You are on page 1of 34

Journal Pre-proof

Insights into orientation-dependent plasticity deformation of


HfNbTaTiZr refractory high entropy alloy: An atomistic investigation

Wei JIAN , Lu REN

PII: S0749-6419(23)00351-0
DOI: https://doi.org/10.1016/j.ijplas.2023.103867
Reference: INTPLA 103867

To appear in: International Journal of Plasticity

Received date: 16 July 2023


Revised date: 22 December 2023

Please cite this article as: Wei JIAN , Lu REN , Insights into orientation-dependent plasticity defor-
mation of HfNbTaTiZr refractory high entropy alloy: An atomistic investigation, International Journal of
Plasticity (2023), doi: https://doi.org/10.1016/j.ijplas.2023.103867

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2023 Published by Elsevier Ltd.


Highlights
 Defect-free HfNbTaTiZr HEA shows a strong correlation between orientation and
tensile property.
 Plasticity deformation of HfNbTaTiZr HEA is affected by crystallographic
orientation.
 Phase transformation under tensile loading contributes to the strengthening in
HfNbTaTiZr HEA.
Title: Insights into orientation-dependent plasticity deformation of HfNbTaTiZr
refractory high entropy alloy: An atomistic investigation

Authors: Wei JIAN1, Lu REN1,2

Authors’ affiliations:
1
School of Mechanical Engineering & Mechanics, Ningbo University, Ningbo 315211,
Zhejiang, China
2
Zhejiang Provincial Key Laboratory of Part Rolling Technology, Ningbo 315211, P.R.
China

Abstract:
HfNbTaTiZr refractory high entropy alloy has emerged as a prospective structural
material suitable for high-temperature applications owing to its remarkable
combination of high strength, good tensile ductility and excellent high-temperature
properties. However, the insufficient understanding of the mechanical responses of
this refractory high entropy alloy has hindered the improvement of performance in alloy
design and potential for engineering applications. Therefore, the orientation-
dependent tensile behaviors of HfNbTaTiZr refractory high entropy alloy are
investigated from nanoscale using molecular dynamics simulations. The simulation
results show that the mechanical responses under uniaxial tension are strongly
correlated to the crystallographic orientation with respect to the loading direction, and
the highest Young’s modulus and yield strength are achieved along [111] direction. In
addition, the deformation twinning and phase transformations are charactered. The
tensile behavior oriented in [001] direction is dominated by the BCC-FCC phase
transformation, while that oriented in [110] and [111] directions by the BCC-HCP
phase transformation. The results reveal the critical role of tensile loading direction
in generating specific crystalline microstructures under high-strain-rate loading
conditions, which can enlighten new design strategy in engineering refractory high
entropy alloys with specific orientations for extreme environments.

Keywords:
Refractory high-entropy alloys; Tensile deformation; Orientation dependence; Phase
transformation; Molecular dynamics simulations

 Corresponding author: renlu@nbu.edu.cn (Lu REN)


1. Introduction
High-entropy alloys (HEAs) are garnering significant interest due to their unique
chemical composition which features the presence of five or more metallic elements in
equiatomic or near-equiatomic proportions with high configurational entropy as well as
unusual materials properties (George et al., 2019). Such a novel alloy design strategy
has paved the way for generating new materials with exceptional properties and
functions, making HEAs highly sought-after in metallurgical research. HEAs prefer
to form simple-phase solid solutions as a result of the high configurational entropy
which contributes to the phase stability and the decrease of Gibbs energy at elevated
temperatures (Luan et al., 2020). The crystalline structures of HEAs generally include
single face-centered cubic (FCC), body-centered cubic (BCC) and hexagonal close-
packed (HCP), and the latter one is relatively rare in current research (Tracy et al., 2017).
Most FCC single-phase HEAs exhibit comparable mechanical properties to nickel-
based alloys or austenitic stainless steels, while lower ultimate tensile strength than that
of second generation advanced high strength steels (George et al., 2020). BCC HEAs
that mainly consist of refractory elements show similar mechanical properties to
transformation-induced plasticity steels and dual-phase steels (George et al., 2020).
Moreover, refractory BCC HEAs maintain superior mechanical properties at high
temperatures, which have been considered as highly promising candidates for practical
applications in hash environment with elevated operating temperatures and demanding
mechanical conditions. They can serve as high melting temperature materials that
simultaneously offer strength and damage tolerance in nuclear reactor systems, various
propulsion systems including aircraft engine components, rocketry, hypersonic vehicles,
engine turbine blades, and surface coating of metal components (El Atwani et al., 2023;
Xiong et al., 2023).

Discovering a combination between room temperature ductility and high temperature


strength is crucial as the majority of refractory HEAs exhibit limited room-temperature
ductility that are unsuitable in usage. In contrast to group VI (such as Cr, Mo and W)
metals that display room-temperature embrittlement, HfNbTaTiZr HEA as a member
of the multi-principal element family from group IV and V have raised great attention
due to its substantial room-temperature tensile ductility and plasticity (Senkov et al.,
2018). HfNbTaTiZr HEAs show not only high yield strength, large plasticity and high
wear resistance but also good bio-corrosion resistance and superior biocompatibility,
which have also been considered as potential metallic biomaterials (Yang et al., 2022).
In addition, they can be used in many diverse ambient-temperature applications,
especially for nuclear reactor systems. Among the involved compositions, Nb and Zr
have low neutron absorption cross section, and the BCC structure offers greater
resistance to irradiation swelling, which are necessary characteristics for nuclear
structural materials (King et al., 2019). While there is a considerable amount of
literature on the HfNbTaTiZr refractory HEAs, it has predominantly focused on the
formability and compression deformation. In prior studies, HfNbTaTiZr HEAs have
been found to display high strain hardening and homogeneous deformation, coupled
with high compression yield strength and ductility (Senkov et al., 2011). The
compressive properties of HfNbTaTiZr HEAs have been evaluated across the
temperature range of 296-1473 K with different strain rate increments, and the results
indicate a correlation between structure evolution observed during compression tests
and compressive properties obtained (Senkov et al., 2012). The designed HfNbTaTiZr
HEAs based on the principles of intrinsic element ductility, maximum atomic size
difference for solid solution strengthening and the valence electron concentration
criterion for ductility have demonstrated high compressive yield strength and
compressive ductility (An et al., 2021). The microstructure and tensile ductility of
HfNbTaTiZr HEAs at room temperature, 800 oC and 1200 oC under uniaxial tensile
tests have been reported recently, showing the complex microstructure evolution and
deformation behaviors which can be controlled by varying temperature ranges (Mills
et al., 2023). The Young’s and shear moduli are found to decrease from 300 K to 1100
K, showing a strong temperature dependence (Laplanche et al., 2019), and the yield
strength of HfNbTaTiZr HEA under uniaxial tensile tests decreases with the
temperature increasing from 77 K to 673 K (Chen et al., 2019b). In addition, tensile
creep behaviors of HfNbTaTiZr samples fabricated using an optical floating zone
technique have been studied, and its solute drag creep is governed by a/2<111> type
dislocations (Liu et al., 2022). Even though these experiments have reported
interesting tensile behaviors of HfNbTaTiZr HEAs, studies on the tensile properties are
relatively scarce, and there is limited information available for HfNbTaTiZr HEAs
under special tensile loading conditions. The corresponding mechanical behaviors of
HEAs are generally correlated to their crystallographic orientations, which have been
identified through experiments or simulations recently (Jha et al., 2023; Li et al., 2023a;
Zhang et al., 2023a). For HfNbTaTiZr HEAs, the in-situ tensile testing has revealed
the orientation dependent plastic localization, where the slip localization behavior with
the appearance of high density of slip bands in grains is discovered to be strongly related
with crystallographic orientations with respect to the loading direction (Charpagne et
al., 2022). In addition, the slip trace analyses have shown the coexistence of a variety
of plastic strain accommodation mechanisms (Charpagne et al., 2022). The
irregularities along the slip trace affect the macroscopic shape of the slip bands, leading
to varied mechanical performance along different crystallographic orientations.
These experimental studies have shown that the BCC refractory HEAs have a strong
orientation dependence on the deformation behaviors under loading. In addition, the
obtained microstructure of HfNbTaTiZr HEAs present preferred orientation after
annealing (Chen et al., 2019a), indicating that the crystal structure can be tuned by
adjusting processing parameters. Therefore, understanding the orientation-dependent
mechanical properties can enrich the knowledge of the structure-property-performance
relationship in HfNbTaTiZr HEAs, and provide new insights for optimization and
design of advanced BCC refractory HEAs. In light of the insufficient experimental
data and limited understanding towards the relationship between crystallographic
orientation and deformation behavior, the investigation of the effect of crystallographic
orientation on the mechanical properties of HfNbTaTiZr HEAs from the atomistic level
is carried out in this study.

Given the detailed description of material behaviors at the atomistic level and effective
prediction of material properties (Borde et al., 2023; Huang et al., 2022), molecular
dynamics (MD) simulations have proven to be an effective and reliable method to
unravel the mechanical behaviors of HfNbTaTiZr HEAs. The chemical short-range
order (SRO) in HEAs, which is an important feature in multi-principal element solid
solutions, can be well characterized in MD simulations (Wu et al., 2021). The
influence of SRO on the ultimate strength and ductility CoCuFeNiPd HEA (Chen et al.,
2021) has been evaluated based on hybrid MD and Monte Carlo simulations. The
initial deformation induced by partial dislocation slip and the continued deformation
governed by twining and perfect dislocation slip with the effect of SRO and grain size
have been discovered in CoCrNi medium entropy alloy (Gupta et al., 2022). It is
found that the twinning mechanisms include the nucleation of a second leading partial
at the grain boundary as the twinning source and the glide of two leading partials along
two different slip planes that intersect within a grain which acts as the intragranular
twin source. As to the BCC crystals, the screw dislocations glide is generally
dominant, especially the a/2<111> screw dislocations (Eleti et al., 2021). The

112 111 deformation twinning has been observed in NbTaTiZr medium entropy

alloy, while 110 110 and 111 110 have been identified as the slip systems in

micropillar experiments and nanoindentation experiments (Csanádi et al., 2019). The


mobilities of screw and edge dislocations in the BCC MoNbTaW refractory HEA across
wide temperature range has been captured with a machine-learning interatomic
potential, and it is observed that the influence of SRO diminishes as the temperature
increases (Yin et al., 2021). The interactions of the nanoscale precipitates with
dislocations which is difficult to observe directly using experimental techniques can be
described with atomistic simulations, and the coherent precipitate combined with the
severe atomic-scale lattice distortion is found to enhance the strength at cryogenic or
elevated temperatures (Li et al., 2020). The deformation-induced martensitic
transformation from FCC to BCC phases in Fe 45Co25Ni10V20 HEA at cryogenic
temperature caused by high-stress concentration at grain boundaries or intersection of
stacking faults is uncovered from MD simulations, which is in good agreement with
experiments (Li et al., 2023b). The computational aided design via MD simulations
is also available for target-oriented improvement in HEAs. By proposing the design
of custom-designed low-angle grain boundary in the nanoscale CuCoNiPdFe FCC
HEAs, the mechanical stability and plastic deformability under external loading has
been well predicted through MD simulations (Zhang et al., 2023b). Considering that
the systematic study in BCC refractory HEA system from nanoscale is still lacking, the
in-depth analysis in HfNbTaTiZr HEA towards the effect of crystallographic orientation
during tensile deformation through MD approach is desired to evaluate the mechanical
responses and explore the phase transformation, which can be helpful in tailoring
microstructure for better strength and ductility.

The objective of this study is to investigate the mechanical behaviors during tensile
deformation along different crystallographic orientations in HfNbTaTiZr HEAs using
MD simulations. The mechanical responses, phase transformation and underlying
mechanisms have been analyzed with the aim to improve the tensile strength and
ductility synergistically in refractory HEA systems. The atomistic model based on
equiatomic BCC HfNbTaTiZr structure is firstly constructed, and the uniaxial tensile
deformation with a fixed strain rate is performed along [001], [110] and [111]
crystallographic orientations. The selection of the three directions is motivated by
previous experimental results that the HfNbTaTiZr crystallites show preferred
orientation of <110> and <100> (Čížek et al., 2018), and the grains are found to
preferentially have <001> or <111> orientations parallel to the loading deformation
(Eleti et al., 2019). In addition, the cross-slip at the scale of lattice has been found to
share the common [111] direction (Charpagne et al., 2022). The tensile properties as
well as atomic arrangement and displacement under tensile loading are characterized to
demonstrate the mechanical responses of HEA systems and the effect of
crystallographic orientation. The simulation results uncover the relationship between
microstructure evolution and mechanical properties in HfNbTaTiZr HEA systems under
tensile deformation, and illustrates how the phase transformation along different
crystallographic orientations affect the corresponding strength and ductility. The
revealed mechanism can provide further insights into the design and engineering of
refractory HEA with enhanced strength-ductility synergy for practical applications.

2. Methods
The MD simulations under tensile deformation of HfNbTaTiZr systems are performed
using a large-scale atomic/molecular massively parallel simulator (LAMMPS)
(Plimpton, 1995). The equiatomic HfNbTaTiZr HEA is a BCC single-phase solid
solution with average grain size at the micrometer scale (Čížek et al., 2018; Eleti et al.,
2019; Mills et al., 2023). The phase transformation generally takes place in the single-
phase structure (Chen et al., 2019a). Therefore, a single-crystal and dislocation-free
structure is adopted to investigate the corresponding mechanical response, which
should satisfy the need to comprehensively analyze the tensile deformation mechanism
and phase transformation in MD simulations. The lattice constant is initially set as
3.40 Å, which follows the measured experimental value for HfNbTaTiZr HEA (Senkov
et al., 2011). For the [001]-oriented model, the simulation box is set by repeating 25,
25 and 50 of the unit cell in x-[100], y-[010] and z-[001] directions. Correspondingly,
the initial size is 88.8 Å × 88.8 Å × 177.5 Å containing 62,500 atoms. For the [110]-
oriented model, the simulation box is set by repeating 15, 15 and 30 of the unit cell in

x-[ 111 ], y-[ 112 ] and z-[110] directions, and the initial size is 92.2 Å × 87.0 Å × 150.6
Å containing 54,000 atoms. For the [111]-oriented model, the simulation box is set

by repeating 15, 15 and 30 of the unit cell in x-[110 ], y-[112 ] and z-[111] directions,

and the initial size is 75.3 Å × 87.0 Å × 184.5 Å containing 54,000 atoms. The
periodic boundary condition is applied for the three directions in all the systems. The
current size ensures the observation of mechanical behaviors and phase transformation,
while benefits the computational efficiency. The detailed verification is shown in the
Supplementary Information. The atoms of the selected five elements with the near-
equiatomic elemental composition constraint of 20% are randomly distributed in the
model. In order to investigate the effect of crystallographic orientations, the HEA
systems are aligned in [001], [110] and [111] directions with z axis respectively.
Hundreds of models are constructed by changing the random seed value in the setting
of random distribution, and calculating the total energy of each optimized model after
energy minimization. The constructed sample with the lowest energy and stable
structure is selected for the subsequent MD simulations. As the arc-melting technique
has been used to synthesize HfNbTaTiZr HEAs (Senkov et al., 2012; Senkov et al.,
2011), the similar preparation process is repeated in order to make the model as close
as possible to the actual situation of as-cast HEA samples. All the models are heated
up from 300 K to 2000 K at a constant rate of 0.007 K/fs, followed by the relaxation
for 1 ns, and then quenched from 2000 K to 300 K at a constant rate of 0.007 K/fs
during annealing. The models are further equilibrated for 10 ns to obtained the final
relaxed structures.

A developed modified embedded-atom method (MEAM) potential has been


successfully applied to predict material properties of HfNbTaTiZr HEA system (Huang
et al., 2021). The MEAM potential has good accuracy with density functional theory
(DFT) or experimental results in characterizing the SRO, the phase transformation and
calculating the properties including cohesive energies, lattice constants, melting points,
thermal expansion coefficients and elastic constants (Huang et al., 2021). Thus, this
developed MEAM potential is selected to describe the interatomic interactions in
HfNbTaTiZr HEA. All the HEA samples are equilibrated by energy minimization
followed by dynamic relaxation for 1 ns under an NPT ensemble with the temperature
of 300 K and the pressure of zero. The equilibrated structure of HEA samples with
three different crystallographic orientations and the corresponding pair distribution
functions are shown in Figure 1(a) and 1(b). The pairwise multicomponent SRO
parameter is used to characterize SRO for the disordered solutions, which is expressed
as  ijm = ( pijm − c j ) / ( ij − c j ) , where pijm is the probability of finding a j-type atom

near the i-type atom in shell m and c j is the concentration of j-type atom in the system;

 ij = 1 when i = j, and  ij = 0 otherwise (Fontaine, 1971). The values of SRO

parameter at the nearest neighbor (m = 1) for each pair in the three HEA systems are
calculated and shown in Figure 1(c). All the parameter values are very close to zero,
indicating that all the constructed HEA models are nearly random solid solutions
without significant local chemical order or segregation. In addition, these constructed
HEA structures are defect free without consideration of the possible oxide layer on the
surface.

After equilibration, the samples were separately stretched along z axis with different
crystallographic orientations at the constant strain rates of 1 × 108 s-1 to reveal the
microstructure evolution. The loading rate in MD simulations is within the ultra-high
strain rate region, which is much higher than that applied in conventional mechanical
experiments according to standard test methods. Such ultra-high strain rate at the
room temperature can be achieved from extreme loading conditions, such as laser-
generated shock or impact loading (Piao et al., 2016; Wilkerson, 2017; Yaakobi et al.,
2005). During tensile loading, the NPT ensemble is adopted to maintain the
temperature of 300 K and zero pressure. The stress of the three samples is calculated
by averaging the stresses of all atoms based on the Virial stress theorem. The
polyhedral template matching (PTM) method that determines structural similarity
based on root-mean-square deviation (Larsen et al., 2016) implemented in OVITO
(Stukowski, 2010) are employed to characterize the atomic structure during plastic
deformation. To characterize the local shear deformation during tensile loading, the
von Mises local shear strain invariant of each atom in the system is calculated using the
equation (Shimizu et al., 2007):

iMises =  xy2 +  yz2 +  zx2 + ( xx −  yy ) + ( xx −  zz ) + ( yy −  zz ) 


1 2 2 2

6 

where  is the component of the local Lagrangian strain matrix for each atom.

Figure 1. (a) Atomistic structures of HfNbTaTiZr HEAs with z axis along [001], [110]
and [111] crystallographic orientation with the snapshots showing different
arrangements of atoms; (b) Pair distribution functions of atoms in the HEA system; (c)

The chemical SRO parameters  ij of all species pairs in the three HEA models. The
1
values of chemical SRO parameter  ij1 are nearly zero, indicating the random mixing

of all the elements in the three systems.

3. Results and discussion


3.1 Mechanical responses during tensile deformation
The stress-strain curves under uniaxial tensile tests along [001], [110] and [111]
crystallographic orientations are shown in Figure 2(a), 2(c) and 2(e). The stress
increases linearly with the strain at the initial elastic stage for all the cases. For the
[001]-oriented sample, the deformation starts to be governed by plastic deformation at
the strain of 3.7 %, and the yield strength at the strain of 4.1 % is measured as 2.43 GPa.
After that the stress declines, and remains fluctuating at about 1.25 GPa within the strain
ranging from 15.0 % to 40.0 %. Then the stress surges with strain larger than 40%
and the ultimate strength at the strain of 47% reaches 4.34 GPa. For the [110]- and
[111]-oriented samples, the stress increases linearly at the beginning, and reaches the
yield strength of 6.69 GPa and 9.49 GPa when the strain reaches to 4.7 % and 6.7 %
respectively, followed by the plastic deform and an instant drop of stress. The stress
raises again at the strain of 7.2 % for the [110]-oriented sample, followed by a drop to
the fluctuant stress level at the strain of 16.8 %. The second peak of the stress at the
strain of 15.4% is much smaller for the [111]-oriented sample, and the stress starts to
drop at a slow rate with the increasing strain. The trends of the stress-stain curves
show a strong orientation dependency in HfNbTaTiZr system. The Young’s modulus
is calculated by fitting the stress-strain curves in the elastic region in the strain range of
around 0 to 3.0 %. The Young’s modulus for the [001]-oriented sample is calculated
as 99.7 GPa, which is close to the experimental measurements ranging from 90.8 GPa
to 112.7 GPa for HfNbTaTiZr HEAs (Fan et al., 2022; Laplanche et al., 2019;
Motallebzadeh et al., 2019). The [110]- and [111]-oriented samples show higher
Young’s modulus of 157.5 GPa and 163.6 GPa respectively, which is much higher than
the experimental values. The discrepancy between experimental and simulation
results might be due to the perfect initial single crystalline structure and the ultra-high
strain rate employed in the MD simulations. It has been pointed out that the Young’s
modulus depends on the deformation strain rate when the strain rate is above the
threshold of 5105 s-1 (Rida et al., 2020). When the strain rate exceeds the threshold
value, the material system is unable to equilibrate completely during deformation
process as the timescale in MD simulations is smaller than the typical relaxation time
of the material system. Such non-equilibrium state during ultrafast deformation has
been confirmed by the analysis of the stress relaxation in crystals (Rida et al., 2019).
In addition, the Young’s modulus of the experimental samples can be affected by other
factors such as actual microstructure or composition ratio, leading to lower value than
the simulation results. Nevertheless, it does not affect the tendency of orientation-
dependent Young’s modulus calculated in this study. Such anisotropy in the Young’s
modulus has already been reported in Nb-Ti-Zr-based alloys in previous study with a
consistent trend, which suggests the structure follows different symmetry breaking
distortion paths in tensile deformation (Tane et al., 2008; Zhang et al., 2011). In the
[110]- and [111]-oriented sample, the Young’s modulus is much higher, and the
difference can be even two times compared to that in the [001]-oriented sample as
measured in Nb-Ta-Ti-Zr single crystals (Tane et al., 2008). Such orientation
dependence of mechanical properties, especially the Young’s modulus, has also been
found in previous experimental studies (Tane et al., 2011; Zhang et al., 2021a; Zhang
et al., 2023a).

The evolution of atomic structures during tensile deformation are also characterized
along with the stress-strain curves, and the fraction of various types of atomic structure
are shown in Figure 2(b), 2(d) and 2(f). When the stress starts to drop for the [001]-
oriented sample, the fraction of FCC- and HCP-type atoms increase simultaneously
during the plastic deformation. The stress fluctuations for the [001]-oriented sample
is caused by the phase transition from FCC to HCP phase. The second peak of the
stress is caused by the reappearance of BCC-type atoms at the strain of 47%. Such
FCC-BCC phase transformation is found to be helpful to increase the strength and
plasticity in HEAs (Wu and Shao, 2023). This phenomenon with two peaks in stress-
strain curves has been observed in the experiments for HfNbTaTiZr system with an
increase of stress in the strain-hardening stage, and the strain-hardening rate curve
showed a typical hump reflecting the high strain-hardening (Eleti et al., 2020a). The
second yield-like point from the strain rate jump has been observed in the true stress-
strain curve for equiatomic HfNbTaTiZr HEA (Mills et al., 2023). Similarly, the drop
of the stress is mainly triggered by the phase transformation from BCC-type to HCP-
type atomic structure, and the increase at the second peak is caused by the slight
increase of the fraction of BCC-type atoms for the [110]- and [111]-oriented samples.
The stress fluctuations are influenced by the appearance of amorphous regions
consisted of different atomic structures. The phase transformation driven by
deformation has been observed in HEA systems in both experiments and simulations
(Fang et al., 2019; Huang et al., 2017; Niu et al., 2018). Especially, similar stress
drops due to phase transformation have been observed in other references from
experiments (Lee et al., 2021) and simulations (Ya-zhou et al., 2022).
3.2 Microstructure evolution and phase transformation
In order to understand the atomic arrangement and to characterize the phase
transformation under tensile deformation, the detailed atomic information involving in
the phase transformation and evolution of structure in the three HEA systems at the
specific strain along the three loading directions are characterized respectively. The
number fractions of all the five elements joining the phase transformation process from
the original BCC to other structures are evaluated as shown in Figure 3. For the [001]-
oriented sample, the fractions of Hf, Ti and Zr atoms in FCC-type and HCP-type atoms
are three highest among the five elements, indicating that Hf, Ti and Zr atoms are the
three main elements in participating the transformation from BCC to FCC and HCP
phases during tensile loading. These atoms are mainly involved in the transition to
FCC phase at the preceding plastic deformation, and to HCP phase at the succeeding
plastic deformation. The refractory elements Hf, Ti and Zr are stable locally in FCC
structure and naturally in HCP phase at ambient conditions (Aguayo et al., 2002; von
Rohr et al., 2016). The volume of BCC phase is generally lower than that of HCP
phase from experimental observation (Huang et al., 2018). During tensile
deformation, the structure is elongated, and the distorted lattice becomes less dense,
which allows these atoms to move easily for BCC-FCC or BCC-HCP phase transitions.
The involvement of Hf, Ti and Zr atoms in the phase transformation has also been
observed from experiments and simulations (Wu et al., 2023). As calculated for the
[110]-oriented sample in Figure 3(c) and 3(d), Hf and Ti atoms are the main element
involving in the process of BCC-HCP phase transformation. This can be attributed to
the reason that Ti and Hf are fully soluble in each other and the BCC-HCP phase
transformation occurs at all concentrations (Yamabe-Mitarai et al., 2022). Similar to
the situation in the [110]-oriented sample, the phase transformations from BCC
structure to FCC and HCP-type atoms in the [111]-oriented sample are mainly induced
by Ti atoms as suggested in Figure 3(e) and 3(f).
Figure 2. The mechanical responses of HfNbTaTiZr HEAs: Stress-strain curves and the
fraction of various types of atomic structure under tensile deformation along (a)(b)
[001], (c)(d) [110] and (e)(f) [111] crystallographic orientations. The atomic structures
at different stages with respect of the stress-strain curves are captured, and the atoms
colored in blue, green and red represent those in BCC, FCC and HCP structures,
respectively. The peaks in the stress-strain curves are highlighted with brown lines,
indicating the corresponding fractions of atomic structure for visual guidance.

The corresponding evolution of BCC structure during deformation process for the
[001]-oriented sample is characterized in Figure 4. After the initiation of plastic
deformation, BCC-type atoms at random positions of the HEA system start to transform
to other structure type with increasing strain, and such transformation also complete
when the strain reaches 25%. Further increase of strain leads to the reappearance of
BCC structure, and the distribution of BCC-type atoms at this stage are similar to that
at the preceding stage, which can be contributed to the generation of second peak in the
stress-strain curve. The continuous loading induces the second phase transformation
from BCC to other atomic structures, causing the drop of stress.
Figure 3. The corresponding changes of atom number fraction for FCC and HCP
structures in the (a)(b) [001]-, (c)(d) [110]-, and (e)(f) [111]-oriented HEA models
during tensile deformation.

Figure 4. The evolution of BCC structure in the HfNbTaTiZr HEA model loaded along
[001] orientation at different stages.

In order to understand the atomic arrangement and to characterize the phase


transformation under tensile deformation, the twin phenomenon and deformed
structures are captured in Figure 5. Although dislocations are initially absent in the
simulation models, dislocations will appear with twinning events at lattice phase
transformation. In the [001]-oriented sample, FCC phases nucleate separately in the
matrix and grow under uniaxial tensile loading. The slip of several planes after the
BCC to FCC phase transformation creates several stacking faults in HCP structure. In
the FCC structure, two adjacent HCP layers denote an intrinsic stacking fault (ISF), and
the inclusion of an FCC layer between two HCP layers denote an extrinsic stacking
fault (ESF) (Guo et al., 2023), which have been highlighted in Figure 5(a) and 5(b).
The sequential expansion of the stacking fault region is observed with an increase in
strain, which leads to the growth of HCP phase. In the [110]-oriented sample, the twin
structure of FCC phase is formed in HCP structure (see Figure 5(c) and 5(d)). In the
[111]-oriented sample, the HCP phases nucleate and grow along the equivalent (110)
plane in band-like structure (see Figure 5(e) and 5(f)). The change of dislocation
density is discussed in the Supplementary Information. The zoomed microstructure
evolution, the centrosymmetry parameter (CSP), the atomic displacement vectors and
the local shear strain for the [001]-oriented HEA system at the specific strains is
characterized in Figure 6. It can be seen that FCC-type atoms appear in the [001]-
oriented sample at the initial plastic deformation, leading to the decrease of stress.
The clusters in FCC structure starts to grow with increasing strain while HCP structure
appears almost simultaneously (see Figure 6(a1)). CSP is used to measure the local
lattice distortion around an atom, which can characterize whether the atom belongs to
a perfect lattice or a local defect. Most HCP atoms possess a high value of CSP (see
Figure 6(a2)), accompanied by random atomic movements at the lattice position (see
Figure 6(a3)). The local shear strain is mainly concentrated in the region of FCC
phase, which can be linked to the strain-induced transformation from BCC to FCC
structure. BCC structure has limited twinning ability due to its high stacking fault
energy (Wang et al., 2020b). As a result, deformation twinning is activated after the
phase transformation to FCC and HCP structures as shown in Figure 6(b1). Further
deformation generates many stacking faults and twins with different atomic
arrangements. Two adjacent HCP atomic layers indicate an intrinsic stacking fault,
while two HCP layers with an FCC atomic layer in between correspond to an extrinsic
stacking fault (Li et al., 2016). Similar phenomena have also been observed in BCC
HEA systems from experiments (Čížek et al., 2018; Gao et al., 2022). It is clear that
the atoms in the center of stacking faults in HCP structure show high value of CSP (see
Figure 6(b2)), but small value of local shear strain (see Figure 6(b4)). This can be
attributed to the distortions suffered by the boundary atoms around the formed stacking
faults (Cao et al., 2020), which can be seen from the directional atomic movements
along the boundary of the stacking fault captured in Figure 6(b3). More and more
HCP-type atoms appear, and they also act as a nucleation site for the regeneration of
BCC structure with the increase of strain (see Figure 6(c1)). The lattice distortion
becomes obvious in the system, and more and more atoms join in the directional
movements with large local shear strain (see Figure 6(c2), 5(c3) and 6(c4)). The
reemerging BCC structure is then turned into HCP structure with further increase of
strain (see Figure 6(d1)), resulting in the decrease of stress. The lattice distortion
becomes even severe with random atomic movements (see Figure 6(c2) and 6(c3)).
The decrease in plasticity and strength of HEA system is mainly determined by the
fraction of FCC structure, and the reappearance of BCC structure may be contributed
to the increase in the yield strength of HEA system. As the tensile strain continues to
increase, the larger shear strain locates in stacking faults formed by FCC structure and
the regeneration zone of BCC structure.

Figure 5. The local structural evolution during deformation in the (a)(b) [001]-, (c)(d)
[110]- and (e)(f) [111]-oriented HfNbTaTiZr HEA systems.
Figure 6. Microstructural evolution during tensile deformation in [001]-oriented
HfNbTaTiZr HEA system. (a1)(b1)(c1)(d1) The atomic configurations of a local region
at different strains; the atoms colored in blue, green and red represent those in BCC,
FCC and HCP structures, respectively. (a2)(b2)(c2)(d2) the corresponding CSP,
(a3)(b3)(c3)(d3) the corresponding atomic displacement vector, and (a4)(b4)(c4)(d4)
the corresponding local shear strain. The color bars are the same in (a2)-(d2), and the
same in (a4)-(d4). In order to provide a clear view of the displacement vector, only
the atoms with the highest atom number fraction in FCC and HCP structures are
characterized (i.e. Hf element for [001]-oriented sample). The insets in (a1)-(d1)
indicate the zoom out of the whole HEA structure.

For the [110]-oriented sample, the phase transformation from BCC to HCP structure
appears after the yielding point, and the BCC phase region decrease in the band shape
with the increasing strain. The HCP phase starts to grow, and the phase boundary
between BCC and the other atomic structure is in a random shape as shown at the strain
of 12.5% in Figure 7(a). Unlike the situation in the [001]-oriented sample, the
structure of the whole system almost turns into HCP structure without the transition to
FCC structure. With further deformation, the stress decreases slowly with randomly
generated FCC and HCP structures. The BCC-type atoms continue to decrease and
are distributed in the HEA system in a disordered manner with fluctuated tensile stress,
as shown at the strain of 45% from Figure 7(a). It can also be seen that the interface
between the BCC structure and the region containing other type of atoms is highly
irregular and incoherent. Such amorphous arrangement without a clear separation
between different crystalline structures contributes to the plasticity and ductility of the
HEA system with specific crystallographic orientation. For the [111]-oriented sample,
the direct drop of the stress after the yielding point is mainly related to the phase
transformation from BCC-type to other type of atoms in random positions. Unlike the
case in the [110]-oriented sample, such phase transformation starts from some local
positions with the transition to HCP-type atoms. The BCC-type atoms start to
decrease in the band shape from the strain of 12.5%, and slowly turn into amorphous
state while keeping the band shape with increasing strain to 45%. Such amorphization
under extreme uniaxial tensile has been discovered in FCC HEA system previously
(Jiang et al., 2022). The hardening of the [110]- and [111]-oriented HEA system is
mainly affected by phase transformation from BCC to HCP structure as well as from
crystalline structure to amorphization transition. Such transformation-induced
hardening from BCC to HCP phase has been observed in TiZrHfVxNbxTax refractory
HEAs (Jung et al., 2021), and the amorphization transition in HEA under extreme
uniaxial tensile that leads to hardening behavior has been found from both experiments
and simulations (Jiang et al., 2022; Zhao et al., 2021).
Figure 7. The evolution of BCC structure in the HfNbTaTiZr HEA model loaded along
(a) [110] and (b) [111] orientation at different stages.

The microstructure evolution and phase transformation under tensile deformation for
[110]- and [111]-oriented samples are shown in Figure 8. As shown in Figure 8(a1)-
7(b4), the main phase transformation in [110]-oriented sample under tensile loading is
from BCC to HCP structure with a high value of CSP in HCP phase, which does not
trigger much local shear strain at the strain of 12.5% (see Figure 8(a2) and 8(a4)). The
atoms almost keep the original atomic alignment with random atomic displacement.
Further increase of tensile strain leads to strong lattice distortion and disordered atomic
arrangement. The local shear strain increases when the amorphous regions start to
form, and is mainly concentrated in these regions, which is clearly demonstrated at the
strain of 25% (see Figure 8(b2) and 8(b4)). The increase of strain leads to the random
movement of atoms, suggesting the initiation of amorphous state formed by FCC- and
BCC-type atoms. The atoms in the random motion directions affect the movement of
surrounding atoms, causing the enlargement of disordered atomic arrangement. It is
noted from Figure 8(b3) that the atomic movement show inconspicuous localized
directionality in the amorphous region as highlighted in the yellow box.

For the [111]-oriented sample, the detailed atomic arrangement in the specified region
captured in Figure 8(c1) display a formed band shape of mainly HCP-type atoms in
amorphous region under tensile loading, which affects the plasticity of this HEA system.
The large lattice distortion appears at the region formed by HCP-type atoms with high
local shear strain. The atomic displacement characterized in Figure 8(c3) shows an
interesting result. The atoms inside the amorphous region move in a random manner,
while the atoms move directionally and oppositely around the phase boundaries
between BCC-type and HCP-type atoms at the strain of 12.5%, suggesting that such
process of phase transformation is caused by the relative interlayer slip. These atoms
affect the motion of surrounding atoms, causing the enlargement of the amorphous
region with increasing strain. The movement of these atoms then turns to several
vortex-like motions in the region with disordered atomic arrangement at the strain of
25%. Such vortex-like motions with certain flow direction cause larger lattice
distortion in the system, causing the enlargement of amorphous bands. Affected by
the vortex-like atomic displacement, the corresponding shear strain are still
concentrated in the amorphous bands within the HEA system, and remain in these
regions even when most of the system turns into amorphous state as can be seen from
the local shear strain map in Figure 8(d4). Therefore, the phase transformation can be
manipulated by loading in the corresponding crystallographic orientation for better
combination of strength-ductility properties in alloy processing.

Figure 8. Microstructural evolution during tensile deformation in the other two


HfNbTaTiZr HEA systems. The atomic configurations of a local region at different
strains for (a1)(b1) [110]- and (c1)(d1) [111]-oriented samples; the atoms colored in
blue, green and red represent those in BCC, FCC and HCP structures, respectively.
The corresponding CSP, atomic displacement vector, and local shear strain for (a2)-(a4),
(b2)-(b4) [110]- and (c2)-(c4), (d2)-(d4) [111]-oriented samples. The color bars are
the same in (a2)-(d2), and the same in (a4)-(d4). In order to provide a clear view of
the displacement vector, only the atoms with the highest atom number fraction in FCC
and HCP structures are characterized (i.e. Ti element for [110]- and [111]-oriented
sample). The insets in (a1)-(d1) indicate the zoom out of the whole HEA structure.

3.3 Nanoscale deformation and phase transformation mechanism


The formation of different phases is correlated with the critical stress generally for
single crystal alloys. The activation of slip systems in single crystal alloy generally
follows the Schmid’s Law, which states that the applied resolved shear stress (ARSS)
is equal to the critical shear stress (CSS). The slip of metals during plastic
deformation is generally activated when the ratio of ARSS to CSS is high. The
Schmidt factor is defined as: S = cosλcosφ, where  denotes the angle between the slip
direction and the tensile loading direction;  denotes the angle between the normal
direction of the slip plane and the tensile loading direction. Table 1 summarizes the
detailed information of the Schmid factor in HfNbTaTiZr systems with several slip
systems. It is apparent from the table that more slip and twin systems can be generated
in the [001]-oriented sample if the values of CSS are equal for all the systems.
However, the breakdown of the Schmid law among BCC metals has been discovered
in both experiments and simulations, and the systems with lower Schmid factor show
more possibility of preference for slip over that with higher Schmid factor (Gröger and
Vitek, 2020; Kraych et al., 2019). The breakdown of Schmid law under tensile
loading is principally due to the effect of shear stresses perpendicular to the slip
direction that change the structure of the dislocation core (Gröger et al., 2008). For
BCC HEAs, the activate slip system identified after tensile deformation has the highest
Schmid factor (Eleti et al., 2020b). Similarly, the activations of slip systems with high
Schmid factors in BCC HEA pillars and cylindrical samples have been observed in
compression deformation process (Huang et al., 2024; Zhang et al., 2021b). The
breakdown of the Schmid law is also captured. It has been observed that a few slip
systems with low Schmid factors have been activated, while the slip trace of the slip
system with high Schmid factors have been absent after nanoindentation, which can be
attributed to the slip selection on a widespread basis and uniformly over various
crystallographic slip planes (Wang et al., 2020a).

Slip systems
Orientation in BCC Mode of Schmid
single crystal Failure factor
Slip Slip
plane direction
{111} <110> Dislocation 0.408
[001] {112} <111> Twinning 0.471
{123} <111> Dislocation 0.463
{110} <111> Dislocation 0.408
[110] {112} <111> Twinning 0.236
{123} <111> Dislocation 0.154
{110} <111> Dislocation 0.272
[111] {112} <111> Twinning 0.314
{123} <111> Dislocation 0.103
Table 1 Schmid factors for several slip systems in BCC HEA samples along [001], [110]
and [111] loading directions. Here, the common slip systems have been selected based
on experimental observation (Charpagne et al., 2022).

To better understand the phase transformation mechanism, the BCC, FCC and HCP
structures obtained in the MD simulations are zoomed and taken out for analysis.
Figure 9 and Figure 10 give the evolution of phase transformation from BCC structure
to FCC structure in the [001]-oriented sample and to HCP structure in the [111]-oriented
sample. The sudden increase of FCC structure at about 20% under uniaxial tensile
loading along [001] direction in Figure 2(b) indicates that there exists a preferred
orientation for the occurrence of BCC-FCC phase transformation. According to the
Bain model, the BCC phase can transform into FCC phase when being uniaxially
stretched along <100> direction (Krasko and Olson, 1989). In the [001]-oriented
sample, the lattice deformation is triggered under tensile loading, and the lattice
constant with the initial value of 0.34 nm starts to decrease (see Figure 9(a) and 9(b)),
while slowly increases along the [001] direction accordingly. The atoms continue to
move until the ratio of the lattice constants along the three directions reaches a certain
value and satisfies the condition for the formation of FCC structure. Such BCC-FCC
phase transformation under ultrahigh stretching tensile stress has been observed in Nb
nanowires (Wang et al., 2018). The involved atoms undergo a slight movement and
rotation during the phase transformation process, as shown in Figure 9(c). The
diffusion coefficient is affected by applied stress and changes with strain. Phase
transformation is governed by atomic diffusion. The rate of change of the diffusion
coefficient with strain is related to the interatomic forces, and such relation can be
interpreted in terms of the interatomic potential-energy functions of the material system
between the diffusion coefficient and the applied stress (Girifalco and Grimes, 1961).
The diffusion coefficients for individual elements are calculated based on the measured
velocity auto-correlation function (VACF) in Figure 9(d). The VACF is defined by
the scalar product of the velocity vectors of a diffusing atom during dynamical process,
and its time-integral is proportional to the diffusion coefficient. The instantaneous
VACF value for each atom is measured during MD simulations. Therefore, the time-
dependent diffusion coefficient D(t) is calculated by integrating the VACF over the
simulation time and divide it by the number of degrees of freedom, which is expressed
as:

1
D ( t ) =  v ( 0 )  v ( t ) dt
d0
where d is the dimensionality of the space, v is the velocity vector of each atom, and
t is the time. Diffusion coefficient helps to describe the diffusion related phenomena
such as stress induced diffusion and phase transformation during tensile deformation.
The change of diffusion coefficient can reveal the main contributing element for the
deformation-induced phase transformation.

The diffusion coefficients of all the elements keep in a relatively low value at the strain
from 0 to 15%, suggesting that the phase transformation is mainly caused by the lattice
distortion. In the [111]-oriented sample, the original symmetry of BCC structure is
more likely to transform into trigonal symmetry of hexagonal lattice. The crystalline
lattice originally featured in BCC structure (see Figure 10(a)) changes into the state in
HCP structure (see Figure 10(b)). The involved atoms move almost linearly when
BCC structure evolves into HCP structure as shown in Figure 10(c). The diffusion
coefficients of Hf, Ti and Zr atoms keep in a relatively high value starting from the
strain of around 10% as illustrated in Figure 10(d), suggesting that the phase
transformation is mainly governed by the atomic diffusion. It should be noted that
diffusivity is affected by the melting point, the thermodynamic factor and the atomic
mobility (Liu et al., 2022; Zhang et al., 2022). The Zr diffusion is temperature
dependent following a linear Arrhenius-type behavior, and its improvement can be
contributed from lattice distortions (Zhang et al., 2022). The calculations using
ThermoCalc show that Zr and Hf have higher intrinsic diffusion coefficients than Ta in
several orders of magnitude at high temperature ranging from 1100 oC to 1250 oC (Liu
et al., 2022).

Figure 9. Schematic diagram of the mechanism of phase transformation (a) from BCC
structure to (b) FCC structure along [001] direction during tensile deformation, and the
green dashed lines in the crystal structure indicate the formed FCC structure after the
phase transformation; (c) The trajectories of selected atoms in the loading process from
(a) to (b), which are colored by grey lines, and the atoms colored in red, blue and green
represent Hf, Nb and Zr elements respectively; (d) The corresponding diffusion
coefficients for individual element in the [001]-oriented HEA model.

Figure 10. Schematic diagram of the mechanism of phase transformation (a) from BCC
structure to (b) HCP structure along [111] direction during tensile deformation, and the
red dashed lines in the crystal structure indicate the formed HCP structure after the
phase transformation; (c) The trajectories of selected atoms in the loading process from
(a) to (b), which are colored by grey lines, and the atoms colored in red, blue, yellow
and pink represent Hf, Nb, Ta and Ti elements respectively; (d) The corresponding
diffusion coefficients for individual element in the [111]-oriented HEA model.

The evolution of the energy in the HEA systems are calculated accordingly to analyze
energy barriers for the phase transformation. To perform the calculations, the groups
of atoms that change from BCC to FCC structure in the [001]-oriented sample and from
BCC to HCP structures in the [111]-oriented sample are selected. The evolution of
the average energy of these atoms during phase transformation in the deformation
process are calculated as shown in Figure 11. The equilibrium values at the initial
stage are -3.24 eV/atom for [001]-oriented sample and -3.42 eV/atom for [111]-oriented
sample. As the tensile deformation proceeds, the energy gradually increases as a
result of the strain energy. As the external strain further increases to 15% in the [001]-
oriented sample, a peak value of -3.18 eV/atom in the energy can be seen, which
corresponds to the almost completion of the phase transformation from BCC to FCC
structure as shown by the snapshot in Figure 11(a). The corresponding energy barrier
of the BCC to FCC structure transformation is ∼60 meV/atom. The peak value in the
[111]-oriented sample is -3.39 eV/atom, showing the initiation of phase transformation
as shown in Figure 11(b). The corresponding energy barrier for the BCC to HCP
structure transformation is ~25 meV/atom, which is lower than that for the BCC to FCC
structure transformation in [001]-oriented sample. Therefore, the BCC structure is
more easily transformed to HCP structure in the [111]-oriented HEA system under
extreme tensile loading condition. Compared to the energy difference of the ground
state and different structures from first-principles calculations from this study (see
Supplementary Information) and another reference (Chen et al., 2022), both values are
reasonable. The results indicate that the phase transformation is sensitive to the
crystallographic orientation. The higher energy barrier in the [001]-oriented sample
is due to more atoms involved in the transformation from BCC phase to FCC phase in
the initial plastic stage as shown in Figure 2(a). BCC phase is found to prefer to
transform into FCC phase through the Bain model when the BCC phase is uniaxially
stretched along [001] direction (Xie et al., 2021). It has been observed from
experiments that FCC phase is soft with relatively low strength while BCC phase is
hard with strain-hardening ability (He et al., 2014; Qin et al., 2019). This is because
FCC phase has more slip systems and thus deforms easier than BCC phase. During
the BCC to FCC phase transformation under tensile deformation, the structure loses its
stability and becomes more sensitive to the stress, leading to the drop of the tensile
strength along [001] orientation.

Figure 11. Energy evolution of the selected region during phase transformation. The
energy barriers are ∼60 meV/atom from BCC to FCC structure in the [001]-oriented
sample, and ∼25 meV/atom from BCC to HCP structure in the [111]-oriented sample.
The blue, green and red colors indicate atoms belonging to BCC, FCC and HCP
structures, respectively.

The simulation results suggest the strong orientation dependence of mechanical


properties on crystallographic orientation in HfNbTaTiZr HEAs, resulting in special
deformation mechanism with different plasticity and ductility. Phase transformation
plays an important role in enhancing the ductility of HEA systems. In the [001]-
oriented sample, the BCC to FCC phase transformation leads to lower strength. The
continuous accumulation of the deformation stacking faults caused by the Shockley
partial dislocations eventually leads to the FCC to HCP phase transformation in the
HEA system under continued tensile loading. The BCC transformation at the later
stage could cause structural dissipation and strain relaxation, which leads to the increase
in strength without the sacrifice of the ductility. In the [110]- and [111]-oriented
sample, the early-stage HCP phase transformation leads to better strength, and the
following amorphous transition could help in the enhancement of ductility, which is
shown in Figure 8 that the atoms characterized in various phases are dispersed randomly
at the shear band for the amorphous structure. The enhancement of strength-ductility
synergy can be achieved through tuning the HEA structure along specific
crystallographic orientation.

4. Conclusions
In this study, the mechanical responses and phase transition mechanism of HfNbTaTiZr
HEAs in different crystallographic orientations under uniaxial tensile deformation have
been investigated using MD simulations under the assumption of dislocation-free
samples and no multiplication features. The simulation results reveal that the
mechanical properties are orientation-dependent with the highest Young’s modulus
observed along with the [111] crystallographic orientation. The characterization of
microstructure evolution demonstrates that the deformation twinning, dislocation
process and phase transformation are affected by the crystallographic orientation with
respect to the loading direction. The crystal structure mainly transforms from BCC
phase into FCC or HCP phase as the tensile strain increases during the plastic
deformation, which is in consistence with experimental observations. In detail, the
atoms maintain their initial BCC structure during the elastic deformation. With the
further increase of tensile strain, the phase transformation undergoes different stages,
including the occurrence, growth and disappearance of BCC, FCC and HCP structures
in different HEA systems, resulting in diverse mechanical responses. The sudden drop
of the stress after reaching the yield stress in the [001]-oriented sample is caused by the
phase transition from BCC to FCC and HCP structures. The phase transformation
from BCC to HCP structure are strongly correlated to the strength and plasticity for the
[111]-oriented sample. In addition, the corresponding energy barrier needed for such
transformation from BCC to FCC or HCP is analyzed. These findings provide an
atomistic insight for comprehending the mechanical response of HfNbTaTiZr refractory
HEAs in high-strain-rate loading conditions, and serve as a valuable reference for
strengthening strategies in alloy design and the application of refractory HEA system
under extreme environments.
Acknowledgement
The authors are grateful to the support from the National Natural Science Foundation
of China (22108316 and 12302179), Natural Science Foundation of Zhejiang Province
(LY24E010001) and Ningbo Municipal Natural Science Foundation (No. 2023J102).
The authors are also thankful to Prof. F.H. Zhou and Prof. R.J. Jiang for their fruitful
discussions.

References
Aguayo, A., Murrieta, G., de Coss, R., 2002. Elastic stability and electronic structure
of fcc Ti, Zr, and Hf: A first-principles study. Physical Review B 65, 092106.
An, Z., Mao, S., Liu, Y., Wang, L., Zhou, H., Gan, B., Zhang, Z., Han, X., 2021. A novel
HfNbTaTiV high-entropy alloy of superior mechanical properties designed on the
principle of maximum lattice distortion. Journal of Materials Science & Technology 79,
109-117.
Borde, M., Dupuy, L., Pivano, A., Michel, B., Rodney, D., Amodeo, J., 2023.
Interaction between 1/2¡110¿{001} dislocations and {110} prismatic loops in uranium
dioxide: Implications for strain-hardening under irradiation. International Journal of
Plasticity, 103702.
Cao, F.-H., Wang, Y.-J., Dai, L.-H., 2020. Novel atomic-scale mechanism of incipient
plasticity in a chemically complex CrCoNi medium-entropy alloy associated with
inhomogeneity in local chemical environment. Acta Materialia 194, 283-294.
Charpagne, M.A., Stinville, J.C., Wang, F., Philips, N., Pollock, T.M., 2022. Orientation
dependent plastic localization in the refractory high entropy alloy HfNbTaTiZr at room
temperature. Materials Science and Engineering: A 848, 143291.
Chen, S.-M., Ma, Z.-J., Qiu, S., Zhang, L.-J., Zhang, S.-Z., Yang, R., Hu, Q.-M., 2022.
Phase decomposition and strengthening in HfNbTaTiZr high entropy alloy from first-
principles calculations. Acta Materialia 225, 117582.
Chen, S., Aitken, Z.H., Pattamatta, S., Wu, Z., Yu, Z.G., Srolovitz, D.J., Liaw, P.K.,
Zhang, Y.-W., 2021. Simultaneously enhancing the ultimate strength and ductility of
high-entropy alloys via short-range ordering. Nature Communications 12, 4953.
Chen, S.Y., Tong, Y., Tseng, K.K., Yeh, J.W., Poplawsky, J.D., Wen, J.G., Gao, M.C.,
Kim, G., Chen, W., Ren, Y., Feng, R., Li, W.D., Liaw, P.K., 2019a. Phase
transformations of HfNbTaTiZr high-entropy alloy at intermediate temperatures.
Scripta Materialia 158, 50-56.
Chen, S.Y., Wang, L., Li, W.D., Tong, Y., Tseng, K.K., Tsai, C.W., Yeh, J.W., Ren, Y.,
Guo, W., Poplawsky, J.D., Liaw, P.K., 2019b. Peierls barrier characteristic and
anomalous strain hardening provoked by dynamic-strain-aging strengthening in a body-
centered-cubic high-entropy alloy. Materials Research Letters 7, 475-481.
Čížek, J., Haušild, P., Cieslar, M., Melikhova, O., Vlasák, T., Janeček, M., Král, R.,

Harcuba, P., Lukáč, F., Zýka, J., Málek, J., Moon, J., Kim, H.S., 2018. Strength
enhancement of high entropy alloy HfNbTaTiZr by severe plastic deformation. Journal
of Alloys and Compounds 768, 924-937.
Csanádi, T., Castle, E., Reece, M.J., Dusza, J., 2019. Strength enhancement and slip
behaviour of high-entropy carbide grains during micro-compression. Scientific Reports
9, 10200.
El Atwani, O., Vo, H.T., Tunes, M.A., Lee, C., Alvarado, A., Krienke, N., Poplawsky,
J.D., Kohnert, A.A., Gigax, J., Chen, W.Y., Li, M., Wang, Y.Q., Wróbel, J.S., Nguyen-
Manh, D., Baldwin, J.K.S., Tukac, O.U., Aydogan, E., Fensin, S., Martinez, E., 2023.
A quinary WTaCrVHf nanocrystalline refractory high-entropy alloy withholding
extreme irradiation environments. Nature Communications 14, 2516.
Eleti, R.R., Bhattacharjee, T., Shibata, A., Tsuji, N., 2019. Unique deformation behavior
and microstructure evolution in high temperature processing of HfNbTaTiZr refractory
high entropy alloy. Acta Materialia 171, 132-145.
Eleti, R.R., Klimova, M., Tikhonovsky, M., Stepanov, N., Zherebtsov, S., 2020a.
Exceptionally high strain-hardening and ductility due to transformation induced
plasticity effect in Ti-rich high-entropy alloys. Scientific Reports 10, 13293.
Eleti, R.R., Stepanov, N., Yurchenko, N., Klimenko, D., Zherebtsov, S., 2021. Plastic
deformation of solid-solution strengthened Hf-Nb-Ta-Ti-Zr body-centered cubic
medium/high-entropy alloys. Scripta Materialia 200, 113927.
Eleti, R.R., Stepanov, N., Zherebtsov, S., 2020b. Mechanical behavior and thermal
activation analysis of HfNbTaTiZr body-centered cubic high-entropy alloy during
tensile deformation at 77 K. Scripta Materialia 188, 118-123.
Fan, X.J., Qu, R.T., Zhang, Z.F., 2022. Remarkably high fracture toughness of
HfNbTaTiZr refractory high-entropy alloy. Journal of Materials Science & Technology
123, 70-77.
Fang, Q., Chen, Y., Li, J., Jiang, C., Liu, B., Liu, Y., Liaw, P.K., 2019. Probing the phase
transformation and dislocation evolution in dual-phase high-entropy alloys.
International Journal of Plasticity 114, 161-173.
Fontaine, D.d., 1971. The number of independent pair-correlation functions in
multicomponent systems. Journal of Applied Crystallography 4, 15-19.
Gao, J., Huang, Y., Hu, X., Wang, S., Rainforth, W.M., Todd, I., Zhu, Q., 2022. An
alternative formation mechanism of {332}BCC twinning in metastable body-centered-
cubic high entropy alloy. Scripta Materialia 217, 114770.
George, E.P., Curtin, W.A., Tasan, C.C., 2020. High entropy alloys: A focused review
of mechanical properties and deformation mechanisms. Acta Materialia 188, 435-474.
George, E.P., Raabe, D., Ritchie, R.O., 2019. High-entropy alloys. Nature Reviews
Materials 4, 515-534.
Girifalco, L.A., Grimes, H.H., 1961. Effect of Static Strains on Diffusion. Physical
Review 121, 982-991.
Gröger, R., Bailey, A.G., Vitek, V., 2008. Multiscale modeling of plastic deformation
of molybdenum and tungsten: I. Atomistic studies of the core structure and glide of
1/2<111> screw dislocations at 0K. Acta Materialia 56, 5401-5411.
Gröger, R., Vitek, V., 2020. Single crystal yield criterion for chromium based on
atomistic studies of isolated 1/2[111] screw dislocations. International Journal of
Plasticity 132, 102733.
Guo, Q., Hou, H., Wang, K., Li, M., Liaw, P.K., Zhao, Y., 2023. Coalescence of
Al0.3CoCrFeNi polycrystalline high-entropy alloy in hot-pressed sintering: a
molecular dynamics and phase-field study. npj Computational Materials 9, 185.
Gupta, A., Jian, W.-R., Xu, S., Beyerlein, I.J., Tucker, G.J., 2022. On the deformation
behavior of CoCrNi medium entropy alloys: Unraveling mechanistic competition.
International Journal of Plasticity 159, 103442.
He, J.Y., Liu, W.H., Wang, H., Wu, Y., Liu, X.J., Nieh, T.G., Lu, Z.P., 2014. Effects of
Al addition on structural evolution and tensile properties of the FeCoNiCrMn high-
entropy alloy system. Acta Materialia 62, 105-113.
Huang, H., Wu, Y., He, J., Wang, H., Liu, X., An, K., Wu, W., Lu, Z., 2017. Phase-
Transformation Ductilization of Brittle High-Entropy Alloys via Metastability
Engineering. Advanced Materials 29, 1701678.
Huang, M., Jiang, J., Wang, Y., Liu, Y., Zhang, Y., Dong, J., Tong, Z., 2024. Unraveling
hot deformation behavior and microstructure evolution, flow stress prediction of
powder metallurgy BCC/B2 Al1.8CrCuFeNi2 HEA. Journal of Alloys and Compounds
972, 172828.
Huang, Q., Zhao, Q., Zhou, H., Yang, W., 2022. Misorientation-dependent transition
between grain boundary migration and sliding in FCC metals. International Journal of
Plasticity 159, 103466.
Huang, S., Li, W., Holmström, E., Vitos, L., 2018. Phase-transition assisted mechanical
behavior of TiZrHfTax high-entropy alloys. Scientific Reports 8, 12576.
Huang, X., Liu, L., Duan, X., Liao, W., Huang, J., Sun, H., Yu, C., 2021. Atomistic
simulation of chemical short-range order in HfNbTaZr high entropy alloy based on a
newly-developed interatomic potential. Materials & Design 202, 109560.
Jha, S., Sharma, A., Dasari, S., Muskeri, S., Banerjee, R., Mukherjee, S., 2023.
Orientation dependent stress-induced martensitic and omega transformations in a
refractory high entropy alloy. Materialia 28, 101741.
Jiang, K., Zhang, Q., Li, J., Li, X., Zhao, F., Hou, B., Suo, T., 2022. Abnormal hardening
and amorphization in an FCC high entropy alloy under extreme uniaxial tension.
International Journal of Plasticity 159, 103463.
Jung, Y., Lee, K., Hong, S.J., Lee, J.K., Han, J., Kim, K.B., Liaw, P.K., Lee, C., Song,
G., 2021. Investigation of phase-transformation path in TiZrHf(VNbTa)x refractory
high-entropy alloys and its effect on mechanical property. Journal of Alloys and
Compounds 886, 161187.
King, D.J.M., Cheung, S.T.Y., Humphry-Baker, S.A., Parkin, C., Couet, A., Cortie,
M.B., Lumpkin, G.R., Middleburgh, S.C., Knowles, A.J., 2019. High temperature, low
neutron cross-section high-entropy alloys in the Nb-Ti-V-Zr system. Acta Materialia
166, 435-446.
Krasko, G.L., Olson, G.B., 1989. Energetics of bcc-fcc lattice deformation in iron.
Physical Review B 40, 11536-11545.
Kraych, A., Clouet, E., Dezerald, L., Ventelon, L., Willaime, F., Rodney, D., 2019. Non-
glide effects and dislocation core fields in BCC metals. npj Computational Materials 5,
109.
Laplanche, G., Gadaud, P., Perrière, L., Guillot, I., Couzinié, J.P., 2019. Temperature
dependence of elastic moduli in a refractory HfNbTaTiZr high-entropy alloy. Journal
of Alloys and Compounds 799, 538-545.
Larsen, P.M., Schmidt, S., Schiøtz, J., 2016. Robust structural identification via
polyhedral template matching. Modelling and Simulation in Materials Science and
Engineering 24, 055007.
Lee, H., Shabani, M., Pataky, G.J., Abdeljawad, F., 2021. Tensile deformation behavior
of twist grain boundaries in CoCrFeMnNi high entropy alloy bicrystals. Scientific
Reports 11, 428.
Li, J., Chen, H., Fang, Q., Jiang, C., Liu, Y., Liaw, P.K., 2020. Unraveling the
dislocation–precipitate interactions in high-entropy alloys. International Journal of
Plasticity 133, 102819.
Li, J., Fang, Q., Liu, B., Liu, Y., Liu, Y., 2016. Mechanical behaviors of AlCrFeCuNi
high-entropy alloys under uniaxial tension via molecular dynamics simulation. RSC
Advances 6, 76409-76419.
Li, W., Chen, S., Aitken, Z., Zhang, Y.-W., 2023a. Shock-induced deformation and
spallation in CoCrFeMnNi high-entropy alloys at high strain-rates. International
Journal of Plasticity 168, 103691.
Li, Y.X., Nutor, R.K., Zhao, Q.K., Zhang, X.P., Cao, Q.P., Sohn, S.S., Wang, X.D., Ding,
S.Q., Zhang, D.X., Zhou, H.F., Wang, J.W., Jiang, J.Z., 2023b. Unraveling the
deformation behavior of the Fe45Co25Ni10V20 high entropy alloy. International
Journal of Plasticity 165, 103619.
Liu, C.-J., Gadelmeier, C., Lu, S.-L., Yeh, J.-W., Yen, H.-W., Gorsse, S., Glatzel, U.,
Yeh, A.-C., 2022. Tensile creep behavior of HfNbTaTiZr refractory high entropy alloy
at elevated temperatures. Acta Materialia 237, 118188.
Luan, H.-W., Shao, Y., Li, J.-F., Mao, W.-L., Han, Z.-D., Shao, C., Yao, K.-F., 2020.
Phase stabilities of high entropy alloys. Scripta Materialia 179, 40-44.
Mills, L.H., Emigh, M.G., Frey, C.H., Philips, N.R., Murray, S.P., Shin, J., Gianola,
D.S., Pollock, T.M., 2023. Temperature-dependent tensile behavior of the HfNbTaTiZr
multi-principal element alloy. Acta Materialia 245, 118618.
Motallebzadeh, A., Peighambardoust, N.S., Sheikh, S., Murakami, H., Guo, S.,
Canadinc, D., 2019. Microstructural, mechanical and electrochemical characterization
of TiZrTaHfNb and Ti1.5ZrTa0.5Hf0.5Nb0.5 refractory high-entropy alloys for
biomedical applications. Intermetallics 113, 106572.
Niu, C., LaRosa, C.R., Miao, J., Mills, M.J., Ghazisaeidi, M., 2018. Magnetically-
driven phase transformation strengthening in high entropy alloys. Nature
Communications 9, 1363.
Piao, M., Huh, H., Lee, I., Ahn, K., Kim, H., Park, L., 2016. Characterization of flow
stress at ultra-high strain rates by proper extrapolation with Taylor impact tests.
International Journal of Impact Engineering 91, 142-157.
Plimpton, S., 1995. Fast Parallel Algorithms for Short-Range Molecular Dynamics.
Journal of Computational Physics 117, 1-19.
Qin, G., Chen, R., Zheng, H., Fang, H., Wang, L., Su, Y., Guo, J., Fu, H., 2019.
Strengthening FCC-CoCrFeMnNi high entropy alloys by Mo addition. Journal of
Materials Science & Technology 35, 578-583.
Rida, A., Micoulaut, M., Rouhaud, E., Makke, A., 2019. Crystals at high deformation
rates displaying glassy behavior. Physica Status Solidi (B) 256, 1800649.
Rida, A., Micoulaut, M., Rouhaud, E., Makke, A., 2020. Understanding the strain rate
sensitivity of nanocrystalline copper using molecular dynamics simulations.
Computational Materials Science 172, 109294.
Senkov, O.N., Miracle, D.B., Chaput, K.J., Couzinie, J.-P., 2018. Development and
exploration of refractory high entropy alloys—A review. Journal of Materials Research
33, 3092-3128.
Senkov, O.N., Scott, J.M., Senkova, S.V., Meisenkothen, F., Miracle, D.B., Woodward,
C.F., 2012. Microstructure and elevated temperature properties of a refractory
TaNbHfZrTi alloy. Journal of Materials Science 47, 4062-4074.
Senkov, O.N., Scott, J.M., Senkova, S.V., Miracle, D.B., Woodward, C.F., 2011.
Microstructure and room temperature properties of a high-entropy TaNbHfZrTi alloy.
Journal of Alloys and Compounds 509, 6043-6048.
Shimizu, F., Ogata, S., Li, J., 2007. Theory of Shear Banding in Metallic Glasses and
Molecular Dynamics Calculations. Materials Transactions 48, 2923-2927.
Stukowski, A., 2010. Visualization and analysis of atomistic simulation data with
OVITO–the Open Visualization Tool. Modelling and Simulation in Materials Science
and Engineering 18, 015012.
Tane, M., Akita, S., Nakano, T., Hagihara, K., Umakoshi, Y., Niinomi, M., Nakajima,
H., 2008. Peculiar elastic behavior of Ti–Nb–Ta–Zr single crystals. Acta Materialia 56,
2856-2863.
Tane, M., Nakano, T., Kuramoto, S., Hara, M., Niinomi, M., Takesue, N., Yano, T.,
Nakajima, H., 2011. Low Young’s modulus in Ti–Nb–Ta–Zr–O alloys: Cold working
and oxygen effects. Acta Materialia 59, 6975-6988.
Tracy, C.L., Park, S., Rittman, D.R., Zinkle, S.J., Bei, H., Lang, M., Ewing, R.C., Mao,
W.L., 2017. High pressure synthesis of a hexagonal close-packed phase of the high-
entropy alloy CrMnFeCoNi. Nature Communications 8, 15634.
von Rohr, F., Winiarski, M.J., Tao, J., Klimczuk, T., Cava, R.J., 2016. Effect of electron
count and chemical complexity in the Ta-Nb-Hf-Zr-Ti high-entropy alloy
superconductor. Proceedings of the National Academy of Sciences 113, E7144-E7150.
Wang, F., Balbus, G.H., Xu, S., Su, Y., Shin, J., Rottmann, P.F., Knipling, K.E., Stinville,
J.-C., Mills, L.H., Senkov, O.N., Beyerlein, I.J., Pollock, T.M., Gianola, D.S., 2020a.
Multiplicity of dislocation pathways in a refractory multiprincipal element alloy.
Science 370, 95-101.
Wang, Q., Wang, J., Li, J., Zhang, Z., Mao, S.X., 2018. Consecutive crystallographic
reorientations and superplasticity in body-centered cubic niobium nanowires. Science
Advances 4, eaas8850.
Wang, X., Wang, J., He, Y., Wang, C., Zhong, L., Mao, S.X., 2020b. Unstable twin in
body-centered cubic tungsten nanocrystals. Nature Communications 11, 2497.
Wilkerson, J.W., 2017. On the micromechanics of void dynamics at extreme rates.
International Journal of Plasticity 95, 21-42.
Wu, Y.-C., Shao, J.-L., 2023. FCC-BCC phase transformation induced simultaneous
enhancement of tensile strength and ductility at high strain rate in high-entropy alloy.
International Journal of Plasticity 169, 103730.
Wu, Y., Yu, W., Shen, S., 2023. Developing a variable charge potential for
Hf/Nb/Ta/Ti/Zr/O system via machine learning global optimization. Materials &
Design 230, 111999.
Wu, Y., Zhang, F., Yuan, X., Huang, H., Wen, X., Wang, Y., Zhang, M., Wu, H., Liu, X.,
Wang, H., Jiang, S., Lu, Z., 2021. Short-range ordering and its effects on mechanical
properties of high-entropy alloys. Journal of Materials Science & Technology 62, 214-
220.
Xiong, W., Guo, A.X.Y., Zhan, S., Liu, C.-T., Cao, S.C., 2023. Refractory high-entropy
alloys: A focused review of preparation methods and properties. Journal of Materials
Science & Technology 142, 196-215.
Ya-zhou, L., Yun, L., Shuo, S., Yan-yu, S., Sheng-peng, H., Xiao-guo, S., Ning, G.,
Wei-min, L., 2022. Molecular dynamics simulation of phase transition and crack
propagation in metastable high entropy alloy. Materials Today Communications 33,
104642.
Yaakobi, B., Boehly, T.R., Meyerhofer, D.D., Collins, T.J.B., Remington, B.A., Allen,
P.G., Pollaine, S.M., Lorenzana, H.E., Eggert, J.H., 2005. EXAFS Measurement of Iron
bcc-to-hcp Phase Transformation in Nanosecond-Laser Shocks. Physical Review
Letters 95, 075501.
Yamabe-Mitarai, Y., Yanao, K., Toda, Y., Ohnuma, I., Matsunaga, T., 2022. Phase
stability of Ti-containing high-entropy alloys with a bcc or hcp structure. Journal of
Alloys and Compounds 911, 164849.
Yang, W., Pang, S., Liu, Y., Wang, Q., Liaw, P.K., Zhang, T., 2022. Design and
properties of novel Ti–Zr–Hf–Nb–Ta high-entropy alloys for biomedical applications.

Intermetallics 141, 107421.


Yin, S., Zuo, Y., Abu-Odeh, A., Zheng, H., Li, X.-G., Ding, J., Ong, S.P., Asta, M.,
Ritchie, R.O., 2021. Atomistic simulations of dislocation mobility in refractory high-
entropy alloys and the effect of chemical short-range order. Nature Communications 12,
4873.
Zhang, C., Wang, X., Xu, M., MacDonald, B.E., Hong, R., Zhu, C., Dai, X., Vecchio,
K.S., Apelian, D., Hahn, H., Schoenung, J.M., Lavernia, E.J., 2021a. Orientation-
dependent superelasticity of a metastable high-entropy alloy. Applied Physics Letters
119.
Zhang, J., Gadelmeier, C., Sen, S., Wang, R., Zhang, X., Zhong, Y., Glatzel, U.,
Grabowski, B., Wilde, G., Divinski, S.V., 2022. Zr diffusion in BCC refractory high
entropy alloys: A case of ‘non-sluggish’ diffusion behavior. Acta Materialia 233,
117970.
Zhang, Q., Huang, R., Zhang, X., Cao, T., Xue, Y., Li, X., 2021b. Deformation
Mechanisms and Remarkable Strain Hardening in Single-Crystalline High-Entropy-
Alloy Micropillars/Nanopillars. Nano Letters 21, 3671-3679.
Zhang, W., Shen, J., Oliveira, J.P., Kooi, B.J., Pei, Y., 2023a. Crystallographic
orientation-dependent deformation characteristics of additive manufactured interstitial-
strengthened high entropy alloys. Scripta Materialia 222, 115049.
Zhang, Y.W., Li, S.J., Obbard, E.G., Wang, H., Wang, S.C., Hao, Y.L., Yang, R., 2011.
Elastic properties of Ti–24Nb–4Zr–8Sn single crystals with bcc crystal structure. Acta
Materialia 59, 3081-3090.
Zhang, Z., Huang, Q., Zhou, H., 2023b. High-entropy alloy nanocrystals with low-
angle grain boundary for superb plastic deformability and recoverability. International
Journal of Plasticity 167, 103679.
Zhao, S., Li, Z., Zhu, C., Yang, W., Zhang, Z., Armstrong, D.E.J., Grant, P.S., Ritchie,
R.O., Meyers, M.A., 2021. Amorphization in extreme deformation of the CrMnFeCoNi
high-entropy alloy. Science Advances 7, eabb3108.

Author statement
Wei Jian: Conceptualization, Formal analysis, Methodology, Software, Validation,
Visualization, Writing-Original Draft, Funding acquisition.
Lu Ren: Formal analysis, Resources, Investigation, Writing- Review & Editing,
Funding acquisition.
Declaration of interests

☒ The authors declare that they have no known competing financial interests or

personal relationships that could have appeared to influence the work reported in this
paper.

☐ The authors declare the following financial interests/personal relationships which may

be considered as potential competing interests:

You might also like