Efectul Aluminiului
Efectul Aluminiului
Efectul Aluminiului
com
ScienceDirect
Acta Materialia 68 (2014) 214–228
www.elsevier.com/locate/actamat
Received 16 August 2013; received in revised form 5 November 2013; accepted 19 January 2014
Available online 23 February 2014
Abstract
The microstructure, phase composition and mechanical properties of the AlMo0.5NbTa0.5TiZr and Al0.4Hf0.6NbTaTiZr high-entropy
alloys are reported. The AlMo0.5NbTa0.5TiZr alloy consists of two body-centered cubic (bcc) phases with very close lattice parameters,
a1 = 326.8 pm and a2 = 332.4 pm. One phase was enriched with Mo, Nb and Ta and another phase was enriched with Al and Zr. The
phases formed nano-lamellae modulated structure inside equiaxed grains. The alloy had a density of q = 7.40 g cm3 and Vickers hard-
ness Hv = 5.8 GPa. Its yield strength was 2000 MPa at 298 K and 745 MPa at 1273 K. The Al0.4Hf0.6NbTaTiZr had a single-phase bcc
structure, with the lattice parameter a = 336.7 pm. This alloy had a density q = 9.05 g cm3, Vickers microhardness Hv = 4.9 GPa, and
its yield strength at 298 K and 1273 K was 1841 MPa and 298 MPa, respectively. The properties of these Al-containing alloys were com-
pared with the properties of the parent CrMo0.5NbTa0.5TiZr and HfNbTaTiZr alloys and the beneficial effects from the Al additions on
the microstructure and properties were outlined. A thermodynamic calculation of the solidification and equilibrium phase diagrams was
conducted for these alloys and the calculated results were compared with the experimental data.
Ó 2014 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Refractory alloys; Phase composition; Crystal structure; Microstructure; Mechanical properties
1. Introduction alloys, in an alloy composition space that has not been pre-
viously explored. While the HEA approach has produced
Multi-principal-element alloys, also known as some stable solid solution body-centered-cubic (bcc) and
high-entropy alloys (HEAs) because of their high entropy face-centered-cubic (fcc) alloys [1,4–9], recent studies have
of mixing of alloying elements, have recently come to the shown that intermetallic phases can form in HEAs. This
attention of the scientific community due to some interest- often is associated with alloying with elements with large
ing and unexpected microstructures and properties [1–3]. differences in atomic radius and large negative enthalpies
The metallurgical strategy is to stabilize the disordered of mixing [4,10,11].
phase relative to impinging ordered intermetallics by max- Several high-entropy refractory alloys with promising
imizing the configurational entropy. One appealing aspect combinations of room temperature and elevated tempera-
of this approach is that the reduction of the Gibbs free ture mechanical properties and oxidation resistance have
energy, by the entropy of formation, increases with an recently been reported. These are MoNbTaW, MoNb-
increase in temperature. Such an approach could be very TaVW [6,7], HfNbTaTiZr [8,9], CrMo0.5NbTa0.5TiZr
useful in developing new high-temperature structural [12,13] and CrxNbTiVyZr [14,15]. The high entropy of mix-
ing and similar atomic radii (e.g. 146 pm) of the alloying
⇑ Corresponding author. Tel.: +1 937 255 4064.
elements resulted in the formation of disordered bcc crystal
E-mail address: oleg.senkov@wpafb.af.mil (O.N. Senkov).
structures in the alloys without Cr. However, the alloys
http://dx.doi.org/10.1016/j.actamat.2014.01.029
1359-6454/Ó 2014 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
O.N. Senkov et al. / Acta Materialia 68 (2014) 214–228 215
with Cr, the atomic radius (rCr = 128 pm) of which is much high-purity argon. The Al0.4Hf0.6NbTaTiZr alloy was
smaller than the atomic radii of other elements, addition- HIPed at 1473 K and 207 MPa for 2 h and then annealed
ally contained a cubic Laves phase, resulting in a consider- at 1473 K for 24 h in continuously flowing high-purity
able decrease in ductility at temperatures below 800 °C argon. During HIP and annealing, the samples were cov-
[12,14,15]. ered with Ta foil to minimize oxidation. The cooling rate
In the present work, the compositions of two earlier after annealing in both cases was 10 K min1. The crystal
reported refractory alloys, HfNbTaTiZr and CrMo0.5- structure was identified with the use of an X-ray diffrac-
NbTa0.5TiZr, have been modified to produce the Al0.4Hf0.6- tometer, Cu Ka radiation and a 2H scattering range of
NbTaTiZr and AlMo0.5NbTa0.5TiZr alloys. Here we study 10–140°. The experimental error in the measurements of
the effect of alloying with Al on the microstructure, compo- the lattice parameters was ±0.5 pm.
sition and mechanical properties of these new refractory Alloy densities were measured with an AccuPyc 1330
HEAs. Aluminum forms a number of binary and ternary V1.03 helium pycnometer. Vickers microhardness was
intermetallic phases with bcc refractory elements. At the measured on polished cross-section surfaces using a 136°
same time, the atomic radius of Al (rAl = 143 pm) is very Vickers diamond pyramid under 500 g load applied for
similar to the atomic radii of the refractory elements 20 s. The microstructure was analyzed with a scanning elec-
(hri = 146 pm), excluding Cr (rCr = 128 pm), which may tron microscope (SEM) Quanta 600F (FEI, North America
affect the formation energy of the intermetallic phases in NanoPort, Hillsboro, Oregon, USA) equipped with back-
the HEAs. Furthermore, it has been well documented that scatter electron (BSE), energy-dispersive X-ray spectros-
additions of Al stabilize the bcc crystal structure in the copy (EDS) and electron backscatter diffraction detectors.
AlxCoCrCuFeNi [1] and AlxCoCrFeMnNi [16] HEAs The experimental error in the measurements of the chemi-
and gradually transform their crystal structure from fcc cal composition was ±0.3 at.%. The average grain/particle
to bcc. It is also expected that alloying with Al will consid- size and the volume fractions of the phases were deter-
erably reduce the density of the refractory HEAs. mined in accordance with ASTM E112 and ASTM E562
standards, using the image analysis software Fovea Pro
2. Experimental procedures 4.0 by Reindeer Graphics, Inc.
Compression tests of rectangular specimens with the
The AlMo0.5NbTa0.5TiZr and Al0.4Hf0.6NbTaTiZr dimensions of 4.7 mm 4.7 mm 7.7 mm were con-
HEAs were prepared by vacuum arc melting of nominal ducted at 298 K, 873 K, 1073 K, 1273 K and 1473 K in a
mixtures of the corresponding elements. Titanium, zirco- computer-controlled Instron (Instron, Norwood, MA)
nium and hafnium were in the form of 3.2 mm diameter mechanical testing machine outfitted with a Brew vacuum
slugs with purities of 99.98%, 99.95% and 99.9%, respec- furnace and silicon carbide dies. Prior to each test, the fur-
tively. Niobium and tantalum were in the form of 1.0 nace chamber was evacuated to 104 N m2. The test
and 2.0 mm wires, and their purities were 99.95% and specimen was then heated to the test temperature at a heat-
99.9%, respectively. Molybdenum was in the form of ing rate of 20 K min1, soaked at the test temperature for
1 mm thick sheet with a purity of 99.99%. Aluminum was 15 min under 5 N controlled load and then compressed to a
in the form of 50–100 mm3 buttons with a purity of 50% height reduction or to fracture, whichever happened
99.999%. Arc melting was conducted on a water-cooled first. A constant ramp speed that corresponded to an initial
copper plate. High-purity molten titanium was used as a strain rate of 103 s1 was used. Room temperature tests
getter for residual oxygen, nitrogen and hydrogen. To were conducted at the same loading conditions but in air.
achieve a homogeneous distribution of elements in the The deformation of all specimens was video-recorded and
alloys, each alloy was re-melted five times, was flipped image correlation software Vic-Gauge (Correlated Solu-
for each melt, and was in a liquid state for 5 min during tions, Inc.) was used to measure strains.
each melting event. The prepared specimens were 12 mm
high, 30 mm wide and 100 mm long and had shiny surfaces, 3. Results
indicating minimal oxidation during vacuum arc melting.
The actual alloy compositions, determined with inductively 3.1. Crystal structure, density and microhardness
coupled plasma-optical emission spectroscopy, are given in
Table 1. The AlMo0.5NbTa0.5TiZr alloy was hot isostati- X-ray diffraction patterns of the annealed cast alloys are
cally pressed (HIPed) at 1673 K and 207 MPa for 2 h and shown in Fig. 1. Two phases, both with the bcc crystal
then annealed at 1673 K for 24 h in continuously flowing structures, are identified in the AlMo0.5NbTa0.5TiZr alloy
Table 1
Chemical compositions (in at.%) of the alloys studied in this work.
Alloy Al Hf Mo Nb Ta Ti Zr
AlMo0.5NbTa0.5TiZr 20.4 – 10.5 22.4 10.1 17.8 18.8
Al0.4Hf0.6NbTaTiZr 7.9 12.8 – 23.0 16.8 18.9 20.6
216 O.N. Senkov et al. / Acta Materialia 68 (2014) 214–228
Fig. 1. X-ray diffraction patterns of the annealed cast alloys: (a) AlMo0.5NbTa0.5TiZr and (b) Al0.4Hf0.6NbTaTiZr.
2500 2500
1000 1000
500 500
0 0
0 20 40 60 0 20 40 60
Engineering Strain (%) Engineering Strain (%)
Fig. 2. Engineering stress–strain compression curves of annealed: (a) AlMo0.5NbTa0.5TiZr and (b) Al0.4Hf0.6NbTaTiZr alloys tested at different
temperatures in air (T = 296 K) and vacuum (T = 1073–1473 K).
Table 4
Compression yield strength, r0.2, maximum strength, rp, elastic modulus, (a)
E, and fracture strain, d, of the AlMo0.5NbTa0.5TiZr alloy at different
temperatures.
T (K) 296 1073 1273 1473 1 2
r0.2 (MPa) 2000 1597 745 250
rp (MPa) 2368 1810 772 275
E (GPa) 178.6 80 36 27
d (%) 10 11 >50 >50 1
Table 5 (b)
Compression yield strength, r0.2, maximum strength, rp, elastic modulus,
E, and fracture strain, d, of the Al0.4Hf0.6NbTaTiZr alloy at different
4
temperatures.
T (K) 296 1073 1273 1473
r0.2 (MPa) 1841 796 298 89
rp (MPa) 2269 834 455 135 3
E (GPa) 78.1 48.8 23.3 4
d (%) 10 >50 >50 >50
2000
AlMo NbTa TiZr
CrMo NbTa TiZr
Al Hf NbTaTiZr Fig. 4. Backscatter electron images of the microstructure of the annealed
Yield Strength (MPa)
1600
HfNbTaTiZr AlMo0.5NbTa0.5TiZr alloy: (a) equiaxed grain structure with dark-color
second-phase particles precipitated at grain boundaries; (b) basket-like
1200 lamellar structure inside grains and at grain boundaries. Numbers and
respective arrows indicate typical regions used for chemical analysis (see
Table 6).
800
0
3.3.1. Annealed condition
296 1073 1273 1473 Representative SEM backscatter images of the AlMo0.5-
Temperature (K) NbTa0.5TiZr alloy after annealing at 1673 K for 24 h and
slow cooling are shown in Fig. 4. The alloy consists of a
Fig. 3. Comparison of the yield strength of the AlMo0.5NbTa0.5TiZr,
CrMo0.5NbTa0.5TiZr, Al0.4Hf0.6NbTaTiZr and HfNbTaTiZr alloys tested
polycrystalline matrix with average grain size of
at T = 296 K, 1073 K, 1273 K and 1473 K. 75 ± 5 lm and second-phase particles (darker material in
Fig. 4a) precipitated at grain boundaries. The volume frac-
tion of these second-phase particles is vf = 11.5 ± 1.5%.
difference decreases with an increase in temperature, and EDS analysis (Table 6) shows that the average composition
no effect from the Al addition is seen for this alloy pair of the matrix grains (identified as #1 in Fig. 4a and Table 6)
at T = 1273 K and 1473 K (see Fig. 3). is very close to the overall composition of the
218 O.N. Senkov et al. / Acta Materialia 68 (2014) 214–228
Table 6
Chemical compositions (in at.%) of the AlMo0.5NbTa0.5TiZr alloy constituents (shown in Fig. 4) in the annealed condition.
ID # Constituent Al Mo Nb Ta Ti Zr
1 Matrix grains 19.9 11.3 22.9 10.6 18.5 16.8
2 Dark large particles (at grain boundaries) 39.5 0.0 10.9 2.3 7.2 40.1
3 Bright nano-lamellar phase 15.4 14.5 25.6 12.8 19.2 12.5
4 Dark nano-lamellar phase 27.6 6.8 19.1 7.2 15.2 24.1
AlMo0.5NbTa0.5TiZr alloy (compare Tables 6 and 1), while of local plastic flow, which are inclined by 90–60° to
the dark particles (identified as #2) are enriched with Al the compression direction, and dark second-phase particles
and Zr. Higher magnification images reveal the presence remain present at grain boundaries (Fig. 6a). The volume
of very fine, basket-like lamellar structure inside the grains fraction of these particles is vf = 9.0 ± 1.0%, i.e. slightly
(Fig. 4b). The average lamellar spacing inside the grains is smaller than before the deformation. A characteristic relief
estimated to be 70 ± 5 nm. The lamellae coarsen at grain forms inside the grains (Fig. 6b). This relief is likely a result
boundaries and reveal the presence of two phases with dis- of interaction of the local material flow with the interface
tinct Z contrasts. Namely, the BSE contrast of one phase is boundaries between two nano-lamellar phases. After the
bright, which is an indication that this phase contains lar- deformation, the lamellar structure inside the matrix grains
ger amounts of heavier elements, and another phase is and at grain boundaries becomes coarser (Fig. 6b and c).
dark, which is an indication that this phase mainly consists Additionally, submicron-sized regions (spots), in which
of lighter elements. The EDS analysis shows that the lamel- the nano-lamellar structure disappears, form inside many
lae of the bright phase (similar to those identified as #3 in grains (Fig. 6d). The chemical composition of the alloy
Fig. 4b) are enriched with Mo, Nb and Ta (relative to the constituents is given in Table 7. After deformation, the
average alloy composition), while the lamellae of the dark average composition of grains (#1 in Fig. 6a and Table 7)
phase (identified as #4) are enriched with Al and Zr does not change and only minor changes in the composi-
(Table 6). These EDS data should, however, be interpreted tion of the dark-color large second-phase particles (#2 in
as an estimate of the differences in composition of these Fig. 6a and Table 7) are detected. At the same time, the
two phases, because X-ray emission volume, at the electron coarsened nano-lamellae of the bright phase (#3 in
beam size of 3–4 nm and accelerating voltage of 15 kV, is Fig. 6b and 6c and in Table 7) become more enriched with
estimated to be above 100 nm in diameter in this alloy Mo and Nb and slightly depleted of Al and Zr (in compar-
[17,18], and the adjacent lamellae of another phase contrib- ison with the annealed state). Coarsened nano-lamellae of
ute to the intensity of the collected EDS peaks. Therefore, the dark phase (#4 in Fig. 6 and Table 7) are more enriched
the measured compositions of the lamellae of the dark and with Al and Zr and depleted of other elements (compare
bright phases are somewhat between the actual composi- Tables 6 and 7) and their composition becomes closer to
tions of these phases. Taking into account that the X-ray the composition of the dark-color large second-phase par-
diffraction shows the presence of only two phases, it is rea- ticles (#2) located at grain boundaries. This observation
sonable to suggest that the large dark particles at grain seems to support our earlier suggestion that the large dark
boundaries and nanometer-sized dark lamellae inside the particles at grain boundaries and nanometer-sized dark
grains are of the same phase and thus have the same com- lamellae inside the grains are the same phase and thus have
position. Because the two phases (bright and dark on the the same composition. The composition of newly formed
BSE images) have the same bcc crystal structure with very gray spots is very close to the average composition of
close lattice parameters (see Table 2) we were unable to grains and differs only by a slightly smaller concentration
identify which phase is bcc1 and which is bcc2. of Mo and a higher concentration of Ti.
After annealing at 1473 K for 24 h, the single-phase Fig. 7 shows the microstructure of the Al0.4Hf0.6NbTa-
Al0.4Hf0.6NbTaTiZr alloy has an equiaxed grain structure TiZr alloy after 50% compression deformation at
(Fig. 5a). The average grain size is 140 ± 10 lm. Slight T = 1273 K. Grains become elongated in the directions of
etching reveals the presence of finer subgrain structure plastic flow (Fig. 7a) and deformation bands crossing some
inside the grains (Fig. 5b). Black dots seen inside the grains grain boundaries can be observed (Fig. 7b). Characteristic
in Fig. 5b correspond to etched sub-grain boundaries. bands and spots with different contrasts are present inside
the deformed grains. These distinctive contrasts inside the
3.3.2. Microstructure after compression deformation at grains are likely caused by different electron channeling
1273 K conditions and indicate the presence of a subgrain structure
The microstructure of the AlMo0.5NbTa0.5TiZr alloy and internal stresses, which lead to slight misorientations
after 50% compression deformation at T = 1273 K is of the regions inside the deformed grains [19]. Fine recrys-
shown in Fig. 6. The alloy constituents present in the tallized grains, the average size of which is 1.3 lm, are
annealed condition are retained after the deformation. formed at grain boundaries of deformed grains (Fig. 7c
The matrix grains are slightly elongated in the directions and d). The chemical compositions of the recrystallized
O.N. Senkov et al. / Acta Materialia 68 (2014) 214–228 219
Fig. 5. (a) Equiaxed grain structure of the annealed Al0.4Hf0.6NbTaTiZr alloy; (b) sub-grain structure inside the grains revealed after slight etching.
2 4
1
3
(a) (b)
4
3
(c) (d)
Fig. 6. Backscatter electron images of the microstructure of the AlMo0.5NbTa0.5TiZr alloy after compression deformation at 1000 °C: (a) low-
magnification image showing deformed grains and dark-color second-phase particles at grain boundaries; (b) a junction of three grains with different fine
morphologies of the two phases inside the grains; (c,d) two-phase basket-like structure inside the grains. Numbers and respective arrows indicate typical
regions used for chemical analysis (see Table 7).
Table 7
Chemical compositions (in at.%) of the AlMo0.5NbTa0.5TiZr alloy constituents (shown in Fig. 6) after compression deformation at 1273 K.
ID # Constituent Al Mo Nb Ta Ti Zr
1 Matrix grains 20.8 11.6 21.9 8.7 18.2 18.8
2 Dark large particles (at grain boundaries) 38.9 0.5 11.6 2.5 7.5 39.0
3 Bright nano-lamellar phase 12.1 18.1 27.1 13.2 18.8 10.6
4 Dark nano-lamellar phase 32.6 4.5 15.3 4.9 12.7 30.0
5 Gray regions (inside grains, Fig. 6d) 21.7 7.2 23.2 7.8 21.3 18.9
220 O.N. Senkov et al. / Acta Materialia 68 (2014) 214–228
Fig. 7. Backscatter electron images of the microstructure of the Al0.4Hf0.6NbTaTiZr alloy after compression deformation at 1000 °C: (a and b) deformed
grains with (a) characteristic channeling contrast bands inside the grains due to crystal lattice distortions and (b) deformation bands crossing a grain
boundary; (c) fine recrystallized grains formed at grain boundaries; (d) a higher magnification image shows the presence of nano-precipitates at the
boundaries of the recrystallized grains.
and non-recrystallized grains are the same as the average above described changes in the compositions of the bright
composition of the alloy (Table 1). A higher magnification and dark particles suggest that the bright-color particles
image (Fig. 7d) shows the presence of nano-precipitates at are the bcc1 phase while the dark-color particles are the
the boundaries of the recrystallized grains. Their small size bcc2 phase. However, additional transmission electron
does not allow chemical analysis. microscopy studies are required to verify this suggestion.
Fig. 8 shows X-ray diffraction patterns of the AlMo0.5- After the compression deformation at 1273 K, the
NbTa0.5TiZr and Al0.4Hf0.6NbTaTiZr alloys after 50% Al0.4Hf0.6NbTaTiZr alloy retains the single-phase bcc
compression deformation at 1273 K. The diffraction peaks structure (Fig. 8b). The lattice parameter of the bcc phase
from the deformed AlMo0.5NbTa0.5TiZr alloy sample before and after deformation is almost the same in this
become sharper and the presence of two bcc phases is alloy (a = 336.7 pm and 337.2 pm, respectively).
clearly seen (Fig. 8a). The lattice parameter of the bcc1
phase slightly decreases from a1 = 326.8 pm in the 3.4. Thermodynamic analysis
annealed condition to a1 = 325.9 pm after deformation,
while the lattice parameter of the bcc2 phase does not The phase diagrams for the studied complex alloys are
change and is a2 = 332.2 pm. These results taken with the currently unavailable. Therefore, an attempt has been
Fig. 8. X-ray diffraction patterns of the (a) AlMo0.5NbTa0.5TiZr and (b) Al0.4Hf0.6NbTaTiZr alloys after 50% compression deformation at 1273 K.
O.N. Senkov et al. / Acta Materialia 68 (2014) 214–228 221
phase starts to form in this alloy at 1391 K. Solidification by 241 K, which increases DTL by 189 K.
Fig. 9. Simulated solidification curves for the (a) AlMo0.5NbTa0.5TiZr, (b) Al0.4Hf0.6NbTaTiZr, (c) CrMo0.5NbTa0.5TiZr and (d) HfNbTaTiZr alloys.
222 O.N. Senkov et al. / Acta Materialia 68 (2014) 214–228
3.4.2. Equilibrium phase diagrams the volume fractions of 0.63, 0.37 and 0.01, respectively,
The simulated equilibrium phase diagrams of the four and the CrMo0.5NbTa0.5TiZr alloy contains two phases,
alloys are given in Fig. 10 and characteristic equilibrium bcc1 and Laves, at the volume fractions of 0.8 and 0.2,
phase transformation temperatures and reaction equa- respectively. In Al0.4Hf0.6NbTaTiZr, the bcc1 phase is
tions are given in Table 9. Equilibrium solidification depleted by Nb and Ta, the bcc2 phase is enriched with
occurs in a much narrower temperature range than NE Nb and Ta and depleted by Al and Hf and the hcp phase
solidification and results in the formation of a disordered is enriched with Hf and Al. In the CrMo0.5NbTa0.5TiZr
bcc1 phase in these alloys. The single-phase bcc1 region is alloy, the bcc1 phase is slightly enriched with Nb and Ti
the widest in HfNbTaTiZr (DTbcc1 = 873 K) and the nar- and depleted by Cr and Zr, while the Laves phase is essen-
rowest in CrMo0.5NbTa0.5TiZr (DTbcc1 = 371 K). In tially a binary Cr2Zr phase. The composition of the bcc1
AlMo0.5NbTa0.5TiZr and Al0.4Hf0.6NbTaTiZr, this region phase in AlMo0.5NbTa0.5TiZr and HfNbTaTiZr corre-
is 660 K and 703 K, respectively. With decreasing temper- sponds to the composition of the respective alloy (see
ature below 1131 K and 1068 K, the bcc1 phase in the Table 10).
AlMo0.5NbTa0.5TiZr alloy is computed to partially trans- At T = 973 K, the AlMo0.5NbTa0.5TiZr alloy contains
form to Ti3Al- and Al2Zr3-based intermetallic phases, three phases, bcc1, Ti3Al and Al2Zr3, at the volume frac-
respectively (Fig. 10a). In the Al0.4Hf0.6NbTaTiZr alloy, tions of 0.64, 0.24 and 0.12, respectively. The bcc1 phase
the bcc1 phase partially transforms to disordered bcc2 is slightly enriched with Mo, Nb and Ta, the Ti3Al-based
(at T 6 1318 K) and hexagonal close packed (hcp, at is essentially ternary Ti2ZrAl phase and Al2Zr3 is the bin-
T 6 1290 K) phases (Fig. 10b). The Laves phase in the ary phase with 40% Al and 60% Zr. The Al0.4Hf0.6NbTa-
CrMo0.5NbTa0.5TiZr alloy is predicted to be present at TiZr alloy also contains three phases, bcc1, bcc2 and hcp,
T 6 1542 K (Fig. 10c). Finally, complete transformation at the volume fractions of 0.44, 0.42 and 0.13, respec-
of the bcc1 phase into disordered bcc2 and hcp phases tively. The bcc1 phase is enriched with Ti and Zr, the
occurs in the HfNbTaTiZr alloy in the temperature range bcc2 phase is enriched with Nb and Ta and the hcp phase
from 1258 K to 992 K, thus predicting two-phase struc- is enriched with Al and Hf. The CrMo0.5NbTa0.5TiZr
ture, bcc2 and hcp, at T 6 992 K. alloy has two phases, bcc1 and Laves, with the volume
The calculated volume fractions and compositions of the fractions of 0.71 and 0.29. The bcc1 phase is enriched
equilibrium phases in the studied alloys at T = 1273 K and with Mo, Nb and Ti, and the Laves phase is essentially
T = 973 K are given in Tables 10 and 11, respectively. At a binary Cr2Zr phase. The HfNbTaTiZr alloy consists
T = 1273 K, the AlMo0.5NbTa0.5TiZr and HfNbTaTiZr of 51% bcc2 and 49% hcp. The bcc2 phase is enriched
alloys are single phase bcc1 structures, the Al0.4Hf0.6NbTa- with Nb and Ta and the hcp phase is enriched with Hf,
TiZr alloy contains three phases, bcc1, bcc2 and hcp, with Ti and Zr (see Table 11).
Fig. 10. Simulated equilibrium phase diagrams for the (a) AlMo0.5NbTa0.5TiZr, (b) Al0.4Hf0.6NbTaTiZr, (c) CrMo0.5NbTa0.5TiZr and (d) HfNbTaTiZr
alloys.
O.N. Senkov et al. / Acta Materialia 68 (2014) 214–228 223
Table 9
Characteristic equilibrium phase transformation temperatures (in K) and respective reaction equations in the studied alloys. Simulated results.
Reaction equation AlMo0.5NbTa0.5TiZr Al0.4Hf0.6NbTaTiZr CrMo0.5NbTa0.5TiZr HfNbTaTiZr
L ! bcc1 (TL) 2046 2238 2877 2290
L ! bcc1 (TS) 1791 2021 1913 2131
bcc1 ! bcc2 – 1318 – 1258
bcc1 ! bcc2 + hcp – 1290 – 1092, 992
bcc1 ! Laves – – 1542 –
bcc1 ! Ti3Al 1131 – – –
bcc1 ! Ti3Al + Al2Zr3 1068 – – –
Table 10
Calculated volume fractions and compositions (in at.%) of equilibrium phases in the studied alloys at T = 1273 K.
T = 1273 K Fraction Al or Cr Hf or Mo Nb Ta Ti Zr
AlMo0.5NbTa0.5TiZr
bcc1 1.00 20 10 20 10 20 20
Al0.4Hf0.6NbTaTiZr
bcc1 0.63 9.4 15.1 15.4 13.9 21.8 24.3
bcc2 0.37 5.2 5.5 28.3 30.9 17.3 12.9
hcp 0.01 22.4 63.1 0.1 0.1 4.8 9.5
CrMo0.5NbTa0.5TiZr
bcc1 0.80 8.4 12.5 25.0 12.5 24.8 16.8
Laves 0.20 66.5 0.0 0.0 0.0 0.6 33.0
HfNbTaTiZr
bcc1 1 20.0 20.0 20.0 20.0 20.0
Table 11
Calculated volume fractions and compositions (in at.%) of equilibrium phases in the studied alloys at T = 973 K.
T = 973 K Fraction Al or Cr Hf or Mo Nb Ta Ti Zr
AlMo0.5NbTa0.5TiZr
bcc1 0.64 14.5 15.5 30.3 14.8 11.8 13.1
Ti3Al 0.24 25.1 0.3 2.3 2.1 51.1 19.2
Al2Zr3 0.12 40.0 60.0
Al0.4Hf0.6NbTaTiZr
bcc1 0.44 8.4 7.2 10.9 6.5 30.5 36.5
bcc2 0.42 2.9 0.7 35.8 40.4 14.1 6.2
hcp 0.13 22.9 64.3 0.0 0.0 3.6 9.2
CrMo0.5NbTa0.5TiZr
bcc1 0.71 1.2 14.0 28.1 14.0 28.1 14.6
Laves 0.29 66.6 0.0 0.0 0.0 0.1 33.4
HfNbTaTiZr
bcc2 0.51 – 3.3 38.3 38.4 15.5 4.5
hcp 0.49 – 37.6 0.8 0.6 24.7 36.3
without any evidence of phase transformation and the Cr2Ta, Cr2Ti and Cr2Zr. The AlZr intermetallic phase has
alloys with X = 0 and X = 0.4 have similar single phase the largest enthalpy of formation, DHmix = 43.7 kJ
bcc crystal structures. Moreover, the average grain size in mol1, followed by Al3Zr2 (42.6 kJ mol1) and Al2Zr3
these alloys is also the same [8]. Therefore, the Al-induced (41.4 kJ mol1). The large negative values mean strong
strengthening mechanism in the AlXHf1–XNbTaTiZr alloy bonding between the two elements, and heat is evolved
system is not related to the phase changes. One may when forming the compound. The enthalpies of formation
suggest that it can be caused by the formation of stronger of other Al–Zr intermetallics and those of Al–Hf are
interatomic bonds, as Al has strong bonding to each of the between 40 and 30 kJ mol1 and Al–Ti intermetallics
alloying elements (see below). The rapid decrease in the are between 30 and 20 kJ mol1 (see Table 13). From
difference between the strengths of the Al0.4Hf0.6NbTaTiZr Table 13, it is found that strong binary compounds form
and HfNbTaTiZr alloys with an increase in temperature in the Al–Zr and Al–Hf binary systems, followed by those
can then be explained by rapid weakening of the Al–TM in the Al–Ti system, then in the Al–Nb binary. In compar-
bonds. The first-principles calculations are, however, ison, Cr2Zr has the strongest bond (with an enthalpy of
required to verify the proposed scenario. formation of 11.3 kJ mol1) among all the binary Laves
At the same time, the AlMo0.5NbTa0.5TiZr alloy con- phases, but it is significantly weaker than the bonds in all
sists of a very fine nano-scale mixture of two (likely coher- the Al–Zr, Al–Hf, Al–Ti, Al–Ta and Al–Nb binary
ent) phases, bcc1 and bcc2, at near equal volume fractions, intermetallic phases. Because more negative enthalpy of
while the parent CrMo0.5NbTa0.5TiZr alloy consists of formation between two elements generally indicates a
three relatively coarse-grained phases, bcc1, bcc2 and higher tendency to form intermetallic phases, a number
Laves, at the volume fractions of 67%, 16% and 17%, of Al-containing intermetallics should be expected in the
respectively [12]. Therefore, the Al-induced strengthening Al-containing refractory HEAs. Surprisingly, the experi-
in this alloy system can be due to both stronger interatomic mental results did not show the formation of intermetallic
bonds and much more developed interface boundaries phases in the AlMo0.5NbTa0.5TiZr and Al0.4Hf0.6NbTa-
between the phases, which impede deformation flow. Due TiZr HEAs. The thermodynamic analysis also does not
to high thermal stability of the two-phase nanostructure, predict intermetallic phases in the Al0.4Hf0.6NbTaTiZr
the AlMo0.5NbTa0.5TiZr alloy retains its high strength at HEA at T P 900 K, in the AlMo0.5NbTa0.5TiZr HEA at
temperatures up to 1473 K, and the difference in the T P 1131 K and in the CrMo0.5NbTa0.5TiZr HEA at
strengths of this alloy and the parent Cr-containing alloy T P 1542 K (see Table 9). The simulation, however, calcu-
increases with an increase in temperature. lates two intermetallic phases, Ti3Al and Al2Zr3, in
Additional advantage from the replacement of Cr with AlMo0.5NbTa0.5TiZr at T < 1131 K and 1068 K, respec-
Al in the refractory HEAs is suppression or complete elim- tively, and a Laves Cr2Zr phase in CrMo0.5NbTa0.5TiZr.
ination of the formation of a brittle, topologically close- It is likely that although the very negative mixing enthal-
packed Laves phase. The formation of such a Laves phase pies favor the formation of intermetallic phases, other
is favored by the presencep offfiffiffiffiffiffiffi
twoffi types of atoms with the parameters of the HEAs, such as mixing entropy, atomic
atom size ratio close to 3=2 1:225 [26]. Among the size difference and electronegativity difference of the alloy-
refractory elements, Cr has the smallest atomic radius ing elements, may favor the formation of solid solution
and forms binary Laves phases with Nb, Ta, Ti and Zr phases.
[27] at the atomic radius ratios of 1.14, 1.14, 1.15 and Zhang et al. and Yang and Zhang [11,30] have recently
1.25, respectively (see Table 12). At the same time, the defined two parameters, dr and X, to predict the composi-
atomic radius of Al is very close to the atomic radii of other tion range of solid solution phase formation in HEAs. The
alloying elements. The small atomic size difference between atomic size difference parameter dr is calculated using the
the alloying elements has recently been shown to be one of following equation:
the necessary criteria favoring the formation of disordered qX ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2
solid solutions and discouraging the formation of interme- dr ¼ 100% ci ð1 ri =rÞ ð1Þ
tallic phases in HEAs [10,11,28,29]. On the other hand, Al
where
P ci is the atomic fraction of element i in the alloy,
has a different crystal structure and forms a number of bin-
r ¼ ci ri is the average atomic radius and ri is the atomic
ary intermetallic phases with refractory elements [27].
radius of element i. The parameter X takes into account the
Therefore, it is worth understanding why the refractory
combined
P effects of the mixing entropy, P DS mix ¼
HEAs containing Al do not form intermetallic phases.
R ci ln ci , mixing enthalpy, DH mix P ¼ 4x ij c i cj , and
Binary intermetallic phases, which are present in the bin-
effective melting temperature, T m ¼ ci T mi , of a HEA
ary alloy systems of the selected alloying elements, are
and is defined as:
listed in Table 13. Their enthalpies of mixing and the tem-
perature ranges of stability are also shown there. It is seen T m DS mix
from this table that Al can form a number of binary inter- X¼ ð2Þ
jDH mix j
metallic phases with Hf, Mo, Nb, Ta, Ti and Zr. Other bin-
ary intermetallic phases that can form between the selected Here xij is a concentration-dependent interaction parame-
alloying elements are five Laves phases, i.e. Mo2Zr, Cr2Nb, ter between elements i and j in a sub-regular solid solution
O.N. Senkov et al. / Acta Materialia 68 (2014) 214–228 225
Table 12
Metallic atomic radius, r, Pauling electronegativity, v, valence electron concentration, V, and melting temperature, Tm, of the elements in the studied alloys
[36].
Element Al Cr Hf Mo Nb Ta Ti Zr
r (pm) 143 128 159 139 146 146 147 160
v 1.61 1.66 1.3 2.16 1.60 1.50 1.54 1.33
V 3 6 4 6 5 5 4 4
Tm (K) 933.5 2180 2506 2896 2750 3290 1941 2128
Table 13
Binary intermetallic phases, their enthalpy of mixing and temperature range of stability in given binary systems.
System Phase DHmix [30] (kJ mol1) Temperature range [27] (K)
Al–Mo Al12Mo 1.45 <934
Al5Mo 2.83 <1008
Al4Mo 3.26 <1403
Al8Mo3 4.05 <1818
AlMo 5.10 1743–1993
AlMo3 3.83 <2423
Al–Nb Al3Nb 13.76 <1873
AlNb2 16.09 <2213
AlNb3 13.54 <2333
Al–Ta Al3Ta 14.51 <1824
Al2Ta3 19.20 <1867
AlTa 18.37 <2043
AlTa2 17.16 <2373
Al–Ti Al3Ti 22.11 <1623
Al2Ti 26.20 <1513
AlTi 29.50 <1753
AlTi3 22.18 <1453
Al–Zr Al3Zr 34.04 <1853
Al2Zr 39.82 <1918
Al3Zr2 42.57 <1868
AlZr 43.70 <1548
Al2Zr3 41.38 <1758
AlZr2 37.97 <1523
AlZr3 31.70 <1261
Al–Hf Al3Hf 29.89 <1863
Al2Hf 35.00 <1923
Al3Hf2 37.46 <1913
AlHf 38.50 <2073
Al2Hf3 36.50 <1863
AlHf2 33.52 <1393
Mo–Zr Mo2Zr 5.67 <2153
Cr–Nb Cr2Nb 6.65 <2043
Cr–Ta Cr2Ta 6.17 <2293
Cr–Ti Cr2Ti 6.47 <1643
Cr–Zr Cr2Zr 11.30 <1946
model [31] and Tmi is the melting temperature of element i. 2.7 kJ mol1 for HfNbTaTiZr and the minimum value
Analyzing a number of non-refractory HEAs, Yang and of 17.0 kJ mol1 for AlMo0.5NbTa0.5TiZr. The CrMo0.5-
Zhang have found that the formation of intermetallic NbTa0.5TiZr HEA, which contains 17% of the Laves
phases in a HEA during solidification is restricted if phase, has DHmix = 5.2 kJ mol1. It can be seen that
X > 1.1 and dr < 6.6% [30]. there is no direct correlation between the DHmix values
Using the atomic radii and melting temperatures of the and the formation of the intermetallic phase. The Al-
elements given in Table 12, as well as xij values from Ref. and Cr-containing alloys have the same values of
[31], the parameters DHmix, DSmix, X and dr have been DSmix = 14.5 J mol1 K1 and HfNbTaTiZr has
calculated for the four refractory HEAs, and their values DSmix = 13.4 J mol1 K1. Thus, there is no correlation
are given in Table 14. DHmix has the maximum value of between DSmix and the formation of intermetallic phases
226 O.N. Senkov et al. / Acta Materialia 68 (2014) 214–228
in these alloys either. At the same time, the combined 4.2. Comparison of thermodynamic calculations with
parameter X varies from X = 1.7 for AlMo0.5NbTa0.5TiZr experimental data
to X = 12.7 for HfNbTaTiZr, while dr is equal to 4.3% for
the Hf-containing alloys, 4.5% for AlMo0.5NbTa0.5TiZr 4.2.1. AlMo0.5NbTa0.5TiZr
and 7.2% for CrMo0.5NbTa0.5TiZr. The X criterion Although the thermodynamic analysis predicts two
(X > 1.1) predicts that these four refractory alloys should phases, bcc1 and Al2Zr3, after NE solidification, and three
form solid solution phases. On the other hand, the dr phases, bcc1, Ti3Al and Al2Zr3, in the equilibrium condi-
criterion (dr > 6.6%) confirms the formation of an inter- tion, the experimental results show the presence of two
metallic phase in the Cr-containing alloy. One can there- bcc phases in the AlMo0.5NbTa0.5TiZr alloy. There is no
fore conclude that the formation of intermetallic phases in experimental evidence for the presence of a hexagonal (Ti3-
the refractory HEAs seems to be more sensitive to the Al) and/or tetragonal (Al2Zr3) phases in this alloy at RT.
atomic size difference than to the values of the enthalpy Moreover, evolution of the microstructure during compres-
of mixing of the alloying elements. sion deformation at 1273 K suggests that the two bcc
Table 14 also shows the values of the electronegativity phases are also present at 1273 K, while the thermody-
difference parameter, dv, and the valence electron concen- namic analysis suggests the existence of a single bcc phase
tration (V) difference parameter, dv, for the studied HEAs. at T > 1131 K. At the same time, the binary Ti–Mo, Nb–Zr
The dv and dv are defined as: and Ta–Zr phase diagrams contain a wide phase separation
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
X range where the high-temperature bcc phase decomposes
2
dv ¼ 100% ci ð1 vi = vÞ ð3Þ into two bcc phases, one of which is rich with Ti and/or
qX ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Zr and another is rich with Mo, Nb, and/or Ta. For
2
dv ¼ 100% ci ð1 V i =V Þ ð4Þ example, in the Ta–Zr system, two bcc phases are present
Here vi and Vi are the Pauling electronegativity and va- at 1073 K 6 T 6 2053 K [33]. It is likely that similar phase
lencePelectron concentration, respectively, of element i, separation also occurs in the multicomponent
P AlMo0.5NbTa0.5TiZr alloy, with one bcc phase enriched
v ¼ ci vi and V ¼
ci V i are the average electronega-
tivity and valence electron concentration of the alloying with Mo, Nb and Ta and another bcc phase enriched with
elements. The vi and Vi values are given in Table 12. Zr. The formation of the low-temperature Ti3Al and Al2-
In accordance with Hume-Rothery rules for binary sub- Zr3 phases is probably restricted by slow diffusion of the
stitutional solid solutions [32], the solvent and solute alloying elements in HEAs [33].
should have the same valency and similar electronegativ-
ity, in addition to small atomic size difference. Therefore, 4.2.2. Al0.4Hf0.6NbTaTiZr
it is interesting to see if these rules also work for com- The thermodynamic analysis predicts a single bcc phase
plex HEAs. The results seem to show no direct correla- after NE solidification, and two bcc phases and one hexag-
tion between dv and the formation of intermetallic onal phase in equilibrium conditions at T 6 1318 K. The
phases. Both the AlMo0.5NbTa0.5TiZr and CrMo0.5- X-ray and microstructural analysis revealed the presence
NbTa0.5TiZr alloys have the same high value of of only one bcc phase in this alloy, both in annealed and
dv = 13.8%, in spite of the fact that the first alloy does in hot deformed conditions. The absence of the low tem-
not form an intermetallic phase. Two Hf-containing al- perature phases can be due to sluggish diffusion of the
loys have noticeably smaller dv = 7.8–8.1%. A similar alloying elements or to the limitation of the current PanTi
conclusion has been drawn earlier for non-refractory thermodynamic database.
HEAs [30]. We also did not find any reasonable correla-
tion between dV and the formation of intermetallic 4.2.3. CrMo0.5NbTa0.5TiZr
phases (see Table 14). The smallest dV value of 11.1% A eutectic-type reaction is predicted at the end of solid-
has the single bcc phase HfNbTaTiZr and the highest ification of this alloy. This result is in fairly good agreement
dV value of 20.9% has the two-phase AlMo0.5NbTa0.5- with the experimental observations of the microstructure
TiZr alloy. The three-phase CrMo0.5NbTa0.5TiZr alloy consisting of the large bcc1 particles (dendrites) embedded
has an intermediate value of dV = 17.0%. in a continuous network of the mixture of the bcc2 and
Table 14
The enthalpy of mixing, DHmix, entropy of mixing, DSmix, effective melting temperature, Tm, and parameter X, as well as atomic size difference, dr,
electronegativity difference, dv, and valence electron concentration difference between the alloying elements, in four HEAs.
Alloy DHmix (kJ mol1) DSmix (J mol1 K1) Tm (K) X dr (%) dv (%) dV (%) Phases
AlMo0.5NbTa0.5TiZr 17.0 14.5 1982 1.7 4.5 13.8 20.9 bcc1 + bcc2
Al0.4Hf0.6NbTaTiZr 6.5 14.5 2397 5.4 4.3 7.8 14.2 bcc
CrMo0.5NbTa0.5TiZr 5.2 14.5 2418 6.8 7.2 13.8 17.0 bcc1 + bcc2 + Laves
HfNbTaTiZr 2.7 13.4 2523 12.7 4.3 8.1 11.1 bcc
O.N. Senkov et al. / Acta Materialia 68 (2014) 214–228 227
Laves phases (inter-dendritic eutectic) [12]. At the same two disordered bcc phases, mainly in the form of
time, the equilibrium diagram predicts only one bcc1 phase spinodal-like nano-lamellar structure, and no intermetallic
and the formation of the Laves phase by a solid-state reac- phases, were present in the AlMo0.5NbTa0.5TiZr alloy.
tion from the bcc1 phase at T 6 1542 K, which is not sup- Partial substitution of Hf with Al in the HfNbTaTiZr
ported by the experiment [12,34]. alloy reduced the alloy density by 9% and increased RT
hardness and yield strength by 29% and 98%, respectively.
4.2.4. HfNbTaTiZr The difference in the yield strength of the HfNbTaTiZr and
Experimental results showed a single bcc phase in this Al0.4Hf0.6NbTaTiZr alloys, however, rapidly disappears
alloy in annealed condition [8]. However, precipitation of with an increase in temperature and the properties of these
fine particles of an unidentified second phase was noticed two alloys were the same at 1273 K and 1473 K. Both
in a narrow temperature range near 1073 K [9]. These alloys had a single-phase bcc structure with the average
experimental results seem to support the results of the ther- grain size of 140 lm.
modynamic analysis of this alloy. Solidification and phase equilibrium conditions of the
studied alloys were calculated using the available PanTie
The comparison of the simulated and experimental (Computherm, LLC) thermodynamic database. Although
results for the four studied alloys indicates satisfactory satisfactory agreements between the experimentally
agreement between the experimentally observed phases observed phases and phases predicted after NE solidifica-
and phases predicted after NE solidification. However, tion were observed, the calculated equilibrium phase dia-
noticeable disagreements of the calculated equilibrium grams of the three alloys, AlMo0.5NbTa0.5TiZr,
phase diagrams with the experimentally observed phase Al0.4Hf0.6NbTaTiZr and CrMo0.5NbTa0.5TiZr, noticeably
compositions of the three alloys, AlMo0.5NbTa0.5TiZr, disagreed with the experimentally observed phase composi-
Al0.4Hf0.6NbTaTiZr, and CrMo0.5NbTa0.5TiZr, is tions. It was concluded that the current PanTie database,
observed. One reason for this is that the alloys, even after which was developed for the Ti-rich alloys, cannot be
24 h annealing at 1673 or 1473 K, were still in NE condi- directly applied to the multi-principal-alloy compositions.
tions because of sluggish diffusion. Another explanation A thermodynamic database covering the full composition
is that the currently available thermodynamic database is range for the Al–Cr–Hf–Mo–Nb–Ta–Ti–Zr system needs
not sufficient to correctly predict phase compositions of to be developed to correctly predict phase equilibria and
these multi-principal-element alloys. Indeed, this database guide the design of refractory HEAs based on this system.
was developed by CompuTherm, LLC for Ti-rich alloys
via extrapolation of the interaction parameters from the Acknowledgements
lower order constituent binary and (some) ternary systems
to higher order interactions [35]. Since it is focused at the Valuable discussions with Drs. Jonathan Miller, Daniel
Ti-rich corner, many phases away from the Ti-rich corner Miracle, Jay Tiley and Fan Zhang are recognized. This
are not well modeled, or even not included in the database. work was supported through the Air Force Research Lab-
This database, therefore, needs further development in oratory Director’s fund and through the Air Force on-site
order to be used in the middle of the composition space Contract No. FA8650-10-D-5226 conducted by UES, Inc.,
for the design of HEAs. Dayton, Ohio.
[14] Senkov ON, Senkova SV, Woodward C, Miracle DB. Acta Mater [24] Huang PK, Yeh JW, Shun TT, Chen SK. Adv Eng Mater 2004;6:74.
2013;61:1545. [25] Li C, Li JC, Zhao M, Jiang Q. J Alloys Compd 2010;S504:S515.
[15] Senkov ON, Senkova SV, Miracle DB, Woodward CF. Mater Sci [26] Stein F, Plam M, Sauthoff G. Intermetallics 2004;12:713.
Eng A 2013;565:51. [27] Massalski TB, Okamoto H. Binary alloy phase diagrams. 2nd
[16] He JY, Liu WH, Wang H, Wu Y, Liu XJ, Nieh TG, et al. Acta Mater ed. Materials Park (OH): ASM International; 1990.
2013. http://dx.doi.org/10.1016/j.actamat.2013.09.037. [28] Zhang Y, Zhou YJ. Mater Sci Forum 2007;561–565:1337.
[17] Friel JJ. X-ray and image analysis in electron microscopy. Princeton [29] Guo S, Liu CT. Prog Nat Sci: Mater Int 2011;21:433.
Gamma-Tech.; 1995. [30] Yang X, Zhang Y. Mater Chem Phys 2012;132:233.
[18] Monte Carlo simulation of electron trajectory in solids, <http:// [31] Takeuchi A, Inoue A. Intermetallics 2010;18:1779.
www.gel.usherbrooke.ca/casino/>. [32] Hume-Rothery W. The structure of metals and alloys. London: Insti-
[19] Newbury DE, Joy DC, Echlin P, Fiori CE, Goldstein JI. Advanced tute of Metals; 1936.
scanning electron microscopy and X-ray microanalysis. New [33] Tsai KY, Tsai MH, Yeh JW. Acta Mater 2013;61:4887.
York: Plenum; 1986. [34] Senkov ON, Senkova SV, Dimiduk DM, Woodward C, Miracle DB.
[20] Cao W, Chen SL, Zhang F, Wu K, Yang Y, Chang YA, et al. J Mater Sci 2012;47:6522.
CALPHAD 2009;33:328. [35] Zhang C, Zhang F, Chen S, Cao W. J Met 2012;64(7):839.
[21] Zhang F, Xie FY, Chen SL, Chang YA, Furrer D, Venkatesh V. J [36] <http://en.wikipedia.org/wiki/Atomic_radii_of_the_elements_(data_
Mater Eng Perf 2005;14(6):717–21. page)>; <https://en.wikipedia.org/wiki/Electronegativity>.
[22] Scheil E. Z Metallkd 1942;34:70.
[23] Porter DA, Easterling KE. Phase transformations in metals and
alloys. 2nd ed. London: Chapman & Hall; 1992.