Sunteți pe pagina 1din 5

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/255745197

Glass formation in a high entropy alloy system


by design
Article in Intermetallics April 2012
DOI: 10.1016/j.intermet.2011.12.006

CITATIONS

READS

19

223

4 authors, including:
Andrew Cunliffe

I. A. Figueroa

Firth Rixson Metals

Universidad Nacional Autnoma de Mxico

4 PUBLICATIONS 24 CITATIONS

62 PUBLICATIONS 286 CITATIONS

SEE PROFILE

SEE PROFILE

I. Todd
The University of Sheffield
120 PUBLICATIONS 911 CITATIONS
SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

Available from: Andrew Cunliffe


Retrieved on: 10 September 2016

Intermetallics 23 (2012) 204e207

Contents lists available at SciVerse ScienceDirect

Intermetallics
journal homepage: www.elsevier.com/locate/intermet

Glass formation in a high entropy alloy system by design


Andrew Cunliffe a, *, John Plummer a, Ignacio Figueroa b, Iain Todd a
a
b

University of Shefeld, Department of Materials Science and Engineering, Sir Robert Hadeld Building, Mappin Street, Shefeld, South Yorkshire S1 3JD, UK
Instituto de Investigaciones en Materiales, UNAM, 04510 Mexico D.F., Mexico

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 3 October 2011
Received in revised form
8 December 2011
Accepted 10 December 2011
Available online 23 January 2012

The alloys (TiZrNbCu)1xNix x 0.125, 0.15, 0.2, 0.25 were produced by vacuum arc melting and
investigated as a potential high entropy alloy and glass forming system. The alloys x 0.125 and 0.15
were capable of forming either amorphous or high entropy alloy phases, increasing the Ni content
destabilised the high entropy alloy phase. This behaviour is explained by considering the enthalpy and
entropy of mixing of the alloys and decomposition products. The unique feature of the high entropy
alloys in this study was the high vacancy formation energies compared to other reported high entropy
alloys and this is believed to be responsible for their ability to form an amorphous phase and their
relative sensitivity to Ni content.
2011 Elsevier Ltd. All rights reserved.

Keywords:
B. Alloy design
B. Thermodynamic and thermochemical
properties
C. Casting
C. Rapid solidication processing
E. Phase stability, prediction

1. Introduction
Two metallic materials receiving extensive interest are bulk
metallic glasses (BMGs) [1] and the more recently reported high
entropy alloys [2,3] (HEAs). Amorphous or glassy alloys have been
extensively studied, resulting in a good understanding of the
essential considerations for obtaining glass formation, with models
to predict glass forming compositions based on the need for a high
density of atomic packing. Inoue proposed three parameters that
promote glass formation by encouraging dense random packing;
a large negative enthalpy of mixing between atom pairs, three or
more components and large differences in atomic radii between the
constituent species [4]. BMG compositions also tend to occur close
to eutectic compositions, where the melting point is suppressed
and so the temperature difference between it and the point at
which the glassy structure is frozen in (termed the glass transition
temperature, Tg) is at a minimum. Mutual frustration of several
possible competing eutectic phases also aids in shifting the nose of
the crystalline phase eld on a time-temperature-transformation
(TTT) diagram to the right, reducing the critical cooling rate for
glass formation [5].
HEAs were dened by Yeh et al. [2] as alloys comprising 5 or
more components in roughly equiatomic proportions between 5 at

* Corresponding author. Tel.: 44 0 114 2225941; fax: 44 0 114 2225943.


E-mail address: andrew.cunliffe@shefeld.ac.uk (A. Cunliffe).
0966-9795/$ e see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.intermet.2011.12.006

% and 35 at% and possessing the ability to form simple solid solution phases with face centred cubic (fcc) or body centred cubic (bcc)
crystal structures. This ability to form simple solid solutions rather
than the expected complex mixture of intermetallic phases (which
are typically observed on devitrication of BMGs) was rst
observed by Cantor et al. [6] in 2004. Yeh [2] later introduced the
phrase high entropy alloy, to describe these systems, with their
unique ability to precipitate solid solution phases from the melt in
preference to intermetallics, which would be expected according to
the Gibbs phase rule and the Hume-Rothery rules for alloying. This
phenomenon forms alloys which, as a result of heavy alloying solid
solution strengthening, have high strengths comparable to those of
BMGs in some instances [2,3,7,8] whilst exhibiting work hardenability [9] and good ductility up to 26% strain to failure [2,3,7e9].
These combinations of high strength and ductility are not realized
in conventional alloys where high yield strength and ductility are
often competing properties and, for these reasons, HEAs are
attracting interest as potential structural materials.
The ability of HEAs to form such simple microstructures is the
key to their properties; the large number of component elements
results in a solid solution, which has sufciently high congurational entropy to overcome the effects of the heat of mixing of the
alloy. The molar congurational entropy (DSconf) of the alloys is
related to the number of equivalent microstates (U), which may
result in an observable macrostate by Boltzmanns equation, which
states Sconf RlnU, where R is the universal gas constant. In the case
of HEAs U is the number of ways of arranging N component species

A. Cunliffe et al. / Intermetallics 23 (2012) 204e207


Table 1
Theoretical elastic moduli (calculated from method in Ref. [25]) and kinetic fragility
index, m (m 12(B/G0.67)) [17].
Alloy

Dr (%)

DHmix (kJ mol1)

G (GPa)

B (GPa)

A-HEA1
A-HEA2
A-HEA3
A-HEA4

12.0
12.15
12.36
12.5

35
40
49
55

43.4
43.9
45.0
46.2

124.5
125.6
127.8
130.0

42
42
42
42

onto the n non-equivalent lattice points of the crystal unit cell


[10,11]. Calculating Sconf in this way it can be seen that (1) the
entropy of the bcc lattice is highest for 5 component HEAs, (2) the
fcc lattice has the highest entropy for 6e10 component HEAs and
(3) for compositions above 10 components an hexagonal close
packed (hcp) lattice would have the greatest entropy. This may be
the reason why no hcp HEAs have been reported although electronic structure undoubtedly plays an important role in the lattice
type [12].
The high Sconf of the HEA solid solutions must be large enough to
overcome the enthalpy of mixing of the metals composing the alloy.
Many of the reported HEA compositions have positive heats of
mixing when calculated using the Miedema model [13] and the
regular melt model [14]. The compositions CoCrCuFeNiAl (fcc) [8]
and WNbMoTaV (bcc) [15] have heats of mixing 18.7 kJ mol1
and 13.9 kJ mol1 respectively and would normally be expected to
segregate, though at high temperature the congurational entropy
of each solid solution of 22.5 J K1mol1 and 19.1 J K1mol1
respectively would make a random solid solution the more stable
state. It may be that compositions with positive heats of mixing and
prone to segregation form HEAs most readily, as the increase in
entropy from a segregated alloy to a random solid solution is
greater than that for a polysubstituted intermetallic to a random
solid solution [16]. Thus if the formation of random solid solution
phases in HEAs depends wholly upon sufciently high entropy of
mixing then a zero or positive value of enthalpy of mixing must be
a requirement of the alloy system, and one that is incompatible
with glass formation. To date, no high entropy alloy system, consisting of 5 or more elements and forming a simple fcc or bcc solid
solution [2] with the ability to form an amorphous phase on rapid
cooling has been reported. Identication of an alloy system which
forms both an HEA solid solution and a glassy phase would allow
novel processing routes for high entropy alloys and create he
possibility of HEA reinforced amorphous alloy composites but
would also force a re-evaluation of the explanation of HEA formation. The compositions (TiZrNbCu)1xNix; x 0.125, 0.15, 0.20, 0.25
full the main requirements of a glass former discussed above, as
calculations using Miedemas model [2,14] give negative DHmix in
the range 35 to 55 kJ mol1, the size mismatch of the alloys

205

varies from 12.0 % to 12.5 % and the kinetic glass index, m, was
calculated using the method of Park et al. [17] to be the intermediate value of 42 (based on theoretical elastic moduli) see Table 1.
The compositions also all full the requirements of an HEA former,
save for the inferred requirement for a positive enthalpy of mixing.
In this work we investigate the extent to which this alloy system
may form an HEA solid solution or amorphous phase depending on
chemistry and cooling rate.
2. Experimental procedures
Four alloys in the (TiZrCuNb)1xNix system with x 0.125, 0.15,
0.20, 0.25, A-HEA1, A-HEA2, A-HEA3 and A-HEA4 respectively and
the reported HEA compositions CoCrNiFeTi [18] and CoCrNiFeAl
[19] were prepared from high purity elements by arc melting in the
presence of a Ti getter. The chamber was evacuated to a pressure
of 103 Pa and backlled with high purity argon. The ingots were
ipped and re-melted 4 times to ensure complete melting and then
suction cast into a water-cooled copper die of 6 mm diameter.
Ribbons with a thickness of approximately 20 microns of each alloy
were generated by melt spinning. The suction cast rods were
investigated by: (1) X-ray diffraction (XRD) using a Siemens D500
diffractometer with a Cuka source, and (2) scanning electron
microscopy (SEM) using an FEI Sirion, with energy dispersive
spectrometry (EDS) capability. The melt-spun ribbons were also
analysed by XRD, with a slow scan rate of 0.2 /min, and by differential scanning calorimetry (DSC) with a Perkin Elmer Diamond
DSC, using gold sample pans and a heating rate of 20 Kmin1.
3. Results and discussion
The A-HEA alloys solidied to form the dendritic microstructure
seen in Fig. 1a and EDS analysis found that the dendrites consisted
mainly of Nb e see Fig. 1b. XRD traces in Fig. 2 conrm that the cast
alloys A-HEA1 and A-HEA2 solidied to form two phases, the rst
corresponding to niobium and the second an unknown bcc HEA
phase with a 2.51 nm. The compositions A-HEA3 and A-HEA4
also precipitated niobium dendrites but the HEA peak is suppressed
in the XRD trace of A-HEA3 and absent in A-HEA4, being replaced
by Ni42(Ti,Zr)58 and Cu10Zr7 signals. This corresponds to increased
Ni content causing the enthalpy of mixing of the alloys to become
more negative. High concentrations of Ni push the equilibrium
between the solid solution and the decomposition products
towards formation of Ni-X intermetallic phases. Heats of mixing
calculated using the Miedema model [13] imply that at these
compositions Ni would react most favourably with Ti and Zr to form
intermetallic phases of the type Ni42(Ti,Zr)58, DHmix 64 kJ mol1.

Fig. 1. (a) Nb Ka EDS map A-HEA3 Nb (green) partitions to dendrites the remaining elements form the matrix phase, (b) back scattered electron image of A-HEA3 microstructure.
(For interpretation of the references to colour in this gure legend, the reader is referred to the web version of this article.)

206

A. Cunliffe et al. / Intermetallics 23 (2012) 204e207

Fig. 2. XRD traces show A-HEA1 & 2 solidify to form a dual phase microstructure
consisting Nb and another bcc HEA phase.

Should this occur, the alloy would then be depleted in Ti and Ni and
the remaining composition has insufcient congurational entropy
to form an HEA random solid solution phase, resulting in the Zr and
Cu reacting to form the intermetallic Cu10Zr7. This appears to be the
mechanism at work here. As a consequence of this, at some value of
x between 0.15 and 0.2 the enthalpy of formation of the intermetallic phase is such that the entropy of a random solid solution is
insufcient to suppress ordering. Yeh [2] states the ability of an HEA
to form a simple solid solution results from the high congurational
entropy of the system. The congurational entropy of the 5
component bcc solid solutions was calculated as described above,
with enthalpy of mixing of the alloys used to estimate the critical
temperature; Tc Hmix/Sconf, for the suppression of intermetallics,
(i.e. when the difference in free energy of the solid solution and the
intermetallic phase DGIM-SS 0). The Tc of each alloy is shown with
Hmix in Table 2. The melting point of each alloy was outside the
range of DTA measurements and therefore known to be greater
than 1723 K; the rule of mixtures (RoM) melting point is shown in
Table 2 for comparison. It is noteworthy that the Tc of A-HEA1,
which does not show signs of ordering in the XRD trace, is significantly lower than the RoM melting temperature of the alloy, whilst
those for A-HEA3 and A-HEA4, which both show a clear tendency
towards ordering have values of Tc greater than the predicted
melting temperature and would therefore be expected to solidify
directly to the ordered structure. This is indeed what was observed.
For A-HEA2 the melting and critical temperatures are very similar
but XRD data shows no sign of ordering. The method of calculation
of the enthalpy of mixing and melting points of the alloys is neglect
of vibrational entropy, resulting in ambiguity in the prediction for
this alloy, being as it is on the cusp of HEA formability. These alloys
are unable to form a high entropy phase when the critical
temperature for HEA formability is higher than the solidus
temperature of the alloy; this supports Yehs explanation of the
HEA phenomenon [2]. In a similar calculation Zhang [20] discussed
a HEA formability parameter, U 1000DSmix/DHmix where if U < 1
a HEA will be formed; 1000 K is approximately 0.6 Tm for the alloys
reported in [20] below which diffusion would be sluggish and the
HEA phase would be metastable, thus Zhangs model predicts

Fig. 3. XRD traces of the A-HEA alloys showing amorphous halo.

whether an alloy can remain in the HEA solid solution phase on


cooling. For the alloys reported here the melting temperature is not
known and so a suitable value of 0.6 Tm to substitute into Zhangs
equation is not available.
The CoCrFeNiTi and CoCrFeNiAl melt-spun ribbons did not form
an amorphous phase during solidication, and instead precipitated
crystals. However, XRD traces of the melt-spun A-HEA alloys reported here (Fig. 3) show amorphous traces, with a characteristic
broad peak between approximately 20 and 30 2q. The DSC data
shows the glass transition and crystallisation start temperatures of
the alloys, Tg and Tx respectively, also conrming glass formation.
These were taken as the onset values and are shown in Table 2. In
general both Tg and Tx increase with increasing Ni content. This is
due to the effect of increasing Ni content on the glass formation
parameters of Inoue [4] as a result of the increasing atomic mist
from 12.0 % to 12.5 % further lowering of the enthalpy of mixing of
the alloys and in A-HEA3 and A-HEA4 competition between intermetallic formation. DSC traces show two crystallisation peaks for AHEA1, A-HEA 2 and A-HEA3 and three peaks for A-HEA4. The
increasing area of the rst crystallisation peak with Ni content
indicates an increase in volume fraction of the intermetallic phase.
The devitricaton products in A-HEA1 and A-HEA2 are the bcc HEA
phase and Nb solid solution and these crystallisation events
correspond to peaks 1 and 2 respectively in the DSC traces of these
alloys in Fig. 4. As Ni content is increased in A-HEA3 and A-HEA4
there is an additional increase in peak 1 area, corresponding to an
increase in the volume fraction of the Ni-rich intermetallic phase.

Table 2
Transition temperatures of the alloys.
Alloy

Tg (K)  5 K

Tx (K)  5 K

DTx (K)

RoM Tm (K)

Tc (K)

A-HEA1
A-HEA2
A-HEA3
A-HEA4

648
651
662
678

721
717
731
743

73
66
69
65

2000
1992
1977
1961

1781
2130
2582
2880

Fig. 4. DSC traces of A-HEA alloys, showing Tg and Txs.

A. Cunliffe et al. / Intermetallics 23 (2012) 204e207

207

Table 3
Vacancy formation energies of various alloy types.
Alloy

Vitralloy1

A-HEA1

A-HEA2

A-HEA3

A-HEA4

CoCrFeNiTi

CoCrFeNiAl

NiTi

DHh/kJ mol1

283

281

280

274

266

225

212

133

The remaining material then crystallises to form a mixture of Nb


solid solution and the Cu10Zr7 phase, causing a broadening and
eventual splitting of the second peak. That devitrication of these
alloys results in the formation of Nb, Ni42Ti58 and Cu10Zr7 indicates
there is a cooling rate effect in the formation of the HEA phase; on
slow cooling from the melt the HEA phase is observed because the
nose of the HEA phase eld on a TTT type diagram occludes any
intermetallic phases but when crystallisation occurs from the
metastable amorphous phase, the alloy decomposes to form binary
intermetallics. Which suggests the HEA phase does not replace
intermetallics as the most stable phase but that it is the most readily
formed when cooling from the randomly distributed melt phase.
The vast majority of reported HEA compositions contain Fe, Ni
and Co which are all able to substitute onto the face centred positions of the L12 g0 phase when the two other common HEA
components Ti and Al are present [12]. Increasing the ratio of Al to
all other elemental constituents (Fe, Co, Ni and Ti) to greater than
a 3:2 ratio disrupts this arrangement, leading to decomposition of
the solid solution and the formation of binary intermetallics [21].
Similar reasoning would also account for the disruption of HEA
formability with increasing Ni in this series of alloys, as Ni promotes
the Ni-Ti/Zr intermetallics and the whole system then decomposes
to binary and ternary intermetallics. This strongly suggests that the
ability of some alloys to form apparent high entropy phases results
from their constituent metals abilities to substitute for one another
in ordered cubic structures. This would allow the formation of
glassy and high entropy phases in a single system. The observation
that there is a kinetic contribution to the formation of HEAs
suggests that they are not the most thermodynamically stable
phase of the system. Indeed, the FeCoCrNiAlCu [22] alloy is known
to precipitate intermetallic phases after heat treatment. If the
entropy of mixing were solely responsible for the formation of
a random solid solution then the microstructure would be more
stable at elevated temperatures and elements would be expected to
remain in solid solution.
Recent work by Pan [23] showed a link between vacancy
formation energy of an alloy, where they devised the gure of merit
DHhDHmixSconf and discovered that there was a strong trend
between the magnitude of this value and the critical diameter of an
amorphous alloy. It was suggested that this link is due to the
relationship between vacancy formation energy and crystallisation
temperature, where the formation of vacancies permits selfdiffusion, promoting crystallisation. The ability to form vacancies
also improves the structural stability of ordered compounds with
respect to chemistry. Vacancy formation energies of the alloys in
this study were calculated in Table 3 using the method in [23],
along with those of reported CoCrFeNiTi and CoCrFeNiAl HEAs,
intermetallic alloys and glass forming alloys. The low vacancy
formation energy of the intermetallic alloys allows the structure to
be maintained despite variations in stoichiometry. It is interesting
that the HEAs thus far reported have lower vacancy formation
energies, suggesting that they are also capable of accommodating
some variation in chemistry whilst still maintaining their structure.
This allows for the large number of degrees of freedom when the
Gibbs phase rule is applied to HEA compositions [6] while,
conversely, the A-HEAs in this study have high vacancy energies
similar to that of the good glass former Vitreloy 1 [24] The high
vacancy binding energy of the A-HEAs compliments the atomic size

variation, negative enthalpy of mixing and multiple components,


enabling them to form amorphous alloys on rapid cooling. These
properties separate them from many of the reported HEA compositions, which are unable to form glassy phases by rapid
solidication.
4. Conclusions
If HEA formability depends only on the entropy of mixing
overcoming the enthalpy of mixing to suppress precipitation of
intermetallics it would be difcult to produce a system capable of
forming both HEA and amorphous phases. Here the (TiZrNbCu)1xNix system has been shown to be capable of doing both,
depending upon cooling rate. This suggests that there is more
determining the formation of HEA phases than simply the entropy
of mixing. A possible mechanism, not at odds with the requirements of a good glass former, is suggested. However, regardless of
mechanism, an alloy system with the capacity to form both a high
entropy solid solution and an amorphous phase opens up the
possibility of novel processing routes for producing HEAs with
nanocrystalline microstructures and metallic glassy alloys toughened with HEA crystallites.
Acknowledgements
A Cunliffe and J Plummer would like to acknowledge the
nancial support of the EPSRC in funding their studentships.
References
[1] Greer AL, Ma E. MRS Bulletin 2007;32:611e5.
[2] Yeh JW, Chen YL, Lin SJ, Chen SK. Materials Science Forum 2007;560:1e9.
[3] Zhang Y, Zhou YJ, Hui XD, Wang ML, Chen GL. Science in China Series 2008;G
51:427e37.
[4] Inoue A. Acta Materialia 2000;48:279.
[5] Waniuk TA, Schroers J, Johnson WL. Applied Physics Letters 2001;78:1213e5.
[6] Cantor B, Chang ITH, Knight P, Vincent AJB. Materials Science and Engineering
A 2004;375e377:213e8.
[7] Wang FJ, Zhang Y, Chen GL. Journal of Alloys and Compounds 2009;478:
321e4.
[8] Tong CJ, Chen YL, Chen SK, Yeh JW, Shun TT, Tsau HC, et al. Metallurguical and
Materials Transactions A 2005;36:881e93.
[9] Varalakshmi S, Kamaraj M, Murty BS. Materials Science and Engineering A;
2008.
[10] Mandle F. Statistical physics. Wiley; 1998.
[11] Stoloff NS, Sikka VK. Physical metallurgy and processing of intermetallic
compounds. Springer; 1995.
[12] Paxton AT, Methfessel M, Pettifor DG. Proceedings of the Royal Society London
A; 1997:1493e513.
[13] Boer FR, Boom R, Mattens WCM, Miedema AR, Niessen AK. Cohesion in metals
transition metal alloys. North Holland Physics Publishing; 1988.
[14] Li A, Zhang X. Acta Metallurgica Sinica (English Letters) 2009;22:219e24.
[15] Senkov ON, Wilkes GB, Miracle DB, Chuang CP, Liaw PK. Intermetallics 2010;
18:1756e8.
[16] Stoloff NS. International Materials Reviews 1989;34:153e83.
[17] Park ES, Na JH, Kim DH. Applied Physics Letters 2007;91:031907.
[18] Zhang KB, Fu ZY, Zhang JY, Wang WM, Wang H, Wang YC, et al. Materials
Science and Engineering A 2009;508:214e9.
[19] Li C, Li JC, Zhao M, Jiang Q. Journal of Alloys and Compounds 2009;475:752e7.
[20] Zhang Y. Materials Science Forum 2010;654e656:1058e61.
[21] Li C, Zhao M, Li JC, Jiang Q. Journal of Applied Physics 2008;104.
[22] Wang YP, Li BS, Fu HZ. Advanced Engineering Materials 2009;11:641e4.
[23] Xuilin J, Pan Y. Materials Science and Engineering A 2008;485:154e9.
[24] Lofer JF. Intermetallics 2003;11:529e40.
[25] Plummer J, Figueroa IA, Hand RJ, Davies HA, Todd I. Journal of Non Crystalline
Solids 2009;355:335e9.

S-ar putea să vă placă și