ArticlePDF Available

Distinct point defect behaviours in body-centered cubic medium-entropy alloy NbZrTi induced by severe lattice distortion

Authors:

Abstract and Figures

The point defect properties of body-centered cubic medium-entropy alloy NbZrTi were studied by first-principles calculations. Due to severe lattice distortion, a significant portion of conventional vacancy and interstitial structures are unstable and require large structural relaxation, indicating an irregular energy landscape with large site-to-site variations. The average vacancy and interstitial formation energy are 0.95 eV± 0.34 eV and 1.92 eV ± 0.39 eV, respectively, much lower than that of Nb (2.77 eV and 4.38 eV). The vacancy migration energy exhibits a wide distribution extending to 0 eV, resulting in preferential vacancy migration through low barrier sites. The interstitial diffusion is slower than that of pure Nb due to the reduction of long <111> diffusion induced by the site-to-site variations in stable interstitial orientations. Ti atoms diffuse much faster than Nb and Zr atoms due to the preferential interstitial binding with Ti. The effect of atomic composition and short-range order on elemental and total interstitial diffusion was also investigated. The obtained first-principles results are important for the development of interatomic potentials for radiation damage studies. When irradiated with 3-MeV Fe ions at 675 ∘C to a peak dose of ∼100 dpa, NbZrTi reduced the void formation at high temperature compared to Nb owing to its higher equilibrium vacancy concentration and closer mobility between vacancies and interstitial atoms.
Content may be subject to copyright.
Distinct point defect behaviours in body-centered cubic medium-entropy alloy
NbZrTi induced by severe lattice distortion
Tan Shia, Zhengxiong Sua, Jing Lia, Chenguang Liuc, Jinxue Yanga, Xinfu Hed, Di Yuna, Qing Penge,,
Chenyang Lua,b,
aSchool of Nuclear Science and Technology, Xi’an Jiaotong University, Xi’an, 710049, China
bState Key Laboratory of Multiphase Flow in Power Engineering, Xi’an Jiaotong University, Xi’an, 710049, China
cCollege of Nuclear Equipment and Nuclear Engineering, Yantai University, Yantai, 264005, China
dChina Institute of Atomic Energy, Beijing, 102413, China
eState Key Laboratory of Nonlinear Mechanics, Institute of Mechanics, Chinese Academy of Sciences, Beijing, 100190, China
Abstract
The point defect properties of body-centered cubic medium-entropy alloy NbZrTi were studied by first-principles
calculations. Due to severe lattice distortion, a significant portion of conventional vacancy and interstitial structures
are unstable and require large structural relaxation, indicating an irregular energy landscape with large site-to-site
variations. The average vacancy and interstitial formation energy are 0.95 eV±0.34 eV and 1.92 eV ±0.39 eV,
respectively, much lower than that of Nb (2.77 eV and 4.38 eV). The vacancy migration energy exhibits a wide
distribution extending to 0 eV, resulting in preferential vacancy migration through low barrier sites. The interstitial
diusion is slower than that of pure Nb due to the reduction of long <111>diusion induced by the site-to-site
variations in stable interstitial orientations. Ti atoms diuse much faster than Nb and Zr atoms due to the preferential
interstitial binding with Ti. The eect of atomic composition and short-range order on elemental and total interstitial
diusion was also investigated. The obtained first-principles results are important for the development of interatomic
potentials for radiation damage studies. When irradiated with 3-MeV Fe ions at 675 °C to a peak dose of 100 dpa,
NbZrTi reduced the void formation at high temperature compared to Nb owing to its higher equilibrium vacancy
concentration and closer mobility between vacancies and interstitial atoms.
Keywords: Medium-entropy alloys, Point defect properties, First-principles calculations, Ion irradiation, Radiation
resistance
1. Introduction
Medium-entropy alloys (MEAs) and high-entropy al-
loys (HEAs) are important groups of multi-principal el-
ement alloys (MPEAs), a class of materials with two
or more principal alloying elements. The entropy-
based definitions originate from the contribution of
high mixing entropy to the formation of solid-solution
phase [1]. MPEAs have demonstrated excellent me-
chanical properties [2, 3] and damage resistance [4]
owing to their unique features of lattice distortion,
sluggish diusion, high-entropy eect and cocktail ef-
fect [1]. Among MPEAs, MEA NbZrTi and NbZrTi-
based refractory HEAs have exhibited great mechani-
Corresponding authors.
Email addresses: pengqing@imech.ac.cn (Qing Peng),
chenylu@xjtu.edu.cn (Chenyang Lu)
cal performance and have shown great promise for ap-
plications as high-temperature structural materials [3,
5, 6]. This group of materials has been recently
considered for potential employment in advanced nu-
clear reactors [7]. The advanced reactors pose se-
vere challenges to the material performance, as mate-
rials need to suer from high operating temperature (up
to 1050 °C), high irradiation damage (up to 150 dis-
placement per atom (dpa)), harsh corrosive environ-
ment, and also ideally a longer service time [8, 9].
Therefore, there is a continuous demand for ideal ma-
terial candidates that can meet these extreme require-
ments. As conventional refractory alloys, NbZrTi-based
refractory MPEAs have excellent mechanical strength
at high temperature [3]. In addition, various types
of NbZrTi-based alloys have also shown good room-
temperature ductility [7], such as NbZrTi [5], NbZr-
Preprint submitted to Acta Materialia March 5, 2022
TiHf [10], NbZrTiHfTa [11, 12], NbZrTiMo0.3V0.3[13],
NbZrTiTaHfMo0.75 [14], (NbZrTiHf)0.95Al0.05 [15], etc.
The body-centered cubic (BCC) structure also has an
intrinsically better resistance to radiation-induced void
swelling [16]. The concept of MPEAs oers a large
compositional space for the material property tuning,
but the actual application of a material is usually a
multifaceted problem where dierent properties need
to be balanced at the stage of alloy design. While the
aspects of alloy production and mechanical properties
of various alloy compositions have been studied previ-
ously [3], the point defect properties have rarely been
explored. As point defect energetics have a great im-
pact on the mechanical properties [17], corrosion be-
haviours [18] and radiation resistance [19], they are
studied in this work with NbZrTi as a representative ma-
terial to show the distinct point defect properties of this
group of materials.
Point defect behaviours have been systematically
studied in face-centered cubic (FCC) structures, in-
cluding Ni-based binary and ternary MPEAs [19,
20], Co-free HEA NiFeMnCr [21] and CoCrNi-based
MPEAs [22], and have also been studied in BCC HEA
VTaCrW [23]. These studies have all shown that point
defect formation and migration energies exhibit a cer-
tain degree of energy fluctuation due to the dierence in
the local chemical environment, and the interstitial dif-
fusion is reduced due to lattice distortion and sluggish
diusion eect. The closer mobility between vacancies
and interstitials is believed to promote the point defect
recombination during irradiation. However, the defect
energy seems to be also alloy dependent. For example,
there is no consistent trend for the relative magnitude of
vacancy formation energy compared to the constituent
elements. In addition, due to the more severe lattice dis-
tortion of BCC MPEAs, the energy distribution is much
wider [23], indicating significantly dierent point defect
properties from their widely studied FCC counterparts.
Recently, there have been controversies regarding
the currently established theory on the radiation resis-
tance of MPEAs. FCC MPEAs, which are more ex-
tensively studied, have shown great radiation resistance
including reduced void swelling [4, 24], delayed de-
fect microstructure growth and less degree of radiation-
induced segregation [25]. It is believed to be attributed
to the enhanced point defect recombination, slower en-
ergy dissipation, extended incubation period of dislo-
cation loop growth and slower diusion of defect clus-
ters [4, 20, 25, 26]. However, Parkin et al. found similar
saturation dose for defect clusters between BCC NbTa-
TiV and V under heavy ion irradiation [27]. If the de-
fect recombination rate is higher and the defect cluster
growth is delayed, the threshold dose should be higher
in MPEAs. The same work also showed that NbTaTiV
with a large lattice distortion did not lead to a greater re-
duction in defect generation compared to its FCC coun-
terparts, whereas the reduction in defect mobility is usu-
ally believed to be associated with the magnitude of
lattice distortion. In addition, by the same method of
molecular dynamics (MD) simulations of collision cas-
cades induced by energetic particles [28], discrepancies
exist on whether enhanced point defect recombination
during the initial displacement cascade is a major con-
tributor to the radiation resistance of MPEAs, which
may be only caused by defect evolution on a longer time
scale [29, 30]. Recently, there has been great progress
in the development of interatomic potentials for defect
and radiation damage studies for BCC MPEAs, includ-
ing MoNbTaVW [31], TiVTa [32, 33] and VTaW [33].
Thus, for NbZrTi-based MPEAs and also other BCC
MPEAs, it is crucial to first understand their unique
point defect behaviours by first-principles method so
that accurate interatomic potentials can be constructed
and their impact on radiation resistance can be accu-
rately characterized.
In this work, we present a first-principles investiga-
tion on the energy properties and migration behaviours
of point defects in NbZrTi. First, the magnitude of
lattice distortion of NbZrTi was characterized, show-
ing more severe lattice distortion compared to FCC and
some BCC MPEAs. Next, vacancy formation energy
and migration energy were systematically calculated,
revealing an unusual energy landscape induced by the
large lattice distortion and chemical complexity. The
interstitial formation energy and migration behaviours
were then studied, along with the impact of atomic com-
position and short-range order (SRO) on interstitial mi-
gration. The obtained first-principles results were then
used to benchmark an interatomic potential based on the
embedding-atom method (EAM) potential. Ion irradia-
tion experiment was performed at 675 °C with 3-MeV
Fe ions at a peak dose of 100 dpa to study the dier-
ence in radiation damage behaviours between pure Nb
and NbZrTi, showing a direct impact of point defect en-
ergetics on the material irradiation resistance.
2. Methodology
2.1. First-principles calculations
The first-principles calculations were performed with
the Vienna ab initio simulation package (VASP) us-
ing the projector-augmented wave method [34]. The
generalized gradient approximation of Perdew-Burke-
Ernzerhof (PBE) formalism was used to describe the
2
exchange-correlation functionals [35]. Semi-core elec-
trons were included in all calculations to describe
short interatomic distances for interstitial configura-
tions, where 13, 12 and 12 electrons were considered
for Nb, Zr and Ti, respectively. The studied system con-
sists of 128 atoms with 42 Nb atoms, 43 Zr atoms, and
43 Ti atoms. The alloy structure was constructed by
the special quasi-random structure (SQS) method [36]
using the Alloy Theoretic Automated Toolkit [37, 38]
to ensure the equal probability of appearance of each
type of atom at the first nearest neighbour (1NN) and
second nearest neighbour (2NN) sites. The Brillouin
zone was sampled with a Monkhorst-Pack mesh of 3
×3×3. This choice of k-point grid was validated by
vacancy and interstitial formation energies of pure Nb,
Zr and Ti (see Fig. S1 of the Supplementary Materi-
als for the comparison of vacancy and interstitial for-
mation energies with dierent k-point meshes) [39–41].
The cutoenergy for the plane-wave basis was set to
400 eV, and the convergence criteria for ionic relax-
ation and self-convergence loop were set to 104eV and
0.01 eV/Å, respectively. The spin polarization was ini-
tially included by using three dierent initial magnetic
moments, but no magnetism was manifested after en-
ergy minimization. Thus, spin polarization was not in-
cluded for all the following calculations of NbZrTi.
First, the system volume and atomic coordinates of
the perfect lattice structure was determined based on
the conjugate gradient optimization. The phonon cal-
culation was performed to verify the structural dynami-
cal stability. Second-order force constants of the super-
cell were calculated with a finite atomic displacement
of 0.01 Å to each atom of the supercell for a total 768
sets of displacement. We maintained a k-point mesh of
3×3×3 and used an additional support grid for the
calculation of augmentation charges. The energy con-
vergence criteria was set to 107eV. The band unfolding
method from Ref. [42, 43] was used to decompose the
phonon modes onto the Brillouin zone of the primitive
BCC cell based on the crystallographic symmetry using
the PHONOPY package [44].
Next, dierent vacancy and interstitial structures
were constructed and relaxed with a fixed system vol-
ume. For the calculation of vacancy formation ener-
gies, 15 vacancies of each type of atom were randomly
selected while it was ensured that the average number
of each element is approximately equal at neighbour-
ing sites. The initial structures for the calculation of
interstitial formation energies include in total 18 <110>
dumbbells, 18 <111>dumbbells, 9 <100>dumbbells
and 9 tetrahedral sites with equal number of Nb, Zr or Ti
atoms being added as the interstitial atoms. The choice
of initial interstitial structures is based on the relative
stability of dierent interstitial structures in pure Nb.
Our simulations show that the final interstitial structures
have weak dependence on the initial structures. In ad-
dition to the random generation of interstitial sites, dif-
ferent <110>and <111>dumbbell orientations at the
same interstitial site was also simulated to study the im-
pact of local chemical environment. The point defect
formation energy was calculated as:
Ef
def =Edef Eper ±µx,(1)
where Eper is the energy of the perfect structure, Edef
is the system energy with a vacancy/interstitial, and µx
is the chemical potential of the removed atom (+) or
added atom (-). The chemical potential of each ele-
ment was determined by the Widom-type substitution
technique [45] as adopted by previous defect studies of
MPEAs [20, 21]. By substituting one element to an-
other and calculating the energy dierence of the two
systems, the chemical potential of each element can be
determined. 20 substitutions were computed in total and
the chemical potentials for Nb, Zr and Ti were deter-
mined to be -10.18 eV ±0.03 eV, -8.39 eV ±0.04 eV
and -7.76 eV ±0.02 eV, respectively. The Bader charge
analysis was performed to determine the charge and vol-
ume of the removed atom [46].
The vacancy migration energies were computed
using the climbing-image nudged elastic band (CI-
NEB) method with a force convergence tolerance of
0.03 eV/Å [47]. In total, 60 migration paths were com-
puted in which the first 30 migration paths used 3 or
5 intermediate images to characterize the shape of the
migration barriers and the rest of the migration paths
used 1 intermediate image to increase the statistics of
the average migration energy. As there is a large varia-
tion among migration barriers of dierent sites, one in-
termediate image has sucient accuracy to evaluate the
average migration energy.
The interstitial migration dynamics behaviours were
studied by the ab initio molecular dynamics (AIMD)
method. In dynamics simulations, a single k-point was
employed with an energy cutoof 400 eV for the plane-
wave basis. A systematic comparison of point defect
formation energies was performed between a k-point
mesh of 1 ×1×1 and 3 ×3×3, showing similar
average formation energies and correct order for dier-
ent types of interstitial pairs (see Table S1 and S2 of
the Supplementary Materials for the comparison of va-
cancy and interstitial formation energies). The system
was controlled by the NVT ensemble using the Nose-
Hoover thermostat. The interstitial migration of dier-
3
ent structures was compared at a temperature of 1200 K.
The results were averaged over four dierent simula-
tions in which the interstitial atom was introduced at a
random site. For each simulation, after an initial equili-
bration time of 3 ps, a total simulation time of 32 ps was
computed with a timestep of 2 fs. The mean square dis-
placement (MSD) was analyzed by the VASP analysis
program VASPKIT [48]:
MS D(m)=1
Na
Na
X
i=1
1
Ntm
Ntm1
X
k=0
(ri(k+m)ri(k))2,(2)
where mis the length of the time window, ri(k) is the
position vector of atom iat a timestep of k,Ntis the
total number of timesteps, and Nais the total number
of atoms. The MSD of the first half of the simulation
time was selected to avoid large statistical variations at
longer time due to limited integration timesteps [19, 49].
The interstitial trajectory was analyzed via the Wigner-
Seitz defect analysis from the Ovito program [50]. The
distribution of atomic positions of the perfect lattice
was also studied with similar settings using the AIMD
method at 300 K, 600 K and 900 K. The NVT ensem-
ble was employed with a timestep of 3 fs for a total of
15 ps. The atom positions after 1000 timesteps were
used to ensure the equilibration of the system.
Lattice Monte Carlo (MC) simulations were per-
formed to obtain the SRO structure of NbZrTi [51]. The
positions of two random atoms of dierent species were
exchanged with an acceptance probability based on the
Metropolis-Hastings sampling [52]. The applied tem-
perature was 500 K. The supercell volume was opti-
mized every 50 swaps. A k-point mesh of 1 ×1×1 was
used during the MC process. The Warren-Cowley SRO
parameter was used to describe the extent of chemical
ordering [51, 53]:
αν
i j =1
pν
i j
cj
,(3)
where pν
i j is the probability of finding element jaround
element iin the νth neighbouring shell, cjis the atomic
concentration of element j. A random solid solution
corresponds to a SRO parameter of zero. Positive and
negative values correspond to a tendency of deceasing
and increasing the number of i,jpairs, respectively. As
shown in Fig. 1, with the increase of MC steps, there is
a general trend of energy decrease and a clear tendency
of elemental ordering. Although it is dicult to obtain
a full convergence with limited MC steps, the variations
of system energy and SRO parameters become more
gentle around 1500 steps. Thus, the SRO structure at
1500 MC steps was used for the following study of in-
terstitial migration.
(b)
(a)
Figure 1: (a) Evolution of accepted total energy relative to the initial
system energy as a function of the Monte Carlo steps. (b) Evolution
of Warren-Cowley SRO parameters of dierent atom pairs in the first
nearest neighbour shell for the accepted Monte Carlo steps.
For comparison among BCC and FCC structures,
equiatomic NiCoCr and NiCoFeCrMn with a FCC
structure and equiatomic VTaTi and VTaW with a BCC
structure were also studied by first-principles calcula-
tions with similar simulation settings. NiCoCr, VTaTi
and VTaW were used to compare the extent of lattice
distortion and NiCoFeCrMn was used to study the pro-
cess of vacancy structural relaxation. For FCC struc-
tures, a supercell of 108 atoms was constructed with the
SQS method. A Monkhorst-Pack grid of 2 ×2×2 was
used to sample the Brillouin zone with an energy cut-
oof 500 eV for the plane-wave basis. Spin polariza-
tion was considered for all calculations to account for
the magnetic properties of the constituent elements. For
each alloy system, three dierent initial paramagnetic
spin moments were considered, and the relaxed struc-
ture with the lowest system energy was selected.
2.2. Irradiation experiment and characterization
The NbZrTi with a nominal composition of
Nb33Zr34 Ti33 (at.%) was prepared by arc melting and
casting by mixing pure metals (purity >99.9 wt.%)
in a vacuum induction furnace. The as-cast alloys
4
were homogenized at 1200 °C for 2 h in an Ar atmo-
sphere to achieve a uniform elemental distribution. The
as-homogenized bulk sample presented a single BCC
structure, as evidenced by the X-ray diraction pattern
(see Fig. S2 of the Supplementary Materials). For com-
parison, Nb with a purity >99.99 wt.% was also pre-
pared for the irradiation experiment. Before irradiation,
all samples were mechanically polished using colloidal
silica and then vibratory polished to remove the stress
layer, resulting in a mirror surface with a roughness of
less than 3 nm. Subsequently, the samples were irra-
diated at 675 °C with 3-MeV Fe ions at a fluence of
1.02 ×1017 ions/cm2. The depth profile of displacement
damage was calculated by the SRIM code [54], assum-
ing a displacement threshold energy of 60 eV, 40 eV and
30 eV for Nb, Zr and Ti, respectively [55]. The peak
dose was calculated to be 99.8 dpa and 129 dpa for Nb
and NbZrTi, respectively, with the Kinchin–Pease op-
tion in the quick calculation mode [56].
After irradiation, a focused ion beam (FIB) lift-out
technique was utilized to prepare the cross-sectional
transmission electron microscope (TEM) samples. The
FIB-induced damage was completely removed from the
TEM samples using “flashing polishing” technique to
avoid artifacts [57]. The characterization of radiation-
induced defect microstructures was carried out on a FEI
Talos-F200X operated at 200 kV.
3. Results
3.1. Local lattice distortion
Local lattice distortion is one of the important prop-
erties of MPEAs, which has a significant impact on
their mechanical properties, electrical and thermal con-
ductivity, and damage tolerance [4, 58, 59]. Based
on first-principles calculations, the lattice constant for
NbZrTi is 3.381 Å, which is in good agreement with
experimental results of 3.36 3.41 Å [5]. The BCC
HEAs tend to have a larger lattice distortion than FCC
NiCoFeCrMn-based alloys due to the larger atomic size
mismatch [60, 61]. It is shown from Fig. 2(b) and (c)
that after structural relaxation, the broadening in the
atomic distance is much wider in NbZrTi than the FCC
MEA NiCoCr. Severe lattice distortion can be seen
by the overlap between the 1NN and 2NN in the ra-
dial distribution function, which is also seen in other
refractory MEAs and HEAs including VTaCrW, HfN-
bZr, HfNbTiZr and HfNbTaTiZr [23, 61].
The extent of lattice distortion of NbZrTi is quantified
here by several commonly used parameters [60, 61, 63],
(a)
(b)
(c)
Figure 2: Radial distribution function of the BCC NbZrTi structure
constructed by the special quasi-random structure method (a) before
and (b) after structural relaxation. The distributions from dierent
atom pairs are overlapped in (a). (c) Radial distribution function of
the FCC equiatomic NiCoCr alloy after structural relaxation.
Table 1: Quantification of the local lattice distortion in NbZrTi. The
atomic radii are taken from Ref. [62].
Atomic size Local atomic distortion Average atomic
mismatch δ(%) parameter α2displacement d(Å)
5.03 0.023 0.171
as listed in Table 1:
δ=v
tn
X
i
ci(1 ri/r)2,r=
n
X
i
ciri,(4)
5
α2=
n
X
ji
cicjri+rj2r
2r,(5)
d=1
NaX
irrelaxed,irunrelaxed,i,(6)
where riand ciare the atomic radius and atomic percent-
age of element i,runrelaxed and rrelaxed are atomic position
vectors in unrelaxed and relaxed structures, Nais the to-
tal number of atoms, and α2adopts the parameter repre-
sentation from Ref. [60] to reflect the pairwise relation-
ship between dierent atom types. δis frequently used
to estimate the lattice distortion due to the atomic size
dierence [63, 64], and α2has shown strong correlation
with dierent alloy properties including phase, intrinsic
elastic strain energy and excess entropy [60]. However,
these two parameters are calculated directly from the
elemental radii, whereas dis determined by the first-
principles method, which should provide a more accu-
rate measure of the lattice distortion. The parameter of
dwas compared among supercells of dierent sizes,
showing that a supercell of 128 atoms is sucient to
represent the lattice distortion (see Fig. S3 of the Sup-
plementary Materials for the comparison of supercell
sizes of 54 atoms, 128 atoms and 250 atoms). The d
parameters for Nb, Zr and Ti are 0.14 Å, 0.15 Å and
0.22 Å, respectively, where Ti has the largest displace-
ment from its ideal lattice site.
By examining the 1NN bond length of dierent atom
pairs, it is shown that Zr atoms tend to have larger
bond length while Nb-containing bond length tends to
be smaller. Ti atom, as the lightest element in NbZrTi,
tends to accommodate these dierences by atomic dis-
placements. By analyzing the projected density of states
(DOS) of dierent elements, it is also shown that Ti
has a large DOS near Fermi level, which generally sug-
gests its inferior lattice instability compared to Nb and
Zr [23, 65]. Thus, large relaxation can further help sta-
bilize Ti atoms in the BCC structure. The supporting
evidence related to bond length and DOS analysis can
be found in Table S3, Fig. S4 and Fig. S5 of the Sup-
plementary Materials. In addition, there is a tendency
of forming larger atomic displacement when there are
fewer Nb atoms in the 1NN. The d-orbitals of a BCC
structure can be classified into the d-egand d-t2gcom-
ponent. The d-t2gcontribution points along the <111>
direction (direction of 1NN) and helps stabilize the BCC
structure, whereas the d-egcontribution points along the
<100>direction (direction of 2NN) [66]. The valence
charge density of NbZrTi is presented in Figure 3(a)) in
the (110) plane to visualize both 1NN and 2NN chemi-
cal bonding. Clearly, Nb atoms have a strong d-t2gcom-
ponent along the <111>direction, which is manifested
by the higher charge density between nearest neigh-
bours. In contrast, the 1NN bonding is relatively weaker
for Ti and Zr atoms. Thus, when fewer Nb atoms are
present in the surroundings, especially with 0 or 1 Nb
atom in the 1NN, the lack of stabilizing element leads
to larger local lattice distortion.
Although local lattice distortion has proved to be es-
sential in stabilizing BCC structures [67, 68], atomic
displacements from chemically-induced lattice distor-
tion could be intermixed with atomic displacements in-
duced by the tendency of structural transformation [68].
Thus, before performing systematic defect calculations,
the phonon dispersion of the perfect lattice was calcu-
lated, as shown in Fig. 3(b). Below zero frequency, there
is little imaginary component (less than 0.05 THz1
in the spectral function), showing the stability of the
studied structure. Significant phonon broadening can
be observed due to the fluctuations of mass and force
constant in the MPEAs and the extent of broadening
is also consistent with other refractory MPEAs [43].
As the structural instability is usually associated with
the transition to HCP and ωphases, three HCP and ω
phase SQS structures were constructed using 96 atoms
and 81 atoms, respectively. If the cell shape was not
fixed, all six configurations relaxed to BCC structure
after energy minimization (see Fig. S6 of the Supple-
mentary Materials for the structural transformation). If
the cell shape was fixed, the energy of HCP phase was
33 meV/atom higher than the BCC phase, whereas the
ωphase still transformed to the BCC phase, showing
the stability of the BCC structure. In addition to the
stability study at 0 K, the atom trajectories from vibra-
tion were also examined at 300 K, 600 K and 900 K
(see Fig. 3(c)). If there is a tendency of phase trans-
formation, the atom positions will exhibit a large devia-
tion from the ideal BCC lattice sites and show a consis-
tent shift along a specific direction for all atoms, as evi-
denced in the case of TiZrHf [68]. For NbZrTi, most of
the atoms vibrate around BCC lattice sites with a circu-
lar distribution profile within the projection plane. Only
few atoms (3 out of 128 atoms) show a deviation from
the ideal lattice site at 300 K, and this localized insta-
bility is greatly reduced at 600 K and 900 K. Based on
these analyses, we show that the constructed SQS struc-
ture is stable with little dynamical instability.
Based on the parameter of d, the extent of local
lattice distortion in NbZrTi is larger than BCC NbTiV
and VTaCrW, and similar to AlNbTiV and HEAs from
the HfNbTaTiZr-based alloy system [23, 61]. Although
point defect properties depend on the physical and elec-
6
[001]
[110]
1.0
0.0
0.2
0.4
0.6
0.8
e3
(a)
(c) 300 K 600 K 900 K
Nb
Zr
2 NN
1 NN
(b)
Figure 3: (a) Valence charge density of NbZrTi and the corresponding atom types in the (110) plane. The partial charge density from Ef- 5 eV
to Efwas selected, where Efis the Fermi energy. (b) Phonon band spectrum of NbZrTi. (c) Spatial distribution of atomic positions of NbZrTi at
a temperature of 300 K, 600 K and 900 K. Atom positions at each timestep are projected in the (001) plane with four atoms aligned in the same
column.
tronic properties of the constituent elements, the local
lattice distortion has a great impact on the structural sta-
bility and energy landscape, which also influences the
point defect behaviours.
3.2. Vacancy
Based on first-principles calculations, the vacancy
formation energies (Ev
f) of Nb, Zr and Ti are
0.97 eV ±0.39 eV, 0.95 eV ±0.36 eV and
0.92 eV ±0.29 eV, respectively, as shown in Fig. 4(a).
There is a wide range of energies for each element and
the average energy is approximately equal. The vacancy
formation energy is much smaller than those of its con-
stituent elements (2.77 eV from BCC Nb, and 2.00 eV
and 2.01 eV from Zr and Ti of HCP structure).
Interestingly, in 24% of the relaxed vacancy struc-
tures, the initial vacancy site was filled by a neighbour-
ing atom after energy minimization. For one instance,
the new vacancy site was subsequently filled by its own
neighbouring atom again. To our best knowledge, this
type of instability of vacancy site was not reported in
other MEAs or HEAs. There is no clear dierence in
the local chemical environment around these unstable
vacancy sites when the number of each element in the
1NN and 2NN shell is assessed. In addition, these sites
are distributed randomly in the studied structure with
no clear spatial aggregation. For comparison, vacancy
relaxation was also performed in FCC HEA NiCoFe-
CrMn and no such instability was identified. By com-
paring the atomic force of each atom in the unrelaxed
vacancy structures (see Fig. 4(b) and (c)), it is found that
the atomic forces have similar distributions and magni-
tudes between NbZrTi and NiCoFeCrMn. One might
expect that the extreme values of the atomic force in the
unrelaxed structure, corresponding to the neighbouring
atoms of the vacancy site, is higher in NbZrTi due to the
larger lattice distortion. However, it is shown here that
the dierence in the initial atomic force is not signifi-
cant between NbZrTi and NiCoFeCrMn, although two
dierent structures are compared. In addition, the un-
stable atom that fills into the vacancy site does not nec-
essarily experience a large initial atomic force, which
could be quite low in some cases (see Fig. 4(b)). The
spontaneous replacement can occur when the path to the
7
(a)
(d) (e)
NbZrTi
NiCoFeCrMn
NiCoFeCrMn
NbZrTi
(b) (c)
Figure 4: (a) Vacancy formation energy of dierent types of atoms in NbZrTi (blue error bar: standard deviation; gray error bar: maximum and
minimum). (b) Atomic force of each atom in NbZrTi in the unrelaxed structure of Zr vacancy, where the unstable atom refers to the neighbouring
atom that fills into the vacancy site during the structural relaxation. In total, 15 Zr vacancy structures were simulated. Results for Nb and Ti
vacancies are also included in Fig. S7 of the Supplementary Materials. (c) Atomic force of each atom in NiCoFeCrMn in the unrelaxed vacancy
structure. (d-e) Average displacement per atom in NbZrTi and NiCoFeCrMn after removal of a vacancy.
vacancy site has no energy barrier regardless of the mag-
nitude of the atomic force. Therefore, it is the unusual
local potential landscape that leads to the spontaneous
replacement of the vacancy site by the neighbouring
atoms. The large lattice distortion changed the poten-
tial energy landscape significantly from a lattice-based
fluctuation, completely suppressing vacancy migration
barriers between certain neighbouring sites, while the
slope of the potential of these unusual local sites were
not greatly enhanced.
By comparing the atomic displacement during the va-
cancy relaxation using Eq. (6), it is shown from Fig. 4(d)
and (e) that NbZrTi experienced much larger structural
adjustment. The average atomic displacement is 6×
larger than that of NiCoFeCrMn. The overall larger de-
gree of structural adjustment may lead to the lower Ev
f
of NbZrTi compared to that of its constituent elements.
One clear evidence is that for unstable vacancy struc-
tures with spontaneous exchange with a neighbouring
atom, the average Ev
fis 0.69 eV ±0.29 eV, which is
lower than the overall average Ev
fof 0.95 eV ±0.34 eV.
The removal of an atom can trigger a large extent of
system relaxation, which lowers the total system en-
ergy and Ev
f. However, there is no strong correlation
between individual Ev
fand the total relaxed distance of
the system because the contribution of atomic displace-
ment to the lowering of the system energy also depends
on the atom locations relative to the vacancy site. The
relationships between Ev
fand the number of each ele-
ment in the 1NN shell, the number of each element in
the 1NN and 2NN shell, Bader volume of the removed
atom and Bader charge of the removed atom all show
weak correlations (see Fig. S8 - Fig. S10 of the Sup-
plementary Materials for the dependence of Ev
fon these
parameters). Further exploration is needed to determine
whether a single parameter or combined parameters can
be used to estimate the individual vacancy formation en-
ergy. Here, we show that due to the large local lattice
distortion and local chemical complexity, the structural
relaxation is a complicated process and depends heavily
on its local environment.
In previous studies of MPEAs, no significant de-
crease of Ev
fwas observed. The average Ev
fof NiFeM-
nCr HEA is 1.96 eV [21], which is comparable to its
constituent elements of Ni [69, 70], Fe [69, 71], Cr [69]
and Mn [71]. In the Ni–containing MPEAs, the aver-
age Ev
fis equal or slightly higher than pure Ni [20]. In
BCC equiatomic VTaCrW, the Ev
falso has a large en-
8
(a) (b)
(c) (d)
Figure 5: (a) Distribution of vacancy migration energies in NbZrTi. (b) Correlation between the forward and backward migration energy. (c)
Vacancy migration energies for paths to all the 1NN sites for three dierent Nb vacancy sites. (d) Representative vacancy migration barriers. The
corresponding transition structures can be found in Fig. S11 of the Supplementary Materials.
ergy spread, and the average Ev
fis slightly lower than
pure W, but higher than pure Ta [23]. Thus, there seems
no clear trend on the relative magnitude of vacancy for-
mation energy compared to pure elements based on pre-
vious studies, and such low Ev
fwas not observed.
First, lower vacancy formation energy in NbZrTi
leads to a higher equilibrium vacancy concentra-
tion [55]:
Cv=exp Sv
f
k!exp Ev
f
kT !,(7)
where Sv
fand Ev
fare vacancy formation entropy and en-
ergy, respectively. For example, for a temperature of
1000 K, the exp (Ev
f/kT ) term is 109higher than that
of pure Nb. Here, energy calculations were performed
at 0 K and we assume that Sv
fdoes not decrease sig-
nificantly relative to its constituent elements [72]. Fur-
ther evaluation of Sv
fis necessary by taking into account
dierent entropy components [73–76]. Recent studies
have shown that for a perfect lattice, the change of vi-
bration entropy is of the same magnitude as that of con-
figurational entropy among dierent BCC MPEAs [43],
and entropy can be important in predicting phase sta-
bility of binary random solid solutions [77]. From the
perspective of Cv, the equivalent average Ev
fshould be
actually lower than the arithmetic mean due to the wide
distribution of Ev
fand the tendency of vacancies to form
near low-energy neighborhoods [73]. In addition to the
impact on Cv, lower vacancy formation energy also in-
dicates that less energy is required to produce a vacancy,
which means that vacancy defects can be more easily
formed.
These two dierences have implications for materials
used for nuclear applications. First, radiation-induced
swelling is a major concern for long-term mechanical
integrity of structural materials. High equilibrium va-
cancy concentration promotes the thermal emission of
vacancies by void, reducing the extent of swelling at
high temperature [55]. Second, it is believed that dur-
ing the initial stage of irradiation when the displacement
cascade takes place, the reduced heat dissipation from
MPEAs can result in enhanced defect recombination,
leading to a lower number of surviving defects after the
thermal spike period [26, 29, 78]. If the defect forma-
tion energies are much lower, the same imparted energy
from the incident radiation can create a higher number
9
of defects during the thermal spike period, which com-
petes with the aforementioned mechanism of enhanced
defect recombination.
Figure 5(a) shows the distributions of vacancy migra-
tion energies (Ev
m) in NbZrTi. All three elements have
similar wide energy distributions with the lower end ex-
tending to nearly 0 eV. This is consistent with previ-
ous calculations of vacancy formation energy with no
migration barrier. There is a weak correlation between
the forward and backward migration energy, showing a
Pearson’s correlation coecient of 0.22 (see Fig. 5(b)).
We need to emphasize that there is a distinct dierence
from FCC MPEAs, which usually have stronger positive
linear correlations between the forward and backward
migration barriers [21, 23].
Figure 5(c) lists the migration barriers to all eight
neighbouring sites at three dierent Nb vacancy sites.
At the same vacancy site, depending on the migration
direction, the energy barrier varies substantially. For in-
stance, the barrier can be 1.4 eV in one direction and
0.1 eV in a dierent direction. At site 3, several mi-
gration paths were unstable, where the final states were
spontaneously relaxed to the initial state. Representa-
tive migration paths were presented in Fig. 5(d): the
path 1 and 2 are conventional barrier profiles with good
symmetry; the path 3 and 4 have large dierences be-
tween forward and backward migration energies; the
path 5 and 6 exhibit irregular barrier shapes. All these
results indicate that due to the large lattice distortion,
the energy landscape could dier significantly at a very
local scale. The site-to-site variation is more promi-
nent compared to FCC MPEAs and even some BCC
MPEAs [23]. For future material modeling at a larger
scale, such as MD and kinetic Monte Carlo (KMC) sim-
ulations, these characteristics need to be taken into ac-
count to reflect the distinct defect behaviours of NbZrTi.
Due to the wide distribution of the vacancy migration
energy, using the arithmetic mean to represent the av-
erage migration energy could potentially bias the actual
migration behaviours. For each vacancy site, assum-
ing only the vacancy jumps through the 1NN, there are
eight possible migration pathways. Since the jump fre-
quency scales with Ev
m/kT , the vacancy will preferen-
tially migrate through sites of low Ev
m. A simple KMC
algorithm was used to estimate the equivalent Ev
m. For
each migration event, the number of each type of atom
in the 1NN shell was randomly generated and Ev
mof
the corresponding atom type was randomly assigned to
these sites based on the first-principles results. This is
supported by the weak correlation among vacancy mi-
gration energies at the same site (see Fig. 5(c)). The mi-
gration event and step time were then determined based
on the Monte Carlo method [79]. Because the corre-
lation between the forward and backward Ev
mis weak
(see Fig. 5(b)), we assume that each jump is indepen-
dent from its previous jump. In total, 106jumps were
computed. Here, we assume that the Debye frequency
and vacancy migration entropy do not vary, and we only
focus on the impact of the variation of Ev
m. The KMC-
equivalent mean was then determined by equating the
total time and total number of jumps of each element.
As shown in Table 2, the arithmetic mean overesti-
mates the average Ev
mof NbZrTi. This is expected be-
cause it is highly likely that the migration takes place
quickly through the low Ev
msites among all eight neigh-
bouring sites. Since the method of arithmetic mean
gives equal weight to large Ev
mvalues, it could give
inaccurate Ev
morder among dierent elements, which
aects the analysis related to elemental diusion and
segregation. The KMC method is less sensitive to ex-
treme values from the limited data set, as each jump
event is determined collectively from eight dierent
Ev
mvalues. Based on the KMC method, the average
Ev
mis approximately 0.35 eV, which is smaller than
0.45 eV from pure Nb. For pure Zr, the Ev
mvalues are
0.54 eV and 0.63 eV for the basal and non-basal migra-
tion, respectively [80]. For pure Ti, the basal and non-
basal Ev
menergies are 0.40 eV and 0.53 eV [81]. Al-
though Zr and Ti have a dierent crystal structure, we
show here that the Ev
mof NbZrTi is smaller than all of
its constituent elements under ambient conditions. We
note that in NbZrTi, there is a small dierence in Ev
m
among Nb, Zr and Ti. Due to the large energy spread of
Ev
m, more statistics are needed to fully characterize the
relative order of Ev
mamong dierent elements.
3.3. Interstitial
Dierent initial structures including <111>dumb-
bells, <110>dumbbells, <100>dumbbells and tetra-
hedral sites were simulated, as described in Section 2.1.
Among all the configurations, 61% of the interstitials
moved to a dierent lattice site by rotating their direc-
tions and replacing other neighbouring atoms. The in-
terstitial instability originates from the irregular energy
landscape induced by the large lattice distortion. In the
work of Zhao et al. [23], unstable interstitial configura-
tions have also been observed in VTaCrW but were not
counted in the calculation of Ei
fdue to the involvement
of antisite exchange. In this work, those configurations
were included in the calculation of Ei
fusing Eq. (1). We
have them considered because of two factors: (1) they
are representative events in NbZrTi, and (2) they have
stable final states based on energy minimization.
10
Table 2: Average vacancy migration energy based on the KMC-equivalent method and the arithmetic mean. The jump percentage of each type of
atom is normalized by the total number of jumps from the KMC method.
Atom type KMC-equivalent mean at 1000 K (eV) Jump percentage (%) Arithmetic mean (eV) Jump percentage (%)
Nb 0.352 33.3 0.526 ±0.506 4.8
Zr 0.348 34.5 0.378 ±0.298 20.0
Ti 0.354 32.2 0.498 ±0.388 6.5
(a) (b) (c)
Figure 6: (a) Interstitial formation energies for dierent final interstitial configurations (blue error bar: standard deviation; gray error bar: maximum
and minimum). (b) Occurrence probabilities of dierent types of final interstitial configurations. (c) Occurrence probabilities of dierent elements
in the final interstitial configurations.
As shown in Fig. 6(a), in the final configurations,
Ti-Ti dumbbell has the lowest formation energy, fol-
lowed by Ti-Nb and Ti-Zr. Other types are not listed
individually due to their limited number of occurrences.
The average Ei
fis 1.92 eV ±0.39 eV, much lower than
4.38 eV of the most stable <111>dumbbell of pure
Nb. For pure Zr and Ti with HCP structure under
ambient conditions, according to dierent studies, the
Ei
fvalues for the most stable interstitial configuration
are in the range of 2.70 eV 3.03 eV and 2.27 eV
2.42 eV, respectively, which are also larger than that of
NbZrTi [39, 40, 80, 82]. The occurrence probabilities
of interstitial configurations with low Ei
fhave a higher
tendency to form (see Fig. 6(b) and (c)). The occurrence
probability here corresponds to static calculations with
the same number of each type of atom being added in
the initial structures. For dynamic simulations, such as
cascade collision simulations for radiation damage stud-
ies, we still expect a higher occurrence probability of
Ti-containing interstitials and a lower probability of Zr-
containing interstitials based on Ei
f. However, we note
that the defect formation dynamics is not reflected from
static calculations.
The preferred dumbbell directions are dierent for
dierent types of atoms, as presented in Table 3. For
example, the Nb-Ti interstitial can form both <110>
Table 3: Preferred directions for dierent types of dumbbell intersti-
tials.
Dumbbell Preferred Occurrence Second-preferred Occurrence
type direction probability direction probability
Nb-Ti <110>9/17 <111>7/17
Ti-Ti <110>13/14 <100>1/14
Zr-Ti <110>9/12 <113>2/12
Nb-Nb <111>5/6<012>1/6
Zr-Nb <111>1/2<110>1/2
Zr-Zr <013>1/1
Table 4: Stable final interstitial configurations at the same initial lat-
tice site with dierent initial orientations for a Nb-Nb dumbbell. The
corresponding final structures are provided in Fig. S12 of the Sup-
plementary Materials (Lc: lattice parameter, Dinitialfinal : distance be-
tween the initial and final interstitial site).
Initial dumbbell Final dumbbell Dinitialfinal Ei
f(eV)
direction configuration
[111], [110], [101], [011] Ti-Ti [010] 3Lc1.82
[111], [011] Ti-Ti [011] 3Lc1.79
[111], [110], [101] Ti-Ti [011] 23Lc1.31
[111] Nb-Ti [111] 1.53Lc1.61
and <111>dumbbells, but Ti-Ti and Nb-Nb pair have
a strong preference for <110>dumbbell and <111>
dumbbell, respectively. Due to the large distortion of
the lattice, unusual dumbbell directions can be formed.
11
For the same initial interstitial site with dierent dumb-
bell directions, the final configurations can be dierent,
as illustrated in Table 4. The location of the final inter-
stitial site can be far away from the initial site and the Ei
f
of the stable interstitials can also be dierent. For nu-
clear materials, the interstitial structures and formation
energies have a great impact on the dislocation loop nu-
cleation, which further impacts the irradiation harden-
ing and defect microstructure evolution [55, 83–86].
The interstitial migration behaviours were studied by
AIMD simulations at a temperature of 1200 K. NbZrTi
has a lower MSD than pure Nb due to the more tor-
tuous migration pathways (see Fig. 7). The diusion
coecients for Nb and NbZrTi are 1.02 ×1010 m2/s
and 3.88 ×1011 m2/s, respectively, according to the
Einstein relation [87]. The interstitial atom in pure
Nb can form long one-dimensional (1D) diusion path
along <111>direction, whereas the interstitial atom in
NbZrTi diuses in a three-dimensional (3D) manner
with more frequent changes of directions. The 3D mi-
gration behaviour leads to a larger cross section of in-
teractions with defects of low mobility, such as vacancy
and vacancy clusters, and enhances the defect recombi-
nation [88]. The jump rate for NbZrTi is actually 9.6%
higher than that of Nb, but due to the 3D spatial migra-
tion behaviour, the diusion coecient is only 38% of
Nb at 1200 K.
The correlation factor fcis used to examine the angu-
lar correlation between consecutive jumps [19, 88]:
fc=1+cos θ
1cos θ,(8)
where θis the angle between two consecutive jumps and
fcis calculated by averaging cosθover all trajectories.
Nb and NbZrTi have a fcof 0.19 and 0.20, respectively,
showing negligible dierence. A value between 0 and
1 means that there is a tendency for the next jump to
return back to the initial location of the previous jump.
Although the average angular correlation between con-
secutive jumps is similar, two mechanisms contribute to
the dierence in diusion coecients between Nb and
NbZrTi. First, interstitial atoms in NbZrTi have their
own preferred dumbbell directions and have a higher
probability to rotate their orientations (see Table 3).
Therefore, it is more dicult to maintain the same 1D
<111>direction. Second, certain lattice sites can act as
“trap sites” for interstitial atoms and make it dicult for
interstitials to escape their surroundings. Based on the
AIMD trajectories, there are six lattice sites in NbZrTi
that an interstitial spends longer time at or returns to
for more number of times compared to the maximum
case from pure Nb. In addition to the stability of these
sites, we also conjecture that as a large fraction of in-
terstitial structures are unstable at their initial sites, the
distance between stable interstitial structures is farther
than the regular atomic spacing from conventional ma-
terials. Once an interstitial is trapped at a stable site,
it is more dicult to escape due to the longer distance
from the neighbouring stable site. However, this needs
further confirmation by simulations at a larger scale.
In NbZrTi, Ti diuses much faster than Nb and Zr,
as shown in Fig. 8. By examining the time fraction and
number fraction of each type of interstitial, we find that
Ti-Ti dumbbell has the highest occurrence probability,
followed by Ti-Nb and Ti-Zr. In contrast, Nb-Zr and
Zr-Zr dumbbells have the lowest probability to form.
This is in good agreement with the order of Ei
f(see
Fig. 6), where Ti-Ti dumbbell has the lowest Ei
fand
Zr-containing interstitials tend to have higher Ei
f. For
each type of interstitials, the time fraction and number
fraction have similar values because dierent interstitial
pairs have similar distributions of dwelling time at each
site (see Fig. S13 of the Supplementary Materials for the
dwelling time distribution of dierent interstitial pairs).
According to the model for radiation-induced segrega-
tion [89], assuming the vacancy diusion behaviours are
identical for dierent alloy elements, a strong binding
of interstitials to a certain type of atom leads to a higher
tendency of its segregation near defect sinks, such as
surface or grain boundary [55]. Here, Ti atom is prefer-
entially associated with the interstitial, which can be ex-
plained by its positive binding energy, i.e., the energy of
converting one type of interstitial to the preferred type
of interstitial [55]. Thus, Ti atom diuses faster through
the interstitial diusion mechanism. The overall segre-
gation behaviour is then determined collectively by the
interstitial and vacancy diusion mechanism [90–92].
Attempts have been made to tune the interstitial dif-
fusion behaviours through changing the elemental com-
position and local atomic bonding. The first attempt
is to reduce Ti composition to reduce the diusion of
Ti. When the composition changes from equiatomic
NbZrTi to Nb0.4Zr0.4Ti0.2, as shown in Fig. 9(a) and (b),
the total diusion coecient is reduced due to the de-
crease of Ti content and the corresponding decrease of
Nb and Ti diusion coecients. The Ti-Ti dumbbell be-
comes less easily formed due to the reduced Ti compo-
sition. Accordingly, the proportions of Ti-Nb and Ti-Zr
dumbbells become higher (see Fig. 9(c)). As the prefer-
ential Ti diusion is promoted by the preferential inter-
stitial binding of Ti-Ti pair, Ti diusion coecient be-
comes lower. Although the fraction of Ti-Nb dumbbell
becomes higher, the Ti atom in Ti-Nb dumbbell is more
12
(a) (b)
NbZrTi
Nb
Figure 7: (a) Mean square displacement of interstitial migration in NbZrTi and Nb at 1200 K. (b) Representative trajectories of interstitial pairs in
NbZrTi and Nb with the same diusion time scale. The interstitial trajectories are unwrapped from the simulation box with periodic boundaries
and are enclosed here in boxes of the same length scale.
(a)
(b)
Figure 8: (a) Mean square displacement of dierent elements in
NbZrTi at 1200 K. (b) Time fraction and number fraction of inter-
stitial sites of dierent dumbbell types.
active and the Nb atom is relatively immobile, which
leads to the lower Nb diusion coecient.
The second attempt is to study the impact of SRO
structure while keeping the same equiatomic composi-
tion. The SRO structure was created by a combined
Metropolis MC and density functional theory (DFT)
scheme, as described in Section 2.1. The overall dif-
fusion coecient is reduced due to the decrease of Ti
and Nb diusion coecients (see Fig. 9(d) and (e)).
Fig. 1(b) shows that the SRO structure has a strong ten-
dency of Nb-Nb clustering and a moderate tendency of
Zr-Ti clustering. Interestingly, the interstitial prefers to
migrate in the Ti- and Zr-enriched region as revealed by
the diusion trajectories, which leads to the lower diu-
sion coecient of Nb. This can also be reflected by the
decrease of Ti-Nb dumbbell fraction and the increase of
Ti-Zr dumbbell fraction (see Fig. 9(f)). As the intersti-
tial is mostly confined in the Ti- and Zr-enriched region,
the long-range diusion becomes less favoured, which
results in the decrease of elemental and total diusion
coecients. The ordering tendencies among dierent
atom pairs all have an impact on the interstitial migra-
tion behaviour. In contrast to these two structural tuning
methods, another attempt by simply reducing the per-
centage of the fast diusion species of Ti atom in the
1NN shell shows no improvement for the reduction of
Ti and total diusion coecients (see Section 2 of Sup-
plementary Materials for the comparison).
3.4. Ion irradiation experiment
To study the impact of point defect energy on
radiation-induced void swelling, ion irradiation exper-
iment was performed at 675 °C with 3-MeV Fe ions at a
fluence of 1.02 ×1017 ions/cm2, corresponding to a peak
dose of 99.8 dpa and 129 dpa in Nb and NbZrTi, respec-
tively. Void swelling is known to be pronounced at in-
termediate homologous temperature (T/Tm) because the
void growth is limited by the low defect mobility at low
temperature and is limited by the high vacancy super-
saturation concentration at high temperature [55]. As
explained in Section 3.2, the vacancy formation energy
13
SRO NbZrTi
(MC+DFT)
Nb0.4Zr0.4Ti0.2
(a) (b) (c)
(d) (e) (f)
Figure 9: (a-b) Comparison of MSD of interstitial migration between NbZrTi and Nb0.4Zr0.4Ti0.2. (c) Number fraction of dierent dumbbell types
in Nb0.4Zr0.4Ti0.2. (d-e) Comparison of MSD between SQS-based NbZrTi and NbZrTi with a SRO structure based on the MC +DFT method.
(f) Number fraction of dierent dumbbell types in NbZrTi with a SRO structure based on the MC +DFT method. The dotted lines in (c) and (f)
represent the results from the SQS-based NbZrTi structure.
0 500 1000 1500 2000 2500 3000 3500 4000
0
40
80
120 dpa
Fe
Depth (nm)
dpa
0.0
1.5
3.0
4.5
Fe (at.%)
20 nm
(e) (f)
(b) (c)
20 nm
20 nm 20 nm
void void
400-520 nm
950-1070 nm
400-520 nm
1100-1220 nm
(a) (d)
0 500 1000 1500 2000 2500 3000 3500 4000
0
40
80
120
Fe (at.%)
dpa
Fe
Depth (nm)
dpa
0.0
1.5
3.0
1 µm1 µm
Figure 10: (a) Cross-sectional STEM-BF image of Nb irradiated at 675 °C. Depth profiles of the dpa and Fe ion concentration predicted by the
SRIM code are shown. (b-c) BF TEM images of void formation in the irradiated regions of pure Nb at two dierent depths acquired at under-focused
condition. (d) Cross-sectional STEM-BF image of NbZrTi irradiated at 675 °C along with the depth profiles of dpa and Fe ion concentration. (e-f)
No formation of voids and visible defect microstructures in the irradiated NbZrTi.
of NbZrTi is much lower than that of pure Nb, the equi-
librium vacancy concentration should be significantly
higher, limiting the void formation at high temperature.
According to previous experiments [93–97], voids can
be formed in Nb by heavy ion irradiation from 650 °C
to 1120 °C, corresponding to a T/Tmof 0.34 to 0.51.
With 3.3-MeV 58Ni+ions, the maximum void swelling
occurred at a temperature of 1010 °C, corresponding to
14
a T/Tmof 0.47 [93]. Based on these previous eorts,
an irradiation temperature of 675 °C was selected be-
cause in addition to the expected void formation in Nb at
this temperature, a T/Tmof 0.48 in NbZrTi [98] is also
in the normal void swelling temperature window [55].
Furthermore, this temperature falls into the operating
temperature range of fourth-generation nuclear reactors
including gas-cooled fast reactor and molten salt reac-
tor [8, 9].
As shown in Fig. 10(a), in the irradiated region of
pure Nb, voids and dislocation loops were formed by
agglomeration of vacancies and interstitial atoms. Dis-
location loops were also formed outside the SRIM-
predicted ion range by diusion of radiation-induced
defects at high temperature. A great number of voids
were observed throughout the irradiation area with an
average diameter of 5.5 nm ±1.7 nm (see Fig. 10(b) and
(c)). In contrast, no voids and other radiation-induced
defect microstructures were observed in NbZrTi, as
shown in Fig. 10(d) (f). The dislocation lines at a
depth range of 1500 nm 2500 nm correspond to pre-
irradiation defects. The energy dispersive spectroscopy
(EDS) mapping also showed that the spatial distribu-
tions of Nb, Zr and Ti are uniform in the irradiated
and unirradiated region, and the atomic composition re-
mains equiatomic (see Fig. S14 and Table S4 of the Sup-
plementary materials for the EDS mapping results). As
expected, the higher equilibrium vacancy concentration
limits the growth of voids, where the vacancy influx to
voids are counterbalanced by the thermal vacancy emis-
sion [55].
In addition, the enhanced recombination of point
defects contribute to the observation of reduced de-
fects. As reported for FCC Ni-based MPEAs and BCC
VTaCrW [20, 23], the recombination probability of
point defects should be higher than pure metals due to
the closer mobility between the vacancy and interstitial
atom. When the point defect recombination is higher,
there are less surviving defects that can further agglom-
erate into large defect microstructures, such as voids and
dislocation loops. This has been considered as one of
the important mechanisms for the enhanced radiation
resistance of HEAs [20, 26]. The same mechanism ap-
plies to NbZrTi because of the lower average vacancy
migration energy (see Table. 2) and lower interstitial dif-
fusion coecient (see Fig. 7), i.e, higher vacancy mobil-
ity and lower interstitial mobility. However, with lower
vacancy and interstitial formation energy, more vacan-
cies and interstitial atoms are created during the ther-
mal spike period. The relevant importance of these two
competing mechanisms needs to be further evaluated.
4. Discussions
Based on first-principles calculations, NbZrTi ex-
hibits distinctive point defect properties from pure
Nb. The average vacancy formation energy is
0.95 eV ±0.34 eV compared to 2.77 eV in Nb. The av-
erage interstitial formation energy is 1.92 eV ±0.34 eV,
whereas the interstitial formation energies in Nb are
4.38 eV 5.12 eV for dierent interstitial configura-
tions. In NbZrTi, the interstitial formation energy has a
strong dependence on the dumbbell atom types, which
can be clearly seen from Fig. 6(a). The dependence on
the dumbbell orientation is insignificant relative to the
site-to-site variation. For example, for a final configura-
tion of <110>dumbbell and <111>dumbbell, the aver-
age energies are 1.83 ±0.37 eV and 2.07 ±0.49 eV, re-
spectively. The dierence could also be attributed to the
preferential formation of <110>dumbbell of the most
stable Ti-Ti pair. According to the KMC-equivalent
method computed at 1000 K (see Table 2), the aver-
age vacancy migration energy is 0.35 eV in NbZrTi,
which is lower than 0.45 eV from pure Nb. Based on
AIMD simulations, the interstitial diusion coecient
in NbZrTi is 38% of that in Nb at 1200 K. In summary,
NbZrTi has much lower vacancy and interstitial forma-
tion energies, higher vacancy mobility and lower inter-
stitial mobility.
The instability of defect structures was observed in
both vacancy and interstitial configurations. In the va-
cancy structures, it was manifested by the spontaneous
replacement of the vacancy site by a neighbouring atom,
whereas in the interstitial structure, it was manifested by
the large relaxation distance of the interstitial atom away
from the initial interstitial site (see Table 4). This can
be attributed to the severe lattice distortion of NbZrTi
that leads to an irregular energy landscape (see Fig.5(d))
and the resulting suppression of migration barriers at
local sites. The lower vacancy formation and inter-
stitial formation energies can also be explained by the
large degree of relaxation resulting from such an irreg-
ular energy landscape. The BCC VTaCrW, which has
only interstitial instability, has a local lattice distortion
parameter dof 0.087 Å [23], which is smaller than
0.171 Å from NbZrTi. It is also reported that BCC
MEA VTaTi and VTaW have lower interstitial forma-
tion energies due to the large local relaxation based on
classical molecular static simulations [33]. For com-
parison, the dvalues of VTaTi and VTaW were cal-
culated by first-principles calculations and were found
to be 0.132 Å and 0.061 Å, respectively, which are
also smaller than that of NbZrTi. By comparing dif-
ferent BCC MPEAs, there seems to be a trend that with
15
the increase of lattice distortion, the structural instabil-
ity extends from interstitial structures to vacancy struc-
tures and leads to the decrease of the corresponding de-
fect formation energy. NbZrTi-based MPEAs, such as
NbZrTiTaHf and NbZrTiHf, all have similar or even
larger dvalues, ranging from 0.16 Å– 0.20 Å [61].
Thus, it is worth exploring further whether these alloys
have similar low point defect formation energies and
large extent of defect structural relaxation.
As explained previously, it was proposed that the
point defect recombination in MPEAs is higher than
pure elements under irradiation due to the closer mo-
bility between vacancies and interstitials [20]. This was
largely attributed to the slower diusion of interstitial
atoms and the overlap between vacancy and interstitial
migration energy distributions. Here, due to the ex-
tremely large distribution of vacancy migration energy
and also the skewed distribution profile towards low en-
ergy (see Fig. 5(a)), the vacancy migration can occur
much more easily because the probability of migration
through a low migration barrier is high. Therefore, the
higher vacancy mobility can further enhance the point
defect recombination under irradiation. It is noted that
higher mobility of vacancy can also promote the cluster-
ing of vacancies by increasing the vacancy arrival rate at
void embryos [55]. Considering also the other distinct
point defect properties, the overall impact on irradiation
resistance could dier significantly from FCC MPEAs
and needs to be further evaluated with MD simulations
and irradiation experiments.
Regarding the diusion capabilities of dierent ele-
ments, Nb, Zr and Ti have similar vacancy migration
energies (see Table 2), and the relative energy dier-
ence cannot be clearly distinguished. On the other hand,
interstitial diusion through Ti is much faster than Zr
and Ti (see Fig. 8) due to the preferential binding of in-
terstitials with Ti atom. The reduction of Ti content in
NbZrTi and the creation of low-energy SRO structure
based on the lattice MC method are both eective in de-
creasing the Ti and total diusion coecients. A slower
diusion coecient is usually associated with more
trapping around certain lattice sites or within certain lo-
cal regions, leading to a lower probability of long-range
diusion. When there is a strong short-range clustering
tendency of atom pairs in NbZrTi, the diusion of in-
terstitials is preferred in regions where Zr and Ti are en-
riched, which limits the long-range interstitial diusion.
The existence of SRO structure has been confirmed ex-
perimentally and has shown great impact on the me-
chanical and radiation properties of MPEAs [99–101].
In this work, a SQS model with no SRO was used to cal-
culate the defect properties, but the actual local ordering
tendency in BCC MPEAs could complicate the defect
properties, which needs additional investigation [102].
The obtained first-principles results are useful for the
validation of interatomic potentials. A systematic com-
parison of point defect properties of NbZrTi, including
vacancy formation energy, vacancy migration energy
and interstitial formation energy, was demonstrated with
an EAM potential with potential parameters taken from
Ref. [103, 104] (see Section 3 of Supplementary Mate-
rials for the detailed analysis). Although a quantitative
agreement was not fully reached, several conclusions or
implications can be drawn from such comparison. First,
accurate defect energies from pure metal elements need
to be guaranteed. For instance, the underestimation of
interstitial formation energy of pure Nb leads to the
underestimation of Nb-containing interstitial formation
energies in NbZrTi. Second, the EAM potential with a
relatively simple formalism is able to capture some im-
portant characteristics of NbZrTi, including lower va-
cancy and interstitial formation energy, unstable initial
vacancy and interstitial structures, as well as wide en-
ergy distribution in vacancy migration energy. Third,
the currently used EAM potential underestimates the
lattice distortion and overestimates the defect formation
energies. This seems aligned with our conjecture that
the decrease of defect formation energies is correlated
with the increase of lattice distortion. A good agreement
on lattice distortion relies on an accurate description of
cross interactions among principal elements.
According to the ion irradiation experiment at
675 °C, we showed that the concept of MPEAs can
greatly modify the point defect properties and lead to a
dierent response to the irradiation temperature. Com-
pared to void swelling in Nb at both an absolute tem-
perature of 675 °C and a homologous temperature of
0.48 [93], negligible radiation-induced defects were ob-
served in NbZrTi under the same condition. The re-
duction in swelling at high temperature is aligned with
the high equilibrium vacancy concentration originating
from the low vacancy formation energy. The increase of
the point defect recombination also plays an important
role in the reduction of radiation-induced defects. The
characteristics of radiation-induced defects at a range
of temperatures and simulations of defect cluster be-
haviours need to be studied in the future to fully un-
derstand the irradiation response of NbZrTi.
5. Conclusions
First-principles calculations were performed to assess
the vacancy and interstitial formation energies and their
16
migration behaviours in body-centered cubic medium-
entropy alloy NbZrTi. Due to the irregular energy land-
scape induced by the large local lattice distortion, un-
usual structural instability was observed in both va-
cancy and interstitial structures. Compared to pure Nb,
NbZrTi exhibits much lower vacancy and interstitial for-
mation energies, higher vacancy mobility and lower in-
terstitial mobility. Because of the wide energy distri-
bution and the preferential vacancy migration through
low-energy sites, the equivalent vacancy migration en-
ergy is lower than pure Nb. The slower interstitial dif-
fusion and the chemically-biased diusion among con-
stituent elements are highly related to the stability of
dierent interstitial configurations. The atomic compo-
sition and chemical short-range order can influence the
elemental and total interstitial diusion behaviour. The
obtained first-principle results are useful for the bench-
mark of classical interatomic potential for defect stud-
ies. The ion irradiation experiment at 675 °C with 3-
MeV Fe ions at a total dose of 100 dpa showed re-
duced void swelling in NbZrTi, which is consistent with
the high equilibrium vacancy concentration induced by
the low vacancy formation energy and enhanced point
defect recombination induced by the closer mobility be-
tween vacancies and interstitials.
Acknowledgement
This work is supported by the National Key Re-
search and Development Program of China under Grant
No. 2019YFA0209900, the National Natural Science
Foundation of China under Grant No. 12075179 and
No. 12105219, the China Postdoctoral Science Founda-
tion under Grant No. 2021M702583, the Nuclear Mate-
rial Technology Innovation Center Project under Grant
No. ICNM 2020 ZH05, the Continuous Basic Scien-
tific Research Project under Grant No. WDJC-2019-
10, and the Innovative Scientific Program of CNNC. Q.
P. would like to acknowledge the support provided by
LiYing Program of the Institute of Mechanics, Chinese
Academy of Sciences through Grant No. E1Z1011001.
In addition, we would like to thank Instrument Anal-
ysis Center of Xi’an Jiaotong University for the TEM
characterization and Lanzhou Heavy Ion Accelerator
National Laboratory for the support on heavy ion ex-
periment.
References
[1] M.-H. Tsai, J.-W. Yeh, High-entropy alloys: a critical review,
Materials Research Letters 2 (3) (2014) 107–123.
[2] E. P. George, W. Curtin, C. C. Tasan, High entropy alloys:
A focused review of mechanical properties and deformation
mechanisms, Acta Materialia 188 (2020) 435–474.
[3] O. N. Senkov, D. B. Miracle, K. J. Chaput, J.-P. Couzinie,
Development and exploration of refractory high entropy al-
loys—a review, Journal of materials research 33 (19) (2018)
3092–3128.
[4] C. Lu, L. Niu, N. Chen, K. Jin, T. Yang, P. Xiu, Y. Zhang,
F. Gao, H. Bei, S. Shi, M.-R. He, I. M. Robertson, W. J. Weber,
L. Wang, Enhancing radiation tolerance by controlling defect
mobility and migration pathways in multicomponent single-
phase alloys, Nature communications 7 (1) (2016) 1–8.
[5] Y.-M. Hu, X.-D. Liu, N.-N. Guo, L. Wang, Y.-Q. Su, J.-J. Guo,
Microstructure and mechanical properties of NbZrTi and NbH-
fZrTi alloys, Rare Metals 38 (9) (2019) 840–847.
[6] X. Duan, Z. Zhang, H. He, Z. Liu, X. Duan, B. Shan, A system-
atic investigation on quaternary NbTiZr-based refractory high
entropy alloys using empirical parameters and first principles
calculations, Modelling and Simulation in Materials Science
and Engineering (2021).
[7] T. Shi, P.-H. Lei, X. Yan, J. Li, Y.-D. Zhou, Y.-P. Wang, Z.-X.
Su, Y.-K. Dou, X.-F. He, D. Yun, et al., Current development
of body-centered cubic high-entropy alloys for nuclear appli-
cations, Tungsten (2021) 1–21.
[8] P. Yvon, F. Carr ´
e, Structural materials challenges for advanced
reactor systems, Journal of nuclear materials 385 (2) (2009)
217–222.
[9] S. J. Zinkle, J. T. Busby, Structural materials for fission & fu-
sion energy, Materials today 12 (11) (2009) 12–19.
[10] Y. Wu, Y. Cai, T. Wang, J. Si, J. Zhu, Y. Wang, X. Hui, A
refractory Hf25Nb25 Ti25Zr25 high-entropy alloy with excellent
structural stability and tensile properties, Materials Letters 130
(2014) 277–280.
[11] O. Senkov, J. Scott, S. Senkova, D. Miracle, C. Woodward, Mi-
crostructure and room temperature properties of a high-entropy
TaNbHfZrTi alloy, Journal of alloys and compounds 509 (20)
(2011) 6043–6048.
[12] O. Senkov, J. Scott, S. Senkova, F. Meisenkothen, D. Miracle,
C. Woodward, Microstructure and elevated temperature prop-
erties of a refractory TaNbHfZrTi alloy, Journal of Materials
Science 47 (9) (2012) 4062–4074.
[13] Y. Wu, Y. Cai, X. Chen, T. Wang, J. Si, L. Wang, Y. Wang,
X. Hui, Phase composition and solid solution strengthening ef-
fect in TiZrNbMoV high-entropy alloys, Materials & Design
83 (2015) 651–660.
[14] S. Zhang, Z. Wang, H. Yang, J. Qiao, Z. Wang, Y. Wu, Ultra-
high strain-rate strengthening in ductile refractory high en-
tropy alloys upon dynamic loading, Intermetallics 121 (2020)
106699.
[15] Y. Wu, J. Si, D. Lin, T. Wang, W. Y. Wang, Y. Wang,
Z. Liu, X. Hui, Phase stability and mechanical properties of
AlHfNbTiZr high-entropy alloys, Materials Science and Engi-
neering: A 724 (2018) 249–259.
[16] B. Singh, J. Evans, Significant dierences in defect accu-
mulation behaviour between fcc and bcc crystals under cas-
cade damage conditions, Journal of Nuclear Materials 226 (3)
(1995) 277–285.
[17] K. Gan, D. Yan, S. Zhu, Z. Li, Interstitial eects on the
incipient plasticity and dislocation behavior of a metastable
high-entropy alloy: Nanoindentation experiments and statis-
tical modeling, Acta Materialia 206 (2021) 116633.
[18] S. Middleburgh, D. King, G. Lumpkin, M. Cortie, L. Edwards,
Segregation and migration of species in the CrCoFeNi high en-
tropy alloy, Journal of alloys and compounds 599 (2014) 179–
182.
17
[19] S. Zhao, Y. Osetsky, Y. Zhang, Preferential diusion in con-
centrated solid solution alloys: NiFe, NiCo and NiCoCr, Acta
Materialia 128 (2017) 391–399.
[20] S. Zhao, G. M. Stocks, Y. Zhang, Defect energetics of con-
centrated solid-solution alloys from ab initio calculations:
Ni0.5Co0.5 , Ni0.5Fe0.5 , Ni0.8Fe0.2 and Ni0.8 Cr0.2, Physical
Chemistry Chemical Physics 18 (34) (2016) 24043–24056.
[21] C. Li, J. Yin, K. Odbadrakh, B. C. Sales, S. J. Zinkle, G. M.
Stocks, B. D. Wirth, First principle study of magnetism and va-
cancy energetics in a near equimolar NiFeMnCr high entropy
alloy, Journal of Applied Physics 125 (15) (2019) 155103.
[22] H. Guan, S. Huang, J. Ding, F. Tian, Q. Xu, J. Zhao, Chemical
environment and magnetic moment eects on point defect for-
mations in CoCrNi-based concentrated solid-solution alloys,
Acta Materialia 187 (2020) 122–134.
[23] S. Zhao, Defect properties in a VTaCrW equiatomic high en-
tropy alloy (HEA) with the body centered cubic (bcc) structure,
Journal of Materials Science & Technology 44 (2020) 133–
139.
[24] K. Jin, C. Lu, L. Wang, J. Qu, W. Weber, Y. Zhang, H. Bei,
Eects of compositional complexity on the ion-irradiation in-
duced swelling and hardening in Ni-containing equiatomic al-
loys, Scripta Materialia 119 (2016) 65–70.
[25] C. Lu, T. Yang, K. Jin, N. Gao, P. Xiu, Y. Zhang, F. Gao, H. Bei,
W. J. Weber, K. Sun, et al., Radiation-induced segregation on
defect clusters in single-phase concentrated solid-solution al-
loys, Acta Materialia 127 (2017) 98–107.
[26] Y. Zhang, G. M. Stocks, K. Jin, C. Lu, H. Bei, B. C. Sales,
L. Wang, L. K. B´
eland, R. E. Stoller, G. D. Samolyuk, et al.,
Influence of chemical disorder on energy dissipation and defect
evolution in concentrated solid solution alloys, Nature commu-
nications 6 (1) (2015) 1–9.
[27] C. Parkin, M. Moorehead, M. Elbakhshwan, J. Hu, W.-Y.
Chen, M. Li, L. He, K. Sridharan, A. Couet, In situ microstruc-
tural evolution in face-centered and body-centered cubic com-
plex concentrated solid-solution alloys under heavy ion irradi-
ation, Acta Materialia 198 (2020) 85–99.
[28] Q. Peng, F. Meng, Y. Yang, C. Lu, H. Deng, L. Wang, S. De,
F. Gao, Shockwave generates <100>dislocation loops in bcc
iron, Nature communications 9 (1) (2018) 1–6.
[29] Y. Lin, T. Yang, L. Lang, C. Shan, H. Deng, W. Hu, F. Gao,
Enhanced radiation tolerance of the Ni-Co-Cr-Fe high-entropy
alloy as revealed from primary damage, Acta Materialia 196
(2020) 133–143.
[30] O. Deluigi, R. Pasianot, F. Valencia, A. Caro, D. Farkas,
E. Bringa, Simulations of primary damage in a high entropy
alloy: Probing enhanced radiation resistance, Acta Materialia
213 (2021) 116951.
[31] J. Byggm¨
astar, K. Nordlund, F. Djurabekova, Modeling re-
fractory high-entropy alloys with ecient machine-learned in-
teratomic potentials: defects and segregation, arXiv preprint
arXiv:2106.03369 (2021).
[32] R. Qiu, Y. Chen, X. Liao, X. He, W. Yang, W. Hu, H. Deng,
Finnis–Sinclair-type potential for atomistic simulation of de-
fects behaviour in V-Ti-Ta ternary system, Journal of Nuclear
Materials (2021) 153231.
[33] S. Zhao, Y. Xiong, S. Ma, J. Zhang, B. Xu, J.-J. Kai, Defect ac-
cumulation and evolution in refractory multi-principal element
alloys, Acta Materialia (2021) 117233.
[34] P. E. Bl¨
ochl, Projector augmented-wave method, Physical re-
view B 50 (24) (1994) 17953.
[35] M. Ernzerhof, G. E. Scuseria, Assessment of the Perdew–
Burke–Ernzerhof exchange-correlation functional, The Journal
of chemical physics 110 (11) (1999) 5029–5036.
[36] A. Zunger, S.-H. Wei, L. Ferreira, J. E. Bernard, Special quasir-
andom structures, Physical review letters 65 (3) (1990) 353.
[37] M. C. Gao, C. Niu, C. Jiang, D. L. Irving, Applications of spe-
cial quasi-random structures to high-entropy alloys, in: High-
Entropy Alloys, Springer, 2016, pp. 333–368.
[38] A. Van De Walle, M. Asta, G. Ceder, The alloy theoretic auto-
mated toolkit: A user guide, Calphad 26 (4) (2002) 539–553.
[39] G. V´
erit´
e, C. Domain, C.-C. Fu, P. Gasca, A. Legris,
F. Willaime, Self-interstitial defects in hexagonal close packed
metals revisited: Evidence for low-symmetry configurations in
Ti, Zr, and Hf, Physical Review B 87 (13) (2013) 134108.
[40] Q. Peng, W. Ji, H. Huang, S. De, Stability of self-interstitial
atoms in hcp-Zr, Journal of Nuclear Materials 429 (1-3) (2012)
233–236.
[41] D. Nguyen-Manh, A. Horsfield, S. Dudarev, Self-interstitial
atom defects in bcc transition metals: Group-specific trends,
Physical Review B 73 (2) (2006) 020101.
[42] Y. Ikeda, A. Carreras, A. Seko, A. Togo, I. Tanaka, Mode de-
composition based on crystallographic symmetry in the band-
unfolding method, Physical Review B 95 (2) (2017) 024305.
[43] F. K¨
ormann, Y. Ikeda, B. Grabowski, M. H. Sluiter, Phonon
broadening in high entropy alloys, npj Computational Materi-
als 3 (1) (2017) 1–9.
[44] A. Togo, I. Tanaka, First principles phonon calculations in ma-
terials science, Scripta Materialia 108 (2015) 1–5.
[45] B. Widom, Some topics in the theory of fluids, The Journal of
Chemical Physics 39 (11) (1963) 2808–2812.
[46] W. Tang, E. Sanville, G. Henkelman, A grid-based Bader anal-
ysis algorithm without lattice bias, Journal of Physics: Con-
densed Matter 21 (8) (2009) 084204.
[47] G. Henkelman, B. P. Uberuaga, H. J´
onsson, A climbing image
nudged elastic band method for finding saddle points and min-
imum energy paths, The Journal of chemical physics 113 (22)
(2000) 9901–9904.
[48] V. Wang, N. Xu, J.-C. Liu, G. Tang, W.-T. Geng, VASPKIT: a
user-friendly interface facilitating high-throughput computing
and analysis using VASP code, Computer Physics Communi-
cations (2021) 108033.
[49] L. Barnard, D. Morgan, Ab initio molecular dynamics simu-
lation of interstitial diusion in Ni–Cr alloys and implications
for radiation induced segregation, Journal of Nuclear Materials
449 (1-3) (2014) 225–233.
[50] A. Stukowski, Visualization and analysis of atomistic simula-
tion data with OVITO–the open visualization tool, Modelling
and Simulation in Materials Science and Engineering 18 (1)
(2009) 015012.
[51] A. Tamm, A. Aabloo, M. Klintenberg, M. Stocks, A. Caro,
Atomic-scale properties of Ni-based FCC ternary, and quater-
nary alloys, Acta Materialia 99 (2015) 307–312.
[52] W. K. Hastings, Monte carlo sampling methods using Markov
chains and their applications (1970).
[53] J. Cowley, An approximate theory of order in alloys, Physical
Review 77 (5) (1950) 669.
[54] J. F. Ziegler, M. D. Ziegler, J. P. Biersack, SRIM–the stop-
ping and range of ions in matter (2010), Nuclear Instruments
and Methods in Physics Research Section B: Beam Interactions
with Materials and Atoms 268 (11-12) (2010) 1818–1823.
[55] G. S. Was, Fundamentals of radiation materials science: metals
and alloys, springer, 2016.
[56] R. E. Stoller, M. B. Toloczko, G. S. Was, A. G. Certain,
S. Dwaraknath, F. A. Garner, On the use of SRIM for comput-
ing radiation damage exposure, Nuclear instruments and meth-
ods in physics research section B: beam interactions with ma-
terials and atoms 310 (2013) 75–80.
[57] C. Lu, K. Jin, L. K. B´
eland, F. Zhang, T. Yang, L. Qiao,
Y. Zhang, H. Bei, H. M. Christen, R. E. Stoller, et al., Direct
18
observation of defect range and evolution in ion-irradiated sin-
gle crystalline Ni and Ni binary alloys, Scientific reports 6 (1)
(2016) 1–10.
[58] D. B. Miracle, O. N. Senkov, A critical review of high entropy
alloys and related concepts, Acta Materialia 122 (2017) 448–
511.
[59] M. C. Gao, J.-W. Yeh, P. K. Liaw, Y. Zhang, et al., High-
entropy alloys, Cham: Springer International Publishing
(2016).
[60] Z. Wang, W. Qiu, Y. Yang, C. Liu, Atomic-size and lattice-
distortion eects in newly developed high-entropy alloys with
multiple principal elements, Intermetallics 64 (2015) 63–69.
[61] H. Song, F. Tian, Q.-M. Hu, L. Vitos, Y. Wang, J. Shen,
N. Chen, Local lattice distortion in high-entropy alloys, Physi-
cal Review Materials 1 (2) (2017) 023404.
[62] I. U. of Crystallography, C. H. Macgillavry, G. D. Rieck,
K. Lonsdale, International tables for X-ray crystallography,
Kynoch Press, 1962.
[63] Y. Zhang, Y. J. Zhou, J. P. Lin, G. L. Chen, P. K. Liaw,
Solid-solution phase formation rules for multi-component al-
loys, Advanced engineering materials 10 (6) (2008) 534–538.
[64] S. Guo, Q. Hu, C. Ng, C. Liu, More than entropy in high-
entropy alloys: Forming solid solutions or amorphous phase,
Intermetallics 41 (2013) 96–103.
[65] S. Huang, Z. Dong, W. Mu, V. Str ¨
om, G. Chai, L. Vitos,
Thermo-elastic properties of bcc Mn-rich high-entropy alloy,
Applied Physics Letters 117 (16) (2020) 164101.
[66] G. Grad, P. Blaha, J. Luitz, K. Schwarz, A. F. Guillermet,
S. Sferco, Electronic structure and chemical bonding eects
upon the bcc to ωphase transition: Ab initio study of Y, Zr,
Nb, and Mo, Physical Review B 62 (19) (2000) 12743.
[67] G. D. Samolyuk, Y. Osetsky, G. M. Stocks, J. Morris, Role
of static displacements in stabilizing body centered cubic high
entropy alloys, Physical Review Letters 126 (2) (2021) 025501.
[68] Y. Ikeda, K. Gubaev, J. Neugebauer, B. Grabowski,
F. K¨
ormann, Chemically induced local lattice distortions ver-
sus structural phase transformations in compositionally com-
plex alloys, npj Computational Materials 7 (1) (2021) 1–8.
[69] P. K. Nandi, M. Valsakumar, S. Chandra, H. Sahu, C. Sun-
dar, Ecacy of surface error corrections to density functional
theory calculations of vacancy formation energy in transition
metals, Journal of Physics: Condensed Matter 22 (34) (2010)
345501.
[70] Y. Gong, B. Grabowski, A. Glensk, F. K¨
ormann, J. Neuge-
bauer, R. C. Reed, Temperature dependence of the Gibbs en-
ergy of vacancy formation of fcc Ni, Physical Review B 97 (21)
(2018) 214106.
[71] A. Schneider, C.-C. Fu, C. Barreteau, Local environment de-
pendence of Mn magnetism in bcc iron-manganese alloys: A
first-principles study, Physical Review B 98 (9) (2018) 094426.
[72] J. Burton, Vacancy-formation entropy in cubic metals, Physical
Review B 5 (8) (1972) 2948.
[73] X. Zhang, M. H. Sluiter, Ab initio prediction of vacancy prop-
erties in concentrated alloys: The case of fcc Cu-Ni, Physical
Review B 91 (17) (2015) 174107.
[74] B. Grabowski, Y. Ikeda, P. Srinivasan, F. K¨
ormann,
C. Freysoldt, A. I. Du, A. Shapeev, J. Neugebauer, Ab ini-
tio vibrational free energies including anharmonicity for mul-
ticomponent alloys, npj Computational Materials 5 (1) (2019)
1–6.
[75] X. Zhang, B. Grabowski, F. K¨
ormann, C. Freysoldt, J. Neuge-
bauer, Accurate electronic free energies of the 3 d, 4 d, and
5 d transition metals at high temperatures, Physical Review B
95 (16) (2017) 165126.
[76] X. Zhang, B. Grabowski, T. Hickel, J. Neugebauer, Calculating
free energies of point defects from ab initio, Computational
Materials Science 148 (2018) 249–259.
[77] A. Manzoor, S. Pandey, D. Chakraborty, S. R. Phillpot, D. S.
Aidhy, Entropy contributions to phase stability in binary ran-
dom solid solutions, npj Computational Materials 4 (1) (2018)
1–10.
[78] Y. Zhang, K. Jin, H. Xue, C. Lu, R. J. Olsen, L. K. Beland,
M. W. Ullah, S. Zhao, H. Bei, D. S. Aidhy, et al., Influence of
chemical disorder on energy dissipation and defect evolution in
advanced alloys, Journal of Materials Research 31 (16) (2016)
2363–2375.
[79] R. LeSar, Introduction to computational materials science: fun-
damentals to applications, Cambridge University Press, 2013.
[80] G. D. Samolyuk, A. V. Barashev, S. I. Golubov, Y. Osetsky,
R. E. Stoller, Analysis of the anisotropy of point defect diu-
sion in hcp Zr, Acta materialia 78 (2014) 173–180.
[81] S. Shang, L. Hector Jr, Y. Wang, Z. Liu, Anomalous energy
pathway of vacancy migration and self-diusion in hcp Ti,
Physical Review B 83 (22) (2011) 224104.
[82] F. Willaime, Ab initio study of self-interstitials in hcp-Zr, Jour-
nal of nuclear materials 323 (2-3) (2003) 205–212.
[83] B. Eyre, R. Bullough, On the formation of interstitial loops in
bcc metals, Philosophical Magazine 12 (115) (1965) 31–39.
[84] B. Wirth, G. Odette, D. Maroudas, G. Lucas, Dislocation loop
structure, energy and mobility of self-interstitial atom clusters
in bcc iron, Journal of nuclear materials 276 (1-3) (2000) 33–
40.
[85] J. Chen, N. Gao, P. Jung, T. Sauvage, A new mechanism of
loop formation and transformation in bcc iron without dislo-
cation reaction, Journal of nuclear materials 441 (1-3) (2013)
216–221.
[86] Y. Zhang, Z. Xiao, X.-M. Bai, Eect of Cr concentration on
1/2<111>to <100>dislocation loop transformation in Fe-Cr
alloys, Journal of Nuclear Materials 543 (2021) 152592.
[87] J. K¨
arger, D. M. Ruthven, Diusion in zeolites and other mi-
croporous solids, Wiley New York, 1992.
[88] N. Anento, A. Serra, Y. N. Osetsky, Atomistic study of multi-
mechanism diusion by self-interstitial defects in α-Fe, Mod-
elling and Simulation in Materials Science and Engineering
18 (2) (2010) 025008.
[89] H. Wiedersich, P. Okamoto, N. Q. Lam, A theory of radiation-
induced segregation in concentrated alloys, Journal of Nuclear
Materials 83 (1) (1979) 98–108.
[90] A. Marwick, Segregation in irradiated alloys: the inverse Kirk-
endall eect and the eect of constitution on void swelling,
Journal of Physics F: Metal Physics 8 (9) (1978) 1849.
[91] T. Allen, J. Busby, G. Was, E. Kenik, On the mechanism of
radiation-induced segregation in austenitic Fe–Cr–Ni alloys,
Journal of nuclear materials 255 (1) (1998) 44–58.
[92] P. Okamoto, L. Rehn, Radiation-induced segregation in binary
and ternary alloys, Journal of Nuclear Materials 83 (1) (1979)
2–23.
[93] B. Loomis, A. Taylor, S. Gerber, Void swelling of Nb and Nb-
1% Zr induced by 58Ni+bombardment, Journal of Nuclear Ma-
terials 56 (1) (1975) 25–37.
[94] B. Loomis, S. Gerber, A. Taylor, Void ordering in ion-
irradiated Nb and Nb-1% Zr, Journal of Nuclear Materials
68 (1) (1977) 19–31.
[95] N. Ghoniem, D. Walgraef, S. Zinkle, Theory and experiment of
nanostructure self-organization in irradiated materials, Journal
of computer-aided materials design 8 (1) (2001) 1–38.
[96] J. L. Brimhall, G. L. Kulcinski, Void formation in ion bom-
barded niobium, Radiation Eects 20 (1-2) (1973) 25–31.
[97] B. Loomis, S. Gerber, Void formation and solute segregation
in ion-irradiated niobium-base alloys, Journal of Nuclear Ma-
19
terials 117 (1983) 224–233.
[98] Y. Lederer, C. Toher, K. S. Vecchio, S. Curtarolo, The search
for high entropy alloys: a high-throughput ab-initio approach,
Acta Materialia 159 (2018) 364–383.
[99] Q. Ding, Y. Zhang, X. Chen, X. Fu, D. Chen, S. Chen, L. Gu,
F. Wei, H. Bei, Y. Gao, et al., Tuning element distribution,
structure and properties by composition in high-entropy alloys,
Nature 574 (7777) (2019) 223–227.
[100] X. Chen, Q. Wang, Z. Cheng, M. Zhu, H. Zhou, P. Jiang,
L. Zhou, Q. Xue, F. Yuan, J. Zhu, et al., Direct observation of
chemical short-range order in a medium-entropy alloy, Nature
592 (7856) (2021) 712–716.
[101] R. Zhang, S. Zhao, J. Ding, Y. Chong, T. Jia, C. Ophus,
M. Asta, R. O. Ritchie, A. M. Minor, Short-range order
and its impact on the CrCoNi medium-entropy alloy, Nature
581 (7808) (2020) 283–287.
[102] S. Zhao, Local ordering tendency in body-centered cubic
(BCC) multi-principal element alloys, Journal of Phase Equi-
libria and Diusion (2021) 1–14.
[103] X. Zhou, H. Wadley, R. A. Johnson, D. Larson, N. Tabat,
A. Cerezo, A. Petford-Long, G. Smith, P. Clifton, R. Martens,
et al., Atomic scale structure of sputtered metal multilayers,
Acta materialia 49 (19) (2001) 4005–4015.
[104] D.-Y. Lin, S. Wang, D. Peng, M. Li, X. Hui, An n-body poten-
tial for a Zr–Nb system based on the embedded-atom method,
Journal of Physics: Condensed Matter 25 (10) (2013) 105404.
20
... High-entropy alloy (HEA) is loosely defined as multicomponent alloys comprised of four or more elements with each element at significant concentration [1]. HEA has gained a lot of interest in recent years because of its potential excellent comprehensive properties, such as good mechanical properties, excellent corrosion resistance, and promising radiation resistance [2][3][4][5][6][7]. In the family of HEA, refractory high entropy alloys (RHEA) indicates the alloys composed of refractory elements from groups IVB, VB and VIB, such as Ti, Zr, Hf, Nb, Ta, Mo and W [8]. Senkov et al. [9] examined 151 different types of RHEA compositions that have been reported in the literatures and discovered that the most studied quaternary systems are Nb-Ti-Zr-Mo (46 compositions), Nb-Ti-V-Mo (42 compositions), Nb-Ti-Zr-V (42 compositions), and Nb-Ti-Zr-Ta (32 compositions). ...
... By positron annihilation experiments, Lu et al. suggested that the concentration of inherent vacancy defects in Ti2ZrHfV0.5Mo0.2 is higher than conventional alloys, which enhances the irradiation tolerance of RHEA [28]. Recent studies also indicated that the enhanced irradiation resistance of BCC RHEAs can be attributed to the lattice distortion and disparity in chemistry [4,27,29,37,45]. ...
... The underlying mechanism under these interesting phenomena deserves further investigation. An increasing number of studies indicated that some RHEAs exhibit distinct irradiation behaviors [4,37,46]. Zhao et al. [37] reported that the distinct defect evolution in VTaTi and VTaW alloys is closely related to the compositional complexity induced chemical fluctuations and associated local lattice distortion. The chemical fluctuations induce a rough defect energy landscape and restrict the defect recombination in a local region, resulting in the formation of discrete and small defect clusters. ...
... Unlike conventional dilute alloys and pure metals, which exhibit single point defect energies or a few discrete values, multi-principal element alloys (MPEAs) have a wide range of migration barriers due to variable local chemical environments. [18][19][20][21] This makes the calculation of point defects energies and the prediction of their relative sequence among constituent elements in MPEAs relatively complex. Several factors, including chemical disorder and constituent species of HEAs, can alter point defect migration energies. ...
... Furthermore, vacancies are known to preferentially migrate through low-energy sites due to the exponential dependence of jump probability on the energy barrier height. 20 Therefore, the distribution of migration barriers at low energies has a higher impact on the overall diffusion behavior. In the NS state, 7.5% of the migration barriers are below 0.5 eV, compared to only 1.7% in the PM/FM state. ...
Article
Full-text available
Randomly mixing ferromagnetic (FM) and antiferromagnetic (AFM) elements in high-entropy alloys (HEAs) can create fluctuating local magnetic moments that influence the energetics of point defects. In this study, we employed first-principles calculations to investigate the influence of magnetic properties on vacancy migration energy in Fe49.5Mn29.4Co10.1Cr10.1C0.9, alongside equiatomic NiCoFeCrMn alloy. By examining structures with paramagnetism, ferromagnetism, and no spin polarization, our study reveals significant impacts of magnetic interactions on vacancy migration barriers, potentially altering the sequence of elemental migration energies if overlooked. In Fe49.5Mn29.4Co10.1Cr10.1C0.9, the order of vacancy migration barriers is Co > Fe > Mn > Cr across all magnetic states, suggesting the dominant roles of atomic properties and inherent chemical bonding. Conversely, the NiCoFeCrMn HEA exhibits a pronounced magnetic state-dependent elemental migration energy order, indicating that magnetic interactions significantly influence vacancy migration behavior in this alloy. In addition, while FM elements generally exhibit higher migration barriers, AFM elements display lower barriers in the investigated Cantor alloys, with notable variations between the studied compositions. These findings underscore the critical role of magnetism in accurate migration energy calculations, which is important for studying chemically biased diffusion and radiation-induced segregation in HEAs.
... These characteristics make HEAs promising structural materials for use in advanced nuclear reactors. Recent reports [16][17][18] on the irradiation resistance of HEAs document significant progress in the field. ...
... Recently, Lu et al. [34] found that the introduction of interstitial N into the FCC-structured NiCoFeCrMn HEA could elevate the chemical short-range order, composition variations, and lattice strain, thereby enhancing the barriers to the diffusion of interstitial atoms and clusters induced by the irradiation of 3 MeV Ni 2+ , which effectively delayed void growth and dislocation loop evolution and improved irradiation resistance. However, studies on the effects of N on the mechanical properties and He bubble behavior of irradiated body-centered cubic (BCC)-structured HEAs are quite limited, especially for refractory high entropy alloys (RHEAs) [9], which have shown fine potential as structural materials under extreme conditions [11,16,18,35]. ...
Article
Interstitial strengthening with nitrogen (N) is one of the effective ways to improve the mechanical properties of HEAs, but the effects of N on the microstructures and mechanical properties of the irradiated HEAs have not been studied extensively. Here, the microstructures and mechanical properties of N-free and N-doped Ti2ZrNbV0.5Mo0.2 HEAs before and after He irradiation were investigated. The results showed that the solid solution strengthening caused by interstitial N improved the yield strength at room temperature and 1023 K without significantly reducing plasticity. N doping significantly promoted the growth, aggregation and wider spatial distribution of He bubbles by enhancing the mobility of He atoms/He-vacancy complexes, with the average size of He bubbles increasing from 10.4 nm in N-free HEA to 31.0 nm in N-doped HEA. In addition, N-doped HEA showed a much higher irradiation hardness increment and hardening fraction than N-free HEA. Contrary to conventional materials doped with N, the introduction of N into Ti2ZrNbV0.5Mo0.2 HEA had adverse effects on its resistance to He bubble growth and irradiation hardening. The results of this study indicated that N doping may not improve the irradiation resistance of HEAs.
... The structural properties and performances of high-entropy alloys are unique and interesting issues that have attracted intensive attention [1][2][3][4][5][6][7][8][9][10]. HEAs consisting of four or more elements with equal atomic concentrations are intriguing materials because of their excellent comprehensive properties as structural materials [2] which contribute to sluggish diffusion [3,7,9], critical internal lattice distortion [8,[11][12][13], and corrosion resistance [14][15][16]. More demands for nuclear energy are met using HEAs, which are considered promising materials for fusion and generation IV fission reactors [17][18][19][20]. ...
... In general, the materials manifest sharply declining performances in hardening and swelling under irradiation [12,[21][22][23][24][25][26][27][28][29]. Introducing several defects sinks can effectively enhance the radiation resistance of materials, such as oxide-dispersion-strengthened steels [30,31] and nanograined polycrystalline alloys [32]. ...
Article
Full-text available
High-entropy alloys (HEAs) attract much attention as possible radiation-resistant materials due to their several unique properties. In this work, the generation and evolution of the radiation damage response of an FeNiCrCoCu HEA and bulk Ni in the early stages were explored using molecular dynamics (MD). The design, concerned with investigating the irradiation tolerance of the FeNiCrCoCu HEA, encompassed the following: (1) The FeNiCrCoCu HEA structure was obtained through a hybrid method that combined Monte Carlo (MC) and MD vs. the random distribution of atoms. (2) Displacement cascades caused by different primary knock-on atom (PKA) energy levels (500 to 5000 eV) of the FeNiCrCoCu HEA vs. bulk Ni were simulated. There was almost no element segregation in bulk FeNiCrCoCu obtained with the MD/MC method by analyzing the Warren–Cowley short-range order (SRO) parameters. In this case, the atom distribution was similar to the random structure that was selected as a substrate to conduct the damage cascade process. A mass of defects (interstitials and vacancies) was generated primarily by PKA departure. The number of adatoms grew, which slightly roughened the surface, and the defects were distributed deeper as the PKA energy increased for both pure Ni and the FeNiCrCoCu HEA. At the time of thermal spike, one fascinating phenomenon occurred where the number of Frenkel pairs for HEA was more than that for pure Ni. However, we obtained the opposite result, that fewer Frenkel pairs survived in the HEA than in pure Ni in the final state of the damage cascade. The number and size of defect clusters grew with increasing PKA energy levels for both materials. Defects were suppressed in the HEA; that is to say, defects were “cowards”, behaving in an introverted manner according to the anthropomorphic rhetorical method.
Article
The characterization of short-range order (SRO) and its influence on performance are a widely debated topics in high-entropy alloys (HEAs). In this work, taking the Co-Fe-Ni-Ti alloy without complicated magnetism as a benchmark of 3d HEAs, we investigate the effect of SRO on local lattice distortion (LLD), general stacking fault energy (GSFE), tensile and shear strength of FCC Co30Fe16.66Ni36.67Ti16.67 via the combination of Monte Carlo (MC) and molecular dynamics (MD). This alloy shows the typical SRO of Ti-X (X=Fe,Co) atomic pairs, while the segregation of Ti-Ti atomic pairs. The SRO has a minor inhibition on LLD. Considering the thermal vibration induced atomic displacement, the degree of LLD increases nonlinearly with increasing temperature. Both the severe LLD and SRO are helpful to tune the GSFE at finite temperature. The SRO enhances the degree of deformation twinning and delays the appearance of the HCP phase but increases the number of HCP-type atoms as the energy buffers. For the polycrystalline systems, SRO promotes the precipitation of BCC phase at grain boundaries and the number of HCP-type atoms in the grain and activates the deformation of slip surfaces. Therefore SRO could play a key role for the outstanding strength and plasticity of Co-Fe-Ni-Ti HEAs.
Article
V-Ti-based refractory high entropy alloys (RHEAs), e.g., VTiTa and VTiTaNb, have been found with outstanding mechanical properties and irradiation resistance, granting them application potential in nuclear engineering. Nonetheless, similar to conventional V-Ti-based alloys such as V-4Ti-4Cr, these alloys are sensitive to C/N/O interstitial impurities. Under irradiation, such impurities may promote the formation of precipitates which may affect the defect evolution, especially at the appearance of helium. Previous studies have observed different helium behavior in the above alloys, including the heterogeneous distribution of helium bubbles in VTiTa, but the origin was not clear. In this work, detailed microstructural characterization is performed to analyze the irradiation induced precipitates and their impact on helium bubble formation in these alloys. Under helium ion irradiation at 700 °C, the TiN precipitates with NaCl structure are formed in all the three alloys. Nonetheless, the interfaces between precipitates and matrix in VTiTa and VTiTaNb are semi-coherent while those in V-4Ti-4Cr are coherent, due to the various lattice parameters of the alloys. Moreover, severe lattice distortion is found also inside the precipitates in the two RHEAs. Consequently, significant heterogeneous nucleation of He bubbles are observed in these two RHEAs but not in V-4Ti-4Cr. Among the three alloys, the growth of precipitates is the fastest in VTiTa, causing the formation of bubble clusters. Furthermore, the growth direction/shape of precipitates affect the shape of bubbles, leading to the {001} cuboids shape in V-4Ti-4Cr and the truncated {110} dodecahedra shape in V-Ti-Ta and V-Ti-Ta-Nb.
Article
The Snoek anelastic internal friction (IF) peak due to stress-induced dipole ordering associated with interstitial atoms was well observed in IF curves of single-principal element alloys (SPEAs) as early as in the 1950s, however, its behavior remains elusive in multi-principal element alloys (MPEAs) to date. Here the temperature-dependent IF spectra of NbTiV0.5Zr refractory MPEAs with varying O contents were presented, and the analysis of these data indicates that for O-doped NbTiV0.5Zr MPEAs, the O addition triggers two anelastic IF peaks (the PO1 and PO2), corresponding to the O-Snoek-type relaxation in random solid solution (RSS) and local chemical ordering (LCO) structures, respectively. The half-peak width of the PO1 is 2–3 times broader than that for SPEAs, associated with the wide O migration energy barrier distribution in MPEAs. These phenomena show a striking contrast to the single and narrow Snoek-type peak observed for SPEAs. Our analysis further reveals an approximate linear correlation between the half-peak width of the Snoek-type peak and the mixing entropy. The introduction of O also shifts grain boundary relaxation peaks (PG) towards higher temperatures. Interestingly, high-temperature anomalous modulus changes were observed for the first time in TiV-based lightweight refractory MPEAs, which is suggested to be originated from the order reduction of the body-centered cubic (bcc) structure at elevated temperatures. Moreover, the 1 at.%O doped NbTiV0.5Zr MPEA with a favorable balance between the peak heights of the PO2 and PG, exhibits outstanding specific yield stress (SYS) among refractory MPEAs while maintaining an elongation as high as 20%. This work provides not only a bridge between the anelastic behavior of SPEAs and MPEAs for upcoming studies and theories concerning bcc MPEAs with interstitial atoms, but contributes to the holistic design of bcc MPEAs with high strength and excellent ductility.
Article
Full-text available
We develop a fast and accurate machine-learned interatomic potential for the Mo-Nb-Ta-V-W quinary system and use it to study segregation and defects in the body-centered-cubic refractory high-entropy alloy MoNbTaVW. In the bulk alloy, we observe clear ordering of mainly Mo-Ta and V-W binaries at low temperatures. In damaged crystals, our simulations reveal clear segregation of vanadium, the smallest atom in the alloy, to compressed interstitial-rich regions such as radiation-induced dislocation loops. Vanadium also dominates the population of single self-interstitial atoms. In contrast, due to its larger size and low surface energy, niobium segregates to spacious regions such as the inner surfaces of voids. When annealing samples with supersaturated concentrations of defects, we find that in complete contrast to W, interstitial atoms in MoNbTaVW cluster to create only small (∼1 nm) experimentally invisible dislocation loops enriched by vanadium. By comparison to W, we explain this by the reduced but three-dimensional migration of interstitials, the immobility of dislocation loops, and the increased mobility of vacancies in the high-entropy alloy, which together promote defect recombination over clustering.
Article
Full-text available
High-entropy alloys greatly expand the alloy design range and offer new possibilities for improving material performance. Based on the worldwide research efforts in the last decade, the excellent mechanical properties and promising radiation and corrosion resistance of this group of materials have been demonstrated. High-entropy alloys with body-centered cubic (BCC) structures, especially refractory high-entropy alloys, are considered as promising materials for high-temperature applications in advanced nuclear reactors. However, the extreme reactor conditions including high temperature, high radiation damage, high stress, and complex corrosive environment require a comprehensive evaluation of the material properties for their actual service in nuclear reactors. This review summarizes the current progress on BCC high-entropy alloys from the aspects of neutron economy and activation, mechanical properties, high-temperature stability, radiation resistance, as well as corrosion resistance. Although the current development of BCC high-entropy alloys for nuclear applications is still at an early stage as the large design space of this group of alloys has not been fully explored, the current research findings provide a good basis for the understanding and prediction of material behaviors with different compositions and microstructures. Further in-depth understanding of the degradation mechanisms and characterization of material properties in response to conditions close to reactor environment are necessary. A critical down-selection of potential candidates is also crucial for further comprehensive evaluation and engineering validation.
Article
Full-text available
Complex concentrated solutions of multiple principal elements are being widely investigated as high- or medium-entropy alloys (HEAs or MEAs)1–11, often assuming that these materials have the high configurational entropy of an ideal solution. However, enthalpic interactions among constituent elements are also expected at normal temperatures, resulting in various degrees of local chemical order12–22. Of the local chemical orders that can develop, chemical short-range order (CSRO) is arguably the most difficult to decipher and firm evidence of CSRO in these materials has been missing thus far16,22. Here we discover that, using an appropriate zone axis, micro/nanobeam diffraction, together with atomic-resolution imaging and chemical mapping via transmission electron microscopy, can explicitly reveal CSRO in a face-centred-cubic VCoNi concentrated solution. Our complementary suite of tools provides concrete information about the degree/extent of CSRO, atomic packing configuration and preferential occupancy of neighbouring lattice planes/sites by chemical species. Modelling of the CSRO order parameters and pair correlations over the nearest atomic shells indicates that the CSRO originates from the nearest-neighbour preference towards unlike (V−Co and V−Ni) pairs and avoidance of V−V pairs. Our findings offer a way of identifying CSRO in concentrated solution alloys. We also use atomic strain mapping to demonstrate the dislocation interactions enhanced by the CSROs, clarifying the effects of these CSROs on plasticity mechanisms and mechanical properties upon deformation.
Book
Emphasising essential methods and universal principles, this textbook provides everything students need to understand the basics of simulating materials behaviour. All the key topics are covered from electronic structure methods to microstructural evolution, appendices provide crucial background material, and a wealth of practical resources are available online to complete the teaching package. Modelling is examined at a broad range of scales, from the atomic to the mesoscale, providing students with a solid foundation for future study and research. Detailed, accessible explanations of the fundamental equations underpinning materials modelling are presented, including a full chapter summarising essential mathematical background. Extensive appendices, including essential background on classical and quantum mechanics, electrostatics, statistical thermodynamics and linear elasticity, provide the background necessary to fully engage with the fundamentals of computational modelling. Exercises, worked examples, computer codes and discussions of practical implementations methods are all provided online giving students the hands-on experience they need.
Article
Refractory multiple principal elemental alloys (MPEAs) hold great promise for structural materials in future nuclear energy systems. Compared to the extensively studied face-centered cubic (FCC) MPEAs, the irradiation resistance of body-centered cubic (BCC) refractory MPEAs is relatively less known. In this work, we study defect accumulation and evolution in two BCC VTaTi and VTaW MPEAs comparatively through atomistic simulations. For this purpose, we have parameterized the Embedded Atom Method (EAM) potential parameters for V metals. Combined with available potential parameters for other elements, we construct average atom models for the considered alloys to elucidate the effects of chemical complexities in BCC MPEAs. Our results based on Frenkel pair accumulation simulations suggest that the major influence of chemical fluctuations in BCC MPEAs is on the clustering behavior of defects, which leads to discrete point defects or small defect clusters, in contrast to the large defect clusters observed in the average atom model. We further show that the diffusion of interstitial clusters exhibits different modes due to chemical complexity. While interstitials show either three-dimensional or one-dimensional diffusion in the average atom model, the mean free path of interstitials in the random alloy is strongly suppressed. These results provide fundamental insights into the irradiation response of BCC MPEAs and pinpoint the critical role of chemical complexity on defect evolution, which lay the basis for the future development of irradiation-resistant structural materials based on BCC MPEAs.
Article
Vanadium (V)-based alloys are potential candidates for structural materials of fusion reactors because of their excellent properties; recently, the V–Ti–Ta alloy has received considerable attention. The irradiation-resistant performance is significant for the development and application of fusion reactor materials. In this study, combining with our previously developed potentials of V and Ta, we developed the V-Ti-Ta interatomic potential based on the Finnis–Sinclair (FS) formalism, which is an important step towards an atomic-scale investigation of V-based alloys. The potential parameters were determined by fitting to a large database of experimental data and density functional theory (DFT) calculations. The formation energies of vacancy and self-interstitial atoms, as well as the pressure-volume relation obtained from the present Ti potential, were well reproduced. The typical defect energies of solute atoms Ti and Ta in the V matrix and Ti and V in the Ta matrix, such as the formation energies of substitutional solute atoms, binding energies between solute atoms and point defects, agree well with DFT results. Thus, the developed potentials are expected to be suitable for atomistic simulations of point defects evolution in the V-Ti-Ta ternary system.
Article
We present the VASPKIT, a command-line program that aims at providing a powerful and user-friendly interface to perform high-throughput analysis of a variety of material properties from the raw data produced by the VASP code. It consists of mainly the pre- and post-processing modules. The former module is designed to prepare and manipulate input files such as the necessary input files generation, symmetry analysis, supercell transformation, k-path generation for a given crystal structure. The latter module is designed to extract and analyze the raw data about elastic mechanics, electronic structure, charge density, electrostatic potential, linear optical coefficients, wave function plots in real space, and etc. This program can run conveniently in either interactive user interface or command line mode. The command-line options allow the user to perform high-throughput calculations together with bash scripts. This article gives an overview of the program structure and presents illustrative examples for some of its usages. The program can run on Linux, MacOS, and Windows platforms. The executable versions of VASPKIT and the related examples, together with the tutorials, are available in its official website vaspkit.com. Program summary Program title: VASPKIT CPC Library link to program files: https://doi.org/10.17632/v3bvcypg9v.1 Licensing provisions: GPLv3 Programming language: Fortran, Python Nature of problem: This program has the purpose of providing a powerful and user-friendly interface to perform high-throughout calculations together with the widely-used VASP code. Solution method: VASPKIT is able to extract, calculate and even plot the mechanical, electronic, optical and magnetic properties from density functional calculations together with bash and python scripts. It can run in either interactive user interface or command line mode.
Article
High Entropy Alloys (HEA) attract attention as possible radiation resistant materials, a feature observed in some experiments that has been attributed to several unique properties of HEA, in particular to the disorder-induced reduced thermal conductivity and to the peculiar defect properties originating from the chemical complexity. To explore the origin of such behavior we study the early stages (less than 0.1 ns), of radiation damage response of a HEA using molecular dynamics simulations of collision cascades induced by primary knock-on atoms (PKA) with 10, 20 and 40 keV, at room temperature, on an idealized model equiatomic quinary fcc FeNiCrCoCu alloy, the corresponding “Average Atom” (AA) material, and on pure Ni. We include accurate corrections to describe short-range atomic interactions during the cascade. In all cases the average number of defects in the HEA is lower than for pure Ni, which has been previously used to help claiming that HEA is radiation resistant. However, simulated defect evolution during primary damage, including the number of surviving Frenkel Pairs, and the defect cluster size distributions are nearly the same in all cases, within our statistical uncertainty. The number of surviving FP in the alloy is predicted fairly well by analytical models of defect production in pure materials. All of this indicates that the origin of radiation resistance in HEAs as observed in experiments may not be related to a reduction in primary damage due to chemical disorder, but is probably caused by longer-time defect evolution.
Article
The recent development of high-entropy alloys (HEAs) has opened a new avenue for alloy design by incorporating multiple principal elements into a simple crystal lattice. In HEAs, short-range chemical ordering may arise due to the different mixing properties of various constituent elements. In this work, we explore the trend of elemental arrangement in five typical body-centered cubic (BCC) multi-principal element alloys (MPEAs), including CrFeV, HfNbZr, TaVW, NbTaV, and TaTiV. Combining the Monte Carlo method and density-functional theory calculations, we calculated the evolution of short-range order (SRO) parameters in these alloys at 500 K. Our results demonstrated that all these MPEAs might develop a certain degree of SRO, depending on the mixing properties of different element pairs. Using a modified quasi-chemical model, we show that the SRO values in these MPEAs can be accurately predicted as long as the mixing enthalpy values between different atomic pairs are known. The presence of SRO has a profound impact on the local structure of MPEAs, which further alters the electronic and elastic properties of MPEAs. Our results provide a theoretical understanding of the basic alloy structure of novel MPEAs, which is an essential step toward tailoring their properties.