You are on page 1of 30

Accepted Manuscript

The Thermo-Mechanical Properties estimation of Fullerene-Reinforced Resin Epoxy


Composites by Molecular Dynamics Simulation- A Comparative Study

F. Jeyranpour, Gh Alahyarizadeh, H. Minuchehr

PII: S0032-3861(16)30098-2
DOI: 10.1016/j.polymer.2016.02.018
Reference: JPOL 18450

To appear in: Polymer

Received Date: 9 December 2015


Revised Date: 5 February 2016
Accepted Date: 6 February 2016

Please cite this article as: Jeyranpour F, Alahyarizadeh G, Minuchehr H, The Thermo-Mechanical
Properties estimation of Fullerene-Reinforced Resin Epoxy Composites by Molecular Dynamics
Simulation- A Comparative Study, Polymer (2016), doi: 10.1016/j.polymer.2016.02.018.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M

A comprehensive and comparative study on the effect of fullerene on the thermo-mechanical


properties of Araldite LY 5052 / Aradur HY 5052 cross-linked resin epoxies, of which one has
D

no fullerene and two have weight fractions of 8.64% (LY/HY/C60 (1) epoxy system) and 15.9%
(LY/HY/C60 (2) epoxy system), were numerically studied. The results indicated that
TE

mechanical properties have considerable improved by adding fullerene to epoxy systems.


LY/HY/C60 (1) epoxy system presents higher mechanical properties than LY/HY and
EP

LY/HY/C60 (2) epoxy systems in almost all temperatures.


C
AC
ACCEPTED MANUSCRIPT

The Thermo-Mechanical Properties Estimation of Fullerene-

Reinforced Resin Epoxy Composites by Molecular Dynamics

Simulation- A Comparative Study

PT
RI
F. Jeyranpour1, Gh. Alahyarizadeh1*, H. Minuchehr1
1
Engineering Department, Shahid Beheshti University, G.C., P.O. Box 1983969411, Tehran, Iran

SC
* Corresponding author* E-mail: g_alahyarizadeh@yahoo.com

Tel: +9821 29904226

U
AN
Abstract: A comparative study was conducted to determine the effects of fullerene on the thermo-
mechanical properties of Araldite LY 5052/Aradur HY 5052 cross-linked resin epoxy by molecular
M

dynamics simulations. Different from ordinary epoxy resins, Araldite LY 5052/Aradur HY 5052 cross-
linked resin epoxy is composed of four compounds that provide different thermal and mechanical
D

properties. The thermo-mechanical properties of three different Araldite LY 5052/Aradur HY 5052 cross-
linked resin epoxies, of which one has no fullerene and two have weight fractions of 8.64% (LY/HY/C60
TE

(1) epoxy system) and 15.9% (LY/HY/C60 (2) epoxy system), were comparatively studied. Simulation
results indicated that the glass transition temperature (Tg) has slight variations for different fullerene-
EP

containing epoxy systems and that the coefficients of thermal expansions decrease by adding fullerene to
the resin epoxy system. Furthermore, mechanical properties such as Young’s, bulk, and shear moduli, as
well as Poisson’s ratio and density, show considerable improvements after the addition of fullerene to the
C

epoxy systems. The best mechanical properties were obtained for the weight fractions of 8.64%; this result
AC

was in agreement with the experimental results. For example, the Young’s modulus of the LY/HY epoxy
system increases by 19.5% and 7.5% in the LY/HY/C60 (1) epoxy system and LY/HY/C60 (2) epoxy
system, respectively. The evolution of the local structures of the different fullerene-containing epoxy
systems were also investigated by studying the radial distribution functions.

Keywords: Fullerene- Araldite LY 5052 / Aradur HY 5052 epoxy resins; Thermo- mechanical properties;

Molecular dynamics simulation.


ACCEPTED MANUSCRIPT

INTRODUCTION
Epoxy resins are thermoset polymers that have broad practical applications [1]. The superior
adhesive and mechanical characteristics of 3D cross-linked epoxy polymers make them suitable
materials for numerous industrial applications such as aerospace, coatings, marine vessels, space
vehicles, adhesives, electronics, automotive, and biotechnology [2, 3, 4]. Epoxy resins constitute

PT
a large class of composites that contain two or more epoxy groups. Epoxy resins are generally
complicated 3D network structures. The epoxies can be reacted with other components in

RI
conjunction with hardeners or curing agents. Hardeners typically have active hydrogens including
amines and anhydrides. The resultant composites exhibit a series of exceptional performances,

SC
such as low creep, high-temperature performance, high modulus, and fracture strength. These
outstanding properties enabled them to be widely used in various applications, such as

U
composites, coatings, and adhesives. Furthermore, the cross-linking procedure contains
irreversible covalent bonds, and cross-linked polymers cannot be re-melted or re-shaped [5].
AN
Araldites are one of the most important classes of epoxies. LY 5052/HY 5052 is a highly used
epoxy system of the araldite epoxy resins because of its highly significant mechanical properties.
M

The system is qualified for applications that need high reliability, particularly in aerospace [6].
Different from ordinary epoxy resins, Araldite LY 5052/Aradur HY 5052 cross-linked resin
D

epoxy is composed of four components that enable the epoxy to provide various epoxy resin
systems with different thermal and mechanical properties. This capability causes Araldite LY
TE

5052/Aradur HY 5052 cross-linked resin epoxy to be considered for potential use in several
applications such as aerospace, industrial composites, aircraft repair, and tooling [7, 8]. The main
EP

advantages of araldite epoxy resins are low viscosity, easy impregnation, reinforcement materials,
ample processing time for the production of large objects, high temperature resistance (glass
C

transition temperature) after ambient cure at 60 °C, after post-cure at 100–120 °C, exceptional
dynamic and mechanical properties after ambient cure for high properties, and after post-cure at
AC

elevated temperatures. In Germany, these products are qualified for use in gliders because their
adequate skin protection is indispensable and are ideal materials for aerospace and other
laminating applications [9].
On the basis of author research, no theoretical or simulation-based study on Araldite LY
5052/Aradur HY 5052 cross-linked resin epoxy is available. The available considerable research
in this area is a few experimental studies that show some important features of this epoxy system,
ACCEPTED MANUSCRIPT

such as density and Young’s modulus. Comparative studies were conducted by Bolasodun et al.
[10] on the cure kinetics of Araldite LY 5052 and hardener 4,4'-diaminodiphenylsulfone. Yusoff
et al. [11] experimentally investigated the influences of infrared radiation on the cure of the two
epoxy resin systems of Araldite LY 5052 and Araldite DLS 772, as well as 4,4'-
diaminodiphenylsulfone.

PT
The epoxy resins are inherently brittle, and their high performance in various applications,
such as in structural adhesives, microelectronic encapsulants, and composite materials, needs a

RI
high-performance epoxy resin. To obtain the enhanced fracture toughness and better
characteristics of epoxy resins, several researchers have proposed new researchable ideas by

SC
adding different fillers and additives. Therefore, attention has been directed at the use of different
fillers and additives such as nanoparticles, graphene, and carbon nanotubes (CNTs) to improve

U
traditional composite systems.
Sprenger [12] investigated surface-modified silica and their distribution in epoxy resin system
AN
for industrial uses. He showed that using nanoparticles with tougheners overcompensates for the
low modulus and loss in strength. Rajabi et al. [13] experimentally studied the effects of the TiO2
M

nanoparticles on the thermal stability and mechanical moduli of the composites. Kochetov et al.
[14] prepared epoxy resins with high thermal and electrical properties by using aluminum oxide
D

and silicon dioxide nanoparticles. They used ultrasonic processing and high shear mechanical
mixing to reach a fine dispersion of the fillers. Epoxy/Al2O3 nanocomposites with different
TE

weight fractions of Al2O3 nanoparticles (1, 2, 3, 4, 5, and 7 wt%) with average particle sizes
smaller than 50 nm were prepared by Kadhim et al. [15]. Xiong et al. [16] and Gou et al. 17] used
EP

molecular dynamics (MD) simulations to predict the mechanical characteristics of CNT-


reinforced epoxy composites. Li et al. [18] also used MD simulations were also used to
characterize the thermal and mechanical properties of multilayer graphene-reinforced resin epoxy
C

composites. They focused on two configurations, namely, parallel and perpendicular graphene
AC

layers to polymer–graphene interface. Amico [19] experimentally showed the effects of the
dispersion of carboxylated single-walled CNTs in epoxy matrices. Harris [20] investigated the
properties of CNTs and described the preparation and properties of CNT-reinforced composites.
Fullerenes, which is another member of carbon-based nano-materials, have significant
physical and chemical properties. They are commercially available at an affordable cost, thus
making them a considerable candidate in mechanics, optics, electrochemistry, gas absorption, and
ACCEPTED MANUSCRIPT

other various industrial fields. Fullerenes can also be used as a modifier to improve the
performance of engineering polymers and carbon-based epoxy resin composites [21]. To the best
of the knowledge of the author, few experimental works have focused on the fullerene-reinforced
epoxy resins. Okonkwo et al. [22] used the fullerene to strengthen the epoxy resin with 1%
weight fraction. Wang et al. [23] added fullerene into epoxy resins to prepare fullerene-containing

PT
epoxy membranes. They studied the effects of the chemical structure of aminated fullerene on the
thermal, mechanical, and optical properties of fullerene-containing epoxy membranes.

RI
The most usual method to obtain the thermal, physical, and mechanical properties of materials
such as epoxy polymers are experimental tests, but they are expensive and time consuming. Thus,

SC
most current studies on this topic use MD simulations. The MD simulations of polymers is a new
method of analysis that can be applied to a wide variety of materials, crystals, and polymer

U
systems. The method proposes an alternative approach for estimating the thermal and mechanical
properties of epoxy resin [5]. In the present study, the comparative investigation on the thermo-
AN
mechanical properties of LY/HY 5052 and fullerene (C60)-containing LY/HY 5052 epoxy resins
with different molecular weight fractions was conducted by MD simulations.
M

MODELING AND SIMULATION DETAILS


D

Molecular Structures of resins and hardeners


TE

Araldite LY 5052 is composed of two components: an epoxy phenol novolak resin


(C35H32O4) with 38–42 wt% composition, and 1,4-butanediol diglycidyl ether (C10H18O4)
with 55–68 wt% composition. The chemical structures of the two components are shown in Figs.
EP

1 (a) and (b). The number of epoxide group at phenol novolak resin is four, and its molecular
weight is 345 g/mol. The functionality of 1,4-butanediol diglycidyl is two, and its molecular
C

weight is 202.3 g/mol [24]. Aradur 5052 (PREV HY5052) hardener also consists of two
AC

polyamine components: IPDA (C10H22N2) with 30–60 wt% and cycloaliphatic diamine
(C15H30N2) with 30–60 wt%. Figs. 1 (c) and (d) show the chemical structures of IPDA and
cycloaliphatic diamine, respectively. This hardener is usually used with Araldite LY5052 resin [7,
8]. Fig. 1 (e) also shows the 3D molecular structure of C60.
ACCEPTED MANUSCRIPT

PT
a

RI
U SC
AN
b
M
D
TE
EP

c
C
AC

d
ACCEPTED MANUSCRIPT

PT
RI
e

SC
Fig.1. Structures of epoxy resin and hardeners. a) Epoxy phenol novalok b) Buthanediol diglycidyl ether
c) IPDA d) Cycloaliphatic diamine, and e) 3D molecular structure of C60

U
IPDA and cycloaliphatic diamine are capable of reacting with four epoxide groups at the most.
AN
Novolac and butanediol diglycidyl ether has four and two epoxide groups, respectively. The
composition ratios of resins and hardeners in LY/HY 5052 epoxy system and the weight fraction
of each component are listed in Table 1. Fig. 2 shows the polymer structure of the LY/HY 5052
M

epoxy system in different situations: without cross-linking and with cross-linking at 250 and 550
K. Fullerene-containing LY/HY epoxy systems with fullerene weight fractions of 8.64% and
D

15.92% are also illustrated in Figs. 3 and 4. Tables 2 and 3 summarize the weight fractions of
TE

these systems.

Table 1. Composition ratios for LY/HY epoxy system and molecular weight of each component.
EP

Molecule No. of molecule Molecular Weight


in cell weight (g/mol) percentage
C

Novalok 20 10320 45.18%


AC

Buthanediol diglycidyl ether 28 5656 24.76%


IPDA 18 3060 13.39%
Cycloaliphatic diamine 16 3808 16.67%
ACCEPTED MANUSCRIPT

Table 2. Composition ratios for the LY/HY/C60 (1) epoxy system and molecular weight of each
component.

Molecule No. of molecule Molecular Weight


in cell weight (g/mol) percentage
Novalok 20 10320 41.27%

PT
Buthanediol diglycidyl ether 28 5656 22.62%
IPDA 18 3060 12.24%
Cycloaliphatic diamine 16 3808 15.23%

RI
Fullerene 3 2160 8.64%

SC
Table 3. Composition ratios for the LY/HY/C60 (2) epoxy system and molecular weight of each

U
component.

Molecule No. of molecule Molecular Weight


AN
in cell weight (g/mol) percentage
Novalok 20 10320 37.99%
M

Buthanediol diglycidyl 28 5656 20.82%


ether
D

IPDA 18 3060 11.26%


Cycloaliphatic diamine 16 3808 14.01%
TE

Fullerene 6 4320 15.92%


C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
a

U SC
AN
M

b
D
TE
C EP
AC

Fig. 2. The polymer structure of LY/HY 5052 (a) without cross-linking, (b) with cross-linking at 250◦K
and (c) 550◦K.
ACCEPTED MANUSCRIPT

PT
RI
SC
a

U
AN
M
D
TE

b
C EP
AC

Fig. 3. The polymer structure of LY/HY 5052/C60 (1) (a) without cross-linking, (b) with cross-linking
at250◦K and (c) 550◦K.
ACCEPTED MANUSCRIPT

PT
RI
a

U SC
AN
M
D

b
TE
C EP
AC

Fig. 4. The polymer structure of LY/HY 5052/C60 (2) (a) without cross-linking, (b) with cross-linking at
250◦K and (c) 550◦K.
ACCEPTED MANUSCRIPT

Cross Linking Mechanism


The molecular structures of the resins and hardeners and the potential curing sites were
introduced in the above section. In each epoxide group, the C–O bond must be broken to form a
reactive –CH2 site (Fig. 5). These sites are capable of being cross-linked to hardeners [2, 5]. In
the combination of resin and hardener, the molecules change their locations under appropriate

PT
states; therefore, the covalent bonds could be formed between C and N atoms. A 100%
convergence will be achieved if all potential covalent bonds are made, although such a

RI
phenomenon is rarely observed in real reactions. In natural conditions, the convergence (cross-
linking density) is typically less than 100% [2, 4, 5]. All possible connections in this epoxy

SC
system are 136 bonds. Fullerenes do not react with resins or hardeners.

U
AN
M

Fig. 5. Conversion of an original epoxide group to a reactive one through breaking the C-O bond [5].

Software and Force-Field


D

In this research, all simulations were performed by the Accelrys Materials Studio 6.0 software
TE

package [25]. The force-field used in all dynamic simulations and minimization processes was
the Condensed-phase Optimized Molecular Potentials for Atomistic Simulation Studies
EP

(COMPASS) force-field. This force-field, which is used for various molecules such as polymers,
most common organics, and small inorganic molecules, is an ab initio force-field. This force-field
is utilized to determine the properties of compounds (for various molecules) such as polymers,
C

common organics, and small inorganic molecules [2, 5]. COMPASS is a new MD simulation
AC

method that was made by modifying the polymer consistent force-field (pcff) computational
model. CFF91, CFF, pcff, and COMPASS force-fields have the same functional form. The atomic
force-field model defines physical systems as assemblies of atoms that are collected by
interatomic forces. Specifically, chemical bonds result from the particular form of the interactions
among atoms that form a molecule. The interaction law is indicated by the potential U(r1,…, rN),
which signifies the potential energy of N interacting atoms as a function of their situations ri=(xi,
ACCEPTED MANUSCRIPT

yi, zi). Given the potential, the force acting upon ith atom is determined by the gradient with
respect to atomic displacements, as shown in Equation (1).

PT
The functional form of the mentioned force-field is shown in Equation (2).

RI
E = ∑  K 2 (b − b 0 ) + K 3 (b − b 0 ) + K 4 (b − b 0 )  + ∑  H (θ − θ ) + H 3 (θ − θ 0 ) + H 4 (θ − θ 0 )  +
2 3 4 2 3 4

b
  θ
 2 0 

SC
(a ) (b )

∑φ {V 1 − cos (φ − φ10 ) +V 2 1 − cos(2φ − φ20 )  +V 3 1 − cos ( 3φ − φ30 )  + } ∑K x2+ ∑∑F (b − b ) (b ′ − b ) + '


bb ′
 
0 0
   
x
1 b b'
x

(c )
U (d ) (e )
AN
Fθθ (θ − θ ) (θ ′ − θ ) + F θ (b − b )(θ − θ ) + ∑∑ (b − b ) [V cos φ +V cos 2φ +V cos 3φ ] +
∑∑ ∑∑
'
′ 0 0 b 0 0 0 1 2 3
θ θ'
θ b φ b
M

(f ) (g ) (h )

∑∑ (b ′ − b ) [V cosφ +V cos 2φ +V cos 3φ ] + ∑∑∑K


'
0 1 2 3
φ θ φ'
φθθ ′cos φ (θ − θ 0 ) (θ ′ − θ 0' ) +
∑ εr
qi q j
+
φ
D

b'
i>j ij

(i ) (j) (k )
TE

 A ij   B ij 
∑  r
i>j 
9 

  r 6  
  ij   ij  
EP

(l ) (2)
C

The energy terms were separated into three classes: the cross-terms, bonded and non-bonded
AC

energy terms. The bonded energy terms are made of (a) the covalent bond stretching, (b) the bond
angle bending, and (c) the torsion angle rotation energy terms of the polymer chain. A Fourier
series function was used to fit the energy of the torsion angle. The out-of-plane (improper) energy
(d) was considered as a harmonic function. The cross interactions terms included the dynamic
variations among the bond stretching, bending, and torsion angle rotation (e–j). The last two
terms, (k) and (l), represent the Coulombic electrostatic and the van der Waals force, respectively.
ACCEPTED MANUSCRIPT

These forces are interactive forces between silica networks and small molecules. The COMPASS
force-field is considered the first ab initio force-field that allows the precise and instantaneous
prediction of gas-phase properties (structural, conformational, and vibrational properties) and
condensed-phase properties (equation of state and cohesive energies) for a broad range of
molecules and polymers. This force-field is also a high value force-field for combining the

PT
parameters of organic and inorganic constituents [34].

RI
SIMULATION AND DATA COLLECTION
The following processes were performed to simulate the cross-linking procedure and to collect

SC
the required data. (1) Cell construction and energy minimization: an amorphous cell composed of
resins and hardeners (Table 1) was first created. The initial density of the cell was 1 g/cm3. The

U
mentioned cell was under periodic boundary condition and was made at room temperature.
Thereafter, 1000-step minimization was conducted to deliver the system to minimum available
AN
energy. (2) Pre-equilibration: 1 ns isothermal–isobaric (NPT) dynamic was needed to deliver the
system to the global equilibrium. The NPT must be conducted at 298 K and 1 atm, and a time
M

step of 1 fs should be used. (3) Addition of C60 with packing method: after preparing cell with
resins and hardeners, C60 molecules were added in two of the epoxy systems. Packing method
D

was used to add these molecules with constant volume cell. Constant volume helps to achieve
better comparison between cells. As shown in Tables 2 and 3, C60 molecules have been added.
TE

Therefore, three epoxy systems were obtained. Each of these systems should be minimized. (4)
Preparing covalent bonds between resins and hardeners: after minimizing, distances between the
EP

pair of atoms that can react were measured. Thereafter, the connection between them within
cross-linking cut-off distance was made. A smart minimization task was performed in each
C

connection. The system was then exposed to high temperature canonical (NVT) dynamics at 500
K and 1 atm. This procedure generates sufficient kinetic energy to the molecules and increases
AC

the opportunity of reacting sites to be posed in the reaction cut-off distance. In this work, the
relatively small cut-off distance of 6 Å restricted the maximum possible density of cross-linking
[5, 26]. (5) Post-equilibration: each final cross-linked structure was equilibrated over 2 ns NPT
dynamics, with the same process as in pre-equilibration level to reach the most stable
construction and to compute the interested properties [5, 29].
ACCEPTED MANUSCRIPT

The processes will be continued by collecting the desired data. Several consecutive NPT
dynamics at various temperatures were made to obtain glass transition temperature and material
properties including densities and cell volume in different temperatures. Mechanical properties at
different temperatures can also be achieved. The NPT processes were at 250 K to 600 K [26].
Each analysis started from the final configuration of the previous analysis [27].

PT
Calculation of Glass transition Temperature and Thermal Properties

RI
The glass transition temperature (Tg) is one of the significant thermal properties of epoxy
resins. In this temperature, the behavior of a polymer shifts from a glassy and brittle state to a

SC
rubbery state. The treating of polymers is influenced by the temperature range determined by Tg.
To determine the transition temperature, a few experimental approaches are available such as

U
thermo-mechanical analysis, dilatometry, differential scanning calorimetry, and dynamic
AN
mechanical analysis. Three usual theoretical methods wereused to estimate Tg by MD
simulations: (a) the property-temperature method, (b) the energy-temperature method, and (c) the
mean squared displacement [9]. In the present research, the first method was used. Moreover, the
M

coefficient of thermal expansion (CTE) was estimated by the slopes of volume change curves.
Their discontinuity concludes the Tg [27].
D
TE

Calculation of Mechanical Properties


Three main methods are used to calculate the mechanical properties by MD simulations: static
(constant−strain minimization), fluctuation formula, and dynamics (constant−stress MD) [2]. In
EP

the current study, the static method was used to calculate the elastic constants of the two cross-
linked epoxy systems. The stress–strain behavior in linear–elastic materials can be described by
C

Hook’s law:
AC

where i, j = 1, 2, 3. and are the stress and strain vectors, respectively. is the six-

dimentional stiffness matrix. The maximum strain in this calculation was set to 0.003. The stress
components were estimated using the so-called virial expression as follows [5]:
ACCEPTED MANUSCRIPT

where V is the volume; mk and uk stand for the mass and velocity of the kth particle, respectively;

PT
rkl is the distance between kth and lst particles, and flk denotes the force exerted on lst particle by
kth particle. Given the static conditions, the first term on the right hand side was omitted. The

RI
elastic modulus were then obtained by the first derivative of the virial stress equations regarding
the strain, . In other words, the full 6 6 stiffness matrix was formed from the slopes of

SC
∂ / ∂ in tension and shear [4].

Lamé coefficients and µ can be calculated by the following equations [26]:

U
AN
M
D
TE

The other elastic modulus and properties can be calculated from the above coefficients:
C EP
AC

where E, K, and G are Young’s, bulk, and shear moduli, respectively; illustrates the Poisson’s
ratio.
ACCEPTED MANUSCRIPT

RESULTS AND DISCUSSION


Glass transition temperature and thermal properties
Tg was estimated on the basis of the discontinuity in the slope of the density–temperature plot. Tg
can also be calculated from the temperature at the break in the volume variation curve, and was

PT
compared with that obtained from previous methods [4].
Fig. 6 shows the glass transition temperature (Tg) values of the three polymer structures
including LY/HY 5052 epoxy system, and fullerene-containing LY/HY 5052 with different

RI
fullerene weight fractions of 8.64% and 15.9%. The values were extracted from the variation of
the epoxy system density as a function of temperature. The density of each epoxy system at the

SC
certain temperature was calculated from the average density of the system in the total simulation
time. The abrupt change in the slope of the density–temperature curve determines the values of

U
Tg, which are ~415–425 K, 420–430 K, and 410–420 K for the epoxy systems with the cross-link
AN
density of 62.5%. The results indicated that the glass transition temperature has a slight variation
for different fullerene-containing epoxy systems. The calculated Tg has good agreement with the
experimental values [29]. The experimental and theoretical calculated Tg for different epoxy
M

systems are summarized in Table 4.


D
TE
C EP
AC

Fig. 6. Density of three epoxy system as a function of temperature with 62.5% cross-link density.
ACCEPTED MANUSCRIPT

Tg was also estimated through the volume change of the cell structure of epoxy system versus
temperature. Similar to the above results, the abrupt change in the slope of the curve determines
Tg. Fig. 7 shows the volume variations of the different epoxy systems with temperature. As
shown in Fig. 7, the values of Tg are estimated as ~430–440 K, ~420–430 K, and ~405–415 K for

PT
the epoxy systems with the cross-link density of 62.5%. The results indicated that the first
method (density–temperature) is closer than the second method to the experimental Tg value of

RI
~400 K [29].

U SC
AN
M
D
TE
EP

Fig. 7. Change in the volume of three cell structures with 62.5% cross-link density as a function of
temperature.
C

Table.4. Glass transition temperature of epoxy systems for cross-link density of 62.5%
AC

Epoxy system MD simulation MD simulation Ref.


(based on density) (based on volume change)
Ly/Hy 415-425 oK 430-440 oK 390-405 oK [44]
Ly/Hy/C60 (1) 420-430 oK 420-430 oK -
Ly/Hy/C60 (2) 410-420 oK 405-415 oK -
ACCEPTED MANUSCRIPT

The volumetric CTE for both glassy and rubbery (below and above the regions of Tg) is found
from the slope of the two curves. Their equations are shown in Fig. 7. The volume CTE α is
defined by the following equation:

PT
RI
SC
where V, P, and T are volume, pressure, and temperature, respectively. β is the linear thermal
expansion coefficient that is correlated to volume CTE. The volumetric CTE and linear thermal

U
expansion coefficients below and above Tg were estimated for three epoxy systems with cross-
link density of 62.5%, which are summarized in Table 5. The simulation results indicated that the
AN
CTEs have decreased by adding fullerene to resin epoxy system.
M

Table 5. Volumetric CTE and linear thermal expansion coefficients of three epoxy systems with 62.5%
cross-link density.

Material property LY/HY LY/HY/C60 LY/HY/C90(2)


D

MD Sim. Ref. (1)


(below Tg) 12.66 - 7.00 8.33
TE

(ppm/oK)
(above Tg) 18.33 - 15.70 10.66
EP

(ppm/oK)
(below Tg) 38.00 - 21.00 25.00
(ppm/oK)
C

(above Tg) 55.00 71.00 [22] 47.10 32.00


(ppm/oK)
AC

Mechanical properties
The estimated mechanical properties of three epoxy systems, including the density, cell volume,
and elastic constants of each system, are listed in Table. 6. When the C60 molecules have been
added at constant volume, the density of epoxy systems increases as the number of C60
molecules increases. The added fullerenes also cause to repel epoxies far away, thus leading to an
ACCEPTED MANUSCRIPT

increase in cell volume. The simulation results also illustrate that the fullerene- containing epoxy
resin systems have higher mechanical properties than the main epoxy resin system. However, the
improvements are accompanied by fluctuations. The results show that the fullerene-containing
epoxy resin system of 8.6% weight fraction has higher elastic moduli than others. This result can
be described by the fact that excess fullerene in fullerene-dispersed epoxy remains as aggregated

PT
particles. The results have good agreement with the experimental achievement by Jiang et al.
[29]. They showed that most fullerenes can be soluble in epoxy. Their suggestion indicated that

RI
the dispersion on the fullerene effects on the mechanical property of epoxy can be expressed by
the mechanism that is not only caused by the agglomerated fullerene nano-particles but also

SC
induced by the dissolved fullerenes [21].

U
Table 6. Mechanical properties of three epoxy systems with cross-link density of 62.5% at room
temperature.
AN
Epoxy System LY/HY LY/HY/C60 LY/HY/C60
(1) (2)
3
~1.07 ~1.12 ~1.16
M

Density (gr/cm ) Sim.


Ref. ~1.09 - -
LY=1.16-1.17 [44]
D

HY=0.93-0.95 [44]
4.27 5.12 4.59
TE

Young Modulus Sim.


(GPa) Ref. 3.20-3.90 [44] - -
Mechanical property

Bulk Modulus Sim. 3.38 3.63 3.18


EP

(GPa) Ref. - - -
Shear Modulus Sim. 1.65 2.03 1.86
(GPa) Ref. - - -
C

Poisson’s Ratio Sim. 0.29 0.25 0.27


AC

Ref. ~0.35 - -
Cell Volume (nm3) Sim. 35969.68 37126.00 39027.21
Ref. - - -

Figs. 8–12 show the mechanical properties of three epoxy systems, including Young’s, shear,
and bulk moduli, as well as densities and cell volume, for the epoxy systems with the cross-link
ACCEPTED MANUSCRIPT

density of 62.5% at four different temperatures. Nearly all of the calculated results for mechanical
properties, such as Young’s modulus, are larger than the experimental results because the
simulated system has no defects and is an ideal situation, whereas several types of defects are
present in real samples that result in the reduced measured properties. As shown in Table 6, the
calculated value of the density is lower than experimental value. By increasing the number of

PT
atoms in the cell, better material properties can be obtained [30]. Obtaining high value of
Poisson’s ratio and elastic modulus at temperature above Tg is expected, which will be around

RI
three orders of magnitude lower than the glassy state (below Tg) [31].

U SC
AN
M
D
TE
EP

Fig. 8. Calculated Young’s Modulus of three epoxy systems with 62.5% cross-link density at room
temperature.
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
Fig. 9. Calculated Bulk Modulus of three epoxy systems with 62.5% cross-link density at room
temperature.
M
D
TE
C EP
AC

Fig. 10. Calculated Shear Modulus of three epoxy systems with 62.5% cross-link density at room
temperature.
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
Fig. 11. Calculated Density of three epoxy systems with 62.5% cross-link density at room temperature.
M
D
TE
C EP
AC

Fig. 12. Calculated Cell Volume of three epoxy systems with 62.5% cross-link density at room
temperature.
ACCEPTED MANUSCRIPT

Local structure
The radial distribution function (RDF), g(r)a–b, which is obtained directly from the
simulations is another parameter to study the effect of various factors, such as cross-linking
density, fillers, and additives, on the local structure of the simulated epoxy resin systems. g(r) is

PT
also defined as the probability of finding a given particle at a distance of r from a reference
particle [32, 37]. The RDF is calculated by the following:

RI
SC
where i and j are the ith and jth atoms of the group A of NA atoms and B of NB atoms,

U
respectively; NAB is the number of atoms common to both groups A and B; angle brackets imply
AN
by an average over different structures. This method has been verified to be an actual way of
recitation of the typical construction of a polymer [2, 32].
The RDF curves of all three models are shown in Fig. 13. The peaks around 0.9–1.1 Å are
M

attributed to the chemical bonds between hydrogen and other atoms. The peaks between 1.4 and
1.45 Å correspond to the C−N and C−O bonds, whereas those at nearly 1.75 Å come from the
D

correlation between hydrogen atoms in the methyl (−CH3) and methylene (−CH2−) assemblies.
TE

The following intra-molecular peaks result from distances between atoms, including hydrogen
and carbon in H−C−C (r=2.16 Å) and carbon atoms in C−C−C sequences (r=2.44 Å) [28]. As
shown in Fig. 13, by increasing the content of fullerene, the total-molecular RDFs decrease for
EP

the direct chemical bonds between hydrogen and other atoms. However, the RDF behavior
changes for other bonds. In the case of hydrogen and other atom bonds, the longer chains will be
C

created by adding fullerene. The chains can also be the result of the growth of polymer networks.
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
Fig. 13. Total radial distribution function for 62.5% cross-link density at room temperature for three epoxy
systems.
M

CONCLUSION
The effects of fullerene as an addition on the thermo-mechanical properties of Araldite LY
D

5052/Aradur HY 5052 cross-linked resin epoxy were numerically studied by MD simulation.


Three different Araldite LY 5052/Aradur HY 5052 cross-linked resin epoxy systems were used in
TE

the simulations, of which one has no fullerene and two have different weight fractions of 8.64%
(LY/HY/C60 (1) epoxy system) and 15.92% (LY/HY/C60 (2) epoxy system). The thermo-
EP

mechanical properties of the considered epoxy systems were comparatively studied. The
simulation results showed that different fullerene-containing epoxy systems have the same values
C

of Tg with slight variations. Moreover, CTEs decrease by adding fullerene. The results also
indicated that mechanical properties such as Young’s, bulk, and shear moduli, as well as
AC

Poisson’s ratio and density, improves considerably by adding fullerene to epoxy systems.
Furthermore, the best mechanical properties were obtained for the weight fractions of 8.64%,
which was in agreement with the experimental results. Finally, the effects of fullerene on the
evolution of the local structures were investigated by studying their RDFs.
ACCEPTED MANUSCRIPT

Acknowledgment

The support from Shahid Beheshti University is gratefully acknowledged.

References:

PT
[1] J. J. Chruściel and E. Leśniak, "Modification of epoxy resins with functional silanes,
polysiloxanes,silsesquioxanes, silica and silicates," Progress in Polymer Science, 2014.

RI
[2] A. Shokuhfar and B. Arab, "The effect of cross linking density on the mechanical properties and
structure of the epoxy polymers: molecular dynamics simulation," J Mol Model, vol. 19, p. 3719–

SC
3731, 2013.

[3] B. Qi, Q. X. Zhang, M. Bannister and Y. -W. Mai, "Investigation of the mechanical properties of
DGEBA-based epoxy resin with nanoclay additives," Composite Structures, vol. 75, p. 514–519,

U
2006.
AN
[4] C. Wu and W. Xu, "Atomistic molecular modelling of crosslinked epoxy resin," Polymer, vol. 47, pp.
6004-6009, 2006.
M

[5] F. Jeyranpour, G. Alahyarizadeh and B. Arab, "Comparative investigation of thermal and mechanical
properties of cross-linked epoxy polymers with different curing agents by molecular dynamics
simulation," Journal of Molecular Graphics and Modelling, vol. 62, p. 157–164, 2015.
D

[6] H. Corporation, "Advanced Material Composites Selector Guide," 2012.


TE

[7] D. A. Jorba and E. L. Thomas, "Foams of graphene, method of makin and materials made thereof".
United States 20 Dec. 2012.
EP

[8] H. Company, "www.silmid.com," Sil-Mid Limited, 2012. [Online]. Available:


https://www.silmid.com/products/hy505219kg-aradur-5052ch-hardener-19kg.aspx.
C

[9] N. Company, "http://www.nedform.com," Nedform, 1999. [Online]. Available:


http://www.nedform.com/magento/en/araldite-ly-5052-aradur-5052.html.
AC

[10] B. Bolasodun, O. Rufai, A. Nesbitt and R. Day, "Comparison of the Isothermal Cure Kinetics of
Araldite LY 5052 / 4 4’ DDS Epoxy System Using a Differential Scanning Calorimetry and a
Microwave Heated Calorimeter," International Journal of Materials Engineering, vol. 4, no. 4, pp.
148-165, 2014.

[11] R. YUSOFF, M. K. AROUA, A. NESBITT and R. J. DAY, "Curing of polymeric composites using
microwave," Journal of Engineering Science and Technology, vol. 2, no. 2, pp. 151 - 163, 2007.
ACCEPTED MANUSCRIPT

[12] S. Sprenger, "Epoxy resins modified with elastomers and surface-modified silica nanoparticles,"
Polymer, vol. 54, pp. 4790-4797, 2013.

[13] L. Rajabi, Z. Mohammadi and A. A. Derakhshan, "Thermal Stability and Dynamic Mechanical
Properties of Nano and Micron-TiO2 Particles Reinforced Epoxy Composites: Effect of Mixing
Method," Iranian Journal of Chemical Engineering, vol. 10, no. 1, pp. 16-29, 2013.

PT
[14] R. Kochetov, T. Andritsch, P. H. Morshuis and J. J. Smit, "Thermal and Electrical Behaviour of
Epoxy-based Microcomposites Filled with Al2O3 and SiO2 Particles".

RI
[15] M. J. Kadhim, A. . K. Abdullah, I. A. Al-Ajaj and A. S. Khalil, "Mechanical Properties of
Epoxy/Al2O3 Nanocomposites," International Journal of Application or Innovation in Engineering
& Management, vol. 2, no. 11, pp. 10-16, 2013.

SC
[16] Q. L. Xiong, S. A. Meguid, Y. Wang and G. J. Weng, "Molecular dynamics and atomistic based
continuum studies of the interfacial behavior of nanoreinforced epoxy," Mechanics of Materials, vol.

U
85, p. 38–46, 2015. AN
[17] J. Gou, B. Fan, G. Song and A. Khan, "Study of affinitties between single-walled nanotube and epoxy
resin using molecular dynamics simulation," International Journal of Nanoscience, vol. 5, no. 1, p.
131–144, 2006.
M

[18] C. Li, A. R. Browning, S. Christensen and A. Strachan, "Atomistic simulations on multilayer


graphene reinforced epoxy composites," Composites: Part A, vol. 43, p. 1293–1300, 2012.
D

[19] C. E. Pizzuto, J. Suave, J. Bertholdi, S. H. Pezzin and L. A. Coelho, "Mechanical and Dilatometric
Properties of Carboxylated SWCNT/Epoxy Composites: Effects of the Dispersion in the Resin and in
TE

the Hardener," Journal of Reinforced Plastics and Composites, vol. 29, no. 4, pp. 524-530, 2010.

[20] P. J. F. Harris, "Carbon nanotube composites," International Materials Reviews, vol. 49, no. 1, pp.
31-43, 2004.
EP

[21] T. Ogasawara, Y. Ishida and T. Kasai, "Mechanical properties of carbon fiber/fullerene-dispersed


epoxy composites," Composites Science and Technology, vol. 69, p. 2002–2007, 2009.
C

[22] H. Corporation, "http://www.compositesone.com," 2010. [Online]. Available:


AC

http://www.compositesone.com/case-study/technical-data-sheet-araldite-ly-8605-aradur-8605-
system/.

[23] X. Wang, L. Mao, M. Luo, P. Fang, Y. Dai and K. M. Liew, "Study of fullerene-containing epoxy
membranes with tunable ultraviolet light-filtering properties," Progress in Organic Coatings, vol. 67,
pp. 398-404, 2010.

[24] A. O. Okonkwo, P. Jagadale, J. E. Garcia Herrera, V. G. Hadjiev, J. . M. Saldana, A. Tagliaferro and


F. . C. Robles Hernandez, "High-toughness/low-friction ductile epoxy coatings reinforced with
ACCEPTED MANUSCRIPT

carbon nanostructures," Polymer Testing, vol. 47, pp. 113-119, 2015.

[25] ""‘BIOVIA Material Studio 6.0, http://accelrys.com/products/materials-studio,"," [Online].

[26] B. Arab and A. Shokuhfar, "Molecular Dynamics Simulation of Cross-Linked Epoxy Polymers: the
Effect of Force Field on the Estimation of Properties," Journal of nano and electronic physics, vol. 5,
no. 1, 2013.

PT
[27] B. Arab and A. Shokuhfar, "Molecular dynamics simulation of cross-linked urea-formaldehyde
polymers for self-healing nanocomposites: prediction of mechanical properties and glass transition

RI
temperature," Journal of molecular modelling, vol. 19, pp. 5053-5062, 2013.

[28] C. Wu and W. Xu, "Atomistic molecular simulations of structure and dynamics of crosslinked epoxy

SC
resin," Polymer, vol. 48, pp. 5802-5812, 2007.

[29] Z. Jiang, H. Zhang, Z. Zhang, H. Murayama and K. Okamoto, "Improved bonding between PAN-
based carbon fibers and fullerene-modified epoxy matrix," Composites Part A, vol. 39, pp. 1762-

U
1767, 2008.
AN
[30] N. B. Shenogina, M. Tsige, S. . M. Mukhopadhyay and S. S. Patnaik, "MOLECULAR MODELING
OF THERMOSETTING POLYMERS: EFFECTS OF DEGREE OF CURING AND CHAIN
LENGTH ON THERMO-MECHANICAL PROPERTIES," in 18TH INTERNATIONAL
M

CONFERENCE ON COMPOSITE MATERIALS, Korea, 2011.

[31] N. B. Shenogina, M. Tsige, S. S. Patnaik and S. M. Mukhopadhyay, "Molecularmodeling of elastic


D

properties of thermosetting polymers using a dynamicdeformation approach," Polymer, vol. 54, no.
13, pp. 3370-3376, 2013.
TE

[32] J. . S. Bermejo and C. M. Ugarte, Macromolecular Theory and Simulations, vol. 18, pp. 317-327,
2019.
EP

[33] J. Meller, "Molecular Dynamics," Encyclopedia of life science, 2001.

[34] H. Sun, Macromolecules, 1995, 28, 701-712.


C
AC
ACCEPTED MANUSCRIPT
Highlights

• Thermo-mechanical properties of the cross-linked epoxies were estimated by MD

simulations.

• The effects of fullerene on the thermo-mechanical properties of LY5052/HY5052

were studied.

PT
• Tg of the epoxy systems calculated through density-temperature, and cell volume

RI
variation.

• Tg has slight variations for different fullerene-containing epoxy systems.

SC
• Higher mechanical properties were observed for Araldite LY 5052 / Aradur HY 5052/

C60.

U
AN
M
D
TE
C EP
AC

You might also like