Vous êtes sur la page 1sur 42

Multiscale modeling of microscale fiber reinforced composites with

nano-engineered interphases
S. I. Kundalwala,*, S. Kumar a,b,# and B. L. Wardle c
a

Department of Mechanical and Materials Engineering, Masdar Institute of Science and Technology, Abu
Dhabi 54224, UAE
b
Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge,
MA 02139-4307, USA
c
Department of Aeronautics and Astronautics, Massachusetts Institute of Technology, Cambridge,
MA 02139, USA

ABSTRACT
This study is focused on the mechanical properties and stress transfer behavior of multiscale
composite containing nano- and micro-scale fillers. A novel concept has been proposed to
exploit the remarkable mechanical properties of carbon nanotubes (CNTs) to improve the stress
transfer through the interphases, enabling their additional functionalities not available otherwise
at the microscale. The distinctive feature of construction of this composite is such that CNTs are
dispersed around the microscale fiber to modify fiber-matrix interfacial adhesion. Accordingly,
models are developed for hybrid composites. First, molecular dynamics simulations in
conjunction with the Mori-Tanaka method are used to determine the effective elastic properties
of nano-engineered interphase layer comprised of CNT bundles and epoxy. Subsequently, a
micromechanical pull-out model is developed for the resulting multiscale composite and its
stress transfer behavior is studied for different orientations of CNT bundles. The current pull-out
model accounts for the radial as well as the axial deformations of the different orthotropic
constituent phases of multiscale composite. The results from the developed pull-out model are
also compared with those of the finite element analyses and are found to be in good agreement.
Our results reveal that the stress transfer characteristics of the multiscale composite is
significantly improved by dispersing CNS around the fiber, particularly, when CNT bundles are
aligned along the axial direction of the fiber.
Keywords: Multiscale composites, Molecular dynamics, Micromechanics, Stress transfer
*

Banting Fellow, presently working at Mechanics and Aerospace Design Laboratory, Department
of Mechanical and Industrial Engineering, University of Toronto, Toronto, Canada

Prepared Manuscript

1. Introduction
The structural performance of a composite under service load is largely affected by the
fiber-matrix interfacial properties. Good interfacial properties are essential to ensure efficient
load transfer from matrix to the reinforcing fibers, which help to reduce stress concentrations and
improve overall mechanical properties of a resulting composite (Zhang et al., 2012). Several
experimental and analytical techniques have been developed thus far to gain insights into the
basic mechanisms dominating the fiber-matrix interfacial characteristics. The former is
considered to be inadequate to capture the physics of stress transfer through the fiber-matrix
interface. The extent and efficiency of stress transfer rests on two facts: (i) efficient stress
transfer from matrix to fiber utilizing the high mechanical properties of fibers in composites and
(ii) toughness of the resulting composite is undeniably dependent on the nature of fiber-matrix
interface. To characterize these issues, the pull-out test is proved to be a very efficient method to
study the stress transfer characteristics of a composite. A number of analytical and computational
two- and three-cylinder pull-out models have been developed to better understand the stress
transfer mechanisms across the fiber-matrix interface (Kim et al., 1992; Kim et al., 1994; Tsai
and Kim, 1996; Quek and Yue, 1997; Fu et al., 2000; Fu and Lauke, 2000; Banholzer et al.,
2005; Meng and Wang, 2015; Upadhyaya and Kumar, 2015). These two models differ in terms
of whether the interphase/interlayer between the fiber and matrix is considered or not. In case of
three-cylinder pull-out model, a thin layer of interphase, formed as a result of physical and
chemical interactions between the fiber and the matrix, is considered. The chemical composition
of such an interphase differ from both the fiber and matrix materials but its mechanical
properties lie between that of the fiber and the matrix (Drzal, 1986; Sottos et al., 1992;
Kundalwal and Meguid, 2015), and such nanoscale interphase has a marginal influence on the
bulk elastic properties of a composite. On the other hand, a relatively thick interphase can be
utilized between the fiber and matrix, especially a third phase made of different material than the
main constituent phases. Such microscale interphase (hereinafter the interlayer) strongly
influences the mechanical and interfacial properties of a composite, where the apparent
reinforcing effect is related to the cooperation of the strong interfacial adhesion strength and the
interlayer serving to inhibit crack propagation or as mechanical damping elements [see Zhang et
al. (2010) and the references therein].

Recently, graphene and CNTs attracted intense research interest because of their
remarkable electro-thermo-mechanical properties, which make them ideal candidates as nanofillers in composite materials (Pal and Kumar, 2015; Cui et al., 2015). Extensive research has
been dedicated to the introduction of graphene and CNTs as the modifiers to the conventional
composites in order to enhance their multifunctional properties (see Fig. 1). For example,
Bekyarova et al. (2007) reported an approach to the development of advanced structural
composites based on engineered multiscale CNT-microscale fiber reinforcement. The CNTcarbon fabric-epoxy composites showed 30% enhancement of the interlaminar shear strength

as compared to that of microscale fiber-epoxy composites. Cho et al. (2007) modified the epoxy
matrix in microscale fiber-epoxy composites with graphite nanoplatelets and reported the
improved in-plane shear properties and compressive strength for the resulting hybrid composite.
Wardle and coauthors (2008; 2012; 2014) grew CNTs on the circumferential surfaces of
microfibers to modify the fiber-matrix interfacial adhesion and reported the improvement in
composite delamination resistance, toughness, Mode I fracture toughness, interlaminar shear
strength, matrix-dominated elastic properties and electrical conductivity. Hung et al. (2009)
fabricated unidirectional composite in which CNTs were directly grown on the circumferential
surfaces of conventional microscale fibers. Compared with the conventional composite, their
tensile test results revealed that the multiscale composite is added with three new material
properties: the CNT joint strength, the CNT-matrix bonding strength, and the CNT tensile
strength. These strengths dictate the failure behavior and the capability of energy dissipation.
Davis et al. (2010) fabricated the carbon fiber reinforced composite incorporating the
functionalized CNTs; as a consequence, they observed significant improvements in the
mechanical properties of tensile strength, stiffness and resistance to failure due to cyclic loadings
resulted for the carbon fiber reinforced epoxy composite. Zhang et al. (2010) deposited CNTs on
the circumferential surfaces of electrically insulated glass fiber surfaces. According to their
fragmentation test results, the incorporation of an interlayer with a small number of CNTs
around the fiber, remarkably improved the interfacial shear strength of the fiber-epoxy
composite. They also found that the semiconductive interlayer results in a high sensitivity of the
electrical resistance to the tensile strain of single glass fiber model composites. The
functionalized CNTs were incorporated by Davis et al. (2011) at the fiber/fabricmatrix
interfaces of a carbon fiber reinforced epoxy composite laminate material; their study showed

improvements in the tensile strength and stiffness, and resistance to tensiontension fatigue
damage due to the created CNT reinforced region at the fiber/fabricmatrix interfaces. A
numerical method is proposed by Jia et al. (2014) to theoretically investigate the pull-out of a
hybrid fiber coated with CNTs. They developed two-step finite element (FE) approach: a single
CNT pull-out from the matrix at microscale and the pull-out of the hybrid fiber at macroscale.
Their numerical results indicate that the apparent interfacial shear strength of the hybrid fiber and
the specific pull-out energy is significantly increased due to the additional bonding of the CNT
matrix interface. A beneficial interfacial effect of the presence of CNTs on the circumferential
surface of the microscale fiber samples is demonstrated by Jin et al. (2014) resulting in an
increase in the maximum interlaminar shear strength (>30 MPa) compared to uncoated samples.
This increase is attributed to an enhanced contact between the resin and the fibers due to an
increased surface area as a result of the CNTs. To improve the interfacial properties of
microscale fiber-epoxy composites, Chen et al. (2015) introduced a gradient interlayer reinforced
by graphene sheets between microscale fibers and matrix using a liquid phase deposition
strategy; due to the formation of this gradient interlayer, 28.3% enhancement in interlaminar
shear strength of unidirectional microscale fiber-epoxy composites is observed with 1 wt%
loading of graphene sheets. Recently, two types of morphologies are investigated by Romanov et
al. (2015): CNTs grown on fibers and CNTs deposited in fiber coatings. The difference in the
two cases is the orientation of CNTs near the fiber interface: radial for grown CNTs and tangent
for CNTs in the coatings.
Suggestive findings in the literature indicate that the use of nano-fillers and conventional
fibers together, as multiscale reinforcements, significantly improves the overall properties of
multiscale composites, which are unseen in conventional composites. As is well known, damage
initiation is progressive with the applied load and that the small crack at the fiber-matrix
interface may reduce the fatigue life of composites. By strengthening the interfacial fiber-matrix
region with nano-fillers, we can increase the damage initiation threshold and long-term reliability
of conventional composites. This concept can be utilized to grade the matrix properties around
the microscale fiber, which may eventually improve the stress transfer behaviour of the
multiscale composite. To the best of our knowledge, there has been no pull-out model to study
the stress transfer characteristics of multiscale composite containing transversely isotropic nanoand micro-scale fillers. This is indeed the motivation behind the current study. The current study

is devoted to the development of a pull-out model for analyzing the stress transfer characteristics
of multiscale composite. CNT bundles are considered as a special case of nanostructures
embedded between the fiber and matrix; the resulting intermediate phase, containing carbon
nanostructures (CNS) and epoxy, is considered as an interlayer. First, we carried out multiscale
study to determine the transversely isotropic elastic properties of an interlayer through MD
simulations in conjunction with the Mori-Tanaka model. Then the determined elastic moduli of
the interlayer are used in the development of three-phase pull-out model. Particular attention is
paid to investigate the effect of orientations of CNS on the stress transfer characteristics of
multiscale composite.

2. Multiscale modeling
For most multiscale composites whose mechanical response and fracture behavior arise
from the properties of the individual constituents at each level as well as from the interaction
between these constituents across the different length scales. As a consequence, different
multiscale modeling techniques have been developed over the last decade to predict the
properties of composites at the microscale level (Tsai et al., 2010; Yang et al., 2012; Alian et al.,
2015a,b). Here, multiscale modeling of a composite is achieved in two consecutive steps: (i)
elastic properties of the CNS comprised of a bundle of CNTs and epoxy molecules are evaluated
using molecular dynamics (MD) simulations; (ii) Mori-Tanaka method is then used to calculate
the bulk effective properties of the nano-engineered interphase layer.

2.1 Molecular modeling


This section describes the procedure for building a series of MD models for the epoxy
and the CNS. The technique for creating an epoxy and CNS is described first, followed by the
MD simulations for determining the isotropic elastic properties of the epoxy material and the
transversely isotropic elastic properties of the CNS. All MD simulations runs are conducted with
large-scale atomic/molecular massively parallel simulator (LAMMPS; Plimpton, 1995). The
consistent valence force field (CVFF; Dauber-Osguthorpe, 1998) is used to describe the atomic
interactions between different atoms. The CVFF has been used by several researchers to model
the CNTs and their composite systems (Tunvir et al., 2008; Li et al., 2012; Kumar et al., 2014).
A very efficient conjugate gradient algorithm is used to minimize the strain energy as a function
of the displacement of the MD systems while the velocity Verlet algorithm is used to integrate

the equations of motion in all MD simulations. Periodic boundary conditions have been applied
to the MD unit cell faces. Determination of the elastic properties of the pure epoxy and the CNS
is accomplished by straining the MD unit cells followed by constant-strain energy minimization.
The averaged stress tensor of the MD unit cell is defined in the form of virial stress (Allen and
Tildesley, 1987); as follows
N

1
mi
= vi 2 + Fi ri
V
2
i=1

(1)

where V is the volume of the unit cell; vi , mi , ri and Fi are the velocity, mass, position and force
acting on the ith atom, respectively.

2.1.1 Modeling of EPON 862-DETDA epoxy


Thermosetting polymers are the matrices of choice for structural composites due to their
high stiffness, strength, creep resistance and thermal resistance when compared with
thermoplastic polymers (Pascault et al., 2002). These desirable properties stem from the threedimensional (3D) crosslinked structures of these polymers. Many thermosetting polymers are
formed by mixing a resin (epoxy, vinyl ester, or polyester) and a curing agent. We used epoxy
material based on EPON 862 resin and Diethylene Toluene Diamine (DETDA) curing agent to
form a crosslinked structure, which is typically used in the aerospace industry. The molecular
structures of these two monomers are shown in Fig. 2. To simulate the crosslinking process, the
potential reactive sites in the epoxy resin can be activated by hydrating the epoxy oxygen atoms
at the ends of the molecule, see Fig. 2(b). The EPON 862-DETDA weight ratio was set to 2:1 to
obtain the best elastic properties (Bandyopadhyay and Odegard, 2012). The initial MD unit cell,
consisting of both activated epoxy (100 molecules of EPON 862) and a curing agent (50
molecules of DETDA), was built using the PACKMOL software (Martnez et al., 2009), as
shown in Fig. 3. The polymerization process usually occurs into two main stages: pre-curing
equilibration and curing of the polymer network. These main steps involved in determining the
elastic moduli of pure epoxy are described as follows:
Step 1 (pre-curing equilibration): The initial MD unit cell was compressed gradually through
several steps from its initial size of 50 50 50 to the targeted dimensions of 39 39

39 . The details of the MD unit cell are summarized in Table 1. At each stage, the atoms

coordinates are remapped to fit inside the compressed box then a minimization simulation was

performed to relax the coordinates of the atoms. The system was considered to be optimized
once the change in the total potential energy of the system between subsequent steps is less than
1.010-10 kcal/mol (Alian et al., 2015b). The optimized system is then equilibrated at room
temperature in the constant temperature and volume canonical (NVT) ensemble over 100 ps by
using a time step of 1 fs.
Step 2 (curing): After the MD unit cell is fully equilibrated in step 1, the polymerization and
crosslinking is simulated by allowing chemical reactions between reactive atoms. Chemical
reactions are simulated in a stepwise manner using a criterion based on atomic distances and the
type of chemical primary or secondary amine reactions as described elsewhere (Bandyopadhyay
and Odegard, 2012). The distance between all pairs of reactive CN atoms are computed and
new bonds are found to be created between all those that fall within a pre-assigned cutoff
distance. We considered this distance to be 5.64 , four times the equilibrium CN bond length
(Varshney et al., 2008; Li and Strachan, 2010). After the new bonds are identified all new
additional covalent terms were created and hydrogen atoms from the reactive C and N atoms
were removed. Then several 50 ps isothermalisobaric (NPT) simulations are preformed until no
reactive pairs are found within the cut-off distance. At the end, the structure is again equilibrated
for 200 ps in the NVT ensemble at 300 K.
After the energy minimization process, the simulation box was volumetrically strained in
both tension and compression to determine the bulk modulus by applying equal strains in the
loading directions along all three axes; the bulk modulus (K) was calculated by:
h
K=
v

(2)

where v and h are the volumetric strain and the averaged hydrostatic stress, respectively. The

average shear modulus was determined by applying equal shear strains on the simulation box in
xy, xz, and yz planes; the shear modulus (G) was calculated by:
ij
G= , ij
ij

where ij and ij denote the averaged shear stress and shear strain, respectively.

(3)

In all simulations, strain increments of 0.25% have been applied along a particular

direction by uniformly deforming or shearing the simulation box and updating the atom
coordinates to fit within the new dimensions. After each strain increment, the MD unit cell was
equilibrated using the NVT ensemble at 300 K for 10 ps. It may be noted that the fluctuations in

the temperature and potential energy profiles are less than 1% when the system reached
equilibrium after about 5 ps (Haghighatpanah and Bolton, 2013) and several existing MD studies
(Frankland et al., 2003; Tsai et al., 2010; Haghighatpanah and Bolton, 2013; Alian et al., 2015b)
used 2 ps to 10 ps time step in their MD simulations to equilibrate the systems after each strain
increment. Then, the stress tensor is averaged over an interval of 10 ps to reduce the effect of
fluctuations. These steps have been repeated again in the subsequent strain increments until the
total strain reached up to 2.5%. Based on the calculated bulk and shear moduli, Youngs modulus
(E) and Poissons ratio () are determined as follows:
E=

9KG
3K + G

and =

3K 2G
2(3K + G)

(4)

The predicted elastic properties of the epoxy using MD simulations are summarized in Table 2
and Youngs modulus is found to be consistent with the experimentally measured modulus of a
similar epoxy (Morris, 2008).

2.1.2 MD simulations of CNS


Despite the great potential of applying CNTs in composite materials, an intrinsic
limitation in directly scaling up the remarkable elastic properties of CNTs to microscale due to
their poor dispersion, agglomeration and aggregation. It is difficult to uniformly disperse CNTs
in the matrix during the fabrication process and the situation becomes more challenging at high
CNT loadings. This is attributed to the fact that CNTs have a tendency to agglomerate and
aggregate into bundles due to their high surface energy and surface area (Dumlich et al., 2011).
Therefore, we consider the epoxy nanocomposite reinforced with CNT bundles, which is more
practical and realistic representation of embedded CNTs. The MD unit cell is constructed to
represent an epoxy nanocomposite containing bundle of thirteen CNTs, as shown in Fig. 4. The
initial distance between the adjacent CNTs in the bundle considered was 3.4 , which is the
intertube separation distance in multi-walled CNTs (Odegard et al., 2003; Qi et al., 2005; Tunvir
et al., 2008; Dumlich et al., 2011). The noncovalent bonded CNT-epoxy nanocomposite system
is considered herein, therefore, the interactions between the atoms of the embedded CNTs and
the surrounding epoxy are solely resulted from non-bonded interactions. These non-bonded
interactions between the atoms are represented by van der Waals (vdW) interactions and

Coulombic forces. The cut-off distance for the non-bonded interaction was set to 14.0
(Haghighatpanah and Bolton, 2013). The unit call is assumed to be transversely isotropic with
the 1axis being the axis of symmetry; therefore, only five independent elastic coefficients are
required to define the elastic stiffness matrix. The cylindrical molecular structure of the CNT is
treated as an equivalent solid cylindrical fiber (Odegard et al., 2003; Tsai et al., 2010) for
determining its volume fraction in the CNS (Frankland et al., 2003),
vCNT

R CNT +

hvdW 2
2 NCNT LCNT
VCNS

(5)

where R CNT and LCNT denote the respective radius and length of a CNT; hvdW is the vdW

equilibrium distance between a CNT and the surrounding epoxy matrix; NCNT is the number of
CNTs in the bundle; and VCNS is the volume of the CNS.

The CNS is constructed by randomly placing the crosslinked epoxy structures around the

CNT bundle. The details of the CNS are summarized in Table 1. Five sets of boundary
conditions have been chosen to determine each of the five independent elastic constants such that
a single property can be independently determined for each boundary condition. The
displacements applied at the boundary of the CNS are summarized in Table 3; in which symbols
have usual meaning. To determine the five elastic constants, the CNS is subjected to five
different tests: longitudinal tension, transverse tension, in-plane tension, in-plane shear and out
of-plane shear. The steps involved in the MD simulations of the CNS are the same as adopted in
the case of pure epoxy. The first row of Table 4 presents the computed effective elastic
properties of the CNS through MD simulations. These properties of the CNS will be used as the
properties of nanoscale fiber in the micromechanical model to determine the effective elastic
moduli of the interlayer at the microscale level (see Fig. 5).

2.2 Effective elastic properties of CNS engineered interphase/interlayer


In this section, the elastic properties of the pure epoxy and the CNS obtained from the
MD simulations are used as an input in the Mori-Tanaka model in order to determine the bulk
elastic properties of the interlayer. Figure 5 shows the schematic cross-section of the three-phase
multiscale composite. Around the microscale fiber, the interlayer is considered to be made of
CNS and epoxy. Here, we consider three different cases of interlayer, in which: (i) CNS are
aligned along the direction of a microscale fiber, (ii) CNS are aligned radially to the axis of a

10

microscale fiber, and (iii) CNS are randomly dispersed. Considering the CNS as a fiber, the
Mori-Tanaka model (Mori and Tanaka, 1973) can be utilized to estimate the effective elastic
properties of the interlayer, as follows (Benveniste, 1987):

in which

1 vm [I] + vCNS A
1
[C] = [Cm ] + vCNS ([CCNS ] [Cm ]) A
1 = [I] + [S E ]([Cm ])1 ([CCNS ] [Cm ])
A

(6)

Where [CCNS ] and [Cm ] are the stiffness tensors of the CNS and the epoxy matrix, respectively;

[I] is an identity tensor; vCNS and vm represent the volume fractions of the CNS and the epoxy
matrix, respectively; and [S E ] indicates the Eshelby tensor. The specific forms of the Eshelby

tensor, for the first two cases, are explicitly given in Appendix A.

In case of epoxy reinforced with randomly oriented CNS in the three-dimensional space,
the following relation can be used to determine the effective elastic properties of the interlayer:
1

in which

2 vm [I] + vCNS A
2
[C] = [Cm ] + vCNS ([CCNS ] [Cm ]) A
2 = [I] + [S E ]([Cm ])1 ([CCNS ] [Cm ])
A

(7)

The terms enclosed with angle brackets in Eq. (7) represent the average value of the term over all
orientations defined by transformation from the local coordinate system of the CNS to the global
coordinate system. The transformed mechanical strain concentration tensor for the CNS with
respect to the global coordinates is given by
ijkl = t ip t jq t kr t ls Apqrs
A

(8)

where t ij are the direction cosines for the transformation and are given in Appendix A.

Consequently, the random orientation average of the dilute mechanical strain

2 can be determined by using the following relation (Marzari and


concentration tensor A
Ferrari, 1992):

2 =
A

0 0

where , , and are the Euler angles.

(, , ) sin ddd
A

0 0

sin ddd

(9)

11

3. Three-phase pull-out model


The axisymmetric RVE of the multiscale composite consists of a microscale fiber
embedded in a compliant matrix having a nano-engineered interphase between them illustrated in
Fig. 5. Each constituent of the composite is regarded as a transversely isotropic linear elastic
continuum material. The cylindrical coordinate system (rz) is considered in such a way that
the axis of the representative volume element (RVE) coincides with the zaxis. The RVE of a
multiscale composite has the radius c and the length L; the radius and the length of the
microscale fiber are denoted by a and Lf, respectively; the inner and outer radii of the interlayer
are a and b, respectively. Within the nano-engineered interphase, CNS are randomly or orderly
distributed in the epoxy matrix. Both interlayer and epoxy matrix are fixed at one end (z = 0) and
a tensile stress, p , is applied on the other end (z = L) of the embedded microscale fiber.

Analytical solutions are obtained for free boundary conditions at the external surface of the
matrix cylinder to model a single fiber pull-out problem. The mode of deformation is
axisymmetric and thus the stress components (r , , z and rz ) and the displacement

components (w, u) are in all three phases function of r and z only. For an axisymmetric
geometry, with cylindrical coordinates r, , and z, the governing equilibrium equations in terms
of normal and the shear stresses are given by:
(k)

(k)

(k)

(k)

r
r
rz
+
+
=0
r
z
r
(k)

(10)

(k)

z
1 (rrz )
+
=0
z
r r

(11)

while the relevant constitutive relations are


(k)

(k)

C11
C(k)

= 12
(k)
(k)
z C13
(k)
rz 0
r
(k)

(k)

(k)

C12

C13

(k)

C22

(k)

C23

(k)

C23

(k)

C33

0 (k)
r
(k)

; k = f, i and m
0 (k)
z

(k) (k)
C66 rz

(12)

in which the superscripts f, i and m denote the microscale fiber, the interlayer and the matrix,
(k)

(k)

respectively. For the kth constituent phase, z and r represent the normal stresses in the z and
(k)

(k)

(k)

r directions, respectively; z , and r are the normal strains along the z, and r, directions,
(k)

(k)

(k)

respectively; rz is the transverse shear stress, rz is the transverse shear strain and Cij are the

12

elastic constants. The strain-displacement relations for an axisymmetric problem relevant to this
RVE are
(k)
z

w (k) (k) u(k) (k) u(k)


w (k) u(k)
(k)
=
, =
, r =
and rz =
+
z
r
r
r
z

(13)

in which w (k) and u(k) represent the axial and radial displacements at any point of the kth phase

along the z and r directions, respectively. The interfacial traction continuity conditions are given
by
i
m
fr r=a = ir r=a ; frz r=a = irz r=a = 1 ; ir r=b, = m
r |r=b ; rz r=b = rz |r=b = 2 ;

w f r=a = w i r=a ; w i r=b = w m |r=b ; uf r=a = ui r=a and ui r=b = um |r=b

(14)

where 1 is the transverse shear stress at the interface between the microscale fiber and the

interlayer while 2 is the transverse shear stress at the interface between the interlayer and the
matrix. The average axial stresses in the different phases are defined as
a

1
fz = 2 fz 2r dr ;

a
0

1
1
iz =

iz 2r dr and
m
m
z =
z 2r dr
2
2
2
(b a )
(c b 2 )
a

(15)

Now, making use of Eqs. (10), (11), (14) and (15), it can be shown that

fz
2

iz
2a
2b

m
2b
z
= 1 ;
= 2

and
=

1
2
z
z
z
a
b a2
b 2 a2
c2 b2 2

(16)

Using the equilibrium equation given by Eq. (11), the transverse shear stresses in the interlayer
and the matrix can be expressed in terms of the interfacial shear stresses 1 and 2 , respectively,
as follows:

a
1
2a
2b
irz = 1 + (a2 r 2 ) 2

1
r
2r
b a2
b 2 a2 2
c2
b
c
m
=

r 2
2 +
rz
2
c b
r
r

(17)
(18)

13

Also, since the RVE is an axisymmetric problem, it is usually assumed (Nairn, 1997) that the
gradient of the radial displacements with respect to the zdirection is negligible and so, from the
constitutive relation given by Eq. (12) and the strain-displacement relations given by Eq. (13)
(k)

(k)

between rz and rz , we can write

1 (k)
w (k)
(k) rz ; i and m
r
C66

(19)

Solving Eq. (19) and enforcing the continuity conditions at r = a and r = b, respectively, the axial
displacements of the interlayer and the matrix along the zdirection can be derived as follows:

wm

w i = waf + A1 1 + A2 2
c
r
= waf + A3 1 + A4 2 + m ln and
C66
b
waf = w f r=a

(20)
(21)
(22)

in which Ai (i = 1, 2, 3 and 4) are the constants of the displacement fields and are explicitly given

in Appendix B. All other constants, Ai , obtained in the course of deriving the pull-out model
herein are also explicitly expressed in Appendix B.

The radial displacements in the three constituent phases can be assumed as (Hashin,
1964)
uf = C1 r, ui = Ai r +

Bi
C3
and um = C2 r +
r
r

(23)

where C1 , Ai , Bi , C2 and C3 are the unknown constants. Invoking the continuity conditions for

the radial displacement at the interfaces r = a and b, the radial displacement in the interlayer can
be augmented as follows:
a2
b2
b2
a2
1
a2
u = 2
r C1 2
r C2 2
r C3
b a2 r
b a2 r
b a2 r
i

(24)

Using displacement fields, constitutive relations and strain-displacement relations, the

expressions for the normal stresses can be written in terms of the unknown constants C1 , C2 and
C3 as follows:
f
fz = C11

waf
f
+ 2C12
C1
z

(25)

14

fr
iz

=
=

f
C12

fz

f
C11

i
C11
f
C11

fz

f
C33

f
2C12

f
C11

C1

i
f
i
i
i
2C12
a2 2C12
C11
2C12
b2
2C12
1
i
2
+

C
+
C
+
C
+
C
A
1
2
3
1
11
f
b a2
b 2 a2
b 2 a2
z
C11

i
+C11
A2

ir =

f
C23

i
C13

2
z

i
i
f
i
a2 b 2
a2
C13
C23
C33
b2
2C12
f

1
+

C1
z
f
f
b 2 a2 r 2
b 2 a2
r2
C11
C11

i
i
C33
C23
b 2 a2
b 2 a2

1
+
+ 1 C2
b 2 a2 r 2
b 2 a2 r 2

i
i
C23
a2
C33
a2
1
2
i
i
+ 2

1
+

+
1
C
+
C
A
+
C
A
3
1
2
13
13
b a2 r 2
b 2 a2 r 2
z
z

m
z =
m
r

m
C11
f
C11

m
C12
f
C11

fz

fz

f
m
2C12
C11
f
C11

f
m
2C12
C12
f
C11

m
m
C1 + 2C12
C2 + C11
A3

1
2
m
+ C11
A4
z
z

m
m )C
m
m
C1 + (C11
+ C12
2 + (C12 C11 )

(26)

(27)

(28)
(29)

C3
1
2
m
m
+ C12
A3
+ +C12
A4
(30)
2
r
z
z

Invoking the continuity conditions fr r=a = ir r=a and ir r=b = m


r |r=b , the following
equations for solving unknown constants C1 , C2 and C3 are obtained:
B11
B21
B31

B12
B22
B32

f
i
C12
C13
0
B13 C1
0
fz

1
2
i
m
B23 C2 = f C13 C12 + A5 A7
+ A 6 A 8
z
C11
m
m
B33 C3
A9
C12
A3 z
C12

(31)

The expressions of the coefficients Bij are presented in Appendix B. Solving Eq. (31), the

solutions of the constants C1 , C2 and C3 can be expressed as:


Ci = bi1
fz + bi2

1
2
+ bi3
; i = 1, 2, 3
z
z

(32)

The expressions of the coefficients bi1 , bi2 and bi3 are evident from Eq. (32), and are not shown
here for the sake of clarity. Now, making use of Eqs. (27), (29) and (32) in the last two equations

15

of (15), respectively, the average axial stresses in the interlayer and the matrix are written as
follows:
iz = A14

fz + A15

1
2
+ A16
z
z

fz + A18
m

z = A17

1
2
+ A19
z
z

(33)

(34)

Now, satisfying the equilibrium of force along the axial (z) direction at any transverse crosssection of the RVE, the following equation is obtained:
a2 p = a2
fz + (b2 a2 )
iz + (c 2 b2 )
m
z

where p is the pull-out stress applied on the fiber end.

(35)

Differentiating the first and last equations of (16) with respect to z, we have
a 2
c 2 b2 2
fz
m
z

and

=
2
2 z 2
2b
z 2

(36)

fz
2
fz
4
+A24 2 + A25
fz A26 p = 0
z 4
z

(37)

fz = A27 sinh(z) + A28 cosh(z) + A29 sinh(z) + A30 cosh(z) + (A26 /A25 )p

(38)

= 12 A24 + (A24 )2 4A25 and = 12 A24 (A24 )2 4A25

(39)

1 =

Using Eqs. (33 36), the governing equation for the average axial stress in the microscale fiber is
obtained as follows:

Solution of Eq. (37) is given by:

where

Substitution of Eq. (38) into the first equation of (16) yields the expression for the microscale
fiber/interlayer interfacial shear stress as follows:
a
1 = [A27 cosh(z) + A28 sinh(z) + A29 cosh(z) + A30 sinh(z)]
2

The stress boundary conditions of the model are given by

(40)

16

fz z=0 = 0,

z |z=0 =

fz z=L = p ,

a2 p
,
c2 b2

z |z=L = 0,

fz

=0
z z=0

and

m
z

=0
z z=0

and

fz

=0
z z=L

m
z

=0
z z=L

(41)
(42)

Utilizing the boundary conditions given by Eq. (41) in Eq. (38), the constants A27 , A28 , A29 and

A30 can be obtained. Similar governing equations for the average axial stress in the matrix and

the interlayer/matrix interfacial shear stress can be derived using the same solution methodology;
such that,
4
m
2
m
z
z
+A36
+ A37
m
z A38 p = 0
4
2
z
z

Solution of Eq. (43) is given by:

z = A39 sinh(m z) + A40 cosh(m z) + A41 sinh(m z) + A42 cosh(m z)


where

+(A38 /A37 )p

m = 12 A36 + (A36 )2 4A37 and m = 12 A36 (A36 )2 4A37

(43)

(44)
(45)

Substitution of Eq. (44) into the last equation of (16) yields the expression for the interlayer/bulk
matrix interfacial shear stress as follows:
2 =

c 2 b2
[A39 m cosh(m z) + A40 m sinh(m z) + A41 m cosh(m z) + A42 m sinh(m z)] (46)
2b

Utilizing the boundary conditions given by Eq. (42) in Eq. (44), the constants A39 , A40 , A41 and
A42 can be obtained.

4. Results and discussion


In this section, the results of the developed pull-out model are compared with the FE
results. Subsequently, the effect of orientation of CNT bundles and their loading on the stress
transfer behavior of the multiscale composite are analysed and discussed. Considering the epoxy
as the matrix phase and the CNS as the reinforcements, the effective elastic properties of the
nanocomposite have been determined by following the micromechanical modeling approach

17

developed in section 2.2. Note that the effective elastic properties of nano-engineered interlayer
are obtained by MD simulations in conjunction with the MT model and are summarized in Table
4. It should be noted that the CNT volume fraction used in the computations (see Table 4) is
either with respect to the CNS volume or interlayer volume. Unless otherwise stated: (i) the
geometrical parameters a, b and c of the pull-out model considered here are as 5 m, 6.5 m and
9 m, respectively; (ii) the value of L for the model as 100 m; and (iii) the CNS, with 5%
loading in the interlayer, are considered to be aligned with the axis of a microscale fiber. Also
the following transversely isotropic elastic properties of the carbon fiber (Honjo, 2007) are used
for both analytical and FE analyses presented herein.
C11 = 236.4 GPa, C12 = 10.6 GPa, C13 = C12 , C22 = 24.8 GPa, C33 = C22 , C44 = 7 GPa, C55 = 25 GPa,

and C66 = C55 .

4.1. Validation of the analytical model by FE method


The main novelty of the three-phase pull-out model derived in this study is that the radial
as well as the axial deformations of the different transversely isotropic constituent phases of the
multiscale composite have been taken into account. Therefore, it is desirable to justify the
validity of the pull-out model derived in section 3. For this purpose, 3D axisymmetric FE model
has been developed using the commercial software ANSYS 14.0. The geometry of the FE model,
and the loading and boundary conditions are selected to represent those of the actual
experimental test. Also the geometric and material properties used in the FE simulations are
identical to those of the analytical model. The microscale fiber, the interlayer and the epoxy
matrix are constructed and meshed with twenty-node solid elements SOLID186. Identical to the
analytical model, a uniformly distributed tensile stress (p ) is applied to the free end of the

microscale fiber along the zdirection at z = L. The boundary conditions are imposed in such
way that the bottom cylindrical surface of the matrix is fixed (z = 0) and the top cylindrical

surface of the matrix is traction-free. Moreover, the lateral cylindrical surface of the matrix is
assumed to be traction-free, to simulate single fiber pull-out problem. The case of fixed lateral
surface of matrix to approximately model a hexagonal array of fibers in the matrix is not
considered as the results are identical to that of the single fiber pull-out model as demonstrated
by Upadhyaya and Kumar (2015). Three- and two-phase FE models have been developed to

18

validate the analytical pull-out model. The three-phase model made of the microscale fiber, the
interlayer containing CNS aligned along the direction of a microscale fiber (see Fig. 5), and the
matrix; the two-phase model consists of only the microscale fiber and matrix. At first, a mesh
convergence study is performed to see the effect of element size on the stress transfer
characteristics. Once the convergence is assured, FE simulations are carried out to validate the
analytical model. Figure 6 shows the comparisons of the average axial stress in the microscale
fiber computed by the analytical pull-out model and the FE pull-out model. It may be observed
that the axial stress decreases monotonously from the pulled fiber end to the embedded fiber end.
Moreover, the results predicted by the analytical models are in great accordance with those of FE
simulations. As is well known, within the framework of linear theory of elasticity, singular stress
field may arise at the corners of biomaterial interfaces between different phases with different
elastic properties (Marotzke, 1994). FE results seem to indicate the existence of a singularity at
the square corner between the fiber and interlayer at the fiber entry. Therefore, in order to ensure
realistic stress conditions, i.e. an unrestricted variation of the stress field near the fiber entry, we
determined and compared the transverse shear stresses (irz ) in the interlayer over a distance of

0.02 fiber diameter away from the fiber-interlayer interface. This technique has been adopted by
several researchers to ascertain the accuracy of the analytical model (Marotzke, 1994;

Upadhyaya and Kumar (2015) and references therein). It may be observed from Fig. 7 that the
analytical model provides a good estimation of the transverse shear stress distribution. It may be
importantly observed from Fig. 7 that the incorporation of interlayer significantly reduces the
maximum value of irz compared to that of the two-phase model of composite. Our comparisons

clearly indicate that the derived pull-out model herein captures the crucial stress transfer
mechanisms of the multiscale composite; therefore, the subsequent results presented herein are
based on the analytical pull-out model.

4.2. Analytical modeling results


In this section, results of parametric study have been presented and the effects of
orientations of CNS in the interlayer on the stress transfer characteristics of the multiscale
composite are discussed. The maximum CNT volume fraction considered in the interlayer is 5%,
so the CNT volume fraction in the composite is much less than 5%. Three different cases have
been considered: (i) interlayer containing aligned CNS (parallel), (ii) interlayer containing CNS,

19

which are radially aligned to the axis of a microscale fiber (perpendicular), and (iii) interlayer
containing randomly dispersed CNS (random). These cases represent the practical situation of
distribution of CNS in the epoxy matrix and may significantly influence the overall properties of
both interlayer and multiscale composite. Figure 8 demonstrates that the orientations of CNS do
not much influence the average axial stress in the microscale fiber. On the other hand, the effect
of orientations of CNS is found to be significantly influence the interfacial shear stress (see Fig.
9, which shows 1 at the interface between microscale fiber and the interphase). It may be
observed that the maximum value of 1 is reduced by 15.5% for the parallel case compared to

homogenous interlayer. This is attributed to the fact that the axial elastic properties of the
interlayer are higher in the parallel case (see Table 4) and hence the interfacial properties of the

resulting multiscale composite improve in comparison to all other cases. It may also be observed
from Fig. 9 that the perpendicular case does not improve the interfacial properties of the
multiscale composite; but the random case marginally enhances the interfacial properties of the
multiscale composite over the pure matrix case. This is due to the fact that the axial elastic
properties of the interlayer in the perpendicular case are matrix dominant and hence its results
are found to be more close to those of the pure matrix case; in the random case, CNS are
homogeneously dispersed in the interlayer and hence its overall elastic properties improve in
comparison to the pure matrix case.
Figure 10 demonstrates the radial stress at the fiber-interlayer interface along the length
of the fiber for the different cases of orientation of CNS. Here, ri = 5 m and 6.5 m represent

the fiber-interlayer interface and the interlayer-matrix interface, respectively. It is evident from
this figure that the radial stress peaks at the loaded fiber end and decays rapidly with increasing
distance from it, changing from the tensile to compressive. This figure also demonstrates that the
magnitude is nearly equal at both the entry and exit ends of the fiber because of the small aspect
ratio of the microscale fiber. Radial stress at the interface at the fibre entry is tensile due to
Poissons contraction of the fibre, while at the exit end it is compressive due to Poissons
contraction of the matrix; this is consistent with the trend observed in several studies (Quek and
Yue, 1997; Upadhyaya and Kumar (2015) and references therein). Such a concentrated stress
distribution close to stress singular point makes the interface region very susceptible to mixed
mode fracture. Among all the cases shown in Fig. 10, the parallel case reduces the maximum
value of radial stress at the loaded fiber end by 42% compared to the pure matrix case. This is

20

very important finding, because the debonding failure for small fiber aspect ratios is mode I
dominant, and such debonding failure can be prevented engineered interlayer. For better
understanding of the results, we present the radial stress distribution in the rz plane of the
interlayer for the four different cases in Fig. 11. It is very clear from these figures that the radial
stress peaks at the loaded fiber end and decays very slowly towards the other interface. Figure 12
shows the transverse shear stress distribution in the rz plane of the interlayer. Closely observing
the slope of shear stress curve in different regions, it is seen that the value of shear stress
increases rapidly towards the other interface at z = 0; this is true for all the four cases but the

parallel case provides better results. It may be observed from Figs. 11 and 12 that the magnitude
of radial stress is relatively lower compared to the magnitude of shear stress. Radial tension at
fiber-matrix interface may initiate debonding at the loaded fiber end if the radial stress exceeds
the interfacial tensile strength and such microscale damages can grow rapidly leading to
macroscale failure of the component (Upadhyaya and Kumar, 2015). The large reduction in the
radial stress at the interface indicates that the nano-engineered of an interlayer significantly
improves the mode 1 fracture toughness of a resulting composite, which is likely to affect the
initial mode of failure of the composite.
So far, in this work, the stress transfer characteristics of the multiscale composite have
been studied by considering the four different cases. It is evident from the previous set of results
that the parallel case improves the stress transfer behavior of multiscale composite; therefore, we
consider the parallel case to further study the role of loading of CNTs on the stress transfer
characteristics of the multiscale composite. Practically, the CNT loading in the interlayer can
vary around the microscale fiber, therefore, the investigation of the effect of variation of CNT

loading on the stress transfer characteristics of the multiscale composite would be an important
study. For such investigation, three discrete values of CNT loading are considered; 5%, 10% and
15% in the interlayer. As can be seen from the tabulated values in Table 4, the elastic properties
of the interlayer improve with the increase in the CNT loading. Once again it is found that the
average axial stress in the microscale fiber is not much influenced by the loading of CNTs in
comparison to the base composite, i.e., without CNTs, as demonstrated in Fig. 13. Figure 14
demonstrate that the effect of CNT loading on the interfacial shear stress between the microscale
fiber and the interlayer. This figure indicates that the incorporation of different types of
interlayer exhibit almost identical trends and the loading of CNTs more than 5% do not much

21

reduce the maximum value of 1 . On the other hand, the effect of CNT loading on the radial

stress at the fiber-interlayer interface along the length of a microscale fiber is found to be
marginal, as shown in Fig. 15. As expected, the interlayer with 15% CNT loading provide the
lower value of radial stress at the ends of the microscale fiber in comparison to all other cases.

5. Conclusions
In this study, we proposed a novel concept to improve the mechanical properties and stress
transfer behaviour of a multiscale composite. This concept involves the introduction of carbon
nanostructures (CNS) around the microscale fibers embedded in the epoxy matrix, resulting in a
multiscale composite with enhanced properties. Using this concept, damage initiating threshold
and the fatigue strength of conventional composites can be greatly improved by strengthening
the interfacial fiber-matrix region. Accordingly, two aspects of the work examined: (i)
determination of the transversely isotropic properties of CNS comprised of CNT bundle and
epoxy matrix through MD simulations in conjunction with the Mori-Tanaka model, and (ii) the
development of three-phase pull-out model for a multiscale composite. Our pull-out model
incorporates different nano- and micro-scale transversely isotropic phases and allows the
quantitative determination of the stress transfer characteristics of the multiscale composites. This
model has also been validated by comparing the predicted results with those of FE simulations.
The developed analytical model was then applied to investigate the effect of orientation of CNS
and their loading on the stress transfer characteristics of the multiscale composite. The following
is a summary of our findings:
1. The incorporation of the interlayer containing CNS and epoxy matrix between the fiber
and the matrix would significantly improve the mode 1 fracture toughness of a resulting
composite, which is likely to affect the initial mode of failure of the composite,
2. Orientations of CNS have significant influence on the stress transfer characteristics of the
multiscale composite; aligned CNS along the axial direction of the microscale fiber is found
be effective in comparison to all other their orientations, and
3. The three-phase pull-out model developed in this study offers significant advantages over the
existing pull-out models and is capable of investigating the stress transfer characteristics of
any multiscale composite containing either aligned or randomly dispersed nanostructures.

22

References
Alian, A.R., Kundalwal, S.I., Meguid, S.A., 2015a. Interfacial and mechanical properties of epoxy
nanocomposites using different multiscale modeling schemes. Composite Structures 131, 545
555. doi: 10.1016/j.compstruct.2015.06.014
Alian, A.R., Kundalwal, S.I., Meguid, S.A., 2015b. Multiscale modeling of carbon nanotube epoxy
composites. Polymer 70, 149160 . doi:10.1016/j.polymer.2015.06.004
Allen, M.P., Tildesley, D.J., 1987. Computer simulation of liquids. Oxford: Clarendon Press.
Bandyopadhyay, A., Odegard, G.M., 2012. Molecular modeling of crosslink distribution in epoxy
polymers. Modelling and Simulation in Materials Science Engineering 20, 045018.
doi:10.1088/0965-0393/20/4/045018
Banholzer, B., Brameshuber, W., Jung, W., 2005. Analytical simulation of pull-out teststhe direct
problem.
Cement
and
Concrete
Composites
27,
93101.
doi:
10.1016/j.cemconcomp.2004.01.006
Bekyarova, E., Thostenson, E.T., Yu, A., 2007. Multiscale carbon nanotube-carbon fiber
reinforcement for advanced epoxy composites. Langmuir 23, 39703974. doi:10.1021/la062743p
Benveniste, Y., 1987. A new approach to the application of Mori-Tanakas theory in composite
materials. Mechanics of Materials 6(2), 147157. doi: 10.1016/0167-6636(87)90005-6
Chen, L., Jin, H., Xu, Z., Li, J., Guo, Q., Shan, M., Yang, C., Wang, Z., Mai, W., Cheng, B., 2015.
Role of a gradient interface layer in interfacial enhancement of carbon fiber/epoxy hierarchical
composites. Journal of Material Science 50, 112121. doi:10.1007/s10853-014-8571-y
Cho, J., Chen, J.Y., Daniel, I.M., 2007. Mechanical enhancement of carbon fiber/epoxy composites
by graphite nanoplatelet reinforcement. Scripta Materialia 56, 685688. doi:
10.1016/j.scriptamat.2006.12.038
Cui, Y., Kundalwal, S. I., Kumar S., 2015. Gas barrier performance of graphene/polymer
nanocomposites. Carbon (revisions submitted).
Dauber-Osguthorpe, P., Roberts, V.A., Osguthorpe, D.J., Wolff, J., Genest, M., Hagler, A.T., 1998.
Structure and energetics of ligand binding to proteins: Escherichia coli dihydrofolate reductasetrimethoprim, a drug-receptor system. Proteins 4, 3147. doi:10.1002/prot.340040106
Davis, D.C., Wilkerson, J.W., Zhu, J., Ayewah, D.O.O., 2010. Improvements in mechanical
properties of a carbon fiber epoxy composite using nanotube science and technology. Composite
Structures 92, 26532662. doi:10.1016/j.compstruct.2010.03.019
Davis, D.C., Wilkerson, J.W., Zhu, J., Hadjiev, V.G., 2011. A strategy for improving mechanical
properties of a fiber reinforced epoxy composite using functionalized carbon nanotubes.
Composites Science and Technology 71, 10891097. doi: 10.1016/j.compscitech.2011.03.014
Drzal, L., 1986. The interphase in epoxy composites. In: Duek K, editor. Epoxy Resins and
Composites II. vol. 75 of Advances in Polymer Science. Springer Berlin Heidelberg, 132. doi:
10.1007/BFb0017913
Dumlich, H., Gegg, M., Hennrich, F., Reich, S., 2011. Bundle and chirality influences on properties
of carbon nanotubes studied with van der Waals density functional theory. Physica Status Solidi
B 248, 25892592. doi: 10.1002/pssb.201100212

23

Frankland, S.J.V., Harik, V.M., Odegard, G.M., Brenner, D.W., Gates, T.S., 2003. The stress-strain
behavior of polymer-nanotube composites from molecular dynamics simulation. Composites
Science and Technology 63(11), 165561. 10.1016/S0266-3538(03)00059-9
Fu, S.-Y., Lauke, B., 2000. Comparison of the stress transfer in single- and multi-fiber composite
pull-out tests. Journal of Adhesion Science and Technology 14(3), 437452. doi:
10.1163/156856100742690
Fu, S.-Y., Yue, C.-Y., Hu, X., Mai, Y.-W., 2000. Analyses of the micromechanics of stress transfer
in single- and multi-fiber pull-out tests. Composites Science and Technology 60, 569579. doi:
10.1016/S0266-3538(99)00157-8
Garcia, E.J., Wardle, B.L., Hart, A.J., Yamamoto, N., 2008. Fabrication and multifunctional
properties of a hybrid laminate with aligned carbon nanotubes grown In Situ. Composites
Science and Technology 68, 20342041. doi:10.1016/j.compscitech.2008.02.028
Haghighatpanah, S., Bolton, K., 2013. Molecular-level computational studies of single wall carbon
nanotubepolyethylene composites. Computational Materials Science 69, 443454. doi:
10.1016/j.commatsci.2012.12.012
Haghighatpanah, S., Bolton, K., 2013. Molecular-level computational studies of single wall carbon
nanotubepolyethylene composites. Computational Materials Science 69, 443454. doi:
10.1016/j.commatsci.2012.12.012
Hashin, Z., Rosen, B. W., 1964. The elastic moduli of fiber-reinforced materials. ASME Journal
Applied Mechanics 31(2), 223232. doi: 10.1115/1.3629590
Honjo, K., 2007. Thermal stresses and effective properties calculated for fiber composites using
actual cylindrically-anisotropic properties of interfacial carbon coating. Carbon 45(4), 865872.
10.1016/j.carbon.2006.11.007
Hung, K.H., Kuo, W.S., Ko, T.H., Tzeng, S.S., Yan, C.F., 2009. Processing and tensile
characterization of composites composed of carbon nanotube-grown carbon fibers. Composites
Part A 40, 12991304. doi:10.1016/j.compositesa.2009.06.002
Jia, Y., Chen, Z., Yan, W., 2014. A numerical study on carbon nanotubehybridized carbon fibre
pullout. Composites Science and Technology 91, 3844. doi: 10.1016/j.compscitech.2013.11.020
Jiang, Q., Tallury, S.S., Qiu, Y., Pasquinelli, M.A., 2014. Molecular dynamics simulations of the
effect of the volume fraction on unidirectional polyimidecarbon nanotube nanocomposites.
Carbon 67, 440448. doi: 10.1016/j.carbon.2013.10.016
Jin, S.-Y., Young, R.J., Eichhorn, S.J. Hybrid carbon fibrecarbon nanotube composite interfaces
Composites Science and Technology 95, 114120. doi: 10.1016/j.compscitech.2014.02.015
Kim, J.K., Baillie, C., Mai, Y.-W., 1992. Interfacial debonding and fiber pull-out stresses. Part I.
Critical comparison of existing theories with experiments. Journal of Material Science 27, 3143
3154. doi:10.1007/BF01116004
Kim, J.K., Lu, S., Mai, Y.-W., 1994. Interfacial debonding and fiber pullout stresses. Part IV.
Influence of interface layer on the stress transfer. Journal of Material Science 29, 554561. doi:
10.1007/BF01162521

24

Kumar, A., Sundararaghavan, V., Browning, A.R., 2014. Study of temperature dependence of
thermal conductivity in cross-linked epoxies using molecular dynamics simulations with long
range interactions. Modelling and Simulation in Materials Science and Engineering 22, 025013.
doi: 10.1088/0965-0393/22/2/025013
Kundalwal, S.I., Meguid, S.A., 2015. Micromechanics modelling of the effective thermoelastic
response of nano-tailored composites. European Journal of Mechanics A/Solids 53, 241253.
doi: 10.1016/j.euromechsol.2015.05.008
Kundalwal, S.I., Ray, M.C., Meguid, S.A., 2014. Shear lag model for regularly staggered short
fuzzy fiber reinforced composite. ASME Journal of Applied Mechanics 81(9), 091001-1
091001-14. doi: 10.1115/1.4027801
Lachman, N., Wiesel, E., de Villoria R.G., Wardle, B.L., Wagner, H.D., 2012. Interfacial load
transfer in carbon nanotube/ceramic microfiber hybrid polymer composites. Composites Science
and Technology 72, 14161422. doi:10.1016/j.compscitech.2012.05.015
Li, C., Medvedev, G.A., Lee, E.-W., Kim, J., Caruthers, J.M., Strachan, A., 2012. Molecular
dynamics simulations and experimental studies of the thermomechanical response of an epoxy
thermoset. Polymer 53, 42224230. doi:10.1016/j.polymer.2012.07.026
Li, C., Strachan, A., 2010. Molecular simulations of crosslinking process of thermosetting
polymers. Polymer 51, 60586070. doi: 10.1016/j.polymer.2010.10.033
Marotzke, C., 1994. The elastic field arising in the single-fiber pull-out test. Composites Science
and Technology 50(3), 393405. doi: 10.1016/0266-3538(94)90027-2
Martnez, L., Andrade, R., Birgin, E.G., Martnez, J.M., 2009. PACKMOL: a package for building
initial configurations for molecular dynamics simulations. Journal of Computational Physics 30,
21572164. doi: 10.1002/jcc.21224
Marzari, N., Ferrari, M., 1992. Textural and micromorphological effects on the overall elastic
response of macroscopically anisotropic composites. Journal of Applied Mechanics 59, 269275.
doi: 10.1115/1.2899516
Meng, Q., Wang, Z., 2015. Theoretical model of fiber debonding and pull-out in unidirectional
hybrid-fiber-reinforced brittle-matrix composites. Journal of Composite Materials 49(14), 1739
1751. doi: 10.1177/0021998314540191
Mori, T., Tanaka, K., 1973. Average stress in matrix and average elastic energy of materials with
misfitting inclusions. Acta Metallurgica 21(5), 571574. doi: 10.1016/0001-6160(73)90064-3
Morris, J.E., 2008. In Nanopackaging: nanotechnologies and electronics packaging. Springer:
Portland, OR, 4950. doi: 10.1007/978-0-387-47325-3
Nairn, J. A., 1997. On the use of shear-lag methods for analysis of stress transfer in unidirectional
composites. Mechanics of Materials 26(2), 6380. doi: 10.1016/S0167-6636(97)00023-9
Odegard, G.M., Gates, T.S., Wise, K.E., Park, C., Siochi, E.J., 2003. Constitutive modeling of
nanotube-reinforced polymer composites. Composites Science and Technology 2003, 63, 1671
1687. doi:10.1016/S0266-3538(03)00063-0
Pal, G., Kumar, S., 2015. Modeling of effective electrical conductivity of short carbon fiber-carbon
nanotube-polymer matrix hybrid composites. Materials and Design (in press).

25

Pascault, J.P., Sautereau, H., Verdu, J., Williams, R.J.J., 2002. Thermosetting polymers. New York:
CRC Press. doi: 10.1002/9781118480793.ch28
Plimpton, S., 1995. Fast parallel algorithms for short-range molecular dynamics. Journal of
Computational Physics 117, 119. doi:10.1006/jcph.1995.1039
Qiu, Y.P., Weng, G.J., 1990. On the application of Mori-Tanakas theory involving transversely
isotropic spheroidal inclusions. International Journal of Engineering Science 28(11), 1121
1137. doi: 10.1016/0020-7225(90)90112-V
Quek, M.Y., Yue, C.Y., 1997. An improved analysis for axisymmetric stress distributions in the
single fibre pull-out test. Journal of Material Science 37, 54575465. doi:
10.1023/A:1018647718400
Romanov, V., Lomov, S. V., Verpoest, I., Gorbatikh, L., 2015. Inter-fiber stresses in composites
with carbon nanotube grafted and coated fibers. Composites Science and Technology 114, 7986.
doi: 10.1016/j.compscitech.2015.04.013
Sottos, N.R., McCullough, R.L., Scott, W.R., 1992. The influence of interphase regions on local
thermal displacements in composites. Composites Science and Technology 44(4), 319332. doi:
10.1016/0266-3538(92)90069-F
Tsai, J.L., Tzeng, S.H., Chiu, Y.T., Characterizing elastic properties of carbon nanotube/polyimide
nanocomposites using multi-scale simulation. Composites: Part B Engineering 2010, 41, 106
115. doi:10.1016/j.compositesb.2009.06.003
Tsai, K.-H., Kim, K.-S., 1996. The micromechanics of fiber pull-out. Journal of the Mechanics and
Physics of Solids 44(7), 11471177. doi: 10.1016/0022-5096(96)00019-1
Tunvir, K., Kim, A., Nahm, S.H., 2008. The effect of two neighboring defects on the mechanical
properties of carbon nanotubes. Nanotechnology 19, 065703. doi:10.1088/09574484/19/6/065703
Upadhyaya, P., Kumar, S., 2015. Micromechanics of stress transfer through the interphase in fiberreinforced
composites.
Mechanics
of
Materials
89,
190201.
doi:
10.1016/j.mechmat.2015.06.012
Varshney, V., Patnaik, S.S., Roy, A.K., Farmer, B.L., 2008. A molecular dynamics study of epoxybased networks: cross-linking procedure and prediction of molecular and material properties.
Macromolecules 41(18), 68376842. doi: 10.1021/ma801153e
Wicks, S.S., Wang, W., Williams, M.R., Wardle, B.L., 2014. Multi-scale interlaminar fracture
mechanisms in woven composite laminates reinforced with aligned carbon nanotubes.
Composites Science and Technology 100, 128135. doi:10.1016/j.compscitech.2014.06.003
Yang, S., Yu, S., Kyoung, W., Han, D.-S., Cho, M., 2012. Multiscale modeling of size-dependent
elastic properties of carbon nanotube/polymer nanocomposites with interfacial imperfections.
Polymer 53, 623633. doi: 10.1016/j.polymer.2011.11.052
Zhang, J., Zhuang, R., Liu, J., Mder, E., 2010. Heinrich G, Gao S. Functional interphases with
multi-walled carbon nanotubes in glass fibre/epoxy composites. Carbon 48, 22732281. doi:
10.1016/j.carbon.2010.03.001

26

Zhang, X., Fan, X., Yan, C., Li, H., Zhu, Y., Li, X., Yu, L., 2012. Interfacial microstructure and
properties of carbon fiber composites modified with graphene oxide. ACS Applied Material
Interfaces 4, 15431552. doi:10.1021/ am201757v

Appendix A: Elements of Eshelby tensor and transformation matrix


In case of CNS are aligned along the direction of a microscale fiber, the elements of the Eshelby
tensor [S E ] are explicitly written as follows (Qiu and Weng, 1990):
E
S2222
E
S2211

=
=

E
S3333
E
S3311

5 4m
=
,
8(1 m )
m
=
,
2(1 m )
4m 1
,
8(1 m )

E
E
S2233
= S3322
=

S2323

E
E
S1313
= S1212
= 14,

3 4m
=
and all other elements are zero
8(1 m )

(A1)

Similarly, in case of CNS are aligned radially to the axis a microscale fiber, the elements of the
Eshelby tensor [S E ] can be written as follows:

E
E
S1111
= S2222
=

5 4m
,
8(1 m )

E
E
= S2211
=
S1122

E
E
S1133
= S2233
=

E
S1212
=

4m 1
,
8(1 i )

m
,
2(1 m )

E
E
S1313
= S2323
= 14,

3 4m
and all other elements are zero
8(1 m )

Direction cosines corresponding to the transformation matrix [Eq. (8)]:


t11 = cos cos sin cos sin,

t12 = sin cos + cos cos sin,

(A2)

27

t13 = sin sin ,

t 21 = cos sin sin cos cos,

t 22 = sin sin + cos cos cos,


t 23 = sin cos,
t 31 = sin sin,

t 32 = cos sin and t 33 = cos

(A3)

Appendix B: Explicit forms of constant coefficients


The constants (Ai ) obtained in the course of deriving the shear lag model in section 3 are
explicitly expressed as follows:

A1 =

1
r
a
r r 2 a2
2
aln
+
a
ln

,
i
a b 2 a2
a
2
C66

A2 =

r r 2 a2
2
a
ln

,
i (b 2
a
2
C66
a2 )
b

aln(ba)
a
b b 2 a2
2
A3 =
+ i
a ln
,
i
2
a
C66
C66 (b 2 a2 )

A4 =
A5 =

b 3 a2 b
b
a a2 b 2
2
+
a
ln

,
i
i
2
b
2C66
(c 2 b 2 ) C66
(b 2 a2 )

i
C13
a
b b 2 a2
2

)
aln(b
a
+
a
ln

,
i
b 2 a2
a
2
C66

A8 = A4 |r=b,
A12 =

A9 = A4 |r=c,

A10

A6 = A2 |r=b,

i
2b2 C12
= 2
,
b a2

A11

A7 = A3 |r=b,

i
2C12
= 2
b a2

i
aC11
b b2 a2 a2 b2 ln(ba)
b 4 a4
2
b
ln

,
i (b 2
a
2
2(b 2 a2 )
8(b 2 a2 )
C66
a2 )

A13 =

i
a2 b3 C11
ln(ba)
,
i
2
C66 (b a2 )2

A14 =

A15 = A9 b12 + A10 b22 + A11 b32 + A12 ,

i
C11
f
C11

+ A9 b11 + A10 b21 + A11 b31 ,

A16 = A9 b13 + A10 b23 + A11 b33 + A13 ,

28

A17 =
A19 =

m
C11
f
C11

m f
2C11
C12

m f
2C11
C12
f
C11

f
C11

b11 +

m
2C12
b21 ,

A18 =

m
m
b13 + 2C12
b23 + C11
A4 ,

m f
2C11
C12
f
C11

m
m
b12 + 2C12
b22 + C11
A3 ,

A20 = a2 + (b2 a2 )A14 + (c 2 b2 )A17 ,

a
A21 = [(b2 a2 )A15 + (c 2 b2 )A18 ],
2

A22 =

c 2 b2
[(b2 a2 )A16 + (c 2 b2 )A19 ],
2b

a A15 A19
A23 =
A18 ,
2 A16

A24 =
A25

1 A14 A19
a2 A19
A21 A19 (c 2 b2 ) A21

A17 + 2

,
A23 A16
(b a2 )A16 A16 A22 (b 2 a2 ) A22

1 A20 A19 A20 (c 2 b2 )


=

+
,
A23 A22 A16 A22 (b 2 a2 )

A31 =

a A18 A20
+ A21 ,
2 A17
A34

c 2 b2 A19 A20

A22 ,
2b
A17

A15 A19
= A16
,
A18

A36 =
A37

A32 =

A26

1 a2
A19 a2 (c 2 b2 )
=

+
,
A23 A22 A16 A22 (b 2 a2 )

A35

A33 = A14

c 2 b2
= A34 (b a )
,
2b
2

A15 A17
,
A18

1 a2 A32
A15 A32 A33

+ (b2 a2 )
+
+ (c 2 b2 ),
A35 A31
A18
A31

1 a2 A20
A20 A33 2
1 a4
A33
2 )
(b
=

+
a
and A38 =

+ a2 (b2 a2 )
(B1)
A35 A17 A31 A17 A31
A35 A31
A31

The constants bij appeared in Eq. (31) are expressed as follows:


i
f
f
i
B11 = C23
C23
C33
C33

B13 =

i
2C33
,
b 2 a2

f
i
f 2
a2 + b2 2C12
C13
)
2(C12

+
,
f
f
b 2 a2
C11
C11

B21 =

B12 =

i
i
f
m f
2C33
a2 2C13
C12
2C12
C12
+

2
2
f
f
b a
C11
C11

i
b2
2C33
,
b 2 a2

29

f
m
i
i
B22 = C11
+ C12
C23
C33

B31 =

f
m
2C12
C12
f
C11

a2 + b 2
,
b 2 a2

B23 =

1
a2 + a2
m
m
i
i
C

11
23
33 2
b 2 12
b a2

f
m
B32 = C11
+ C12
and B33 =

1 m
m)
(C C11
c 2 12

Table 1 Parameters used for MD simulations


Parameter

Epoxy

Bundle of 13 CNTs

CNT type

(5, 5)

Number of CNTs

13

Length of a CNT ()

100

CNT volume fraction

13.77%

RVE dimensions (3)

393939

7575100

Total number of EPON 862 molecules

100

800

Total number of DETDA molecules

50

400

Total number of atoms

6250

61050

Table 2 Elastic moduli of the epoxy material

Present MD simulations
Experimental (Morris, 2008)

Youngs modulus Poissons


(GPa)
ratio
3.5
0.36
3.43

(B2)

30

Table 3 Effective elastic coefficients of the CNS and corresponding displacement fields
Elastic coefficients

Applied
strains

Applied
displacement

C11

11 = e

u1 = ex1

C22 = C33

22 = e

23 = e2

C44

C55 = C66

K 23 =

C22 + C23
2

13 = e2
22 = 33

u2 = ex2
e

u2 = 2 x3 ,
e

u3 = 2 x2
e

u1 = 2 x3 ,
e

u3 = 2 x1

u2 = ex2 ,
u3 = ex3

Table 4 Effective elastic properties of the CNS and the interlayer

RVE

Modeling
Technique

CNT Volume
Fraction

(GPa)

(GPa)

(GPa)

(GPa)

(GPa)

(GPa)

CNS

MD simulations

18.96%
loading in the
CNS

108.35

3.5

3.5

6.7

1.48

1.83

Interlayer containing
aligned CNS

MD simulations in
conjunction with
Mori-Tanaka model

15%
loading in the
interlayer

86.93

3.46

3.46

6.51

1.39

1.44

Interlayer containing
aligned CNS

MD simulations in
conjunction with
Mori-Tanaka model

10%
loading in the
interlayer

60.21

3.4

3.4

6.3

1.39

1.56

Interlayer containing
aligned CNS

MD simulations in
conjunction with
Mori-Tanaka model

5%
loading in the
interlayer

32.92

3.36

3.36

6.08

1.33

1.42

Interlayer containing
CNS, which are
radially aligned to the
axis of a microscale
fiber

MD simulations in
conjunction with
Mori-Tanaka model

5%
loading in the
interlayer

6.08

3.4

3.36

32.92

1.42

1.33

Interlayer containing
randomly dispersed
CNS

MD simulations in
conjunction with
Mori-Tanaka model
and then transforming
the properties

5%
loading in the
interlayer

10.3

4.93

4.93

10.3

2.9

2.9

31

Fig. 1. Schematic illustration for the preparation of carbon fiber-epoxy composite containing
graphene nanosheets [Chen 2015].

32

Fig. 2. Molecular and chemical structures of (a) EPON 862, (b) activated EPON 862 and (c)
DETDA curing agent. Colored with gray, red, blue, and white are carbon, oxygen, nitrogen, and
carbon atoms, respectively.

33

Fig. 3. Snapshot of atomic configurations of EPON 862-DETDA epoxy.

34

Fig. 4. MD unit cell, representing CNS, comprised of EPON 862-DETDA epoxy and a bundle of
thirteen CNTs.

35

Fig. 5. Conceptual illustration of the multiscale composite: (a) three-phase pull-out model, (b)
nanocomposite RVE comprised of CNS and epoxy, and (c) microscale RVE representing
interlayer.

36

Fig. 6. Comparison of normalized axial stress in the fiber along its normalized length predicted
by analytical model with those of FE simulations. For three-phase pull-out model: CNT bundles
are parallel to the microscale fiber and their volume fraction in the interlayer is 5%.

Fig. 7. Comparison of normalized transverse shear stress along the normalized length of the
interlayer, at r = 5.1 m, predicted by analytical model with those of FE simulations. For threephase pull-out model: CNT bundles are parallel to the microscale fiber and their volume fraction
in the interlayer is 5%.

37

Fig. 8. Normalized axial stress in the fiber along its normalized length for different orientations
of CNS (for 5% CNT loading in the interlayer).

Fig. 9. Normalized interfacial shear stress along the normalized length of the fiber for different
orientations of CNS (for 5% CNT loading in the interlayer).

38

Fig. 10. Normalized iinterfacial radial stress at r = a over the normalized length of the microscale
fiber for different orientations of CNS (for 5% CNT loading in the interlayer).

39

0.04

0.02

(b)

(a)
0.02
i

r/ p

r/ p

0.01

-0.02
6.5

-0.01
6.5
1

6
5

ri

ri

z/Lf

z/Lf

0.04

(c)

(d)
0.02

0.04

r/ p

0.5

5.5

0.08

r/ p

0.5

5.5

0.06

0.02

0
-0.02
6.5

-0.02
6.5
1

6
0.5

5.5

ri

z/Lf

0.5

5.5

rm

z/Lf

Fig. 11. Surface plots depicting the variation of radial stress in the rz plane of: (a) parallel, (b)
perpendicular, (c) random and (d) pure matrix.

40

0.2

0.1

(a)
rz/ p

rz/ p

(b)

0.15

0.05

0.1
0.05
0

-0.05
6.5

-0.05
6.5
1

6
0.5

5.5
5

ri

ri

z/Lf

z/Lf

0.4

(c)

0.15
0.1

0.05

0.2
0.1

-0.05
6.5

-0.1
6.5
1

0.5

5.5

ri

(d)

0.3

rz/ p

0.5

5.5

0.2

rz/ p

z/Lf

0.5

5.5

ri

z/Lf

Fig. 12. Surface plots depicting the variation of transverse shear stress in the rz plane of
interlayer and matrix thicknesses along the axial direction: (a) parallel, (b) perpendicular, (c)
random and (d) pure matrix.

41

Fig. 13. Normalized axial stress in the microscale fiber along its length for different CNT
loadings where CNS are parallel to the microscale fiber.

Fig. 14. Normalized interfacial shear stress along the length of the microscale fiber for different
CNT loadings when CNS are parallel to the microscale fiber.

42

Fig. 15. Normalized interfacial radial stress at r = a along the length of the microscale fiber for
different CNT loadings where CNS are parallel to the microscale fiber.

Vous aimerez peut-être aussi