ArticlePDF Available

Characteristic strengthening mechanisms in body-centered cubic refractory high/medium entropy alloys

Authors:

Abstract and Figures

The present work reports characteristic strengthening mechanisms of HfNbTaTi, HfNbTiZr, HfNbTi, HfTaTi, and NbTiZr equi-atomic refractory medium entropy alloys (RMEAs). The alloys were processed by high-pressure torsion and subsequent annealing to obtain microstructures with average grain sizes ranging from several hundred nanometers to several tens of micrometers. Their mechanical properties were evaluated by tensile tests at room temperature. Precise Hall–Petch relationships of the RMEAs were acquired based on the tensile yield strength data. Small slopes of the Hall–Petch relationships and weak grain refinement strengthening were clarified to be attributed to their low elastic modulus. In addition, the friction stresses in the RMEAs were higher than those of conventional BCC metals and alloys. By comparing the experimental data and a theoretical model, it was suggested that interaction between severe lattice distortion and elastic field of edge dislocations can largely contribute to the high friction stress of the RMEAs at room temperature.
Content may be subject to copyright.
Scripta Materialia 231 (2023) 115442
1359-6462/© 2023 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Characteristic strengthening mechanisms in body-centered cubic refractory
high/medium entropy alloys
Qian He
a
, Shuhei Yoshida
a
,
b
,
*
, Nobuhiro Tsuji
a
,
b
a
Department of Materials Science and Engineering, Kyoto University, 606-8501 Kyoto, Japan
b
Elements Strategy Initiative for Structural Materials (ESISM), Kyoto University, 606-8501 Kyoto, Japan
ARTICLE INFO
Keywords:
Refractory high/medium entropy alloys
Strengthening mechanisms
Grain renement
HallPetch relationships
Dislocations
ABSTRACT
The present work reports characteristic strengthening mechanisms of HfNbTaTi, HfNbTiZr, HfNbTi, HfTaTi, and
NbTiZr equi-atomic refractory medium entropy alloys (RMEAs). The alloys were processed by high-pressure
torsion and subsequent annealing to obtain microstructures with average grain sizes ranging from several
hundred nanometers to several tens of micrometers. Their mechanical properties were evaluated by tensile tests
at room temperature. Precise HallPetch relationships of the RMEAs were acquired based on the tensile yield
strength data. Small slopes of the HallPetch relationships and weak grain renement strengthening were
claried to be attributed to their low elastic modulus. In addition, the friction stresses in the RMEAs were higher
than those of conventional BCC metals and alloys. By comparing the experimental data and a theoretical model,
it was suggested that interaction between severe lattice distortion and elastic eld of edge dislocations can
largely contribute to the high friction stress of the RMEAs at room temperature.
Refractory elements, such as Zr, Hf, V, Nb, Ta, Cr, Mo, W, and Ti,
belong to the groups IVB, VB, and VIB in the periodic table of elements
[1] and possess relatively high melting points. Refractory high entropy
alloys (RHEAs) and refractory medium entropy alloys (RMEAs) are new
classes of refractory alloys consisting of ve or more and three or four
(near-) equi-atomic refractory elements, respectively. Most RHEAs/R-
MEAs have been reported to exhibit single-phase body-centered cubic
(BCC) crystal structure after rapid quenching from high temperatures
above their β-phase solvus temperatures. RHEAs/RMEAs typically show
high yield strengths at a wide range of temperatures [24]. However,
owing to the lack of systematic investigations, the reasons for the high
strengths of RHEAs/RMEAs have been unknown. In the previous studies
on RHEAs/RMEAs, as-cast ingots having complex microstructures
composed of grains with a heterogeneous size distribution, second-phase
precipitates, and elemental segregation were used to characterize their
mechanical properties [5]. However, our recent study has veried that
microstructures largely affect the mechanical properties of RHEAs/R-
MEAs, and as-cast materials with complex microstructures are not
suitable for studying their strengthening mechanisms [6].
In the present study, polycrystals of RMEAs (HfNbTaTi, HfNbTiZr,
HfNbTi, HfTaTi, and NbTaTi, which were family of HfNbTaTiZr RHEA)
with BCC single phase composed of simple equiaxed microstructures
were used to quantitatively reveal characteristics of their strengthening
mechanisms. The polycrystal specimens of the RMEAs with a wide range
of mean grain sizes were fabricated and their mechanical properties
were measured by tensile tests at room temperature (RT). Precise
HallPetch relationships of the RMEAs were obtained, and their friction
stresses (fundamental resistance to dislocation glide in solid solutions)
and the HallPetch slopes were systematically compared. Finally, the
characteristics of the strengthening mechanisms in RHEAs/RMEAs with
BCC structures were discussed based on the contribution of solid solu-
tion strengthening and grain renement strengthening.
Ingots of equi-atomic HfNbTaTi, HfNbTiZr, HfNbTi, HfTaTi, and
NbTiZr RMEAs were produced by vacuum arc-melting of pure metals.
Each arc-melted button was ipped and remelted ve times to homog-
enize elements distribution. Cold rolling with a 90% reduction in
thickness and subsequent homogenization at 1150 C for 24 h under an
Ar atmosphere followed by water quenching were applied to eliminate
complex microstructures in the as-cast materials. Discs with a diameter
of 10 mm and thickness of 0.8 mm were cut, and they were heavily
deformed by high-pressure torsion (HPT) with ve rotations (maximum
equivalent strain of 196 at the edge) at a speed of 0.2 rpm (rotation per
minute) under 7.5 GPa. To obtain the RMEAs with BCC single phase with
a wide range of average grain size, all the RMEAs were subsequently
* Corresponding author at: Department of Materials Science and Engineering, Kyoto University, 606-8501 Kyoto, Japan.
E-mail address: yoshida.shuhei.5s@kyoto-u.ac.jp (S. Yoshida).
Contents lists available at ScienceDirect
Scripta Materialia
journal homepage: www.journals.elsevier.com/scripta-materialia
https://doi.org/10.1016/j.scriptamat.2023.115442
Received 6 December 2022; Received in revised form 27 February 2023; Accepted 20 March 2023
Scripta Materialia 231 (2023) 115442
2
annealed at temperatures ranging from 750 C to 1250 C (corre-
sponding to the BCC single phase region in Fig. S1 in the supplementals)
for short periods (30120 s).
The annealed disks were mechanically polished, and then electro-
polishing was applied in a solution of 10 vol.% HClO
4
and 90 vol.%
C
2
H
5
OH at 20 V and 30 C for 40 s. Field emission scanning electron
microscopy (FE-SEM) (JEOL, JSM-7800F) equipped with a back-
scattered electron (BSE) detector was used to characterize the micro-
structures observed from the normal direction of the disks. The line
intercept method was applied to the SEM-BSE images to determine the
mean grain sizes.
Tensile tests were conducted on a universal tensile test machine
(SHIMADZU, AG-100 kN Xplus) at a quasi-static strain rate of 8.3 ×10
4
s
1
at RT. Tensile test specimens with a gauge dimension of 2 mm ×1
mm ×1 mm were cut from the annealed disks. A CCD video camera
extensometer (SVS625MFCP) was used throughout the tensile tests to
measure the displacement of the gauge section accurately, and the strain
was calculated by the digital image correlation (DIC) technique using
Vic-2D software (Correlated solutions co.) [7,8]. Our previous studies
[9,10] on FCC high/medium entropy alloys have already veried that
the mechanical properties (such as yield strength and Youngs modulus)
obtained by this method using the small-sized specimens are equivalent
to those measured by using standard-sized specimens [11,12]. Tensile
tests were repeated three times for each alloy to ensure the statistical
reliability.
Fig. 1 shows representative SEM-BSE micrographs of the (a)
HfNbTaTi, (b) HfNbTiZr, (c) HfNbTi, (d) HfTaTi, and (e) NbTiZr RMEA
after the HPT and subsequent annealing process. All the alloys showed
single-phase solid solutions of BCC structure. Their mean grain sizes
were ranging from ultra-ne grain (UFG) sizes (e.g., 0.16
μ
m in the
HfNbTiZr RMEA) to ne grain (e.g., 2.4
μ
m in the HfNbTiZr RMEA) and
coarse grain sizes (e.g., 33
μ
m in the HfNbTiZr RMEA). Interestingly,
UFG microstructures, which have been rarely reported in BCC metals
and alloys, could be achieved in all the RMEAs in our study. (See the
supplementals for discussions on the grain growth kinetics.)
Fig. 2 shows tensile nominal stress - nominal strain curves of the (a)
HfNbTaTi, (b) HfNbTiZr, (c) HfNbTi, (d) HfTaTi, and (e) NbTiZr RMEAs
with wide ranges of average grain sizes. It was found that all the RMEAs
showed similar mechanical properties, such as high yield strength (>
700 MPa), limited uniform elongation (<2%), and large total elongation
(2040%). Usually, yield strengths of materials increase, and ductility
decreases with decreasing their average grain sizes. However, as shown
in Fig. 2, changes in the yield strength of each RMEA with varying mean
grain sizes were found to be abnormally small, suggesting that the grain
renement strengthening effect in the RMEAs was signicantly weak
compared with conventional materials.
Grain renement strengthening in polycrystalline metallic materials
is generally described by the Hall-Petch relationship [13,14]:
σ
Y=
σ
0+kHPd1
2,(1)
where
σ
Y is the yield strength,
σ
0 is the friction stress, kHP is the constant
so-called HallPetch slope, and d is the mean grain size of materials. The
σ
0 values can be interpreted as the stress necessary for dislocations to
overcome the energy barrier of the crystal lattice determined from the
Peierls potential and interaction with solutes, for initiating slips on a
large scale with the assistance of thermal activation.
In order to quantitively characterize the grain renement strength-
ening effect of the RMEAs, the yield stress of the (a) HfNbTaTi, (b)
HfNbTiZr, (c) HfNbTi, (d) HfTaTi, and (e) NbTiZr RMEAs were plotted
as a function of reciprocal square roots of the mean grain sizes in Fig. 3.
Note that the yield stress of the materials was determined as the 0.2%
proof stress. The HallPetch relationships of each RMEA were obtained
by tting the data using Eq. (1), and the equations are indicated in Fig. 3.
The obtained
σ
0 and kHP of the RMEAs are summarized in Table 1
together with other parameters.
Fig. 4(a) plots the relationship between the experimentally-
measured
σ
0 and kHP of the HfNbTaTi, HfNbTiZr, HfNbTi, HfTaTi, and
NbTiZr RMEAs together with reference values of pure BCC, hexagonal
close-packed (HCP) metals, and other BCC alloys. The
σ
0 of the
HfNbTaTiZr RHEA and all the RMEAs obtained here were much higher
than those of most other materials but comparable to those of pure W
and some beta-Ti alloys with a high concentration of alloying elements
such as Ti
5
Al
5
Mo
5
V
1
Cr
1
Fe
1
, as well as those of RMEAs homogenized at
high temperature shown in our previous paper [6]. At the same time, it
was found that all the RMEAs exhibited lower kHP than other materials
except for interstitial free (IF) steel and pure Ti. To reveal the origin of
the low kHP (i.e., weak grain renement strengthening) in the RMEAs,
the relationship between the kHP and Youngs modulus (E) of the
RMEAs, pure BCC, HCP, face-centered cubic (FCC) metals, and FCC
Fig. 1. Representative SEM-BSE micrographs of the HfNbTaTi RMEA annealed at (a1) 1050
C for 30 s, (a2) 1200 C for 30 s, and (a3) 1250 C for 120 s, HfNbTiZr
RMEA annealed at (b1) 750 C for 30 s, (b2) 900 C for 30 s, and (b3) 1200 C for 30 s, HfNbTi RMEA annealed at (c1) 950 C for 30 s, (c2) 1050 C for 30 s, and (c3)
1250 C for 30 s, HfTaTi RMEA annealed at (d1) 1000 C for 30 s, (d2) 1150 C for 30 s, and (d3) 1200 C for 60 s, and NbTiZr RMEA annealed at (e1) 750 C for 30 s,
(e2) 850 C for 40 s, and (e3) 1200 C for 30 s, all after the HPT deformation. Their mean grain sizes (d), determined by the intercept method, are indicated in
each gure.
Q. He et al.
Scripta Materialia 231 (2023) 115442
3
HEAs / MEAs [5,10,17] are plotted in Fig. 4(b). Note that the E of the
RMEAs were determined based on the slope in the elastic stage in the
stress-strain curves shown in Fig. 2. It was found that the kHP and E
roughly showed a positive correlation. The reasons can be understood
based on the equations of kHP in several grain renement strengthening
models, summarized in Table S1 in the supplementals. In all the models,
it is obvious that the kHP is a function of the shear modulus (G) and E.
Hence, it could be concluded that the weak grain renement strength-
ening (i.e., low kHP) in the RMEAs was mainly attributed to their low E.
In BCC metals and alloys, their plastic deformation is generally
controlled by the motion of screw dislocations rather than edge dislo-
cations, particularly at RT and low temperatures, due to the non-planar
core structures [28,29]. For evaluating the solid solution strengthening
in conventional BCC dilute alloys, the classical Suzuki model [30] has
been successfully used to simulate solute-dislocation interactions at the
core of a
2[111] screw dislocations. However, the core structure can vary
along the dislocation lines depending on the local chemical environment
in RHEAs/RMEAs, since different elements are distributed in their
crystal lattices [31]. Accordingly, the nucleation and migration barriers
of kinks can vary locally due to the modulation of the local core struc-
tures. Thus, it is not appropriate to simply apply the Suzuki model to
RHEAs/RMEAs [16,32,33]. Recently, some researchers have pointed out
that cross kinks (kinks extended on different slip planes) can sponta-
neously form in screw dislocations of BCC HEAs and MEAs depending on
the local chemical environment. Obviously, the energy barriers for the
glide motion of such cross-kinked screw dislocations in BCC HEAs could
be signicantly higher than those in conventional BCC metals and al-
loys. Maresca et al. [16] proposed a theoretical strengthening model
considering such cross-kinked screw dislocations based on some pa-
rameters extracted from atomistic simulations. However, quantitative
comparison between this theoretical model and our experimental data is
not easy technically.
On the other hand, several analytical solid solution strengthening
models applicable to BCC HEAs and MEAs, such as Toda-Caraballo
model [34] and Maresca-Curtin model [16], have been proposed,
assuming that their plastic deformation is controlled by edge disloca-
tions similar to FCC alloys. The Maresca-Curtin model [16] gives an
analytical expression of the friction stress,
σ
0(T,˙
ε
), at a nite
Fig. 2. Tensile nominal stress (s) nominal strain (e) curves of the (a) HfNbTaTi, (b) HfNbTiZr, (c) HfNbTi, (d) HfTaTi, and (e) NbTiZr RMEAs with various average
grain sizes obtained from the tensile tests at room temperature. Mean grain sizes of the specimens (d) are also indicated in each graph.
Q. He et al.
Scripta Materialia 231 (2023) 115442
4
temperature, T, and strain rate, ˙
ε
, assuming that the elastic interaction
between edge dislocations and lattice distortion owing to the difference
in atomic sizes of solute atoms is dominant. The expressions are as
follows:
σ
0(0 K) = 0.040M
α
1
3
μ
1+
ν
1
ν
4
3ncnΔV2
n
b62
3
,(2)
Fig. 3. Hall-Petch relationships of the (a) HfNbTaTi, (b) HfNbTiZr, (c) HfNbTi, (d) HfTaTi, and (e) NbTiZr RMEAs. Equations of the HallPetch relationships, which
were achieved by tting the data points based on Eq. (1), are given as insets in each graph.
Table 1
List of materials parameters of the RMEAs in this study. The lattice constants and Burgers vectors were calculated by the Vegards law [15]. The grain size range of the
materials used in the present study is also given. Youngs modulus values were obtained from the elastic stages in the tensile stress strain curves shown in Fig. 2. The
HallPetch slopes and the experimental friction stresses were determined from the Hall-Petch relationships shown in Fig. 3, and the theoretical friction stresses were
calculated by the Maresca-Curtin model [16].
Alloys Lattice constant
(Å)
Burgers vector
(Å)
Grain size range
(
μ
m)
Youngs modulus
(GPa)
Hall-Petch slope
(MPa
μ
m
1/2
)
Experimental friction stress
(MPa)
Theoretical friction stress
(MPa)
HfNbTaTi 3.363 2.912 0.3449.4 100 197 797 758
HfNbTiZr 3.449 2.987 0.16033 66.7 99 751 683
HfNbTi 3.363 2.912 0.51219.2 77.4 151 680 623
HfTaTi 3.363 2.913 0.78812.1 69.6 202 752 511
NbTiZr 3.373 2.921 0.20821.6 66.7 155 711 618
Q. He et al.
Scripta Materialia 231 (2023) 115442
5
ΔEb=2.00
α
1
3
μ
b31+
ν
1
ν
2
3ncnΔV2
n
b61
3
,(3)
σ
0(T,˙
ε
) =
σ
0(0 K)1kT
ΔEb
ln
˙
ε
0
˙
ε
2
3,(4)
where
σ
0(0K)is the friction stress for the glide motion of edge dislo-
cations at 0 K, M is the Taylor factor (=2.733 for BCC metals with
random texture),
α
is a dislocation line tension constant (=0.0833 in
RHEAs/RMEAs),
μ
and
ν
are the shear modulus and the Poisson ratio,
respectively, n is the number of constituent elements, cn (n =1, 2, , N)
is the concentration of constituent elements in an N-component alloy, b
is the magnitude of the Burgers vector of the alloy, ΔEb is the predicted
energy barrier for dislocation glide, k is the Boltzmann constant, and ˙
ε
0
is the so-called reference strain rate (=10
4
s
1
). ΔVn is the mist vol-
ume, which is the degree of lattice distortion originating from the dif-
ferences in atomic sizes of alloying elements. The mist volume for the
type-n solute in an N-component alloy can be expressed as:
ΔVn=
Valloy
cn
N
m
cm
Valloy
cm
,(5)
where Valloy is the average atomic volume dened as Valloy =a3/4 (a:
lattice constant). In the present study, ΔVn of various BCC alloys were
calculated based on Eq. (5), assuming that the values of Valloy follow
Vegards law, similar to previous studies [10]. Theoretical values of the
σ
0 in the alloys were calculated by Eqs. (2) - (4) by using materials
constants extracted from literature [3541].
Fig. 5 shows a comparison between the experimental
σ
0 acquired
from the HallPetch relationships and the theoretical values of
σ
0
calculated by the Maresca-Curtin model, of which values are also sum-
marized in Table 1. It was found in Fig. 5 that in the highly-concentrated
BCC alloys (RHEAs, RMEAs, and binary alloys with a solute content of
more than 25 at.%) the experimental and calculated values are in good
agreement with each other. In contrast, the data points for dilute binary
and ternary BCC alloys largely deviate from the predictions (diagonal
broken line). To bridge between the experimental and theoretical
calculation results for different alloys, we will discuss based on the
difference in the temperature dependence of the strength contribution of
screw and edge dislocations interacting solute atoms.
In any BCC alloy,
σ
0 at 0 K is determined by the contribution of screw
dislocations because they have a much higher critical resolved shear
stress (higher
σ
0(0K)) than that for edge dislocations. At nite tem-
peratures (T),
σ
0 of both screw and edge dislocations decreases because
of thermal activation, and the reduction rate of
σ
0 can be calculated by
differentiating Eq. (4) by T, as shown below.
∂σ
0
T˙
ε
=
σ
0(0 K)k
ΔEb
ln
˙
ε
0
˙
ε
2
3
T1
3(6)
The higher
σ
0(0K)in screw dislocations results in a larger reduction in
the contribution to
σ
0, compared to the case of edge dislocations with
lower
σ
0(0K)and smaller reduction rates. Accordingly, at high tem-
peratures, there can be a crossover of contributions between screw and
edge dislocations, and edge dislocations become dominant instead of
screw dislocations. Particularly in highly-concentrated BCC alloys,
σ
0(0K)for screw dislocations can be signicantly higher than that in
dilute alloys owing to the formation of cross-kinks as described above.
Thus, the contribution of screw dislocations to
σ
0 drops very quickly
with increasing temperatures, and there can be a crossover of the con-
tributions (edge dislocations become dominant) even at near RT. Our
systematic comparison between experimental and theoretical values of
the
σ
0 in various BCC alloys shown in Fig. 5 suggests that, at RT, the
glide motion of screw dislocations controls the process of plastic
deformation in pure BCC metals and dilute BCC alloys, while the
contribution of edge dislocations can become comparable or dominant
in high-alloy systems, such as RHEAs/RMEAs. Similar ideas have been
veried in recent studies [16,32,33,4244]. For example, Han et al.
observed dislocation structures in HfNbTiZr RMEA (one of the present
materials) deformed by nanoindentation at RT and found mixed char-
acter of both screw and edge suggesting that the resistances for screw
and edge dislocations were comparable [42]. Detailed observation of
dislocation structures in the RMEAs is our ongoing work and will be
presented elsewhere in the future. Above all, it can be concluded that
elastic interaction between edge dislocations and alloying elements
having different sizes play an essential role in the solid solution
strengthening of RHEAs/RMEAs at RT. We believe this is one of the
essential characteristics of high-alloy BCC solid solutions.
In conclusion, the present work claried the characteristic
strengthening mechanisms in RHEAs/RMEAs with BCC single phase in
Fig. 4. (a) Relationships between the friction stresses,
σ
0, and the HallPetch
slopes, kHP, of the present BCC RHEAs/RMEAs, pure BCC [17], HCP metals
[17], IF steel [18], ferrite steel [19], and other BCC alloys [2022] represented
by lled red pentagrams, hollow orange triangles, hollow black triangles, and
blue triangles, respectively. (b) Relationship between the HallPetch slopes,
kHP, and Youngs modulus, E, of the present RMEAs, pure BCC, HCP, FCC metals
[17,2327], and FCC HEAs/MEAs [10] represented by lled red pentagrams,
hollow green triangles, hollow blue triangles, hollow black triangles, and hol-
low orange triangles, respectively.
Q. He et al.
Scripta Materialia 231 (2023) 115442
6
terms of grain renement strengthening and solid solution strength-
ening effects. The small HallPetch slopes (i.e., weak grain renement
strengthening effect) in the RMEAs could be attributed to their low
elastic modulus. The experimental values of the friction stresses in the
RMEAs were compared with the Maresca-Curtin model. In contrast to
conventional dilute BCC alloys, it was suggested that edge dislocations
could largely contribute to the high strength of BCC RHEAs/RMEAs at
RT due to solute-edge dislocation interactions, as well as screw
dislocations.
Declaration of Competing Interest
The authors declare that they have no known competing nancial
interests or personal relationships that could have appeared to inuence
the work reported in this paper
Acknowledgments
This work was supported by the Elements Strategy Initiative for
Structural Materials (ESISM, No. JPMXP0112101000), JST CREST (No.
JPMJCR1994), the Grant-in-Aid for Scientic Research on Innovative
Area "High Entropy Alloys" (No. JP18H05455), the Grant-in-Aid for
Early-Career Scientists (No. JP22K14501), the Grant-in-Aid for Research
Activity Start-up (No. JP21K20487), and the Grant-in-Aid for JSPS
Research Fellow (No. JP18J20766), all through the Ministry of Educa-
tion, Culture, Sports, Science and Technology (MEXT), Japan. Q.H. was
nancially supported by China Scholarship Council (CSC) for studying
in Kyoto University. All the supports are greatly appreciated.
Supplementary materials
Supplementary material associated with this article can be found, in
the online version, at doi:10.1016/j.scriptamat.2023.115442.
References
[1] O.N. Senkov, D.B. Miracle, K.J. Chaput, J.P. Couzinie, Development and
exploration of refractory high entropy alloys - a review, J. Mater. Res. 33 (2018)
30923128, https://doi.org/10.1557/jmr.2018.153.
[2] D.B. Miracle, O.N. Senkov, A critical review of high entropy alloys and related
concepts, Acta Mater. 122 (2017) 448511, https://doi.org/10.1016/j.
actamat.2016.08.081.
[3] O.N. Senkov, S.V. Senkova, D.B. Miracle, C. Woodward, Mechanical properties of
low-density, refractory multi-principal element alloys of the Cr-Nb-Ti-V-Zr system,
Mater. Sci. Eng. A 565 (2013) 5162, https://doi.org/10.1016/j.
msea.2012.12.018.
[4] O.N. Senkov, G.B. Wilks, J.M. Scott, D.B. Miracle, Mechanical properties of
Nb25Mo25Ta 25W25 and V20Nb20Mo 20Ta20W20 refractory high entropy alloys,
Intermetallics 19 (2011) 698706, https://doi.org/10.1016/j.
intermet.2011.01.004.
[5] C.C. Juan, M.H. Tsai, C.W. Tsai, W.L. Hsu, C.M. Lin, S.K. Chen, S.J. Lin, J.W. Yeh,
Simultaneously increasing the strength and ductility of a refractory high-entropy
alloy via grain rening, Mater. Lett. 184 (2016) 200203, https://doi.org/
10.1016/j.matlet.2016.08.060.
[6] Q. He, S. Yoshida, H. Yasuda, N. Tsuji, Effect of elemental combination on
microstructure and mechanical properties of quaternary refractory medium
entropy alloys, Mater. Trans. 61 (2020) 577586, https://doi.org/10.2320/
matertrans.MT-MK2019003.
[7] A. Weidner, H. Biermann, Review on strain localization phenomena studied by
high-resolution digital image correlation, Adv. Eng. Mater. 23 (2021), 2001409,
https://doi.org/10.1002/adem.202001409.
[8] H. Schreier, J.J. Orteu, M.A. Sutton, Image correlation for shape, motion and
deformation measurements: basic concepts, theory and applications, 2009. doi:
10.1007/978-0-387-78747-3.
[9] S. Yoshida, T. Bhattacharjee, Y. Bai, N. Tsuji, Friction stress and Hall-Petch
relationship in CoCrNi equi-atomic medium entropy alloy processed by severe
plastic deformation and subsequent annealing, Scr. Mater. 134 (2017) 3336,
https://doi.org/10.1016/j.scriptamat.2017.02.042.
[10] S. Yoshida, T. Ikeuchi, T. Bhattacharjee, Y. Bai, A. Shibata, N. Tsuji, Effect of
elemental combination on friction stress and Hall-Petch relationship in face-
centered cubic high /medium entropy alloys, Acta Mater. 171 (2019) 201215,
https://doi.org/10.1016/j.actamat.2019.04.017.
[11] Z. Wu, H. Bei, G.M. Pharr, E.P. George, Temperature dependence of the mechanical
properties of equiatomic solid solution alloys with face-centered cubic crystal
structures, Acta Mater. 81 (2014) 428441, https://doi.org/10.1016/j.
actamat.2014.08.026.
[12] Z. Wu, Y. Gao, H. Bei, Thermal activation mechanisms and Labusch-type
strengthening analysis for a family of high-entropy and equiatomic solid-solution
alloys, Acta Mater. 120 (2016) 108119, https://doi.org/10.1016/j.
actamat.2016.08.047.
Fig. 5. Comparison between the experimental values of the friction stress and
theoretical values of the friction stress determined by the Maresca-Curtin
model in various BCC metals and alloy. The present materials (the RMEAs)
are represented by lled red pentagrams. Other RHEAs, RMEAs, and binary
alloys with solute contents of more than 25 at% were represented by lled
blue circles. Pure BCC metals are represented by hollow orange circles. Dilute
BCC alloys such as Fe-xCr [35], Ti-xMo [38], Ti-xMn [37], Ti-xTa [40], and
Ti-xNb-yMn [36] are represented by hollow black, violet, green, wine, and
pink open circles, respectively. (The values of x and y indicate the atomic
fraction of the solute elements in alloys.) The experimental data and calcu-
lated data were equal on the diagonal black broken line.
Q. He et al.
Scripta Materialia 231 (2023) 115442
7
[13] E.O. Hall, The deformation and ageing of mild steel: III Discussion of results, Proc.
Phys. Soc. Sect. B 64 (1951) 747753, https://doi.org/10.1088/0370-1301/64/9/
303.
[14] N.J. Petch, The cleavage strength of polycrystals, J. Iron Steel InstInst. 174 (1953)
2528.
[15] O.N. Senkov, D.B. Miracle, Effect of the atomic size distribution on glass forming
ability of amorphous metallic alloys, Mater. Res. Bull. 36 (2001) 21832198,
https://doi.org/10.1016/S0025-5408(01)00715-2.
[16] F. Maresca, W.A. Curtin, Mechanistic origin of high strength in refractory BCC high
entropy alloys up to 1900 K, 2020. doi:10.1016/j.actamat.2019.10.015.
[17] Z.C. Cordero, B.E. Knight, C.A. Schuh, Six decades of the HallPetch effect a
survey of grain-size strengthening studies on pure metals, Int. Mater. Rev. 61
(2016) 495512, https://doi.org/10.1080/09506608.2016.1191808.
[18] K. Takeda, N. Nakada, T. Tsuchiyama, S. Takaki, Effect of interstitial elements on
Hall-Petch coefcient of ferritic iron, ISIJ Int. 48 (2008) 11221125, https://doi.
org/10.2355/isijinternational.48.1122.
[19] S. Takaki, Review on the Hall-Petch relation in ferritic steel, Mater. Sci. Forum.
654656 (2010) 1116, https://doi.org/10.4028/www.scientic.net/MSF.654-
656.11.
[20] E.V. Naydenkin, I.V. Ratochka, O.N. Lykova, I.P. Mishin, Evolution of the structural
phase state, deformation behavior, and fracture of ultrane-grained near-β
titanium alloy after annealing, J. Mater. Sci. 55 (2020) 92379244, https://doi.
org/10.1007/s10853-020-04468-y.
[21] J. Il Kim, H.Y. Kim, T. Inamura, H. Hosoda, S. Miyazaki, Effect of annealing
temperature on microstructure and shape memory characteristics of Ti-22Nb-6Zr
(at%) biomedical alloy, Mater. Trans. 47 (2006) 505512, https://doi.org/
10.2320/matertrans.47.505.
[22] W.L. Wang, X.L. Wang, W. Mei, J. Sun, Role of grain size in tensile behavior in
twinning-induced plasticity β Ti-20V-2Nb-2Zr alloy, Mater. Charact. 120 (2016)
263267, https://doi.org/10.1016/j.matchar.2016.09.016.
[23] A.M. James, M.P. Lord, Macmillans Chemical and Physical Data, Macmillan,
London, 1992.
[24] G.W.C. Kaye, T.H. Laby, Tables of Physical and Chemical Constants, Longman,
London, 1993.
[25] G.V. Samsonov, Handbook of the Physicochemical Properties of the Elements, IFI-
Plenum, New York, 1968.
[26] D.R. Lide, Chemical Rubber Company Handbook of Chemistry and Physics, CRC
Press, Florida, 1998.
[27] H. Ellis, Nufeld Advanced Science Book of Data, Longman, London, 1972.
[28] D.R. Trinkle, C. Woodward, Erratum: the chemistry of deformation: how solutes
soften pure metals, Science 311 (2006) 177, https://doi.org/10.1126/
science.311.5758.177.
[29] D. Rodney, J. Bonneville, Dislocations. Physical Metallurgy Fifth Edition, Elsevier,
2014, pp. 15911680, https://doi.org/10.1016/B978-0-444-53770-6.00016-2.
[30] A.S. Parasnis, Dislocations in solids, Acta Crystallogr. Sect. A Found. Crystallogr.
45 (1989) 499500, https://doi.org/10.1107/s0108767388014606.
[31] B. Akdim, C. Woodward, S. Rao, E. Antillon, Predicting core structure variations
and spontaneous partial kink formation for ½<111>screw dislocations in three
BCC NbTiZr alloys, Scr. Mater. 199 (2021), 113834, https://doi.org/10.1016/j.
scriptamat.2021.113834.
[32] B. Chen, S. Li, H. Zong, X. Ding, J. Sun, E. Ma, Unusual activated processes
controlling dislocation motion in body-centered-cubic high-entropy alloys, Proc.
Natl. Acad. Sci. U.S.A. 117 (2020) 16199161206, https://doi.org/10.1073/
pnas.1919136117.
[33] C. Lee, F. Maresca, R. Feng, Y. Chou, T. Ungar, M. Widom, K. An, J.D. Poplawsky,
Y.-.C. Chou, P.K. Liaw, W.A. Curtin, Strength can be controlled by edge dislocations
in refractory high-entropy alloys, Nat. Commun. 12 (2021) 613, https://doi.org/
10.1038/s41467-021-25807-w.
[34] I. Toda-Caraballo, P.E.J. Rivera-Díaz-Del-Castillo, Modelling solid solution
hardening in high entropy alloys, Acta Mater. 85 (2015) 1423, https://doi.org/
10.1016/j.actamat.2014.11.014.
[35] F Bergner, M Hern´
andez-Mayoral, C Heintze, MJ Konstantinovi´
c, L Malerba,
C Pareige, TEM Observation of Loops Decorating Dislocations and Resulting Source
Hardening of Neutron-Irradiated Fe-Cr Alloys, Metals 10 (1) (2020) 147.
10.3390/met10010147.
[36] Z. Chen, Y. Liu, H. Jiang, M. Liu, C.H. Wang, G.H. Cao, Microstructures and
mechanical properties of Mn modied, Ti-Nb-based alloys, J. Alloys Compd. 723
(2017) 10911097, https://doi.org/10.1016/j.jallcom.2017.06.311.
[37] P.F. Santos, M. Niinomi, H. Liu, K. Cho, M. Nakai, Y. Itoh, T. Narushima, M. Ikeda,
Fabrication of low-cost beta-type Ti-Mn alloys for biomedical applications by metal
injection molding process and their mechanical properties, J. Mech. Behav.
Biomed. Mater. 59 (2016) 497507, https://doi.org/10.1016/j.
jmbbm.2016.02.035.
[38] Y. Takemoto, I. Shimizu, A. Sakakibara, M. Hida, Y. Mantani, Tensile behavior and
cold workability of Ti-Mo alloys, Mater. Trans. 45 (2004) 15711576, https://doi.
org/10.2320/matertrans.45.1571.
[39] Q. Wang, C. Han, T. Choma, Q. Wei, C. Yan, B. Song, Y. Shi, Effect of Nb content on
microstructure, property and in vitro apatite-forming capability of Ti-Nb alloys
fabricated via selective laser melting, Mater. Des. 126 (2017) 268277, https://doi.
org/10.1016/j.matdes.2017.04.026.
[40] Y.L. Zhou, M. Niinomi, Ti-25Ta alloy with the best mechanical compatibility in Ti-
Ta alloys for biomedical applications, Mater. Sci. Eng. C 29 (2009) 10611065,
https://doi.org/10.1016/j.msec.2008.09.012.
[41] J.M. Rosenberg, H.R. Piehler, Calculation of the taylor factor and lattice rotations
for bcc metals deforming by pencil glide, Metall. Trans. 2 (1971) 257259, https://
doi.org/10.1007/BF02662666.
[42] W.H.YZ Yin, Y. Lu, T.P. Zhang, Nanoindentation avalanches and dislocation
structures in HfNbTiZr high entropy alloy, Scr. Mater. 227 (2023), 115312,
https://doi.org/10.1016/j.scriptamat.2023.115312.
[43] C. Lee, G. Kim, Y. Chou, B.L. Musico, M.C. Gao, K. An, G. Song, Y.-.C. Chou,
V. Keppens, W. Chen, P.K. Liaw, Temperature dependence of elastic and plastic
deformation behavior of a refractory high-entropy alloy, Sci. Adv. 6 (2020)
eaaz4748.
[44] S. Yin, Y. Zuo, A. Abu-odeh, H. Zheng, S.P. Ong, M. Asta, R.O. Ritchie, Atomistic
simulations of dislocation mobility in refractory high-entropy alloys and the effect
of chemical short-range order, Nat. Commun. (2021) 114, https://doi.org/
10.1038/s41467-021-25134-0.
Q. He et al.
... Therefore, an analysis of all publications on these two compositional systems of RHEAs, including recent ones [19][20][21][22][23][24], shows that neither experimental data [19][20][21][22] nor theoretical predictions [17,18,24] and machine learning [23] have been instrumental in discovering a new method for producing multi-phased RHEAs through severe plastic deformation at room temperature. ...
... Therefore, an analysis of all publications on these two compositional systems of RHEAs, including recent ones [19][20][21][22][23][24], shows that neither experimental data [19][20][21][22] nor theoretical predictions [17,18,24] and machine learning [23] have been instrumental in discovering a new method for producing multi-phased RHEAs through severe plastic deformation at room temperature. ...
Article
Full-text available
For the first time, the refractory high-entropy alloys with equiatomic compositions, HfNbTaTiZr and HfNbTiZr, were synthesized directly from a blend of elemental powders through ten revolutions of high-pressure torsion (HPT) at room temperature. This method has demonstrated its effectiveness and simplicity not only in producing solid bulk materials but also in manufacturing refractory high-entropy alloys (RHEAs). Unlike the melting route, which typically results in predominantly single BCC phase alloys, both systems formed new three-phase alloys. These phases were defined as the Zr-based hcp1 phase, the α-Ti-based hcp2 phase, and the Nb-based bcc phase. The volume fraction of the phases was dependent on the accumulated plastic strain. The thermal stability of the phases was studied by annealing samples at 500 °C for one hour, which resulted in the formation of a mixed structure consisting of the new two hexagonal and cubic phases.
... The unusually large Hall-Petch coefficient is due to the additional coherency stress caused by Ni-Al nano-clustering. In addition, it was reported that the stacking fault energy [23] or elastic modulus [24] of alloys dominated the variation in the k y values of specific alloy systems. ...
... The large k y value of Mo 0.475 HEA cannot be completely ascribed to the large LFSS. In addition to the LFSS, the stacking fault energy [23] and shear modulus [24] of alloys are other two key factors determining the magnitude of Hall-Petch coefficient. It is expressed as below: ...
Article
Full-text available
The Hall-Petch strengthening coefficient of face-centered cubic alloys has the potential to increase. Herein, we experimentally determined an unconventionally large Hall-Petch coefficient equal to 1100 MPa·µm1/2 and a large lattice friction stress for novel Co0.95Cr0.8Fe0.25Ni1.8Mo0.475 high-entropy alloys (Mo0.475 HEAs). Detailed microstructural characterizations showed Mo segregation at grain boundaries (GBs) and no apparent nano-clustering in the matrix. With the increased Mo content, Mo segregating at GBs is an unexpected outcome. The unconventionally large Hall-Petch coefficient is ascribed to the newly generated effect due to Mo segregation at GBs, besides the factors in the grain interior, e.g. the increased solid-solution strengthening. This study is dedicated to enriching the diversity of Hall-Petch strengthening mechanisms for the development of high-strength materials.
... Let us now spend some time in discussing such literatures. Consistently with the predictions of C(x; 300K)/Q 0 (N spe ), all the members of the first group have been reported to grow as a single phase in either as-cast, homogenized or deposited conditions (46)(47)(48)(49)(50)(51). For instance, a comprehensive recent work from He et al. (46) has shown that no phase transformation can be found for TiZrNb, Ti-HfNb, TiHfTa, TiZrHfNb and TiHfTaNb, even after a series of heat treatments, including firstly homogenization at 1150 • C for 24 hours, then heavily deformation by high-pressure torsion, and lastly annealing at various temperatures in the 750-1250 • C range. ...
... Consistently with the predictions of C(x; 300K)/Q 0 (N spe ), all the members of the first group have been reported to grow as a single phase in either as-cast, homogenized or deposited conditions (46)(47)(48)(49)(50)(51). For instance, a comprehensive recent work from He et al. (46) has shown that no phase transformation can be found for TiZrNb, Ti-HfNb, TiHfTa, TiZrHfNb and TiHfTaNb, even after a series of heat treatments, including firstly homogenization at 1150 • C for 24 hours, then heavily deformation by high-pressure torsion, and lastly annealing at various temperatures in the 750-1250 • C range. These experimental results confirm the robust stability of such compositions. ...
Preprint
Full-text available
Understanding the thermodynamics of multi-principal-element alloys (MPEAs), although crucial for their design, remains an elusive task. The configurational entropy, Sconf, is a critical thermodynamic quantity determining stability, but its calculation for real materials poses a hard computational challenge. Strong of a highly efficient cluster expansion, constructed on density functional theory data,and of an advanced sampling technique, we are able to compute Sconf over the huge configurational and compositional space of the prototype bcc Ti-Zr-Hf-Nb-Ta system. In particular, we explore around 200,000 atomic arrangements across more than 200 compositions. The main features of the computed Sconf-vs-temperature curves can be captured by the definition of two characteristic temperatures, which are then used to define a dimensionless descriptor. This, together with Sconf, enable us to rank alloys according to their (meta- )stability across a broad composition range, going from equiatomic to nonequiatomic, and even to the regions covered by conventional alloys. Such classification is informed and validated against hundreds of experimental results. Our analysis allows us to revise the classification scheme of alloys into high-entropy, medium-entropy and low-entropy and, in general, sheds light into the thermodynamic origin of their metastability, ultimately helping in the design and development.
... where Δσ f is the lattice friction stress, commonly replaced by Peierls-Nabarro stress. k coefficient of 151-202 MPa/μm 1/2 was reported for TiNbHf and TiTaHf MPEAs [64]. Since they are most like the present RHEA, k (200 MPa/μm 1/2 ) was adopted. ...
... The mechanical properties of bcc metals are traditionally understood to be dictated by the mobility of 1 2 111 screw dislocations [6], although recent work has also highlighted a role of edge dislocations in certain concentrated systems [7,8], especially at elevated temperatures at which the barriers to screw dislocation motion experience significant thermal softening [7]. Deformation at ambient conditions is nonetheless expected to be influenced by screw dislocation glide, which, in pure bcc metals, involves the nucleation [9] and propagation of pairs of kinks in the dislocation line [10][11][12]. ...
Article
Dislocation-mediated deformation mechanisms in body-centered-cubic solid solutions are expected to be influenced by spatial fluctuations in screw dislocation core energies. In refractory high-entropy alloys, the formation of chemical short-range order has been demonstrated to decrease the heterogeneity of this energy landscape, narrowing the distribution of dislocation core energies. It is, however, unclear if multicomponent compositionally complex systems display any unique effects or if these results are more generally applicable. To answer this question, this study computationally investigates how system chemistry, compositional complexity, and the presence of various degrees of chemical short-range order affect the distribution of screw dislocation energies in binary and ternary subsystems of the NbMoTaW alloy.We report the calculated averages and variances for the diffuse antiphase boundary energy and the dislocation core energy with various degrees of chemical short-range order. While short-range order negligibly affects the average core energies, their distributions are notably narrowed in some, but not all, systems, primarily depending on chemistry rather than the number of components.
... Thus, the different tendency between hardness and elastic modulus might be related to the applied load during hardness measurements, while the ultrasonic testing does not involve the application of mechanical loads. This hypothesis is also in agreement with the higher precision for determination of elastic properties reported by dynamic methods compared to mechanical tests [16,20,[22][23][24] and b) elastic modulus [20,[25][26][27][28][29] of the TNZT alloys of this study (blue and green dashed lines) with other RHEAs from the literature. ...
Article
Full-text available
Energy efficiency is motivating the search for new high-temperature (high-T) metals. Some new body-centered-cubic (BCC) random multicomponent “high-entropy alloys (HEAs)” based on refractory elements (Cr-Mo-Nb-Ta-V-W-Hf-Ti-Zr) possess exceptional strengths at high temperatures but the physical origins of this outstanding behavior are not known. Here we show, using integrated in-situ neutron-diffraction (ND), high-resolution transmission electron microscopy (HRTEM), and recent theory, that the high strength and strength retention of a NbTaTiV alloy and a high-strength/low-density CrMoNbV alloy are attributable to edge dislocations. This finding is surprising because plastic flows in BCC elemental metals and dilute alloys are generally controlled by screw dislocations. We use the insight and theory to perform a computationally-guided search over 107 BCC HEAs and identify over 106 possible ultra-strong high-T alloy compositions for future exploration. The strength in BCC high-entropy alloys is associated with the type of mobile dislocations. Here the authors demonstrate by means of an ample array of experimental techniques that edge dislocations can control the strength of BCC high-entropy alloys.
Article
Full-text available
Refractory high-entropy alloys (RHEAs) are designed for high elevated-temperature strength, with both edge and screw dislocations playing an important role for plastic deformation. However, they can also display a significant energetic driving force for chemical short-range ordering (SRO). Here, we investigate mechanisms underlying the mobilities of screw and edge dislocations in the body-centered cubic MoNbTaW RHEA over a wide temperature range using extensive molecular dynamics simulations based on a highly-accurate machine-learning interatomic potential. Further, we specifically evaluate how these mechanisms are affected by the presence of SRO. The mobility of edge dislocations is found to be enhanced by the presence of SRO, whereas the rate of double-kink nucleation in the motion of screw dislocations is reduced, although this influence of SRO appears to be attenuated at increasing temperature. Independent of the presence of SRO, a cross-slip locking mechanism is observed for the motion of screws, which provides for extra strengthening for refractory high-entropy alloy system.
Article
Full-text available
Herein, a comprehensive review on the application of digital image correlation (DIC) for investigations of strain localization phenomena in the field of materials science and engineering (MSE) is provided. The Review is divided into three parts. In Part 1, the method of DIC is reviewed in terms of basic principles, correlation algorithms, and sources of errors. Part 2 is focused on the application of DIC in the field of MSE. This part is subdivided into the application of DIC in 1) combination with optical microscopy and 2) in combination with scanning electron microscopy (SEM). In Part 3, results of high‐resolution DIC obtained in combination with SEM are presented for high‐strength austenitic stainless steels—transformation‐induced plasticity steels and twinning‐induced plasticity steels—exhibiting microscopic strain localization phenomena. These investigations are supplemented by electron backscattered diffraction for interpretation of strain fields. Although DIC allows for both 2D and 3D calculations of displacement and strain fields, herein, 2D digital correlation is referred to only, as in particular with SEM only in‐plane displacements are calculated so far.
Article
Full-text available
Single-phase solid-solution refractory high-entropy alloys (HEAs) show remarkable mechanical properties, such as their high yield strength and substantial softening resistance at elevated temperatures. Hence, the in-depth study of the deformation behavior for body-centered cubic (BCC) refractory HEAs is a critical issue to explore the uncovered/unique deformation mechanisms. We have investigated the elastic and plastic deformation behaviors of a single BCC NbTaTiV refractory HEA at elevated temperatures using integrated experimental efforts and theoretical calculations. The in situ neutron diffraction results reveal a temperature-dependent elastic anisotropic deformation behavior. The single-crystal elastic moduli and macroscopic Young's, shear, and bulk moduli were determined from the in situ neutron diffraction, showing great agreement with first-principles calculations, machine learning, and resonant ultrasound spectroscopy results. Furthermore, the edge dislocation-dominant plastic deformation behaviors, which are different from conventional BCC alloys, were quantitatively described by the Williamson-Hall plot profile modeling and high-angle annular dark-field scanning transmission electron microscopy.
Article
Full-text available
Significance This work demonstrates dislocation dynamics in body-centered-cubic (BCC) high-entropy alloys (HEAs). The local composition inhomogeneity in these concentrated solutions leads to an unconventional rate-controlling mechanism that governs dislocation mobility. The activated process becomes nanoscale detrapping, of kinks on screw dislocations and of trapped segments of edge dislocations, presenting an activation barrier of similar magnitude for both dislocation types. Their sluggish mobility explains the elevated strength and strain hardening in BCC HEAs.
Article
Full-text available
In the paper, we study the evolution of the structural phase state, deformation behavior, and fracture of ultrafine-grained near-β titanium alloy Ti–5Al–5Mo–5V–1Cr–1Fe after annealing in the temperature range 773–1073 K (0.4–0.55 Tm). The duration of the strain-hardening stage under tensile conditions is shown to be almost independent of the annealing temperature. The character of the strain-softening stage is largely determined by the alloy structure formed after annealing. It is found that annealing at the temperatures 773 and 873 K does not change the deformation behavior of the ultrafine-grained alloy under tension at room temperature. Deformation and fracture of the specimens localize in shear bands. Recrystallization occurring at the annealing temperature 973 and 1073 K leads to the transition of grain boundaries to a more equilibrium state and to a sharp decrease in strength and softening rate of the titanium alloy. It also affects the neck formation prior to fracture, giving a developed neck. Fracture surfaces are indicative of ductile dimple fracture of the alloy in all states. The dimple size depends on the size of structural elements after heat treatments. Based on the experimental data, the σ0 value and Hall–Petch coefficient k are determined. In the investigated grain size range (0.17–1.25 μm) the values are found to be, respectively, 680 MPa and 0.36 MPa m1/2.
Article
Full-text available
The present work reports the effect of elemental combination on microstructure and mechanical properties of quaternary refractory medium entropy alloys (RMEAs) having equi-atomic compositions. As-cast RMEAs ((1) HfNbTaTi, (2) HfNbTaZr, (3) HfNbTiZr, (4) HfTaTiZr, and (5) NbTaTiZr) were fabricated by vacuum arc-melting of pure elements under Ar atmosphere, homogenization was then performed at 1150°C for 24 hours with Ar atmosphere. Firstly, microstructures of both as-cast and homogenized RMEAs were observed by SEM-BSE. Three kinds of microstructures consisting of annealed grains (AG), granular morphology (GM) and dendritic morphology (DM) were found to be distributing along solidification direction in the as-cast RMEAs. Inter-dendritic segregation in the as-cast RMEAs was characterized by SEM-EDX. At the same time, grain boundary precipitates were found in the as-cast (2) HfNbTaZr and (4) HfTaTiZr alloys. After homogenization at 1150°C, a fraction of AG greatly increased while that of DM largely decreased. It was also found that the degree of segregation was largely reduced after homogenization. In addition, grain boundary precipitates having equiaxed morphology and HCP structure were observed in the homogenized (2) HfNbTaZr alloy. Subsequently, tensile tests of both as-cast and homogenized RMEAs were performed at room temperature (RT) to characterize mechanical properties of the RMEAs. After homogenization, ductility of (1) HfNbTaTi, (3) HfNbTiZr, (4) HfTaTiZr, and (5) NbTaTiZr alloys was highly improved while (2) HfNbTaZr alloy still showed early brittle fracture. Better ductility of the homogenized (1) HfNbTaTi, (3) HfNbTiZr, (4) HfTaTiZr, and (5) NbTaTiZr alloys could be attributed to the elimination of inter-dendritic segregation as well as grain boundary precipitates through homogenization. Fullsize Image
Article
Strain localization phenomena are well known for numerous materials occurring under various loading conditions on different length scales. The digital image correlation (DIC) is a full‐field measuring technique which allows to study these phenomena with high lateral and strain resolution. In this review, article number 2001409 by Anja Weidner and Horst Biermann, an overview on developments and applications of DIC is provided and case studies on the application of high‐resolution DIC in combination with scanning electron microscopy are presented.
Article
Recent interest in chemically-complex solid-solution alloys has produced a number of new refractory BCC alloys with superior high-temperature properties. Preliminary atomistic studies show that, unlike simple BCC metals, these alloys produce equilibrium (screw) dislocations spread on varying glide planes along their length. This observation suggests that under load such defects produce kinks on different glide planes leading to a distribution of pinning points that significantly increases high-temperature strength. In order to validate this model a first-principle approach is developed to characterize these sub-nanoscale structures. We find significant spreading of the dislocation onto varying {110} planes (partial kinks) in NbTiZr and Nb17Ti33Zr50, while Nb50Ti33Zr17 produces a straight-compact dislocation as found in simple BCC metals. Chemical analysis around the dislocation indicate that the partial kinks form in response to increasing (decreasing) Ti and Zr (Nb) compositions. These results validate a growing body of work on understanding the hardening mechanisms in chemically-complex alloys.