Alexander
Li
,
Nayla
Qureshi
and
Vivek
Maheshwari
*
Department of Chemistry, The Waterloo Institute for Nanotechnology, University of Waterloo, Waterloo, Ontario, Canada. E-mail: vmaheshw@uwaterloo.ca
First published on 19th August 2025
High entropy alloys have garnered significant research interest in the field of electrochemistry in recent years due to their unique catalytic properties. These materials, characterized by a multi-principal element composition, have demonstrated superior catalytic performance and enhanced stability compared to traditional catalysts. However, the inherent complexity of high entropy alloys poses significant challenges, particularly in two key areas. First, this complexity makes studying and fully understanding the material's properties and behavior difficult. Second, the synthesis of nanoscale high entropy alloys is often complex, costly, and differs substantially from their bulk counterparts. This review begins by discussing the core principles that govern the unique characteristics of high entropy alloys. The central portion of the review focuses on the latest methodologies for synthesizing nanoscale high entropy alloys and briefly tabulates the catalytic performance of these materials. In the concluding section, we examine the recent studies on the formation mechanisms of high entropy alloy nanoparticles, with a particular focus on wet-chemistry synthesis methods conducted under mild conditions. We hope this review will help researchers better understand high entropy alloys and high entropy alloy synthesis methods for electrocatalysis.
Achieving this transition, however, is challenging as there are still several significant issues, one of which stems from electrocatalyst materials, which face several significant limitations that hinder their practical application.
1. Reliance on precious metals: A significant drawback of current high-performance catalyst materials is their heavy dependence on precious noble metals such as platinum, gold, and silver. The high cost and limited availability of these metals make them economically impractical and unsustainable for large-scale industrial applications.
2. Suboptimal performance of existing electrocatalysts: the efficiency of current electrocatalysts remains inadequate, contributing to the high cost of green hydrogen production, which ranges from $4–$6 USD per kilogram compared to gray hydrogen's more economical $0.80–$2.70 USD per kilogram.2 This disparity is primarily due to electricity costs, which account for around 90% of green hydrogen's total production expenditure.3 To address this challenge, it is crucial to develop advanced catalysts capable of achieving higher efficiencies at higher current densities. While research has showcased promising electrolysis systems with high efficiencies, their industrial counterparts lag significantly behind, highlighting a substantial gap between laboratory success and real-world application.
3. Limited stability: The durability of current catalysts proposed in research falls short of meeting the demands of industrial-scale applications. These catalysts struggle to maintain consistent performance under the challenging conditions of industrial electrolysis systems. For example, they must endure extreme operating environments, including high pH levels of 20–30 wt% KOH (4–7 M),4 varying temperatures from below 80 °C for transportation applications to over 650 °C for solid oxide electrolyzer cells and high current densities of up to 2 A cm−2.5 Developing catalysts capable of withstanding these challenges is essential for advancing industrial hydrogen production.
4. Selectivity challenges: Catalysts for CO2 reduction and ammonia synthesis frequently encounter significant selectivity challenges. In the case of CO2 reduction, selectivity is often hindered by competition with the hydrogen evolution reaction, as the CO2-to-CO conversion occurs at a potential of −0.1 V vs. standard hydrogen electrode (SHE). For ammonia synthesis, the difficulty stems from the remarkable stability of N2, requiring substantial energy input to break the triple bond. This demand for high activation energy typically necessitates elevated temperatures and large overpotentials,6 which further intensifies HER competition. As a result, achieving the desired product often requires additional purification or post-processing steps, increasing production costs.
To address these issues, there has been growing interest in developing electrocatalysts using high entropy alloys (HEAs).7–14 HEAs are alloys that typically contain five or more elements with atomic ratios falling between 5 and 35%. HEAs were initially discovered by two independent groups around the same time in 2004.15,16 A prior rationale at the time was that, according to Hume–Rothery rules, the synthesis of this class of material should not be possible. However, Yeh et al.15 found that by introducing more elements into a system, the configurational entropy (ΔSconf) increases. ΔSconf becomes ∼1.6 times the gas constant when the system has five elements at equimolar ratios. As a result, when making an alloy of five or more elements, configurational entropy can be greater than the enthalpies of compound formation, thus deterring the formation of intermetallic phases. Additionally, the high configurational entropy inherent in these materials contributes significantly to their thermodynamic stability. This entropy effect reduces the Gibbs free energy, making the resulting material inherently stable. With about 60 usable elements in the periodic table, the number of possible new, equiatomic, five-element HEAs exceeds five million. Therefore, HEAs are very interesting for catalytic applications due to the unique heterogeneous surfaces created by the diverse elemental combinations. As a result of the random distribution of elements across the surface and throughout, HEA materials possess significant lattice distortion, creating unique local electronic structures. Additionally, the potential diversity of neighboring elements introduces numerous synergistic effects, which enhance the material's ability to optimize adsorbate interactions and stabilize reaction intermediates. With the vast space of possible combinations, HEAs offer the potential of finding catalysts with specific synergistic effects tailored for improving the performance of various electrochemical processes. Due to all these effects, there has been a surge of interest in using HEAs in electrocatalysis. HEAs have already demonstrated superior stability and performance compared to traditional multielement catalysts for HER, OER, ORR, HOR, and CO2 reduction reactions. The focus of this review is on the synthesis methods for HEAs that use mild conditions, and the challenges associated with their formation and characterization. The key electrochemical parameters of HEAs for HER, OER, ORR, CO2 reduction, and ammonia synthesis are tabulated in the review to illustrate their performance.
ΔSmix = −R![]() | (1) |
In mixtures containing 3, 5, 6, 9, and 13 elements, the configurational entropy values are observed to be 1.10R, 1.61R, 1.79R, 2.20R, and 2.57R, respectively.15 Yeh and colleagues proposed that the mixing entropy in multi-component systems is substantial enough to drive the Gibbs free energy of mixing (ΔGmix) to a favorable level, effectively overcoming the Gibbs free energy associated with intermetallic formation (ΔGinter). The high entropy effect has been demonstrated in multiple studies. For instance, in an X-ray diffraction (XRD) investigation by Tsai et al.,8 researchers systematically increased the number of elements in the alloy, progressing from a binary to a septenary system, while observing phase transformations via XRD (Fig. 1).
![]() | ||
Fig. 1 The XRD patterns of a series of alloys starting from the standard low entropy to mid entropy and eventually to high entropy alloy via sequential elemental addition. All the alloys have one or two major phases that have simple structures. Reproduced under terms of the CC-BY license.8 Copyright (2014), M.-H. Tsai, J.-W. Yeh, published by Taylor & Francis. |
The results revealed that, despite the increasing complexity, the phases in quinary, senary, and septenary alloys remain unexpectedly straightforward, primarily forming two major phases with simple body-centered cubic (BCC) and face-centered cubic (FCC) structures. However, it should be noted that the septenary alloy included minor intermetallic phases that were not observable in the XRD data. This simplicity in structure challenges conventional expectations that increasing component complexity would lead to a plethora of binary and ternary compounds.
It is essential to recognize, however, that the high entropy effect alone does not guarantee the formation of a solid solution phase. Numerous studies report the emergence of intermetallics upon the addition of certain elements. Ultimately, the formation of HEAs is governed by the competition between the entropy of mixing and the enthalpy of intermetallic formation.17–21 This balance dictates whether a stable solid solution or intermetallic phases will dominate in the alloy.
![]() | ||
Scheme 1 Example of BCC lattice distortion present in high entropy alloys compared to traditional alloys. Reproduced under terms of the CC-BY license.22 Copyright (2019), W. Dong, Z. Zhou, M. Zhang, Y. Ma, P. Yu, P. K. Liaw, G. Li, published by MDPI. |
To further analyze this phenomenon, Tsai applied a seven-bond model to calculate local energy fluctuations impacting diffusion. This model revealed that the average potential energy difference between lattice sites in Co–Cr–Fe–Mn–Ni alloys is 50% higher than in Fe–Cr–Ni alloys (Fig. 2). As a result, occupation times in lower-energy sites are approximately 1.73 times greater than in higher-energy sites.23 This extended occupation contributes to the slower overall diffusion in HEAs.
![]() | ||
Fig. 2 A potential energy plot illustrating the energy variation during the migration of a Ni atom. In pure metals, the mean difference (MD) in potential energy after each migration is zero, whereas HEAs exhibit the largest energy fluctuation. Reproduced from ref. 23 Copyright (2013), with permission from Elsevier. |
Following the initial study, numerous recent investigations have sought to explore this effect in greater detail. Wang et al.24 conducted an insightful study employing HEA sandwiches with varying atomic ratios of FCC CoCrCuFeNi HEAs.
This approach allowed for the precise measurement of diffusion within individual layers and across the interfaces between HEA layers (Scheme 2). By utilizing these varied atomic ratios, concentration gradients were generated at the interfaces, facilitating the study of diffusion behavior. Their results highlighted a notable finding: sluggish diffusion was observed exclusively at the layer interfaces, while diffusion within the layers themselves did not exhibit sluggish characteristics. Additional studies indicate that diffusion in HEAs does not correlate with factors like configurational entropy or potential energy fluctuations, suggesting that the sluggish diffusion characteristics may not be intrinsic to HEAs but instead emerge from complex interactions among various elements.20,21,25–27 Consequently, sluggish diffusion cannot be assumed for every HEA combination. Nevertheless, sluggish diffusion has often been cited as a key factor contributing to the enhanced stability and corrosion resistance of nanoscale HEA catalysts.28 One approach researchers have used to demonstrate sluggish diffusion is through capturing transmission electron microscopy (TEM) images of HEA nanoparticles after stability testing. These studies have shown that HEAs exhibit reduced or limited in situ surface reconstruction compared to traditional platinum-on-carbon counterparts. HEAs have also demonstrated minimal surface reconstruction even under more challenging conditions such as increased overpotentials, high current densities, and varying electrolyte concentrations.29–38 However, a more thorough investigation is required to determine whether the sluggish diffusion effect is the cause of reduced surface reconstruction in HEA-based electrocatalysts.
![]() | ||
Scheme 2 Prepared HEA sandwiches, with varying HEA compositions at various locations within the structure. Reproduced from ref. 24 Copyright (2018), with permission from Elsevier. |
1. ΔSmix should be maximized, and typically, HEAs are classified as alloys with ΔSmix greater than or equal to 1.5R. Alloys with entropies below 1.5R and greater than 1R are typically considered mid-entropy alloys. However, it should be noted that HEAs are not materials with maximized entropy, as metallic glasses and high-entropy metallic glasses can exhibit even greater entropy. The equation for ΔSmix is listed below.
![]() | (2) |
2. Mixing enthalpy (ΔHmix) should be greater than −15 kJ mol−1 and less than 5 kJ mol−1.40 As values outside these ranges can result in intermediate phases/intermetallics. ΔHmix is given by the following equation:
![]() | (3) |
The binary mixing enthalpy is given as ΔHmixij between the ith and the jth elements. While ci and cj are the respective atomic fractions of the ith and jth elements.
3. Both earlier thermodynamic parameters of ΔHmix and ΔSmix can be combined to form Ω, a unitless parameter, denoted by the following equation:41
![]() | (4) |
![]() | (5) |
4. The atomic radii of the component elements of an HEA should be similar to form a solid solution phase. Zhang et al.,40 therefore proposed a parameter, δ, representing the average lattice mismatch of the constituent elements. The maximum lattice mismatch must be below 6.6% otherwise phase separation or intermetallic formation could occur. The following equation denotes the average lattice mismatch:
![]() | (6) |
![]() | (7) |
5. Valence electron concentration (VEC) is a valuable predictor of crystal phase and stability in HEAs.42 When VEC exceeds 8, the formation of an FCC phase is likely. For VEC values between 6.87 and 8, a mixed BCC-FCC phase may be observed. Finally, if VEC falls below 6.87, BCC phase is more likely to form. However, it should be noted that certain HEA compositions containing Mn deviate from the expected trends predicted by VEC calculations.42
![]() | (8) |
6. The constituent elements in an HEA composition should have similar electronegativities.43,44 Significant electronegativity differences can promote the formation of intermetallic compounds, destabilizing the phase and inhibiting the formation of a single solid-solution phase. While no universally accepted threshold exists for determining solid-solution formation, minimizing electronegativity differences χ is generally recommended to enhance phase stability.
![]() | (9) |
![]() | (10) |
7. Ye et al.45 proposed the parameter ϕ to account for the effects of dense atomic packing and atomic size mismatch, which influence the enthalpy of formation and can lead to excessive ΔSmix. They identified a critical ϕ value of 20 as the boundary between single-phase and multi-phase HEA formation: Values below 20 generally yield multi-phase HEAs, whereas values above 20 typically yield single-phase HEAs.
![]() | (11) |
![]() | (12) |
8. Λ, a geometric parameter proposed by Singh et al.,47 combines thermodynamic and geometric considerations to predict phase formation in HEAs. Λ values over 0.96 predict single-phase HEA formation, values between 0.96 and 0.24 predict multi-phase formation, and values below 0.24 predict the formation of intermetallic compounds.
![]() | (13) |
9. Wang et al.48 proposed a parameter, γ, to more accurately account for atomic packing misfit compared to conventional atomic size mismatch calculations. Their analysis showed that the primary contributions to atomic packing misfit can be effectively simplified by considering the solid angles of atomic packing for the largest and smallest constituent atoms. A critical value of 1.175 was identified, where values below this threshold favored the formation of stable solid-solution phases. In contrast, values above 1.175 tended to produce intermetallic compounds or metallic glasses.
![]() | (14) |
![]() | (15) |
![]() | (16) |
The above-mentioned empirical rules serve as a general guideline for HEA composition selection. However, it should be noted that simply following the rules mentioned above will not guarantee HEA formation for a given elemental composition. Additionally, there is still much debate regarding the definition and classification of HEAs. Therefore, the empirical guidelines for HEA formation are still undergoing ongoing research, which can be read in more detail in the following review papers.7,39,49,50
Hume–Rothery VEC calculations alone are insufficient to fully predict the final crystal phases of HEAs. Numerous studies have sought to improve phase prediction by incorporating additional thermodynamic and computational approaches. For example, Caraballo et al.51 investigated the role of lattice strain, analyzing both interatomic spacing mismatch (sm) and bulk modulus mismatch (km). Their method not only enabled more accurate predictions of HEA phase formation but also successfully distinguished between FCC and BCC structures. A single solid-solution phase formation is predicted to occur when sm < 1 and km < 8. More importantly, they identified a critical km of 4. Where km < 4 predicts FCC formation, while 4 < km < 8 predicts BCC formation.
Another notable extension of Hume–Rothery approach was presented by Poletti et al.,52 who combined itinerant electron concentration with VEC to refine phase predictions. Their analysis revealed that, consistent with previous observations, lower VEC values generally correspond to BCC formation, while higher VEC values favor FCC. Furthermore, they found that high itinerant electron concentrations, particularly when coupled with low VEC, also promote BCC formation (Fig. 3).
![]() | ||
Fig. 3 VEC and itinerant electron (e/a) plot of various multielement systems. Highlighted correlation of lower VEC and high e/a predicts BCC phase formation. Reproduced from ref. 52 Copyright (2014), with permission from Elsevier. |
Aside from general phase predictions, active research has been directed towards the development of phase diagrams for HEAs. Like conventional metals and alloys, HEAs can undergo crystal phase transitions or form mixed-phase structures through composition tuning, mechanical stress, or temperature variations during synthesis and annealing. Their behavior often parallels that of traditional alloy systems described by phase diagrams. However, the high compositional complexity of HEAs means that comprehensive phase diagrams are still under development.53–58
Like traditional metals and alloys, HEAs have also demonstrated polymorphism. Notably, Zhang et al.59 were the first to report polymorphism in HEAs by applying hydrostatic compression to the FCC CoCrFeMnNi Cantor alloy at room temperature, inducing a phase transition from FCC to HCP. However, it should be noted that to achieve the phase transition, a pressure of ∼41Gpa was needed, which can only be achieved using a diamond anvil cell.
Temperature effects on HEA formation remains an area of active research. The temperatures used during synthesis and annealing significantly affect not only phase formation but also the resulting crystal structure and phase purity of HEAs. In addition to temperature, cooling rate has been found to play a critical role in determining crystal phase composition and HEA formation. For example, Al0.5CoCrFeNi has been reported to form either FCC + BCC or FCC + BCC + B2 phases depending on the cooling rate.60 However, careful consideration must be given to cooling conditions, as it is generally observed that rapid cooling suppresses the formation of intermetallic compounds and promotes the retention of solid solution phase.
In summary, the formation and stability of crystal phases in HEAs are not governed solely by composition. They are also heavily influenced by processing and synthesis conditions, including temperature, mechanical stress, and cooling rate. These factors can induce phase transitions, polymorphism, and the emergence of mixed-phase structures, even in systems with identical elemental compositions. As comprehensive phase diagrams for HEAs are still in development, understanding the interplay between thermodynamics and kinetics remains essential. Fine-tuning these conditions provides a powerful approach for tailoring the structural and catalytic properties of HEAs.
Early work by Lee et al.65 investigated the effect of aluminum content on the corrosion behavior of AlxCrFe1.5MnNi0.5 in 0.5 M H2SO4. In H2SO4, aluminum typically forms an Al2O3-rich passive layer. However, this layer is known to be porous and non-uniform, allowing continued ingress of H+ ions and subsequent localized attack beneath the oxide film. Tests showed the corrosion potential (ECorr) decreased, while both the corrosion current (Icorr) and the passivation current (Ipass) increased as the aluminum content increased. Further electrochemical impedance spectroscopy (EIS) tests confirmed this by showing a decrease in charge transfer resistance (Fig. 4a and b).
![]() | ||
Fig. 4 EIS results for CrFe1.5MnNi0.5 alloy in 0.5 M H2SO4 solution. Nyquist (a) and Bode (b) plots along with (c) equivalent electrical circuit. Scattered points represent raw experimental data points, and solid lines represent model fit. Reproduced from ref. 65 Copyright (2008), with permission from Elsevier. |
Interestingly, a second capacitive loop was observed during EIS measurements (Fig. 4c), highlighting the presence of an adsorbed layer and charge transfer processes at the metal–electrolyte interface. The adsorptive complexes on the interface were a result of Al dissolution by the following mechanism:
Al(s) + H2O → Al(OH)ad + H+ + e− | (17) |
Al(OH)ad + 5H2O + H+ → Al3+ + 6H2O + 2e− | (18) |
Another HEA composition, AlxCoCrFeNi, was also tested in the same environment, and comparable results were observed. Increasing the aluminum content led to decreased corrosion resistance. While the observed trends align with conventional understanding of aluminum corrosion in acidic media, it should be noted that as the aluminum content increased, the corrosion resistance decreased. The decrease in corrosion resistance is primarily attributed to increased oxide layer porosity and changes in crystal structure. Although the oxide layer becomes thicker with higher aluminum content in the HEA, it also becomes more porous, allowing greater ingress of H+ ions and thus accelerating corrosion. Additionally, the increased aluminum content induces a phase transition in the alloy, shifting from a mixed FCC/BCC structure to a predominantly BCC phase. Therefore, changes in elemental composition and the resulting changes in crystal structure must be carefully considered. Both the crystal structure and surface behavior play a critical role in determining the corrosion behavior of high entropy alloys.
Another crucial factor that can increase corrosion is galvanic corrosion at the interfaces between mixed-phase HEAs. Ren et al.66 investigated the corrosion behavior of copper additions in a CuxCrFeNiMn HEA composition tested in 1 M H2SO4.
Both Cu and Cr play significant roles in corrosion resistance. Cr forms a passivation layer, enhancing corrosion resistance, while Cu promotes the formation of single FCC phases, reducing galvanic corrosion. However, as Cu content increases, inter-dendritic (ID) FCC + BCC phases begin to form (Fig. 5a and b). This results in Cu-rich, Cr-depleted zones alongside Cr-rich, Cu-depleted zones, leading to galvanic corrosion between the ID and dendrite-rich regions (Fig. 5b).
![]() | ||
Fig. 5 Scanning electron microscope (SEM) images of CuCrFeNiMn alloy systems: (a) CuCr2Fe2Ni2Mn2, (b) Cu2CrFe2NiMn2. Reproduced from ref. 66 Copyright (2011), with permission from Wiley. |
When designing a HEA catalyst, it is essential to thoughtfully select the alloying elements, as the electrolyte conditions significantly impact stability and performance requirements. In acidic environments, such as those involving electrolytes like H2SO4, protective passivation films can degrade rapidly, leading to accelerated oxidation and metal dissolution. Conversely, in alkaline environments, using electrolytes like KOH, corrosion rates may decrease due to the formation of protective oxide layers. However, these oxide layers, while reducing corrosion, may not always enhance catalytic performance and could even hinder it, depending on the specific application. Mixed-phase HEAs, in particular, warrant closer investigation due to their susceptibility to galvanic corrosion at phase interfaces, which can compromise their corrosion stability.61,62,65 Consequently, both phase uniformity and the potential galvanic corrosion between the catalyst and the catalyst support should be carefully evaluated to optimize corrosion resistance. Additionally, passivation films formed by elements such as Cr and Al can affect catalytic performance by altering the active site area/properties, which must be considered in the design process.
For many transition metal-based catalysts, activity is closely tied to the filling of bonding and antibonding states within the d-band. Pt, a benchmark catalyst for HER, has been alloyed with various metals to adjust its lattice parameters. When alloyed with elements of varying atomic sizes, compressive or tensile strain can be introduced into the platinum lattice. This strain shifts the d-band center, improving adsorption and desorption properties and thus enhancing catalytic performance. However, applying these same principles to HEAs introduces significant complexity. In these multicomponent systems, lattice parameters can often be approximated using Vegard's law, which estimates the effective lattice constant as a weighted average of the constituent elements. Maruta et al.67 demonstrated that lattice parameters predicted by Vegard's law show good linear correlation with experimentally measured values across a range of HEA compositions (Fig. 6).
![]() | ||
Fig. 6 Plot of calculated and measured lattice constants as well as d-band centres. Linear correlation is observed between calculated and measured lattice constants. Reproduced from ref. 67 Copyright (2022), with permission from RSC publishing. |
In addition to structural measurements, Maruta et al. also investigated the d-band electronic structure of various HEAs. Their results revealed a lack of correlation between theoretically calculated and experimentally measured d-band centers (Fig. 6). More importantly, the position of the d-band center did not correlate with HER activity in HEAs (Fig. 7a and b). This suggests that factors beyond bulk electronic structure, such as local atomic environments, orbital hybridization, or synergistic (“cocktail”) effects, may play a more dominant role in determining catalytic behavior. Furthermore, recent studies challenge the assumption that HEAs are entirely random solid solutions. Evidence now shows that HEAs can exhibit both long-range disorder and short-range chemical ordering, which can significantly influence their structural stability and catalytic performance.68
![]() | ||
Fig. 7 (a) Overpotential of various HEA composition at 50 mA cm−2. (b) Turn over frequency compared with d-band centre. Reproduced from ref. 67 Copyright (2022), with permission from RSC publishing. |
Crystal phase control has been extensively utilized as a strategy to enhance the catalytic and structural properties of metal and metal alloy-based catalysts. Phase transitions alter atomic packing and surface site symmetry, which in turn can dramatically affect electronic structure. In some cases, this enables catalysis in materials previously considered inert or drastically improves catalytic performance. For instance, Liu et al.69 synthesized a metastable HCP layer of Pt on nickel nanocrystals, resulting in a significant enhancement in mass activity for alkaline HER. Despite the potential performance advantages, the role of the crystal phase in HEA-based electrocatalysts remains largely underexplored. Studies have reported enhanced catalytic activity in both single-phase and dual-phase HEAs.70,71 However, achieving controlled phase transitions in HEAs often requires compositional tuning, making it challenging to isolate the effect of crystal structure on catalytic performance. As a result, more systematic investigations are needed to clarify the influence of the crystal phase on electrocatalytic behavior in HEAs.
Another powerful strategy to enhance catalytic performance is facet engineering. For the HER, Pt has been widely studied, with different crystal facets (e.g., Pt(111), Pt(100)) known to exhibit distinct hydrogen binding energies.72 More recently, edge sites have been identified as even more catalytically active, prompting efforts to design nanostructures that expose these high-index sites to maximize activity and Pt utilization.73 In contrast, most HEA catalysts synthesized to date are limited to simple spherical morphologies, with few methods available to fabricate nanostructures with well-defined or complex geometries.74–76 Recently, Hsiao et al.76 synthesized HEAs with different facets by utilizing cubic and octahedral Pd seeds with {100} facets and {111} facets, respectively. It was found that with the same HEA composition, the {100} facets showed significantly better HER and ethanol oxidation reaction (EOR) performance compared to both commercial pt as well as the {111} facets. Highlighting the fact that more catalytic performance can be obtained in HEAs through facet Engineering. Overall, these findings highlight a critical need for the development of advanced synthesis techniques capable of producing HEAs with tailored nanostructures and exposed active facets.
In conclusion, while crystal lattice engineering has shown great promise in enhancing the catalytic performance of conventional metals and alloys, its full potential in HEAs remains largely untapped. The inherent complexity of HEAs arising from their multi-element compositions and structural variability poses significant challenges for isolating and understanding lattice-driven effects. To move the field forward, more systematic and targeted studies are urgently needed to unravel the interplay between lattice parameters, crystal phase, and surface structure. Advancing our understanding of these areas will be critical for designing the next generation of HEA-based electrocatalysts.
In conventional transition-metal catalysts, DFT calculations often rely on d-band theory to optimize adsorbate bonding. Experimentally, however, HEAs exhibit electronic structures that differ dramatically from those of conventional alloys. HEAs display a widening of the d-band center rather than a simple shift. Hence, methods for traditional alloys are not applicable for the prediction of HEA properties.
The exceptional catalytic performance of HEAs stems from their diverse local density of states and the variety of coordination sites available on their surfaces. These unique properties enable more optimized binding sites and foster interactions between neighboring atoms, which can stabilize reaction intermediates. This intricate synergy is typically associated with the so-called “cocktail effect” frequently referenced in academic papers and reviews. This intricacy also complicates the DFT based calculations for HEAs. However, finding new parameters for predicting the performance of HEAs and methods to effectively model their behavior will be crucial for screening and selection of HEAs for specific electrochemical applications.
The CTS method is performed in two main steps. First, metal salt precursors are mixed and loaded onto conductive carbon supports, such as carbon nanofibers (CNFs) (Fig. 8a). The material is subjected to a brief but intense thermal shock via Joule heating, achieved by passing a large electrical current through the support material (Fig. 8b). This process rapidly decomposes the metal salt precursors and drives the fusion of metallic elements into single-phase nanoparticles (Fig. 8c). Using this method, PtCoNiFeCu, PtPdCoNiFe, PtCoNiFeCuAu, PtPdCoNiFeCuAu, and PtPdCoNiFeCuAu HEA nanoparticles were synthesized, all displaying uniform mixing and single-phase solid solution (Fig. 8d).
![]() | ||
Fig. 8 (a) On left, microscopy images of microsized precursor salt particles on the carbon nanofiber (CNF) support before thermal shock. On right, TEM images of salt loaded CNF after thermal shock resulting in well-dispersed (PtNi) nanoparticles. (b) On left, schematic of the set-up and on right, temperature change during the 55 ms thermal shock. The sample image during the shock is shown in the inset. (c) On left, low-magnification and single-particle elemental maps of PtNi CNF sample. On right, high-angle annular dark-field (HAADF) image, and corresponding atomic maps for a binary PtNi alloy. (d) Individual elemental maps of PtPdNiCoFeAuCuSn HEA-NP, scale bar, 10 nm. Adapted from ref. 77 Copyright (2018), with permission from the American Association for the Advancement of Science. |
Key parameters influencing the composition and size distribution of the nanoparticles include the defect concentration of the carbon nanofiber (CNF) support, shock duration, and cooling rates. In Yao et al.'s77 study, the CNF supports were prepared by carbonizing electrospun polyacrylonitrile fibers. Crucially, the carbonization temperature significantly affected the defect concentration in the supports, which in turn influenced particle dispersion. Higher carbonization temperatures (e.g., 1273 K) resulted in lower defect concentrations, leading to smaller and more uniform particle distributions. Additionally, shorter shock durations, such as those below 55 ms, produced smaller nanoparticles with tighter size distributions (Fig. 9a and b). The cooling rate is also a key factor to keep in mind for this synthesis method. At slower cooling rates, phase separation becomes evident, disrupting the uniformity of the nanoparticles (Fig. 9c and d). Rapid cooling is, therefore, essential to retaining the high-entropy mixing state and achieving single-phase solid-solution nanoparticles from this method (Fig. 9e). Catalytic testing revealed that HEA-NPs synthesized using this method, particularly quinary PtPdRhRuCe nanoparticles, demonstrated exceptional catalytic activity for ammonia oxidation, achieving ∼100% ammonia conversion and >99% nitrogen oxide selectivity at relatively low operating temperatures (700 °C). Operation stability was evaluated for 30 hours of continuous operations, and no signs of degradation were observed.
![]() | ||
Fig. 9 (a) TEM images (scale bars: 100 nm) of PtNi NPs on CNFs, illustrating the effect of thermal shock duration (5 ms, 55 ms, 1 s, and 10 s) on particle size and dispersity, accompanied by the corresponding size distribution graph (b). (c and d) Elemental maps, HAADF, and annular bright-field (ABF) images (scale bar: 10 nm) of AuNi NPs, highlighting the impact of varying cooling rates. (e) Time–temperature-transformation (TTT) diagram depicting the influence of cooling rate on the kinetic formation of metallic glass, HEA, and phase-separated structures. Adapted from ref. 77 Copyright (2018), with permission from the American Association for the Advancement of Science. |
Despite its remarkable capabilities, the CTS method has certain limitations, primarily its restriction to synthesizing nanoparticle morphologies and its reliance on conductive support materials. This reliance on supports capable of withstanding extreme temperatures and electrical currents without degrading significantly narrows the range of materials that can be used. Consequently, this limitation restricts the diversity of support–catalyst systems that can be explored, hindering broader investigations into support–catalyst interactions. As an additional consideration, the synthesis of HEAs on carbon support must consider the possibility of carbide formation. Yao et al.77 did not report any carbide formation. Nonetheless, one of the key advantages of CTS is its simplicity and scalability, offering exceptionally high heating and cooling rates, which make it a cost-effective and efficient method for HEA NP synthesis.
Maulana et al.28 utilized microwave heating to synthesize IrFeCoNiCu HEA nanoparticles for acidic OER in 0.1 M HClO4. Metal salt precursors were drop-cast onto carbon paper and subjected to microwave heating for 10–11 seconds at 1200 W. The resulting nanoparticles exhibited a size distribution ranging from 20 nm to 200 nm, with an average size between 20 nm and 50 nm. This broad distribution was primarily attributed to the heterogeneity of the carbon paper substrate and the batch-to-batch inconsistencies due to the household microwave oven used in the process.
Despite these limitations, the synthesized IrFeCoNiCu HEA nanoparticles demonstrated remarkable catalytic performance for OER. The overpotential to achieve 10 mA cm−2 was approximately 302 mV, accompanied by a low Tafel slope of 58 mV dec−1. Furthermore, the catalyst's stability was notable, with only a 60 mV increase in overpotential after 12 hours of continuous operation at 10 mA cm−2.
Achieving a narrow size distribution requires careful consideration of the interaction between the microwave source and the substrate. Additionally, similar to CTS, the range of substrates compatible with microwave heating is limited, restricting the scope of catalyst studies that can be conducted using this synthesis method. Furthermore, the use of carbon-based substrates introduces the possibility of carbide formation, which must be carefully considered. However, in the highlighted study, no mention of carbide formation was made.
The synthesis process involves dissolving precursor metal salts in ethylene glycol, followed by the addition of X-72 carbon support, a material commonly used with commercially available Pt/C catalysts. The HEA precursor salts were mixed with 10 wt% X-72, calculated relative to the final metallic weight, and then subjected to ultrasonication at 750 watts and 20 kHz under ambient conditions for 10 minutes. Post-sonication, the nanoparticles display a twin-phase structure, with both single and dual-phase FCC peaks observed in XRD analysis.
To achieve a single-phase HEA structure, calcination under N2 is required. When the calcination temperature is increased from 500 °C to 700 °C, the dual FCC peaks observed in the XRD patterns gradually disappear, resulting in a stable, single-phase alloy structure (Fig. 10a–c). Calcination temperature did not affect the morphology or size distribution of the nanoparticles. PtAuPdRhRu nanoparticles demonstrated excellent performance for HER in 1 M KOH conditions, with an overpotential of 190 mV to reach a current density of 30 mA cm−2 and a Tafel slope of 62 mV dec−1. The catalyst maintained stability during continuous operation at 100 mA cm−2 for 8 hours.
![]() | ||
Fig. 10 XRD patterns of synthesized HEA-NPs/carbon prepared via the ultrasonication-assisted wet chemistry method: (a) before calcination under N2, (b) after calcination under N2 at 500 °C for 2 hours, and (c) at 700 °C for 2 hours. Additionally, XRD patterns of (d) PtAuPdRh/carbon and (e) PtAuPd/carbon, both synthesized using the same method and calcined under N2 at 700 °C for 2 hours. Reproduced from ref. 78 Copyright (2022), with permission from Wiley. |
The method, while innovative, faces some limitations that could be addressed for broader adoption. Access to high-power ultrasonication equipment is a challenge, as such devices are not widely available in many research laboratories. Furthermore, the current approach is not optimized for large-scale applications, making scalability a critical area for improvement.
![]() | ||
Fig. 11 (a) FMBP experimental setup. (b) HEA-NP synthesis via FMBP for homogeneous structures and fixed-bed pyrolysis (FBP) for phase-separated structures. (c) Simulation of the time required for precursors/GO (20 mg, 3 wt%) to reach 923 K during the FMBP process, with the metal precursors/GO positioned in a quartz boat at the center. (d) (Scale bar: 10 nm) HAADF-STEM images of denary (MnCoNiCuRhPdSnIrPtAu) HEA-NPs on GO synthesized via FMBP with a 3 wt% loading on GO. (e) (Scale bar: 0.5 nm) HR-STEM image of denary HEA-NPs, with Fourier transform analysis confirming an FCC crystal structure. (f) (Scale bar: 10 nm) Elemental maps of denary HEA-NPs (10 wt% loading on GO), showing an equal atomic ratio of elements. Reproduced under terms of the CC-BY license.79 Copyright (2020), S. Gao, S. Hao, Z. Huang, Y. Yuan, S. Han, L. Lei, X. Zhang, R. Shahbazian-Yassar, J. Lu, published by Springer Nature. |
The versatility of FMBP was demonstrated through the synthesis of various HEA combinations, including a denary (10-element) HEA comprising Mn, Co, Ni, Cu, Rh, Pd, Sn, Ir, Pt, and Au on graphene oxide (Fig. 11d–f). In addition, the method was tested using several support materials, synthesizing CuPdSnPtAu HEA NPs on Al2O3, zeolite, carbon black, and graphene oxide.
Temperature effects were also investigated, comparing traditional furnace pyrolysis to FMBP for non-HEA NiPdPt nanoparticles. In traditional pyrolysis, samples required approximately 31 minutes to reach 923 K from room temperature, with a heating rate of 20 K min−1. In contrast, FMBP achieved 923 K within 5 seconds. The rapid heating rate of FMBP had a significant effect on phase separation. Traditional furnace pyrolysis yielded phase-separated NiPdPt NPs with a Pd-rich core, whereas FMBP avoided such separation, leading to more uniform nanoparticles (Fig. 11b).
The applicability of the FMBP synthesis method was tested using FeCoPdIrPt HEA NPs as a catalyst for HER. The resulting catalyst demonstrated excellent HER performance, achieving a low overpotential of 42 mV to reach a current density of 10 mA cm−2 and a low Tafel slope of 58 mV dec−1. Moreover, the catalyst exhibited outstanding stability, with negligible degradation after 150 hours of continuous operation at 10 mA cm−2.
An excellent demonstration of solvothermal synthesis in HEA catalyst preparation is provided by a study conducted by Chen et al.,74 where convex cube-shaped Pt34Fe5Ni20Cu31Mo9Ru HEA NPs were synthesized (Fig. 12). Using a one-pot solvothermal method, the researchers combined cetyltrimethylammonium chloride (CTAC), oleylamine, metal acetylacetonates, carbon nanotubes (CNTs), and glucose (Scheme 3). The mixture was heated under stirring at 220 °C for varying reaction durations. The study revealed fascinating insights into the morphology of the nanoparticles: those synthesized with a 1-hour reaction time exhibited a cubic shape (Fig. 12k and i), whereas extending the reaction to 2 hours produced convex cubes. The convex cube nanoparticles, characterized by high-index facets (Fig. 12a–c), demonstrated superior catalytic performance for both HER and OER. Specifically, these nanoparticles achieved remarkably low overpotentials of 10 mV and 259 mV to reach a current density of 10 mA cm−2 for HER and OER, respectively, in alkaline conditions (1 M KOH). Furthermore, they exhibited exceptionally low Tafel slopes of 27 mV dec−1 for HER and 39 mV dec−1 for OER, indicating highly efficient kinetics. The stability of the catalysts was equally impressive, as chronoamperometric testing showed only a 5% decrease in HER current and an 8% decrease in OER current after 40 hours of continuous operation.
![]() | ||
Fig. 12 TEM image of convex cube-shaped Pt34Fe5Ni20Cu31Mo9Ru-2 h at (a) low and (b) high magnification. (c) Fourier transformed electron diffraction pattern of (b). (d) HAADF-TEM image of Pt34Fe5Ni20Cu31Mo9Ru and (e–j) the corresponding elemental mappings. TEM image of Pt34Fe5Ni20Cu31Mo9Ru-1 h at a (k) low and (l) high magnification. (m) Fourier transformed electron diffraction pattern for (l). Adapted from ref. 74 Copyright (2022), with permission from Wiley. |
![]() | ||
Scheme 3 Schematic of the synthesis of convex cube-shaped Pt34Fe5Ni20Cu31Mo9 HEA on CNT-support. Reproduced from ref. 74 Copyright (2022), with permission from Wiley. |
In summary, solvothermal synthesis offers a simple and accessible route to producing HEA NPs with diverse morphologies, while maintaining compatibility with a variety of conductive supports. However, a notable challenge associated with this method lies in the use of high-boiling-point organic solvents. Residual solvents or organic capping ligands can be difficult to remove and may significantly impact the active surface area of the nanoparticles, thereby affecting their performance in catalytic applications.
A notable application of hydrothermal synthesis was demonstrated by Kang et al.,75 who developed mesoporous PtPdRhRuCu HEA nanospheres for HER catalysis. The synthesis employed self-assembled micelles of the diblock copolymer PEO-b-PMMA, which served as a structural framework for HEA nanoparticle formation (Fig. 13a). The resulting nanospheres had an average size of approximately 128 nm (Fig. 13), with uniform mesopores of around 23 nm in diameter (Fig. 14c). Local variations in the elemental mixing were observed around the pores (Fig. 14a and c). The choice of surfactant and reducing agent significantly influenced the final morphology of the nanoparticles. For example, the surfactants F127 and PVP were compared, with F127 producing reduced mesoporosity and PVP leading to irregular nanostructures. Similarly, the reducing agent played a crucial role: strong reducing agents like dimethylamine borane caused significant nanoparticle agglomeration, while (milder) formic acid resulted in irregular pore structures.
![]() | ||
Fig. 13 (a) Synthesis schematic of PtPdRhRuCu Mesoporous multimetallic nanospheres (MMNs). (b) SEM and (c) HAADF–STEM images of PtPdRhRuCu MMNs (scale bar: 100 nm), (c) inset: the corresponding selected area electron diffraction (SAED) pattern (scale bar: 5 nm−1). (d) HRTEM (scale bar: 2 nm) with inset showing Fast Fourier Transform (FFT) patterns. (e) Atomic elemental maps (scale bar: 50 nm), (f) line-scan, (g) normalized atomic compositional profile, and (h) powder XRD of PtPdRhRuCu MMNs. (h) PtPdRhRuCu MMNs are compared to FCC Pt. Reproduced under terms of the CC-BY license.75 Copyright (2023), Y. Kang, O. Cretu, J. Kikkawa, K. Kimoto, H. Nara, A. S. Nugraha, H. Kawamoto, M. Eguchi, T. Liao, Z. Sun, T. Asahi, Y. Yamauchi, published by Springer Nature. |
![]() | ||
Fig. 14 (a) PtPdRhRuCu MMN edge pore HAADF–STEM image and Energy-Dispersive X-ray Spectroscopy (EDS) maps (scale bar: 10 nm). (b) ΔSmix value at the selected area found in (a) (1–4). (c) Elemental distribution of mesopore in PtPdRhRuCu MMN. Reproduced under terms of the CC-BY license.75 Copyright (2023), Y. Kang, O. Cretu, J. Kikkawa, K. Kimoto, H. Nara, A. S. Nugraha, H. Kawamoto, M. Eguchi, T. Liao, Z. Sun, T. Asahi, Y. Yamauchi, published by Springer Nature. |
Temperature variations also affected the synthesis outcomes, particularly influencing the Ru content in the nanoparticles. The chain length of the block copolymer PEO-b-PMMA was another critical factor, specifically the PMMA chain, which acted as a sacrificial template to form the mesopores. When the PEO chain length was held constant, varying the PMMA chain length showed a clear correlation with pore size in the resulting mesoporous structures. For example, PEO10000-b-PMMA5500 produced the smallest pore sizes, around 8 nm, while PEO10500-b-PMMA22000 resulted in significantly larger pores, measuring up to 41 nm. After extensive testing, PEO10500-b-PMMA18000, which formed pores approximately 23 nm in size, delivered the best performance for HER under alkaline conditions (1 M KOH). This composition achieved impressively low overpotentials of 10 mV to reach a current density of 10 mA cm−2 and exhibited a Tafel slope of 87 mV dec−1. Stability tests revealed that PtPdRhRuCu nanospheres maintained stable operation for 140 hours across varying current densities from 10 to 100 mA cm−2.
The challenges associated with hydrothermal synthesis are comparable to those of solvothermal methods, but certain issues are more pronounced. In particular, nanoparticle agglomeration is more significant when water is used as the solvent, necessitating the use of stabilizing or capping agents to maintain particle dispersion and stability. Additionally, the milder reaction conditions characteristic of hydrothermal synthesis amplify the influence of both the reducing agents and the stabilizing agents on the final morphology and properties of the nanoparticles. These factors must be carefully optimized to achieve the desired catalytic performance and structural stability.
Composition | Electrolyte | Overpotential at 10mA cm−2 | Tafel slope (mV dec−1) | Synthesis method | Stability | Ref. |
---|---|---|---|---|---|---|
a Overall water splitting. | ||||||
TiNbTaCrMo | Seawater | 960 mV at 50 mA cm−2 | 96.33 | Arc-discharge | Stable for 10 hours at 863 mV overpotential | 80 |
Ni20Fe20Mo10Co35Cr15 | 1 M KOH | 172 mV | 66 | Arc-melting | Stable for 8 hours at 100 mA cm−2 0.5 M H2SO4 | 71 |
RhRuPtPdIr | 0.5 M H2SO4 | 58 mV | 42 | Atomic layer deposition | 7.5% current density decrease after 20 hours | 81 |
FeCoNiMnCr | 1 M KOH | 190 mV | 64 | Single-step pyrolysis | 15% current density decrease after 6 hours | 30 |
FeCoNiCuMn | 1 M KOH | 115 mV | 94.2 | CTS | Stable for 50 hours at 50 mA cm−2 | 82 |
PtNiCoFeCu | 1 M KOH | 7 mV | 58 | CTS | Stable for 500 h maintaining 100 mA cm−2 | 83 |
PtIrFeNiCoCe | 0.1 M KOH | 193 mV | 91.1 | CTS | Stable for over 5000 cycles | 84 |
AlNiCoIrMo | 0.5 M H2SO4 | 275 mV | 33.2 | De-alloying | Stable for 48 hours at 1.52 V![]() |
29 |
AlAgAuCoCuFeIrNiPdPtRhRuMoTi | 0.5 M H2SO4 | 32 mV | 30.1 | De-alloying | Stable for 10 hours at 10 mA cm−2 | 85 |
AlAgAuCoCuFeIrNiPdPtRhRu | 0.5 M H2SO4 | 42 mV | 38.3 | De-alloying | N/A | 85 |
AlNiCoRuMo | 1 M KOH | 24.5 mV | 30.3 | De-alloying | N/A | 86 |
PtPdRhIrNi | 1 M KOH | 55 mV | 44.8 | De-alloying | 21 mV increase after 24 hours | 87 |
ZnNiCoIrMn | 0.1 M HClO4 | 50 mV at 50 mA cm−2 | 30.6 | De-alloying | 16.8 mV increase after 100 hours at 10 mA cm−2 | 88 |
CuCrFeNiCoP | 1 M KOH | 365 mV at 100 mA cm−2 | 118 | Electrodeposition | Stable for 48 hours at 100 mA cm−2![]() |
89 |
FeCoNiCuMn | 1 M KOH | 281 mV at 100 mA cm−2 | 53 | Electrospun nanofiber | Stable over 20 hours at 170 mA cm−2 | 90 |
PtCoCuFeMnNi | 0.5 M H2SO4 | 70 mV | 47 | Electrodeposition | Stable for 27 hours at 10 mA cm−2 | 91 |
CoFeLaNiPt | 0.1 M KOH | 555 mV | N/A | Electrodeposition | Stable for 10 hours at 377 mV overpotential | 92 |
FeCoPdIrPt | 1 M KOH | 42 mV | 82 | Fast-moving bed pyrolysis | Stable for 150 hours at 10 mA cm−2 | 79 |
PtZrNbFeCuTaMoHfBiWZnSnPdNi | 1 M KOH | 18 mV | 30.7 | Hydrothermal | 3 mV decrease after 264 hours at 100 mA cm−2 | 93 |
NiCoFePtRh | 0.5 M H2SO4 | 27 mV | 30.1 | Hydrothermal | 8.4% decrease in current density after 100 hours | 31 |
FeCoNiCuIr | 0.1 M HClO4 | 71 mV | 41.7 | Hydrothermal | 2 mV overpotential increase after 48 hours 100 mA cm−2 | 32 |
PtPdRhRuCu | 1 M KOH | 10 mV | 87 | Hydrothermal | Stable for 140 hours running at varying current densities from 10 to 100 mA cm−2 | 75 |
FeCoNiCuPtIr | 1 M KOH | 21 mV | 54.5 | Laser | Negligible increase after 24 hours at 10 mA cm−2![]() |
94 |
FeCoNiCrN | 1 M KOH | 161.8 mV | 124.5 | Laser | Stable for 10 hours at 100 mA cm−2![]() |
95 |
PtIrCuNiCr | 1 M KOH | 300 mV![]() |
N/A | Laser/solvothermal | Stable for 72 hours at 200 mA cm−2![]() |
96 |
CuNiFeCoCrTi | 1 M KOH | 117.11 mV | 95.32 | Mechanical alloying | Stable for 500 hours at 10 mA cm−2 | 97 |
FeCoNiCuPd | 1 M KOH | 29 mV | 47.2 | Magnetron sputtering | Stable for 24 hours at 100 mA cm−2![]() |
98 |
CoCrFeNiAl | 0.5 M H2SO4 | 73 mV | 39.7 | Mechanical alloying | Stable for 12 hours at 10 mA cm−2 | 99 |
IrPdPtRhRu | 1 M KOH | 60 mV | 42 | Plasma reduction | Stable for 6 hours at 100 mA cm−2 | 100 |
FeNiCoMnVOx | 1 M KOH | 89 mV | 88 | Plasma regulated synthesis | Stable for 100 hours at 10 mA cm−2 | 101 |
NiCoFeMnCrP | 1 M KOH | 220 mV | 94.5 | Sol–gel | Stable for 24 hours at 1.55 V![]() |
102 |
Pt34Fe5Ni20Cu31Mo9Ru | 1 M KOH | 20 mV | 27 | Solvothermal | 5% decrease in current density after 40 hours | 74 |
IrPdPtRhRu | 1 M HClO4 | 6 mV | N/A | Solvothermal | 50% decrease in current density after 6 hours | 103 |
PdPtRhIrCu | 1 M KOH | 15 mV | 37 | Solvothermal | Stable for 20 hours at 10 mA cm−2 | 34 |
Pt18Ni26Fe15Co14Cu27 | 1 M KOH | 11 mV | 30 | Solvothermal | Stable for 10 hours maintaining 10 mA cm−2 | 35 |
Pt28Mo6Pd28Rh27Ni15 | 1 M KOH | 9.7 mV | 25.9 | Solvothermal | 6.9 mV increase after 30 hours 10 mA cm−2 | 104 |
IrPdRhMoW | 0.5 M H2SO4 | 15 mV | 35 | Solvothermal | Stable for 100 hours at 100 mA cm−2![]() |
36 |
PdFeCoNiCu | 1 M KOH | 18 mV | 39 | Solvothermal | Negligible increase after 15 days at 10 mA cm−2 | 37 |
PdMoGaInNi | 0.5 M H2SO4 | 13 mV | 179.8 | Solvothermal | No increase after 12 hours at 10 mA cm−2 | 38 |
PtAuPdRhRu | 1 M KOH | 190 mV at 30 mA cm−2 | 62 | Ultrasonic | Stable for 8 hours at 100 mA cm−2 | 78 |
Composition | Electrolyte | Overpotential at 10 mA cm−2 | Tafel slope (mV dec−1) | Synthesis method | Stability | Ref. |
---|---|---|---|---|---|---|
a Overall water splitting. | ||||||
FeCoNiCuCr | 1 M KOH | 330 mV | 94 | Ball milling | Stable for 25![]() |
105 |
FeCoNiMnCr | 1 M KOH | 350 mV | 95 | Single-step pyrolysis | 15% decrease in current density after 6 hours | 30 |
(MgFeCoNiZn)O | 1 M KOH | 350 mV at 100 mA cm−2 | 44.9 | CTS | Stable for 20![]() |
106 |
PtIrFeNiCoCe | 0.1 M KOH | 485 mV | 119.1 | CTS | Stable for 5000 cycles | 84 |
AlNiCoIrMo | 0.5 M H2SO4 | 255 mV at 20 mA cm−2 | 55.2 | De-alloying | 11.5 mV increase after 7000 cycles | 29 |
AlNiCoIrMo | 0.5 M H2SO4 | 275 mV![]() |
55.2 | De-alloying | Stable for 48 hours at 1.52 V![]() |
29 |
AlAgAuCoCuFeIrNiPdPtRhRuMoTi | 0.5 M H2SO4 | 274 mV | 121.1 | De-alloying | N/A | 85 |
AlAgAuCoCuFeIrNiPdPtRhRu | 0.5 M H2SO4 | 258 mV | 84.2 | De-alloying | Stable for 40![]() |
85 |
MnNiCuCoVFeMoPdPtAuRuIr | 1 M KOH | 263 mV | 82.5 | De-alloying | Stable for 100 hours at 1.49 V | 107 |
AlNiCoRuMo | 1 M KOH | 150 mV | 54.5 | De-alloying | — | 86 |
AlNiCoFeMo | 1 M KOH | 240 mV | 46 | De-alloying | Stable for 2000 cycles | 108 |
ZnNiCoIrMn | 0.1 M HClO4 | 237 mV | 46 | De-alloying | Maintaining 1.52 V vs. RHE for 100 hours at 10 mA cm−2 | 88 |
CuCrFeNiCoP | 1 M KOH | 423 mV at 100 mA cm−2 | 70.7 | Electrodeposition | Stable for 48 hours at 100 mA cm−2![]() |
89 |
FeCoNiCuMn | 1 M KOH | 386 mV at 200 mA cm−2 | 69 | Electrospun nanofiber | Stable for 20 hours at 230 mA cm−2 | 90 |
CoFeLaNiPt | 0.1 M KOH | 377 mV | N/A | Electrodeposition | Stable for 10 hours at 557 mV overpotential | 92 |
FeCoNiIrRu | 0.5 M H2SO4 | 241 mV | 153 | Electrospinning method | Stable for 14 hours at 10 mA cm−2 | 109 |
NiCoNiZnV | 1 M KOH | 274 mV | 59 | Hydrothermal | Stable for 100 hours at 1.5 V vs. RHE | 110 |
MgMnFeCoNi | 1 M KOH | 354 mV at 100 mA cm−2 | 46.8 | Hydrothermal | Stable for 10 hours at 10 mA cm−2 | 33 |
FeCoNiCuPtIr | 1 M KOH | 255 mV | 61.7 | Laser | Negligible increase after 24 hours at 10 mA cm−2![]() |
94 |
FeCoNiCrN | 1 M KOH | 269.7 mV | 42.5 | Laser | Minimal overpotential increase after 50![]() |
95 |
PtIrCuNiCr | 1 M KOH | 300 mV![]() |
N/A | Laser/solvothermal | Stable for 72 hours at 200 mA cm−2 | 96 |
FeCoNiCuPd | 1 M KOH | 194 mV | 39.8 | Magnetron sputtering | Stable for 36 hours at 100 mA cm−2 | 98 |
IrFeCoNiCu | 0.1 M HClO4 | 302 mV | 58 | Microwave-assisted shock | Less than 60 mV increase in overpotential after 12 hours at 10 mA cm−2 | 28 |
NiCoFeMnCrP | 1 M KOH | 270 mV | 52.5 | Sol–gel | Stable for 24 hours at 1.55 V vs. RHE | 102 |
Pt34Fe5Ni20Cu31Mo9Ru | 1 M KOH | 259 mV | 39 | Solvothermal | 8% reduction in current density after 40 hours | 74 |
IrPdRhMoW | 0.5 M H2SO4 | 188 mV | 60 | Solvothermal | stable for 100 hours at 100 mA cm−2![]() |
36 |
MnFeCoNiCu | 1 M KOH | 263 mV | 43 | Solvothermal | Stable for 24 hours maintaining a current density of 10 mA cm−2 | 111 |
AlCrCuFeNi | 1 M KOH | 250 mV | 77.5 | De-alloying | Stable for 35 hours at 290 mV overpotential | 112 |
Composition | Electrolyte | Half-wave potential | Mass activity A mg−1 noble metal | Synthesis method | Stability | Ref. |
---|---|---|---|---|---|---|
Pt34Fe5Ni20Cu31Mo9Ru | 0.1 M HClO4 | 0.87 V | 11.4 A mg−1 Pt | Solvothermal | 11% current density reduction after 40 hours | 74 |
PtZrNbFeCuTaMoHfBiWZnSnPdNi | 0.1 M KOH | 0.86 V | N/A | Solvothermal | 4.8% current density reduction after 120 hours | 93 |
AlNiCoRuMo | 1 M KOH | 0.875 V | N/A | De alloying | Stable for 20![]() |
86 |
AlCuNiPtMn | 0.1 M KOH | 0.954 V | N/A | De alloying | 9.1% current density reduction after 164 hours | 113 |
AlCuNiPtMn | 0.1 M HClO4 | 0.945 V | 0.3466 A mg−1 Pt | De alloying | N/A | 113 |
PtPdFeCoNi | 1 M KOH | 0.85 V | N/A | CTS | 29% current density reduction after 15 hours | 114 |
HfZrLaVCeTiNdGdYPd | 0.1 M KOH | 0.85 V | 0.49 A mg−1 Pd | CTS | 14% current density reduction after 100 hours | 115 |
PdCuPtNiCo | 0.1 M KOH | 0.83 V | 0.176.1 A mg−1 Pt | Furnace pyrolysis | 5 mV increase half-wave potential after 10![]() |
116 |
PtPdRhFeCoNi | 0.1 M KOH | 0.85 V | N/A | Laser | half wave increased after 5000 cycles | 117 |
CoFeNiCuPd | 0.1 M KOH | 0.9 V | 2.037 A mg−1 Pd | Annealed on mesoporous material | 10 mV decrease half-wave potential after 10![]() |
118 |
Composition | Solution | Onset potential | Reaction | Mass activity A mg−1 noble metal | Synthesis method | Stability | Ref. |
---|---|---|---|---|---|---|---|
PtZrNbFeCuTaMoHfBiW ZnSnPdNi | 1 M KOH + 1 M CH3OH | 395 mV | MOR | 12.6 A mg−1Pt | Solvothermal | 28.7% decrease in current density after 10![]() |
93 |
Pt18Ni26Fe15Co14Cu27 | 1 M KOH + 1 M CH3OH | 391 mV | MOR | 10.96 A mg−1Pt | Solvothermal | Stable for 5000 seconds at 0.65 vs. RHE | 35 |
PtRhBiSnSb | 1 M KOH + 1 M C2H5OH | N/A | EOR | 15.558 A mg−1Pt+Rh | Solvothermal | Stable for 20![]() |
119 |
PtRhBiSnSb | 1 M KOH + 1 M CH3OH | N/A | MOR | 19.529 A mg−1Pt+Rh | Solvothermal | Stable for 20![]() |
119 |
NiCoFePB | 1 M KOH + 1 M CH3OH | 1.27 V | MOR | 15.04 A mg−1Pt | Sol–gel synthesis | 5% decrease in current density after 3600 seconds | 120 |
PdPtCuPbBi | 1 M KOH + 1 M C2H5OH | 300 mV | EOR | 18.21 A mg−1Pd+Pt | Template-assisted method | maintained 2.56 A mg−1Pd+Pt for 20![]() |
121 |
Composition | Electrolyte | Faradaic efficiency | Yield | Synthesis method | Stability | Ref. |
---|---|---|---|---|---|---|
NiCoNiZnV | 0.05 M H2SO4 + 1 M KOH | 14.75% at −1.45 V vs. RHE | 42.76 μg hmgcat−1 at −1.45 V vs. RHE | Hydrothermal synthesis | N/A | 110 |
(Co0.2Ni0.2Zn0.2Mg0.2Cu0.2)Fe2O4 | 0.1 M KNO3 + 1 M KOH | 98.1% at −0.5 V vs. RHE | 2.1 mmol hcm−2 at −0.5 V vs. RHE | Sol–gel synthesis | Stable for 50 cycles | 122 |
Despite its importance, the role of temperature during the initial reduction phase remains underexplored. Reported initial reduction temperatures range from room temperature to as high as 280 °C. Similarly, annealing times and temperatures exhibit significant variation. While some methods bypass the need for annealing, others require temperatures ranging from 350 °C to 700 °C for durations exceeding two hours. Notably, HEA-NPs have been successfully synthesized under remarkably mild conditions, such as reduction at room temperature and without annealing. Even when annealing is utilized, the required temperatures are substantially lower than the melting points of the precursor metals. In contrast, bulk-phase HEA synthesis typically demands the complete melting of all precursor metals.
These mild reaction conditions suggest the involvement of additional mechanisms that are not yet fully understood and are less prominent at the bulk scale. Factors such as atomic diffusion and reducing agent interactions require closer examination, as atomic diffusion is significantly accelerated at elevated temperatures and becomes more pronounced at the nanoscale. Additionally, the complex interactions and kinetics between reducing agents and precursor materials can influence mixing and composition at various stages of the synthesis process. A deeper understanding of these factors could be crucial in determining the temperatures required for nanoscale HEA formation and may challenge the current explanation of entropy-driven material formation.
For example, a study by Chen et al.74 demonstrated the synthesis of convex cube-shaped Pt34Fe5Ni20Cu31Mo9Ru HEA nanocrystals via a solvothermal method under mild conditions (220 °C for 2 hours). Demonstrating the feasibility of the synthesis as well as demonstrating outstanding electrocatalytic performance. The study demonstrated unique nanocrystal morphologies compared to many other HEA synthesis methods, with two specific shapes observed: cube-shaped structures after one hour of heating and converged cube-shaped structures after two hours. These time-dependent studies provide valuable insights into the mechanism at the nanoscale. More detailed investigations in future studies, including intermediate characterization and exploration of temperature and time effects, would help clarify the mechanisms underlying HEA formation at the nanoscale. This deeper understanding could also aid in developing new synthesis strategies and addressing challenges such as preventing intermetallic formation and phase separation.
When using a seed-mediated approach for HEA synthesis, several key factors must be carefully considered. First, the seed particles must be monodispersed to enhance shell uniformity. Second, the choice of supporting ligands or capping agents is crucial, as they influence shell deposition. Lastly, the final HEA stoichiometry is determined by the relative sizes of the seed and shell. Kar et al.,123 synthesized various HEA systems using a combination of seed-mediated synthesis and thermal annealing. By systematically varying the seed and shell compositions, they explored different metal combinations, including Pd, Cu, Pt, Ni, Co, Au, Ag, and Sn. Their study provided a comprehensive analysis of how the interplay between seed and shell compositions influences core–shell formation, ultimately shaping the final HEA composition.
The study begins by investigating the role of metals with different redox potentials in core–shell formation. Using monodispersed PdCu, PdAg, PdCo, and PdAu seeds, the researchers examined how galvanic replacement influences shell growth when Pd is paired with metals of varying reduction potentials. Through TEM (Fig. 15a, b and f, g), STEM-EDS elemental mapping (Fig. 15c and h), and linescan analysis (Fig. 15d and i), they determined that for a core–shell structure to form, the core metals must have a higher redox potential than the shell metals. When synthesizing PdPtCuNiAu HEAs using either monodispersed PdCu or AuCu seeds, they observed that Au, having a higher reduction potential than both Pd and Cu, significantly influenced elemental distributions due to galvanic replacement. Specifically, when PdCu was used as the seed and Au was in the shell, both Pd and Cu exhibited depletion in elemental maps and linescan data (Fig. 15c–e).
![]() | ||
Fig. 15 TEM images of PdCu seed (a) and PdCu seeds with PtNiAu shell (b). EDS Map (c), linescan (d), and atomic% break down (e) of PdCu@PtNiAu core shell NPs (b). TEM images of AuCu seed (f) and AuCu seeds with PdPtNi shell (g). EDS Map (h), linescan (i), and atomic% break down (j) of AuCu@PdPtNi core shell NPs (g). Adapted from ref. 123 Copyright (2023), with permission from Springer Nature. |
Conversely, in PdCoPtNiCu systems with PdCo seeds (Fig. 16a–e), only Co was depleted (Fig. 16e), as it had the lowest redox potential among the core metals. Lattice mismatch effects were further investigated using PdSnPtNiCu HEA systems. Sn favors a BCC structure, whereas the remaining constituent elements predominantly adopt an FCC structure, resulting in inherent lattice mismatch. Consequently, when Sn was used to form the shell on PdCu seeds, it exhibited both homogeneous nucleation and heterogeneous nucleation, and growth, on the PdCu seeds (Fig. 16g). Therefore, metals that tend to hinder alloy phase formation or have poor wettability should go in the seed of the core–shell NPs. The study emphasizes the critical importance of accounting for redox potential and lattice mismatch when designing HEAs, as these factors directly influence the structural and compositional outcomes of seed-mediated HEA synthesis. Furthermore, the findings encourage further investigation, as the observed effects are closely related to key empirical parameters governing HEA formation. Further research could focus on exploring VEC and electronegativity effects on HEA NP formation.
![]() | ||
Fig. 16 TEM images of PdCo seed (a) and PdCo seeds with PtNiCu shell (b). EDS Map (c), linescan (d), and atomic% break down (e) of PdCo@PtNiCu core shell NPs (b). TEM images of PdCu seed (f) and PdCu seeds with PtNiSn shell (g). EDS Map (h), linescan (i), and atomic% break down (j) of PdCu@PtNiSn core shell NPs (g). Adapted from ref. 108 and 123 Copyright (2023), with permission from Springer Nature. |
The co-reduction mechanism refers to the simultaneous reduction of all the elements in an HEA precursor mixture, facilitating their homogeneous mixing and the subsequent formation of HEA NPs. Co-reduction as a proposed mechanism is particularly interesting as an additional annealing step is not mandatory in some reported cases. However, even in co-reduction processes, annealing may still be employed to enhance phase uniformity, such as achieving a well-defined FCC or BCC phase single-phase HEA.
Sun et al.,124 investigated the co-reduction of Au, Ag, Pt, Cu, and Pd metal salts using poly(ethylene glycol) methyl ether thiol (mPEG–SH) ligands and being reduced by sodium borohydride (NaBH4). The resulting single and polycrystalline nanoparticles ranged in diameters from 0.5 nm to 3 nm. HAADF-STEM images revealed the coexistence of sub-nanometer, non-crystalline metal clusters alongside the synthesized nanoparticles (Fig. 18f). The authors subsequently varied the metal-to-ligand ratios used during synthesis to further investigate the HEA formation (Fig. 17a–c). It was observed that a high metal-to-ligand ratio (5:
1) resulted in the formation of phase-separated core–shell nanoparticles (Fig. 17c and f). The authors highlight that thiol ligands react with metal ions to form metal thiolates, which exhibit similar reduction rates across all metals thiolates. However, if insufficient thiol ligands are present, the remaining metal ions will have varying reduction rates.
![]() | ||
Fig. 17 AuAgCuPtPd HEA NPs synthesized with different metal![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
![]() | ||
Fig. 18 Time-lapsed LPTEM of AuAgCuPtPd HEA NP formation with metal![]() ![]() ![]() ![]() ![]() ![]() |
However, a 1:
1 metal-to-ligand ratio was identified as optimal due to the smaller particle size and improved size distribution compared to higher ratios. However, the resulting parameter changes still did not fully explain the presence of sub-nano clusters. Subsequently, liquid phase TEM (LPTEM) using aqueous medium and reduction by electron beam was used to study the HEA NP formation. The use of aqueous medium with electron beam reduction had similar reaction kinetics to that of NaBH4 when using mPEG-SH ligands. Given the spatial resolution limitations of the experimental setup, metal-to-ligand ratios of 1
:
1.5 and 1
:
2 were selected to produce larger nanoparticles, ensuring clearer imaging and more accurate analysis (Fig. 18a and b). Using LPTEM, nanoparticle growth was tracked in real-time, revealing the formation of large aggregates (10–20 nanoparticles per cluster) (Fig. 18a–c). Interestingly, despite this aggregation, the nanoparticles exhibited independent, isotropic growth, maintaining clear interparticle separation before reaching their final size of 5–10 nm within 30–60 seconds.
Further HAADF-STEM images of the samples after electron beam reduction detected sub-nanometer metal clusters (Fig. 18f), suggesting that HEA formation deviates from classical nucleation theory and Lamer's model. In a follow-up experiment to investigate the role of supersaturation, the researchers returned to the original synthesis method using NaBH4 as the reducing agent and analyzed its effect on HEA formation by varying its concentration. The results showed that higher NaBH4 levels led to increased nanoparticle size and aggregation, supporting previous reports that NaBH4 can displace ligands.
Based on these findings, the authors proposed a non-classical nucleation mechanism for HEA formation. In their model, NaBH4 rapidly reduces all metal salts, generating small, ligand-stabilized metal clusters. Subsequently, borohydride and hydride ions displace the ligands, triggering cluster aggregation and ultimately forming HEA (Scheme 4). The study highlights the complex interactions between supporting ligands and reducing agents and their impact on HEA formation. Further research on the complexity of HEA formation mechanisms will help in developing a comprehensive understanding of the synthesis process.
![]() | ||
Scheme 4 Schematic of HEA formation due to NaBH4 and ligand interactions. Reproduced from ref. 124 Copyright (2023), with permission from RSC publishing. |
Recent comprehensive investigations into the formation mechanisms of HEA NPs by Dey et al.,125 utilized a solvothermal synthesis approach to explore the time evolution of HEA formation across seven distinct systems. Through detailed investigations, the study revealed that different systems follow unique formation pathways depending on their elemental composition. The study initially focused on noble metals Pd, Pt, Rh, and Ir combined with other transition metals and a post-transition metal, Sn. For the NiPdPtRhIr system, the synthesis began with the formation of Pd-rich seeds that served as templates for the gradual incorporation of other metals, ultimately achieving equimolar compositions (Fig. 19). This seed-mediated mechanism was analyzed by collecting and evaluating samples at various time intervals using TEM and EDS. Subsequent composition changes with further transition metal additions resulted in the same seed-mediated formation pathway for NiFeCoPdPt, NiFePdPtIr, and NiFeCoPdPtIr systems.
![]() | ||
Fig. 19 EDS maps of NiPdPtRhIr HEA NP formation at various time intervals (1, 2.5, 5, 7.5, and 10 minutes), accompanied by bar charts illustrating the corresponding elemental composition changes. Reproduced from ref. 125 Copyright (2023), with permission from American chemical society. |
The most intriguing findings emerged when noble metals were combined with the p-block element Sn. In the SnPdPtRhIr system, the metals underwent simultaneous co-reduction and mixing, deviating from the seed-mediated growth observed in other systems. However, the order of precursor addition was crucial. If the Sn precursor was injected first, followed by the remaining metal precursors then PdSn rich nanoparticles were formed initially (seed mediated growth). In contrast, injecting Sn after the other precursors led to co-reduction. Interestingly, Sn alone did not induce a color change in the reaction flask, indicating that the chosen reducing agent could not reduce Sn salts in the absence of other metal precursors. Therefore, it is crucial to study the interaction between reducing agents and individual metal precursors. As it not only reveals possible redox interactions between precursor metals but also highlights potential causes of proportional discrepancies, even when equiatomic ratios are used.
The role of Sn in HEA formation was also further investigated in the study using a NiSnPdPtIr system. Interestingly characteristics of both pathways were observed for this system.
Overall, these studies underscore the significant influence of elemental composition on the formation pathways of HEA NPs. The distinct mechanisms of seed-mediated growth, co-reduction, and hybrid pathways, highlight the complexity of HEA NPs synthesis. In particular, the role of elements such as Sn in modulating reduction dynamics suggests new opportunities for tailoring HEA structures through strategic precursor selection.
This review discussed the synthesis of HEAs under mild conditions, which presents several advantages over more conventional top-down methods. Notably, many of the discussed synthesis techniques facilitate the study and characterization of reaction intermediates as well as catalyst–support interactions. One of the most intriguing aspects of these mild synthesis conditions is the significantly lower temperatures required compared to their top-down counterparts. Both seed-mediated and co-reduction formation pathways demonstrate substantially reduced temperatures, challenging the prevailing notion that HEAs can only form under high-temperature conditions.
Overall, the outlook for HEAs in electrocatalysis appears promising. Given the relatively early stage of the field, many fundamental questions regarding material properties, formation mechanisms, and catalytic performance remain unanswered. We emphasize the need for more rigorous research in several key areas of nanoscale HEA synthesis.
First, further investigation into reduction methods is crucial, as few studies have systematically examined the impact of reduction rates independent of temperature or ligand interactions. Second, a deeper understanding of initial reduction conditions is needed to explore their potential for substituting or modifying annealing requirements. Faster diffusion rates and enhanced atomic mobility during reduction could lower the temperatures and durations required for final annealing, improving overall efficiency. Finally, continued research into the catalytic properties of HEAs will be essential for guiding the design of next-generation catalysts and accelerating the development of low-cost, noble-metal-free alternatives.
The commercialization of nanoscale HEAs will ultimately depend on the development of scalable, cost-effective synthesis methods with reduced energy demands. Advancing these aspects will be key to unlocking their full potential for sustainable catalysis applications.
This journal is © The Royal Society of Chemistry 2025 |