Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Materials and Design 139 (2018) 498–511

Contents lists available at ScienceDirect

Materials and Design

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

Compositional variation effects on the microstructure and properties of a


refractory high-entropy superalloy AlMo0.5NbTa0.5TiZr
O.N. Senkov a,⁎, J.K. Jensen b, A.L. Pilchak a, D.B. Miracle a, H.L. Fraser b
a
Air Force Research Laboratory, Materials and Manufacturing Directorate, Wright-Patterson AFB, OH 45433, United States
b
Center for the Accelerated Maturation of Materials, Department of Materials Science and Engineering, The Ohio State University, 1305 Kinnear Rd., Columbus, OH 43212, United States

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• A BCC-based analog to superalloy


microstructures is sought with a ductile
BCC matrix and coherent ordered B2
nano-precipitates.
• AlMo0.5NbTa0.5TiZr has a brittle B2 ma-
trix, coherent BCC nano-precipitates and
coarse HCP particles at grain boundaries.
• Sequential composition changes removed
BCC and HCP phases in some alloys but
the B2 phase remained continuous.
• Deformation inverted the microstructure
into a continuous BCC matrix and discrete
B2 precipitates.

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

Article history: An AlMo0.5NbTa0.5TiZr baseline alloy was shown earlier to have good high temperature strength but poor ductility
Received 14 September 2017 below 600 °C due to coarse intermetallic grain boundary particles and a continuous ordered B2 matrix phase.
Received in revised form 19 October 2017 Systematic composition changes intended to remove the deleterious microstructural features and to improve
Accepted 17 November 2017
mechanical properties were explored in the present work. The baseline alloy and the new alloys studied here,
Available online 21 November 2017
AlMo0.5NbTa0.5TiZr0.5, AlNbTa0.5TiZr0.5, Al0.5Mo0.5NbTa0.5TiZr and Al0.25NbTaTiZr, all had an ordered B2 matrix
Keywords:
crystal structure. Additionally, coherent BCC nanoscale precipitates were present at a high volume fraction inside
High entropy alloy the B2 matrix grains in AlMo0.5NbTa0.5TiZr, Al0.5Mo0.5NbTa0.5TiZr and Al0.25NbTaTiZr, and/or coarse, grain-
Superalloy boundary particles existed in AlMo0.5NbTa0.5TiZr and AlMo0.5NbTa0.5TiZr0.5. The mechanical properties were
Microstructure assessed with microhardness and compression testing at 25 °C and 1000 °C. Al0.5Mo0.5NbTa0.5TiZr showed the
Mechanical properties highest hardness (Hv = 6.4 GPa) and strength (σ0.2 = 2350 MPa) at 25 °C and modest strength (σ0.2 =
579 MPa) at 1000 °C. AlMo0.5NbTa0.5TiZr0.5 had the highest strength (σ0.2 = 935 MPa) at 1000 °C, but was brittle
at 25 °C. High-temperature deformation produced a desirable microstructure in Al0.5Mo0.5NbTa0.5TiZr and
Al0.25NbTaTiZr alloys consisting of a continuous BCC phase and discontinuous B2 nano-precipitates. The
relationships between the composition, microstructure, and properties were identified and discussed.
© 2017 Elsevier Ltd. All rights reserved.

1. Introduction

High entropy alloys (HEAs) are one of the most recent developments
⁎ Corresponding author. in material science, which open a vast, unexplored area of alloy compo-
E-mail address: oleg.senkov.ctr@us.af.mil (O.N. Senkov). sitions and the potential to influence solid solution phase stability

https://doi.org/10.1016/j.matdes.2017.11.033
0264-1275/© 2017 Elsevier Ltd. All rights reserved.
O.N. Senkov et al. / Materials and Design 139 (2018) 498–511 499

through control of configurational entropy [1–3]. Using the HEA con- comprised of the nano-scale interpenetrating phases, by adjusting the
cept, several refractory high entropy alloys (RHEAs) have recently composition of the base AlMo0.5NbTa0.5TiZr superalloy. Based on the
been reported as promising alternatives to Ni-based superalloys large number of intermetallic phases that form in the Al-Zr binary sys-
[4–11]. RHEAs with a low density and high strength are among the tem and the large electronegativity difference compared to other con-
most promising recent achievements in this field [10–17]. The reported stituent elements in the HEA composition, the effects of reducing Al
densities of these new RHEAs are in the range of 5.9 to 8.4 g/cm3 and and Zr on the HEA microstructure were first investigated and character-
their specific strengths are often superior to Ni-based superalloys ized. In addition, the effect of removing Mo on the alloy phase composi-
[11–14]. Some of these alloys consist of very fine, nano-scale mixtures tion, strength and ductility was also explored.
of disordered BCC and ordered B2 phases with similar lattice parameters
[11,18,19], which may be responsible for their high strengths. Among 2. Experimental procedures
these RHEAs, AlMo0.5NbTa0.5TiZr has impressive yield strength of
745 MPa and good compressive ductility (N50% height reduction) at HEA ingots with the compositions AlMo0.5NbTa0.5TiZr (base compo-
1000 °C. Detailed analysis of the microstructure and crystal structure sition, alloy #1), AlMo0.5NbTa0.5TiZr0.5 (reduction of Zr by ½ molar ratio,
of this alloy using advanced transmission electron microscopy (TEM) alloy #2), AlNbTa0.5TiZr0.5 (reduction of Zr by ½ molar ratio and removal
and atom probe tomography (APT) techniques revealed a nano-scale, of Mo, alloy #3), Al0.5Mo0.5NbTa0.5TiZr (reduction of Al by ½ molar ratio,
modulated mixture of an ordered B2 matrix and coherent, cuboidal pre- alloy #4), and Al0.25NbTaTiZr (reduction of Al by ¾ molar ratio and re-
cipitates of a disordered BCC phase [18,19]. Based on these microstruc- placement of Mo with Ta, alloy #5) were prepared by vacuum arc melt-
tural features, which are typical to superalloys, complex composition ing in an inert gas atmosphere. The samples were then hot isostatic
and the high strength at temperatures from 25 °C to 1200 °C, this pressed (HIP'd) for 2 h at 1400 °C and 207 MPa to remove any porosity
two-phase alloy was named a “refractory high entropy superalloy” [19]. and annealed in a flowing argon furnace at 1400 °C for 6 h followed by a
Despite showing exceptionally good high-temperature properties, furnace cool to room temperature with the initial cooling rate of 20 °C/
the AlMo0.5NbTa0.5TiZr superalloy has very limited ductility at temper- min.
atures ≤ 600 °C, which has been explained by the presence of brittle, A Bruker D2 Phaser X-ray diffractometer, at Cu Kα radiation, was
coarse intermetallic precipitates at grain boundaries [19,20]. It was used to identify the crystal structure of main phases in the alloys. The
found that these grain-boundary (GB) precipitates have a hexagonal microstructure was characterized using scanning electron microscopy
(P63/mcm) crystal structure and are predominantly composed of Al (SEM) and transmission electron microscopy (TEM) techniques. SEM
and Zr with an overall composition (in at.%) of ~Al37Mo2Nb8Ta2Ti9Zr42 was equipped with backscattered electron (BSE) and electron backscat-
[20]. Based on model γ/γ′ superalloys, an ‘ideal’ microstructure of re- ter diffraction (EBSD) detectors. BSE and EBSD image processing and
fractory superalloys would consist of two nanometer-scale, ordered/ analysis was performed using the Materials Image Processing and Auto-
disordered coherent phases, one of which is a disordered BCC matrix mated Reconstruction (MIPAR) software package [21].
and another is ordered B2 precipitates, while the deleterious hexagonal Samples for TEM analysis were extracted at site specific locations
intermetallic grain-boundary phase is completely removed. In some from the bulk sample using an FEI Helios NanoLab 600 Dual-Beam fo-
cases, heat treatment and/or hot working processing are used to remove cused ion beam (DB-FIB) instrument. The FIB lift-out and thinning pro-
intermetallic particles from grain boundaries, but annealing at cedures are described elsewhere [18]. The thinned FIB lamellae were
1000–1400 °C and water quenching of AlMo0.5NbTa0.5TiZr revealed characterized using conventional bright-field (BF) and dark-field (DF)
this phase to be stable over a large range of temperatures [20]. TEM imaging in an FEI CM200. Scanning transmission electron micros-
In this work, an attempt has been made to remove the coarse, GB in- copy (STEM) high-angle annular dark-field (HAADF) micrographs
termetallic phase, while maintaining the equiaxed grain microstructure were taken with an FEI Probe-Corrected Titan 80-300 STEM operating

(a) (b)
AlMo0.5NbTa0.5TiZr0.5 AlNbTa0.5TiZr0.5
Intesity (a.u.)

Intesity (a.u.)

BCC, a = 326.9 pm BCC, a = 330.6 pm

20 40 60 80 100 120 20 40 60 80 100 120


(Degrees)
(c) (d)
Al0.5Mo0.5NbTa0.5TiZr Al0.25NbTaTiZr
Intesity (a.u.)

Intesity (a.u.)

BCC, a = 328.3 pm BCC, a = 330.0 pm

20 40 60 80 100 120 20 40 60 80 100 120

Fig. 1. X-ray diffraction patterns of (a) AlMo0.5NbTa0.5TiZr0.5, (b) AlNbTa0.5TiZr0.5, (c) Al0.5Mo0.5NbTa0.5TiZr, and (d) Al0.25NbTaTiZr.
500 O.N. Senkov et al. / Materials and Design 139 (2018) 498–511

Table 1
Lattice parameters of the BCC phase, a, alloy density, ρ (measured vs. calculated), Vickers micro-hardness, Hv, and room temperature compression properties (Young's modulus, E, yield
stress, σ0.2, fracture stress, σf, and fracture strain, εf) of the studied RHEAs.

Alloy no. Alloy ID a ρ Hv E σ0.2 σf εf


pm g/cm3 GPa GPa MPa MPa %

1 AlMo0.5NbTa0.5TiZr 326.9/330.4 7.4/7.31 5.8 ± 0.1 122 2197 2370 4.1


2 AlMo0.5NbTa0.5TiZr0.5 326.9 7.4/7.32 6.2 ± 0.1 133 – 1320 1.0
3 AlNbTa0.5TiZr0.5 330.6 7.1/7.02 5.0 ± 0.1 124 1352 1357 1.3
4 Al0.5Mo0.5NbTa0.5TiZr 328.3 8.1/7.96 6.4 ± 0.1 132 2350 2460 3.2
5 Al0.25NbTaTiZr 330.0 9.4/9.27 4.9 ± 0.1 118 1745 1830 3.8

at 300 kV and a camera length of 115 mm. The convergence semiangle alloy) retained a single-phase BCC structure, but with an increased lat-
of the electron beam was 11.4 mrad and the inner and outer collection tice parameter a = 330.6 pm (Fig. 1b). According to the X-ray diffrac-
angles of the HAADF detector were 34.8 and 230.0 mrad, respectively. tion analysis, the reduced Al content alloy (Al0.5Mo0.5NbTa0.5TiZr) also
X-ray energy dispersive spectroscopy (XEDS) was performed using an contained a single BCC phase with a lattice parameter a = 323.3 pm
FEI Image-Corrected Titan 60-300 STEM equipped with a Super-X™ (Fig. 1c), which is the average value of the lattice parameters of the
XEDS detector system. The TEM was operated with an accelerating BCC and B2 phases in the base AlMo0.5NbTa0.5TiZr alloy. Finally, the
voltage of 300 kV and an electron beam convergence semiangle of alloy with a reduced Al and Mo content replaced with Ta
25 mrad. Cliff-Lorimer quantification and analysis of Super-X™ XEDS (Al0.25NbTaTiZr) has a BCC crystal structure with a = 330.0 pm, similar
data was performed using standardless native k-factors in the Bruker to the Al, Zr-rich phase in the base alloy.
Esprit software. The alloy density was measured using a helium pycnometer and the
A Vickers microhardness unit was used to measure room tempera- results are given in Table 1. The density values calculated using the rule
ture microhardness. Ten measurements were collected and averaged of mixtures of pure elements, Eq. (1), are also given in Table 1 for com-
for each sample using a 1000 g load applied for 15 s. Room temperature parison.
compression tests were conducted using a servo-hydraulic MTS
machine operating at a constant ram speed of 0.008 mm/s and silicon X X
ρcalc ¼ ci M i = ci V i ð1Þ
carbide dies. A thin (~50 μm) teflon foil was used as a lubricating mate-
rial between the die and sample contacting surfaces. The initial sample
dimensions were 5 mm × 5 mm × 8 mm, giving an initial strain rate of In Eq. (1), ci, Mi and Vi are the atomic fraction (taken from Table 2),
0.001 s−1. Sample deformation was monitored and synced with the atomic mass, and atomic volume of element i, respectively. The mea-
load output from the test frame using a Video Displacement and sured density is always higher than the calculated density, which is an
Image Correlation System (Correlated Solutions Inc.), which also pro- indication that the average atomic volume of the elements in each of
vided a visual record of individual specimen deformation response. the alloys is smaller than the average atomic volume of the pure ele-
Compression tests at 1000 °C were conducted in a computer- ments, i.e. the elements experience some constraints (shrinkage) in
controlled Instron mechanical testing machine outfitted with a Brew vac- the studied alloys. The atomic volume reduction is ~1% in these alloys.
uum furnace and silicon carbide dies covered by boron-nitride powder to The atomic volume reduction is an indication of formation of some
reduce friction. The specimen dimensions were 5 mm × 5 mm × 8 mm. degree of covalency in the bonds between the elements and may also
Prior to each test, the furnace chamber was evacuated to 10−6 Torr. The indicate ordering.
sample was then heated to the test temperature over a period of
~45 min and soaked at 1000 °C for 15 min under 22 N load control, and 3.2. Microstructure of the annealed alloys
then compressed to fracture or to a 55% height reduction, whichever
occurred first, at a constant ram speed of 0.008 mm/s that provided 3.2.1. AlMo0.5NbTa0.5TiZr
an initial strain rate of 0.001 s− 1. Load versus frame displacement re- The microstructure of the base AlMo0.5NbTa0.5TiZr alloy was report-
sponses were registered without specimens and these displacement ed in detail elsewhere [11,12,18,19]. SEM-BSE images show the micro-
values were subtracted from the frame displacements measured structure consisted of equiaxed grains, with low-angle subgrain
during the specimen compression at every registered load value. boundaries inside the grains and coarse intermetallic particles at grain
This procedure allowed measurements of deformation of the speci-
mens only and excluded the elastic responses of the dies, machine
Table 2
frame and load cell. Chemical composition (in at.%) of the studied alloys and phases in these alloys determined
using SEM/EDS analysis.
3. Results and discussion
Alloy constituent Al Mo Nb Ta Ti Zr

3.1. X-ray diffraction analysis and alloy density AlMo0.5NbTa0.5TiZr


Alloy 17.6 10.1 21.1 10.6 20.9 19.7
Grains 16.7 11.1 21.7 11.0 22.9 16.6
X-ray diffraction (XRD) patterns of the four modified alloys after HIP GB particles 36.6 2.5 8.1 2.5 8.6 41.7
and annealing at 1400 °C are given in Fig. 1. The diffraction peaks in
AlMo0.5NbTa0.5TiZr0.5
these alloys correspond to a single-phase BCC crystal structure, with Alloy 20.7 12.7 22.3 13.6 20.2 10.4
the lattice parameter changing with composition (Table 1). XRD of the Grains 20.6 12.3 22.9 13.3 22.5 8.4
base AlMo0.5NbTa0.5TiZr alloy conducted in previous work clearly GB particles 37.7 1.2 12.2 3.6 9.9 35.3
showed the presence of two BCC phases [11,19]. Detailed TEM and AlNbTa0.5TiZr0.5
APT phase analysis in AlMo0.5NbTa0.5TiZr revealed that one is a disor- Alloy 25.8 0 24.8 13.5 24.1 11.8
dered BCC (A2) phase enriched with Nb and Ta with the lattice param-
Al0.5Mo0.5NbTa0.5TiZr
eter a = 326.9 pm, and another is an ordered B2 phase enriched with Al Alloy 9.8 11.2 24.5 13.6 21.4 19.5
and Zr with a = 330.4 pm [18,19]. Reducing the amount of Zr in the
AlMo0.5NbTa0.5TiZr0.5 alloy resulted in a single BCC crystal structure Al0.25NbTaTiZr
Alloy 5.0 0 26.0 28.9 21.1 19.0
with a = 326.9 pm (Fig. 1a). Additional removal of Mo (AlNbTa0.5TiZr0.5
O.N. Senkov et al. / Materials and Design 139 (2018) 498–511 501

Fig. 2. AlMo0.5NbTa0.5TiZr: (a) Low-magnification SEM BSE micrograph showing grains, a subgrain structure inside the grains and dark intermetallic particles at grain boundaries. (b) High
magnification BSE image revealing a two-phase, basket-weave nano-lamellar structure inside the grains.

boundaries (GBs) (Fig. 2a). The chemical composition of the alloy, ma- to that from the base alloy, AlMo0.5NbTa0.5TiZr. It consists of equiaxed
trix grains and GB particles is given in Table 2. The GB particles are grains and dark-contrast second-phase GB particles (Fig. 4a). However,
rich in Al and Zr while the matrix grains are slightly depleted with unlike AlMo0.5NbTa0.5TiZ, the higher magnification BSE images do not
these elements, relative to the alloy composition. The volume fraction show contrast modulations inside the grains (Fig. 4b), suggesting that
of the GB particles was estimated to be ~6–12%, depending on the loca- the matrix grains are likely single-phase. XEDS analysis reveals that
tion in the alloy. A detailed analysis using converged beam electron dif- the GB particles have high concentrations of Al and Zr and their compo-
fraction (CBED) revealed that the GB intermetallic particles have an sition is similar to the GB particles in the base alloy (Table 2). The com-
ordered hexagonal crystal structure (space group P63/mcm) [20]. A position of the grains is only slightly different from the composition of
very fine, basket-weave nano-lamellar structure was identified at high the alloy.
magnifications inside the grains (Fig. 2b). It consists of continuous chan- A STEM-HAADF micrograph taken from the interface between a
nels of an ordered B2 phase and cuboidal and/or plate-like disordered darker contrast GB precipitate phase and a lighter contrast intragranular
A2 (BCC) phase between the channels (Fig. 3). These phases are coher- (matrix) phase is shown in Fig. 5. The intragranular region is a single
ent with each other and exhibit continuous lattice registry with a cube- phase. A selected area diffraction analysis revealed that this phase has
on-cube orientation relationship described by 〈100〉BCC//〈100〉B2, an ordered B2 crystal structure. For example, the (100) super-lattice re-
{001}BCC//{001}B2 [18]. The precipitate dimensions in two orthogonal flections indicating ordering are clearly seen in the [001] zone axis of the
directions parallel to the precipitate edges vary from ~ 10 to ~ 55 nm B2 (intragranular) phase (inset in Fig. 5). A low-angle boundary
and from ~25 to ~55 nm, respectively. The width of the B2 channels is consisting of an array of dislocation can also be seen inside the matrix
~ 7.0 ± 1.0 nm. The volume fraction of the disordered BCC phase was grain in Fig. 5.
estimated to be 62 ± 2% [19]. Super-X™ XEDS composition maps were acquired from the region of
the AlMo0.5NbTa0.5TiZr0.5 thinned foil indicated by the HAADF micro-
3.2.2. AlMo0.5NbTa0.5TiZr0.5 graph in Fig. 5. The XEDS spectra were only used to describe general
SEM-BSE micrographs of the annealed AlMo0.5NbTa0.5TiZr0.5 alloy trends in composition. As shown in the elemental maps, Fig. 6, the GB
are shown in Fig. 4. The low magnification microstructure looks similar particle has a higher concentration of Al and Zr and reduced

Fig. 3. STEM-HAADF images of the nano-phase structure present inside the grains of AlMo0.5NbTa0.5TiZr. (a) Cuboidal and plate-like precipitates of a disordered BCC phase are separated by
continuous channels of an ordered B2 phase. (b) The fast Fourier transforms (shown inside the red squares) reveal an ordered B2 structure for the dark channels and a disordered A2 (BCC)
structure for the cuboidal precipitates.
502 O.N. Senkov et al. / Materials and Design 139 (2018) 498–511

Fig. 4. AlMo0.5NbTa0.5TiZr0.5: (a) Low-magnification BSE micrograph showing grains and dark intermetallic particles at grain boundaries. (b) High magnification BSE image revealing a
single-phase structure inside the grains.

concentrations of other elements relative to the matrix (B2) phase, a second phase with a lower average atomic number, and no voids or
which is in agreement with SEM/EDS (Table 2). The regions on both inter-granular cracks were detected along GBs.
sides of a subgrain boundary inside the matrix show the same composi- A STEM-HAADF micrograph of a GB region showing two adjacent
tion as the matrix phase, further indicating a single B2 phase matrix matrix grains and a darker contrast, second phase along the GB in
microstructure. shown in Fig. 8. The matrix grains are single phase with an ordered B2
crystal structure, based on the inset selected area diffraction pattern of
3.2.3. AlNbTa0.5TiZr0.5 a [001] zone axis of the B2 phase. Element ordering is revealed by the
SEM BSE images of annealed AlNbTa0.5TiZr0.5 are shown in Fig. 7. Low presence of faint (100) super-lattice reflections. XEDS composition
magnification micrographs of this Mo-free alloy reveal a coarse-grained maps from the same region are also shown in Fig. 8. Qualitatively, the
(mm size scale), equiaxed single-phase microstructure and complete darker contrast GB phase was rich in Al and Zr compared with the B2
absence of coarse GB precipitates typical in AlMo0.5NbTa0.5TiZr and intra-granular phase. The overall composition of the GB phase in this
AlMo0.5NbTa0.5TiZr0.5 RHEAs (Fig. 7a). Higher magnification micro- modified alloy is similar to the composition of the ordered hexagonal
graphs show no evidence of the modulated two-phase structure inside phase in the base alloy. This Al-Zr intermetallic is surrounded by a re-
the equiaxed grains. Fig. 7b shows a triple point and grain boundaries be- gion deficient in Al and Zr and higher in Ta, which appears as a bright-
tween three large grains. A dark contrast phase with a thin, serrated contrast layer in the HAADF images. Despite the reduction in Zr and
morphology is present uniformly along the grain boundaries. The dark Mo concentration, the microstructure still contains a thin layer of the
contrast and jaggedness of the grain boundary look similar to inter- secondary intermetallic Al-Zr phase along high-angle grain boundaries.
granular cracks. However, a TEM FIB sample extracted and thinned
along a similar grain boundary verified that the contrast was caused by 3.2.4. Al0.5Mo0.5NbTa0.5TiZr
The effect on the microstructure by reducing Al was very different
from reducing Zr. SEM BSE micrographs of Al0.5Mo0.5NbTa0.5TiZr (Al
was decreased by half relative to the base alloy) are shown in Fig. 9.
The low magnification BSE image shows an equiaxed grain microstruc-
ture with needle-like features inside the grains (Fig. 9a). A lamellar
eutectic-like two-phase microstructure is present along high- and
low-angle GBs (Fig. 9b). High magnification images of the grain interior
revealed an orthogonally aligned nanoscale microstructure of two dis-
tinct phases (Fig. 9c), nearly identical to the base alloy nanoscale micro-
structure. Subgrain boundaries within the equiaxed grains are also
present and look similar to those in the AlMo0.5NbTa0.5TiZr base alloy.
The needle-like features inside grains consist of three phases with dis-
tinct Z contrasts (Fig. 9d). The core of the needles consists of a bristle
of a dark phase surrounded by a bright phase. This core is surrounded
by another dark phase, which also contains fine spherical bright parti-
cles at the boundary with the matrix (Fig. 9d). XEDS chemical analysis
of the bright- and dark- contrast phases inside the needle-like features
and at grain/subgrain boundaries (i.e. in the locations containing coarse
particles of these phases) shows that the bright phase is enriched in Mo,
Nb and Ta, while the dark phase is rich in Al and Zr. The dark core bristle
is heavily rich in Zr.
STEM-HAADF micrographs of Al0.5Mo0.5NbTa0.5TiZr are shown in
Fig. 10. An edge of a needle-like precipitate surrounded by the two-
phase nanoscale microstructure is seen at the top of Fig. 10a. The nano-
scale microstructure consists of a higher average atomic number phase
of rounded cuboidal precipitates aligned along 〈100〉 orientations in a
Fig. 5. AlMo0.5NbTa0.5TiZr0.5: STEM-HAADF image acquired along the [001] zone axis of
the intragranular B2 phase bordering an ordered intermetallic GB particle. The selected
lower average atomic number continuous matrix (Fig. 10b). The inset
area diffraction pattern (SADP) shown in the insert was taken from the area indicated SADP shows (100) superlattice reflections indicating sublattice ordering
by the arrow. in at least one of the nanoscale phases. The nanoscale precipitate phase
O.N. Senkov et al. / Materials and Design 139 (2018) 498–511 503

Fig. 6. Super-X™ XEDS compositional maps acquired at the boundary between the intermetallic precipitate and the intragranular phase in annealed AlMo0.5NbTa0.5TiZr0.5.

was of the same size scale of the disordered BCC cuboidal precipitates in in Fig. 12. The lower magnification BSE images reveal a microstructure
the base alloy microstructure but was not as periodic, and only a single consisting of equiaxed grains and a well-developed subgrain structure
precipitate morphology was present. inside the grains (Fig. 12a). No coarse GB particles are present. Higher
XEDS elemental maps acquired near the needle-like precipitate are magnification micrographs show a two-phase nanoscale interpen-
shown in Fig. 11. The lighter contrast phase (Fig. 10) has a composition etrating microstructure, with the two phases coarsened at grain/
higher in Mo, Nb, and Ta compared to the darker contrast matrix phase subgrain boundaries (Fig. 12b), similar to the base alloy.
rich in Al and Zr. The partitioning of elements in the nanoscale phases is A STEM-HAADF micrograph of a thin foil, which was extracted from
similar to the as-received base alloy microstructure in which BCC pre- a grain interior by FIB, is shown in Fig. 13. The nanoscale microstructure
cipitates (rich in Mo, Nb, and Ta) are present in a continuous B2 matrix consists of brighter contrast cuboidal and plate-like precipitates in a
(rich in Al and Zr). Bright-field and dark-field micrographs of the nano- darker contrast continuous matrix. The lighter contrast precipitates co-
scale microstructure revealed that, in similar fashion to the base RHEA, alesce together to form larger structures but remain aligned along cube
the precipitate phase is a disordered solid solution A2 (BCC) phase face directions. Complimentary BF and DF micrographs were acquired
while the continuous matrix is an ordered B2 phase. The phases in the with the electron beam parallel to the [001] zone axis of the nanoscale
needle-like precipitate (at the top and in the center of the XEDS scan) microstructure and are shown in Fig. 14. The inset [001] SADP taken
have the same composition as the nano-scale intra-granular phases, with a large selected area aperture over both phases is consistent with
with the exception of the acicular feature in the center, which has a a BCC crystal structure. The contrast observed in the dark-field micro-
high concentration of Zr (Fig. 11). It is possible that bristles of almost graph indicates that the continuous channels of the matrix phase are
pure Zr precipitated inside grains in the annealed alloy, which are nuclei an ordered B2 structure while the cuboidal precipitates are a disordered
for the formation of the needle-like two-phase precipitates. BCC solid solution phase. While the morphology of the BCC precipitates
is slightly different from the periodic plate and cube structure in the
3.2.5. Al0.25NbTaTiZr base RHEA composition, the phases present in the nanoscale micro-
The Al0.25NbTaTiZr RHEA has even less Al than the previous alloy and structure are the same. A region of the nanoscale microstructure and
all Mo is replaced with Ta. SEM/BSE micrographs of this alloy are shown grain boundary was compositionally mapped using Super-X™ TEM

Fig. 7. SEM BSE micrographs of AlNbTa0.5TiZr0.5 showing (a) equiaxed grain structure and (b) a secondary phase present at grain boundaries.
504 O.N. Senkov et al. / Materials and Design 139 (2018) 498–511

Fig. 8. STEM-HAADF micrograph of grain boundary precipitates between two B2 phase regions with associated SADP taken from the area indicated by the arrow (HAADF) and super-X™
XEDS compositional maps acquired across a high angle grain boundary in annealed AlNbTa0.5TiZr0.5.

XEDS. Similar to the nanoscale microstructure of the base alloy, the or- 3.3. Mechanical properties
dered continuous matrix is rich in Al and Zr while the disordered precip-
itate phase has a higher concentration of Nb and Ta (no Mo in this 3.3.1. Hardness and compression properties at room temperature
modified composition). The coarser precipitates at subgrain/grain Room temperature Vickers hardness and compression properties of
boundaries, similar to those shown in Fig. 12b, have the same composi- the alloys are given in Table 1 and typical stress-strain curves are
tion and the BCC and B2 crystal structures as in the respective intra- shown in Fig. 15a. The base AlMo0.5NbTa0.5TiZr alloy has rather high
granular nanoscale phases. hardness (Hv = 5.8 GPa), yield (σ0.2 = 2197 MPa) and fracture stress

Fig. 9. BSE images of Al0.5Mo0.5NbTa0.5TiZr showing (a) equiaxed grains with needle-like precipitates inside the grains, (b) eutectic-like microstructure at grain/subgrain boundaries,
(c) nanoscale interpenetrating two phase microstructure and a low-angle boundary, and (d) a needle-like particle.
O.N. Senkov et al. / Materials and Design 139 (2018) 498–511 505

Fig. 10. STEM-HAADF micrographs of the Al0.5Mo0.5NbTa0.5TiZr alloy showing the two phase nanoscale microstructure along with (a) a coalesced precipitate and (b) an inset SADP
acquired from the two phase region.

(σf = 2370 MPa), but relatively low fracture strain (εf = 4.1%). A de- These results indicate that Al and Zr reductions strengthen the alloys
crease in the amount of Zr (AlMo0.5NbTa0.5TiZr0.5) increases Hv to while Mo reductions soften the alloys considerably. Therefore, the room
6.2 GPa but makes the alloy brittle and fracture occurs during elastic temperature strength of the base alloy can be improved by increasing
loading at a relatively low fracture stress (σf = 1320 MPa, Table 1). An the amount of Mo and/or by decreasing the amount of Al. The results
additional removal of Mo (AlNbTa0.5TiZr0.5) slightly improves compres- also show that a decrease in Zr makes the alloy more brittle. This simple
sion ductility, but this alloy has low σ0.2 = 1352, σf = 1357 MPa and analysis does not address the influence of compositional changes on mi-
Hv = 5.0 GPa. A decrease in the amount of Al in Al0.5Mo0.5NbTa0.5TiZr crostructure, which will be considered in the Discussion.
RHEA slightly decreases the compression ductility (εf = 3.2%) but no-
ticeably increases σ0.2 (= 2350 MPa), σf (= 2460 MPa) and Hv 3.3.2. Compression properties at 1000 °C
(=6.4 GPa) relative to the base alloy. Finally, Al0.25NbTaTiZr, which has During compression deformation at 1000 °C all five RHEAs had good
even lower Al and completely replaced Mo with Ta, drops σ0.2 to ductility and showed no evidence of fracture when deformation
1745 MPa, σf to 1830 MPa and Hv to 4.9 GPa, while having εf = 3.8%. stopped upon reaching a compression strain of 55%. Typical engineering
The Young's modulus E is highest (≈132–133 GPa) for stress vs. engineering strain curves for these alloys are shown in Fig. 15b
AlMo0.5NbTa0.5TiZr0.5 and Al0.5Mo0.5NbTa0.5TiZr RHEAs. These are about and the characteristic properties are given in Table 3. After plastic yield-
10 GPa higher than the Young's modulus of the base, AlMo0.5NbTa0.5TiZr, ing, all the alloys show a rapid stress increase to a peak value (σp)
alloy. Alloys without Mo (AlNbTa0.5TiZr0.5 and Al0.25NbTaTiZr) have re- followed by a flow softening stage. This flow softening stage switched
duced Young's moduli of 124 GPa and 118 GPa, respectively (Table 1). to a steady-state stage with a nearly constant flow stress in the

Fig. 11. XEDS compositional maps acquired at a lamellar precipitate surrounded by a two phase nanoscale microstructure in Al0.5Mo0.5NbTa0.5TiZr.
506 O.N. Senkov et al. / Materials and Design 139 (2018) 498–511

Fig. 12. BSE images of Al0.25NbTaTiZr showing (a) equiaxed grains and low-angle grain boundaries inside the grains and (b) a two phase interpenetrating nanoscale microstructure inside
the grains and coarser particles at a subgrain boundary.

AlMo0.5NbTa0.5TiZr and AlMo0.5NbTa0.5TiZr0.5 alloys, or replaced by a 3.3.3. Microstructure after compression deformation at 1000 °C
hardening stage in the AlNbTa0.5TiZr0.5 alloy, or continued up to about After 55% compression deformation at 1000 °C, grains retained near-
30% compression strain in the Al0.5Mo0.5NbTa0.5TiZr and Al0.25NbTaTiZr equiaxed shapes and their average size was almost the same as before
alloys (Fig. 15b). A stress increase after ~ 30% compression strain, ob- deformation (Figs. 16–19). Characteristic electron channeling contrast
served in all samples, was due to friction between the contact surfaces features in the form of bands, spots and/or protrusions were found on
of the specimens and the dies. the BSE images of many deformed grains (e.g. Figs. 16a–b, 17a, 18a,
The AlMo0.5NbTa0.5TiZr0.5 RHEA (with a reduced amount of Zr) was 19a). These channeling contrast features reveal the presence of internal
the strongest one at 1000 °C, having σ0.2 = 935 MPa and σp = stresses and local rotations of the crystal lattice inside the deformed
1019 MPa. This alloy also showed a steady-state flow, with the steady- grains [22]. The matrix grains in alloys with reduced Zr retain their
state flow stress of ~910 MPa. These stress values are noticeably higher single-phase B2 structure (Fig. 16b). A few former GB precipitates in
than those shown by the base alloy. Additional removal of Mo resulted the AlMo0.5NbTa0.5TiZr0.5 alloy are located inside the grains (Fig. 16a),
in a significant decrease in σ0.2 to 535 MPa and σp to 542 MPa in the which could indicate a high mobility of grain boundaries and their
AlNbTa0.5TiZr0.5 RHEA. The Al0.5Mo0.5NbTa0.5TiZr alloy (with a reduced unpinning from GB particles during deformation at 1000 °C.
amount of Al) has σ0.2 = 579 MPa and σp = 646 MPa. The Submicron-sized, dynamically recrystallized grains are present at the
Al0.25NbTaTiZr RHEA, which has even less Al and no Mo, shows the boundaries of deformed grains in the AlNbTa0.5TiZr0.5 alloy (Fig. 17b).
smallest strength at 1000 °C: σ0.2 = 366 MPa, σp = 408 MPa. After Dynamic recrystallization broke up a continuous layer of the Al-Zr rich
reaching the peak stress at εp ≈ 1.6–2.0% these two alloys started losing phase observed at grain boundaries before deformation (Fig. 7b). After
strength continuously, and after 20% compression strain they had a flow deformation, this secondary phase is present in the form of fine, individ-
stress σ20 = 312 MPa and 162 MPa, respectively. It can be clearly seen ual particles mixed with the recrystallized grains (Fig. 17b).
that Mo has a strong strengthening effect in these RCCAs. The presence The alloys with reduced Al (Al0.5Mo0.5NbTa0.5TiZr and
of Al is likely also beneficial for the high-temperature strength, while Zr Al0.25NbTaTiZr) retained their two-phase nano-structure inside grains
has an opposite effect. (Figs. 18 and 19). However, in both the alloys the nano-scale precipi-
tates coarsened during deformation and the BCC phase (bright in the
BSE images) became continuous, while formerly continuous B2
channels broke into separate elongated particles (Figs. 18d and 19c).
No noticeable coarsening of the needle-like features common to the
annealed Al0.5Mo0.5NbTa0.5TiZr alloy (see Fig. 9) occurred during defor-
mation at 1000 °C, they became hardly visible on the low-magnification
images (Fig. 18a) but could be identified at high magnifications as long
chains of elongated B2 particles covered by a layer of the BCC phase
(Fig. 18c).

3.4. Discussion

3.4.1. Alloys with a reduced amount of Zr


The results of this work show that a ½ decrease in the molar ratio of
Zr relative to the base alloy does not remove the hexagonal GB phase
from the RHEA microstructure. Moreover, the average particle size
and the volume fraction of this phase are not affected. However, the
nano-scale disordered BCC precipitates disappear and the matrix grains
have an ordered single-phase B2 crystal structure. The B2 phase has a
composition close to the overall composition of the modified alloy
(AlMo0.5NbTa0.5TiZr0.5). The fully ordered state explains the lack of
ductility of this alloy at room temperature and the highest plastic flow
stress among the studied alloys at 1000 °C (see Fig. 15).
Fig. 13. STEM-HAADF micrograph of Al0.25NbTaTiZr showing a nanoscale two phase When Mo is additionally removed (AlNbTa0.5TiZr0.5), the coarse GB
microstructure. particles disappear and are replaced by a thin, continuous GB layer of
O.N. Senkov et al. / Materials and Design 139 (2018) 498–511 507

Fig. 14. TEM (a) BF and (b) DF images of a nano-phase structure of Al0.25NbTaTiZr acquired with the transmitted and (100) superlattice reflection, respectively, as indicated by the circles in
the SADP insets.

the ordered hexagonal phase. This Mo-free alloy is essentially a single- nano-scale morphologies. Additionally, Al0.5Mo0.5NbTa0.5TiZr RHEA
phase ordered B2 crystal structure. Mo is present in the hexagonal GB contains needle-like features that are evenly distributed in the alloy.
phase in a very small amount and removing this element from the These features consist of coarser BCC lamellae nucleated on Zr-rich bris-
alloy was not expected to affect the nature of the GB phase. Mo may fa- tles. This indicates that reducing Al in Al0.5Mo0.5NbTa0.5TiZr produces Zr
cilitate formation of the GB particles by restricting solubility of Al in the supersaturation and forms Zr-rich bristles in a BCC solid solution phase
matrix phase. When Mo is removed, the solubility of Al in the B2 phase present at high temperatures. These bristles become nucleation sites for
may increase, leaving no extra Al to form the coarse (Al, Zr)-rich GB par- a lower-temperature, Mo-Nb-Ta-rich BCC phase leading to local coars-
ticles. However, the amount of Al in AlNbTa0.5TiZr0.5 alloy probably still ening relative to homogeneously nucleated nano-scale particles in
exceeds the maximum solubility, which results in the formation of a other regions. After substituting Mo with Ta and an additional Al de-
thin, Al-Zr-rich layer at grain boundaries. The removal of Mo slightly im- crease in the Al0.25NbTaTiZr RHEA, Zr-rich bristles do not form and the
proves ductility of this alloy (relative to the completely brittle needle-like features disappear. This RHEA has the most homogeneous
AlMo0.5NbTa0.5TiZr0.5) at room temperature, which may indicate that and refined microstructure among the five studied alloys. This suggests
the coarse GB particles do deteriorate RT ductility. The removal of Mo that Mo reduces the solubility of not only Al but also Zr in the high-
also considerably decreases the plastic flow stress at 1000 °C, indicating temperature BCC phase. Both the Al0.5Mo0.5NbTa0.5TiZr and
that Mo is a strong solid solution strengthening element in this alloy Al0.25NbTaTiZr RHEAs have high strength and some ductility at room
system. temperature, although substitution of Mo with Ta decreases the RT
strength by ~ 600 MPa and the 1000 °C strength by ~ 150–200 MPa.
3.4.2. Alloys with a reduced amount of Al The most dramatic effect from decreasing the amount of Al (and Mo re-
Reducing the amount of Al in Al0.5Mo0.5NbTa0.5TiZr and moval) is observed during compression testing at 1000 °C. These alloys
Al0.25NbTaTiZr RHEAs is much more effective for removing the ordered soften continuously after reaching the peak stress at the beginning of
hexagonal GB phase. This phase is absent in both of the Al-lean alloys, plastic deformation, and the flow stress drops by ~50% after 20% plastic
supporting our earlier suggestion that excessive Al is likely responsible strain. The flow softening is likely due to coarsening and coalescence of
for the formation of this GB phase. Similar to the base alloy, those the BCC nano-scale particles (compare Fig. 18 with Fig. 9 and Fig. 19
with the reduced Al consist of two coherent phases: an ordered (Al with Fig. 12). As a result, these BCC nano-particles transform to a contin-
and Zr rich) B2 matrix and a disordered (Nb, and Ta rich) BCC phase uous matrix phase, while the formerly continuous channels of the or-
with spherical (in Al0.5Mo0.5NbTa0.5TiZr) or cuboidal (in Al0.25NbTaTiZr) dered B2 phase fragment and become embedded inside the BCC

(a) (b)
2500 1200
#4 T = 1000oC
#2
Engineering Stress (MPa)
Engineering Stress (MPa)

#1 1000
2000

800 #1
1500 #2 #5
600 #3
#3
1000 #4
400

500 200
#5
T = 25 oC
0 0
0 1 2 3 4 5 0 10 20 30 40 50 60
Engineering Strain (%) Engineering Strain (%)

Fig. 15. Engineering stress vs. engineering strain curves obtained during (a) room temperature (T = 25 °C) and (b) elevated temperature (T = 1000 °C) compression testing of
AlMo0.5NbTa0.5TiZr (alloy #1), AlMo0.5NbTa0.5TiZr0.5 (#2), AlNbTa0.5TiZr0.5 (#3), Al0.5Mo0.5NbTa0.5TiZr (#4) and Al0.25NbTaTiZr (#5).
508 O.N. Senkov et al. / Materials and Design 139 (2018) 498–511

Table 3 forms, while Zr-rich alloys also experience an eutectoid transformation


Compression properties of the compositionally modified RHEAs at 1000 °C (Young's mod- BCC → HCP + Laves at 738 °C [25].
ulus, E, yield stress, σ0.2, peak stress, σp, peak strain at peak stress, εp, stress at 10% strain,
σ10, and stress at 20% strain, σ20).
Al forms numerous binary intermetallic compounds with all the ele-
ments in the studied alloy systems; however, none of these binary com-
Alloy no. Alloy ID E σ0.2 σp εp σ10 σ20 pounds has an ordered B2 crystal structure. B2 ordering has, however,
GPa MPa MPa % MPa MPa
been observed in ternary or higher order Al-Ti-X and Al-Zr-X systems,
1 AlMo0.5NbTa0.5TiZr [11] 70 745 792 1.9 735 727 where X is Mo, Nb, Ta or the combination of these elements [26–28].
2 AlMo0.5NbTa0.5TiZr0.5 76 935 1019 2.0 907 912
Using this information on phase relationships, it is suggested that
3 AlNbTa0.5TiZr0.5 53 535 542 1.4 433 483
4 Al0.5Mo0.5NbTa0.5TiZr 78 579 646 2.0 461 312 Nb, Ta and Zr are responsible for the formation of the two-phase,
5 Al0.25NbTaTiZr 63 366 408 1.6 288 162 nano-scale microstructure inside the matrix grains. The matrix grains
represent a high-temperature BCC phase, which, during cooling, sepa-
rates into two BCC phases due to the presence of the miscibility gap.
matrix (see Figs. 18d and 19c). Eventually, the microstructure contain- One phase is rich in Nb and Ta and another phase is rich in Zr. Ti
ing the continuous disordered BCC matrix is softer and more ductile seems to be neutral in this reaction and is distributed almost evenly be-
than that with continuous ordered B2 matrix phase. tween these two phases. Mo prefers the Nb and Ta rich phase due to its
complete solubility with these elements. On the other hand, Al prefers
3.4.3. The mechanism of formation of nanometer-scaled B2/BCC the Zr-rich BCC phase due to a higher solubility and the highest negative
microstructure enthalpy of mixing between Al and Zr, relative to those with other
The mechanism of formation of the coherent, two-phase nano- alloying elements [29]. The presence of Nb and Ta in the Al-Zr rich
scaled microstructure inside the matrix grains of this refractory alloy BCC phase allows the disorder-order transformation (BCC → B2) during
system has not been well understood [12,18,19]. Results of the present cooling, without formation of other Al-containing phases. It is likely that
work suggest that Zr is somehow responsible for the formation of this a decrease in the amount of Al in the Al0.5Mo0.5NbTa0.5TiZr and
two-phase nano- structure. Decreasing the concentration of this ele- Al0.25NbTaTiZr RHEAs decreases the disorder-order transformation
ment suppresses the phase separation resulting in an ordered single- temperature below 1000 °C, which explains the reduced strength and
phase B2 crystal structure. On the other hand, Al does not seem to noticeable deformation softening in these Al lean alloys during defor-
play important role in the phase separation, although its presence is mation at 1000 °C. A decrease in the amount of Zr moves the alloy com-
likely responsible for the formation of the ordered B2 phase [23,24]. position outside the miscibility gap and the Zr-rich BCC phase does not
The refractory elements Mo, Nb and Ta are completely miscible in form. In this case, the interaction of Al with other elements in the high-
both liquid and solid states, and in the solid state they form continuous temperature BCC phase results in ordering and complete transforma-
BCC solid solutions [25]. Ti also forms continuous BCC solid solutions tion of this phase in the ordered B2 phase upon cooling. It is likely that
with these three refractory elements at high temperatures, but a two- ordering in this case occurs at temperatures above 1000 °C, because
phase range, BCC + HCP, can be present below 880 °C with the refracto- these Zr-lean alloys show high strength, no softening and no evidence
ry elements suppressing the HCP solvus temperature. In the Ti-Zr sys- of the phase transformations during deformation at 1000 °C.
tem, a continuous BCC solid solution region is present at high
temperatures and an HCP solid solution region is present at lower tem- 3.4.4. Mechanical properties
peratures. Zr forms continuous BCC solid solutions with Nb and Ta at Single-phase B2 RHEAs show the highest (AlMo0.5NbTa0.5TiZr0.5)
high temperatures. At lower temperatures, a two-phase, BCC + HCP re- and next-to-lowest (AlNbTa0.5TiZr0.5) yield strengths at 1000C, suggest-
gion is present at the concentrations of Nb and Ta below 19% and. 6.5%, ing that this phase can be intrinsically strong at high temperatures, and
respectively. At higher Nb and Ta concentrations, i.e. 19–92.2% Nb or that solid solution hardening (Mo-induced in this case) plays a signifi-
6.5–91% Ta, a miscibility gap is present in the temperature range of cant role. However, these are the most brittle alloys and show essential-
620–977 °C in Nb-Zr or 800–1760 °C in Ta-Zr systems, where the ly no plastic deformation in room temperature compression. The B2
high-temperature BCC phase separates in two BCC phases, one of phase typically deforms either by ½ 〈111〉 or by 〈100〉 dislocations.
which is rich with Nb or Ta and another with Zr [25]. Finally, both Nb- The former provides 5 independent strains needed for plasticity in poly-
Zr and Ta-Zr experience eutectoid reactions (at 620 °C or 800 °C, respec- crystalline materials and the latter do not. The present results suggest
tively) during which the Zr-rich BCC phase transforms to Zr-rich HCP that the B2 phase studied here may deform by 〈100〉 dislocations at
and Nb- or Ta-rich BCC phases. In the Mo-Zr system, Zr has the maxi- room temperature. However, some B2 compounds that deform by ½
mum solubility of ~10% in BCC-Mo and Mo has a maximum solubility 〈111〉 slip can still be brittle [30], and so TEM studies are required
of ~41% in BCC-Zr. Beyond these solubility limits, a Laves Mo2Zr phase to determine which type of dislocations operate in these RHEAs.

Fig. 16. Microstructure of AlMo0.5NbTa0.5TiZr0.5 after 55% compression at 1000 °C.


O.N. Senkov et al. / Materials and Design 139 (2018) 498–511 509

Fig. 17. Microstructure of AlNbTa0.5TiZr0.5 after 55% compression at 1000 °C: (a) heavily deformed grains and (b) dynamically recrystallized grains and second-phase particles (dark) at a
grain boundary.

The two-phase intragranular microstructure in the base become continuous during deformation at 1000 °C. This transition is
AlMo0.5NbTa0.5TiZr RHEA and in the Al-lean alloys Al0.5Mo0.5NbTa0.5TiZr accompanied by extensive compressive ductility and a continuous drop
and Al0.25NbTaTiZr can be thought of as analogous to the γ/γ′ microstruc- in flow stress, which could result from the rather high volume fraction
ture in superalloys. In those microstructures, a relatively soft and ductile of the BCC phase (see Fig. 18d). These results also indicate that the low
disordered FCC continuous phase (γ) surrounds nanometer-sized, coher- ductility B2/BCC structure can be transformed in a more ductile BCC/B2
ent precipitates of the ordered FCC Ni3Al intermetallic phase. It might structure by extensive hot working, such as forging and/or rolling. The
therefore be expected that the BCC/B2 (or β/β′) microstructures in concept of designing a BCC/B2 analog to the γ/γ′ microstructure of
these three alloys may have potential to provide a balance of strength superalloys thus requires further study to provide a balance of room
and ductility. Supporting this view, these alloys have the highest com- temperature ductility and high temperature strength.
pressive strengths at room temperature and the base alloy is the second
strongest at 1000 °C. However, the present microstructures are reversed 4. Summary and conclusions
from the γ/γ′ stereotype – the strong but brittle B2 phase is continuous
and the disordered BCC precipitates are discontinuous. This may explain Compositional adjustments of the base AlMo0.5NbTa0.5TiZr refractory
the poor room temperature ductility of these alloys. The BCC phase does high entropy superalloy were made to improve mechanical properties

Fig. 18. Microstructure of Al0.5Mo0.5NbTa0.5TiZr after 55% compression at 1000 °C: (a) Equiaxed grain structure, (b) subgrain and nano-phase structures inside a grain, (c) needle-like
features consisting of a B2 core and a BCC shell and (d) a high magnification image illustrating that the BCC phase (bright) is a matrix phase and the B2 phase is discontinuous after
the compression deformation.
510 O.N. Senkov et al. / Materials and Design 139 (2018) 498–511

Fig. 19. Microstructure of Al0.25NbTaTiZr after 55% compression at 1000 °C: (a) Equiaxed grain structure, (b) a junction of three grains, (c) nano-phase structures inside a grain.

by modifying the phase content and microstructure. For this, four new Analysis of available phase diagrams suggests that Nb, Ta and Zr are
alloys were produced by reducing the amounts of Zr, Al and/or Mo. likely responsible for the decomposition of a high-temperature BCC
These are AlMo0.5NbTa0.5TiZr0.5, AlNbTa0.5TiZr0.5, Al0.5Mo0.5NbTa0.5TiZr phase in two BCC phases during cooling due to a miscibility gap in the
and Al0.25NbTaTiZr. The microstructure, microhardness and compression studied alloy system. When the amount of Zr is reduced, the alloy com-
properties of the cast alloys were studied in a 1400 °C annealed position is moved outside the miscibility gap and no phase separation
condition. occurs. Strong interaction of Al with Zr in the presence Nb and Ta is like-
The microstructure of the base AlMo0.5NbTa0.5TiZr RHEA consists of ly responsible for the BCC → B2 ordering of a Zr-rich phase upon cooling.
equiaxed grains of an ordered B2 matrix phase, coherent, nanometer- The ordered intermetallic B2 phase is strong when Mo is present, but
sized precipitates of a disordered BCC phase (~60% by volume) inside is brittle at room temperature causing brittleness or low ductility of the
the grains and coarse particles (~10%) of an ordered hexagonal phase studied RHEAs. In the two-phase B2/BCC RHEAs the disordered BCC
at GBs. The alloy has Hv = 5.8 GPa, σ0.2 = 2197 MPa, εf = 4% at 25 °C phase becomes continuous during deformation at 1000 °C, and this
and σ0.2 = 735 MPa, εf N 55% at 1000 °C. transformation is accompanied by extensive compressive ductility and
Reducing Zr eliminates the BCC phase in AlMo 0.5NbTa0.5TiZr0.5 a progressive drop in flow stress. It is suggested that the transformed
and additional removal of Mo also eliminates the ordered hexago- microstructure consisting of the disordered BCC matrix and ordered
nal phase in AlNbTa0.5 TiZr 0.5 , making the latter RHEA a single- B2 nanometer-sized precipitates would have room temperature ductil-
phase B2 structure. Both alloys are brittle (εf b 1.5%) at 25 °C and ity sufficient for engineering applications. Additional work is suggested
ductile (εf N 55%) at 1000 °C. AlMo0.5NbTa0.5TiZr0.5 has exception- to engineer two-phase BCC/B2 microstructures with a balance of room
ally high strength (σ 0.2 = 935 MPa) at 1000 °C. Removing Mo temperature tensile ductility and high temperature strength.
decreases the 1000 °C strength by 400 MPa (σ0.2 = 535 MPa) in
AlNbTa0.5TiZr0.5. Acknowledgements
Decreasing Al produces a two-phase Al0.5Mo0.5NbTa0.5TiZr
consisting of the ordered B2 matrix and coherent nano-meter-sized Work by ONS was supported through the Air Force on-site contract
spherical BCC precipitates. A further decrease in Al and replacement of FA8650-15-D-5230 managed by UES, Inc., Dayton, Ohio and through
Mo with Ta retained the two-phase structure in Al0.5NbTaTiZr, however, the Air Force Research Laboratory, Materials and Manufacturing Direc-
with cuboidal-shaped coherent nano-meter-sized BCC precipitates. torate, Lab Director's funds. JKJ acknowledges the financial support of
Al0.5Mo0.5NbTa0.5TiZr has σ0.2 = 2350 MPa, εf = 3.2% at 25 °C and the Dayton Area Graduate Studies Institute (DAGSI) program of the
σ0.2 = 579 MPa, εf N 55% at 1000 °C. The alloy strength decreases con- State of Ohio and the Air Force Research Laboratory, grant number
siderably, by ~600 MPa at 25 °C and 210 MPa at 1000 °C, when Mo is re- RX0(1)-OSU-13-4.
placed with Ta, so that Al0.25NbTaTiZr has σ0.2 = 1745 MPa, εf = 3.8% at
25 °C and σ0.2 = 366 MPa, εf N 55% at 1000 °C. Both these alloys soften References
during deformation at 1000 °C due to microstructural transformations
resulting in the disordered BCC phase becoming a continuous matrix [1] B.S. Murty, J.-W. Yeh, S. Ranganathan, High Entropy Alloys, Butterworth-
Heinemann, London, UK, 2014.
and the ordered B2 phase to transform to discrete particles inside the [2] D.B. Miracle, O.N. Senkov, A critical review of high entropy alloys and related con-
BCC matrix. cepts, Acta Mater. 122 (2017) 448–511.
O.N. Senkov et al. / Materials and Design 139 (2018) 498–511 511

[3] H.Y. Diao, R. Feng, K.A. Dahmen, P.K. Liaw, Fundamental deformation behavior in [17] N.Y. Yurchenko, N.D. Stepanov, S.V. Zherebtsov, M.A. Tikhonovsky, G.A. Salishchev,
high-entropy alloys: an overview, Curr. Opinion Solid State Mater. Sci. 21 (2017) Structure and mechanical properties of B2 ordered refractory AlNbTiVZrx (x =
252–266. 0–1.5) high-entropy alloys, Mater. Sci. Eng. A 704 (2017) 82–90.
[4] O.N. Senkov, G.B. Wilks, J.M. Scott, D.B. Miracle, Mechanical properties of [18] J.K. Jensen, B.A. Welk, R.E.A. Williams, J.M. Sosa, D.E. Huber, O.N. Senkov, G.B.
Nb25Mo25Ta25W25 and V20Nb20Mo20Ta20W20 refractory high entropy alloys, Inter- Viswanathan, H.L. Fraser, Characterization of the microstructure of the composition-
metallics 19 (2011) 698–706. ally complex alloy AlMo0.5NbTa0.5TiZr, Scr. Mater. 121 (2016) 1–4.
[5] O.N. Senkov, C.F. Woodward, Microstructure and properties of a refractory [19] O.N. Senkov, D. Isheim, D.N. Seidman, A.L. Pilchak, Development of a refractory high
NbCrMo0.5Ta0.5TiZr alloy, Mater. Sci. Eng. A 529 (2011) 311–320. entropy superalloy, Entropy 18 (2016) 102.
[6] O.N. Senkov, J.M. Scott, S.V. Senkova, F. Meisenkothen, D.B. Miracle, C.F. Woodward, [20] J.K. Jensen, Characterization of a High Strength, Refractory High Entropy Alloy,
Microstructure and elevated temperature properties of a refractory TaNbHfZrTi AlMo0.5NbTa0.5TiZrPhD Dissertation The Ohio State University, Columbus, OH,
alloy, J. Mater. Sci. 47 (2012) 4062–4074. USA, 2017.
[7] Y.D. Wu, Y.H. Cai, X.H. Chen, T. Wang, J.J. Si, L. Wang, Y.D. Wang, S.D. Hui, Phase com- [21] J.M. Sosa, D.E. Huber, B. Welk, H.L. Fraser, Development and application of MIPAR™:
position and solid solution strengthening in TiZrNbMoV high-entropy alloys, Mater. a novel software package for two- and three-dimensional microstructural charac-
Des. 83 (2015) 651–660. terization, Integr. Mater. Manuf. Innov. 3 (2014) 10.
[8] H. Jiang, H. Zhang, T. Huang, Y. Lu, T. Wang, T. Li, Microstructure and mechanical [22] D.E. Newbury, D.C. Joy, P. Echlin, C.E. Flori, J.I. Goldstein, Advanced Scanning Electron
properties of Co2MoxNi2VWx eutectic high-entropy alloys, Mater. Des. 109 (2016) Microscopy and X-ray Microanalysis, Plenum Press, New York, 1986.
539–546. [23] Y. Yamabe-Mitarai, H. Harada, Design of quaternary Ir-Nb-Ni-Al refractory superal-
[9] Z.D. Han, N. Chen, S.F. Zhao, L.W. Fan, G.N. Yang, Y. Shao, K.F. Yao, Effect of Ti addi- loys, Metall. Mater. Trans. A 31 (2000) 173–178.
tions on mechanical properties of NbMoTaW and VNbMoTaW refractory high entro- [24] J.P. Neumann, Y.A. Chang, H. Ipser, On the relationship between the enthalpy of for-
py alloys, Intermetallics 84 (2017) 153–157. mation and the disorder parameter of intermetallic phases with the B2 structure,
[10] O.N. Senkov, S.V. Senkova, D.B. Miracle, C. Woodward, Mechanical properties of low- Scr. Mater. 10 (1976) 917–922.
density, refractory multi-principal element alloys of the Cr–Nb–Ti–V–Zr system, [25] H. Okamoto, Phase Diagrams for Binary Alloys, 2nd ed. ASM International, Materials
Mater. Sci. Eng. A 565 (2013) 51–62. Park, Ohio, 2010.
[11] O.N. Senkov, C. Woodward, D.B. Miracle, Microstructure and properties of [26] D.M. Cupid, O. Fabrichnaya, O. Rios, F. Ebrahimi, H.J. Seifert, Thermodynamic re-
aluminum-containing refractory high-entropy alloys, JOM 66 (2014) 2030–2042. assessment of the Ti–Al–Nb system, Int. J. Mater. Res. 100 (2009) 218–233.
[12] O.N. Senkov, S.V. Senkova, C. Woodward, Effect of aluminum on the microstructure [27] D.H. Hou, H.L. Fraser, The ordering scheme in N-B aluminides with the B2 crystal-
and properties of two refractory high entropy alloys, Acta Mater. 68 (2014) structure, Scr. Mater. 36 (1997) 617.
214–228. [28] Y.V. Lysova, M. Materials Science International Team, Al-Ta-Zr Ternary Phase Dia-
[13] N.D. Stepanov, D.G. Shaysultanov, G.A. Salishchev, M.A. Tikhonovsky, Structure and gram Evaluation. Phase Diagrams, Crystallographic and Thermodynamic data:
mechanical properties of a light-weight AlNbTiV high entropy alloy, Mater. Lett. 142 Datasheet from MSI Eureka in SpringerMaterials, http://materials.springer.com/
(2015) 153–155. msi/literature/docs/sm_msi_r_10_022102_01MSI Materials Science International
[14] N.D. Stepanov, N.Y. Yurchenko, V.S. Sokolovsky, M.A. Tikhonovsky, G.A. Salishchev, Services GmbH, 1993.
An AlNbTiVZr0.5 high-entropy alloy combining high specific strength and good duc- [29] A. Takeuchi, A. Inoue, Mixing enthalpy of liquid phase calculated by Miedema's
tility, Mater. Lett. 161 (2015) 136–139. scheme and approximated with sub-regular solution model for assessing forming
[15] C.-M. Lin, C.-C. Juan, C.-H. Chang, C.-W. Tsai, J.-W. Yeh, Effect of Al addition on me- ability of amorphous and glassy alloys, Intermetallics 18 (2010) 1779–1789.
chanical properties and microstructure of refractory AlxHfNbTaTiZr alloys, J. Alloys [30] K. Vedula, FeAl and Fe3Al, in: J.H. Westbrook, R.L. Fleischer (Eds.), Intermetallic Com-
Compd. 624 (2015) 100–107. pounds: Principles and Practice, John Wiley & Sons, Ltd., West Sussex, England 1995,
[16] H. Chen, A. Kauffmann, B. Gorr, D. Schliephake, C. Seemuller, J.N. Wagner, H.J. Christ, pp. 199–209.
M. Heilmaier, Microstructure and mechanical properties at elevated temperatures of
a new Al-containing refractory high-entropy alloy Nb-Mo-Cr-Ti-Al, J. Alloys Compd.
661 (2016) 206–215.

You might also like