Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Polymer 70 (2015) 149e160

Contents lists available at ScienceDirect

Polymer
journal homepage: www.elsevier.com/locate/polymer

Multiscale modeling of carbon nanotube epoxy composites


A.R. Alian, S.I. Kundalwal, S.A. Meguid*
Mechanics and Aerospace Design Laboratory, Department of Mechanical and Industrial Engineering, University of Toronto, Toronto, Ontario, M5S 3G8,
Canada

a r t i c l e i n f o a b s t r a c t

Article history: In this article, a multiscale modeling technique is developed to determine the effective elastic moduli of
Received 25 February 2015 CNT-reinforced epoxy composites containing either well-dispersed or agglomerated carbon nanotubes
Received in revised form (CNTs). Two aspects of the work are accordingly examined. In the first, molecular dynamics simulations
1 June 2015
are carried out to determine the atomic-level elastic properties of a representative volume element (RVE)
Accepted 7 June 2015
Available online 20 June 2015
comprised of either epoxy polymer or transversely isotropic CNT-epoxy composite. To study the effect of
agglomeration of CNTs on the bulk elastic properties of the nanocomposite, CNT bundles of different
sizes were considered. A constant-strain energy minimization method is used to determine the elastic
Keywords:
Multiscale modeling
coefficients of the RVEs. In the second, the Mori-Tanaka method is used to scale up the properties of the
Molecular dynamics atomic structure to the microscale level, and the outcome is used to investigate the effect of orientations
Micromechanics and agglomeration of CNTs on the bulk elastic properties of the nanocomposite. Our results reveal that as
Nanocomposites the number of CNTs in the bundle increases, the effective elastic properties of the nanocomposite
Elastic properties decrease at the same CNT volume fraction.
© 2015 Elsevier Ltd. All rights reserved.

1. Introduction and optical properties. Gou et al. [9] investigated the interfacial
bonding of CNT-epoxy composites using a combination of compu-
Owing to their remarkable mechanical and physical properties, tational and experimental methods. The uniform dispersion and
CNTs [1] have emerged as one of the most promising re- good interfacial bonding of CNTs in the epoxy resin resulted in a
inforcements to tailor the properties of polymer based nano- 250e300% increase in the storage modulus with the addition of
composites [2e5]. Due to these exceptional properties, it is 20e30 wt% of CNTs. Li et al. [10] studied the mechanical properties
believed that few weight percentages of CNTs can significantly of CNTereinforced epoxy composite. At 5 wt% of CNTs, Young's
improve the mechanical properties of CNT-based composites. A modulus of the nanocomposite increased from 4 to 7 GPa. Meguid
significant number of experimental and numerical studies have and Sun [11] showed that the homogeneous dispersion of CNTs in
been conducted to study the effective elastic properties of CNT- the epoxy matrix can improve the tensile and shear strength of the
reinforced polymer composites. An earlier attempt was made by resulting synthesized nanocomposite. However, at higher CNT
Schadler et al. [6] to measure the mechanical properties of CNT- concentrations the mechanical properties of the nanocomposite
reinforced polymer composite. They reported that Young's were found to deteriorate due to the formation of CNT agglomer-
modulus in tension and compression can be improved by as much ates, which act as stress concentrators. Qu et al. [12] fabricated CNT
as 20% and 45% at 5 wt% of CNTs. Allaoui et al. [7] achieved sig- composite films with a specifically designed amine-terminated
nificant enhancement in the mechanical properties of CNTerein- polyimide and obtained high quality, optically transparent, and
forced epoxy composite. In their study, Young's modulus and the homogeneous films.
yield strength were doubled for the nanocomposite with 1 and 4 wt Additionally, molecular dynamics (MD) simulations have also
% of multi-walled CNTs, respectively, compared to the pure epoxy been used to study the mechanical properties of CNT-reinforced
matrix. Park et al. [8] synthesized a polyimide composite reinforced composites. For example, Frankland et al. [13] used MD simula-
with CNTs, and reported improved mechanical, thermal, electrical tions to calculate the longitudinal and transverse Young's moduli of
polymer nanocomposite reinforced by long and short CNTs. Griebel
and Hamaekers [14] estimated the elastic moduli of single walled
carbon nanotube (SWCNT)-PE composites using MD simulations.
* Corresponding author.
E-mail address: meguid@mie.utoronto.ca (S.A. Meguid). Their MD results show an excellent agreement with the results

http://dx.doi.org/10.1016/j.polymer.2015.06.004
0032-3861/© 2015 Elsevier Ltd. All rights reserved.
150 A.R. Alian et al. / Polymer 70 (2015) 149e160

predicted by rule-of-mixtures. Al-Ostaz et al. [15] carried out MD polyethylene, polypropylene and PMMA as a matrix phase
simulations to estimate the elastic properties of SWCNT, interfacial [13,14,17]. On the other hand, MD studies utilizing thermoset
bonding, polyethylene matrix and composites with aligned and polymer as a matrix phase are limited in the literature. To the best
randomly distributed SWCNTs. Grujicic et al. [16] studied the effect of our knowledge, there has been no multiscale modeling study for
of chemical functionalization on the mechanical properties of investigating the influence of CNT agglomeration on the effective
multi-walled CNT-vinyl ester epoxy composites using MD simula- elastic properties of CNT-reinforced epoxy composites. It is there-
tions. Introduction of covalent bonds between the CNT and the fore the objective of this paper to develop a multiscale model to
polymer results in significant improvements in the transverse determine the bulk elastic properties of CNT-reinforced epoxy
elastic properties of the nanocomposite. Yang et al. [17] developed a composite. In the current study, CNT bundles are considered as a
multiscale model considering the CNT size and weakened bonding special case of CNT agglomerates. The proposed multiscale model is
effects of the CNT-polymer interface on the effective elastic stiff- developed in two steps. In the first, MD simulations were used to
ness of CNT-polymer composite by using MD simulations and a determine the transversely isotropic elastic properties of RVEs
modified multi-inclusion micromechanics model. Recently, Khare containing either an individual CNT or a bundle of CNTs embedded
et al. [18] used amido-amine functionalized CNTs that form cova- in epoxy. In the second, the determined elastic moduli of the RVEs
lent bonds with cross-linked epoxy matrices to investigate the role were used to determine the influence of orientations and
of the CNT-matrix interphase in the enhancement of mechanical agglomeration of CNTs on the bulk elastic properties of the nano-
and thermal properties in these nanocomposites. composite using Mori-Tanaka model.
Despite the remarkable elastic properties of CNTs, the reported
results in the literature show a limited enhancement in the effec- 2. Multiscale modeling procedure
tive elastic properties of their nanocomposites. There exist
numerous reasons for this discrepancy; including, poor dispersion In this Section, a multiscale model is developed in two steps. In
of the CNTs, agglomeration and aggregation leading to inferior the first, we use MD simulation to determine the isotropic stiffness
properties, poor interfacial properties resulting in limited stress tensor of the RVE filled with the epoxy material and the trans-
transfer and load sharing. Other parameters that influence the versely isotropic stiffness tensor of the RVEs representing either an
resulting properties of nanocomposite include the quality of CNTs individual CNT or a bundle of CNTs embedded in an epoxy matrix.
used, their purities, geometry and aspect ratios as well as orienta- In the second, considering the nanocomposite RVE as the rein-
tion within the nanocomposite [19]. For instance, micro-Raman forcement and the epoxy polymer as the matrix phase, the effective
spectroscopy was used by Ajayan et al. [20] to measure the local elastic moduli of the nanocomposite are scaled up from the nano-
elastic behavior of individual SWCNT bundles in epoxy-based scale to the microscale using the Mori-Tanaka method. Fig. 1 shows
nanocomposite. They noticed that the efficiency of the stress the steps involved in the hierarchical multiscale model.
transfer and hence the enhancement of the mechanical properties
is lower than expected due to the sliding of the CNTs in the 2.1. MD simulations
agglomerated bundles. More recently, in order to investigate how
the CNT content and arrangement affect the characteristics of the In the present study, MD simulations were conducted to
system, molecular models of unidirectional nanocomposite determine the equivalent elastic properties of the composite con-
comprised of different volume fractions of CNTs were studied by stituents at the atomic scale. MD simulations offer an appropriate
Jiang et al. [21]. and effective means to deal with large systems and relatively longer
Due to the technical difficulties inherent in the preparation and simulation times. All MD simulations runs were conducted with
testing of CNT-reinforced polymer composites, different multiscale large-scale atomic/molecular massively parallel simulator
modeling techniques have been developed over the last decade to (LAMMPS) [29] by using the consistent valence force field (CVFF)
predict the properties of these nanocomposites at the microscale [30]. This force field has been used by other researchers to predict
level. In general, multiscale modeling of CNT-reinforced polymer the mechanical properties of CNTs and epoxy polymers [31e33].
composites is achieved in two consecutive steps. At the nanoscale, Conjugate gradient algorithm was used to minimize the total po-
the atomic structures of the CNT and the surrounding epoxy matrix tential energy of the initial configurations, while velocity
cannot be considered as a continuum medium, and the bulk elastic Verlet algorithm was used to integrate the equations of motion in
properties can no longer be obtained using the traditional contin- all MD simulations. Periodic boundary conditions were imposed on
uum mechanics approaches. Therefore, in the first step the prop- all directions of the MD unit cells. The noncovalent bonded CNT-
erties of an effective fiber representing long or short CNT embedded epoxy nanocomposite system is considered in the current study.
inside a unit cell surrounded by polymer molecules are evaluated We did not consider any functionalization between the embedded
using either atomistic based continuum (ABC) technique or mo- CNTs and the surrounding epoxy structures. Therefore, the in-
lecular dynamics (MD) simulation. In the second step, micro- teractions between the atoms of the embedded CNTs and the sur-
mechanics method or finite element technique can then be used to rounding epoxy are solely resulted from non-bonded interactions.
calculate the bulk effective properties of the composite at the In the simulations, the non-bonded interactions between the atoms
microscale [5,17,22e25]. are represented by van der Waals (vdW) interactions and
The performance of CNT-polymer composites is greatly influ- Coulombic forces. The cut-off distance for the non-bonded inter-
enced by the orientations and agglomeration of CNTs in polymers action was set to 14.0 Å [34]. Determination of the atomic-level
[11,19,20,22,25e28]. Due to the high-surface-to-volume ratio, CNT elastic properties of the pure epoxy and the nanocomposite RVE
agglomerates in the synthesized composite leads to nanocomposite was accomplished by straining the MD unit cells followed by
with very low elastic properties. On the other hand, the alignment constant-strain energy minimization. The averaged stress tensor of
of the dispersed CNTs was found to enhance the elastic modulus the MD unit cell is defined in the form of virial stress [35]; as
along the alignment direction [19]. However, limited work has been follows
reported on MD simulations of CNT bundle-based composites
N  
compared to the researches on well-dispersed CNT-reinforced 1 X mi 2
composites. In addition, most of the MD studies on CNT-based s¼ vi þ Fi ri (1)
V 2
composites utilized thermoplastic polymers, such as i¼1
A.R. Alian et al. / Polymer 70 (2015) 149e160 151

Fig. 1. Modeling steps involved in the developed multiscale model.

where V is the volume of the RVE; vi, mi, ri and Fi are the velocity, were remapped to fit inside the compressed box, then a minimiza-
mass, position and force acting on the ith atom, respectively. tion simulation was performed to relax the coordinates of the atoms.
The system was considered to be optimized once the change in the
2.1.1. Modeling of pure epoxy total potential energy of the system between subsequent steps is less
To model the surrounding matrix, we used a specific two- than 1.0  1010 kcal/mol. The optimized system was then equili-
component epoxy material based on a diglycidyl ether of bisphe- brated at room temperature in the constant temperature and volume
nol A (DGEBA) epoxy resin and triethylene tetramine (TETA) curing canonical (NVT) ensemble over 100 ps by using a time step of 1 fs.
agent, which is typically used in the aerospace industry (see Fig. 2).
During the curing process, the hydrogen atoms in the amine groups 2.1.1.2. Step 2 (density adjustment). The compressed system was
of the hardener (TETA) react with the epoxide groups of the resin equilibrated for another 200 ps in the isothermaleisobaric (NPT)
(DGEBA) forming covalent bonds, which result in a highly cross- ensemble at 300 K and 1 atm to generate an epoxy system with the
linked epoxy structure [36], as depicted in Fig. 3. The resin/curing correct density and to reduce the induced residual stresses due to
agent weight ratio in the epoxy polymer was set to 2:1 in order to the volume reduction. This equilibration step resulted in an equil-
achieve the best elastic properties [25]. The cross-linked polymer ibrated amorphous structure with an average density of 1.0 g/cm3.
structure consisted of 80 DGEBA molecules cross-linked with 40 At the end, the structure is again equilibrated for 200 ps in the NVT
molecules of curing agent TETA (see Fig. 3a). The cross-linked ensemble at 300 K.
structure was utilized to form a 3D structure, as shown in Fig. 3b.
This cross-linked epoxy structure was then used to build the epoxy 2.1.1.3. Step 3 (elastic constants). The simulation box was volu-
system in the subsequent MD simulations for both neat epoxy and metrically strained in both tension and compression to determine
CNT-epoxy composite. The neat epoxy model was generated by the bulk modulus by applying equal strains along all three axes. The
randomly placing 5 cross-linked structures in a cubic simulation bulk modulus was calculated by:
box of size 150 Å  150 Å  150 Å to form a system containing 400
sh
chains of DEGBA and 200 chains of TETA [37,38], as shown in Fig. 3c. K¼ (2)
The total number of atoms in the simulation box is 25300. In the
εv
current work, the volumetric bulk modulus (K) and shear modulus
where εv and sh are the volumetric strain and the averaged hy-
(G) of the cross-linked epoxy were determined by applying the
drostatic stress, respectively. The average shear modulus was
volumetric and three-dimensional shear strains, respectively [38].
determined by applying equal shear strains on the simulation box
The main steps involved in determining the elastic moduli of pure
in xy, xz, and yz planes. The shear modulus was calculated by:
epoxy were as follows:
tij
G¼ ; isj (3)
2.1.1.1. Step 1 (volume reduction). The simulation box was com- gij
pressed gradually through 25 steps from its initial size of
150 Å  150 Å  150 Å to the targeted dimensions of where tij and gij denote the averaged shear stress and shear strain,
61 Å  61 Å  61 Å (see Fig. 3d). At each stage, the atoms coordinates respectively.

Fig. 2. Molecular structures of (a) epoxy resin (DGEBA) and (b) curing agent.
152 A.R. Alian et al. / Polymer 70 (2015) 149e160

Fig. 3. Steps used in the preparation of the polymer matrix (a) cross-linked structure, (b) distorted structure in the 3D unit cell, (c) five 80:40 cross-linked structure randomly placed
in a simulation box of size 150 Å  150 Å  150 Å, and (d) the equilibrated system in a simulation box of size 61 Å  61 Å  61 Å after 25 volume reduction steps.

In all simulations, strain increments of 0.25% were applied along thin epoxy layer at the CNT-epoxy interface [43]. This ultra-thin
a particular direction by uniformly expanding or shearing the layer consists of a highly packed crystalline polymer, which has
simulation box and updating the atoms coordinates to fit within the higher elastic properties than the amorphous bulk polymer
new dimensions. After each strain increment, the MD unit cell was [43e45]. Therefore, in order to obtain the actual CNT-epoxy prop-
equilibrated using the NVT ensemble at 300 K for 10 ps. It may be erties, the size of the RVE must be large enough to incorporate the
noted that the fluctuations in the temperature and potential energy interface layer. Armchair (5, 5) CNTs were used in the present work
profiles are less than 1% when the system reached equilibrium after for both individual and agglomerated CNTs reinforcements. The
about 5 ps [34] and several existing MD studies [5,13,21,34,39,40] cylindrical molecular structure of the CNT is treated as an equiva-
used the 2 pse10 ps time step in their MD simulations to equili- lent solid cylindrical fiber [5,22,46,47] for determining its volume
brate the systems after each strain increment. Then, the stress fraction in the nanocomposite RVE,
tensor is averaged over an interval of 10 ps to reduce the effect of
fluctuations. These steps were repeated again in the subsequent
f n NCNT p ðDCNT þ hvdW Þ2 LCNT
deformation increments. The procedure was stopped when the vCNT y (5)
total strain reached up to 2.5%. Based on the calculated bulk and
4Vcell
shear moduli, Young's modulus (E) and Poisson's ratio (n) were
where DCNT and LCNT denote the respective diameter and length of a
determined as follows:
CNT; hvdW is the vdW equilibrium distance between a CNT and the
9KG 3K  2G surrounding polymer matrix; NCNT is the number of CNTs in the
E¼ and n¼ (4) bundle; fn is a factor based on the shape of the bundle of CNTs; and
3K þ G 2ð3K þ GÞ
Vcell is the volume of the RVE.
The predicted elastic properties of the epoxy using MD simu- In order to determine the thickness of the interface layer, we
lations are summarized in Table 1. These moduli are consistent with performed MD simulations for a system consisting of a CNT sur-
the experimentally measured moduli of a similar epoxy [41,42]. rounded by epoxy structures. We built epoxy structure consists of
80 DEGBA and 40 TETA molecules then these chains were equil-
2.1.2. CNT-epoxy interface layer thickness ibrated. A certain number of equilibrated epoxy chains were then
The structure of the epoxy matrix at the vicinity of the CNT distributed around the embedded CNT. The initial size of the
surface differs from the bulk epoxy due to the formation of an ultra- periodic RVE was 150 Å  150 Å  150 Å. Subsequently, the
A.R. Alian et al. / Polymer 70 (2015) 149e160 153

Table 1
Elastic moduli of the epoxy material.

Young's modulus (GPa) Shear modulus (GPa) Bulk modulus (GPa) Poisson's ratio

MD simulations 3.2 1.1 4.8 0.39


Experimental work [41] 2.9 1.07 3.3 0.35
Experimental work [42] 2.6 0.96 3.0 0.35

volume of the simulation box was gradually reduced to 160


60 Å  60 Å  76 Å and equilibrated using the same steps as

RDF of epoxy atoms (Atoms/nm )


adopted in the pure epoxy case without any further chaining 140

3
between the epoxy chains. Therefore, the developed epoxy layer
near the CNT surface resulted solely from the nonbonded in- 120
teractions. This approach does not address how the thickness of
the interface layer changes with the epoxy chain growth during 100
reaction curing, but we can determine the average interface layer
thickness when the MD system reaches an equilibrium state. 80
Fig. 4 shows the simulation box of the nanocomposite system
prior to and post equilibration. Fig. 5 shows the radial distribution 60
function (RDF) of the epoxy atoms that surround the embedded
CNT after the equilibration. The variation of the RDF along the 40
radial direction represents the change of the epoxy structure in
the vicinity of the embedded CNT. It may be observed from Fig. 5 20
that the RDF of epoxy atoms is zero at the radial distance of
0
0.56 nm and reaches its maximum value of 160 atoms/nm3 at the 0 0.25 0.5 0.75 1 1.25 1.5 1.75 2 2.25
radial distance of 0.77 nm. Then, it starts to fluctuate around an Radial distance from CNT axis (nm)
average value of 110 atoms/nm3. This result indicates that the
value of the vdW equilibrium distance hvdW, is ~2.75 Å and the Fig. 5. RDF of the epoxy atoms around the embedded CNT.
thickness of CNT-epoxy matrix interface layer is ~3.0 Å. The ob-
tained values of the interface thickness and the equilibrium
the bulk elastic properties of their nanocomposites. For such an
separation distance were used to select the appropriate RVE sizes
investigation, three RVEs were constructed to represent an
and in calculating the actual volume fraction of the embedded
epoxy matrix containing a: (i) single CNT, (ii) bundle of three
CNTs in the RVE.
CNTs, and (iii) bundle of seven CNTs. Such RVEs are shown in
Fig. 6. The initial distance between the adjacent CNTs in the
2.1.3. MD simulations of CNT-reinforced epoxy composite bundle was taken to be 3.4 Å (see Fig. 7), which is the intertube
It is difficult to uniformly disperse CNTs in the matrix during separation distance in multi-walled CNTs. The RVEs are assumed
the fabrication process and the situation becomes more chal- to be transversely isotropic with the 3eaxis being the axis of
lenging at high CNT loadings. This is attributed to the fact that symmetry. Therefore, only five independent material constants
CNTs have a tendency to agglomerate and aggregate into bundles are required to fully define the elastic stiffness matrix. The
due to their high surface energy and surface area [48]. Therefore, constitutive relationship of the transversely isotropic RVE is
it is necessary to investigate the influence of CNT bundling on given by:

Fig. 4. The simulation box used to study the interface layer: (a) before volume reduction, and (b) after equilibration.
154 A.R. Alian et al. / Polymer 70 (2015) 149e160

Fig. 6. MD unit cells containing a: (a) single CNT, (b) bundle of three CNTs and (c) bundle of seven CNTs.

in the MD simulations of the RVEs are the same as adopted in the


2 3
C11 C12 C13 0 0 0 case of pure epoxy. The boundary and loading conditions that have
2 3 6 72 3 been applied to the RVE to determine the corresponding five in-
s11 6 C12 C11 C13 0 0 0 7 ε11
6 7 dependent elastic coefficients of the RVE are listed in Table 3.
6 s22 7 6 76 ε22 7
6 7 6 C13 C13 C33 0 0 0 76 7 A series of MD simulations were carried out to determine the
6 s33 7 6 76 ε33 7
6 7 6 76 7
6 s23 7 ¼ 6 0 0 0 C44 0 0 76 ε23 7 elastic properties of transversely isotropic RVEs reinforced with
6 7 6 76 7
4 s13 5 6 74 ε13 5 either an individual CNT or bundles of CNTs. Fig. 8 demonstrates
6 0 0 0 0 C44 0 7
s12 6 7 ε12 close views of the equilibrated RVEs. Table 4 summarizes the
4 5
ðC11  C12 Þ outcome of the MD simulations. It may be observed from the re-
0 0 0 0 0
2 sults that the elastic properties of the RVEs are significantly higher
(6) than those of the neat epoxy. It is also clear from the results that the
CNT agglomeration reduces the reinforcing effect of the embedded
where sij and εij are the respective stress and strain components CNTs, which eventually degrades the bulk elastic moduli of the
with (i, j ¼ 1, 2, 3, 4, 5, 6) and Cij represents the elastic coefficients of nanocomposites. The axial elastic coefficients (C33) of the RVEs
the RVE. A series of MD simulations were carried out to determine containing bundles of three and seven CNTs decreased by some 21%
the elastic moduli of the RVEs. The simulation box in each case was and 38.5%, respectively, as compared with the RVE containing an
constructed by randomly placing the cross-linked epoxy structures individual CNT. The CNT agglomeration is also found to affects the
around the CNT bundle. The size of the RVE in each case was transverse elastic coefficients of the RVEs. For example, the trans-
adjusted in such a way that the CNT volume fraction in it remains verse elastic coefficient (C11) of the RVEs containing bundles of
constant at 6.5%, irrespective of the number of CNTs, as shown in three and seven CNTs decreased by 11.0% and 22.9%, respectively,
Fig. 6. The details of the three RVEs are summarized in Table 2. To compared with the RVE containing an individual CNT. Each RVE will
determine the five elastic constants, the RVEs were subjected to five be used as an effective fiber in the micromechanical model to
different tests: longitudinal tension, transverse tension, in-plane calculate the effective elastic moduli of the nanocomposite at the
tension, in-plane shear and out of-plane shear. The steps involved microscale level (see Fig. 1).

Fig. 7. Geometrical configurations of the bundle of CNTs: (a) bundle of three CNTs and (b) bundle of seven CNTs.
A.R. Alian et al. / Polymer 70 (2015) 149e160 155

Table 2
Parameters used in the RVE.

Parameter Single CNT Bundle of three CNTs Bundle of seven CNTs

CNT type (5, 5) (5, 5) (5, 5)


Number of CNTs 1 3 7
Length of a CNT (Å) 73 73 76
CNT volume fraction 6.5% 6.5% 6.5%
RVE dimensions (Å3) 31  31  76 55  55  76 88  88  76
Total number of DEGBA molecules 120 320 960
Total number of TETA molecules 60 160 480
Total number of atoms 7836 21116 62028

Table 3 in which the mechanical strain concentration tensor [A] is given by


Effective elastic coefficients of the RVEs and corresponding displacement fields.

Elastic coefficients Applied strains Applied displacement h h i h i i1


1
C11 ε11 ¼ e u1 ¼ ex1 ½A ¼ ½I þ SRVE ð½Cm Þ CRVE  ½Cm  (8)
C33 ε33 ¼ e u3 ¼ ex3
C44 ε23 ¼ e/2 u2 ¼ 2e x3 , u3 ¼ 2e x2
C66 ε12 ¼ e/2 u1 ¼ 2e x2 , u2 ¼ 2e x1
where [Cm] and [CRVE] are the stiffness tensors of the epoxy matrix
K12 ¼ C11 þC
2
12
ε11 ¼ ε22 u1 ¼ ex1, u2 ¼ ex2
and the RVE, respectively; [I] is an identity matrix; vm and vRVE
represent the volume fractions of the epoxy matrix and the RVE,
respectively; and [SRVE] indicate the Eshelby tensor. The specific
2.2. Micromechanics model
form of the Eshelby tensor for the RVE inclusion given by Qiu and
Weng [51] is utilized herein.
In this section, the elastic moduli of the pure epoxy and the
It may be noted that the elastic coefficient matrix [C] directly
nanocomposite RVE obtained from the MD simulations are used as
provides the values of the effective elastic properties of the nano-
an input to the micromechanical model in order to determine the
composite, where the RVE is aligned with the 3eaxis. In case of
bulk elastic properties of the nanocomposite. With this process
random orientations of CNTs, the terms enclosed with angle
firmly established in several studies [5,17,22e25], the Mori-Tanaka
brackets in Eq. (7) represent the average value of the term over all
model [49,50] can be developed by utilizing the transversely
orientations defined by transformation from the local coordinate
isotropic elastic properties of the nanocomposite RVE and the
system of the RVE to the global coordinate system. The transformed
isotropic elastic properties of the pure epoxy matrix. In case of two-
mechanical strain concentration tensor for the RVEs with respect to
phase composite, where the inhomogeneity is randomly orientated
the global coordinates is given by
in the three-dimensional space, the following relation can be used
to determine the effective stiffness tensor [C] of the
h i  
nanocomposite: ~
A ijkl ¼ tip tjq tkr tls Apqrs (9)
Dh i E 
m
½C ¼ ½C  þ vRVE CRVE  ½Cm  ½A½vm ½I þ vRVE 〈½A〉1
(7) where tij are the direction cosines for the transformation and are
given by

Fig. 8. Snapshots of the transverse cross-sections of the MD unit cells containing a: (a) single CNT, (b) bundle of three CNTs and (c) bundle of seven CNTs.

Table 4
Material properties of the nanocomposite RVE containing either an individual CNT or its bundle.

RVE CNT volume fraction in RVE (vCNT) C11 (GPa) C12 (GPa) C13 (GPa) C33 (GPa) C44 (GPa) C66 (GPa)

Single CNT 6.5% 11.8 7.1 5.6 47.35 3.16 2.35


Bundle of three CNTs 10.5 6.6 4.5 37.4 1.89 1.95
Bundle of seven CNTs 9.1 5.9 3.5 29.2 1.48 1.6
156 A.R. Alian et al. / Polymer 70 (2015) 149e160

t11 ¼ cosf cosj  sinf cosg sinj; t12 ¼ sinf cosj þ cosf cosg sinj;
t13 ¼ sinj sing; t21 ¼ cosf sinj  sinf cosg cosj; t22 ¼ sinf sinj þ cosf cosg cosj;
t23 ¼ sing cosj; t31 ¼ sinf sing; t32 ¼ cosf sing and t33 ¼ cosg

Consequently, the random orientation average of the dilute properties of the representative fiber, comprising a CNT, the sur-
mechanical strain concentration tensor [A] can be determined by rounding epoxy and their interface, they calculated the bulk elastic
using the following equation [52]: properties of the nanocomposite using large-scale hybrid Monte
Carlo finite element simulations. The concept of the representative
Z p Z p Z p=2 h
i
~ ðf; g; jÞ sing dfdgdj fiber is similar to that of the nanocomposite RVE used in the current
A
p 0 0 study. Thus, to validate the present multiscale model, the results of
〈½A〉 ¼ Z p Z p Z p=2 (10)
the developed multiscale model in the current work are compared
sing dfdgdj with those shown in Ref. [25] for CNT-reinforced adhesives con-
p 0 0
taining randomly oriented CNTs, as shown in Fig. 10. The results
where f, g, and j are the Euler angles as shown in Fig. 9. It may be predicted by the present multiscale and the ABC models are in good
noted that the averaged mechanical strain concentration tensors agreement, validating the multiscale model developed in this
given by Eqs. (8) and (10) are used for the cases of aligned and study. The difference in the elastic moduli between those provided
random orientations of CNTs, respectively, in Eq. (7). in Ref. [25] and the current study is attributed to the lower values of
E (1.14 GPa) and G (0.41 GPa) of the epoxy used. In our work, the
respective elastic moduli E and G for the epoxy were 3.2 GPa and
3. Results and discussion
1.1 GPa. In addition, a uniform distribution of the polymer nodal
density (99 nodes/nm3) around the embedded CNT is assumed in
In this Section, the results of the developed multiscale model are
Refs. [25], while the polymer nodal density obtained from the MD
compared with the existing results predicted by the atomistic
simulations is found to vary along the radial direction, as shown in
based continuum (ABC) model developed by Wernik and Meguid
Fig. 5. At this juncture, it is important to note that the present
[25]. Subsequently, the effect of orientation and agglomeration of
multiscale model represents the real nanocomposite structure with
CNTs on the bulk elastic properties of the nanocomposite are
the minimum assumptions compared to that of the ABC model
analyzed and discussed.
which model the atomic structure as an equivalent beam and truss
elements. Thus, it can be inferred from these comparisons that the
3.1. Comparisons with existing ABC results present multiscale model can be reliably applied to determine the
effective elastic properties of nanocomposites.
Recently, Wernik and Meguid [25] developed a novel multiscale
ABC approach to determine the elastic properties of CNT-reinforced 3.2. Effect of orientation and agglomeration of CNTs
epoxy composites. In their model, the governing atomistic consti-
tutive laws were integrated within the continuum framework by Unless otherwise stated, the bulk elastic properties of the
modeling the CNT as a space-frame-like structure where the adja- nanocomposite are relating to the aligned nanocomposite RVEs
cent atomic bonds are considered as a load-bearing beam element. along the 3-axis. The maximum CNT volume fraction (VCNT)
Nodes were used to connect the beam elements to form the CNT considered in the nanocomposite is 5% because CNT concentrations
structure and truss elements were used to represent the vdW in- above this loading are not normally realized [25]. Let us first
teractions at the CNT-polymer interface. After obtaining the elastic

14
E (Present Multiscale Model)
12 E (ABC Model [26])
G (Present Multiscale Model)
G (ABC Model [26])
10
E and G (GPa)

0
0 1 2 3 4 5
CNT Volume Fraction (%)

Fig. 10. Comparison of Young's (E) and shear (G) moduli of the nanocomposite con-
Fig. 9. Relationship between the local coordinates (1, 2, 3) of the RVE and the global taining randomly oriented CNTs predicted by the present multiscale model with those
coordinates (1000, 2000 , 3000 ) of the bulk composite. of ABC model [25].
A.R. Alian et al. / Polymer 70 (2015) 149e160 157

demonstrate the effect of agglomeration of CNTs on the effective 6.5


elastic properties of the aligned CNT-reinforced epoxy composite. Single CNT
Considering the epoxy as the matrix phase and the RVE as the 6 Bundle of 3 CNTs
reinforcement, the effective elastic coefficients of the nano- Bundle of 7 CNTs
composite were determined by following the micromechanical 5.5
modeling approach developed in Section 2.2. The effective elastic
moduli of the nanocomposite are related to the effective elastic

E11 (GPa)
5
stiffness components as follows:
4.5
2C223
E33 ¼ C33  (11)
C12 þ C11
4
 
ðC11  C12 Þ C11 C33 þ C12 C33  2C223 3.5
E11 ¼ E22 ¼ (12)
C11 C33  C223
3
0 1 2 3 4 5
ðC  C12 Þ CNT Volume Fraction (%)
G12 ¼ 11 (13)
2 Fig. 12. Variation of the effective transverse Young's modulus (E11) of the aligned CNT-
reinforced epoxy composite with the CNT volume fraction.
G13 ¼ G23 ¼ C44 (14)

C11 þ C12 CNTs are found to be ~12% and ~23% in comparison to the nano-
K12 ¼ (15) composite reinforced with individual CNTs, respectively.
2
Figs. 13 and 14 show the variations of the effective axial shear
where E33, E11, G23, G12 and K12 are the effective axial Young's, modulus (G23) and the transverse shear modulus (G12) of the
transverse Young's, axial shear, transverse shear and bulk moduli of nanocomposite with the CNT volume fraction, respectively. The
the nanocomposite. CNT bundling (agglomeration) significantly reduced the effective
Figs. 11 and 12 show the variations of the effective axial Young's axial shear modulus of the composites from 2.55 to 1.7 GPa (i.e.,
modulus (E33) and the transverse Young's modulus (E11) of the around ~ 33% reduction in case of bundle of three CNTs) and from
nanocomposite against the CNT volume fraction (VCNT), respec- 2.55 to 1.4 GPa (i.e., around ~ 45% reduction in case of bundle of
tively. Both moduli exhibit similar trends; with increasing CNT seven CNTs), at 5 vol% of CNTs. On the other hand, the agglomer-
volume fraction, both moduli increase significantly. Significant ation of three CNTs has a marginal effect on the transverse shear
enhancements are observed in the values of E33 and E11 compared modulus (G12) while agglomeration of seven CNTs has significant
with the pure epoxy matrix, but it is more prominent in the case of effect on the same and decreases its value up to 26%, at 5 vol% of
the axial Young's modulus (E33). It may also be observed that the CNTs. Fig. 15 demonstrate that the CNT bundling has marginal effect
effect of agglomeration of CNTs on these constants becomes more on the effective bulk moduli (K12) of the nanocomposite. It is worth
pronounced at higher CNT volume fraction (>2%). At 5 vol% of CNTs, noting that the results predicted by our multiscale model as
the nanocomposite containing bundles of three and seven CNTs, depicted in Figs. 11e15 show clearly that the agglomeration of CNTs
the values of E33 are decreased by ~20% and ~37% as compared with significantly affects the elastic moduli and has a marginal influence
the nanocomposite reinforced with individual CNTs, respectively. on the bulk modulus of the nanocomposite. This happens because
For the same CNT loading, the percentage reductions in the values the interfacial properties of the nanocomposites deteriorate with
of E33 of the nanocomposite containing bundles of three and seven increasing bundle size. The presence of the bundle leads to a

35 2.6
Single CNT Single CNT
30 Bundle of 3 CNTs 2.4 Bundle of 3 CNTs
Bundle of 7 CNTs Bundle of 7 CNTs
2.2
25
2
E33 (GPa)

G23 (GPa)

20
1.8
15
1.6
10
1.4

5 1.2

0 1
0 1 2 3 4 5 0 1 2 3 4 5
CNT Volume Fraction (%) CNT Volume Fraction (%)

Fig. 11. Variation of the effective axial Young's modulus (E33) of the aligned CNT- Fig. 13. Variation of the effective axial shear modulus (G23) of the aligned CNT-
reinforced epoxy composite with the CNT volume fraction. reinforced epoxy composite with the CNT volume fraction.
158 A.R. Alian et al. / Polymer 70 (2015) 149e160

2.4
Single CNT
2.2 Bundle of 3 CNTs
Bundle of 7 CNTs
2
G12 (GPa)

1.8

1.6

1.4

1.2

1
0 1 2 3 4 5
CNT Volume Fraction (%)

Fig. 14. Variation of the effective transverse shear modulus (G12) of the aligned CNT-
Fig. 16. Variation of the effective Young's modulus (E) of the nanocomposite con-
reinforced epoxy composite with the CNT volume fraction.
taining randomly oriented CNTs with the CNT volume fraction.

reduction of the interface area between the CNTs and the matrix
nanocomposite improve in comparison to the aligned case. These
(see Fig. 8).
findings are also consistent with the previously reported findings
In the previous set of results (Figs. 11e15), the effect of
[22,25]. It may also be observed that both E and G decrease with the
agglomeration of CNTs on the effective properties of the aligned
increase in the number of CNTs in the bundle, and this effect be-
CNT-reinforced epoxy composite is studied. Practically, the orien-
comes more pronounced at higher volume fractions of CNTs.
tations of the CNT reinforcement in the polymer matrix can vary
over the volume of the synthesized nanocomposite. Therefore,
studying the properties of nanocomposites reinforced with 4. Conclusions
randomly oriented CNTs is of a great importance. For such inves-
tigation, CNTs or their bundles are considered to be randomly In this article, we develop a multiscale model to determine the
dispersed in the epoxy matrix over the volume of the nano- bulk elastic properties of CNT-reinforced epoxy composites. Two
composite. As expected, this case provides the isotropic elastic aspects of the work were examined. First, MD simulations were
properties for the resulting nanocomposite. Figs. 16 and 17 illus- carried out to determine the transversely isotropic properties of
trate the variations of the effective Young's (E) and shear (G) moduli nanocomposite RVEs containing either an individual CNT or CNT
of the nanocomposite with the CNT loading, respectively. These bundles embedded in an epoxy matrix. The isotropic elastic prop-
results clearly demonstrate that the randomly dispersed CNTs erties of the pure epoxy were also obtained from the MD simula-
improve the effective Young's and shear moduli of the nano- tions. Second, the developed RVEs were, in turn, used with
composite over those of the transverse Young's modulus (E11) and analytical micromechanical technique of the Mori-Tanaka type to
the shear moduli (G12 and G23) of the nanocomposite reinforced determine the bulk elastic properties of the nanocomposite. This
with aligned CNTs. This is attributed to the fact that the CNTs are model was then validated by comparing the predicted results with
homogeneously dispersed in the epoxy matrix in the random case those of a hybrid atomistic-based continuum model. The developed
and hence the overall elastic properties of the resulting

8.5
Single CNT
8 Bundle of 3 CNTs
Bundle of 7 CNTs
7.5
K12 (GPa)

6.5

5.5

5
0 1 2 3 4 5
CNT Volume Fraction (%)

Fig. 15. Variation of the effective bulk modulus (K12) of the aligned CNT-reinforced Fig. 17. Variation of the effective shear modulus (E) of the nanocomposite containing
epoxy composite with the CNT volume fraction. randomly oriented CNTs with the CNT volume fraction.
A.R. Alian et al. / Polymer 70 (2015) 149e160 159

model was applied to investigate the influence of the orientation Mech. Eng. 193 (2004) 1773e1788, http://dx.doi.org/10.1016/j.cma.2003.
12.025.
and agglomeration of the dispersed CNTs on the elastic properties
[15] A. Al-Ostaz, G. Pal, P.R. Mantena, A. Cheng, Molecular dynamics simulation of
of CNT-epoxy composites. The following is a summary of our SWCNTepolymer nanocomposite and its constituents, J. Mater. Sci. 43 (2008)
findings: 164e173, http://dx.doi.org/10.1007/s10853-007-2132-6.
[16] M. Grujicic, Y.P. Sun, K.L. Koudel, The effect of covalent functionalization of
carbon nanotube reinforcements on the atomic-level mechanical properties of
1. The presence of CNT agglomerates in the nanocomposite de- poly-vinyl-ester-epoxy, Appl. Surf. Sci. 253 (2007) 3009e3021, http://
grades its effective elastic properties, dx.doi.org/10.1016/j.apsusc.2006.06.050.
2. The effective elastic properties of the nanocomposites decrease [17] S. Yang, S. Yu, W. Kyoung, D.-S. Han, M. Cho, Multiscale modeling of size-
dependent elastic properties of carbon nanotube/polymer nanocomposites
with the increase of the CNT agglomerate size, with interfacial imperfections, Polymer 53 (2012) 623e633, http://dx.doi.org/
3. Randomly oriented CNTs or their bundles have significant in- 10.1016/j.polymer.2011.11.052.
fluence on the elastic properties of the nanocomposite [18] K.S. Khare, F. Khabaz, R. Khare, Effect of carbon nanotube functionalization
on mechanical and thermal properties of cross-linked epoxycarbon
considered, Nanotube nanocomposites: role of strengthening the interfacial interactions,
4. The transverse Young's modulus, the axial shear and the trans- ACS Appl. Mater. Interf. 6 (2014) 6098e6110, http://dx.doi.org/10.1021/
verse shear moduli of nanocomposites reinforced with aligned am405317x.
[19] J.M. Wernik, S.A. Meguid, Recent developments in multifunctional nano-
CNTs are less than those of the randomly dispersed case, and composites using carbon nanotubes, ASME Appl. Mech. Rev. 63 (2010)
5. The multiscale model developed in this study is capable of 050801, http://dx.doi.org/10.1115/1.4003503.
determining the effective elastic properties of any advanced [20] P.M. Ajayan, L.S. Schadler, C. Giannaris, A. Rubio, Single-walled carbon
nanotube-polymer composites: strength and weakness, Adv. Mater. 12 (2000)
nanocomposite containing either aligned or randomly dispersed
750e753, http://dx.doi.org/10.1002/(SICI)1521-4095(200005)12:10<750::
CNTs. AID-ADMA750>3.0.CO;2-6.
[21] Q. Jiang, S.S. Tallury, Y. Qiu, M.A. Pasquinelli, Molecular dynamics simulations
of the effect of the volume fraction on unidirectional polyimideecarbon
Acknowledgments nanotube nanocomposites, Carbon 67 (2014) 440e448, http://dx.doi.org/
10.1016/j.carbon.2013.10.016.
[22] G.M. Odegard, T.S. Gates, K.E. Wise, C. Park, E.J. Siochi, Constitutive modeling
The authors are grateful for the financial support provided by of nanotube-reinforced polymer composites, Compos. Sci. Technol. 63 (2003)
NSERC and the Discovery Accelerator Supplement in support of this 1671e1687, http://dx.doi.org/10.1016/S0266-3538(03)00063-0.
[23] S.A. Meguid, J.M. Wernik, Z.Q. Cheng, Atomistic-based continuum represen-
research.
tation of the effective properties of nano-reinforced epoxies, Int. J. Solids
Struct. 47 (2010) 1723e1736, http://dx.doi.org/10.1016/j.ijsolstr.2010.03.009.
[24] J.M. Wernik, S.A. Meguid, Multiscale modeling of the nonlinear response of
References nanoereinforced polymers, Acta Mech. 217 (2012) 1e16, http://dx.doi.org/
10.1007/s00707-010-0377-7.
[1] S. Iijima, Helical microtubules of graphitic carbon, Nature 354 (1991) 56e58, [25] J.M. Wernik, S.A. Meguid, Multiscale micromechanical modeling of the
http://dx.doi.org/10.1038/354056a0. constitutive response of carbon nanotube ereinforced structural adhesives,
[2] M.M.J. Treacy, T.W. Ebbesen, J.M. Gibson, Exceptionally high Young's modulus Int. J. Solids Struct. 51 (2014) 2575e2589, http://dx.doi.org/10.1016/j.ijsolstr.
observed for individual carbon nanotubes, Nature 381 (1996) 678e680, 2014.03.009.
http://dx.doi.org/10.1038/381678a0. [26] S.H. Park, P.R. Bandaru, Improved mechanical properties of carbon nanotube/
[3] C. Li, T.W. Chou, A structural mechanics approach for the analysis of carbon polymer composites through the use of carboxyl-epoxide functional group
nanotubes, Int. J. Solids Struct. 40 (2003) 2487e2499, http://dx.doi.org/ linkages, Polymer 51 (2010) 5071e5077, http://dx.doi.org/10.1016/
10.1016/S0020-7683(03)00056-8. j.polymer.2010.08.063.
[4] L. Shen, J. Li, Transversely isotropic elastic properties of single-walled carbon [27] F. Puch, C. Hopmann, Morphology and tensile properties of unreinforced and
nanotubes, Phys. Rev. B 69 (2004) 045414, http://dx.doi.org/10.1103/ short carbon fibre reinforced Nylon 6/multiwalled carbon nanotube-com-
PhysRevB.69.045414. posites, Polymer 55 (2014) 3015e3025, http://dx.doi.org/10.1016/
[5] J.L. Tsai, S.H. Tzeng, Y.T. Chiu, Characterizing elastic properties of carbon j.polymer.2014.04.052.
nanotube/polyimide nanocomposites using multi-scale simulation, Compos. [28] S. Gong, Z.H. Zhu, S.A. Meguid, Carbon nanotube agglomeration effect on
Part B Eng. 41 (2010) 106e115, http://dx.doi.org/10.1016/j.compositesb. piezoresistivity of polymer nanocomposites, Polymer 55 (2014) 5488e5499,
2009.06.003. http://dx.doi.org/10.1016/j.polymer.2014.08.054.
[6] L.S. Schadler, S.C. Giannaris, P.M. Ajayan, Load transfer in carbon nanotube [29] S. Plimpton, Fast parallel algorithms for short-range molecular dynamics,
epoxy composites, Appl. Phys. Lett. 73 (1998) 3842e3844, http://dx.doi.org/ J. Comput. Phys. 117 (1995) 1e19, http://dx.doi.org/10.1006/jcph.1995.1039.
10.1063/1.122911. [30] P. Dauber-Osguthorpe, V.A. Roberts, D.J. Osguthorpe, J. Wolff, M. Genest,
[7] A. Allaoui, S. Bai, H.M. Cheng, J.B. Bai, Mechanical and electrical properties of a A.T. Hagler, Structure and energetics of ligand binding to proteins: Escherichia
MWNT/epoxy composite, Compos. Sci. Technol. 62 (2002) 1993e1998, http:// coli dihydrofolate reductase-trimethoprim, a drug-receptor system, Proteins 4
dx.doi.org/10.1016/S0266-3538(02)00129-X. (1998) 31e47, http://dx.doi.org/10.1002/prot.340040106.
[8] C. Park, Z. Ounaies, K.A. Watson, R.E. Crooks, J. Smith Jr., S.E. Lowther, [31] K. Tunvir, A. Kim, S.H. Nahm, The effect of two neighboring defects on the
J.W. Connell, E.J. Siochi, J.S. Harrison, T.L. Clair, Dispersion of single wall mechanical properties of carbon nanotubes, Nanotechnology 19 (2008)
carbon nanotubes by in situ polymerization under sonication, Chem. Phys. 065703, http://dx.doi.org/10.1088/0957-4484/19/6/065703.
Lett. 364 (2002) 303e308, http://dx.doi.org/10.1016/S0009-2614(02)013 [32] C. Li, G.A. Medvedev, E.-W. Lee, J. Kim, J.M. Caruthers, A. Strachan, Molecular
26-X. dynamics simulations and experimental studies of the thermomechanical
[9] J. Gou, B. Minaie, B. Wang, Z. Liang, C. Zhang, Computational and experimental response of an epoxy thermoset, Polymer 53 (2012) 4222e4230, http://
study of interfacial bonding of single-walled nanotube reinforced composites, dx.doi.org/10.1016/j.polymer.2012.07.026.
Comput. Mater. Sci. 31 (2004) 225e236, http://dx.doi.org/10.1016/ [33] A. Kumar, V. Sundararaghavan, A.R. Browning, Study of temperature depen-
j.commatsci.2004.03.002. dence of thermal conductivity in cross-linked epoxies using molecular dy-
[10] X.D. Li, H.S. Gao, W.A. Scrivens, D.L. Fei, X.Y. Xu, M.A. Sutton, A.P. Reynolds, namics simulations with long range interactions, Model. Simul. Mater. Sci.
M.L. Myrick, Nanomechanical characterization of single-walled carbon Eng. 22 (2014) 025013, http://dx.doi.org/10.1088/0965-0393/22/2/025013.
nanotube reinforced epoxy composites, Nanotechnology 15 (2004) [34] S. Haghighatpanah, K. Bolton, Molecular-level computational studies of single
1416e1423, http://dx.doi.org/10.1088/0957-4484/15/11/005. wall carbon nanotubeepolyethylene composites, Comput. Mater. Sci. 69
[11] S.A. Meguid, Y. Sun, On the tensile and shear strength of nano-reinforced (2013) 443e454, http://dx.doi.org/10.1016/j.commatsci.2012.12.012.
composite interfaces, Mater. Des. 25 (2004) 289e296, http://dx.doi.org/ [35] D. Qi, J. Hinkley, G. He, Molecular dynamics simulation of thermal and me-
10.1016/j.matdes.2003.10.018. chanical properties of polyimideecarbon-nanotube composites, Model. Simul.
[12] L. Qu, Y. Lin, D.E. Hill, B. Zhou, W. Wang, X. Sun, A. Kitaygorodskiy, M. Suarez, Mater. Sci. Eng. 13 (2005) 493e507, http://dx.doi.org/10.1088/0965-0393/13/
J.W. Connell, L.F. Allard, Y.-P. Sun, Polyimide functionalized carbon nanotubes: 4/002.
synthesis and dispersion in nanocomposite films, Macromolecules 37 (2004) [36] L. Vignoud, L. David, B. Sixou, G. Vigier, Influence of electron irradiation on the
6055e6060, http://dx.doi.org/10.1021/ma0491006. mobility and on the mechanical properties of DGEBA/TETA epoxy resins,
[13] S.J.V. Frankland, V.M. Harik, G.M. Odegard, D.W. Brenner, T.S. Gates, The Polymer 42 (2001) 4657e4665, http://dx.doi.org/10.1016/S0032-3861(00)
stressestrain behavior of polymerenanotube composites from molecular 00791-6.
dynamics simulation, Compos. Sci. Technol. 63 (2003) 1655e1661, http:// [37] http://nanoengineer-1.software.informer.com/.
dx.doi.org/10.1016/S0266-3538(03)00059-9. [38] A. Bandyopadhyay, P.K. Valavala, T.C. Clancy, K.E. Wise, G.M. Odegard, Mo-
[14] M. Griebel, J. Hamaekers, Molecular dynamics simulations of the elastic lecular modeling of crosslinked epoxy polymers: the effect of crosslink
moduli of polymerecarbon nanotube composites, Comput. Methods Appl.
160 A.R. Alian et al. / Polymer 70 (2015) 149e160

density on thermomechanical properties, Polymer 52 (2011) 2445e2452, [45] J.N. Coleman, U. Khan, W.J. Blau, Y.K. Gun'ko, Small but strong: A review of the
http://dx.doi.org/10.1016/j.polymer.2011.03.052. mechanical properties of carbon nanotubeepolymer composites, Carbon 44
[39] S.J.V. Frankland, V.M. Harikb, G.M. Odegard, D.W. Brenner, T.S. Gates, The (2006) 1624e1652, http://dx.doi.org/10.1016/j.carbon.2006.02.038.
stressestrain behavior of polymerenanotube composites from molecular [46] E.T. Thostenson, T.W. Chou, On the elastic properties of carbon nanotube
dynamics simulation, Compos. Sci. Technol. 63 (2003) 1655e1661, http:// based composites: modeling and characterization, J. Phys. D: Appl. Phys. 36
dx.doi.org/10.1016/S0266-3538(03)00059-9. (2003) 573e582, http://dx.doi.org/10.1088/0022-3727/36/5/323.
[40] M. Bohle n, K. Bolton, Molecular dynamics studies of the influence of single [47] S.I. Kundalwal, M.C. Ray, Effective properties of a novel continuous fuz-
wall carbon nanotubes on the mechanical properties of Poly(vinylidene zyefiber reinforced composite using the method of cells and the finite
fluoride), Comput. Mater. Sci. 68 (2013) 73e80, http://dx.doi.org/10.1016/ element method, Eur. J. Mech. e A/Solids 36 (2012) 191e203, http://
j.commatsci.2012.10.010. dx.doi.org/10.1016/j.euromechsol.2012.03.006.
[41] J.D. Littell, C.R. Ruggeri, R.K. Goldberg, G.D. Roberts, W.A. Arnold, [48] H. Dumlich, M. Gegg, F. Hennrich, S. Reich, Bundle and chirality influences on
W.K. Binienda, Measurement of epoxy resin tension, compression and shear properties of carbon nanotubes studied with van der Waals density functional
stress-strain curves over a wide range of strain rates using small test speci- theory, Phys. Status Solidi B 248 (2011) 2589e2592, http://dx.doi.org/
mens, J. Aerosp. Eng. 21 (2008) 162e173, http://dx.doi.org/10.1061/(ASCE) 10.1002/pssb.201100212.
0893-1321(2008)21:3(162)). [49] T. Mori, K. Tanaka, Average stress in matrix and average elastic energy of
[42] nchez, J. Rieumont, Mechanical
F.G. Garcia, B.G. Soares, V.J.R.R. Pita, R. Sa materials with misfitting inclusions, Acta Metall. 21 (1973) 571e574, http://
properties of epoxy networks based on DGEBA and aliphatic amines, J. Appl. dx.doi.org/10.1016/0001-6160(73)90064-3.
Polym. Sci. 106 (2007) 2047e2055, http://dx.doi.org/10.1002/app.24895. [50] Y. Benveniste, A new approach to the application of Mori-Tanaka's theory in
[43] C.A. Cooper, S.R. Cohen, A.H. Barber, H.D. Wagner, Detachment of nanotubes composite materials, Mech. Mater. 6 (1987) 147e157, http://dx.doi.org/
from a polymer matrix, Appl. Phys. Lett. 81 (2002) 3873e3875, http:// 10.1016/0167-6636(87)90005-6.
dx.doi.org/10.1063/1.1521585. [51] Y.P. Qui, G.J. Weng, On the application of Mori-Tanaka's theory involving
[44] J.N. Coleman, M. Cadek, R. Blake, V. Nicolosi, K.P. Ryan, C. Belton, transversely isotropic spheroidal inclusions, Int. J. Eng. Sci. 28 (1990)
A. Fonseca, J.B. Nagy, Y.K. Gun'ko, W.J. Blau, High performance nanotube- 1121e1137, http://dx.doi.org/10.1016/0020-7225(90)90112-V.
reinforced plastics: understanding the mechanism of strength increase, [52] N. Marzari, M. Ferrari, Textural and micromorphological effects on the overall
Adv. Funct. Mater. 14 (2004) 791e798, http://dx.doi.org/10.1002/adfm. elastic response of macroscopically anisotropic composites, J. Appl. Mech. 59
200305200. (1992) 269e275, http://dx.doi.org/10.1115/1.2899516.

You might also like