Você está na página 1de 31

A Continuum Model to Study Interphase Effects on

Elastic Properties of CNT/GS-Nanocomposite


Ashish Srivastava*, Dinesh Kumar$
*
Ph.D. Scholar, Mechanical Engineering Department, Malaviya National Institute of
Technology, Jaipur, India
Email: ashish.memech@gmail.com, Ph. +919784277269
$
Assistant Professor, Mechanical Engineering Department, Malaviya National Institute of
Technology, Jaipur, India
Email : dkumar.mech@mnit.ac.in, Ph. +919549654562

Abstract: In the present study, initially a continuum level model of interphase region in a carbon
nanotube (CNT)/graphene sheet (GS) based nanocomposite is established. Cohesive zone model
through Lennard Jones (i.e., LJ) potential is utilized to model the interphase zone in terms of its
thickness and elastic modulus, for different matrix materials containing CNTs of various radii
(including graphene - a CNT of infinite radius). Thereafter, a finite element based study is
conducted to characterize and compare the CNT- and GS-reinforced nanocomposites with and
without interphase, using the method of representative volume element (RVE). Based on the
study, it is concluded that the thickness as well as elastic modulus of the interphase zone is
significantly affected by the matrix material of nanocomposite which is in contrast to the general
assumption, made in the literature, of constant thickness of interphase zone for different matrix
materials. In the case of nanocomposite with small radius CNTs, slightly higher, noticeable only
at narrow scale, elastic modulus of interphase zone is obtained than the nanocomposite with
large radius CNTs, irrespective of matrix materials. For metal nanocomposites, interphase zone
results in loss of all stiffness properties of the resulting CNT/GS nanocomposite, and this loss is
more prevalent in GS nanocomposite for its out-of-plane stiffness properties. But, on the
contrary, in the case of polyethylene (i.e., PE) nanocomposite, the elastic modulus of interphase
causes enhanced stiffness properties of the resulting nanocomposite, as compared to the perfectly
bonded nanocomposite, for CNT and GS reinforcements.

Keywords: Carbon nanotube (CNT), Graphene sheet (GS), Elastic properties, Cohesive zone,
Lennard Jones (LJ) potential, Representative volume element (RVE).

1. Introduction

Carbon nanotube (CNT), a member of fullerene family, is 1D cylindrical structure of


hexagonally arranged carbon atoms. Because of high aspect ratio, low density, high strength and
stiffness, CNTs are considered as a great reinforcing agents for light weight nanocomposites [1].
Graphene sheet (GS), a 2D structure of hexagonally arranged carbon atoms and another member
of fullerene family, was first produced in 2004 by Andrew Geim and Konstantin Novoselov [2].
GSs are also found to possess extra-ordinary physical and mechanical properties, like CNTs, and
hence, can be a very effective reinforcing element to produce GS based nanocomposites. GSs
display enhanced in-plane values of thermal conductivity, mechanical stiffness, fracture strength
properties because of its high surface area and its stronger interfacial bonding with polymers, and
greater ease in its production and handling than CNTs [3]. The effects of GS and CNT on the
performance of nanocomposite are judged only on the case-by-case basis depending on the
intended applications. For instance, CNTs are more effective as mechanical reinforcement due to
their fibrous structure, which would give advantage of short-fiber reinforcement, and in case of
emitters because of their cylindrical shape, whereas GSs perform better in case of composites
processed through high pressure techniques and in tribological applications [4].

The attractive mechanical properties of CNT/GS can be imparted into the nanocomposites by
combining the nanofillers with different kinds of matrix materials. Because of miniature size of
nanofillers (CNT/GS), the proper characterization of nanocomposites having enhanced
mechanical properties has been a challenging task either experimentally or analytically. Many
researchers have developed experimental and simulation methods to predict the behavior of
nanocomposites. There are enormous challenges associated with experimental characterization
of nanocomposites, such as very expensive fabrication of nanocomposite and highly time
consuming process [5–8], whereas computational simulations play a vital role in material
characterization of nanocomposites owing to its versatile nature (for various size and
configuration of nanofillers and loading pattern) as well as because of less computational time
and cost. Furthermore, the analysis results of computational simulation can also be used as a
guide to plan and conduct experiments.

Nomenclature
: Cohesive energy
σ: Van-der-Waals radius
ε: Bond energy at equilibrium distance
r: Variable distance between a point in the matrix material from surface of
nanofillers
: Area density of CNT/GS
: Volume density of matrix material
R: Outer radius of CNT
ℎ : Equilibrium distance between the nanofiller and matrix material
: Volume fractions of CNT & GS into the RVE
a and b : Sides of square RVE
w and t : Width and thickness of graphene sheet
and : Internal and external radii of the CNT
: Volume fraction of interphase zone
L: Length of RVE and nanofillers (CNT/GS)
, and : Displacements of RVE in x, y and z directions, respectively
: Constant value of displacement applied in x-direction on x = L face of the
RVE
: Constant value of displacement applied in y-direction on y = a face of the
RVE
: Constant value of displacement applied in z-direction on z = b face of the
RVE
: Constant value of displacement applied in y-direction on z = b face of the
RVE
: constant value of displacement applied in x-direction on y = a face of the
RVE
, and : Stress, strain and stiffness tensors, respectively
and ̅ : Volumetric average of stress and strain of RVE
: Effective stiffness tensor
, and : Young's moduli of nanocomposite in the x, y and z-direction
, and : Shear moduli of nanocomposite corresponding to x-y, x-z and y-z planes
: Young's modulus of nanofillers (i.e., CNT/GS)
: Young's modulus of matrix materials

Computational simulations can be broadly classified into two widely used simulation techniques:
molecular simulation and continuum simulation. Plethora of publications are available in the
literature based upon molecular simulation to characterize nanocomposites [9,10], but such
approaches have limitations of small length and time scales which restrict their applicability to
small number of molecules/atoms. On the other hand, continuum methods are not constrained on
the length and time scales, and are suitable for the study of nanocomposites at relatively less
computational time and cost. To solve continuum problems, finite element method (FEM) is a
very efficient tool and has been used very frequently by many researchers to solve problems at
smaller length and time scales as well because of its high computational capabilities and
accuracy [11–16].

Extremely large aspect ratios of CNT/GS tend to transfer high load from the matrix to nanofillers
which in turn depends upon the strength of interfacial region between nanofiller and matrix
material, that in turn is found to influence the mechanical behaviour of nanocomposites [17–19].
Two kinds of model are often used to simulate the interfacial region between reinforcements and
matrix materials in composite materials: the interface model wherein displacement and/or stress
discontinuities are presumed to exist at an interface, and the interphase model which describes
interface region as a layer, called an interphase zone [20]. The two interface models, namely
linear- spring model and interface stress model, are two-phase models since the interface region
occupies zero volume fraction in the composite material [21,22]. Whereas the interphase model
is a three-phase model consisting of the reinforcement, the interphase and the matrix material
with perfect bonding at both the interfaces: matrix/interphase and interphase/reinforcement, and
with the elastic modulus of the interphase zone different from those of the matrix and
reinforcement [23]. Numerous studies have been performed by the researchers around the world
to discuss the effect of interfacial bonding on the effective mechanical behavior of
nanocomposite materials. In 2003, Frankland et al. [24] performed molecular dynamics
simulation of the vdW interfacial interaction between the polymer and the CNTs using Lennard-
Jones (LJ) potential and concluded that the long CNTs increase the stiffness of the composite
more effectively than short CNTs because of their high aspect ratio. In a study by Tsai et al. [23],
a three-phase model of the CNT nanocomposites was introduced by modeling the non-bonded
gap between the CNT and polyimide as a separate phase and calculated the elastic stiffness of the
interphase by the non-bonded energy (i.e., LJ potential). Shokrieh and Rafiee [25] used the
COMBIN39 element of FEM based software ANSYS to model the interphase (treated using
vdW interactions) and developed an equivalent long fiber for predicting the mechanical
properties of the CNT/polymer nanocomposite and concluded that micromechanics based
method, such as rule of mixtures (ROM), cannot capture the difference between the micro and
nanoscale. The effects of CNT length and diameter, interphase thickness and the cut-off distance
of LJ potential on the interfacial shear strength between the CNT and matrix material was
investigated by Wernik et al. [26] by simulating a nanotube pull-out experiment using multiscale
computational model and observed that CNTs of smaller diameter, smaller nanotube embedded
length, and lower interfacial thickness favor high interfacial shear strength. Liu et al. [27] used
boundary element methods (BEM) to developed a cohesive interphase model between the CNT
and polymer in the nanocomposite. Multiwall carbon nanotube (MWCNT) based nanocomposite
having matrices ranging from soft polymers to stiff ceramics with weak interphase region was
modeled by Joshi and Upadhyay [21] to predict the elastic properties of the resulting
nanocomposite. Mohammadpour et al. [28] studied the large strain behavior of nanocomposite
using FEM based software ANSYS by adopting the non-linear material properties of CNT and
matrix material and predicted the Young's modulus and ultimate strength of the nanocomposite
under axial load.

Very few publications are available in literature describing the mechanical behavior of GS based
nanocomposite including the effect of interphase zone. Parashar and Mertiny [29] predicted the
buckling behavior of GS based nanocomposite by considering the interphase zone through vdW
interactions between the GS and polymer and reported a significant improvement in the buckling
strength of the nanocomposite as compared to that of neat polymer. Awasthi et al. [30] studied
the load transfer between GS and polymer matrix and reported that mechanical interactions
between GS and polymer chains are stiffer than those amongst the polymer chains.

A generalized equation of interface potential, based on vdW forces between SWCNT (single
wall CNT)/ MWCNT (multi-wall CNT)/ GS and polymer matrix, in terms of area density of
CNT/GS and volume density of polymer was given by Jiang et al. [31], and derived an
expression for net cohesive stresses between the nanofillers and matrix material. Tan et al. [32]
extended the work of Jiang et al. [31] to derive a nonlinear cohesive law for the interface
between CNT and polymer and, concluded that though at smaller strains CNTs improve the
mechanical behavior of the nanocomposite, but at relatively higher strains such improvements
disappears because of debonding of CNTs with matrix materials. In another study on the
cohesive energy of interface zone, Zhao et al.[33] discussed the effects of size, spacing and
crossing angles of the nanofillers (i.e., CNT and GS) on the effective potential of interface zone,
and found that smaller diameter CNTs results in higher cohesive energy than large diameter
CNT, and the equilibrium distances of interface region increase with increase in CNT radius.
Zhang et al. [34] derived analytical solution on interphase for large diameter carbon nanotube
reinforced composite with functionally graded variation using cohesive zone model proposed by
Jiang et al. [31] to connect CNTs and the interface.

From the above literature, it is evident that the interfacial region between the nanofillers and
matrix material affects significantly the effective mechanical properties of the nanocomposite,
and thus, needs proper attention. In most of the studies considering interphase zone, the thickness
of the interphase zone is assumed to be constant, without any physical or experimental basis, for
different matrix materials in various continuum based FEM simulations [21, 28–30]; whereas it
is taken equivalent to the vdW gap (i.e., equilibrium distance) between the nanofiller and matrix
material in the studies based on molecular dynamics [31, 32]. Therefore, it is required to develop
a continuum level model of interphase zone based on physical and experimental bases. Based on
these observations, the aim of present study is twofold. First, to develop a continuum level model
of interfacial region as a separate layer termed as interphase zone in a CNT/GS-reinforced
nanocomposite based on the physical background and the experimentally determined
intermolecular parameters. A cohesive zone model (as introduced by Jiang et al. [31]), through
LJ potential, is used to model interphase zone in terms of its thickness and elastic modulus, for
different matrix materials containing CNT of various radii (including CNT of infinite radius i.e.,
GS). Perfect bonding is assumed between matrix/interphase and interphase/reinforcement
interfaces. Second, an FEM based study is conducted to characterize and compare CNT- and GS-
reinforced nanocomposites, with and without interphase, using an RVE method.

2. Characterizing Interphase Zone

The non-bonded interactions between the nanofillers (CNT/GS) and matrix materials can be
described by the vdW force based cohesive zone model established, through L-J potential, by
Jiang et al. [31] for predicting the cohesive stresses (tensile and shear) between the nanofillers
(CNT/GS) and matrix materials. The same model has been used in the present study to predict
the linear elastic behavior of interfacial region modeled as interphase zone to subsequently study
its effect on the resulting nanocomposite of different matrix materials.

Cohesive energy between the matrix and GS, per unit area of GS, [31], is given as:

= − , (1)

and, cohesive energy between the matrix and CNT, per unit length of CNT, is given as

= 2 − + − ; (2)
where, , a measure of how strongly two atoms attract each other, is the well depth of the LJ
potential curve of a hypothetical material to mimic the effect of interphase region; , also
referred as vdW radius, represents the distance at which the intermolecular potential between the
two atoms vanishes and it measures how close two non-bonding atoms can approach. Further,
the area density, of the CNT/GS is the number of carbon atoms per unit surface area of
CNT/GS and it is evaluated using: = , wherein represents equilibrium bond length

prior to deformation; represents the volume density of matrix material (i.e., number of atoms
per unit matrix volume), and; r and R represent the variable distance between a point in the
matrix material from surface of nanofillers, and the outer radius of CNT, respectively.

Since in the present study, interfacial interaction between nanofiller and matrix material is
simulated through a hypothetical interphase zone, thus intermolecular interfacial parameters can
be utilized to model the interphase zone, as utilized by Shokrieh, & Rafiee, [25] and Joshi &
Upadhay [21] to model the interphase zone. Lorentz-Berthelot (LB) mixing rule can be utilized
to determine interatomic potential parameters, as mentioned by Fan [40]. In the current study as
well, intermolecular potential parameters (i.e., and ) for the assumed hypothetical interphase
zone between the atoms of CNT/GS and matrix material are calculated using Lorentz-Berthelot
(LB) mixing rule [41,42]. LB mixing rule estimates the intermolecular potential parameters
between the pairs of non-identical atoms, such as C-Mg, Ni-Mg, using the intermolecular
potential parameters between the pairs of identical atoms. For instance, Zhou et al. [43] studied
the stress-strain behavior of nickel coated CNT-reinforced magnesium nanocomposite by using
MD and utilized LB mixing rule to predict the intermolecular potential parameters between
nickel-magnesium, carbon-magnesium and magnesium-nickel atoms. Similarly, Choi et al. [44]
& Silvestre et al. [45] employed the LB mixing rule to estimate the intermolecular potential
parameters between atoms of CNT and that of aluminum matrix material subjected to tensile
loading and compressive loading respectively. Moreover, Yani [46] simulated the CNT-polymer
nanocomposite with the aid of LB mixing rule. Very recently, Rezaei et al. [47] reported the
mechanical behavior of CNT-reinforced metallic glass nanocomposite and utilized the same rule
to obtain the values of intermolecular potential parameters between the carbon atoms of CNT
and copper & zirconium atoms of metallic glass matrix; and Lu et al. [48] reported a comparative
study between the elastic properties of two dimensional GS & zero dimensional fullerene-
reinforced polymer nanocomposite.

The LB mixing rule is given by:

= , (3)
= . (4)

Interphase potentials, as given by Eqs. (1) & (2), are considered as the total strain energy of the
assumed isotropic interphase material. The stiffness of this isotropic interphase material can be
predicted by below Eq. (5), as used by Sears and Batra [49],
= , (5)
where, strain (S) at any distance is taken as, = , wherein, ℎ represents the equilibrium
distance between the nanofiller (CNT/GS) and matrix material (i.e., the value of r at
equilibrium), and it is evaluated by minimizing the interphase potentials by satisfying: ⁄ =
0. The equilibrium distances calculated by the minimization of interphase potentials represented
by Eqs. (1) & (2) are evaluated to be almost same and is given by:

ℎ = 0.8584 . (6)

The elastic modulus of the interphase is evaluated for the equilibrium position (i.e., = ℎ ), and
the expression for the stiffness of the interphase between GS and matrix material is derived as:

=8 ℎ − . (7)

Similarly, the stiffness of the interphase between CNT and matrix material can be evaluated from
the following derived expression:

= ℎ 24 − +9 − . (8)

It is necessary to mention here that E obtained from Eq. (8) for the interphase between CNT and
matrix material will have unit of force per unit length and hence, needs to be divided by average
circumference [i.e., by 2 + , obtained by average of outer circumference of CNT (i.e.,
2 ) and inner circumference of matrix material (i.e., 2 ( + ℎ ))].

3. Modeling Approach
In order to predict the elastic behavior of the nanocomposite, a square representative volume
element (RVE) consisting of reinforcement (CNT/GS), interphase material and surrounding
matrix material is modeled using FEM based software COMSOL Multiphysics. The method of
RVE is well established and has been used by many researchers [12,14] for evaluating stiffness
properties of composite materials at micro and nano scales with the suitable application of
periodic boundary conditions. The thickness of the interphase zone in the RVE is considered as
the equilibrium distance between the nanofillers and matrix materials, given by Eq. (6). Below
subsections further describe the modeling of RVE, the periodic boundary conditions and the
homogenization technique to predict the effective (i.e., averaged over the volume of RVE) elastic
properties of nanocomposite material.

3.1. Representative Volume Element (RVE)

In a nanocomposite material, the actual nanofillers are randomly distributed across the volume.
But for simplicity reasons, most researchers have assumed a periodic arrangement of nanofillers
for which a RVE can be isolated [13,21,56]. In the current study, the distribution of nanofillers
within the matrix material is assumed to be periodic, and thus the concept of representative
volume elements (RVEs) can be applied to model the nanocomposite. Continuum mechanics
based 3-D representative volume elements (i.e., cylindrical, square and hexagonal) proposed by
Liu & Chen [12], in conjunction with FEM, have been widely used by different authors
[14,50,51] to evaluate the elastic constants of nanocomposites. In a subsequent study by Chen &
Liu [13], it was found that, out of the three proposed RVE, the cylindrical RVE tends to
overestimate the effective Young's moduli of the nanocomposite material because of the
overestimation of the volume fraction of the CNT in a matrix material, and it was also shown
that the predictions made by using the square RVE are more accurate. Whereas, different authors
have also used cylindrical and hexagonal RVEs [14,15] to make predictions of mechanical
properties of nanocomposite. Therefore, based on the study by Chen & Liu [13], a square RVE is
used in the present study to predict the elastic behavior of nanocomposite material, including the
effect of the interfacial region. The non-bonded interfacial region between CNT/GS and matrix
material is simulated as a hypothetical interphase zone and the interfaces between
matrix/interphase and interphase/reinforcement are assumed to be perfect. Interphase zone
modeled between the nanofiller and matrix material for a given volume fraction of reinforcement
in the matrix material is utilized to evaluate the effect of vdW interaction between nanofiller and
matrix material on the elastic properties of nanocomposite material. The expressions for
calculating volume fraction of nanofillers in the matrix material are given below.

The volume fractions (the subscript N means nanofiller) of CNT & GS into the RVE are
given by the following equations.

For CNT: = , and for GS: = ; (9)

whereas, the volume fraction of interphase (the subscript I stands for interphase zone) in the
CNT reinforced RVE and in the GS reinforced RVE are given by the following equations.

( )( )
For CNT: = , and for GS: = , (10)

where, and are the inner and outer radii of the CNT, respectively, and w and t represent the
width and thickness of the GS, respectively. Eqs. (9 & 10) are used to evaluate the side of square
RVE (i.e., a) having interphase zone as shown in Fig. 1. (= + ℎ ) represents the outer
radius of interphase zone.
Fig. 1. Square RVE with interphase zone (a) CNT reinforcement (b) GS reinforcement.

3.2. Boundary Conditions

Boundary conditions play a vital role in the accurate computation of effective elastic moduli by
simulating the actual deformation within the nanocomposite and are thus required to be specified
precisely. In the present analysis, the displacement boundary condition is applied on the RVE for
different loading conditions by judicious use of symmetry and periodicity conditions, as derived
by Sun and Vaidya [52] and are used for the prediction of elastic properties of different
nanocomposites using FEM based commercial package COMSOL Multiphysics. Following
subsections contain the displacement boundary conditions applied to the finite element model of
RVE for normal loading, transverse shear loading and longitudinal shear loading cases to
calculate various elastic moduli.

3.2.1. Boundary conditions for normal loading

Following periodic displacement boundary conditions on the RVE are applied to calculate
longitudinal and transverse moduli and Poison's ratios:

For calculating 1, 12 and 13 (refer Fig. 2):

(0, , ) = ( , 0, ) = ( , , 0) = 0; ( , , )= , (11)

where, , and represent displacement components in x, y and z directions, respectively,


and δ1 is the constant value of displacement applied in x-direction on x = L face of the RVE.

For calculating and :

(0, , ) = ( , 0, ) = ( , , 0) = 0; ( , , )= , (12)
where δ2 is the constant value of displacement applied in y-direction on y = a face of the RVE.

Similarly, for calculating :

(0, , ) = ( , 0, ) = ( , , 0) = 0; ( , , )= , (13)

where is the constant value of displacement in z-direction on z = b face of the RVE.

Fig. 2. Typical RVE under normal loading in x-direction (for calculating , and ).

Fig. 3. RVE under transverse shear loading (for calculating ).


3.2.2. Boundary conditions for transverse shear loading
For calculating 23 (refer Fig. 3):

( , , 0) = ( , , 0) = ( , , 0) = 0; ( , , )= , (14)

where is the constant value of displacement applied in y-direction on z = b face of the RVE, as
shown in Fig. 3.

3.2.3. Boundary conditions for longitudinal shear loading

For calculating 13 (refer Fig. 4):

( , 0, ) = ( , 0, ) = ( , 0, ) = 0 ; ( , , )= , (15)

where is the constant value of displacement applied in x-direction on y = a face of the RVE, as
shown in Fig. 4.

Similar to 13 , the boundary conditions for calculating 12 would be as follows:

( , , 0) = ( , , 0) = ( , , 0) = 0 ; ( , , )= , (16)

where is the constant value of displacement in x-direction on z = b face of the RVE

Fig. 4. RVE under longitudinal shear loading (for calculating and ).

3.3. Homogenization method for evaluating average stress and strain over the RVE

At a nanoscale, an RVE consisting of three phases- matrix material, nano reinforcement:


CNT/GS, and interphase zone - is actually a heterogeneous composite medium that is to be used
for evaluating effective (i.e., average) material properties of the nanocomposite that is considered
to be homogeneous at larger scales (i.e., micro and meso scales). Therefore, it is required to use
some homogenization techniques to find a globally homogeneous medium equivalent to the
original heterogeneous composite medium at nanoscale and to reduce the non-uniform stress and
strain fields within the heterogeneous material obtained from the finite element analysis of RVE
to the volume-averaged stress and strain. Following paragraphs explains a numerical
homogenization procedure to determine the effective moduli that describe the 'average' material
properties of the actual heterogeneous nanocomposite.

Individual phases have isotropic material properties, and it is assumed that the constitutive law in
the matrix and the reinforcement is given by the following generalized Hooke's Law:

= , , , , = 1, 2, 3, (17)

where, , and are the coefficients of stress tensor, linear strain tensor and stiffness
tensor, respectively.

The FEM analysis of the RVE would yield the above mentioned stress and strain fields within
the heterogeneous material. The effective (i.e., averaged) stiffness coefficients of nanocomposite
(at micro or macro scale) can be calculated from

= ̅ , (18)

where, refers to the effective stiffness tensor, and and ̅ are the volume-averaged
components of stress and strain tensors calculated over the volume of the RVE using following
volumetric integral expressions as:

= ∫ ( , , ) , (19)

̅ = ∫ ( , , ) , (20)

where, represents the volume of RVE.

In the present study, COMSOL Multiphysics software is used to carry out the FEM analysis of
the RVE, and all finite element based calculations, required to determine homogenized material
properties of nanocomposites such as volume-average of stress and the strain components, are
also done using the especial features of the tool.

The RVE is meshed with tetrahedron elements using physics-controlled meshing feature of
COMSOL with fine enough mesh near to CNT/GS (as shown in Figs. 5 & 6) to deliver
converged FEM results. Various periodic boundary displacement conditions for different loading
cases, as discussed in Section 3.2 and shown in Figs. (2-4), are applied to yield the stress and
strain fields within the actual heterogeneous RVE.
Fig. 5. FEM model of a square RVE of CNT reinforced in a matrix
Fig. 6. FEM model of a square RVE of graphene reinforced in a matrix

After calculating the volume-averaged stress and strain components, the relevant nanocomposite
moduli (i.e., effective/averaged stiffness coefficients) can be obtained from average stresses and
average strains. The stiffness tensor can be written as:

̇ ̇
= , = = = ; (for pure normal strain states) (21)

̇ ̇
= , ( = ) ≠ ( = ); (for pure shear strain states) (22)

and, the Poison's ratio is given by:

̇ ̇
=− . (23)
̇̇

Therefore, Eqs. (21-23) are used, respectively, to calculate the effective Young's moduli, shear
moduli and Poison's ratios of the nanocomposite.

4. Analytical approach

To validate the procedure, to evaluate elastic properties of nanocomposite, mentioned in Section


3.0, the evaluated axial modulus of nanocomposite is compared with the widely used analytical
technique, rule of mixtures (ROM) [13–15] under constant strain condition, and it is given as:

= + + (1 − − ) , (24)

where, is the axial elastic modulus of nanocomposite; , and represent the elastic
modulus of nanofiller, interphase material and matrix material, respectively, and and
represent the volume fractions of nanofiller and interphase material, respectively.

5. Present Study

In the present study, initially the elastic modulus of interphase zone between the nanofillers and
matrix material is predicted. Stiffness of interphase material is calculated using Eqs. (8 & 9) for
the cases of GS and CNT nanocomposites, respectively, by substituting the values of vdW radius
(i.e., ), the bond energy of the interphase zone (i.e., ), the area density of carbon atoms in
CNT/GS (i.e., ) and the mass density of matrix material (i.e., ). The values of and are
obtained by widely used LB mixing rule given in Eqs. (3 & 4), respectively. In Eqs. (3 & 4), the
bond energy and equilibrium distance between a pair of carbon atoms are taken as = 0.002390
eV and = 0.3415 nm, respectively, whereas, the values of and between a pair of atoms
of various matrix materials are tabulated in Table 1. The value of the area density of carbon
atoms in CNT/GS is taken as 3.8177 × 10 [26], and the values of mass density (i.e., )
of various matrix materials can be evaluated by the ratio of matrix mass density to the mass of
per unit atom (i.e., amu) of the matrix, as given in Table 1.

Table 1
Potential parameters (i.e., and ), density and atomic mass of different metals
L-J Potential parameters Gold, Au Aluminum, Al Copper, Cu Iron, Fe Magnesium, Mg
σm(A0)[35] 1.8700 2.8500 2.3380 4.1400 3.0030
*
εm[eV] [35] 0.0375 0.0050 0.0041 0.0007 0.0012
3
density (g/cm )[40] 19.3200 2.7020 8.9600 7.8600 1.7380
**
atomic mass (amu) [40] 196.96655 26.981539 63.5460 55.8450 24.3050
*
1 = 1.60217657 × 10 Joules
**
1 = 1.660540199 × 10 Kg

For the case of interphase between the nanofiller and polyethylene matrix material, the required
LJ potential parameters between the polyethylene unit molecule -CH2- and carbon atoms of
CNT/GS (i.e., for − pair) are taken as: = 0.004656 eV and = 0.3825 nm [53]. The
volume density of polyethylene polymer ( ), a ratio of polyethylene mass density (i.e., 0.71 ×
103 kg/m3) to the mass of –CH2–unit (i.e., 2.3 ×1026 kg), is calculated as 3.1 ×1028 m-3 [31].

After evaluating the stiffness properties of interphase materials for different matrices, the square
RVE containing CNT/GS, interphase material and matrix is modeled and analyzed using FEM
based tool COMSOL Multiphysics to predict the effective material properties of the resulting
CNT- and GS-based nanocomposites for different volume fractions of reinforcements (i.e.,
CNT/GS).

The dimensions and elastic properties of the CNT and GS are given below.

For CNT: Length = 100 nm, Thickness of CNT t = 0.34 nm, Young modulus (CNT) = 1
TPa, Poisson's ratio υ = 0.3. Chiralities of CNTs taken are (5,5), (10,10), (20,20) and (50,50).

For GS: Length = 100 nm, Thickness of graphene, t = 0.34 nm, Young modulus (GS) = 1
TPa, Poisson's ratio υ = 0.3.

Elastic moduli for different matrices considered are listed in Table 2, whereas the Poisson's ratio
for each matrix material is taken as 0.3.

Table 2
Elastic moduli of different matrix materials
Gold, Aluminum, Copper, Iron, Magnesium, Polyethylene,
Au Al Cu Fe Mg PE
E [GPa] 78 70 130 211 45 3.4
For a given volume fraction of the reinforcement, Eq. (9) is used to model the RVE with
interphase of thickness ℎ , evaluated using Eq. (6), for CNT and GS reinforced nanocomposites.
To model a square RVE for GS reinforced nanocomposites, the width of GS is taken as 0.8a,
where a represents the side of square RVE. Further, to compare the effect of interphase zone
around the surfaces of GS and CNT on the elastic behavior of resulting nanocomposite, the width
of GS is taken equal to the periphery of CNT (i.e., = 2 ), for the same volume fraction;
where, is the mean of inner and outer radii of a CNT.

After the characterization of interphase between nanofillers and different matrix materials for its
elastic modulus, and the modeling of RVE, the stiffness properties of the CNT/GS reinforced
nanocomposites are predicted through finite element analysis of the formed RVE with an
objective to investigate the effects of perfect and imperfect bonding between nanofillers and
different matrix materials on the elastic behaviour of CNT/GS nanocomposite.

6. Results and discussions

In this section, initially the equilibrium distance between CNT/GS and different matrix material
(i.e., thickness of interphase zone, ℎ ) are evaluated and discussed in Section 5.1 and thereafter,
stiffness property of hypothetical interphase material between CNT/GS and different matrix
materials are evaluated in Section 5.2 for various radii of CNT (including GS, i.e., CNT of
infinite radius). Finally, the effect of interphase on the stiffness properties of the CNT and GS
reinforced nanocomposites are studied in Section 5.3.

6.1 Interphase thickness (h0) between the nanofiller and different matrix materials

Stiffness of the nanocomposite is affected significantly by the load transfer efficiency of the
interphase region which in turn depends on the bond energy between the nanofillers and matrix
materials, and the thickness of interphase region [21,36]. The thickness of interphase (i.e.,
equilibrium distance between the reinforcement and matrix material) is calculated using Eq. (6),
which is found to vary with respect to vdW radius (i.e., ) of the interphase material. According
to the LB rule [as given by Eq. (3)], vdW radius of interphase depends on vdW radii of atoms of
CNT/GS ( ) and that of matrix material ( ). Since, in the present study vdW radius between
the carbon atoms ( ) of CNT/GS remains constant, thus the equilibrium distance of the
interphase depends solely on the vdW radius of the atoms/molecules of matrix materials (i.e.,
) in the case of metal matrix.

Equilibrium distances (ℎ ) obtained between the nanofillers and different matrix materials are
plotted in Fig. 7. It can be seen from Fig. 7 that higher equilibrium distance is obtained for the
nanocomposites with interphase of large vdW radius (i.e., ) as in the cases of polyethylene and
iron matrices. In case of metal matrix, the maximum and minimum values of the equilibrium
distances (ℎ ) are obtained for iron and gold matrix materials, respectively; whereas, in the case
of polyethylene matrix, possessing higher value of vdW radius of interphase zone than
considered metal matrix materials, the thickness of interphase zone is maximum.

Gold, Au
0.35 Aluminum, Al
Copper, Cu
Iron, Fe
0.30 Magnesium, Mg
Polyethylene, PE

0.25
Equilibrium distance (ho), nm

0.20

0.15

0.10

0.05

0.00
Matrix materials

Fig. 7. Equilibrium distance (h0) between the nanofiller and different matrices
6.2 Stiffness of the interphase material

The elastic modulus of the interphase region between the CNT and different matrix materials is
predicted for different values of radius of CNT (i.e., R = 0.35, 0.69, 1.39, 3.46 nm and ∞), and
the predicted values are shown in Table 3. It can be seen from Table 3 that for a given matrix
material, the effect of radius of the CNT on elastic modulus of interphase zone is not
pronounced, and it would be noticeable only if compared on a narrow scale (as reflected in Fig. 8
drawn for gold matrix). Table 3 also depicts a slight higher elastic modulus of interphase zone
for small radius CNT (i.e., 0.35 nm) than large radius CNT, irrespective of matrix material. This
finding can be attributed to the fact that the bond energy at equilibrium distance between CNT
and matrix material is high for small radius CNT, and it decreases as the radius of CNT is
increased to that of GS (i.e., R = ∞), as shown in Fig. 9. Thus, CNTs of small radius are found to
be better reinforcement than CNTs of large radius and this finding is in good concurrence with
the experimental study by Nam et al. [55] about the mechanical properties of CNT/epoxy
nanocomposite.

It can also be observed from Table 3 that in the case of PE matrix, the elastic modulus of
interphase zone is significantly more than that of pure PE matrix material. This finding is in good
concurrence with the similar findings by Zhang et al. [37] and Rafiee & Pourazizi [38],
respectively, about higher strength and elastic modulus of interphase zone between CNT and
polymer matrix than that of normal polymer matrix material. Table 3 also depicts that for a given
radius of CNT, the elastic modulus of interphase zone is significantly affected by the matrix
material, and the maximum and minimum values being for gold and magnesium matrix
materials, respectively. It is necessary to mention here that for a given matrix material, the elastic
modulus of interphase zone would be a function [through Eqs. (7 & 8)] of the intermolecular
potential parameters (i.e., and ), and the volume density of matrix material (i.e., ).
Therefore, the combined effect of these variables will fix the value of elastic modulus of the
interphase zone, as shown in Table 4.
Table 3
Comparison of the elastic modulus of interphase (GPa) between the CNT and the different matrix materials, for different values of
radius of CNT.
Elastic modulus of interphase (GPa)
Radius of
Gold, Aluminum, Copper, Iron, Magnesium, Polyethylene,
CNT, R (nm)
Au Al Cu Fe Mg PE
3.18850992 8.96585181
∞ (i.e., GS) 16.57921024 8.68539066 9.33778876 6.64201399
9.33779843
3.46 16.57922604 8.68540042 6.64202292 3.18851359 8.96586401
9.33781164
1.39 16.57924780 8.68541367 6.64203482 3.18851855 8.96588026
9.33783133
0.69 16.57928054 8.68543320 6.64205194 3.18852586 8.96590357
9.33786190
0.35 16.57933212 8.68546307 6.64207715 3.18853697 8.96593780

Table 4
Combined value of the intermolecular potential parameters (i.e., and ) of interphase zone for various matrix materials, and the
volume density of matrix material (i.e., ) to study its effect on the elastic modulus of interphase zone.
Matrix material (eV) (nm) (m-3)
Gold, Au 0.021141 0.264250 5.91×1028 0.008725×1028
Aluminum, Al 0.003456 0.313250 6.03×1028 0.002045×1028
Copper, Cu 0.003130 0.287650 8.49×1028 0.002199×1028
Iron, Fe 0.001293 0.377750 8.47×1028 0.001563×1028
Magnesium, Mg 0.001693 0.320900 4.31×1028 0.000751×1028
Polyethylene, PE 0.004656 0.382500 3.10 ×1028 0.002112×1028
16.57935 R = 0.35 nm
R = 0.69 nm
R = 1.39 nm
R = 3.46 nm
Elastic modulus of interphase (GPa) R = Infinity
16.57930

16.57925

16.57920

16.57915

Radius of CNT (nm)

Fig. 8. Effective elastic modulus of interphase between gold matrix and CNT of various radii.

0.15
R=0.35nm
R=0.69nm
0.1 R=1.39nm
R=3.46nm
Graphene
0.05
Cohesive energy (J/m2)

-0.05

-0.1

-0.15

-0.2
0.18 0.19 0.2 0.21 0.22 0.23 0.24 0.25 0.26 0.27 0.28
Inter-atomic distance, r (nm)

Fig. 9. Cohesive energy of interphase (between the GS/CNT as nanofiller and gold matrix) per unit surface area
of CNT/GS.
6.3 Evaluation of Stiffness properties of nanocomposite

6.3.1 Validation

To validate the procedure, as discussed in Section 3, employed in the present paper to predict
the elastic properties of nanocomposite material, the stiffness properties of CNT-reinforced
nanocomposites are compared with the results obtained through ROM and also with the
available results- experimental and Molecular Dynamics (MD) simulation- in literature, as
given in Table 5. It is to mention that in literature MD simulations techniques are used to
predict axial modulus of anisotropic nanocomposites having a single CNT, and the same value
is compared with the obtained value using FEM in Table 5. Whereas the nanocomposite tested
experimentally will be more or less isotropic in nature having same elastic modulus in all
directions. To compare the elastic modulus of isotropic nanocomposite predicted
experimentally in the literature with that obtained through FEM, Tsai-Pagano method, as used
by other researchers as well [58–60], is utilized to obtain the corresponding elastic modulus of
an isotropic nanocomposite using the longitudinal and transverse modulii of RVE of the
nanocomposite from FEM analysis [refer Section 3]. Tsai-Pagano provide the following
expression to evaluate the elastic modulus:

= + , (25)

where, E represents the effective Young's modulus of the isotropic nanocomposite material;
& are the longitudinal and transverse Young's modulus of single CNT-reinforced RVE.
While comparing the results with those in the literature, the elastic modulii of CNT & matrix
materials and volume fraction are taken same as in the corresponding literature.
As observed in Table 5, a good agreement can be seen in the compared values of elastic
modulus of nanocomposite. The same technique is utilized in Section 6.3.2 of the present study
to characterize CNT and GS-reinforced nanocomposite materials and to conduct a comparative
study between them.
Table 5 Comparison of elastic modulii (in GPa) of nanocomposite.
FEM ROM Reference
** %age
* ** deviation in
Matrix material * [using Tsai-Pagano *
corresponding
method, i.e., (using MD) (Experimental)
values
Eq.(25)]
Gold, Au 150.8086 72.0709 101.5975 150.7992 156.0000 [61] - 3.3278
92.1123 73.1583 80.2660 92.1106 97.0100 [44] - 5.3170
Aluminum, Al
89.3836 70.9970 77.8919 89.3845 - 78.0000 [62] 0.1386
85.0582 56.6983 67.3332 85.0534 76.8000 [71] - 9.7088
Copper, Cu
229.8354 37.2721 109.4833 229.8386 - 105.9000 [64] 3.3836
Iron, Fe 235.0966 189.8166 206.7966 235.1403 - 222.2222 [65] 7.2324
Magnesium, Mg 52.9626 33.5903 40.8549 52.9587 - 38.6000 [66] 5.5192
Polyethylene, PE 51.2027 4.4790 22.0003 51.1998 50.0000 [24] - 2.4054
* **
Represents axial modulus of CNT-reinforced nanocomposites; Represents Young's modulus of randomly-dispersed CNT-reinforced nanocomposites with
isotropic properties.
6.3.2 Stiffness properties of nanocomposite

In this section, the effect of bonding (i.e., perfect and imperfect) between CNT/GS
reinforcement and different matrix materials on stiffness properties of the resulting
nanocomposite are presented and discussed. The stiffness properties of nanocomposite are
obtained from FEM analysis of 3D nanoscale RVE. In the case of imperfect bonding between
CNT/GS and matrix material, a interphase zone of thickness equal to equilibrium spacing
between atoms of nanofiller and matrix materials is considered.

Elastic properties of CNT/GS reinforced nanocomposite are given in Tables (6 &7). It can be
observed from Tables 6 & 7 that axial modulus of CNT- or GS- reinforced nanocomposite is
increased as compared to that of pure matrix material, for perfect and imperfect bonding
conditions as well as for all matrix materials. Among the considered matrices, the maximum
and minimum enhancements, as compared to pure matrix material, in axial modulus of
nanocomposite are found for PE (i.e., 564.67 %) and iron (i.e., 7.48 %) matrix materials,
respectively, for perfect bonding condition. For better visualization of the effect of bonding
(perfect and imperfect) on axial modulus, the FEM results are also plotted in the form of bar
chart in Fig. 10. It can be noticed from Fig. 10 that imperfect bonding results in a loss of axial
modulus of metal matrix nanocomposite; whereas, in the case of PE nanocomposite, increased
stiffness of the interphase zone, as compared to pure PE matrix, results in positive effect on
the axial modulus of the PE nanocomposite.

Other (than axial) stiffness properties of CNT- and GS-reinforced nanocomposites, with and
without perfect bonding, are also shown in Tables 6 and 7, respectively. It can be noted from
Tables 6 and 7 that CNT- and GS-reinforced nanocomposites show transversely isotropic and
orthotropic behavior, respectively. Further, GS reinforcement is found to provide better in-
plane stiffness properties (i.e., and ) than CNT reinforcement, but poorer out-of-plane
stiffness properties (i.e., , & ), irrespective of matrix material. Better in-plane
stiffness properties in the case of GS reinforcement than CNT can be attributed to the 2-D
planar geometry with good biaxial stiffness properties of GS.

Further, the imperfect bonding causes a loss in other stiffness properties of CNT as well as GS
reinforced metal matrix nanocomposite, as shown in Tables 6 and 7. This loss is more
prevalent for out-of-plane stiffness properties of GS nanocomposite than CNT nanocomposite.
For instance, in the cases of CNT- and GS-iron nanocomposites, imperfect bonding leads to
almost 17.24 and 48.37 % reductions in transverse shear modulus ( ), respectively, as
compared to the perfectly bonded CNT- and GS-iron nanocomposites, whereas reductions in
in-plane shear modulus (i.e., ) are found to be only 13.56 and 7.42 % for CNT- and GS-
iron nanocomposites, respectively. It is important, however, to note from Tables 6 and 7 that
in the case of imperfectly bonded PE nanocomposite, the more elastic modulus of interphase
than the corresponding matrix material causes enhanced stiffness properties of the resulting
nanocomposite, as compared to the perfectly bonded nanocomposite, for both kind of
reinforcements. For instance, enhancements in the transverse shear modulus ( ) of
imperfectly bonded CNT- and GS-iron nanocomposites, as compared to the perfectly bonded
nanocomposite, are obtained as 2.83 and 3.18 % respectively; whereas, the obtained values for
in-plane shear modulus (i.e., ) are 2.55 and 8.69 %, respectively. From Tables 6 & 7, it can
also be concluded that among all the stiffness properties, the highest enhancement, in
comparison to the matrix material, is obtained in axial modulus, for CNT as well as GS
nanocomposite, and both types of bonding (i.e., perfect and imperfect).

300
Matrix material
CNT/GS nanocomposite perfect bonding
CNT nanocomposite imperfect bonding
250
GS nanocomposite imperfect bonding
Axial elastic modulus (GPa)

200

150

100

50

0
Au Al Cu Fe Mg PE
Matrix materials
Fig. 10. Comparison of FEM results of axial elastic modulus of 2% CNT- or GS-reinforced nanocomposite, with
and without perfect bonding, with that of the corresponding matrix material.

7. Conclusions

In this paper, an initial study is conducted to characterize a hypothetical interphase material


assumed to be filled in the region between nanofillers (i.e., CNT/GS) and different matrix
materials for its thickness and elastic modulus, using cohesive zone model. Thereafter, an
investigation is carried to study the effects of perfect and imperfect (i.e., with and without
interphase material) bonding on the elastic behaviour of CNT/GS reinforced nanocomposite,
using the method of representative volume element, for different matrix materials, including
metal and polymer matrices.
Based on the above results and discussions, the following important conclusions can be
drawn.

 Thickness of the interphase zone is found to vary with the matrix material of the
nanocomposite which is in contrast to the general assumption of constant thickness of
interphase zone for different matrix materials made by many researchers.
 In the case of nanocomposite with small radius CNTs, slightly higher, noticeable only
at narrow scale, elastic modulus of interphase zone is obtained than the large radius
CNTs, irrespective of matrix materials.
 The elastic modulus of interphase zone of a nanocomposite is significantly affected by
the matrix material, and for a given matrix material, the value of elastic modulus of
interphase zone will depend of the intermolecular potential parameters of interphase
(i.e., and ) and the matrix volume density (i.e., ).
 As compared to CNT nanocomposite, GS nanocomposite possess lower out-of-plane
stiffness properties, whereas in-plane stiffness properties are higher, irrespective of
matrix material, and type of bonding between nanofiller and matrix material.
 Interphase (i.e., imperfect bonding) in CNT- as well as GS-reinforced metal-matrix
nanocomposite results in loss of all stiffness properties, and this loss is more prevalent
for out-of-plane stiffness properties of GS nanocomposite.
 In the case of imperfectly bonded PE nanocomposite, the more elastic modulus of
interphase than the corresponding matrix material causes enhanced stiffness properties
of the resulting nanocomposite, as compared to the perfectly bonded nanocomposite,
for both kind of reinforcements.
Table 6
Comparison of elastic moduli (in GPa) of CNT-reinforced nanocomposite with perfect and imperfect bonding between CNT and different matrices ( = 0.02).
CNT Au Al Cu Fe Mg PE

P* Im** P Im P Im P Im P Im P Im

1 96.4412 95.8055 88.6001 87.6221 147.4010 145.5303 226.7808 222.2645 64.0992 63.3601 23.3319 23.3321
= 3 83.1481 79.2719 74.8482 69.2042 136.6482 125.0098 218.9716 192.9207 48.7511 43.6098 3.9712 4.0561
= 32.4546 30.8597 29.2128 26.6464 53.2349 48.0471 84.9617 73.4361 18.9799 16.4960 1.4691 1.5065
32.1902 29.7901 28.9426 25.6272 53.1266 46.2533 85.2380 70.5419 18.7360 15.9136 1.4687 1.5101
= 0.2999 0.2999 0.2999 0.2999 0.2999 0.2999 0.2999 0.2999 0.3000 0.3001 0.2999 0.2999
0.3390 0.3512 0.3419 0.3527 0.3265 0.3381 0.3163 0.3210 0.3540 0.3542 0.4074 0.4062
*
Perfect bonding
**
Imperfect bonding

Table 7
Comparison of elastic moduli (in GPa) of GS-reinforced nanocomposite with perfect and imperfect bonding between GS and different matrices ( = 0.02).
GS Au Al Cu Fe Mg PE

P* Im** P Im P Im P Im P Im P Im

1 96.4424 94.6799 88.5975 86.5051 147.3994 143.6256 226.7787 218.2742 64.1003 62.6351 23.3320 23.5607
91.1224 86.8620 82.8071 76.0488 144.0474 134.2668 224.8486 206.3365 56.2283 48.3737 5.4623 5.9061
82.2826 75.9515 74.0863 62.9271 135.2184 106.6338 217.0469 130.6120 48.3142 36.7057 3.9224 4.0367
35.4866 33.9639 32.3166 29.9351 55.7073 52.3631 86.6692 80.2369 22.1815 19.2071 2.3307 2.5332
30.5985 27.8458 27.4652 22.8134 50.9337 39.0684 82.5153 49.7453 17.6671 13.1906 1.3362 1.3749
30.6896 27.4615 27.7108 22.0125 51.3542 36.8412 82.7321 42.7131 17.8298 12.3428 1.3410 1.3836
0.2999 0.2999 0.2999 0.2999 0.2999 0.3000 0.2999 0.3000 0.2999 0.2999 0.2999 0.2999
0.2999 0.2999 0.2999 0.2999 0.2999 0.3000 0.2999 0.3000 0.2999 0.2999 0.2999 0.2999
0.3068 0.3070 0.3080 0.3087 0.3028 0.3016 0.3011 0.2974 0.3151 0.3174 0.3915 0.3914
*
Perfect bonding
**
Imperfect bonding
Acknowledgements

The present research work is carried out by utilizing the computation research facilities at the
Material Research Center (MRC), Malaviya National Institute of Technology (MNIT) Jaipur.
The authors would like to kindly acknowledge the MRC, MNIT Jaipur.

References

[1] S. Iijima, Helical microtubules of graphitic carbon, Nature. 354 (1991) 56–58.
[2] A.K. Geim, K.S. Novoselov, The rise of graphene, Nature. 6 (2007) 183–191.
[3] S. Stankovich, D.A. Dikin, G.H.B. Dommett, K.M. Kohlhaas, E.J. Zimney, E.A. Stach,
R.D. Piner, S.T. Nguyen, R.S. Ruoff, Graphene-based composite materials., Nature. 442
(2006) 282–286.
[4] W. Choi, J. Lee, Graphene synthesis and applications, CRC Press Taylor & Francis
Group, 2012.
[5] H.J. Choi, J.H. Shin, D.H. Bae, The effect of milling conditions on microstructures and
mechanical properties of Al / MWCNT composites, Compos. Part A. 43 (2012) 1061–
1072.
[6] A.M.K. Esawi, K. Morsi, A. Sayed, M. Taher, S. Lanka, Effect of carbon nanotube
(CNT) content on the mechanical properties of CNT-reinforced aluminium composites,
Compos. Sci. Technol. 70 (2010) 2237–2241.
[7] M. Hashempour, A. Vicenzo, F. Zhao, M. Bestetti, Effects of CVD direct growth of
carbon nanotubes and nanofibers on microstructure and electrochemical corrosion
behavior of 316 stainless steel, Mater. Charact. 92 (2014) 64–76.
[8] D. Poirier, R. Gauvin, R.A.L. Drew, Structural characterization of a mechanically milled
carbon nanotube / aluminum mixture, Compos. Part A. 40 (2009) 1482–1489.
[9] A. Al-ostaz, G. Pal, P.R. Mantena, A. Cheng, Molecular dynamics simulation of
SWCNT – polymer nanocomposite and its constituents, J. Mater. Sci. 43 (2008) 164–
173.
[10] J. Cho, C.T. Sun, A molecular dynamics simulation study of inclusion size effect on
polymeric nanocomposites, Comput. Mater. Sci. 41 (2007) 54–62.
[11] Y.J. Liu, X.L. Chen, Continuum models of carbon nanotube-based composites using the
boundary element method, Electron. J. Bound. Elem. 1 (2003) 316–335.
[12] Y.J. Liu, X.L. Chen, Evaluations of the effective material properties of carbon nanotube-
based composites using a nanoscale representative volume element, Mech. Mater. 35
(2003) 69–81.
[13] X.L. Chen, Y.J. Liu, Square representative volume elements for evaluating the effective
material properties of carbon nanotube-based composites, Comput. Mater. Sci. 29
(2004) 1–11.
[14] P. Joshi, S.H. Upadhyay, Evaluation of elastic properties of multi walled carbon
nanotube reinforced composite, Comput. Mater. Sci. 81 (2014) 332–338.
[15] U.A. Joshi, P. Joshi, S.P. Harsha, S.C. Sharma, Evaluation of the mechanical properties
of CNT based composites using hexagonal RVE, J. Nanotechnol. Eng. Med. 1 (2010) 1–
7.
[16] D. Kumar, A. Srivastava, Elastic properties of CNT- and graphene-reinforced
nanocomposites using RVE, Steel Compos. Struct. 21 (2016) 1085–1103.
[17] G.I. Giannopoulos, I.G. Kallivokas, Mechanical properties of graphene based
nanocomposites incorporating a hybrid interphase, Finite Elem. Anal. Des. 90 (2014)
31–40.
[18] S. Herasati, L. Zhang, Interphase effect on the macroscopic elastic properties of non-
bonded single-walled carbon nanotube composites, Compos. Part B Eng. 77 (2015) 52–
58.
[19] Y. Zare, Effects of interphase on tensile strength of polymer/CNT nanocomposites by
Kelly-Tyson theory, Mech. Mater. 85 (2015) 1–6.
[20] H.L. Duan, X. Yi, Z.P. Huang, J. Wang, A unified scheme for prediction of effective
moduli of multiphase composites with interface effects. Part I: Theoretical framework,
Mech. Mater. 39 (2007) 81–93.
[21] P. Joshi, S.H. Upadhyay, Effect of interphase on elastic behavior of multiwalled carbon
nanotube reinforced composite, Comput. Mater. Sci. 87 (2014) 267–273.
[22] M.M. Shokrieh, R. Rafiee, Stochastic multi-scale modeling of CNT/polymer
composites, Comput. Mater. Sci. 50 (2010) 437–446.
[23] J.L. Tsai, S.H. Tzeng, Y.T. Chiu, Characterizing elastic properties of carbon
nanotubes/polyimide nanocomposites using multi-scale simulation, Compos. Part B. 41
(2010) 106–115.
[24] S.J. V Frankland, V.M. Harik, G.M. Odegard, D.W. Brenner, T.S. Gates, The stress–
strain behavior of polymer–nanotube composites from molecular dynamics simulation,
Compos. Sci. Technol. 63 (2003) 1655–1661.
[25] M.M. Shokrieh, R. Rafiee, On the tensile behavior of an embedded carbon nanotube in
polymer matrix with non-bonded interphase region, Compos. Struct. 92 (2010) 647–652.
[26] J.M. Wernik, B.J. Cornwell-Mott, S. A. Meguid, Determination of the interfacial
properties of carbon nanotube reinforced polymer composites using atomistic-based
continuum model, Int. J. Solids Struct. 49 (2012) 1852–1863.
[27] Y.J. Liu, N. Nishimura, D. Qian, N. Adachi, Y. Otani, V. Mokashi, A boundary element
method for the analysis of CNT/polymer composites with a cohesive interface model
based on molecular dynamics, Eng. Anal. Bound. Elem. 32 (2008) 299–308.
[28] E. Mohammadpour, M. Awang, S. Kakooei, H.M. Akil, Modeling the tensile stress-
strain response of carbon nanotube/polypropylene nanocomposites using nonlinear
representative volume element, Mater. Des. 58 (2014) 36–42.
[29] A. Parashar, P. Mertiny, Representative volume element to estimate buckling behavior
of graphene/polymer nanocomposite., Nanoscale Res. Lett. 7 (2012) 515.
[30] A.P. Awasthi, D.C. Lagoudas, D.C. Hammerand, Modeling of graphene-polymer
interfacial mechanical behavior using molecular dynamics, Model. Simul. Mater. Sci.
Eng. 17 (2009) 15002.
[31] L.Y. Jiang, Y. Huang, H. Jiang, G. Ravichandran, H. Gao, K.C. Hwang, B. Liu, A
cohesive law for carbon nanotube/polymer interfaces based on the van der Waals force,
J. Mech. Phys. Solids. 54 (2006) 2436–2452.
[32] H. Tan, L.Y. Jiang, Y. Huang, B. Liu, K.C. Hwang, The effect of van der Waals-based
interface cohesive law on carbon nanotube-reinforced composite materials, Compos.
Sci. Technol. 67 (2007) 2941–2946.
[33] J. Zhao, J.W. Jiang, Y. Jia, W. Guo, T. Rabczuk, A theoretical analysis of cohesive
energy between carbon nanotubes, graphene and substrates, Carbon 57 (2013) 108–119.
[34] Y. Zhang, J. Zhao, Y. Jia, T. Mabrouki, Y. Gong, N. Wei, T. Rabczuk, An analytical
solution on interface debonding for large diameter carbon nanotube-reinforced
composite with functionally graded variation interphase, Compos. Struct. 104 (2013)
261–269.
[35] M.R. Ayatollahi, S. Shadlou, M.M. Shokrieh, Multiscale modeling for mechanical
properties of carbon nanotube reinforced nanocomposites subjected to different types of
loading, Compos. Struct. 93 (2011) 2250–2259.
[36] D.C. Hammerand, G.D. Seidel, D.C. Lagoudas, Computational micromechanics of
clustering and interphase effects in carbon nanotube composites, Mech. Adv. Mater.
Struct. 14 (2007) 277–294.
[37] A. Hernández-Pérez, F. Avilés, Modeling the influence of interphase on the elastic
properties of carbon nanotube composites, Comput. Mater. Sci. 47 (2010) 926–933.
[38] Z. Yuan, Z. Lu, M. Chen, Z. Yang, F. Xie, Interfacial properties of carboxylic acid
functionalized CNT/Polyethylene composites: A molecular dynamics simulation study,
Appl. Surf. Sci. 351 (2015) 1043–1052.
[39] A.R. Alian, S.I. Kundalwal, S. A. Meguid, Interfacial and mechanical properties of
epoxy nanocomposites using different multiscale modeling schemes, Compos. Struct.
131 (2015) 545–555.
[40] J. Fan, Mutiscale analysis of deformation and failure of materials, John Wiley & Sons
Ltd., 2011.
[41] A. White, Intermolecular potentials of mixed systems : Testing the Lorentz-Berthelot
mixing rules with ab initio calculations, 2000.
[42] D. Boda, D. Henderson, The effects of deviations from Lorentz–Berthelot rules on the
properties of a simple mixture, Mol. Phys. An Int. J. Interface Between Chem. Phys. 106
(2008) 2367–2370.
[43] X. Zhou, S. Song, L. Li, R. Zhang, Molecular dynamics simulation for mechanical
properties of magnesium matrix composites reinforced with nickel-coated single-walled
carbon nanotubes, J. Compos. Mater. (2015) 1–10.
[44] B.K. Choi, G.H. Yoon, S. Lee, Molecular dynamics studies of CNT-reinforced
aluminum composites under uniaxial tensile loading, Compos. Part B. 91 (2016) 119–
125.
[45] N. Silvestre, B. Faria, J.N. Canongia Lopes, Compressive behavior of CNT-reinforced
aluminum composites using molecular dynamics, Compos. Sci. Technol. 90 (2014) 16–
24.
[46] Y. Yani, Molecular dynamics simulation of nanocomposite materials, Iowa State
University, 2009.
[47] R. Rezaei, M. Shariati, H. Tavakoli-Anbaran, C. Deng, Mechanical characteristics of
CNT-reinforced metallic glass nanocomposites by molecular dynamics simulations,
Comput. Mater. Sci. 119 (2016) 19–26.
[48] C.T. Lu, A. Weerasinghe, D. Maroudas, A. Ramasubramaniam, A Comparison of the
elastic properties of graphene- and fullerene-reinforced polymer composites: The role of
filler morphology and size, Sci. Rep. 6 (2016) 31735.
[49] A. Sears, R.C. Batra, Macroscopic properties of carbon nanotubes from molecular-
mechanics simulations, Phys. Rev. B. 69 (2004) 235406.
[50] U.A. Joshi, S.C. Sharma, S.P. Harsha, A multiscale approach for estimating the chirality
effects in carbon nanotube reinforced composites, Phys. E Low-Dimensional Syst.
Nanostructures. 45 (2012) 28–35.
[51] S. Paunikar, S. Kumar, Effect of CNT waviness on the effective mechanical properties
of long and short CNT reinforced composites, Comput. Mater. Sci. 95 (2014) 21–28.
[52] C.T. Sun, R.S. Vaidya, Prediction of composite properties from a representative volume
element, Compos. Sci. Technol. 56 (1996) 171–179.
[53] W.B. Lu, J. Wu, J. Song, K.C. Hwang, L.Y. Jiang, Y. Huang, A cohesive law for
interfaces between multi-wall carbon nanotubes and polymers due to the van der Waals
interactions, Comput. Methods Appl. Mech. Eng. 197 (2008) 3261–3267.
[54] M.M. Shokrieh, R. Rafiee, Prediction of mechanical properties of an embedded carbon
nanotube in polymer matrix based on developing an equivalent long fiber, Mech. Res.
Commun. 37 (2010) 235–240.
[55] T.H. Nam, K. Goto, Y. Yamaguchi, E.V.A. Premalal, Y. Shimamura, Y. Inoue, K.
Naito, S. Ogihara, Effects of CNT diameter on mechanical properties of aligned CNT
sheets and composites, Compos. Part A Appl. Sci. Manuf. 76 (2015) 289–298.
[56] Y. Zhang, X. Zhuang, J. Muthu, T. Mabrouki, M. Fontaine, Y. Gong, T. Rabczuk, Load
transfer of graphene / carbon nanotube / polyethylene hybrid nanocomposite by
molecular dynamics simulation, Compos. Part B. 63 (2014) 27–33.
[57] R. Rafiee, R. Pourazizi, Influence of CNT functionalization on the interphase region
between CNT and polymer, Comput. Mater. Sci. 96 (2015) 573–578.
[58] T. Stern, A. Teishev, G. Marom, Composites of polyethylene reinforced with chopped
polyethylene fibers: Effect of transcrystalline interphase, Compos. Sci. Technol. 57
(1997) 1009–1015.
[59] T. Kunanopparat, P. Menut, M.H. Morel, S. Guilbert, Plasticized wheat gluten
reinforcement with natural fibers: Effect of thermal treatment on the fiber/matrix
adhesion, Compos. Part A Appl. Sci. Manuf. 39 (2008) 1787–1792.
[60] A. Srivastava, D. Kumar, Postbuckling of nanocomposite plate reinforced with
randomly oriented and dispersed cnts modeled through RSA technique, Int. J. Multiscale
Comput. Eng. 14 (2016) 585–606.
[61] S.H. Yang, Z.X. Wei, Molecular dynamics study of mechanical properties of carbon
nanotube-embedded gold composites, Phys. B. 403 (2008) 559–563.
[62] Z.Y.L.B.L. Xiao, W.G.W.Z.Y. Ma, Effect of Carbon nanotube orientation on
mechanical properties and thermal expansion coefficient of carbon nanotube-reinforced
aluminum matrix composites, Acta Met. Sin. (Engl. Lett.). 27 (2014) 901–908.
[63] S. Bashirvand, A. Montazeri, New aspects on the metal reinforcement by carbon
nanofillers: A molecular dynamics study, Mater. Des. 91 (2016) 306–313.
[64] W.M. Daoush, B.K. Lim, C.B. Mo, D.H. Nam, S.H. Hong, Electrical and mechanical
properties of carbon nanotube reinforced copper nanocomposites fabricated by
electroless deposition process, Mater. Sci. Eng. A. 514 (2009) 247–253.
[65] J.Y. Suh, D.H. Bae, Mechanical properties of Fe-based composites reinforced with
multi-walled carbon nanotubes, Mater. Sci. Eng. A. 582 (2013) 321–325.
[66] E. Carreño-Morelli, J. Yang, E. Couteau, K. Hernadi, J.W. Seo, C. Bonjour, L. Forró, R.
Schaller, Carbon nanotube/magnesium composites, Phys. Status Solidi. 201 (2004) R53-
54.

Você também pode gostar