetics. In this section, we will describe the various Fe/Ce-based catalysts proposed in the literature.
Zhang et al. [83] synthesized a series of MOFs with Ce/La MOFs and investigated the effect of Lewis’s base sites, composed of rare earth metals, on the synergistic ORR activity of porous Fe-Nx active sites (Figure 5a). The Le/Ce-doped MOFs were synthesized by a simple co-precipitation method, and then the synthesized MOFs were immersed in the Fe3+ solution. The resulting Fe-MOFs were then pyrolyzed to obtain the xLa-CeNC-Fe. Analysis of the diffraction peaks of Ce-NC-Fe with La3+ show that the diffraction peaks shift to a lower 2theta angle, indicating the successful incorporation of La3+ (0.116 nm), whose ionic radius is larger than the Ce4+ (0.097 nm). Both SEM and TEM images show CNT structures, obviously due to the ability of Fe to graphitize the carbon during the pyrolysis process. The ORR analysis of the 0.5La-CeNC-Fe catalyst showed a well-defined cathodic reduction peak in the cyclic voltametric analysis, with a peak potential of 0.886 V vs. RHE, much higher than the Pt/C catalyst of 0.875 V. The catalyst of 0.5La-CeNC-Fe with an E1/2 of 0.870 V exhibits significantly better ORR activity than that of other catalysts and is slightly better than that of Pt/C with an E1/2 of 0.862 V. The K-L plots indicate that the 0.5La-CeNC-Fe catalyst exhibited an ‘n’ of 3.94, indicating almost a direct four-electron reduction of O2 to H2O (Figure 5b–i).
In addition to the role of Ce as a co-catalyst, O2 buffering agent, Ce plays an important role in spatially confining and stabilizing Fe atoms [84]. Thus, Ce can greatly improve the ORR/OER catalyst’s stability and catalytic activity when added to metal-based hybrid electrocatalysts. Wang et al. [85] synthesized a two-dimensional ZIF assisted with a NaCl molten salt synthesis method to create a Ce/Fe-NC/Fe3C-P nanosheet with a hierarchical porous structure (Figure 5j). Thanks to the abundance of molten salt around the ZIF nanosheets, the precursor’s shape was kept intact. At the same time, adding molten salt of sodium chloride to a carbon matrix can improve electron transport, increase the catalyst’s durability, and raise the degree of local graphitization and graphitic N dopants. Both Fe-ZIF and Ce/Fe-ZIF exhibit an elliptical nanoflake structure, and the Fe-ZIF-derived Fe-NC/Fe3C catalyst exhibited structures that resembled carbon nanotubes (CNTs), with a significant amount of agglomeration. On the other hand, the Ce-ZIF-pyrolyzed Ce/Fe-N-C/Fe3C-P catalyst displayed a very small number of CNT-like structures. This is because the addition of Ce causes the Ce/Fe-ZIF nanoflakes to shrink and link with each other during the pyrolysis process, resulting in the formation of a coralloid structure. According to this, Ce is responsible for regulating the electronic structure of Fe and preventing the growth of carbon nanotubes. Furthermore, it was found that the ratio of molten salt NaCl:Ce/Fe-ZIF also plays a role in preserving the elliptical nanoflake structures. Obtaining nanosheets with a mass ratio of 1:1 between NaCl and Ce/Fe-ZIF results in significant agglomeration. The elliptical nanoflake structure of Ce/Fe-ZIF was integrally preserved when the ratio increased to 3:1. As the ratio increased to 5:1, the morphology became nearly identical to that of 3:1. HR-TEM images of the Ce/Fe-NC/Fe3C-P catalyst show the homogeneous distribution of C, Fe, and Ce, along with the Fe3C nanoparticles, meaning that Fe particles are surrounded by the N-doped carbon shells which act as a catalytic active site to facilitate the ORR/OER [86,87,88,89,90]. The core–shell structure is crucial for avoiding NP agglomeration and boosts catalyst durability by halting Fe3C nanoparticle leaching in extreme electrochemical environments. Interestingly, it is also found that the Ce also regulates the porosity of the catalysts. According to the pore size distribution analysis, micropores are more prevalent in Fe-NC/Fe3C, but mesopores appear after Ce incorporation in the Ce/Fe-NC/Fe3C-P catalyst. This could be because the preexisting Ce controlled the electrical structure of Fe and prevented the growth of CNTs, in contrast to Fe-NC/Fe3C [91]. In addition, when NaCl was introduced during the synthesis process, it was found that well-balanced micro and mesopores were identified in the catalyst [92]. The role of micro and mesopores includes hosting the edge, highly active ORR sites in the form of Fe-Nx-C and providing a pathway for the reactants and products to the ORR active sites, respectively [93]. The XPS analysis reveals the presence of definite Ce-Nx bonding active sites along with pyridinic-N, Fe−Nx, pyrrolic-N, and graphitic-N, which can act as ORR active sites. There is a noticeable change in the binding energy of Fe2p3/2 between the Fe-NC/Fe3C and Ce/Fe-NC/Fe3C-P catalysts, which is an intriguing finding. This might be because Ce doping allows charges to migrate between Fe and N elements, among others. The charge transfer effect, according to earlier research, lowers the d-band center value, which in turn reduces the adsorption energy of oxygen-containing intermediates in electrochemical reactions and increases the ORR electrocatalytic efficiency [94]. The electrochemical ORR studies indicate that the Ce/Fe-NC/Fe3C-P catalyst showed the highest ORR activity, followed by Ce/Fe-NC/Fe3C, 20% Pt/C, and Fe-NC/Fe3C. The Ce/Fe-NC/Fe3C-P catalyst’s outstanding performance in ORR is likely connected to Ce species, which effectively lower the energy barrier of the rate-limiting step and promote more efficient transport of oxygen-containing intermediates containing H2O2, OH, and *OH radicals during the ORR. When applied as a cathode catalyst in Zn–air batteries and in PEMFC environments, the Ce/Fe-NC/Fe3C-P catalyst delivered a powder density of 184 and 347 mW cm−2, respectively (Figure 5k,l).
Figure 5. (a) Schematic representation of 0.5La-CeNC-Fe. (b) XPS deconvoluted peaks of Ce3+ and (c) obtained Ce3+ content. (d) CV and (e) LSV curves of various 0.5La-CeNC-Fe catalysts. (f) LSVs recorded at various rotations per minute. (g) K–L plots. (h) Number of electrons. (i) Tafel plots of 0.5La-CeNC-Fe catalysts (Reproduced with permissions from Ref. [83]). (j) Schematic synthesis of Ce/Fe-NC/Fe3C-P catalyst. (k) LSV curves of various Ce/Fe-NC/Fe3C-P catalysts in 0.1 M HClO4. (l) PEMFC polarization curves of Ce/Fe-NC/Fe3C-P catalyst (Reproduced with permissions from Ref. [85]).
Figure 5. (a) Schematic representation of 0.5La-CeNC-Fe. (b) XPS deconvoluted peaks of Ce3+ and (c) obtained Ce3+ content. (d) CV and (e) LSV curves of various 0.5La-CeNC-Fe catalysts. (f) LSVs recorded at various rotations per minute. (g) K–L plots. (h) Number of electrons. (i) Tafel plots of 0.5La-CeNC-Fe catalysts (Reproduced with permissions from Ref. [83]). (j) Schematic synthesis of Ce/Fe-NC/Fe3C-P catalyst. (k) LSV curves of various Ce/Fe-NC/Fe3C-P catalysts in 0.1 M HClO4. (l) PEMFC polarization curves of Ce/Fe-NC/Fe3C-P catalyst (Reproduced with permissions from Ref. [85]).
Nanomaterials 15 00600 g005
In another study, Zenghui et al. [95] synthesized a unique three-dimensional (3D) star-like carbon material electrocatalyst made of Fe and Ce, synthesized by a ZIF-8 structures with Zn2+ and 2-MIM and the Fe3+ and Ce3+ ions then incorporated into the ZIF-8 structures in the presence of a CTAB surfactant, and pyrolyzed at 910 °C, followed by acid etching (Figure 6a). SEM and TEM measurements prove that the FeCeNC catalyst has a star shape, where each corner is of the same size and all corners are oriented in different directions in three dimensions. While the synthesis of the FeCeNC catalyst looks almost very similar to the simple, traditional ZIF-8 synthesis, the only difference is the use of the CTAB surfactant; therefore, the evolution of the 3D star-shaped FeCeNC catalyst could be primarily attributed to the use of CTAB (Figure 6b–k). The ORR activity of FeCeNC catalyst was evaluated in 0.1 M KOH solution. The order of ORR activity is as follows: NC < CeNC < FeNC < FeCeNC catalysts with a half-wave potential of 0.796, 0.799, 0.803, and 0.855 V vs. RHE, respectively. The ORR activity of the final and optimized FeCeNC catalyst is also higher than the commercial catalysts Pt/C by 25 mV. However, the number of electrons transferred per O2 molecule is slightly lower than the catalysts discussed so far, with an ‘n’ value of 3.7 derived from the K-L plots. Thanks to the strong bond between Fe, Ce, and the coordinating N atoms (Fe-N and Ce-N), FeCeNC also showed remarkable durability, with only a 2 mV decay after 1000 cycles. When assembled with FeCeNC as a cathode catalyst with rechargeable ZAB, it delivered a power density of 169.2 mW cm−2, much higher than the Pt/C catalyst which only delivered 88.0 mW cm−2. Furthermore, after 80 h of cycling, the charge–discharge efficiency of the ZAB battery dropped by just 6.8%, demonstrating excellent durability thanks to the higher power density and FeCeNC catalyst (Figure 6l–q).
In another study, Kumar et al. [96] synthesized highly efficient ZIF-derived FexCey@N–C composite catalysts for ORR. In this study, a simple ZIF-8 synthesis in the presence of Ce3+ and Fe3+ with ZIF-8 and MWCNTs, which were then pyrolyzed at 900 °C, generated highly defect-rich and disordered structured electrocatalysts of a FexCey@N–C (x:y =1:1) composition, which offer a greater tendency to adsorb O2 and promote the ORR kinetics which contains Fe-Nx and Ce-Nx active sites for ORR. In particular, it was well documented that Fe-Nx is the most accepted form of catalytic active site of Mx-Ny-C, and Ce-Nx takes a dual role as an ORR active site and a scavenger of free radical that are generated from Fe-Nx sites, as a result of the 2+2 electron reduction reaction, and mitigate catalyst degradation from reactive oxygen species (ROS) generated nearby the iron centers.
The XRD analysis of the catalysts reveals both metallic Fe and Ce and Fe-Ce alloy phases. XPS analysis reveals several important reasons contributing to the ORR activity of the FexCey@N–C catalyst, especially in Fe3p and Ce3d XPS spectra, as discussed below. The XPS peaks at 710.8 and 723.1 eV could potentially be indicative of the Fe3+ and Fe2+ species present in the octahedral places. The exclusive presence of divalent Fe is indicated by the satellite peak at approximately 716.0 eV. The adsorption of O2 intermediates is directly impacted by the movement of Fe in or out of the N4 plane, caused by changes in potential which correspond to changes in its oxidation state from +2 to +3. The structural distortion caused by substituting Fe3+ for Zn2+ in ZIF-8, as shown in experimental studies, is less stable because of the smaller ion radius (49–55 pm) [97]. Initial reduction of Fe3+ to Fe2+ in active sites improves ORR kinetics, which in turn helps to displace adsorbed H2O with O2 and facilitates electron transfer from Fe2+ to O2 [98,99]. Thus, by supplying evenly distributed Fe-Nx moieties and easing the adsorption of O2 intermediates, Fe2+ improves the ORR electrocatalytic activity. The Ce3d XPS spectra show the co-existence of Ce3+3d3/2 (3d10 4f1) and Ce4+ (3d104f0). In particular, the presence of Ce3+3d3/2 (3d10 4f1) with the 4f1 orbital and a lone pair of electrons is an important aspect of the Ce-based catalysts. Ce 4f1 electrons a have high affinity to interact with transition metal atoms such as Fe and help in electron transfer from Ce → Fe, further enhancing the interaction of O2 on Fe active sites, due to the 4f1 localized electron transfer to d-orbitals of Fe active site due to their difference in their electronegativity values [100]. This hybridization of 4f-3d orbitals reduces the band gap and enhances the conduction band dispersion, facilitating efficient ORR. To briefly conclude, the Fe2+/3+ and Ce3+/4+ redox active couples promote enhanced ORR kinetics. As expected, a clear demonstration of the synergistic effects of Fe2+/3+ and Ce3+/4+ on the ORR activity has been established in the RDE studies in 0.1 M KOH solution. The ORR onset potential of 0.92 V is recorded for the Ce@NC catalyst, whereas the Fe1Ce1@NC catalyst ORR onset potential was found to be around 0.981 V, with a half-wave potential of 0.867 V vs. RHE, higher than that of 30 mV from the Pt/C catalyst.
From the discovery of single-atom catalysts, research on embedding two or more atoms of different properties, creating dual-atom and multiple-atom catalysts, has become popular. Due to complexity of the multi-atom-based catalysts, dual-atom-based catalytic systems have recently attracted tremendous attention due to the easier gauge of the synergistic effects between the different metal atoms. Yang et al. [101] demonstrated the effect of Fe and Ce dual-metal catalysts (Fe-Ce-SAD/HPNC) to modulate the d-f orbital hybridization with an optimal O-binding energy to create an excellent ORR catalyst. It was found that, by hybridizing d-f orbitals, the d-band center of the Fe atom decreases, thus making the transfer of electrons to the adsorbed OH intermediates easier, therefore helping in accelerating the rate-determining step and thus the ORR (Figure 7a–e). The XANES study of the Fe-Ce-SAD/HPNC revealed that Fe atoms were positively charged, with Ce in the form of a valence between +3 and +4. Furthermore, the Ce L3-edge XANES spectrum suggests that Ce can exist in the form of isolated atoms, not bonded to Fe. However, the EXAFS spectrum indicates that Ce is bonded to -N due to Ce−N scattering, indicating that catalyst have atomically dispersed Ce-N sites. The ORR studies indicate that the Fe-Ce-SAD/HPNC catalyst showed a positive ORR onset potential of 0.91 V, a half-wave potential of 0.81 V, and a Jk of 32.7 mA cm−2 at 0.75 V vs. RHE.
In addition, the Fe-Ce-SAD/HPNC catalyst exhibited the highest double-layer capacitance of 53 mF cm−2 among all the catalysts in the study, indicating that the Fe-Ce-SAD/HPNC catalyst possess a higher surface area and hence high ORR activity. Furthermore, the reactive oxygen species (ROS) that include H2O or H2O2 and OH formed during the ORR process is one of the major issues that elevates the degradation of several components; therefore, it is important to quantify the ROS of the catalysts [102]. The qualitative and quantitative analysis of ROS is generally performed by the UV spectroscopy analysis with the help of 2,20-azinobis(3-ethylbenzthiazoline-6-sulfonate) (ABTS) as a substrate, that reacts with ROS formed during the ORR to form a green-colored solution upon reacting with ROS, with an absorption maxima at 417 nm. Interestingly, the Fe-Ce-SAD/HPNC catalysts showed lower intensity of the peak at 417 nm, along with the visual appearance of a light-colored solution, indicating that the Fe-Ce-SAD/HPNC catalyst forma less ROS during ORR. In addition to the RDE studies, the authors also evaluated the practical application of the Fe-Ce-SAD/HPNC catalyst in PEMFC conditions. The Fe-Ce-SAD/HPNC catalyst delivered a powder density of 0.771 W cm−2, with a catalyst loading of 3 mg cm−2, at a pressure of 1 bar H2-O2 and 0.498 W cm−2 in H2-O2 (Figure 7f–k).
The ORR mechanism of Fe-Ce-SAD/HPNC catalysts is assessed using density functional theory analysis by constructing a model of an active site composed of N4−Fe−Ce−N6−C and Fe−N4−C active sites for understanding the differences in the ORR mechanism [101]. The distance between Fe and Ce atoms was adjusted to 0.385 nm. The obtained Bader charge analysis reveals charge transfers of 0.81 and 1.10e for Fe-SAD HPNC and Fe-Ce-SAD/HPNC for the adsorbed *OH, confirming that the higher electron density around the Fe-Ce-SAD/HPNC catalyst is due to efficient electron transfer from Ce to Fe, enhancing the adsorption of O2. The density of states (DOS) further reveals that the d-band center of Fe in Fe-Ce-SAD/HPNC catalyst significantly decreased and that more occupied orbitals start to appear near the Fermi level due to f-d orbital interaction. The decreased d-band center means that more and more electrons are being added from Fe to the ORR intermediates such as -OH, which, therefore, reduces the adsorption strength between the ORR intermediate and the Fe-Ce-SAD/HPNC catalyst surface. The obtained d-band center from the DOS values relative to the Femi levels are −1.7 and −2.6 eV, respectively. The calculated free energy values were derived for various ORR intermediates such as *OOH, *O, and *OH. It is seen that there is a significant difference in the RDS of Fe-SAD HPNC and Fe-Ce-SAD/HPNC catalysts. For FeCe-SAD/HPNC catalysts, it is *O + H+ + e− → *OH, whereas for Fe-SAD/HPNC, it was *OH + H+ + e− → H2O, indicating that Ce 4F weakens the adsorption strength between the Fe active sites and the ORR intermediates, shifting the reaction mechanism with *OH desorption to *O > *OH. Therefore, it can be inferred that in Fe-Ce dual-atom-based catalysts, Fe is the main active center and rare earth Ce 4f electrons help to enhance the charge transfer and reduce the d-band center and hence enhance the ORR kinetics.
In a study, it was found that, in the process of Pt anchoring onto CeO2 through a high temperature treatment, the Pt atoms are tightly trapped on the ceria surface. Rather than diffusing int to the bulk phase of ceria and DFT, studies further indicate the stabilization effect of ceria for various metals such as Ni/Pd/Pt; essentially, CeO2 acts as dispersing agent. Based on these assumptions, Li et al. [103] explored the possibilities of stabilizing atomically dispersed Fe via ceria confining and trapping. With the CeO2 confining and trapping strategy, an Fe content as high as 4.6 wt% has been achieved. The Ce3+ and Fe3+ were adsorbed on the polypyrrole nanowires through electrostatic attraction. The metal ions are attached to the polypyrrole via backbone N atoms and the resulting material is then thermally treated to synthesize Ce/Fe-NCNW. Among various compositions of Ce3+ and Fe3+, the atomic ratio of 1:1 Ce3+ and Fe3+ is found to be optimal in terms of ORR activity. The Ce/Fe-NCNW catalyst showed an excellent half-wave potential of 0.915 V, higher than Pt/C by 49 mV, and a 3.3-times higher kinetic current density than Pt/C in 0.1 M KOH electrolyte, with a HO2 yield of 2.5%. In a fuel cell comprising an alkaline membrane, the Ce/Fe-NCNW catalyst delivered a power density of 496 mW cm−2, with a catalyst loading of 1.0 mg cm−2 at 30 pis pressure. The catalyst synthesized without Ce showed lower ORR activity, confirming the definite role of ceria in ORR catalysis. The structural analysis of the Ce/Fe-NCNW catalyst indicates no visible agglomeration of Fe, whereas nanoparticles of CeO2 were observed. A high content of atomically dispersed Fe (4.6 wt.%) in Ce/Fe-NCNW is achieved by the bonding of Fe atoms to O in the lattice during heat treatment and the effective suppression of agglomeration of isolated Fe atoms, made possible by ceria’s spatial confinement and strong trapping in this process. In addition, the synergistic effect of Fe and Ce further improves O2 adsorption and reduction kinetics, leading to enhanced ORR activity.
Though Fe-N-C-based catalysts are sought for as the best alternatives to the traditional Pt/C catalysts, they suffer from poor stability due to surface carbon oxidation, demetallation, protonation of the N-groups, and anion adsorption. One of the greatest limitations of Fe-N-C-based catalysts is the formation of reaction oxygen species (ROS) such as OH, which irreversibly destroys the carbon support and membrane. Carbon support corrosion leads to the demetallation or loss of FeNx active sites. To mitigate the effect of ROS, Chu et al. [104] designed atomically dispersed Fe and Ce atoms on the microporous carbon derived from the MOFs. While synthesizing the Fe/Ce through ZIF-8 strategy, the presence of Ce3+ ions could dynamically alter the growth of ZIF-8 and significantly alters (decreases) the particle size. This feature may be linked to the comparatively reduced growth rate of ZIF-8 in the presence of Ce, resulting in a smaller particle size and enhanced encapsulation of Fe ions within the framework. Consequently, in addition to the potential radical scavenging effect, the introduction of Ce enhances catalytic activity through effective Fe incorporation (augmenting active site density) and ZIF-8 particle size reduction (enhancing Fe utilization). DFT studies suggest that the reaction of H2O2 decomposition to O2 and H2O are similar on both FeCe-N-C and Fe-N-C catalysts. However, a significant difference is observed in the potential energy diagram of the FeCe-N-C catalyst, in which Ce sites catalyzing the *O to O2 step are thermodynamically faster, which is beneficial to reduce ROS conversion into H2O. Furthermore, the Gibbs free energy profiles suggests that the FeCe-N-C catalyst showed the lowest energy barrier for facilitating the adsorbed *OH species from the Fe site. Exploration of the electronic density distribution through Badar charge analysis reveals lower adsorption energy for oxygen species on the FeCe-N-C catalyst due to electron delocalization between Fe and Ce. The independent active sites of CeNx are highly active towards catalyzing OOH* (to yield O2*) and O* (to yield OH*). This DFT analysis clearly reveals the participation of Ce on the radical scavenging of ROS, thereby protecting the Fe-Nx active sites. The experimentally synthesized catalysts showed that FeCe-N-C catalysts exhibited enhanced ORR activity in the order of Fe,Ce-N-C > Fe-N-C >> Ce-N-C with the half-wave potential of 0.808, 0.791, and 0.454 V vs. RHE, respectively, in 0.1 M HClO4. Experimental and direct evidence of the FeCe-N-C catalyst could be deduced from the RRDE experiments and estimation of H2O2, which is found to be <1% in all the potential range, suggesting that the FeCe-N-C catalyst is a great choice for the ORR. In addition, the effect of Ce radical scavenging can also be translated from the stability test, in which the FeCe-N-C catalyst presented excellent stability with just a loss of 22 mV after 30,000 potential cycles, better than the Fe-N-C catalyst and Pt/C catalyst, suggesting that Ce radical scavenging helps to enhance the stability of the catalysts. The experiment and theoretical analyses indicate that Ce single sites promote catalysis, scavenging ROS and promoting the 4e ORR process by decomposing H2O2 (Figure 8a–h).
In another study, there is a similar approach to improving electrocatalytic activity and stability by introducing CeO2 as an antioxidant agent, presented by Luo et al. [105]. An interesting in situ Raman spectroscopy analysis revealed the gradual appearance of Raman shift at 1166 and 1520 cm−1, corresponding to the O2¯ and *OOH intermediates as the potential rises from 1.0 to 0.1 V for CeO2@Fe-NC, whereas these Raman shift peaks are barely visible in the case of Fe-NC catalysts, indicating a definite role of Ce in enhancing ORR activity. Theoretical simulations indicate a strong electronic interaction between Fe-Nx moieties and CeO2. The Gibbs free energy for the RDS of *OH was found to be 0.39 eV for the CeO2@Fe-NC catalyst, whereas it was 1.01 eV for the Fe-NC catalyst, indicating that the presence of CeO2 enhances the ORR activity significantly. Similar observations were drawn from the experimental studies, in which the CeO2@Fe-NC catalyst showed a half-wave potential of 0.89 V vs. RHE. The number of electrons transferred per O2 molecules was found to be four, suggesting that CeO2@Fe-NC ORR proceeds by a direct four-electron pathway. The H2O2 UV-Vis absorption spectra shows that H2O2 begins to decompose at 80 s for the CeO2@Fe-NC catalyst, whereas it is 225 sec in case of the Fe-NC catalyst, indicating that CeO2 played a crucial role in eliminating H2O2 during ORR (Figure 8i–m). In another study [106], it was found that Ce incorporation into CL-Fe/(Fe,Ce)xOy-NC enhances the concentration of active sites, including pyridinic nitrogen, graphitic nitrogen, and metal-chelated nitrogen. The CL-Fe/(Fe,Ce)xOy-NC shows both a micro- and mesoporous nature, with a specific surface area of 891 m2 g−1. XPS analysis reveals that presence of Ce3+ (5s25p64f1) can interact positively with the intermediate p electrons of the reaction intermediates, promoting charge transfer and proton coupling, whereas the Ce in the form of Ce4+ with the electronic configuration of Ce4+(5s25p6) is completely occupied, and the internal electronic structure is closed, resulting in poor catalytic activity. Therefore, the electrocatalyst with high Ce3+ results in enhanced ORR activity. The resulting CL-Fe/(Fe,Ce)xOy-NC exhibits a E1/2 of 0.861V vs. RHE and acceptable stability.
In a unique study, Zhang et al. [107] introduced cerium oxide cyanamide (Ce2O2CN2) into the Fe-N-C active sites due to the presence of unique (N=C=N)2−, which can enhance the conduction of the carbon support through a conjugation effect. Since the (N=C=N)2− group has a lower electronegativity than metal oxides such as CeO2, it is possible to alter the distribution of electronic clouds, which in turn affects the material’s charge transfer capability. Therefore, the cerium oxide cyanamide can act as promoter to improve the catalytic activity and stability of M–N–C catalysts. Based on these assumptions, a heterostructure of Fe–N–C@Ce2O2CN2 has been synthesized. As expected, the Fe–N–C@Ce2O2CN2 catalyst showed remarkable electrocatalytic performance and methanol tolerance for the ORR, surpassing that of Fe-N-C and commercial Pt/C, due to the surface oxygen density modulation ability of the Ce4+/Ce3+ redox pair and the high density of the Fe-Nx active sites. The Fe–N–C@Ce2O2CN2 catalyst outperforms the Pt/C catalyst in terms of ORR activity, with a half-wave potential (E1/2) of 0.89 V vs. RHE in a 0.1 M KOH solution. In addition, in Zn–air batteries, the catalyst shows a much higher peak power density (119.35 mW cm2) than commercial Pt/C (80 mW cm2).
In brief, Ce is shown to be in charge of Fe’s electronic structure regulation as well as for spatially confining and stabilizing Fe atoms. Because 4f1 (Ce3+) localized electron transfer to d-orbitals of the Fe active site is due to their difference in electronegativity values in electrons, transfer from Ce → Fe further enhances the interaction of O2 on Fe active sites. Ce 4f1 of (Ce3+) has great affinity for interacting with transition metal atoms such as Fe. By lowering the band gap and improving conduction band dispersion, this hybridization of 4f-3d orbitals promotes effective ORR. Therefore, it is important to have a high ratio of Ce3+ in the catalyst. It was observed that hybridizing d-f orbitals reduces the d-band center of the Fe atom, so facilitating easy transfer of electrons to the adsorbed OH intermediates and accelerates the rate-determining step and hence the ORR. Consequently, it can be deduced that in Fe-Ce dual-atom-based catalysts, Fe is the main active center and rare earth Ce 4f electrons help to improve the charge transfer and hence reduce the d-band center and thus enhance the ORR kinetics.

5. Co/CeO2/C Catalysts for ORR

Due to their lower radical oxygen species production, Co-N-C catalysts have gained tremendous attention as an alternative to of Fe-N-C. Since Co-N-C electrocatalysts are inactive for Fenton-like reactions and stable under harsh conditions, they have been pursued as ORR catalysts. Due to modified catalyst electron structure, O2 storage capacity, and shift in the d-band center, combining Co and Ce could improve ORR activity. Furthermore, the O2 buffering (store/release mechanism of oxygen) capacity of CeO2 can further be tuned by choosing alternative Ce precursors to the traditional Ce nitrates.
Xia et al. [108] found that by adding a Ce2(OH)4SO4.2H2O precursor during the nucleation process of ZIF, 2D-hexagonal-leaf-like ZIF lamellae (ZIF-L) can be obtained which could generate more oxygen vacancies than by using traditional Ce precursors such as Ce nitrates (Figure 9a–g). The leaf-like ZIF-L structures are formed by careful control of the 2-MIM and Co3+ and Zn2+ ions. In general, the amine N-groups from the 2-MIM interact with the H atoms of H2O, forming H-bonds and bridging 2MIM-N-H bonds, making it further extended with other molecules of N-2MIM through sodalite layers, leading to the formation of leaflike structures. Based on this assumption, the leaf-like 2D ZIF-like structures have been observed to have a unique, smooth surface. Although the surfaces of the leaf-like lamellae become rough and numerous nanoparticles (white dots) are embedded in the surfaces of the Ce-HPCNs, the pristine leaf-like morphology of the Ce-HPCN precursor is preserved in the resulting Ce-HPCNs, as anticipated. The HR-TEM analysis shows clear leaf-like structures, with Co and CeO2 nanoparticles in close contact with each other. The RDE studies further showed that the ORR onset potential of 0.923 V and a half-wave potential of 0.831 V for Ce-HPCN are higher than the Pt/C catalysts, due to ther unique leaf-like morphology, highly 2D open pore structure, higher content of pyridinic-N and graphitic-N sites, and the synergistic effect of Co-Nx, along with the oxygen buffering properties of CeO2 (Figure 9h). The O2 buffering capacity of Ce is further determined from the O2 temperature-programmed desorption (O2-TPD) spectra, which clearly show a high-intensity O2 desorption peak, indicating the enhanced adsorption energy of O2 for the sample of Ce–HPCNs [109]. All of these findings point to the possibility that doped CeO2 could significantly improve O2 adsorption capacity and the interaction between Co-Nx active sites, leading to outstanding ORR performance.
Mixed metal oxides are known examples of synergistic catalysts, in which two types of dissimilar metal oxides are known to improve electrochemical properties. For instance, a heterostructure of the CO3O4-MnO2/C catalyst delivered enhanced ORR activity, owing to a direct four-electron reduction of O2 as a result of covalent coupling between Co3O4-MnO2 [110]. Similarly, the Pd@PdO-CO3O4 heterostructure catalysts delivered high ORR/OER activity [111]. In the same way, coupling CO3O4 with a rare earth CeO2 could be the best choice, owing to their abundant oxygen vacancies and oxygen storage capacity, along with reversible surface ion exchange capacity, leading to enhanced ORR activity. Li et al. [112] synthesized Co3O4@Z67-NT@CeO2 by a sequential synthesis process in which CeO2 nanoparticles were introduced onto the surface of Co3O4@Z67-NT by a hydrothermal synthesis method. The XRD analysis revealed characteristic diffraction peaks of both Co3O4 and CeO2, indicating the crystalline nature of bi-metal oxides in the catalysts. The XPS analysis showed the presence of high pyridinic-N, which is responsible for enhancing electronic conductivity, and onset potential for ORR, whereas graphitic-N content is known to enhance the limiting current density [113,114]. The XPS analysis of Co3O4@Z67-NT@CeO2 shows a high proportion of Ce2+, which are responsible for high ORR activity, along with Ce3+ ions. Furthermore, the electrostatic interaction between Co and Ce is ascertained from the positive shift in the binding energy of Ce3+. The relative atomic composition of Ce3+ and Ce4+ hint at the reason behind the ORR activity of Co3O4@Z67-NT@CeO2 catalysts. It was found that the relative abundance of Ce3+ is about 36.91%, indicating that the ORR activity of the Co3O4@Z67-NT@CeO2 catalyst is due to the role of charge compensation by Ce3+, which will generate oxygen vacancies to promote the oxygen adsorption on the catalyst surface [115]. Furthermore, the continuous availability of a higher valance state of Co is possible due to the lattice oxygen provided by the CeO2, keeping the catalytic cycle happening continuously. The ORR half-wave potential of Co3O4@Z67-NT@CeO2 is found to be higher than Co3O4@Z67-NT by 20 mV, being 0.88 V vs. RHE in 0.1 M KOH electrolyte, indicating that CeO2 guarantees enhanced activity.
Figure 9. (a,b) SEM and (cg) TEM images and elemental mapping of different elements of Ce–HPCN. (h) ORR curves of Ce–HPCN catalysts Inset of (e) shows the SAED picture of Co and CeO2. (f) ORR LSV curves of 3D-Co–Zn/C, HPCNs, and Ce–HPCNs catalysts @ 1600 rpm (Reproduced with permissions from Ref. [108]). (i) Schematic representation of synthesis of Ce@Co3O4/CNFs. (jm) SEM and TEM images and elemental mapping of Ce@Co3O4/CNFs catalysts. (n) LSV curves. (o) K-L plots. (p) Number of electrons. (q) Tafel plots of Ce@Co3O4/CNFs catalyst (Reproduced with permissions from Ref. [116]).
Figure 9. (a,b) SEM and (cg) TEM images and elemental mapping of different elements of Ce–HPCN. (h) ORR curves of Ce–HPCN catalysts Inset of (e) shows the SAED picture of Co and CeO2. (f) ORR LSV curves of 3D-Co–Zn/C, HPCNs, and Ce–HPCNs catalysts @ 1600 rpm (Reproduced with permissions from Ref. [108]). (i) Schematic representation of synthesis of Ce@Co3O4/CNFs. (jm) SEM and TEM images and elemental mapping of Ce@Co3O4/CNFs catalysts. (n) LSV curves. (o) K-L plots. (p) Number of electrons. (q) Tafel plots of Ce@Co3O4/CNFs catalyst (Reproduced with permissions from Ref. [116]).
Nanomaterials 15 00600 g009
In another study that examined the effect of rare earth metal oxide, CeO2 was combined with Co3O4, gaining the advantage of CeO2 as a redox couple and oxygen buffering agent. In a three-step strategy, Ce is introduced into the carbon nanofibers loaded with Co3O4 (Ce@Co3O4/CNFs) using an electrospinning (Figure 9i) method [116]. The XRD analysis reveals almost no visible CeO2 peaks, indicating that either Ce is doped into the Co3O4 or dispersed on the catalyst due to low Ce content. This is further confirmed when Ce atoms have no effect on the polyhedral morphologies of MOF–fiber composites. TEM images show the highly dispersed, tiny Co3O4 nanoparticles evenly dispersed on the nanofiber surface, ensuring abundant exposed active sites for ORR (Figure 9j–m). The XPS analysis revealed that Co3+/(Co3+ + Co2+) molar ratio increases in Ce@Co3O4/CNFs, indicating the possibility of Ce atoms being in close contact with Co3O4 or inserted into the crystal structure of Co3O4. Furthermore, the significantly increased Co3+ species infer that the electronic structure of Co in Co3O4 can be easily fine-tuned by introducing the Ce species, owing to their redox coupling reaction between Co3+/Co2+ and Ce3+/Ce4+. Moreover, the presence of a high proportion of Ce3+ is found to be beneficial for electrocatalytic activity [117]. The ORR curves clearly establish the synergistic effect of Co3O4 and Ce on the ORR activity as the Ce@Co3O4/CNF catalyst shows almost similar activity as that of the commercial Pt/C catalyst (Figure 9n–q). The Ce@Co3O4/CNFs ZAB exhibits remarkable electrocatalytic activity, which can be attributed to the combination of Co3O4 and CeO2 as well as the high conductivity of CNFs.
Metal oxides such as Co3O4-based spinel oxides, which are a kind of p-type semiconductor, have been explored as efficient ORR catalysts due to their unique crystal structure in which Co2+ occupies tetrahedral sites and Co3+ occupy octahedral sites [118,119]. CeO2, a rare earth based on Ce, is an n-type semiconductor that is an interesting metal oxide due to its high ionic/electronic conductivity and oxygen storage capacity and can influence the oxygen reduction catalysis [120]. Because of the strong electronic coupling effect, it is believed that hybridizing p-type Co3O4 and n-type CeO2 will increase the activity of oxygen catalysis. By combining n and p-type materials, engineers can design and fine-tune an electrical structure that is advantageous for optimizing surface properties for enhanced ORR. Guo et al. [121] synthesized a Co3O4/CeO2 heterostructure in situ embedded in Co/N-doped carbon nanofibers through a facile electrospinning process (Figure 10a–j). The bimetallic ZIF grows in a tiled pattern due to the attractive force of metal ions on the surface of the nanofibers, which facilitates the cross-linking with 2-MIM. The following in situ secondary growth method is then used to synthesize regular ZIF nanoparticles. The subsequent hydrolysis of Ce3+ in the presence of oxygen from the surrounding environment yields CeO2 product, which is then loaded onto the surface of the fiber and appears yellow. After that, annealing in an Ar environment yields a composite of Co3O4/CeO2@Co/N-CNF and an air environment yields a composite of the same composition. A distinct heterointerface structure was observed in the synthesized Co3O4/CeO2@Co/N-CNF.
At the interface, there was a distinct boundary between the two phases, with distinct groups of lattice fringes with interplanar spacings of 0.286 and 0.271 nm, representing the (220) and (200) planes of Co3O4 and CeO2, respectively. The heterointerface between Co3O4 and CeO2 is a clear indication of the strong interfacial electronic interaction between them. The XPS analysis reveals the presence of Co in the form of metallic, Co2+ and Co3+ and Ce in the form of Ce3+/Ce4+ oxidation states. The Ovac calculated from the O1s spectra; it was found that the Ovac ratio is found to be higher for the Co3O4/CeO2@Co/N-CNF catalyst compared to Co3O4@Co/N-CNF, indicating the significance of CeO2 in creating the Ovac in the catalyst. The unsaturated coordination environment is conducive to electron delocalization, and it has been generally reported that oxygen vacancies can accelerate the charge transfer rate by promoting more carriers to pass through. Also, oxygen vacancies can control the adsorption energy of the intermediate and enhance oxygen adsorption, which can lead to more efficient ORR kinetics.
According to the DFT studies by Li et al. [122], the oxygen adsorption on the defected oxygen vacancy metal oxides is ascribed to less filling of the antibonding states. In other words, the presence of oxygen defects moves the d-band center towards the Fermi level, indicating enhanced bond strength and adsorption of oxygen to the metal oxide surface. Therefore, the more oxygen vacancies, the better the electrocatalytic properties of the catalysts. The Co3O4/CeO2@Co/N-CNF catalyst was found to have more oxygen vacancies; therefore, it can deliver better ORR kinetics. Accordingly, the Co3O4/CeO2@Co/N-CNF catalyst presented a higher onset and half-wave potential of 0.88 V and 0.75 V, respectively, higher than the counter catalysts such as the Co3O4@Co/N-CNF and CeO2/N-CNF catalysts (Figure 10k–m).
Metal–organic framework-derived catalysts benefit from structural and morphological aspects such as the derived catalysts generally exhibiting large specific surface area and excellent porosity [123,124]. However, one of the persistent problems with MOF-derived catalysts is structural collapse and their susceptibility to agglomeration during the pyrolysis process [125]. Polymer surface functionalization of the ZIFs stands out to be one of the best strategies for mitigating structural collapse and agglomeration [126,127]. Among several polymers, conducting polymers such as Polyacrylinitrile and polypyrrole (PPy) are very beneficial for enhancing the ORR activity [128,129]. Tang et al. [130] described an efficient strategy in which Co-ZIFs are surface-functionalized with PPy precursor to develop hierarchical CeCo-Nx-CC catalysts for ORR (Figure 10n–p). The synthesis process involves three steps, starting with ZIFs synthesis from Co2+ and 2-MIM, after which Ce3+ ions are doped into the Ce-ZIFs by a solution incorporation strategy; finally, the surface functionalization of Co-Ce-ZIFs is conducted by polymerization of PPy. Further heat treatment of the Co-Ce-ZIF@PPy leads to the target catalyst CeCo-NxCC. In this work, interestingly, the role of Ce3+ as structural adjusting agent is established. The authors claim that the ZIF-67 particles appeared to be spherical initially, but thereafter, with the addition of Ce, promoted the formation of a dodecahedral structure. The TEM results reveal that the CeCo-NxCC-700 catalyst displays a dodecahedral hollow frame morphology; as expected, the dodecahedral structure remains un-collapsed, indicating the importance of PPy functionalization for structural tuning. The XRD and XPS analysis reveal the possibility of metal–Nx species in the catalyst as potential ORR catalytic centers. During RDE studies, the CeCo-NxCC-700 catalyst showed excellent ORR activity, with an onset potential of 0.95 V and a half-wave potential of 0.85 V vs. RHE, much higher than the catalyst prepared without PPy coating and Ce incorporation. This is further confirmed by the high ECSA value of the CeCo-NxCC-700 catalyst, which is 8.32 mF cm−2 higher than the catalyst without PPy coating and Ce incorporation, indicating that both PPy and Ce incorporation helped in enhancing the ORR active sites. In addition to the enhanced ORR activity, the CeCo-NxCC-700 catalyst presented excellent stability with a loss of just 10 mV in half-wave potentials when compared to the 20 mV loss in half-wave potential for Pt/C, for 8000 cycles, suggesting that N-doped carbon derived from the PPy helps to mitigate the nanoparticle agglomerations during a potential cycling test (Figure 10q–s).
Figure 10. (a) Schematic representation of Co3O4/CeO2@Co/N-CNF catalyst synthesis. (bj) SEM and TEM images of Co3O4/CeO2@Co/N-CNF catalyst and the colored elemental mapping images. (k) LSV (l) histogram representing Eonset, E1/2, and JL. (m) Tafel plots of Co3O4/CeO2@Co/N-CNF catalysts (Reproduced with permissions from Ref. [121]). (n,o) STEM elemental mapping and HRTEM images of the CeCo-NxCC-700 catalysts. (p) Digital micrographs of Ce(111), Co(111), and Co(200) on image (o) of the CeCo-NxCC-70 catalyst. (q,r) ORR LSV curves recorded in 0.1 M KOH electrolyte of CeCo-NxCC catalysts. (s) Histogram representing the onset, half-wave, and JL of CeCo-NxCC catalysts (Reproduced with permissions from Ref. [130]).
Figure 10. (a) Schematic representation of Co3O4/CeO2@Co/N-CNF catalyst synthesis. (bj) SEM and TEM images of Co3O4/CeO2@Co/N-CNF catalyst and the colored elemental mapping images. (k) LSV (l) histogram representing Eonset, E1/2, and JL. (m) Tafel plots of Co3O4/CeO2@Co/N-CNF catalysts (Reproduced with permissions from Ref. [121]). (n,o) STEM elemental mapping and HRTEM images of the CeCo-NxCC-700 catalysts. (p) Digital micrographs of Ce(111), Co(111), and Co(200) on image (o) of the CeCo-NxCC-70 catalyst. (q,r) ORR LSV curves recorded in 0.1 M KOH electrolyte of CeCo-NxCC catalysts. (s) Histogram representing the onset, half-wave, and JL of CeCo-NxCC catalysts (Reproduced with permissions from Ref. [130]).
Nanomaterials 15 00600 g010
In another study, Ce doping is found to be beneficial, not only for increasing the ORR activity but also in enhancing the stability of the transition metal catalysts, such as the Co@NC catalyst. A DFT study by Liu et al. [131] demonstrated that for the Co@NC catalyst, excessive energy is required for the adsorption of O2, which also requires a higher amount of heat adsorption for three elementary reactions (*O*OH → *O → *OH → * + OH), indicating that Co alone is not ideal for efficient ORR kinetics. When Ce atoms are introduced Co/Ce@NC, it was found that the kinetics enhances due to the involvement of Ce 4f electrons. It was found that with the introduction of Ce, optimal adsorption of the O2 intermediates and quick description of the products have been observed. In addition, the addition of Ce was found to have a stability effect on the irreversible aggregation of atomic cobalt sites into Co nanoparticles, with the help of assessing the binding energy (ΔEb) between Co/Ce single atoms and carbon-based defect vacancies. In general, higher ΔEb values represent the stable atomic positioning of the metallic active sites in the carbon system. The obtained ΔEb values are 7.66 and 9.25 eV for Co@NC and Ce@NC catalysts, respectively. This clearly indicates that Ce atoms have high stability. In addition, cohesive energiy (ΔEc) is the energy required to form metal-to-metal bonds; for both Co and Ce, this is around 5.4 eV [132,133]. For Co atoms, the ΔEb is lower than the Ce; the Co atoms are prone to undergo agglomeration, whereas for the high ΔEb of Ce with stronger binding energy, Ce-doping into the catalysts increases the distance between the Co atoms and hence acts as a dispersing agent and stabilizing agent, effectively mimicking the Co aggregation and hence increasing the overall stability of the electrocatalyst. Based on the DFT assumption, the authors synthesized the Co/Ce@NC catalyst and investigated the effect of ORR and Ce’s role as an ORR-enhancing element (Figure 11a–d). The SEM and TEM images provide evidence of the presence of dodecahedron structures of the Co/Ce@NC catalyst along with the CNTs grown on the surface. It was also seen that some of the Co nanoparticles are encapsulated by carbon in the form of CNTs. XRD results show Co present in the metallic state. XPS results show the presence of Ce in the form of Ce3+/Ce4+. Peaks around 896 and 915 eV in the high-resolution Ce 3d XPS spectrum of Co/Ce@NC are ascribed to the Ce3+ and Ce4+ species. Interconversion between Ce4+ and Ce3+ leads to enhanced charge redistribution on Co, which results in enhanced ORR kinetics. The Co/Ce@NC catalyst in RDE studies delivered a half-wave potential of 0.80 V, higher than the Co@NC at 0.77 V, along with the lowest Tafel slopes of 64 mV dec−1. Furthermore, the Co/Ce@NC catalyst generated the lowest % of peroxide at 3.17%, with a n = 3.94 indicating excellent ORR activity (Figure 11e–j).
Another family of electrocatalysts that is abundant in nature and has great catalytic activity are transition metal oxides, phosphides, and sulfides, which can include elements like nickel, iron, and cobalt. Co9S8 has recently drawn a lot of interest in the energy field due to its catalytic activity towards OER. Its ORR catalysis still requires improvement though, because of the disadvantage of poor conductivity, low surface area, and its susceptibility to aggregation during the electrochemical reaction [134,135]. Because of its rich oxygen vacancies and multivalent electron transfer, CeO2 may effectively encourage electrocatalytic reactions [136,137]. Sun et al. [138] presented a strategy to synthesize a heterostructured Co9S8/CeO2/Co-NC catalyst as an excellent strategy to improve the catalytic performance of Co9S8. The resulting Co9S8/CeO2/Co-NC electrocatalyst demonstrated a positive half-wave potential for ORR (0.875 V vs. RHE), nearly affecting ORR activity with equal effect to the commercial Pt/C catalyst, thanks to the special heterostructure of Co9S8/CeO2 and its synergistic effects with Co. This study emphasizes the effect of Ce as a structural directing agent. The SEM morphology of Co-ZIF shows a spherical 3D structure, whereas Co/Ce-ZIF shows a 2D sheet morphology, suggesting the phase transition ability of Ce atoms in the ZIF structures. After pyrolysis, the SEM images show numerous nanotubes grown on the surface of 2D Co/Ce-ZIF. After a sulfidation process, a 2D Co9S8/CeO2/Co-NC electrocatalyst, i.e., a “senbei”-like nanosheet structure was obtained. The TEM measurements of the 2D Co9S8/CeO2/Co-NC catalyst show well-dispersed metallic Co and Co9S8 nanoparticles in close conjunction with CeO2, forming a compact heterostructured catalyst. XPS analysis clearly showed that CeO2 affects the charge distribution on the surface of Co9S8 due to the observable shift in the XPS analysis by 0.17 eV for Co9S8/CeO2/Co-NC when compared to Co9S8/Co-NC. The ratio of Ce3+/Ce4+ is an important indicator of the oxygen buffering capacity of CeO2, due to its ability to interexchange between the oxidation states. The Ce3+/Ce4+ ratio for Co9S8/CeO2/Co-NC is ∼0.27, that of CeO2/Co-NC is∼0.25. Electron interaction in the heterostructure may enhance the conversion from Ce4+ to Ce3+, as indicated by the rise in the Ce3+/Ce4+ ratio [139]. Unsaturated chemical bonds are formed inside CeO2 as a result of Ce3+, creating a charge imbalance. The formation of numerous oxygen vacancies as a result of charge compensation improves ORR performance and encourages oxygen adsorption on the catalyst surface [140,141]. The RDE studies show that the Co9S8/CeO2/Co-NC catalyst exhibited a half-wave potential of 0.875 V, much higher than the CeO2/Co-NC (0.83 V) and Co9S8/Co-NC catalysts (0.85 V). The high half-wave potential values of the Co9S8/CeO2/Co-NC catalyst indicates the heterostructured morphology and presence of CeO2 and Co9S8 and that the synergistic effect between the three components of metallic Co, CeO2, and Co9S8 plays a significant role in enhancing ORR activity. The enhanced ORR activity is further ascertained from the lowest Tafel slope of 56 mV dec−1 and the average number of electrons of 4.0 and 3.8 electrons from K-L plots and RRDE measurements, respectively. The % of peroxide yield of Co9S8/CeO2/Co-NC catalyst is 8.3%, which is the lowest of the CeO2/Ce-NC (23%) and Co9S8/Co-NC (9.8%) catalysts. Furthermore, the Co9S8/CeO2/Co-NC catalyst also showed admirable stability under potential cycling conditions, with a loss of just 10 mV in the half-wave potential, whereas Pt/C lost about 50 mV. In chronoamperometric analysis at 0.8 V for 30,000 s, the Co9S8/CeO2/Co-NC catalyst lost just 9% of the relative current, whereas Pt/C lost about 25% of the relative current, indicating that the Co9S8/CeO2/Co-NC catalyst possess higher ORR activity.
To conclude briefly, it was found that the relative abundance of Ce3+ is the prime factor responsible for enhancing the ORR activity of Co-based catalysts due to its role in charge compensation which generates oxygen vacancies to promote the oxygen adsorption on the Co-N-C catalyst surface. Furthermore, the continuous availability of a higher valance state of Co is possible due to the lattice oxygen provided by Ce/CeO2, keeping the catalytic cycle occurring continuously. From the significantly increased Co3+ species observed in Co-based catalysts, it can be inferred that the electronic structure of Co in Co3O4 can be easily fine-tuned by introducing the Ce species, owing to their redox coupling reaction between Co3+/Co2+ and Ce3+/Ce4+. Because of the strong electronic coupling effect, it is believed that hybridizing p-type Co3O4 and n-type CeO2 will increase the activity of oxygen catalysis. By combining n-type and p-type materials, engineers can design and fine-tune an electrical structure that is advantageous for optimizing surface properties for enhanced ORR, which is experimentally observed. The presence of oxygen defects in CeO2 moves the d-band center towards the Fermi level, indicating an enhanced bond strength and adsorption of oxygen to the metal oxide surface, thereby promoting the Co-based catalysts to exhibit improved ORR performance.

6. Atomically Dispersed Ce-Based and CeO2 Catalysts for ORR

The role of Ce in enhancing electrocatalytic activity is clearly seen in the case of the Pt/CeOx, Fe/CeO2/C, and Co/CeO2/C catalysts for ORR. However, several studies have shown that pristine CeO2 in combination with N-doped carbon itself could possess admirable ORR activity. This is because, due to the abundant oxygen vacancies and quick transition between Ce3+ and Ce4+, Ce itself can act as ORR active site. In addition, when compared to Ce in the form of CeO2 nanoparticles, the atomically dispersed Ce catalysts, commonly known as single-atom catalysts (SACs), have gained tremendous popularity in recent years due to their homogeneous dispersion, well-defined active sites, efficient metal atom utilization and unique co-ordination environment in the catalysts. This section describes the recent trends in the intrinsic catalytic activity of CeO2 in combination with N-doped carbon and atomically dispersed Ce-N4-C catalyst ORR activity.
Yu et al. [142] demonstrated the use of a CeO2@ZIF-8-derived CeO2@N-doped carbon as an ORR catalyst. The synthesis process included the CeO2 encapsulation of ZIF-8 in the presence of surfactant PVP, leading to the trapping of the CeO2 nanoparticles inside the ZIF-8 structures. The pyrolysis temperature was found to have a direct influence on the ORR activity of CeO2@N-doped carbon as an ORR catalyst. The RDE studies clearly show the increase of half-wave potentials with the increase in pyrolysis temperatures, with optimum ORR activity achieved with the pyrolysis temperature of 900 °C for the catalyst CeO2@N-C-900, with an onset potential of 1.003 V and halfwave potential of 0.908 V vs. RHE in 0.1 M KOH electrolyte, surpassing the commercial Pt/C catalyst. In addition, the CeO2@N-C-900 catalyst showed an average number of electrons transferred per O2 molecule be 3.97, confirming the excellent ORR activity of the CeO2@N-C-900 catalyst. However, the CeO2@N-C-900 catalyst delivered a slightly lower ORR performance in acidic electrolytes. In addition to the enhanced ORR activity, the CeO2@N-C-900 catalyst also presented excellent durability, with almost no activity loss in 0.1 KOH and about 21 mV loss of activity in 0.1 M HClO4 electrolytes. In another study, Wen et al. [143] investigated the effect of CeO2 nanoparticle grown on N-doped carbon (CeO2–CN) and its ORR activity. The CeO2–CN nanoparticles were synthesized by a solid-state synthesis process, in which previously synthesized ZIF-8 nanoparticles are placed in a gate motor and mixed with Ce-nitrate and NaCl precursors and the resulting mixture is ground, followed by calcination and NaCl removal with H2O, leading to the formation of CeO2–CN electrocatalyst. Similarly to other studies, the authors claim that the enhanced ORR activity may be attributed to the content of Ce3+, evaluated from the XPS analysis, which described the role of Ce3+ in compensating for the oxygen vacancies, therefore indicating that the content of Ce3+ is directly proportional to the oxygen vacancies in the catalysts [144,145]. The CeO2–CN-800 catalyst showed a higher content of Ce3+; therefore, higher ORR activity can be expected. In RDE studies, the CeO2–CN-800 catalyst exhibited an onset potential of 0.90 V and a half-wave potential of 0.84 V, nearing the standard values of Pt/C which are 0.92 V and 0.85 V, respectively. In another study, in an ORR test conducted in alkaline media, cerium carbide with nitrogen-doped carbon (CeCx-NC) displayed a four-electron mechanism and a half-wave potential close to Pt/C [146].
When compared to Ce in the form of CeO2 nanoparticles, atomically dispersed Ce catalysts, commonly known as single-atom catalysts (SACs), have gained tremendous popularity in recent years due to their homogeneous dispersion, well defined active sites, efficient metal atom utilization, and unique co-ordination environment in the catalysts [147,148]. In addition, the SACs with M-N-C active sites are certainly suitable candidates as electrocatalyst due to their high electronic conductivity due to the presence of N, which exhibits excellent resistance to poisoning spices and admirable stability [149,150,151]. The best M-N-C-based catalyst so far is composed of Fe-N-C, with the potential to replace the Pt-based catalysts [152,153,154,155]. However, when Fe-N-C catalysts are formed during the high-temperature pyrolysis process, Fe co-ordinates with N (Fe-Nx), with a variety of co-ordination numbers ranging from x = 2–6, leading to the formation of Fe-N2 and Fe-N5 in addition to Fe-N4. There is enough evidence regarding the electronic structure and mechanistic studies indicating that Fe-N4 sites are the most active and desirable ORR active sites that catalyze a direct four-electron reduction of O2 to H2O, whereas other forms of Fe-coordination sites such as Fe-N2 and Fe-N5 catalyze by a mixture of four and 2+2 O2 reduction pathways, which lead to the formation of H2O and H2O2 as the final products.
H2O2 is an undesirable product that reacts with the Fe2+ ions, leading to the formation of OH radicals that react with other catalyst components and membranes, potentially degrading them [156]. In this regard, Ce is a very interesting element which has the potential to mimic the disadvantages of Fe-Nx-based catalysts due to the fact that Ce has a radical scavenging capacity. Ce SACs have been found to have high-spin state configurations that are potentially the active sites for ORR [156]. However, the synthesis of the higher dispersed single-atom catalysts composed of Ce is challenging due to their high surface and migration energies, leading to the formation of Ce nanoparticles during the high-temperature pyrolysis step. Furthermore, due to its high affinity towards oxygen, Ce forms Ce-oxides [157]. Therefore, controlling the synthesis conditions is a key to obtaining the Ce SACs. Liu et al. [158] proposed a unique synthesis strategy in which atomic-level dispersion of Ce achieved Ce loading of as high as 15.9% with a unique catalytic active site composed of Ce-N4O2 SACs as an excellent ORR catalyst, along with superb radical scavenging ability. A unique iodine-induced metal phase modulation strategy was developed with the help of MET-6, which is a zinc-triazolate metal–organic framework (MOF) derivative. With this unique synthesis strategy, the formation of Ce nanoparticles is greatly avoided by achieving an excellent dispersion of the Ce atoms with an atomic percentage as high as 15.9%. For the detailed synthesis process, we ask the readers to refer to the experimental section of the article. The RDE studies show that Ce-N4O2 SACs revealed a half-potential of 0.85 V, Tafel slope of 76 mV dec1, and a Jk of 8.82 mA cm−2 at 0.85 V, surpassing the Pt/C catalyst. The radical scavenging capacity of Ce-N4O2 SACs catalyst is assessed with UV-vis absorption spectra (Figure 12a,b). The Ce-N4O2 SACs and Ce-NPs-NC SACs displayed lower absorbance at 417 nm, indicating that Ce-N4O2 SACs can effectively scavenge the reactive oxygen species. The ORR mechanism on the Ce-N4O2 SACs catalyst is assessed by a theoretical simulation. Two possible atomic models were considered, one of which involved placing the CeN4O2 planarly and other involving placing it at the curved location of the graphene-like carbon support. Both the planar and curved Ce-N4O2 showed dynamic stability, and both the models displayed effective electronic charge redistribution, among which curved Ce-N4O2 active sited displayed a greater charge density than the planar one. Badar charge analysis reveals 1.96 and 2.21 electrons being added from the N atoms to the Ce in planar and curved Ce-N4O2 active sites, respectively. In a free energy analysis, both planar and curved Ce-N4O2 active sites showed a downhill trend for the ORR intermediates suggesting the spontaneity of the ORR. The curved structure’s computed ORR over-potential under alkaline conditions was 0.28 eV, which was less than the planar site’s (0.40 eV) (Figure 12c–f). This displays the advantageous role that graphene curvature plays in maximizing ORR activity [159].
As previously mentioned, new research indicates that single-atom catalysts that are atomically dispersed have a different reactivity than catalysts based on nanoparticles. Instead of looking into Ce in the form of CeO2 nanoparticles, it is worthwhile looking into atomically dispersed Ce in N-doped carbons that are strictly confined in the carbon matrix. This is because the electrocatalytic activity can be greatly increased when the concentration of Ce species decreases from the nanoscale to the single atomic level, due to the enhanced surface free energy that aids in O2’s adsorption and subsequent reduction. The catalyst’s stability can be increased by fixing the cerium atoms with the aid of four coordinated -N atoms to create Ce-N4-C. This helps to reduce the leaching of cerium atoms. In a study, Peera et al. [160] described Ce-N4-C catalyst synthesis by a simple ZIF-based strategy and found that the Ce-N4-C catalyst possessed almost similar ORR kinetics as that of Pt/C. The successful insertion of Ce into the ZIF’s crystal lattice was confirmed from the X-ray diffraction analysis, in which the diffraction peak at 7.46° shifts to the higher 2θ angles for Ce-ZIF precursors; this is due to the ionic radius of Ce3+/Ce4+, which is larger than the Zn2+ ions (Figure 12g,h). The Ce-N4-C catalyst derived from the pyrolysis of the Ce-ZIFs presented an extraordinary BET surface area of 905 m2/g. The TEM images show the very small amount of Ce atoms in the atomically dispersed form. The XPS analysis shows the co-existence of Ce4+/Ce3+ in almost equal proportions, which guarantees the altering oxidation states during the ORR. In addition, the Ce4+ was 12% more abundant compared to Ce3+. While Ce3+ is good at scavenging OH radicals, Ce4+ is good at oxidizing H2O2. Therefore, the overall ORR catalytic activity is improved by the synergistic effect of the Ce4+/Ce3+ redox couple [161]. The optimized Ce-N-C-3 catalyst showed the best ORR activity, with a half-wave potential of 0.89 V vs. RHE, which is just 10 mV lower than the Pt/C catalyst, indicating that the Ce-N-C-3 catalyst could be a promising alternative to the Pt/C catalyst. Interestingly, the Ce-N-C-3 catalyst also showed the lowest H2O of just 4%, the lowest among several Ce-based catalysts presented in this review. The authors in this study also performed the peroxide test to confirm the ability of Ce4+/Ce3+’s redox couple scavenging capacity towards peroxide/hydroperoxide radicals. In a specialized test, H2O2 was added to the 0.1 M KOH solution under chronoamperometric conditions and current profiles were recorded for Pt/C and Ce-N-C-3 catalysts. The Pt/C catalyst showed current spiking with the addition of increasing levels of H2O2, whereas the Ce-N-C-3 catalyst showed an almost negligible rise in the currents, suggesting that the Ce-N-C-3 catalyst successfully deactivated or scavenged the H2O2/HO/hydroperoxyl radical (HO2•/−) (Figure 12i–k). These findings show that Ce-based catalysts are effective at catalyzing OH scavenging reactions, which reduces carbon support corrosion, accelerates the dissolution of particles and atomic atoms, breaks down the ionomer in the catalyst layer, and damages PEM/AEM membranes in fuel cells.
Though the Ce has excellent properties, such as its ability to interchange the oxidation states between Ce3+/Ce4+, it is challenging to obtain Ce in an atomically dispersed state due to its wide range of possible oxidation states from 0 to +4; therefore, Ce has the tendency to form several phases such as oxides, carbides, and metallic Ce during the high-temperature pyrolysis. Furthermore, controlling the porous nature of the catalyst is paramount to host a high density of atomically dispersed Ce atoms, which helps in exposing the high volume of the ORR active sites. Zhu et al. [162] presented highly dispersed Ce atom catalysts with a high coordination number for excellent ORR. The highly electrocatalytically active Ce SAS/HPNC catalyst was derived by a combined MOF and hard-template synthesis strategy by using SiO2 as a structural assisting agent, followed by acid leaching and a gas-migration process (Figure 13a). This study involves pyrolysis of the Ce-ZIF-8@SiO2 precursor at a relatively high temperature of 1150 °C, slightly higher than in serval studies described in the previous sections. Authors describe that, during a 1150 °C in flowing of N2, the surface cerium atoms of CeO2 were initially volatilized to generate a potential volatile cerium species which could be conveyed and captured by the defective nitrogen-rich porous carbon, resulting in an atomically dispersed Ce-SAS/HPNC catalyst. The Ce-SAS/HPNC catalyst showed excellent morphology with a highly porous rhombododecahedron single-particle (Figure 13b–d). Authors systematically established a relation between the Ce and Zn precursors and the obtained catalyst morphology. It was found that Ce’s precursor (with respect to Zn precursor) with 10–20 wt.% resulted in the traditional rhombododecahedron morphology of ZIF-8 in the Ce SAS/HPNC catalyst, whereas, with ZIF-8 and SiO2, a Ce precursor as high as 50 wt% has been utilized. With the SiO2 template strategy, the Ce-SAS/HPNC catalyst resulted in a surface area of 1095 m2 g−1 from an unusually low surface area of 61 m2 g−1 for 20Ce SAS/NC. This indicates that the adopted SiO2 hard-template strategy could enhance the surface area 10 times more and help host the high wt% of the Ce precursor and hence encourage the high density of Ce-N4-C catalysts for enhanced ORR activity. The FT-EXAFS analysis of the Ce-SAS/HPNC catalyst revealed the presence of Ce-N/O bonding at the 2.09 Å, which falls in between the bonding peaks of Ce-Cl (2.05 Å) and Ce-Cl (2.05 Å); therefore, there is a high probability that Ce might exist in the form of Ce-N/O. The ORR analysis of the Ce-SAS/HPNC catalyst in 0.1 M HClO4 delivered a high ORR onset potential of 1.04 V and a half-wave potential of 0.862 V vs. RHE, which are almost similar to commercial the Pt/C catalyst of 20 wt% of Pt. In addition, the Ce-SAS/HPNC catalyst also delivered an admirable performance in realistic working fuel cell conditions. The Ce-SAS/HPNC catalyst delivered a peak power density of 475 and 525 mW cm−2 in H2-O2 with 1 and 2 bar pressure. At 2 bar pressure, the Ce-SAS/HPNC catalyst showed a current density of 0.473 A cm−2 at 0.6 V (Figure 13e–h).
Though the atomically dispersed Ce atoms exhibited excellent ORR activity, there is still room to further enhance the ORR activity. The shielding effect of the inner electron shell of cerium results in the localization of the outermost 4f orbital electrons. Reshaping the coordination and geometrical structure of the active sites could lead to the transition of electrons from 4d104f1 to 4d84f3, leading to the formation of high-spin Ce active sites that enhance the unpaired f electrons to occupy the anti-π orbitals of O2 and generate optimal binding strength with ORR intermediates. To understand the importance of spin effects on ORR catalysts, it is important to understand the simple ORR process. The ORR process begins with gaseous O2 adsorbance on the catalyst surface and for the adsorbance of the electrons from the catalytic active site, say, a metallic site, to anti-bonding orbitals of O2 to occur. The high electron density of O2 weakens the O=O bond, allowing the O=O bond to break [163]. This entire reaction proceeds in several steps, with multiple reaction intermediates such as O*, OH*, and OOH*. During these steps, O2 changes its paramagnetic nature to a diamagnetic nature [164]. For efficient ORR, the catalyst surface with high spin is generally considered more favorable for O2 adsorption, as it helps to facilitate the spin-state matching of the adsorbed O2. This enhances the rapid electron transfer and ORR kinetics [165]. Therefore, the spin-state modulation of the catalyst surface could be an effective strategy to tune the ORR activity of the catalysts. In this regard, Ce spin state modulation is possible due to its nearly empty orbitals being able to accommodate high-spin electrons [166]. Based on this assumption, Zhao et al. [100] proposed a carboxylate mediated synthesis approach to tune the CeNCs catalyst toward ORR. The synthesized CeNC-40 showed an accordion-like stacked multilayer shape, with large pores between the layers to facilitate exposure of more active sites and mass transport along with an atomically dispersed, high density of Ce atoms inside the N-doped carbon matrix (Figure 13j–l). The ICP-OES analysis shows that a Ce atomic wt% as high as 7.33 is achieved without any agglomerations along with a surface area of CeNC-40 catalyst with 607 m2 g−1 with micro and mesopores in the catalysts.
The XPS analysis showed Ce-Nx moieties, suggesting an electronic interaction between Ce and Nx/OH ligands, which helps in re-distributing the electronic density around the Ce atoms and hence effects the local electron and spin configuration of the Ce in the catalyst matrix. As mentioned earlier, the ratio of Ce3+/Ce4+ is an important deciding factor of ORR. This is due to the Ce4+ (5s25p6)’s fully occupied electrons that are shielded with inner orbitals, whereas Ce3+ (5s25p64f1) has localized f-electrons which interact easily with the adsorbed O2 p electrons, thus promoting the effective absorption of ORR intermediates. Therefore, the atomic ratio of Ce3+/Ce4+ is an important deciding factor for Ce-based ORR catalysts. The Ce L3-edge X-ray absorption near edge structure (XANES) verified that the CeNC-40 catalyst contains Ce, predominantly in the form of Ce3+, which is the most active site of the catalyst. When applied as an ORR catalyst, CeNC-40 showed excellent halfwave potentials of 0.90, 0.80, and 0.7 V in alkaline, acidic, and neutral electrolytes, respectively (Figure 13m,n). The ORR activity of the CeNC-40 catalyst is gauged through theorical simulations based on the co-ordination environment derived from the EXAFS fitting result, suggesting that the coordination number of Ce-N/O is six. The theoretical simulations suggest that O2 adsorption occurs on the Ce-N/O active site by an end-on mode with the O-O bond length increased by 0.04Å, with a final bond length of 2.92Å on the Ce-O active site. This effective adsorption leads to d-p-f orbital hybridization and coupling, resulting in formation of *OH as RDS with an energy barrier of 0.2eV. The high-spin states of Ce are estimated by measuring magnetic susceptibility and magnetic moment of the CeNC-40 catalyst which are 5.7 and 5, respectively, which are significantly higher than those of the NC, which are 0.4 and 0, which confirms the high-spin sate of the Ce in the catalyst. PDOS analysis shows the electron insertion from the dz2 → dx2y2 of Ce into 4f orbitals of Ce, which causes the high-spin sate of Ce (Figure 13o–s). This electron transfer shifts the Ce electronic configuration to 4d104f1 → 4d84f3. The increased magnetic moment 1.95 of the Ce-NC-40 suggests that the adsorption and activation of O2 is elevated, with change in the spin-state of Ce promoting a direct four-electron reduction of O2.
Ce SAs with optimized local coordination microenvironments via heteroatoms are anticipated to significantly enhance ORR performance. The introduction of N atoms into the carbon matrix, together with Ce-Nx active sites, are found to enhance the ORR activity. Introducing multiple heteroatoms having electron donation/withdrawing ability could create multiple electron delocalization pathways to synergistically enhance ORR activity. In a study by Du et al. [157], introducing single-atom Ce catalysts doped with C, N, and P in the carbon matrix is found to enhance the ORR activity, surpassing the effect of the Pt/C catalyst. The experimental validation and DFT insights clearly show enhanced ORR activity for Ce SAs/PSNC compared to the Ce SAs/NC catalyst (Figure 14a–f). DFT analysis of the Ce SAs/PSNC and Ce SAs/NC catalysts clearly distinguishes the contribution of P and S on the electronic distribution near Fermi energy (Ef). It was found that the binding orbitals and electronic distribution on the Ce SAs/PSNC catalyst are larger than on the Ce SAs/PSNC catalyst due to additional contributions from the p orbitals’ S and P atoms, therefore enhancing electron delocalization and the electron transfer efficiency in Ce SAs/PSNC compared to the Ce SAs/NC catalyst. With P and S atoms in equal ratio, it was found that Ce 4f orbital electron density was closest to the Ef, whereas for the Ce SAs/NC catalyst there is a definite gap. From DFT analysis, it was found that an applied potential of 0.94 V is required for Ce SAs/PSNC to initiate the ORR, whereas it is 0.77 V for the Ce SAs/NC catalyst, suggesting that tuning the carbon support with different heteroatoms of electron donation or electron withdrawing properties can alter the electronic charge polarization and local change and spin densities favoring the ORR. The assumptions made from DFT studies have been clearly established from the experimental point of view. The Ce SAs/PSNC catalyst is synthesized from the poly(cyclotriphospazeneco-4,4-sulfonyldiphenol) (PZS) polymerization on the ZIF-8, onto which a gas-migration strategy was applied to vaporize atomic Ce from CeO2 at the temperature of 1150 °C to be incorporated into NSP-doped carbon. In the ORR electrocatalytic performance, the Ce SAs/PSNC catalyst exhibited a half-wave potential of 0.90 V vs. RHE, better than the Ce SAs/NC, PSNC-C, and Pt/C catalysts, along with the highest obtained kinetic current, surpassing several non-precious metal-based electrocatalysts for ORR and enhanced stability with a loss of just 4 mV after 5000 potential cycles (Figure 14a–f).
In a unique study, Liu et al. [167] demonstrated the role of the porous structure of the catalyst and its interaction with Ce-N4 active sites which certainly affect the active site accessibility and enhances the interaction with O2. Atomically dispersed Ce SACs were synthesized via pore-confinement-pyrolysis of metal Ce/phenanthroline complexes. The detailed experimental and theoretical analysis of the Ce SA/MC catalyst revealed that pore-coupling with CeN4 active sites can accelerate the electron transfer, resulting in lower free energy for the RDS step during ORR. The theoretical simulations suggest clear charge redistribution in N-C and CeN4 active sites. In the case of the Ce SA/MC catalyst, an apparent electron transfer from CeN4 to mesoporous carbon was noted. The Bader charge analysis of CeN4 and Ce SA/MC catalysts reveals that 1.22–1.26 e were transferred from Ce/C active sites to four coordinating N atoms of CeN4, whereas 1.22–1.39 e have been transferred to four coordinating N atoms of CeN4 in a mesoporous Ce SA/MC catalyst, suggesting that the porous nature of Ce SA/MC catalyst helps in enhanced charge transfer to the active sites. The potential free energy diagram for the RDS shows that an energy barrier of 0.24 eV was noticed for the Ce SA/MC catalyst, whereas it was 0.36 eV for simple CeN4. These results indicate that the mesoporous nature of the catalyst certainly helps with efficient charge transfer and ORR kinetics. Similar assumptions were experimentally derived for Ce SA/MC catalysts. The Ce SA/MC catalysts exhibited a half-wave potential of 0.845 V vs. RHE, similar to the commercial Pt/C catalyst, with an electron transfer number of 3.88 and loss of 9 mV half-wave potential in a stability test by potential cycling for 5000 cycles.
The DFT analysis by Jang et al. [44] revealed the effect of the type of oxygen vacancies in the CeO2 (111) on the OH and OOH radical scavenging efficacy. In terms of CeO2 radical scavenging capacity, the generation of Ce3+ is synchronized with the desorption (i.e., vacancy formation) and adsorption of oxygen on the ceria surface. The vacancies created are either in triangular or linear patterns. Jang et al. [R6] investigated the effect of the CeO2 vacancies on the radical scavenging mechanism. The DFT studies modeled with OH and OOH show that the negatively polarized O atom of OH and OOH interacts with the high-spin state Ce and this interaction involves the spin change (decrease) on the Ce atoms. Radical oxygen would be more drawn to the cerium atoms surrounding the triangular defect, increasing the magnitude of the binding energy for stronger adsorption since it has a high-spin state and prefers to associate with other high-spin state atoms. For instance, the OOH adsorption results in a spin decrease from 1.10 to 0.86 and for OH it is 1.08 to 0.82 on the triangular defect site, lower than the spin decrease in the linear defect sites. The OH and OOH binding energies are found to be −4.54 and −3.20 eV for triangular defects, better than the values on the linear defects. The PDOS analysis reveals that the 4f band of Ce on the defective surface would play a crucial role in forming a bond with the radical species. Multiple dense peaks for the triangular defect arise from the 4f peak above the Fermi level, moving down to the below the Fermi level (−0.5 to −0 eV). More enhanced OOH radical scavenging ability is attributed to these 4f peaks around the −0.5−0 eV region in the triangular defect rather than in the linear defect. Consequently, our PDOS calculations imply that the oxygen vacancy generates the electronic spin localization on Ce atoms at the defect site, therefore generating a downshift of the Ce 4f band to below the Fermi level. The extreme scavenging capacity of ceria depends on such a change in the Ce 4f band. When developing radical scavengers against OH and OOH to improve the durability of polymer electrolyte membranes in fuel cells, it is advised to use ceria with mostly triangular defects, considering their stability and radical scavenging capacity.
In short conclusion, the pristine CeO2 could alone possess admirable ORR activity. In addition, the atomically dispersed Ce atoms in the form of Ce-N4-C active sites have shown excellent electrocatalytic activity. In this regard, using MET-6, a zinc triazolate metal–organic framework derivative, is a praising strategy which could introduce Ce atoms with atomic levels as high as 15.9% atomic percentage, successfully avoiding the generation of Ce nanoparticles. Most single-atom Ce-based catalyst studies conclude a similar point with regard to the Ce3+/Ce4+ atomic ratio, which suggests that Ce3+ is good at scavenging OH radicals and that Ce4+ is good at oxidizing H2O2. The ratio of Ce3+/Ce4+ is an important deciding factor of ORR. This is due to the Ce4+ (5s25p6)’s fully occupied electrons that are shielded with inner orbitals, whereas Ce3+ (5s25p64f1) has localized f-electrons which interact easily with the adsorbed O2 p electrons, thus promoting the effective absorption of ORR intermediates. Therefore, the atomic ratio of Ce3+/Ce4+ is an important deciding factor for Ce-based ORR catalysts. In addition, it was found that the increased magnetic moment of the Ce-NC-catalysts suggests that the adsorption and activation of O2 is elevated with a change in the spin-state of Ce, promoting a direct four-electron reduction of O2. Therefore, in the Pt-, Co-, and Fe-based catalysts, the CeO2 could itself be an ORR active site and, in close proximity to Pt, CoN4, and FeN4 active sites, it can act as a co-catalyst in modifying the electronic nature of the major active site due to its excellent oxygen storage capacity, enhancing stability and scavenging radical species. Table 1 shows the ORR half-cell and fuel cell/Zn–air battery performance data of Ce-based catalysts [168,169,170,171,172] derived from metal–organic framework precursors.

7. Conclusions

Rare earth metal-based catalysts with 4f electronic shells are popular in electrocatalysis. With its rich oxygen vacancies and affinity and facile exchange of Ce3+ and Ce4+ oxidation states, Ce is especially attractive for electrochemical reactions like oxygen reduction reaction (ORR). CeO2 also interacts strongly with noble metals like Pt and non-precious metals, enhancing metal/metal oxide interactions and promoting ORR by optimizing reaction intermediate adsorption. Most importantly, Ce is one of the best radical scavenging agents for fuel cell catalysts because it can interexchange Ce3+/Ce4+ oxidation states. CeO2 is stable in acidic conditions and improves carbon corrosion resistance in acidic electrolytes. Additionally, ceria promotes noble metal dispersion and improves O2 adsorption and ORR kinetics. Ceria-modified carbon compounds are effective ORR direct four-electron reduction catalysts. After a detailed analysis of the literature, it is found that Ce-based catalysts show excellent ORR activity. In the case of Pt/CeO2, the presence of Ce3+ was found to enhance the metallic ability of Pt nanoparticles and effectively inhibit the oxidation of Pt nanoparticles. The mass and specific activities of the Pt/CeOx/C catalysts are shown to be 10 times greater than the commercial Pt/C catalyst. The triple-phase interface structures formed from the Pt, CeOx, and C, are known to be responsible for the strong metal–support interactions (SMSI), contributing to the enhanced stability of Pt nanoparticles on the CeOx/C catalyst support. The Pt/CeO2-NC catalyst delivered a power density of 1.08 W cm−2, higher than Pt/NC and Pt/C catalysts. The enhanced ORR activity, stability, and fuel cell performance of the Pt/CeO2-NC catalyst was attributed to several causes, such as an effective three-phase interface structure.
In addition to the role of Ce as a co-catalyst and O2 buffering agent, Ce plays an important role in spatially confining and stabilizing Fe atoms. Ce doping allows charges to migrate between Fe and N elements, affecting the charge imbalance in the catalyst. This charge transfer effect lowers the d-band center value, which in turn reduces the adsorption energy of oxygen-containing intermediates in electrochemical reactions and increases the ORR electrocatalytic efficiency. The Ce 4f1 electrons have a high affinity for interacting with the transition metal atoms such as Fe and help in electron transfer from Ce → Fe, further enhancing the interaction of O2 on the Fe active sites, due to 4f1 localized electron transfer to the d-orbitals of the Fe active site due to their difference in electronegativity values, therefore enhancing the ORR activity of the Fe-N-C catalysts, as seen from the literature discussed inside the manuscript. The synergistic effect of ceria not only prevails in Fe-based ORR catalysts, but also in Co-based catalysts for ORR. The buffering (store/release mechanism of oxygen) capacity of CeO2 can be further tuned by choosing alternative Ce precursors to the traditional Ce nitrates. It was found that with a Ce2(OH)4SO4·2H2O precursor, the catalyst could generate more oxygen vacancies than by using traditional Ce precursors such as Ce nitrates. CeO2 also acts as a coupling metal oxide in the heterostructure catalysts with other transition metal oxides, exhibiting synergistic effects such as the heterostructure of the CO3O4-MnO2/C catalyst, delivering enhanced ORR activity, owing to a direct four-electron reduction of O2 as a result of covalent coupling between CO3O4-MnO2. Because of the strong electronic coupling effect, it is believed that hybridizing p-type Co3O4 and n-type CeO2 will increase the activity of oxygen catalysis. By combining n- and p-type materials, engineers can design and fine-tune an electrical structure that is advantageous for optimizing surface properties for enhanced ORR. Ce doping is found to be beneficial, not only for increasing the ORR activity but also for enhancing the stability of the transition metal catalysts such as the Co@NC catalyst. The addition of Ce was found to have a stability effect on the irreversible aggregation of atomic cobalt sites into Co nanoparticles, with the help of assessing the binding energy (ΔEb) between Co/Ce single atoms and carbon-based defect vacancies. Ce doping into the catalysts increases the distance between the Co atoms and hence acts as the dispersing agent and the stabilizing agent, effectively mimicking the Co aggregation and hence increasing the overall stability of the electrocatalyst. Atomically dispersed Ce atoms in the N-doped carbon showed excellent ORR activity. This is because, due to the abundant oxygen vacancies and quick transition between Ce3+ and Ce4+, the Ce itself can act as ORR active site. Ce SACs were found to have high-spin state configurations that are potentially active sites for ORR. The Ce-N-C catalysts have been shown to possess very high surface areas, as high as 900 m2/g. The Ce3+ ↔ Ce4+ redox couples’ ability to scavenge the peroxide/hydroperoxide radicals is confirmed from the chronoamperometric conditions and oxidation current analysis profiles of Pt/C and Ce-N-C catalysts, which show a strong scavenging ability of Ce-based SACs. These results indicate that the rare earth Ce-based catalysts are particularly promising for ORR, either in an atomically dissevered state or in combination with transition metal atoms, or with Pt that results in electrocatalytically active and stable catalysts for ORR in fuel cells and metal–air batteries.

8. Future Perspectives

1. Pt/CeOx/C catalysts were found to enhance the stability of the catalyst. However, there are only a few studies on the fuel cell performance analysis of the Pt/CeOx/C catalysts. At this moment, most of the studies conclude either with RDE studies and/or with single-cell performance. These catalysts could be the best alternative to the traditional Pt/c catalyst, aiming to achieve best durability. However, care should be taken while optimizing the CeO2 content due to its poor electronic conductivity property.
2. Fe/Co-Ce-based catalysts have the potential to be alternative catalysts for ORR. However, most of the catalysts’ ORR activity is analyzed in a traditional three-electrode system, whose output might not always be translated to realistic fuel cells and Zn–air battery applications. Therefore, it is highly essential to gauge the activity of the catalysts in terms of deliverable power density in single cell/fuel cell stacks in order to really appreciate their performance. While several studies have performed Zn–air battery performance studies, fuel cell performance studies are rare and insufficient to include their use in realistic applications, including catalyst loading optimizations, catalyst thickness optimization, and catalyst ink optimizations.
3. There is enough evidence on the effect of Ce3+/Ce4+ redox couple on the radical scavenging in fuel cells, especially CeO2 as additive to the nafion membrane. However, the Fe/Co-Ce-based catalyst’s radical scavenging ability in the catalyst layer is just being studied. The authors believe it is essential to investigate the ability of Fe/Co-Ce-based catalysts and their radical scavenging capacity in realistic fuel cell/Zn–air battery conditions by employing in situ studies to understand the stability of the catalysts.
4. At present, there is a lot of ambiguity in comparing Fe/Co-Ce and Ce SACs with traditional Pt/C activity, because most of the studies express the superiority of the Fe/Co-Ce and Ce SACs in terms of half-wave potentials. In order to truly compare the superiority, it is essential to express the activity of the catalyst, as recommended by the Department of Energy (DoE) regulations, such as kinetic current, mass activity of the catalysts, and the stability tests, as recommended by the DoE.
5. Few studies have investigated the role of ceria in the preparation process, which not only spatially confines Fe atoms, but also traps and stabilizes Fe atoms, resulting in high atomically dispersed Fe contents in the catalysts. However, there is no clear understanding or discussion of the confining mechanism and the kinetic control of stabilizing the Fe atoms. Also, it is essential to validate the confining mechanism by theoretical simulations.
6. From the extensive experimental output, it is concluded that the catalyst having a high content of Ce3+ is an important aspect for ORR. However, the precise control and abundance of a specific oxidation state in the catalyst is not well understood in the studies. Therefore, efforts need to be undertaken through novel catalyst synthesis strategies and processing strategies
7. Through DFT analysis, it is derived that CeO2 with a triangular defective structure is the most energetically favored, considering for structure, stability, and radical scavenging capacity. However, there is no experimental evidence regarding CeO2 with a triangular defective structure. Synthesizing CeO2 with triangular and linear defects and experimentally validating them still needs to be explored.
8. Single-atom Ce-based catalysts are only just being explored as excellent ORR catalysts, intrinsically or in combination with transition metal atoms. In particular, the high-spin states of Ce spin mechanisms are very interesting. However, at this moment, DFT studies provide very limited understanding. It is essential to investigate more on the effect of electronic spin states and the effect of magnetic moments on the ORR in detail.
9. Fe/Co-ZIF synthesis in the presence of Ce always alters the kinetic and growth of ZIF, which has been observed experimentally, where one can see the slow appearance of the milky product/precipitate of ZIFs. From this point of view, kinetic studies are required to understand the growth kinetics of Ce-ZIFs.

Author Contributions

Conceptualization, S.G.P.; methodology, S.G.P.; validation, S.G.P.; writing—original draft preparation, S.G.P.; writing—review and editing, S.W.K. All authors have read and agreed to the published version of the manuscript.

Funding

The research was supported by the National Research Foundation of Korea (NRF), funded by the Korean Government, Ministry of Science and ICT (MSIT) (No. 2021R1F1A1046648), Republic of Korea.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bhuiyan, M.M.H.; Siddique, Z. Hydrogen as an alternative fuel: A comprehensive review of challenges and opportunities in production, storage, and transportation. Int. J. Hydrogen Energy 2025, 102, 1026–1044. [Google Scholar] [CrossRef]
  2. Zhang, Q.; Li, W.; Peng, J.; Xue, L.; He, G. Cold plasma activated Ni0/Ni2+ interface catalysts for efficient electrocatalytic methane oxidation to low-carbon alcohols. Green Chem. 2024, 26, 7091–7100. [Google Scholar] [CrossRef]
  3. Li, L.; Tang, X.; Wu, B.; Huang, B.; Yuan, K.; Chen, Y. Advanced architectures of air electrodes in zinc–air batteries and hydrogen fuel cells. Adv. Mater. 2024, 36, 2308326. [Google Scholar] [CrossRef]
  4. He, Y.; Yang, J.; Wang, Y.; Jia, Y.; Li, H.; Liu, Y.; Liu, L.; Tan, Q. Atomically dispersed dual-metal ORR catalyst with hierarchical porous structure for Zn–Air batteries. ACS Appl. Mater. Interfaces 2024, 16, 12398–12406. [Google Scholar] [CrossRef]
  5. Li, B.; Jiang, S.; Fu, Q.; Wang, R.; Xu, W.; Chen, J.; Song, B. Tailoring Nanocrystalline/Amorphous Interfaces to Enhance Oxygen Evolution Reaction Performance for FeNi-Based Alloy Fibers. Adv. Funct. Mater. 2025, 35, 2413088. [Google Scholar] [CrossRef]
  6. Poudel, M.B.; Balanay, M.P.; Lohani, P.C.; Sekar, K.; Yoo, D.J. Atomic Engineering of 3D Self-Supported Bifunctional Oxygen Electrodes for Rechargeable Zinc-Air Batteries and Fuel Cell Applications. Adv. Energy Mater. 2024, 14, 2400347. [Google Scholar] [CrossRef]
  7. Alhakemy, A.Z.; Wang, G.; Chen, K.; Hassan, A.E.; Wen, Z. Boosting Zn-air battery performance: Fe single-atom anchored on F, N co-doped carbon nanosheets for efficient oxygen reduction. J. Alloys Compd. 2025, 1010, 177166. [Google Scholar] [CrossRef]
  8. Jithul, K.P.; Tamilarasi, B.; Pandey, J. Electrocatalyst for the oxygen reduction reaction (ORR): Towards an active and stable electrocatalyst for low-temperature PEM fuel cell. Ionics 2024, 30, 6757–6787. [Google Scholar] [CrossRef]
  9. Saeidfar, A.; Kirlioglu, A.C.; Gursel, S.A.; Yesilyurt, S. Role of catalyst layer composition in the degradation of low platinum-loaded proton exchange membrane fuel cell cathodes: An experimental analysis. J. Power Sources 2025, 625, 235676. [Google Scholar] [CrossRef]
  10. Peera, S.G.; Koutavarapu, R.; Akula, S.; Asokan, A.; Moni, P.; Selvaraj, M.; Balamurugan, J.; Kim, S.O.; Liu, C.; Sahu, A.K. Carbon nanofibers as potential catalyst support for fuel cell cathodes: A review. Energy Fuels 2021, 35, 11761–11799. [Google Scholar] [CrossRef]
  11. Deng, X.; Ma, L.; Wang, C.; Ye, H.; Cao, L.; Zhan, X.; Tian, J.; Tong, X. Recent Progress in Materials Design and Fabrication Techniques for Membrane Electrode Assembly in Proton Exchange Membrane Fuel Cells. Catalysts 2025, 15, 74. [Google Scholar] [CrossRef]
  12. Zhou, Z.; Zhang, H.J.; Feng, X.; Ma, Z.; Ma, Z.F.; Xue, Y. Progress of Pt and iron-group transition metal alloy catalysts with high ORR activity for PEMFCs. J. Electroanal. Chem. 2024, 959, 118165. [Google Scholar] [CrossRef]
  13. Ziehl, T.J.; Li, J.; Sun, S.; Zhang, P. New Insights into the ORR Catalysis on Pt Alloy Nanoparticles from an Element Specific d-Band Analysis. J. Phys. Chem. Lett. 2024, 15, 8306–8314. [Google Scholar] [CrossRef] [PubMed]
  14. Peera, S.G.; Lee, T.G.; Sahu, A.K. Pt-rare earth metal alloy/metal oxide catalysts for oxygen reduction and alcohol oxidation reactions: An overview. Sustain. Energy Fuels 2019, 3, 1866–1891. [Google Scholar] [CrossRef]
  15. Wang, H.; Lin, R.; Cai, X.; Liu, S.; Zhong, D.; Hao, Z. Transition metal dissolution control in Pt-alloy catalyst layers for low Pt-loaded PEMFCs for improving mass transfer. Int. J. Heat Mass Transf. 2021, 178, 121615. [Google Scholar] [CrossRef]
  16. Li, X.; Wang, D.; Zha, S.; Chu, Y.; Pan, L.; Wu, M.; Liu, C.; Wang, W.; Mitsuzaki, N.; Chen, Z. Active sites identification and engineering of MNC electrocatalysts toward oxygen reduction reaction. Int. J. Hydrogen Energy 2024, 51, 1110–1127. [Google Scholar] [CrossRef]
  17. Zhao, T.; Wang, J.; Wei, Y.; Zhuang, Z.; Dou, Y.; Yang, J.; Li, W.-H.; Wang, D. From Lab-Scale to Industrialization: Atomically MNC Catalysts for Oxygen Reduction Reaction. Energy Environ. Sci. 2025. [Google Scholar] [CrossRef]
  18. Man, X.; Chang, Y.; Jia, J. Rare-earth-based catalysts for oxygen reduction reaction. Mol. Catal. 2024, 565, 114389. [Google Scholar] [CrossRef]
  19. Zheng, B.; Fan, J.; Chen, B.; Qin, X.; Wang, J.; Wang, F.; Deng, R.; Liu, X. Rare-earth doping in nanostructured inorganic materials. Chem. Rev. 2022, 122, 5519–5603. [Google Scholar] [CrossRef]
  20. Jiang, Y.; Fu, H.; Liang, Z.; Zhang, Q.; Du, Y. Rare earth oxide based electrocatalysts: Synthesis, properties and applications. Chem. Soc. Rev. 2024, 53, 714–763. [Google Scholar] [CrossRef]
  21. Li, C.; Wang, P.; He, M.; Yuan, X.; Fang, Z.; Li, Z. Rare earth-based nanomaterials in electrocatalysis. Coord. Chem. Rev. 2023, 489, 215204. [Google Scholar] [CrossRef]
  22. Gao, Q.; Wang, H.; Liu, C.; Luo, L.; Li, X.; Jiang, Q.; Ji, Y.; Zheng, T.; Xia, C. Regulatory Mechanisms and Applications of Rare Earth Elements-Based Electrocatalysts. Chin. J. Chem. 2025, 43, 205–218. [Google Scholar] [CrossRef]
  23. Wang, X.; Tang, Y.; Lee, J.-M.; Fu, G. Recent advances in rare-earth-based materials for electrocatalysis. Chem Catal. 2022, 2, 967–1008. [Google Scholar] [CrossRef]
  24. Cheng, H.; Guan, Y.; Wang, F.P.; Yang, T.; Huang, L.; Tang, J.; Hu, R. Oxygen-vacancies-rich CeOx/TbCoP nanoparticles for enhanced electrocatalytic oxygen evolution reaction. J. Alloys Compd. 2024, 972, 172820. [Google Scholar] [CrossRef]
  25. Xu, K.; Liu, J.C.; Wang, W.W.; Zhou, L.L.; Ma, C.; Guan, X.; Wang, F.R.; Li, J.; Jia, C.J.; Yan, C.H. Catalytic properties of trivalent rare-earth oxides with intrinsic surface oxygen vacancy. Nat. Commun. 2024, 15, 5751. [Google Scholar] [CrossRef]
  26. Ahn, S.Y.; Jang, W.J.; Shim, J.O.; Jeon, B.H.; Roh, H.S. CeO2-based oxygen storage capacity materials in environmental and energy catalysis for carbon neutrality: Extended application and key catalytic properties. Catal. Rev. 2024, 66, 1316–1399. [Google Scholar] [CrossRef]
  27. Zhang, C.; Jia, J.; Chang, Y.; Guo, S.; Jia, M. The corrosion resistant Pt5Ce–CeO2 structure provides significant oxygen reduction catalysis. Int. J. Hydrogen Energy 2024, 51, 1169–1175. [Google Scholar] [CrossRef]
  28. Xia, W.; Zhao, J.; Wang, T.; Song, L.; Gong, H.; Guo, H.; Gao, B.; Fan, X.; He, J. Anchoring ceria nanoparticles on graphene oxide and their radical scavenge properties under gamma irradiation environment. Phys. Chem. Chem. Phys. 2017, 19, 16785–16794. [Google Scholar] [CrossRef]
  29. Yu, Y.; Wang, X.; Gao, W.; Li, P.; Yan, W.; Wu, S.; Cui, Q.; Song, W.; Ding, K. Trivalent cerium-preponderant CeO2/graphene sandwich-structured nanocomposite with greatly enhanced catalytic activity for the oxygen reduction reaction. J. Mater. Chem. A 2017, 5, 6656–6663. [Google Scholar] [CrossRef]
  30. Yuan, X.; Ge, H.; Liu, X.; Wang, X.; Chen, W.; Dong, W.; Huang, F. Efficient catalyst of defective CeO2−x and few-layer carbon hybrid for oxygen reduction reaction. J. Alloys Compd. 2016, 688, 613–618. [Google Scholar] [CrossRef]
  31. Ashmath, S.; Wu, H.; Peera, S.G.; Lee, T.-G. Metal–Organic Framework-Derived Rare Earth Metal (Ce-NC)-Based Catalyst for Oxygen Reduction Reactions in Dual-Chamber Microbial Fuel Cells. Catalysts 2024, 14, 506. [Google Scholar] [CrossRef]
  32. Cai, D.; Yang, Z.; Tong, R.; Huang, H.; Zhang, C.; Xia, Y. Binder-Free MOF-Based and MOF-Derived Nanoarrays for Flexible Electrochemical Energy Storage: Progress and Perspectives. Small 2024, 20, 2305778. [Google Scholar] [CrossRef] [PubMed]
  33. Fan, Y.; Wang, W.; Chen, Y.; Xu, Z.; Cai, D.; Xu, M.; Tong, R. Cobalt-containing ZIF-derived catalysts for Zn–air batteries. Mater. Chem. Front. 2024, 8, 2394–2419. [Google Scholar] [CrossRef]
  34. Meng, S.; Li, G.; Wang, P.; He, M.; Sun, X.; Li, Z. Rare earth-based MOFs for photo/electrocatalysis. Mater. Chem. Front. 2023, 7, 806–827. [Google Scholar] [CrossRef]
  35. Molavi, H. Cerium-based metal-organic frameworks: Synthesis, properties, and applications. Coord. Chem. Rev. 2025, 527, 216405. [Google Scholar] [CrossRef]
  36. Reed, K.; Cormack, A.; Kulkarni, A.; Mayton, M.; Sayle, D.; Klaessig, F.; Stadler, B. Exploring the properties and applications of nanoceria: Is there still plenty of room at the bottom? Environ. Sci. Nano 2014, 1, 390–405. [Google Scholar] [CrossRef]
  37. Miles, M.H. Evaluation of electrocatalysts for water electrolysis in alkaline solutions. J. Electroanal. Chem. Interfac. Electrochem. 1975, 60, 89–96. [Google Scholar] [CrossRef]
  38. Melchionna, M.; Fornasiero, P. The role of ceria-based nanostructured materials in energy applications. Mater. Today 2014, 17, 349–357. [Google Scholar] [CrossRef]
  39. Xia, J.; Zhao, H.; Huang, B.; Xu, L.; Luo, M.; Wang, J.; Luo, F.; Du, Y.; Yan, C.-H. Efficient optimization of electron/oxygen pathway by constructing ceria/hydroxide interface for highly active oxygen evolution reaction. Adv. Funct. Mater. 2020, 30, 1908367. [Google Scholar] [CrossRef]
  40. Kaiser, S.K.; Chen, Z.; Akl, D.F.; Mitchell, S.; Pérez-Ramírez, J. Single-atom catalysts across the periodic table. Chem. Rev. 2020, 120, 11703–11809. [Google Scholar] [CrossRef]
  41. Zhang, S.; Saji, S.E.; Yin, Z.; Zhang, H.; Du, Y.; Yan, C.-H. Rare-earth incorporated alloy catalysts: Synthesis, properties, and applications. Adv. Mater. 2021, 33, 2005988. [Google Scholar] [CrossRef] [PubMed]
  42. Zhang, Q.; Peng, J.; Xiong, H.; Jiang, S.; Li, W.; Fu, X.; Shang, S.; Xu, J.; He, G. Cold plasma-activated Ni-Co tandem catalysts with CN vacancies for enhancing CH4 electrocatalytic to methyl formate. Appl. Catal. B Environ. Energy 2025, 362, 124759. [Google Scholar] [CrossRef]
  43. Tian, Y.; Cai, Y.; Chen, Y.; Jia, M.; Hu, H.; Xie, W.; Li, D.; Song, H.; Guo, S.; Zhang, X. Accessing the O Vacancy with Anionic Redox Chemistry Toward Superior Electrochemical Performance in O3 type Na-Ion Oxide Cathode. Adv. Funct. Mater. 2024, 34, 2316342. [Google Scholar] [CrossRef]
  44. Lawler, R.; Cho, J.; Ham, H.C.; Ju, H.; Lee, S.W.; Kim, J.Y.; Choi, J.I.; Jang, S.S. CeO2(111) surface with oxygen vacancy for radical scavenging: A density functional theory approach. J. Phys. Chem. C 2020, 124, 20950–20959. [Google Scholar] [CrossRef]
  45. Sun, H.; Tian, C.; Fan, G.; Qi, J.; Liu, Z.; Yan, Z.; Cheng, F.; Chen, J.; Li, C.-P.; Du, M. Boosting activity on Co4N porous nanosheet by coupling CeO2 for efficient electrochemical overall water splitting at high current densities. Adv. Funct. Mater. 2020, 30, 1910596. [Google Scholar] [CrossRef]
  46. Wang, C.; Lv, X.; Zhou, P.; Liang, X.; Wang, Z.; Liu, Y.; Wang, P.; Zheng, Z.; Dai, Y.; Li, Y.; et al. Molybdenum nitride electrocatalysts for hydrogen evolution more efficient than platinum/carbon: Mo2N/CeO2@ nickel foam. ACS Appl. Mater. Interfaces 2020, 12, 29153–29161. [Google Scholar] [CrossRef]
  47. Peng, W.; Zhao, L.; Zhang, C.; Yan, Y.; Xian, Y. Controlled growth cerium oxide nanoparticles on reduced graphene oxide for oxygen catalytic reduction. Electrochim. Acta 2016, 191, 669–676. [Google Scholar] [CrossRef]
  48. Zhang, N.; Jiang, R. Interfacial engineering of metal/metal oxide heterojunctions toward oxygen reduction and evolution reactions. ChemPlusChem 2021, 86, 1586–1601. [Google Scholar] [CrossRef]
  49. Yook, S.H.; Kim, H.Y.; Kim, S.J.; Choi, S.; Kwon, T.; Cho, H.; Kim, J.M.; Yoon, K.R.; Jo, S.; Lee, S.Y.; et al. Boosting antioxidation efficiency of nonstoichiometric CeOx nanoparticles via surface passivation toward robust polymer electrolyte membrane fuel cells. Chem. Eng. J. 2022, 432, 134419. [Google Scholar] [CrossRef]
  50. Oh, H.; Son, B.; Shanmugam, S. Cerium-based perovskite mixed metal oxide as the radical scavenger for PEM fuel cells operating under low humidity conditions. ACS Appl. Mater. Interfaces 2023, 15, 28093–28105. [Google Scholar] [CrossRef]
  51. Zhang, N.; Yan, H.; Li, L.; Wu, R.; Song, L.; Zhang, G.; Liang, W.; He, H. Use of rare earth elements in single-atom site catalysis: A critical review. J. Rare Earths 2021, 39, 233–242. [Google Scholar] [CrossRef]
  52. Schauermann, S.; Hoffmann, J.; Johánek, V.; Hartmann, J.; Libuda, J.; Freund, H.-J. Catalytic activity and poisoning of specific sites on supported metal nanoparticles. Angew. Chem. Int. Ed. 2002, 41, 2532–2535. [Google Scholar] [CrossRef]
  53. Liu, L.; Corma, A. Metal catalysts for heterogeneous catalysis: From single atoms to nanoclusters and nanoparticles. Chem. Rev. 2018, 118, 4981–5079. [Google Scholar] [CrossRef]
  54. Yang, X.-F.; Wang, A.; Qiao, B.; Li, J.; Liu, J.; Zhang, T. Single-atom catalysts: A new frontier in heterogeneous catalysis. Acc. Chem. Res. 2013, 46, 1740–1748. [Google Scholar] [CrossRef]
  55. Boronin, A.I.; Slavinskaya, E.M.; Figueroba, A.; Stadnichenko, A.I.; Kardash, T.Y.; Stonkus, O.A.; Fedorova, E.A.; Muravev, V.V.; Svetlichnyi, V.A.; Bruix, A.; et al. CO oxidation activity of Pt/CeO2 catalysts below 0 °C: Platinum loading effects. Appl. Catal. B Environ. 2021, 286, 119931. [Google Scholar] [CrossRef]
  56. Li, Z.; Ji, S.; Liu, Y.; Cao, X.; Tian, S.; Chen, Y.; Niu, Z.; Li, Y. Well-defined materials for heterogeneous catalysis: From nanoparticles to isolated single-atom sites. Chem. Rev. 2019, 120, 623–682. [Google Scholar] [CrossRef]
  57. Zhang, P.; Liu, Y.; Liu, S.; Zhou, L.; Wu, X.; Han, G.; Liu, T.; Sun, K.; Li, B.; Jiang, J. Precise Design and Modification Engineering of Single-Atom Catalytic Materials for Oxygen Reduction. Small 2024, 20, 2305782. [Google Scholar] [CrossRef]
  58. Shudin, H.; Eguchi, R.; Ueda, S.; Ankit, S.; Hashimoto, A.; Abe, H. Heteroepitaxial Interface of Pt//CeO2 Nanoparticles for Enhanced Catalysis in Oxygen Reduction Reaction (ORR). J. Mater. Chem. A 2025, 13, 9660–9664. [Google Scholar] [CrossRef]
  59. Masuda, T.; Fukumitsu, H.; Fugane, K.; Togasaki, H.; Matsumura, D.; Tamura, K.; Nishihata, Y.; Yoshikawa, H.; Kobayashi, K.; Mori, T.; et al. Role of cerium oxide in the enhancement of activity for the oxygen reduction reaction at Pt–CeOx nanocomposite electrocatalyst-an in situ electrochemical X-ray absorption fine structure study. J. Phys. Chem. C 2012, 116, 10098–10102. [Google Scholar] [CrossRef]
  60. Luo, Y.; Calvillo, L.; Daiguebonne, C.; Daletou, M.K.; Granozzi, G.; Alonso-Vante, N. A highly efficient and stable oxygen reduction reaction on Pt/CeOx/C electrocatalyst obtained via a sacrificial precursor based on a metal-organic framework. Appl. Catal. B Environ. 2016, 189, 39–50. [Google Scholar] [CrossRef]
  61. Lei, L.; Li, F.; Jiang, X.; Yang, H.; Yu Xia, B. Corrosion and Protection of Electrocatalysts toward Advanced Energy Technologies. Adv. Funct. Mater. 2024, 34, 2405726. [Google Scholar] [CrossRef]
  62. Zhao, Z.-G.; Guo, P.; Shen, L.-X.; Liu, Y.-Y.; Zhang, Z.-Y.; Tu, F.-D.; Ma, M.; Liu, X.-W.; Zhang, Y.-L.; Zhao, L.; et al. Triple-phase interfacial engineering Pt-CeO2-nitrogen-doped carbon electrocatalysts for proton exchange membrane fuel cells. Appl. Surf. Sci. 2023, 609, 155302. [Google Scholar] [CrossRef]
  63. Lu, Q.; Wang, Z.; Tang, Y.; Huang, C.; Zhang, A.; Liu, F.; Liu, X.; Shan, B.; Chen, R. Well-controlled Pt–CeO2–nitrogen doped carbon triple-junction catalysts with enhanced activity and durability for the oxygen reduction reaction. Sustain. Energy Fuels 2022, 6, 2989–2995. [Google Scholar] [CrossRef]
  64. Xu, F.; Cheng, K.; Yu, Y.; Mu, S. One-pot synthesis of Pt/CeO2/C catalyst for enhancing the SO2 electrooxidation. Electrochim. Acta 2017, 229, 253–260. [Google Scholar] [CrossRef]
  65. Xiao, L.; Bao, W.; Zhang, J.; Yang, C.; Ai, T.; Li, Y.; Wei, X.; Jiang, P.; Kou, L. Interfacial interaction between NiMoP and NiFe-LDH to regulate the electronic structure toward high-efficiency electrocatalytic oxygen evolution reaction. Int. J. Hydrogen Energy 2022, 47, 9230–9238. [Google Scholar] [CrossRef]
  66. Ye, X.; Wang, H.; Lin, Y.; Liu, X.; Cao, L.; Gu, J.; Lu, J. Insight of the stability and activity of platinum single atoms on ceria. Nano Res. 2019, 12, 1401–1409. [Google Scholar] [CrossRef]
  67. Chen, A.; Lu, J.; Zhu, H.; Zhang, H.; Zeng, S.; Zheng, L.; Liang, H.-P. Construction of highly durable electrocatalysts by pore confinement and anchoring effect for the oxygen reduction reaction. New J. Chem. 2022, 46, 7253–7262. [Google Scholar] [CrossRef]
  68. D’Urso, C.; Oldani, C.; Baglio, V.; Merlo, L.; Aricò, A.S. Fuel cell performance and durability investigation of bimetallic radical scavengers in Aquivion® perfluorosulfonic acid membranes. Int. J. Hydrogen Energy 2017, 42, 27987–27994. [Google Scholar] [CrossRef]
  69. Ji, S.; Zhang, C.; Guo, R.; Jiang, Y.; He, T.; Zhan, Q.; Li, R.; Zheng, Y.; Li, Y.; Dai, S.; et al. Effect of interfacial interaction on electrocatalytic activity and durability of Pt-based core–shell nanocatalysts. ACS Catal. 2024, 14, 11721–11732. [Google Scholar] [CrossRef]
  70. Mao, G.; Zhou, Q.; Wang, B.; Xiong, Y.; Zheng, X.; Ma, J.; Fu, L.; Luo, L.; Wang, Q. Modulating d-Orbital electronic configuration via metal-metal oxide interactions for boosting electrocatalytic methanol oxidation. J. Colloid Interface Sci. 2025, 677, 657–665. [Google Scholar] [CrossRef]
  71. Ramani, S.; Sarkar, S.; Vemuri, V.; Peter, S.C. Chemically designed CeO2 nanoboxes boost the catalytic activity of Pt nanoparticles toward electro-oxidation of formic acid. J. Mater. Chem. A 2017, 5, 11572–11576. [Google Scholar] [CrossRef]
  72. Xu, H.; Wang, A.-L.; Tong, Y.-X.; Li, G.-R. Enhanced catalytic activity and stability of Pt/CeO2/PANI hybrid hollow nanorod arrays for methanol electro-oxidation. ACS Catal. 2016, 6, 5198–5206. [Google Scholar] [CrossRef]
  73. Du, C.; Gao, X.; Cheng, C.; Zhuang, Z.; Li, X.; Chen, W. Metal organic framework for the fabrication of mutually interacted PtCeO2C ternary nanostructure: Advanced electrocatalyst for oxygen reduction reaction. Electrochim. Acta 2018, 266, 348–356. [Google Scholar] [CrossRef]
  74. Montini, T.; Melchionna, M.; Monai, M.; Fornasiero, P. Fundamentals and catalytic applications of CeO2-based materials. Chem. Rev. 2016, 116, 5987–6041. [Google Scholar] [CrossRef]
  75. Du, J.S.; Bian, T.; Yu, J.; Jiang, Y.; Wang, X.; Yan, Y.; Jiang, Y.; Jin, C.; Zhang, H.; Yang, D. Embedding Ultrafine and High-Content Pt Nanoparticles at Ceria Surface for Enhanced Thermal Stability. Adv. Sci. 2017, 4, 1700056. [Google Scholar] [CrossRef]
  76. Jia, Q.; Ghoshal, S.; Li, J.; Liang, W.; Meng, G.; Che, H.; Zhang, S.; Ma, Z.-F.; Mukerjee, S. Metal and metal oxide interactions and their catalytic consequences for oxygen reduction reaction. J. Am. Chem. Soc. 2017, 139, 7893–7903. [Google Scholar] [CrossRef]
  77. Wang, W.; Dong, Y.; Yang, Y.; Chai, D.; Kang, Y.; Lei, Z. CeO2 overlapped with nitrogen-doped carbon layer anchoring Pt nanoparticles as an efficient electrocatalyst towards oxygen reduction reaction. Int. J. Hydrogen Energy 2018, 43, 12119–12128. [Google Scholar] [CrossRef]
  78. Chen, J.; Li, Z.; Chen, Y.; Zhang, J.; Luo, Y.; Wang, G.; Wang, R. An enhanced activity of Pt/CeO2/CNT triple junction interface catalyst prepared by atomic layer deposition for oxygen reduction reaction. Chem. Phys. Lett. 2020, 755, 137793. [Google Scholar] [CrossRef]
  79. Brummel, O.; Waidhas, F.; Faisal, F.; Fiala, R.; Vorokhta, M.; Khalakhan, I.; Dubau, M.; Figueroba, A.; Kovács, G.; Aleksandrov, H.A.; et al. Stabilization of small platinum nanoparticles on Pt–CeO2 thin film electrocatalysts during methanol oxidation. J. Phys. Chem. C 2016, 120, 19723–19736. [Google Scholar] [CrossRef]
  80. Lykhach, Y.; Kozlov, S.M.; Skála, T.; Tovt, A.; Stetsovych, V.; Tsud, N.; Dvořák, F.; Johánek, V.; Neitzel, A.; Mysliveček, J.; et al. Counting electrons on supported nanoparticles. Nat. Mater. 2016, 15, 284–288. [Google Scholar] [CrossRef]
  81. Xie, Z.; Yan, B.; Kattel, S.; Lee, J.H.; Yao, S.; Wu, Q.; Rui, N.; Gomez, E.; Liu, Z.; Xu, W.; et al. Dry reforming of methane over CeO2-supported Pt-Co catalysts with enhanced activity. Appl. Catal. B Environ. 2018, 236, 280–293. [Google Scholar] [CrossRef]
  82. Tao, L.; Shi, Y.; Huang, Y.-C.; Chen, R.; Zhang, Y.; Huo, J.; Zou, Y.; Yu, G.; Luo, J.; Dong, C.-L.; et al. Interface engineering of Pt and CeO2 nanorods with unique interaction for methanol oxidation. Nano Energy 2018, 53, 604–612. [Google Scholar] [CrossRef]
  83. Zhang, Z.; Liu, S.; Li, X.; Qin, T.; Wang, L.; Bo, X.; Liu, Y.; Xu, L.; Wang, S.; Sun, X.; et al. Lewis-basic lanthanide metal-organic framework-derived versatile multi-active-site synergistic catalysts for oxygen reduction reaction. ACS Appl. Mater. Interfaces 2018, 10, 22023–22030. [Google Scholar] [CrossRef] [PubMed]
  84. Bruix, A.; Lykhach, Y.; Matolínová, I.; Neitzel, A.; Skála, T.; Tsud, N.; Vorokhta, M.; Stetsovych, V.; Ševčíková, K.; Mysliveček, J.; et al. Maximum noble-metal efficiency in catalytic materials: Atomically dispersed surface platinum. Angew. Chem. Int. Ed. 2014, 53, 10525–10530. [Google Scholar] [CrossRef]
  85. Wang, L.; Liu, X.; Su, K.; Liu, W.; Niu, F.; Li, X.; Yue, H.; Dong, H.; Yang, S.; Yin, Y. Ce-Doped Fe–N–C/Fe3C Nanosheets as an Efficient Oxygen Electrocatalyst under Alkaline and Acidic Media. ACS Appl. Nano Mater. 2024, 7, 22855–22864. [Google Scholar] [CrossRef]
  86. Eissa, A.A.-S.; Kim, N.H.; Lee, J.H. Rational design of a highly mesoporous Fe–N–C/Fe3C/C–S–C nanohybrid with dense active sites for superb electrocatalysis of oxygen reduction. J. Mater. Chem. A 2020, 8, 23436–23454. [Google Scholar] [CrossRef]
  87. Yang, C.C.; Zai, S.F.; Zhou, Y.T.; Du, L.; Jiang, Q. Fe3C-Co nanoparticles encapsulated in a hierarchical structure of N-doped carbon as a multifunctional electrocatalyst for ORR, OER, and HER. Adv. Funct. Mater. 2019, 29, 1901949. [Google Scholar] [CrossRef]
  88. Liu, S.; Li, L.; Yang, T.; Wang, E.; Yu, X.; Hou, Y.; Du, Z.; Cao, S.; Chou, K.-C.; Hou, X. Enhanced overall water splitting by morphology and electronic structure engineering on pristine ultrathin metal-organic frameworks. J. Mater. Sci. Technol. 2025, 220, 92–103. [Google Scholar] [CrossRef]
  89. Yang, L.; Yang, T.; Wang, E.; Yu, X.; Wang, K.; Du, Z.; Cao, S.; Chou, K.-C.; Hou, X. Bifunctional hierarchical NiCoP@ FeNi LDH nanosheet array electrocatalyst for industrial-scale high-current-density water splitting. J. Mater. Sci. Technol. 2023, 159, 33–40. [Google Scholar] [CrossRef]
  90. Wang, B.; Liu, Q.; Yuan, A.; Shi, Q.; Jiang, L.; Yang, W.; Yang, T.; Hou, X. A facile and green strategy for mass production of dispersive FeCo-rich phosphides@ N, P-doped carbon electrocatalysts toward efficient and stable rechargeable Zn-air battery and water splitting. J. Mater. Sci. Technol. 2024, 182, 1–11. [Google Scholar] [CrossRef]
  91. Esconjauregui, S.; Whelan, C.M.; Maex, K. The reasons why metals catalyze the nucleation and growth of carbon nanotubes and other carbon nanomorphologies. Carbon 2009, 47, 659–669. [Google Scholar] [CrossRef]
  92. Qian, Y.; An, T.; Birgersson, K.E.; Liu, Z.; Zhao, D. Web-like interconnected carbon networks from NaCl-assisted pyrolysis of ZIF-8 for highly efficient oxygen reduction catalysis. Small 2018, 14, 1704169. [Google Scholar] [CrossRef] [PubMed]
  93. Huang, X.; Shen, T.; Zhang, T.; Qiu, H.; Gu, X.; Ali, Z.; Hou, Y. Efficient oxygen reduction catalysts of porous carbon nanostructures decorated with transition metal species. Adv. Energy Mater. 2020, 10, 1900375. [Google Scholar] [CrossRef]
  94. Wang, L.; Zhai, J.-J.; Jiang, K.; Wang, J.-Q.; Cai, W.-B. Pd–Cu/C electrocatalysts synthesized by one-pot polyol reduction toward formic acid oxidation: Structural characterization and electrocatalytic performance. Int. J. Hydrogen Energy 2015, 40, 1726–1734. [Google Scholar] [CrossRef]
  95. Bi, Z.; Wang, Y.; Chen, J.; Zhang, X.; Zhou, S.; Wang, X.; Wågberg, T.; Hu, G. Three-dimensional star-like mesoporous nitrogen-doped carbon anchored with highly dispersed Fe and Ce dual-sites for efficient oxygen reduction reaction in Zn-air battery. Colloid Interface Sci. Commun. 2022, 49, 100634. [Google Scholar] [CrossRef]
  96. Kumar, R.; Mooste, M.; Zekker, I.; Käärik, M.; Leis, J.; Kikas, A.; Treshchalov, A.; Kozlova, J.; Otsus, M.; Aruväli, J.; et al. Tailoring highly efficient ZIF-derived FexCey@ N–C composite catalysts for oxygen reduction reaction in H-shape microbial fuel cells. Int. J. Hydrogen Energy 2025, 98, 793–806. [Google Scholar] [CrossRef]
  97. Reichel, F.; Clegg, J.K.; Gloe, K.; Gloe, K.; Weigand, J.J.; Reynolds, J.K.; Li, C.G.; Aldrich-Wright, J.R.; Kepert, C.J.; Lindoy, L.F.; et al. Self-assembly of an imidazolate-bridged FeIII/CuII heterometallic cage. Inorg. Chem. 2014, 53, 688–690. [Google Scholar] [CrossRef]
  98. Sgarbi, R.; Kumar, K.; Jaouen, F.; Zitolo, A.; Ticianelli, E.A.; Maillard, F. Oxygen reduction reaction mechanism and kinetics on MN x C y and M@ NC active sites present in model MNC catalysts under alkaline and acidic conditions. J. Solid State Electrochem. 2021, 25, 45–56. [Google Scholar] [CrossRef]
  99. Ramaswamy, N.; Tylus, U.; Jia, Q.; Mukerjee, S. Activity descriptor identification for oxygen reduction on nonprecious electrocatalysts: Linking surface science to coordination chemistry. J. Am. Chem. Soc. 2013, 135, 15443–15449. [Google Scholar] [CrossRef]
  100. Zhao, Y.; Wang, H.; Li, J.; Fang, Y.; Kang, Y.; Zhao, T.; Zhao, C. Regulating the spin-state of rare-earth Ce single atom catalyst for boosted oxygen reduction in neutral medium. Adv. Funct. Mater. 2023, 33, 2305268. [Google Scholar] [CrossRef]
  101. Yang, B.; Yu, H.; Jia, X.; Cheng, Q.; Ren, Y.; He, B.; Xiang, Z. Atomically dispersed isolated Fe–Ce dual-metal-site catalysts for proton-exchange membrane fuel cells. ACS Appl. Mater. Interfaces 2023, 15, 23316–23327. [Google Scholar] [CrossRef] [PubMed]
  102. Mazzapioda, L.; Moscatelli, G.; Carboni, N.; Brutti, S.; Navarra, M.A. Super Hygroscopic Non-Stoichiometric Cerium Oxide Particles as Electrode Component for PEM Fuel Cells. ChemElectroChem 2023, 10, e202300168. [Google Scholar] [CrossRef]
  103. Li, J.-C.; Maurya, S.; Kim, Y.S.; Li, T.; Wang, L.; Shi, Q.; Liu, D.; Feng, S.; Lin, Y.; Shao, M. Stabilizing single-atom iron electrocatalysts for oxygen reduction via ceria confining and trapping. ACS Catal. 2020, 10, 2452–2458. [Google Scholar] [CrossRef]
  104. Chu, Y.; Luo, E.; Wei, Y.; Zhu, S.; Wang, X.; Yang, L.; Gao, N.; Wang, Y.; Jiang, Z.; Liu, C.; et al. Dual single-atom catalyst design to build robust oxygen reduction electrode via free radical scavenging. Chem. Catal. 2023, 3, 100532. [Google Scholar] [CrossRef]
  105. Luo, H.; Wang, J.; Zhang, S.; Sun, B.; Chen, Z.; Ren, X.; Luo, Z.; Han, X.; Hu, W. In Situ Symbiosis of Cerium Oxide Nanophase for Enhancing the Oxygen Electrocatalysis Performance of Single-Atom Fe─ N─ C Catalyst with Prolonged Stability for Zinc–Air Batteries. Small 2024, 20, 2400357. [Google Scholar] [CrossRef]
  106. Wu, J.; Chen, J.; Li, R.; Chen, H.; Shu, C.; Jin, R.; Zhou, J.; Huang, Y.; Guo, C.; Si, Y.; et al. Boosting the activity of nitrogen-doped carbon catalyst by the guided formation of active sites and enhanced scavenging on reactive oxygen species under the regulation of cerium. J. Alloys Compd. 2024, 1008, 176603. [Google Scholar] [CrossRef]
  107. Zhang, X.; Zhang, J.; Zeng, S.; Wang, H.; Bai, Y.; Hou, H.; Zhang, K.; Li, B.; Zhang, G. Coupling of cerium oxide cyanamide with Fe–N–C for enhanced oxygen reduction reaction. New J. Chem. 2023, 47, 6058–6065. [Google Scholar] [CrossRef]
  108. Xia, W.; Li, J.; Wang, T.; Song, L.; Guo, H.; Gong, H.; Jiang, C.; Gao, B.; He, J. The synergistic effect of Ceria and Co in N-doped leaf-like carbon nanosheets derived from a 2D MOF and their enhanced performance in the oxygen reduction reaction. Chem. Commun. 2018, 54, 1623–1626. [Google Scholar] [CrossRef]
  109. Yang, C.; Yu, X.; Heißler, S.; Nefedov, A.; Colussi, S.; Llorca, J.; Trovarelli, A.; Wang, Y.; Wöll, C. Surface faceting and reconstruction of ceria nanoparticles. Angew. Chem. Int. Ed. 2017, 56, 375–379. [Google Scholar] [CrossRef]
  110. Cui, C.; Du, G.; Zhang, K.; An, T.; Li, B.; Liu, X.; Liu, Z. Co3O4 nanoparticles anchored in MnO2 nanorods as efficient oxygen reduction reaction catalyst for metal-air batteries. J. Alloys Compd. 2020, 814, 152239. [Google Scholar] [CrossRef]
  111. Hu, T.; Wang, Y.; Zhang, L.; Tang, T.; Xiao, H.; Chen, W.; Zhao, M.; Jia, J.; Zhu, H. Facile synthesis of PdO-doped Co3O4 nanoparticles as an efficient bifunctional oxygen electrocatalyst. Appl. Catal. B Environ. 2019, 243, 175–182. [Google Scholar] [CrossRef]
  112. Li, X.; You, S.; Du, J.; Dai, Y.; Chen, H.; Cai, Z.; Ren, N.; Zou, J. ZIF-67-derived Co3O4@carbon protected by oxygen-buffering CeO2 as an efficient catalyst for boosting oxygen reduction/evolution reactions. J. Mater. Chem. A 2019, 7, 25853–25864. [Google Scholar] [CrossRef]
  113. Wang, H.; Maiyalagan, T.; Wang, X. Review on recent progress in nitrogen-doped graphene: Synthesis, characterization, and its potential applications. ACS Catal. 2012, 2, 781–794. [Google Scholar] [CrossRef]
  114. Lai, L.; Potts, J.R.; Zhan, D.; Wang, L.; Poh, C.K.; Tang, C.; Gong, H.; Shen, Z.; Lin, J.; Ruoff, R.S. Exploration of the active center structure of nitrogen-doped graphene-based catalysts for oxygen reduction reaction. Energy Environ. Sci. 2012, 5, 7936–7942. [Google Scholar] [CrossRef]
  115. Chen, Y.; Zhao, S.; Liu, Z. Influence of the synergistic effect between Co–N–C and ceria on the catalytic performance for selective oxidation of ethylbenzene. Phys. Chem. Chem. Phys. 2015, 17, 14012–14020. [Google Scholar] [CrossRef]
  116. Sun, X.; Xu, T.; Sun, W.; Bai, J.; Li, C. Ce-doped ZIF-67 derived Co3O4 nanoparticles supported by carbon nanofibers: A synergistic strategy towards bifunctional oxygen electrocatalysis and Zn-Air batteries. J. Alloys Compd. 2022, 898, 162778. [Google Scholar] [CrossRef]
  117. Zhong, H.; Estudillo-Wong, L.A.; Gao, Y.; Feng, Y.; Alonso-Vante, N. Oxygen vacancies engineering by coordinating oxygen-buffering CeO2 with CoOx nanorods as efficient bifunctional oxygen electrode electrocatalyst. J. Energy Chem. 2021, 59, 615–625. [Google Scholar] [CrossRef]
  118. Jiang, Y.; Deng, Y.-P.; Fu, J.; Lee, D.U.; Liang, R.; Cano, Z.P.; Liu, Y.; Bai, Z.; Hwang, S.; Yang, L.; et al. Interpenetrating triphase cobalt-based nanocomposites as efficient bifunctional oxygen electrocatalysts for long-lasting rechargeable Zn–air batteries. Adv. Energy Mater. 2018, 8, 1702900. [Google Scholar] [CrossRef]
  119. Hutchings, G.S.; Zhang, Y.; Li, J.; Yonemoto, B.T.; Zhou, X.; Zhu, K.; Jiao, F. In situ formation of cobalt oxide nanocubanes as efficient oxygen evolution catalysts. J. Am. Chem. Soc. 2015, 137, 4223–4229. [Google Scholar] [CrossRef]
  120. Wang, Y.; Wang, J.; Mohamed, Z.; Huang, Q.; Chen, T.; Hou, Y.; Dang, F.; Zhang, W.; Wang, H. A free-standing CeO2/Co3O4 nanowires electrode featuring a controllable discharge/charge product evolution route with enhanced catalytic performance for Li-O2 batteries. Appl. Mater. Today 2020, 19, 100603. [Google Scholar] [CrossRef]
  121. Guo, S.; Wang, J.; Sun, Y.; Peng, L.; Li, C. Interface engineering of Co3O4/CeO2 heterostructure in-situ embedded in Co/N doped carbon nanofibers integrating oxygen vacancies as effective oxygen cathode catalyst for Li-O2 battery. Chem. Eng. J. 2023, 452, 139317. [Google Scholar] [CrossRef]
  122. Li, J.; Shu, C.; Liu, C.; Chen, X.; Hu, A.; Long, J. Rationalizing the effect of oxygen vacancy on oxygen electrocatalysis in Li–O2 battery. Small 2020, 16, 2001812. [Google Scholar] [CrossRef] [PubMed]
  123. Yin, C.-S.; Leng, Y.; Yang, X.; Min, C.-G.; Ren, A.-M. Unraveling the mechanism of pyrrole and N-defect regulating CoN4 single atom catalysts as a pH-universal bifunctional electrocatalyst for OER and ORR. Appl. Surf. Sci. 2024, 643, 158605. [Google Scholar] [CrossRef]
  124. Peng, L.; Peng, X.; Zhu, Z.; Xu, Q.; Luo, K.; Ni, Z.; Yuan, D. Efficient MnO and Co nanoparticles coated with N-doped carbon as a bifunctional electrocatalyst for rechargeable Zn-air batteries. Int. J. Hydrogen Energy 2023, 48, 19126–19136. [Google Scholar] [CrossRef]
  125. Li, B.-Q.; Zhao, C.-X.; Chen, S.; Liu, J.-N.; Chen, X.; Song, L.; Zhang, Q. Framework-porphyrin-derived single-atom bifunctional oxygen electrocatalysts and their applications in Zn–air batteries. Adv. Mater. 2019, 31, 1900592. [Google Scholar] [CrossRef]
  126. Shen, S.; Sun, Y.; Sun, H.; Pang, Y.; Xia, S.; Chen, T.; Zheng, S.; Yuan, T. Research progress in ZIF-8 derived single atomic catalysts for oxygen reduction reaction. Catalysts 2022, 12, 525. [Google Scholar] [CrossRef]
  127. Chen, Y.; Ji, S.; Zhao, S.; Chen, W.; Dong, J.; Cheong, W.-C.; Shen, R.; Wen, X.; Zheng, L.; Rykov, A.I.; et al. Enhanced oxygen reduction with single-atomic-site iron catalysts for a zinc-air battery and hydrogen-air fuel cell. Nat. Commun. 2018, 9, 5422. [Google Scholar] [CrossRef]
  128. Wu, M.; Liu, Z.; Wang, G.; Liu, H.; Ma, J. Electrostatic spinning strategy to prepare cage-like PAN-fiber network-wrapped Co–N–C structures for the oxygen reduction reaction. ACS Appl. Energy Mater. 2022, 5, 14155–14163. [Google Scholar] [CrossRef]
  129. Lee, K.; Zhang, L.; Lui, H.; Hui, R.; Shi, Z.; Zhang, J. Oxygen reduction reaction (ORR) catalyzed by carbon-supported cobalt polypyrrole (Co-PPy/C) electrocatalysts. Electrochim. Acta 2009, 54, 4704–4711. [Google Scholar] [CrossRef]
  130. Tang, S.; Li, X.; Li, H.; Qi, H.; Wang, J. Design and assembly of ZIF@ PPy-derived hierarchical structure CeCo-NxCC electrocatalyst towards oxygen reduction reaction. Int. J. Hydrogen Energy 2024, 68, 170–180. [Google Scholar] [CrossRef]
  131. Wu, K.; Wang, D.; Fu, Q.; Xu, T.; Xiong, Q.; Peera, S.G.; Liu, C. Co/Ce-MOF-derived oxygen electrode bifunctional catalyst for rechargeable zinc–air batteries. Inorg. Chem. 2024, 63, 11135–11145. [Google Scholar] [CrossRef] [PubMed]
  132. Farkaš, B.; de Leeuw, N.H. Towards a morphology of cobalt nanoparticles: Size and strain effects. Nanotechnology 2020, 31, 195711. [Google Scholar] [CrossRef] [PubMed]
  133. Podloucky, R.; Glötzel, D. Band structure, cohesive properties, and Compton profile of γ- andα-cerium. Phys. Rev. B 1983, 27, 3390–3405. [Google Scholar] [CrossRef]
  134. Feng, X.; Jiao, Q.; Liu, T.; Li, Q.; Yin, M.; Zhao, Y.; Li, H.; Feng, C.; Zhou, W. Facile synthesis of Co9S8 hollow spheres as a high-performance electrocatalyst for the oxygen evolution reaction. ACS Sustain. Chem. Eng. 2018, 6, 1863–1871. [Google Scholar] [CrossRef]
  135. Long, J.; Gong, Y.; Lin, J. Metal–organic framework–derived Co9S8@ CoS@ CoO@ C nanoparticles as efficient electro-and photo-catalysts for the oxygen evolution reaction. J. Mater. Chem. A 2017, 5, 10495–10509. [Google Scholar] [CrossRef]
  136. Arunchander, A.; Peera, S.G.; Giridhar, V.V.; Sahu, A.K. Synthesis of cobalt sulfide-graphene as an efficient oxygen reduction catalyst in alkaline medium and its application in anion exchange membrane fuel cells. J. Electrochem. Soc. 2016, 164, F71. [Google Scholar] [CrossRef]
  137. Qian, J.; Wang, T.; Zhang, Z.; Liu, Y.; Li, J.; Gao, D. Engineered spin state in Ce doped LaCoO3 with enhanced electrocatalytic activity for rechargeable Zn-Air batteries. Nano Energy 2020, 74, 104948. [Google Scholar] [CrossRef]
  138. Yinqing, S.; Guan, Y.; Wu, X.; Li, W.; Li, Y.; Sun, L.; Mi, H.; Zhang, Q.; He, C.; Ren, X. ZIF-derived “senbei”-like Co9S8/CeO2/Co heterostructural nitrogen-doped carbon nanosheets as bifunctional oxygen electrocatalysts for Zn-air batteries. Nanoscale 2021, 13, 3227–3236. [Google Scholar]
  139. Cao, C.; Xie, J.; Zhang, S.; Pan, B.; Cao, G.; Zhao, X. Graphene-like δ-MnO2 decorated with ultrafine CeO2 as a highly efficient catalyst for long-life lithium–oxygen batteries. J. Mater. Chem. A 2017, 5, 6747–6755. [Google Scholar] [CrossRef]
  140. Hou, Y.; Wang, J.; Hou, C.; Fan, Y.; Zhai, Y.; Li, H.; Dang, F.; Chou, S. Oxygen vacancies promoting the electrocatalytic performance of CeO2 nanorods as cathode materials for Li–O2 batteries. J. Mater. Chem. A 2019, 7, 6552–6561. [Google Scholar] [CrossRef]
  141. Li, J.; Zhou, H.; Zhuo, H.; Wei, Z.; Zhuang, G.; Zhong, X.; Deng, S.; Li, X.; Wang, J. Oxygen vacancies on TiO2 promoted the activity and stability of supported Pd nanoparticles for the oxygen reduction reaction. J. Mater. Chem. A 2018, 6, 2264–2272. [Google Scholar] [CrossRef]
  142. Yu, Y.; Gao, L.; Liu, X.; Wang, Y.; Xing, S. Enhancing the Catalytic Activity of Zeolitic Imidazolate Framework-8-Derived N-Doped Carbon with Incorporated CeO2 Nanoparticles in the Oxygen Reduction Reaction. Chem.-A Eur. J. 2017, 23, 10690–10697. [Google Scholar] [CrossRef] [PubMed]
  143. Wen, X.; Chang, Y.; Jia, J. Evaluating the growth of ceria-modified N-doped carbon-based materials and their performance in the oxygen reduction reaction. Nanomaterials 2022, 12, 3057. [Google Scholar] [CrossRef]
  144. Ghosh, D.; Parwaiz, S.; Mohanty, P.; Pradhan, D. Tuning the morphology of CeO2 nanostructures using a template-free solvothermal process and their oxygen reduction reaction activity. Dalton Trans. 2020, 49, 17594–17604. [Google Scholar] [CrossRef]
  145. Jiang, L.; Yao, M.; Liu, B.; Li, Q.; Liu, R.; Lv, H.; Lu, S.; Gong, C.; Zou, B.; Cui, T.; et al. Controlled synthesis of CeO2/graphene nanocomposites with highly enhanced optical and catalytic properties. J. Phys. Chem. C 2012, 116, 11741–11745. [Google Scholar] [CrossRef]
  146. Milikić, J.; Fuentes, R.O.; Tasca, J.E.; Santos, D.M.F.; Šljukić, B.; Figueiredo, F.M.L. Nickel-doped ceria bifunctional electrocatalysts for oxygen reduction and evolution in alkaline media. Batteries 2022, 8, 100. [Google Scholar] [CrossRef]
  147. Liu, K.; Fu, J.; Lin, Y.; Luo, T.; Ni, G.; Li, H.; Lin, Z.; Liu, M. Insights into the activity of single-atom Fe-NC catalysts for oxygen reduction reaction. Nat. Commun. 2022, 13, 2075. [Google Scholar] [CrossRef]
  148. Xing, G.; Tong, M.; Yu, P.; Wang, L.; Zhang, G.; Tian, C.; Fu, H. Reconstruction of highly dense Cu− N4 active sites in electrocatalytic oxygen reduction characterized by operando synchrotron radiation. Angew. Chem. Int. Ed. 2022, 61, e202211098. [Google Scholar] [CrossRef]
  149. Peng, L.; Yang, J.; Yang, Y.; Qian, F.; Wang, Q.; Sun-Waterhouse, D.; Shang, L.; Zhang, T.; Waterhouse, G.I.N. Mesopore-rich Fe–N–C catalyst with FeN4–O–NC single-atom sites delivers remarkable oxygen reduction reaction performance in alkaline media. Adv. Mater. 2022, 34, 2202544. [Google Scholar] [CrossRef]
  150. Qin, Y.; Guo, C.; Ou, Z.; Xu, C.; Lan, Q.; Jin, R.; Liu, Y.; Niu, Y.; Xu, Q.; Si, Y.; et al. Regulating single-atom Mn sites by precisely axial pyridinic-nitrogen coordination to stabilize the oxygen reduction. J. Energy Chem. 2023, 80, 542–552. [Google Scholar] [CrossRef]
  151. Xu, C.; Guo, C.; Liu, J.; Hu, B.; Dai, J.; Wang, M.; Jin, R.; Luo, Z.; Li, H.; Chen, C. Accelerating the oxygen adsorption kinetics to regulate the oxygen reduction catalysis via Fe3C nanoparticles coupled with single Fe-N4 sites. Energy Storage Mater. 2022, 51, 149–158. [Google Scholar] [CrossRef]
  152. Yang, L.; Zhang, X.; Yu, L.; Hou, J.; Zhou, Z.; Lv, R. Atomic Fe–N4/C in flexible carbon fiber membrane as binder-free air cathode for Zn–air batteries with stable cycling over 1000 h. Adv. Mater. 2022, 34, 2105410. [Google Scholar] [CrossRef] [PubMed]
  153. Wang, Q.; Feng, Q.; Lei, Y.; Tang, S.; Xu, L.; Xiong, Y.; Fang, G.; Wang, Y.; Yang, P.; Liu, J.; et al. Quasi-solid-state Zn-air batteries with an atomically dispersed cobalt electrocatalyst and organohydrogel electrolyte. Nat. Commun. 2022, 13, 3689. [Google Scholar] [CrossRef]
  154. Zhao, S.-N.; Li, J.-K.; Wang, R.; Cai, J.; Zang, S.-Q. Electronically and geometrically modified single-atom Fe sites by adjacent Fe nanoparticles for enhanced oxygen reduction. Adv. Mater. 2022, 34, 2107291. [Google Scholar] [CrossRef]
  155. Cui, X.; Gao, L.; Lei, S.; Liang, S.; Zhang, J.; Sewell, C.D.; Xue, W.; Liu, Q.; Lin, Z.; Yang, Y. Simultaneously crafting single-atomic Fe sites and graphitic layer-wrapped Fe3C nanoparticles encapsulated within mesoporous carbon tubes for oxygen reduction. Adv. Funct. Mater. 2021, 31, 2009197. [Google Scholar] [CrossRef]
  156. Li, G.; Zheng, W.; Li, X.; Luo, S.; Xing, D.; Ming, P.; Li, B.; Zhang, C. Application of the Ce-based radical scavengers in proton exchange membrane fuel cells. Int. J. Hydrogen Energy 2024, 74, 17–30. [Google Scholar] [CrossRef]
  157. Yin, L.; Zhang, S.; Sun, M.; Wang, S.; Huang, B.; Du, Y. Heteroatom-driven coordination fields altering single cerium atom sites for efficient oxygen reduction reaction. Adv. Mater. 2023, 35, 2302485. [Google Scholar] [CrossRef]
  158. Liu, J.; Jin, Y.; Jin, R.; Liu, Y.; Ma, Z.; Guo, C.; Lei, Y.; Sun, L.; Chen, H.; Si, Y.; et al. In situ Regulating the phase structure of Ce-based catalytic sites to boost the performance of zinc-air batteries. Nano Energy 2024, 129, 110030. [Google Scholar] [CrossRef]
  159. Zhang, C.; Huang, N.; Zhai, Z.; Liu, L.; Chen, B.; Yang, B.; Jiang, X.; Yang, N. Bifunctional Oxygen Electrocatalyst of Co4N and Nitrogen-Doped Carbon Nanowalls/Diamond for High-Performance Flexible Zinc–Air Batteries. Adv. Energy Mater. 2023, 13, 2301749. [Google Scholar] [CrossRef]
  160. Peera, S.G.; Liu, C.; Reddy, P.S.P.; Das, S.K.; Sahu, A.K.; Siddiqui, M.R.; Sarkar, I.J.R.; Kumar, S.; Bhatnagar, A. Enhanced oxygen reduction reaction with rare-earth metal-based Ce-NC catalyst. Int. J. Hydrogen Energy 2025, 107, 83–95. [Google Scholar] [CrossRef]
  161. Xu, K.; Pei, S.; Zhang, W.; Han, Z.; Guan, P.; Wang, L.; Zou, Y.; Ding, H.; Ma, X.; Xu, C.; et al. Ce (III)-terephthalic acid metal-organic frameworks as highly efficient· OH radical scavengers for fuel cells and investigation of its antioxidation mechanism. Mater. Today Energy 2023, 31, 101195. [Google Scholar] [CrossRef]
  162. Zhu, M.; Zhao, C.; Liu, X.; Wang, X.; Zhou, F.; Wang, J.; Hu, Y.; Zhao, Y.; Yao, T.; Yang, L.-M.; et al. Single atomic cerium sites with a high coordination number for efficient oxygen reduction in proton-exchange membrane fuel cells. ACS Catal. 2021, 11, 3923–3929. [Google Scholar] [CrossRef]
  163. Yang, G.; Zhu, J.; Yuan, P.; Hu, Y.; Qu, G.; Lu, B.-A.; Xue, X.; Yin, H.; Cheng, W.; Cheng, J.; et al. Regulating Fe-spin state by atomically dispersed Mn-N in Fe-NC catalysts with high oxygen reduction activity. Nat. Commun. 2021, 12, 1734. [Google Scholar] [CrossRef]
  164. Sun, Y.; Sun, S.; Yang, H.; Xi, S.; Gracia, J.; Xu, Z.J. Spin-related electron transfer and orbital interactions in oxygen electrocatalysis. Adv. Mater. 2020, 32, 2003297. [Google Scholar] [CrossRef]
  165. Li, Z.; Zhuang, Z.; Lv, F.; Zhu, H.; Zhou, L.; Luo, M.; Zhu, J.; Lang, Z.; Feng, S.; Chen, W.; et al. The marriage of the FeN4 moiety and MXene boosts oxygen reduction catalysis: Fe 3d electron delocalization matters. Adv. Mater. 2018, 30, 1803220. [Google Scholar] [CrossRef]
  166. Wang, X.; Li, M.; Wang, P.; Sun, D.; Ding, L.; Li, H.; Tang, Y.; Fu, G. Spin-selective coupling in Mott–Schottky Er2O3-Co boosts electrocatalytic oxygen reduction. Small Methods 2023, 7, 2300100. [Google Scholar] [CrossRef]
  167. Liu, J.; Guo, C.; Sun, L.; Liu, Y.; Chen, H.; Shu, C.; Dai, J.; Xu, C.; Jin, R.; Li, H.; et al. Unraveling the Electron Transfer Effect of Single-Metal Ce-N4 Sites via Mesopore-Coupling for Boosted Oxygen Reduction Activity. Small 2024, 20, 2305615. [Google Scholar] [CrossRef]
  168. Yu, Y.; Peng, X.; Ali, U.; Liu, X.; Xing, Y.; Xing, S. Facile route to achieve bifunctional electrocatalysts for oxygen reduction and evolution reactions derived from CeO2 encapsulated by the zeolitic imidazolate framework-67. Inorg. Chem. Front. 2019, 6, 3255–3263. [Google Scholar] [CrossRef]
  169. Huang, Y.; Liu, Y.; Deng, Y.; Zhang, J.; He, B.; Sun, J.; Yang, Z.; Zhou, W.; Zhao, L. Enhancing the bifunctional activity of CoSe2 nanocubes by surface decoration of CeO2 for advanced zinc-air batteries. J. Colloid Interface Sci. 2022, 625, 839–849. [Google Scholar] [CrossRef]
  170. Sun, S.; Xue, Y.; Wang, Q.; Huang, H.; Miao, H.; Liu, Z. Cerium ion intercalated MnO2 nanospheres with high catalytic activity toward oxygen reduction reaction for aluminum-air batteries. Electrochim. Acta 2018, 263, 544–554. [Google Scholar] [CrossRef]
  171. Sun, L.; Zhou, L.; Yang, C.; Yuan, Y. CeO2 nanoparticle-decorated reduced graphene oxide as an efficient bifunctional electrocatalyst for oxygen reduction and evolution reactions. Int. J. Hydrogen Energy 2017, 42, 15140–15148. [Google Scholar] [CrossRef]
  172. Ganguly, D.; Jayakumar, N.; Gopalan, N.K. Unveiling a synthetic strategy for a cerium-based hetero-phase hybrid electrocatalyst for near-zero peroxide yield in the ORR. Sustain. Energy Fuels 2025, 9, 778–786. [Google Scholar] [CrossRef]
Figure 1. The unique properties of the Ce/CeO2-based electrocatalysts for ORR.
Figure 1. The unique properties of the Ce/CeO2-based electrocatalysts for ORR.
Nanomaterials 15 00600 g001
Figure 2. Advantages of CeO2 and Ce-N4-C for ORR.
Figure 2. Advantages of CeO2 and Ce-N4-C for ORR.
Nanomaterials 15 00600 g002
Figure 3. TEM images of SMOF-3 (a,b) and LSVs curves (c,d) of various SMOF catalysts at 900 rpm (Reproduced with permissions from Ref. [60]). (e) The schematic representation of Pt/CeO2-NC catalyst synthesis. Polarization curves (f), durability (g), and polarization curves recorded at specific time intervals (h), power density retention curves of Pt/CeO2-NC and Pt/C catalysts (i). (j) Diagram representing the role of CeO2 (Reproduced with permissions from Ref. [62]).
Figure 3. TEM images of SMOF-3 (a,b) and LSVs curves (c,d) of various SMOF catalysts at 900 rpm (Reproduced with permissions from Ref. [60]). (e) The schematic representation of Pt/CeO2-NC catalyst synthesis. Polarization curves (f), durability (g), and polarization curves recorded at specific time intervals (h), power density retention curves of Pt/CeO2-NC and Pt/C catalysts (i). (j) Diagram representing the role of CeO2 (Reproduced with permissions from Ref. [62]).
Nanomaterials 15 00600 g003
Figure 4. Schematic representation (a). HR-TEM images (b). CV curves recorded in N2 atmosphere in 0.1 M HClO4 electrolyte (c). LSV curves recorded in O2 atmosphere in 0.1 M HClO4 electrolyte at 1600 rom (d,e) and Tafel curves of Pt-CeO2-C catalysts at 0.9 V (Reproduced with permissions from Ref. [73]).
Figure 4. Schematic representation (a). HR-TEM images (b). CV curves recorded in N2 atmosphere in 0.1 M HClO4 electrolyte (c). LSV curves recorded in O2 atmosphere in 0.1 M HClO4 electrolyte at 1600 rom (d,e) and Tafel curves of Pt-CeO2-C catalysts at 0.9 V (Reproduced with permissions from Ref. [73]).
Nanomaterials 15 00600 g004
Figure 6. (a) Schematic representation of the synthesis process of the FeCeNC catalyst with CTAB as surfactant. (bg) HR TEM images. (hk) Elemental mapping for C, N, Fe, and Ce atoms in FeCeNC catalyst. (l) LSV curves. (m) Relationship between E1/2 and Jk. (n) Tafel plots. (o) LSVs recorded at different rpms. (p) K–L plots. (q) Stability curves of FeCeNC catalysts (Reproduced with permissions from Ref. [95] open access).
Figure 6. (a) Schematic representation of the synthesis process of the FeCeNC catalyst with CTAB as surfactant. (bg) HR TEM images. (hk) Elemental mapping for C, N, Fe, and Ce atoms in FeCeNC catalyst. (l) LSV curves. (m) Relationship between E1/2 and Jk. (n) Tafel plots. (o) LSVs recorded at different rpms. (p) K–L plots. (q) Stability curves of FeCeNC catalysts (Reproduced with permissions from Ref. [95] open access).
Nanomaterials 15 00600 g006
Figure 7. (a) Schematic of FeCe-SAD/HPNC synthesis. (b) Charge density. (c) Free energy diagram. (d) PDOS. (e) ORR pathway of FeCe-SAD/HPNC catalysts. (f) Typical representation of PEM cathode. (g) PEM polarization curves. (h) Impedance curves of various compositions of FeCe-SAD/HPNC catalysts. (i) PEM polarization curves of Fe-SAD/HPNC, Ce-SAD/HPNC, and FeCe-SAD/HPNC cathodes. (j) Stability curves at constant current and (k) various current density of FeCe-SAD/HPNC catalyst (Reproduced with permissions from Ref. [101]).
Figure 7. (a) Schematic of FeCe-SAD/HPNC synthesis. (b) Charge density. (c) Free energy diagram. (d) PDOS. (e) ORR pathway of FeCe-SAD/HPNC catalysts. (f) Typical representation of PEM cathode. (g) PEM polarization curves. (h) Impedance curves of various compositions of FeCe-SAD/HPNC catalysts. (i) PEM polarization curves of Fe-SAD/HPNC, Ce-SAD/HPNC, and FeCe-SAD/HPNC cathodes. (j) Stability curves at constant current and (k) various current density of FeCe-SAD/HPNC catalyst (Reproduced with permissions from Ref. [101]).
Nanomaterials 15 00600 g007
Figure 8. (a) LSV curves of FeCe-N-C catalysts in O2 saturated 0.1 M HClO4. (b) Current density and TOF (measured at 0.8 V vs. RHE) of FeCe-N-C, and other recently reported M-N-C catalysts. (c) LSV curves of various FeCe-N-C catalysts before and after accelerated degradation test (ADT) for 30,000 with theoretical calculations of the superior performance of FeCe-N-C catalysts. (d) Optimized atomic structures for the main process of Fenton reactions with different sites. (e) Gibbs free energy profiles for ORR at U = 1.23 V. (f) Relationship between the ORR overpotential, the binding energy of O*, and the partial charges of Fe, calculated by Bader charges. The cyan and yellow sections illustrate electron consumption and cumulation, respectively. (g) The mechanism schemes on FeN4 sites of FeCe-N-C at U = 0 V. (h) Radical scavenging mechanism Ref. [104]. (i) Schematic illustration for construction of CeO2@Fe–NC based on ZIF-8. (j) ORR curves at 1600 rpm in 0.1 m KOH. (k) Activity comparison of CeO2@Fe-NC with many representative catalysts and Raman characterizations of the CeO2@Fe-NC and Fe-NC catalysts. (l,m) In situ Raman spectroscopy of CeO2@Fe-NC in O2-saturated 0.1 m KOH and in situ Raman characterization of the Fe-NC and FeCe-NC catalyst in O2-saturated 0.1 M KOH Ref. [105].
Figure 8. (a) LSV curves of FeCe-N-C catalysts in O2 saturated 0.1 M HClO4. (b) Current density and TOF (measured at 0.8 V vs. RHE) of FeCe-N-C, and other recently reported M-N-C catalysts. (c) LSV curves of various FeCe-N-C catalysts before and after accelerated degradation test (ADT) for 30,000 with theoretical calculations of the superior performance of FeCe-N-C catalysts. (d) Optimized atomic structures for the main process of Fenton reactions with different sites. (e) Gibbs free energy profiles for ORR at U = 1.23 V. (f) Relationship between the ORR overpotential, the binding energy of O*, and the partial charges of Fe, calculated by Bader charges. The cyan and yellow sections illustrate electron consumption and cumulation, respectively. (g) The mechanism schemes on FeN4 sites of FeCe-N-C at U = 0 V. (h) Radical scavenging mechanism Ref. [104]. (i) Schematic illustration for construction of CeO2@Fe–NC based on ZIF-8. (j) ORR curves at 1600 rpm in 0.1 m KOH. (k) Activity comparison of CeO2@Fe-NC with many representative catalysts and Raman characterizations of the CeO2@Fe-NC and Fe-NC catalysts. (l,m) In situ Raman spectroscopy of CeO2@Fe-NC in O2-saturated 0.1 m KOH and in situ Raman characterization of the Fe-NC and FeCe-NC catalyst in O2-saturated 0.1 M KOH Ref. [105].
Nanomaterials 15 00600 g008
Figure 11. Free energy images of (a) Co NPs and (c) Co/Ce@NC and the reaction intermediates of (b) Co (111) and (d) Co@NC. (e) CV. (f) LSV of (0.1 M KOH electrolyte). (g) K-L plots. (h) i-t curves. (i) Methanol sensitivity. (j) Histograms showing the relationship between ORR E1/2 and OER potentials at 10 mA cm−2 (Reproduced with permissions from Ref. [131]).
Figure 11. Free energy images of (a) Co NPs and (c) Co/Ce@NC and the reaction intermediates of (b) Co (111) and (d) Co@NC. (e) CV. (f) LSV of (0.1 M KOH electrolyte). (g) K-L plots. (h) i-t curves. (i) Methanol sensitivity. (j) Histograms showing the relationship between ORR E1/2 and OER potentials at 10 mA cm−2 (Reproduced with permissions from Ref. [131]).
Nanomaterials 15 00600 g011aNanomaterials 15 00600 g011b
Figure 12. (a) LSV curves. (b) Comparison of half-wave potentials, Jk of the various CeN4O2 catalysts, and charge density around different N and Ce atoms of (c) planar and (d) curved CeN4O2. PDOS of (d) planar and (e) curved CeN4O2. (f) Free energy diagram of planar and curved CeN4O2 (Reproduced with permissions from Ref. [158]). (g) XRD pattens of Ce-ZIFs with different concentration of Ce. (h) Zoomed version of the Ce-ZIFs. (i) LSV curves of various Ce-N-C catalysts. (j) Ce-N-C, Pt/C catalysts. (k) Chronoamperometric curves of Pt/C and Ce-N-C-3 catalysts in response to the addition of H2O2 (Reproduced with permissions from Ref. [160]).
Figure 12. (a) LSV curves. (b) Comparison of half-wave potentials, Jk of the various CeN4O2 catalysts, and charge density around different N and Ce atoms of (c) planar and (d) curved CeN4O2. PDOS of (d) planar and (e) curved CeN4O2. (f) Free energy diagram of planar and curved CeN4O2 (Reproduced with permissions from Ref. [158]). (g) XRD pattens of Ce-ZIFs with different concentration of Ce. (h) Zoomed version of the Ce-ZIFs. (i) LSV curves of various Ce-N-C catalysts. (j) Ce-N-C, Pt/C catalysts. (k) Chronoamperometric curves of Pt/C and Ce-N-C-3 catalysts in response to the addition of H2O2 (Reproduced with permissions from Ref. [160]).
Nanomaterials 15 00600 g012
Figure 13. (a) Schematic synthesis of Ce SAC/HPNC (bd) SEM and TEM images of Ce SAC/HPNC catalysts. (e) LSV curves of various Ce SAC/HPNCs. (f) Half-wave potential and Jk of various Ce SAC/HPNC catalysts. (g,h) Fuel cell characterization of the various Ce SAC/HPNC catalysts (Reproduced with permissions from Ref. [162]). (i) Typical representation of the synthesis of CeNC-40 catalyst. (jl) TEM images of the CeNC-40 catalyst. (m,n) LSV curves of the CeNcs-40 catalysts in 0.1 M KOH and 0.1 M HClO4. (o) Free energy of the ORR on CeNCs-40. (p) Magnetic susceptibility and (q) effective magnetic moment of CeNC-40 catalyst. (r) Schematic representation of the electronic configuration of the high spin-state cerium atom. (s) Illustration of the catalytic mechanism of the ORR (Reproduced with permissions from Ref. [100]).
Figure 13. (a) Schematic synthesis of Ce SAC/HPNC (bd) SEM and TEM images of Ce SAC/HPNC catalysts. (e) LSV curves of various Ce SAC/HPNCs. (f) Half-wave potential and Jk of various Ce SAC/HPNC catalysts. (g,h) Fuel cell characterization of the various Ce SAC/HPNC catalysts (Reproduced with permissions from Ref. [162]). (i) Typical representation of the synthesis of CeNC-40 catalyst. (jl) TEM images of the CeNC-40 catalyst. (m,n) LSV curves of the CeNcs-40 catalysts in 0.1 M KOH and 0.1 M HClO4. (o) Free energy of the ORR on CeNCs-40. (p) Magnetic susceptibility and (q) effective magnetic moment of CeNC-40 catalyst. (r) Schematic representation of the electronic configuration of the high spin-state cerium atom. (s) Illustration of the catalytic mechanism of the ORR (Reproduced with permissions from Ref. [100]).
Nanomaterials 15 00600 g013
Figure 14. (a) Schematic of the synthesis procedure of Ce SAs/PSNC ORR activity of different catalysts. (b) Polarization curves in O2-saturated 0.1 m KOH solution. (c) Comparison of the Jk and E1/2 values between Ce SAs/PSNC and the catalysts reported recently. (d) LSV of Ce SAs/PSNC before and after 5000 cycles. (e) Adsorption energy of O2 and H2O. (f) The reaction energy trends of ORR formation for Ce SAs/PSNC and Ce SAs/NC (Reproduced with permission from Ref. [157]).
Figure 14. (a) Schematic of the synthesis procedure of Ce SAs/PSNC ORR activity of different catalysts. (b) Polarization curves in O2-saturated 0.1 m KOH solution. (c) Comparison of the Jk and E1/2 values between Ce SAs/PSNC and the catalysts reported recently. (d) LSV of Ce SAs/PSNC before and after 5000 cycles. (e) Adsorption energy of O2 and H2O. (f) The reaction energy trends of ORR formation for Ce SAs/PSNC and Ce SAs/NC (Reproduced with permission from Ref. [157]).
Nanomaterials 15 00600 g014
Table 1. ORR half-cell and fuel cell data of Ce-based catalysts derived from metal–organic framework precursors.
Table 1. ORR half-cell and fuel cell data of Ce-based catalysts derived from metal–organic framework precursors.
CatalystORR Onset Potential (V vs. RHE)
a HClO4
b KOH
ORR Half-Wave Potential (V vs. RHE)
a HClO4
b KOH
a Number of Electrons (n)
(from K–L Plots)
b Tafel Slope (mV/dec)
Number
(RRDE) of Electrons (n)
and % of a H2O2
b H2O-
Potential Loss E1/2 (mV)
Number Cycles
b Loss in E1/2
Power Density
a PEM
b AEM
c Zn–Air Battery
mW/cm2
Ref.
CeGSb 0.92b 0.81a 4
b 111
n = 3.9
b 5
NRNR29
Ce-N-C-3NRa 0.68a 3.91
b 97
NRa 5000
b 70
NR31
Pt/CeO2-NCNRa 0.922NRa n = 3.94
a 0.5%
a 10,000
b 10
1.08 W cm−262
0.5La-CeNC-FeNRb 0.870a 3.94
b 86.8
a n = 3.90−3.96
b 2.26−5.55%
a 1000
b 7
NR83
Ce/Fe-NC/Fe3C-Pa 0.94
b 0.98
b 0.87
b 0.78
b 167n = 3.96–3.99
a 6%,
b 3%
a 10,000
b 14
a 347
c 184
85
FeCeNCNRb 0.855a 3.7
b 55
a NRa 1000
b 2
c 169.2105
Fe1Ce1@N–Cb 0.981b 0.867a 4
b 60
a n = 3.87
b 7.1
a 1000
b 40
NR106
CeNCa 1.05
b 1.05
a 0.90
b 0.80
a ~4
b 80
n = 4.0
a 0.93–
1.67%
b 0.41–0.70%
a 10,000
b 8
b 165100
FeCe-SAD/HPNCa 0.91a 0.81b 69a n = 4
b <1.2
NRa H2/O2: 771 H2/air: 498101
Ce/Fe-NCNWNRb 0.915a 68b <2.5a 5000
b 19
b 496103
Ce-HPCNb 0.923b 0.831b 91b 5.6%a 1000
b no loss
NR108
Co3O4@Z67-N700@CeO2NRb 0.86n = 4
b 66.8
n = 3.82–3.92
b 4.10 to 8.55%
a 5000
b 6
NR112
Ce@Co3O4/CNFsNRb 0.81n = 4
b 85.8
NRNRc 97.7116
Co3O4/CeO2@Co/N-CNFb 0.88b 0.75a 127.6NRNRNR121
CeCo-NxCCb 0.95b 0.85n = 3.7n = 3.98
b 9
a 8000
b 10.1
c 114130
Co9S8/CeO2/Co-NCNRb 0.875n = 4.0n = 3.80
b 8.3
a 2000
b 10
c 164138
CeO2@N-C-900b 1.003b 0.908a n = 3.97NRa 3000
b no loss
NR142
CeO2–CN–800b 0.90b 0.84a n = ~4n = ~4
b 10
NRc 65143
Ce-N4O2NRb 0.87a n = ~4
b 76
NRa 30,000
b 11
b 180159
Ce-N-CNRb 0.89a n = 3.94
b 62
n = ~4
b 4
a 10,000
b 190
NR160
Ce SAS/HPNCa 1.04a 0.862a 66NRNRH2/O2 525 H2/air: 250 W162
CeO2/Co@N–Cb 0.998b 0.934a n = 4NRNRc 103168
CeO2@CoSe2-NCsNRb 0.76a n = 4
b 86
n = 3.8–3.9
b 9
NRc 153169
4.8% Ce-MnO2/Cb 0.872b 0.783n = 3.97
b 90
n = 4
b 1.0–2.7
NR348
(Al–air)
170
CeO2/rGOb 0.91NRa n = 3.3–3.5NRNRNR171
CeCNCFb 0.86b 0.70n = 3.99b 0.06a 3000
b 25
NR172
NR = Not reported.
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

Peera, S.G.; Kim, S.W. Rare Earth Ce/CeO2 Electrocatalysts: Role of High Electronic Spin State of Ce and Ce3+/Ce4+ Redox Couple on Oxygen Reduction Reaction. Nanomaterials 2025, 15, 600. https://doi.org/10.3390/nano15080600

AMA Style

Peera SG, Kim SW. Rare Earth Ce/CeO2 Electrocatalysts: Role of High Electronic Spin State of Ce and Ce3+/Ce4+ Redox Couple on Oxygen Reduction Reaction. Nanomaterials. 2025; 15(8):600. https://doi.org/10.3390/nano15080600

Chicago/Turabian Style

Peera, Shaik Gouse, and Seung Won Kim. 2025. "Rare Earth Ce/CeO2 Electrocatalysts: Role of High Electronic Spin State of Ce and Ce3+/Ce4+ Redox Couple on Oxygen Reduction Reaction" Nanomaterials 15, no. 8: 600. https://doi.org/10.3390/nano15080600

APA Style

Peera, S. G., & Kim, S. W. (2025). Rare Earth Ce/CeO2 Electrocatalysts: Role of High Electronic Spin State of Ce and Ce3+/Ce4+ Redox Couple on Oxygen Reduction Reaction. Nanomaterials, 15(8), 600. https://doi.org/10.3390/nano15080600

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