You are on page 1of 36

Progress in Materials Science 120 (2021) 100754

Contents lists available at ScienceDirect

Progress in Materials Science


journal homepage: www.elsevier.com/locate/pmatsci

Multicomponent high-entropy Cantor alloys


B. Cantor *
Department of Materials, University of Oxford, Parks Road, Oxford OX1 3PH, UK
Brunel Centre for Advanced Solidification Technology (BCAST), Brunel University, Kingston Road, Uxbridge, London UB8 3PH, UK

A R T I C L E I N F O A B S T R A C T

Keywords: Multicomponent high-entropy Cantor alloys are single-phase and near-single-phase face-centred
Multicomponent materials cubic alloys, occupying an enormous region of multicomponent phase space. They were discov­
High-entropy alloys ered by Cantor and co-workers in the late 1970s/early 1980s, but have not been studied inten­
Cantor alloys
sively until the last decade or so. This review describes: the extensive range and complexity of
multicomponent phase space, including the prevalence of single (or relatively few) phases and the
paucity of intrinsically new multicomponent compounds; the thermodynamics of multicompo­
nent single-phase materials such as the Cantor alloys, and the extent to which they are stabilised
by a high configurational entropy; the multiplicity and complexity of local atomic configurations
and associated lattice strains in multicomponent single-phase high-entropy materials such as the
Cantor alloys; the effect of multiple atomic configurations and lattice strains on the rate of atomic
diffusion and on the creation and motion of dislocations; and the resulting excellent mechanical
properties, equal to and sometimes exceeding the very best high strength steels, nickel superalloys
and hard ceramics, with enormous potential for future further enhancement and optimisation.

1. Introduction

1.1. Materials and alloys

All materials are alloys. In other words they consist of mixtures of one or more different starting materials or components. In this
review we will use the words materials and alloys largely interchangeably.
We have used alloying to manufacture our materials since about 5000 years ago (~3000 BC). This was the end of the stone age,
when weapons and tools were made by mechanical shaping from naturally occurring stones, and the beginning of the bronze age [1],
when much better weapons and tools were made from bronze, i.e. copper alloyed with a few percent of tin. Smelting (reduction) of
copper ores was already in use to make copper ornaments and jewellery, but the key technological development was alloying with tin
to make a much stronger, castable material. The tradition of alloying to create our materials continued into the iron age [1], beginning
about 3000 years ago (~1000 BC), when the higher temperatures needed to smelt iron ores were achieved. This time the key tech­
nological development was the deformation and (in a later period) partial re-oxidation of the smelted cast iron to control the amount
and distribution of small amounts of carbon alloying addition, making even stronger weapons and tools. And the tradition of alloying
to create our materials has continued further into modern times with the silicon age, beginning in the latter half of the 20th century,

* Address: Department of Materials, University of Oxford, Parks Road, Oxford OX1 3PH, UK; and Brunel Centre for Advanced Solidification
Technology (BCAST), Brunel University, Kingston Road, Uxbridge, London UB8 3PH, UK.
E-mail address: briancantor11@gmail.com.

https://doi.org/10.1016/j.pmatsci.2020.100754
Received 3 July 2020; Received in revised form 20 September 2020; Accepted 8 October 2020
Available online 13 October 2020
0079-6425/© 2020 Published by Elsevier Ltd.
B. Cantor Progress in Materials Science 120 (2021) 100754

when the discovery of semiconductors underpinned the development of digital computing and telecommunication technologies, via
doping, i.e. micro-alloying, to obtain precisely controlled electrical properties in silicon chips [2].

1.2. Conventional alloying strategy

Conventional alloying strategy for developing materials has, for the last 5000 years, been to select the main component based on a
primary property requirement, and to use alloying additions to confer secondary properties. This strategy has led to the development of
many successful materials based on a single main component with a mixture of different alloying additions to provide a balance of
required in-service properties. Typical examples include: high temperature nickel superalloys, with nickel as the main component
selected for its high melting point, chromium added for corrosion resistance, and titanium and aluminium added for high strength [3];
corrosion-resistant stainless steels, based on iron as a main component for its formability, with chromium added for corrosion resis­
tance, and nickel to prevent brittleness [4,5]; and high strength aluminium alloys, using aluminium as a main component for its
lightness, with copper and magnesium additions for high strength [6].
Of course, some materials are very high purity, but even these materials are alloys, with important admixtures of impurities, for
example: gemstone diamonds are high purity carbon, typically > 99.5% purity, with up to 0.5% nitrogen, oxygen and hydrogen
impurities that control the colour and shape of the diamond crystals [7]; the nominal highest grade of gold for bullion, coins and
jewellery is 24 karat, with up to 0.1% of silver and copper impurities that prevent it being too soft (though the very highest purity gold
is 99.9999% (six nines) Canadian mint coinage) [8]; the highest-purity mild steel is typically ~99.95% (three nines) pure iron to make
it soft and highly formable, with ~0.05% levels of carbon and manganese that allow it to be hardened after forming to shape [5]; and
semiconductor-grade silicon is purified to 99.9999% (six nines), with parts per million of dopants such as boron (n-type) and phos­
phorus (p-type) to control accurately its semiconducting properties [9].
Conventional alloying strategy has led, therefore, to an enormous amount of theoretical and experimental knowledge about ma­
terials based on one component with relatively dilute alloying additions of one or two (or, occasionally, three or four) other com­
ponents. On the other hand, it has led to little or no knowledge about materials containing several main components in approximately
equal proportions. In summary, we can say that information and understanding is highly developed about materials close to the
corners and edges of a hyper-dimensional multicomponent phase diagram, but that very little is known about alloys in the centre of the
diagram.

1.3. Multicomponent alloying strategies

In 2004, two influential papers were published suggesting a new and completely different set of strategies for developing new
materials [10,11]. Independently, my own research group in the UK [10] in the 1980s, and a research group led by Professor Jien-Wei
Yeh in Taiwan [11] in the 1990s, had both begun to explore the manufacture and properties of alloys consisting of several principal
components in equal or near-equal proportions, i.e. alloys in the centre of multicomponent phase space, well away from the corners
and edges of multicomponent phase diagrams. These new materials are called multicomponent [10] or high-entropy [11] alloys or
materials (or, sometimes, multiple-principal-element, compositionally-complex or complex concentrated alloys or materials). Both pieces of
work were overturning millennia-old ideas of how to make materials, and both were achieved despite lack of support and widespread
disbelief in the value of such an unusual and innovative approach. Both papers were met at first with limited or no attention.
Nevertheless, by the end of the decade hundreds of papers were being published using this novel approach to manufacturing new
materials, and by the end of the next decade (i.e. by now) this had escalated to thousands of papers, as scientists worldwide saw the
enormous potential for finding and developing new materials with exciting and valuable new properties [12]. The two key papers by
Cantor et al. [10] and Yeh et al. [11] had a total of 8 citations in the first two years after their publication (2004 and 2005). Together
these two papers have now received a total of almost 5000 citations. Total publications and total citations in the field have now
reached almost 10,000 and almost 200,000 respectively, with both figures rising rapidly [12]. A recent review of the field [13] stated
that this new approach had already led to the development of “eight major new categories of materials”, despite having still “only
scratched the surface” of the vast potential for new materials.

1.4. Discovering multicomponent Cantor alloys

During the 1970s, I was working in the Materials Science section of the School of Engineering at the University of Sussex, and in
1979 and 1980 I suggested to my colleagues, students and research funders that it would be interesting to study multicomponent
materials, i.e. alloys made from equiatomic mixtures of many different components. I argued that no-one seemed to have considered
doing this before, and it would be bound to lead to some unusual new materials with potentially valuable properties. The uniform
response was negative, treating the proposal as risible, more of a joke than a realistic suggestion. Common comments were: “It’s a crazy
idea”; “Why would anyone want to do that?”; and “Do something more sensible”. In the end, I persuaded a young undergraduate
student, Alain Vincent, to study two alloys for his undergraduate project, one containing sixteen components in equal proportions and
the other containing twenty components in equal proportions, made by simple melting of the individual components in a crucible,
followed by natural air-cooling and solidification.
Both alloys solidified with a complex structure of several different phases, as shown for the twenty-component material in the
optical micrograph in Fig. 1. We were surprised, however, to discover by energy-dispersive X-ray analysis in a scanning electron
microscope that one of the phases consisted predominantly of five components – chromium, manganese, iron, cobalt and nickel - in

2
B. Cantor Progress in Materials Science 120 (2021) 100754

approximately equal proportions [10,14]. We were even more surprised when X-ray diffraction on a new alloy made up of just those
five elements in equal proportions showed that the material had a single-phase face-centred cubic (fcc) A1 structure. This was not
expected, given that only one of the five components (nickel) is fcc in its pure form at room temperature, with chromium and iron being
body-centred cubic (bcc) A2, cobalt being hexagonal close-packed (hcp) A3, and manganese having its own complex bcc α-Mn A12
structure [15]. The fcc single-phase alloy discovered by Alain Vincent with the composition1 CrMnFeCoNi or Cr20Mn20Fe20Co20Ni20 is
now well known as the Cantor alloy. Alain Vincent wrote up his final year undergraduate project and obtained his bachelor’s degree.
However, his results were not of sufficient quality to be published in the scientific literature, though his “Part II” undergraduate
dissertation is, I believe, lodged in the University of Sussex library [14]. I left Sussex at the beginning of the 1980s to work for General
Electric in the States, and did not at that time take the project further forward.
Many years later, in 1995, after returning to England to work in the Department of Materials at the University of Oxford, I
persuaded another undergraduate student, Peter Knight, to do his undergraduate project on the same topic. History repeated itself. He
reproduced the same results, extended them (as had Vincent) to other multicomponent materials consisting of five or more components
in equiatomic and non-equiatomic proportions, and also extended the manufacturing process to rapid solidification by melt spinning,
thus developing a range of modified Cantor alloys based on CrMnFeCoNi with additions of titanium, vanadium, copper, germanium
and niobium, showing that they also consist of simple single-phase fcc, single-phase bcc, or two-phase fcc + bcc precipitate or eutectic
structures. He then wrote up his dissertation and obtained his bachelor’s degree, but again the results were not of sufficient quality for
publication, though his Part II dissertation is similarly lodged in the University of Oxford Department of Materials Library [16]. Finally,
towards the end of the 1990s, one of the senior post-docs in my (by then very large) research group, Isaac Chang, agreed to repeat the
work yet again, in his spare time. He reproduced the previous results and did (at last!) an excellent and professional job. We presented
the results in 2002 at the Rapidly Quenched and Metastable Materials conference (RQ11) in Oxford [17] and the work was finally
published in 2004 in the research journal Materials Science and Engineering [10], fully 25 years after the first idea and 24 years after
the first results were obtained. We proposed, following my original speculation 25 years earlier, that such extended multicomponent
solid solution and solid solution + precipitate structures would have unusual and potentially valuable mechanical and other properties
[10].
Instead of using Vincent’s [14], Knight’s [16] and Chang’s [10,17] approach of making up equiatomic component mixtures and
then identifying the resulting phases, a research student of mine at Oxford, Ki-Buem Kim showed in the 2000s that another way of
accessing multicomponent phase space was to use equiatomic substitution, i.e. to take a known material of interest and then replace one
or more of the components by an equiatomic mix of chemically similar components. He studied amorphous (TiZrHf)1-x-y (NiCuAg)x Aly
alloys with x = 20–70at% and y = 10–20 at% [18-21], based on the well-known Zr40Cu50Al10 bulk metallic glass [22], using
equiatomic substitution for both Zr and Cu components. A wide range of the alloys were amorphous as-solidified, and crystallised on
subsequent heating to form an extended single MgZn2-type C14 Laves-phase field in multicomponent phase space. The same
equiatomic-substitution technique has been used more recently to manufacture a wide range of multicomponent or high-entropy
compounds, for example mixed monocarbides MC such as (TiZrNbHfTa)C [23], diborides MB2 such as (TiZrNbHfTa)B2 and (TiZr­
MoHfTa)B2 [24], and mono-oxides MO such as (MgZnCoNiCu)O [25], all of which also occupy large single-phase fields in multi­
component phase space. Shortly after Chang’s and Kim’s work, I left to become Vice-Chancellor of the University of York and later the
University of Bradford, both in the North of England, and again the project was not taken further forward.

1.5. Discovering high-entropy alloys

In the meantime a similar train of events was taking place in Taiwan. In the early/mid 1990s, Professor Jien-Wei Yeh developed
independently the idea of alloys containing multiple principal elements in equal or near-equal proportions [26]. He also had to struggle
to get his early research undertaken and completed [26]. After many years of effort, he finally published, also in 2004, in the research
journal Advanced Engineering Materials [11], concentrating on equiatomic alloys of chromium, iron, cobalt, nickel and copper with
varying aluminium content, i.e. Alx(CrFeCoNiCu) with x = 0–3. The alloys were again manufactured by simple melting and casting,
and again had surprisingly simple structures, with single-phase fcc for low Al content (x < 0.5 or about 8–9 at%), single-phase bcc for
high Al content x > 2.5 or about 33%), and mixed fcc + bcc for intermediate Al content in the vicinity of equiatomic AlCrFeCoNiCu.
Professor Yeh was the first to use the concept of high entropy to explain the surprising formation in multicomponent materials of
relatively few phases and very extended solid solubility [11], and he coined the term high-entropy alloys [11], which has since become
widely accepted and has been widely influential in promoting the field. The same year, 2004, he and his research colleagues published
a number of other papers on multicomponent high-entropy alloys [27-30], exploring the effect on microstructure and mechanical
properties of using different manufacturing methods, such as spray coating [27] and reactive DC sputtering [30], and varying alloy
concentration, with additions of boron [28], iron, titanium and vanadium [29]. They demonstrated a range of impressive mechanical
and other properties [27-30].
A few years later, Professor Yeh predicted that multicomponent materials would exhibit four “core” properties [31,32]:

1. high entropy;

1
Note that in this review we will identify the chemical composition of multicomponent alloys: (1) by listing the components according to their
atomic number, i.e. in ascending position in the periodic table; and (2) by using either atomic ratios or atomic percent, so an equiatomic five-
component alloy made from components A, B, C, D and E is described as either ABCDE or A20B20C20D20E20, whichever is most convenient.

3
B. Cantor Progress in Materials Science 120 (2021) 100754

Fig. 1. Optical micrograph showing a multiphase microstructure in the first equiatomic 20-component alloy
MgAlSiMnCrFeCoNiCuZnGeNbMoAgCdSnSbWPbBi.

2. lattice distortion;
3. sluggish diffusion; and
4. cocktail effects.

The cocktail effect had been proposed previously by Ranganathan [33], and is not really a separate property, more an observation
that in multicomponent systems we can expect to see emergent properties arising from substantial cross or synergistic effects from the
multiplicity of components [13]. The cocktail effect can be characterised as a warning not to expect that multicomponent material
properties can be easily approximated by a simple linear superposition of the effects of adding each component individually.
Unlike my own research group, Professor Yeh and his colleagues continued working on high-entropy alloys throughout the 2000s
and 2010s, making a string of further contributions and helping to develop the field into its current strong position.

1.6. The range of this review

A key discovery of the early work by Cantor et al [10] and Yeh et al [11] is that there is an enormous region in multicomponent
phase space consisting of materials with a single-phase or near-single-phase fcc structure, made up mostly (but not exclusively) of
transition metal components, loosely based around the original Cantor alloy CrMnFeCoNi. This region of Cantor alloys includes a wide
variety of (1) single-phase fcc materials made by a combination of addition to, variation from, replacement of and/or reduction of the
set of original Cantor alloy components, and (2) two-phase precipitate-reinforced fcc materials made by the addition of one or more
components just beyond the solubility limit (or limits) of the single-phase region. To take just a few individual examples from a
potentially enormous number of different alloys:

1. Addition: CrMnFeCoNiCu, i.e. addition of copper to make an equiatomic six-component single-phase fcc material [10], and
CrMnFeCoNiAlx with x < 0.5, i.e. addition of aluminium up to half the amount of the other components, or less than about 8–9 at%,
to make a non-equiatomic six-component single-phase fcc material [34];
2. Variation: CrMnFeCoxNi and CrMnFeCoNix with x = 0–2, i.e. variation of the amount of cobalt or nickel from 0–33.3 at% to make
non-equiatomic five-component single-phase fcc materials [35];
3. Replacement: CrFeCoNiCu and TiCrFeCoNi, i.e. replacing manganese with copper or titanium respectively to make equiatomic five-
component single-phase fcc materials [36,37];
4. Reduction: CrFeCoNi and CrCoNi, i.e. removing either manganese alone or both managanese and iron, to make equiatomic four-
component and three-component single-phase fcc materials respectively, so-called medium-entropy alloys with less than five
components [38];
5. Precipitation: CrMnFeCoNiAlx with x > 0.5, i.e. addition of aluminium above its fcc solubility limit of x = 0.5, half the amount of
the other components, or above about 9 at%, to make a non-equiatomic six-component two-phase material with a structure of fcc
matrix + bcc precipitates [34], and CrMnFeCoNi(NbC)x with x ~ 0.1 at%, i.e. addition of small amounts of niobium carbide to
make a non-equiatomic seven-component two-phase material with a structure of fcc + NbC precipitates [39].

This review describes some general features of the history, manufacture, structure and properties of multicomponent and high-
entropy alloys. It then concentrates on describing some aspects of the structure and properties of the variety of Cantor alloys out­
lined above, i.e. alloys in or near the extensive region of single-phase fcc structure in multicomponent phase space based on the original
Cantor alloy CrMnFeCoNi, including materials with single-phase fcc and two-phase fcc + precipitate structures in both medium-
entropy and high-entropy alloys.
There have been many previous reviews of multicomponent and high-entropy alloys, too numerous to cover fully in this review.

4
B. Cantor Progress in Materials Science 120 (2021) 100754

Particularly important previous reviews that we will draw on to some extent in this review include the books by Murty, Yeh and
Ranganathan [40,41] and Gao et al. [42], and the reviews by Cantor [43], Tsai and Yeh [44], Zhang et al. [45] and Miracle and Senkov
[13]. The recent Murty et al. book [41] and the recent Miracle and Senkov review [13] are particularly comprehensive and up-to-date.
This review will not aim to replicate their work, and will concentrate instead on a number of underlying general principles that I
believe have not received sufficiently careful attention, including how many materials and phases there are in multicomponent ma­
terials, the variety of different local atomic environments in multicomponent single-phase solid-solution Cantor alloys, and some
aspects of the thermodynamic, transport and other properties of single-phase and two-phase precipitate-reinforced Cantor alloys. Key
themes will be: (1) the astronomical number of potentially valuable and still largely (so far) uninvestigated multicomponent materials;
(2) the complexity and variability of local atomic environments in multicomponent single-phase fcc Cantor alloys; (3) the thermo­
dynamics of multicomponent single-phase fcc Cantor alloys and the extent to which they are stabilised by entropy; (4) the complexity
of transport mechanisms in Cantor alloys such as diffusion via vacancy flow, and deformation via dislocation slip; and (5) the resulting
range of potentially valuable mechanical, chemical, electrical and other properties in Cantor alloys.

1.7. Terminology

Professor Yeh invented the term high-entropy alloys [11], and this has become the accepted terminology for the field. However, this
term has been criticised, and there has been some dispute about the extent to which different multicomponent materials do or do not
exhibit low, medium or high entropy [13,40-45], though the terms medium-entropy alloys and high-entropy alloys as defined by the
number of components being below or above five respectively, have become accepted and embedded in the literature [13,40-45].
There are fairly complex issues that bear on the accuracy of the term high-entropy alloys, associated with: (1) the distinction between
configurational and non-configurational entropy; (2) the relative effects of enthalpy or internal energy vis-a-vis entropy; and (3) the
entropic effects of changing the number of components and the number of phases. These are all discussed more fully below, in the
section on the thermodynamics of Cantor alloys.
I prefer to use the term multicomponent alloys. However, this term has also been criticised [13,40-45] as including a few (as it
happens only a very few, because there aren’t very many) conventional multicomponent alloys such as modern high temperature
nickel superalloys, which can contain as many as six or seven consciously-engineered alloying additions, albeit all or almost all at
relatively low alloying levels. I don’t see this, however, as any kind of problem or indeed criticism. I take the view that, as a scientific
and technological community, we have until recently been very unadventurous in our alloying strategy, and that this applies as much
to materials with a single principal component containing large numbers of relatively low-level alloying additions as it does to ma­
terials with large numbers of principal components in equal or near-equal proportions. Our key lack of adventure has been our
reluctance to use a multiplicity of components. In other words, I believe that we should be more pro-active in investigating all regions of
unexplored (and relatively unexplored) multicomponent phase space, including:

1. Multicomponent materials with large numbers of principal components in an extended single-phase field, forming (therefore) a
single-phase structure with relatively high levels of configurational entropy – we call these high-entropy alloys (including the
original Cantor alloy [10]);
2. Multicomponent materials with fewer principal components in an extended single-phase field, also forming (therefore) a single
phase structure with somewhat lower levels of configurational entropy – we call these medium-entropy alloys;
3. Multicomponent materials with large numbers of principal components just outside an extended (as well as in some cases fairly
limited) single-phase field, with precipitate-reinforced structures – we call these precipitate-reinforced high-entropy alloys (including
the fcc + bcc alloys discovered initially by Cantor et al [10] and Yeh et al [11]);
4. Multicomponent materials far from any single-phase field, with multiple phases consisting of mixed eutectic, peritectic and/or
other polyphase structures – we call these multiphase alloys;
5. Multicomponent materials with one (or occasionally two or three) principal components and large numbers of relatively dilute
alloying additions – we call these multicomponent dilute alloys.

There has been extensive investigation of category (1), because of the surprising discovery and novelty of extended single-phase
solid-solutions in multicomponent phase space [13,40-45]. There has recently been substantial and growing investigation of cate­
gories (2) and (3), as it has become increasingly clear that there are major opportunities for conventional-style metallurgical/materials
engineering of the microstructure and properties of single-phase solid solutions in multicomponent materials, by manipulating the
number of components, their proportions, and the effects of thermomechanical treatments [13,40-45]. But there has been little or no
investigation of categories (4) and (5), in the first case because of their complexity and in the second case because of their similarity to
(albeit only a few) conventional materials. It is, of course, quite clear that the term multicomponent alloys is wider than and includes as a
special case high-entropy alloys, i.e. some multicomponent alloys are high-entropy, but others are not.
Alternative terms such as multiple-principal-element alloys, compositionally-complex and complex concentrated alloys have also been
proposed [13,40-45], but in my view these are rather clumsy, and do not really add any clarity of meaning. This review will stick to the
well-known and easily understood terms multicomponent alloys for the overall range of alloying options as described above, and high-
entropy alloys for extended single-phase and near-single-phase solid solutions that do have significant levels of configurational entropy
of mixing.

5
B. Cantor Progress in Materials Science 120 (2021) 100754

2. How many materials are there?

We can start by asking an important question. How many materials are there? In order to answer this question we need to consider
two subsidiary questions. First, how many components can we use? And second, what counts as a different material? When we have
answered these two questions we can work out the answers to a number of other important questions. How can we calculate the total
number of all different possible materials? How can we represent all the different possible materials in multicomponent phase space?
And what fraction of all possible materials do we know anything about?

2.1. How many components can we use?

All alloys or materials are, of course, made up of many different components, which in standard thermodynamics can be either
elements and/or compounds, corresponding to atomic and/or molecular species respectively. The use of compounds is, however,
something of a convenience for the starting point of our manufacturing processes, because we can often access compounds more
readily than the elements from which they are made. In fact, however, all molecules are made up of component atoms so that, for our
purposes, we can concentrate exclusively on atomic species and elemental components without any loss of generality.
The total number of known elements and, therefore, the total number of potential components c from which we can manufacture
different materials is c = 118, ranging from hydrogen to oganesson, with atomic numbers Z = 1–118, completing the first seven rows of
the periodic table (the last four elements were discovered in 2015 and named in 2016) [46]. Elements with atomic numbers Z = 1–94
occur naturally, and elements with a higher atomic number Z = 95–118 have only been made synthetically by high-energy
bombardment with fundamental particles to initiate nuclear reactions in a particle accelerator [47]. If we exclude row 7 (Z =
87–118: i.e. Fr, Ra, the actinides and the trans-actinides) as being too radioactive, and column 18 (Z = 2, 10, 18, 36, 54 and 86: i.e. the
noble gases He, Ne, Ar, Kr, Xe and Rn), the number of usable elements from which we can manufacture different materials is reduced to
c = 80. We could be much more (probably over-) restrictive, and leave out periods 1 and 2, and columns 15 – 17, which would reduce
further the number of available elements from which to manufacture different materials, down to c = 64. This would, however, exclude
important components including anionic compound-formers such as fluorine and chlorine, interstitial elements and covalent
compound-formers such as carbon, nitrogen and boron, light metals such as lithium and beryllium, and semi-metals such as selenium
and tellurium. We could be even more (certainly very over-) restrictive and stick to transition metals in periods 4 – 6 only, which would
reduce even further the number of elements from which we can manufacture different materials, down to c = 44.

2.2. What counts as a different material?

Let each material to be specified to an accuracy of x%, i.e. materials that differ in composition by more than x% are considered to be
different, whereas those that differ in composition by less than x% are considered to be the same. All materials are, of course, man­
ufactured to a certain specification, which represents partly our ability to manufacture materials of a given composition accurately,
and partly the importance of achieving that composition accurately in order to end up with a material with a particular set of desired
properties. The specification represents effectively the error in composition we are prepared to allow ourselves during manufacture of
the material, and still achieve the required property range. Specification to an accuracy of x% means that there are effectively n = 100/
x different compositions available for each of the c components. Changing the composition of a material by x% for any of its c
components is equivalent to moving in multicomponent phase space to an adjacent one of the n composition points along the
composition axis for that particular component, and is a significant enough change to correspond to manufacturing a different
material.
Most high-quality materials are specified to at least x = 0.1% [48,49], though some relatively poorly-controlled and impure
materials are only specified to x = 1%. However, many high-performance materials require specification to x = 0.01% or even 0.001%
[48,49]. And in some cases, components are specified to the nearest part per million (ppm) or 0.0001%, for instance interstitial im­
purities in steels and dopants in semiconductors [9].

2.3. How do we calculate how many materials there are?

The number of different materials N that can be manufactured from c components with a specification of x%, i.e. n = 100/x
composition points for each of the c components, is given by the law of combinations with repetition [50] as the binomial coefficient2:

2
Note that one of my previous publications [43] quoted a different and incorrect formula for N as N = (100/x)c− 1 = nc− 1 . Our problem is to
calculate how many ways we can pick n = 100/x atoms from a total of c components. For our calculation, it doesn’t matter in which order we select
different atoms to make up a particular composition, so we are dealing with a combination not a permutation (order matters in a permutation but
not in a combination); and any particular composition can contain multiple atoms of each component, so we are dealing with a combination with
repetition and not a combination without repetition [50]. The incorrect expression above overestimates the number of different materials, since the
number of possible permutations is always greater than the equivalent number of possible combinations – this is because several different ordered
permutations represent the same combination.

6
B. Cantor Progress in Materials Science 120 (2021) 100754

( )
c+n− 1 (c + n − 1)!
N= =
n (c − 1)!n!
This expression can be simplified somewhat by using Stirling’s approximation for large numbers:
lnN = (c + n − 1)ln(c + n − 1) − (c − 1)ln(c − 1) − nlnn

1)ln(c+n− 1)− (c− 1)ln(c− 1)− nlnn }


∴N = e{(c+n−

2.4. Multicomponent phase space

Essentially the full range of possible materials is defined in multicomponent phase space by the set of all n composition points in a
multicomponent phase diagram consisting of c components, i.e. all points that are separated by at least x% in all c dimensions within a
c-cornered hyper-polyhedron. It is hard to draw this hyper-polyhedron, but it is not too difficult to conceptualise it as an extension of
our familiar binary, ternary and quaternary phase diagrams, in which all possible compositions and different materials are represented
as composition points along a line or within a triangle or a tetrahedron respectively. Our lack of knowledge about materials in the
centre of multicomponent phase space is shown schematically for ternary and quaternary systems in Fig. 2.

2.5. How many materials are there?

In summary the range of possible values for the number of components we can use when manufacturing different materials is c =
44–80, and for the material specification is x = 10− 6–10− 2% corresponding to a number of composition points for each component of n
= 108–102 respectively. A very conservative estimate in calculating the number of different materials might be c = 60, x = 0.1% and n
= 1000. Table 1 shows the number of different materials that can be manufactured, in the form of log10 (N) as a function of c, x and n,
using the law of combinations with repetition [50]. For c = 60, x = 0.1% and n = 1000, the total number of different materials is N =
10100 = 1 googol. This is an enormous number of potential materials that we can manufacture. For comparison, the number of atoms in
the galaxy is estimated as approximately 1068 [51], the number of atoms in the universe is estimated as approximately 1080 [52-54]
and the size of the universe is estimated as approximately 9.3 billion light years or 1036 nm [51]. And, of course, less conservative,
though still not unreasonable, estimates for c, x and n indicate an even more enormous number of potential materials that we can
manufacture.

2.6. How many materials do we know?

We can ask another important question: how many materials do we know anything about? Or, to put it another way, how many
materials have we manufactured, studied and/or considered for use in any product? The number of known materials NK is almost
certainly no bigger than the number of all possible binary and ternary materials, since: (1) there are some, though relatively few, binary
systems for which the phase structure and any other information is sketchy at best; (2) there are many ternary systems for which little
or nothing is known in the centre of the phase diagram; and (3) we still know very little about the vast majority of possible quaternary
and/or higher order systems. The number of all possible binary and ternary materials is the same as the number of all possible ternary
materials NT , since the set of all ternary materials contains the set of all binary materials, along the edges of all the ternary phase
diagrams. In other words, the number of known materials is less than the number of ternary materials NK < NT .

Fig. 2. Schematic ternary and quaternary phase diagrams showing regions of well-known materials near the corners and edges and regions of
poorly-known or unknown materials in the centre.

7
B. Cantor Progress in Materials Science 120 (2021) 100754

Table 1
The number of different materials N, in the form log10 (N), that can be manufactured from c components, with a material specification of x %,
corresponding to a number of composition points n = 100/x for each component.
c= 40 50 60 70 80

x (%) n
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

With c components the number of different ternary systems is c(c − 1)(c − 2), and the number of materials in each ternary system is
given (again) by the law of combinations with repetition [50], this time for just three components. We can use the same conservative
estimate for the material specification of x = 0.1%, corresponding to n = 1000 composition points for each component. The number of
known materials NK is then given by:
( )
3+n− 1
NK < NT = c(c − 1)(c − 2)
n
( )
n+2 c(c − 1)(c − 2)(n + 2)!
= c(c − 1)(c − 2) =
n 2!n!

60 × 59 × 58 × 1002 × 1001
= = 1.02968 × 1011 ≈ 1011
2
In other words the number of known materials is NK < 1011 and the fraction fK of all materials that are known is infinitesimally
small:
89
fK = NK /N < 10−

3. How many phases are there in multicomponent materials?

3.1. The Gibbs phase rule

We can ask another important question: how many phases are there in multicomponent materials? For any material at equilibrium,
the Gibbs phase rule [52,53] gives a relationship between the number of components c, the number of phases p, and the number of
thermodynamic degrees of freedom f:
p+f =c+2
For a given temperature and pressure, e.g. at room temperature and atmospheric pressure, the number of degrees of freedom is
reduced by two, and the Gibbs phase rule becomes [55,56]:
p+f =c
The number of degrees of freedom cannot be less than zero, so the number of phases in a material at equilibrium and at room
temperature and atmospheric pressure is given by [55,56]:
p=c− f ≤c
In other words, for a material at equilibrium, at room temperature, and at atmospheric pressure, the number of phases can have any
value between one and the number of components, c ≥ p ≥ 1. And for a material not at equilibrium, the number of phases can be
higher, p ≥ 1 (including p > c).

3.2. How many phases are there in different materials?

As expected from and as allowed by the Gibbs phase rule, the number of phases found in equilibrated binary materials is usually one
or two [57], and under non-equilibrium conditions is sometimes higher, three or occasionally four [3-5]. Similarly, the number of
phases found in equilibrated ternary materials is usually one, two or three [58], and under non-equilibrium conditions is sometimes
higher, four or occasionally five [3-5]. In general, therefore, p is fairly evenly spread across the range 1 to c for small c in binary and
ternary materials
In multicomponent materials, however, it is surprisingly common to find only one or two phases, with sometimes three or four
phases, and only very rarely more than three or four phases. We have already seen this in the early work, in some of the more recent
studies, and in the previous reviews of multicomponent materials mentioned so far in this review [10,11,14,16-21,23-25,27-45]. The
recent book by Murty et al [41], for instance, lists 220 separate reports of single-phase fcc multicomponent materials and 157 separate
reports of single-phase bcc multicomponent materials. In general, therefore, p ≉ c, and p is not fairly evenly spread across the range 1 to

8
B. Cantor Progress in Materials Science 120 (2021) 100754

c for large values of c in multicomponent materials.


In order to understand the reasons for and the implications of the smaller number of phases found in multicomponent materials
compared to binary or ternary materials, we need to consider several interlinked and non-independent factors: (1) the number of single
phases in a material system; (2) the number of compounds in a material system and their stoichiometry; (3) the extent of inter-
solubility between the different components in a material system; and (4) underpinning (1)-(3), the entropy and heat (or enthalpy)
of mixing in the different phases in a material system. We will consider (1)-(3) in the rest of this section, and consider (4) in the context
of a more general discussion of the thermodynamics of multicomponent materials in a later section.

3.3. How many single-phase materials are there in binary and ternary systems?

First let us consider the number of single-phase materials in binary and ternary systems. A scan of standard binary and ternary
phase diagram compilations [57,58] shows that the number of single-phase materials ns can vary from 0 to 100%, with an average ns
of, very roughly, 10–20%, i.e. 0 ≤ ns ≤ 100% and ns ≈ 10 − 20% in binary and ternary systems. The factors that affect the number of
single-phase materials in binary and ternary systems are: (1) the extent of terminal solid solubility; (2) the presence or otherwise of
compound phases; and (3) the stoichiometric or non-stoichiometric nature of any compound phases. The number of single-phase
materials is small when the terminal solid solubility is low, when there are one or more binary or ternary compounds, and when
these compounds are highly stoichiometric, i.e. form line compounds with a fixed composition and a narrow single-phase field. In
contrast, the number of single-phase materials is high when there is substantial terminal solid solubility, when there are no or few
compounds, and/or when any compounds are non-stoichiometric, with a considerable spread in composition and a correspondingly
wide single-phase field.
Terminal solid-solubility levels in binary systems are described by the Hume-Rothery rules [59], which say that solubility increases
when:

1. Solvent and solute atoms have similar atomic size.


2. Solvent and solute atoms have similar valence.
3. Solvent and solute atoms have similar electronegativity.

Conversely, the propensity for compound formation in binary systems is described by the inverse of the Hume-Rothery rules, i.e.
compound formation is promoted when:

1. Solvent and solute atoms have different atomic size.


2. Solvent and solute atoms have different valence.
3. Solvent and solute atoms have different electronegativity.

In other words, when a solvent and solute obey the Hume-Rothery rules, they exhibit extended or complete terminal solid solu­
bility, and limited or no formation of compounds; and, conversely, when a solvent and solute do not obey the Hume-Rothery rules, they
exhibit limited or no terminal solid solubility, and the formation of one or more compounds, which itself further restricts the extent of
terminal solid solubility [59].
Notice that the first Hume-Rothery rule is associated with the effect of different solvent and solute atomic size on lattice strain,
causing an increase in the internal energy and enthalpy of the solution, thus limiting the solubility; and the second and third Hume-
Rothery rules are associated with chemical interactions between solvent and solute atoms, causing compound formation, thus again
limiting the solubility. In some formulations, the Hume-Rothery rules are concatenated to two rather than three rules [56]:

1. Solvent and solute atoms have similar atomic size.


2. Solvent and solute atoms have similar electron/atom ratio.

This is because valence and electronegativity can both be regarded as influencing chemical interactions between solvent and solute
atoms via differences in electron/atom ratio [59].
The number of single-phase materials in a binary system is equal to the sum of the two terminal solid-solution composition ranges if
either there are no compounds formed, or any compounds that do form are highly stoichiometric with fixed compositions. In that case,
outside the terminal solid-solution composition ranges, binary materials are two-phase mixtures of the terminal solid solutions and/or
the compounds. On the other hand, the number of single-phase materials in a binary system is enhanced beyond the sum of the two
terminal solid-solution composition ranges if any compounds that form are non-stoichiometric with wide and variable compositions,
restricting the composition ranges of any two-phase materials.
More or less similar considerations apply to ternary systems. In other words, the number of single-phase materials in a ternary
system is equal to the sum of the three terminal solid-solution composition ranges if either there are no compounds formed, or any
compounds that do form are highly stoichiometric with fixed compositions. In that case, outside the terminal solid-solution compo­
sition ranges, ternary materials are two-phase or three-phase mixtures of the terminal solid solutions and/or compounds. On the other
hand, the number of single-phase materials in a ternary system is enhanced beyond the sum of the terminal solid-solution composition
ranges if any compounds that form are non-stoichiometric with wide and variable compositions, restricting the composition ranges of
any two-phase or three-phase materials.

9
B. Cantor Progress in Materials Science 120 (2021) 100754

3.4. How many single-phase materials are there in multicomponent systems?

Next let us consider the number of single-phase materials in multicomponent systems. The number of single phases will depend on
(1) whether terminal solid solubility is enhanced or diminished with multiple solutes in a given single solvent; (2) whether the number
of compounds increases or decreases in a multicomponent system; and (3) whether non-stoichiometry and variability of composition
increases or decreases in a multicomponent system.
The effect of having multiple solutes on the terminal solid solubility in a given single solvent has not been investigated to any great
extent. We can ask the question: is the terminal solid solubility of one solute in a given solvent promoted, restricted or unaffected by the
presence of multiple other solutes? To put it another way: is the terminal solid solubility for multiple solutes simply additive of the
individual binary solubilities, or are there cross or synergistic effects that enhance or reduce the terminal solid-solubility composition
ranges in multicomponent systems? Is the extent of the terminal solid-solution range adjacent to the corners of the multicomponent
phase diagram roughly constant (or not) along the binary axes and in multi-solute directions between the binary axes? Some recent
work on multiple solutes in Al alloys indicates that in this particular case terminal solid solubility is not strongly enhanced or restricted,
i.e. the number of single-phase terminal solid-solution materials is not strongly increased or reduced when there are multiple solutes in
a multicomponent system [60,236]. However, the presence of a large multicomponent Cantor alloy fcc single-phase field and similar
large multicomponent bcc and hcp single-phase fields, with structures similar to many single elemental components, suggests the
opposite, and that Al may be a special case.
If the result for Al proves to be the case more generally, then the propensity for more single-phase materials and fewer multiphase
materials in multicomponent systems is not caused by an increase in terminal solid solubility. This means that the propensity for more
single-phase materials and fewer multiphase materials in multicomponent systems must be caused by either a reduction in compound
formation and/or an increase in their non-stoichiometry and the width of their composition ranges. Whether terminal solid solubility is
enhanced (as suggested by the wide range of Cantor and other multicomponent single-phase alloys with fcc, bcc and hcp structures) or
not (as suggested by the provisional results for multicomponent dilute solutions in Al), it nevertheless seems likely that reduced
compound formation and increased non-stoichiometry in compounds are both important factors, as described in the next sub-section.

3.5. How many compounds are there in binary, ternary and multicomponent systems?

There is a rich array of different binary and ternary compounds with different crystal structures, different point and lattice sym­
metries (i.e. different point and space groups), and different sublattices [61,62]. The compilation by Westbrook and Fleischer [62], for
instance, reports over 200 distinct compound crystal structures. However, there have been very few, if any, reports of new multi­
component compounds with unique crystal structures and multiple sublattices, different and distinct from existing well-known binary
and ternary compounds [13,37-45,57,58,61,62]. This is quite a striking observation, which has not been much remarked on previ­
ously. After all, when alloy structures were extended by high speed solidification and vapour deposition techniques, a wide range of
new binary and ternary crystal compounds was discovered, but this has not been replicated by extension of alloy structures into
multicomponent phase space.
The paucity of distinctive multicomponent compounds seems to be something to do with the fact that we live in a three-dimensional
world [237]. The different crystal structures of different binary and ternary compounds represent the various different ways that two
or three different-sized atoms and molecules can be packed into 3-D space, allowing for any spatial effects of directionality in bonding.
And strong ionic or covalent bonds are intrinsically 2D in nature, with each bond resulting from the electrostatic attraction between
individual atoms in a molecule. Given the relatively restricted size range of component atoms, there is only a relatively limited number
of ways that spatial packing can be achieved, with only a limited number of different sublattices for the different atoms and molecules.
Most binary and ternary compounds have no more than two or three sublattices in their crystal structure. Increasing the number of
components to make an intrinsically different multicomponent compound would require increasing the number of different sublattices
to accommodate the different atomic species, and this is just not possible in our limited 3D space.
In any case and for whatever reason, the number of different compounds does not scale with the number of components, i.e. the
number of different compounds does not keep increasing as we increase the number of components above two or three in a multi­
component system. Instead, the additional components are accommodated as substitutional solutes on the most appropriate sublattice
(s) in the terminal solid solutions and/or binary and ternary compounds, increasing their range of non-stoichiometry. Thus a

Table 2
Equiatomic and non-equiatomic 6-component single-phase fcc Cantor alloys made by adding components to the original Cantor alloy CrMnFeCoNi.
Alloying addition Solubility (at%) Process route Reference

Al 8 Arc melting [34]


Al 8 Arc melting [63]
C 1 Arc melting [64]
Ti 2 Casting + high-pressure torsion [65]
V 16 Melt spinning [10]
V 4.8 Arc melting [66]
Cu 16 Melt spinning [10]
Cu 4.8 Arc melting [67]
Nb 16 Melt spinning [10]

10
B. Cantor Progress in Materials Science 120 (2021) 100754

combination of fewer compounds and increased solubility and non-stoichiometry jointly contribute to the observed increasing number
of single phase materials in multicomponent systems, alongside the effect (except for Al) of enhanced terminal solid solubility.

4. Single-phase fcc Cantor alloys

4.1. General characteristics of the single-phase fcc region

As explained previously, there is an enormous region in multicomponent phase space consisting of materials with a single-phase fcc
structure, made up mostly (but not exclusively) of transition metal components, loosely based around the original Cantor alloy
CrMnFeCoNi, including materials made by a combination of addition to, variation from, replacement of and/or reduction of the set of
original Cantor alloy components. The book by Murty et al [41] gives extensive detail about the wide range of different alloy structures
that have been manufactured in multicomponent materials, and we will draw on their documented information to explore initially the
extent of the single-phase fcc Cantor alloy region in multicomponent phase space. Tables 2–5 list some of the different compositions
and manufacturing methods that have been used to produce single-phase fcc modified Cantor alloys made, respectively, by addition to,
variation from, replacement of, and reduction of the set of original Cantor alloy components. And Tables 6–8 list some of the different
compositions and manufacturing methods that have been used to produce single-phase fcc modified Cantor alloys by a combination of
some of the different effects of variation from, replacement of, and reduction of the set of original Cantor alloy components. Tables 2–8
give only a selection of the alloys that have been manufactured, and fuller details and much more extensive lists are provided by Murty
et al [41]. Tables 2–8 do not, for instance, include any of the non-equiatomic 6- and 7-component single-phase fcc alloys that have been
made by a combination of addition to, as well as replacement of, and variation from the set of original Cantor alloy components.
It is abundantly clear from Tables 2–8 that, as suggested previously, there is a vast region of single-phase fcc materials in the middle
of multicomponent phase space, which can be accessed by a wide variety of processing methods, and which contains literally millions
and millions of different single-phase fcc materials. This region has the following characteristics:

1. It is largely based around components from the middle and late part of the first transition period, i.e. Cr, Mn, Fe, Co, Ni and Cu, as
represented by the original Cantor alloy CrMnFeCoNi;
2. It extends to some extent towards the earlier part of the first transition period, with additions of components such as Ti and V;
3. It extends to some extent towards other transition periods, with additions of components such as Nb, Mo, Pd, Pt and Au;
4. It extends to some extent towards main period components such as Al and Si;
5. It extends to some extent towards ternary and quaternary sub-systems of the original Cantor alloy, such as CrCoNi and CrFeCoNi.

As well as the original Cantor alloy CrMnFeCoNi, therefore, this single-phase fcc region includes:

1. Equiatomic 6-component alloys such as CrMnFeCoNiCu, VCrMnFeCoNi and CrMnFeCoNiNb, obtained by adding Cu, V or Nb [10];
2. Equiatomic 5-component alloys such as CrFeCoNiCu and TiCrFeCoNi obtained by replacing Mn with either Cu [81,82] or Ti
[74,75]; and MnFeCoNiCu obtained by replacing Cr with Cu [78];
3. Equiatomic 4-component alloys such as CrFeCoNi [38,85,86] and CrMnCoNi [84], obtained by removing Mn or Fe, and CrCoNiCu
[96] and FeCoNiCu [88,89], obtained by removing Mn and Fe and adding Cu;
4. Equiatomic 3-component alloys such as CrCoNi [38,84] and FeNiCo [38,85,86], obtained by removing Mn and Fe or Cr and Mn,
and CoNiCu [72], obtained by adding Cu and removing Cr, Mn and Fe;
5. Equiatomic 5-component alloys such as TiFeCoNiCu and VFeCoNiCu obtained by replacing both Cr and Mn with either Ti and Cu
[76] or V and Cu [77]; CrFeNiCuMo obtained by replacing both Mn and Co with Cu and Mo [83]; and MnFeNiCuPt obtained by
replacing both Cr and Co with Cu and Pt [79];
6. Non-equiatomic 5-component alloys such as (CrFeCoNi)-11at%Cu [107,108], (CrMnFeNi)-11at%Cu [111,112] and CrFeCoNi-10at
%Mo [113];
7. Non-equiatomic 6-component alloys such as (CrMnFeCoNi)-2at%Ti [65];
8. Non-equiatomic 5- and 6-component alloys containing aluminium such as (CrFeCoNi)-11at%Al [101] and (CrMnFeCoNi)-8at%Al
[34,63].

We will explore in the next few sub-sections several specific features of this single-phase fcc region that have been studied recently

Table 3
Non-equiatomic 5-component single-phase fcc Cantor alloys made by varying component composition in the original Cantor alloy CrMnFeCoNi.
Alloy composition Composition range x (at%) Process route Reference

(CrFeCoNi)86 Mn14 14 Arc melting [68]


(CrFeCoNi)100-x Mnx 0–11 Arc melting [69]
(CrMnFeNi)100-x Cox 5–20 Arc melting [70]
CrMnFeCo15Ni25 – Arc melting [68]
(CrMnFeCo)100-x Nix 0–92 Casting [71]
CrMnFeCoNix 15–43 Arc melting [35]

11
B. Cantor Progress in Materials Science 120 (2021) 100754

Table 4
Equiatomic 5-component single-phase fcc Cantor alloys made by replacing components in the original
Cantor alloy CrMnFeCoNi.
Alloy composition Process route Reference

AlCoNiCuZn Mechanical alloying [72]


AlCoNiCuZn Mechanical alloying [73]
TiCrFeCoNi Arc melting [74]
TiCrFeCoNi Mechanical alloying [75]
TiFeCoNiCu Arc melting [76]
VFeCoNiCu Casting [77]
MnFeCoNiCu Arc melting [78]
MnFeNiCuPt Casting [79]
NiCuPdPtAu Arc melting [80]
CrFeCoNiCu Arc melting [81]
CrFeCoNiCu Sputtering [82]
CrFeNiCuMo Arc melting [83]

Table 5
Equiatomic 3- and 4-component single-phase fcc Cantor alloys made by reducing the number of
components in the original Cantor alloy CrMnFeCoNi.
Alloy composition Process route Reference

CrCoNi Mechanical alloying [84]


CrCoNi Casting [38]
MnFeNi Casting [38]
MnCoNi Casting [38]
FeNiCo Casting [38]
FeNiCo Mechanical alloying [85]
FeNiCo Casting [86]
FeNiCo Casting [87]
FeNiCo Arc melting [76]
FeNiCo Mechanical alloying [88]
FeNiCo Mechanical alloying [89]
CrFeCoNi Casting [86]
CrFeCoNi Arc melting [90]
CrFeCoNi Arc melting [91]
CrFeCoNi Arc melting [92]
CrFeCoNi Mechanical alloying [85]
CrFeCoNi Casting [38]
CrFeCoNi Casting [93]
CrMnCoNi Arc melting [84]
MnFeCoNi Mechanical alloying [89]
MnFeCoNi Arc melting [94]

Table 6
Equiatomic 3- and 4-component single-phase fcc Cantor alloys made by a combination of reducing the
number of components and replacing components in the original Cantor alloy CrMnFeCoNi.
Alloy composition Process route Reference

CoNiCu Mechanical alloying [72]


AlTiNiCu Casting [95]
NiCuPdPt Arc melting [80]
NiCuPdAu Arc melting [80]
NiCuPtAu Arc melting [80]
NiPdPtAu Arc melting [80]
CuPdPtAu Arc melting [80]
CrCoNiCu Mechanical alloying [96]
FeCoNiCu Mechanical alloying [88]
FeCoNiCu Mechanical alloying [89]
CoNiCuZn Mechanical alloying [72]

12
B. Cantor Progress in Materials Science 120 (2021) 100754

Table 7
Non-equiatomic 3- and 4-component single-phase fcc Cantor alloys made by a combination of reducing the number of components and varying
component composition in the original Cantor alloy CrMnFeCoNi.
Alloy composition Composition range x (at%) Process route Reference

Al0.5FeNiCu 14 Splat quenching [97]


Al0.3FeNiCu 9 Laser deposition [98]
(CrFeCo)100-xNix 20–40 Arc melting [35]
(CrCoNi)100-xFex 0–17 Laser deposition [99]
Mn10 Fe30Co30Ni30 10 Mechanical alloying [100]

Table 8
Non-equiatomic 5-component single-phase fcc Cantor alloys made by a combination of replacing components and varying component composition in
the original Cantor alloy CrMnFeCoNi.
Alloy composition Composition range x (at%) Process route Reference

Al0.5CrFeCoNi 11 Arc melting [101]


Al0.3CrFeCoNi 7 Arc melting [102]
AlxCrFeCoNi 0–14 Arc melting [103]
Al0.5CrFeCoNi 11 Casting [77]
AlxCrFeCoNi 0–11 Arc melting [104]
AlxSixFeCoNi 0–8 Arc melting [105]
Al0.2Si0.2FeCoNi 6 Casting [106]
CrFeCoNiCu0.5 11 Arc melting [107]
CrFeCoNiCu0.5 11 Arc melting [108]
CrFeCoNiCux 0–27 Casting [109]
Mn25Fe30Co5Ni25Cu15 – Arc melting [78]
Mn1.75 Fe0.25 CoNiCu – Casting [110]
CrMnFe2Ni2Cu2 – Arc melting [111,112]
Cr2Mn2Fe2Ni2Cu 11 Arc melting [111,112]
CrMn2FeNi2Cu – Arc melting [111,112]
Cr15Fe20Co35Ni20Mo10 – Arc melting [113]
Cr12.5Fe20Co42.5Ni20Mo5 – Arc melting [113]

in more detail: the basic microstructural features of Cantor alloys; non-equiatomic variations of the Cantor alloy composition
CrMnFeCoNi; medium-entropy alloys based on quaternary CrFeCoNi; and aluminium-containing alloys (CrMnFeCoNi)Alx.

4.2. Cantor alloy microstructure

As indicated originally by Cantor et al [10], modified Cantor alloys are dendritic as-solidified, as shown in Fig. 3, with a typical
〈1 0 0〉 solidification texture [114] and dendritic microsegregation patterns consisting of relatively minor composition variations be­
tween the dendritic spines and interdendritic regions [115]. Dendritic spines are enriched in Cr, Fe and Co, and depleted in Mn and Ni;
and interdendritic regions are correspondingly depleted in Cr, Fe and Co, and enriched in Mn and Ni [10,115]. The spatial scale of the
dendritic microsegregation is typically on the order of 5-50μ m depending on the solidification speed, with a pseudo-binary local
composition ranging from approximately (CrFeCo)70(MnNi)30 in the dendritic spines to approximately (CrFeCo)50(MnNi)50 in the
interdendritic regions. Differential scanning calorimetry [115] indicates that the liquidus and solidus temperatures are Tl ≈ 1340 ◦ C
(1613 K) and Ts ≈ 1290 ◦ C (1563 K), i.e. there is a solid–liquid thermal gap of ΔT = 50 ◦ C (50 K) at the Cantor alloy composition
CrMnFeCoNi (corresponding to a pseudo-binary composition of (CrFeCo)60(MnNi)40 or co = 40 at%). From the Scheil equation
[116,117], the interdendritic microsegregation corresponds, therefore, to a distribution coefficient of k ≈ 0.75, with pseudo-binary
initial and final solidification compositions of kco = 30 at% and co /k = 53 at% [115].
Detailed atom-probe studies show that heat treatment for an hour at 1100 ◦ C (1373 K) removes all dendritic microsegregation, and
that heat-treated single-phase fcc Cantor alloys are then true random solid solutions at the atomic level [115]. The random nature of
multicomponent solid solutions and the corresponding local atomic distributions and short-range order are discussed in more detail in
a later section of this review on the thermodynamics of Cantor alloys. Deformation and heat treatment of as-solidified Cantor alloys
removes the dendritic microstructure and microsegregation patterns, leading to a homogeneous microstructure consisting of relatively
large equiaxed grains containing occasional annealing twins, as shown in Fig. 4, with grain sizes typically 5-500μ m depending on the
annealing temperature and time [118]. These as-solidified and heat-treated microstructural features have been demonstrated by many
other investigations of the Cantor alloy CrMnFeCoNi and modified Cantor alloys across the single-phase fcc region in multicomponent
phase space [119-121].

4.3. Cantor alloy composition variations CraMnbFecCodNie

Bracq et al [122] have recently studied quite extensively the effect on microstructure and mechanical properties of varying the
composition of the original Cantor alloy components, by changing independently each of the different components over a wide range.

13
B. Cantor Progress in Materials Science 120 (2021) 100754

Fig. 3. Scanning electron micrograph with energy-dispersive X-ray mapping showing dendritic Fe segregation in the as-cast multicomponent single-
phase fcc Cantor alloy CrMnFeCoNi (after Cantor et al [10]).

Fig. 4. Optical micrographs and transmission electron micrographs showing the microstructure of the multicomponent single-phase fcc Cantor alloy
CrMnFeCoNi after homogenisation, rolling and recrystallisation to produce grain sizes of (a) 4.4 μm; (b) 50 μm; and (c) 155 μm (after Otto et
al [118]).

They investigated alloys with the general composition CraMnbFecCodNie by changing the composition of each component Cr, Mn, Fe,
Co and Ni one at a time along the five different main composition axes, i.e. five different sets of alloys with compositions (MnFe­
CoNi)100-xCrx, (CrFeCoNi)100-xMnx, (CrMnCoNi)100-xFex, (CrMnFeNi)100-xCox, and (CrMnFeCo)100-xNix, with x = 0–100%. A total of 24
different alloys were manufactured by induction melting followed by gravity casting into 3 mm rods, cutting into slices, heat treating
for 6 days under argon at a temperature of 1100 ◦ C (1373 K), and then rapid cooling to room temperature. Bracq et al concentrated on
varying Cr, Mn, Fe and Co, since a similar investigation had already been undertaken previously for Ni by Laurent-Brocq et al
[123,124]. Taking these results together, single-phase fcc alloys are found for modified Cantor alloys CraMnbFecCodNie with
composition ranges 0–25% Cr, 0–50% Mn, 0–50% Fe, 10–50% Co and 10–100% Ni. This demonstrates clearly the wide region of

14
B. Cantor Progress in Materials Science 120 (2021) 100754

single-phase fcc alloys in the quinary system CraMnbFecCodNie, based around the original Cantor alloy CrMnFeCoNi, which extends to
the surfaces and edges of the quinary phase diagram to include the equiatomic quaternary alloys MnFeCoNi, CrFeCoNi and CrMnCoNi
with 0% Cr, Mn and Fe respectively, as well as to include ultimately single-component pure Ni (and at high temperature pure Fe and Co
too).
The single-phase fcc composition ranges determined by Bracq et al [122] agree well with other (more limited) studies of individual
modified Cantor alloys [35,68-71]. They also agree well with CALPHAD calculations of single-phase fcc stability at the annealing
temperature of 1100 ◦ C (1373 K), but not at room temperature, where CALPHAD predicts that all the alloys should become multiphase
[122]. This is discussed further in Section 5.3.

4.4. Medium-entropy alloys based on quaternary CrFeCoNi

Equiatomic quaternary CrFeCoNi is a medium-entropy single-phase fcc alloy with very similar general as-solidified and heat-
treated microstructural features to those described above for the original Cantor alloy CrMnFeCoNi [38,86,90-93]. There is an
enormous region of single-phase fcc materials surrounding quaternary single-phase CrFeCoNi, made by a combination of addition to,
variation from, replacement of and/or reduction of the set of original quaternary alloy components. This is effectively part of the even
larger region of single-phase fcc materials surrounding the original Cantor alloy CrMnFeCoNi.
Some examples of materials made by varying the quaternary alloy components and/or including additional early transition metal
components, i.e. with the generic composition (CraFebCocNid)Xx, are listed in Table 9. Early transition metal components can dissolve
in the single-phase fcc structure up to a solubility limit that depends on the alloying component, beyond which the alloys become two-
phase mixtures of fcc and an intermetallic sigma or Laves phase. With increasing alloying addition the as-solidified materials show a
classic transition [116,117] from dendritic solid solution to pseudo-binary hypoeutectic dendrite plus interdendritic eutectic, full
eutectic and finally hyper-eutectic microstructure [125-127]. In the solid solution region and just beyond, deformation and heat
treatment removes dendritic and interdendritic eutectic microstructures and microsegregation patterns, leading to a homogeneous
microstructure of relatively large equiaxed grains containing occasional annealing twins, with grain sizes 5–500 μm depending on the
annealing temperature and time [129,130]. Alloys beyond the solubility limit can be heat treated to develop classic precipitate-
reinforced fcc-matrix microstructures [129,130].

4.5. Aluminium-containing alloys (CrMnFeCoNi)Alx

The Cantor alloy CrMnFeCoNi with additions of Al, i.e. a series of alloys with compositions (CrMnFeCoNi)100-xAlx, have been
studied in detail by He et al [34] manufactured by arc melting followed by drop casting into 10x10x50mm moulds, and by Hsu et al
[132] manufactured by RF magnetron sputtering; and the modified medium-entropy Cantor alloy CrFeCoNi with additions of Al, i.e. a
series of alloys with compositions (CrFeCoNi)100-xAlx, has been studied in detail by Cieslak et al [104] manufactured by arc melting
followed by air cooling and then heat treating for 3 days at 1000 ◦ C (1273 K) sealed in vacuum, and by sintering from pure metallic
powders for 2 weeks at 1000 ◦ C (1273 K) sealed in vacuum.
The Al-containing alloys are fcc for compositions ≤8–9 at% Al, fcc + bcc for compositions between 8 and 16at% Al, and bcc for
compositions ≥16at% Al. With increasing Al alloying addition the as-solidified materials again show a classic transition [116,117]
from dendritic solid solution to pseudo-binary hypoeutectic dendrite plus interdendritic eutectic, full eutectic and finally hyper-
eutectic microstructure [34,104]. In mixed fcc + bcc alloys there is microsegregation between fcc dendrites enriched in Cr and Fe,
and interdendritic bcc enriched in Ni and Al [104]. On cooling to room temperature the interdendritic or eutectic bcc phase shows
partial ordering to CsCl-type B2 ordered bcc, similar to the well-known binary intermetallic compounds NiAl, FeAl and CoAl [61,62],
indicating a transition from high temperature random solid solution bcc to a multicomponent compound (CraMnbFecCodNie)Al. In the
solid solution and two-phase hypoeutectic and eutectic regions, deformation and heat treatment removes dendritic and interdendritic
eutectic microstructures and microsegregation patterns, leading to a homogeneous microstructure of relatively single-phase or duplex
grains, with grain sizes depending on the annealing temperature and time [34,104]. Alloys in the vicinity of the solubility limit can be
heat treated to develop reinforced fcc-matrix microstructures containing bcc B1 or ordered bcc B2 precipitates [129,130]. These
microstructural features have been seen in a variety of other Al-containing modified Cantor alloys such as (CrMnFeCoNi)Al0.5 [101],
(CrMnFeCoNi)Al0.75 [133], (CrFe2CoNi)Al0,75 [134], and (CrFeCoNi2.1)Al [135,136].

5. Thermodynamics of Cantor alloys

5.1. The free energy of multicomponent solutions

In a multicomponent single-phase solid solution, the molar Gibbs free energy of the material ΔG at a temperature T is given by
[55,56,137]:
∑ ∑
ΔG = xi μi + ΔGmix = xi μi + ΔHmix − TΔSmix
i i

where i is the i’th component, xi and μi are, respectively, the molar fractions and chemical potentials (or molar free energies) of the
components, and ΔGmix , ΔHmix and ΔSmix are, respectively, the free energy of mixing, heat (or enthalpy) of mixing and entropy of

15
B. Cantor Progress in Materials Science 120 (2021) 100754

Table 9
Equiatomic and non-equiatomic materials made by adding components to single-phase fcc alloys in the quaternary CrFeCoNi system, i.e. with generic
compositions (CraFebCocNid)Xx.
Alloy Composition range x (at%) Structure Reference

CrFeCoNiV 20 fcc +σ [125]


CrFeCoNiNb 20 fcc + C14
CrFeCoNiTa 20 fcc + C14
CrFeCoNiTax 2–15 fcc + C14 [126]
CrFeCoNiMox 0–7 fcc [127]
CrFeCoNiMox 13 fcc + CrFe2.3MoNi
CrFeCoNiMo0.2 5 fcc [131]
CrFeCoNiCux 0–20 fcc [107]
CrFeCoNiCux 0–27 fcc [109]
Cr5Fe20Co35Ni20Ti20 20 fcc [128]
Cr5Fe20Co20Ni35Ti20 20 fcc
Cr5Fe35Co20Ni20Ti20 20 fcc
Cr20Fe30Co17Ni30Mo2 W1 3 fcc [129]
Cr12.5Fe20Co42.5Ni20Mo5 5 fcc [113]
Cr15Fe20Co35Ni20Mo10 10 fcc
Cr15Fe21Co28Ni28Al4Ti4 8 fcc [130]

mixing. The first term on the right hand side of the first expression for ΔG is the free energy of the unmixed components, and the second
term is the free energy change caused by mixing them to form the single-phase solid solution.
If the solid solution is ideal, there is no interaction energy between the different atoms, which act in the solid solution effectively
like a packing of non-interacting billiard balls. The heat of mixing is zero, i.e. ΔHmix = 0, and the entropy of mixing is given by the
configurational entropy of a random distribution of the different atoms within the solid solution:

ΔSmix = − R xi lnxi
i

The overall ideal free energy of the material is then [55,56,137]:


∑ ∑
ΔG = xi μi + RT xi lnxi
i i

If the solid solution is regular, there are small pairwise, three-way, four-way etc. interaction energies ωij , ψ ijk , χ ijkl etc. between the
different atoms, so the heat of mixing is not zero, i.e. ΔHmix = f(ωij , ψ ijk , χ ijkl ⋯) ∕
= 0, but atomic interactions are insufficiently strong to
affect the entropy of mixing significantly, which is still, therefore, given by the configurational entropy of a random distribution of the
different atoms within the solid solution. The overall regular free energy of the material is then [55,56,137]:
∑ ∑
ΔG = xi μi + ΔHmix (ωij , ψ ijk , χ ijkl ⋯) + RT xi lnxi
i i

If the solid solution is non-regular, there are larger interaction energies ωij , ψ ijk , χ ijkl etc. between the different atoms, so the entropy
of mixing is no longer given only by the configurational entropy of a random distribution of the different atoms within the solid
solution. The overall non-regular free energy of the material is then [55,56,137]:
∑ ∑
xs
ΔG = xi μi + ΔHmix (ωij , ψ ijk , χ ijkl ⋯) + RT xi lnxi − TΔSmix (ωij , ψ ijk , χ ijkl ⋯)
i i

where ΔSxsmix = f’(ωij , ψ ijk , χ ijkl ⋯) ∕


= 0 is the excess entropy of mixing over and above the random configurational entropy. Notice that
there are two contributions to the excess configurational entropy ΔSxs mix , which together can be positive or negative:

1. excess configurational entropy of mixing arising from a non-random distribution of the different atoms, caused by their in­
teractions; and
2. excess vibrational entropy of mixing arising from changes in the vibrational spectra of the different atoms caused by their
interactions.

Miracle and Senkov [13] have pointed out that excess vibrational entropy of mixing can often be more significant than configu­
rational entropy of mixing.
There have been many theories, analyses and approximations used to assess and calculate values of the heat of mixing ΔHmix and
the excess entropy of mixing ΔSxs mix for different types of interatomic interactions ωij and ψ ijk in binary and ternary solid solutions,
allowing for factors such as long-range order, short-range order, spinodal demixing and differences in molar volume [55,137-139]. In
general, however, these thermodynamic calculations become increasingly unwieldy, difficult to use and, therefore, unhelpful as the
number of components increases in a multicomponent system. For instance, the total of number of different interaction energies in a
single-phase solid solution with c components is:

16
B. Cantor Progress in Materials Science 120 (2021) 100754

c(c − 1) c(c − 1)(c − 2) c(c − 1)(c − 2)(c − 3)


+ + +⋯
2 3×2 4×3×2

which is 11 or 26 for 4-component or 5-component materials respectively, and all of these interaction energies need to be estimated, as
well as their independent effects on the internal energy (or enthalpy) and atomic vibrational spectra of the material.
More details of the thermodynamics of multicomponent materials are given in the book by Murty et al [41] and the review by
Miracle and Senkov [13].

5.2. Are multicomponent materials high entropy?

In an equiatomic multicomponent single-phase solid solution with c components, the molar fractions of all the components are the
same, xi = 1/c for all i, and the ideal entropy of mixing, ideal free energy of mixing and overall ideal free energy are then [13,41]:

ΔSmix = − R xi lnxi = − Rln(1/c)
i

ΔGmix = − TΔSmix = RTln(1/c)



ΔG = xi μi + RTln(1/c)
i

Table 10 shows the ideal entropy of mixing ΔSmix and the corresponding ideal free energy of mixing ΔGmix at 1000 K in an
equiatomic multicomponent single-phase solid solution for different numbers of components ranging from c = 2–10. The ideal entropy
of mixing in J/K/mol and the ideal free energy of mixing at 1000 K in kJ/mol are numerically equal and opposite, the first positive and
the second negative. The ideal entropy of mixing increases from approximately 5 J/K/mol with 2 components to approximately 20 J/
K/mol with 10 components; and the ideal free energy of mixing at 1000 K correspondingly decreases from approximately − 5 kJ/mol
with 2 components to approximately − 20 kJ/mol for 10 components. Differentiating:
dΔGmix dΔSmix RT
=− T =− → 0 as c → ∞
dc dc c

so the ideal entropy of mixing continues to increase and the ideal free energy of mixing continues to decrease, but both do so only
slowly3, for further increases in the number of components beyond 10.
Table 10 explains why Professor Yeh and his group coined the term high-entropy alloys [11] for multicomponent materials such as
the original Cantor alloy and, as explained in a previous section, this has become the accepted name for the field [13,37-45]. In fact the
term high-entropy alloy applies very well to the original Cantor alloy, which is an equiatomic 5-component near-random solid-solution
material [115], close to ideal, with a free energy of mixing dominated, therefore, by the configurational entropy of mixing.
We have seen previously that the region of modified Cantor alloys, i.e. of a single-phase fcc solid solution, is very large in
multicomponent phase space [13,37-45]. Moreover it is surrounded by similarly large regions of other single-phase solid solutions,
such as bcc with Al additions, bcc with some early transition metal additions, and intermetallic sigma and Laves phases with other early
transition metal additions [13,37-45]. The enormous number of materials in these large single-phase regions of multicomponent phase
space all have substantial configurational entropy from the mixing of many components on a single-phase terminal or compound solid-
solution structure. And this high configurational entropy of mixing acts together with the paucity of multicomponent compounds,
described previously, to increase intersolubility between the different components and stabilise the single-phase solid-solutions, which
explains why they are found to be so prevalent and extended in multicomponent phase space [13,37-45]. The term high-entropy alloys

Table 10
Ideal entropy and Gibbs free energy of mixing ΔSmix and ΔGmix versus number of components c.
Number of components c ln (1/c) Ideal entropy of mixing ΔSmix (J/K/mol) Ideal free energy of mixing ΔGmix at 1000 K (kJ/mol)

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

3
For instance, the ideal entropy of mixing is ~25 J/K/mol with 20 components, ~28 j/K/mol with 30 components and ~31 J/K/mol with 40
components.

17
B. Cantor Progress in Materials Science 120 (2021) 100754

applies well to all these materials. Furthermore, many other multicomponent materials are multiphase mixtures of such extended solid
solutions, e.g. fcc + bcc, fcc + hcp, fcc + σ, etc., which also have substantial configurational entropy of mixing from the sum of the
individual configurational entropies of mixing of the individual phases [13,37-45]. Once again, the term high-entropy alloys applies
well to all these materials.
However, not all multicomponent materials have a free energy that is dominated by a high configurational entropy of mixing, since:

1. many multicomponent solid solutions have substantial excess vibrational entropies of mixing ΔSxs
mix [13] and/or heats of mixing
ΔHmix [140-142], larger than and, therefore, dominating over the configurational entropy of mixing; and
2. many multicomponent materials are multiphase with phases of limited stoichiometry range [13,37-45], i.e. not extended solid
solutions with only limited configurational entropies of mixing.

The term high-entropy alloys is less appropriate for these materials.


To develop the first of these points further, consider Table 11, which shows heats of mixing and interaction energies in the ten
different binary alloys that can be made from the five components of the original Cantor alloy [140,141]. The interaction energies in
the binary alloys are a reasonable first-order approximation to the pairwise interaction energies ωij in the Cantor alloy; and the heat of
mixing in the Cantor alloy is to a first approximation given by a weighted sum of the heats of mixing of the binary alloys [140]. Notice
that the binary interaction energies and heats of mixing are all relatively low, which is why the 5-component single-phase fcc Cantor
alloy is stable in the first place, and also why it is a near-random, near-ideal solid solution [115]. Even in this case, however, the binary
heat of mixing for Mn-Ni is 14 kJ/mol, comparable with (and in fact just slightly larger than) the contribution to the free energy of
mixing associated with a 5-component configurational entropy of mixing of 13.38 kJ/mol at 1000 K, as shown in Table 10. In many
other multicomponent materials, the pairwise interaction energies and binary heats of mixing are larger than for Mn-Ni [140-142],
which is why Otto et al [142] found that other equiatomic 5-component alloys such as TiCrMnFeNi, VCrMnCoNi, CrMnFeCoCu and
MnFeCoNiMo are not single-phase fcc solid solutions. In these cases, a positive heat of mixing is dominant over the negative effect (on
the free energy) of a high configurational entropy of mixing, and demixing into two (usually) or more phases is preferred to a single-
phase solid solution.
In summary, there is an enormous number of multicomponent materials. Many of them have structures that are strongly influenced
by and stabilised by the configurational entropy of mixing and are, therefore, appropriately designated as high-entropy alloys. However,
many others have structures that are not strongly influenced by or stabilised by the configurational entropy of mixing, and are,
therefore, less appropriately designated as high-entropy alloys.

5.3. Stability of Cantor alloys

The third law of thermodynamics defines an absolute scale for entropy, with the entropy of any material equal to zero at absolute
zero temperature, S = 0 at T = 0 [55,56,137]. This means that the equilibrium structure for all materials consisting of a single
component, either a single element or a single compound is a perfect single crystal at absolute zero; and the equilibrium structure for
all materials that are alloys, i.e. a combination of more than one component, is a mixture of one or more perfect crystals. In other
words, all solid solution alloys when cooled to room temperature at equilibrium must either demix fully or order fully. Very frequently,
however, atomic diffusion is too slow to allow this to take place.
In a multicomponent single-phase solid solution such as the original Cantor alloy or any of the many other single-phase fcc modified
Cantor alloys described in earlier sections, we have seen that the molar Gibbs free energy of the material ΔG at a temperature T is given
quite generally by [55,56,137]:
∑ ∑
xs
ΔG = xi μi + ΔHmix (ωij , ψ ijk , χ ijkl ⋯) + RT xi lnxi − TΔSmix (ωij , ψ ijk , χ ijkl ⋯)
i i

Many of the observed multicomponent solid solutions are near-ideal or near-regular, so the heat of mixing ΔHmix and the excess
entropy of mixing ΔSxs mix are quite small, and the free energy of mixing is dominated at high temperature by a high multicomponent
configurational entropy of mixing, which stabilises a multicomponent solid solution. Stabilisation of a solid solution weakens, how­
ever, at low temperature, as the entropy terms in the free energy tend towards zero. At sufficiently low temperature, therefore, even a
small heat of mixing begins to dominate, and any solid solution becomes unstable relative to a mixture of one or more elemental
components and/or stoichiometric compounds, though whether demixing takes place depends upon whether or not the temperature is
still high enough to sustain significant levels of atomic diffusion. This is why, as remarked previously, CALPHAD calculations predict
an fcc solid solution at high temperature and a multiphase mixture at low temperature in a number of modified Cantor alloys [122].
The thermodynamic switch from a dominant configurational entropy of mixing at high temperature to a dominant heat of mixing at
low temperature explains two important microstructural features of multicomponent single-phase fcc modified Cantor alloys:

1. The range of single-phase fcc alloy compositions is broadened by using metastable high-speed processing methods such as melt
spinning or vapour deposition, compared to more conventional processing methods such as arc melting or ingot casting [13,41];
2. Very prolonged heat treatment at intermediate temperatures can lead to precipitation of other phases within the fcc phase as a
matrix [143,144].

In summary, there has been some discussion about whether the multicomponent single-phase fcc structures of the original Cantor

18
B. Cantor Progress in Materials Science 120 (2021) 100754

Table 11
Heats of mixing and interaction energies in the different binary alloys that can be made from the five components of the original
Cantor alloy [140,141].
Binary alloy Heat of mixing ΔHmix (kJ/mol) Interaction energy ωij (eV/atom)

Cr-Mn − 3.8 − 0.12


Cr-Fe 5.9 0.19
Cr-Co 5.8 0.18
Cr-Ni 6.4 0.20
Mn-Fe − 4.7 − 0.15
Mn-Co 2.2 0.07
Mn-Ni − 14.0 − 0.43
Fe-Co − 1.3 − 0.04
Fe-Ni − 4.6 − 0.14
Co-Ni 0 0

alloys and associated modified versions of it are truly stable [143,144]. The answer seems clear:

1. Multicomponent single-phase fcc Cantor alloys are thermodynamically stable at high temperature [122];
2. Multicomponent single-phase fcc Cantor alloys become thermodynamically unstable at sufficiently low temperatures [122], but
may still remain stable if precipitation and/or demixing is kinetically precluded [13,41].

5.4. Local atomic distributions in Cantor alloys

Fig. 5 shows a schematic random distribution of 5 equiatomic components on a two-dimensional square lattice, taken from the book
by Murty et al [40,41], demonstrating in very simplified form two important features of multicomponent single-phase fcc materials
such as the original Cantor alloy and the many other modified Cantor alloys described in earlier sections:

1. There is a very large number of different local atomic environments or configurations with, for example, different A atoms surrounded
by a wide range of different distributions of the other A, B, C, D and E atoms;
2. There are significant local atomic distortions and lattice strains, caused by the different sizes of the different atoms and their varied
local atomic environments.

We deal with the first of these issues in the present sub-section, and the second in the next sub-section.
Consider a general lattice with a single atom at each lattice point. Each lattice point and, therefore, each atom has n1 first near
neighbours. The cluster size of each atom together with its n1 first near neighbours is n1 + 1 atoms. The number of different clusters N1 of
n1 + 1 atoms in a multicomponent equiatomic single-phase material with c components is given by the law of permutations with
repetition4,5 [50]:

N1 = cn1 +1
If we extend the cluster size out to include second near neighbours, the number of different clusters N2 of n2 + n1 + 1 atoms is given
similarly by [47]:

N2 = cn2 +n1 +1
Each atom in a single-phase fcc material has n1 = 12 first and n2 = 6 second near neighbours [145,146]. Table 12 shows the
numbers of different local clusters N1 and N2 out to first and second near neighbours respectively, i.e. for cluster sizes of n1 + 1 = 13
and n2 + n1 + 1 = 19 respectively, in equiatomic multicomponent single-phase fcc materials such as the Cantor alloys, for different
numbers of components c ranging from c = 3 to 6. The number of different clusters, i.e. the number of different local atomic

4
Our problem is to calculate how many ways we can pick n1 + 1 (or, including second near neighbours, 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). 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 [50].
5
The number of different clusters also depends on whether or not we are concerned with cluster orientation. If clusters of different orientation are
to be regarded as different, the number of different clusters is as given in the main text. If cluster orientation is not significant, the number of
different clusters is smaller than the number given in the main text, which should be divided by the number of self-similarity operations s associated
with the crystal symmetry, i.e. N1 = (cn1 +1 )/s (or N2 = (cn2 +n1 +1 )/s including second near neighbours), with s = 24 for a cubic crystal [146,147].
This smaller number is appropriate when we are dealing with properties that are independent of orientation e.g. density. The number given in the
main text is more appropriate when we are dealing with properties that depend on orientation, e.g. dislocation slip, diffusion or magnetisation,
when the material is responding to an oriented vector of applied stress, concentration gradient or magnetic field respectively. In later sections, we
will be more concerned with properties of the latter type, so we have used the orientation-dependent formulation in the main text.

19
B. Cantor Progress in Materials Science 120 (2021) 100754

Fig. 5. Schematic representation of five components distributed randomly on a square lattice (after Murty et al [40,41]).

Table 12
Number of different local atomic clusters N1 and N2 out to first and second near neighbours
respectively for a multicomponent single-phase fcc material with c components.
Number of components c N1 N2
6
3 1.5 × 10 1.2 × 109
7
4 6.7 × 10 2.7 × 1011
5 1.2 × 109 1.9 × 1013
6 1.3 × 1010 6.1 × 1014

environments, is very large (perhaps surprisingly so) in all cases. For the original Cantor alloy with 5 components, there is a total of
more than a billion different local atomic environments including first near neighbours only; and there is a total of almost twenty
trillion local atomic environments including first and second near neighbours (the latter number is probably more relevant, since 1st
and 2nd near neighbour atomic interactions are both usually significant in fcc materials). Taking, therefore, the cube root of the latter
number, this means that a piece of the alloy sufficiently large to include all possible local atomic configurations out to second near
neighbours, i.e. a piece big enough to average reasonably over all local atomic environments and, therefore, to represent fully the alloy
and its properties, must have a linear dimension equivalent to 2.7 × 104 adjacent clusters or be approximately 27 μm in size (taking the
cluster dimension very roughly as 1 nm). In other words, individual grains smaller than 27 μm in the Cantor alloy (or any other 5-
component single-phase solid solution) can be expected to have properties that deviate from the alloy mean, and a piece of the
alloy with a grain size below 27 μm will have properties that vary from grain to grain6. This is clearly a very different situation than that
which is found in a conventional material consisting of either a single component or a single main component with one or more dilute
alloying additions. The large number of different local atomic environments in multicomponent single-phase fcc solid solutions (and
indeed in other crystal structures) is expected to play an important role in properties that depend strongly on local atomic interactions,
such as vacancy migration and diffusion, or dislocation slip and plastic flow.

5.5. Lattice distortions in Cantor alloys

Table 13 shows the 12-coordinated metallic Goldschmidt atomic radii for some of the components that have been used to
manufacture multicomponent single-phase fcc Cantor alloys [147]. The maximum difference in atomic size Δr and the maximum
atomic misfit δ amongst the five different components of the original Cantor alloy are:
Δr = rCr − rNi = 128 − 125 pm = 3 pm

6
Notice that for an equiatomic 6-component alloy, the number of different local atomic environments rises to more than 600 trillion, with a
representative piece of the material needing to be ~84μm in size. Notice also that including third near neighbours increases the cluster size to 43
atoms, and the 5-component number of different local atomic environments to ~1030 , with a representative piece of the material needing to be ~10
m in size (though 3rd near neighbour atomic interactions are usually rather small in fcc materials).

20
B. Cantor Progress in Materials Science 120 (2021) 100754

Table 13
Goldschmid atomic radii r for components used to manufacture Cantor
alloys [147].
Component Goldschmid atomic radius r (pm)

Cr 128
Mn 127
Fe 126
Co 125
Ni 125
Al 143
Cu 128

rCr − rNi 128 − 125


δ= = = 2.4%
rNi 125
Including Al, the maximum size difference and maximum misfit are larger:
Δr = rAl − rNi = 143 − 125 pm = 18 pm

rAl − rNi 143 − 125


δ= = = 14.4%
rNi 125

which explains in general terms why the maximum solubility of Al in the Cantor alloy is only 8 at% [34,63], approximately half the
equiatomic level of 16.67 at% for a 6-component alloy. And including Cu instead of Al, there is no change in the maximum difference in
atomic size and maximum misfit, which explains why the maximum solubility of Cu in the Cantor alloy is higher, reaching the
equiatomic level of 16.67%, although only when the material is manufactured by melt spinning.7
We might, therefore, expect to find lattice distortions on a scale of up to approximately 3 pm in the original Cantor alloy, and higher
in some of the modified Cantor alloys such as those containing Al. More detailed analyses of the expected lattice distortion take account
of the measured lattice parameter of the Cantor alloy fcc solid solution, use a variety of different definitions of atomic misfit and lattice
distortion, and use more sophisticated ab initio calculations of the Cantor alloy structure, but all lead to similar though slightly higher
predicted local lattice distortions of 3.3–6.6 pm [148-151].
Lattice distortions have been measured in multicomponent solid solutions such as the Cantor alloys by a variety of different
techniques, including synchrotron X-ray diffraction (XRD) [148], neutron diffraction [152] and extended X-ray fine-structure analysis
(EXAFS) [149]. Measurements are difficult because of the need to separate on the one hand static local atomic displacements caused by
lattice distortions and on the other hand dynamic displacements caused by thermal vibrations and local variations in composition and
bulk lattice strain. In the end, the best adjusted measurements give values of lattice distortion ranging from 2.2 to 4.8 pm, in reasonable
agreement with the predicted values of 3–6.6 pm. Other techniques such as atom-probe field-ion microscopy (APFIM), and high-
resolution transmission electron microscopy (HREM) have been used on other multicomponent alloys8 [153,154].
Song et al [150] and Owen and Jones [155] have provided detailed reviews of theoretical calculations and experimental mea­
surements of lattice distortion in multicomponent materials. Egami et al have shown recently [238] that local lattice distortions can be
affected as much by charge transfer as by atomic size differences.

5.6. Multicomponent Hume-Rothery rules

The Hume-Rothery rules for solubility in binary alloys have been described in a previous section. There have been many attempts to
develop extended versions of the Hume-Rothery rules suitable for application to multicomponent materials [156-170], and detailed
discussion of these attempts is provided in the book by Murty et al [41] and the review by Miracle and Senkov [13]. The basic objective
is to develop a predictive model, or at least predictive criteria, for the formation of extended solid solutions in multicomponent
materials. As with binary alloys, these attempts rely on parameterising two different basic effects, both of which act to limit inter­
solubility between the different components: firstly the lattice strain caused by different atomic sizes; and secondly the propensity for
compound formation associated with chemical interactions between the different atoms. In very general terms, and again as with
binary alloys, large differences in atomic size and valency more or less preclude extensive intersolubility. However, in multicomponent
systems many different methods can be used and indeed have been proposed for measuring the effect of differences in atomic size and
chemical activity.
The most common parameter used to determine the atomic size difference between the different components is the atomic misfit δ

7
However, the relationship between single-phase fcc stability in multicomponent materials and the range of individual component atomic sizes is
not simple and straightforward, as discussed in more detail in the next sub-section on multicomponent Hume-Rothery rules.
8
In the 5-component single-phase bcc solid solution alloy VNbMoTaW, the maximum atomic size difference and maximum atomic misfit are Δr =
7pm and δ = 5% [148], somewhat larger than in the Cantor alloy, and again in reasonable agreement with a lattice distortion of 9.5 pm measured
from the variation in lattice fringe widths in HREM [155]

21
B. Cantor Progress in Materials Science 120 (2021) 100754

calculated on a weighted root mean square basis:


√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅

xi (ri − r)2
i
δ=
r

where xi and ri are the molar fraction and atomic radius of the i‘th component and r = i xi ri is the average atomic radius [13,41,152].
And the most common parameter used to determine the chemical interaction between the different components is the heat of mixing
ΔHmix [13,41] (though it is worth noticing that the heat of mixing strictly speaking includes a combination of both size and chemical
interaction effects). Fig. 6 shows a typical Cartesian plot of the heat of mixing ΔHmix versus the atomic misfit δ, onto which is mapped a
range of different multicomponent alloys and their structures [165]. For approximately 10 kJ/mol > ΔHmix > − 10 kJ/mol and
δ < 6 − 7%, multicomponent materials tend to form single phase solid solutions; for approximately ΔHmix < − 10 kJ/mol and
δ > 6 − 7%, multicomponent materials tend to form amorphous alloys; and for ΔHmix ≈ − 10 kJ/mol and δ ≈ 6 − 7%, there is an
intermediate zone where multicomponent materials tend to form either an intermetallic compound or a mixture of a solid solution and
an intermetallic compound [165].
Recent attempts to predict the structure and properties of multicomponent materials have tried to use advanced materials
modelling techniques including: semi-empirical thermodynamic calculations such as CALPHAD [171-175]; molecular dynamics
methods, based on ab initio first-principles calculations of atomic energies [176,177]; and multi-parameter extrapolation methods
using artificial intelligence and machine learning algorithms [178-183]. While these methods are not unhelpful and have had some
degree of success, they are still severely limited by two basic problems: (1) most datasets are still largely derived from measurements
on binary and ternary alloys; and (2) most theoretical understanding relies on dilute solution approximations [171-183].

6. Properties of Cantor alloys

6.1. Diffusion in Cantor alloys

Sluggish diffusion in multicomponent solid-solution materials was first proposed by Professor Yeh and his co-workers in their
seminal paper initiating the field of (and defining the term) high-entropy alloys [11], to explain relatively slow spinodal decomposition
and precipitation of ordered bcc B2 nanoparticles in an initial fcc matrix in Alx(CrFeCoNiCu) alloys at intermediate Al contents of 8-
11at%. Sluggish diffusion was then incorporated by Professor Yeh as one of his four proposed “core” properties of high-entropy alloys
[31,32]. The first direct study of diffusion in a high-entropy alloy was by Tsai et al [140], using three pseudo-binary diffusion couples
between alloys with compositions based on the single-phase fcc Cantor alloy CrMnFeCoNi, but with composition variations of two
components in each couple. In general, the extension of Fick’s second law to a multicomponent system with c components, leads to a
total of c2 interdiffusion coefficients Dij between components i and j, each of which is composition dependent:

Dij = f (xk : k = 1to c)


For the 5-component Cantor alloy, there are, therefore, 25 interdiffusion coefficients each of which depends on 5 component

compositions, all of which can vary in any general composition field (though, of course, only four are independent since i xi = 1 at
each point). This large number of variables makes it fiendishly difficult to deconvolute general composition profiles and obtain useful
diffusion coefficients. Tsai et al [140] simplified the problem by keeping three component compositions constant in each of the three
diffusion couples, so that in each case there was effectively a binary diffusion couple between the other two components. The diffusion
profiles could then be analysed using Darken’s equations with a Matano-Boltzmann interface [184]. Moreover, Tsai et al [140] used

Fig. 6. Heat of mixing ΔHmix versus atomic misfit δ showing regions of single-phase solid solutions, amorphous alloys and intermetallic compounds
(after Guo et al [159]).

22
B. Cantor Progress in Materials Science 120 (2021) 100754

Fig. 7. Intrinsic diffusivities Di for i = Cr, Mn, Fe, Co and Ni in the multicomponent single-phase fcc Cantor alloy CrMnFeCoNi on Arrhenius plots
versus inverse homologous temperature Tm /T, compared with similar diffusivities in four stainless steels Fe-15Cr-20Ni, Fe-15Cr-45Ni, Fe-22Cr-45Ni
and Fe-15Cr-20Ni-Si, as well as self diffusion in pure Fe, Co and Ni (after Tsai et al [140]).

alloys with relatively small composition differences, so that the motion of the Matano-Boltzmann interface was negligible and the
composition dependence of interdiffusion could be ignored. In this case, the directly measured interdifusion coefficients are equal to
the intrinsic diffusivities of the diffusing components Dij = Dji = Di = Dj .
Fig. 7 shows Tsai et al’s results for the intrinsic diffusivities Di of the five different components in the Cantor alloy, plotted on
Arrhenius plots as a function of inverse homologous temperature Tm /T where Tm is the melting point, compared with similar intrinsic
diffusivities in four Fe-Cr-Ni steels, as well as self-diffusion coefficients in pure Fe, Co and Ni. In all cases, diffusivities in the Cantor
alloy are between half and one order of magnitude lower than comparable diffusivities in the steels and pure elements. Tsai et al’s work
[140] has been influential in supporting and promoting the concept of sluggish diffusion as one of the core properties of high-entropy
alloys.
Vaidya et al [185] made the first measurements of tracer diffusion coefficients for diffusion of the radioactive isotope 63Ni in the
original single-phase fcc Cantor alloy CrMnFeCoNi and the single-phase fcc medium-entropy modified Cantor alloy CrFeCoNi [185],
and have followed this up with similar measurements of tracer diffusion coefficients for the radioactive isotopes 51Cr, 54Mn, 59Fe and
57
Co in the same two single-phase fcc Cantor alloys CrMnFeCoNi and CrFeCoNi [186]. Tracer diffusion experiments are more difficult
to perform than interdiffusion experiments between dissimilar alloys, but the resulting tracer diffusion profiles are much more
straightforward to analyse compared with the complexity of deconvoluting interdiffusion profiles. The resulting tracer diffusivities D*i
are not, therefore, prone to the uncertainties associated with complex deconvolution algorithms. In their conclusions, Vaidya et al
[185,186] throw doubt on the concept of sluggish diffusion in multicomponent solid solutions. Despite this, their results show un­
equivocally that, although the component tracer diffusivities in the multicomponent alloys are not significantly slower than in similar
conventional alloys and pure metals when compared at the same absolute temperature, they are in line with Tsai et al’s results [140] in
being approximately half to one order of magnitude slower, when compared on a homologous temperature basis, as shown for Cr, Mn,
Fe and Co in Fig. 8. Comparing the two multicomponent alloys themselves shows that diffusivities in the 5-component Cantor alloy
CrMnFeCoNi are not always slower than in the 4-component modified alloy CrFeCoNi, i.e. that diffusivity is not simply proportional to
the number of components.
There have been a number of other studies of interdiffusion in multicomponent materials [187-195] and a detailed review of all the
results obtained so far [196], with substantial divergence between studies claiming to support [187-192] and oppose [193-195] the
concept of sluggish diffusion as a core property of high-entropy alloys. As a good example, the excellent study by Dabrowa et al [193]
investigated a total of 21 diffusion couples between dissimilar alloys close to the single-phase fcc original Cantor alloy CrMnFeCoNi,
the single-phase fcc medium-entropy modified Cantor alloys CrFeCoNi, CrMnCoNi and MnFeCoNi and single-phase fcc Al0.5(CrFe­
CoNi), and used a complex deconvolution algorithm, involving many detailed assumptions and approximations, fitted thermodynamic
parameters and a learning-optimisation process, to analyse the interdiffusion profiles and extract intrinsic diffusivities for all the
components in the five alloys. Like Vaidya et al [184,185], Dabrowa et al [193] concluded that diffusion in multicomponent alloys is
not sluggish, but their results again show quite clearly that component intrinsic diffusivities in the Cantor alloy and some of the 4-

23
B. Cantor Progress in Materials Science 120 (2021) 100754

Fig. 8. Tracer diffusivities Di for i = Cr, Mn, Fe and Co in the multicomponent single-phase fcc Cantor alloy CrMnFeCoNi and the modified single-
phase fcc Cantor alloys CrCoMn0.5FeNi and CrCoFeNi on Arrhenius plots versus inverse homologous temperature Tm /T, compared with similar
diffusivities in the stainless steel Fe-15Cr-20Ni, as well as self-diffusion in pure Fe, Co and Ni (after Vaidya et al [186]).

component medium-entropy alloys are slower than comparable conventional alloys. To make the point, in their conclusion Dabrowa et
al [193] say “sluggish diffusion in HEAs [high-entropy alloys] is just an artifact”, but towards the end of their discussion they say for each of
the components one at a time that the intrinsic diffusivity in “CoCrFeMnNi is significantly slower than most of the other systems [i.e.
including the medium-entropy alloys, conventional alloys and pure elements]” on a homologous temperature basis. They ascribe
slower diffusion in some of the multicomponent alloys to the presence of manganese.
Different authors have undoubtedly reached different conclusions concerning the sluggishness or otherwise of diffusion in
multicomponent materials. But these differences are not as stark as they seem. Much of the difference arises from different definitions
of “sluggish diffusion”: how slow does diffusion have to be to be “sluggish”?; do we mean that all multicomponent materials show
sluggish diffusion?; does “sluggishness” mean lower diffusivity or lower activation energy?; are we claiming that diffusivity is inversely
proportional to the number of components?; or just that there is a greater likelihood of lower diffusion rates in multicomponent
materials? Early investigators perhaps over-emphasised the effect of reduced diffusion rates, and later investigators have perhaps been
over-enthusiastic in pointing this out. It is quite clear from all the studies that:

1. Component diffusivities in multicomponent high-entropy alloys are not slower than conventional alloys at an equivalent absolute
temperature [185-192];
2. Component diffusivities are not simply proportional to the number of components in an alloy [185,186,193];
3. Component diffusivities in the Cantor alloy and some other multicomponent high- and medium-entropy alloys are slower than
comparable conventional alloys and pure elements by half to one order of magnitude when compared on a homologous temper­
ature basis [140,185,186,193].
4. There is a range of diffusivities for the different components;
5. Activation energies are broadly similar across different multicomponent and conventional alloys, in the range 200–300 kJ/mol
[140,185-195].

And in fact it is not surprising that:

24
B. Cantor Progress in Materials Science 120 (2021) 100754

1. Diffusion is somewhat slower in some multicomponent alloys at a comparable homologous temperature, notably the original
Cantor alloy, because of the lattice distortions that are present;
2. The effect is not that large, given the relatively small scale of the lattice distortions (as explained in a previous section, large lattice
distortions would lead to a breakdown of the solid-solution structure);
3. Because the effect is small, it can only be seen when comparisons are on a homologous rather than absolute temperature basis;
4. Diffusivities vary from one multicomponent material to another, because of differences in the extent of lattice distortion.
5. Diffusivities are not proportional to the number of components, since different components lead to different levels of lattice
distortion
6. Slower diffusion caused by Mn in the Cantor alloys is entirely compatible with the general proposition that multicomponent
materials sometimes exhibit slower diffusion rates – it is effectively part of the explanation of the effect, since Mn enhances lattice
distortions.

In summary, there is a small but not insignificant propensity for sluggish diffusion in multicomponent materials such as the Cantor
alloys, caused by an enhanced probability of local lattice distortions from the multiplicity of components. This effect is not that large,
depends on the particular components in the alloy, is not directly proportional to the number of components, and only manifests itself
compared with conventional materials at an equivalent homologous temperature.

6.2. Dislocation slip in Cantor alloys

In the last few years there have been quite a few detailed studies of deformation, fracture and mechanical properties in the original
Cantor alloy CrMnFeCoNi, and a number of other modified Cantor alloys, notably 4-component and 3-component medium-entropy
single-phase fcc alloys such as CrFeCoNi and CrCoNi, and Al-containing single-phase fcc alloys in the system AlxCrMnFeCoNi
[118,122,197-200,202-213]. There are excellent recent reviews of this work by Li et al [198], George et al [199] and Basu and de
Hosson [200].
Otto et al [118] manufactured billets of the Cantor alloy CrMnFeCoNi by arc melting, drop casting under argon, solution treating
for 48 h at 1200 ◦ C (1473 K), cold rolling by 87%, and annealing for 1 h at 800, 1000 and 1200 ◦ C (1073, 1273 and 1423 K
respectively), to develop homogeneous equiaxed microstructures with grain sizes of 4.4, 50 and 155μ m respectively as shown in Fig. 4,
in order to study the mechanical properties of the alloy under tensile loading conditions. Fig. 9 shows typical tensile test curves from
specimens with 4.4 and 155μ m grain sizes at temperatures ranging from − 196 to 800 ◦ C (77 to 1073 K). The yield strength, ultimate
tensile strength and ductility of the large-grain-size Cantor alloy all increase, from ~100 to 400 MPa, ~150 to 1 GPa, and ~30 to 100%
respectively, as the testing temperature is decreased progressively from 800 to − 196 ◦ C (1073 to 77 K), in all cases showing prolonged
and substantial work hardening, leading to extended ductility and a large increment of the ultimate tensile strength over the yield
strength [118]. Similarly, the yield strength, ultimate tensile strength and ductility of the small-grain-size Cantor alloy also all increase,
in this case at somewhat higher levels, from ~250 to 600 MPa, ~500 to 1.5 GPa, and ~40 to 100% respectively, as the testing
temperature is decreased progressively from 600 to − 196 ◦ C (873 to 77 K), again in all cases showing prolonged and substantial work
hardening, leading to extended ductility and a large increment of the ultimate tensile strength over the yield strength [118]. The only
exception to this general behaviour is the alloy with 4.4 μm grain size tested at 800 ◦ C (1073 K), which shows no work hardening, with
an ultimate tensile strength equal to the yield strength, and slow deformation and tearing from the yield point to final failure, caused by
extensive grain boundary deformation as might be expected for this special case of particularly small grain size and high testing
temperature.
Fig. 10 shows transmission electron micrographs of typical dislocation structures generated at 20 and − 196 ◦ C (293 and 77 K) at

Fig. 9. Tensile stress–strain curves at temperatures between − 196 and 800 ◦ C (77 and 1073 K) for the multicomponent single-phase fcc Cantor alloy
CrMnFeCoNi as-homogenised, rolled and recrystallised with grain sizes of (a) 4.4μ m and (b) 155μ m (after Otto et al [118]).

25
B. Cantor Progress in Materials Science 120 (2021) 100754

Fig. 10. Transmission electron micrographs showing dislocation structures in the multicomponent single-phase fcc Cantor alloy CrMnFeCoNi at 20
and − 196 ◦ C (273 and 77 K) at tensile strains ∊ between ~4 and 20% (after Laplanche et al [197]).

strains of between ~4 and 20%, from another investigation of tensile deformation in the multicomponent single-phase fcc Cantor alloy
CrMnFeCoNi by Laplanche et al [197]. In many ways, deformation in the Cantor alloy takes place similarly to that seen in low stacking-
fault energy (SFE) conventional fcc alloys and pure fcc metals [198-201], i.e.:

26
B. Cantor Progress in Materials Science 120 (2021) 100754

1. Initial yield takes place by simple planar slip of ½〈1 1 0〉 dislocations, often separated into Shockley partials, on {1 1 1} slip planes
[118,197-201];
2. Dislocations pile up at grain boundaries and other obstacles, initiating secondary slip, creating dislocation intersections and
increasing the number of obstacles to slip, leading to the formation of extensive dislocation tangles and forest dislocations
[118,197-201];
3. Dislocation multiplication and interactions lead to classical parabolic Taylor work hardening, with a square root dependence of
flow stress on dislocation density, and a roughly inverse parabolic dependence on plastic strain [198-201];
4. There is a significant, temperature-independent Hall-Petch effect for the flow stress at different grain sizes, with a Hall-Petch
coefficient of k ≈ 500 MPa/μm1/2 [118,198-202]; and
5. There is relatively insignificant strain-rate dependence of deformation and flow stress at normal low strain rates [199,203], but a
stronger effect at high strain rates [204,205].

There are also, however, important differences compared with conventional fcc alloys and pure fcc metals:

1. Initial yield strengths are higher, because of the presence of local lattice distortions associated with multiple components, which act
as pinning centres of strain, resisting the creation of dislocations and impeding their subsequent motion [118,197,199];
2. Dislocation lines are wavy on a near-atomic scale, again because of the presence of local lattice distortions associated with multiple
components, as they bow out between the pinning centres of strain under the action of the applied tensile stress [118,197,199];
3. A variable stacking-fault energy of approximately 15–30 mJ/m2 has been measured by X-ray diffraction [206] and by transmission
electron microscope observation of the separation of partials [207],with stacking-fault energy and partial-dislocation separation
both fluctuating by up to a factor of 2x from point-to-point along each dislocation line, again because of local lattice distortions
associated with multiple components, which directly affect the local energy penalty associated with changes in stacking sequence;
4. Work hardening is high and extends to high levels of plastic strain, again partly because of lattice distortions from multiple
components and their associated pinning effect impeding dislocation flow, with an additional contribution from deformation
twinning at high strains, high strain rates and/or low temperatures [118,197-200];
5. Necking is delayed and overall ductility enhanced, because of the combination of high initial yield strength and high and sustained
work hardening rate, since necking instability sets in according to the Considère criterion when the work-hardening rate falls below
the flow stress dσ /d∊ < σ [198];
6. Strength and ductility both increase on cooling to sub-ambient temperatures, because of the onset of extensive twinning as a major
secondary deformation mode [197].

The general features of deformation described above have been demonstrated quite widely for a variety of multicomponent single-
phase fcc Cantor alloys [118,122,197-200,202-213].
The overall deformation behaviour at normal strain rates (with the exception of the effect of grain boundary deformation at very
small grain sizes and very high temperatures) can, therefore, be described by a flow stress σ f as a function of plastic strain ∊ that
depends on an initial dislocation-free, large-grain-size yield stress σy plus a temperature- and strain-rate-dependent Taylor work-
hardening term σwh , and a temperature- and strain-rate-independent Hall-Petch grain-boundary-hardening term σ gb :

σf = σy (T, ∊)
˙ + σwh (∊, T, ∊)
˙ + σgb (d)

= σy + αMGbρ½ + kd − ½

Fig. 11. Predicted values and experimental measurements of the initial large-grain-size dislocation-free yield stress σ y (=σalloy ) for (a) the qua­
ternary single-phase fcc modified Cantor alloy CrFeCoNi and (b) the multicomponent single-phase fcc Cantor alloy CrMnFeCoNi as a function of
temperature; and (c) a range of other binary, ternary and quaternary equiatomic single-phase fcc alloys at room temperature (after Varvenne et
al [208]).

27
B. Cantor Progress in Materials Science 120 (2021) 100754

where α and k are constants, M is the Taylor factor converting applied tensile stress to resolved shear stress on the primary slip
plane, G is the shear modulus, b is the Burgers vector, ρ is the dislocation density and d is the grain size [198-200].
Varvenne et al [208,209] have developed a detailed model for the initial yield strength σ y , treating each atom as a solute in a matrix
represented by a suitable average structure over the variety of different components, and the ramifications and applications of this
model are discussed in considerable detail by George et al [199]. They stress that: (1) the average matrix structure must be taken as that
of the pure random concentrated solid solution rather than an average of the pure components; (2) the Peierls-Nabarro force is small
and can be ignored; (3) solute effects on the shear modulus are relatively limited; and (4) the main solute effect comes from solute
atoms acting as centres of strain in the lattice, i.e. acting as barriers to nucleation and motion of dislocations [199,208,209]. Dislo­
cations are wavy on a near-atomic scale, effectively being pinned at various points along their length by the strongest centres of lattice
distortion caused by the variation in local atomic size and local atomic distributions associated with multiple components. The
resulting initial yield strength σ y depends on a weighted root mean square sum of misfits of the individual solute atoms δi :
∑ 2/3 [ ]
KG( i xi δ6i ) RT ∊˙ o
σy = 2
1− ln
b ΔE ∊˙

where K is a constant, ΔE is the activation barrier for dislocation motion between pinning points, and ∊˙ o is a reference strain rate. The
first term is the initial yield strength at absolute zero, and the second term gives the temperature- and strain-rate-dependence ac­
cording to an Arrhenius law for activating dislocation motion between pinning points. The overall expression for the flow stress is then:
∑ 2/3 [ ]
KG( i xi δ6i ) RT ∊˙ o
σf = 1 − ln + αMGbρ1/2 + kd − 1/2
b2 ΔE ∊˙
Fig. 11 shows excellent agreement between predicted values of σ y and experimental results for the multicomponent single-phase fcc
Cantor alloy CrMnFeCoNi and the quaternary single-phase fcc modified Cantor alloy CrFeCoNi at different temperatures, as well as for
a range of different quaternary, ternary and binary equiatomic modified Cantor alloys at room temperature [199,208,209]. This
agreement is particularly impressive since the model has no adjustable parameters. Similar agreement has also been achieved sub­
sequently for other high- and medium-entropy single-phase fcc alloys such as (CrMnFeCo)xNi1-x [124,199], Al0.1CrMnFeCoNi [210],
CuNiRhPdIrPt [211,212] and CoNiV [199]. Recent experiments have also shown reasonable agreement between the model’s pre­
dictions for σ y and nanoindentation hardness measurements obtained from a total of 24 modified Cantor alloys, varying the con­
centration of each of the components one at a time [122]. The predicted activation volume from Varvenne et al’s model is
approximately 300b3 at room temperature, with a linear size of 6b or 7b, i.e.: (1) smaller than typical activation volumes in con­
ventional fcc alloys and pure metals, which correspond to the separation of dislocation intersections with forest dislocations along the
slip plane; but (2) in good agreement with experimental values for Cantor alloys [213]; and (3) in reasonable agreement with the
expected separation of particular solute atoms and local atomic configurations acting as pinning centres of strong lattice strain.
In summary, deformation in multicomponent single-phase fcc Cantor alloys is similar to conventional fcc alloys and pure metals,
with higher yield strengths, ultimate tensile strengths, work hardening rates, and ductilities, all caused essentially by the local lattice
distortions and strains associated with multiple components, acting to restrict dislocation generation and subsequent motion.

7. Mechanical properties of Cantor alloys

We have already seen that the original Cantor alloy CrMnFeCoNi can exhibit exceptional mechanical properties, particularly at low
temperatures, reaching an ultimate tensile strength of ~1 GPa and ductility of ~100% at − 196 ◦ C (77 K) with a fine-scale equiaxed
grain size of 4.4 μm, as described in the last section and as shown in Fig. 9. Moreover, the fracture toughness of the Cantor alloy has
been measured as ~220 MPa/√m with an equiaxed grain size of 6 μm [214], and the room-temperature high-cycle fatigue strength has
been measured as approaching 300 MPa after 107 cycles with an ultra-fine grain size of 0.65 μm [215]. These results have led to a
substantial body of research investigating: (1) other mechanical properties of the Cantor alloy itself, CrMnFeCoNi, such as
compression, torsion, wear and impact resistance; as well as (2) extensions to other single-phase fcc modified Cantor alloys, such as
medium-entropy alloys with fewer components, and variations in alloy composition; and (3) alloying additions and thermomechanical
treatments designed to enhance strength and other mechanical properties using, for instance, precipitation hardening, grain boundary
hardening, enhanced twinning response, transformation hardening, and surface treatment [198,199]. Much of this research is
described and assessed in considerable detail in the recent reviews by Li et al [198] and George et al [199]. We will not, therefore, try to
provide an extended overview of the work to date, and instead will describe the mechanical properties of one or two particularly
interesting materials. In passing, it is worth noting that Cantor alloys have also shown improvements in other properties such as ra­
diation and corrosion resistance, which we will not discuss here, since they are less well developed, and are covered in the recent
review by Miracle and Senkov [13] and the book by Murty et al [41].
Gludovatz et al studied the mechanical properties of the equiatomic ternary medium-entropy modified Cantor alloy CrCoNi [216].
At room temperature, CrCoNi with an equiaxed grain size of ~5 μm has a yield strength and fracture toughness comparable to the
Cantor alloy CrMnFeCoNi, but exhibits significantly higher ultimate tensile strength and ductility. At cryogenic temperatures the
mechanical properties of CrCoNi are even more impressive, with a yield strength, ultimate tensile strength, ductility and fracture
toughness of ~650 MPa, 1.3 GPa, 90% and 270 MPa/√m respectively. The excellent toughness at low temperatures of CrCoNi, and
indeed of CrMnFeCoNi, is caused by intense nanoscale deformation twinning, a mechanism known as twinning-induced plasticity

28
B. Cantor Progress in Materials Science 120 (2021) 100754

(TWIP) [216], with resistance to crack growth coming from bridging ligaments of material behind the crack tip [217]. Deformation
twinning is used similarly to increase work hardening and fracture toughness in TWIP steels and other alloys [218,219]. Overall,
Gludovatz et al [216] comment as follows: “in terms of…crack-initiation and crack-growth toughnesses, the NiCoCr medium-entropy alloy
represents one of the toughest materials in any materials class ever reported”.
It is now well established that Cantor alloys have low stacking-fault energies, typically 15–30 mJ/m2 at room temperature, which
decrease even further on cooling to cryogenic temperatures [68,206,220]. Lower stacking-fault energy corresponds to lower twinning
energy, which is why Cantor alloys are also able to enhance their ductility, work hardening and resistance to necking and fracture by
adding twinning as a deformation mode (TWIP) to conventional dislocation slip and multiplication [118,197,216]. However, lower
stacking-fault energy also corresponds to a lower energy barrier for the transformation from fcc to hcp ∊-martensite, since stacking
faults, twins and hcp ∊-martensite represent different but related changes to the stacking sequence in the underlying fcc structure. This
provides the possibility of another deformation mode known as transformation-induced plasticity (TRIP) [118,197,216]. Li et al [221-
223] studied modified Cantor alloys without Ni, with lower Cr and Co, and with varying Mn and Fe concentrations, in the system
Cr10MnxFe80-xCo10, to develop materials with a mixed fcc/hcp structure, so that deformation could be further enhanced by utilising
TRIP via the fcc-hcp∊-martensite phase transformation. An alloy of composition Cr10Mn30Fe50Co10 with a grain size of 4.5 μm was
shown to have a yield strength, ultimate tensile strength and ductility of 350 MPa, 900 MPa and 80% respectively [221]. The same
research group have since taken this type of alloy design further, by adding interstitial carbon and using thermomechanical treatment
to develop bimodal and trimodal grain-size structures [224,225]. A cast alloy of composition Cr10Mn30Fe49.5Co10C0.5 was hot rolled
50% at 900 ◦ C (1173 K), homogenised for 2 h at 1200 ◦ C (1473 K), water-quenched, cold rolled 67%, tempered for 3–10 min at 400 ◦ C
(673 K), and finally annealed for 3–10 min at 650–750 ◦ C (923–1023 K). The resulting alloy microstructures consisted of complex
bimodal and trimodal mixtures of highly deformed μm- and nm-sized grains and twins. The yield strength, ultimate tensile strength and
ductility at room temperature were 1.3 MPa, 1.5 MPa and 14% respectively before annealing, and 600 MPa, 900 MPa and 60% after
annealing [224]. Deformation took place, as expected, by a complex mixture of dislocation slip, twinning and martensite trans­
formation, all contributing to the resulting excellent combinations of strength and ductility [224]. Like twinning-induced plasticity
(TWIP), transformation-induced plasticity (TRIP) has also been used in conventional steels and alloys [8,198,199].
Seol et al [226] doped both the original equiatomic single-phase fcc Cantor alloy CrMnFeCoNi and a modified single-phase fcc alloy
with lower Mn and Co content (CrFe)0.8(MnCo)0.2 with boron in order to strengthen the grain boundaries. The modified alloy was
selected because of its particularly low stacking-fault energy and, therefore, a deformation mechanism dominated by twinning rather
than dislocation slip [226]. Boron is known to segregate to grain boundaries in conventional steels and other alloys, leading to two
major effects, both of which improve the mechanical properties: (1) reduced grain boundary mobility and, therefore, lower final grain
size after homogenisation and recrystallisation; and (2) increased grain boundary cohesion and strength [227-229]. Seol et al [226]
found that the addition of as little as 30 ppm boron led to considerable increments in yield strength and ultimate tensile strength,
without detriment to ductility. The Cantor alloy CrMnFeCoNi with 30 ppm boron had a room temperature yield strength, ultimate
tensile strength and ductility of 850 MPa, 1.04 GPa and almost 60% respectively. Despite these substantial improvements in me­
chanical properties, the overall deformation behaviour was, however, largely unchanged, indicating that the effect of boron was
caused directly by the expected grain boundary and grain-size hardening effects rather than by a change in deformation mode [226].
Many nickel superalloys and high-strength stainless steels have structures consisting of an fcc matrix, which are hardened by a
combination of suitable additional alloying components and thermomechanical processing so as to develop a microstructure of
coherent L12 γ’-Ni3(Al,Ti) precipitates within the fcc matrix, which confer enhanced room temperature as well as high temperature
mechanical properties [3,4,8]. The same alloying strategy has recently begun to be used for a variety of modified Cantor alloys [230-
235]. Table 14 gives details of alloy compositions, thermomechanical treatments and resulting room temperature yield strengths,
ultimate tensile strengths and ductilities. Careful homogenisation, deformation, recrystallisation and ageing treatments can be used to
develop room temperature yield strengths above 1GPa and ultimate tensile strengths above 1.5 GPa, allied with ductilities above 50%.
Microstructures typically show an fcc matrix with dense arrays of coherent L12 γ’-Ni3(Al,Ti) and sometimes other precipitates in the
size range 10–40 nm, which are clearly very effective in restricting easy dislocation flow, producing strong precipitation hardening,
and preventing the onset of premature fracture [229-234]. There have been fewer mechanical property measurements at elevated
temperatures, but Li et al [235] quote yield and ultimate tensile strengths in the range 550–620 MPa, 700–950 MPa and 28–40%
respectively for testing temperatures in the range 600–700 ◦ C (873–973 K).
Li et al [198] have followed Gludavatz et al [214] and plotted the range of mechanical properties that have been measured for
multicomponent and high entropy materials, including single-phase fcc and multiphase fcc-based Cantor alloys, on an Ashby map of
fracture toughness versus yield strength, in order to compare them with a wide range of other materials such as high-strength steels,
nickel superalloys, hard ceramics, and metallic and oxide glasses, as shown in Fig. 12. The range of mechanical properties of the high-
entropy alloys, notably the single-phase and multiphase fcc-based Cantor alloys, exceed those of all the other materials, in particular in
exhibiting exceptionally high fracture toughnesses at very high yield strengths. As Li et al [198] say: “the CrMnFeCoNi alloy represents
one of the toughest materials reported to date, with a plane-strain fracture toughness, K1c , that exceeds 200 MPa/√m and with an outstanding
tensile strength exceeding 1 GPa”. George et al [199] have also plotted the range of mechanical properties that have been measured for
multicomponent and high-entropy materials, including single-phase fcc and multiphase fcc-based Cantor alloys, on a slightly different
Ashby map of tensile tensile strength versus elongation, again to compare them with the highest strength steels and other metallic
alloys, as shown in Fig. 13. George et al say only that: “[t]he observed mechanical properties are well within those observed in martensitic
steels, advanced high strength steels, nickel-based alloys etc.” but, as can be seen on Fig. 13, the very highest combinations of strength and
ductility are once again shown by some of the fcc-based high-entropy Cantor alloys.
In summary, multicomponent single-phase and near-single-phase fcc Cantor alloys have mechanical properties that are comparable

29
B. Cantor Progress in Materials Science 120 (2021) 100754

Table 14
Mechanical properties of Cantor alloys precipitation hardened with nm-sized coherent L12 γ’-Ni3(Al,Ti) precipitates (H = homogenised, CR = cold
rolled, RX = recrystallised, YS = yield strength, UTS = ultimate tensile strength, ∊ = ductility).
Alloy Condition YS (MPa) UTS (MPa) ∊ (%) Ref

(CrFeCoNi)94Al4Ti2 H ~200 503 ~69 [230]


(CrFeCoNi)94Al4Ti2 H/CR/RX/aged1073K 645 1094 39
(CrFeCoNi)94Al4Ti2 H/CR/aged923K 1005 1273 17
CrFeCoNi2 Cu0.2Al0.2 H/CR/aged973K 720 1050 31 [231]
CrFeCoNi2 Cu0.2Al0.2 H/CR/aged1073K 460 700 32
(Cr10Fe9Co17.5Ni49)Al7.5Ti7 as-cast 859 1044 14.7 [232]
(Cr10Fe9Co17.5Ni48)Al7.5Ti7Mo1 as-cast 899 1064 14.3
(FeCoNi)86Al7Ti7 not given 1100 1500 50 [233]
(FeCoNi)86Al8Ti6 not given 1000 1250 35
(Cr15Fe21Co28Ni28)Al4Ti4 H/CR/RX/aged1073K 855 1031 27 [234]
(Cr9Fe27Co27Ni27)Al8Ti1Nb1 H/CR/RX/aged1168K [235]
H/CR/RX/aged1168K then 993 K 863 1286 28.2

Fig. 12. Ashby map of fracture toughness Kc versus yield strength σ y showing the mechanical properties of multicomponent and high-entropy
materials including single-phase fcc and multiphase fcc-based Cantor alloys compared with high-strength steels, nickel superalloys, hard ce­
ramics and metallic and oxide glasses (after Gludovatz et al [214], reprinted with permission from AAAS).

Fig. 13. Ashby map of tensile strength versus elongation showing the mechanical properties of multicomponent and high-entropy materials
including single-phase fcc and multiphase fcc-based Cantor alloys compared with high-strength steels and other metallic alloys (2nd and 3rd AHSS,
DP and TRIP are 2nd and 3rd generation advanced high-strength steels, dual-phase steels and transformation-induced plasticity steels respectively)
(after George et al [199]).

30
B. Cantor Progress in Materials Science 120 (2021) 100754

to, and in some cases exceed, those of the very strongest and toughest conventional materials, including high strength alloy steels,
nickel superalloys and advanced engineering ceramics. And we have only just begun, in the last two or three years, to try to optimise
alloy compositions and thermomechanical treatments, with an enormous range of potential compositions and treatments yet to be
explored. Indubitably, as George et al [199] go on to say: “HEAs [high-entropy alloys] provide a unique advantage in that they offer a vast
compositional space in which the different mechanisms may be tuned and optimised more effectively to develop new alloys with superior
properties”.

8. Conclusions

Multicomponent fcc-based high-entropy Cantor alloys were discovered and identified as a potentially exciting new class of ma­
terials at the end of the 1970s and the beginning of the 1980s [10]. For complex reasons their discovery was not published until the
mid-2000s [10,11], and research has only become intensive in the last 5–10 years [12,13,40-44]. During this last decade we have made
much progress, but it is very clear that we have still only just scratched the surface. The early researchers perhaps over-exaggerated
some of the special features of these new materials, but on the other hand later researchers have probably overdone their zeal in
showing the limits to these special features. The truth is somewhere between the two extremes. It was never likely, for instance, that a
large number of components would automatically guarantee high configurational entropy, formation of a single phase, slow diffusion,
or extreme resistance to dislocation flow and fracture, leading to high strength and toughness. By the same token, however, using a
large number of components not only expands our potential range of materials dramatically, but also undoubtedly makes it more
likely, depending of course on the properties of the particular components involved, that high entropy will stabilise a single-phase
structure and that local atomic strains will resist rapid diffusive atomic motion and dislocation creation and movement. We have
reached a more mature position, where we can see the the merits of multicomponent high-entropy Cantor alloys, as well as their
massive potential, at the same time as not being too carried away, and realising that their development still requires a lot of further
hard work and intensive research. The number of potentially useful materials goes up exponentially with increasing number of
components, but so too does their complexity and the corresponding difficulty of making progress, given the relative weakness of our
theoretical understanding and experimental databases that are so firmly built on single-component and dilute-solution assumptions
and simplifications.
Overall, we can make the following conclusions about multicomponent materials and the manufacture, structure and properties of
multicomponent high-entropy single-phase and near-single-phase fcc Cantor alloys:

1. There are literally billions and billions of multicomponent materials, most of which have not yet been studied.
2. Multicomponent materials can be represented in multicomponent phase space as points in a hyperdimensional polyhedron.
3. Single-phase structures are surprisingly common in multicomponent materials, for a variety of reasons, including high
configurational entropy, enhanced intersolubility, and a paucity of intrinsically new multicomponent compounds.
4. Many multicomponent materials are high-entropy single-phase or multiphase alloys, but many are not.
5. Multiphase materials and multicomponent dilute alloy materials have only been investigated to a limited extent.
6. There is an enormous single-phase fcc field in multicomponent phase space, based around the original Cantor alloy CrMnFe­
CoNi, and including modified Cantor alloys that are formed by adding, removing, replacing and varying the composition of the
alloy components.
7. There are significant, but not exceptionally severe, lattice distortions in multicomponent single-phase fcc Cantor alloys, the
extent of which depends on the particular components present in the material.
8. There is an enormous number of different local atomic configurations in multicomponent single-phase fcc Cantor alloys.
9. Atomic diffusion in single-phase fcc Cantor alloys is similar to conventional single-phase fcc alloys and pure fcc metals, but is
somewhat slower because of the greater likelihood of local lattice distortions associated with the wide range of different local
atomic configurations.
10. Dislocation slip in multicomponent single-phase fcc Cantor alloys is similar to conventional single-phase fcc alloys and pure fcc
metals, but with higher yield stresses, higher and more extended work hardening rates, and delayed necking, all arising from
local lattice distortions associated with the wide range of different local atomic distributions, which restrict dislocation
nucleation and subsequent motion.
11. Plastic flow in single-phase fcc Cantor alloys is described well by an initial large-grain-size, dislocation-free yield strength
determined by solute strengthening, plus increments from temperature-dependent parabolic Taylor work hardening, and
temperature-independent Hall-Petch grain boundary strengthening.
12. Single-phase fcc and two-phase fcc + precipitate Cantor alloys exhibit combinations of strength, ductility and fracture
toughness close to and sometimes exceeding the very best high strength steels and nickel superalloys, with considerable po­
tential for further improvement and optimisation.
13. Cantor alloys exhibit a range of other potentially valuable properties, such as resistance to corrosion and radiation damage.

Despite intense recent interest in multicomponent and high-entropy materials, the scientific and technical community continues to
be relatively unambitious in investigating the depths of multicomponent phase space. There have still been very few studies of ma­
terials with more than five components. The pioneering work by Professor Yeh’s group and my own have lifted researchers’ eyes
beyond conventional binary, ternary and dilute alloys, but not all that far. Is it possible to have a single-phase fcc material where on
average every atomic near neighbour is different (i.e. with 12 or more components)? Is it possible to have five, six or seven phases

31
B. Cantor Progress in Materials Science 120 (2021) 100754

solidifying together in a complex multiple eutectic reaction? Is it possible to dissolve (say) ten or fifteen impurity elements collectively
in a single solvent matrix? How would these more extreme structures affect transport properties such as atomic diffusion and electrical
conductivity, or mechanical properties such as strength and toughness? We just don’t know. We have learnt a lot about multicom­
ponent materials in the last decade, notably though not exclusively about single-phase fcc Cantor alloys. I look forward with great
expectation to at least another decade of similarly exciting major developments and insights into the variety of possible materials that
we can make, and their properties and applications.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

References

[1] Jones A, Prehistoric Europe: theory and practice. Oxford: Wiley-Blackwell; 2008.
[2] Browne J. Seven elements that have changed the world. London: Weidenfeld and Nicholson; 2013.
[3] Reed RC. The superalloys: fundamentals and applications. Cambridge: Cambridge University Press; 2006.
[4] Farrar JCM. The alloy tree: a guide to low-alloy steels, stainless steels and nickel-base alloys. Cambridge: Woodhead; 2004.
[5] Davis JR, editor. Aluminium and aluminium alloys. Metals Park Ohio: ASM International; 1993.
[6] Field JE, editor. The properties of natural and synthetic diamond. Cambridge: Academic Press; 1992.
[7] Grimwade M. The metallurgy of gold. Interdisc Sci Rev 1992;17:371–81.
[8] Bhadeshia H, Honeycombe R. Steels: microstructure and properties. 4th ed. Oxford: Butterworth-Heineman; 2017.
[9] Grovenor CRM. Microelectronic materials. New York: Taylor and Francis; 1998.
[10] Cantor B, Chang ITH, Knight P, Vincent AJB. Microstructural development in equiatomic multicomponent alloys. Mater Sci Eng, A 2004;375:213–8.
[11] Yeh JW, Chen SK, Lin SJ, Gan JY, Chin TS, Shun TT, et al. Nanostructured high entropy alloys with multiple principal elements: novel alloy design concepts
and outcomes. Adv Eng Mater 2004;6:299–303.
[12] Web of Science publication and citation count for “high entropy alloys” and “multicomponent alloys”.
[13] Miracle DB, Senkov ON. A critical review of high-entropy alloys and related concepts. Acta Mater 2017;122:448–511.
[14] Vincent AJB. A study of three multicomponent alloys. Part II thesis. University of Sussex Library; 1981.
[15] Barrett CS, Massalski TB. Structure of metals. Oxford: Butterworth-Heinemann; 1980.
[16] Knight P. The study of multicomponent alloy phases. Part II thesis. Department of Materials Library, University of Oxford; 1995.
[17] Cantor B, Chang ITH, Knight P, Vincent AJB. Microstructural development in equiatomic multicomponent alloys. In: O’Reilly KAQ, Warren P, Schumacher P,
Cantor B, editors. 11th International conference on rapidly quenched and metastable materials. Amsterdam: Elsevier; 2004. p. 213–8.
[18] Kim KB, Warren PJ, Cantor B. Formation of metallic glasses in novel (Ti33Zr33Hf33)((100-x-y)(Ni50Cu50)(x)Al(y) alloys. Mater Trans 2003;44:411–3.
[19] Kim KB, Warren PJ, Cantor B. Metallic glass formation in multicomponent (Ti, Zr, Hf, Nb)-(Ni, Cu, Ag)-Al alloys. J Non-Cryst Solids 2003;317:17–22.
[20] Kim KB, Warren, Cantor B. Crystallization behaviour of novel (TiZrHf)100-x-y (NiCu)x Aly with x = 48-55. J Metast Nanocryst Mater 24–25;657–60.
[21] Kim KB, Warren PJ, Cantor B. Glass-forming ability of novel multicomponent (Ti33Zr33Hf33)-(Ni50Cu50)-Al alloys developed by equiatomic substitution.
Mater Sci Eng, A 2004;375:317–21.
[22] Inoue A. Stabilization of metallic supercooled liquid and bulk metallic glasses. Acta Mater 2000;8:279–306.
[23] Zhou J, Zhang J, Zhang F, Niu B, Lei L, Wang W. High-entropy carbide: a novel class of multicomponent ceramics. Ceram Int 2018;44:22014–8.
[24] Gild J, Zhang Y, Harrington T, Jiang S, Hu T, Quinn MC, et al. High-entropy metal diborides: a new class of high-entropy materials and a new type of ultrahigh
temperature ceramics. Science Reports 2016;6:37946.
[25] Rost CM, Sachet E, Borman T, Moballegh A, Dickey EC, Hou D, et al. Entropy-stabilized oxides. Nature. Communications 2015;6:8485.
[26] Yeh JW. Private communication.
[27] Huang PK, Yeh JW, Shun TT, Chen SK. Multi-principal-element alloys with improved oxidation and wear resistance for thermal spray coating. Adv Eng Mater
2004;6:74–8.
[28] Hsu CY, Yeh JW Chen SK, Shun TT. Wear resistance and high-temperature compression strength of Fcc CuCoNiCrAl0.5Fe alloy with boron addition. Metall
Mater Trans 2004;A35:1465–9.
[29] Yeh JW, Chen SK, Gan JY, Lin SJ, Chin TS, Shun TT, et al. Formation of simple crystal structures in Cu-Co-Ni-Cr-Al-Fe-Ti-V alloys with multiprincipal metallic
elements. Metall Mater Trans A 2004;35:2533–6.
[30] Chen TK, Shun TT, Yeh JW, Wong MS. Nanostructured nitride films of multi-element high-entropy alloys by reactive DC sputtering. Surf Coat Technol 2004;
188:193–200.
[31] Yeh JW. Recent progress in high entropy alloys. Annale de Chimie Science des Matériaux 2006;31:633–48.
[32] Yeh JW. Alloy design strategies and future trends in high-entropy alloys. J Metals 2013;65:1759–71.
[33] Ranganathan S. Alloyed pleasures: multimetallic cocktails. Curr Sci 2003;85:1404–6.
[34] He JY, Liu WH, Wang H, Wu Y, Liu XJ, Nieh TJ, et al. Effects of Al addition on structural evolution and tensile properties of the FeCoNiCrMn high-entropy alloy
system. Acta Mater 2014;62:105–13.
[35] Zhu ZG, Ma KH, Wang Q, Shek CH. Compositional dependence of phase formation and mechanical properties in three CoCrFeNi-(Mn/Al/Cu) high entropy
alloys. Intermetallics 2016;79:1–11.
[36] Qin G, Wang S, Chen R, Gong X, Wang L, Su Y, et al. Microstructures and mechanical properties of Nb-alloyed CoCrCuFeNi high entropy alloys. J Mater Sci
2018;34:365–9.
[37] Zhang KB, Fu ZY, Zhang JY, Wang WM, Wang H, Wang YC. Microstructure and mechanical properties of CoCrFeNiTiAlx high-entropy alloys. Mater Sci Eng, A
2009;508:214–9.
[38] Laplanche G, Gadauad P, Bärsch C, Demtröder K, Reinhart C, Schreuer J, et al. Elastic moduli and thermal expansion coefficients of medium-entropy
subsystems of the CrMnFeCoNi high-entropy alloy. J Alloy Compd 2018;746:244–55.
[39] Abbasi E, Dehghani K. Hot tensile properties of CoCrFeMnNi(NbC) compositionally complex alloys. Mater Sci Eng, A 2020;772:138771.
[40] Murty BS, Yeh JW, Ranganathan S. High-entropy alloys. Amsterdam: Elsevier; 2014.
[41] Murty BS, Yeh JW, Ranganathan S, Bhattacharjee PP. High-entropy alloys. Amsterdam: Elsevier; 2019.
[42] Gao MC, Yeh JW, Liaw PK, Zhang Y, editors. High-entropy alloys: fundamentals and applications. Basel: Springer; 2016.
[43] Cantor B. Multicomponent and high entropy alloys. Entropy 2014;16:4749–68.
[44] Tsai MH, Yeh JW. High-entropy alloys: a critical review. Mater Res Lett 2014;2:107–23.
[45] Zhang Y, Zuo TT, Tang Z, Gao MC, Dahmen K, Liaw PK, et al. Microstructures and properties of high entropy alloys. Prog Mater Sci 2014;61:1–93.
[46] Scerri ER. The periodic table: its story and its significance. Oxford: OUP; 2019.
[47] Schädel M, Shaugnessy D, editors. The chemistry of superheavy elements. 2nd ed. Heidelberg: Springer-Verlag; 2014.
[48] Worldwide guide to equivalent irons and steels, 5th ed. Metals Park Ohio: ASM International; 2006.

32
B. Cantor Progress in Materials Science 120 (2021) 100754

[49] Worldwide guide to equivalent non-ferrous metals and alloys, 4th ed. Metals Park Ohio: ASM International; 2001.
[50] Koh KM, Chen CC. Principles and techniques in combinatorics. ch 1. Singapore: World Scientific; 1992. p. 1–50.
[51] https://www.quora.com/How-many-atoms-are-there-in-the-milky-way-galaxy.
[52] Barrow JD. The constants of nature. ch 5. New York: Vintage Books; 2002. p. 77–97.
[53] https://www.thoughtco.com/number-of-atoms-in-the-universe-603795.
[54] https://educationblog.oup.com/secondary/maths/numbers-of-atoms-in-the-universe.
[55] Gaskell DR, Laughlin DE. Introduction to the thermodynamic of materials. 6th ed. Boca Raton: CRC Press; 2017.
[56] Cantor B. The equations of materials. Oxford: Oxford University Press; 2020.
[57] Massalski TB, editor. Binary alloy phase diagrams. Ohio: ASM International; 1990.
[58] Villars P, Prince A, Okamoto H. Handbook of ternary alloy phase diagrams. Ohio: ASM International; 1995.
[59] Mizutani U. Hume-Rothery rules for structurally complex alloy phases. Boca Raton: CRC Press; 2019.
[60] Hou N, Mendis C, Chang ITH and Fan ZY. Private communication.
[61] Villars P, Calvert LD. Pearson’s handbook of crystallographic data for intermetallic phases. Ohio: ASM International; 1991.
[62] Westbrook JH, Fleischer RL, editors. Intermetallic compounds: vol 1: crystal structures of intermetallic compounds. Wiley-Blackwell; 2000.
[63] Pauzi SSM, Darham W, Ramli R, Harun MK, Talari MK. Effect of Zr addition on microstructure and properties of FeCrNiMnCoZrx and Al0.5FeCrNiMnCoZrx high
entropy alloys. Trans Indian Inst Met 2013;66:305–8.
[64] Li J, Gao B, Tang S, Liu B, Liu Y, Wang Y, et al. High temperature deformation behaviour of carbon-containing FeCoCrNiMn high entropy alloy. J Alloy Compd
2018;747:571–9.
[65] Shahmir H, Nili-Ahmadabadi M, Shafiee A, Langdon T. Effect of a minor titanium addition on the superplastic properties of a CoCrFeNiMn high-entropy alloy
processed by high pressure torsion. Mater Sci Eng, A 2018;718:468–76.
[66] Stepanov ND, Shaysultanov DG, Salischev GA, Tikhonovsky MA, Oleynik EE, Tortika AS, et al. Effect of V content on microstructure and mechanical properties
of the CoCrFeMnNiVx high entropy alloys. J Alloy Compd 2015;628:170–85.
[67] Xian X, Lin L, Zhong Z, Zhang C, Chen C, Song K, et al. Precipitation and its strengthening of Cu-rich phase in CrMnFeCoNiCux high-entropy alloys. Mater Sci
Eng 2018;A713:134–40.
[68] Liu SF, Wu Y, Wang HT, He J, Liu JB, Chen CX, et al. Stacking fault energy of face-centred-cubic high entropy alloys. Intermetallics 2018;93:269–73.
[69] Christofidou KA, Pickering EJ, Orsatti P, Mignanelli PM, Slater TJA, Stone HJ, et al. On the influence of Mn on the phase stability of the CrMnxFeCoNi high
entropy alloys. Intermetallics 2018;92:84–92.
[70] Zhu ZG, Ma KH, Yang X, Shek CH. Annealing effect on the phase stability and mechanical properties of (FeNiCrMn)100-xCox high entropy alloys. J Alloys
Comp 2017;695:2945–50.
[71] Brocq ML, Perrière L, Pirès R, Champion Y. From high entropy alloys to diluted multi-component alloys: range of existence of a solid solution. Mater Des 2016;
103:84–9.
[72] Varalakshmi S, Kamaraj M, Murty BS. Formation and stability of equiatomic and nonequiatomic nanocrystalline CuNiCoZnAlTi high-entropy alloys by
mechanical alloying. Metall Mater Trans A 2010;41:2703–9.
[73] Mohanty S, Gurao NP, Biswas K. Sinter ageing of equiatomic AlCoCuZnNi high entropy alloy via mechanical alloying. Mater Sci Eng, A 2014;617:211–8.
[74] Zhang KB, Fu ZY, Zhang JY, Wang WM, Wang H, Wang YC. Microstructure and mechanical properties of CoCrFeNiTiAl high-entropy alloys. Mater Sci Eng, A
2009;508:214–9.
[75] Mishra RK, Shahi RR. Phase evolution and magnetic characteristics of TiFeNiCr and TiFeNiCrM (M=Mn, Co) high entropy alloys. J Magn Magn Mater 2017;
442:218–23.
[76] Wang X, Xie H, Jia L, Lu Z. Effect of Ti, Al and Cu addition on structural evolution and phase constitution of FeCoNi system equimolar alloys. Mater Sci Forum
2012;724:335–8.
[77] Zhang Y, Zhou YJ, Lin JP, Chen GL, Liaw PK. Solid-solution phase formation rules for multi-component alloys. Adv Eng Mater 2008;10:534–8.
[78] Tazuddin A, Gurao NP, Biswas K. In the quest of single phase multi-component multi-principal high entropy alloys. J Alloy Compd 2017;697:434–42.
[79] Takeuchi A, Wada T, Zhang Y. MnFeNiCuPt and MnFeNiCuCo high-entropy alloys designed based on L1o structure in Pettifor map for binary compounds.
Intermetallics 2017;82:107–15.
[80] Freudenberger J, Rafaja D, Geissler D, Giebeler L, Ullrich C, Kauffman A, et al. Face centred cubic multi-component equiatomic solid solutions in the Au-Cu-Ni-
Pd-Pt system. Metals 2017;7:135.
[81] Qin G, Wang S, Chen R, Gong X, Wang L, Su Y, et al. Microstructures and mechanical properties of Nb-alloyed CoCrCuFeNi high-entropy alloys. J Mater Sci
Technol 2018;34:365–9.
[82] Wu ZF, Wang XD, Cao QP, Zhao GH, Li JX, Zhang DX, et al. Microstructure characterization of AlxCoCrCuFeNi (x=0 and 2.5) high-entropy alloy films. J Alloy
Compd 2014;609:137–42.
[83] Li C, Li JC, Zhao M, Jiang Q. Effect of alloying elements on microstructure and properties of multiprincipal elements high-entropy alloys. J Alloy Compd 2009;
475:752–7.
[84] Adomoko NK, Kim JH, Hyun YT. High temperature oxidation behaviour of low-entropy alloy to medium- and high-entropy alloys. J Therm Anal Calorim 2018;
133:13–26.
[85] Mane RB, Panigrahi BB. Effect of alloying order on non-isothermal sintering kinetics of mechanically alloyed high entropy alloy powders. Mater Lett 2018;217:
131–4.
[86] Guo S, Ng C, Wang Z, Liu CT. Solid solutioning in equiatomic alloys: limit set by topological instability. J Alloy Compd 2014;583:410–3.
[87] Singh AK, Subramaniam A. On the formation of disordered solid solutions in multi-component alloys. J Alloy Compd 2014;587:113–9.
[88] Praveen S, Murty BS, Kottada RS. Alloying behaviour in multi-component AlCoCrCuFe and NiCoCrCuFe high entropy alloys. Mater Sci Eng, A 2012;534:83–9.
[89] Praveen S, Murty BS, Kottada RS. Phase evolution and densification behaviour of nanocrystalline multicomponent high entropy alloys during spark plasma
sintering. Journal of Metals 2013;65:1797–804.
[90] Chen D, Tong Y, Li H, Wang J, Zhao YL, Hu A, et al. Helium accumulation and bubble formation in FeCoNiCr alloy under high fluence He+ implantation. J Nucl
Mater 2018;501:208–16.
[91] Huang T, Jiang L, Zhang C, Jiang H, Lu Y, Li T. Effect of carbon addition on the microstructure and mechanical properties of CoCrFeNi high entropy alloy. Sci
China Technol Sci 2018;61:117–23.
[92] Vaidya M, Pradeep KG, Murty BS, Wilde G, Divinski SV. Bulk tracer diffusion in CoCrFeNi and CoCrFeMnNi high entropy alloys. Acta Mater 2018;146:211–24.
[93] Lin Q, Liu J, An X, Wang H, Zhang Y, Liao X. Cryogenic-deformation-induced phase transformation in a FeCoCrNi high-entropy alloy. Mater Res Lett 2018;6:
236–43.
[94] Zuo T, Gao MC, Ouyang L, Yang X, Cheng Y, Feng R, et al. Tailoring magnetic behaviour of CoFeMnNix (x=Al, Cr, Ga and Sn) high entropy alloys by metal
doping. Acta Mater 2017;130:10–8.
[95] Fazakas E, Zadorozhnyy V, Louzguine-Luzgin DV. Effect of iron content on the structure and mechanical properties of Al20Ti20Ni20Cu20 and
(AlTi)60-xNi20Cu20Fex (x=15, 20) high-entropy alloys. Appl Surf Sci 2015;358:549–55.
[96] Durga A, Hari Kumar KC, Murty BS. Phase formation in equiatomic high entropy alloys: CALPHAD approach and experimental studies. Trans Indian Inst Met
2012;65:375–80.
[97] Bashev VF, Kushnerov OI. Structure and properties of cast and splat quenched high entropy AlCuFeNiSi alloys. Phys Met Metall 2017;118:239–47.
[98] Chao Q, Guo T, Jarvis T, Wu X, Hodgson P, Fabijanic D. Direct laser deposition cladding of AlxCoCrFeNi high entropy alloys on a high -temperature stainless
steel. Surface Coatings Technology 2017;332:440–51.
[99] Cai Z, Chen Y, Manladan SM, Luo Z, Gao F, Li L. Influence of dilution rate on the microstructure and properties of FeCrCoNi high-entropy alloy coating. Mater
Des 2018;142:124–37.

33
B. Cantor Progress in Materials Science 120 (2021) 100754

[100] Wu B, Chen W, Jiang Z, Chen Z, Fu Z. Influence of Ti addition on microstructure and mechanical behaviour of an fcc-based Fe30Ni30Co30Mn10 alloy. Mater Sci
Eng, A 2016;676:492–500.
[101] Lin CM, Tsai HL. Evolution of microstructure, hardness and corrosion properties of high entropy Al0.5CoCrFeNi alloy. Intermetallics 2011;19:288–94.
[102] Ma SG, Zhang SF, Gao MC, Liaw PK, Zhang Y. A successful synthesis of the CoCrFeNiAl0.3 single-crystal high-entropy alloy by Bridgman solidification. J Metals
2013;65:1751–8.
[103] Tian F, Delczeg L, Chen N, Varga LK, Shen J, Vitos L. Structural stability of NiCoFeCrAlx high-entropy alloy from ab initio theory. Phys Rev B: Condens Matter
2013;88:085128–32.
[104] Cieslak J, Tobola J, Berent K, Marciszko M. Phase composition of AlxFeNiCrCo high entropy alloys by sintering and arc-melting methods. J Alloy Compd 2018;
740:264–72.
[105] Zhang Y, Zuo TT, Cheng YQ, Liaw PK. High entropy alloys with high saturation magnetization, electrical resistivity and malleability. Science Reports 2013;3:
1–7.
[106] Zuo TT, Ren SB, Liaw PK, Zhang Y. Processing effects on the magnetic and mechanical properties of FeCoNiAl0.2 Si0.2 high entropy alloy. Int J Miner Metall
Mater 2013;20:549–55.
[107] Hsu YJ, Chiang WC, Wu JK. Corrosion behaviour of FeCoNiCrCux high-entropy alloys in 3.5% sodium chloride solution. Mater Chem Phys 2005;92:112–7.
[108] Dahlborg U, Cornide J, Dahlborg MC, Hansen TC, Leong Z, Dominguez LA, et al. Crystalline structures of some high entropy alloys obtained by neutron and X-
ray diffraction. Acta Phys Pol A 2015;128:552–6.
[109] Cai Y, Manladan SM, Luo Z. Tribological behaviour of the double FeCoNiCrCux middle entropy alloy coatings. Surf Eng 2019;35:14–21.
[110] Bridges D, Zhang S, Lang S, Gao M, Yu Z, Feng Z, et al. Laser brazing of a nickel-based superalloy using a NiMnFeCoCu high entropy filler material. Mater Lett
2018;215:11–4.
[111] Ren B, Liu ZX, Li DM, Shi L, Cai B, Wang MX. Effect of elemental interaction on microstructure of CuCrFeNiMn high entropy alloy system. J Alloy Compd 2010;
493:148–53.
[112] Ren B, Liu ZX, Cai B, Wang MX, Shi L. Ageing behaviour of a CuCrFeNiMn high-entropy alloy. Mater Des 2012;33:121–6.
[113] Ming K, Bi X, Wang J. Precipitation strengthening of ductile CrFeCoNiMo alloys. Scr Mater 2017;137:88–93.
[114] Kim JH, Lim KR, Won JW, Na YS, Kim HS. Mechanical properties and deformation twinning behaviour of as-cast CoCrFeNiMn high-entropy alloy at low and
high temperatures. Mater Sci Eng, A 2018;712:108–13.
[115] Laurent-Brocq M, Akhatova A, Perrière L, Chebini S, Sauvage X, Leroy E, et al. Insights into the phase diagram of the CrMnFeCoNi high entropy alloy. Acta
Mater 2015;88:355–65.
[116] Flemings MC. Solidification processing. New York: McGraw-Hill; 1974.
[117] Glicksmann ME. Principles of solidification. New York: Springer; 2011.
[118] Otto F, Diouhý A, Somsen C, Bei H, Eggeler G, George EP. The influence of temperature and microstructure on the tensile properties of a CoCrFeNiMn high-
entropy alloy. Acta Mater 2013;61:5743–55.
[119] Kim SW, Kim JH. In-situ observations of deformation twins and crack propagation in a CoCrFeNiMn high-entropy alloy. Mater Sci Eng 2018;A718:113–9.
[120] Jang MJ, Praveen S, Sung HJ, Bae JW, Moon J, Kim HS. High temperature tensile deformation of hot rolled CrMnFeCoNi high-entropy alloy. J Alloy Compd
2018;730:242–8.
[121] Luo H, Li Z, Mingers AM, Raabe D. Corrosion behaviour of an equiatomic CoCrFeMnNi high-entropy alloy compared with 304 stainless steel in sulphuric acid
solution. Corros Sci 2018;134:131–9.
[122] Bracq G, Laurent-Brocq M, Varvenne C, Perrière Curtin WA, Joubert J-M, Guillot I. Combining experiments and modelling to explore the solid solution
strengthening of high and medium entropy alloys. Acta Mater 2019;177:266–79.
[123] Laurent-Brocq M, Perrière L, Pirès R, Champion Y. From high entropy alloys to diluted multi-component alloys: range of existence of a solid-solution. Mater
Des 2016;103:84–9.
[124] Laurent-Brocq M, Perrière L, Pirès R, Prima F, Vermaut P, Champion Y. From diluted solid solutions to high entropy alloys: on the evolution of properties with
composition of multi-component alloys. Mater Sci Eng, A 2017;696:228–35.
[125] Fan AC, Li JH, Tsai MH. On the phase constituents of three CoCrFeNiX (X=V, Nb, Ta) high-entropy alloys after prolonged annealing. J Alloy Compd 2020;823:
153524.
[126] Ai C, Wang G, Liu L, Guo M, He F, Zhou J, et al. Effect of Ta addition on solidification characteristics of CoCrFeNiTax eutectic high entropy alloys.
Intermetallics 2020;120:106769.
[127] Dai C, Zhao T, Du C, Liu Z, Zhang D. Effect of molybdenum content on the microstructure and corrosion behaviour of FeCoCrNiMox high-entropy alloys.
J Mater Sci Technol 2020;46:64–73.
[128] Mishra RK, Shahi R. A systematic approach for enhancing magnetic properties of CoCrFeNiTi-based high entropy alloys via stoichiometric variation and
annealing. J Alloy Compd 2020;821:153534.
[129] Tawancy HM. On the tensile strength of medium entropy FeNiCrCoMoW alloy with high microstructural stability. Mater Sci Eng, A 2020;781:139239.
[130] Peng H, Hu L, Li L, Zhang W. Ripening of L12 nanoparticles and their effects on the mechanical properties of NiCoFeCrAlTi high-entropy alloys. Mater Sci Eng,
A 2020;772:138803.
[131] Tang L, Yan K, Cai B, Wang Y, Liu B, Kabra S, et al. Deformation mechanisms of FeCoCrNiMo0.2 high entropy alloy at 77 and 15K. Scr Mater 2020;178:166–70.
[132] Hsu YC, Li CL, Hsueh CH. Effects of Al addition on microstructure and mechanical properties of CoCrFeMnNiAlx high entropy alloy films. Entropy 2020;22.
article number 2.
[133] Lin L, Xian X, Zhong Z, Chen C, Zhu Z, Wu Y, et al. A multiphase CrMnFeCoNiAl0.75 high entropy alloy with high strength at intermediate temperature.
Intermetallics 2020;120:196744.
[134] Wang Q, Ma Y, Jiang BB, Li XN, Shi Y, Dong C, et al. A cuboidal B2 nanoprecipitation-enhanced body-centered-cubic alloy Al0.7CoCrFe2Ni with prominent
tensile properties. Scr Mater 2016;120:85–9.
[135] Lu Y, Dong Y, Guo S, Jiang L, Kang H, Wang T, et al. A promising new class of high-temperature alloys: eutectic high-entropy alloys. Science Reports 2014;4.
article number 6200.
[136] Shi P, Ren W, Zheng T, Ren Z, Hou X, Peng J, et al. Enhanced strength–ductility synergy in ultrafine-grained eutectic high-entropy alloys by inheriting
microstructural lamellae. Nature Commun 2019;10. article number 489.
[137] Hillert M. Phase equilibria, phase diagrams and phase transformations: their thermodynamic basis. Cambridge: Cambridge University Press; 2008.
[138] Andersson J, Guillermet AF, Hillert M. A compound energy model of ordering in a phase with sites of different coordination numbers. Acta Metall 1975;34:
437–45.
[139] Muggianu YM, Gambino M, Bros JP. Enthalpies of formation of liquid alloys bismuth-gallium-tin at 273K – choice of an analytical representation of integral
and partial thermodynamic functions of mixing for this ternary system. J Chim Phys Phys- Chim Biol 1975;72:83–8.
[140] Tsai KY, Tsai MH, Yeh JW. Sluggish diffusion in CoCrFeMnMi high-entropy alloys. Acta Mater 2013;61:4887–97.
[141] Hultgren R. Selected values of the thermodynamic properties of binary alloys. Cleveland: ASM International; 1973.
[142] Otto F, Yang Y, Bei H, George EP. Relative effects of enthalpy and entropy on the phase stability of equiatomic high-entropy alloys. Acta Mater 2013;61:
2628–38.
[143] Otto F, Diouhý A, Pradeep KG, Kubenová M, Raabe D, Eggeler G, et al. Decomposition of the single-phase high-entropy alloy CrMnFeCoNi after prolonged
anneals at intermediate temperatures. Acta Mater 2016;112:40–52.
[144] Park N, Lee BJ, Tsuji N. The phase stability of equiatomic CoCrFeMnNi high-entropy alloy: comparison between experiment and calculation results. J Alloy
Compd 2017;719:188–93.
[145] Philips FC. An introduction to crystallography. New York: Wiley; 1973.
[146] Kelly A, Knowles KM. Crystallography and crystal defects. New York: Wiley; 2012.

34
B. Cantor Progress in Materials Science 120 (2021) 100754

[147] Greenwood NN, Earnshaw A. Chemistry of the elements (2nd edition). Butterworth-Heinemann; 1997.
[148] Okamoto NL, Yuge K, Tanaka K, Inui H, George EP. Atomic displacement in the CrMnFeCoNi high entropy alloy: a scaling factor to predict solid solution
strengthening. Am Instit Phys 2016;6 [number 125008].
[149] Oh HS, Ma D, Leyson GP, Grabowski B, Park ES, Körmann Raabe D. Lattice distortions in the FeCoNiCrMn high entropy alloy studied by theory and
experiment. Entropy 2016;18:321–9.
[150] Song H, Tian F, Hu QM, Vitos L, Wang Y, Shen J, Chen N. Local lattice distortions in high-entropy alloys. Phys Rev Mater 2018;1 [number 023404].
[151] Yen CC, Huang GR, Huang YC, Yeh HW, Luo DJ, Hsieh KT, Huang EW, Yeh JW, Lin SJ, Wang CC, Kuo CL, Chang SY, Lo YC. Lattice distortion effect on elastic
anisotropy of high entropy alloys. J Alloys Comp 2020;818 [number 152876].
[152] Owen LR, Pickering EJ, Playford HY, Stone HJ, Tucker MG, Jones NG. An assessment of the lattice strain in the CrMnFeCoNi high-entropy alloy. Acta Mater
2017;122:11–8.
[153] Zou Y, Maiti S, Steurer W, Spolensk R. Size-dependent plasticity in an Nb25Mo25Ta25W25 refractory high-entropy alloy. Acta Mater 2014;65:85–97.
[154] Zou Y, Maiti S, Steurer W, Spolenak R. Size-dependent plasticity in a NbMoTaW refractory high-entropy alloy. Acta Mater 2014;65:85–97.
[155] Owen LR, Jones NG. Lattice distortions in high-entropy alloys. J Mater Res 2018;33:2954–69.
[156] Zhang Y, Zhou YJ, Lin JP, Chen GL, Liaw PK. Solid-solution phase for- mation rules for multi-component alloys. Adv Eng Mater 2008;10:534–8.
[157] Yang X, Zhang Y. Prediction of high-entropy stabilized solid-solution in multi-component alloys. Mater Chem Phys 2012;132:233–8.
[158] Zhang Y, Yang X, Liaw PK. Alloy design and properties optimization of high-entropy alloys. Journal of Metals 2012;64:830–8.
[159] Guo S, Hu Q, Ng C, Liu CT. More than entropy in high-entropy alloys: forming solid solutions or amorphous phase. Intermetallics 2013;41:96–103.
[160] Takeuchi A, Amiya K, Wada T, Yubuta K, Zhanb W, Makino A. Entropies in alloy design for high entropy and bulk glassy alloys. Entropy 2013;15:3810–21.
[161] Ren MX, Li BX, Fu HZ. Formation condition of solid solution type high-entropy alloy. Trans Nonferr Metals Soc China (English Edition) 2013;23:991–5.
[162] Poletti MG, Battezzati L. Electronic and thermodynamic criteria for the occurrence of high entropy alloys in metallic systems. Acta Mater 2014;75:297–306.
[163] Salishchev GA, Tikhonovsky MA, Shaysultanov DG, Stepanov ND, Kuznetsov AV, Kolodiy IV, et al. Effect of Mn and V on structure and mechanical properties
of high-entropy alloys based on CoCr- FeNi system. J Alloy Compd 2014;591:11–21.
[164] Nong ZS, Zhu JC, Cao Y, Yang XW, Lai ZH, Liu Y. Stability and structure prediction of cubic phase in as cast high entropy alloys. Mater Sci Technol 2014;30:
363–9.
[165] Guo S. Phase selection rules for cast high entropy alloys: an overview. Mater Sci Technol 2015;31:1223–30.
[166] Wang Z, Huang Y, Yang Y, Wang J, Liu CT. Atomic size effect and solid solubility of multicomponent alloys. Scr Mater 2015;94:28–31.
[167] Ye YF, Wang Q, Lu J, Liu CT, Yang Y. The generalized thermodynamic rule for phase selection in multicomponent alloys. Intermetallics 2015;59:75–80.
[168] King DJM, Middleburgh SC, Mc Gregor AG, Cortie MB. Predicting the formation and stability of single phase high-entropy alloys. Acta Mater 2016;104:172–9.
[169] Senkov ON, Miracle DB. A new thermodynamic parameter to to predict formation of solid solutions or intermetallic phases in high entropy alloys. J Alloy
Compd 2016;658:603–7.
[170] Caraballo IT, Rivera-Diaz-del-Castillo PEJ. A criterion for the formation of high entropy alloys based on lattice distortion. Intermetallics 2016;71:76–87.
[171] Yang S, Lu J, Xing F, Zhang LJ, Zhong Y. Revisit the VEC rule in high entropy alloys with high-throughput CALPHAD approach and its applications for
materials design – a case study with AlCoCrFeNi. Acta Mater 2020;192:11–9.
[172] Gorsse S, Senkov ON. About the reliability of CALPHAD predictions in multicomponent systems. Entropy 2018;20. article number 899.
[173] Gao MC, Zhang C, Gao P, Zhang F, Ouyang LZ, Widom M, et al. Thermodynamics of concentrated solid solution alloys. Curr Opin Solid State Mater Sci 2017;
21:238–51.
[174] Chen HL, Mao H, Chen Q. Database development and Calphad calculations for high entropy alloys: challenges, strategies and tips. Mater Chem Phys 2018;210:
279–90.
[175] Choi WM, Jung S, Jo YH, Lee S, Lee BJ. Design of new fcc high entropy alloys by thermodynamic calculation. Met Mater Int 2017;23:839–47.
[176] Feng R, Liaw PK, Gao MC, Widom M. First-principles prediction of high-entropy-alloy stability. npj Comput Mater 2017;3.
[177] Ikeda Y, Grabowski B, Körmann F. Ab initio phase stabilities and mechanical properties of multicomponent alloys: a comprehensive review for high entropy
alloys and compositionally complex alloys. Mater Charact 2019;147:464–511.
[178] Pei ZR, Yin JQ, Hawk JA, Alman DE, Gao MC. Machine-learning informed prediction of high-entropy solid solution formation: beyond the Hume-Rothery rules.
npj Comput Mater 2020;6:issue 1.
[179] Choudhury A, Konnur T, Chattopadhyay PP. Structure prediction of multi-principal element alloys using ensemble learning. Eng Comput 2019;37:1003–22.
[180] Li Y, Guo WL. Machine learning model for predicting phase formations of high-entropy alloys. Phys Rev Mater 2019;3:095005.
[181] Huang W, Martin P, Zhuang HL. Machine learning phase prediction of high-entropy alloys. Acta Mater 2019;169:225–36.
[182] Wen C, Zhang Y, Wang C. Machine learning assisted design of high entropy alloys with desired property. Acta Mater 2019;170:109–17.
[183] Zhang Y, Wen C, Wang C. Phase prediction in high entropy alloys with a rational selection of materials descriptors and machine learning models. Acta Mater
2020;185:528–39.
[184] Christian JW. The theory of transformations in metals and alloys. 3rd ed. Oxford: Pergamon; 2002.
[185] Vaidya M, Trubel S, Murty BS, Wilde G, Divinsky S. Ni tracer diffusion in CoCrFeNi and CoCrFeMnNi high entropy alloys. J Alloy Compd 2016;688:994–1001.
[186] Vaidya M, Pradeep KG, Murty BS, Wilde G, Divinsky SV. Bulk tracer diffusion in CoCrFeNi and CoCrFeMnNi high entropy alloys. Acta Mater 2018;146:211–24.
[187] Dabrowa J, Kucza W, Cieslak G, Kulik T, Danielewski M, Yeh JW. Interdiffusion in the FCC-structured Al-Co-Cr-Fe-Ni high entropy alloys: experimental studies
and numerical simulations. J Alloy Compd 2016;674:455–62.
[188] Beke DL, Erdélyi G. On the diffusion in high-entropy alloys. Mater Lett 2016;164:111–3.
[189] Kulkarni K, Chauhan GPS. Investigations of quaternary interdiffusion in a constituent system of high entropy alloys. American Institute of Physics. Advances
2015;vol 5 097162.
[190] Chen W, Zhang L. High-throughput determination of interdiffusion coefficients for Co-Cr-Fe-Mn-Ni high-entropy alloys. J Phase Eq Diff 2017;38:457–65.
[191] Li Q, Chen W, Zhong J, Zhang L, Chen Q, Liu Z. On sluggish diffusion in fcc AlCoCrFeNi high-entropy alloys: an experimental and numerical study. Metals
2018;8. article number 16.
[192] Wang R, Chen W, Zhoung J, Zhang L. Experimental and numerical studies on the sluggish diffusion in face centered cubic Co-Cr-Cu-Fe-Ni high-entropy alloys. J
Mater Sci Technol;34:1791–8.
[193] Dabrowa J, Zajusz M, Kucza W, Cieslak G, Berent K, Czeppe T, et al. Demystifying the sluggish diffusion effect in high entropy alloys. J Alloy Compd 2019;783:
193–207.
[194] Zhang C, Zhang F, Jin K, Bei H, Chen S, W. X Cao WX, Zhu J, Lv D. Understanding of the elemental diffusion behavior in concentrated solid solution alloys. J
Phase Equilibr Diffus 2017;38:434–44.
[195] Vaidya M, Pradeep KG, Murty BS, Wilde G, Divinski S. Radioactive isotopes reveal a non sluggish kinetics of grain boundary diffusion in high entropy alloys.
Sci Rep 2017;7. article number 12293.
[196] Dabrowa J, Danielewsky M. State-of-the-art diffusion studies in the high entropy alloys. Metals 2020;10:347–91.
[197] Laplanche G, Kostka A, Horst OM, Eggeler G, George EP. Microstructure evolution and critical stress for twinning in the CrMnFeCoNi high-entropy alloy. Acta
Mater 2016;118:152–63.
[198] Li Z, Zhao S, Ritchie RO, Meyers MA. Mechanical properties of high-entropy alloys with emphasis on face-centered cubic alloys. Prog Mater Sci 2019;102:
296–345.
[199] George EP, Curtin WA, Tasan CC. High entropy alloys: A focused review of mechanical properties and deformation mechanisms. Acta Mater 2020;188:435–74.
[200] Basu I, de Hosson JTM. Strengthening mechanisms in high-entropy alloys: fundamental issues. Scr Mater 2020. in press.
[201] Honeycombe RWK. The Plastic Deformation of Metals. 2nd ed. Maidenhead: Edward Arnold; 1984.
[202] Liu W, Wu Y, He J, Nieh T, Lu Z. Grain growth and the Hall-Petch relationship in a high-entropy FeCrNiCoMn alloy. Scr Mater 2013;68:526–9.
[203] Gali A, George EP. Tensile properties of high- and medium-entropy alloys. Intermetallics 2013;39:74–8.

35
B. Cantor Progress in Materials Science 120 (2021) 100754

[204] Kang M, Won JW, Kwon JB, Na YS. Intermediate strain rate deformation behavior of a CoCrFeMnNi high-entropy alloy. Mater Sci Eng, A 2017;707:16–21.
[205] Park JM, Moon J, Bae JW, Jang MJ, Park J, Lee S, et al. Strain rate effects of dynamic compressive deformation on mechanical properties and microstructure of
CoCrFeMnNi high-entropy alloy. Mater Sci Eng, A 2018;719:155–63.
[206] Zaddach AJ, Niu C, Koch CC, Irving DL. Mechanical properties and stacking fault energies of NiFeCrCoMn high-entropy alloy. J Met 2013;65:1780–9.
[207] Okamoto NL, Fujimoto S, Kambara Y, Kawamura M, Chen ZMT, Matsunoshita H, et al. Size effect, critical resolved shear stress, stacking fault energy, and solid
solution strengthening in the CrMnFeCoNi high-entropy alloy. Sci Rep 2016;6. article number 35863.
[208] Varvenne C, Luque A, Curtin WA. Theory of strengthening in fcc high entropy alloys. Acta Mater 2016;118:164–76.
[209] Varvenne C, Leyson GPM, Ghazisaeidi M, Curtin WA. Solute strengthening in random alloys. Acta Mater 2017;124:660–83.
[210] Varvenne C, Curtin WA. Strengthening of high entropy alloys by dilute solute additions: CoCrFeNiAlx and CoCrFeNiMnAlx alloys. Scr Mater 2017;138:92–5.
[211] Sohn S, Liu Y, Liu J, Gong P, Prades-Rodel S, Blatter A, et al. Noble metal high entropy alloys. Scr Mater 2017;126:29–32.
[212] Yin B, Curtin WA. First-principles-based prediction of yield strength in the RhIrPdPtNiCu high-entropy alloy. npj Comput Mater 2019;5.
[213] Laplanche G, Bonneville J, Varvenne C, Curtin WA, George EP. Thermal activation parameters of plastic flow reveal deformation mechanisms in the
CrMnFeCoNi high-entropy alloy. Acta Mater 2018;143:257–64.
[214] Gludovatz B, Hohenwarter A, Catoor D, Chang EH, George EP, Ritchie RO. A fracture-resistant high-entropy alloy for cryogenic applications. Science 2014;345:
1153–8.
[215] Tian YZ, Sun SJ, Lin HR, Zhang ZF. Fatigue behavior of CoCrFeMnNi high-entropy alloy under fully reversed cyclic deformation. J Mater Sci Technol 2019;35:
334–40.
[216] Gludovatz B, Hohenwarter A, Thurston KVS, Bei H, Wu Z, George EP, et al. Exceptional damage-tolerance of a medium-entropy alloy NiCoCr at cryogenic
temperatures. Nature. Communications 2016;7. article number 10602.
[217] Zhang Z, Mao MM, Wang J, Gludovatz B, Zhang Z, Mao SX, et al. Nanoscale origins of the damage tolerance of the high-entropy alloy CrMnFeCoNi. Nature.
Communications 2015;6. article number 10143.
[218] Bintu A, Vincze G, Picu CR, Lopes AB, Grácio JJ, Barlat F. Strain hardening rate sensitivity and strain rate sensitivity in TWIP steels. Mater Sci Eng, A 2015;629:
54–9.
[219] Byun T, Hashimoto N, Farrell K. Temperature dependence of strain hardening and plastic instability behaviors in austenitic stainless steels. Acta Mater 2004;
52:3889–99.
[220] Huang S, Li W, Lu S, Tian FY, Shen J, Holmstrom E, Vitos L. Temperature dependent stacking fault energy of FeCrCoNiMn high entropy alloy. Scripta
Materialia 2015;108:44–7.
[221] Li Z, Pradeep KG, Deng Y, Raabe D, Tasan CC. Metastable high-entropy dual-phase alloys overcome the strength–ductility trade-off. Nature 2016;534:227–30.
[222] Li Z, Kormann F, Grabowski B, Neugebauer J, Raabe D. Ab initio assisted design of quinary dual-phase high-entropy alloys with transformation-induced
plasticity. Acta Mater 2017;136:262–70.
[223] Li Z, Tasan CC, Pradeep KG, Raabe D. A TRIP-assisted dual-phase high-entropy alloy: grain size and phase fraction effects on deformation behavior. Acta Mater
2017;131:323–35.
[224] Su J, Raabe D, Li Z. Hierarchical microstructure design to tune the mechanical behavior of an interstitial TRIP-TWIP high-entropy alloy. Acta Mater 2019;163:
40–54.
[225] He ZF, Jia N, Ma D, Yan HL, Li ZM, Raabe D. Joint contribution of transformation and twinning to the high strength-ductility combination of a FeMnCoCr high
entropy alloy at cryogenic temperatures. Mater Sci Eng, A 2019;759:437–47.
[226] Seol JB, Bae JW, Li Z, Han JC, Kim JG, Raabe D, et al. Boron doped ultrastrong and ductile high-entropy alloys. Acta Mater 2018;151:366–76.
[227] Wu R, Freeman AJ, Olson GB. First principles determination of the effects of phosphorus and boron on iron grain boundary cohesion. Science 1994;15:376–80.
[228] Lejĉek P, Hofmann S, Paidar V. Segregation based classification of [100] tilt grain boundaries in α-iron and its consequences for grain boundary engineering.
Acta Mater 2003;51:3951–63.
[229] Raabe D, Herbig M, Sandlobes Y, Li Y, Tytko D, Kuzmina M, et al. Grain boundary segregation engineering in metallic alloys: a pathway to the design of
interfaces. Current Opinions in Solid State. Mater Sci 2014;18:253–61.
[230] He JY, Wang H, Huang HL, Xu XD, Chen MW, Wu Y, et al. A precipitation-hardened high-entropy alloy with outstanding tensile properties. Acta Mater 2016;
102:187–96.
[231] Wang ZG, Zhou W, Fu LM, Wang JF, Luo RC, Han XC, et al. Effect of coherent L12 nanoprecipitates on the tensile behavior of a fcc-based high-entropy alloy.
Mater Sci Eng, A 2017;696:503–10.
[232] Zhang L, Zhou Y, Jin X, Du X, Li B. Precipitation-hardened high entropy alloys with excellent tensile properties. Mater Sci Eng, A 2018;732:186–91.
[233] Yang T, Zhao YL, Tong Y, Jiao ZB, Wei J, Cai JX, et al. Multicomponent intermetallic nanoparticles and superb mechanical behaviors of complex alloys. Science
2018;362:933–7.
[234] Peng H, Hu L, Li L, Zhang W. Ripening of L12 nanoparticles and their effects on mechanical properties of Ni28Co28Fe21Cr15Al4Ti4 high-entropy alloys. Mater Sci
Eng, A 2020;772:138803.
[235] Li Z, Fu L, Peng J, Zheng H, Ji X, Sun Y, et al. Improving mechanical properties of an FCC high-entropy alloy by γ′ and B2 precipitates strengthening. Mater Sci
Charact 2020;159:109989.
[236] Kube SA, Schroers J. Metastability in high entropy alloys. Scr Mater 2020;186:392–400.
[237] Mackay AL. On complexity. Crystallogr Rep 2001;46:524–6.
[238] Egami T, Odbadrakh K, Oh H. to be published.

Brian Cantor is Professor of Materials at Oxford and Brunel Universities. He is also Chair of the World Technology Universities Network, a Trustee of the Science
Museums Group and an Editor of Elsevier’s premier review journal, Progress in Materials Science. He is acknowledged as a world expert on materials manufacturing, has
published over 300 books and papers and is on the ISI List of Most Cited Researchers. He invented the field of High Entropy Alloys and discovered “Cantor alloys”. He was
awarded a CBE for services to higher education in 2013. He was educated at Manchester Grammar School and Christ’s College, Cambridge University. In the recent past,
he has been Vice-Chancellor of the Universities of Bradford and York, Head of Mathematical and Physical Sciences at Oxford, a research scientist at GE Labs in the USA, a
consultant for Alcan, NASA and Rolls-Royce, and a Vice-President of the Royal Academy of Engineering. He has worked at other universities such as Sussex, North­
eastern, Banaras, Washington State and IISc Bangalore, and chaired or been on boards such as the Marshall Aid Commission, UK Universities Pensions Forum, the
National Science and Media Museum, the Kobe Institute, Oxford Innovation, York Science Park, and Bradford, Leeds and York Local Economic Partnerships (LEPs). He
founded and built up the Begbroke Science Park, the Heslington East Campus, the Hull-York Medical School, the Wolfson Centre for Applied Health Research, the Digital
Health Enterprise Zone, the York Neuroimaging Centre, the Institute for Effective Education, the Centre for Applied Human Rights, the National Science Learning Centre,
and the World Technology Universities Network. He was awarded the Ismanam Prize, the York Lifetime Achievement Award, and the Rosenhain and Platinum Medals of
the Institute of Materials. He is an Honorary Professor at Shenyang, Zhejiang and Nanjing Universities, the Chinese Institute of Materials Research and the Indian
Institute of Sciences Bangalore; an Honorary Member of the Indian Institute of Metals; a Member of Academia Europea and the World Technology Network; and a Fellow
of the Institute of Materials, the Institute of Physics, the Chartered Management Institute and the Royal Academy of Engineering.

36

You might also like