LX - Li JAC 2023

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Alloys and Compounds 947 (2023) 169650

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


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

Shock compression and spall damage of dendritic high-entropy alloy


CoCrFeNiCu ]]
]]]]]]
]]

⁎ ⁎
L.X. Li a, X.Y. Liu a, J. Xu a, S.C. Hu a, Y. Cai a, L. Lu b, J.C. Cheng c, Y. Tang d, , C. Li b, , N.B. Zhang b,
S.N. Luo b
a
The Peac Institute of Multiscale Sciences, Chengdu, Sichuan, PR China
b
School of Materials Science and Engineering, Southwest Jiaotong University, Chengdu, Sichuan, PR China
c
Institute for Advanced Study, Chengdu University, Chengdu, Sichuan, PR China
d
College of Aerospace Science and Engineering, National University of Defense Technology, Changsha, Hunan, PR China

a r t i cl e i nfo a bstr ac t

Article history: Plate impact experiments are conducted on a dendritic dual face-center cubic phase high-entropy alloy
Received 14 January 2023 (HEA) CoCrFeNiCu, consisting of Cu-lean dendritic (DR) and Cu-rich interdendritic regions (ID). Free surface
Received in revised form 3 March 2023 velocity histories are obtained along with microstructure characterizations. The Hugoniot equation of state
Accepted 11 March 2023
and spall strength at different shock stresses are determined. Dislocation slip is the main deformation mode
Available online 15 March 2023
for CoCrFeNiCu HEA, and the dislocation density of the Cu-rich interdendritic regions increases more sig­
nificantly during shock compression. Both ductile and brittle damage modes are observed. With increasing
Keywords:
CoCrFeNiCu impact velocity, the Cu-rich interdendritic regions with severe strain localizations and more defects provide
Dendritic structure more damage nucleation sites. The spall strength increases firstly, reaches its maximum at a peak shock
Hugoniot stress of 5.8 GPa, and then decreases. Molecular dynamics simulations reveal that the abnormal dependence
Spall damage of spall strength on peak shock stress is a result of strain localization and thermal softening in the Cu-rich
Microstructure regions.
© 2023 Elsevier B.V. All rights reserved.

1. Introduction [13], transformation-induced plasticity (TRIP) [14,15], and synergy


between dendritic (DR) and interdendritic regions (ID) [16], con­
High-entropy alloys (HEAs) have received considerable atten­ tribute considerably to the improvement of their mechanical prop­
tions for their enormous compositional space and excellent me­ erties. CoCrFeNiCu under quasi-static loading exhibits high strength
chanical and chemical properties, for example, high strength, high and hardness via dendritic segregation strengthening and grain re­
ductility, high fracture toughness, and high fatigue resistance and finement [7,10,17], as well as good ductility due to nanoscale de­
corrosion resistance [1–5]. formation twinning in the dendritic matrix [7,8]. In addition,
The study on HEAs has expanded from HEAs with a single phase dendritic structures also give rise to high wear resistance via self-
to those with dual phases, multiple phases and other special phase lubricating of the interdendritic ductile phase [18,19].
structures, such as dendritic structures. As a representative dendritic Molecular dynamics (MD) simulations were used to investigate
dual face-center cubic (FCC) phase HEA, CoCrFeNiCu has large po­ the tensile response [20], the relationship between shear stress and
sitive mixing enthalpies between Cu and other elements [6]. One dislocation velocity [21], wear performance and protection ability
phase is the hard Cu-lean matrix, and the other is the soft Cu-rich [22], and the shock response, dynamic failure [23,24] and the phase
interdendritic phase in CoCrFeNiCu [7,8]. transition [23] in the CoCrFeNiCu HEA. However, the effects of
Previous studies demonstrated that dendritic structures can dendritic structure or phase separation were not adequately con­
overcome the strength–ductility tradeoff of HEAs via dendritic seg­ sidered on dynamic mechanical properties.
regation strengthening and grain refinement [9–12]. Moreover, such Previous experimental studies on HEAs were mostly conducted
deformation mechanisms as twinning-induced plasticity (TWIP) quasi-static loading, and only a very few types of HEAs were in­
vestigated under high strain rate loading [25–28]. For CoCrFeNiCu or
other dendritic HEAs, there exist essentially no experiments on

shock compression or spall damage under planar impact. The ex­
Corresponding authors.
E-mail addresses: tangyu15@zju.edu.cn (Y. Tang), cli@swjtu.edu.cn (C. Li).
perimental data are lacking including the equation of state (EOS), the

https://doi.org/10.1016/j.jallcom.2023.169650
0925-8388/© 2023 Elsevier B.V. All rights reserved.
L.X. Li, X.Y. Liu, J. Xu et al. Journal of Alloys and Compounds 947 (2023) 169650

Hugoniot elastic limit (HEL) and spall strength. Furthermore, there


are two open questions for such HEAs: whether or not the dendritic
structure can improve dynamic mechanical properties like HEL and
spall strength, and how the hard Cu-lean matrix and the soft Cu-rich
interdendritic phase respond under high strain rate loading.
In this work, plate impact experiments including Hugoniot
equation of states (EOS) and spallation-recovery experiments are
conducted on a representative dendritic HEA, CoCrFeNiCu, to in­
vestigate the effect of dendritic segregation on the shock compres­
sion and damage properties. Postmortem samples are characterized
with scanning electron microscopy (SEM) and electron backscatter
diffraction (EBSD). Moreover, classical MD simulations are con­
ducted to investigate damage mechanisms and help interpret ex­
periments. The Cu-rich interdendritic regions accommodate more
plastic deformation than the Cu-lean dendritic matrix. Both the
ductile and brittle fracture modes are observed. With increasing
peak stress, the spall strength increases firstly and then decreases
due to thermal softening in the Cu-rich interdendritic regions. The
impact properties of the dendritic HEA including deformation and
spallation and underlying mechanisms reported here improve our
understanding of relevant dendritic structure materials which can
benefit micro-structural design and applications of HEAs at dynamic
extremes.

2. Material and methodology

2.1. Material
Fig. 2. SEM micrograph and corresponding elemental distributions of the as-cast
CoCrFeNiCu.
An equimolar CoCrFeNiCu HEA is provided by Beijing Yanbang New
Materials Co. Ltd. The as-cast alloy has a density ρ0 = 8.356 g cm−3,
Table 1
longitudinal sound speed CL = 5.652 km s−1, transverse sound speed Compositions of the as-cast CoCrFeNiCu with dual FCC phases and corresponding
CT = 3.037 km s−1, bulk sound speed CB = 4.433 km s−1, and Poisson’s lattice parameters.
ratio ν = 0.297.
EDS scan fraction (at. %) lattice parameter
The phase and elemental analyses of the as-received alloy are
Co Cr Fe Ni Cu (nm)
conducted with an X-ray diffractometer (the Cu Kα radiation) and an
FEI Quanta 250 FEG-SEM equipped with an energy dispersive X-ray Cu-lean phase scan 27.4 21.4 21.8 20.5 8.9 0.358
spectrometer (EDS), respectively. The grain orientations are char­ Cu-rich phase scan 2.8 1.6 2.2 6.9 86.5 0.361
Whole sample scan 24.5 20.0 19.9 19.6 16.0 –
acterized with the SEM equipped with an Oxford EBSD detector and
the HKL channel 5 OIM software, at 20 kV. Before the EBSD analysis,
the samples are ground sequentially with 1 μm and 0.3 μm alumina corresponding lattice parameters are obtained within 0.5 % from the
particles. The as-received alloy is composed of two face-center cubic X-ray diffraction pattern via the Bragg’s law [29], and are close to the
(FCC) phases with different compositions (Figs. 1 and 2; Table 1). The literature values [18,19]. The Cu-lean dendritic regions and the Cu-
rich interdendritic regions are shown in Fig. 2. The Cu-lean dendritic
regions have a mean grain size of 40–80 μm. The molar fraction of Cu
in the Cu-rich interdendritic regions is more than 86.5 %, as opposed
to 8.9 % in the dendritic regions.

2.2. Experiments

Schematic setups for two types of plate impact experiments (the


Hugoniot EOS and spallation-recovery) are shown in Fig. 3. Laser
Doppler velocimeter (LDV), similar to photon Doppler velocimeter
(PDV) [30], is used to measure the particle velocity of the free surface
of the driver plate or the sample. A magnetic induction velocimeter
system or an optical beam block velocimeter system is used to
measure the projectile velocity with an accuracy of 0.5 %, and pro­
vides a trigger signal for LDV.
The Hugoniot EOS experiments are conducted on a 28-mm bore
diameter single-stage gas gun (Fig. 3a). The flyer plates and the
driver plates are made of annealed pure oxygen-free high-con­
Fig. 1. X-ray diffraction curve of the as-cast CoCrFeNiCu. Inset: enlarged view. DR: Cu- ductivity (OFHC) copper, and the sabots are made of polycarbonate
lean FCC dendritic region; ID: Cu-rich FCC interdendritic region. (PC) with a low shock impedance. The flyer plates, driver plates and

2
L.X. Li, X.Y. Liu, J. Xu et al. Journal of Alloys and Compounds 947 (2023) 169650

Fig. 4. Initial atomic configuration of the CoCrFeNiCu HEA. Atoms are color-coded
according to local structure type. Impact direction: along the x-axis.

(corresponding to ∼ 6 × 106 atoms), and then the elemental per­


centages in grain boundaries are adjusted to match the Cu-rich
composition [36]. The modified grain boundary regions are treated
as the interdendritic regions, as in the experiments.
Shock direction is along the longest dimension (the x-axis), from
left to right. During dynamic loading, periodic boundary conditions
are applied except for the impact direction. The piston method [37]
is adopted for shock, release and spall simulations with the micro­
Fig. 3. Schematic setups for (a) Hugoniot equation of state experiments and (b) canonical ensemble. A small region on the left is initially set as the
spallation-recovery experiments. 1: gun barrel; 2: polycarbonate sabot; 3: flyer plate; piston with a piston velocity up = 0.9 km s−1, and drives the rest of
4: magnetic induction velocimeter or optical beam block velocimeter; 5: sample
holder; 6: driver plate; 7: sample; 8: optical fiber connected to a laser Doppler ve­
the atoms to up. For spallation, the piston is removed after shock
locimeter (LDV); 9: vacuum/target chamber; 10: recess for release waves; 11: mo­ compression, yielding a release fan from the piston side. When the
mentum trapping ring; 12: thin turning mirror; 13: lens; 14: soft material. shock wave reaches the free surface, it is reflected as another release
fan propagating leftwards. The interaction with these two release
samples are ground and polished to mirror-like surfaces, and their fans induces tension and spallation. The time step for integrating the
thicknesses are 2 mm, 2 mm, and 3 mm, respectively. A total of 9 LDV equation of motion is 1 fs. Prior to shock loading, the configuration is
probes are used for velocimeter. Four single-mode fibers are set 90∘ relaxed with the constant-pressure-temperature ensemble under
apart on a circle to monitor the motion of the free surface of the Cu three-dimensional periodic boundary conditions at 300 K and zero
driver plate. A central fiber and three surrounding fibers set 120∘ pressure.
apart on a circle monitor the free surface movement of the sample.
The outermost optical fiber is used to measure the impact velocity 3. Results and Discussions
directly with an accuracy of 0.2 %.
Spallation-recovery experiments are conducted on a 14-mm bore Six Hugoniot EOS experiments and seven spallation-recovery
diameter single-stage gas gun (Fig. 3b). Disk-shaped flyer plates and experiments are conducted. The experimental parameters and re­
samples are ground and polished to 1 mm and 2 mm, respectively, sults are listed in Tables 2 and 3, respectively. The EOS results are
with mirror-like surfaces. A momentum trapping ring is used to shown in Fig. 5. The free surface velocity histories for the spallation
mitigate rogue radial release waves [31,32]. Different from the EOS shots are shown in Fig. 6.
experiments, the Doppler-shifted signal beam reflected from the free
surface of a moving sample is mixed with a non-Doppler-shifted 3.1. Hugoniot equation of state
reference beam. The impacted samples are “soft-recovered" with
soft materials. The microstructure of the postmortem samples is also Let subscripts 0, 1, and 2 denote the initial state, the Hugoniot
characterized with SEM and EBSD in the center region of a sample elastic limit (HEL), and the plastic shock state in the sample, re­
with the minimum effect of edge release. spectively. The stress along the shock direction (σ), density (ρ), shock
wave velocity (us), and particle velocity (up) at shock states i = 1 and
2.3. MD simulations 2 satisfy the following Hugoniot jump conditions [38],

i i 1 = i 1 (upi upi 1)(usi upi 1), (1)


MD simulations of shock compression and spallation of poly­
crystalline CoCrFeNiCu HEA are performed as a complement to the
experiments. The embedded-atom-method (EAM) potential [33] is
i (us i )
u pi = i 1 (us i
upi ).
1 (2)
used to describe the atomic interactions, which has been success­ σ0 and up0 are 0 for the initial state, and ρ0 = 8.356 g cm−3.
fully used to simulate the shock response [34] and phase transition The flyer plate velocity (uf) and corresponding free surface ve­
[23] in the CoCrFeNiCu HEA. The large-scale atomic/molecular locity history (ufs) are obtained from the central LDV probe. The
massively parallel simulator (LAMMPS) [35] is employed for the MD Lagrangian shock velocities (denoted with superscript L) for the
simulations. elastic precursor and the plastic shock are determined from the
The CoCrFeNiCu configuration (Fig. 4) is first created as a Cu-lean sample thickness and corresponding transit times, and further cor­
polycrystalline model with dimensions of 1000 nm × 24 nm × 24 nm rected for wave front tilt [30,39]. Then the Lagrangian shock

3
L.X. Li, X.Y. Liu, J. Xu et al. Journal of Alloys and Compounds 947 (2023) 169650

Table 2
Summary of experimental parameters and results for the Hugoniot equation of state experiments. df: flyer plate thickness; ds: sample thickness; uf: flyer plate velocity; up1:
particle velocity at HEL; us1: elastic shock wave velocity; σHEL: stress at HEL; up2 : particle velocity for the plastic shock; us2 : plastic shock wave velocity; εH: peak volumetric strain;
σH: peak shock stress. Numbers in parentheses denote uncertainties in the last 1 or 2 digits.

Shot ds uf up1 us1 σHEL up2 us2 εH σH


No. (mm) (km s−1) (km s−1) (km s−1) (GPa) (km s−1) (km s−1) (GPa)

EOS1 3.044(2) 0.318(1) 0.010(1) 5.751(55) 0.481(7) 0.154(1) 4.639(15) 0.033(1) 6.083(13)
EOS2 3.044(2) 0.406(1) 0.015(1) 5.715(59) 0.716(10) 0.197(1) 4.731(15) 0.041(1) 7.910(16)
EOS3 3.047(2) 0.502(1) 0.013(1) 5.786(56) 0.604(8) 0.244(1) 4.811(16) 0.050(1) 9.921(20)
EOS4 3.044(2) 0.608(1) 0.015(1) 5.845(57) 0.733(10) 0.296(1) 4.917(17) 0.060(1) 12.266(24)
EOS5 3.040(2) 0.705(1) 0.017(1) 5.712(54) 0.787(11) 0.344(1) 4.979(17) 0.069(1) 14.398(29)
EOS6 3.046(2) 0.916(1) 0.017(1) 5.691(54) 0.808(11) 0.450(1) 5.114(18) 0.088(1) 19.273(38)

Table 3
Summary of experimental parameters and results for spallation-recovery experi­
ments. df: flyer plate thickness; ds: sample thickness; uf: flyer plate velocity; σH: peak
shock stress; : tensile strain rate; Δu: pullback velocity; σsp: spall strength; ar: re-
acceleration.

Shot df ds uf σH Δu σsp ar
No. (mm) (mm) (m s−1) (GPa) (105 s−1) (m s−1) (GPa) (108 m s−2)

SP1 1.041 1.988 0.193 4.071 0.67 120.54 2.502 1.41


SP2 1.004 2.009 0.222 4.684 0.73 122.58 2.545 2.67
SP3 1.003 2.002 0.275 5.819 0.79 123.79 2.570 3.63
SP4 1.004 1.960 0.309 6.555 0.82 119.46 2.480 3.70
SP5 1.002 2.008 0.355 7.563 0.86 118.30 2.456 3.77
SP6 1.007 2.008 0.460 9.911 0.96 115.08 2.389 3.78
SP7 1.007 2.005 0.606 13.285 1.07 110.32 2.290 3.83

velocities are converted to the Eulerian shock velocities. For the


elastic shock, us1 = usL1. For the plastic shock, us2 = ( 0 1) usL2 + up1.
1
up1 can be approximated as 2 ufs at the HEL. up2 can then be deduced
with the impedance match method [40]. The stress continuity is
satisfied across the interface between the Cu driver plate (denoted
with superscript c) and the sample, i.e.,
c c c
2 = 0 us up . (3)
c
For Cu, usc
= 3.91 + 1.51upc
(km s ), −1
= 8.955 g cm 0 [41], −3

and upc = uf up2.


The elastic precursor and HEL can be identified from re­
presentative free surface velocity histories of the Hugoniot EOS
shots. The HEL of CoCrFeNiCu is relatively low due to the Cu segre­
gation (the Cu-rich phase), similar to Cu [42,43] and other Cu-con­
taining composites [44,45].
The values of us1, up1, and σ1 are listed in Table 2. The average
1 2
value of yield stresses (σy) is 0.398 GPa, calculated as y = 1 1.
Given us1, up1, us2 , and uimp measured experimentally, up2 , σ2, and
ρ2 can be obtained from Eq. (1)–(3) (Table 2). In the following dis­
cussion, we also use us, up, and σH to denote us2 , up2 , and σ2, re­
spectively, unless stated otherwise. We also introduce the
volumetric strain (assuming compressive strain is positive), εH
= 1 − ρ0∕ρ2. The experimental values of us–up for the plastic shock
and σH–εH are plotted in Fig. 5.
Fig. 5. (a) Shock velocity versus particle velocity plot (us − up), and (b) peak shock
There normally exists a linear us–up relation for plastic shocks in
stress versus volumetric strain plot (σH − εH) for the as-cast CoCrFeNiCu. CB is the bulk
metals, i.e., us = C0 + λup. Here, C0 and λ are material parameters. sound speed at ambient condition. σHEL is the tress at the Hugoniot elastic limit.
Linear fitting to the experimental data yields
us = 4.41(2) + 1.63(8) up (km s 1), (4)
material [46], assuming an ideal mechanical mixing of the con­
for the CoCrFeNiCu HEA. tributions from individual elements, i.e., C0 = ∑miC0i, and λ = ∑miλi,
As an alternative, the mixture method is occasionally used as a where mi is the mass fraction of the ith element, and C0i and λi are
first-order estimate of the us–up relation for a multi-element the known Hugoniot parameters for corresponding elements [41].

4
L.X. Li, X.Y. Liu, J. Xu et al. Journal of Alloys and Compounds 947 (2023) 169650

Fig. 7. Spall strength σsp as a function of peak stress σH.

With increasing shock stress, shock-induced dislocation density


Fig. 6. Free surface velocity histories, ufs(t), for spallation-recovery experiments.
Numbers on the curves denote impact velocities uimp in m s−1.
increases, leading to increasing hindrance to dislocation slip, and
thus increased spall strength [57,58]. BMGs and brittle materials
The us–up relation for CoCrFeNiCu obtained via the mixture method have a poorer capability to accommodate plastic deformation in the
is us = 4.46 + 1.45up. The mixture method yields a reasonable esti­ form of dislocation or twinning. For CoCrFeNiCu with dual FCC
mate of the us–up relation. phases, experiments [17,59] and simulations [36,60] show excellent
plasticity mainly via dislocation slip.
As shown below, impact-induced strain localization occurs
3.2. Spallation mostly in the Cu-rich interdendritic regions, providing damage nu­
cleation sites and reducing the spall strength. In addition, ar in­
For spallation experiments, the free surface velocity histories, ufs creases considerably from 1.41 × 108 m s−2 to 3.63 × 108 m s−2 with
(t), are obtained (Fig. 6), labeled with A–G. Segments AB, CD, DE, EF peak stress increasing from 4.1 GPa to 5.8 GPa, and then remains
and FG represent the elastic precursor, plastic shock, shock plateau, nearly unchanged at higher stress (Table 3).
release and spall pullback, respectively.
Spall strength (σsp) is normally estimated from ufs(t) with the
acoustic method as [47]. 3.3. Deformation induced by shock compression

1 The SEM micrographs and corresponding kernel average mis­


sp 0 CL u CL
,
1+ orientation (KAM) maps for the as-received sample and the impact
CB (5)
surface of the postmortem sample of shot SP5 are shown in Fig. 8.
where Δu = ufs∣D − ufs∣E. Moreover, the tensile strain rate can be es­ For the as-received sample, the KAM map (Fig. 8b) indicates
timated as [48]. apparent residual dislocations within the interdendritic regions. Due
to the differences of physical and mechanical properties between the
1 dufs (t) two FCC phases, the mismatch in stress/strain can readily occur at
.
2CB dt EF (6) the DR-ID interface, where geometrically necessary dislocations
(GNDs) are nucleated. Then such dislocations propagate into the Cu-
The fracture rate during spallation [49] can be characterized with the rich phase with low stacking fault energy (SFE) [61,62].
re-acceleration (the pullback, segment FG), After shock compression, deformation is more severe within the
interdendritic regions. The average KAM values (ζav) increase from
dufs (t)
ar = . 0.11∘ to 0.28∘, and from 0.34∘ to 0.89∘ for the dendritic and inter­
dt FG (7) dendritic regions, respectively. ζav is usually correlated with the GND
density (ρGND) [63,64]:
The experimental parameters and results of eight spallation-re­
covery shots are summarized in Table 3. 2 av
Spall strength normally increases monotonically with increasing GND = ,
µb (8)
shock stress. However, spall strength of the CoCrFeNiCu HEA in­
creases (4.1–5.8 GPa) and then abnormally decreases (5.8–13.3 GPa) where μ is the the EBSD scan step size, b is the Burgers vector, b
with increasing peak stress (Fig. 7), similar to steel [50–52], Zr-based = a∕2 < 110 > , and a is the lattice parameter for the dendritic re­
bulk metallic glasses (BMGs) [53], and brittle materials (SiC, SiN, and gions or the interdendritic regions in Table 1. ρGND increases from
S2 glass composites) [54–56]. 0.9 × 1015 m−2 to 2.2 × 1015 m−2, and from 2.6 × 1015 m−2 to 7.0 × 1015

5
L.X. Li, X.Y. Liu, J. Xu et al. Journal of Alloys and Compounds 947 (2023) 169650

Fig. 8. (a) and (b) SEM micrographs and corresponding KAM maps for the unshocked sample and the impact surface of the postmortem sample (shocked to 7.6 GPa), respectively.
Shock direction: inward. (c) Corresponding KAM statistics for different regions. ζav: average KAM value.

m−2 for the dendritic and interdendritic regions, respectively. The detailed characterizations (like transmission electron microscopy)
interdendritic regions accommodate more plastic deformation, are needed.
leading to a significant increase in ζav and ρGND.
For dual-FCC-phase CoCrFeNiCu, the predominant deformation 3.4. Damage characterization
mechanism is dislocation slip [17,36]. The deformation localization
in the Cu-rich regions can be attributed to two reasons. Firstly, dis­ SEM characterizations are conducted on the postmortem sam­
locations nucleate at DR-ID interfaces due to lattice distortion, and ples from the spallation-recovery experiments. For shots SP6 and
then preferentially propagate into the lower SFE Cu-rich regions. The SP7, full spallation occurs and the postmortem samples each become
DR-ID interfaces can hinder dislocation emission from the Cu-rich two separate disks. The cross-sections of the central areas for shots
regions to the Cu-lean regions, further enhancing multiple slips in SP1–5 and the fractograph of the spall surface for SP6 are shown
the Cu-rich phase [65]. The DR-ID interfaces can promote serve in Fig. 9.
plastic deformation in the phase with lower SFE and restrict de­ With increasing shock stress (Fig. 9a), the degree of spall da­
formation in the other phase; deformation localization and spall mage increases, showing increasing number of isolated voids, void
void nucleation in the Cu-rich regions of the Cu-Ta and Cu-Nb sys­ coalescence cracks, increasing length of large cracks (the length of a
tems, and in the Ag-rich regions of the Cu-Ag system [57,66,67]. large crack is several times of the grain size), and the formation of
Secondly, the relatively small, abundant DR-ID interfaces can provide main cracks (7.6 GPa). Voids nucleate at the DR-ID interfaces and
more dislocation nucleation sites, and thus, the dislocation density the interior of the dendritic or interdendritic regions (Fig. 9b). In­
in the Cu-rich interdendritic regions increases significantly [68]. tragranular voids in the dendritic and interdendritic regions are
In addition, shock-induced deformation twins are not found in quasi-spherical, while voids at the interfaces are quasi-spherical or
dendritic regions as shown in the EBSD mapping of the impact elliptical due to the resolved shear stress around the DR-ID inter­
surface for σH = 7.6 GPa. Generally, the nucleation and growth of faces [36]. Under high shock stress, the two kinds of cracks which
deformation twins are facilitated by long pulse duration and high are formed via void coalescence (Fig. 9c and d) or crack tip opening
peak stress during shock compression [26,69]. The absence of the (Fig. 9e) coexist, and the cracks formed via void coalescence ac­
deformation twins can be attributed to the not high enough shock count for the majority. Damage characterizations suggest a mix
stress. Or the size of deformation twins are so small that other more mode of predominant ductile fracture and secondary brittle

6
L.X. Li, X.Y. Liu, J. Xu et al. Journal of Alloys and Compounds 947 (2023) 169650

Fig. 9. (a) SEM micrographs of the spall regions for shots SP1–5. (b) Magnified views of voids nucleated at three typical locations: dendritic region (DR), interdendritic region (ID)
and DR-ID interfaces, delimited by red, blue and black dotted circles, respectively. (c)–(f) Magnified views of two kinds of cracks formed via void coalescence and crack tip opening,
delimited by the rectangles in (a). (g) Magnified view and corresponding binarised image of the region delimited by the dotted rectangle in (a). Lc: crack length; Wc: crack width.
(h) SEM micrographs of the fracture surface for shot SP6. (i) and (j) Magnified views of the regions delimited by the rectangles in (h). Shock direction is upward for (a)–(g) and
inward for (h)–(g).

fracture. In Fig. 9f, cracks form and link in the interdendritic re­
gions, partially encircling a dendritic region.
As shown in Fig. 9h–j, the micrograph of the spall surface consists
of ductile dimples, voids and cracks of different sizes. The magnified
view of a large crack (Fig. 9j) exhibits relatively flat fracture surfaces,
sparse tearing ridges (the dotted ellipse), and obvious intergranular
grain protrusions, identified as a brittle crack. The protrusions have a
mean size of 50–60 μm comparable to the size of the Cu-lean regions
(Fig. 2), indicating that the Cu-lean regions are relatively intact
around the large crack. Since large cleavage planes or “river pat­
terns" are absent, and dimples are widely distributed on fracture
surfaces, ductile damage is also dominant in this case.
Spall strength is dictated by the initial stage of damage nuclea­
tion and growth [70,71]. To understand the features of damage nu­
cleation, the distributions of void nucleation sites under different
shock stresses are quantified in Fig. 10. The void types are dis­
tinguished by void geometric center [65], and few confused void
sites are not included in the statistical analysis. For each shot, more
than 200 voids are analyzed. For the morphology observed by SEM is Fig. 10. Fractions of microvoids at different nucleation regions as a function of peak
two-dimensional, some statistical voids nucleated at the inter­ stress.
dendritic regions may be expansion of nucleation at the DR-ID

7
L.X. Li, X.Y. Liu, J. Xu et al. Journal of Alloys and Compounds 947 (2023) 169650

As shown in Fig. 11, the crack density (ρc, crack number per unit
area) and mean crack area (Ac, total crack threshold area divided by
crack number) are quantified. Mean crack lengths (Lc) and mean
length-to-width ratios (Lc/Wc) are also included. As σH increases
from 4.1 GPa to 5.8 GPa, ρc firstly increases, while Ac, Lc and Lc/Wc
remain unchanged. At higher stresses, due to the coalescence of
crack, ρc changes slightly, while Ac, Lc and Lc/Wc increase con­
siderably.
At low impact velocities, dislocations can impede the growth of
the damage due to strain hardening, thus leading to the increasing in
spall strength (magenta region in Fig. 7) [77]. As impact velocity
increases, the Cu-rich interdendritic regions accommodate more
plastic deformation (Fig. 8) due to their lower stacking fault energy.
Then, severe plastic deformation (such as highly tangled dislocations
[78]) in interdendritic regions leads to strain localization. As the
degree of strain localization reaches a critical value, these strain
localization regions can act as nucleation sites [65] of interdendritic
voids. Via providing more void nucleation sites [51], the shock-in­
duced strain localization results in decreased spall strength (the
cyan region in Fig. 7).
Previous studies on the dendritic structure or other complex
phase structures demonstrated that precipitates or phase segrega­
tion can have a positive effect on yield strength or ductility under
quasi-static loading [11,79,80]. In this study, the dendritic structure
appears to have a negative effect on spall strength for CoCrFeNiCu
HEA under high strain rate loading.

3.5. MD simulations

To quantify the microstructure and damage evolution during


shock compression and spallation, the atomic configurations and
corresponding centro-symmetry parameter (CSP) [81] maps are
plotted in Fig. 12a–d at different instants. The atoms with CSP ≤ 4,
Fig. 11. (a) Overall density (ρc) and mean area of cracks (Ac) as a function of peak 4 < CSP ≤ 15 and CSP > 15 are assigned structure types of FCC, dis­
stress. (b) Mean crack length (Lc) and mean aspect ratio (Lc∕Wc) as a function of peak location/stacking fault, and void surface, respectively. In addition,
stress.
local kinetic energy (Ek) maps and normal stress profiles along the
shock direction are obtained via two- and one-dimensional binning
analysis [78,82], respectively (Fig. 12e–h).
interfaces into interdendritic regions. This statistical result is only The normal stress profile at 26 ps (Fig. 12g) shows the two release
used to reflect a certain trend with not rigorous numerical value. fans propagating toward each other. Within the compression area
With increasing peak stress, the proportion of interdendritic in­ (Fig. 12c), the Cu-rich regions contain more defects, consistent with
tragranular voids increases, while the proportion of DR-ID interface the KAM results in Fig. 8b. At 44 ps (Fig. 12h), the two release fans
voids decreases accordingly. As for the dendritic intragranular voids, encounter and their interactions induce release and an evolving
the proportion remains a very low level ( ∼ 10 %), indicating that the tensile region in the sample, and consequently spallation. Compared
dendritic regions are not the preferred sites for damage nucleation. to the Cu-lean regions, the Cu-rich regions are the preferred sites for
In general, DR-ID interfaces serve as the major weaker links and void nucleation (Fig. 12d). As shown in Fig. 12g and h, the atomic
preferred damage nucleation sites [72,73]. However, at high peak kinetic energy in the Cu-rich regions is higher than that in the Cu-
stress, the Cu-rich interdendritic regions serve as an important lean regions under shock compression and spallation. The impact-
secondary damage nucleation site for the as-cast CoCrFeNiCu HEA. induced temperature rise induces the decrease of critical stress for
To analyze the damage evolution (Fig. 11), cracks are extracted void nucleation with increasing spall temperature [83,84], providing
from the binarized SEM micrographs via threshold and top-hat more potential void nucleation sites in the Cu-rich regions, and
segmentation [74,75]. Crack length Lc and crack width Wc are leading to the decrease in spall strength. Thus, the shock-induced
quantified via gyration tensor analysis [52,76] as shown in Fig. 9g. strain localization (Fig. 12d) and thermal softening (Fig. 12f) [85–87]
For each shot, the area for statistics is 2.1 × 1.4 mm2. together contribute to the current deformation mechanisms.

8
L.X. Li, X.Y. Liu, J. Xu et al. Journal of Alloys and Compounds 947 (2023) 169650

Fig. 12. (a–f) Atomic configurations of multigrained CoCrFeNiCu during shock compression and spallation at different instants, with color-coding based on (a, b) local structure
type, (c, d) centro-symmetry parameter (CSP) and (e, f) atomic kinetic energy. (g, h) Corresponding normal stress (σxx) profiles along the shock direction.

4. Conclusions • The MD simulations show that pronounced strain localization


and thermal softening in the Cu-rich regions leads to the de­
A dendritic CoCrFeNiCu HEA consisting of a harder Cu-lean crease in spall strength at high peak shock stress.
dentritic phase and a softer Cu-rich interdendritic phase has been
investigated with plate impact loading and MD simulations. Our
main conclusions are as follows. CRediT authorship contribution statement

• The Hugoniot equation of state is obtained as us L. X. Li: Investigation, Visualization, Data curation, Writing -
= 4.41(2) + 1.63(8)up (km s−1), with ρ0 = 8.356 g cm−3. The yield original draft. X. Y. Liu: Resources. J. Xu: Investigation, Methodology.
stress is 0.398 GPa. With increasing peak shock stress, the spall S. C. Hu: Investigation, Methodology. Y. Cai: Data curation,
strength increases and then decreases, peaking at σH = 5.8 GPa Resources. L. Lu: Validation, Supervision, Resources. J. C. Cheng:
with σsp = 2.57 GPa. Methodology. Y. Tang: Methodology. C. Li: Data curation. N. B.
• The predominant deformation mechanism is dislocation slip. The Zhang: Validation, Supervision, Writing - review & editing. S. N.
Cu-rich interdendritic regions accommodate more plastic de­ Luo: Validation, Supervision, Writing - review & editing, Project
formation due to their lower stacking fault energy and administration, Funding acquisition, Resources.
smaller size.
• Both the ductile (dimples and void coalescence) and brittle
(tearing ridges and crack tip opening) fracture modes are ob­ Data Availability
served. At high peak stress (above ∼ 5.8 GPa), the interdendritic
regions with pronounced strain localization and more defects Data will be made available on request.
provide more void nucleation sites.

9
L.X. Li, X.Y. Liu, J. Xu et al. Journal of Alloys and Compounds 947 (2023) 169650

Declaration of Competing Interest [22] J. Li, L. Dong, X. Dong, W. Zhao, J. Liu, J. Xiong, C. Xu, Study on wear behavior of
FeNiCrCoCu high entropy alloy coating on Cu substrate based on molecular
dynamics, Appl. Surf. Sci. 570 (2021) 151236.
The authors declare that they have no known competing fi­ [23] H. Xie, Z. Ma, W. Zhang, H. Zhao, L. Ren, Phase transition in shock compressed
nancial interests or personal relationships that could have appeared high-entropy alloy FeNiCrCoCu, Int. J. Mech. Sci. 238 (2023) 107855.
to influence the work reported in this paper. [24] S.K. Singh, A. Parashar, Effect of lattice distortion and nanovoids on the shock
compression behavior of (Co-Cr-Cu-Fe-Ni) high entropy alloy, Comp. Mater. Sci.
209 (2022) 111402.
Acknowledgments [25] Z.J. Jiang, J.Y. He, H.Y. Wang, H.S. Zhang, Z.P. Lu, L.H. Dai, Shock compression
response of high entropy alloys, Mater. Res. Lett. 4 (4) (2016) 226–232.
[26] N.B. Zhang, J. Xu, Z.D. Feng, Y.F. Sun, J.Y. Huang, X.J. Zhao, X.H. Yao, S. Chen, L. Lu,
This work was sponsored in part by Sichuan Science and S.N. Luo, Shock compression and spallation damage of high-entropy alloy Al0.1
Technology Program (Grant No. 2022YFG0033), Natural Science CoCrFeNi, J. Mater. Sci. Technol. 128 (2022) 1–9.
Foundation of Sichuan Province (Grant Nos. 2022NSFSC0345 and [27] N.B. Zhang, Z.J. Tang, Z.H. Lin, S.Y. Zhu, Y. Cai, S. Chen, L. Lu, X.J. Zhao, S.N. Luo,
Deformation and damage of heterogeneous-structured high-entropy alloy
2023NSFSC1284) and National Natural Science Foundation of China
CrMnFeCoNi under plate impact, Mater. Sci. Eng. A (2022) 143069.
(Grant Nos. 12102491 and 11627901). [28] S.P. Zhao, Z.D. Feng, L.X. Li, X.J. Zhao, L. Lu, S. Chen, N.B. Zhang, Y. Cai, S.N. Luo,
Dynamic mechanical properties, deformation and damage mechanisms of eu­
References tectic high-entropy alloy AlCoCrFeNi2.1 under plate impact, J. Mater. Sci. Technol.
134 (2023) 178–188.
[29] W.H. Bragg, W.L. Bragg, The reflection of x-rays by crystals, Proc. R. Soc. Lond. 88
[1] B. Cantor, I.T.H. Chang, P. Knight, A.J.B. Vincent, Microstructural development (605) (1913) 428–438.
in equiatomic multicomponent alloys, Mater. Sci. Eng. A 375–377 (2004) [30] O.T. Strand, D.R. Goosman, C. Martinez, T.L. Whitworth, W.W. Kuhlow, Compact
213–218. system for high-speed velocimetry using heterodyne techniques, Rev. Sci.
[2] B. Gludovatz, A. Hohenwarter, D. Catoor, E.H. Chang, E.P. George, R.O. Ritchie, A Instrum. 77 (8) (2006) 083108.
fracture-resistant high-entropy alloy for cryogenic applications, Science 345 [31] N.K. Bourne, G. Gray, Computational design of recovery experiments for ductile
(6201) (2014) 1153–1158. metals.
[3] Y. Zhang, T.T. Zuo, Z. Tang, M.C. Gao, K.A. Dahmen, P.K. Liaw, Z.P. Lu, [32] R. Vignjevic, K. Hughes, T. De Vuyst, N. Djordjevic, J.C. Campbell, M. Stojkovic,
Microstructures and properties of high-entropy alloys, Prog. Mater. Sci. 61 (2014) O. Gulavani, S. Hiermaier, Lagrangian analysis led design of a shock recovery
1–93. plate impact experiment, Int. J. Impact Eng. 77 (2015) 16–29.
[4] E.P. George, W.A. Curtin, C.C. Tasan, High entropy alloys: a focused review of [33] D. Farkas, A. Caro, Model interatomic potentials and lattice strain in a high-en­
mechanical properties and deformation mechanisms, Acta Mater. 188 (2020) tropy alloy, J. Mater. Res. 33 (19) (2018) 3218–3225.
435–474. [34] S. Liu, G. Feng, L. Xiao, Y. Guan, W. Song, Shock-induced dynamic response in
[5] W. Li, D. Xie, D. Li, Y. Zhang, Y. Gao, P.K. Liaw, Mechanical behavior of high- single and nanocrystalline high-entropy alloy FeNiCrCoCu, Int. J. Mech. Sci. 239
entropy alloys, Prog. Mater. Sci. 118 (2021) 100777. (2023) 107859.
[6] S. Praveen, B.S. Murty, R.S. Kottada, Alloying behavior in multi-component [35] S. Plimpton, Fast parallel algorithms for short-range molecular dynamics, J.
AlCoCrCuFe and NiCoCrCuFe high entropy alloys, Mater. Sci. Eng. A 534 (2012) Comput. Phys. 117 (1) (1995) 1–19.
83–89. [36] H. Liu, C. Peng, X. Li, S. Wang, L. Wang, The effect of phase separation on the
[7] C. Du, L. Hu, Q. Pan, K. Chen, P. Zhou, G. Wang, Effect of Cu on the strengthening mechanical behavior of the Co–Cr–Cu–Fe–Ni high-entropy alloy, Materials 14
and embrittling of an FeCoNiCr-xCu HEA, Mater. Sci. Eng. A 832 (2022) 142413. (21) (2021) 6523.
[8] H. Zheng, R. Chen, G. Qin, X. Li, Y. Su, H. Ding, J. Guo, H. Fu, Microstructure [37] B. Holian, Atomistic computer simulations of shock waves, Shock Waves 5 (3)
evolution, Cu segregation and tensile properties of CoCrFeNiCu high entropy (1995) 149–157.
alloy during directional solidification, J. Mater. Sci. Technol. 38 (2020) 19–27. [38] L. Davison, Fundamentals of Shock Wave Propagation in Solids, Springer Science
[9] B. Li, Y. Wang, M. Ren, C. Yang, H. Fu, Effects of Mn, Ti and V on the micro­ & Business Media, 2008.
structure and properties of AlCrFeCoNiCu high entropy alloy, Mater. Sci. Eng. A [39] D.H. Dolan, Accuracy and precision in photonic Doppler velocimetry, Rev. Sci.
498 (1–2) (2008) 482–486. Instrum. 81 (5) (2010) 053905.
[10] X. Xian, L. Lin, Z. Zhong, C. Zhang, C. Chen, K. Song, J. Cheng, Y. Wu, Precipitation [40] A.C. Mitchell, W.J. Nellis, Shock compression of aluminum, copper, and tantalum,
and its strengthening of Cu-rich phase in CrMnFeCoNiCux high-entropy alloys, J. Appl. Phys. 52 (5) (1981) 3363–3374.
Mater. Sci. Eng. A 713 (2018) 134–140. [41] S.P. Marsh, LASL Shock Hugoniot Data, University of California Press, 1980.
[11] Z. Li, K.G. Pradeep, Y. Deng, D. Raabe, C.C. Tasan, Metastable high-entropy dual- [42] R. Chau, J. Stölken, P. Asoka-Kumar, M. Kumar, N.C. Holmes, Shock Hugoniot of
phase alloys overcome the strength–ductility trade-off, Nature 534 (7606) (2016) single crystal copper, J. Appl. Phys. 107 (2) (2010) 023506.
227–230. [43] A.C. Mitchell, W.J. Nellis, J.A. Moriarty, R.A. Heinle, N.C. Holmes, R.E. Tipton,
[12] X.W. Liu, L. Liu, G. Liu, X.X. Wu, D.H. Lu, J.Q. Yao, W.M. Jiang, Z.T. Fan, W.B. Zhang, G.W. Repp, Equation of state of Al, Cu, Mo, and Pb at shock pressures up to 2. 4
The role of carbon in grain refinement of cast CrFeCoNi high-entropy alloys, TPa (24 Mbar), J. Appl. Phys. 69 (5) (1991) 2981–2986.
Metall. Mat. Trans. A 49 (6) (2018) 2151–2160. [44] W.Z. Han, E.K. Cerreta, N.A. Mara, I.J. Beyerlein, J.S. Carpenter, S.J. Zheng,
[13] J.T. Fan, L.J. Zhang, P.F. Yu, M.D. Zhang, G. Li, P.K. Liaw, R.P. Liu, A novel high- C.P. Trujillo, P.O. Dickerson, A. Misra, Deformation and failure of shocked bulk
entropy alloy with a dendrite-composite microstructure and remarkable com­ Cu–Nb nanolaminates, Acta Mater. 63 (2014) 150–161.
pression performance, Scr. Mater. 159 (2019) 18–23. [45] Y. Yang, C. Wang, X. Chen, H. Hu, K. Chen, Y. Fu, Effects of the phase interface on
[14] Y. Jung, K. Lee, S.J. Hong, J.K. Lee, J. Han, K.B. Kim, P.K. Liaw, C. Lee, G. Song, spallation damage nucleation and evolution in multiphase alloy, J. Alloy. Compd.
Investigation of phase-transformation path in TiZrHf(VNbTa)x refractory high- 740 (2018) 321–329.
entropy alloys and its effect on mechanical property, J. Alloy. Compd. 886 (2021) [46] M.A. Meyers, Dynamic behavior of materials john wiley & sons Inc N. Y. (1994)
161187. 31–40.
[15] T. Zhang, R.D. Zhao, F.F. Wu, S.B. Lin, S.S. Jiang, Y.J. Huang, S.H. Chen, J. Eckert, [47] V. Romanchenko, G. Stepanov, Dependence of the critical stresses on the loading
Transformation-enhanced strength and ductility in a FeCoCrNiMn dual phase time parameters during spall in copper, aluminum, and steel, J. Appl. Mech. Tech.
high-entropy alloy, Mater. Sci. Eng. A 780 (2020) 139182. Phys. 21 (4) (1980) 555–561.
[16] R. Fan, L. Wang, L. Zhao, L. Wang, S. Zhao, Y. Zhang, B. Cui, Synergistic effect of Nb [48] B. Arman, S.N. Luo, T.C. Germann, T. Çağın, Dynamic response of Cu46 Zr54 me­
and Mo alloying on the microstructure and mechanical properties of CoCrFeNi tallic glass to high-strain-rate shock loading: Plasticity, spall, and atomic-level
high entropy alloy, Mater. Sci. Eng. A 829 (2022) 142153. structures, Phys. Rev. B 81 (14) (2010) 144201.
[17] S.M. Oh, S.I. Hong, Microstructural stability and mechanical properties of [49] G. Kanel, A. Utkin, Estimation of the spall fracture kinetics from the free-surface
equiatomic CoCrCuFeNi, CrCuFeMnNi, CoCrCuFeMn alloys, Mater. Chem. Phys. velocity profiles, in: AIP Conference Proceedings, Vol. 370, American Institute of
210 (2018) 120–125. Physics, 1996, 487–490.
[18] E. Zhou, D. Qiao, Y. Yang, D. Xu, Y. Lu, J. Wang, J.A. Smith, H. Li, H. Zhao, P.K. Liaw, [50] W. Wang, H. Zhang, M. Yang, P. Jiang, F. Yuan, X. Wu, Shock and spall behaviors of
F. Wang, A novel Cu-bearing high-entropy alloy with significant antibacterial a high specific strength steel: effects of impact stress and microstructure, J. Appl.
behavior against corrosive marine biofilms, J. Mater. Sci. Technol. 46 (2020) Phys. 121 (13) (2017) 135901.
201–210. [51] C. Li, K. Yang, X.C. Tang, L. Lu, S.N. Luo, Spall strength of a mild carbon steel:
[19] A. Verma, P. Tarate, A.C. Abhyankar, M.R. Mohape, D.S. Gowtam, V.P. Deshmukh, effects of tensile stress history and shock-induced microstructure, Mater. Sci.
T. Shanmugasundaram, High temperature wear in CoCrFeNiCux high entropy Eng. A 754 (2019) 461–469.
alloys: the role of Cu, Scr. Mater. 161 (2019) 28–31. [52] K. Yang, Y. Chen, X. Liu, C. Li, H. Chen, J.Y. Huang, S.N. Luo, Spall properties and
[20] T. Gao, H. Song, B. Wang, Y. Gao, Y. Liu, Q. Xie, Q. Chen, Q. Xiao, Y. Liang, damage mechanisms of a low-alloy steel fabricated via laser powder bed fusion,
Molecular dynamics simulations of tensile response for FeNiCrCoCu high-en­ Mater. Sci. Eng. A 840 (2022) 142910.
tropy alloy with voids, Int. J. Mech. Sci. 237 (2023) 107800. [53] F. Yuan, V. Prakash, J.J. Lewandowski, Spall strength and Hugoniot elastic limit of
[21] Y. Shen, D.E. Spearot, Mobility of dislocations in FeNiCrCoCu high entropy alloys, a zirconium-based bulk metallic glass under planar shock compression, J. Mater.
Model. Simul. Mater. SC 29 (8) (2021) 085017. Res. 22 (2) (2007) 402–411.

10
L.X. Li, X.Y. Liu, J. Xu et al. Journal of Alloys and Compounds 947 (2023) 169650

[54] P. Bartkowski, D.P. Dandekar, Spall strengths of sintered and hot pressed silicon [70] E.M. Bringa, S. Traiviratana, M.A. Meyers, Void initiation in fcc metals: effect of
carbide, in: AIP Conference Proceedings, Vol. 370, AIP, Seattle, Washington loading orientation and nanocrystalline effects, Acta Mater. 58 (13) (2010)
(USA), 1996, 535–538. 4458–4477.
[55] D. Nathenson, V. Prakash, D.P. Dandekar, Dynamic response of silicon nitride under [71] V. Lubarda, M. Schneider, D. Kalantar, B. Remington, M. Meyers, Void growth by
combined pressure and shear impact, in: Proceedings of the 2005 SEM Annual dislocation emission, Acta Mater. 52 (6) (2004) 1397–1408.
Conference and Exposition on Experimental and Applied Mechanics, 2005. [72] D.L. Steinbrunner, D.K. Matlock, G. Krauss, Void formation during tensile testing
[56] L. Tsai, V. Prakash, Dynamic response and spall strength of S2-glass fiber re­ of dual phase steels, MTA 19 (3) (1988) 579–589.
inforced polymer composites, in: Proceedings of the 2005 SEM Annual [73] S.K. Yerra, G. Martin, M. Véron, Y. Bréchet, J.D. Mithieux, L. Delannay, T. Pardoen,
Conference and Exposition on Experimental and Applied Mechanics, 2005. Ductile fracture initiated by interface nucleation in two-phase elastoplastic
[57] J. Chen, S.N. Mathaudhu, N. Thadhani, A.M. Dongare, Correlations between dis­ systems, Eng. Fract. Mech. 102 (2013) 77–100.
location density evolution and spall strengths of Cu/Ta multilayered systems at the [74] J. Serra, P. Soille, Mathematical Morphology and Its Applications to Image
atomic scales: The role of spacing of KS interfaces, Materialia 5 (2019) 100192. Processing, vol. 2 Springer Science & Business Media, 2012.
[58] K. Mackenchery, R.R. Valisetty, R.R. Namburu, A. Stukowski, A.M. Rajendran, [75] G.V. Tcheslavski, et al., Morphological image processing: gray-scale morphology,
A.M. Dongare, Dislocation evolution and peak spall strengths in single crystal ELEN 4304 (2010) 5365.
and nanocrystalline Cu, J. Appl. Phys. 119 (4) (2016) 044301. [76] J.C. Cheng, H.W. Chai, G.L. Fan, Z.Q. Li, H. Xie, Z.Q. Tan, B.X. Bie, J.Y. Huang,
[59] P. Yu, Y. Zhuang, J.P. Chou, J. Wei, Y. Lo, A. Hu, The influence of dilute aluminum S.N. Luo, Anisotropic spall behavior of CNT/2024Al composites under plate im­
and molybdenum on stacking fault and twin formation in FeNiCoCr-based high pact, Carbon 170 (2020) 589–599.
entropy alloys based on density functional theory, Sci. Rep. 9 (1) (2019) 10940. [77] S.V. Razorenov, G.I. Kanel, G.V. Garkushin, O.N. Ignatova, Resistance to dynamic
[60] J. Liu, Molecular dynamic study of temperature dependence of mechanical deformation and fracture of tantalum with different grain and defect structures,
properties and plastic inception of CoCrCuFeNi high-entropy alloy, Phys. Lett. A Phys. Solid State 54 (4) (2012) 790–797.
384 (22) (2020) 126516. [78] S.N. Luo, Q. An, T.C. Germann, L.-B. Han, Shock-induced spall in solid and liquid
[61] J.P. Hirth, J. Lothe, T. Mura, Theory of dislocations, J. Appl. Mech. 50 (2) (1983) Cu at extreme strain rates, J. Appl. Phys. 106 (1) (2009) 013502.
476. [79] A.J. Ardell, Precipitation hardening, Metall. Trans. A 16 (12) (1985) 2131–2165.
[62] K. Kang, J. Wang, I.J. Beyerlein, Atomic structure variations of mechanically stable [80] J. He, H. Wang, H. Huang, X. Xu, M. Chen, Y. Wu, X. Liu, T. Nieh, K. An, Z. Lu, A
fcc-bcc interfaces, J. Appl. Phys. 111 (5) (2012) 053531. precipitation-hardened high-entropy alloy with outstanding tensile properties,
[63] C. Moussa, M. Bernacki, R. Besnard, N. Bozzolo, About quantitative EBSD analysis Acta Mater. 102 (2016) 187–196.
of deformation and recovery substructures in pure Tantalum, in: IOP Conference [81] C.L. Kelchner, S. Plimpton, J. Hamilton, Dislocation nucleation and defect struc­
Series: Materials Science and Engineering, Vol. 89, IOP Publishing, 2015, 012038. ture during surface indentation, Phys. Rev. B 58 (17) (1998) 11085.
[64] L. Yang, Z. Chen, X. Ma, D. Zhong, X. Zhao, L. Xiao, X. Fang, Improvement of [82] S.N. Luo, T.C. Germann, T.G. Desai, D.L. Tonks, Q. An, Anisotropic shock response
strength and ductility in a gradient structured Ni fabricated by severe torsion of columnar nanocrystalline Cu, J. Appl. Phys. 107 (12) (2010) 123507.
deformation, Mater. Sci. Eng. A 826 (2021) 141980. [83] M. Xiang, H. Hu, J. Chen, Y. Liao, Molecular dynamics studies of thermal
[65] M. Cheng, C. Li, M.X. Tang, L. Lu, Z. Li, S.N. Luo, Intragranular void formation in dissipation during shock induced spalling, J. Appl. Phys. 114 (12) (2013)
shock-spalled tantalum: mechanisms and governing factors, Acta Mater. 148 123509.
(2018) 38–48. [84] S. Razorenov, A. Bogatch, G. Kanel, A. Utkin, V. Fortov, D. Grady, Elastic-plastic
[66] J. Chen, S.N. Mathaudhu, N. Thadhani, A.M. Dongare, Unraveling the role of in­ deformation and spall fracture of metals at high temperatures, AIP Conf. Proc.,
terfaces on the spall failure of Cu/Ta multilayered systems, Sci. Rep. 10 (1) (2020) Vol. 429, American Institute of Physics, 1998, pp. 447–450.
208. [85] Y. Guan, W. Song, Y. Wang, S. Liu, Y. Yu, Dynamic responses in shocked Cu-Zr
[67] N. Gupta, M.I. Baskes, S.G. Srinivasan, The role of interface structure in spallation nanoglasses with gradient microstructure, Int. J. Plast. 149 (2022) 103154.
of a layered nanocomposite, JOM 63 (9) (2011) 74–77. [86] Y. Wang, X. Zeng, X. Yang, T. Xu, Shock-induced spallation in single-crystalline
[68] J. Jiang, T. Britton, A. Wilkinson, Evolution of dislocation density distributions in tantalum at elevated temperatures through molecular dynamics modeling,
copper during tensile deformation, Acta Mater. 61 (19) (2013) 7227–7239. Comp. Mater. Sci. 201 (2022) 110870.
[69] Z.H. Zeng, X.Z. Li, C. Li, L. Lu, H. Zhang, S.N. Luo, Deformation twinning in a mild [87] X. Tian, J. Cui, K. Ma, M. Xiang, Shock-induced plasticity and damage in single-
steel: loading dependence and strengthening, Mater. Sci. Eng. A 751 (2019) crystalline Cu at elevated temperatures by molecular dynamics simulations, Int.
332–339. J. Heat. Mass Tran. 158 (2020) 120013.

11

You might also like