Designing VNbMoTa

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Scripta Materialia 191 (2021) 131–136

Contents lists available at ScienceDirect

Scripta Materialia
journal homepage: www.elsevier.com/locate/scriptamat

Designing Vx NbMoTa refractory high-entropy alloys with improved


properties for high-temperature applications
M. Wang a,b, Z.L. Ma a,b,∗, Z.Q. Xu a,b, X.W. Cheng a,b,∗
a
School of Materials Science and Engineering, Beijing Institute of Technology, Beijing 100081, China
b
National Key Laboratory of Science and Technology on Materials Under Shock and Impact, Beijing 100081, China

a r t i c l e i n f o a b s t r a c t

Article history: Poor high/medium-temperature phase stability and/or low room-temperature ductility are currently bot-
Received 16 July 2020 tlenecks of refractory high-entropy alloys (RHEAs) that restrict their high-temperature applications. Here,
Revised 14 September 2020
we explored novel Vx NbMoTa RHEAs with vanadium concentrations of 0–25 at.%. Vx NbMoTa RHEAs ex-
Accepted 14 September 2020
hibit a single body-centered cubic (BCC) structure with unprecedented phase stability at a wide temper-
ature range from solidus down to 350°C. The grain structure of Vx NbMoTa can be substantially refined
Keywords: when increasing the V concentration since higher V contents induce stronger solutal effects which con-
Refractory high-entropy alloys tribute to higher growth restriction factors. Equimolar VNbMoTa exhibits the yield strength of 811 MPa
Grain Refinement at 10 0 0°C that is superior to most RHEAs reported by far, and this alloy also shows excellent room-
Phase stability
temperature ductility with the fracture strain > 25% and no strain-softening at high temperature, which
Mechanical properties
is rarely seen in many popular RHEAs. These exceptional performances of VNbMoTa enable it to be a very
promising material for high-temperature applications.
© 2020 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

For a long time, conventional alloy designs are constrained to this bottleneck is to design basket-weave nano-scale microstruc-
the model of combining a/few principal component(s) with sev- tures analogous to Nickel-based superalloys. As recently reviewed
eral minor elements added to tailor properties. Since this strat- by Miracle et al. [15], this type of RHEA is named as refractory
egy only explores limited compositional space, the development high entropy superalloys (RSAs) which exhibit disordered A2 + or-
and application of alloys are largely restricted. Different from con- dered B2 basket-weave nano-structure and can show compressive
ventional alloys, high-entropy alloys (HEAs) that consist of at least yield strengths up to 935 MPa at 10 0 0°C and high RT compres-
four principal elements explore a wide composition space in the sive ductility (>50%). However, some RSAs suffer from the dis-
interior regions of hyper-dimensional multicomponent phase di- solution of B2 phases during heat treatment above 600°C, which
agrams and, therefore, open up a new dimension for the explo- degrades their high-temperature strength [15]. Another approach
ration of new materials and new properties [1,2]. Inspired by this to maintain both RT ductility and high-temperature strength is
novel alloy design strategy, refractory high-entropy alloys (RHEAs) to design single-phase disordered solid solution RHEAs and adjust
were also fabricated by employing refractory elements from Group their properties via alloying. This is focused in this study. Single-
IV, V, and VI, and sometimes with non-refractory additions such phase RHEAs are usually more ductile than multi-phase RHEAs
as Ni, Co, Al, or Si to tune the mechanical properties [3–6]. Since and can exhibit good high-temperature phase stability [16], and
alloy softening generally occurs at temperatures higher than 0.5- therefore high strength. However, it is still challenging to simul-
0.6Tm , RHEAs featuring high melting points can exhibit higher taneously achieve good RT ductility and high strength and good
strength at elevated temperatures (> 800°C) compared with widely phase stability at high temperatures. For example, WNbMoTa al-
used Nickel-based superalloys such as Inconel 718 and Haynes 230 loy which is one of the most classical RHEAs initially developed
[7–10], which enables RHEAs to be very promising alternatives by Senkov et al. exhibits a single-phase structure and the ex-
for high-temperature applications. Nevertheless, most RHEAs ex- ceptionally high yield strength (405 MPa) at 1600°C, but it only
hibit lower ductility at room temperature (RT) and/or poor phase has 2.1% fracture strain at RT [7]. Until now, only TiZrHf alloyed
stability at high temperatures [7,11–14]. One approach to tackle with Ta, Nb, and Nb-Ta reveal single-phase structures at RT and
tensile fracture strains ~15–30% that are comparable to Ni-based
superalloys, but these alloys are relatively weak above 10 0 0 °C

Corresponding authors. and show poor phase stabilities at high/medium temperature
E-mail addresses: z.l.ma@bit.edu.cn (Z.L. Ma), chengxw@bit.edu.cn (X.W. Cheng). [16,17].

https://doi.org/10.1016/j.scriptamat.2020.09.027
1359-6462/© 2020 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
132 M. Wang, Z.L. Ma, Z.Q. Xu et al. / Scripta Materialia 191 (2021) 131–136

Fig. 1. (a) Pseudo-binary phase diagram of V-NbMoTa [24]. (b) XRD patterns. (c) Microstructure and (d) EDS mappings of V1.0 RHEA. (e) Segregation ratios K. (f) TEM-SAED
patterns of Vx NbMoTa (x = 0.25, 0.5, 0.75, and 1.0) RHEAs.

Based on the classical WNbMoTa, this study will further adjust Table 1
Lattice constants of Vx NbMoTa measured by XRD
the composition to explore novel single-phase RHEAs with good
and TEM.
RT ductility and high strength and good phase stability at high
temperatures. It is suggested that Mo and W contribute to the im- Alloys V0.25 V0.5 V0.75 V1.0
proved high-temperature performance of WNbMoTa RHEA but are aXRD (Å) 3.241 3.227 3.214 3.204
also responsible for the room temperature brittleness [13, 18–21]. aTEM (Å) 3.266 3.244 3.228 3.194
Yao et al. [22] replaced W in WNbMoTa with V and found that
VNbMoTa RHEA can exhibit the yield strength of 1.5 GPa and the
compressive fracture strain of 21% at RT, which are both higher
than those of WNbMoTa (σ y = 1.1 GPa, ε f = 2.1%). Binglun Yin ray spectroscopy (EDS), and an Oxford HKL Nordlys II Electron
et al. [23] proposed that V can be the prime element for strength- Backscattered Diffraction (EBSD). TEM samples were prepared by
ening HEAs, and HEAs containing V have higher strengths com- mechanical milling and ion thinning and tested with a transmis-
pared with their counterparts without V. Besides, the pseudo- sion electron microscope (TEM, FEI Tecnai G2 F20). RT and high-
binary phase diagram of V-NbMoTa (calculated in the current temperature (800°C and 1000°C) mechanical properties of V1.0 al-
study with the Thermo-Calc and the TCHEA1 database [24]) in loy were measured at a constant strain rate of 10−3 s−1 using
Fig. 1(a) indicates that alloys containing 0–30 at.% V have a wide an Instron 5569 universal testing machine and a dynamic ther-
temperature range (from 2300 to 500°C) exhibiting only a single mal simulation testing machine (Gleeble-3500) respectively. CAL-
BCC phase, which implies that alloys in this composition range culation of PHAse Diagrams (CALPHAD) was performed using the
would have good phase stabilities (i.e. only a single phase) at Thermo-Calc and the TCHEA1 database to model the required ther-
high temperatures. Therefore, based on these benefits of vanadium, modynamic properties of the alloy.
Vx NbMoTa RHEAs with different V concentrations are fabricated Typical XRD patterns in Fig. 1(b) show that all as-cast
and explored in this study. Influences of the V contents on mi- Vx NbMoTa RHEAs exhibit the BCC crystal structure. But the XRD
crostructures and phase stabilities of Vx NbMoTa are studied, and pattern moves slightly to higher 2θ when increasing the V con-
mechanical properties at different temperatures are compared with centration, as typically shown by the enlarged (110) peaks at the
RHEAs reported by far. upper right corner of Fig. 1(b). This peak movement indicates the
Vx NbMoTa (x = 0.25, 0.5, 0.75, 1.0) RHEAs, hereinafter referred decreased lattice constant (Table 1) as estimated by the Bragg’s
to as V0.25, V0.5, V0.75, and V1.0, were prepared with high pu- Law (2dsinθ = λ), which should essentially attribute to the smaller
rity (>99.9 at. %) V, Nb, Mo, and Ta using vacuum arc melting in atomic radius of V (rv =1. 346 Å) compared with the other three
water-cooled copper crucible under argon atmosphere. Alloys were elements (rTa = 1.467 Å, rNb = 1.468 Å, rMo = 1.4 Å) [25], which
remelted at least eight times to ensure the compositional homo- leads to lattice shrinkage.
geneity. Crystal structures were examined with a Rigaku Ultima III Vx NbMoTa RHEAs exhibit dendritic microstructures as typically
X-ray diffractometer using Cu-Kα radiation and the scanning rate shown by V1.0 in Fig. 1(c). Fig. 1(d) shows the EDS map of the red
of 5°/min from 20° to 100°. Microstructures of as-cast alloys were rectangle area is in Fig. 1(c). Compositions of the dendrite and in-
studied using a scanning electron microscope (SEM, Hitachi S4800) terdendrite regions of Vx NbMoTa RHEAs are specified in Table 2.
equipped with an Oxford Instruments INCA Energy-dispersive X- The segregation ratio of the element in dendritic and interden-
M. Wang, Z.L. Ma, Z.Q. Xu et al. / Scripta Materialia 191 (2021) 131–136 133

Table 2
EDS measured compositions in dendritic and interdendritic regions in Vx NbMoTa RHEAs. Mean values and standard deviations are given
based on at least 5 measurements for each case. Segregation ratios (K) are calculated using mean values.

Alloys V Nb Mo Ta

V0.25 Nominal composition (at. %) 7.69 30.77 30.77 30.77


Dendrite (at. %) 5.32 ± 0.79 32.62 ± 1.19 30.22 ± 1.40 31.84 ± 1.60
Interdendrite (at. %) 12.99 ± 1.53 37.35 ± 1.23 26.62 ± 0.98 23.04 ± 1.91
K = cdr /cid 0.41 0.87 1.14 1.38
V0.5 Nominal composition (at. %) 14.29 28.57 28.57 28.57
Dendrite (at. %) 10.12 ± 1.37 29.70 ± 1.05 27.31 ± 2.13 32.87 ± 2.71
Interdendrite (at. %) 20.14 ± 1.96 30.54 ± 2.24 25.13 ± 1.57 24.19 ± 1.47
K = cdr /cid 0.50 0.97 1.09 1.36
V0.75 Nominal composition (at. %) 20 26.67 26.67 26.67
Dendrite (at. %) 14.45 ± 1.71 27.84 ± 0.54 26.51 ± 1.75 31.20 ± 1.58
Interdendrite (at. %) 30.28 ± 2.28 28.59 ± 1.73 21.16 ± 2.19 19.97 ± 1.13
K = cdr /cid 0.48 0.97 1.25 1.56
V1.0 Nominal composition (at. %) 25 25 25 25
Dendrite (at. %) 18.53 ± 1.92 25.83 ± 0.44 25.24 ± 0.85 30.40 ± 1.67
Interdendrite (at. %) 36.51 ± 3.05 27.18 ± 1.43 17.39 ± 1.96 18.92 ± 2.02
K = cdr /cid 0.51 0.95 1.45 1.61

dritic regions is given by K = cdr /cid . K > 1 or K < 1 indicates to develop in the liquid ahead of growing crystals in which fur-
that the element is segregated into the dendritic or interdendritic ther nucleation events can be triggered [32–33]. A measure of how
region, and the more the K value deviates from 1, the more se- rapidly constitutional undercooling develops in the early stages of
vere the segregation of the element. As shown in Fig. 1(e), it is growth is the growth restriction factor Q, which is defined as
clear that high melting-point elements Ta and Mo are prone to    
segregate in the dendritic region, and the lower melting-point ele- ∂ (TCS ) ∂T
Q= =− (1)
ment V is segregated into the interdendritic region. Nb is nearly ∂ fS f →0
∂ fS
f S →0
S
evenly distributed in the alloy. This phenomenon highlights the
influence of melting points on segregations and it is typical in where TCS is the constitutional undercooling and defined by
casted alloys. Segregations of these elements in the alloy should TCS = TL − TA . TL is the liquidus temperature and TA is the ac-
also attribute to their different mutual solid solubilities. The bi- tual temperature. fS is the mass fraction of solid. For Vx NbMoTa
nary phase diagrams [26] of these four elements indicate that i) RHEAs, TL , TA , and fS can be calculated with the Thermo-Calc
V cannot form complete solid solutions with Ta and Mo but it is and the TCHEA1 database [24]. Fig. 2(b) shows the TCS - fS
completely soluble in Nb above 300 K, ii) Nb forms complete solid curves of Vx NbMoTa RHEAs. The initial slope of each curve is ob-
solutions with the other three elements at room temperature, and tained by fitting 5-10 closely spaced sampling points in the range
iii) Ta and Mo show complete solid solubility. These are consistent 0 < fS < 0.01 using a parabolic function (TCS = a + b · fS + c · fS2 ),
with the nearly zero segregation ratio of Nb and mutual segrega- and the growth restriction factor Q equals the value b [34]. Fig. 2(c)
tions of Ta and Mo in the dendritic region, and also indicate that shows the result of the calculation using V0.25 as an example. The
the pseudo-binary V-NbMoTa can be comparable to the pseudo- calculated Q values of Vx NbMoTa RHEAs are shown in Fig. 2(d). It
binary V-MoTa. Therefore, it is anticipated that increasing the V can be seen that values of Q increase significantly as the V content
content would expand the mushy zone (e.g. Fig. 1(a)) and thus in- increases, which means that V can be considered to be the solute
duce increased segregations of Ta and Mo as evidenced by Fig. 1(e). in the pseudo-binary system V-NbMoTa that effectively limits the
The increased degree of element segregation with the expansion growth of grains. The solutal effect of V is also supported by the
of the mushy zone is also commonly found in other RHEAs (e.g. segregation of V shown in Table 2 and Fig. 1(e), i.e. V is rejected
WNbMoTa and WNbMoTaV [27]). Although dendritic and interden- into the liquid as the solid grows, corresponding to its higher con-
dritic regions show different compositions, they have the same tent in the interdendritic region. Schematics in Fig. 2(e) illustrate
crystal structure and approximate lattice constants as indicated by the mechanism of the grain refinement induced by V. In the al-
the XRD results. Fig. 1(f) shows the selected area electron diffrac- loy of higher V concentration, a shorter growth distance (Rg ) of
tion (SAED) patterns of Vx NbMoTa RHEAs which confirm that they the solid would generate a steeper concentration profile of V in
are all single-phase BCC structures. The lattice constants calculated the liquid ahead of the interface, which induces a higher constitu-
from SAED (Table 1) also show a decreasing trend as increasing the tional undercooling and therefore, triggers more nucleation events.
V content, which is consistent with the XRD results. Meanwhile, the distance (w∗ ) from the dendrite tip (i.e. the solid-
Fig. 2(a) summarizes the grain sizes of as-cast Vx NbMoTa liquid interface) to where the free growth of other nuclei could oc-
RHEAs. It is clear that as increasing the V content from V0.25 cur (i.e. reaching the critical undercooling Tfg ) is much shorter for
to V1.0, the grain size of the alloy continuously decreases from the alloy of higher V concentration, which suppresses the growth
~830 μm to ~250 μm, indicating that increasing the V content of the grain and contributes to the grain refinement.
can significantly refine grains. Similarly, Senkov et al. have also re- It was found that the grain size can be related to the solute
ported that adding V into WNbMoTa can reduce the grain size from content by a semi-empirical relationship d = a + b/Q [35], where
200 μm to 80 μm [27]. d is the grain size and a and b are constants. Using the measured
Two theoretical paradigms, i.e. nucleant catalyst and solutal ef- mean grain sizes and calculated Q, a and b can be calculated as -
fect, account for the grain refinement in alloys [28]. It is believed 100 and 24750 respectively. The relationship between Q and the V
that both nucleation and the effect of the solute on the growth of content c in mass fraction (as shown in Fig. 2(d)) can be described
dendrites and the constitutionally undercooled zone in front of the by a parabolic function Q = 15.77 + 277c + 850c2 if we extend the
liquid-solid interface are important for grain refinement [29,30]. data to higher and lower V concentrations, as shown in the Supple-
Solute restricts crystal growth by requiring an interface to cool as mentary Information. Therefore, combing these two equations, the
it grows [31]. It also causes a region of constitutional undercooling relationship between the V content and the grain size of Vx NbMoTa
134 M. Wang, Z.L. Ma, Z.Q. Xu et al. / Scripta Materialia 191 (2021) 131–136

Fig. 2. (a) Grain sizes and (b) TCS -fS curves of Vx NbMoTa (x = 0.25, 0.5, 0.75, 1.0) RHEAs. (c) Calculation of the growth restriction factor Q using V0.25 RHEA as an example.
(d) Q in Vx NbMoTa as a function of V contents. (e) Schematic diagrams show how constitutional undercooling (Tcs ) influences the nucleation and grain growth in alloys
of high and low V contents. The blue, green, and orange curves correspond to profiles of the V concentration, liquidus temperature (TL ), and actual temperatures (TA ) in the
liquid in front of the solid-liquid interface (i.e. the dendrite tip) respectively. The dashed circle indicates the envelope beyond which the constitutional undercooling exceeds
the critical undercooling Tfg . Therefore, the nucleants within the envelope (red particles) are inactivated and only those out of the envelope can be triggered.

RHEAs can be described by the Eq. 2: ements in RHEAs that are commonly reported [8] and also evi-
denced by segregations in dendrites as shown in Fig. 3(d)–(f), and
d = −100 + 24750/(15.77 + 277c + 850c ) 2
(2)
ii) the inaccuracy of the CALPHAD simulation for the incomplete
Maintaining the phase stability in a broad temperature range thermodynamic description (full assessment) of all binary and, es-
during a prolonged service time is crucial for applications of RHEAs pecially, ternary systems that essential for building higher-order
at high temperatures [16]. Some RHEAs have a single-phase struc- systems [42]. Microstructures of V1.0 alloy after heat treatments
ture at the as-cast condition but phase transformation can occur still exhibit dendrites (Fig. 3(d) and (e)) similar to Fig. 1(c), and
during heat treatment [5,13,36]. For instance, HfNbTaTiZr has a sin- Fig. 3(f) shows the segregation ratios of all elements are weakened
gle BCC structure at the as-cast condition and after heat treatments but not eliminated, which could all attribute to the sluggish diffu-
(10 0 0°C~140 0°C), but it decomposes into two phases (BCC matrix sion featured by this alloy.
and BCC Ta-Nb rich precipitates) or three phases (BCC matrix, BCC Fig. 4(a) compares yield strengths of 67 RHEAs reported by
Ta-Nb rich precipitates, and HCP Zr-Hf rich precipitates) after an- far [9,43,44]. V1.0 alloy in this study is represented by the red
nealing at 50 0–90 0°C [16,36–40]. After annealing above 1400°C, dot symbol. It can be seen that the yield strength of V1.0 at
this alloy was also reported to decompose into three phases (BCC) RT (1256 MPa) is not outstanding compared with other RHEAs.
matrix, FCC1 Zr-Hf-Ti rich precipitates, and FCC2 Zr-Hf-Ti rich pre- However, as the temperature increases, the yield stress of V1.0
cipitates) [41]. Fig. 1(a) shows that V-NbMoTa (0–30 at %) alloys exceeds those of most other RHEAs, indicating that V1.0 al-
can exhibit a single BCC phase at a wide temperature range (2300– loy has stronger resistance to high-temperature softening com-
500°C), which indicates that this alloy should possess good phase pared with other RHEAs. At 10 0 0°C, the yield strength of V1.0
stability. To test this anticipation, V1.0, which has the highest V (811 MPa) is only lower than four RHEAs (in the blue rectan-
contents and the highest predicted secondary-phase precipitation gle), i.e. MoNbTaVW [7], HfMoTaTiZr [13], HfMoNbTaZr [44], and
temperature (451°C as shown in Fig. 3 (a)) among the four stud- AlMo0.5 NbTa0.5 TiZr0.5 [45]. Nevertheless, Fig. 4(b) (at the upper
ied alloys was heat-treated at 1250°C and 350°C (i.e. ~100°C lower right corner of (a)) shows that V1.0 alloy has the RT fracture strain
than the predicted precipitation temperature). The heat-treatment >25% that is superior to other four RHEAs, especially MoNbTaVW
profiles are shown in Fig. 3 (b). For the 1250°C case, samples were and AlMo0.5 NbTa0.5 TiZr0.5 (fracture strains ~2%), indicating V1.0
held at 1250°C for 1 day and quenched to room temperature. For possesses much better RT ductility/workability. Fig. 4(c) further
the 350°C heat treatment, samples were first soaked at 1250°C for compares the stress-strain curves of the five alloys at 10 0 0°C.
3 days, then cooled (10°C/min) to 800°C and held for 7days, and fi- MoNbTaVW exhibits the highest yield stress and flow stress but
nally cooled to 350°C and held for 30 days. The previous two steps much lower ductility. AlMo0.5 NbTa0.5 TiZr0.5 , HfMoTaTiZr, and Hf-
at higher temperature are to promote atomic motion to facilitate MoNbTaZr show high yield strength and good ductility but ex-
the phase transformation (if there is any), since it would be ki- hibit significant strain-softening after yielding (or as increasing the
netically inert for the alloy to undergo phase transformation if it strain), which is obviously inferior compared with V1.0 that fea-
was directly held at 350°C. Fig. 3(c) shows XRD patterns of V1.0 tures the ultimate strength as high as ~1.4 GPa and no strain soft-
at different conditions, indicating that as-cast and heat-treated al- ening.
loys all exhibit the same single BCC phase structure without phase In summary, novel Vx NbMoTa RHEAs with varied V concen-
transformations, though the secondary phase is predicted to form trations were explored in this study. Vx NbMoTa alloys exhibit a
at 350°C (Fig. 3(a)). The absence of the phase transformation at single-phase BCC structure and possess unprecedented phase sta-
350°C is probably due to i) the low diffusivities of refractory el- bility in a wide temperature range from 350°C to the solidus
M. Wang, Z.L. Ma, Z.Q. Xu et al. / Scripta Materialia 191 (2021) 131–136 135

Fig. 3. (a) Solidification path of V1.0 RHEA. (b) Heat treatment profiles. (c) XRD patterns and (f) segregation ratios of V1.0 RHEA at different conditions. SEM images of V1.0
RHEA after heat treatments at (d) 1250°C and (e) 350°C.

Fig. 4. (a) Yield strengths of V1.0 RHEA and 67 RHEAs reported by far at different temperatures. Data are from references [9,43,44]. (b) Yield strengths at 10 0 0°C versus
fracture strains at RT of alloys in the blue rectangle in (a). Stress-strain curves (10 0 0°C and 10−3 s−1 ) of the five alloys are further compared in (c).

temperature. The grain structure of the alloy can be signif- Acknowledgments


icantly refined as increasing the V concentration due to the
strong solutal effect of V that contributes to higher growth We acknowledge National Natural Science Foundation of China
restriction factors at higher V contents. Compared with most (NSFC) grant number 51804032 and China Postdoctoral Science
RHEAs reported by far, equimolar VNbMoTa exhibit superior high- Foundation grant number 2018M640076 for financial supports.
temperature strength and excellent room-temperature ductility,
and it even shows no stain-softening (at high temperature up to Supplementary materials
10 0 0°C) that is rarely seen in other popular RHEAs, highlight-
ing that this alloy could be very promising for high-temperature Supplementary material associated with this article can be
applications. found, in the online version, at doi:10.1016/j.scriptamat.2020.09.
027.

References
Declaration of Competing Interest
[1] Y. Zhang, T.T. Zuo, Z. Tang, M.C. Gao, K.A. Dahmen, P.K. Liaw, Z.P. Lu, Prog.
The authors declare that they have no known competing finan- Mater. Sci. 61 (2014) 1–93.
[2] E.P. George, D.R.R. O, R.O. Ritchie, Nat. Rev. Mater. 4 (2019) 515–534.
cial interests or personal relationships that could have appeared to [3] H. Jiang, L. Jiang, Y.P. Lu, T.M. Wang, Z.Q. Cao, T.J. Li, Mater. Sci. Forum 816
influence the work reported in this paper. (2015) 324–329.
136 M. Wang, Z.L. Ma, Z.Q. Xu et al. / Scripta Materialia 191 (2021) 131–136

[4] M. Zhang, X. Zhou, J. Li, J. Mater. Eng. Perform. 26 (2017) 3657–3665. [24] Thermo-Calc, 2015.
[5] O.N. Senkov, S.V. Senkova, C. Woodward, Acta Mater. 68 (2014) 214–228. [25] W.B. Pearson, in: The crystal chemistry and physics of metals and alloys, Wiley
[6] N.N. Guo, L. Wang, L.S. Luo, X.Z. Li, R.R. Chen, Y.Q. Su, J.J. Guo, H.Z. Fu, J. Alloy. Interscience, New York, 1972, p. 151.
Compd 660 (2016) 197–203. [26] All phase diagrams: http://www.crct.polymtl.ca/fact/documentation/FS_All_
[7] O.N. Senkov, G.B. Wilks, J.M. Scott, D.B. Miracle, Intermetallics 19 (2011) PDs.htm
698–706. [27] O.N. Senkov, G.B. Wilks, D.B. Miracle, C.P. Chuang, P.K. Liaw, Intermetallics 18
[8] O.N. Senkov, D.B. Miracle, K.J. Chaput, J. Couzinie, J. Mater. Res. 33 (2018) (2010) 1758–1765.
3092–3128. [28] M. Esaton, D. StJohn, Metall. Mater. Trans. A 30A (1999) 1613–1623.
[9] J.P. Couzinié, O.N. Senkov, D.B. Miracle, G. Dirras, Data in Brief 21 (2018) [29] Z.L. Ma, J.W. Xian, S.A. Belyakov, C.M. Gourlay, Acta Mater. 150 (2018) 281–294.
1622–1641. [30] H. Shang, Z.L. Ma, S.A. Belyakov, C.M. Gourlay, J. Alloy. Compd 715 (2017)
[10] O.N. Senkov, S. Gorsse, D.B. Miracle, Acta Mater. 175 (2019) 394–405. 471–485.
[11] O.N. Senkov, C. Woodward, D.B. Miracle, JOM 66 (2014) 2030–2042. [31] I. Maxwell, A. Hellawell, Acta Mater. 23 (1975) 229–237.
[12] O.N. Senkov, C.F. Woodward, Mater. Sci. Eng. A 529 (2011) 311–320. [32] D.H. StJohn, M. Qian, M.A. Easton, P. Cao, Acta Mater. 59 (2011) 4907–4921.
[13] C. Juan, M. Tsai, C. Tsai, C. Lin, W. Wang, C. Yang, S. Chen, S. Lin, J. Yeh, Inter- [33] D. Shu, B. Sun, J. Mi, P.S. Grant, Acta Mater. 59 (2011) 2135–2144.
metallics 62 (2015) 76–83. [34] R. Schmid-Fetzer, A. Kozlov, Acta Mater. 59 (2011) 6133–6144.
[14] Z. Guo, A. Zhang, J. Han, J. Meng, J. Mater. Sci. 54 (2019) 5844–5851. [35] M. Easton, D. StJohn, Metall. Mater. Trans. A 36 (2005) 1911–1920.
[15] D.B. Miracle, M. Tsai, O.N. Senkov, V. Soni, R. Banerjee, Scr. Mater. 187 (2020) [36] O.N. Senkov, J.M. Scott, S.V. Senkova, F. Meisenkothen, D.B. Miracle, C.F. Wood-
445–452. ward, J. Mater. Sci. 47 (2012) 4062–4074.
[16] S.Y. Chen, Y. Tong, K.K. Tseng, J.W. Yeh, J.D. Poplawsky, J.G. Wen, M.C. Gao, [37] B. Schuh, B. Völker, J. Todt, N. Schell, L. Perrière, J. Li, J.P. Couzinié, A. Hohen-
G. Kim, W. Chen, Y. Ren, R. Feng, W.D. Li, P.K. Liaw, Scr. Mater. 158 (2019) warter, Acta Mater. 142 (2018) 201–212.
50–56. [38] O.N. Senkov, J.M. Scott, S.V. Senkova, D.B. Miracle, C.F. Woodward, J. Alloy.
[17] N.D. Stepanov, N.Y. Yurchenko, S.V. Zherebtsov, M.A. Tikhonovsky, G.A. Sal- Compd. 509 (2011) 6043–6048.
ishchev, Mater. Lett. 211 (2018) 87–90. [39] O.N. Senkov, S.L. Semiatin, J. Alloy. Compd. 649 (2015) 1110–1123.
[18] Z.D. Han, N. Chen, S.F. Zhao, L.W. Fan, G.N. Yang, Y. Shao, K.F. Yao, Intermetallics [40] O.N. Senkov, A.L. Pilchak, S.L. Semiatin, Metall. Mater. Trans. A 49 (2018)
84 (2017) 153–157. 2876–2892.
[19] H.W. Yao, J.W. Qiao, J.A. Hawk, H.F. Zhou, M.W. Chen, M.C. Gao, J. Alloy. Compd [41] C. Yang, K. Aoyagi, H. Bian, A. Chiba, Mater. Lett. 254 (2019) 46–49.
696 (2017) 1139–1150. [42] S. Gorsse, O. Senkov, Entropy 20 (2018) 899.
[20] Z.D. Han, H.W. Luan, X. Liu, N. Chen, X.Y. Li, Y. Shao, K.F. Yao, Mater. Sci. Eng. [43] W. Guo, B. Liu, Y. Liu, T. Li, A. Fu, Q. Fang, Y. Nie, J. Alloy. Compd 776 (2019)
A 712 (2018) 380–385. 428–436.
[21] H.W. Yao, J.W. Qiao, M.C. Gao, J.A. Hawk, S.G. Ma, H.F. Zhou, Y. Zhang, Mater. [44] K. Tseng, C. Juan, S. Tso, H. Chen, C. Tsai, J. Yeh, Entropy 21 (2019).
Sci. Eng. A 674 (2016) 203–211. [45] O.N. Senkov, J.K. Jensen, A.L. Pilchak, D.B. Miracle, H.L. Fraser, Mater. Design
[22] H. Yao, J. Qiao, M. Gao, J. Hawk, S. Ma, H. Zhou, Entropy 18 (2016) 189. 139 (2018) 498–511.
[23] B. Yin, F. Maresca, W.A. Curtin, Acta Mater. 188 (2020) 486–491.

You might also like