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

Materials Science and Engineering A 487 (2008) 574–585

Fabrication and characterization of TiO2–epoxy nanocomposite


Amit Chatterjee a,b,∗ , Muhammad S. Islam a
a Center for Composite Materials, University of Delaware, Newark DE-19716, DE, USA
b Navel Materials Research Laboratory, DRDO, Ambernath (E) 421506, India

Received 27 August 2007; received in revised form 15 November 2007; accepted 15 November 2007

Abstract
A systematic study has been conducted to investigate the matrix properties by introducing nanosize TiO2 (5–40 nm, 0.5–2% by weight) fillers
into an epoxy resin. Ultrasonic mixing process, via sonic cavitations, was employed to disperse the particles into the resin system. The thermal,
mechanical, morphology and the viscoelastic properties of the nanocomposite and the neat resin were measured with TGA, DMA, TEM and Instron.
The nano-particles are dispersed evenly throughout the entire volume of the resin. The nanofiller infusion improves the thermal, mechanical and
viscoelastic properties of the epoxy resin. The nanocomposite shows increase in storage modulus, glass transition temperature, tensile modulus,
flexural modulus and short beam shear strength from neat epoxy resin. The mechanical performance and thermal stability of the epoxy nanocom-
posites are depending on with the dispersion state of the TiO2 in the epoxy matrix and are correlated with loading (0.0015–0.006% by volume). In
addition, the nanocomposite shows enhanced flexural strength. Several reasons to explain these effects in terms of reinforcing mechanisms were
discussed.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Epoxy; TiO2 ; Nanocomposite

1. Introduction [2]. A new approach aiming to overcome this basic problem is


related to the nanotechnology and uses of fillers in the nanometer
High performance polymer composite materials are nowa- scale. Nanoparticles embedded in polymer matrix have attracted
days has wide application under hard working conditions. The increasing interest because of the unique mechanical, optical,
materials must provide unique mechanical and thermal proper- electrical and magnetic properties displayed by nanocomposites.
ties combined with a low specific weight and high resistance to Due to nanometer size of these particles, their physicochemi-
degradation in order to ensure safety and economic efficiency. cal characteristics differ significantly from those of molecular
One of the composite materials for technical applications may be and bulk materials [3,4]. Nanoparticle reinforced polymers, syn-
represented by thermoset polymer matrix, epoxy resins, which ergistically combine the properties of both the host polymer
already covers along some of the demanded properties. Epoxy matrix and the discrete nanoparticles [4]. The approach demon-
resins are a class of versatile thermosetting polymers, which are strates the potential to change characteristics of thermosetting
widely used in structural composites, adhesives, surface coatings and thermoplastic polymers fundamentally to improve their
and electronic circuit board laminates [1]. However, the poly- general performance [2,5]. Nanocomposites include more than
mer matrix must withstand high mechanical loads; it is usually one solid-phase, where at least one-phase dimension is in the
reinforced with fillers. The addition of different fillers favorably nanometer range. Typically, all solid phases, may be amorphous,
stiffens the material and may also increase the strength under semicrystalline or crystalline, are in the 1–20 nm range [6]. The
certain load conditions. But the micro-fillers have a detrimental mechanical properties of the nanocomposite are expected to be
effect on certain important properties such as impact resistance much higher compared to the micro-filler because of the par-
ticle size. The absolute number of particles would rise rapidly
by a factor of 1000 after imaginary changing the particle size

by a factor of only 10, e.g. 1 ␮m to 100 nm [7]. The resultant
Corresponding author at: Gentex Corporation, 324 Main Street, Carbondale,
PA-18407, USA.
composite properties should be more improved than expected
E-mail address: chatterjeamit@yahoo.com (A. Chatterjee). from the individual components because the nanoparticles pro-

0921-5093/$ – see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2007.11.052
A. Chatterjee, M.S. Islam / Materials Science and Engineering A 487 (2008) 574–585 575

vide a very high specific surface area. This area can generate a capacitors, solid oxide fuel cell, UV protection and waste
new material behavior, which is widely determined by interfacial water purification. So, epoxy–TiO2 nanocomposite may have
interactions, offering unique properties and a completely new wide application, particularly for bio-terrorism, organic photo-
class of materials [8]. The fundamental change in the material voltaic, fire retardant composite, composite nanomenbrane, etc.
properties arises becomes obvious, if one considers for exam- Nanocomposites using thermoplastic polymers are well known
ple a small crack propagating through the material, which is for improving the mechanical, electrical, thermal and insulating
interacting with many nanoparticles instead of crossing only a properties. However, nanocomposite using thermosetting poly-
few microparticles. So, higher energy consumption would be mers have not been studied so extensively [15,16] particularly
expected. It was already verified by several research groups that using TiO2 . The influence of nanoclay on the mechanical and
nanoparticles are reinforced effectively into the thermoplastic surface properties of epoxies has been characterized [15–17].
and thermosetting polymer matrices [9] and that enhances the The influence of TiO2 on epoxy resin, SC79 is reported in this
properties. Specially, the reinforcement covers improvement of study for the first time. We focused on the performance opti-
the flexural modulus without loosing flexural strength. This mization of composites containing TiO2 nano-particles and to
effect is at the same time accompanied by improvements in understand the role of nanoparticles. The nanosized TiO2 par-
fracture toughness and impact energy which, however, depend ticles have been infused through sonic cavitations in an epoxy
strongly on the filler volume content. The nano-filler induced resin matrix. An ultrasonic mixing procedure was developed
micromechanical deformation processes can be responsible for to uniformly disperse the particles into the resin system. The
this behavior [10]. mechanical and thermal performance of the nano-filled polymer
However, the unique nanocomposite effect can only be effec- matrix as a function of the nanopartcicle content (0.5–2%) was
tive, if the nanoparticles are well dispersed in the surrounding evaluated. The expected overall properties of nanocomposites
polymer matrix. It is reported that a considerable amount of may open the way towards new applications of high perfor-
improvement in mechanical properties can be achieved using mance polymers, leading to an innovative product development
very low amount of nanopartcicles loadings [1–10]. A neg- in the automotive industry, electronics for coatings and many
ligible contribution made by the interphase provides diverse other applications.
possibilities of performance tailoring and is able to influence
the properties of the matrices to a much greater extent under 2. Experimental
rather low nanofiller loading. Significant improvement in the
tensile performance of polypropylene composites has also been 2.1. Materials and methods
reported in terms of stiffening, strengthening and toughening
with a low filler content of about 0.5%. The mechanical per- The epoxy resin used in this investigation has two part,
formance significantly depends on the dispersion state and the part A (diglycidylether of bisphenol A 60–70%, aliphatic
microstructural homogeneity of the nanoparticles [11]. The dis- diglycidylether 10–20%) and part B (cycloaliphatic 85–90%,
persion method for the nanoparticle incorporation in polymer proprietary amine 10–15%, toughener 10–20%) of Applied Pol-
matrix must be very vital footstep and key point to receive the eramic Inc., Benecia, CA, USA. The nanoparticles used here is
desired material properties. TiO2 of 5 nm size supplied by Nanostructured & Amorphous
Nanoparticles were generally introduced into polymer matrix Materials Inc., Los Alamos, NM, USA. The properties of the
using various techniques. The various dispersion processes are particles are given in Table 1.
necessary in order to transfer the nanoparticles from the agglom-
erated state into a homogeneously dispersed state. The direct
2.2. Fabrication
incorporation using chemical methods and by application of
high shear forces during mechanical powder dispersion pro-
There are various techniques such as melt mixing, solution
cess [12–14] are the popular one. Ultrasound vibration was
mixing; shear mixing and mechanical stirring to infuse nanopar-
used to disperse nanoparticles, which also helps to improve the
ticles into a polymer are available. Acoustic cavitations is one of
dispersion state of nanoparticles. Chemical methods are able
the efficient ways to disperse nanoparticles into virgin materials
to generate individual and non-agglomerated nanoparticles “in
[18]. In this case, the application of alternating acoustic pres-
situ” within a thermosetting/thermoplastic polymer. However,
sure above the cavitation threshold creates numerous cavities
an additional chemical treatment of the nanoparticle surface
in the liquid. Some of these cavities oscillate at a frequency of
may further enhance the composite’s properties by improving
the applied field (usually 20 kHz) while the gas content inside
the filler-matrix coupling quality [12–14].
these cavities remain constant. However, some other cavities
The TiO2 is extensively used in the industries such as addi-
tives in plastics, agglomerates for thermal sprays, air/fuel ratio
controller in automobile, attenuation of ultraviolet light, cat- Table 1
Mechanical properties of TiO2
alysts and catalyst supports, demilitarization of chemical and
biological warfare agents, electrode materials in lithium bat- Comp. strength (MPa) 800–1000
teries, energy converter in solar cells, gas sensors, inorganic Hardness (kg f mm−2 ) 980
Tensile strength (MPa) 350
membranes, photo catalytic degradation of bacteria and grime, Ten. Mod. (GPa) 200–300
photochemical degradation of toxic chemicals, piezoelectric
576 A. Chatterjee, M.S. Islam / Materials Science and Engineering A 487 (2008) 574–585

grow intensely under tensile stresses while yet another portion Flexural properties of the materials were obtained according
of these cavities which are not completely filled with gas, start to ASTM D-790. Samples length to depth ratio was kept higher
to collapse under the compression stresses of the sound wave. than 16:1. Flexural load was employed in an Instron under three
In the later case, the collapsing cavity generates tiny particles of point bend fixture and the loading rate was applied according to
debris and the energy of the collapsed one is transformed into the specimen dimension as defined in ASTM D-790.
pressure pulses. It is note worthy that the formation of the debris Dynamic mechanical analysis was conducted in a TA
further facilitates the development of cavitation. It is assumed dynamic mechanical analyzer (model DMA 2980) at a frequency
that acoustic cavitation in liquids develops according to a chain of 1 Hz using dual cantilever clamp with amplitude of 15 ␮m.
reaction. Therefore, individual cavities on real nuclei develop The samples were ramped from room temperature to 250 ◦ C
so rapidly that within a few microseconds an active cavitation using a ramp rate of 2 ◦ C/min. The storage modulus and loss
region is created close to the source of the ultrasound probe. modulus were measured. The loss modulus peak is used for
The development of a cavitation processes in the ultrasonically measuring the Tg of the resin. The nanocomposite samples were
processed melt creates favorable conditions for the intensifica- directly cut into rectangular shapes 60 mm × 13 mm × 3 mm for
tion of various physio-chemical processes. Acoustic cavitation DMA analysis.
accelerates heat and mass transfer processes such as diffusion, Thermogravimetric analysis (TGA Q-500, TA Inc.) was used
wetting, dissolution, dispersion and emulsification. Recently is to measure the degradation properties of the nanocomposites.
has been reported that there is a clear acceleration of polymer The experimental parameters used for TGA included a heating
reaction under ultrasound in both catalyzed and uncatalysed rate of 10 ◦ C/min, room temperature to 900 ◦ C and a flow rate
reactions [19]. In the present investigation ultrasonic mixing of helium purging gas is 100 ml/min.
was employed to infuse TiO2 nanoparticles into the epoxy Tensile properties of the materials were obtained according to
resin. ASTM D-638. Samples were prepared according to the ASTM
Titanium nanoparticles were added to the epoxy part A at standard. The bidirectional stain gages were mounted in the sam-
varying percentage in a plastic beaker and mixed by hand as an ples to measure the axial and the transverse stain. The tensile
initial dispersion. The mixer was then placed for the sonication load was employed in an Instron. The loading rate was applied
process. Vibracell ultrasonic processor shown in Fig. 1 was used according to the specimen dimension as defined in ASTM D-
for dispersion. The maximum power output is 750 watts and is 638. The short beam shear testing was conducted using ASTM
applied at a frequency of 20 kHz using a 12.5 mm diameter tita- D-2344. The sample preparation, conditioning, testing and the
nium alloy tip. An amplitude of 15 ␮m is applied with a pulse (9 s analysis was done as defined in ASTM D-2344.
on,1 s off) for about 20 min. During sonication, heat was builds
up. In order to avoid temperature rise during sonication, exter- 3. Results and discussion
nal cooling was employed by submerging the mixing beaker in
a mixture of ice and water. The equivalent amount of part B 3.1. Morphology
was added in the mixture, and mixing was continued for another
10 min using a high speed mechanical stirrer. After degassing, Dispersion of nanoparticles in the epoxy resin is very diffi-
the resin mixture was poured into a mold. The sample was cured cult because of the Van Der Waal’s force between the particles.
for about 28 h at room temperature and postcured at 93 ◦ C for Resin viscosity plays a significant role for the uniform disper-
about 8 h. The postcured resin was removed from the mould and sion of nano-particle. We have chosen a low viscosity VARTM
samples were prepared for various tests discussed below. epoxy resin (300 cp) for this work that helps to disperse the
Transmission electron microscopy was conducted in JEM- particles uniformly. To optimize the dispersion the sonication
200 FX. Thin sections (70 nm) of the specimen were obtained by time was varied from 0 to 120 min. The best result was obtained
microtome with diamond knife for TEM analysis. The filament using a total mixing time of 30 min. Further increment of disper-
voltage was set at 200 KV to make a bright field image of the sion time increases the temperature and viscosity of the resin.
nanostructure. The resin starts to cure that decreases the dispersion of the
nano-particles. Transmission electron microscopy (TEM) analy-
sis was conducted to verify the size and dispersion quality of the
nanocomposite. Fig. 2 is the TEM image of an agglomeration of
TiO2 nanoparticles prior to adding to the epoxy. Fig. 3 shows the
TEM picture of the 1% TiO2 –epoxy nanocomposite. In Fig. 3
it is shown that the particles have been effectively dispersed in
the epoxy. The nano-particles are mostly circular in shape. The
size varies from 5 to 10 nm. Most of the particles are distributed
uniformly. However, some residual agglomeration of particles
remains (20–50 nm) (Fig. 3). Van Der Waal force between the
particles is the leading reason for weak dispersion of the nano-
particles in the epoxy resin matrix. The effect of loading on
dispersion is also observed. Increasing on loading creates non-
Fig. 1. Ultrasonic Vibracell. uniformity that clearly observed in the TEM pictures. However,
A. Chatterjee, M.S. Islam / Materials Science and Engineering A 487 (2008) 574–585 577

Fig. 4. Loading effect on flexure properties of epoxy–TiO2 nanocomposite.

matrix. The particle–particle interaction is high for smaller size.


Fig. 2. TEM image of powder TiO2 (5 nm). The applied ultrasonic energy and time of dispersion are unable
to separate the all particle. So, the particles agglomerate and
eventually form lumps that we have seen here. Efforts are con-
tinuing to focus on improving more homogenous dispersion of
the particles.

3.2. Mechanical performance of nanocomposite

Two types of mechanical properties, flexural and tensile were


performed to evaluate the stiffness and strength of the materials.
Five samples of each type were tested for measuring the flexural
properties of neat epoxy resin and nanocomposite samples using
ASTM-D790. In the case of 1% loading, the 5 nm TiO2 –epoxy
nanocomposite shows 5% increase in modulus and an increase in
flexural strength (Table 2). For the different amount of loading,
the modulus is observed to decrease as increasing the loading to
2% for the 5 nm particle. A typical stress versus displacement
behavior from the flexural test is shown in Fig. 4. The energy
absorption behavior of the neat resin and the nanocomposite are
also plotted and presented in Fig. 5. One percent infusion the
strain, stiffness and energy absorption of the nanocomposite is
increased. However, further increment of % loading decreases
Fig. 3. TEM image of TiO2 –epoxy nanocomposite (5 nm, 1% TiO2 ).
the mechanical properties of the nanocomposites from the neat
resin. The influence of rigid micro-particulate fillers on the
1% loading gives better dispersion with homogeny. Size effect stress–stain behavior of polymers is well known. Micro-fillers
of nano-particle also plays a big role for uniform dispersion. commonly increase the stiffness, but on the other side they may
Larger particle size improves dispersion. Smaller size has more have a detrimental effect on the flexural stain to break [20]. The
Van Der Waal force between the particles and that decrease the flexural strength of microparticle filled composites is known to
dispersion capability of the TiO2 nano-particles in the epoxy be reduced with rising filler content [20–21]. The results mea-

Table 2
Mechanical properties for TiO2 –epoxy Nanocomposite
Materials Flexure strength (KSi) Flexure mod. (KSi) Tensile strength (PSi) Tensile mod. (KSi) SBS strength (KSi) % Elongation to break (T)

Neat resin 18.0 ± 0.58 487 ± 9.72 7321 ± 755 418 ± 122 2.41 ± 0.1 4.28
1.0% TiO2 18.9 ± 0.28 514 ± 9.64 5424 ± 728 456 ± 60 2.81 ± 0.05 3.22
2.0% TiO2 17.3 ± 0.69 440 ± 1.92 5024 ± 498 476 ± 45 2.15 ± 0.05 2.92
578 A. Chatterjee, M.S. Islam / Materials Science and Engineering A 487 (2008) 574–585

Fig. 5. Effect of nanoparticle loading on energy absorption of nanocomposite. Fig. 6. Effect of weight percent loading on the stain at break of the TiO2 –epoxy
nanocomposite.

sured for the nanocomposites in this study offer an apparent


conflict with the known behavior. Flexural properties increase Another important aspect, which is more related to the dis-
with addition of nano-particulates (Table 2). persion state of the nanoparticles, may additionally play a
Some important characteristic of composites have to be con- significant role. For the case that relatively large agglomerates
sidered in order to explain this phenomenon. The quality of the would remain in the matrix, a propagation crack could encounter
interface in composites, i.e. the static adhesion strength as well a stress concentration locally and then easily induce the ini-
as the interfacial stiffness, usually plays a very important role tiation of the final failure. The embrittling effects occur only
in the materials’ capability to transfer stresses and elastic defor- at higher filler contents, where more agglomerates are likely
mation from the matrix to the fillers [22]. This is especially to be formed. The improved properties at low filler contents,
true for nanocomposites, because they impart a high portion of testify lower stress concentrations and therefore, the remain-
interface. If the filler matrix interaction is poor, the particles are ing nanoparticle clusters should be small. This demonstrates
unable to carry any part of the external load. In that case, the the importance of a homogeneous distribution of nanoparticles
strength of the composite cannot be higher than that of the neat within the matrix. The more nanoparticles are in contact with the
polymer matrix. If the bonding between the fillers and matrix is polymer, the higher is a probability of interaction. Large contact
instead strong enough, the yield strength of a particulate com- areas benefit an efficient stress transfer between both parts and
posite can be higher than that of the matrix polymer [23]. In the result in improved flexural stiffness, strength and failure stain.
same way corresponds a high interfacial stiffness and to a high Visual examination on the fracture surfaces of nanocomposite
composite modulus. Hence, the gradual increase in stiffness and by SEM methods shows detailed information on the cause and
flexural strength, as observed for the nanocomposites, reveals location of failure. Some specimens were chosen to study the
the stresses are efficiently transferred via interface. The stain at acting deformation mechanisms responsible for the reinforce-
break usually declines with rising filler content. Low filler con- ment. Several mechanisms of crack are applicable for filled
tent can cause dramatic drop in the fracture strain. The composite epoxy composites [24]. The crack may pass through particles
is partly filler and part matrix. Due to rigid nature of the fillers, if the fillers are weak, known as trans-particulate fracture, or
the deformation comes from the polymer. The actual deforma- they may pass around if the particles are string enough. The
tion experienced only by the polymer matrix is much larger failure may occur by interfacial debonding or by cohesive fail-
than the measured deformation of the sample with the result ure of the matrix. Some of the mechanisms are proven true for
that the polymer reaches the failure strain limit at a lower total nanocomposites in the present study. The SEM was also done for
deformation. Hence, total composite strain-to-break decreases. the neat resin and the nanocomposite samples. The neat epoxy
However, it is surprisingly observed that for the nanocomposites matrix resin system shows brittle behavior characterized by
in the present study, the stain-to-break behaves contrary to con- large smooth areas, large hyperbolic markings, ribbons and frac-
ventionally filled composites (Fig. 6). It tends rather to higher ture steps in the direction of crack propagation. In contract the
values as long as the filler content does not surpass 1% loading nanocomposite surfaces are rougher structured, clearly reveals
by weight (Fig. 6). This increase suggests that the nanoparti- many hyperbolic markings opening in the direction of crack
cles are able to introduce additional mechanisms of failure and propagation. The similar observation was reported by scientists
energy consumption without blocking matrix deformation. Par- and those tallies well with our findings [2]. It is already reported
ticles may induce matrix yielding under certain conditions and that the void generation in the nanocomposite are less compared
may furthermore act as stoppers to crack growth by pinning the to the neat resin system [25]. The reduction in void content in the
creaks [24]. Nevertheless, if the fillers exceed 1% loading by matrix increases the strength of the nanophase composite. With
weight the failure stain undergoes a slight decay. Such a reduc- incremental of particle loading above 1%, the number of par-
tion proposes that the large number of fillers now dominate, and ticles increases, that cause particle–particle interaction rather
they reduce the matrix deformation by restraining mechanically. than the particle–matrix interaction. So, the particles begin to
A. Chatterjee, M.S. Islam / Materials Science and Engineering A 487 (2008) 574–585 579

agglomerate and form lumps which eventually affect the Van


Der Waals interaction between the polymer chains and that may
cause decrease in properties.
The volume fractions for the 5 nm of 0.5, 1, 1.5, and 2%
by weight are 0.0015, 0.003, 0.0045, and 0.006, respectively.
Based on typical properties of Ti02 given in Table 1, and consid-
ering composite micromechanics [26], we have calculated the
modulus of the nanocomposite having different volume frac-
tions. The values for 0.5, 1, 1.5 and 2% by weight of 5 nm TiO2
are 4.88E5, 4.9E5, 4.91E5 and 4.93E5 PSI. Our experimental
value for the 0.003 volume fraction is about 5% higher than the
calculated value. The increase in the modulus is reflected due
to the nano-effect which directly influences the bulk material
properties.
Nanoparticle infusion has been typical with the doping of
nanoclays or nanoparticles in polymeric matrices [15,16,27–29].
This is quite understandable from the fact that a highly stiffer
material is being mixed with relatively softer epoxy resin. A rule
of mixture calculation based on volume fraction and individual
stiffness would certainly reflect the trend [30,31]. However, our
result did reflect the improvement in modulus and strength of Fig. 7. Tensile modulus of neat resin and nanocomposite.
the nanocomposite. In the case of strength, the influence of par-
ticles will also be strongly dependent on the interfacial adhesion
between the particle surface and the epoxy molecule discussed then it propagate towards the thickness finally it flailed at the
previously. No special sizing was used in the present study to tension side. For both of the cases failure is identical, but from
promote adhesion. Results in Table 2 shows that the particles the experiment it is observed that the crack initiation occurred at
significantly reduce flexural strength except in the case of the higher load in nanocomposite compare to the neat resin. Since
1% loading of 5 nm particles. Overall improvements might be the particle size is very small, the surface area is higher. It offers
possible with surface treatments. In addition, the large reduction more interfacial surface interaction, which is strong in this case,
for the 2% of 5 nm particles may be a consequence of a higher and the stress transfer through this strong interfacial bonding;
level of agglomerated particles that serve as defects sites. The delay the crack propagation resulting a higher strength.
mechanical load imposed on the nanocomposites is effectively As started earlier, the investigation began with the infusion
transferred to the TiO2 from the matrix. It is well known that of nanoparticles into epoxy matrix. The tensile properties were
good interfacial adhesion between fillers and polymer matrix measured. The tensile responses of these systems are shown in
restricted the molecular motions of the polymer networks that Fig. 7 (Table 2). It is seen in Fig. 7 that epoxy nanocomposite
may cause the improvement of the flexural properties. However improve the tensile properties. A batch of epoxy matrix was also
after 1% of the filler content, this particle gets agglomerated mixed mechanically for equal length of time without any par-
while processing the sample. Therefore, the dispersion becomes ticle infusion to evaluate the effect of sonication on the epoxy.
poor and the flexure properties decreases. As described in the It is observed that sonicated epoxy has better (10%) in stiffness
literature [32,33], modulus is a bulk property of the materials and strength (5%) than the mechanically mixed epoxy system.
that depends on the geometry, particle size distribution and con- The reason for such change in stiffness of the matrix due to
centration of the filler. Higher particle loading increases Van Der ultrasound irradiation comes from the fact that the ultrasound
Waal force between the particles are high and that may lower the irradiation enhances the homogeneity of the reaction mixture of
dispersion. It will reduce the surface area and lower any inter- resin of the epoxy matrix. The epoxy resin used here is the con-
action between the matrix and filler. Effective load transfer will tent the mixture of diglycidyl ether of bisphenol A and aliphatic
be low, resulting the lower modulus value. The flexural strength diglycidyl ether. The ultrasound irradiation helps in molecu-
is known to be reduced with raising the filler content [34]. The lar mixing of these components together and the formation of
static adhesion strength as well as the interfacial stiffness usually reactive species which ultimately leads to increase the cross-
plays a significant role to transfer stress and elastic deformation linking in the polymer when mixed with amine. This effect is
from the matrix to the fillers [35]. This is especially true for more prominent when each of the highly reactive surfaces of
nanocomposite, because they impart a high portion of interface. the nanoparticle is effectively coated with the resin and reacted
If filler matrix interaction is poor, the particles are unable to with the amine. When the matrix is sonicated with TiO2 nanopar-
carry out any part of external load. In this case the strength can- ticles in it, it is observed that the enhancement in stiffness has
not be higher than the neat composite. If the bonding between the significantly increased to around 15% over the neat resin system.
matrix and filler is strong then the strength would be higher than The corresponding improvement however, over the sonicated-
that of neat composite. If we consider the failure of flexure load, epoxy system in stiffness is 10%, which we believe due to the
crack is generally initiate at the tension side along the width nanoparticle infusion along.
580 A. Chatterjee, M.S. Islam / Materials Science and Engineering A 487 (2008) 574–585

Fig. 8. Relation between the tensile strength and flexure strength of the neat and
TiO2 –epoxy nanocomposites.
Fig. 9. Behavior of storage modulus with temperature of the neat epoxy resin
and the TiO2 –epoxy nanocomposites.
The tensile strength is decreased than the neat resin sys-
tem. However, the modulus is increased compared to neat epoxy
system (Table 2). The decreased in the strength of the compos- gain in mechanical properties is due to the reduction of void
ites with addition of the particulate fillers is also observed by content [25] and that have discussed before. This reduction in
other scientists [7,8]. The strong filler/matrix adhesion would void content of the matrix provided strength for the nanophased
lead to enhanced strength of the nanocomposite. Accordingly composites. However, the increases in loading above 1%, the
to their consideration [7,8], it is known that the improvement particles started to lumps, greater than the size of voids. Those
of the tensile modulus of the composites exhibited in Fig. 5, lumps act as impurities and act as the source of failure. So,
should also be interpreted as the improvement of the interfacial the properties decreases for higher loading and have discussed
interaction. Especially epoxy and nano-TiO2 is incorporated, earlier.
the physiochemical entanglement between the matrix polymer
and the nano-particles guarantees effective interfacial bond- 3.3. Dynamic mechanical thermal properties
ing over the whole filler content range of interests. To obtain
the relation between the flexural and tensile strength we have Dynamical mechanical tests over a wide range of tempera-
plotted both the values (Fig. 8). A parabolic relation is found, ture were performed to see the physical and chemical structural
Y = −3.3683X2 + 122.18X − 1100.6, where Y = tensile strength changes of the polymers and the nanocomposites. The glass tran-
in KSi and X = flexure strength, KSi. The correlation factor, sitions or secondary transitions and yield information about the
R2 = 1. However, this is a preliminary correlation, developed morphology of polymers were determined. Experimental results
based on three data point. Additional study will be helpful to of dynamic tests conducted on nanocomposites and neat resin
verify the correlation and is ongoing. are presented in Figs. 9 and 10 of a temperature range from
From engineering point of view, elongation-to-break is an RT to 200 ◦ C. The complex moduli for all composites contain-
important parameter describing the rapture behavior of compos- ing fillers are pushed to a higher level relative to the neat resin
ite materials. The addition of mineral particulates into polymer system. On the other hand the modulus decreases with the tem-
used to lower it, even through the matrix has high impact tough- perature rising, but they retain a higher stiffness than the matrix
ness [9]. The same trend is follows here also Table 2. The when the temperature exceeds 200 ◦ C. The effect of TiO2 con-
neat resin has 4.27% compared to 1% nanocomposite 3.12%.
It implies that in principle matrix polymer to be more involved
more plastic deformation than the nanocomposites. The chain
flexibility of the crosslinked polymer on the nanoparticles is
restricted that causes the decreases in the elongation to break
of the nanocomposite. In contrast to tensile strength of the
composites that needs strong filler-matrix bonding, the micro-
deformation mechanism involved in elongation-to-break of the
composites depends upon extensionality of the nanoparticu-
lates agglomerates in the composites. The increment in the
loading causes more agglomeration confirmed by the TEM anal-
ysis. Balanced viscoelasticity of the grafting polymers on the
nanoparticles might thus be necessary for the improvement of
the elongation-to-break of the composites. As started previously
the volume fractions for the 5 nm of 1% by weight is 0.003%.
With such a low volume fraction it is very difficult to visual- Fig. 10. Behavior of loss modulus with temperature of the neat epoxy resin and
ize 10% increase in stiffness of the nanoinfused system. The the TiO2 –epoxy nanocomposites.
A. Chatterjee, M.S. Islam / Materials Science and Engineering A 487 (2008) 574–585 581

Fig. 11. Variation of storage modulus with % of loading of the TiO2 .

tent at percentage of loading on the storage modulus of epoxy act as additional virtual network nodes. Hence, the enhancement
resin is shown in Fig. 11. With increase in loading the modulus of this Tg is due to the incorporation of nanoparticles. In most
increases up to 1% loading. At 1% TiO2 dispersion in epoxy of the cases crosslink density is the key factor of controlling
shows the highest value of storage modulus among all the mate- the Tg for normal thermoset polymer systems. With increasing
rials. This result is similar to the flexural response, which is the crosslink density Tg increases. The enhanced glass transi-
mentioned earlier. Because of the high modulus of TiO2 and the tion temperature is related to the restriction effect of the TiO2
improved interfacial adhesion between nanoparticles and poly- on segmental motion. The TiO2 may chemically react with the
mer by physiochemical interaction, mechanical load imposed epoxy–amine system. Interfacial layers formed during the curing
on the nanocomposites transfers through the stronger interfacial process will exhibit a significantly different segmental dynam-
surface to the strongest nanoparticles effectively. The storage ics from the bulk resin. Therefore, TiO2 –epoxy-nanocomposite
modulus starts decreasing from 1.5% of loading. The decreas- has exhibited a higher Tg than the corresponding neat resin sys-
ing in modulus may be the poor dispersion of nano-particles tem. Smaller the particle size, the more surface reactivity that
with the matrix resin as have discussed previously. amplify the restriction of the resin networks. Therefore, the Tg
Fig. 12 is representing the Tg of neat and nanophased compos- has increased for the smaller particles. With increasing the load-
ites. During the DMA analysis with increasing the percentage ing the Tg increases and optimum conditions were achieved at
of nanoparticles loading, Tg of the materials increases linearly. 1% by weight loading. Further increase in the loading causing
At 1% loading of nanoparticles (5 nm) the epoxy resin give the poor dispersion that decreases the Tg (Fig. 12). An increased
highest Tg . After that it starts decreasing. So, with increasing the in Tg has reported for nanocomposite using the epoxy network
fillers contents the Tg of nanocomposites tends to move to higher systems [21,25,36,37], and tallies well with our findings.
values relative to the Tg of the neat system (118 ◦ C) Fig. 10. The The peak temperature in the loss modulus tends to shift to
nanoparticles obsequiously influence the Tg . The increases in Tg higher temperatures with increasing loading. On the other hand,
may be attributed to a loss in the mobility of chain segments of the restriction effect of segmental motion is also manifest by
the epoxy systems resulting from the nanoparticle/matrix inter- the change in peak width of the loss modulus. According to the
action. Impeded chain mobility is possible if the nanoparticles coupling model theory suggested by Roland and Ngai [38], the
are well dispersed in the matrix. The particle surface-to-surface segmental motion of polymers below Tg is a cooperative pro-
distance should then be relatively small and chain segment cess and has to overcome the resistance from the surrounding
movement may be restricted. Good adhesion of nanoparticles segments in order to accomplish the transformation between
with the surrounding polymer matrix would additional benefit configurations. As more segments are restricted by the pres-
the dynamic modulus by hindering molecular motion to some ence of TiO2 , the activation threshold for the motion of some
extend. The hard particles incorporated into the polymer would segments becomes higher. As a consequence, the loss modulus

Fig. 12. Tg values of neat resin and nanocomposites at different % of TiO2 loading.
582 A. Chatterjee, M.S. Islam / Materials Science and Engineering A 487 (2008) 574–585

peak during the glass transition trends to become wider with


increasing restriction effect. The restricted segmental motion
observed herein not only causes an increase in the Tg and the
flexural modulus but also improves the dimensional stability of
the polymeric materials. For the polymeric materials, dimen-
sional stability is closely related to the chemical structure of
the molecular chains and the topological structure formed. To
change in the stiffness of primary chains, one can adjust the
response of segment motions to the change in surroundings tem-
perature. The reduction of Tg above 1% loading may be because
of two reasons, the adsorption of matrix resin to the surface of the
Fig. 14. Variation of decomposition temperature with respect to weight loss for
nanoparticles that shifted the non-stoichiometric mixture. The
neat epoxy resin and the epoxy–TiO2 nanocomposites.
non-stoichiometry epoxy–amine system shifted the Tg to lower
side as has been reported by several authors [39,40] and thus a
Table 3
decrease in the Tg is observed. The second and most probable TGA data of epoxy nanocomposites
cause is the lack of uniform dispersion that has discussed previ-
ously. The agglomerated particles decrease the crosslinking and % of wt. remaining Temperature (◦ C)
ultimately the Tg . TiO2 (%)

0 0.50 1 2
3.4. Thermal analysis of nanocomposite 90 342 349 355 352
80 351 360 364 363
The thermo gravimetric analysis (TGA) was done to under- 70 360 366 372 370
stand the thermal stability of the nanophase composite system. 60 369 375 380 378
The real time characteristic curves were generated using the Char yield % at 650 ◦ C 1.17 1.47 4.4 2.43
Universal Analysis Data Acquisitation System. Data from a
typical TGA of the neat resin and the manocomposites are
shown in Fig. 13. TGA graphs show that increases in TiO2 load- The thermal stability drop down to 327 ◦ C (5% weight loss).
ing increases the thermal stability of the SC79 resin. The 5% Nano-particle loading was (0.5–2%) increased and measured
weight loss for the neat SC79 resin, 0.5, 1, and 2% TiO2 –SC79- the thermal stability. At 1% loading the highest thermal stability
nanocomposite are occurred at 295, 300, 317 and 319 ◦ C, is observed. Further increment in particle loading decreases the
respectively (Fig. 14). Researchers are commonly considered thermal stability (Table 3) (Fig. 13).
50% weight loss as an indicator for structural destabilization The increase in thermal stability may be the higher crosslink
[36]. So, in the present study 50% of the total weight loss is density of the epoxy-nanocomposite than the neat epoxy-resin
considered as the structural destabilization point of the system. system. The TGA measurement shows at 550 ◦ C the neat
It is clearly observed that the neat resin sample without TiO2 resin has 5% char content. However, for 1% TiO2 –SC79-
nanoparticles is stable up to 171 ◦ C (1% weight loss (Table 3)), nanocomposite system about 20% weight is present at 550 ◦ C,
where as the 1% loading the composite was stable up to 198 ◦ C. about ∼15% higher weight retention than the other systems

Fig. 13. TGA responses of the neat resin and the epoxy–TiO2 nanocomposites.
A. Chatterjee, M.S. Islam / Materials Science and Engineering A 487 (2008) 574–585 583

around that high temperature (Fig. 13). Generally, thermoset


polymer having higher crosslink density shows higher maximum
decomposition temperature [39]. This may be another cause to
increase the decomposition temperature of the nanocomposite.
The 2% nanocomposite shows the lower decomposition tem-
perature than the 1% loading (Fig. 14). The crosslink density
is maximized when the complete stoichiometry of the epoxy is
maintained. As soon as the stoichiometry of epoxy matrix was
broken with an addition of nano-loading [39], the crosslink den-
sity was reduced and lower crosslink density resulted in lower
decomposition temperature as observed in Figs. 13 and 14. So, Fig. 15. Kinetics of thermal decomposition of epoxy nanocomposites (0.5%
the decomposition temperature is lowered for the 2% loading. TiO2 by wt) by using Coats–Redfern equation.
When the TiO2 loading is very low (0.5 wt.%), the nanocom-
posite dispersed uniformly and develops a barrier to heat and
oxygen in the epoxy matrix due to ceramic nature of the particles. Coats–Dfern equation [41,42] is one of the well-known models:
The retardant effects of the composite to heat and oxygen in the    
g(α) AR/φE E
epoxy matrix are strengthened when the TiO2 loading increases. ln = − (3)
Since the number of the particles is increasing with loading. Fur- T 2 1 − 2RT/E RT
ther enhancement of loading (>1%) causes particle-to-particle
where (g(α) = 1 − (1 − α)1−n /1 − α).
interaction domination than the particle-to-matrix interaction,
If the order is equal to 1
discussed previously. The defect due to agglomeration increases
with loading and shows the less retardant effects to heat and g(α) = ln (1 − α)
oxygen, but till higher than that of the neat resin system. The
reason for lowering of the temperature above 1% by weight of    
ln (1 − α) AR/φE E
TiO2 –epoxy systems is due to (macroscopically) simple colliga- ln = − (4)
T 2 1 − 2RT/E RT
tive thermodynamics effect of an impurity on a bulk solution.
Microscopically, it may be seen as the result of the perturba- For the decomposition of thermosets, the mechanism of random
tion that the TiO2 introduces to the three-dimensional structure scission of molecules dominates. Therefore, the decomposi-
of the polymer. This perturbation weakens the Van Der Waals tion could be treated as one order reaction. The curves of
interaction between the polymer chains. This affects the stability (g(α) = ln (1 − α) versus 1000/T were computed with conver-
for the polymer, and that is reflected in lowering of the thermal sion. The activation energy was calculated from the slope of
stability. This perturbation begins at a point when the number the line. Fig. 15 shows the Coats–Dfern curve treated with one
of particle reaches a certain level and particle-to-particle inter- order reaction. The apparent activation energies are tabulated in
action initiates leading to agglomeration of particles into lumps, Table 4.
which acts as an impurity in the system [25]. The TEM picture Energy of activation for decomposition of epoxy-
confirms the results that the agglomeration of the nanoparticle nanocomposite is calculated from TGA curves by the
with increasing the particle loading in our mixing system. integral method of Horowitz and Metzger [43] according to the
To evaluate the thermal stability more detailed, nonisothermal following equation:
thermal degradation kinetics were calculated. For an n-order     
1 Ea θ
reaction, the reaction rate can expressed as ln ln = (5)
1−α RTmax
2

dα where α is the decomposed fraction, Ea the activation energy for


= k(1 − α)n (1)
dt decomposition, Tmax the temperature at maximum rate of weight
loss, θ = T − Tmax and R the universal gas constant. From the
where α is the apparent conversion, k the rate constant and n is plot of ln [ln (1/(1 − α))]versus θ (Fig. 16), the activation energy
the reaction order. In dynamic mode, the rate expression can be for decomposition was calculated from the slope of the straight
written as line in Eq. (5). The calculated activation energy as a function
of TiO2 loading is presented in Fig. 17. The activation energy
 
dα A −E
= exp (1 − α)n (2) Table 4
dT φ RT
Kinetic data of epoxy nanocomposites by using Coats–Redfern equation
% TiO2 loading by wt. Ea (kJ/mol)
where α is the conversion at temperature T, dependent on the
heating rate φ, E the apparent activation energy and A is the 0 (neat resin) 186
Arrhenius frequency factor. Different kinetics expressions could 0.5 197
1 205
be obtained by integrating the above equation with different 2 191
approximation treatments. Among those kinetics equations, the
584 A. Chatterjee, M.S. Islam / Materials Science and Engineering A 487 (2008) 574–585

Fig. 18 shows the schematic diagram of Doyle’s proposition [44]


for determining the IPDT of a major factor on thermal stabilities
of the samples and the IPDT is calculated as follows:

(A1 + A2 )3
IPDT = (Tf − Ti ) + Ti (6)
(A1 + A2 + A3 )2

where Ti and Tf are the initial and final experimental tempera-


tures, respectively. The relationship between IPDT and the TiO2
Fig. 16. Kinetics of thermal decomposition of epoxy nanocomposites (0.5% loading is shown in Fig. 17. The IPDT clearly increases with
TiO2 ) by using Horowitz–Metzger method. nano-loading.
The changes in activation energy of the nanocomposites may
be the structural evolution. In minimum loading, the particle-
matrix distribution is homogeneous in nature. The ceramic
particles block the heat as it is insulator as described previously.
Heat and oxygen movement paths are lingering due to the pres-
ence of the ceramic particles. So, the retardant effects to heat
and oxygen are observed and that enlarged further with load-
ing, since the number of nano-particles increases (up to 1%).
Additional increment of loading causes detrimental effect due
to the agglomeration of particles that has occurred, confirmed by
TEM. Agglomerated particles are non-uniformly distributed in
the matrix and are less effective to blocking the heat and oxygen,
is still higher than the neat epoxy system.
The IPDT trend is same as for the evolution of the activa-
tion energy. This may be the same physic natures of activation
energy and IPDT. The activation energy describes the fast
decomposition process of the resin network, ignoring the ini-
tial decomposition and char yield formation (Table 3). IPDT,
however, reflects the whole stability of the sample, including
Fig. 17. Effect of TiO2 loading on activation energy (calculated by using the initial, fast decomposition and final (char forming) steps.
Horowitz–Metzger method, Coats–Redfern method) and integral procedural As the TiO2 in the epoxy matrix mainly contributes the retar-
decomposition temperature (IPDT). dant effects to heat and oxygen and char formation, the IPDT
increases with TiO2 loading and maximum at 1% loading.
is a function of nano-loading (Table 4). The activation energy The activation energy, also reaches a maximum at 1% loading
reaches a maximum at 1% TiO2 loading. because the dispersion state of the TiO2 is the main concern
In addition, thermal stability has also been characterized by in determining the fast decomposition process of the epoxy
given integral procedural decomposition temperature (IPDT). resin.

Fig. 18. The schematic diagram of the Doyle’s method for determining the IPDT.
A. Chatterjee, M.S. Islam / Materials Science and Engineering A 487 (2008) 574–585 585

4. Conclusions [12] M.Z. Rong, M.Q. Zhang, Y.X. Zheng, H.M. Zeng, R. Walter, K. Friedrich,
J. Mater. Sci. Lett. 19 (2000) 1159.
[13] C. Becker, H. Krug, H. Schmidt, Mater. Res. Soc. Symp. Proc. 435 (1996)
An ultrasonic cavitations process has been employed to
237.
manufacture the epoxy–TiO2 -nanocomposite. The particles are [14] G. Carotenuto, L. Nicolais, X. Kuang, Z. Zhu, Appl. Comp. Mater. 2 (1995)
dispersed the entire volume of the resin. Resin properties have 385.
been influenced with percent of loading of the nanoparticles. [15] P.B. Messersmith, E.P. Giannelis, Chem. Mater. 6 (1994) 1719.
The tensile modulus, flexural modulus, storage modulus, Tg and [16] P. Kelly, A. Akelah, S. Qutubuddin, A.J. Moet, J. Mater. Sci. 29 (1994)
2274.
thermal stability were increased with 1% TiO2 of 5 nm size.
[17] V. Nigam, D.K. Setua, G.N. Mathur, K.K. Kar, J. Appl. Polym. Sci. 93
Nanoparticle infusion changes the morphology of the resin sys- (2004) 2201.
tems that increases the Tg of the bulk matrix, the mechanical and [18] G.I. Eskin, Ultrason. Sonochem. 1 (2001) 319.
thermal properties of the nanocomposites. Nano-particle incor- [19] G.J. Price, Ultrason. Sonochem. 10 (2003) 277.
poration enhances the activation energy of the nanocomposites [20] L. Nielsen, R. Landel, Mechanical Properties of Polymers and Composites,
Marcel Decker, New York (USA), 1994.
and that level off at 1% loading than decreases with increas-
[21] L. Nicolais, M. Narkis, Polym. Eng. Sci. 11 (1971) 194.
ing loading. The IPDT for the epoxy TiO2 -nanocomposite is [22] M. Zhang, H. Zeng, L. Zhang, L. Lin, G. Lin, R.K.Y. Li, Polym. Compos.
increased with nano-loading and the trend is dissimilar from the 1 (1993) 357.
literature [41]. We have observed the same trend for the IPDT [23] C.L. Wu, M.Q. Zhang, M.Z. Rong, K. Friedrick, Comp. Sci. Technol. 62
and activation energy for the epoxy TiO2 -nanocomposite. The (2002) 1327.
[24] A.C. Roulin-Moloney, Fractography and Failure Mechanisms of Polymer
study reveals that the nanofiller infusion improves the thermal,
and Composites, Elsevier Applied Science, London, 1986.
mechanical and viscoelastic properties of the epoxy resin. [25] N. Chisholm, H. Mahfuz, V.K. Rangari, A. Ashfaq, S. Jeelani, Comp. Struct.
(2004) (web published).
Acknowledgements [26] L.A. Carlsson, J.W. Gillespie, in: J.M. Whitney, R.L. Mc Cullough (Eds.),
Micromechanical Materials Modelling, vol. 2, Technomic Publishing Co.
Inc., 1990, p. 49.
The authors would like to express their gratitude and sincere
[27] N. Chisholm, H. Mahfuz, S. Jeelani, Proceedings of the ASME Interna-
appreciation to Jack W. Gillespie, Jr. Director CCM, University tional Congress & Exposition, Louisiana, New Orleans, 2002, p. 17.
of Delaware, for his endless help and encouragement. Finan- [28] E. Reynaud, C. Gauthier, J. Perez, Rev. Metall. 96 (2) (1999) 169.
cial support from Center for Composite Materials, University of [29] L.W. Chun, Q.Z. Ming, Z.R. Min, F. Klaus, Compo. Sci. Technol. 62 (2002)
Delaware is highly appreciated. 1327.
[30] S. Wu, Polymer 60 (1988) 248.
[31] C.Q. Yang, J.F. Mouldev, W. Iateley, J. Adhes. Technol. 2 (1988) 11.
References [32] S. Wu, Macromolecule 18 (10) (1985) 2023.
[33] S. Wu, Polym. Eng. Sci. 30 (1990) 753.
[1] D. Ratna, A.K. Banthia, P.C. Deb, J. Appl. Polym. Sci. 80 (2001) 1792. [34] I. Nicolais, M. Narkis, Polym. Eng. Sci. 11 (1971) 194.
[2] B. Wetzel, F. Haupert, F. Zhang, Comp. Sci. Technol. 63 (2003) 2055. [35] M. Zhang, H.L. Zeng, G. Lin, R.K.Y. Li, Polym. Compos. 1 (1993) 357.
[3] M. Antonietti, C. Giltner, Angew. Chem. Int. 36 (1997) 910. [36] V.K. Rangari, Y.U. Koltypin, A. Gedanken, J. Appl. Polym. Sci. 86 (2002)
[4] G. Schmid, Chem. Rev. 92 (1992) 1709. 160.
[5] R. Mulhaupt, T. Engelhardt, N. Schall, Plast. Eur. 91 (2001) 63. [37] C.L. Wu, M.Q. Zhang, M.Z. Rong, K. Friedrick, Comp. Sci. Technol. 65
[6] S. Komarneni, J. Mater. Chem. 2 (1992) 1219. (2005) 635.
[7] R.P. Singh, M. Zhang, D. Chan, J. Mater. Sci. 37 (2002) 371. [38] C.M. Roland, K.L. Ngai, Macromolecules 24 (1991) 5315.
[8] B. Wetzel, F. Haupert, K. Friedrich, M.Q. Zhang, M.Z. Rong, Proceedings [39] H. Miyagawa, L.T. Drzal, Polymer 45 (2004) 5163.
of the 13th ECCM Brugge, Belgium, 2001. [40] J. Palmese, University of Delaware, Dept. of Chemical Engineerings, Ph.D.
[9] M. Hussain, A. Nakahira, K. Niihara, Mater. Lett. 26 (1996) 185. Thesis, 1990.
[10] C.B. Ng, L.S. Schadler, R.W. Siegel, Nanostruct. Mater. 12 (1999) [41] B. Guo, D. Jia, C. Cai, Eur. Polym. J. 40 (2004) 1743.
507. [42] A.W. Coats, J.P. Redfern, Nature 201 (1964) 68.
[11] M.Z. Rong, M.Q. Zhang, H. Liu, H.M. Zeng, B. Wetzel, K. Friedrich, Ind. [43] H.H. Horowitz, G. Metzger, Anal. Chem. 35 (1963) 1464.
Lubr. Tribol. 53 (2001) 72. [44] C.D. Dolye, Anal. Chem. 33 (1961) 77.

You might also like