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

Journal of Molecular Modeling (2018) 24:178

https://doi.org/10.1007/s00894-018-3716-6

ORIGINAL PAPER

Effect of temperature on elastic properties of CNT-polyethylene


nanocomposite and its interface using MD simulations
Akhileshwar Singh 1 & Dinesh Kumar 1

Received: 8 January 2018 / Accepted: 12 June 2018


# Springer-Verlag GmbH Germany, part of Springer Nature 2018

Abstract
This paper investigates the effect of temperature on the elastic modulus of carbon nanotube-polyethylene (CNT-PE) nanocom-
posite and its interface using molecular dynamics (MD) simulations, by utilizing the second-generation polymer consistent force
field (PCFF). Two CNTs—armchair and zigzag—were selected as reinforcing nano-fillers, and amorphous PE was used as the
polymer matrix. For atomistic modelling of the nanocomposite, the commercially available code Materials Studio 8.0 was used
and all other MD simulations were subsequently performed using the open source code Large-Scale Atomic/Molecular
Massively Parallel Simulator (LAMMPS). To obtain the elastic modulus of the nanocomposite, stress-strain curves were drawn
at different temperatures by performing uniaxial deformation tests on the nanocomposite material, whereas the curvatures of the
interfacial interaction energy vs. strain curves were utilized to obtain Young’s modulus of the interface. In addition, the glass
transition temperatures of the polymer matrix and nanocomposites were also evaluated using density-temperature curves. Based
on the results, it is concluded that, irrespective of temperature condition, a nanocomposite reinforced with CNT of larger chirality
(i.e., armchair) yields a higher value of Young’s modulus of the nanocomposite and its interface. It was also found that, at the
phase transition (from a glassy to a rubbery state) temperature (i.e., glass transition temperature), Young’s moduli of the polymer
matrix, nanocomposite, and its interface drop suddenly. The results obtained from MD simulations were verified with results
obtained from continuum-based rule-of-mixtures.

Keywords CNT . Polyethylene . Nanocomposite . Interface . Molecular dynamics . Young’s modulus

Introduction filler, and a few weight-percentage of CNTs reinforcement in a


polymer matrix can significantly improve the mechanical
The discovery of lightweight carbon nanotubes (CNTs) [1] in properties of the resulting nanocomposite. A significant num-
1991 stimulated the attention of researchers across the world ber of experimental and computational studies have been con-
in the fields of science and engineering because of their ex- ducted to investigate the effective elastic properties of CNT-
ceptional physical and chemical properties. For example, reinforced polymer nanocomposites (CRPN).
CNTs are found to possess Young’s modulus, thermal and In the last two decades, many experimental works [7–11]
electrical conductivities in the ranges of 1–5 TPA [2–5], have demonstrated that reinforcement of CNTs in a polymer
6000 W/mK [5] and 200,000 S cm−1 [6], respectively. matrix invokes the enhancement of mechanical properties of
Combined with these remarkable properties, a very high as- the nanocomposite. But due to the difficulty and high cost
pect ratio of CNTs, varying from few hundreds to several involved in experimental works at nanoscales, continuum-
thousand on the nanoscale, makes them promising reinforcing and atomistic-based computational approaches have been
used as powerful tools to predict the properties of CNTs and
other nano-fillers as well as their nanocomposites. Atomistic
* Dinesh Kumar
dkumar.mech@mnit.ac.in
simulations can accurately capture the real nature of the atom-
ic interactions, but are computationally expensive compared
Akhileshwar Singh with continuum-based techniques. Various atomistic-based
akhileshwarsingh@live.com
techniques used are: ab initio, tight bonding molecular dynam-
1
Department of Mechanical Engineering, Malaviya National Institute ic (TBDM), density functional theory (DFT) and molecular
of Technology, Jaipur 302017, India dynamic (MD). Of these, the first three are based on quantum
178 Page 2 of 11 J Mol Model (2018) 24:178

mechanics, which need a solution of Schrödinger wave equa- angle have higher interfacial interaction energy. Yang et al.
tion for each atom, whereas the latter (MD) requires solving [23] developed a multiscale approach to consider the effect
Newton’s second law of motion and provide an acceptable of CNT size and non-bonded interaction at the interface on the
level of accuracy with lesser run time than ab initio, TBDM, effective elastic stiffness of CRPN using MD simulations and
and DFT. a modified multi-inclusion model using continuum
There are many research articles [12–20] predicting the micromechanics. The load transfer behavior of graphene/
mechanical properties of CRPN using MD simulation. Han CNT-PE hybrid nanocomposite was investigated by Zhang
and Elliott [12] investigated the elastic properties of a CRPN et al. [24], and it was found that the peak strength of traction
and found longitudinal Young’s modulus of CRPN equal to curve for PE matrix is much lower than the interface of PE/
94.6 GPa and 138.9 GPa for 12% and 17% volume fractions graphene and graphene/graphene. Arash et al. [25] proposed
of CNT, respectively. Al-ostaz et al. [13] studied the effect of an approach to predict Young’s modulus of the interfacial
aligned- and randomly oriented CNT-reinforced polymer ma- region between CNT-reinforced PMMA nanocomposites
trix and concluded that the clustering of CNTs reduces the and also presented a three-phase micromechanical model for
transverse elastic properties as compared to single-CNT rein- predicting the elastic properties of the nanocomposite. Very
forced nanocomposite. Mandal et al. [15] investigated the ef- recently, Srivastava and Kumar [26] developed a continuum
fect of temperature on the elastic properties of CRPN and level model to characterize interphase region and found that
reported that, as the temperature increases from 100 K to elastic properties of the nanocomposite are enhanced by con-
400 K, the stiffness of CRPN decreases. Rouhi et al. [17] sidering the effect of interface for CRPN.
investigated the influence of CNT reinforcement on the elastic Moreover, Young’s modulus and high thermal conductivity
properties CNT-reinforced poly(ethylene oxide) nanocompos- of CNTs can significantly modify the thermo-mechanical
ite and found dense polymer matrix around CNT, and it was properties of the CNT-polymer nanocomposite. Further, these
also shown that zigzag-CNT reinforcement provides higher properties of CNT, as well as the properties of the resulting
Young’s modulus to the nanocomposite. The effects of single nanocomposite, such as elastic moduli [27] and load transfer
adatom (SA) and Stone-Wales (SW) defects in CNT on the characteristic [28, 29], are greatly influenced by the high-
elastic properties of polypropylene matrix nanocomposite was temperature environments, especially when the temperature
studied by Lv et al. [18]. Chawla et al. [21] conducted MD involved is higher than the glass transition temperature (Tg)
simulations to study the effect of pull-out of CNT from PE of the nanocomposite. The Tg of the nanocomposite is influ-
matrix on the mechanical properties of the nanocomposite and enced by many parameters, such as the degree of polymeriza-
found a decrease in Young’s modulus of nanocomposite with tion [28] and interfacial interaction between CNT and polymer
pull-out displacement. Ozer and Cabuk [20] computed the matrix [14, 29]. Several studies have also reported the effect of
structural parameters, and mechanical and electronic proper- CNT reinforcement in a polymer matrix on the thermo-
ties of SbXI (X = S, Se, Te) crystals using DFT , and found mechanical response and Tg of the CNT-polymer nanocom-
that SbXI crystals are mechanically stable, slightly stiff, elas- posite. For instance, Qi [30] used the density-temperature
tically anisotropic and ductile in nature. Recently, Alian and curve to find the Tg of the CNT-polyimide nanocomposite,
Meguid [19] predicted the elastic moduli of the large model and showed that CNT reinforcement reduces the softening
(i.e., up to ∼9 × 105 atoms) of CNT-reinforced representative effects of temperature on mechanical properties. Wei et al.
volume elements, considering different parameters of CNTs [28] investigated the thermal expansion and diffusion coeffi-
such as different lengths, curvatures, and bundle sizes to sim- cients of characteristics of CNT-PE composites and found that
ulate actual nanocomposites. CNTs reinforcement to the polymer matrix enhances the Tg,
Interfacial interactions between the nanofiller and the ma- diffusion, and thermal expansion coefficients of the composite
trix play a significant role in enhancing the bulk properties of system. Li and Strachan [31] characterized the thermo-
the resulting nanocomposite, but this area is less explored mechanical response of pure thermoset polymer composed
because of the experimental limitations in probing material of epoxy EPON862 and curing agent DETDA, and it was
behavior near interface at the nanoscale. Therefore, atomistic found that mechanical response and Tg of the depend on the
simulations through MD can be applied to obtain nanoscale degree of polymerization.
interfacial information to incorporate subsequently into It is evident from the above literature study that there are
continuum-level modeling for the theoretical prediction of number of experimental and analytical (MD/continuum) re-
the bulk properties of the nanocomposite. There are very search articles on the mechanical characterization of the
few publications [22–27] on the prediction of the mechanical CRPN and the interface, as well as few investigations involv-
behavior of the interface using atomistic-based simulations. ing temperature effects on the properties of the nanocompos-
Zheng et al. [22] studied the influence of chirality on the ite. However, there is no comprehensive study on the
interfacial interaction energy of CNT-reinforced polymethyl temperature-dependent elastic properties of the CNT-
methacrylate (PMMA) and found that CNTs with larger chiral reinforced nanocomposite including its interfacial region.
J Mol Model (2018) 24:178 Page 3 of 11 178

Fig. 1 Atomistic unit cell of


carbon nanotube (CNT)

Further, there is a scarcity of any investigations on the effect of Methods


Tg on the elastic properties of nanocomposite and its interface.
Therefore, the objective of the present study was to explore Atomistic modeling
the elastic modulus of a periodic atomistic model of nanocom-
posite (CNT-reinforced polyethylene) and its interface (be- In the present study, two CNTs—armchair type with chiral vector
tween CNT and PE matrix) for different temperatures using (5, 5) and zigzag type with chiral vector (9, 0)—having nearly
MD simulations. The Tg values of the polymer and the nano- same diameters of 6.78 Å and 7.05 Å, respectively, were selected
composite were also evaluated using density-temperature as reinforcing fillers, and amorphous PE was used as the polymer
curves. The atomistic modeling was done using Material matrix. For simplicity and generic representation of polymer ma-
Studio 8.0 [32], and further MD simulations were performed trix, a single chain of PE, with 500 repeating units of –CH2– in
using Large-Scale Atomic/Molecular Massively Parallel each chain, was chosen as a building block to model the actual
Simulator (LAMMPS) [33]. The present report is organized polymer. Further, a periodic atomistic model of CNT-reinforced
as follows: In Section Atomistic modeling, a general proce- polyethylene (i.e., CRPE) was constructed to avoid the end-effect
dure for modeling an atomistic model of the polymer/ of CNT. The modeling procedure of all atomistic models (i.e.,
nanocomposite is discussed. The force field and its parameters armchair and zigzag CNT-reinforced PE nanocomposites, and
used in the study (i.e., PCFF) are explained in Force field. pure bulk PE matrix) was similar and performed using a commer-
Section Simulation procedure describes the equilibration and cial MD simulation package: Material Studio 8.0. However, for
simulation procedures carried out using LAMMPS. The meth- the sake of brevity, only the modeling details of the armchair-CNT
od to determine elastic modulus of the material system is nanocomposite is discussed in the following paragraph. The poly-
provided in Present Study. In the section Validation Study, a mer consistent force field (PCFF) [34] was used throughout the
validation study is conducted with a view to validate the meth- analysis, which is considered in section Force field.
odology followed for MD simulations in the current work. Initially, a pristine armchair CNT was placed at the center of a
Finally, the results are presented and discussed in Results periodic unit cell (60 Å × 60 Å × 51.65 Å) with its length aligned
and discussion. along the z-direction, as shown in Fig. 1, and, subsequently, PE

Fig. 2a,b Atomistic unit cells. a


CNT-reinforced polyethylene PE chains
(PE) nanocomposite. b Pure PE
matrix
CNT

(a) (b)
178 Page 4 of 11 J Mol Model (2018) 24:178

Fig. 3a–d Schematic


representation of polymer
consistent force field (PCFF)
interaction. a Bond interaction. b
Angle interaction. c Dihedral
interaction. d Non-bonded (a)
interaction (b)

(c) (d)

chains, were packed into the unit cell to form CNT-PE model with however, the non-bonded interaction energy is considered be-
a target density of 0.8 g/cm3, as shown in Fig. 2a. In the process, cause of van der Waals interaction only, which is defined by
the total potential energy of the nanocomposite was minimized the Lennard-Jones (LJ) 9–6 inter-atomic potential, for a po-
using the steepest descent method [35] with energy convergence tential cutoff radius (i.e., rcutoff) of 10 Å and having interatom-
tolerance of 0.001 kcal mol−1. A similar modeling procedure was ic potential parameters given in Table 1. More detailed infor-
followed to prepare the atomic model for the pure PE matrix, as mation and description of the second generation PCFF can be
shown in Fig. 2b. found in the literature [34].

Force field
Simulation procedure
The second generation PCFF was used in the present study to
describe the covalent interactions of atoms in the CNT and the After building the initial configuration of the unit cell, MD simu-
polymer as well as the non-covalent interaction at CNT–PE lations were performed with LAMMPS utilizing the PCFF force
matrix interface. The schematic representations of bonded and field. The potential energy of the CNT-PE model (i.e., unit cell)
non-bonded interactions of PCFF are shown in Fig. 3. Based was minimized using the conjugate gradient method, before an-
on the PCFF force field, the total potential energy (i.e., Epot) of nealing it to 600 K for 100 ps and subsequently quenching to
an atomistic system includes different components, as given in 100 K in 200 ps by following the NVT ensemble using the
the following equation: Nose-Hoover thermostat. Thereafter, the material system is equil-
ibrated under NPT ensemble for 200 ps at a temperature of 100 K
E pot ðr1 ; r2 …rn Þ ¼ Ebond þ E angle þ Edihedral þ E improper and pressure of 1 atm in all directions, so that the volume and
þ E non−bonded: ð1Þ potential energy of the material become stable and relaxed. The
final equilibrated density of the polymer matrix relaxed around the
value of 0.92 g cm−3, which is lower than the experimental value
where the bonded interaction energy consists of: Ebond, of 0.95 g cm−3 [36]. The difference is because MD simulations
Eangle, Edihedral, Eimproper terms, and the non-bonded interac- generally consider material free from defects, and the veracity of
tion energy (Enon−bonded) in PCFF includes van der Waals the parameterization in the force field. A similar procedure was
(vdW) and electrostatic interactions. In the present work, followed to generate the equilibrated models for the temperatures
(i.e., quenched temperature) of 300 K, 400 K, and 500 K.
Table 1 Non-bonded Lennard-Jones (LJ) 9–6 interatomic potential After the equilibration procedure, MD simulations
parameters were performed at NVT ensemble using the set-up
LJ 9–6 interactiona ε (kcal mol−1) σ (Å) shown in Fig. 4 under uniaxial deformation. While
keeping the lower group of the nanocomposite fixed,
ðC−CÞsp2 0.064 4.010 the upper group is subjected to a constant velocity of
H−H 0.020 2.995 1 × 10−4 Å fs−1, also shown in Fig. 4. In order to get
ðC−CÞsp3 0.054 4.010 the elastic properties of the nanocomposite from the
a
stress-strain curves obtained during uniaxial deforma-
C−C and H–H represent interactions between carbon and hydrogen
tion, the stress (virial) was calculated using the follow-
atoms, respectively, and the subscripts sp2 and sp3 specify the type of
hybridization ing relation [37, 38]:
J Mol Model (2018) 24:178 Page 5 of 11 178

Fig. 4 Molecular dynamics (MD) Constant Velocity


simulation set-up for the
nanocomposite under uniaxial
deformation
Upper group of atoms moved
rigidly during simulation

Middle group of atoms are


free to move in any direction
during simulation

Lower group of atoms are


held fixed during simulation

nanocomposite by following the procedure discussed in


∑N mK vKi vKj ∑K f Kj rKi
N
σij ¼ K þ ; ð2Þ Atomistic modeling. Two CNTs—armchair type with chiral
V V vector (5, 5) and zigzag type with chiral vector (9, 0)—having
where V is the volume of the CNT-PE model, N is a diameters of 6.78 Å and 7.05 Å, respectively, were selected as
number of atoms in the model, vKi is the ith component reinforcing fillers to make the nanocomposite. Before
of velocity of Kth atom, rKi is the ith component of the conducting the study, a simulation procedure was performed
position vector of Kth atom, and fKj is jth component of using LAMMPS, described in Simulation procedure, for
force applied on Kth atom. The strain is calculated using equilibration and, subsequently, conducting the uniaxial de-
the following relation [27]: formation, is verified by comparing the result (i.e., elastic
modulus) for CNT-reinforced PE matrix nanocomposite with
ðNo:of stepsÞ  timestep  Velocity
ɛ¼ : ð3Þ that reported by Mahboob and Islam [39]. During the uniaxial
Length of CNT deformation of the nanocomposite, the stresses were calculat-
ed on the atomistic scale using virial stress tensor defined by
Eq. (2). Further, for predicting Young’s modulus of the inter-
Present study facial region, the interfacial interaction energy (i.e., Eint,
a function of applied strain ɛ) between CNT and PE matrix
In the present study, MD simulations were conducted to pre- is calculated after each step of the applied constant velocity
dict the elastic constant of the nanocomposite and its interfa- and a polynomial curve of the second degree is fitted to close-
cial region (between CNT and PE matrix) using second gen- ly interpolate the Eint data points with respect to applied strain.
eration PCFF, as described in the section Force field. Material The Young’s modulus of the interface region at different tem-
Studio 8.0 was used for atomistic modeling of the peratures was then calculated using the obtained expression.

Fig. 5 a Simulation model of


CNT-PE matrix nanocomposite,
reinforced with 9.7% volume
fraction of CNT. b Stress-strain
curve of the nanocomposite

(a) (b)
178 Page 6 of 11 J Mol Model (2018) 24:178

Glassy Glassy
state state

Rubbery Rubbery
state state

(a) (b)
Fig. 6a,b Temperature-density plots. a Bulk PE matrix, b CNT (zigzag and armchair)-reinforced PE nanocomposites

Thereafter, Young’s modulus of the nanocomposite at differ- Results and discussion


ent temperatures, taken in the range of 100 K–500 K, was
obtained. The Tg values of the polymer and nanocomposite In this section, initially, the Tg values of bulk PE matrix and
are also evaluated using density-temperature curves. Finally, CNT-reinforced nanocomposite are evaluated and discussed
for the verification of the results obtained from MD simula- in the section Glass transition temperature. Young’s moduli of
tions, a comparison of results with the continuum-based rule- PE matrix and CNT, interface, and the resulting nanocompos-
of-mixtures (ROM) is made. ite are calculated in Elastic moduli of pure PE matrix and
CNT, Elastic modulus of the interfacial region of the nano-
composite and Elastic modulus of the nanocomposite, respec-
Validation study tively. In addition, the influence of phase transition (from
glassy to a rubbery state) temperature (i.e., Tg) on Young’s
To validate the methodology of the present study, a CNT [of moduli of the polymer matrix, nanocomposite, and its inter-
chirality (6,6)] -polyethylene nanocomposite is modeled to face is also discussed in these sections. Finally, in Comparison
compare its Young’s modulus with that reported by of MD results with continuum-based ROM, the obtained
Mahboob and Islam [39]. The lattice parameters of the unit Young’s moduli of PE matrix, CNT, interface and the resulting
cell (30 Å × 30 Å × 51.65 Å) and volume fraction (9.7%) nanocomposite from MD simulations are verified using
used were the same as in the reference paper [39]. The slope of continuum-based ROM.
the linear elastic region (taken up to 3% strain) of the obtained
stress-strain curve shown in Fig. 5b was used to predict
Glass transition temperature
Young’s modulus of the nanocomposite. Young’s modulus
obtained at 300 K from the simulation procedure followed in
MD simulations provide an effective approach for evaluating
the present paper was 107 GPa, which is in good agreement
the Tg of the bulk polymer matrix and its nanocomposite. To
with the value of 103 GPa reported by Mahboob and Islam
evaluate Tg, the equilibrated structure of the bulk PE matrix/
[39].
nanocomposite at 500 K was slowly cooled down with a rate

Table 3 Initial values of interfacial interaction energy Eint (kcal mol−1)


Table 2 Elastic modulus of PE matrix at different temperatures
of nanocomposite at different temperatures
Temperature (K) Density (g cm−3) Young’s modulus (GPa)
CNT type (chirality) Temperature (K)
100 0.92 2.05
100 300 400 500
300 0.85 1.20
400 0.77 0.89 Zigzag (9,0) −353.95 −323.49 −259.30 −230.10
500 0.72 0.72 Armchair (5, 5) −366.77 −335.60 −269.15 −244.92
J Mol Model (2018) 24:178 Page 7 of 11 178

matrix. Hence, the results obtained are consistent with the


literature. The influence of Tg, evaluated in this section, on
the Young’s moduli of the polymer matrix, nanocomposite,
and its interface is discussed in detail in the following sections.

Interfacial region
Elastic moduli of pure PE matrix and CNT

To evaluate the elastic modulus of the PE matrix, the PE matrix


was equilibrated by following the procedure discussed in
Simulation procedure at different temperatures, and relaxed to
obtain mass density in the range of 0.92–0.72 g cm−3. The ob-
tained values of elastic modulus and density of the PE matrix at
Fig. 7 An atomistic model of the nanocomposite featuring interfacial different temperatures are given in Table 2. The values of elastic
region modulus of PE matrix evaluated at 100 K and 300 K (i.e.,
2.05 GPa and 1.20 GPa, respectively) are found to be in good
of 0.5 K ps−1 from 500 K to 50 K at NPT ensemble. The correspondence with the value of 2 GPa reported by Awasthi et
variation of density is monitored over the wide range of tem- al. [44] at 100 K and the value of 1.22 GPa by Al-ostaz et al. [13]
perature values, and is plotted in the form of a density- at 300 K. An expected observation made from the Table 2 is that
temperature plot in Fig. 6a,b along with estimated error bars, the highest and lowest values of Young’s modulus for PE matrix
obtained from the density-temperature data using segmented are found at the highest and lowest temperatures, respectively.
regression (R2 > 0.99). The Tg values of the bulk PE matrix Alternatively, it can be stated that an increase in temperature
and the nanocomposite are estimated at ~295 K and ~320 K, would result in softening of the matrix material. Further, a sig-
respectively, obtained at the intersection points of linear fits to nificant drop in elastic modulus of the PE matrix was noted when
the initial (i.e., glassy-state) and final (i.e., rubbery-state) re- the temperature changes from 100 K to 300 K and, this indicates
gions of density-temperature data points, as shown in Fig. the phase transition of PE matrix from a glassy state to rubbery
6a,b. It is also important to mention that the Tg of the nano- state, as shown in Fig. 6a.
composite is found to be same for both armchair- and zigzag- The elastic properties of the two CNTs—armchair type
CNT reinforcements. As reported in several research studies with chiral vector (5, 5) and zigzag type with chiral vector
[28, 40–43], the Tg of long-chained PE falls in the range of (9, 0)—having nearly the same diameters at 6.78 Å and 7.05
250 K–350 K, which depends on the molecular weights and Å, respectively, were also determined using MD simulations
degrees of polymerization in the PE matrix. Wei et al. [28] also at a temperature of 300 K. The values of Young’s modulus of
found that the CNT reinforcement in PE matrix enhances the the armchair- and zigzag-CNTs were found to be nearly same
Tg of the nanocomposite in comparison to the bulk polymer (i.e., 1056.95 GPa) calculated at 300 K.

(a) (b)
Fig. 8 Variations in interfacial interaction energy at different temperatures with respect to applied strain. a Armchair-CNT nanocomposite, b zigzag-
CNT nanocomposite
178 Page 8 of 11 J Mol Model (2018) 24:178

energy for an armchair-CNT nanocomposites than zigzag-


CNT nanocomposites is found to be in good concurrence with
results reported by Zheng et al. [22].
Thereafter, the interfacial interaction energy Eint corre-
sponding to the applied strain ε (up to 5%) is calculated after
each step of applied constant velocity at different tempera-
tures. Based on the variation of Eint with respect to applied
strain ɛ, a polynomial curve of second-degree is fitted to close-
ly interpolate the data points (Eint, ɛ) at various temperatures.
The corresponding Eint–ɛ curves are shown in Fig. 8a,b for
armchair- and zigzag-CNT nanocomposites, respectively. The
associated error estimates, for the fitted curves, are obtained
from the Eint–ɛ data using regression analysis, where the
values of the coefficient of determination (R2) for different
curves are as follows: R2 > 0.99 for 100 K and 300 K, R2 >
0.97 for 400 K, and R2 > 0.96 for 500 K.
Thereafter, the interface Young’s modulus is calculated
using the following relation:
 
Fig. 9 Radial distribution function (RDF) of PE atoms of the polymer 1 ∂2 E int ðεÞ
Y int ¼ ; ð5Þ
matrix around the carbon atoms of the embedded CNT at different V int ∂ε2
temperatures
where, Eint(ε) represents the interfacial interaction energy as a
Elastic modulus of the interfacial region function of strain, and Vint is the volume of the interface region
of the nanocomposite (between CNT and PE matrix), which is obtained using the
following equation.
The interfacial interaction energy between the CNT and PE h i
matrix is governed by the non-bonded interaction energy (i.e., V int ¼ πð rCNT þ t int Þ2 −πr2CNT L; ð6Þ
here EvdW interaction energy only). The interfacial interaction
energy is calculated as the difference between the total poten- where L is the length of the CNT-PE model along the axis of
tial energy of nanocomposite (ENC) and the sum of potential the CNT, rCNT is the radius of the CNT, and tint represents the
energies of the CNT (ECNT) and polymer matrix (Em), as giv- thickness of the interfacial region. The thickness of the inter-
en in the following relation: facial region (also called as vdW separation distance) is ob-
tained from the radial distribution function (RDF) curve. The
E int ¼ E NC −E m −E CNT : ð4Þ
variation of RDF for PE atoms around the CNT in the periodic
The initial values of interfacial interaction energy between unit cell is plotted using VMD [44] in Fig. 9. It can be ob-
CNT and PE matrix before the uniaxial deformation of the served from the Fig. 9 that the RDF of PE atoms is zero until
nanocomposite are represented in Table 3, at different temper- the radial distance of ~2.7 Å from CNT. This value indicates
atures. The negative value of interaction energy shows the the approximate thickness of the interface region between
existence of attractive interactions between CNT and PE ma- CNT and PE matrix.
trix (Fig. 7). It can be seen from the Table 3 that, irrespective Finally, the values of Young’s modulus calculated at differ-
of temperature, armchair-CNT nanocomposites have more in- ent temperatures for the interfacial region of the nanocompos-
teraction energy than zigzag-CNT nanocomposites; moreover, ite are given in Table 4. It can be noted from Table 4 that the
an increase in temperature results in a decrease of Eint for both highest value of Young’s modulus of the interface is obtained
the nanocomposites. The above finding of higher interaction as 28.02 GPa at a temperature of 100 K for armchair-CNT
nanocomposite, whereas the lowest value of 10.11 GPa is

Table 4 Elastic modulus (in GPa)


of nanocomposite and its CNT (chirality) Interface Nanocomposite
interface at different temperatures
100 K 300 K 400 K 500 K 100 K 300 K 400 K 500 K

Zigzag (9, 0) 25.02 21.94 15.20 10.11 32.12 29.59 21.01 17.33
Armchair (5, 5) 28.02 25.57 17.16 11.75 34.94 32.95 23.96 21.00
J Mol Model (2018) 24:178 Page 9 of 11 178

(a) (b)
Fig. 10 Stress-strain curves for a armchair, and b zigzag nanocomposite at different temperatures

evaluated for zigzag-CNT nanocomposite at a temperature of nanocomposite shows higher value than zigzag-CNT nano-
500 K. It is also important to highlight from Table 4 that, composite at all temperatures, and this observation found
because of the transition of the state of nanocomposite from good in concurrence with the result presented by Zuberi and
glassy to rubbery around Tg (i.e., 320 K), a significant drop in Esat [45] that the larger chirality angle of CNT reinforcement
Young’s modulus of the interface is also observed between yields greater Young’s modulus of the nanocomposite.
300 K and 400 K. Further, it can be noted from Fig. 10a,b that the stress-strain
data is grouped into two distinct categories, one with 100 K
Elastic modulus of the nanocomposite and 300 K, and the other with 400 K and 500 K. The Tg (i.e., ~
320 K) of the nanocomposite, as marked in Fig. 6b, lies be-
Two types of nanocomposites made by reinforcing armchair- tween these two distinct categories.
CNT [with chiral vector (5, 5)] and zigzag-CNT [with chiral
vector (9, 0)] into the pure PE polymer matrix were consid- Comparison of MD results with continuum-based
ered. The slope of the linear elastic region (taken up to 3% ROM
strain) of the obtained stress-strain curve for a particular nano-
composite at different temperatures taken in the range 100 K– In this section, an indirect validation of the above findings of
500 K was used to predict Young’s modulus of the nanocom- the elastic properties of the CNT, polymer, interface, and
posite. The stress-strain curves drawn at different temperatures nanocomposite obtained from the MD simulations done
are shown in Fig. 10a,b for armchair- and zigzag-CNT nano- through continuum-based ROM is presented. ROM is based
composites, respectively, and the corresponding evaluated on the theory of continuum mechanics and assumes that there
values of elastic modulus are given in Table 4. It can be seen is perfectly bonding between the three-phases of the nano-
from the Table 4 that Young’s modulus of the armchair-CNT composite (i.e., CNT, polymer matrix and its interface), as
shown in Fig. 11. By considering the compatibility equations
for strains and equilibrium equations for stresses, Young’s
modulus of a three-phase nanocomposite is expressed mathe-
matically as [26]:
Interface PE matrix Y NC ¼ V CNT
f Y CNT þ V int
f Y int þ V f Y m ;
m
ð7Þ
Solid CNT
where, YNC, YCNT, Yint and Ym represent the elastic moduli of
the nanocomposite, CNT, interface and the polymer matrix,
respectively, and V CNT
f , V int m
f and V f are the volume fractions
of CNT, interfacial region and the polymer matrix in the nano-
Fig. 11 Schematic of three-phase continuum-based model of CNT- composite, respectively. Further, the volume fractions of CNT,
reinforced PE matrix featuring interfacial region interface and polymer must satisfy the following relationship:
178 Page 10 of 11 J Mol Model (2018) 24:178

Table 5 Verification of molecular dynamics (MD) simulation results through rule-of-mixtures (ROM)

CNT (chirality) Volume fraction Young’s modulus (GPa)

V CNT
f V int
f MD simulation ROM

CNT Interface Nanocomposite Nanocomposite

Zigzag (9, 0) 3.00 3.29 1056.95 21.94 29.59 33.49


Armchair (5, 5) 2.89 3.20 1056.95 25.57 32.95 32.44

evaluated. The Tg values of the polymer and the nanocompos-


V CNT
f þ V int
f þ V f ¼ 1:
m
ð8Þ ite were also evaluated using density-temperature curves.
Finally, the results obtained from MD simulations were veri-
fied with ones obtained from continuum-based ROM.
Based on the basic micromechanics assumption (i.e., con-
Based on the above results and discussion, the following
tinuum reinforcement) of ROM, a hollow CNT has to be
important conclusions and findings are highlighted:
modeled first as a solid CNT, and then the volume fraction,
V CNT
f of CNT is computed. The equivalent radius of solid & Young’s modulus of nanocomposite and its interfacial re-
CNT (rsCNT) is computed by considering the cross-sectional gion is sensitive to the temperature variation, i.e., an in-
area of the hollow CNT equal to that of the solid CNT, as crease in temperature results in a decrease of Young’s
reported by Rafiee and Madhavi [27]. The equations for radius modulus of the nanocomposite and its interfacial region.
and volume fraction of solid CNT are derived as: & To utilize the full potential of CNT reinforcement for
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi yielding a high value of Young’s modulus of the resulting
rsCNT ¼ 2rCNT t CNT ; ð9Þ
nanocomposite and its interface at all temperatures, CNTs
πr2sCNT of larger chirality (i.e., armchair-type CNT) are to be used.
V CNT ¼ ; ð10Þ
f
Acs & CNT reinforcement in the PE matrix enhances the Tg of
the material, and in the temperature range containing the
where tCNT and rCNT are the wall-thickness and radius of hol- Tg of the nanocomposite, Young’s modulus of nanocom-
low CNT, respectively, and Acs represents the cross-sectional posite and its interface drop suddenly.
area (50.2 Å × 50.2 Å) of the CNT-PE model (i.e., cross-
sectional area of the final CNT-PE model after NPT
equilibration).
It can be observed from Table 5 that the results obtained
from the continuum ROM for Young’s modulus of the nano- References
composite in the axial direction, evaluated using Eq. (7), are
reasonably close to the MD simulations results obtained at 1. Iijima S (1991) Helical microtubules of graphitic carbon. Nature
300 K. Therefore, the values of Young’s modulus of CNT, 354:56–58
2. Lourie O, Wagner HD (1998) Evaluation of Young’s Modulus of
polymer, interface, and nanocomposite obtained from the carbon nanotubes by micro-Raman spectroscopy. J Mater Res 13:
MD simulations were validated through the continuum- 2418–2422
based ROM. 3. Salvetat J-P, Bonard J-M, Thomson NH (1999) Mechanical prop-
erties of carbon nanotubes. Appl Phys A Mater Sci Process 69:255–
260
4. Yakobson BI, Avouris P (2001) Mechanical properties of carbon
Conclusions nanotubes. In: Dresselhaus MS, Dresselhaus G, Avouris Ph (eds)
Carbon nanotubes. Springer, Berlin, pp 287–327
MD simulations were conducted in this study to predict the 5. Allaoui A, Bai S, Cheng HM, Bai JB (2002) Mechanical and elec-
trical properties of a MWNT/epoxy composite. Compos Sci
elastic properties of a periodic atomistic model of CNT- Technol 62:1993–1998
reinforced PE nanocomposite and its interfacial region (be- 6. Ebbesen TW, Lezec HJ, Hiura H et al (1996) Electrical conductivity
tween CNT and PE matrix) using second generation PCFF. of individual carbon nanotubes. Nature 382:54–56
The two CNTs—armchair type with chiral vector (5, 5) and 7. Qian D, Dickey EC, Andrews R, Rantell T (2000) Load transfer and
zigzag type with chiral vector (9, 0)—were selected as rein- deformation mechanisms in carbon nanotube-polystyrene compos-
ites. Appl Phys Lett 76:2868–2870
forcing fillers for the nanocomposite. Further, the influence of 8. Thostenson ET, Chou T-W (2002) Aligned multi-walled carbon
temperature, taken in the range of 100 K– 500 K, on the elastic nanotube-reinforced composites: processing and mechanical char-
modulus of the interfacial region and the nanocomposites was acterization. J Phys D Appl Phys 35:L77
J Mol Model (2018) 24:178 Page 11 of 11 178

9. Cadek M, Coleman JN, Barron V et al (2002) Morphological and 26. Srivastava A, Kumar D (2017) A continuum model to study inter-
mechanical properties of carbon-nanotube-reinforced semicrystal- phase effects on elastic properties of CNT/GS-nanocomposite.
line and amorphous polymer composites. Appl Phys Lett 81:5123– Mater Res Express 4:25036
5125 27. Rafiee R, Mahdavi M (2016) Characterizing nanotube-polymer in-
10. Gijny FH, Wichmann MHG, Fiedler B, Schulte K (2005) Influence teraction using molecular dynamics simulation. Comput Mater Sci
of different carbon nanotubes on the mechanical properties of ep- 112:356–363
oxy matrix composites—a comparative study. Compos Sci Technol 28. Wei C, Srivastava D, Cho K (2002) Thermal expansion and diffu-
65:2300–2313 sion coefficients of carbon nanotube-polymer composites. Nano
11. Moniruzzaman M, Winey KI (2006) Polymer nanocomposites con- Lett 2:647–650
taining carbon nanotubes. Macromolecules 39:5194–5205 29. Velasco-Santos C, Martínez-Hernández AL, Fisher FT et al (2003)
12. Han Y, Elliott J (2007) Molecular dynamics simulations of the Improvement of thermal and mechanical properties of carbon nano-
elastic properties of polymer/carbon nanotube composites. tube composites through chemical functionalization. Chem Mater
Comput Mater Sci 39:315–323 15:4470–4475
13. Al-Ostaz A, Pal G, Mantena PR, Cheng A (2008) Molecular dy- 30. Qi D, Hinkley J, He G (2005) Molecular dynamics simulation of
namics simulation of SWCNT-polymer nanocomposite and its con- thermal and mechanical properties of polyimide–carbon-nanotube
stituents. J Mater Sci 43:164–173 composites. Model Simul Mater Sci Eng 13:493–507
14. Khare KS, Khabaz F, Khare R (2014) Effect of carbon nanotube 31. Li C, Strachan A (2011) Molecular dynamics predictions of thermal
functionalization on mechanical and thermal properties of cross- and mechanical properties of thermoset polymer EPON862/
linked epoxy-carbon nanotube nanocomposites: role of strengthen- DETDA. Polymer 52:2920–2928
ing the interfacial interactions. ACS Appl Mater Interfaces 6:6098– 32. Material Studio 8.0. Accelrys Inc, San Diego
6110 33. Plimpton S (1995) Fast parallel algorithms for short-range molecu-
15. Mandal A, Singh SP, Prasad R (2016) Dynamic mechanical char- lar dynamics. J Comput Phys 117:1–19
acterization of CNT–PP nanocomposites. J Mol Model 22:1–7 34. Sun H, Mumby SJ, Maple JR, Hagler AT (1994) An ab initio
CFF93 all-atom force field for polycarbonates. J Am Chem Soc
16. Sharma S, Chandra R, Kumar P, Kumar N (2016) Molecular dy-
116:2978–2987
namics simulation of functionalized SWCNT–polymer composites.
35. Fletcher R, Powell MJD (1963) A rapidly convergent descent meth-
J Compos Mater 21998316628973
od for minimization. Comput J 6:163–168
17. Rouhi S, Alizadeh Y, Ansari R (2016) On the elastic properties of
36. Peacock A (2000) Handbook of polyethylene: structures: proper-
single-walled carbon nanotubes/poly (ethylene oxide) nanocom-
ties, and applications. CRC, Boca Raton
posites using molecular dynamics simulations. J Mol Model 22:41
37. Nayebi P, Zaminpayma E (2016) A molecular dynamic simulation
18. Lv Q, Wang Z, Chen S et al (2017) Effects of single adatom and
study of mechanical properties of graphene–polythiophene com-
Stone-Wales defects on the elastic properties of carbon nanotube/
posite with Reax force field. Phys Lett A 380:628–633
polypropylene composites: a molecular simulation study. Int J
38. Frankland SJV, Harik VM, Odegard GM et al (2003) The stress-
Mech Sci 131–132:527–534
strain behavior of polymer-nanotube composites from molecular
19. Alian AR, Meguid SA (2018) Large-scale atomistic simulations of dynamics simulation. Compos Sci Technol 63:1655–1661
CNT-reinforced thermoplastic polymers. Compos Struct 191:221– 39. Mahboob M, Zahabul Islam M (2013) Molecular dynamics simu-
230 lations of defective CNT-polyethylene composite systems. Comput
20. Ozer T, Cabuk S (2018) First-principles study of the structural, Mater Sci 79:223–229
elastic and electronic properties of SbXI (X= S, se, Te) crystals. J 40. Hendra PJ, Jobic HP, Holland-Moritz K (1975) Low temperature
Mol Model 24:66 crystallization in polyethylene and the value of Tg J Polym Sci Part
21. Chawla R, Sharma S (2017) Molecular dynamics simulation of C Polym Lett 13:365–368
carbon nanotube pull-out from polyethylene matrix. Compos Sci 41 Lam R, Geil PH (1981) Amorphous linear polyethylene: annealing
Technol 144:169–177 effects J Macromol Sci Part B Phys 20:37–58
22. Zheng Q, Xue Q, Yan K et al (2008) Influence of chirality on the 42 Koyama A, Yamamoto T, Fukao K, Miyamoto Y (2001) Molecular
interfacial bonding characteristics of carbon nanotube polymer dynamics studies on local ordering in amorphous polyethylene. J
composites. J Appl Phys 103:44302 Chem Phys 115:560–566
23. Yang S, Yu S, Kyoung W et al (2012) Multiscale modeling of size- 43. Brown D, Clarke JHR, Okuda M, Yamazaki T (1994) The prepa-
dependent elastic properties of carbon nanotube/polymer nanocom- ration of polymer melt samples for computer simulation studies. J
posites with interfacial imperfections. Polymer 53:623–633 Chem Phys 100:6011–6018
24. Zhang Y, Zhuang X, Muthu J et al (2014) Composites: part B load 44. Awasthi AP (1989) An atomistic study of the mechanical behavior
transfer of graphene / carbon nanotube / polyethylene hybrid nano- of carbon nanotubes and nanocomposite interfaces. J Chem Inf
composite by molecular dynamics simulation. Compos Part B 63: Model 53:160
27–33 45. Zuberi MJS, Esat V (2015) Investigating the mechanical properties
25. Arash B, Wang Q, Varadan VK (2014) Mechanical properties of of single walled carbon nanotube reinforced epoxy composite
carbon nanotube/polymer composites. Sci Rep 4:6479 through finite element modelling. Compos Part B 71:1–9

You might also like