Open Access Article
Brian Cantorab
aDepartment of Materials, University of Oxford, Parks Road, Oxford OX1 3PH, UK. E-mail: brian.cantor@materials.ox.ac.uk
bBrunel Advanced Solidification Centre (BCAST), Brunel University London, Uxbridge, UB8 3PH, UK. E-mail: brian.cantor@brunel.ac.uk
First published on 9th September 2025
Multicomponent phase space is enormous and contains a vast number of complex new materials. Despite intensive investigation in the last decade and a half however, we are only slowly making progress towards understanding these new materials. This paper attempts to summarise some of the fundamental discoveries we have made about the geography of multicomponent phase space and the wide range of complex new materials that we have found within it. This paper discusses briefly the following topics: the size and shape of multicomponent phase space and the range of single- and multiple-phase fields that it contains; the (initially) surprising presence of many large near-ideal single-phase solid-solution phases, stabilised by a high configurational entropy of mixing; the extensive and wide-ranging variation of local nanostructure and associated mechanical and electronic lattice strain that permeates throughout high-entropy solid-solution phases; and some of the unusual, exciting and valuable properties that are then produced within multicomponent and high-entropy materials. Many of the results discussed have been obtained from the fcc Cantor alloys (based on the original Cantor alloy, equiatomic fcc CrMnFeCoNi) and the bcc Senkov alloys (based on the original Senkov alloy, equiatomic VNbMoTaW), two groups of multicomponent high-entropy single-phase materials that have been particularly widely studied. Similar behaviour is also found in other multicomponent high-entropy single-phase materials, though these have not been studied so intensively. In comparison with multicomponent high-entropy single-phase materials, rather little is known about multicomponent multiphase materials that have also not been studied so intensively.
There was antipathy and some disdain amongst scientific colleagues and research funders in the 1980s, 1990s and 2000s about the idea of investigating multicomponent and high-entropy materials, and Yeh and myself both received considerable resistance from our respective academic establishments against pursuing our early experiments, which partly explains the lengthy delays in both cases between our first research work being undertaken and then finally being published.14,15 For similar reasons, our first published papers on multicomponent and high-entropy materials in the early 2000s, including the two seminal papers mentioned above, did not initially receive much (or even any) significant scientific attention, with only a few citations in the first five or ten years following their publication. Since then, however, scientific interest has mushroomed rapidly, with citations increasing dramatically and multi-million dollar research programmes being initiated in many countries worldwide.3,16,17 The discovery of multicomponent high-entropy materials has opened up the possibility of developing many new materials with exciting new properties and, after an initial delay, scientists have, not surprisingly, begun to realise their potential and have been increasingly active in investigating different multicomponent materials and trying to understand their structures and properties.3,16 The Web of Science gives a total, up to the present time, of almost fifty thousand research papers on multicomponent or high-entropy alloys or materials, and Google Scholar gives a total of approximately ten and fifteen thousand citations, respectively, for each of the two seminal papers by Cantor and Yeh mentioned above. And the number of scientific papers on multicomponent materials and the number of citations to them are both continuing to rise rapidly.
The last decade and a half in the 2010s and 2020s has seen, therefore, a fairly large body of scientific investigation into the manufacture and resulting structures and properties of multicomponent and high-entropy materials.3,16,17 It turns out, however, that multicomponent phase space is vast, and contains a truly enormous number of materials, running into many trillions, the majority of which have still never been made, let alone investigated in any detail.3,16,17 Until we began to study multicomponent materials seriously in the 2010s and 2020s, virtually all the materials that we had ever made had been based, as mentioned above, on either a single component or binary mixtures of different components, with only occasionally a concentrated ternary addition, and with higher order (quaternary, quinary, etc.) additions never much above very dilute levels. To put it another way, all our materials have until recently been at or very close to the corners and binary edges and occasionally the ternary faces of multicomponent phase space, and we have until recently avoided exploring the inner regions within the main body of multicomponent phase space where most of the compositions of the trillions of different and new multicomponent materials can be found. Unfortunately, it also turns out that our thermodynamic, classical and quantum mechanical, and statistical theories of the structure and properties of materials have all been devised for relatively pure materials and/or simple mixtures with only one or two relatively dilute alloying additions that can then be treated as relatively small and independent linear perturbations of the properties of the main component in the material.3,18 Because of the approximations required to make these theories tractable as well as the multiplicity of parameters needed to make calculations based on them, our fundamental theories are not easily extrapolated to the complexity of multicomponent materials, which often exhibit non-linear behaviour because of the interactions between multiple components in concentrated proportions.3,18 This means that in order to explore multicomponent phase space and the different materials within it, there is no real alternative to massive amounts of experimental investigation to provide new data that will ultimately allow us to develop more complex and more definitive theories of multicomponent material structures and properties. At present, however, the size of multicomponent phase space is so large that, despite intensive research over the last decade and a half, we have still explored only a small fraction of the totality of all possible materials that could be made, so we are still a long way from developing a detailed scientific understanding of the range and extent of all the different multicomponent materials and their structures and properties.3,18
Despite these difficulties, after considerable false starts and with considerable initial misunderstandings, we have finally begun to understand some of the main features of the geography of multicomponent phase space and of the structures and properties of some of the different multicomponent materials within it,3,16,17 and it is the purpose of this paper to try to summarise briefly where we have got to. I do this by asking a series of simple and obvious questions (that I am frequently asked) and either giving the answer or answers that we have finally been able to come up with, often after a lot of hard work to clear away initial misconceptions, or alternatively indicating the extent of our continued lack of understanding and the need for further work (a conclusion that is still fairly common, essentially because of the enormity of multicomponent phase space and the complexity of the structure and properties of the multicomponent materials within it).
Calculating the total number of materials in a given alloy system is a matter of combinatorial maths.22,23 If one out of n = 100/x atoms is changed from one kind to another, the composition changes by x at%, the material goes out of specification, and it is effectively changed into a different material. We need, therefore, to work out how many different ways we can pick n = 100/x atoms from c different kinds of atom (i.e. c different components), with the order of picking being irrelevant, i.e. mathematically a combination rather than a permutation, and with multiple picking of atoms of each type allowed, i.e. mathematically with repetition rather than without repetition. For c components and n = 100/x composition points for each component, the total number of materials that can be made N is given then by the law of combinations with repetition:22,23
We can generate the whole of multicomponent phase space, containing all the possible materials that could conceivably be made, by using all the elements in the periodic table, shown in Fig. 1, as our set of starting materials or components. This gives us a total of c = 118 components with which to manufacture different materials. We can reduce this to c = 80, by removing all the elements in row seven because they are radioactive with short lifetimes, and also removing all the elements in column eighteen because they are noble gases and chemically inert. Alternatively, we might want to be more restrictive and take (say) c = 50 or 60, because some of the remaining elements are highly reactive and difficult to handle, such as lithium, fluorine or arsenic, though all these elements are used in important materials. Most engineering materials are specified to x = 0.1%, though lower-grade materials are often specified to no better than x = 1% or 5%, and some high-performance materials require a greater degree of specification, down to x = 0.01%, or 0.001%19–21 (or even, in some special cases, to the nearest part per million (ppm) or x = 0.0001%).
Table 1 shows the total number of materials that can be made using different assumptions for the number of components c, the specification x, and the number of composition points for each component n = 100/x. Taking c = 80, x = 0.1% and n = 1000 gives the total number of materials that can be made as:
ln N = 1079 ln 1079 − 79 ln 79 − 1000 ln 1000 ≈ 282 |
| ∴N = 10123 |
| x (at%) | n | c = 40 | c = 50 | c = 60 | c = 70 | c = 80 |
|---|---|---|---|---|---|---|
| 5 | 20 | 16 | 18 | 19 | 21 | 22 |
| 1 | 102 | 36 | 41 | 46 | 50 | 53 |
| 0.1 | 103 | 72 | 86 | 100 | 111 | 123 |
| 0.01 | 104 | 111 | 134 | 157 | 179 | 200 |
| 0.001 | 105 | 150 | 183 | 216 | 248 | 279 |
Being extremely conservative and taking c = 60 and x = 1% or even 5% gives the total number of materials that can be made as N = 1046 or 1019 respectively. In all cases, the number of possible materials that could be made is enormous. Many multicomponent material compositions will, of course, be difficult to manufacture and not be of much use. Nevertheless, it is quite clear that, whatever values are used for c, x and n, there is an extremely large number of different materials populating multicomponent phase space.
We have studied enough multicomponent materials to be able to say definitively that many but not all of them are high entropy materials.3,16 In many cases, multicomponent materials have been found to exhibit no more than one or two random or near-random solid-solution phases:3,16,17 in these cases, the configurational entropy of mixing is high enough to suppress chemical reactions between the different components and the resulting formation of stoichiometric fixed-composition chemical compounds. These are high-entropy materials. In many other cases, however, multicomponent materials have been found to exhibit mixtures of multiple phases:3,16 in these cases, the configurational entropy of mixing is not high enough to suppress chemical reactions between the different components and the resulting formation of stoichiometric fixed-composition chemical compounds. These are not high-entropy materials.
The Cantor alloys are the set of multicomponent materials that have a single-phase face-centred cubic (fcc) solid-solution structure, named after the first such material that was discovered, the original Cantor alloy CrMnFeCoNi;3,4,16,24 similarly, the Senkov alloys are the set of multicomponent materials that have a single-phase body-centred cubic (bcc) solid-solution structure, named after the first such material that was discovered, the original Senkov alloy VNbMoTaW.3,16,25–27 Very many different Cantor alloys have been manufactured, probably several hundred different materials with different compositions, but all with the same single-phase fcc random or near-random solid-solution structure, and there are many more yet to be made;3,16,24 similarly, very many different Senkov alloys have also been manufactured, again probably several hundred different materials with different compositions, but all with the same single-phase random or near-random (in this case) bcc solid-solution structure, and again there are many more yet to be made.3,16,26,27 There have also been somewhat fewer, but still a reasonable number, of different multicomponent materials manufactured so far with different compositions but all with the same single-phase hexagonal close-packed (hcp) random or near-random solid-solution structure.3,16 And similar behaviour has also been found in many compounds, i.e. somewhat fewer have been made so far, but there are still very many different multicomponent materials with different compositions but the same compound structure.3,16 This includes intermetallic compounds such as the multicomponent monoaluminides MAl that have a single-phase B2 CsCl-type ordered bcc structure with all the aluminium atoms on one sublattice and a mixture of different metallic atoms distributed randomly on the other sublattice; ionic compounds such as the multicomponent mono-oxides MO that have a single-phase rock-salt ordered simple-cubic structure with all the oxygen atoms on the anion sublattice and a mixture of different metallic cations distributed randomly or near-randomly on the cation sublattice; and covalent compounds such as the multicomponent diborides MB2 that have a single-phase C32 TiB2-type hexagonal structure, with a mixture of different metallic atoms distributed randomly or near-randomly on the lattice, each covalently bonded by two boron atoms.3,16 Table 2 lists some of the compounds that have been shown to extend over a range of different compositions in multicomponent phase space.
| Phase | Typical compositions |
|---|---|
| Rock salt | (Li6.5Ti0.5MnNb)(O1.7F0.3) |
| (MgCoNiCuZn)O | |
| (TiVZrNb)C; (TiZrNbHfTa)C | |
| Fluorite | (ZrYHfCe)O2; (ZrYHfCeGd)O2 |
| (TiZrSnCeHf)O2 | |
| (Mo0.5CePrNdSmGd)O2 | |
| Pyrochlore | (SmEuTbDyLu)2Zr2O7 |
| (LaCeNdSmEu)2Zr2O7 | |
| Perovskite | (TiMnZrSnHf)SrO3; (TiGeZrSnSn)SrO3 |
| (TiZrHfNbSn)(SrBa)O3 | |
| (YLaNdSmGd)(CrMnFeCoNi)O3 | |
| Spinel | (AlMnFeCoNi)3O4; (AlCrMnFeNi)3O4; (CrMnFeCoNi)3O4; (CrMnFeNiZn)3O4 |
| (CrMnFeCoNi)Fe2O4 | |
| Monoborides | (VCrNbMoTa)B; (VCrNbMoW)B (CrMnFeCoMo)B; (CrNiMoTaW)B |
| Boro-carbo-nitrides | (TiZrNbHfTa)(CN); (TiZrNbTaW)(BCN) (TiZrNbHfTa)(BCN); (ZrNbHfTaW)(BCN) |
| Hexaborides | (YNdSmEuYb)B6; (YCeSmErYb)B6 |
| Disilicides | (TiZrNbMoW)Si2; (TiNbMoTaW)Si2 |
A sufficiently large number of Cantor alloys have been investigated to allow a rough estimate of the extent of the single-phase fcc Cantor-alloy phase field from the different compositions that have been found to adopt a single-phase random or near-random fcc solid-solution structure. The maximum solubilities that have been found for individual components in single-phase fcc Cantor alloys3,16,24 are shown in Table 3 and Fig. 2. Similarly, a sufficiently large number of Senkov alloys have been investigated to allow a rough estimate of the extent of the single-phase bcc Senkov-alloy phase field from the different compositions that have been found to adopt a single-phase random or near-random bcc solid-solution structure. The maximum solubilities that have been found for individual components in single-phase bcc Senkov alloys3,16,26,27 are shown in Table 4 and Fig. 3. Solubility levels of any particular component vary, of course, with temperature, material composition and manufacturing method. Maximum solubilities such as those shown in Tables 3 and 4 are, therefore, almost certainly underestimates, since only a limited number of temperatures, material compositions and manufacturing methods have been investigated to date. Nevertheless, as can be seen in Tables 3 and 4, the multicomponent single-phase fcc Cantor and bcc Senkov alloy fields are both very large and complex in shape, each containing enormous numbers of different material compositions. It is likely that this is also the case for many other single-phase fields such as hcp, B2 monoaluminides, rock-salt mono-oxides, C32 diborides, and other phases3,16 such as those listed in Table 2. These single-phase fields are separated in multicomponent phase space by multiple-phase fields,3,16 and it is almost certain that these also extend over wide ranges of different compositions. There is clearly a major experimental job to be done in mapping out the size and shape of all the different single-phase and multiple-phase fields in multicomponent phase space.
| Solute max. at% | |||||||
|---|---|---|---|---|---|---|---|
| Au | Pd | Pt | Zn | Nb | V | Ti | Mo |
| 20 | 20 | 20 | 20 | 20 | 20 | 20 | 10 |
| Solute max. at% | ||||||
|---|---|---|---|---|---|---|
| Cu | Al | Cr | Mn | Fe | Co | Ni |
| 100 | 8 | 25 | 50 | 50 | 50 | 100 |
![]() | ||
| Fig. 2 Approximate maximum composition ranges for different components in single-phase fcc Cantor alloys. | ||
| Solute max. at% | |||||||
|---|---|---|---|---|---|---|---|
| Re | Si | Al | Hf | Zr | Ti | W | Ta |
| 8 | 2 | 30 | 30 | 30 | 30 | 100 | 100 |
| Solute max. at% | |||
|---|---|---|---|
| Mo | Nb | Cr | V |
| 100 | 100 | 100 | 100 |
![]() | ||
| Fig. 3 Approximate maximum composition ranges for different components in single-phase bcc Senkov alloys. | ||
When alloying additions are increased beyond fairly low levels, the resulting materials often become, as just explained, relatively difficult to process, because of the formation of larger and larger quantities of brittle compounds. This seems to have led, as also just explained, to a strong and widespread tendency to avoid adding too much or too many alloying additions. We have now, however, studied enough multicomponent materials to know that this effect is not always maintained and indeed can often be reversed when we continue to add increasingly higher numbers and higher concentrations of alloying elements.3,16 When enough alloying additions are made in sufficiently high concentrations, the tendency to form compounds is often, though not always, suppressed, leading instead to the formation in many cases of single-phase random solid-solution phases.3,16 To summarise: in a large number of multicomponent materials, compound formation is suppressed, leading to a single-phase random solid-solution phase that is often easy to process; in a large number of other multicomponent materials, however, compound formation is not suppressed, leading to the formation of a mixture of multiple compound phases, which is often not easy to process.
, where R is the gas constant, and a minimum value when the material forms a mixture of pure components and stoichiometric fixed-composition compounds of ΔScmix = 0.‡‡ The excess entropy of mixing ΔSxsmix comes from any other changes in disorder in the material and, like the heat of mixing, also depends on the set of pairwise, three-way, four-way, etc. interaction energies between the different atoms in the material, i.e. ΔSxsmix = f′(ωij, ψijk, χijkl…), where f and f′ are different functions. Overall, the Gibbs free energy is given, therefore, by:33–35The first term on the right hand side is the free energy of the unmixed components and is independent of the final structure of the material; the sum of the last three terms adds up to the free energy of mixing, all of which are dependent on the final structure of the material. When a given set of components are mixed together to form a material, it can adopt a variety of different structures, and the final structure is the one with the lowest free energy G, which is the same, therefore, as the one with the lowest free energy of mixing ΔGmix.33–35
The heat of mixing ΔHmix usually favours compound formation, since it is determined by the interaction energies and the resulting chemical bonding between the different atoms in the material; on the other hand, the entropy of mixing ΔSmix usually favours solid-solution formation, since its configurational component is determined by the disorder of mixing the different atoms closely together in the material. As mentioned in the introduction, most of our conventional materials have almost all been made from one, or sometimes two, main components, with relatively few alloying elements added in relatively dilute amounts. In conventional materials, therefore, the free energy of mixing ΔGmix and the overall free energy G of the various possible material structures are usually dominated by the heat of mixing ΔHmix. This means that compounds are formed readily when the interaction energies are positive (i.e. attractive, so that chemical bonds can form between the different atoms), so that compound formation increases with increasing number and concentration of the alloying elements, and the entropy of mixing ΔSmix is relatively small, so that random or near-random solid solutions are only favourable at high temperatures.
The situation is different, however, in multicomponent materials, because the entropy of mixing can be much higher.3,4,11,16 The maximum configurational entropy of mixing in any given alloy system is at the equiatomic composition, i.e.
, where c is the number of components, and all xi values are the same (xi = 0.5 in a binary alloy, 0.33 in a ternary alloy, 0.25 in a quaternary alloy, etc.). Table 5 shows the resulting maximum ideal§§ entropy of mixing ΔSmix = ΔScmix = −R
ln(1/c) and the corresponding ideal free energy of mixing ΔGmix = −TΔScmix = RT
ln(1/c) at 1000 K in such an equiatomic multicomponent material for different numbers of components ranging from c = 2 to 50. The same results are shown graphically in Fig. 4. The entropy of mixing in J K−1 mol−1 and the free energy of mixing at 1000 K (1273 °C) in kJ mol−1 are numerically equal and opposite, the first positive and the second negative. Differentiating:
ln(1/c) and ΔGmix = RT
ln(1/c) versus number of components c in an equiatomic multicomponent material (taking R = 8.3145 J K−1 mol−1)
| Number of components c | ln(1/c) | Ideal entropy of mixing ΔSmix (J K−1 mol−1) | Ideal free energy of mixing ΔGmix at 1000 K (kJ mol−1) |
|---|---|---|---|
| 2 | −0.69 | 5.76 | −5.76 |
| 3 | −1.10 | 9.13 | −9.13 |
| 4 | −1.39 | 11.53 | −11.53 |
| 5 | −1.61 | 13.38 | −13.38 |
| 6 | −1.79 | 14.90 | −14.90 |
| 7 | −1.95 | 16.18 | −16.18 |
| 8 | −2.08 | 17.29 | −17.29 |
| 9 | −2.20 | 18.29 | −18.29 |
| 10 | −2.30 | 19.12 | −19.12 |
| 20 | −3.00 | 24.91 | −24.91 |
| 30 | −3.40 | 28.28 | −28.28 |
| 40 | −3.69 | 30.67 | −30.67 |
| 50 | −3.91 | 32.53 | −32.53 |
An entropy of mixing below (approximately) the gas constant R, i.e. ΔSmix ≲ 8.3 J K−1 mol−1 or c ≲ 3, is lower than most of the interaction energies between the different atoms in a material, and is not, therefore, large enough to suppress compound formation in most materials;3,16 on the other hand, an entropy of mixing above (approximately) the gas constant R, i.e. ΔSmix ≳ 8.3 J K−1 mol−1 or c ≳ 3, is higher than some or all of the interaction energies between the different atoms in a material, and is becoming high enough, therefore, to suppress compound formation in an increasing number of materials as the number of components increases.3,16 Overall, therefore, compound formation is favoured when there are only low concentrations of a few alloying elements, since the configurational entropy of mixing is too low to outweigh chemical bonding effects, but compound formation becomes increasingly disfavoured when there are high concentrations (e.g. at equiatomic compositions) of many alloying elements (i.e. c ≳ 3), since the configurational entropy is then becoming high enough to outweigh chemical bonding effects. This explains why multicomponent materials were avoided for many years, when relatively small increases in alloying numbers and concentrations led to excessive compound formation, and the materials were found to be difficult to manufacture. It also explains our recent and, at first, surprising but now well-established discovery of large single-phase solid-solution regions in multicomponent phase space for equiatomic (and near-equiatomic or, at least, highly concentrated) compositions with five or more components, as discussed in previous sections.
It is worth pointing out, however, that an entropy of mixing even as high as (approximately) three or four times the gas constant R, i.e. ΔSmix ≳ 25–35 J K−1 mol−1 or c ≳ 20, is not larger than the interaction energies between the atoms in a material when they are very different chemically, and the resulting chemical bonds between them are very strong.36 It seems likely, therefore, that when the number of components is increased to more than (say) ten or twelve, i.e. c ≳ 10–12, the inevitable chemical differences and interaction energies between an increasingly wide range of different elements become too large, and compound formation again takes over. This has not yet been fully established, but it would explain why extensive solid-solution formation has been discovered for multicomponent materials when most of the components are relatively similar chemically, e.g. for single-phase fcc Cantor alloys containing predominantly late transition elements, or single-phase bcc Senkov alloys containing predominantly early transition elements.
Many experimenters have manufactured the original Cantor alloy via many different methods and subjected it to a wide range of different mechanical and thermal treatments, in almost all cases finding it to have a single-phase fcc structure that was easy to manufacture and apparently highly stable.3,16 There was a certain degree of schadenfreude, therefore, when it was first reported that small compound precipitates had been found after extended heat treatment of the original Cantor alloy.37,38 However, decomposition of the solid solution via precipitation only took place after annealing for several years in a fairly narrow temperature range, and it should not really have been a surprise. It is quite obvious from the functional form of the free energy of a material that entropy effects must become vanishingly small as the temperature decreases (since the second term in ΔGmix = ΔHmix − TΔSmix goes to zero). In fact the third law of thermodynamics says that all materials at absolute zero must form mixtures of perfect single crystals, i.e. mixtures of pure components and fixed-composition compounds. It is similarly quite obvious from the functional form of the free energy of a material that entropy effects must become completely dominant as the temperature increases (since the second term in ΔGmix = ΔHmix − TΔSmix increases without bound). In fact, all materials must become fully mixed atomically at sufficiently high temperatures. All materials will, therefore, in principle separate and decompose from an atomically mixed solution into a mixture of pure components and fixed-composition compounds (sometimes, though not always by precipitation) as they are cooled from high temperature. The only question is how strong are the compound-forming chemical interaction energies between the different atoms in the material relative to the solution-forming configurational entropy of mixing them intimately together; how high, therefore, is the critical temperature above which the equilibrium structure of a material is a solution and below which its equilibrium structure is a mixture of separated phases? In many cases, the chemical interaction energies are strong, much greater than the configurational entropy, and the critical temperature is well above the melting point, so that no solid-solution phase can form, but a liquid solution is often formed at high temperatures above the melting point. In many other cases, the chemical interaction energies are less strong, i.e. of a similar size to or somewhat lower than the configurational entropy, and the critical temperature is then below the melting point, so that a solid solution is formed at high temperatures, but it separates into a mixture of pure components and compounds as the material is cooled down to room temperature.3,16 In some other cases, however, the chemical interaction energies are so much lower than the configurational entropy that the critical temperature is far below the melting point, approaching or even below room temperature, and a solid-solution again forms at high temperatures, but it does not separate into a mixture of pure components and compounds because the critical temperature is too low for significant atomic diffusion to take place.3,16 To put it another way, solutions are always thermodynamically stable at high temperatures; they sometimes decompose (by precipitation) during cooling to room temperature; but they also sometimes fail to decompose during cooling to room temperature and are instead kinetically stabilised at low temperatures. Whether decomposition takes place depends on the relative strength of the chemical-bonding interaction energies and the configurational entropy of the solution, with the balance between them determining the critical temperature at which decomposition takes place.
We now know37,38 that the original Cantor alloy CrMnFeCoNi has a critical temperature approximately equal to 800 °C, so that a single-phase fcc solid solution is thermodynamically stable at high temperatures T ≳ 800 °C; precipitation takes place very slowly during annealing for many years in a band of intermediate temperatures 450 °C ≲ T ≲ 750 °C; and a single-phase fcc solid solution is kinetically stable at low temperatures T ≲ 400 °C. Similar behaviour is expected for single-phase solid solutions in many different multicomponent materials, i.e. thermodynamic stability at high temperatures, kinetic stability at low temperatures, and (sometimes) decomposition into multiple phases at intermediate temperatures after very extended heat treatment.
In an ideal or regular solid solution, the atoms are distributed at random on the crystal lattice, and the probability of finding an atom of one of the components at any particular lattice point is proportional to its molar fraction, independent of the surrounding environment, i.e. independent of its neighbouring atoms. In a non-regular solid solution, however, the probability of finding an atom of one of the components at any particular lattice point is not proportional to its molar fraction, and is not independent of its neighbouring atoms. To be more specific, the probability pijn of two atoms of components i and j being nth near neighbours is given by pijn = xixj in an ideal or regular solid solution, and by pijn ≠ xixj in a non-regular solid solution, where xi and xj are the molar fractions of the i'th and j'th components, respectively.39 Non-ideal and non-regular solid solutions exhibit short-range order, with an increased probability for either association or disassociation of any particular pair of component atom types i and j in any particular near-neighbour shell n, driven by the non-zero pairwise, three-way, four-way, etc. interaction energies between the different atoms ωij, ψijk, χijkl, etc. that contribute to the heat of mixing and the corresponding deviation from a random distribution of atoms. Instead of using the probabilities pijn themselves, short-range order effects are usually quantified39 by the Warren–Cowley short-range order (SRO) parameters αijn = 1 − (pijn/xixj). In an ideal or regular solution, the SRO parameters αijn are all zero, and in a non-regular solution they are either positive or negative for, respectively, the disassociation or association of atoms of components i and j as nth near neighbours.
Many multicomponent high-entropy materials are ideal or very near-ideal single-phase solid solutions. The original Cantor alloy is often found to be ideal or very close to ideal, with as far as we can tell a completely random distribution of the five component atoms (Cr, Mn, Fe, Co and Ni) and with no detectable short-range order. This can be seen in the atom probe (AP) measurements in Fig. 5, showing an even distribution of each of the five component atoms in CrMnFeCoNi, with statistical analysis confirming a random pattern in each case.40 Similar results have also been found in other multicomponent materials, such as the single-phase rock-salt-structured mixed Rost oxide (MgCoNiCuZn)O. This can be seen41 in the high-angle annular dark-field (HAADF) image and associated high-resolution energy-dispersive X-ray (EDX) maps in Fig. 6, again showing an even distribution of each of the five cations (Mg, Co, Ni, Cu and Zn) in (MgCoNiCuZn)O. There are also many multicomponent high-entropy materials that are non-ideal single-phase solid solutions. The simpler equiatomic three-component single-phase fcc modified Cantor alloy CrCoNi is considerably less homogeneous than the original five-component Cantor alloy, and contains small ∼1 nm-sized domains of short-range ordering with enhanced Cr–Co and Cr–Ni and depleted Cr–Cr pairs in the 1st near-neighbour shell, i.e. with negative α1CrCo and α1CrNi, and positive α1CrCr, as shown in the high-resolution transmission electron microscope (HRTEM) images42 in Fig. 7. Another example can be seen in the atom-probe (AP) measurements43 in Fig. 8 from a quaternary single-phase bcc modified Senkov alloy ZrNbTaHf, showing short-range clustering of Zr and Hf atoms, i.e. negative α1ZrZr, α1HfHf and α1ZrHf and corresponding positive α1ZrNb, α1ZrTa, α1NbHf and α1TaHf, that builds up progressively with annealing time at 1800 °C (2073 K), beginning with small nm-sized regions of SRO and finally forming a 3-D phase-separated network on a scale of ∼10 nm.
![]() | ||
| Fig. 5 Atom probe measurements from a recrystallised specimen of the single-phase fcc Cantor alloy CrMnFeCoNi, showing a random atomic-scale distribution of all five components: (a) composition maps (each point is a single atom); (b) composition profiles for Mn and Fe; and (c) random Gaussian composition distributions for Mn and Fe (after Laurent-Brocq et al.40). Reproduced with permission from Elsevier from ref. 40. | ||
![]() | ||
| Fig. 6 HAADF-STEM atomic image of the single-phase multicomponent oxide (MgCoNiCuZn)O, and corresponding atomic-scale energy-dispersive X-ray (EDX) maps for each of the five cations, showing even intensities from a near-random distribution of cations on the cation sub-lattice (after Rost et al.41). | ||
![]() | ||
| Fig. 7 HRTEM of modified Cantor alloy CrCoNi after homogenisation: (a) and (b) filtered dark-field images using blue and red selected-area diffraction apertures, respectively; (c) diffraction pattern; (d) filtered Gaussian fitting showing nm-sized domains; and (e) domain size distribution (after Zhang et al.42). Reproduced with permission from Springer Nature from ref. 42. | ||
![]() | ||
| Fig. 8 AP composition profiles for Zr (purple) and Hf (green) from modified Senkov alloy ZrNbTaW: (a) and (b) as-cast, and after annealing at 1800 °C for (c) and (d) 6 hours, (e) and (f) 1 day, and (g) and (h) 4 days (after Maiti and Steurer43). Reproduced with permission from Elsevier from ref. 43. | ||
As mentioned in the last section, all high-entropy solid solutions must in principle decompose to form a multiphase structure on cooling to room temperature, driven by the interaction energies between the different atoms in the material. But many will not do so, when the interaction energies are relatively weak and the configurational entropy is relatively high, so that the critical temperature for decomposition is too low for significant atomic diffusion to take place.3,16,36 All high-entropy solid solutions must also in principle exhibit short-range ordering, typically at temperatures in the vicinity of the critical temperature for decomposition (with stable short-range order forming just above the critical temperature, and transitory short-range order forming as a precursor to precipitation just below the critical temperature), again driven by the interaction energies between the different atoms in the material.3,16,36 But similarly many will not do so, again when the critical temperature is too low for significant atomic diffusion to take place.3,16,36 Some multicomponent materials, like the original five-component Cantor alloy CrMnFeCoNo, have a low critical decomposition temperature and are, therefore, stable as a high-entropy near-ideal random solid solution, rarely exhibiting short-range order and decomposing only with difficulty after very prolonged heat treatment.37,38,40 Other multicomponent materials, however, like the three-component modified Cantor alloy CrCoNi or the four-component modified Senkov alloy ZrNbTaHf, have a high critical decomposition temperature and are, therefore, relatively unstable, frequently exhibiting significant short-range order and decomposing easily and rapidly during cooling to room temperature or after relatively short subsequent heat treatment.42,43
Consider a general lattice with a single atom at each lattice point, and with n1, n2 and n3 first, second and third near neighbours respectively. The cluster size of each atom together with its first near neighbours is n1 + 1 atoms, increasing to n2 + n1 + 1 atoms and then to n3 + n2 + n1 + 1 atoms including second and then third as well as first near neighbours. The number of different clusters N1 of n1 + 1 atoms, N2 of n2 + n1 + 1 atoms, and N3 of n3 + n2 + n1 + 1 atoms in a multicomponent equiatomic single-phase material with a random (ideal or regular) arrangement of c components is given22 by the law of permutations with repetition:||||
| N1 = cn1+1 |
| N2 = cn2+n1+1 |
| N3 = cn3+n2+n1+1 |
These are orientation-dependent cluster numbers (i.e. equivalent clusters with a different orientation are regarded as different), as is appropriate when we are dealing with properties that depend on orientation, such as dislocation slip, diffusion or magnetisation, when the material is responding to an oriented vector of applied stress, concentration gradient or magnetic field, respectively. If cluster orientation is not significant, however, the number of different clusters is smaller because of the number of self-similarity operations s associated with the crystal symmetry:45,46
| N1 = (cn1+1)/s |
| N2 = (cn2+n1+1)/s |
| N3 = (cn3+n2+n1+1)/s |
Table 6 shows the numbers of oriented and non-oriented local atomic clusters N1, N2 and N3 for multicomponent equiatomic single-phase fcc Cantor alloys with c components, taking for fcc n1 = 12, n2 = 6 and n3 = 24, so the cluster sizes are 13, 19 and 43 atoms, respectively.3,44 Table 7 shows equivalent numbers for multicomponent equiatomic single-phase bcc Senkov alloys, taking for bcc n1 = 8, n2 = 6 and n3 = 12, so the cluster sizes are 9, 15 and 27 atoms, respectively.3,44 The same results are shown graphically in Fig. 9 and 10 for fcc and bcc respectively. It is obvious from Tables 6 and 7 and Fig. 9 and 10 that the numbers of different local atomic clusters in multicomponent single-phase materials are very large indeed. Many but not all of the properties of a material are determined by interactions between first and second near-neighbour atoms, and many but not all are orientation-dependent, so we concentrate here on the number of oriented, first and second near-neighbour, local atomic clusters, which is just under twenty trillion (1.9 × 1013) for the original single-phase fcc Cantor alloy CrMnFeCoNi and just over thirty billion (3.1 × 1010) for the original single-phase bcc Senkov alloy, both with five components. With six components, the number of oriented, first and second near-neighbour, local atomic clusters increases to just over six hundred trillion (6.1 × 1014) for fcc Cantor alloys and almost half a trillion (4.7 × 1011) for bcc Senkov alloys; and with eight components, it is well over a quadrillion (1.4 × 1017) for fcc Cantor alloys and over thirty trillion (3.5 × 1013) for bcc Senkov alloys. The number of local atomic clusters is, of course, very much larger again if we include third near neighbours and/or increase the number of components to (say) ten or more, as shown in Tables 6 and 7 and Fig. 9 and 10. The number in all cases will be somewhat reduced, of course, if there is significant short-range ordering.
| Number of components c | Oriented clusters | Non-oriented clusters | ||||
|---|---|---|---|---|---|---|
| 1st | 2nd | 3rd | 1st | 2nd | 3rd | |
| 3 | 1.6 × 106 | 1.2 × 109 | 3.3 × 1020 | 6.4 × 104 | 4.8 × 107 | 1.3 × 1019 |
| 4 | 6.7 × 107 | 2.7 × 1011 | 7.7 × 1025 | 2.7 × 106 | 1.1 × 1010 | 3.1 × 1024 |
| 5 | 1.2 × 109 | 1.9 × 1013 | 1.1 × 1030 | 4.8 × 107 | 7.6 × 1011 | 4.4 × 1028 |
| 6 | 1.3 × 1010 | 6.1 × 1014 | 2.9 × 1033 | 5.2 × 108 | 2.4 × 1013 | 1.2 × 1032 |
| 8 | 5.5 × 1011 | 1.4 × 1017 | 6.8 × 1038 | 2.2 × 1010 | 6.0 × 1015 | 2.7 × 1037 |
| 10 | 1013 | 1019 | 1043 | 4.0 × 1011 | 4.0 × 1017 | 4.0 × 1041 |
| 20 | 8.2 × 1016 | 5.2 × 1024 | 8.8 × 1055 | 3.3 × 1015 | 2.1 × 1023 | 3.5 × 1054 |
| 50 | 1.2 × 1022 | 1.9 × 1032 | 1.1 × 1073 | 4.8 × 1020 | 7.6 × 1030 | 4.4 × 1071 |
| Number of components c | Oriented clusters | Non-oriented clusters | ||||
|---|---|---|---|---|---|---|
| 1st | 2nd | 3rd | 1st | 2nd | 3rd | |
| 3 | 2.0 × 104 | 1.4 × 107 | 7.6 × 1012 | 8.0 × 102 | 5.6 × 105 | 3.0 × 1011 |
| 4 | 2.6 × 105 | 1.1 × 109 | 1.8 × 1016 | 1.0 × 104 | 4.4 × 107 | 7.2 × 1014 |
| 5 | 2.0 × 106 | 3.1 × 1010 | 7.5 × 1018 | 8.0 × 104 | 1.2 × 109 | 3.0 × 1017 |
| 6 | 1.0 × 107 | 4.7 × 1011 | 1.0 × 1021 | 4.0 × 105 | 1.9 × 1010 | 4.0 × 1019 |
| 8 | 1.3 × 108 | 3.5 × 1013 | 2.4 × 1024 | 5.2 × 106 | 1.4 × 1012 | 9.6 × 1022 |
| 10 | 109 | 1015 | 1027 | 4.2 × 107 | 4.2 × 1013 | 4.2 × 1025 |
| 20 | 5.1 × 1011 | 3.3 × 1019 | 1.3 × 1035 | 2.0 × 1010 | 1.3 × 1018 | 5.2 × 1033 |
| 50 | 2.0 × 1015 | 3.1 × 1025 | 7.5 × 1045 | 8.0 × 1013 | 1.2 × 1024 | 3.0 × 1044 |
The number of local atomic clusters is very large indeed, and this has an important effect on the spatial consistency of the properties of multicomponent high-entropy single-phase solid solutions such as the fcc Cantor and bcc Senkov alloys.3,44 Taking the cube root of the number of local atomic clusters gives the size of a piece of the material sufficiently large to include all possible local atomic configurations, i.e. a piece of the material big enough to average reasonably over all the different local atomic environments and, therefore, big enough to represent fully the material and its properties. The linear dimension of such a piece of material is
clusters or ∼8 μm for oriented, first and second near-neighbour clusters in the original single-phase fcc Cantor alloy, and
clusters or ∼1 μm in the original single-phase bcc Senkov alloy, in both cases taking the atomic separation and, therefore, the cluster size as ∼0.3 nm. In other words, the properties of the original fcc Cantor and bcc Senkov alloys vary from grain to grain if their polycrystalline grain size is below ∼8 μm or ∼1 μm, respectively. To put it another way, the grain size needs to be above ∼8 μm or ∼1 μm, respectively, to have a material with consistent properties. And for equiatomic ten-component single-phase fcc Cantor or bcc Senkov alloys, the grain size needs to be above
clusters ≈ 0.7 mm or
clusters ≈ 30 μm, respectively.
This is clearly a very different situation from that which is found in conventional materials consisting of either a single component or a single main component with one or more dilute alloying additions. The extremely large number of different local atomic environments and atomic clusters in multicomponent single-phase fcc and bcc solid solutions such as the Cantor and Senkov alloys (and in other crystal structures such as multicomponent intermetallic and ceramic compounds) plays, therefore, an important role in material properties that depend strongly on local atomic interactions, such as vacancy migration and diffusion, or dislocation slip and plastic flow.3,44 The extremely large number of different local atomic environments and atomic clusters in these materials also makes it extremely difficult, in fact almost impossible, to determine their structure and properties with any degree of confidence by fundamental techniques such as ab initio molecular dynamics or quantum mechanical modelling that are limited to no more than a thousand or two atoms at best.
, where summation is over all the components i = 1 to n, xi = 0.2 is the mole fraction of the i'th component, and n = 5 is the number of components. The values of Δr, δmax and δav remain unchanged when Cu is included in a modified CrMnFeCoNiCux Cantor alloy, but they are somewhat larger when Al instead of Cu is included in a modified AlxCrMnFeCoNi Cantor alloy: Δr = 18 pm, δmax = 14.4, and δav = 1.4%, for the maximum Al solubility with xAl = 8% and xi≠Al = 18.4%. We might expect, therefore, to find lattice distortions on a scale of somewhat under ∼1 pm or 1% on average, rising to a maximum of ∼3 pm or ∼3% in the original Cantor alloy, and a bit higher in some of the modified Cantor alloys, such as those containing Al. These calculations are rather approximate since they ignore effects such as short-range order, non-ideality, relaxation, and electronic distortions. More detailed analyses have used measured lattice parameters, different definitions of atomic misfit and lattice distortion, and ab initio calculations, but all lead to similar predicted local lattice distortions48–51 of Δr ≈ 3.3–6.6 pm.
| Component | Goldschmidt atomic radius r (pm) |
|---|---|
| Cr | 128 |
| Mn | 127 |
| Fe | 126 |
| Co | 125 |
| Ni | 125 |
| Al | 143 |
| Cu | 128 |
Lattice distortions have been measured in multicomponent solid solutions via synchrotron X-ray diffraction (XRD),48 neutron diffraction (ND)52 and extended X-ray absorption fine-structure (EXAFS) analysis.49,53 The measurements are difficult because of the need to separate static displacements caused by lattice distortions, dynamic displacements caused by thermal vibrations, and bulk lattice strain. Sophisticated fitting software is used with high-quality diffraction data to separate out different effects within the measured broadening of diffraction peaks. Okamoto et al.48 used single-crystal synchrotron X-ray diffractometry to measure atomic displacements in the original Cantor alloy of 4.8 and 7.7 ± 0.5 pm at 25 K and 300 K, respectively, indicating a static displacement (at 25 K) of Δr ≈ 4.8 pm and a thermal displacement (at 300 K) of 7.7 − 4.8 = 2.9 pm. Owen et al.52 used neutron powder diffractometry to measure partial distribution functions in the original Cantor alloy, with different peak widths between the Cantor alloy and pure Ni corresponding to static distortions of Δr ≈ 2 ± 0.5 pm and the underlying peak width giving thermal distortions of ∼18 ± 0.5 pm. Oh et al.49 used EXAFS spectra from the original Cantor alloy to obtain mean elemental distortions of +0.1%, +0.5%, −0.1%, −0.4% and −0.1% for Cr, Mn, Fe, Co and Ni respectively,*** correlating roughly with the Goldschmidt atomic radii in Table 8, and with maximum distortions up to ∼3% of rav (=126.2 pm), i.e. Δr ≈ 3.8 pm.
Table 9 shows experimental and calculated results for local distortions Δr in the original fcc Cantor alloy CrMnFeCoNi. In conclusion, there are small but significant local lattice distortions that fluctuate randomly over large distances (because of the enormous number of different local atomic cluster configurations), with an average value of just under 1 pm (corresponding to just under ∼1% strain), reaching a maximum value of ∼2–6 pm (corresponding to ∼2–6% strain). Similar measurements and calculations indicate, as expected, somewhat smaller lattice distortions in simpler fcc Cantor alloy compositions such as CrFeCoNi,50,53–55 but somewhat larger lattice distortions in modified Cantor alloy compositions containing additional components such as VFeCoNi56 and FeNiCoCrPd.55
| Method | Maximum local atomic distortion Δr (pm) | |
|---|---|---|
| Calculations | Goldschmidt radii | 3.0 |
| Ab initio | 3.3–6.6 | |
| Experiments | Synchrotron XRD | 4.8 |
| Neutron diffraction | 2.0 | |
| EXAFS | 3.8 |
Table 10 shows 12-coordinated metallic Goldschmidt atomic radii r for some of the components that have been used to manufacture multicomponent single-phase bcc Senkov alloys.47 The maximum difference in atomic size amongst the five different components in the original Senkov alloy VNbMoTaW is Δr = rNb − rV = 10 pm, the corresponding maximum atomic misfit is δmax = (rNb − rV)/rV = 7.4%, and the root-mean-square (RMS) average atomic misfit is
%. The values of Δr, δmax and δav remain unchanged when either Al or Ti are included in modified AlxVNbMoTaW or TixVNbMoTaW Senkov alloys, respectively, but they are somewhat larger again when either Zr or Hf is included in modified VNbZrxMoTaW or VNbMoHfxTaW Senkov alloys, respectively, for an equiatomic six-component alloy with xi = 16.67%. We might expect, therefore, to find lattice distortions on a scale of ∼4 pm or ∼4% on average, rising to a maximum of ∼7–10 pm or ∼10% in the original Senkov alloy, and still higher in some of the modified Senkov alloys, such as those containing Zr or Hf. These calculations are again rather approximate since they ignore effects such as short-range order, non-ideality, relaxation, and electronic distortions. Other analyses of the lattice distortion have again used ab initio calculations, leading to similar predicted local lattice distortions of Δr ≈ 5–15 pm.50
| Component | Goldschmidt atomic radius r (pm) |
|---|---|
| V | 135 |
| Nb | 145 |
| Mo | 145 |
| Ta | 145 |
| W | 135 |
| Al | 143 |
| Ti | 140 |
| Zr | 155 |
| Hf | 155 |
Zou et al.57 used single-crystal X-ray diffractometry to measure average atomic displacements in the original Senkov alloy, indicating static and dynamic lattice distortions of Δr ≈ 7.4 and 6.1 ± 0.5 pm, respectively. Lattice distortions on a scale of tens of pm have also been observed by Zou et al.57 in the same alloy via high-resolution transmission electron microscopy (HRTEM), with [100] zone axis images containing distorted {110} planes, as shown in Fig. 11. Guo et al.58 used a combination of synchrotron X-ray diffractometry (XRD) and time-of-flight (TOF) neutron diffractometry to obtain radial distribution functions from a modified ternary Senkov alloy ZrNbHf, with an overlap of the first and second near-neighbour peaks corresponding to lattice distortions of ∼9.5 pm, and similar combined XRD and ND results for a quaternary modified Senkov alloy ZrNbTaHf showed static and dynamic thermal distortions of 13.7 and 2.4 pm, respectively.59
![]() | ||
| Fig. 11 HRTEM images of the equiatomic quaternary NbMoTaW bcc Senkov alloy: (a) bright-field (BF) image with [100] zone axis; (b) corresponding inverse fast-Fourier transform; (c) and (e) enlarged images of the boxes in (b); and (d) and (f) traces corresponding to (c) and (e) to show the lattice distortions (after Zou et al.57). Reproduced with permission from Elsevier from ref. 57. | ||
Table 11 shows experimental and calculated results for maximum local distortions Δr in bcc Senkov alloys. In conclusion, there are again significant local lattice distortions in the bcc Senkov alloys, somewhat larger than those found in the fcc Cantor alloys, that again fluctuate randomly over large distances (because of the enormous number of different local atomic cluster configurations), with average values of ∼4 pm (corresponding to ∼4% strain), reaching maximum values of ∼5–20 pm (corresponding to >5% strain).
| Method | Alloy | Maximum local atomic distortion Δr (pm) | |
|---|---|---|---|
| Calculations | Goldschmidt radii | VNbMoTaW | 10.0 |
| AlVNbMoTaW | |||
| TiVNbMoTaW | |||
| Goldschmidt radii | VNbZrMoTaW | 20.0 | |
| VNbMoHfTaW | |||
| Ab initio | TiVNb | 5–15 | |
| AlTiVNb | |||
| TiZrNbTaHf | |||
| Experiments | XRD | NbMoTaW | 4.8 |
| Synchrotron XRD & ND | ZrNbHf | 9.5 | |
| Synchrotron XRD & ND | ZrNbTaHf | 13.7 |
It is highly likely that similar substantial local atomic strains are to be found in all multicomponent high-entropy single-phase solid-solutions, extending over large distances, and influencing many of their properties.
Atomic motion in a concentrated multicomponent high-entropy solid solution, such as an fcc Cantor alloy or a bcc Senkov alloy, is more complex because of the wide variety of different local atomic structures and associated lattice distortions surrounding the vacancies, and a correspondingly wide range of different vacancy formation and migration energies ΔEv and ΔEm, depending upon where the vacancy is and what exactly are its surrounding atoms.3,44 There is an enormous number of different local atomic clusters surrounding each individual type of atom in a multicomponent solid solution (as discussed previously and as shown in Tables 6 and 7 and Fig. 9 and 10): and there is a similarly enormous number of different local atomic clusters surrounding the vacancies in a multicomponent solid solution or, to put it in another way, there is an enormous number of different vacancy structures.3,44
There is little or no lattice distortion in a single-component pure material, and the variation of energy as an individual vacancy hops from one lattice site to another consists of a series of identical jumps, as shown schematically in Fig. 12, from identical wells at each of the lattice sites, all with the same well energy Ew, over identical saddle points midway between each of the lattice sites, all with the same saddle-point energy Es. In a concentrated multicomponent high-entropy solid solution, however, the situation is very different, because of the large number of vacancy structures and associated local lattice distortions at different lattice points, the corresponding range of vacancy formation and migration energies ΔEv and ΔEm, and the corresponding spread of well and saddle-point energies Ew and Es. The variation in energy of an individual vacancy as it hops from one lattice site to another consists instead, therefore, of a series of variable jumps, as also shown schematically in Fig. 12, from variable wells at the different lattice sites, all with different well energies Ew, over variable saddle points midway between each of the lattice sites, all with different saddle-point energies Es.
The overall rate of atomic diffusion and, therefore, the diffusion coefficient in concentrated multicomponent high-entropy solid-solution materials is affected in a variety of ways44,62 by the variation in vacancy formation and migration energies ΔEv and ΔEm, and the corresponding spread of well and saddle point energies Ew and Es. High-energy barriers in some places make it particularly difficult for a vacancy to hop from one site to the next; low energy barriers in other places make it particularly easy for a vacancy to hop from one site to the next; and a vacancy can be trapped in some places, hopping backwards and forwards between adjacent or near-adjacent lattice points, or going round and round in circles. The overall impact of these different effects is, not surprisingly, quite complex, and strictly speaking, diffusion is no longer truly random spatially, since the lattice points are no longer all identical. We have not yet made much experimental or theoretical progress towards developing a detailed understanding of this complex range of different vacancy behaviours and the resulting overall rates of atomic diffusion. Thomas and Patala62 calculated the expected spread of well and saddle-point energies Ew and Es for vacancies hopping between adjacent lattice sites at 1000 °C in the original five-component single-phase fcc Cantor alloy CrMnFeCoNi, using a nudged elastic band (NEB) molecular dynamics (MD) method with a modified embedded atom method (MEAM) potential within the Sandia Labs LAMMPS software. The diffusion coefficient D was found to be fairly sensitive to the spread of energy values during the hopping process, varying between a half and five times a reference diffusion coefficient D* in an equivalent material with the same barrier energy of 0.81 eV, but with constant well and saddle-point energies. The exact results obtained via these calculations should be treated with considerable caution (as with other modelling results), since they average over only about three thousand vacancy hops, well below the total number of almost twenty trillion different vacancy structures (including first and second near neighbours). Nevertheless, the results show clearly that diffusion is slower when there is a wide spread of well energies (high σw) because vacancies can become trapped in the lowest energy wells, but is faster when there is a wide spread of saddle-point energies (high σs) because vacancies can take migration paths that simply avoid the highest energy barriers.
There has been considerable discussion in the previous literature about whether or not diffusion is slower in multicomponent high-entropy solid-solution single phases relative to pure materials,63–67 somewhat confused by the difficulties of deciding what are appropriate comparison materials, and the complexity of measuring diffusion coefficients from multicomponent material diffusion couples. There is, however, little doubt that diffusion is quite slow in many but not all cases,3,63,64 often by up to about a half or one order of magnitude, with correspondingly slow diffusion-controlled processes such as precipitation and recrystallisation,3,16,24 almost certainly caused by the wide variety of local atomic structures as described above. Clearly we need to develop more detailed theories of diffusion in multicomponent high-entropy solid-solution materials such as the fcc Cantor and bcc Senkov alloys, based on non-Fickian atomic percolation through complex structural landscapes with varying lattice-point and saddle-point energies.
| Δτ = τo + τwh + τss + τgb + τph |
Dislocation motion in a concentrated multicomponent high-entropy solid-solution such as an fcc Cantor alloy or a bcc Senkov alloy is more complex, because of the wide variety of different local atomic structures and associated lattice distortions surrounding atoms along the length of the core of a dislocation3,44 (as discussed previously and as shown in Tables 6 and 7 and Fig. 9 and 10), and a correspondingly wide variation in the dislocation core energy and line tension Ec and ΔEd. The atomic structure varies widely from point to point along a dislocation line, and also from time to time at any given point on a dislocation line as it moves under the action of an applied stress.3,44 Once again, we have not yet made much experimental or theoretical progress towards developing a detailed understanding of this complex range of different local dislocation structures, the resulting dislocation motions and interactions, and, therefore, the overall plasticity in multicomponent high-entropy materials.
Dislocations in fcc Cantor alloys are in some ways found to be similar to those observed previously in pure fcc metals and dilute fcc binary alloys with a relatively low stacking fault energy.70–72 Unlike pure fcc metals and dilute binary fcc alloys, however, the dislocations in multicomponent single-phase fcc Cantor alloys are wavy rather than straight on a near-atomic scale, with a wide variation along the dislocation line in the separation of the two Shockley partials (and, therefore, the width of the stacking fault between them and corresponding stacking fault energy), and a much higher shear stress needed for slip along the {111} planes.70–76 An example of a wavy dislocation and measurements of the varying partial separation and corresponding stacking fault energy73 are shown in Fig. 13 for the original fcc Cantor alloy CrMnFeCoNi. These effects are all caused by the variation in local atomic structures along the dislocation lines as described above, with the associated local lattice strains acting as pinning centres for the dislocations. Okamoto et al.73 used high-resolution weak-beam imaging to make a large number of measurements of the separation between partials along the length of dislocation lines in the original fcc Cantor alloy CrMnFeCoNi, which were found to be variable in the range d = 3–8 nm, corresponding to stacking fault energies in the range γsf = 25–35 mJ m−2. Similar wavy dislocations with varying separation of partials along the dislocation line have also been seen in a number of multicomponent single-phase fcc Cantor alloys using a variety of transmission electron microscope techniques and also via atomistic ab initio modelling techniques.74–79
![]() | ||
| Fig. 13 (a) High-resolution weak-beam image of a ½〈110〉 dislocation in the original fcc Cantor alloy CrMnFeCoNi showing a wavy dislocation line and separation into two ⅙〈112〉 Shockley partials; and (b) partial separation distance d in a large number of dislocations as a function of dislocation orientation θ, ranging from pure screw (θ = 0°) to pure edge (θ = 90°) (after Okamoto et al.73). | ||
Dislocation structures in bcc Senkov alloys are also found to be similar in some ways to those observed previously in pure bcc metals and dilute bcc binary alloys.80–83 Unlike pure bcc metals and dilute binary bcc alloys, however, the dislocations in multicomponent single-phase bcc Senkov alloys are, like the fcc Cantor alloys, wavy rather than straight on a near-atomic scale, and require a much higher shear stress for slip along the different planes.80–83 These effects are again, like the fcc Cantor alloys, caused by the variation in local atomic structures along the dislocation lines, with the associated local lattice strains acting as pinning centres for the dislocations.
There has been some success in explaining the strength of multicomponent high-entropy solid-solution materials such as the fcc Cantor alloys and bcc Senkov alloys, based on averaging the pinning effects of varying local atomic structures along the dislocation lines.84–89 Clearly, we still need to investigate experimentally what must be a much wider range of varying dislocation behaviour, and thus develop a more detailed theory of overall dislocation dynamics, as dislocations move about and interact within a complex landscape of varying local atomic structure and associated varying dislocation energies.
As already mentioned, a much higher shear stress is needed for dislocation slip in multicomponent high-entropy solid-solution single-phase materials such as the fcc Cantor alloys and bcc Senkov alloys. This is caused by the variation in local atomic structures along the dislocation lines, with the associated local lattice strains acting as pinning centres for the dislocations.3,44 Effectively this corresponds to raising the lattice friction stress because of the dense number of pinning centres created by a high concentration of many different types of solute atoms. The shear flow stress can be rewritten as:3,44
and with an outstanding tensile strength exceeding 1 GPa”. Much research is now underway to improve even further the mechanical properties of these multicomponent high-entropy materials via additional strengthening mechanisms such as reducing the grain size, work hardening via severe plastic deformation, and alloying to develop precipitation hardening.3,16,44,91–98
![]() | ||
| Fig. 14 Schematic Ashby map of fracture toughness Kc versus yield strength σy for multicomponent high-entropy alloys compared with high-strength steels, nickel superalloys, hard ceramics, and metallic and oxide glasses (redrawn14 from Gludovatz et al.90 and Li et al.91). | ||
Direct observation of grain-boundary structures is not easy, even in pure materials and simple dilute binary alloys, because of the complex imaging conditions created by the atomic or molecular disorganisation in grain-boundary regions,46,68 complicating the use of standard high-resolution imaging techniques such as transmission electron microscopy (TEM), X-ray diffraction (XRD) and atom probe tomography (AP). And the problems associated with the direct observation of grain-boundary structures are only made worse by the chemical complexity found in multicomponent high-entropy materials. Not surprisingly, therefore, there have been relatively few attempts to investigate multicomponent high-entropy grain-boundary structures by direct observation,3,16,99 and there have also been relatively few attempts to use atomistic modelling techniques,3,16,100–102 which are more straightforward technically, but always suffer, as remarked previously, from not being able to handle anywhere near the very large number of different local atomic clusters that are present in multicomponent high-entropy materials.
A common method of controlling grain size is by using a mixture of deformation and heat treatment to homogenise and then recrystallise the material.103,104 There have been quite a few studies of recrystallisation, grain growth and texture in the original equiatomic five-component single-phase fcc Cantor alloy CrMnFeCoNi105–111 and its equiatomic single-phase fcc quaternary (CrFeCoNi, MnFeCoNi and CrMnCoNi) and ternary (CrFeNi, CrCoNi, MnFeNi, MnCoNi and FeCoNi) subsystems.110,112–114 Recrystallisation and grain growth in these multicomponent materials are found to be considerably slower than in fcc pure metals and binary alloys, essentially caused by the wide range of local atomic structures and corresponding lattice strains, as shown in Fig. 6 and Table 9, slowing the rates of atomic diffusion during heat treatment.105–107,110,113 Recrystallisation temperatures correspond typically to homologous temperatures relative to the melting point of Tr/Tm = 0.6–0.75, depending on the alloy composition and the extent of deformation,113 considerably higher than the more typical values in conventional pure metals and binary alloys of Tr/Tm = 0.3–0.5. The resulting final recrystallised grain sizes are, as a consequence, very small and slow to coarsen, often no larger than 1–5 μm after 1 hour at annealing temperatures of 700–800 °C (973–1073 K)105–107,110,113 and much smaller than found in fcc nickel,114 independent of the extent of the initial deformation treatment.105–107,110,113
All of our understanding of the electronic behaviour of crystalline materials comes from being able to find approximate solutions to the time-dependent Schrödinger equation:121–123
, ℏ = h/2π, h is Planck's constant, Ψ(ri) is the wave function for the set of constituent fundamental particles (electrons, protons and neutrons) i at positions ri in the material, E is the energy of the atom and
is the Hamiltonian given by:
and
are kinetic and potential energy operators given respectively on the right-hand side by the sum of the individual kinetic energies of all the constituent fundamental particles i and the sum of the individual electrostatic attractions or repulsions of all pairs of constituent fundamental particles i and j, mi and ri are the mass and position of the i'th particle, e is the charge on the electron, zi = −1, +1 and 0 for electrons, protons and neutrons, respectively, and ∇i2 = ∂/∂xi2 + ∂/∂yi2 + ∂/∂zi2 is the Laplacian operator for the i'th particle. Solving this equation requires a large number of simplifying approximations,121–123 including the following: concentrating on stationary states to remove time-dependence, i.e. solving the simpler time-independent (instead of time-dependent) Schrödinger equation
;122,123 using the Born–Oppenheimer approximation124 to concentrate on the valence electrons only, by assuming the nucleus and inner electrons of each atom form a relatively massive clamped central ion that is effectively fixed in space; using the Hohenberg–Kohn theorem125 to concentrate exclusively on the ground state of the electrons, with the energy a function then of the electron density ρ(r) rather than the set of all electron positions {ri}, which is why the methodology is often called density functional theory (DFT); treating the valence electrons as independent by re-writing the overall wave function
i.e. as a product of a set of individual wave functions ψi(ri), one for each valence electron i;122,123 and using Bloch's theorem,126 to re-write further the set of wave functions ψi(ri) as a simpler set of plane waves ϕi(ri) = eikri modulated by a function ui(ri) = ui(ri + R) which has the periodicity of the crystal lattice:| ψi(ri) = ϕi(ri)ui(ri) = eikriui(ri) |
The mechanical strains in multicomponent high-entropy materials act as pinning points, making atomic diffusion difficult by restricting vacancy migration and dislocation slip. Similarly, the electronic strains in multicomponent high-entropy materials act as scattering points, making electronic motion difficult and leading to relatively low electrical conductivity.3,44,127–133 Because of the disruption of lattice periodicity described above, it is very difficult to determine more complex electronic behaviour, and we clearly need more experimental and theoretical studies to explore the resulting electronic, optical and magnetic properties.
The techniques that have been employed to explore multicomponent phase space have included the following: semi-empirical multicomponent Hume-Rothery rules;3,16,18,134–136 thermodynamic (Calphad) modelling;3,16,18,137–140 atomistic modelling using density functional theory (DFT), molecular dynamics (MD) and Monte-Carlo (MC) methods;3,16,18,123,141 and statistical machine-learning (ML) algorithms such as support vector machines, decision trees and neural networks.3,16,18,142–147 Overall, there has been, therefore, quite a large number of detailed studies using multicomponent Hume-Rothery rules,148–162 thermodynamic modelling,163–172 atomistic modelling173–194 and machine learning,195–212 all aiming to predict the structures and properties of a variety of different multicomponent and high-entropy materials. In summary, all this work shows that valuable information can be obtained, but that the results are somewhat variable and not in general sufficiently accurate for reliable ab initio predictions of either the structures or properties of new multicomponent and high-entropy materials. Hume-Rothery methods try to correlate material structures with parameters calculated from the atomic radii, valencies, electronegativities, bulk modulii, melting points and enthalpies and entropies of the component atomic species.3,16,18,134–136 There are, not surprisingly, clear but rough correlations, with solid solutions being favoured for materials where the components have similar values of atomic radius, valency and electronegativity, and where the heats of mixing are small and the entropies of mixing are large, but the results are not good enough to guarantee that ab initio predictions are correct.3,18 Calphad-style thermodynamic modelling is based on the fundamental equation discussed previously for the Gibbs free energy of a material:3,16,18,137–140
Footnotes |
| † This paper uses the term alloy to mean any mixture of more than one starting material or component. The term alloy is used most commonly to refer to metallurgical materials, but here it is used more generally for any mixture of materials, whether metallurgical, ceramic, semiconductor, polymeric or of any other type. Since all materials are alloys, as explained in the main text, the two terms material and alloy are fairly synonymous but not entirely, since referring to different alloys usually means referring (only) to different compositions of the mixed components, whereas referring to different materials often means referring to different compositions, but can also mean referring to the same composition made under different manufacturing conditions and with different resulting microstructures. Thus, for instance, martensitic and ferritic steels with the same composition are the same alloy but different materials; and, similarly, amorphous and devitrified soda-lime glasses with the same composition are also the same alloy but different materials. |
| ‡ This paper draws a distinction between multicomponent and high-entropy materials. The term “multicomponent material” is taken to mean any material containing three or more components, i.e. with c ≥ 3, where c is the number of components in the material. This is logical and natural, in line with the normal definition of multiple as meaning more than one or two. The term high-entropy material is more restrictive and is taken to mean any multicomponent material, i.e. any material containing three or more components with c ≥ 3, where in addition the components are all or almost all in one or, at most, two random or near-random solid-solution phases. This is again more logical and natural, in line with the implied normal definition of high entropy as meaning a configurational entropy of mixing in the material sufficiently large to stabilise one or at most two random solid-solution phases, rather than forming a mixture of several or many stoichiometric, fixed-composition or near-fixed-composition compounds. This is different, but simpler and less arbitrary, than the common usage of the terms high-entropy material and medium-entropy material to mean any material containing high concentrations of more than five components, i.e. c ≥ 5, and between three and five components, i.e. 5 ≥ c ≥ 3, respectively. Overall the definitions used in this paper are, as already indicated, more logical and more natural than some common practice, and also more useful, since they carefully separate the different concepts of multiplicity of components and entropy in any given material. |
| § The term Cantor alloy means any multicomponent material with a single-phase random or near-random solid-solution face-centred cubic (fcc) structure. There are very many Cantor alloys with different compositions, as discussed later in the main text. |
| ¶ This paper adopts the convention of naming multicomponent and high-entropy materials by listing their components in ascending atomic number rather than alphabetically, i.e. the original Cantor alloy is represented as CrMnFeCoNi rather than CoCrFeMnNi. This has the advantage of emphasising more clearly the chemical nature of the different components. This paper also adopts the convention of naming multicomponent and high-entropy materials by using subscripts for the atomic percent rather than weight percent of the different components, but truncated by leaving out the subscripts when the components are present in equiatomic proportions, i.e. the original Cantor alloy is represented as CrMnFeCoNi, truncated from Cr20Mn20Fe20Co20Ni20. The paper deviates from the ascending atomic number convention when it contradicts a well-established usage. Thus, the well-known and important ordered body-centred cubic (bcc) multicomponent monoaluminide is represented as (CrMnFeCoNi)Al rather than Al(CrMnFeCoNi). |
| || In this paper, the term concentrated component is taken to mean that a linear dilute solution approximation is not likely to be valid, i.e. that the different atoms of the component, when added in solution to the material, are not so widely separated that they can be treated as affecting the material independently or one at a time. In a solid or liquid material, atoms exhibit strong interactions chemically when they are first or second near neighbours, so a concentrated solute addition is typically above about one atom in twenty, i.e. >5%, depending somewhat on the structure of the host material. |
| ** In this paper, the term high-entropy material is taken to mean that the configurational entropy of the material is large enough to suppress or largely suppress the formation of stoichiometric fixed-composition compounds, leading instead to the formation of one or sometimes two extended solid-solution phases in which the atoms are randomly or near-randomly distributed across the lattice or sublattice points in the crystal structure of the material. |
| †† This does not mean that all single-phase or multiple-phase structures occupy large regions in multicomponent space, but we have certainly discovered that many of them do. |
| ‡‡ In this paper, the configurational entropy of mixing ΔScmix is (for convenience, though somewhat unconventionally) taken to be determined by (only) the configurational entropy of mixing different atoms in solid-solution phases in the material; and other contributions to the configurational entropy of mixing, such as the entropy of distributing defects throughout the material (again for convenience, though somewhat unconventionally) are absorbed into the excess entropy of mixing ΔSxsmix alongside other contributions to the excess entropy of mixing, such as, for instance, changes in disorder arising from changes in the vibrational, electronic or magnetic states of the atoms in the material. |
§§ In an ideal material, there are no chemical interactions between the different atoms in the material, the heat of mixing ΔHmix and excess entropy of mixing ΔSxsmix are both, therefore, equal to zero, and the ideal free energy of mixing is determined by the configurational entropy of mixing ΔGmix = −RTΔScmix = RT ln(1/c) for an equiatomic material. |
| ¶¶ An ideal solution is one in which the interaction energies between the different atoms in the material are zero, so the heat of mixing is zero and the atoms are completely randomly mixed; a near-ideal or regular solution is one in which the interaction energies between the different atoms in the material are not zero but are very small, so the heat of mixing is not zero but the atoms are still completely randomly mixed; a non-regular solution is one in which the interaction energies between the different atoms in the material are not zero and are not small, so the heat of mixing is not zero, the atoms are not completely randomly mixed, and the solution exhibits short-range order. |
| |||| We want to calculate how many ways we can pick n1 + 1 (or, including second or second and third near neighbours, n2 + n1 + 1 or n3 + n2 + n1 + 1) atoms from a total of c components. Let the atomic sites in the cluster be labelled 1, 2, 3, etc. up to n1 + 1 (or n2 + n1 +1 or n3 + n2 + n1 + 1). Different sets of the c components distributed across the different atomic sites are clearly different, but different arrangements of any given set of c components across the different atomic sites are also different. In other words, we need an ordered permutation rather than a non-ordered combination. And we need a permutation with repetition since each of the different component atoms can be picked more than once.22 |
| *** Positive and negative values represent increased and decreased bond lengths respectively, corresponding to local atomic expansion and compression strains, respectively. |
††† Notice that i in the exponential is the imaginary number: , and that elsewhere i is a label for different electrons: {i} = i, j, k, etc. |
| This journal is © The Royal Society of Chemistry 2025 |