You are on page 1of 18

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/249353913

S2Glass/Epoxy Polymer Nanocomposites: Manufacturing,


Structures, Thermal and Mechanical Properties

Article  in  Journal of Composite Materials · October 2003


DOI: 10.1177/002199803035186

CITATIONS READS

234 1,105

4 authors:

Anwarul Haque Mohammad Shamsuzzoha


University of Alabama University of Alabama
34 PUBLICATIONS   480 CITATIONS    166 PUBLICATIONS   1,057 CITATIONS   

SEE PROFILE SEE PROFILE

F. Hussain Derrick Dean


Mitsubishi Chemicals Alabama State University
13 PUBLICATIONS   2,031 CITATIONS    104 PUBLICATIONS   3,824 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Al alloy development View project

All content following this page was uploaded by Mohammad Shamsuzzoha on 11 August 2015.

The user has requested enhancement of the downloaded file.


S2-Glass/Epoxy Polymer Nanocomposites:
Manufacturing, Structures, Thermal
and Mechanical Properties

A. HAQUE1,* AND M. SHAMSUZZOHA2


1
Department of Aerospace Engineering & Mechanics
2
Department of Metallurgical and Materials Engineering
The University of Alabama, Tuscaloosa, AL – 35487-0280, USA

F. HUSSAIN AND D. DEAN


Tuskegee University Center for Advanced Materials
Tuskegee, Al – 36088, USA

(Received September 16, 2002)


(Revised February 27, 2003)

ABSTRACT: This paper is primarily focused in studying the effects of nanoclay


particles such as montmorillonite on improving mechanical and thermal properties
of fiber reinforced polymer matrix composite materials. Basic correlations between
polymer morphology, strength, modulus, toughness, and thermal stability of
thermoset nanocomposites were investigated as a function of layered silicate
content. S2-glass/epoxy–clay nanocomposites were manufactured through an
affordable vacuum assisted resin infusion method (VARIM). The nanocomposites
are formed during polymerization when the adsorbing monomer separates the clay
particles into nanometer scales. Transmission electron microscopy (TEM) and wide
angle X-ray diffraction(WAXD) were used to characterize the morphology of the
dispersed clay particles. The thermal properties such as onset of decomposition and
glass transition temperatures were determined by Thermo Gravimetric Analysis
(TGA) and Dynamic Modulus Analyzer (DMA). Mechanical properties such as
interlaminar shear strength, flexural properties and fracture toughness are also
determined for both conventional S2-glass/epoxy composites and S2-glass fiber
reinforced nanocomposites.
The results show significant improvements in mechanical and thermal properties
of conventional fiber reinforced composites with low loading of organo silicate
nanoparticles. By dispersing 1% by weight nanosilicates, S2-glass/epoxy–clay
nanocomposites attributed to almost 44, 24 and 23% improvement in interlaminar
shear strength, flexural strength and fracture toughness in comparison to conventio-
nal S2-glass/epoxy composites. Similarly, the nanocomposites exhibit approximately
26 C higher decomposition temperatures than that of conventional composites. This
improved properties of fiber reinforced polymer nanocomposites are achieved mostly
due to increased interfacial surface areas, improved bond characteristics and
intercalated/exfoliated morphology of the epoxy–clay nanocomposites. The TEM

*Author to whom correspondence should be addressed. E-mail: ahaque@coe.eng.ua.edu

Journal of COMPOSITE MATERIALS, Vol. 37, No. 20/2003 1821


0021-9983/03/20 1821–18 $10.00/0 DOI: 10.1177/002199803035186
ß 2003 Sage Publications
1822 A. HAQUE ET AL.

observations provide evidence of detailed morphology of the polymer layered-clay


nanocomposites.

KEY WORDS: nanocomposites, TEM, XRD, thermal analysis, mechanical


properties, fracture.

INTRODUCTION

OLYMER NANOCOMPOSITE IS a new class of material in which resin materials is usually


P tailored by inorganic nanoparticles to exhibit improved mechanical, thermal, and fire
resistance properties [1–4]. Such enhancement in the properties of nanocomposites occurs
mostly due to their unique phase morphology and improved interfacial properties [5].
The types of nanomaterials that are most widely used for this purpose include: silica
particles, layered silica clays, electro-spun polymeric nanofibers and nanotubes. Due to
dispersion of nanoscale inorganic particles into organic resin materials, three distinct
morphologies such as phase separated, intercalated and exfoliated regions are typically
observed in nanocomposites structures. Among those, intercalated and exfoliated
structures are mostly desirable for improving the performance of the material. In phase
separated structure, polymer is unable to intercalate between the layered silicate sheet and
the properties remain almost same as typical microcomposites. In such case, the individual
sheets in the silicate are seen in agglomerated form due to strong surface attraction force
between clay layers and unable to separate between each other. In an intercalated system,
single extended polymer chain is intercalated between the silicate layers resulting a well-
ordered multilayer morphology built-up with alternating polymeric and inorganic layers.
The third category is exfoliated structure in which silicate layers are completely and
uniformly dispersed in inorganic polymer matrix. A schematic of the above morphologies
are shown in Figure 1.
The development of various morphologies of nanocomposites basically depends upon
the surface treatment of the nanoparticles and specific dispersion or processing techniques.

Phase separated Intercalated Exfoliated


(microcomposite) (nanocomposite) (nanocomposite)

Figure 1. Schematic of the typical structure of polymer nanocomposite.


S2-Glass/Epoxy Polymer Nanocomposites 1823

Various processing and surface treatment routes are implemented in order to reduce
particle–particle attraction, promoting an expansion of the distance between each layer.
The introduction of organic onium ions with long alkyl chains on the gallery surface of the
layered silicate reduces the interaction force between the individual sheets and separates
them further away. The particles can also be separated further either by absorbing
monomer into the gallery prior to polymerization or in the case of high polymer by
employing shearing force using an extrusion compounder. The individual silicate sheets
are hydrophilic and thus are not compatible with the hydrophobic organic matrix
polymer. The application of surfactants such as alkyl ammonium salt through the ion
exchange process makes the inorganic hydrophilic nanoparticles compatible with the
hydrophobic organic resin system.
In recent years, the use of organoclays as precursors to nanocomposite formation has
been extended into various polymer systems including epoxy, polyurethens, polyimides,
nitrile, polyesters etc [6–8]. In the early 1990, the researchers at Toyota motor company
first demonstrated that the properties of nylon-6, containing a homogenous dispersion of
2–7 percent by weight of nanometer sized clay sheets, showed excellent mechanical,
thermal and barrier properties compared with nonreinforced nylon [9]. Such improvement
in overall properties of nanocomposites is achieved due to high aspect ratio and high
strength of these nanoscaled inorganic particles. Following Toyota’s lead, a number of
other researchers also began investigation with various types of polymer nanocomposites
[10–13]. It is well known that epoxy nanocomposites have significant importance due to
their wide range of applications. Up to now, most of the work that has been carried out
related to polymer nanocomposites basically consists of layered silicate in polymer matrix.
There is very little information available about continuous fiber reinforced polymer
nanocomposites. Moreover, the manufacturing of these fiber reinforced polymer
nanocomposites in a cost effective manner is another challenge. In this paper, emphasis
is given in manufacturing S2-glass fiber reinforced epoxy–clay nanocomposites through a
cost effective resin infusion process. The structures, fracture behavior, thermal and
mechanical properties of these nanocomposites are investigated. These polymer matrix
nanocomposites are suitable for lightweight and high temperature structural applications
specifically under high impact loading conditions.

EXPERIMENTAL WORK

Materials and Processing

Two types of nanocomposites such as epoxy–clay and S2-glass/epoxy–clay were


prepared for this investigation. In the first category, a desired amount (500–700 gm) of SC-
15 epoxy (Part-A) resin was mixed with various amounts (1, 2, 5 and 10% by weight) of
nanoclay particles in an Erlenmeyer flask. SC-15 epoxy formulation is based on two parts,
Part A (bisphenol-A and bisphenol-F blend) and Part B (hardener cyclo aliphatic amine)
manufactured by Applied Poleramic INC, USA. The suggested mixing ratio is 100 A : 30 B
by weight. It is a very low viscosity (300 cps) toughened epoxy resin specifically developed
for Vacuum Assisted Resin Infusion Molding(VARTM). This resin is suitable for
structural and ballistic applications. The nanoclay particles used in this study were
organically modified montmorillonite minerals of type 1.28E supplied by Nanocor. It is a
layered structure consisting primarily a single layer of aluminum octahedron sandwiched
1824 A. HAQUE ET AL.

between two layers of silicon tetrahedron. Each layered sheet is slightly less than one
nanometer thin with surface dimensions extending to about a micron (1000 nm). The
aspect ratio is approximately 1000 and the surface area of the clay is in the range of 750
square meters per gram. The montmorillonite clay particles have been surface treated to
convert the surface from hydrophilic to organophilic. This was accomplished through an
onium-ion substitution reaction with the surface sodium ions. A quaternary onium ion
was used as the compatibilizing agent. The intercalant (surface treatment) ionically bonds
to the clay surface which is compatible with host epoxy resin. The mixture of SC-15 resin
(Part-A) and nanoclay was stirred at 25 C for an hour with a mechanical stirrer to perform
clay dispersion into polymer matrix. An ultrasonic liquid processor (Model VC 750 by
Sonic) with a probe tip diameter 13 mm was used for 3–4 min to complete the final
dispersion. This probe with tip diameter 13 mm resulted high intensity of cavitations and
the amplitude was within 80–100 mm. Conversely, probe with larger tip diameter
eventually produces lower intensity but the energy is released over a greater area. Once
the nanoclay is properly dispersed in epoxy resin (A), cyclo aliphatic amine hardener (B)
was mixed in 100 A : 30 B ratio by weight. A portion of this mixture was poured into a
mold for curing at 60 C for 2 h. The post curing process was carried out at 100 C for a
period of 5 h. This resulted epoxy–clay nancomposites through in situ polymerization
process.
In the second category, S2-glass/epoxy–clay polymer nanocomposite was manufactured
using plain weave S2-glass fabric and epoxy–clay resin system. A cost effective vacuum
assisted resin infusion molding (VARIM) process was used to manufacture S2-glass/
epoxy–clay nanocomposites. A schematic of the VARIM process is shown in Figure 2.
The dry S2-glass woven preforms were cut into plies and stacked in several layers on an
open mold with a peeling cloth in between the fabric and the mold. A separate peeling

Vacuum
pump
Resin trap

Resin tank
Ejection line

Resin feed line

Distribution media

Teflon peel-off ply


Preform
Vacuum bag
Vacuum seal
Moulding tool

Figure 2. Schematic of VARIM process.


S2-Glass/Epoxy Polymer Nanocomposites 1825

cloth and distribution mesh were placed over the preform followed by a vacuum plastic
bag. One resin-ejection line and one resin feed line were attached to the mold. The resin
feed line was connected to a resin tank and the ejection line was connected to a vacuum
pump through a resin trap. Prior to resin injection, vacuum was applied to the mold
for two hours to debulk the perform. The mixture of nanoclay dispersed resin and
hardener was poured into a resin tank immediately before resin injection. The process of
dispersing nanoclay into epoxy resin and the mixing ratios of nanoclay, epoxy and
hardener were similar as described earlier. The amount of clay particles mixed with
the epoxy resin was restricted up to 10% by weight in order to maintain the resin viscosity
sufficiently low (500 cps) specifically for VARIM process. After debulking S2-glass
perform, the resin injection port was opened and immediately uncured epoxy–clay resin
started to infuse the perform under vacuum pressure. Once the preform was completely
wetted by the epoxy–clay resin, the injection line was closed by a stopper and the preform
was left for curing under vacuum at room temperature for 24 h. The post curing was done
in an oven at 100 C for 2 h. The samples of both epoxy–clay and S2-glass/epoxy–clay
nanocomposites were prepared for identifying structure, thermal and mechanical
properties. In a similar manner conventional S2-glass/epoxy composite laminate was
also manufactured for comparison with S2-glass/epoxy–clay composites.

Methodology

The thermal stability of the nanocomposites was studied using thermo gravimetric
analyzer (TGA) in which the sample mass loss due to volatilization of degraded by-
products was monitored as a function of temperature. This operation was carried out on
10–15 mg sample at a temperature range 40–1000 C (with 10 C/min. heating rate) using a
TA Instrument 2950 TGA system. The glass transition temperature was measured
following dual cantilever beam method (ASTM standard D 4092) using a TA Instruments
2980 Dynamic Mechanical Analyzer (DMA). The bar specimen with dimensions
44.4 mm  6.4 mm  1.6 mm was used as a beam and gripped between two clamps. An
alternating strain was applied at the center of the dual-cantilever beam with a frequency
of 1 Hz. The amplitude and the heating rate during the alternating strain were 10 mm and
4 C/min, respectively. The resulting stress during the alternating strain condition of the
beam was monitored which consists of both in-phase and out-of-phase components. These
stress components were then used to calculate storage modulus, G1 and loss modulus, G2
of the polymer. Finally, the tan , which is a ratio of G2/G1 was calculated as a function of
temperatures. This ‘tan ’ primarily represents the ratio of the dissipation energy to the
energy stored in the specimen during the alternating cycle. At Tg, a significant drop in the
storage modulus, G1 occurs which eventually provide maximum values for ‘tan ’.
X-ray scattering of epoxy–clay nanocomposites was done using a Rigaku RU200
generator with Cuk radiation (1.54 A) at an accelerating voltage of 40 kV/170 ma. Data
was collected on phosphor image plates and digitized through a scanner. The cylindrical
specimens of 3 mm diameter were cut by an ultrasonic disc cutter for transmission electron
microscopic (TEM) examination. The grinding of the specimen was done with a 600-grade
silicon carbide paper to about 200 mm and then dimpled by mechanical polishing. The
discs were then thinned by ion milling to a thickness of less than 100 nm. The thin foils
thus prepared were examined with a 200 keV TEM (Hitachi H 8000) to determine
structures of the nanocomposites.
1826 A. HAQUE ET AL.

The flexural and interlaminar shear properties of fiber reinforced composites were
determined using three point bend (ASTM D 790) and short beam shear (ASTM D 2344)
test methods. The tests were carried out in a servohydraulic testing machine at a stroke
rate of 0.12 mm/min. The span to depth ratio for flexural and short beam shear test
specimen was maintained at 32 and 5, respectively. The fracture toughness was determined
from the notched bend tests (NBT) using the following relationship [14].

pffiffiffi
K1 ¼ Y c,

where, K1 is the stress intensity factor;  is the applied stress, c is the initial crack length,
w is the specimen depth; Y is the calibration factor ¼ 1.93  3.07(c/w) þ 14.53(c/w)2
25.11(c/w)3 þ 25.80(c/w)4.
The fracture toughness test of fiber reinforced composites was done using a notched
bend specimen under three-point loading in a sevohydraulic MTS machine. In case of neat
epoxy resin and epoxy–clay nanocomposites, the flexural properties and fracture
toughness were determined using a minimat test system with a much smaller loading
capacity (1 kN) for better accuracy in the test results. The fracture surface of the
composites were also evaluated by both optical and JOEL Scanning Electron Microscopy.

RESULTS AND DISCUSSIONS

Structures

Transmission electron microscopy (TEM) was applied to characterize the hierarchial


morphology present in the nanocomposites. Figure 3 shows the bright field TEM
micrographs typically representing the microstructures in epoxy–clay nanocomposites.

Figure 3. Bright field TEM micrographs of epoxy–clay nanocomposites showing both intercalated and phase
separated morphology.
S2-Glass/Epoxy Polymer Nanocomposites 1827

The interesting microstructural feature to be noted in these micrographs is the pattern that
is formed by the constituent polymer and silicate phases. The silicate phase, which appears
the morphology of more contrasted layer lines, is stacked alternately with less contrasted
polymer chains to form a lamellar type pattern (intercalated) representing the basic
microstructure of the epoxy–clay nanocomposites. However, the micrograph also contains
regions at which the microstructure is not strictly of lamellar morphology, but rather is
made of a group of differently oriented parallel silicate and polymer layers that are
bundled together (phase separated). These characteristics of the microstructure probably
owe to their origin in the manufacturing process. It is possible that the chemical mixture
employed to the clay particles for the development of the nanocomposites lacked an equal
distribution of alkyllamonium ions. As a result a normal lamellar type microstructure
(intercalated) is formed during polymerization for only that part of the solution which
contains an equal distribution of alkyllamonium ions. While the remaining part of the
solution that lacked alkyllamonium ions resulted the microstructure of the other
morphology (phase separated) during a similar polymerization process. Moreover, TEM
micrographs reveal that at higher loading rate (5 and 10%) the dispersion of nanoparticles
results more agglomeration (phase separated) and less lamellar (intercalated) structures. It
is to be noted that development of various morphologies of nanocomposites basically
depends upon the surface treatment of nanoparticles and specific dispersion process
parameters. In our study the dispersion process parameters (mechanical stirrer time and
speed, ultrasonic dispersion time and intensity) were maintained same for all the loading
rates (1–10%). Possibly the dispersion process parameters are required to be adjusted
depending upon the loading rates (1–10%) in order to achieve perfect intercalated or
exfoliated structures. The perfect intercalated and exfoliated structures provide more
interfacial surface areas than phase separated (agglomerated) structures. The overall
structural integrity of epoxy-nanocomposites enhances due to these increased interfacial
surface areas. As a result improved thermal (Tdec) and mechanical properties are mostly
observed in nanocomposites with low clay loading (1%). The micrographs also reveal the
presence of well faceted nano-sized particles embedded in the background of lamellar
matrix. Selected area diffraction patterns taken from such particle is shown in Figure 4.
Analysis of these diffraction patterns established that these particles are to be crystal.

Figure 4. Diffraction pattern of clay particle.


1828 A. HAQUE ET AL.

Figure 5. Bright field TEM micrographs of S2-glass/epoxy–clay nanocomposites.

50000

E poxy-2% clay
35.08A Epoxy-5%clay
40000 Epoxy-10%clay
Intensity (Counts/sec)

30000

20000 34.90A

35.20A

10000

0
0 2 4 6 8 10 12
Two-theta (degrees)

Figure 6. Wide angle X-ray diffraction pattern of epoxy–clay nanocomposites.

Elaborate TEM observations on a number of specimens taken from these composites also
revealed that these nano sized clay particles (which were employed in the development of
the nanocomposites with an aim to improve mechanical properties) appear to isolate
individually and distribute randomly in the matrix of the composites. Figure 5 shows the
TEM micrographs of S2-glass/epoxy–clay nanocomposites. The presence of isolated
nanoclay particles are clearly evident at the interface region.
The interaction between clay particles and the polymer matrix results interculation of
the polymer between the layers of the clay, which increases the clay interlayer distance.
The extent of the dispersal can be determined by Wide Angle X-Ray Scattering (WAXS)
method. Bragg’s equation is typically used to measure the interlayer distance using the
diffraction peak and its position in the WAXS patterns. It is typically shown as n ¼ 2d
sin , where is the diffraction angle, d is the interlayer and  is the wavelength (1.54 A).
In Figure 6 the WAXS data is displayed as the intensity of the reflections (in counts or
S2-Glass/Epoxy Polymer Nanocomposites 1829

arbitrary units) versus the scattering angle, 2 . The clay should show a peak near 6 , due to
a basal spacing of 15 Å. As the nanoclay particles are dispersed into epoxy resin and cured,
the clay interlayer spacing increases to 35 Å, indicating an intercalated morphology of the
system. These results are seen to be consistant that has been observed in TEM
micrographs. Moreover, the WAXS data also shows almost similar interlayer spacing for
all the three epoxy–clay nanocomposites system (2, 5 and 10% nanoclay particles)
considered in this study.

Thermal Analysis

Figure 7 shows TGA scan in the form of weight loss versus temperature for S2-glass/
epoxy and S2-glass/epoxy–clay nanocomposite containing various amounts of nanoclay
particles. This provides information about the thermal stability of the nanocomposites in
comparison to conventional composites. Table 1 shows the quantitative values of onset of
decomposition temperatures for various composite system. A gradual weight loss is
observed at 175 C for both conventional composites and nanocomposites. This initial
weight loss is possibly due to volatilization of unreacted small molecules such as
accelerator/hardener etc. which continues for some period at a very slow rate. As the
actual decomposition of the cured composite materials begins at elevated temperatures,
the weight loss occurs at a faster rate. The intersection point between the initial slope and

Figure 7. TGA plot of weight loss vs. temperature for S2-glass/epoxy–clay nanocomposites.

Table 1. Glass transition temperature(Tg) and onset of decomposition


temperature(Tdec) for S2-glass/epoxy–clay nanocomposites.
Sample Montmorrilonite wt.% Tg ( C) Tdec ( C)
S2 glass-epoxy Neat cured 100 241.37
1.0 103 267.07
2.0 104 260.67
5.0 — 251.33
1830 A. HAQUE ET AL.

the slope followed after the actual decomposition basically provide primary onset of
decomposition temperature, Tdec. This Tdec was directly computed by the TGA
instrument. The sample containing 1% clay provided a higher primary onset of
decomposition (267 C) temperature than any other system considered in this study. The
onset of decomposition temperatures of the specimens containing 2 and 5% clay particles
are seen to be 260 and 251 C, respectively. This study indicates that although the initial
weight loss for both conventional S2-glass/epoxy and S2-glass/epoxy–clay nanocompo-
sites commences at the same temperature, the decomposition rate for the conventional
composites from between Tdec, 241 C and 330 C is comparatively higher than that of
nanocomposites. The nature of the plots also show that the decomposition rate beyond
Tdec for all the nanocomposites are almost constant up to 330 C. Beyond that temperature
the decomposition rates for nanocomposites are seen to be increased. The conventional
composites show almost 10% weight loss at 330 C and the decomposition rate is seen to
be almost constant between Tdec (241 C) and 550 C. At temperatures beyond 600 C, all
the system maintained almost constant weights, which primarily represent weight of the
totally decomposed composite system. It is clearly evident from Figure 7 that the mass loss
of totally decomposed nanocomposites with 1% clay particle is comparatively 5% less
than the conventional composites. All the above discussion eventually shows improved
thermal stability of S2-glass/epoxy–clay nanocomposites.
The tan  versus the temperature plots calculated from the DMA study of the
conventional S2-glass/epoxy and S2-glass/epoxy–clay nanocomposite containing 1 and
2% silicates are shown in Figure 8. The shift of the tan  peak to higher temperature
indicates an increase in glass transition temperature, Tg for the nanocomposite. The Tg, for
both cured S2-glass/epoxy–clay nanocomposites and uncured epoxy–clay are provided in
Tables 1 and 2. The epoxy–clay system without post curing exhibited almost 20  C less Tg
than that of post cured S2-glass/epoxy composites. Although the Tg of the post cured
epoxy–clay system is expected to be higher. The effects of nanoclay particles in enhancing
Tg value are observed to be similar in both epoxy–clay and S2-glass/epoxy–clay system.
The tan  peak for S2-glass/epoxy conventional composites is observed to be 100 C, where

Figure 8. DMA study showing tan  vs. temperature plot for S2-glass/epoxy–clay nanocomposites.
S2-Glass/Epoxy Polymer Nanocomposites 1831

Table 2. Glass transition temperatures of uncured epoxy–clay


nanocomposites.
Sample Glass Transition Temperature ( C)
Epoxy 82.56
Epoxy 1% clay 83.79
Epoxy 2% clay 86.23
Epoxy 5% clay 81.47
Epoxy 10% clay 80.37

as the same for the S2-glass/epoxy–clay nanocomposites containing 2% clay is seen to be


104 C. Such increases in Tg in the order of few degrees (4 C) are also reported for other
epoxy nanocomposite system [14] and are usually attributed to restricted relaxation
motion near the interface of the inorganic–organic nanocomposites. The chemical bonding
at the interface of the nanoclay layer and the epoxy resin basically reduces the relaxation
mobility of the polymer at the interface region. All the composites show higher storage
modulus, E00 and slightly reduced loss modulus, E0 exhibiting an increased tan  below the
glass transition temperature, Tg. At Tg, tan  reaches to a maximum value due to
significant drop in loss modulus, E0 . Beyond the glass transition temperature, Tg, the
storage modulus, E00 decreases significantly and thereby, a decrease in tan  is observed.

Mechanical Properties and Fracture Behavior

Figure 9 shows interlaminar shear stress versus displacement plots for both conventional
S2-glass/epoxy and S2-glass/epoxy–clay nanocomposites. An enhanced interlaminar shear
failure stress is clearly observed for the case of nanocomposites with 1 and 2% nanosilicate
particles. Table 3 shows the average interlaminar shear strength data determined from five
specimens. S2-glass/epoxy–clay nanocomposites containing 1 and 2% nanosilicate
particles show almost 44 and 20% higher interlaminar shear strength (33 MPa, 28 MPa)
than conventional S2-glass/epoxy composites (23 MPa). It is to be noted that interlaminar
shear strength is a matrix dominant property and such improvement in interlaminar shear
strength of fiber reinforced composites is mostly due to improved property of epoxy–clay
nanocomposites. This property enhancement is achieved due to increased interfacial areas,
improved bond characteristics and unique phase morphology of the epoxy–clay
nanocomposites. Figure 10 shows the SEM micrograph of the fractured surface of the
short beam shear test specimen. Matrix agglomeration and coarse surface morphology are
observed in the fiber–matrix interface region of the nanocomposite specimen resulting
enhanced interfacial bond strength. In other case, the conventional composites show very
clean interfacial morphology indicating comparatively less bond strength. Figure 11 shows
flexural load versus deflection plot for both conventional S2-glass/epoxy composites and
S2-glass/epoxy–clay nanocomposites containing 1 and 2% nanoclay particles. The plots
clearly show enhanced flexural properties for nanocomposites specimens. The nature of
the stress and deflection plot in Figure 11 is seen to be partially nonlinear immediately
after the initial loading until the final failure occurs. The modulus data is calculated from
the initial stage of the load–displacement plot.
The flexural strength and modulus data for S2-glass/epoxy composites are presented in
Table 4. It is clearly seen from Table 4 that the S2-glass/epoxy nanocomposites containing
1832 A. HAQUE ET AL.
40

30
Shear stress (MPa)

20

10
S2 glass epoxy composite
Nanocomposite - 1% clay
Nanocomposite - 2% clay

0
0 0.5 1.0 1.5 2.0
Displacement (mm)

Figure 9. Interlaminar shear stress vs. displacement plot for both conventional S2 glass/epoxy and S2-glass/
epoxy–clay nanocomposites. (a) S2-glass/epoxy; (b) S2-glass/epoxy–nanocomposites.

Table 3. Interlaminar shear strength S2-glass/epoxy–clay nanocomposites.


Average Interlaminar
Montmorrilonite Shear Strength
Sample (wt.%) (MPa) % increase
S2 glass-epoxy 0 23.05 –
S2 glass-epoxy 1 33.21 44.07
S2 glass-epoxy 2 27.67 20.04

Figure 10. SEM micrograph of fractured surface under interlaminar shear failure.

1% nanoclay particles provide 24% higher flexural strength (370 MPa) and 17% higher
flexural modulus (32.52 GPa) than that of conventional composite (298 MPa, 27.76 GPa
average value). Although this flexural failure strength of the composite is a fiber dominant
property, the matrix has an influence in the overall properties of the composites. As a
S2-Glass/Epoxy Polymer Nanocomposites 1833

400

300
Load (N)

200

100
S2 glass epoxy composite
Nanocomposite 1% clay
Nanocomposite -2% clay

0
0 5 10 15 20 25

Displacement (mm)

Figure 11. Load vs. displacement plot for S2-glass/epoxy and S2-glass/epoxy nanocomposites in
bending test.

Table 4. Flexural test data for S2 glass/epoxy-clay nanocomposites.


Average Average % Increase
Montmorrilonite Flexural Flexural
Sample (wt.%) Strength (MPa) Modulus (GPa) Flexural Strength Modulus
S2 glass-epoxy 0 298.51 27.25 – –
S2 glass-epoxy 1.0 370.19 30.47 24.01 14%
S2 glass-epoxy 2.0 331.77 28.48 11.14 6%

Table 5. Flexural test data for epoxy–clay nanocomposites.


Average Average % Increase
Montmorrilonite Flexural Strength Flexural Modulus
Sample (wt.%) (MPa) (GPa) Flexural Strength Modulus
Epoxy 0 129 2.78
Epoxy 1.0 157 3.5 21 25
Epoxy 2.0 149 4.94 15 76
Epoxy 5.0 71.8 2.2 38 20
Epoxy 10.0 45.3 1.81 64 34

result the enhancement in flexural properties of fiber reinforced nanocomposites


is achieved due to improved properties of epoxy–clay nanocomposites and also for
unique interfacial fiber matrix bond characteristics. Table 5 shows the similar data for the
epoxy–clay system without fiber reinforcement. The results show both strength and
modulus enhancement for epoxy–clay nanocomposites with 1% montmorrilonite clay
particles. Figure 12 shows the SEM micrographs of the fractured surface for S2-glass/
epoxy and S2-glass/epoxy–clay nanocomposites. The fiber surface morphology of
1834 A. HAQUE ET AL.

(a) (b)

(c) (d)
Figure 12. OM and SEM micrographs of fractured surface under bending load: (a) S2-glass/epoxy;
(b) S2-glass/epoxy; (c) S2-glass/epoxy–clay; (d) S2-glass/epoxy–clay.

conventional S2-glass/epoxy composites is observed to be very clean and significant resin


cracking and agglomeration are not visible at the interface region. In other case, the
nanocomposites show significant resin cracking at the interface and the surface
morphology is seen to be comparatively coarse indicating enhanced interfacial bond
strength.
Figures 13 and 14 show force versus displacement plots of notched fracture toughness
specimens for epoxy resins, epoxy–clay nanocomposites, conventional S2-glass/epoxy
composites and S2-glass/epoxy–clay nanocomposites. In all these plots higher failure loads
are clearly seen for composites with 1 and 2% nanosilicates. Significant reduction of
failure load is observed for epoxy–clay composites with increased clay loading (5 and
10%). This is mostly due to enhanced agglomeration of clay particles. Table 6 shows
fracture toughness data for epoxy–clay nanocomposites (without fiber reinforcement). An
enhanced fracture toughness in the range of 28–32% is observed for epoxy–clay
nanocomposites (with 1 and 2% clay particles) in comparison to neat epoxy resin. The
fracture toughness data for S2-glass fiber reinforced composites is shown in Table 7. In
this case, the fiber reinforced nanocomposites containing 1–2% nanoclay particles also
provided 20–23% higher fracture toughness than the conventional fiber reinforced
composites. This increased fracture toughness in fiber reinforced nanocomposites is
mostly achieved due to epoxy/clay system with improved fracture toughness. The
fractured surface at the notched section of both S2-glass/epoxy and S2-glass/epoxy–clay
nanocomposites are shown in Figure 15. Enhanced matrix agglomeration with significant
matrix cracking and fragmentations are clearly seen at the interface region. This type of
morphology primarily indicates strong interfacial bonding and prolonged fracture process.
S2-Glass/Epoxy Polymer Nanocomposites 1835
150

neat epoxy
epoxy 1% clay
epoxy 2% clay
epoxy 10% clay
epoxy 5% clay

100
Load (Newton)

50

0
0.10 0.25 0.40 0.55 0.70 0.85
Displacement (mm)

Figure 13. Force vs. displacement plot of fracture toughness test: neat epoxy resin and epoxy–clay
nanocomposites.

800

600
Load (Newton)

400

200 S2 glass epoxy composite


Nanocomposite - 1% clay
Nanocomposite - 2% clay

0
0 0 .2 0 .4 0 .6 0 .8

Displacement (mm)

Figure 14. Force vs. displacement plot of fracture toughness test: S2-glass/epoxy composites and S2-glass/
epoxy–clay nanocomposites.

Table 6. Fracture toughness test data for epoxy–clay nanocomposites.


Sample Montmorrillonite % Peak Load (N) Average KQ (MPa mm1/2) % Increase
Epoxy 0 92 57.99 –
Epoxy 1 121 76.79 32.40
Epoxy 2 117 74.46 28.40
Epoxy 5 67 44.97 22
Epoxy 10 56 40.98 29.33
1836 A. HAQUE ET AL.

Table 7. Fracture toughness test data for S2 glass/epoxy–clay nanocomposites.


Sample Montmorrillonite% Peak Load (N) Average KQ (MPa mm1/2) % Increase
S2 glass epoxy 0 578.87 4135.47 –
S2 glass epoxy 1 779.12 5088.26 23.04
S2 glass epoxy 2 671.65 4984.25 20.52

(a)

(b)
Figure 15. SEM micrographs of the fractured surface of notched bend specimens: (a) S2-glass/epoxy;
(b) S2-glass/epoxy–clay nanocomposites.

In the case of conventional composites the interfacial morphology is seen to be


comparatively clean with comparatively less matrix agglomeration than that of
nanocomposites.

CONCLUSIONS

Fiber reinforced polymer nanocomposites have been manufactured using an affordable


vacuum assisted resin infusion process. TEM observation of epoxy–clay system reveals
that silicate phase is stacked alternately with polymer chain forming basic lamellar type
microstructures of the nanocomposites (intercalated). However, the micrographs also
show regions where it is not strictly of lamellar morphology, rather a group of differently
oriented parallel silicate and polymer layers bundled together (phase separated). The on-
set temperature of decomposition (Tdec) of nanocomposites is seen to be increased
approximately 20–22 C and simultaneously, the glass transition temperature (Tg) is also
observed to be slightly increased (3–4 C) than the conventional composites.
The mechanical properties of fiber reinforced polymer nanocomposites have been eva-
luated. Interlaminar shear and flexural strength of S2-glass/epoxy–clay nanocomposites
are seen to be increased almost 44 and 24%, respectively by adding 1% clay into the
S2-Glass/Epoxy Polymer Nanocomposites 1837

system. The fracture toughness of the same is also seen to be increased almost 23%.
The results indicate that the thermomechanical properties mostly increase at low clay
loadings(1–2% by weight) but decreases at higher clay loadings (5 wt.% and above).
The increased properties at low loading are possibly due to several factors: (1) enhanced
matrix properties due to lamellar structures; (2) synergistic interaction between the matrix,
clay and fibers and (3) enhanced matrix–fiber adhesion promoted by the clay. The clays
are also presumed to decrease the CTE mismatch, significantly reducing residual stresses
and leading to higher quality laminates. An increased interfacial bonding, matrix
agglomeration and coarse morphology are observed from the fractured surface of low
loading nanocomposites. The degradation of properties at higher clay loadings is believed
to be caused by phase-separated structures and also by defects in the cross-linked
structures. Further work is necessary in order to achieve fully exfoliated structure in
clay–epoxy nanocomposites.

ACKNOWLEDGEMENT

The authors thank Air Force Research Laboratory for their cooperation in performing
WAXD. Financial support received from the National Science Foundation through
CREST program and Army Research Office is gratefully acknowledged.

REFERENCES

1. Alexander, M. and Dubois, P. (2000). Polymer-layered Silicate Nanocomposites: Preparation,


Properties and Uses of a New Class of Materials, Materials Science & Engineering, 28: 1–63.
2. Morgan, A., Gilman, J. Kashiwagi, T. and Jackson, C. (2000). Flammability of Polymer-Clay
Nanocomposites, In: Proceedings of Fire Safety Developments, Non-Halogen FR’s, Standards and
Regulations, National Institute of Standards and Technology, pp. 25–39, March 12–15,
Washington.
3. Lagly, G. (1999). Introduction: From Clay Mineral-Polymer Interactions to Clay Mineral-
Polymer Nanocomposites, Applied Clay Sci, p. 15.
4. Wang, Z and Pinnavaia, T. (1998). Hybrid Organic-Inorganic Nanocomposites: Exfoliation of
Magadiite Nanolayers in an Elastomeric Epoxy, Chem. Mater, 10: 1820–1826.
5. Novak, B.M. (1993). Advanced Materials, 5: 422.
6. Giannelis, E.P. (1992). New Strategy for Synthesizing Polymer-Ceramic Nanocomposites, J.
Minerals, Metals & Materials Soc, 44(3): 28–30.
7. Kormann, X., Lindberg, H. and Berglund, L.A. (2001). Synthesis of Epoxy-Clay
Nanocomposites: Influence of the Nature of the Clay on the Structure, Polymer, 42: 1303–1310.
8. Chen, C. and Curliss, D. (2001). Resin Matrix Composites: Organoclay-Aerospace Epoxy
Nanocomposites: Influence of the Nature of the Clay on the Structure, Polymer, 42: 1303–1310.
9. Usuki, A., Kawasumi, M., Kozima, Y. and Okada, A.J. (1993). Material. Res., 8: 1174.
10. Fisher, H., Gielgens, L. and Koster, T. (1998). Nanocomposites from Polymers and Layered
Minerals: TNO-TPD Report.
11. Le Beron, P.C., Wang, Z. and Pinnavia, T.J. (1999). Applied Clay Science, 15: 11.
12. Okada, A. and Usuki, A. (1995). The Chemistry of Polymer-Clay Hybrid, Materials Science and
Engineering, C3: 109–115.
13. Messersmith, B. Phillip and Giannelis, P. Emmanuel, (1994). Synthesis and Characterizationof
Layered Silicate-Epoxy Nanocomposites, Chem. Mater. pp. 1719–1725.
14. Agarwal, B.D. and Broutman, J.L. (1980). Analysis and Performance of Fiber Composites, John
Wiley and Sons, New York, USA.

View publication stats

You might also like