You are on page 1of 18

Query Form

SREP
Journal: 41598 [Art. ID: 30601]

Journal: SREP
Author: The following queries have arisen during the editing of your manuscript. Please answer queries by making the
requisite corrections at the appropriate positions in the text.

Query Details Required Author’s Response


AQ1 Please confirm if the author names are presented accurately and in the correct sequence
(given name, middle name/initial, family name) for authors A. S. M. Jannatul Islam,
Md. Sayed Hasan, Md. Sherajul Islam and amend if necessary.
AQ2 Kindly check and confirm the layout of Tables 1 and 2.

Journal : SREP 41598 Article No : 30601 Pages : 1 MS Code : 30601 Dispatch : 27-2-2023
www.nature.com/scientificreports

OPEN Crystal orientation‑dependent


tensile mechanical behavior
and deformation mechanisms

F
of zinc‑blende ZnSe nanowires

O
A. S. M. Jannatul Islam 1, Md. Sayed Hasan 1, Md. Sherajul Islam 1*, A. G. Bhuiyan 1,

O
Catherine Stampfl 2 & Jeongwon Park 3,4

PR
AQ1 Crystal deformation mechanisms and mechanical behaviors in semiconductor nanowires (NWs), in
particular ZnSe NWs, exhibit a strong orientation dependence. However, very little is known about
tensile deformation mechanisms for different crystal orientations. Here, the dependence of crystal
orientations on mechanical properties and deformation mechanisms of zinc-blende ZnSe NWs are
explored using molecular dynamics simulations. We find that the fracture strength of [111]-oriented
D
ZnSe NWs shows a higher value than that of [110] and [100]-oriented ZnSe NWs. Square shape ZnSe
NWs show greater value in terms of fracture strength and elastic modulus compared to a hexagonal
TE
shape at all considered diameters. With increasing temperature, the fracture stress and elastic
modulus exhibit a sharp decrease. It is observed that the {111} planes are the deformation planes at
lower temperatures for the [100] orientation; conversely, when the temperature is increased, the
{100} plane is activated and contributes as the second principal cleavage plane. Most importantly, the
EC

[110]-directed ZnSe NWss show the highest strain rate sensitivity compared to the other orientations
due to the formation of many different cleavage planes with increasing strain rates. The calculated
radial distribution function and potential energy per atom further validates the obtained results.
This study is very important for the future development of efficient and reliable ZnSe NWs-based
RR

nanodevices and nanomechanical systems.

It has recently been demonstrated e­ xperimentally1,2 and ­theoretically3,4 that the mechanical behavior in semicon-
ductor nanowires (NWs) exhibits a strong orientation dependence. Significant anisotropic mechanical properties
CO

along different crystal orientations are observed in NW s­ ystems5–7. Other physical properties, including electrical
and thermal conductivity, piezoelectric polarization, refraction index, surface reactivity, and bandgap can be
precisely altered by controlling the crystal growth ­directions2,8–10. There are differences among the side surfaces
in wurtzite/zinc-blende NWs that grow in different directions. These differences can significantly alter the defor-
mation mechanism over perfect and partial dislocation slipping and deformation ­twinning11–15. Moreover, when
N

crystal orientation is coupled with temperature and strain rate, different deformation mechanisms occur in NWs
due to competition between global and local deformations, activation variation of different planes, and influ-
U

ence disparity of inter-planar d­ istances16–21. Thus, a deeper understanding of the crystal orientation-dependent
properties is essential for the production and functionality of NWs in their real-world applications.
On the other hand, the unique applications, microscopic physics, and fabrication of nanoscale devices, such as
laser diodes, photodetectors, field effect transistors (FETs), and solar cells, make semiconductor NWs extremely
promising in these research fields. In particular, zinc-blend ZnSe N ­ Ws22 have garnered substantial attention as
next-generation nanoelectronic materials owing to their outstanding performance with mechanical flexibility,
transmittance, conductivity, and inexpensive s­ ynthesis23–26. In addition, the direct bandgap (~ 2.7 eV), high
absorption coefficient, suitable electronegativity, and unique nonlinear properties of zinc-blende ZnSe make
it potentially an important material for blue lasing applications, optical waveguides, thermoelectric systems,

1
Department of Electrical and Electronic Engineering, Khulna University of Engineering and Technology,
Khulna 9203, Bangladesh. 2School of Physics, The University of Sydney, Sydney, NSW 2006, Australia. 3Department
of Electrical and Biomedical Engineering, University of Nevada, Reno, NV 89557, USA. 4School of Electrical
Engineering and Computer Science, University of Ottawa, Ottawa, ON K1N 6N5, Canada. *email: sheraj_kuet@
eee.kuet.ac.bd

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 1


Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023
Vol.:(0123456789)
www.nature.com/scientificreports/

light-emitting diodes, nanosensors, nanoactuators, nanoresonators, and magnetic information s­ torage27–37.


However, despite considerable effort to assess the electronic, thermal, and optical characteristics of zinc-blende
ZnSe NWs, both in ­theory23,24,26,38,39 and in e­ xperiments40–46, to the best of our knowledge, there is no study on
their mechanical properties. In particular, the tensile deformation mechanisms for different crystal orientations
have not been reported in the literature. Furthermore, the mechanical strength variation under diverse external
variables such as temperature and strain rate coupled with orientation is also unknown.
In this work, we have employed molecular dynamics (MD) simulations to expose the influence of crystal
orientations such as [100], [110], and [111] at different temperatures (ranging from 100 to 600 K) and strain rates
(varying from 1 × ­108 ­s−1 nm to 1 × ­1010 ­s−1) on the tensile mechanical properties and deformation mechanisms
of zinc-blende ZnSe NWs. To describe the atomistic interaction of zinc-blende ZnSe NWs systems, the classical
Stillinger–Weber (SW) potential developed by Zhou et al.47 was utilized. In addition, we have calculated the radial
distribution function and potential energy per atom to elucidate the orientation-dependent mechanical behav-
ior and deformation mechanisms. This work clarifies the dependence of crystal orientation on the mechanical
properties of ZnSe NWs and brings practical applications of ZnSe-based nanodevices one step further.

Computational details

F
The tensile mechanical characteristics and deformation mechanisms of zinc-blende ZnSe NW were explored
using MD simulation using the LAMMPS p ­ rogram48. To integrate the usual Newton equations of motion in time,

O
the Velocity Verlet approach with a time step of 1 fs was used. The Stillinger–Weber (SW) potential developed by
Zhou et al.47 was used for defining the interatomic interactions between atoms of the Zn–Se, Zn–Zn, and Se–Se
systems. The NW models are created by forming a rectangular box of zinc-blende ZnSe with a = 5.6676 Å as

O
the lattice constant and then creating NWs from a defined rectangular pillar using the Atomsk t­ ool49. The study
considers loading along three crystal orientations: [100], [110], and [111]. We considered rectangular nanopil-
lar structures consisting of ~ 23,520 to 23,800 atoms with a dimension of ~ (34.35 nm × 3.97 nm × 3.97 nm) by

PR
maintaining a length-to-width ratio of 8.57:116,50. Figure 1 shows the [100], [110], and [111] crystal-oriented
zinc-blende ZnSe NW models with a cross-sectional area of ~ 15.76 ­nm2. The tensile mechanical characteristics
of the different crystal-oriented zinc-blende ZnSe NWs were examined in relation to different temperatures and
strain rates. Temperatures were varied from 100 to 600 K, while strain rates varied from 1 × ­108 ­s−1 to 1 × ­1010 ­s−1.
To guarantee adequate space for the NWs to deform freely, a relatively large surface of 10 nm by 10 nm in
D
the x–y plane was used. To further examine the impact of shape and diameter on the mechanical behavior of
[111]-oriented ZnSe NWs, we also considered hexagonal structure in addition to square structure. The thickness
TE
and diameter for both shapes varied from 4 to 16 nm. Square and hexagonal-shaped ZnSe NWs with different
thicknesses and diameters are shown in Fig. S1 (Supporting Information).
The conjugate gradient method was used to perform the energy minimization. After that, the system was
relaxed in the NPT ensemble (constant atom number, pressure, and temperature) at a certain temperature with
zero pressure applied in the axial direction for 40 ps. Finally, the system is thermally relaxed for 40 ps using a
EC

canonical (NVT) ensemble. In these procedures, a Nose–Hoover thermostat is used to control the temperature.
The periodic boundary condition was used in the axial direction, while the fixed boundary conditions were
used in the other two directions. We examined the energy conservation of zinc-blende ZnSe NWs in the NVE
ensemble (constant atom number, volume, and energy) to verify the reliability of the time step used in our MD
RR

simulations. As demonstrated in Fig. S2a–c (Supporting Information), the total energy of NWs considered in
this work ([111], [110], and [100]-oriented) remains almost constant over the NVE simulations, demonstrating
that the time step used in our MD simulations is appropriate. Furthermore, the lattice constant of zinc-blende
ZnSe NWs was discovered to be 5.6682 Å in the x direction, which agrees well with the experiments by Karzel
et al.51 (5.6676 Å) and first-principles calculations by Okoye et al.38 (5.666 Å), Khenata et al.52 (5.624 Å), and
CO

Wang et al.53 (5.633 Å). This result indicates that the SW potential utilized in this work may adequately explain
atom interactions in the ZnSe crystal. Moreover, to explore the deformation mechanisms, all visualizations are
done with the OVITO p ­ ackage54.
After the energy minimization and structural relaxation, we employed uniaxial tensile loading along the
length direction of the NWs at a constant strain rate of 1­ 09 ­s−1. This is a favorable strain rate for MD simulations
N

because of the computational restrictions, and it has been used effectively in several previous ­studies16,50. Because
a considerably lower strain rate is being used experimentally, different strain rates were also considered to explore
the influence of strain rate. In our models, the atomic stresses are estimated using the Virial stress ­theorem55,
U

which is acquired as the arithmetic mean of the native stresses on all atoms and has the resulting expression:
 
1  1
σvirial (r) = −mi u̇i ⊗ u̇i + rij ⊗ fij (1)
� i 2 j�=i

where Ω denotes the total volume of the NW, ­mi denotes the mass of the atom i, u̇i denotes the velocity com-
ponent of atom i, ⊗ is the cross product, rij denotes the distance between atom i and j and fij is the force exerted
by atom j on atom i.

Results and discussion


In group II–VI and III–V binary semiconductor NW systems, due to the crystal polarity of the material and the
combination of low energy and high symmetry of the [111] direction, the ZnSe NWs tend to grow in a zinc-
­ irection56. However, in recent years, non-[111]-oriented semiconductor
blende cubic structure along the [111] d
NWs have attracted increasing interest in terms of fundamental research and promising applications due to their
outstanding crystal quality and distinctive physical ­properties57–59. Introducing numerous growth techniques,

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 2

Vol:.(1234567890) Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023
www.nature.com/scientificreports/

F
O
O
PR
D
TE
EC
RR

Figure 1.  Initial structure of the (a) [100]-directed, (b) [110]-directed and (c) [111]-directed zinc-blende
ZnSe NW with a dimension of ~ 34 nm × 3.97 nm × 3.97 nm. The cross-sectional area of all the NWS models is
CO

~ 15.76 ­nm2.

including chemical vapor deposition and the vapor–liquid–solid growth mode of low-pressure metal–organic
N

vapor-phase epitaxy, the selective growth of stacking fault free ZnSe NWs along [100], [110], and [112] crystal
orientations is ­reported56,60. By maintaining a cross connection among the NWs diameter, length, and growth
temperatures, 99% of the total ZnSe NWs are shown to be grown either along [100], [110], [111], or [112] growth
U

directions and are verified by photoluminescence spectroscopy, X-ray diffraction, and transmission electron
­microscopy59,61. Understanding the mechanical properties and deformation behavior of ZnSe NWs grown along
the [100], [110], and [111] directions is thus critical for developing and optimizing their applications in various
fields, such as nanoelectronics and nanoelectromechanical systems.
Firstly, we investigate the significance of crystal orientation on the stress–strain response behavior of the
zinc-blende ZnSe NWs. Figure 2a reveals the room temperature stress–strain plots of a square cross-section zinc-
blende ZnSe NWs subjected to uniaxial tensile loading at a strain rate of ­109 ­s−1 in the three crystal orientations
[100], [110], and [111]. As can be observed, the curves are mostly made up of elastic and plastic phases. We can
see from these curves that at 300 K, [111]-directed zinc-blende ZnSe has the highest fracture stress behavior,
whereas the [100] loading direction has the lowest. Analogous types of orientation-dependent fracture stress
behavior were also reported for CdSe, CdTe, InP, Ni-Co, and SiGe ­NWs16–18,62,63. Conversely, the [100]-oriented
zinc-blende ZnSe displays more significant fracture strain than the other two crystal directions. The reason
behind this greater fracture strain required to deform the [100]-oriented NW will be discussed later in this sec-
tion. The area underneath the stress–strain behavior can be used to calculate fracture toughness, which is the
extent of energy required prior to f­ ailure16. The exact value of the fracture toughness of bulk ZnSe might differ
depending on several factors, including grain size, crystal structure, and the presence of defects or impurities. On

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 3


Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023 Vol.:(0123456789)
www.nature.com/scientificreports/

F
O
O
PR
D
TE
Figure 2.  (a) Stress–strain response and (b) potential energy/atom of 15.76 n
­ m2 zinc-blende ZnSe NWs for
different crystal orientations at 300 K.
EC

the other hand, the fracture toughness of ZnSe NWs can depend on the size and shape of the NW, the crystalline
direction, and the orientation of the material with respect to the applied s­ tress16,50. This study found that due to
RR

the favorable crystal structure, [111]-directed ZnSe NW has the highest fracture toughness, while the [100] load-
ing direction has the lowest value. Afterward, using a graphical technique, we calculated the elastic modulus for
all NWs systems based on the slope of the stress–strain curve in the linear region (up to ≤ 1% strain). The small
strain region ensures that the configuration follows Hook’s law and that linear elastic destruction is achieved.
CO

The estimated elastic modulus for three crystal orientations of zinc-blende ZnSe NWs at 300 K with a thickness
of ~ 4 nm is found to be 50.08 GPa, 72.03 GPa, and 74.46 GPa, respectively, along the [100], [110], and [111]
orientations. The [111] direction has the maximum elastic modulus of ~ 74.46 GPa, indicating that the ZnSe
NW has high stiffness and is difficult to deform. In the [111] loading direction, the maximum fracture stress and
elastic modulus are caused by the lowest surface energy of the materials. Previous research has shown that the
N

[111] crystal orientation has the longest surface atomic distance and thus the smallest concentration of surface
atoms, allowing it to have the lowest surface energy, followed by the [110] and [100] crystal ­orientations64,65.
Hence, when a tensile force is applied in the [111] direction, the [111] surface with the lowest energy has the
U

most significant capacity to resist f­ racture64,65. To explain the effect of NW orientations on the stress–strain
performance, we consider the zoomed-in view of the surface of the zinc-blende ZnSe NWs along the three ori-
entations, as shown in Fig. 1a–c. From the zoomed-in view, it can be seen that the [100]-directed NW surface
shows a very dense structure compared to the [110] and [111]-directed NW surfaces. Therefore, the surface
atomic distance of the [100]-oriented NW will show a smaller value than the other two directions. Moreover, to
signify the surface atomic distance of different NWs, we have calculated the potential energy per atom of [111],
[110], and [100]-oriented ZnSe NWs, which are shown in Fig. 2b. From the figure, it has been noticed that as
[111]-oriented NW shows very low surface density compared to the other two directions, the potential energy
per atom at zero strain shows a lower value for that direction, which signifies its superior structural stability and
strength, including fracture strength and elastic modulus, compared to the other directions.
Moreover, zinc-blende ZnSe NWs with [111] orientation can be grown in different shapes with thickness
ranges from 8 to 30 nm or a­ bove1,2,56,60,61,66 using molecular-beam epitaxy and the vapor–liquid–solid mecha-
nisms. To make the next generation of nanotechnology more enduring and lightweight, researchers and experi-
mentalists are also working hard to develop smaller diameter N ­ Ws2. Hence, in this study, we have also studied
the mechanical behavior such as fracture stress and elastic modulus of [111]-oriented square and hexagonal-
shaped ZnSe NWs with thickness varying from ~ 4 to ~ 16 nm. The stress–strain relationship of square and

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 4

Vol:.(1234567890) Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023
www.nature.com/scientificreports/

hexagonal-shaped ZnSe NWs at 300 K for four different thicknesses or diameters throughout the tensile process
is shown in Fig. 3a,b, respectively. The results show that for both types of shapes, the ultimate tensile strength
increases as the thickness or diameter of the NWs increases; this outcome is consistent with other varieties of
semiconductor ­NWs5,16,67–69. In particular, as the thickness increases from ~ 4 to ~ 16 nm, the ultimate tensile
strength increases from 12.24 to 16.12 GPa for square-shaped ZnSe NWs. On the other hand, the ultimate ten-
sile strength increases from 3.81 to 4.08 GPa for hexagonal-shaped ZnSe NWs with the increase in thickness
from ~ 4 to ~ 16 nm. For both square and hexagonal-shaped ZnSe NWs with different thicknesses, the elastic
moduli are then determined using stress–strain curves with (strain ≤ 1%) linear regression and are shown in AQ2
Table 1. We compared our computed elastic modulus of zinc-blende ZnSe NWs at 300 K with previous pub-
lished ­studies52,53,70,71, as shown in Table 1, to validate the interatomic potential and computational approach
employed in this work. The outcomes show the possibility of correctly predicting the mechanical characteristics
of the ZnSe nanostructures using the SW interatomic potential. Previous research discovered that elasticity is
size-dependent2 and is restricted to NWs with diameters less than 20 nm. Therefore, our work also shows a
similar size dependency, which confirms our investigation. Most importantly, the square-shaped ZnSe NWs for
all considered thicknesses show greater mechanical behavior than hexagonal-shaped ZnSe NWs. The potential

F
O
O
PR
D
TE
EC
RR
CO
N
U

Figure 3.  Effect of thickness or diameter on the stress–strain response curves of [111]-oriented zinc-blende
ZnSe NWs in (a) square shape and (b) hexagonal shape at 300 K. Thickness-dependent potential energy/atom of
[111]-oriented zinc-blende ZnSe NWs in (c) square and (d) hexagonal shapes at 300 K.

Elastic modulus (GPa) found in this study


Diameter/thickness (nm) Square shape NWs Hexagonal shape NWs Elastic modulus (GPa) found in literatures
4 74.46 47.59
8 89.49 60.35
81.052, 82.853, 85.970, 88.871
12 92.44 67.12
16 94.75 67.92

Table 1.  Calculated elastic modulus of [111]-oriented zinc-blende ZnSe NWs at 300 K.

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 5


Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023 Vol.:(0123456789)
www.nature.com/scientificreports/

energy per atom at zero strain is shown in Fig. 3c,d, which explains why this variation occurs. For similar thick-
ness or diameter, the potential energy per atom at zero strain shows greater negativity for square-shaped NWs
compared to hexagonal shapes; verify our calculations. Moreover, since the elastic modulus and fracture stress
both rise with increasing thickness at a certain temperature, it is evident that the thickness of the ZnSe NW
has a substantial impact on its tensile mechanical characteristics. The de-cohesion effect of the surface atoms,
which is caused by the large surface area to volume ratio of NWs, is the cause of their size-dependent mechanical
­characteristics68,72,73. Furthermore, mechanical behavior variation is caused by NWs lattice defects, which are
primarily caused by surface atoms. The surface atoms will move inward to the core as the surrounding coordinates
decrease, resulting in lattice ­defects16. However, NWs with greater thickness are less susceptible to surface defects
compared to NWs with smaller thickness, and hence, when the thickness increases, mechanical characteristics
such as the elastic modulus and fracture stress of ZnSe NW show an increasing nature.
Temperature is a powerful external parameter that can significantly affect the NW’s mechanical behavior. The
brittle or ductile behavior of semiconductor NWs can be significantly affected by the temperature and loading
­conditions5,16,67–69,74. At high temperatures, the thermal energy can cause the bonds between atoms to weaken,
making the material more prone to plastic deformation and resulting in a more ductile response. On the other
hand, low temperatures can make the material more brittle, with a lower amount of plastic deformation before

F
fracture. Here, we conducted several MD simulations for numerous temperatures to explore the temperature
dependence of the mechanical behavior and deformation mechanism of zinc-blende ZnSe NWs. Figure 4a–c

O
depicts the stress–strain plots of zinc-blende ZnSe NWs for various crystal orientations at temperatures rang-
ing from 100 to 600 K. As the figures demonstrate, when the temperature is increased from 100 to 600 K, the
fracture stress and the fracture strain decrease sharply. Furthermore, the peak in the stress versus strain curve

O
becomes smooth rather than abrupt, and the stress variation in the plastic distortion part turns is minimal.
Therefore, zinc-blende ZnSe NW exhibits conventional brittle behavior and no temperature for the transition
from brittle to ductile is identified. The stress–strain curves at various temperatures are used to calculate the

PR
ultimate fracture strength and elastic modulus (shown in Fig. 5a,b). At a temperature of 100 K, the ZnSe NW
has a maximum fracture stress of ~ 10.46 GPa, ~ 12.82 GPa, and ~ 14.79 GPa, along the [100], [110], and [111]
directions, respectively. The fracture stress drops to ~ 5.65 GPa, ~ 7.56 GPa, and ~ 9.29 GPa along the [100], [110],
and [111] directions, respectively, as soon as the temperature rises to 600 K. The fracture stress of the zinc-blende
ZnSe NW decreases by 45.98%, 41.02%, and 37.24% along the three directions when the temperature is increased
D
TE
EC
RR
CO
N
U

Figure 4.  Stress–strain response of (a) [100], (b) [110], and (c) [111]-oriented 15.76 ­nm2 zinc-blende ZnSe
NWs at different temperatures.

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 6

Vol:.(1234567890) Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023
www.nature.com/scientificreports/

F
O
O
PR
D
TE
EC

Figure 5.  (a) Fracture stress and (b) Elastic modulus of 15.76 ­nm2 zinc-blende ZnSe NWs for different crystal
loading directions with different temperatures.
RR

to 600 K. Like fracture stress, the estimated elastic modulus also shows a decreasing trend with increasing tem-
perature. As shown in Fig. 4b, when the temperature increases from 100 to 600 K, the elastic modulus decreases
from ~ 61.27 to ~ 42.99 GPa, ~ 81.67 to ~ 63.56 GPa, and ~ 84.69 to ~ 67.37 GPa along the [100], [110], and [111]
directions, respectively. The elastic modulus of the zinc-blende ZnSe NW decreases by 29.84%, 22.16%, and
CO

20.45% along the three directions when the temperature is raised to 600 K. Because of the different effects, such
as thermal expansion, greater atomic mobility, and fast diffusion of free volume at high temperatures, the elastic
modulus decreases as temperature rises. Moreover, Fig. 5a shows that the rate of decrease in fracture stress with
temperature is most significant for the [100] orientation and lowest for the [111] orientation. Similarly, for the
three crystal orientations, the rate of reduction in elastic modulus with temperature is also greatest for the [100]
N

orientation and lowest for the [111] orientation. In a later section, we will look at the deformation mechanisms of
different crystal-oriented NWs that cause this reducing trend variation. The potential energy is a valid measure
to quantify the reducing trend of mechanical strength for different orientations, which we have already shown
U

and explained in Fig. 2b. Furthermore, the total energy per atom prior to strain can be used to determine why
mechanical strength decreases with increasing temperature. Hence, the total energy per atom curves of [111],
[110], and [100]-oriented zinc-blende ZnSe NWs for six different temperatures before tensile deformation are
presented in Fig. S2a–c (Supporting Information). Before applying tensile force, 40,000 MD steps are used for
equilibration. However, the figure shows only the first 20,000 MD steps for all considered temperatures. As can
be seen from Fig. S2, zinc-blende ZnSe NWs for all different crystal orientations show stable behavior after 1000
MD steps at all temperatures. Furthermore, as the kinetic energy of atoms increases with temperature, so does
our calculated total energy per atom curve, with a higher value found for 600 K for all orientations, indicating
that the structures are less stable than for zinc-blende ZnSe NWs at 100 K. Hence, lower strain energy is sufficient
to deform the NW when the temperature increases.
The radial distribution function (RDF), g(r), shows how the density of neighboring atoms around a particular
atom varies as a function of distance. The distribution around that atom is computed by the distance between the
specific target atom and neighboring a­ toms17,75. The temperature has a significant effect on the g(r) function. Due
to the increase in temperature, there is substantial lattice vibrations in the atomic arrangements, and the separa-
tion distances of various pairs formed in the systems also show changing behavior. We have estimated the RDF

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 7


Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023 Vol.:(0123456789)
www.nature.com/scientificreports/

at three different temperatures to qualitatively elucidate the temperature-dependent mechanical performance.


Figure S3a–c (Supporting Information) depicts the temperature-dependent RDFs of Zn–Zn, Zn–Se, and Se–Se
pairs of the ZnSe system. The g(r) peaks are extremely tall and thin at a low temperature of 100 K, indicating a
solid crystal structure that is well-ordered and compact. However, the particles become more energetic when the
temperature rises and oscillate from their equilibrium position. As a result, the height of the g(r) peaks drops. At
the same time, the breadth expands, indicating that the number of ordered crystal structures reduces while the
number of disordered structures increases as temperature increases. As a result, when the structure is at a higher
temperature, a smaller amount of uniaxial force is required to break it than when it is at a lower temperature.
This form of RDF function with increasing temperature thus denotes the cause of mechanical strength loss at
high temperatures, which is consistent with previous ­research17.
Deformation mechanisms at different strain values can be used for quantitative and qualitative explanations
of mechanical ­behavior76–79. Firstly, to explore the effects of crystal orientations, the stress propagation and defor-
mation pattern of ZnSe NWs at room temperature for three different orientations ([111], [110], and [100]) are
shown in Figs. 6, 7, and 8, respectively. According to the findings, the crack initiation of [100]-oriented ZnSe NWs
(shown in Fig. 8) starts at a higher strain value compared to that of [111] and [110]-oriented ZnSe NWs, which
proves our finding of Fig. 2a. Now the question arises: although the developed stress along the [100] orientation

F
is lower compared to the [111] and [110] orientations, why does this orientation show a higher strain during its
deformation? We have already explained that the surface energy of [111]-oriented ZnSe NWs is lower compared

O
to [110] and [100] orientations, and hence [111] exhibits better fracture stress as well as elastic modulus. Never-
theless, the atomic distance between the atoms along the [111] direction (shown in Fig. 1c) is large compared to
the [100] direction, and hence, a lesser extent of uniaxial tensile strain compared to the [100] direction facilitates

O
the crack initiation. Finally, the structure deforms in a very short time. Furthermore, it has been reported that
at room temperature or lower temperatures, due to its fully activated nature over other planes, the {111} plane
can dominate the cleavage nucleation when a force is applied to its direction. Similar results for the mechanical

PR
properties of [111]-directed NWs were also reported for CdSe, CdTe, and InP N ­ Ws16,50,62.
Using a similar reasoning as above, the surface energy of [100]-oriented ZnSe NWs is very high compared to
the [111] and [110] directions, and hence [100] shows lower fracture stress as well as elastic modulus. Here, as
the atomic distance between the atoms along the [100] direction is smaller than the other orientations (shown
in Fig. 1a), the nucleation of a crack along the {100} plane would demand a considerable amount of energy.
D
Moreover, as the bonding shows strong rigidity, it needs a more significant tensile strain to deform the structure.
As shown in Fig. 8, it can be noticed that at room temperature, [100]-directed NWs need a large strain value of
TE
EC
RR
CO
N
U

Figure 6.  Tensile deformation profiles of [111] directed 15.76 ­nm2 zinc-blende ZnSe NWs for various strain
levels at a temperature of 300 K. Dislocation slipping induced cleavage plane along an angle of 45° with the
applied tension is shown in the zoomed-in view.

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 8

Vol:.(1234567890) Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023
www.nature.com/scientificreports/

F
O
O
PR
D
TE
Figure 7.  Tensile deformation profiles of [110] directed 15.76 ­nm2 zinc-blende ZnSe NWs for various strain
levels at 300 K. Dislocation slipping induced cleavage plane along an angle of 45° with the applied tension is
EC

shown in the zoomed-in view.

21.875% to initiate the structure’s deformation compared to other orientations owing to the dominance of the
smaller interatomic distance, producing an enormous attractive electrostatic force amid the p ­ lanes50. At room
RR

temperature along the [100] orientation, due to the dominance of its smaller interplanar distance, the {100}
plane was not fully activated, and hence the {111} cleavage plane activated with the applied tension and pro-
duced the result of lower fracture stress with higher strain. Similar results along [100]-oriented NWs mechanical
behavior can also be found in the literature for CdTe and SiGe ZB NWs mechanical p ­ roperties62,67. Moreover,
CO

for zinc-blende NWs, the principal source of native deformations that create NW collapse is the nucleation and
transmission of dislocations through partial slip, full slip, or twinning. As the energy barrier for the nucleation
of dislocation is lower at edges than on faces or in the bulk, the disruption originates from one of the edges of
the NW, circulates through the cross-section, and finally forms one of the local deformations. For [110] and
[111]-oriented NWs, it has been noticed that dislocation slip is initiated at the left or right edges of the structures.
N

One type of cleavage plane creates a deformation angle of 45° with the direction of the tensile force.
As with the increase in temperature, all the different orientations show a reducing trend in mechanical
behavior, and especially the [100]-oriented ZnSe NW shows a greater reduction. Here, we have only considered
U

the stress propagation and deformation pattern of [100]-oriented ZnSe at 600 K to understand the effects of
increasing temperatures. The deformation profiles of [100]-oriented ZnSe NWs at 600 K are depicted in Fig. 9.
The fracturing process of [100]-oriented ZnSe NW began with the initiation of a crack at 21.50% strain and ended
with the ultimate failure at 21.875% strain at 300 K temperature. When the structure is at 600 K, the crack initi-
ates with a strain value of 18.50%. It happens because the zinc-blende ZnSe crystal’s atomic linkages experience
more thermal fluctuations at higher temperatures, which eventually cause the chemical bonds to w ­ eaken16,50.
Moreover, as the influence of thermal vibrations is very high at higher temperatures, a minimal strain is sufficient
to disrupt the bond and create the void, destroying the pristine N­ Ws16,50. After fracture initiation, the crack begins
to branch, and the major atomic displacement occurs in the loading direction. The initial crack generation and
eventual collapse happens at nearly the same strain level, suggesting a brittle failure manner. At 300 K tempera-
ture we noticed that at 21.625% strain, fractures began to form at a single location of the NWs, (as illustrated by
red blocks in Fig. 8), and bonds began to break at around 21.875% strain. This happened because the structure
only experiences a modest thermal vibration at low temperatures, which prevents the distortion and stress from
rapidly spreading from the initial bond rupture in the NW to the entire arrangement and the lack of activation
in the {100} plane. However, when the temperature is considerably high and increased to 600 K, the magnitude
and the rate of atomic vibration increases, and atoms can travel more effortlessly from their equilibrium location,

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 9


Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023 Vol.:(0123456789)
www.nature.com/scientificreports/

F
O
O
PR
D
TE

Figure 8.  Tensile deformation profiles of [100] directed 15.76 ­nm2 zinc-blende ZnSe NWs for various strain
levels at 300 K. The zoomed-in view shows the dislocation slipping induced {111} cleavage plane along an angle
EC

of 45° with the applied tension.

which increases the mean interatomic distance and activates the {100} plane strongly. As a result, at a greater
temperature, the comparative importance of interplanar distance reduces. Cracks initiate at two different places
RR

(denoted by red square-shaped dotted blocks) due to the combined effects of the {111} plane (making an angle of
45° with the direction of tension) and the {100} plane (making an angle of 90° with the direction of tension)16,50.
Because of the activation of two types of planes as the temperature increases, the fracture stress and elastic
modulus of zinc-blende ZnSe NWs rapidly decreases, and the NWs collapse at very low strain levels.
CO

It is worth noting that MD simulations are computationally intensive, and simulations at lower strain rates
need very long computational t­ imes50,80. Therefore, in recent times, most of the tensile tests using MD simulation
have been done at strain rates of 1­ 08 to ­1010 ­s−1, which are very popular, computationally efficient, and can predict
the experimental description of the considered s­ ystems2,5,16,67. Besides, high strain rates in MD simulations allow
to study the mechanical behavior of materials under extreme loading conditions that may be difficult or impos-
N

sible to obtain ­experimentally80–82. It has been reported that due to the variation of strain rates from low to high,
there is a change mainly in the plastic phases of the stress–strain curves, which is responsible for the variation of
ultimate tensile strength and toughness. In the plastic phase of the NWs, different failure mechanisms, includ-
U

ing partial slipping, dislocation slipping, twinning, reorientation, necking, amorphization, etc., are thus induced
with varying strain ­rates83. Moreover, the difference in strain rates between the simulations and experiments
can indeed affect the observed brittle or ductile behavior of the semiconductor NWs. In general, smaller strain
rates can lead to more gradual and plastic deformation, which may result in a more ductile response character-
ized by a larger amount of plastic deformation before failure. However, the smaller strain rate-based brittle or
ductile behavior of semiconductor NWs can be influenced by the size and shape of the NWs, crystal structure
and orientation, the presence of defects and impurities, and the temperature and loading conditions.
In this study, we have considered strain rates from ­108 ­s−1 to ­1010 ­s−1 at 300 K to explore their consequences
on the mechanical behavior of ZnSe NWs for three crystal orientations. Figure 10 shows the effects of numerous
strain rates varying from 1­ 08 ­s−1 to 1­ 010 ­s−1 on the stress–strain performance of [111], [110], and [100]-oriented
zinc-blende ZnSe with a cross-sectional area of ~ 15.76 ­nm2. Before reaching the maximum stress point, the
material’s stress–strain performance is independent of strain rates, meaning that strain rates do not influence
elastic modulus. However, for zinc-blende ZnSe NWs, when the strain rate declines, both the fracture stress and
fracture strain drop. This drop results from temperature fluctuation and stress relaxation being favored at a slower
strain rate because the atoms have more time to respond. As a result, the bond breaks down earlier because the
atoms can cross the energy barrier at a lower tension. On the other hand, insufficient time to relax at a higher

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 10

Vol:.(1234567890) Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023
www.nature.com/scientificreports/

F
O
O
PR
D
TE
Figure 9.  Tensile deformation profiles of [100] directed 15.76 ­nm2 zinc-blende ZnSe NW for various strain
levels at 600 K. The zoomed-in view shows the dislocation slipping induced {100} cleavage plane along an angle
EC

of 90° with the applied tension.

strain rate creates rapid fluctuations in atoms, which cause cracks to start to nucleate from different regions of the
NW simultaneously and propagate instantly. Therefore, slight distortions in the stress–strain plot corresponding
RR

to a strain rate of 1­ 010 ­s−1 are found. Moreover, when the strain rate increases, both rupture strength and strain
rise, which is consistent with prior observations of comparable zinc-blende N ­ Ws16,50,84. Table 2 lists the failure
stress and elastic modulus parameters obtained from Fig. 10. Because the slopes of the stress–strain curves in
the elastic area during deformation coincide for different strain rates, the elastic modulus is determined to be
CO

strain rate independent. The relationship between the ultimate fracture strength of zinc-blende ZnSe NW along
a particular crystal orientation and the strain rate can be related by the Arrhenius equation ­of77:
 
1 Q
ε̇ = Aσ m exp − (2)
RT
N

where ε̇ signifies the strain rate, σ signifies the fracture strength, Q signifies the activation energy, R is the uni-
versal gas constant, T is the deformation temperature, m is the strain-rate sensitivity, and A is a constant. By
U

using natural logarithms on both sides and assuming that the temperature remains constant throughout the
deformation, Eq. (2) can be further simplified as:
1 Q
ln ε̇ = ln (A) + ln (σ ) − (3)
m RT
Using the slopes of ln(σ) and ln(ε̇), the strain-rate sensitivity m along the different crystal orientations can be
obtained using the following relationship:
∂ ln(σ )
m= (4)
∂ ln(ε̇)
Here, in this work, the strain-rate sensitivity m for [111], [110], and [100] orientations of zinc-blende ZnSe NW
was found to be 0.0176, 0.029, and 0.0209, respectively; as shown in Fig. 11. Hence, the strain rate has a crucial
effect on the [110]-oriented zinc-blende ZnSe NW. This high sensitivity along the [110] orientation may have
occurred due to its deformation mechanisms.

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 11


Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023 Vol.:(0123456789)
www.nature.com/scientificreports/

F
O
O
PR
D
TE
EC

Figure 10.  Stress–strain relationship of (a) [111], (b) [110], and (c) [100]-oriented ZnSe NWs for different
strain rates at 300 K.
RR

Fracture strength
(GPa)
Strain rate ­(s−1) [100] [110] [111]
CO

1 × ­108 7.25 8.49 11.87


1 × ­109 7.75 9.55 12.24
5 × ­109 7.99 9.65 12.66
1 × ­1010 7.94 9.73 12.89
Elastic modulus (GPa)
N

Strain rate ­(s−1) [100] [110] [111]


1 × ­108 50.20 72.01 74.20
U

1 × ­109 50.08 72.03 74.46


5 × ­109 50.30 72.20 74.52
1 × ­1010 50.40 72.37 74.64

Table 2.  Calculated room temperature fracture stress and elastic modulus of three different crystal-oriented
zinc-blende ZnSe NWs at different strain rates.

To explain the strain rate sensitivity along different crystal orientations, we first looked at the potential energy
curves for three different crystal orientations in this section. Figure S4a–c (Supporting Information) show the
potential energy per atom curves for [100], [110], and [111]-directed zinc-blende ZnSe NWs as a function of
strain for different strain rates. As seen in the figures, for all three different crystal orientations, the potential
energy curve for different strain rates shows a similar nature before reaching the final rupturing stage. From
Fig. S4b, it can be noticed that for the [110]-crystal oriented ZnSe NW with different strain rates, the fracture

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 12

Vol:.(1234567890) Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023
www.nature.com/scientificreports/

F
O
Figure 11.  The strain-rate sensitivity of zinc-blende ZnSe NWs along three crystal orientations at 300 K.

O
PR
strain at which the potential energy curve shows a sharp decrease deviates considerably compared to the other
two directions. Hence, this [110]-oriented NW demonstrates the most significant strain rate sensitivity. In addi-
tion, we have also explored the deformation mechanisms of the three different crystal-oriented ZnSe NWs for
three different strain rates for a qualitative and quantitative explanation of strain rate sensitivity. Figure 12a–c
depict the final stage deformation mechanisms of [100], [110], and [111]-oriented zinc-blende ZnSe NWs at
D
strain rates of 1­ 08 ­s−1, ­109 ­s−1, and 1­ 010 ­s−1, respectively. These figures show the variation in deformation behaviors
for different strain rates. For simplicity, we have just elucidated the [110]-oriented NWs deformation. Figure 12b
shows two dislocations slipping at two different locations of the NW when the strain rate was ­108 ­s−1 for the [110]
TE
orientation. When the strain rate was increased to 1­ 09 ­s−1, however, only one cleavage plane was observed in the
NW, which is in stark contrast to the findings along the other two orientations shown in Fig. 12a,c. Further, a
combination of different cleavage planes is noticed when the strain rate increases to its highest value of 1­ 010 ­s−1.
They are homogeneously nucleated using dislocation slipping and deformation twinning in several places of the
EC

NW with a cascading connection. As a result, when the strain rate changes, [110]-oriented NWs exhibit highly
sensitive behavior when compared to other NWs oriented along [100] and [111] orientations.

Conclusions
RR

In conclusion, the tensile mechanical behavior and deformation mechanism of zinc-blende ZnSe NWs have
been extensively studied at the atomistic level using MD simulations. The fracture stress, fracture strain, and
elastic modulus of zinc-blende ZnSe NWs has been investigated for different crystal orientations, temperatures,
and strain rates. Although zinc-blende ZnSe NW exhibits conventional brittle behavior, no temperature for the
transition from brittle to ductile is identified. Square-shaped ZnSe NWs show greater value in terms of fracture
CO

strength and elastic modulus compared to hexagonal shapes at all considered diameters. We also found that the
elastic modulus and ultimate tensile strength of the NWs are significantly negatively correlated with tempera-
ture. In terms of ultimate strength, elastic modulus, and fracture toughness, the [111] loading direction offers
the highest values at all temperatures, whereas the [100] loading direction yields the lowest values. The NWs
show the highest fracture strain under [100] loading directions. It is noticed for [100] orientation that the {111}
N

planes are the deformation planes at smaller temperatures; conversely, when the temperature is increased to
a higher value, the {100} plane is activated and contributes as the second principal cleavage plane. The elastic
moduli of different crystal orientations at different strain rates show nearly constant values. Finally, it has been
U

discovered that the [110]-direction exhibits a greater strain rate sensitivity than the other two orientations. When
the strain rate increases to its highest value of ­1010 ­s−1, a combination of different cleavage planes is noticed.
They are homogeneously nucleated through dislocation slipping and deformation twinning in several places
of the NW with a cascading connection. This work offers a comprehensive understanding of the mechanical
properties and fracture mechanisms of zinc-blende ZnSe NWs, which are useful to effectively design future
nanoelectromechanical systems.

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 13


Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023 Vol.:(0123456789)
www.nature.com/scientificreports/

F
O
O
PR
D
TE
EC

Figure 12.  Variation in the deformation mechanisms of zinc-blende ZnSe NWs for three different tensile strain
rates along the (a) [111], (b) [110], and (c) [100]-orientation.
RR

Data availability
The datasets used and/or analyzed during the current study available from the corresponding author on reason-
CO

able request.

Received: 11 January 2023; Accepted: 27 February 2023


N

References
1. Firestein, K. L. et al. Crystallography-derived Young’s modulus and tensile strength of AlN nanowires as revealed by in situ trans-
mission electron microscopy. Nano Lett. 19, 2084–2091 (2019).
U

2. Bernal, R. A. et al. Effect of growth orientation and diameter on the elasticity of GaN nanowires. A combined in situ TEM and
atomistic modeling investigation. Nano Lett. 11, 548–555 (2011).
3. Mandal, T., Maiti, P. K. & Dasgupta, C. Mechanical properties of ZnS nanowires and thin films: Microscopic origin of the depend-
ence on size and growth direction. Phys. Rev. B Condens. Matter Mater. Phys. 86, 024101 (2012).
4. Wang, Z., Zu, X., Yang, L., Gao, F. & Weber, W. J. Atomistic simulations of the size, orientation, and temperature dependence of
tensile behavior in GaN nanowires. Phys. Rev. B Condens. Matter Mater. Phys. 76, 045310 (2007).
5. Wang, Z., Zu, X., Gao, F. & Weber, W. J. Atomistic simulations of the mechanical properties of silicon carbide nanowires. Phys.
Rev. B Condens. Matter Mater. Phys. 77, 224113 (2008).
6. Tao, Y. & Degen, C. L. Growth of magnetic nanowires along freely selectable hkl  crystal directions. Nat. Commun. 9, 1–7 (2018).
7. Zhang, Q., Li, Q. K. & Li, M. Key factors affecting mechanical behavior of metallic glass nanowires. Sci. Rep. 7, 1–8 (2017).
8. Kuykendall, T. et al. Crystallographic alignment of high-density gallium nitride nanowire arrays. Nat. Mater. 3, 524–528 (2004).
9. Li, W. et al. Performance modulation of α-MnO2 nanowires by crystal facet engineering. Sci. Rep. 5, 1–8 (2015).
10. Boubenia, S. et al. A facile hydrothermal approach for the density tunable growth of ZnO nanowires and their electrical charac-
terizations. Sci. Rep. 7, 1–10 (2017).
11. Jeong, B., Kim, J., Lee, T., Kim, S. W. & Ryu, S. Systematic investigation of the deformation mechanisms of a γ-TiAl single crystal.
Sci. Rep. 8, 1–12 (2018).
12. Ji, C. & Park, H. S. The coupled effects of geometry and surface orientation on the mechanical properties ofmetal nanowires.
Nanotechnology 18, 305704 (2007).

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 14

Vol:.(1234567890) Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023
www.nature.com/scientificreports/

13. Zhuo, X. R. & Beom, H. G. Effect of side surface orientation on the mechanical properties of silicon nanowires: A molecular
dynamics study. Crystals 9, 102 (2019).
14. Liu, Q. & Shen, S. On the large-strain plasticity of silicon nanowires: Effects of axial orientation and surface. Int. J. Plast. 38, 146–158
(2012).
15. Yang, Z., Zheng, L., Yue, Y. & Lu, Z. Effects of twin orientation and spacing on the mechanical properties of Cu nanowires. Sci.
Rep. https://​doi.​org/​10.​1038/​s41598-​017-​10934-6 (2017).
16. Chowdhury, E. H., Rahman, M. H., Jayan, R. & Islam, M. M. Atomistic investigation on the mechanical properties and failure
behavior of zinc-blende cadmium selenide (CdSe) nanowire. Comput. Mater. Sci. 186, 110001 (2021).
17. Dong, C. et al. Effects of crystallographic orientation, temperature and void on tensile mechanical properties of Ni–Co single
crystal nanopillars. J. Alloys Compd. 870, 159476 (2021).
18. Cao, H., Chen, W., Rui, Z. & Yan, C. Effects of cross-sectional area and aspect ratio coupled with orientation on mechanical proper-
ties and deformation behavior of Cu nanowires. Nanotechnology 33, 365702 (2022).
19. Chang, L., Zhou, C. Y., Liu, H. X., Li, J. & He, X. H. Orientation and strain rate dependent tensile behavior of single crystal titanium
nanowires by molecular dynamics simulations. J. Mater. Sci. Technol. 34, 864–877 (2018).
20. Wang, D. et al. Where, and how, does a nanowire break?. Nano Lett. 7, 1208–1212 (2007).
21. Ho, D. T., Im, Y., Kwon, S., Earmme, Y. Y. & Kim, S. Y. Mechanical failure mode of metal nanowires: Global deformation versus
local deformation. Nat. Publ. Gr. https://​doi.​org/​10.​1038/​srep1​1050 (2015).
22. Jie, J., Zhang, W., Bello, I., Lee, C. S. & Lee, S. T. One-dimensional II–VI nanostructures: Synthesis, properties and optoelectronic
applications. Nano Today 5, 313–336 (2010).

F
23. Zhang, X. et al. ZnSe nanowire/Si p–n heterojunctions: Device construction and optoelectronic applications. Nanotechnology 24,
395201 (2013).

O
24. Wang, J., Qiao, Y., Wang, T. & Chen, K. Catalyst/surfactant co-assisted colloidal synthesis and optical properties of ultrathin/
ultralong ZnSe nanowires. J. Cryst. Growth 509, 54–59 (2019).
25. Yuan Cai, B. et al. The size-dependent growth direction of ZnSe nanowires. Adv. Mater. 18, 109–114 (2006).
26. Philipose, U. et al. Structure and photoluminescence of ZnSe nanostructures fabricated by vapor phase growth. J. Appl. Phys. 101,

O
014326 (2007).
27. Tekcan, B. et al. Semiconductor nanowire metamaterial for broadband near-unity absorption. Sci. Rep. 12, 1–8 (2022).
28. Roy, A., Mead, J., Wang, S. & Huang, H. Effects of surface defects on the mechanical properties of ZnO nanowires. Sci. Rep. 7, 1–8

PR
(2017).
29. Zamil, M. Y., Islam, M. S., Stampfl, C. & Park, J. Tribo-piezoelectricity in group III nitride bilayers: A density functional theory
investigation. ACS Appl. Mater. Interfaces 14, 20856–20865 (2022).
30. Tao, R., Ardila, G., Montès, L. & Mouis, M. Modeling of semiconducting piezoelectric nanowires for mechanical energy harvesting
and mechanical sensing. Nano Energy 14, 62–76 (2015).
31. Sivan, A. K. et al. Plasmon-induced resonant effects on the optical properties of Ag-decorated ZnSe nanowires. Nanotechnology
31, 174001 (2020). D
32. Hou, L. et al. Bound magnetic polaron in Zn-rich cobalt-doped ZnSe nanowires. Nanotechnology 29, 055707 (2018).
33. Wisniewski, D., Byrne, K., De Souza, C. F., Fernandes, C. & Ruda, H. E. Enhancement of transport properties in single ZnSe
TE
nanowire field-effect transistors. Nanotechnology 30, 054007 (2018).
34. Zhang, X., Wu, D. & Geng, H. Heterojunctions based on II-VI compound semiconductor one-dimensional nanostructures and
their optoelectronic applications. Crystals 7, 307 (2017).
35. Zhang, Q., Li, H., Ma, Y. & Zhai, T. ZnSe nanostructures: Synthesis, properties and applications. Prog. Mater. Sci. 83, 472–535
(2016).
EC

36. Li, Z., Sun, Q., Yao, X. D., Zhu, Z. H. & Lu, G. Q. Semiconductor nanowires for thermoelectrics. J. Mater. Chem. 22, 22821–22831
(2012).
37. Long, Y. Z., Yu, M., Sun, B., Gu, C. Z. & Fan, Z. Recent advances in large-scale assembly of semiconducting inorganic nanowires
and nanofibers for electronics, sensors and photovoltaics. Chem. Soc. Rev. 41, 4560–4580 (2012).
38. Okoye, C. M. I. First-principles study of the electronic and optical properties of zincblende zinc selenide. Phys. B Condens. Matter
RR

337, 1–9 (2003).


39. Kremer, R. K., Cardona, M., Lauck, R., Siegle, G. & Romero, A. H. Vibrational and thermal properties of ZnX (X=Se, Te): Density
functional theory (LDA and GGA) versus experiment. Phys. Rev. B Condens. Matter Mater. Phys. 85, 035208 (2012).
40. Chin, P. T. K., Stouwdam, J. W. & Janssen, R. A. J. Highly luminescent ultranarrow Mn doped ZnSe nanowires. Nano Lett. 9,
745–750 (2009).
41. Valdez, L. A., Caravaca, M. A. & Casali, R. A. Ab-initio study of elastic anisotropy, hardness and volumetric thermal expansion
CO

coefficient of ZnO, ZnS, ZnSe in wurtzite and zinc blende phases. J. Phys. Chem. Solids 134, 245–254 (2019).
42. Wang, Y. G., Zeng, Y. P., Qu, B. H. & Zhang, Q. L. Enhancement of current carrying capacity of the strained ZnSe nanowire. J.
Appl. Phys. 109, 104311 (2011).
43. Zhang, X., Meng, D., Tang, Z., Hu, D. & Ma, D. Surface-dominated negative photoresponse of phosphorus-doped ZnSe nanowires
and their detecting performance. J. Mater. Sci. Mater. Electron. 27, 11463–11469 (2016).
44. Zhang, X. et al. Surface induced negative photoconductivity in p-type ZnSe:Bi nanowires and their nano-optoelectronic applica-
N

tions. J. Mater. Chem. 21, 6736–6741 (2011).


45. Wang, Y. G. et al. Current-voltage characteristics of in situ graphitization of hydrocarbon coated on ZnSe nanowire. J. Phys. Chem.
C 114, 12839–12849 (2010).
U

46. Tian, L. et al. Ultrafast carrier dynamics, band-gap renormalization, and optical properties of ZnSe nanowires. Phys. Rev. B 94,
165442 (2016).
47. Zhou, X. W. et al. Stillinger-Weber potential for the II-VI elements Zn-Cd-Hg-S-Se-Te. Phys. Rev. B Condens. Matter Mater. Phys.
88, 085309 (2013).
48. Plimpton, S. Fast parallel algorithms for short-range molecular dynamics. J. Comput. Phys. 117, 1–19 (1995).
49. Hirel, P. Atomsk: A tool for manipulating and converting atomic data files. Comput. Phys. Commun. 197, 212–219 (2015).
50. Pial, T. H., Rakib, T., Mojumder, S., Motalab, M. & Akanda, M. A. S. Atomistic investigations on the mechanical properties and
fracture mechanisms of indium phosphide nanowires. Phys. Chem. Chem. Phys. 20, 8647–8657 (2018).
51. Karzel, H. et al. Lattice dynamics and hyperfine interactions in ZnO and ZnSe at high external pressures. Phys. Rev. B 53, 11425
(1996).
52. Khenata, R. et al. Elastic, electronic and optical properties of ZnS, ZnSe and ZnTe under pressure. Comput. Mater. Sci. 38, 29–38
(2006).
53. Wang, H. Y., Cao, J., Huang, X. Y. & Huang, J. M. Pressure dependence of elastic and dynamical properties of zinc-blende ZnS and
ZnSe from first principle calculation. Condens. Matter Phys. 15 (2012).
54. Stukowski, A. Visualization and analysis of atomistic simulation data with OVITO—The Open Visualization Tool. Model. Simul.
Mater. Sci. Eng. 18, 015012 (2010).
55. Tsai, D. H. The virial theorem and stress calculation in molecular dynamics. J. Chem. Phys. 70, 1375–1382 (1979).
56. Chan, S. K., Cai, Y., Wang, N. & Sou, I. K. Control of growth orientation for epitaxially grown ZnSe nanowires. Appl. Phys. Lett.
88, 013108 (2006).

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 15


Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023 Vol.:(0123456789)
www.nature.com/scientificreports/

57. Yan, X. et al. Non-111 -oriented semiconductor nanowires: Growth, properties, and applications. Nanoscale https://​doi.​org/​10.​
1039/​D2NR0​6421A (2023).
58. Zhang, K. et al. Selective growth of stacking fault free 100  nanowires on a polycrystalline substrate for energy conversion applica-
tion. ACS Appl. Mater. Interfaces 12, 17676–17685 (2020).
59. Fonseka, H. A. et al. Nanowires grown on InP (100): Growth directions, facets, crystal structures, and relative yield control. ACS
Nano 8, 6945–6954 (2014).
60. Ohno, Y., Shirahama, T., Takeda, S., Ishizumi, A. & Kanemitsu, Y. Fe-catalytic growth of ZnSe nanowires on a ZnSe(001) surface
at low temperatures by molecular-beam epitaxy. Appl. Phys. Lett. 87, 043105 (2005).
61. Li, X., Ni, J. & Zhang, R. A thermodynamic model of diameter- and temperature-dependent semiconductor nanowire growth. Sci.
Rep. 7, 1–8 (2017).
62. Munshi, M. A. M., Majumder, S., Motalab, M. & Saha, S. Insights into the mechanical properties and fracture mechanism of
Cadmium Telluride nanowire. Mater. Res. Express 6, 105083 (2019).
63. Xu, W. & Dávila, L. P. Effects of crystal orientation and diameter on the mechanical properties of single-crystal M ­ gAl2O4 spinel
nanowires. Nanotechnology 30, 055701 (2018).
64. Zhang, J. M., Ma, F., Xu, K. W. & Xin, X. T. Anisotropy analysis of the surface energy of diamond cubic crystals. Surf. Interface
Anal. 35, 805–809 (2003).
65. Ma, B., Rao, Q. H. & He, Y. H. Effect of crystal orientation on tensile mechanical properties of single-crystal tungsten nanowire.
Trans. Nonferrous Met. Soc. China 24, 2904–2910 (2014).
66. Yang, Z. X. et al. Approaching the hole mobility limit of GaSb nanowires. ACS Nano 9, 9268–9275 (2015).

F
67. Rahman, M. H., Chowdhury, E. H. & Islam, M. M. Understanding mechanical properties and failure mechanism of germanium-
silicon alloy at nanoscale. J. Nanopart. Res. 22, 1–12 (2020).

O
68. Kang, K. & Cai, W. Size and temperature effects on the fracture mechanisms of silicon nanowires: Molecular dynamics simulations.
Int. J. Plast. 26, 1387–1401 (2010).
69. Kuo, J. K., Huang, P. H., Wu, W. T. & Lu, C. M. Mechanical and fracture behaviors of defective silicon nanowires: Combined effects
of vacancy clusters, temperature, wire size, and shape. Appl. Phys. A Mater. Sci. Process. 114, 1247–1256 (2014).

O
70. Ferahtia, S., Saib, S. & Bouarissa, N. Computational studies of mono-chalcogenides ZnS and ZnSe at high-pressures. Results Phys.
15, 102626 (2019).
71. Casali, R. A. & Christensen, N. E. Elastic constants and deformation potentials of ZnS and ZnSe under pressure. Solid State Com-

PR
mun. 108, 793–798 (1998).
72. Liang, H., Upmanyu, M. & Huang, H. Size-dependent elasticity of nanowires: Nonlinear effects. Phys. Rev. B Condens. Matter
Mater. Phys. 71, 241403 (2005).
73. Huang, P. H., Fang, T. H. & Chou, C. S. The coupled effects of size, shape, and location of vacancy clusters on the structural defor-
mation and mechanical strength of defective nanowires. Curr. Appl. Phys. 11, 878–887 (2011).
74. Wang, N., Cai, Y. & Zhang, R. Q. Growth of nanowires. Mater. Sci. Eng. R Rep. 60, 1–51 (2008).
D
75. Islam, A. S. M. J. et al. Vacancy-induced thermal transport and tensile mechanical behavior of monolayer honeycomb BeO. ACS
Omega 7, 4525–4537 (2022).
76. Islam, A. S. M. J. et al. Anisotropic mechanical behavior of two dimensional silicon carbide: Effect of temperature and vacancy
TE
defects. Mater. Res. Express 6, 125073 (2019).
77. Islam, A. S. M. J., Hasan, M. S., Islam, M. S. & Park, J. Chirality, temperature, and vacancy effects on mechanical behavior of
monolayer zinc-sulfide. Comput. Mater. Sci. 200, 110824 (2021).
78. Islam, A. S. M. J., Akbar, M. S., Islam, M. S. & Park, J. Temperature- and defect-induced uniaxial tensile mechanical behaviors
and the fracture mechanism of two-dimensional silicon germanide. ACS Omega https://​doi.​org/​10.​1021/​ACSOM​EGA.​1C016​91
EC

(2021).
79. Islam, A. S. M. J., Islam, M. S., Hasan, M. S., Akbar, M. S. & Park, J. Tensile mechanical behavior and the fracture mechanism in
monolayer group-III nitrides XN (X = Ga, In): Effect of temperature and point vacancies. ACS Omega 7, 14678–14689 (2022).
80. Ramachandramoorthy, R., Bernal, R. & Espinosa, H. D. Pushing the envelope of in situ transmission electron microscopy. ACS
Nano 9, 4675–4685 (2015).
RR

81. El Chlouk, Z. G., Shehadeh, M. A. & Hamade, R. F. The effect of strain rate and temperature on the mechanical behavior of Al/Fe
interface under compressive loading. Metall. Mater. Trans. A Phys. Metall. Mater. Sci. 51, 2573–2589 (2020).
82. Lee, S. et al. Reversible cyclic deformation mechanism of gold nanowires by twinning–detwinning transition evidenced from
in situ TEM. Nat. Commun. 5, 1–10 (2014).
83. Wang, D. et al. Where, and how, does a nanowire break? 1–5 (2007).
84. Tsuzuki, H., Rino, J. P. & Branicio, P. S. Dynamic behaviour of silicon carbide nanowires under high and extreme strain rates: A
CO

molecular dynamics study. J. Phys. D Appl. Phys. 44, 055405 (2011).

Author contributions
A.S.M.J.I. performed the simulations under the supervision of M.S.I. and A.G.B.; M.S.I. conceived the work;
M.S.H. performed the data curation; C.S. and J.P. reviewed and wrote the manuscript. All authors discussed,
N

analyzed the data, and contributed during the writing of the manuscript.

Competing interests
U

The authors declare no competing interests.

Additional information
Supplementary Information The online version contains supplementary material available at https://​doi.​org/​
10.​1038/​s41598-​023-​30601-3.
Correspondence and requests for materials should be addressed to M.S.I.
Reprints and permissions information is available at www.nature.com/reprints.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 16

Vol:.(1234567890) Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023
www.nature.com/scientificreports/

Open Access This article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the
Creative Commons licence, and indicate if changes were made. The images or other third party material in this
article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder. To view a copy of this licence, visit http://​creat​iveco​mmons.​org/​licen​ses/​by/4.​0/.

© The Author(s) 2023

F
O
O
PR
D
TE
EC
RR
CO
N
U

Scientific Reports | _#####################_ | https://doi.org/10.1038/s41598-023-30601-3 17


Journal : SREP 41598 Article No : 30601 Pages : 17 MS Code : 30601 Dispatch : 27-2-2023 Vol.:(0123456789)

You might also like