Tong Wu‡
a,
Xiaoyi Zhang‡a,
Ziyu Yanga,
Zhilin Chena,
Yihao Longb,
Liang He*b,
Changsong Dai*c,
Jibing Chen
d and
Hui Tang
*a
aSchool of Materials and Energy, University of Electronic Science and Technology of China, Chengdu, 611731, China. E-mail: tanghui@uestc.edu.cn
bSchool of Mechanical Engineering, State Key Laboratory of Intelligent Construction and Healthy Operation and Maintenance of Deep Underground Engineering, Sichuan University, Chengdu 610065, P. R. China. E-mail: hel20@scu.edu.cn
cSchool of Chemistry and Chemical Engineering, Harbin Institute of Technology, Harbin 150001, P. R. China. E-mail: changsd@hit.edu.cn
dSchool of Mechanical Engineering, Wuhan Polytechnic University, Wuhan 430023, China
First published on 6th May 2025
High-entropy materials (HEMs), due to their exceptional physicochemical performance, which includes a unique electronic structure, outstanding catalytic performance, and remarkable electrochemical stability, are considered to be promising catalysts for applications such as water-splitting, underscoring their potential in electrocatalysis. Given the significant potential for their development and promising future applications for HEMs as electrocatalysts, research in this field is rapidly expanding. However, despite numerous innovative advancements, comprehensive summaries of HEMs as electrocatalysts are still lacking. This review summarizes the synthesis, characterization, and applications of HEMs in electrocatalysis. We discussed the synthesis of high-entropy catalysts from three perspectives: dry synthesis, wet synthesis, and rapid energy-based synthesis. Subsequently, the employment of advanced characterization techniques is discussed, along with electronic structure analysis and DFT calculations, to evaluate the high-entropy catalysts. Additionally, we summarized the exploration of the applications of these catalysts in electrocatalysis, focusing primarily on hydrogen evolution, oxygen evolution, and oxygen reduction. Finally, we provided a summary of the review's contents and presented insights into mechanistic research, material synthesis, applications of these, and future development prospects, with the goal of offering valuable suggestions for the future synthesis and applications of these.
The concept of high entropy was first introduced for the development of HEAs, with the earliest reports dating back to 2004.16,17 Before this formal publication, similar materials were studied under different names.18 Following the introduction of the high-entropy concept, these materials rapidly became a focal point of research. HEMs were found to have superior physical properties, and their catalytic potential was subsequently recognized. According to recent reports, substantial research efforts have been directed toward HEMs.19–22 As demonstrated in Fig. 1, since being proposed in 2004, significant progress has been made in the research and development of HEMs.
![]() | ||
Fig. 1 The development of HEMs (cited from refs. 23–27). |
HEMs exhibit remarkable physical properties, such as high strength, superior hardness, and enhanced wear resistance, as well as exceptional chemical properties like corrosion resistance, thermal stability, and catalytic activity.23–25 The multiple components in HEMs interact to provide ligand and strain effects, influencing the d-band structure and the electronic structure of the active sites. This unique, adjustable electronic structure holds great promise in the field of catalysis.26 The core characteristics of HEMs, HEAs, HEOs, and HEMSs are high configurational entropy and significant lattice distortion due to their multi-component nature.27–29 Despite these similarities, differences exist among these materials, stemming from their distinct compositions and structural variations.
Due to their unique composition and special electronic structure, HEMs have shown great potential in energy-related applications such as water-splitting and fuel cells.30,31 Water-splitting and fuel cells are two important fields that have a wide range of applications in energy conversion and storage, and are considered to be the core components of future clean energy systems. With the progress of green and low-carbon science and technology, as well as the maturity of the market, water decomposition and fuel cells are expected to achieve large-scale applications in the next few years, providing strong support for the global economic and social transformation to green and low-carbon.
For instance, many studies have demonstrated the superior performance of HEAs in the catalysis of the hydrogen evolution reaction (HER), the oxygen evolution reaction (OER), and fuel cells. Additionally, other types of HEMs, including HEOs and HEMSs, have shown outstanding performance in the oxygen reduction reaction (ORR) and HER/OER, respectively.32–36 Thus, high-entropy electrocatalysts hold significant research potential and applications in new energy resources.
Recently, numerous exceptional HEMs have been employed as catalysts, exhibiting noteworthy catalytic properties in diverse reactions. Comprehensive reviews have been conducted, covering the design, synthesis, applications, and calculations of HEMs, thereby enhancing our understanding and offering guidance for further development. Building upon existing research efforts, this review presents a thorough overview of recent advancements in the synthesis of high-entropy catalysts, with a focus on practical applications.
This is achieved by examining various synthesis pathways employed for various types of HEMs and elucidating the underlying principles governing their formation. Furthermore, this review describes the distinctive characteristics exhibited by HEMs and explores their applications in water-splitting and fuel cell-related reactions (primarily HER, OER, and ORR).37–39 Finally, an analysis of both the prospects and challenges associated with employing HEMs as catalysts is provided. The ultimate objective of this review is to offer valuable insights and guidance for synthesizing highly efficient high-entropy catalysts. And all the topics covered in this review are demonstrated in Fig. 2.
This review distinguishes itself by introducing a dry–wet rapid energy-based synthesis framework that uniquely elevates rapid energy-based synthesis as an independent category, emphasizing emerging technologies’ critical role in fabricating nanoscale HEMs—a departure from conventional classifications based solely on material types or singular synthesis routes. It establishes a comprehensive research architecture for high entropy electrocatalysis through a closed-loop synthesis–characterization–application–theory analytical approach, integrating fundamental exploration with practical implementation. These original perspectives not only strengthen theoretical foundations for material research but also delineate engineering-oriented pathways, signifying a paradigm shift in HEMs from conventional fabrication toward systematic investigations encompassing precision design, advanced characterization, and scenario-specific optimization.
Take HEAs for example, HEAs can be defined in two primary ways: component-based and entropy-based. The component-based definition characterizes HEAs as materials containing five or more elements in relatively high concentrations (5–35 at%).40 In contrast, the entropy-based definition identifies HEAs through their mixed configurational entropy. While it appears more rigorous to define HEMs by configurational entropy, the component-based definition is empirical. However, the threshold for configurational entropy can vary across different multicomponent systems.41
The entropy-based definition identifies HEAs via the mixed configuration entropy (S). The mixed configuration entropy of HEAs is able to be depicted by the following eqn (1).
![]() | (1) |
As such, S of HEAs with equal molar ratios for metallic elements in the liquid state or the solid solution state can be simplified as eqn (2).
S = R![]() | (2) |
For an alloy with the number of elemental components ≥5, the alloy with mixed configuration entropy S ≥ 1.5R refers to a HEA. Particularly, the alloy with S ≥ 1.36R is also identified as a HEA for a quaternary alloy.19,42 These two definitions of HEAs cover a wide range of alloys based on composition and entropy, and in most cases overlap.
For oxides, configuration entropy S ≥ the experience threshold 1.5R, it can be generally regarded as the formation of HEO. In addition, HEO usually exhibits entropy stability under the premise of high configurational entropy, but entropy stability is not a necessary condition. Other HEMs such as HEMS and HEPI basically refer to configuration entropies greater than or equal to 1.5R, or entropy stability is achieved when the components are not fewer than five.43 For example, HEMSs usually feature homogeneously mixed multi-metallic elements (≥5) in a sulfide structure.44
In addition to the entropy value, the chemical bond properties of HEMs are also closely related. The stability of HEMs is intrinsically linked to their chemical bonding characteristics, which differ markedly between HEAs and HEOs. In HEAs, metallic bonding—defined by delocalized electron clouds and electrostatic interactions between cations and free electrons—imparts isotropic mechanical properties, high ductility, and exceptional electron mobility. This bonding underpins their structural stability, as random solid-solution phases (e.g., FCC or BCC) accommodate multiple principal elements within shared lattice sites. Local lattice distortions, though significant, are mitigated by the electron cloud's ability to buffer strain energy, suppressing elemental segregation. Furthermore, the high configurational entropy (ΔSconfig) of HEAs dominates their Gibbs free energy (ΔG = ΔH − TΔS), favoring single-phase solid solutions over intermetallic compounds at elevated temperatures. In contrast, HEOs derive stability from a hybrid bonding framework: predominantly ionic interactions between metal cations and O2− anions, with partial covalent contributions from high-valent cations.45 The rigid O2− sublattice acts as a stabilizing scaffold, dispersing cationic lattice distortions electrostatically while maintaining structural integrity.46 Unlike HEAs, HEOs leverage both configurational entropy and strong ionic/covalent bonding energies to offset positive mixing enthalpy (ΔHmix), ensuring thermodynamic stability across wide temperature ranges. This synergy also grants HEOs superior corrosion resistance in Cl−-rich environments, where the O2− sublattice effectively blocks Cl− penetration, a critical limitation for HEAs. Mechanistically, metallic bonding in HEAs prioritizes entropy-driven stabilization and mechanical resilience, whereas HEOs excel under extreme chemical/thermal conditions due to their ionic–covalent hybrid nature.47–49 Fig. 3(a–c) gives an example of how element selection is made, and finally HEMs are obtained with outstanding stabilization.
![]() | ||
Fig. 3 (a) Schematic diagram of elemental selection and calculations (cited from ref. 33). (b) Schematic illustrating the mechanism of high-entropy spinel oxides for decoupling oxygen reduction and evolution reactions (cited from ref. 45). (c) Carbon-supported high-entropy oxide (HEO) nanoparticles as stable electrocatalysts (cited from ref. 46). |
According to the Gibbs free energy equation, materials with high-entropy will experience a significant reduction in Gibbs free energy at elevated temperatures. This decrease in free energy enhances the overall stability of the system. Additionally, the high-entropy results in a disordered distribution of the constituent elements within the material, which helps to prevent phase separation.
The random distribution of multiple elements in HEMs leads to high configurational entropy, contributing to material stabilization and enhanced thermodynamic stability. The complex constituent elements and significant atomic size differences, coupled with random atomic distribution, result in internal structures that differ markedly from crystalline materials, causing severe lattice distortion.53,54 The serious lattice distortion in HEAs is considered for raising the energy barrier of atomic diffusion. Because of lattice distortion, mass diffusion inside the material is hindered, which also contributes to the formation of HEAs.55,56 Compared with conventional alloys or ideal lattices, the single-phase crystals formed produce great lattice distortion, greatly changing the atomic environment of each atom, and the electronic structure is changed, thus altering the catalysis performance in various reactions.
Besides, in the high-entropy environment, the strong local electron interactions between the atoms of different elements will change the electron density and thus the catalytic activity. Due to the unique binding energy distribution, HEA nanoparticles can be readily tuned to obtain the desired surface properties for optimal catalytic performance. Interactions between different atoms and irregular atomic arrangements affect the diffusion of atoms in HEAs, with each vacancy in the lattice surrounded by a different atom. As a result, the lattice potential energy at different locations exhibits significant differences, which results in a high diffusion activation energy that inhibits the diffusion of atoms.
The “cocktail effect” in high-entropy alloys (HEAs) refers to the synergistic interplay of multiple constituent elements that collectively yield properties surpassing those of single-metal or binary systems, driven by three interlinked mechanisms: electronic structure modulation, lattice distortion-induced adsorption optimization, and multisite cooperative catalysis.8,57 At the electronic level, the coexistence of elements with varying electronegativities and atomic radii induces charge redistribution, as exemplified by shifts in the d-band center of transition metals (e.g., Ni, Fe, or Co), which directly regulates the adsorption strength of reaction intermediates (e.g., *H, *O, or OOH). Concurrently, severe lattice distortions arising from atomic size mismatch generate localized strain fields and disordered coordination environments, which optimize adsorption energies by creating heterogeneous active sites with tailored binding strengths.58 Furthermore, the multisite synergy enables parallel reaction pathways, collectively accelerating kinetics in complex reactions like overall water splitting. Such cooperative effects, absent in single-element catalysts, are further amplified by entropy-stabilized phase homogeneity, which prevents elemental segregation under operational conditions.
As HEMs typically consist of five or more constituent elements, the process of selecting constituent elements and adjusting the proportions of each element becomes crucial.
HEMs are usually composed of five or more elements mixed in close to equal molar ratios. This multi-alloying effect can lead to enhanced surface activity, as the synergies of different elements may provide diverse reaction paths. By changing the proportion of the elements in a HEM, it is theoretically possible to precisely adjust the physical and chemical properties of the catalyst.
In HEMs, the coexistence of multiple elements can significantly affect the density of electron states (DOS) and thus the position of the Fermi levels.59 This helps to optimize the adsorption strength of the reactants and products and improve the catalytic efficiency. In addition, by adjusting the component elements, it is possible to design HEMs with specific band gaps or conductivity properties, thereby optimizing their performance in electrocatalysis.60
Moreover, due to the uneven distribution of different elements on the surface of HEAs, a large number of heterogeneous active sites will be formed, which may have higher activity and selectivity. The HEMs are prepared as nanoparticles or thin films to increase the specific surface area and improve the catalytic efficiency. Specific morphologies can also be constructed using self-assembly techniques or template methods to further enhance the catalytic activity.
The Sabatier principle posits that for a catalyst to be effective, it must have an optimal binding strength for the reactants. If binding is too weak, the reactants will not adhere to the catalyst surface long enough to react. Conversely, if binding is too strong, the reactants may not be able to desorb after the reaction, leading to catalyst deactivation.61,62 In designing catalysts, especially for processes like hydrogenation and oxidation, understanding the Sabatier principle helps researchers select materials that can achieve the desired balance of binding energies.63,64
Under guidance from the volcano map, the selection of appropriate elements for catalyst synthesis helps to achieve a balance to obtain the best binding energy for improved catalytic efficiency.
Due to the inherent multi-component nature of HEMs, a broad selection of raw materials and diverse preparation strategies are available, offering extensive flexibility in synthesis approaches. In this review, we categorize synthesis methods as dry synthesis and wet synthesis based on whether a solution system is involved in the synthesis process. Given the particularities of deposition techniques such as magnetron sputtering in dry synthesis, we further discussed dry synthesis by dividing them into sputtering and non-sputtering methods.
By examining raw material choices, the underlying principles of each synthesis method, and describing the product morphology and properties, the advantages and disadvantages of different approaches are compared. This comprehensive evaluation aims to provide an objective assessment to guide researchers in selecting appropriate methods for synthesizing high-performance HEM catalysts tailored to specific applications.
For instance, Liu et al. successfully synthesized a porous NiCoFeMoMn HEA exhibiting exceptional electrochemical performance using powder metallurgy techniques.71 The precursor of high-entropy NiCoFeMoMn alloy ribbons was prepared by arc melting and single-roller melt spinning. The nano-porous NiCoFeMoMn catalyst was subsequently fabricated via a one-step electrochemical dealloying process, which significantly increased the number of active sites. This alloy demonstrated high catalytic activity for the HER and impressive efficiency for the OER in alkaline solutions. As shown in Fig. 4(a and b), DFT calculations indicate that the effective performance of the NiCoFeMoMn alloy is attributed to the synergistic effects of the alloying elements on surface electron density between the single atom (SA) and un-SAs. This research contributes to the development of cost-effective HEA catalysts.
![]() | ||
Fig. 4 (a) Colored ΔGH* and ΔEH2O comparisons of SA and un-SA. (b) The ELF maps of SA and un-SA (cited from ref. 72). (c) Atomic-level design and the preparation process for the O-HEA electrode. (d) Multilevel structure of the O-HEA integrated electrode. The inset shows the size of the bulk O-HEA electrode (cited from ref. 73). (e) Multistage mechanical alloying strategy and microstructure design concept (cited from ref. 74). |
Besides, Chen et al. developed a series of oxygen micro-alloyed HEAs (O-HEAs) via a metallurgy approach.72 Fig. 4c outlines the design and preparation strategy for O-HEAs, while Fig. 4d showcases island-like Cr2O3 microdomains within the HEA matrix. Fig. 4e further illustrates the electrochemical applications of various samples. Their findings revealed that a bulk O-HEA composed of (CrFeCoNi)97O3 exhibited a remarkable electrocatalytic performance for the OER, which was attributed to the formation of island-like Cr2O3 microdomains, leaching of Cr3+, and structural amorphization at the interfaces of these domains.
In addition to melting and subsequent treatment such as cutting and corrosion after mixing the metal powder directly, mechanical alloying through ball milling during the mixing process represents a viable strategy. Liu et al. employed multistage mechanical alloying to construct nanocrystalline FeCoNiCr0.4Cu0.2 HEA powders characterized by large aspect ratios and thin intergranular amorphous layers.73 Although this product was not utilized for electrocatalysis, its uniform element distribution, excellent manufacturability, and the ability to adjust the crystal phase and crystallinity through varying grinding times provided a valuable approach for preparing high-entropy catalysts.
For instance, Wang et al. successfully deposited HEA thin films on carbon fiber cloth using pulsed DC reactive magnetron sputtering.74 As illustrated in Fig. 5a, prior to deposition, the target was cleaned by Ar+ bombardment to remove surface impurities. Subsequently, FeCoNiCuPd HEA thin films were deposited onto the carbon fiber cloth through pulsed DC reactive magnetron sputtering of Fe/Co/Ni/Cu targets.
![]() | ||
Fig. 5 (a) Schematic illustration of key FeCoNiCuPd thin film fabrication steps (cited from ref. 75). (b) Schematic illustration of the experimental study platform of the HEA electrocatalysis model surface (cited from ref. 76). |
In a related study, Chida et al. proposed an experimental platform enabling the vacuum synthesis of atomic-level-controlled single-crystal HEA surfaces.75 This platform provides indispensable insights into understanding the microstructural intricacies crucial for electrocatalysis. Specifically, it elucidates the complex interplay between surface microstructures of multi-component alloys and their catalytic behaviors (Fig. 5b).
The dry synthesis methods discussed above offer robust pathways for fabricating advanced high-entropy catalysts with tailored properties suited to specific applications. These methodologies not only ensure homogeneous mixing controlled composition but also enhance environmental sustainability, making them invaluable tools for advancing the field of electrocatalysis beyond traditional boundaries.
![]() | ||
Fig. 6 (a) The traditional synthesis method and (b) the low-temperature plasma strategy for HEOs (cited from ref. 77). (c) Schematic showing the formation process for 10-HEA NWs (cited from ref. 78). (d) Schematic illustration of the electrocatalyst preparation steps (cited from ref. 79). |
In addition to hydrothermal and solvothermal methods, which are also commonly employed for synthesizing high-entropy catalyst precursors, Sun et al. synthesized a series of Pt-based high-entropy metallic nanowires via a solvothermal method as demonstrated in Fig. 6c.77 This technique uniformly mixes multiple elements at low temperatures (180–220 °C), forming high-entropy nanowires with controllable structures and compositions. It can be extended to the preparation of 17 types of high-entropy nanowires. Compared with their low-entropy counterparts, lattice distortion in these nanowires alters the strain distribution and electronic structure, exhibiting excellent catalytic performance in the hydroxide oxidation reaction (HOR) and the HER.
By heating precursors in a solvent at low temperatures, it is possible to synthesize stable nanoparticles with small particle sizes that are well dispersed.78,80,81 Adjusting the heating time and temperature along with solution components allows for tuning of the crystal morphology, size, pore size, and functionalization degree; this is a significant advantage over melting methods that struggle with nanoscale preparation. This is a good solution to the shortcomings of melting and other methods that are difficult to use to prepare nanoscale materials, so it can be used for the preparation of nanoscale catalysts.
Building on solvothermal methods, coating metal precursors with organic frameworks offers another convenient approach for preparing high-entropy catalysts. Fan et al., for instance, synthesized MOF precursors on carbon cloth via solvothermal techniques followed by thermal treatment under a reducing atmosphere to obtain HEA catalysts.
However, the solvothermal process is not always necessary, Wang et al. reported an easy-to-scale method for synthesizing advanced HEA (CoNiCuMnAl)/C nanoparticles from polymetallic MOFs.82 As shown in Fig. 6d, the MOF precursor of the catalyst was obtained by fully stirring the organometallic salt and the organic ligand in the solvent and forming a precipitate during the reaction. The face-centered cubic structure of the HEA core was coated in an ultra-thin carbon shell and deposited on Ni foam, exhibiting a superior OER performance.
The sol–gel method can achieve molecular-level homogeneity in a very short time. During the formation of the gel, the reactants are likely to be evenly mixed at the molecular level.79 Due to the solution reaction step, it is easy to uniformly and quantitatively incorporate certain elements, resulting in uniform doping at the molecular level. This is beneficial for synthesizing high-entropy catalysts. Additionally, the sol–gel method requires only a low synthesis temperature and it is generally believed that component diffusion in the sol–gel system occurs at the nanometer scale. Thus, the reaction proceeds easily and requires relatively low temperatures for synthesis.
Using this method, Tang et al. synthesized a high-entropy perovskite cobaltate consisting of five equimolar metals (Mg, Mn, Fe, Co, and Ni) in the B-site as an electrocatalyst for the OER,83 as shown in Fig. 7(a and b). The high-entropy cobaltate demonstrates a low overpotential, outperforming its other counterparts.
![]() | ||
Fig. 7 (a) Schematic demonstrating the structure of high-entropy perovskite oxide nanoparticles with uniformly dispersed elements. (b) Configurational entropy as a function of the number of cations in the B-site in the material system (cited from ref. 83). (c) Schematic illustration of selective Zn dealloying through a vapor dealloying process (cited from ref. 31). |
While synthesizing HEO particles via the sol–gel method is straightforward and feasible, special consideration must be given to prevent agglomeration of oxides during reduction when synthesizing HEAs. As demonstrated in Fig. 7c, Kwon et al. synthesized an Ir-based electrocatalyst, which was designed based on the HEA platform ZnNiCoIrX with two elements (X: Fe and Mn) and prepared by the sol–gel method.26 Instead of reducing oxide powder produced by the gel directly, they annealed nitrate gels under a flow of H2/Ar gas to avoid potential agglomeration during annealing.
In addition, ion exchange is currently a widely used method for preparing high-entropy catalysts. Miao et al. developed a ZnFeNiCuCoRu–O HEO catalyst that exhibited exceptional activity and ultra-stability for the OER over the full pH range via ion exchange.84 Their synthesis strategy involves using a MOF as a template to create HEO catalysts with polyhedral shapes and hollow structures, incorporating up to 10 different metal elements. This approach underscores the significance of the ion-exchange method for producing highly stable and active hollow-structured HEO catalysts, which are crucial for efficient energy conversion and storage devices.
Carbothermal shock is a facile method for synthesizing multi-metal nanoparticles, including HEM nanoparticles (HEM-NPs). In this approach, the preparation involves loading metal precursors onto a conductive carrier, such as a carbon substrate. Once prepared, the carrier is connected to a power source, and nanoparticles are generated by running a short pulse of electrical current through the sample to raise the temperature instantaneously. The rapid rise and fall in temperature during carbothermal shock make it an innovative way to synthesize nanoparticles.
Yao et al. synthesized extreme HEA nanoparticles containing 15 elements using carbothermal shock.85 Fig. 8a demonstrates that their method involves Joule heating of precursor-loaded carbon nanofiber films. By adhering to alloying criteria and employing high-entropy designs coupled with high temperatures, they achieved a record 15-element HEA nanoparticle. This process overcame immiscibility in strongly repulsive combinations and induced metal reduction of easily oxidized elements, thereby extending the range of synthesizable HEAs and demonstrating the great potential of HEMs. Similarly, Abdelhafiz et al. synthesized a type of non-noble metal HEO catalyst in situ on carbon fiber through rapid Joule heating and quenching.86 Their synthesis protocol started with drop-casting multi-metal chloride salts dissolved in ethanol at a concentration of 50 mM onto carbon fiber paper. Quenching occurred within fractions of a second, forming solid HEA nanoparticles.
![]() | ||
Fig. 8 (a) Schematic showing the ultra-fast rapid Joule heating and spontaneous cooling process within mS pulse intervals (cited from ref. 25). (b) Schematic showing the mechanism of morphological changes in LP532 (cited from ref. 87). (c) Nanodroplet-mediated electrodeposition overview for controlling NP stoichiometry and microstructure (cited from ref. 88). (d) The FMBP strategy for the synthesis of HEA-NPs (cited from ref. 89). |
Laser ablation is a surface modification technique where cladding material is added to the surface of the substrate, which is irradiated by a laser beam with a high energy density. This process forms a cladding layer on the substrate's surface that is metallurgically bonded and exhibits special physical, chemical, or mechanical properties through rapid melting, expansion, and solidification. The cladding layer has a low dilution rate, fewer pores, good metallurgical bonding with the matrix, and properties such as high hardness, good corrosion resistance, wear resistance, and stable quality. As presented in Fig. 8b, Rawat et al. synthesized Al-rich non-equiatomic HEA NPs by ablating the Al40(SiCrMnFeNiCu)60 (at%) target in deionized water.87 This work investigated and discussed structural, compositional, and morphological changes in Al40(SiCrMnFeNiCu)60 NPs, and put forward a possible formation mechanism for Cu–Ni enriched HEA NPs. Besides, Wang et al. synthesized a library of HEA and ceramic nanoparticles by laser scanning ablation.90 Their work presented an easy-to-adopt strategy for developing HEA and HEC NPs.
The electrochemical deposition method involves connecting the power supply to the anode and cathode poles of the electrolyte to form a circuit in a water-soluble or organic-soluble electrolyte. Under the influence of an electric field, an electrochemical reaction occurs, causing ions to precipitate as dense pure metals or alloys onto a substrate through a redox reaction, thereby creating the desired coating. This method avoids high temperatures during synthesis, making it a relatively straightforward synthesis strategy. For example, Glasscott et al. presented a generalized strategy to electro-synthesize HEMG-NPs with up to eight equimolar components by confining multiple metal salt precursors in water nanodroplets emulsified in dichloroethane.91 As shown in Fig. 8c, HEMG-NPs were electrodeposited on highly ordered pyrolytic graphite (HOPG) or glassy carbon substrate electrodes from a water-in-oil emulsion system.88 Additionally, Chang et al. developed a high-entropy FeCoNiMnW alloy in situ on carbon paper using a pulsed current electrodeposition method.92
In previously used synthesis methods like carbothermal shock and electrochemical shock, these methods required conductive carriers, which limited carrier selection significantly. Gao et al., based on the common wet impregnation method, proposed a fast-moving bed pyrolysis (FMBP) approach for synthesizing HEA nanoparticles (HEA-NPs).89 In this work, a facile FMBP synthesis strategy was developed for HEA-NPs with up to ten elements by ensuring mixed precursors were pyrolyzed at high temperatures simultaneously. This resulted in highly dispersed HEA-NPs on supports, as detailed in Fig. 8d.
These methods outlined above highlight the versatility and efficiency of rapid energy-based synthesis techniques for producing advanced HEMs with potential applications in catalysis, coatings, and other fields requiring superior material properties. Finally, the synthesis methods mentioned above are summarized in Table 1.
Synthesis method | Advantage | Limitation | Applicable scope |
---|---|---|---|
Powder metallurgy | Precise composition control; high mechanical stability; tunable porosity via post-sintering dealloying/etching | High equipment cost; challenges in nanomaterial synthesis | High-strength bulk catalysts for high-temperature reactors |
PVD | Ultra-thin film uniformity; precise thickness control | Weak film–substrate adhesion at elevated temperatures; high cost | Ultrathin catalytic coatings |
Magnetron sputtering | High-density films; scalable deposition | Impurities from sputtering gases or low-purity targets; target consumption | Nanoscale multilayer films |
Hydrothermal | High crystallinity; synthesis of metastable phases (e.g., spinel HEOs) | Batch size limitation (<100 mL); high-pressure reactor required | Phase-pure HEOs |
Solvothermal | Size-controlled nanoparticles; excellent dispersion | Toxic/organic solvents; high cost | HEA NPs |
MOF | Ultrahigh surface area; atomic-level dispersion | Low stability in non-carbonized forms; complex synthesis | Porous carbon composite catalyst |
Sol–gel | Molecular homogeneity; low-temperature processing; mesoporous structures via supercritical drying | Time-consuming drying; gel shrinkage | 150 nm |
Ion-exchange | Tailored morphologies (e.g., hollow/core–shell structures) | Impurity risks; strict condition control | Synthesis of catalyst with specific morphology |
Carbon thermal shock | Ultrafast synthesis (<1 s); immiscible metal integration | Limited to conductive substrates (e.g., carbon cloth); small-scale only | HEA NP libraries for high-throughput HER/OER screening |
Laser ablation | Purity >99.9%; ligand-free surfaces; colloidal NP synthesis | Low yield; high energy consumption | HEA |
Electrochemical deposition | Atomic-level thickness control (e.g., monolayer deposition); ambient conditions | Slow deposition rates (improved via pulsed techniques) | Corrosion-resistant coatings, microelectronic interconnects |
Fast-moving bed | Ultra-fine powders; instant precursor aerosolization and pyrolysis | Lab-scale only; precise atmospheric control required | Ultrafine HEOs |
Overall, the synthesis of HEMs faces inherent limitations tied to method-specific constraints. Solid-state approaches like powder metallurgy and magnetron sputtering suffer from high equipment costs, nanomaterial scalability challenges, and impurity incorporation (e.g., sputtering gases), which can be mitigated via cost-effective sintering alternatives, high-purity targets, or reactive gas optimization. Vapor-phase techniques (PVD, electrodeposition) struggle with weak film–substrate adhesion and slow deposition rates, necessitating interfacial engineering (e.g., adhesion layers) or pulsed electrodeposition protocols. Solution-based methods (hydrothermal, solvothermal, MOF-derived) are hindered by batch size restrictions, toxic solvents, and instability in non-carbonized forms; scalable reactors, green solvent alternatives, and hybrid carbonization strategies offer promising solutions. Emerging methods such as carbothermal shock and laser ablation are limited by substrate dependency and low yields, requiring substrate versatility (e.g., ceramic supports) or energy-efficient pulsed modulation. Techniques like sol–gel and flash pyrolysis demand accelerated drying processes (e.g., microwave-assisted supercritical drying) and precise atmospheric control to minimize gel shrinkage and phase inhomogeneity. Cross-method synergies—e.g., integrating ion exchange's morphology control with electrodeposition's atomic precision or coupling MOF-derived confinement with flash pyrolysis—could unlock tailored catalytic architectures. Future advancements should prioritize scalable, energy-efficient protocols with in situ characterization to resolve metastable phase dynamics, ensuring industrial viability for applications in energy conversion, environmental catalysis, and high-throughput systems.
Atomic electron tomography (AET) and atom probe tomography (APT) are advanced analytical techniques used for investigating the atomic-level structure and composition of materials. AET, based on transmission electron microscopy (TEM), captures a series of 2D projection images of a sample from multiple angles, which are then reconstructed into a 3D representation using computational algorithms. It provides atomic-scale spatial resolution, enabling a detailed analysis of nanostructures, interfaces, and crystal defects. APT, on the other hand, utilizes field evaporation of atoms from the sample surface in a high electric field, with a time-of-flight mass spectrometer (TOF-MS) detecting the evaporated ions. This technique allows for the 3D reconstruction of the sample's atomic arrangement and provides precise elemental composition at the atomic level. Both techniques are pivotal in materials science, particularly in the study of nanomaterials, alloys, semiconductors, and catalytic systems.
For instance, as shown in Fig. 9a, to further investigate the surface composition and understand the mechanisms of activation and deactivation of the Cantor alloy during the OER, atom APT was employed to analyze the elemental distribution and composition of the surface oxides following 70, 150, and 2000 CV cycles. The results indicate that the Cantor alloy electrocatalyst surface exhibits its most OER-active state after the 70th cycle, a slightly deactivated state after the 150th cycle, and becomes inactive after 2000 cycles.103 APT provides a precise visualization of the arrangement of constituent elements in high-entropy catalysts, including their spatial distribution, atomic coordination, and potential atomic-scale defects or interfaces within the catalyst. Due to complex alloying and segregation behaviors at the atomic level in high-entropy catalysts, which typically consist of multiple elements in near-equiatomic compositions, APT is particularly valuable for resolving these intricate structural details.
![]() | ||
Fig. 9 (a) 2D concentration profiles of O, Cr, Mn, Fe, Co, and Ni plotted from a 5 nm thick slice of APT data of the Cantor alloys after 70, 150, and 2000 CV cycles under OER conditions (cited from ref. 103). (b) High-magnification HAADF-STEM image of a 15-HEA nanoparticle, on which a 4D-STEM dataset of this particle was acquired. (c) Subsequently generated virtual ADF image in which each pixel corresponds to a diffraction pattern. Three selected diffraction patterns at various locations within this particle are shown. (d) Bragg vector map showing an overlay of all diffraction patterns within the particle. (e) Strain map of the 15-HEA nanoparticle calculated from the variation in the lattice, showing the localized strain and inhomogeneity in the nanoparticle due to extreme mixing (cited from ref. 25). |
Four-dimensional scanning transmission electron microscopy (4D-STEM) extends traditional STEM by incorporating additional dimensions related to the electron angular distribution scattering. This technique simultaneously captures both spatial and angular information about electron scattering in a sample. One key advantage is its ability to capture dynamic processes and time-resolved phenomena at the nanoscale. By acquiring sequential diffraction patterns over time, researchers can observe structural changes, phase transitions, and dynamic events in real-time.
Fig. 9e showcases an atomically resolved high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) image of a singular 15-component high-entropy alloy nanoparticle with an accompanying 4D-STEM dataset acquired for this particle.85 Each pixel corresponds to a diffraction pattern used to derive local structural information such as lattice constants and strain deformation. The resultant Bragg vector map indicates single crystallinity with an FCC structure at the single-particle level. By combining high spatial resolution with the capability to capture dynamic processes, 4D-STEM offers valuable insights into the material's structure, properties, and behavior at the nanoscale.
Besides all the characterization methods mentioned above, Su et al. visualized the entire formation process for a high-entropy fluorite oxide from a polymeric precursor using atomic-resolution in situ gas-phase scanning transmission electron microscopy.104 This approach provides a reference for exploring growth processes and formation mechanisms while aiding better design strategies for synthesizing high-entropy catalysts.
X-ray absorption near edge structure (XANES) and extended X-ray absorption fine structure (EXAFS) are spectroscopic techniques utilizing synchrotron radiation sources to probe local atomic structures and electronic environments at the atomic scale.105,106 As demonstrated in Fig. 10(a–i), these techniques provide detailed coordination environment information that is essential for advancing our understanding of material applications across diverse scientific fields.
![]() | ||
Fig. 10 (a) HAADF-STEM image of Fe@PCN-900a. (b) Normalized Fe K-edge XANES spectra and (c) EXAFS spectra of Fe@PCN-900a, FePc, Fe foil, FeO and Fe2O3. Wavelet transform of Fe K-edge EXAFS of (d) Fe@PCN-900a, (e) Fe foil and (f) FePc (cited from ref. 106). (g) XANES spectra at the Pd K-edge. (h) The k3-weighted Fourier transforms of Pd K-edge EXAFS spectra, and (i) wavelet transforms from experimental data for Pd1@HEFO, PdO, and Pd foil (cited from ref. 105). (j) VB spectra obtained by HAXPES and (k) DOS profiles calculated by DFT for the NPs of NM-HEA, Pt, and Au (cited from ref. 12). |
Hard X-ray photoelectron spectroscopy (HAXPES) is an advanced analytical technique that probes the electronic structure of materials using high-energy X-rays. Unlike traditional XPS, which typically utilizes soft X-rays with energies up to a few keV, HAXPES employs hard X-rays with energies ranging from tens to hundreds of keV. This higher energy enables deeper penetration into the material, allowing for the study of buried interfaces and bulk properties with enhanced sensitivity and resolution.
As illustrated in Fig. 10(j and k), HAXPES can reveal detailed electronic structures such as valence bands and d-band centers in high-entropy nanoparticles.12 These insights are closely related to the adsorption and binding energies of key reaction intermediates, helping to rationalize their catalytic activity.107
For instance, to understand the fundamental mechanism behind enhanced HER catalytic activity, Wang et al. conducted DFT calculations on the basal plane of a catalyst. The geometry was optimized using refined PXRD data.110 As shown in Fig. 11(a–d), their results demonstrated that the enhanced HER activity in high-entropy MPCh3 originated from the synergistic effects of abundant active sites provided by the high-entropy strategy. Optimized sulfur sites on the edge and phosphorus sites on the basal plane offer more active sites for hydrogen adsorption. Additionally, manganese sites introduced on the edges act as efficient centers for water dissociation.
![]() | ||
Fig. 11 (a) Basal-plane models of P sites (P1–P3) and S sites (S1–S9) in Co0.6(VMnNiZn)0.4PS3. (b) HER free-energy diagram of corresponding sites in (a). (c) Calculated charge-density difference of the P1 site for Co0.6(VMnNiZn)0.4PS3. The red and blue regions refer to electron accumulation and depletion, respectively. (d) Calculated reaction energy of water dissociation for Co0.6(VMnNiZn)0.4PS3 and CoPS3, including Co, V, Mn, Ni, and Zn sites (cited from ref. 110). (e–k) Density functional theory calculations for the structural configuration and PDOSs (cited from ref. 111). |
Similarly, Li et al. applied periodic DFT calculations to explore the HER and the methanol oxidation reaction (MOR) performances in HEAs.111 As shown in Fig. 11(e–k), they compared the projected density of states (PDOS) of HEAs with slightly different stoichiometries and found highly similar electronic structures. This indicates that slight variations in HEA stoichiometry have a minimal effect on their electronic structure. By comparing different atomic arrangements, they identified an HEA structure with a slightly nickel- and copper-enriched surface as the most stable lattice model. This model exhibited subtle distortion after relaxation, which was indicative of good durability for electrocatalysis. Moreover, adsorption studies revealed active electron transfer from HEA to water and methanol during HER and MOR processes respectively, ensuring stable adsorption and facilitating subsequent reactions. The existence of a linear correlation between the transformation of intermediates throughout the MOR process, which ensured optimal binding strength and superior MOR performance in HEAs.
DFT calculations have become indispensable for deciphering the catalytic mechanisms of HEMs, bridging atomic-scale electronic interactions with macroscopic performance. At the core of this understanding lies electronic structure modulation, where DFT reveals how compositional complexity tailors catalytic behavior. To address HEMs’ vast compositional space, DFT synergizes with machine learning (ML) by providing datasets (e.g., formation energies, adsorption descriptors) that train ML models to predict phase stability and screen optimal compositions. For example, ML models trained on DFT-calculated d-band centers and metal–oxygen bond strengths accelerate the discovery of OER-active HEOs. However, challenges persist in modeling metastable phases and dynamic surface reconstructions under operational conditions, necessitating advanced ab initio molecular dynamics (AIMD) and hybrid functional approaches. By integrating electronic insights, active site hierarchies, and reaction energetics, DFT not only decodes HEMs’ catalytic superiority but also guides the rational design of next-generation catalysts for energy conversion and storage.
Overall, DFT has enabled researchers to elucidate the electronic structures of the catalysts synthesized effectively. Moving forward, it will play an even greater role in guiding catalyst design and synthesis.
In the adsorption step, protons (H+) are transferred from solution to the surface of the electrode and adsorbed on it, while electrons are transferred from the electrode to the proton.115 Desorption steps involve two possible paths:116 (1) the Tafel step (hydrogen–hydrogen recombination mechanism); two adsorbed hydrogen atoms combine to form hydrogen gas; and (2) the Heyrovsky step (proton–electron coupling transfer mechanism); an adsorbed hydrogen atom combines with a proton and an electron to form hydrogen gas.
In acidic media, the HER usually follows the mechanism described above, while in alkaline media, the difference is that the proton donor is obtained by the dissociation of water molecules, due to the extremely low proton concentration in alkaline media. Obviously, the HER mechanism in alkaline electrolytes is much more complex than in acidic electrolytes. Thus, the corresponding electrocatalytic kinetics of the HER is two to three orders of magnitude slower than that of acid media.117
Therefore, electrocatalysts with outstanding catalytic performances are essential to enhance HER efficiency.118–120 Currently, the HER primarily utilizes Pt-based precious metal catalysts because of their low overpotentials and high exchange current density.121,122 Among these, Pt/C is the most widely used catalyst for the HER. However, the scarcity and high cost of noble metals limit the widespread applications of Pt-based electrocatalysts. Thus, exploring alternative catalysts such as HEMs becomes significant.
As shown in Fig. 12a, Gu et al. reported a Turing structuring strategy to activate and stabilize superthin metal nanosheets by incorporating high-density nanotwins, which synergistically reduced the energy barrier of water dissociation and optimized the hydrogen adsorption free energy for the hydrogen evolution reaction.123
![]() | ||
Fig. 12 (a) Schematic diagram of the prepared Turing PtNiNb and corresponding crystallographic characterization (cited from ref. 123). (b) Schematic illustration of the full-active-site catalytic mechanism of HEANCs for the overall OHzS system (cited from ref. 124). (c) Schematic diagram of the multi-site synergistic effect in PtFeCoNiCuCr@HCS for efficient HER (cited from ref. 125). (d) Schematic diagram of the PEMWE electrolyzer (cited from ref. 126). |
Regarding non-noble metal high-entropy catalysts, Li et al. developed a quaternary FeCoNiCu HEA through dealloying the HEA, which was used as a water-splitting electrocatalyst.124 In alkaline electrolytes, this self-supported HEA-derived porous electrode exhibited a superior HER performance with a low overpotential of 42.2 mV to achieve a current density of 10 mA cm−2, along with a low Tafel slope of 31.7 mV dec−1. Additionally, it demonstrated excellent stability under a high current density of 500 mA cm−2. In this work, the synergistic effect of polymetallic atoms optimizes the HER performance through the following mechanisms: first, the alloying of Fe, Co, Ni with Cu reconstructs the coordination environment, effectively modulating the d-band electronic structure of Cu sites, which reduces the H* adsorption energy barrier and optimizes the HER kinetic process. Specifically, strong electron coupling between the s-orbitals of Cu and H near the Fermi level significantly lowers the energy barrier during H* reduction. Meanwhile, the d-orbitals of Fe/Co/Ni contribute greater electronic state density, further refining the electronic structure of active sites through orbital hybridization. This polymetallic synergy ultimately reduces both the water molecule adsorption energy and the Gibbs free energy of H* adsorption, establishing a more favorable reaction pathway for the HER.
For noble metal-based high-entropy catalysts, notable work has shown that they can approach or even exceed the catalytic performance of commercial Pt/C for the HER while reducing the amounts of precious metals required. Xia et al. synthesized an HEANC/C catalyst that achieved an ultrahigh mass activity of 12.85 A mg−1 noble metals at −0.07 V and an overpotential of only 9.5 mV for achieving 10 mA cm−2 in the alkaline HER.125
In the HEANC system, Pt and Ru atoms exhibit electron-rich states, while Ni and Co atoms with lower electronegativity transfer electrons to Pt/Ru sites, inducing a pronounced electron redistribution effect. This electronic structure modulation effectively optimizes the d-band center position, not only creating abundant active sites for the HER but also significantly reducing the reaction energy barrier. Fig. 12(b–d) further reveals the catalytic mechanism and active sites of the HER, demonstrating the actual process taking place during the reaction. Experimental data reveal that the system demonstrates a superior water adsorption energy (Ead (H2O)) during the Volmer step and more favorable hydrogen adsorption free energy (ΔGH*) in the Tafel step compared to pure Pt surfaces. Notably, the super-active sites exhibit ΔGH* values approaching 0 eV, which remarkably enhance the adsorption–desorption kinetics of hydrogen intermediates.
In Table 2, we provide a comparison and analysis of other recent high-entropy catalysts for the HER (unless otherwise specified, overpotential is always relative to a current of 10 mA cm−2).
Name | Synthesis method | Morphology | Size | Tafel slope | Overpotential | Electrolyte | Ref. |
---|---|---|---|---|---|---|---|
Pt18Rh16Mo15Ir11Co9Ru6Fe6Mn5Cr5 HEA NWs | Reduction–diffusion | Nanowire | N.A. | N.A. | 24 mV | 0.1 M KOH | 77 |
FeCoNiCu | Metallurgy and dealloying | Porous | 12.2 nm | 31.7 mV dec−1 | 42.2 mV | 1 M KOH | 124 |
FeCoNiCuPd | Magnetron sputtering | Film | 2 mm thick | 47.2 mV dec−1 | 29.7 mV | 1 M KOH | 74 |
PdPtCuNiP | Flexible as-spun and dealloying | Nanosponge-like architecture | Less than 10 nm | 37.4 mV dec−1 | 32 mV | 1 M KOH | 126 |
PdFeCoNiCu/C | Oil-phase synthesis | Particle | About 30 nm | 39 mV dec-1 | 18 mV | 1 M KOH | 127 |
np-12HEA | Powder metallurgy | Porous | 5–10 nm (grain size) | 29.5 mV dec−1 | 21 mV | 1 M KOH | 128 |
P doped Ni30Co30Cr10Fe10Al18W2 | Powder metallurgy | Porous | N.A. | 32.6 mV dec−1 | 70 mV | 1 M KOH | 129 |
Pt(Co/Ni)MoPdRh | Solvothermal | Nanoflower | N.A. | 26.8 mV dec−1 | 16.5 mV | 1 M KOH | 130 |
CoZnCdCuMnS@CF | Hydrothermal | Nanowire | N.A. | 98.5 mV dec−1 | 173 mV | 1 M KOH | 131 |
CoFeNiCrMnP/NF | Electrodeposition | Nanosheet | N.A. | 48 mV dec−1 | 51 mV (100 mA cm−2) | 1 M KOH | 132 |
CuAlNiMoFe | Alloying/dealloying | Nanoporous Monolithic | About 400 nm | ≈60 mV dec−1 | 56 mV (100 mA cm−2) | 1 M KOH | 133 |
np-UHEA14 | One-step dealloying | Nanoporous ribbons | 30–40 mm thick and 2 mm wide | 30.1 mV dec−1 | 32 mV | 0.1 M HClO4 | 134 |
NiCoMoPtRu HEANC | Co-reduction and annealing | Nanoclusters | 1.48 nm | 29.8 mV dec−1 | 9.5 mV | 1 M KOH | 125 |
ZnNiCoIrMn | Sol–gel and dealloying | Spherical | ∼200 nm | 30.6 mV dec−1 | 50 mV (50 mA cm−2) | 26 | |
Pt28Mo6Pd28Rh27Ni15 NCs | Wet chemistry | nanosheets | N.A. | 25.9 mV dec−1 | 9.7 mV | 1 M KOH | 135 |
IrPdPtRhRu HEA NPs | One-pot polyol process | Nanoparticles | 5.5 ± 1.2 nm | N.A. | 25 mV | 1 M KOH | 136 |
Pt4FeCoCuNi | Impregnation | Nanoparticles | 5 nm | 31 mV dec−1 | 20 mV | 1 M KOH | 137 |
Ni14Co14Fe14Mo6Mn52 | One-step dealloying and melting and single-roller melt spinning | Nanoporous HEA foil | ∼200–300 nm | 29 mV dec−1 | 14 mV, 150 mV (1000 mA cm−2) | 1 M KOH | 71 |
FeCoNiAlTi HEI | One-step chemical dealloying | N.A. | N.A. | 40.1 mV dec−1 | 88.2 mV | 1 M KOH | 138 |
Co0.6(VMnNiZn)0.4PS3 | Traditional solid-state reaction and ultrasonic peeling | Nanosheets | ∼400–500 nm | 65.5 mV dec−1 | 65.9 mV | 1 M KOH | 110 |
FeCoNiMnRu | Electrospinning technique and graphitization process | Nanoporous particles | 14.2 ± 9.1 nm | N.A. | 71 mV (100 mA cm−2) | 1 M KOH | 139 |
NiCoFePtRh NP | Co-reduction | Nanoporous particles | ∼1.68 nm | 30.1 mV dec−1 | 27 mV | 0.5 M H2SO4 | 140 |
PtCoNiRuIr | High-temperature liquid shock | Nanoporous particles | ∼3.24 nm | 34.2 mV dec−1 | 18 mV | 0.5 M H2SO4 | 141 |
Pt18Ni26Fe15Co14Cu27/C | Oil phase | Nanoparticles | ∼3.4 nm | 30 mV dec−1 | 11 mV | 1 M KOH | 111 |
![]() | ||
Fig. 13 (a) Conventional OER mechanism involving proton and electron transfers on the surface metal centers (cited from ref. 148). (b) Schematic comparison of the OER mechanism over the surface of FeCoNiAlMo high-entropy alloy (cited from ref. 149). (c) Free energy diagrams of the OER at U = 0 and 1.23 V on Ir0.7Ru0.14Ni0.08Mo0.08O2 and IrO2 models (cited from ref. 150). |
HEMs provide a promising path to discover new OER catalysts beyond noble metals. For example, Wang et al. prepared FeCoNiPB, FeCoPB, FeNiPB, and CoNiPB nanomaterials using a one-step chemical reduction method.151 The ORR efficiency of FeCoNiPB, containing three transition metals, was compared with samples containing two transition metals. In 1.0 M KOH solution, FeCoNiPB achieved a current density of 10 mA cm−2 at an overpotential of 235 mV, superior to those of FeCoPB (285 mV), FeNiPB (261 mV), CoNiPB (330 mV), and commercial RuO2 (316 mV).
Notably, at a high current density of 100 mA cm−2, FeCoNiPB maintained a low overpotential of 306 mV, better than most reported electrocatalysts. The mass activity of the FeCoNiPB catalyst was also impressive at 1983 mA mg−1 at 1.7 V vs. RHE, significantly higher than FeCoPB (441 mA mg−1), FeNiPB (1334 mA mg−1), CoNiPB (786 mA mg−1), and RuO2 (218 mA mg−1). Analysis after 40 h of oxygen evolution showed that self-reconstruction ensured its high efficiency and stable catalytic performance during the OER.
The catalyst establishes a moderately oxidized surface microenvironment through the coexistence of metallic (Fe0, Co0, Ni0) and oxidized states (Fe2+/Fe3+, Co2+/Co3+, Ni2+/Ni3+), providing abundant active sites for the OER. Nonmetallic B and P engage in directional electronic interactions with the metals; B donates electrons to Fe/Co/Ni, while P withdraws electrons from the metals. This synergistic electronic effect optimizes the d-band center position, reduces the adsorption free energy of *OOH intermediates, and lowers the OER overpotential. Concurrently, B/P doping enhances the amorphous phase formation capability, with the disordered structure regulating the surface adsorption behavior. During prolonged OER operation, the catalyst undergoes self-reconstruction; initial nanoparticles gradually transform into layered architectures, with edges developing highly active (FeCoNi)OOH crystalline domains. This reconstructed surface exposes numerous coordinatively unsaturated sites; Fe3+ in reconstructed (FeCoNi)OOH acts as the primary active center, with its orbital electron occupancy approaching the ideal value, thereby enhancing the lattice oxygen-mediated (LOM) pathway efficiency.
Additionally, Qiao et al. synthesized a high-entropy phosphate catalyst (HEPi), specifically CoFeNiMnMoPi, which demonstrated superior catalytic activity for the OER.152 The HEPi catalyst achieved an overpotential of 270 mV at 10 mA cm−2, significantly lower than both its HEO counterpart (350 mV) and the benchmark IrOx catalyst (340 mV). Moreover, this HEPi catalyst can be applied to other reactions as well. Their synthesis strategy is efficient and straightforward, offering a new approach to discover various polyanionic materials for energy and catalysis applications.
In Table 3, we compare other recent high-entropy catalysts for the OER alongside those already described to provide further insights into their performance metrics (unless otherwise specified, overpotential is always relative to a current of 10 mA cm−2).
Materials | Synthesis method | Morphology | Size | Tafel slope | Electrolyte | Overpotential | Ref. |
---|---|---|---|---|---|---|---|
FeCoNiMo HEA/C | High temperature reduction | Particle | 8 ± 0.3 nm | N.A. | 1 M KOH | 250 mV | 148 |
CoCrFeNiMo | Powder metallurgy | Porous | N.A. | 59.0 mV dec−1 | 1 M KOH | 220 mV | 149 |
FeCoNiCuPd film/CFC | Magnetron sputtering | Film on carbon fiber cloth | 2 mm | 39.8 mV dec−1 | 1 M KOH | 194 mV | 74 |
(CrFeCoNi)97O3 bulk O-HEA | Metallurgy | Bulk | N.A. | 29 mV dec−1, | 1 M KOH | 196 mV | 72 |
La0.6Sr0.4Co0.2Fe0.2Mn0.2Ni0.2Mg0.2O3 | Sol–gel | Particle | 40 nm | 45 mV dec−1, | 1 M KOH | 320 mV | 83 |
FeCoNiMnPd | Synergistic confinement | Particle | 150 nm | 96 mV dec−1 | 1 M KOH | 390 mV | 150 |
FeCoMoPB | Chemical reduction | Amorphous nanoplate | 20 nm | 38 mV dec−1 | 1 M KOH | 239 mV | 153 |
FeCoNiCrMo | Powder metallurgy | Porous | N.A. | 38.5 mV dec−1 | 1 M KOH | 303 mV (100 mA cm−2) | 154 |
HE–MOFs | MOF | Nanosheet | N.A. | 59 mV dec−1 | 1 M KOH | 274 mV | 155 |
FeCoNiPB | Chemical reduction | N.A. | 53 mV dec−1 | 1 M KOH | 235 mV | 151 | |
np-UHEA12 | One-step dealloying | Nanoporous ultra-HEA ribbons | 30–40 mm thick | 84.2 mV dec−1 | 1 M KOH | 258 mV | 134 |
CoNiCuMnAl@C | Precipitation and pyrolysis | Spherical nano-particles | N.A. | 35.6 mV dec−1 | 1 M KOH | 215 mV | 82 |
ZnFeNiCuCoRu-O HEOs | Ion exchange | Hollow skeleton structure | ≈200–250 nm | 56 mV dec−1 | 1 M KOH | 170 mV | 84 |
La0.6Sr0.4Co0.2Fe0.2Mn0.2Ni0.2Mg0.2O3 | Sol–gel | Near-spherical | ≈70–100 nm | 45 mV dec−1 | 1 M KOH | 320 mV | 83 |
(Cr0.2Mn0.2Fe0.2Co0.2Ni0.2)3O4 NFs | Electrospinning and calcination | Oxide nanofibers | Over 10 nm | 41 mV dec−1 | 1 M KOH | 360 mV | 156 |
CoFeNiMoWTe NHECGs | Hierarchical hybrid | Nanosheet arrays | ≈10 nm thickness | 66.8 mV dec−1 | 0.5 M H2SO4 | 373 mV | 157 |
NiFeCoMnAl oxide | Electrodeposition and dealloying | Nanoporous | N.A. | 47.62 mV dec−1 | 1 M KOH | 190 mV | 158 |
FeCoNiMoAl | As-cast melt-extracted and cooling | Core–shell | N.A. | 39.8 mV dec−1 | 1 M KOH | 223 mV | 159 |
NiFeXO4 | Eutectic dealloying | Nanowires | Diameter 200 nm and 10 μm long | 53.3 mV dec−1 | 1 M KOH | 195 mV | 160 |
PtPdAuAgCuIrRu | Alloying–dealloying | Nanoclusters | 1.5–2 nm | 49.3 mV dec−1 | 1 M KOH | 240 mV | 161 |
FeCoNiCrMo | Arc melting | Plate | 0.5 × 10 × 14 mm | 38.5 mV dec−1 | 1 M KOH | 303 mV (100 mA cm−2) | 154 |
Ni14Co14Fe14Mo6Mn52 | Single-roller melt spinning | Nanoporous | 200–300 nm diameter | 37 mV dec−1 | 1 M KOH | 243 mV | 71 |
AlCrCuFeNi | Vacuum induction melting | Nanoporous particles | N.A. | 77.5 mV dec−1 | 1 M KOH | 270 mV | 162 |
FeCoNiCuMn/CNFs | Polymer fiber nanoreactor | 3D network nanofibers | 12.9 ± 6.6 nm | 69 mV dec−1 | 1 M KOH | 386 mV (200 mA cm−2) | 163 |
FeCoNiMnRu | Electrospinning and graphitization | Nanoporous particles | 14.2 ± 9.1 nm | 67.4 mV dec−1 | 1 M KOH | 308 mV (100 mA cm−2) | 139 |
(CrFeCoNiMo)3O4 | N.A. | Nanosheets | N.A. | 37.0 mV dec−1 | 1 M KOH | 255.3 mV | 164 |
(FeCoNiCrMn)3O4 | Electrospinning | N.A. | N.A. | 71.38 mV dec−1 | 1 M KOH | 318.4 mV | 165 |
(CrMnFeCoNi)3O4 | Electrospun | Fiber | N.A. | 41 mV dec−1 | 1 M KOH | 360 mV | 156 |
(FeCoNiCuZn)O | Sol–gel | Particle | 200 nm–1 mm | 64.5 mV dec−1 | 1 M KOH | 323 mV | 33 |
(Fe0.2Co0.2Ni0.2Cr0.2Mn0.2)3O4 | Solution combustion | Porous | 10 nm | 50.27 mV dec−1 | 1 M KOH | 275 mV | 166 |
To address these challenges, He et al. synthesized FeCoNiMoW HEA nanoparticles using a solution-based low-temperature approach.173 Linear sweep voltammetry (LSV) curves obtained from FeCoNiMoW at various rotation rates (ranging from 400 to 2500 rpm) show that at 1600 rpm, the half-wave potential is 0.71 V with an onset potential of 0.83 V. In the FeCoNiMoW catalytic system, the hybridization between Ni 3d orbitals and O 2p orbitals generates new bonding orbitals, leading to a significant reduction in antibonding orbital electron occupancy and increased electron density in bonding orbitals, thereby effectively enhancing metal–oxygen bond stability. This orbital modulation effect originates from the synergistic electronic interactions between 3d transition metals (Fe/Co/Ni) and 4d/5d high-period metals (Mo/W); this constitutes the key mechanism for improving the catalytic activity for the oxygen reduction reaction (ORR). Density functional theory calculations reveal substantially reduced activation energy barriers for critical ORR steps in this multimetallic alloy: the energy barriers for the first step (O2 → HOO) and third step (O → *HO) are merely 0.11 eV and 0 eV, respectively. These values demonstrate superior kinetic characteristics compared to FeCoNi (0.45 eV, 0.20 eV) and FeCoNiMo (0.30 eV, 0.06 eV) systems. The theoretical predictions align well with experimental observations, including the enhanced half-wave potential (ΔE1/2 = +78 mV) and near-ideal electron transfer number, confirming the optimized reaction pathway through the multimetallic synergy.
In another study, Jin et al. designed and synthesized ultrasmall HEA nanoclusters (∼2 nm) loaded on HEO nanowires for Zn–air batteries.161 Notably, both HEA nanoclusters and HEO nanowires can be tuned separately. This type of HEA@HEO catalyst is bifunctional, demonstrating excellent performance in both the ORR and OER. The HEO is highly active for the OER, while the HEA clusters are responsible for high ORR activity, resulting in a record-low ΔE of 0.61 V. Compared with Pt@HEO, HEA@HEO exhibited improved ORR activity with significantly less Pt usage, reducing synthesis costs and making it more feasible for large-scale applications.161 This HEA system demonstrates a remarkable activity enhancement mechanism. The precise regulation of oxygen adsorption energy (ΔEO) constitutes the core factor. ΔEO on HEA surfaces is 0.2 eV lower than that of Pt (111), with a deviation of less than 0.01 eV from the theoretical optimal adsorption energy (0.2 eV weaker than Pt). This near-ideal adsorption strength significantly accelerates the ORR kinetics. This characteristic originates from the synergistic electronic effects of multiple metallic components (Pt, Pd, Au, Ag, Ir, Ru), which reconstruct the surface electronic structure to effectively reduce oxygen adsorption barriers. Furthermore, electronic structure analysis reveals that the d-band center of Pt in HEA shifts downward to −1.818 eV, showing a notable negative displacement compared to pure Pt (111) (−1.789 eV). This multi-metallic interface-induced d-band center downshift weakens the O–Pt bonding strength, accelerates the desorption kinetics of ORR intermediates, and ultimately achieves a comprehensive improvement in catalytic efficiency.
As shown in Fig. 14a, Huang et al. proposed a simple microenvironment regulation strategy to modulate solvent polarity and nanoparticle–support interactions within precursors for carbon thermal shock pyrolysis.174 They found that reducing solvent polarity and enhancing particle–support affinity could jointly control the nanoparticle size, ultimately achieving a size of approximately 2.68 nm with a Pt loading of ∼10 wt%. This approach improves the catalytic performance and reduces the preparation cost of the catalyst. Additionally, Zou et al. synthesized self-supporting HEA-O (oxygen-doped high-entropy alloys) using a rapid Joule heating method.175 As demonstrated in Fig. 14b, the O-doped HEAs (HEA-O) exhibited exceptional performance and stability in water electrolysis and zinc–air batteries, remaining stable for over 1600 h and capable of reassembly after zinc consumption.
![]() | ||
Fig. 14 (a) PtFeCoNiMn high-entropy alloy supported on carbon via the CTS method (cited from ref. 176). (b) Schematic illustration of the synthesis process and mechanism of HEA-O (cited from ref. 177). (c) Schematic of the design of the dual-active-center HEA catalyst. (d) Schematic diagram of wide-temperature FSZABs (cited from ref. 178). |
Qiu et al. proposed a novel dual-active center alloying strategy to achieve efficient bifunctional oxygen catalysis and further employed the high-entropy effect to regulate the structure and performance of the catalyst,179 as shown in the schematic in Fig. 14c. Notably, the resulting HEA catalyst demonstrates outstanding catalytic activity for both the OER and ORR, with a peak power density of 136.53 mW cm−2 and an energy density of 987.9 mA h gZn−1, surpassing most previously reported bifunctional oxygen electrocatalysts. Furthermore, the assembled flexible rechargeable zinc–air battery (ZAB) shows an excellent performance even at an ultralow temperature of −40 °C (Fig. 14d).
As summarized in Table 4, we provide a comparison and analysis of other recent high-entropy catalysts for the ORR alongside those already described.
Materials | Synthesis method | Morphology | Size | Halfwave potential | Test conditions | Ref. |
---|---|---|---|---|---|---|
FeCoNiCuPd OHEA-mNC | In situ construct | 2D mesoporous | 10 nm | 0.9 V | Alkaline | 180 |
Pt(FeCoNiCuZn)3/C | Impregnation–reduction | Nanoparticle | 12 nm | 0.80 V | Acidic | 181 |
NMnFeCoNiCu HESA | Wet chemical | Single atom | N.A. | 0.87 V | Alkaline | 176 |
Pt4FeCoCuNi | Sulfur-anchoring | Particle | 5.1 nm | 0.943 V | Acidic | 137 |
AlCoFeMoCr/Pt | Top-down | Particle | 1–2 nm | 0.88 V | Alkaline | 177 |
PtIrFeCoCu | Chemically reduced | Particle | 6 nm | 0.894 V | Acidic | 178 |
CrMnFeCoNi | Solvothermal | Particle | 170 nm | 0.78 V | Alkaline | 182 |
FeCoNiMoW | Solvothermal | Particle | 35 ± 20 nm | 0.71 V | Alkaline | 173 |
HEA-NPs-(14) | Alloying | Particle | 5 nm | 0.86 V | Alkaline | 183 |
FeCoNiCuMn | Movable printing | N.A. | N.A. | 0.887 V | Alkaline | 184 |
Pt4FeCoCuNi | Impregnation method | Nanoparticle | 5 nm | 0.943 V | Alkaline | 137 |
PtPdAuAgCuIrRu | Alloying–dealloying | Nanoclusters | 1.5–2 nm | 0.89 V | Alkaline | 161 |
AlNiCoRuMo | Alloying–dealloying | Nanowire | ∼20–100 nm | 0.875 V | Alkaline | 185 |
6-HENs/PC | Anchoring–carbonization | Nanoparticle | 2.2 nm | 0.898 V | Alkaline | 186 |
10-HEO/C | Far-equilibrium synthesis | Nanoparticle | ∼7 nm | 0.85 V | Alkaline | 28 |
FeCoNiOx@IrPt | One-step synthesis | Nanoparticles | ∼5 nm | 0.83 V | Alkaline | 187 |
PtPdRhRuIrFeCoNi MMNCs | Thermal shock | Nanoparticles | N.A. | 0.85 V | Alkaline | 188 |
FeCoNiMoW | Colloidal-based approach | Nanoparticles | 35 ± 20 nm | 0.71 V | Alkaline | 173 |
PtPdFeCoNi | High-temperature injection | Nanoparticles | 12 ± 4 nm | 0.92 V | Alkaline | 82 |
HEA@Pt | In situ growth | Nanoparticles | 23 nm | 0.85 V | Alkaline | 189 |
FeCoNiMnCrO | Electric dipole transition | Nanoparticles | N.A. | 0.87 V | Alkaline | 190 |
Li-HEO | Ball milling | Particle | 2 μm | N.A. | N.A. | 191 |
(Pr0.25La0.25Nd0.25Ca0.25)2NiO4+δ | Sol–gel | Particle | N.A. | N.A. | N.A. | 192 |
Ru@HEPO | Sol–gel | Particle | 100 nm | N.A. | N.A. | 193 |
Cha et al. employed a xenon lamp flash-induced photothermal shock method to rapidly synthesize multielement HEA-NPs on CNF supports.195 By subjecting the CNF surface to instantaneous heating followed by ultrafast cooling, PtIrFeNiCoCe hexanary solid-solution NPs were synthesized, effectively suppressing phase separation governed by conventional thermodynamics. The millisecond-level non-equilibrium process enabled large-area uniform coating (6 × 6 cm2), circumventing alloying theory limitations. Synergy between high-entropy effects and ultrafast kinetics facilitated atomic-scale homogenization: Pt/Ir provided efficient HER active sites, Fe/Ni/Co optimized *OOH intermediate adsorption for the OER, and Ce acted as an electronic modulator to inhibit metal dissolution. The CNF support enhanced electron transfer and mass diffusion through its high conductivity and photothermal-induced surface defects, while maximizing active site exposure. A symmetric electrolyzer achieved overall water splitting in 0.1 M KOH with a total overpotential of 777 mV@10 mA cm−2, surpassing most reported HEA catalysts. Long-term stability tests revealed >95% performance retention after 5000 cycles at 10 mA cm−2, with SEM/EDS confirming negligible NP aggregation or elemental leaching and hypochlorite byproduct concentration of <0.03 ppm. Theoretically, this work demonstrates that the interplay between ultrafast heating/cooling and high entropy is pivotal for phase-separation suppression and atomic-level mixing, establishing a novel paradigm for designing efficient and stable nano-catalytic systems.
Park et al. synthesized AuRuIrPdPt high-entropy alloys (HEAs) with floral and hairy spherical morphologies by irradiating metal salt solutions for 90 s using a continuous-wave CO2 laser at 30%, 60%, and 90% of 25 W power.197 Raman spectroscopy revealed the emergence of Pt–O, Pd–O, and Ru–H bond signatures under applied potential, indicating that Pt and Pd facilitated water dissociation, while Ru served as the primary active site for H* adsorption/desorption. The synergistic interplay between Ru, Pd, and Pt significantly reduced the energy barrier for the HER Volmer step. The optimized HEA-60 catalyst exhibited exceptional HER activity with overpotentials of 37 mV, 34 mV, and 45 mV at 10 mA cm−2 in alkaline, simulated seawater (1 M KOH + 0.5 M NaCl), and natural seawater electrolytes, respectively, surpassing commercial Pt/C (52 mV@alkaline). An overall water-splitting (OWS) electrolyzer assembled with HEA-60 as the cathode and IrO2 as the anode achieved a cell voltage of only 1.62 V at 10 mA cm−2 in natural seawater, maintaining 80% faradaic efficiency after 60 min of operation. The electrolyzer demonstrated remarkable durability with negligible voltage decay and minimal hypochlorite byproduct generation (<0.05 ppm), highlighting its corrosion-resistant design. Besides, Xie et al. synthesized FeNiCoCrRu high-entropy alloy nanoparticles (HEA NPs) in situ on carbon paper via CO2 laser-induced rapid conversion of metal precursors,198 forming a single-phase solid-solution structure with pronounced lattice distortion and electron transfer, which collectively enhanced the catalytic performance. In alkaline seawater electrolyte (1 M KOH + 0.5 M NaCl), the material exhibited exceptional bifunctional activity, requiring overpotentials of only 52 mV for the hydrogen evolution reaction (HER) and 320 mV for the oxygen evolution reaction (OER) at 100 mA cm−2. The assembled FeNiCoCrRu||FeNiCoCrRu electrolyzer achieved overall seawater splitting voltages of 1.594 V, 1.683 V, and 1.808 V at current densities of 10, 50, and 100 mA cm−2, respectively, outperforming the conventional Pt/C||RuO2 system (1.746 V@100 mA cm−2). Notably, faradaic efficiencies reached 99.6% for H2 and 97.7% for O2. The electrolyzer demonstrated remarkable durability, with a minimal voltage increase of 0.153 V after 3050 h of continuous operation at 250 mA cm−2, accompanied by negligible hypochlorite generation (<0.1 ppm) and metal dissolution (<0.5 μg L−1), highlighting its superior corrosion resistance. Electronic structure analysis revealed a downward shift of the Ru 3p binding energy, indicative of electron transfer from 3d transition metals (Fe, Co, Ni, Cr) to Ru atoms, which optimized the adsorption free energy of hydrogen/oxygen intermediates and accelerated the reaction kinetics. Furthermore, the electrolyzer maintained its stable operation at 100 mA cm−2 under varying electrolyte concentrations and temperatures, underscoring its broad applicability.
To fully realize the potential of HEMs in electrocatalysis and other fields, integrating research on these materials with emerging technologies presents both challenges and opportunities. By capitalizing on advancements in artificial intelligence, advanced characterization methods, nanotechnology, additive manufacturing, and sustainable practices, the innovation and performance of high-entropy catalysts can be significantly accelerated. The following perspectives outline proposed strategies for advancing the development of HEMs in catalysis, offering guidance and direction for future research and development endeavors.
Afterall, utilize computational simulations, such as DFT, to pre-screen and identify optimal combinations of reactants in order to forecast their behavior in specific reactions. According to experimental data, the composition ratio is constantly adjusted to find the optimal performance combination. This task is complicated by the vast compositional space available for HEMs, making it difficult to predict which combinations will perform best. To address this challenge, integrating computational tools, such as machine learning and simulation-based screening methods, can significantly enhance the efficiency of material selection. In particular, DFT and other simulation techniques can be used to pre-screen compositions and predict the behavior of different material combinations in specific catalytic reactions.
In addition to their potential as standalone electrocatalysts, HEMs can be integrated with emerging technologies, such as nanotechnology, artificial intelligence, and advanced manufacturing techniques, to further enhance their performance and broaden their application scope. The synergy between these complementary fields presents both challenges and exciting opportunities. By leveraging advancements in these areas, it is possible to accelerate the development and application of high-entropy catalysts in various sectors, including energy conversion and storage.
Despite some challenges remaining unresolved, it is undeniable that the research space and application prospects for HEMs are burgeoning. Collective efforts from scientific researchers worldwide have propelled rapid advancements in this field, paving the way for large-scale applications across diverse domains. As investigations continue to unravel the multifaceted properties and functionalities of HEMs, their significance in energy fields becomes increasingly apparent.
In conclusion, the interdisciplinary approach for HEM research, integrating theoretical modeling, experimental innovation, and technological advancements, will likely propel HEMs into a leading position in the field of electrocatalysis. With ongoing research, HEMs hold the potential to significantly advance energy conversion technologies, reduce the reliance on precious metals, and contribute to a more sustainable future in catalysis. The continued exploration of these materials promises to unlock new functionalities and improve the efficiency of various catalytic processes, making them indispensable in addressing global energy challenges.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5qi00538h |
‡ These authors contributed equally to this work. |
This journal is © the Partner Organisations 2025 |