ArticlePDF Available

Extraordinary impact resistance of carbon nanotube film with crosslinks under micro-ballistic impact

Authors:

Abstract and Figures

The crosslinks of carbon nanotubes (CNT) film has been demonstrated to owing the ability to reinforce the quasi-static mechanical properties. But it is unclear whether crosslinks improve the ballistic impact resistance of CNT film. Here, we investigated the impact resistance of CNT film with crosslinks by combining micro-ballistic impact experiments with coarse-grained molecular dynamics (CGMD) simulations. The impact resistance is quantitatively characterized in terms of the specific penetration energy. Meanwhile, the effective enhancement of impact resistance contributed to the crosslinks is directly observed in the experiment. CGMD simulations are employed to unveil the corresponding mechanisms in terms of deformation behavior, energy dissipation mode, and failure behavior. Our results indicate that with the increase of crosslink density, the energy dissipation mode of the CNT film transforms from bending-dominated to stretching-bending-dominated due to enhanced interaction between the adjacent CNTs. This leads to a transformation of perforated morphology from cascaded interfaces sliding to crosslink-restricted deformation with crosslinks. Our simulations also indicate that the length, bending stiffness of CNTs, and film's thickness play essential roles in the impact resistance of CNT film at various crosslink densities. These results provide a feasible strategy to improve the protective performance of CNT film.
Content may be subject to copyright.
Research Article
Extraordinary impact resistance of carbon nanotube lm with
crosslinks under micro-ballistic impact
Kailu Xiao
a
,
b
,
1
, Xudong Lei
a
,
b
, Yuyu Chen
a
,
b
,QiAn
c
,
d
, Dongmei Hu
e
,
1
, Chao Wang
a
,
**
,
Xianqian Wu
a
,
*
, Chenguang Huang
a
,
b
a
Institute of Mechanics, Chinese Academy of Sciences, Beijing, 100190, China
b
School of Engineering Science, University of Chinese Academy of Sciences, Beijing, 100049, China
c
Department of Chemical and Materials Engineering, University of Nevada, Reno, NV, 89557, USA
d
Nevada Institute for Sustainability, University of Nevada, Reno, NV, 89557, USA
e
Key Laboratory of Multifunctional and Smart Systems, Suzhou Institute of Nano-Tech and Nano-Bionics, Chinese Academy of Sciences, Suzhou, 215123,
China
article info
Article history:
Received 21 October 2020
Received in revised form
23 December 2020
Accepted 2 January 2021
Available online 8 January 2021
Keywords:
CNT lms
Crosslinks
Ballistic impact
Specic penetration energy
Energy dissipation
abstract
The crosslinks of carbon nanotubes (CNT) lm has been demonstrated to owing the ability to reinforce
the quasi-static mechanical properties. But it is unclear whether crosslinks improve the ballistic impact
resistance of CNT lm. Here, we investigated the impact resistance of CNT lm with crosslinks by
combining micro-ballistic impact experiments with coarse-grained molecular dynamics (CGMD) simu-
lations. The impact resistance is quantitatively characterized in terms of the specic penetration energy.
Meanwhile, the effective enhancement of impact resistance contributed to the crosslinks is directly
observed in the experiment. CGMD simulations are employed to unveil the corresponding mechanisms
in terms of deformation behavior, energy dissipation mode, and failure behavior. Our results indicate that
with the increase of crosslink density, the energy dissipation mode of the CNT lm transforms from
bending-dominated to stretching-bending-dominated due to enhanced interaction between the adjacent
CNTs. This leads to a transformation of perforated morphology from cascaded interfaces sliding to
crosslink-restricted deformation with crosslinks. Our simulations also indicate that the length, bending
stiffness of CNTs, and lms thickness play essential roles in the impact resistance of CNT lm at various
crosslink densities. These results provide a feasible strategy to improve the protective performance of
CNT lm.
©2021 Elsevier Ltd. All rights reserved.
1. Introduction
Advanced materials with light-weight and superior impact
resistance are essential for impact engineering of bullet-proof body
armor [1,2] and debris-proof spacecraft shielding [3,4]. In recent
years, carbon nanotube (CNT) lm with nano-network structures
has attracted great attention due to its excellent multifunctional
properties such as superior exibility, tunable pore density, and
large surface-to-volume and strength-to-weight ratios [5e9]. In
addition, the excellent energy dissipation behavior of CNT lm
under impact makes it an ideal protective material.
Extensive research has been dedicated to determining the me-
chanical properties of CNT lms under quasi-static loadings, with
the emphasis generally focused on the type of CNT and the mi-
crostructures. Sakurai et al. [10] investigated the mechanical
strength of CNT lm comprised of millimeter-length single-walled
CNT (SWCNT). The results showed that the strength of the SWCNT
lm with ber lengths of 1500
m
m was more than twice (45 vs.
19 MPa) that of CNT lm with ber lengths of 350
m
m. Ma et al. [11]
observed that the Poissons ratio of CNT lm changes from negative
to positive during uniaxial tensile loadings due to the variation of
angles between CNTs and the elongation of initial-bent CNTs.
Cranford et al. [12] numerically tuned Youngs modulus of CNT lm
over a range of 0.2e3.1 GPa by manipulating the type and density of
the CNT based on the developed coarse-grained molecular dy-
namics (CGMD) model. Xie et al. [8] investigated the relationship
*Corresponding author.
** Corresponding author.
E-mail addresses: wangchao@lnm.imech.ac.cn (C. Wang), wuxianqian@imech.
ac.cn (X. Wu).
1
These authors contributed equally to this work.
Contents lists available at ScienceDirect
Carbon
journal homepage: www.elsevier.com/locate/carbon
https://doi.org/10.1016/j.carbon.2021.01.009
0008-6223/©2021 Elsevier Ltd. All rights reserved.
Carbon 175 (2021) 478e489
between microstructure and mechanical properties of CNT lm
through CGMD simulations. Their results showed that the me-
chanical properties of CNT lm could be enhanced by inhibiting the
inter-tube sliding during the loading process.
Such the mechanical properties of CNT lm as Youngs modulus
and strength do not exhibit performance superior to advanced
metals [13,14] and bers [15,16]. However, recent studies indicate
that the CNT lm exhibits an excellent capacity for energy dissi-
pation under dynamic loadings due to the disordered and tunable
nano-network structures [17], making it great potential to be the
next generation of impact protective materials. Wang et al. [18]
employed CGMD simulations to reveal that brittle fracture of the
CNTs and detachment or sliding-induced ductile failure of vdW
interfaces between the CNTs render abundant channels to dissipate
impact energy. Another CGMD study by Chen et al. [19] also showed
that the CNT lm possesses extremely high efciency to dissipate
impact energy within its deformation limits during high-velocity
plate-impact.
The aforementioned studies on the dynamic behavior of CNT
lm indicate that the interactions between CNTs in the lm
involving bending, sliding, and bundling play a signicant role in
the energy dissipation process. An intriguing question, therefore,
arises: if reinforcing the interaction by introducing crosslinks be-
tween adjacent CNTs, can the impact resistance of the CNT lm be
improved dramatically? It is well known that both physical cross-
links [20,21] and chemical crosslinks [22,23] can signicantly
improve the mechanical properties of CNT lm [17,24e26], e.g. the
increase of quasi-static tensile strength by 4e7 times. Regarding
the mechanical properties under quasi-static loadings, one could
expect a superior impact resistance for the CNT lm with crosslinks.
However, up to now, no direct experimental evidence or simulation
result has yet been reported to prove this fascinating expectation.
In this study, the effects of crosslinks on the micro-ballistic
impact resistance of CNT lm are investigated by experiments
and CGMD simulations. Firstly, micro-ballistic impact experiments
showed the improvement of impact resistance of CNT lms by
introducing crosslinks. Then a series of CGMD simulations are
conducted to obtain the specic penetration energy of the CNT
lms with crosslinks concerning the crosslink density. The superior
impact energy dissipation mechanisms of CNT lms in terms of
dynamical responses, energy dissipation modes, and failure be-
haviors are investigated by simulations. The failure characteristics
involving the change of broken morphology around the penetration
hole and the shrinkage effect of the post-mortem penetration hole
of CNT lm after adding crosslinks are compared qualitatively be-
tween experiment and simulation results. The key parameters that
affect the energy dissipation ability of the CNT lm with crosslinks
such as length, bending stiffness of CNT, and thickness of CNT lm
are also investigated in the simulations. This study provides solid
evidence that the impact resistance of CNT lm can be enhanced by
introducing crosslinks. The strategy is proposed for designing a CNT
lm with further improved protective performance.
2. Experimental method and computational model
2.1. Micro-ballistic impact experiments
To explore the impact performance of CNT lms with and
without crosslinks, a laser-induced micro-particle impact test
(LIPIT) platform, which was originally developed by Lee et al. [27]
and further improved by Hassani-Gangaraj et al. [28,29], was built
as illustrated in Fig. 1(a). In LIPIT experiments, high-pressure
plasma under the connement of a 4-mm-thick BK7 glass is
generated through the interaction between the focused pulse-laser
(1064 nm wavelength, 10 ns FWHM,
F
2 mm) and a 40-
m
m-thick
aluminum lm, resulting in the rapid swelling of a 100-
m
m-thick
polydimethylsiloxane (PDMS) lm attached closely to the
aluminum (Al) lm. During that process, a steel projectile with a
diameter of 360
m
m placed at the back surface of the PDMS lm is
accelerated and then impacts the boundary-xed CNT lms with an
average speed of ~100 m/s determined by the high-speed photog-
raphy (specialised-imaging SIMD16). The residual velocity was also
measured to obtain the impact resistance of the CNT lm. Both the
CNT lms in the pristine state (density ~0.42 g/cm
3
) and the CNT
lm with polyvinyl alcohol (PVA) as crosslinks (density ~0.53 g/
cm
3
) were tested to investigate the effects of crosslinks. Compared
with weak vdW connections between CNT bers, the crosslink is
generally dened as enhanced connections between CNT chains
[25], which can be realized by adding a small amount of PVA in
experiments that strengthens the weak vdW interfaces. The coarse-
grained molecular dynamics (CGMD) simulation aims to investigate
the enhancement effect of the crosslinks as observed in experi-
ments implemented by adding extra bonds between CNT chains. It
is to be noted that a ceiling of weight and spatial dimensions of CNT
lm exists for crosslinks constructed in this experiment due to the
saturation threshold after soaking and solvent evaporation. The
post-mortem fractured lms were characterized by Scanning
Electron Microscope (SEM) to understand the dynamic deforma-
tion and failure behavior.
2.2. Coarse-grained models for CNT lms
A CGMD model is constructed based on Refs. [30,31] to simulate
the realistic CNT systems. Each 100-nm-length CNT chain consists
of 100 beads, and a coarse-grained bead is mapped by a (5, 5)
SWCNTs with the same mass. The bonding between two adjacent
beads and the angle interaction from three successive beads in a
certain CNT chain are represented by springs to reect the
stretching and bending properties. The inter-tube interaction be-
tween the pairs of beads is depicted by the long-ranged vdW
interaction. Therefore, the deformation energy of the coarse-
grained system can be calculated by stretching energy of intrinsic
bonds E
s
, stretching energy of extra bonds of crosslinks E
c
, bending
energy E
b
, and vdW energy E
v
as given by
E
total
¼E
s
þE
c
þE
b
þE
v
;(1a)
E
s
¼X
bonds
k
s
ðrr
0
Þ
2
2;(1b)
E
c
¼X
crosslinks
k
s
ðll
0
Þ
2
2;(1c)
E
b
¼X
angles
k
4
ð44
0
Þ
2
2;(1d)
E
v
¼X
pairs
4εhð
s
=rÞ
12
ð
s
=rÞ
6
i;(1f)
where k
s
is stretching stiffness related to the current bond length r,
and the length in the equilibrium state r
0
is 10 Å. ldenotes the
added bond length, and the length in the equilibrium state l
0
is
10 Å. k
4
denotes bending stiffness related to the current angle of the
triplet 4, and the angle in the equilibrium state 4
0
is 180
.eand
s
represent the depth of the potential well and zero energy inter-tube
distance, respectively. All parameters used in CGMD simulations
are calibrated from full-atom MD simulation and depicted in
Table 1. The cutoff value applied to non-bonding pair interaction is
K. Xiao, X. Lei, Y. Chen et al. Carbon 175 (2021) 478e489
479
4 nm. The bond breaks once the tensile strain exceeds 24%, which
has been veried in the previous study [18]. Pairs of beads within
distance 0.8e1.2 nm are chosen by a detection program and then
selected randomly to generate crosslinks. Without loss of general-
ity, the inter-tube crosslinks adopted here are assumed to be the
same as the intra-tube bonds.
To simulate the experimental system as well as possible, we
constructed the coarse-grained numerical model according to the
geometry scaling relation [32] as listed in Table 2, where L,t, and D
denote the length of an individual CNT chain, the thickness of the
CNT lm and the diameter of projectile, respectively. It can be seen
that the dimensionless in-plane size
h
¼L/D and the dimensionless
thickness
c
¼t/D in our simulation are almost in the same order as
that in the experiments, implying the same deformation mecha-
nisms between the experiments and the simulations.
It should be noted that we study the enhancement effect of PVA
by adding some bonds between adjacent CNTs, and additional PVA
beads are not introduced in the numerical model. The rationality of
this treatment is based on the following experimental evidences:
rstly, the length of the PVA chain is much smaller than that of
CNTs, ~0.15
m
m vs. ~300
m
m; secondly, the volume fraction of PVA is
relatively small, ~6.8%, and the CNT lm can still be viewed as a
porous material. Therefore, it is appropriate to adopt the bond
model in our simulation to simulate the enhanced connections
between neighbor CNTs by PVA and ignore the volume effect of PVA
molecules. This crosslink strategy not only captures the enhanced
effect of crosslinks by PVA, but also effectively improves compu-
tational efciency. Moreover, this scheme has been successfully
adopted to study the effect of crosslinks on the mechanical de-
formations [8,33], viscoelastic properties [34], and fracture behav-
iors [35] of CNT lms.
2.3. Micro-ballistic impact model
All CGMD simulations are performed by adopting the large-scale
atomic/molecular massively parallel simulator (LAMMPS) [36] and
the results are visualized in the Open Visualization Tool (Ovito)
software [37]. Fig. 1(b) shows the CNT lm with crosslinks, in which
the CNT chains and crosslinks are pointed and color-coded by green
and red, respectively. The crosslinksdensity,
r
,isdened as the
average number of crosslinks per coarse-grained CNT chain [35].
The crosslinks are distributed uniformly in the CNT lm as shown in
Fig.1(c) for the system with
r
¼20. The maximum crosslink density,
which is caused by the limitation of adjacent beads distance, is
determined as
r
max
¼35 in the simulation conguration, the value
of which is consistent with the ceiling we observed experimentally.
The density of the CNT lm at the equilibrium state is 0.28 g/cm
3
,
which is comparable with the magnitude of density adopted in the
experiments.
After the CNT lms with crosslinks are prepared, a spherical
diamond projectile with a diameter of 20 nm is constructed and
added to the system as shown in Fig. 1(d). The initial distance be-
tween projectile and CNT lm is about 10 nm to avoid the inter-
action during the relaxation stage. Here, the projectile is considered
as a rigid body since the projectile is much stronger than CNT lm
and no broken of the projectile is observed in experiments. The CNT
lm is periodic-boundary conditions along with the non-impact
directions, i.e. XeY directions, and free-standing in the impact di-
rection, i.e. Z direction. About 2 nm are set as xed-boundary at the
Fig. 1. (a) Schematic of laser-induced projectile impact experimental tests. (b) Cross-linked CNT lm in simulation with
r
¼20 in the top view and the enlarged view represents
crosslinks between two nanotubes (CNTs are shown in green and crosslinks are shown in red). (c) Morphology of enlarged CNT lm in simulation as marked in (b) by blue rectangle
to illustrate the uniform distribution of crosslinks. (d) Micro-projectile impact model in simulation (CNT lm is shown in green, and the projectile is illustrated in orange). (e) Total
energy for the simulation system during the equilibrium NVT ensemble for
r
¼20. (A colour version of this gure can be viewed online.)
Table 1
Parameters of the force eld for the coarse-grained (5, 5) SWCNT model [12].
Parameters k
s
r
0
k
4
4
0
s
e
Value 1000 10 14300 180
9.35 15.1
Units kcal/mol/Å
2
Å kcal/mol/rad
2
eÅ kcal/mol
Table 2
Dimensions of experiments and CGMD simulations.
Methods LtD
h
¼L/D
c
¼t/D
Experiments 300e360
m
m10
m
m 360
m
m 1.2e1 0.03
CGMD simulations 20e100 nm 4e5nm 20nm 1e5 0.2e0.25
K. Xiao, X. Lei, Y. Chen et al. Carbon 175 (2021) 478e489
480
peripheries along X and Y direction during the impact process. The
interaction between the CNT lm and the projectile is considered as
12-6 Lennard-Jones potential described by Eq. (1d), and the pa-
rameters are the same as the inter-tube interactions. The whole
system is equilibrated at 300 K for 10 ns using the NVT ensemble
with a Langevin thermostat. The total energy uctuation is less
than 0.1% after equilibrium as given in Fig. 1(e). After that, the ca-
nonical ensemble (i.e. NVE) is adopted for the impact tests, and a
time step of 1 fs is applied to ensure the stability of the simulation.
A series of impact tests from non-perforation to perforation are
conducted with impact velocities, v
i
, ranging from 8 to 12 km/s.
3. Results and discussion
3.1. Experimental results
To examine the dynamical responses of CNT lm with crosslinks,
the LIPIT tests on CNT lm without and with crosslinks were per-
formed. The energy change of the projectile before and after
impact,
D
E
k
, equals the absorption energy of the membrane during
the impact process, E
p
. The energy loss of air drag can be ignored.
Here,
D
E
k
¼mðv
2
i
v
2
r
Þ=2 and E
p
¼ð
r
A
s
tÞv
2
i
=2þE
d
, where the rst
term of the latter equation represents the kinetic energy transfer
from the projectile to the lm within the deformation area A
s
.m
denotes the mass of the projectile, v
i
and v
r
represent the impact
velocity and residual velocity of the projectile, tis the thickness of
the lm. E
d
is the other possible energy dissipation channels like
propagation of cracks. The specic penetration energy, E
*
p
,is
dened as the normalized absorption energy of CNT lms by the
mass of lm within the impact region, and it is adopted to evaluate
the energy dissipation ability of CNT lm [27]. The specic pene-
tration energy can be written as E
*
p
¼E
p
=
r
A
s
h¼v
2
i
=2þE
*
d
, where
r
denotes the density of the lm, and E
*
d
is the delocalized pene-
tration energy served to evaluate the outward propagation ability
of stress wave and is material-dependent.
The E
p
of CNT lms with and without crosslinks are about
5.5 10
4
J and 4.1 10
4
J, respectively, under v
i
¼100 m/s,
implying that the crosslinks can effectively enhance the energy
absorption capability (i.e. 34% improvement after adding cross-
links) and improve the ballistic resistance of CNT lm. Corre-
sponding, the E
*
p
of CNT lms with and without crosslinks are
determined to be 0.81 MJ/kg and 1.0 MJ/kg, respectively, which are
much higher than other bulletproof materials such as Al (0.12 MJ/
kg) [38], steel (0.1 MJ/kg) [39], Kevlar armor (0.5 MJ/kg) [40], and
PMMA (0.12 MJ/kg) [41]. This phenomenon indicates the great
potential of CNT lm as an advanced bulletproof material. It is
worth noting that E
*
p
of CNT lm with crosslinks is slightly lower
than the pristine CNT lm due to the extra mass of the PVA adhe-
sion, i.e. 27% weight increment after adding crosslinks. Because no
extra mass is introduced, one could expect extremely higher E
p
and
E
*
p
of CNT lm with crosslinks introduced by other methods such as
e-beam irradiation [22] instead of polymer adhesions.
After the micro-ballistic impact tests, the post-mortem frac-
tured morphologies for the CNT lms with and without crosslinks
were examined as given in Fig. 2(a) and (b), respectively. The
cascaded interface sliding mode is observed for CNT lm without
crosslinks, in which abundant CNT bers on the periphery of the
penetration hole are pulled out after impact due to the weak vdW
interfaces between CNT chains, leading to a uffy microstructure as
marked in Fig. 2(a). In contrast, a much smooth prole as illustrated
in Fig. 2(b) is formed for the CNT lm with crosslinks due to the PVA
crosslinks between the adjacent CNT chains that strongly restrict
the sliding behavior of the CNT chains. The interesting trans-
formation of damage mode from the cascaded interface sliding to
the crosslink-restricted deformation after adding crosslinks implies
the mechanism change of energy dissipation, which contributes to
the improvement of the impact resistance of the CNT lm with
crosslinks. Therefore, a series of CGMD simulations need to be
conducted to thoroughly investigate the effects of crosslinks and
the corresponding microscopic energy dissipation mechanism
qualitatively.
3.2. Crosslink-density-dependent protective performance
To further understand the effect of crosslinks and elucidate this
transition of damage mode, the CGMD simulations were per-
formed. The residual velocity, v
r
, of the projectile and the specic
penetration energy, E
*
p
, of the CNT lm with respect to the crosslink
density are given in Fig. 3. As shown in Fig. 3(a), for v
i
¼12 km/s, all
the CNT lms with even the largest crosslink density
r
¼35 are
perforated as indicated by the inserted microstructure, and v
r
de-
creases almost linearly with the increase of
r
. Correspondingly, the
E
*
p
increases almost linearly from 88 MJ/kg to 102.5 MJ/kg at
v
i
¼12 km/s, namely an increase of 16.5% as we increase the
crosslink density
r
from 0 to 35. While decreasing the impact ve-
locity to v
i
¼8 km/s, only the CNT lms with the crosslink density
r
20 are perforated as given by the inserted microstructures. For
r
10, v
r
decreases nearly linearly with
r
, and the slope is
approximately same as v
i
¼12 km/s, leading to the linear increase
of E
*
p
from 34.2 MJ/kg to 37.4 MJ/kg with respect to
r
as depicted in
Fig. 3(b). The CNT lm is perforated by the projectile for
r
20. For
Fig. 2. SEM images of the broken morphologies for (a) the pristine CNT lm and (b) the CNT lm with crosslinks, respectively. (A colour version of this gure can be viewed online.)
K. Xiao, X. Lei, Y. Chen et al. Carbon 175 (2021) 478e489
481
r
>20, v
r
and E
*
p
almost remain constant with respect to
r
because
the CNT lms are not perforated. All the phenomena could be
observed in movies in Supplementary Materials. Since v
i
¼8 km/s is
almost the critical impact velocity v
c
[42] for
r
¼20, v
r
decreases
fast to almost 0 m/s for 10 <
r
20. Correspondingly, E
*
p
rapid in-
creases to 50 MJ/kg as
r
increases to 20. The corresponding defor-
mation mechanism for perforation to non-perforation is much
different as well. For the case of perforation, it is the local tensile
failure of the lm near the impact region that dissipates the impact
energy. However, for the case of non-perforation, the stretching
and bending of almost the whole CNT lm contribute to the dissi-
pation of the impact energy. This enhancement after adding
crosslinks is consistent with the experiments that the energy ab-
sorption capacity has been improved after adding PVA. The specic
penetration energy in the simulations is nearly two orders of
magnitude larger than that in the experimental results, which is
ascribed to the different sizes of CNT lms as adopted in experi-
ments and simulations. Both the high specic impact kinetic energy
v
2
i
=2 and delocalized penetration energy E
*
d
at high impact velocity
contribute to the high E
*
p
in the simulations. The delocalized
penetration energy E
*
d
of the CNT lms without crosslinks is esti-
mated to be 3 and 18 MJ/kg at 8 and 12 km/s, respectively, based on
the simulation. In addition, E
*
d
increases to about 30.5 MJ/kg in
simulation while increasing the crosslink density to
r
¼35 at
v
i
¼12 km/s. It is, therefore, reasonable to believe that moreenergy
dissipation channels could be triggered at high impact velocity and
crosslink density, which will be discussed in the next sections.
3.3. Deformation behavior of CNT lms
As illustrated by experiments, the impact-induced deformation
mode of CNT lms changes after adding crosslinks. To better un-
derstand the strengthening mechanism of crosslinks, the
morphology evolutions of in-plane fracture of the CNT lms with
r
¼0 and 20 at v
i
¼12 km/s were analyzed and shown in Fig. 4.At5
ps, the perforation occurs with a perforated hole as marked by the
dashed blue circles accompanying with some broken bonds as
marked in red. The microstructure evolutions for
r
¼0 and 20 are
different. For
r
¼0, the perforated hole with uffy-edge is formed
due to the cascaded interface sliding as shown in Fig. 4(a). As the
density of crosslinks increases to 20, however, the perforated hole
with smooth edges is observed due to the strong interactions be-
tween adjacent CNT chains that suppress the interfacial sliding of
CNT bers as given in Fig. 4(b). The damage morphologies obtained
from simulations agree very well with the experimental observa-
tion as depicted in the insets of Fig. 4 at 5 ps. Furthermore, the
perforated hole of the CNT lm without crosslink, i.e.
r
¼0, expands
continually after impact from 10 to 20 ps. In contrast, it begins to
shrink with the addition of crosslinks (e.g.
r
¼20) after impact due
to the recovery of elastic deformation, as shown in Fig. 4. This
shrinkage effect is also observed in experimental results by
measuring the size of the post-mortem penetration hole in the
insets of Fig. 4 at 20 ps. Experimentally, for the CNT lm without
PVA, the size of the penetration hole is ~510
m
m, which is much
larger than the projectile diameter (~360
m
m). It decreases to
~300
m
m after adding PVA and is smaller than the projectile
diameter. This shrinkof the penetration hole with crosslinks can
be well understood by the constrained sliding effects of the CNT
chains introduced by the PVA, making the deformation of the CNT
lms around the penetration hole recover partially after penetra-
tion. Besides, the number of broken bonds, N
b
, increases from 56 to
73 with the increase of
r
from 0 to 20, indicating a transformation
of damage mode from vdW interfaces sliding to covalent bonds
breaking of intra-chains.
As depicted in Fig. 5(a), the out-of-plane displacement, z, the
diameter of conic deformation, d
c
, and the diameter of the perfo-
rated hole, d
b
, are extracted to quantitatively describe the inuence
of crosslinks on the dynamical responses of the CNT lms at
v
i
¼12 km/s. Fig. 5(b) indicates that the out-of-plane displacement
increases linearly with time for
r
¼0 after the perforation occurs at
5 ps due to the inertial effects. With the increase of
r
, the out-of-
plane displacement after impact decreases quickly, implying that
the added bonds can effectively restrict the sliding of CNT chains
and further increase the out-of-plane bending stiffness of the CNT
lms. The snapshots of structure morphologies as shown in the
inset of Fig. 5(b) also clearly characterize the less deformation
localization of the CNT lm with
r
¼20 compared to
r
¼0. As
shown in Fig. 5(c), during the expansion of the conic deformation,
the diameter of the conic region, d
c
, increases with the crosslink
density
r
and provides more area to dissipate the impact energy,
indicating a faster conic velocity at a larger
r
. A relatively larger
basal diameter yet a smaller out-of-plane displacement is observed
after adding crosslinks. The diameter of the perforated hole, d
b
,of
the CNT lm has the largest value for
r
¼0 and increases linearly
with time throughout the observation duration as shown in
Fig. 5(d). However, when the crosslink density,
r
, increases to 10
and 20, d
b
is smaller than that in
r
¼0 and tends to decrease after
penetration, indicating the transformation from continued sliding
to partial recovering after adding crosslinks. Therefore, crosslinks
Fig. 3. (a) Residual projectile velocities and (b) the specic penetration energy of the CNT lms with increased crosslink densities
r
at v
i
¼12 km/s and 8 km/s obtained from the
CGMD simulations. (A colour version of this gure can be viewed online.)
K. Xiao, X. Lei, Y. Chen et al. Carbon 175 (2021) 478e489
482
among the CNTs dramatically increase the interaction between the
tubes and signicantly strengthen the weak interfaces. The increase
of the difference between d
c
and d
b
,d
c
-d
b
, indicates more defor-
mation delocalization with the increased density of crosslink and
therefore more area participated in dissipating the impact energy,
resulting in a relatively high specic penetration energy.
3.4. Energy dissipation mode and failure mechanisms
It is vital to reveal the energy dissipation mode and failure
mechanisms of CNT lms with various crosslink densities by nu-
merical simulations to help design high-performance CNT lms
experimentally. The change of potential energy,
D
E, can corre-
spondingly be decomposed into the change of bending energy
D
E
b
,
Fig. 4. Perforated morphologies of CNT lms for (a)
r
¼0 and (b)
r
¼20, respectively, at v
i
¼12 km/s from simulations, which match well with the experiments result as given in the
insets. The marked red beads represent the broken bonds caused by the impact event. (A colour version of this gure can be viewed online.)
Fig. 5. Simulated deformation behavior of CNT lms. (a) Schematic of deformation behavior. (b), (c) and (d) Histories of out-of-plane displacement, diameter of conic deformation,
and diameter of penetration hole for
r
¼0, 10, and 20, respectively. (A colour version of this gure can be viewed online.)
K. Xiao, X. Lei, Y. Chen et al. Carbon 175 (2021) 478e489
483
stretching energy
D
E
s
, and vdW energy
D
E
v
. Here, the ratio of the
D
E
s
to the
D
E
b
, i.e.
D
E
s
/
D
E
b
, is used to explore the major energy
dissipation mode [35,43]. For v
i
¼8 km/s,
D
E
s
/
D
E
b
is smaller than 1
for
r
¼0 and 10 during the whole impact process as depicted in
Fig. 6, indicating the energy dissipation mode is dominated by the
bending of CNT bers, which is consistent with a previous study by
Wang et al. [18]. For
r
¼20,
D
E
s
/
D
E
b
initially increases quickly and
exceeds 1 after 2.5 ps, and then reaches the maximum value at 5 ps
when the perforation occurs. After that, it decreases constantly and
becomes smaller than 1, indicating the stretching-bending two-
stage dominant energy dissipation mode for the CNT lms with a
larger crosslink density. A critical crosslinks density,
r
c
¼10, for
transformation from the one-stage bending to the two-stage
stretching-bending dominant mode, is observed under v
i
¼8 km/
s. Once the stretching-bending-dominated energy dissipation is
triggered after adding crosslinks, there are more additional chan-
nels to dissipate the impact energy and therefore enhance the
specic penetration energy E
*
p
. As the impact velocity v
i
increases to
12 km/s, the plastic deformation of CNT lms implemented by
sliding and bending of the CNT bers is hugely restricted due to the
high loading rate.
D
E
s
/
D
E
b
always increases initially above 1 and
then decreases constantly no matter how many crosslinks are
added, indicating a stable stretching-bending two-stage energy
dissipation mode. As not much difference was observed for other
crosslink densities, only the results for
r
¼0 and 20 under
v
i
¼12 km/s are displayed in Fig. 6. This impact-velocity-dependent
dissipation mode is not only of signicance to reveal the process of
energy dissipation of CNT lms, but also important for optimizing
the design of protective materials with network structure at
different impact conditions.
The evolutions of microstructure morphologies of the CNT lms
with
r
¼0, 10, and 20 as shown in Fig. 7 are carefully investigated at
different scales to reveal the failure mechanism. At the small scale
as shown in Fig. 7(a), the CNT bers in the system bend severely for
r
¼0 and separate from others after impact due to the cascaded
interface sliding compared to its original state. However, the CNT
bers with
r
¼10 and 20 are still bonded together and bend slightly
after impact. At the medium scale as given in Fig. 7(b), a bundle of
molecules is selected and color-coded according to the local virial
stress along the impact direction. The microstructure morphol-
ogies, as well as the local atom stress in both original state and after
impact, are given for comparison. For all three crosslink densities
r
¼0, 10, and 20, the molecules on the edge of the conic defor-
mation region beneath the projectile experience a transition from
the relatively relaxed state with near-zero stress distribution to a
hybrid state with most tensile and some compressive stress
induced by impact. Fig. 7(a) and (b) show that both bending and
stretching of bers exist in such a network structure under impact
regardless of the impact velocity and crosslink density. With the
increase of crosslink density, the bending effect is relatively
restricted, whereas the stretching effect dominates the deforma-
tion and provides more channels to dissipate energy. At the large
scale as shown in Fig. 7(c), perforated holes are eventually formed
due to the continuous breaking of covalent bonds as marked by the
red beads. The size of the penetration hole decreases with
increasing the crosslinksdensity. In addition, the large-
deformation regions near the periphery of the perforated holes
also appear.
3.5. Effects of length, bending stiffness, and thickness of CNT lm
Three key factors, length and bending stiffness of constituent
CNTs, and thickness of CNT lm were investigated to illuminate
their effects on the ballistic resistance of CNT lms. Fig. 8(a) shows
the relationship between the specic penetration energy E
*
p
and the
dimensionless length
h
¼L/D at v
i
¼8 km/s, where Land Dare the
length of CNTs and the projectile diameter, respectively. E
*
p
is barely
dependent on the length of constituent CNT bers for
r
¼0 and 10
Fig. 6. Histories of
D
E
s
/
D
E
b
of the CNT lms with various crosslink densities at v
i
¼8 km/s and 12 km/s based on CGMD simulation results. (A colour version of this gure can be
viewed online.)
K. Xiao, X. Lei, Y. Chen et al. Carbon 175 (2021) 478e489
484
because the energy dissipation is mainly implemented by the inter-
tube sliding for low crosslink densities. This is consistent with the
previous nding that the energy dissipation efciency of CNT lms
without crosslinks varies scarcely with the length of the CNTs [19].
However, E
*
p
increases drastically with respect to
h
for
r
¼20,
indicating that the energy dissipation capability can be effectively
Fig. 7. Evolution of microstructure morphologies of CNT lms from simulations with
r
¼0, 10, and 20 at (a) small scale showing the bending degree decrease with
r
, (b) medium
scale indicating the hybrid state with most tensile and some compressive stress, and (c) large scale demonstrating the formation of penetration hole and large-deformation region,
respectively. The marked red beads in (c) represent the broken bonds. (A colour version of this gure can be viewed online.)
Fig. 8. Dependence relationship of the simulated specic penetration energy E*
pon (a) the ratio
h
for the lms with various
r
¼0, 10, and 20; and (b) the bending stiffness
l
for the
lm with
r
¼8. (A colour version of this gure can be viewed online.)
K. Xiao, X. Lei, Y. Chen et al. Carbon 175 (2021) 478e489
485
enhanced by increasing the length of CNTs only if enough crosslinks
are introduced to generate more inter-tubes bridges and restrain
the inter-tubes sliding.
To understand the effects of bending stiffness, a series of
bending stiffness was studied by altering the k
4
in Table 1 with a
factor
l
from 0.1 to 10 [30,35]at
r
¼8 and v
i
¼6 km/s to reect the
CNT lms from perforation to non-perforation with the increase of
bending stiffness. As shown in Fig. 8(b), E
*
p
increases quickly with
respect to
l
for
l
1.5, and then it increases slowly with the in-
crease of
l
for 1.5 <
l
5. For
l
>5, E
*
p
is saturated, implying that
there is a critical bending stiffness k
4c
, corresponding to
l
¼5, for
the present CNT lm system. The insets of Fig. 8(b) show the
hanging CNT chains around the edge of the perforated hole exhibit
folding characteristics for
l
¼1, and the CNT essentially becomes
more rigid with the increment of bending stiffness with increasing
l
(e.g. for
l
¼5). This result is consistent with the study by Myl-
vaganam et al. [19], which demonstrates that the single CNT with
larger radii, i.e. larger stiffness, own more capability to resist faster
bullet impact through full-atom MD simulations. It is to be noted
this critical bending stiffness is dependent on the crosslink density,
length of CNT bers, and impact velocity based on the above re-
sults. The coupling effects of these parameters on the protective
performance will be performed in the future to guide the optimi-
zation of the CNT lms.
Besides, the thickness of the lm is a crucial issue that affects the
ballistic resistance of lm materials. To investigate the effect of
thickness, two relatively thick (i.e. t¼10 nm and 15 nm) CNT lms
for crosslinksdensity
r
¼10 at v
i
¼12 km/s were also investigated
as given in Fig. 9.Fig. 9(a) shows that the 5-nm-thick CNT lm is
fully perforated and exhibits apparent out-of-plane displacement.
With the increase of thickness (see Fig. 9(b) and (c)), the impact
energy is not adequate to perforate the CNT lm and the projectile
bounces back after impact. The morphologies of the CNT lms with
thicknesses of 10 and 15 nm do not show noticeable change after
the impact, implying the nominal elastic deformation of the CNT
lms. The change of the potential energy,
D
E, as shown in Fig. 10(a),
can be decomposed into the changes of bending energy,
D
E
b
,
stretching energy,
D
E
s
, and vdW energy,
D
E
v
.FromFig. 10(a) and (b),
it clearly show that the change of vdW interface energy is negli-
gible, and stretching and bending of the CNT chains are the domi-
nant energy dissipation methods for all the thicknesses. Fig. 10(c)
shows the ratio of the number of broken bonds to the total bond
number in the system, n, to characterize the damage degree of the
CNT lm. The ndecreases quickly from 5.80
to 0.23
with
increasing the thickness from 5 nm to 10 nm. Then it keeps almost
at the same level with continually increasing the thickness to
15 nm. As shown in Fig. 10(d), E
*
p
increases with respect to thick-
ness, t, of the CNT lm, implying that more impact energy can be
dissipated by the thick lms. Three movies are provided in Sup-
plementary materials to show the dynamical responses for three
thicknesses with
r
¼10 at v
i
¼12 km/s. Details are omitted here for
simplicity.
According to the simulation results as shown in Figs. 9 and 10,
there is not obvious transition of energy dissipation mode for the
CNT lms with increasing the thickness from 5 nm to 15 nm.
However, for the thin CNT lm with a thickness of 5 nm, the in-
plane tension induced stretching and bending provide the mainly
impact energy dissipation channels, whereas both the local
compression and in-plane tension induced stretching and bending
of CNT chains contribute to the impact energy dissipation for the
relatively thick CNT lms with thicknesses of 10 and 15 nm.
In addition to introducing more crosslinks, we have four
methods to improve the impact resistance of CNT lm based on the
results in this paper: the rst is to add crosslinks to enhance the
Fig. 9. Simulated morphologies of the CNT lms with thicknesses of (a) 5 nm, (b) 10 nm, and (c) 15 nm for
r
¼10 during impact. (A colour version of this gure can be viewed
online.)
K. Xiao, X. Lei, Y. Chen et al. Carbon 175 (2021) 478e489
486
inter-tube connections; the second is to increase the bending
stiffness of CNTs, which can be realized by employing multiple
walled carbon nanotubes with larger diameter; the third is to uti-
lize long CNTs, which has also been studied by Bai et al. [44]; the
last one is to increase the thickness of CNT lm with densication
procedure, as the study by Gao et al. [45]. It is noted that the
regulation schemes abovementioned should also be applicable to
enhance the anti-penetration ability of other CNT assemblies like
CNT aerogels [46e48].
4. Conclusions
The impact resistance of CNT lms with and without crosslinks
was investigated by micro-ballistic impact experiments and CGMD
simulations. The CGMD simulations qualitatively reproduced two
typical experimental phenomena: the rst one is the broken
morphology around the penetration hole of the CNT lm changing
from the uffystate to the smoothstate after adding the PVA;
the second one is the shrinkage of the perforation hole after adding
PVA. The main conclusions can be further summarized as follows.
1. The introduction of crosslinks can efciently enhance the ef-
ciency of the energy dissipation of CNT lm due to the enhanced
interaction between adjacent CNTs by adding covalent bonds.
2. A transition from a system where the energy dissipation is
dominated by the bending of the CNTs, to one in which the
stretching-bending-dominated, is observed with increased
density of crosslinks, implying more energy dissipation
channels can be activated at a high impact velocity and a high
crosslink density.
3. For the high crosslinksdensity, both increase the length of in-
dividual CNT and increase the thickness of CNT lm can effec-
tively improve the energy dissipation capability.
4. The increase of bending stiffness of CNTs can effectively increase
the impact resistance of CNTs. However, for a given crosslink
density, a critical bending stiffness is observed for the saturation
of the protective performance of CNT lm.
Methods
Materials. The continuous CNT lms were achieved by densi-
fying CNT aerogels with almost awless and high-crystallinity CNTs
(diameter and length within the range of 0.7e15 nm and
300e360
m
m respectively). The high-crystallinity CNTs were
fabricated by oating catalyst chemical vapor deposition (FCCVD)
method and collected at the end of the corundum tube layer by
layer. The thickness of the CNT lms was controlled by the number
of winding layers. In this work, 10-
m
m-thick lm was obtained by 10
layers of CNT aerogels, in which the CNT bundles were stacked
randomly to form nano-network structures. The volume fraction of
PVA is ~6.8%, which is negligible compared to that of the CNT chains
and can be considered as an enhancement of CNT lm to serves as
crosslinks. The basic morphologies and quality of the adopted CNT
lms are characterized by SEM, Transmission Electron Microscopy
(TEM), and Raman spectroscopy, as shown in Fig. 11.Fig. 11(a) and
(b) show the prepared CNT lms without and with PVA as
Fig. 10. (a) Potential energy histories, (b)
D
E
s
/
D
E
b
, (c) ratio of broken bonds, and (d) specic penetration energy, E*
p, of the CNT lms with various thicknesses based on the CGMD
simulations. (A colour version of this gure can be viewed online.)
K. Xiao, X. Lei, Y. Chen et al. Carbon 175 (2021) 478e489
487
crosslinks, respectively, in which some obvious PVA enhanced re-
gions have been marked by white circles in Fig.11(b). After PVA was
added, the binder effect will generate between CNT chains to
enhance the weak vdW interfaces. The CNT in the experiments is a
mixture of SWCNT and multi-wall CNT (MWCNT), and the diameter
of the CNT ranges from 0.7 to 15 nm as shown in Fig. 11(c). Raman
spectroscopy is employed to investigate the crystallization degree
of the CNT lm as shown in Fig. 11(d). The intensity ratio I
D
/I
G
is
used to characterize the structural integrity of the CNT lm. For the
CNT lm without PVA, its crystallinity is relatively high. However,
the nominal crystallinity of the CNT lm decreases due to the
appearance of the PVA.
Fabrication of Numerical Sample. The CNT lm in the CGMD
model is achieved by depositing 100 CNTs with initial curvatures on
a rigid Lennard-Jones (LJ) wall placed at the bottom of the simu-
lation box under body forces. The isothermal-isobaric ensemble
(NPT) with Langevin thermostat at 300 K and Berendsen barostat at
0 Pa is applied during preparing the CNT lm. The time step is set to
be 10 fs. Once all the CNTchains reach almost equilibrium positions,
the rigid wall and the body forces are removed, and the
equilibrium-state pristine CNT lm is obtained after a further
500,000 time-steps (i.e. 5 ns).
CRediT authorship contribution statement
Kailu Xiao: Conceptualization, Methodology, Validation, Data
curation, Formal analysis, Writing - original draft, preparation.
Xudong Lei: Methodology, Investigation, Data curation. Yuyu
Chen: Methodology, Investigation, Data curation. Qi An: Validation,
Writing - review &editing. Dongmei Hu: Methodology, Validation,
Data curation, Formal analysis. Chao Wang: Conceptualization,
Validation, Methodology, Writing - review &editing. Xianqian Wu:
Conceptualization, Methodology, Formal analysis, Writing - review
&editing, Funding acquisition. Chenguang Huang: Supervision,
Funding acquisition.
Declaration of competing interest
The authors declare that they have no known competing
nancial interests or personal relationships that could have
appeared to inuence the work reported in this paper.
Acknowledgments
This work was supported by the National Natural Science
Foundation of China (Grant Nos. 11672315, 11972348, and
11772347), Science Challenge Project (Grant No. TZ2018001), and
the Strategic Priority Research Program of Chinese Academy of
Sciences (Grant Nos. XDB22040302, and XDB22040303).
Appendix A. Supplementary data
Supplementary data to this article can be found online at
https://doi.org/10.1016/j.carbon.2021.01.009.
Fig. 11. The CNT lm in (a) pristine state (i.e. without crosslinks) and (b) after adding PVA (i.e. with crosslinks), respectively, in which some obvious PVA enhanced regions have been
marked by white circles). (c) TEM image illustrated the number of CNT walls. (d) Raman spectroscopy results indicated the quality of the CNT lm. (A colour version of this gure can
be viewed online.)
K. Xiao, X. Lei, Y. Chen et al. Carbon 175 (2021) 478e489
488
References
[1] S. Zangana, J. Epaarachchi, W. Ferdous, J. Leng, A novel hybridised composite
sandwich core with Glass, Kevlar and Zylon bres einvestigation under low-
velocity impact, Int. J. Impact Eng. 137 (2020) 103430.
[2] S. Khatiwada, C.A. Armada, E.V. Barrera, Hypervelocity impact experiments on
epoxy/ultra-high molecular weight polyethylene ber composites reinforced
with single-walled carbon nanotubes, Procedia Eng 58 (2013) 4e10.
[3] S. Shrestha, B.D. Dunn, Plasma Electrolytic Oxidation and Anodising of
Aluminium Alloys for Spacecraft Applications, Surface Engineering of Light
Alloys, Woodhead Publishing, 2010, pp. 603e641.
[4] B.G. Cour-Palais, J.L. Crews, A multi-shock concept for spacecraft shielding, Int.
J. Impact Eng. 10 (1) (1990) 135e146.
[5] M. Zhang, S. Fang, A.A. Zakhidov, S.B. Lee, A.E. Aliev, C.D. Williams, et al.,
Strong, transparent, multifunctional, carbon nanotube sheets, Science 309
(5738) (2005) 1215.
[6] J. Xu, Y. Li, Y. Xiang, X. Chen, A super energy mitigation nanostructure at high
impact speed based on buckyball system, PloS One 8 (5) (2013) e64697-
e64697.
[7] G. Lu, T. Yu, Energy Absorption of Structures and Materials, Woodhead Pub-
lishing, 2003, pp. 174e214.
[8] B. Xie, Y. Liu, Y. Ding, Q. Zheng, Z. Xu, Mechanics of carbon nanotube net-
works: microstructural evolution and optimal design, Soft Matter 7 (21)
(2011), 10039-0.
[9] N. Anzar, R. Hasan, M. Tyagi, N. Yadav, J. Narang, Carbon nanotube - a review
on Synthesis, Properties and plethora of applications in the eld of biomedical
science, Sensors Inter 1 (2020) 100003.
[10] S. Sakurai, F. Kamada, D.N. Futaba, M. Yumura, K. Hata, Inuence of lengths of
millimeter-scale single-walled carbon nanotube on electrical and mechanical
properties of buckypaper, Nanoscale Res. Lett. 8 (1) (2013) 546.
[11] Y.J. Ma, X.F. Yao, Q.S. Zheng, Y.J. Yin, D.J. Jiang, G.H. Xu, et al., Carbon nanotube
lms change Poissons ratios from negative to positive, Appl. Phys. Lett. 97 (6)
(2010).
[12] S.W. Cranford, M.J. Buehler, In silicoassembly and nanomechanical charac-
terization of carbon nanotube buckypaper, Nanotechnology 21 (26) (2010)
265706.
[13] W. Wang, Y. Zhang, L. Xu, X. Li, Mechanical properties of high-strength Q960
steel at elevated temperature, Fire Saf. J. 114 (2020) 103010.
[14] X. Liu, Z. He, J. Ye, L. Yan, S. Li, Y. Tang, Study on dynamic mechanical behavior
of Q460JSC and HQ600 high strength steel, J. Constr. Steel Res. 173 (2020)
106232.
[15] J.J. Andrew, S.M. Srinivasan, A. Arockiarajan, H.N. Dhakal, Parameters inu-
encing the impact response of ber-reinforced polymer matrix composite
materials: a critical review, Compos. Struct. 224 (2019) 111007.
[16] T.-L. Chu, C. Ha-Minh, A. Imad, A numerical investigation of the inuence of
yarn mechanical and physical properties on the ballistic impact behavior of a
Kevlar KM2®woven fabric, Compos. B Eng. 95 (2016) 144e154.
[17] C. Wang, B. Xie, Y. Liu, Z. Xu, Mechanotunable microstructures of carbon
nanotube networks, ACS Macro Lett. 1 (10) (2012) 1176e1179.
[18] S. Wang, E. Gao, Z. Xu, Interfacial failure boosts mechanical energy dissipation
in carbon nanotube lms under ballistic impact, Carbon 146 (2019) 139e146.
[19] H. Chen, L. Zhang, J. Chen, M. Becton, X. Wang, H. Nie, Energy dissipation
capability and impact response of carbon nanotube buckypaper: a coarse-
grained molecular dynamics study, Carbon 103 (2016) 242e254.
[20] A. Satti, A. Perret, J.E. Mccarthy, K. GkY, Covalent crosslinking of single-walled
carbon nanotubes with poly(allylamine) to produce mechanically robust
composites, J. Mater. Chem. 20 (37) (2010) 7941.
[21] S. Qu, X. Jiang, Q. Li, L. Gao, G. Zhou, D. Zhang, et al., Developing strong and
tough carbon nanotube lms by a proper dispersing strategy and enhanced
interfacial interactions, Carbon 149 (2019) 117e124.
[22] S. Wang, Z. Liang, B. Wang, C. Zhang, High-strength and multifunctional
macroscopic fabric of single-walled carbon nanotubes, Adv. Mater. 19 (9)
(2007) 1257e1261.
[23] T. Filleter, R. Bernal, S. Li, H.D. Espinosa, Ultrahigh strength and stiffness in
cross-linked hierarchical carbon nanotube bundles, Adv. Mater. 23 (25) (2011)
2855e2860.
[24] C. Wang, Y. Li, L. Tong, Q. Song, K. Li, J. Li, et al., The role of grafting force and
surface wettability in interfacial enhancement of carbon nanotube/carbon
ber hierarchical composites, Carbon 69 (2014) 239e246.
[25] P. Lv, Y.Y. Feng, P. Zhang, H.M. Chen, N. Zhao, W. Feng, Increasing the inter-
facial strength in carbon ber/epoxy composites by controlling the orientation
and length of carbon nanotubes grown on the bers, Carbon 49 (14) (2011)
4665e4673.
[26] C. Velasco-Santos, A.L. Martinez-Hernandez, V.M. Castano, Carbon nanotube-
polymer nanocomposites: the role of interfaces, Compos. Interfac. 11 (8e9)
(2005) 567e586.
[27] J.H. Lee, P.E. Loya, J. Lou, E.L. Thomas, Dynamic mechanical behavior of
multilayer graphene via supersonic projectile penetration, Science 346 (6213)
(2014) 1092e1096.
[28] M. Hassani-Gangaraj, D. Veysset, K.A. Nelson, C.A. Schuh, Melting can hinder
impact-induced adhesion, Phys. Rev. Lett. 119 (17) (2017).
[29] M. Hassani-Gangaraj, D. Veysset, K.A. Nelson, C.A. Schuh, Melt-driven erosion
in microparticle impact, Nat. Commun. 9 (1) (2018) 5077.
[30] S. Cranford, H. Yao, C. Ortiz, M.J. Buehler, A single degree of freedom lollipop
model for carbon nanotube bundle formation, J. Mech. Phys. Solid. 58 (3)
(2010) 409e427.
[31] M.J. Buehler, Mesoscale modeling of mechanics of carbon nanotubes: self-
assembly, self-folding, and fracture, J. Mater. Res. 21 (11) (2011) 2855e2869.
[32] Q.-M. Tan, Fundamental Principles of Dimensional Analysis, Dimensional
Analysis, Springer Berlin Heidelberg, Berlin, Heidelberg, 2011, pp. 7e16.
[33] C. Wang, E. Gao, L. Wang, Z. Xu, Mechanics of network materials with
responsive crosslinks, Compt. Rendus Mec. 342 (5) (2014) 264e272.
[34] C. Wang, S. Chen, Viscoelastic properties of randomly entangled carbon
nanotube networks under cyclic tension loading, Comput. Mater. Sci. 119
(2016) 46e51.
[35] T. Yang, C. Wang, Z. Wu, Crosslink-tuned large-deformation behavior and
fracture mode in buckypapers, Carbon 159 (2020) 412e421.
[36] S. Plimpton, Fast Parallel Algorithms for Short-Range Molecular Dynamics,
Academic Press Professional, Inc., 1995.
[37] A. Stukowski, Visualization and analysis of atomistic simulation data with
OVITOethe Open Visualization Tool, Model. Simul. Mater. Sc. 18 (1) (2009),
015012.
[38] N.K. Gupta, M.A. Iqbal, G.S. Sekhon, Experimental and numerical studies on
the behavior of thin aluminum plates subjected to impact by blunt- and
hemispherical-nosed projectiles, Int. J. Impact Eng. 32 (12) (2006) 1921e1944.
[39] J. Dean, C.S. Dunleavy, P.M. Brown, T.W. Clyne, Energy absorption during
projectile perforation of thin steel plates and the kinetic energy of ejected
fragments, Int. J. Impact Eng. 36 (10) (2009) 1250e1258.
[40] B.L. Lee, T.F. Walsh, S.T. Won, H.M. Patts, J.W. Song, A.H. Mayer, Penetration
failure mechanisms of armor-grade ber composites under impact, J. Compos.
Mater. 35 (18) (2001) 1605e1633.
[41] L.H. Abbud, A.R.A. Talib, F. Mustapha, H. Tawque, F.A. Najim, Behaviour of
transparent material under high velocity impact, Inter, J. Mech. Mater. E 5 (1)
(2010).
[42] S. Leigh Phoenix, P.K. Porwal, A new membrane model for the ballistic impact
response and V50 performance of multi-ply brous systems, Int. J. Solid
Struct. 40 (24) (2003) 6723e6765.
[43] G.A. Buxton, N. Clarke, Bending to stretchingtransition in disordered net-
works, Phys. Rev. Lett. 98 (23) (2007) 238103.1e238103.4.
[44] Y. Bai, R. Zhang, X. Ye, Z. Zhu, H. Xie, B. Shen, et al., Carbon nanotube bundles
with tensile strength over 80 GPa, Nat. Nanotechnol. 13 (7) (2018) 589e595.
[45] E. Gao, W. Lu, Z. Xu, Strength loss of carbon nanotube bers explained in a
three-level hierarchical model, Carbon 138 (2018) 134e142.
[46] M.A. Worsley, S.O. Kucheyev, J.H.S. Jr, A.V. Hamza, T.F. Baumann, Mechanically
robust and electrically conductive carbon nanotube foams, Appl. Phys. Lett. 94
(7) (2009) 787.
[47] J. Liu, W. Gong, Y. Yao, Q. Li, J. Jiang, Y. Wang, et al., Strengthening carbon
nanotube bers with semi-crystallized polyvinyl alcohol and hot-stretching,
Compos. Sci. Technol. 164 (2018) 290e295.
[48] D. Hu, W. Gong, J. Di, D. Li, R. Li, W. Lu, et al., Strong graphene-interlayered
carbon nanotube lms with high thermal conductivity, Carbon 118 (2017)
659e665.
K. Xiao, X. Lei, Y. Chen et al. Carbon 175 (2021) 478e489
489
... Similarly, each bead in the coarse-grained graphene represents a graphene flake with a side length of 2.5 nm, as shown in Fig. 1b-IV. The validity of the method and model has been verified by a series of studies by Wang et al. [18,40,41], Xie et al. [42], Xiao et al. [43], and our group [22,25,31]. ...
... However, the microscopic structures observed in the experiment can be qualitatively simulated, and the microscopic deformation mechanism can be revealed. This method has been used in a series of studies by Wang et al. [25,53,56], Xie et al. [57], Xiao et al. [58,59], and our group [42][43][44]. ...
... However, the microscopic structures observed in the experiment can be qualitatively simulated, and the microscopic deformation mechanism can be revealed. This method has been used in a series of studies by Wang et al. [25,53,56], Xie et al. [57], Xiao et al. [58,59], and our group [42][43][44]. ...
Article
Graphene-filling significantly improves the compressive properties of carbon nanotube (CNT) foam (CF). However, the role of graphene in the graphene-filled CNT foam (GFCF) and the corresponding microscopic deformation mechanism, are still unclear. Here, coarse-grained numerical models of pure CF and GFCF are constructed based on molecular dynamics, and the role of graphene in GFCF and corresponding microscopic deformation mechanisms under compression are investigated. It is found that the compressive modulus of GFCF (5.3 MPa) is much larger than that of pure CF (0.7 MPa). The filling of graphene inhibits the aggregation of CNTs, enhances the dispersion of CNTs, impedes the rearrangement of CNTs, and ultimately improves the compressive modulus of the whole material by improving the bending ability of CNTs. It is further found that the compressive modulus of GFCF can be increased to a maximum of 7.4 MPa as the number of graphene flakes increases to 300, but remains almost the same as the graphene thickness increases. The results in this paper deepen the understanding of the microscopic mechanisms of GFCF and provide scientific guidance for the application of CNT and graphene-based materials.
... Wang et al. [19] studied the dissipation of CNN materials under impaction and pointed out that the brittle fracture of carbon nanotubes and ductile fracture of van der Waals interfaces between the nanotubes in forms of detachments or sliding are the channels to dissipate mechanical energy. Later, Xiao et al. [20] studied the similar impaction behavior of CNN materials and observed a mode transition from bending-dominated dissipation to stretching-bending-dominated one by tuning the fraction of strong crosslinkers. Very recently, Yu et al. [21] found that double-crosslinked CNNs can nicely achieve cooperative energy dissipation with minimal structural damage by applying CGMD simulations. ...
Article
Friction behavior of carbon nanotube networks (CNNs) is ubiquitous in applications, however, is poorly investigated. Here, we employ coarse-grained molecular dynamics (CGMD) simulations to investigate friction behaviors and microscopic mechanism of CNNs. We first give a phase diagram to determine the initial contacting state of “supported” or “trapped” for an indenter in CNNs. In a “supported” state, friction force is mainly originated from the local surface adhesion of CNNs which experience stable elastic deformation in friction process. In contrast, in a “trapped” state, friction force is mainly activated by nonlocal reconstitution of carbon nanotubes (CNTs) accompanied with irreversible bond breaking. Furthermore, with an increased normal pressure, the friction force keeps nearly constant for a “supported” indenter while it increases linearly for a “trapped” one; the friction force is linearly or nonlinearly related to the sliding velocity of an indenter in a supported or trapped state. Importantly, there is a critical crosslink density, above which, the coefficient of friction is greatly decreased due to enhanced integrity of CNNs. These results provide a profound understanding of friction deformation behavior of CNNs, which is of great significance for optimal design in practical applications.
Article
Full-text available
Recent remarkable advances in the field of nanotechnology has been achieved in the last few years especially in the fabrication of sensors that have wide number of applications. Nanomaterials are the foundation of nanotechnology that are measured on nanoscale. Carbon nanotubes (CNTs) are tube-like materials that are made up of carbon with a diameter calculating on a nanometer scale. They are originated from graphite sheet and these graphite layers seems similar to a rolled up non-stop unbreakable hexagonal like mesh structure and the carbon molecules appears at the apexes of the hexagonal structures. Depending upon the number of carbon layers, carbon nanotubes can be single-walled carbon nanotubes (SWCNTs), double-walled carbon nanotubes (DWCNTs) and multi-walled carbon nanotubes (MWCNTs). Carbon nanotubes (CNTs) can be fabricated by three main methods i.e., chemical vapor deposition, electric arc method and laser deposition method. Carbon nanotubes exhibit various characteristic properties such as high elasticity, high thermal conductivity, low density and they are chemically more inert etc. Due to these interesting properties, carbon nanotubes have played a significant role in the field of nanotechnology, electronics, optics and other fields of materials science. Carbon nanotubes are being positively applied in drug delivery, sensing, water treatment etc. Functionalization of their surface can result in highly soluble materials, which can be further derivatized with active molecules, making them compatible with biological systems. Surface functionalization enables adsorption or attachment of various molecules or antigens, which subsequently can be targeted to the desired cell population for immune recognition or a therapeutic effect. In this review, properties of carbon nanotubes and their clinical applications such as medical diagnostics and drug delivery are being discussed. Here, antibacterial as well as antifungal activity of carbon nanotubes are also being reviewed.
Article
Full-text available
Impact-induced erosion is the ablation of matter caused by being physically struck by another object. While this phenomenon is known, it is empirically challenging to study mechanistically because of the short timescales and small length scales involved. Here, we resolve supersonic impact erosion in situ with micrometer- and nanosecond-level spatiotemporal resolution. We show, in real time, how metallic microparticles (~10-μm) cross from the regimes of rebound and bonding to the more extreme regime that involves erosion. We find that erosion in normal impact of ductile metallic materials is melt-driven, and establish a mechanistic framework to predict the erosion velocity.
Article
High-strength steel has been gradually applied in practical engineering and achieved good economic and social benefits. Due to the threat of terrorist attacks and accidental explosions, high-strength steel structures need to have high blast resistance. However, the studies on the mechanical properties and constitutive equations of high-strength steels under high strain rate are insufficient. The research on the anti-explosion properties of high-strength steel structures lacks material models and constitutive parameters. Therefore, this paper investigated the quasi-static and dynamic mechanical properties of Q460JSC and HQ600 high strength steels. A series of quasi-static tensile, high-speed tensile and SHTB tests were conducted to obtain the material properties under different strain rate in the range of 0.00025 s⁻¹ to 5000 s⁻¹. Combined with finite element simulation by LS-DYNA, the strain hardening relationship of specimens after necking was treated by inverse extrapolation method, and the true stress-strain curves at different strain rates were obtained. The effects of strain rate on materials strength and plasticity were analyzed. The quasi-static and dynamic constitutive parameters of these two high strength steels were fitted by Ludwik criterion and modified Johnson-Cook model, respectively. The results show that the yield strength, ultimate strength and fracture strain of these two kinds of high strength steels all increase with the increase of strain rate, but the uniform elongation is not sensitive to strain rate. The change of strength is not obvious in the strain rate range of 10³. The inverse extrapolation method is effective in dealing with the true stress-strain curve after necking, and more accurate results can be obtained by continuously optimizing the strain hardening index. The improved Ludwik criterion and the new Johnson-Cook model correction formula can well describe the stress-strain relationship of these two high-strength steel materials in the range of 10⁻³ s⁻¹ to 10³ s⁻¹, which provide basis for the investigation on anti-explosion properties of high-strength steel structures.
Article
This study experimentally investigates the temperature-induced deterioration in the mechanical properties of high-strength Q960 steel. High-strength steel has numerous advantages compared with conventional mild-strength steel and has gained wide applications in engineering structures, particularly in high-rise and long-span buildings. The fundamental issue in the fire-resistance design of high-rise or long-span steel buildings is the evaluation of mechanical properties under elevated temperature. There is little information on the temperature-induced mechanical-property deterioration of Q960 steel because of the limited research on this subject. To bridge this knowledge gap, a series of tensile coupon tests on Q960 steel was carried out under various temperatures from 20 to 900 °C to obtain the corresponding stress–strain relationships, yield strength, ultimate strength, and elastic modulus at elevated temperatures. The test results were compared with the results of other high-strength steel types such as Q460, Q690, and S960 steels reported in previous studies as well as the mechanical properties recommended in applicable design codes or standards. The comparison showed that the effects of temperature on the mechanical properties among different types of high-strength steels are different. New predictive equations for evaluating the material properties of Q960 steel at elevated temperatures were presented.
Article
Strong physical and/or chemical inter-tube crosslinks play a vital role in carbon nanotube buckypapers and composites. However, both underlying mechanisms and regulation patterns remain poorly understood. Here, we employed the coarse-grained molecular dynamics simulations to investigate the nonlinear large-deformation behavior and the fracture mode of crosslinked buckypapers by considering both intrinsic intra- and inter-tube bond-breaking. A critical crosslink density ρc is found to divide the deformation mode of buckypaper into two regimes in uniaxial tension, i.e., the bending-dominated mode at ρ < ρc and the bending-stretching-bending three-stage one at ρ > ρc. This transition is attributed to the stress concentration and the intrinsic bond-breaking in large tensile deformations. In uniaxial compression, it is always bending-dominated, which is independent of crosslinks and compressive strain. Furthermore, there exists another critical crosslink density controlling the ductile-brittle transition of fracture mode of the strongly-crosslinked buckypaper, which is explained from the viewpoint of the collective hierarchical microstructural evolution. This study provides a profound understanding of crosslink effect on the buckypaper, which is of great significance for the optimization design and further practical applications of the promising material.
Article
A novel fibre composite sandwich core has been introduced in this study. Trapezoidal corrugated core glass-fibre sandwich structures were hybridised using Kevlar and Zylon fibres to improve the dynamic impact performance. The composite cores were fabricated with four layers of glass fibre and one of the layers was replaced either by Kevlar or Zylon fibre to create hybrid composite core (Glass-to-Kevlar or Glass-to-Zylon ratio 75:25). The impact behaviour, damage mode, specific absorbed energy, and residual strength after the impact of the composite sandwiches were investigated using a low-velocity impact test with 30 J, 40 J and 50 J kinetic energy level. The experimental results revealed that the hybridised sandwiches with high-performance fibre are performing extremely well when subjected to impact energy above the threshold limit. The observations during the experimental work and numerical simulation have confirmed that Glass-Kevlar and Glass-Zylon hybridisation can eliminate severe core rupture by minimising stress concentration and provide high specific energy without increasing structural weight. Moreover, the loss of strength and stiffness of trapezoidal corrugated core sandwich structures after an impact event can be minimised up to 56% and 69%, respectively using Glass-Zylon hybridisation technique. Furthermore, an empirical relationship for predicting the residual strength of the composite core sandwich is proposed.
Article
The damage of fiber reinforced polymer matrix composite materials induced by impact load is one of the most critical factors that restrict extensive use of these materials. The behavior of composite structures under transient impact loading and the ways to enhance their characteristics to withstand this type of dynamic loading might be of specific significance in the aerospace sector and other applications. This paper critically reviews the important parameters from the published literature influencing the impact resistance and the damage mechanics of fiber-reinforced composite materials. Firstly, the paper reviews the influence of impact velocity on various failure modes. Following this, a comprehensive review on the four key parameters specifically material, geometry, event and the environmental-related conditions that affect the structural behavior of fiber reinforced polymer matrix composites to impact loading is discussed. The review further outlines areas to improve the impact damage characteristics of composites and then conclude with a summary of the discussion on the future work relating to the most influencing parameters.