You are on page 1of 10

International Conference on Challenges and Opportunities in Mechanical Engineering, Industrial Engineering and Management Studies

(ICCOMIM - 2012), 11-13 July, 2012

356

The Effect of Aluminium Oxide Filler on the


Wear Resitance of Glass Fabric Reinforced
Epoxy Composites
B.R. Raju, B. Suresha and R.P. Swamy
Abstract--- In this study the abrasive wear behaviour of glass fabric reinforced epoxy (G-E) and Aluminium
oxide filled G-E (Al2O3-G-E) composites have been carried out by using a rubber wheel abrasion test (RWAT) rig.
Samples of G-E with 5, 7.5 and 10 wt% content of Al2O3 were tested under different loads and abrading distances.
Also, conventional weighing, determination of wear volume, specific wear rate, and examination of the worn surface
morphological features by scanning electron microscopy (SEM) were done. The results showed varied responses
under different abrading distance because of the inclusion of different wt% of Al2O3 filler loading. Further, the test
results show that glass fabric reinforcement obviously improves the strength of epoxy and glass fabric- Al2O3
exhibits a synergistic effect on the wear resistance and reinforcing epoxy simultaneously. Wear of G-E composite
was found to be mainly due to a micro-cracking and fiber fracture mechanisms. It was found that the micro-cracking
mechanism had been caused by progressive surface damage. Further, it was also noticed that G-E composite wear is
reduced to a greater extent by addition of the Al2O3, in which the wear was dominated by micro-plowing/microcutting mechanisms instead of micro-cracking. Selected mechanical properties such as hardness, tensile strength,
and elongation at fracture were analyzed for investigating wear property correlations. The worn surface
morphology was performed to investigate the wear mechanisms involved during abrasion process.
Keywords--- Glass Fabric Reinforced Epoxy Composite, Al2O3 Filler, Abrasive Wear, Worn Surface Morphology

I.

INTRODUCTION

LASS fiber reinforced polymer matrix composites have been extensively used in various fields such as
aerospace industries, automobiles, marine, and defense industries [1]. Their main advantages are good
corrosion resistance, lightweight, dielectric characteristic, and better damping characteristics than metals. Fabric
reinforced and particulate filled polymer composites have become attractive because of their wide spread
applications and low cost. A possibility that the incorporation of both particles and fibers in polymer could provide a
synergism in terms of improved properties and performance has not been adequately explored so far. However, some
recent reports suggest that by incorporating filler particles into the matrix of fiber reinforced composites, synergistic
effects may be achieved in the form of higher modulus and reduced material cost, yet accompanied with decreased
strength and impact toughness [2].
One part of composite material for engineering applications may be represented by a thermosetting polymer
matrix, e.g. an epoxy resin, which already covers alone some of the demanded properties. Diglycidyl of bisphenol A
(DGEBA) type epoxy resin being the most widely matrix for innumerable applications, owing to its well balanced
chemical, adhesive, thermal and processing characteristics. However, the high coefficient of linear expansion, low
thermal conductivity and limited mechanical properties of epoxies limit their use in mechanical and tribological
applications. Recently many attempts were made to develop epoxy resin composites modified by fibers and fillers to
improve the mechanical and tribological performance of the epoxy matrix [3-5]. Bahadur and Zheng [6], in their
studies on short glass fiber (SGF) reinforced polyester by varying SGF up to 60 wt% described the effect of glass
fiber on mechanical properties of the composites. Epoxy resins are superior to polyesters in resisting moisture and

B.R. Raju, Department of Mechanical Engineering, PES Institute of Technology and Management, Shivamogga, India. E-mail:
rajusrujan@gmail.com
B. Suresha, Department of Mechanical Engineering, The National Institute of Engineering, Mysore-570 008, India
R.P. Swamy, Department of Mechanical Engineering, University B.D.T. College of Engineering, Davangere -577004, India
PAPER ID: MEP22

ISBN 978-93-82338-04-8 | 2012 Bonfring

International Conference on Challenges and Opportunities in Mechanical Engineering, Industrial Engineering and Management Studies
(ICCOMIM - 2012), 11-13 July, 2012

357

other environmental influences and offer lower shrinkage and better mechanical properties. However, the structural
application of epoxy resin and its composites is usually limited owing to the relatively poor thermal stabilities and
load carrying capacity. In order to enhance the wear resistance and thermal stability, many research studies have
been carried out. One of these is the fiber reinforcement into the epoxy matrix. Different synthetic fibers and hard
ceramic or metal particles have been tried as fillers in the epoxy matrix [7-9]. Kim et al. [10] reported that the
damage could occur during the fabrication process, storage, service, transport, and maintenance. They are
susceptible to mechanical damage when they are subjected to effects of tension, compression, and flexure, which
can lead to interlayer delamination. The increase of external load favors the propagation of delamination through the
interlayer leading to the catastrophic failure of the component. Another work reported by Unal and Mimaroglu [11]
evaluated mechanical properties of Nylon-6 by incorporating one or a combination of more than one filler by
varying the weight percent. They observed that the tensile strength and modulus of elasticity of Nylon- 6 composites
increased with increase in filler weight percent. Osmani [12] evaluated the mechanical properties of alumina filled
glass-epoxy (G-E) composites and reported that tensile and shear strengths decreased with increase in alumina
content and flexural strength and modulus increased. Suresha et al. [13] studied the mechanical and sliding wear
behavior of SiC filled G-E composites and concluded that the SiC filler addition improved the mechanical as well as
wear resistance of G-E composites. Wetzel et al. [14] evaluated the mechanical and tribological properties of epoxy
filled with nano-Al2O3 and microCaSiO3 hybrid composites. They concluded that incorporation of micro-and nanoscale particles improved the mechanical as well as tribological properties.
Among the wear types, the abrasive wear situation encountered in industries connected with power, automobile,
pumps handling industrial fluids, and earth moving equipment has been received increasing attention. Glass/carbon
fibers are the best known and most widely used reinforcing fibers in advanced polymeric composites. Reports
related to application of polymer matrix composites on mechanical and tribological components such as gears, cams,
wheels, and impellers are cited in literature. The importance of the tribological properties convinced various
researchers to study the wear behavior and to improve the wear resistance of polymers and fiber-reinforced
polymeric composites. Chand et al. [15] reported the three-body abrasive wear behavior of short glass fiber
reinforced polyester composite. Chand and Gautam [16] reported the influence of load on the abrasion of fly ash,
glass fiber- reinforced composites. Suresha et al. [17-19] investigated the abrasive wear behavior of epoxy/vinyl
ester filled with or without particulate filler and glass/carbon fabrics. An investigation on the wear behavior of the
composite materials with epoxy matrix filled with hard powders was reported by Visconti et al. [20]. Studies were
also made on the effect of various fillers on the abrasive wear behavior of polymeric materials [21] and polymer
composites [22-24].
A survey of the literature indicates that the effect of fiber/filler on the abrasion of polymer/composites is a
complex and unpredictable phenomenon [25]. This is because physical effects such as fiber fragmentation,
debonding, pullout, etc., affect the behavior of composite material subject to the abrasive wear process. It is also
difficult to predict their relative contribution because various other mechanisms and influencing factors are involved
(ploughing, cutting and cracking of the matrix). Although, a good amount of work has been reported on mechanical
and three-body wear behavior of polymer matrix composites as discussed earlier in this section, no literature could
be cited on the mechanical and abrasive wear aspect of G-E filled with Al2O3 particles. For sliding elements under
tribological loading, however, the industrial acceptability of polymers is often limited by their thermal behaviors.
The degradation of their mechanical properties at elevated temperature restricts the possibility to apply of these
materials at high sliding speed and loading conditions. That is why high temperature resistant polymers are generally
preferred for such tribological applications. Therefore an attempt has, therefore, been made to study the three-body
abrasive wear behavior of G-E composite filled with different proportions of very fine Al2O3 particles (particle size
5 m) under two different loads and for different abrading distances (250-1000 m). The abrasive wear behavior has
been quantified in terms of wear volume and specific wear rate. The different wear mechanisms under different
abrasive wear conditions have also been reported.

II.

EXPERIMENTAL DETAILS

2.1 Materials and Preparation Method


Woven glass plain weave fabrics made of 360 g/m2, containing E-glass fibers of diameter of about 12 m have
been employed. The epoxy resin (LAPOX L-12) was mixed with the hardener (K-6, supplied by ATUL India Ltd.,
Gujarat, India) in the ratio 100:12 by weight. The filler chosen was aluminium oxide modified using a silane
coupling agent and the average particle size is 5 m. The details of the composites properties are shown in the
Table 1.

ISBN 978-93-82338-04-8 | 2012 Bonfring

International Conference on Challenges and Opportunities in Mechanical Engineering, Industrial Engineering and Management Studies
(ICCOMIM - 2012), 11-13 July, 2012

358

As regards to the processing, on a Teflon sheet, E-glass woven fabric was placed over which the epoxy matrix
system consisting of epoxy and hardener was smeared. Dry hand lay up technique was employed to fabricate the
composites. The stacking procedure consists of placing the fabric one above the other with the resin mix well spread
between the fabrics. A porous Teflon film was again used to complete the stack. To ensure uniform thickness of the
sample, a 3 mm spacer was used. The mould plates were coated with release agent in order to aid the ease of
separation on curing. The cast of each composite after 12 h of impregnation and dried for 2 h at 100 oC followed by
compression moulding at a temperature of 390 oC and pressure of 7.35 MPa. The slabs so prepared measured 250
mm 250 mm 3 mm by size. To prepare different wt. % of particulate filled G-E composites, besides the epoxy
hardener mixture additional wt% filler particles were included to form the resin mix. The details of the composites
(including wt% of the constituents) made are shown in Table 2. The weight percent of the glass fiber in the
composite is 60 wt%. Mechanical and abrasion test samples were prepared according to ASTM standard from the
cured sheet using abrasive cut-off machine.
Table 1: Physical & Mechanical Properties of the Constituents Selected or the Present Work
Property

Epoxy

Glass fibers

Al2O3 filler

Density (g/cm3)

1.15

2.54

3.69

Tensile strength (MPa)

110

3400

260-300

Tensile Modulus (GPa)

4.1

72.3

300-400

Table 2: Composite Selected for the Present Study

Sample name (Designation)

Epoxy (wt.%)

Al2O3 (wt.%)

Glass fabric reinforced epoxy (G-E)

40

-------

Aluminium oxide filled G-E (5 Al2O3-G-E)

35

Aluminium oxide filled G-E (7.5 Al2O3-G-E)

32.5

7.5

Aluminium oxide filled G-E (10 Al2O3-G-E)

30

10

2.2 Characterisation
Density of the composites was determined by using a high precision electronic balance (Mettler Toledo, Model
AX 205) using Archimedes principle. Hardness (Shore-D) of the samples was measured as per ASTM D2240, by
using a Hiroshima make hardness tester (Durometer). Five readings at different locations were noted and average
value is reported. Tensile properties were measured using a Universal testing machine in accordance with the ASTM
D-3039 procedure at a cross head speed of 5 mm/min and a gauge length of 50 mm. The tensile strength and
modulus were determined from the stress-strain curves. Five samples were tested in each set and the average value
was reported. A minimum of four samples were tested in each case. The tensile tests were carried out on a fully
automated Lloyd LR-20 kN Universal testing machine connected to a computer with DAPMAT software.
The three-body abrasive wear tests were conducted using a dry sand/rubber wheel abrasion tester as per ASTM
G-65. The details of the samples preparation and wear testing procedure have been described elsewhere [17].
The experiments were carried out at two different loads (23 and 36 N) under different abrading distances (250 m
to 1000 m). Densities of the composites were determined using a high precision weighing balance by using
Archimedess principle. The wear was measured by the loss in weight, which was then converted into wear volume
using the measured density data. After the wear test, the sample was again cleaned. The specific wear rate (Ks) was
calculated from the equation:

ISBN 978-93-82338-04-8 | 2012 Bonfring

International Conference on Challenges and Opportunities in Mechanical Engineering, Industrial Engineering and Management Studies
(ICCOMIM - 2012), 11-13 July, 2012

359

(2)
Where V is the volume loss, L is the load and D is the abrading distance.

III.

RESULTS AND DISCUSSION

3.1 Effect of Filler Loading on Density


The measured densities of the samples are listed in Table 3. Comparing the results it was observed that inclusion
of ceramic fillers into G-E showed higher density. The density of Al2O3 filled G-E is 2.19 which is highest when
compared to other composites. This is because of the filler Al2O3 has higher density. The densities of all
microparticles filled G-E is higher than the density of unfilled G-E composites.
3.2 Effect of Filler Loading on Hardness
By using Duro-hardness tester, the hardness of the composites is measured; the values recorded are given in
Table 3. The hardness of G-E composite increased with increase of Al2O3 filler loading. From Table 3, it can be seen
that the Al2O3 filler greatly increased the hardness of G-E, which can be attributed to the higher hardness and more
uniform dispersion of Al2O3 filler. The higher hardness is exhibited by the 10 wt% Al2O3 filled G-E compared to
other mirocomposites. The table shows that for a 10 wt% increase in Al2O3 content there is 11% increase in
hardness. The increase in Al2O3 content results in an increase in brittleness of the composite. Hence this results in an
increase in hardness value of the composite. Particulate filled G-E composites with sufficient surface hardness are
resistant to in-service scratches that can compromise fatigue strength and lead to premature failure. Therefore, under
an indentation loading, microparticles would undergo elastic rather than plastic deformation, as compared to unfilled
G-E composites. The improvement in hardness with incorporation of filler can be explained as follows: under the
action of a compressive force, the thermoset matrix phase and the solid fiber and filler phase will be pressed
together, touch each other and offer resistance. Thus the interface can transfer load more effectively although the
interfacial bond may be poor. This results in enhancement of hardness of Al2O3 filled G-E composites.
Table 3: Physico-Mechanical Properties of G-E and Al2O3 Filled G-E Composites
Sample code

G-E

5 Al2O3-G-E

7.5 Al2O3-G-E

10 Al2O3-G-E

Density (g/cm3)

1.984

2.12

2.23

2.3

Tensile strength, (MPa)

254

324

343

352

Tensile modulus, E(GPa)

8.34

10.6

11.26

11.55

Elongation, e (mm)

7.1

6.4

6.2

5.9

Hardness (Shore-D)

63

66

69

72

3.3 Tensile Properties


The typical load-deformation curves of unfilled and particulate filled G-E campsites are shown in Fig. 1 and the
measured mechanical test results are listed in Table 3. It is clear from this table that the tensile strength is increasing
with the increase in Al2O3 content. As the investigation is mainly focused on filler content rather than G-E
composite, taking the 7.5 wt% Al2O3 loading into G-E there is about 25% increase in tensile strength. Because of all
added constituents are brittle in nature in comparison to epoxy therefore this is reflected by the mechanical
properties of the composite as a hybrid one. Thus there is an increase in tensile strength of Al2O3 filled G-E
composite with the increase in Al2O3 content. This could be attributed the uniform dispersion of silane treated Al2O3
filler in G-E. Addition of a coupling agent can provide covalent bonding between epoxy/ Al2O3 hybrid materials,
there by enhancing the mechanical properties of the composites. The surface modified Al2O3 can interact with the
fiber surface and hydrogen bonding increases and leads to the better interaction with glass fiber and epoxy. Addition
of ceramic filler increases the effective mechanical interlocking, which in turn increases the frictional force between
the fiber and matrix. It can also be seen from Table 3 that the tensile modulus of Al2O3 filled G-E composites
increases with increase in wt% of Al2O3 content. It is clear from this table that the elongation at break is decreasing

ISBN 978-93-82338-04-8 | 2012 Bonfring

International Conference on Challenges and Opportunities in Mechanical Engineering, Industrial Engineering and Management Studies
(ICCOMIM - 2012), 11-13 July, 2012

360

with the increase in Al2O3 content. This table shows that for a 10 wt% increase in Al2O3 content there is 11%
decrease in percentage elongation at break. The increase in Al2O3 content results in an increase in brittleness of the
composite. Hence this results in a decrease of the percentage of elongation at break. As the wt. fraction of Al2O3
filler increase, the tensile modulus of the G-E composites increases, but at the same time the system becomes more
brittle. The increase in the tensile strength with wt. fraction of filler is attributed to the high modulus of ceramic
filler which is dispersed uniformly in the fabric layers of G-E composites. Youngs modulus is mainly dependent on
the matrix deformation of the composite and increases as the slope of load-deformation curve at the initial stage and
is practically not much influenced by the interfacial strength between fiber and the matrix . Generally, the addition
of ceramic fillers and glass fiber reduces the elongation at break because of the lower elongation at break values of
ceramic fillers and glass fiber compared to that of epoxy matrix. The addition of Al2O3 particles causes a dispersion
of these particles in the matrix which impede to the propagation of failure along the loading direction. Thus the
failure would propagate easily in those directions where the dispersoid concentration is less leading to increased
tensile strength, tensile modulus, and lower elongation.
14

Load, KN

12
10
8

G-E

5Al2O3-G-E

7.5Al2O3

10Al2O3-G-E

0
0

Displacement, mm
Figure 1: Typical Load v/s Displacement Curves G-E and Al2O3 Filled G-E Samples
3.4 Abrasive Wear Volume and Specific Wear Rate
The three-body abrasive wear results of the unfilled G-E composite indicate that a poor wear performance as
compared to Al2O3 filled G-E composites.
In unfilled G-E, the wear results are poor compared to those obtained in particulate filled G-E. A G-E composite
sample is fixed to a sample holder (as indicated in Fig. 2).
This way the mounted GE sample is made to come in contact with a rotating rubber wheel. In this case, the
abrasion started through contact with the softer phase (epoxy). This contributed to severe matrix damage as reported
in Ref. [5,14]. In Ref. [14], it is suggested that, under low stress three-body abrasive wear, the sand particles
behaved in one of the following ways. First, from free fall, the sand particles gained energy from the rubber wheel
and then struck the composite surface, which would result in the formation of pits. Second, the abrasive particles
were embedded in the rubber wheel, transforming the three-body abrasion into multi-pass two-body abrasion. Third,
the particles roll at the interface, causing plastic deformation to the composite. It has been reported that the presence
of glass fibers in the interface reduces the wear rate [5]. In the present work, during abrasion, in the initial stage of
abrasion, the particle penetrated the soft outer layer of the composite (matrix) due to the lower hardness. Once the
matrix layer was removed, the harder phase of the composite (fibers) was exposed to the rubbing area, which acted
as a protector, leading to a reduction in the removal of material. However, breakage, debonding and pull-out of
fibers have been observed at longer abrading distances.

ISBN 978-93-82338-04-8 | 2012 Bonfring

International Conference on Challenges and Opportunities in Mechanical Engineering, Industrial Engineering and Management Studies
(ICCOMIM - 2012), 11-13 July, 2012

361

Figure 2: Dry Sand/Rubber Wheel Abrasion Test Apparatus


Figure 4a and b show the wear volume loss with abrading distance for different loads. It is evident from these
figures that irrespective of the type of samples used, there is a linear trend of wear volume loss. The Al2O3 filled GE composite exhibited considerably lower wear volume loss than that of G-E composite. This is attributed to the
glass fiber reinforcement in epoxy decrease the abrasive wear resistance due to debonding and tearing at the
glass/matrix interface. Similar behaviour was observed at a higher load of 36 N.Figure 4a and b shown as
histograms, the comparative abrasive wear performance of G-E and Al2O3-G-E composites at 23 N and 36 N loads
respectively. The specific wear rate data reveals that initially the specific wear rate tends to decrease with increasing
abrading distance and further it strongly depends on the applied load for both samples. Also observed is the earlier
noted fact that G-E composite exhibits the highest specific wear rate. It is interesting to note that for Al2O3-G-E
system, the specific wear rate is on the lower side.
The specific wear rate of unfilled and Al2O3 filled glass fabric reinforced epoxy (Al 2O3-G-E) composites versus
the applied load (24 and 36 N) at 200 rpm rotational speed are shown in Fig.3a-b. In general, increase in the applied
load decreases the specific wear rate (Fig. 4a-b ). An increase in load causes a decrease in the wear of particulate
filled G-E composites. Similar results of decreasing wear with increasing load of fiber reinforced polymers are
reported. This behaviour is due to the formation of ridges on the surface of the polymer during sliding on an abrasive
medium [26]. The material from the grooves is not removed but rather displaced towards the sides to form the
ridges. Hence quantification of wear losses either by dimension loss or mass loss does not give the right picture of
the actual loss. Hence there seems to be a reduction in the wear loss with increasing load. Moreover, It was observed
that wear performance of G-E enhanced due inclusion of Al2O3 filler. However the higher loading of fillers in G-E
composite (10% Al2O3) exhibit lower specific wear rate at all the loads and abrading distances. The maximum
specific wear rates of unfilled G-E and Al2O3 filled G-E are 2.2 x 10-11 and 1.69 x 10-11 m3/Nm respectively. The
reason for improvement in wear characteristics may be attributed to the type, surface treatment and filler loading. In
this work, the decrease in the specific wear rate with an increase in the applied load is consistent with other
published works [2325]. However, other reported works [7, 12, 13] showed no influence of the applied load on the
wear rate of the composite. The reduction in the wear rate with increase in applied load could be due to the fact that
the apparent contact area was greatly increased at higher applied loads compared to that at lower applied loads.
Increase in the apparent contact area allows a large number of sand particles to encounter the interface and share the
stress. This, in turn, leads to a steady state or reduction in the wear rate.
The main reason for reinforcing fibers into polymers is to improve their mechanical properties, but the effects on
wear rate are not invariably beneficial. The correlations of wear volume loss with selected mechanical properties
such as (e) factor (where is the ultimate tensile strength and e is the ultimate elongation), hardness (H) have been
reported for single-pass studies of polymers without fillers and composites [15, 16]. Lancaster [15] stated that the
product of and e is a very important factor which controls the abrasive wear behavior of composites. Generally
fiber/filler reinforcement increases the tensile strength () of neat polymer, they usually decrease the ultimate
elongation (e) and hence the product (e) may become smaller than that of neat polymer. Hence, reinforcement
usually leads to deterioration in the abrasive wear resistance

ISBN 978-93-82338-04-8 | 2012 Bonfring

International Conference on Challenges and Opportunities in Mechanical Engineering, Industrial Engineering and Management Studies
(ICCOMIM - 2012), 11-13 July, 2012

(a)

362

(b)

2.5

2.5
2

Specific wear rate (m3 /Nm)

Specific wear rate (m3/Nm)

Figure 3: Wear Volume Loss of Unfilled & Al2O3 Filled G-E Composites at (a)23 N & (b)36 N
The model proposed by Ratner et al. [16] states that the rate of material removal is inversely proportional to the
product of stress and strain at rupture. In the present work, for the composites tested, the wear volume loss increased
with increase in e factor (Table 3) and the results are contradicting the reported literature [15, 16].

1.5

1.5

G-E

10SiO2
Al2O3

0.5
0
250

500

750

1000

Abrading distance (m)

G-E
1

10SiO2

0.5

10Al2O3

0
250

500

750

1000

Abrading distance (m)

(a)

(b)

Figure 4: Specific Wear Rate of Unfilled and Al2O3 Filled G-E Composites at (a) 23 N and (b) 36 N
3.5 Worn Surface analysis
To ascertain the wear mechanisms of unfilled and Al2O3 filled G-E composites, the worn surface of the samples
during abrasion were observed using SEM, and the micrographs are given in Figures 5 and 6 respectively.
Figure 5(a) shows that some deeper furrows with various width presented at the worn surface of unfilled G-E
composite. At higher magnification, a more damage to the epoxy matrix can be observed [Fig. 5(a)]. However,
apparently different phenomenon for the composite is that, a lot of GF embedded in the dark transfer film, which is
surrounded by the melted blend matrix, and the length is less than 50 m. This length is much shorter than that of
unrubbed sample [Fig. 5(a)], indicating that the glass fiber was fractured during sliding under the frictional shear and
normal load. This result indicates poor adhesion of the matrix to the fibers as several clean fibers appear on the
abraded surface. The higher wear rate in GE composite may be attributed to lower matrix ductility and poorer
fibermatrix adhesion. The abraded surface shows evidence of debonding at each fibermatrix is not sufficient to
result in the removal of fiber. The fracture of fiber is due to abrasion and transverse bending by sharp abrasive
particles, resulting in fragments of fibers torn from the matrix (Fig. 5(a)). Once again, the micrograph shows poor
adhesion between fiber and matrix.

ISBN 978-93-82338-04-8 | 2012 Bonfring

International Conference on Challenges and Opportunities in Mechanical Engineering, Industrial Engineering and Management Studies
(ICCOMIM - 2012), 11-13 July, 2012

363

Fig. 6 (a & b) shows the photomicrograph of abraded surface of Al 2O3 filled G-E composites tested at a load of
36 N. At lower abrading distance (Fig. 6b), severe matrix and fiber destructions were the characteristic features in
the micrographs. Also the evidence of fiber cutting, small voids left by debonded fibers and fatigue damage of
matrix can be seen on the surface. It is thought that the observed damaged fibers are the result of surface fatigue due
to repeated abrasion by silica sand particles. The crack propagation through the fiber and interfacial debonding were
also observed at because of the brittle nature of glass fibers, which fractures due to repeated abrasion by silica sand
particles. The interfacial debonding is less because of fine Al2O3 dispersed uniformly in G-E composite and hence,
the fibers are less exposed to the abrasive medium resulted in the lower wear volume loss as compared to unfilled GE composites.

(a)

(b)

Figure 5: SEM Photomicrographs of Unfilled G-E Composite at 36 N :(a)1000m & (b)250m

(a)

(b)

Figure 6: SEM Photomicrographs of 10%Al2O3 Filled G-E Composite at 36 N : (a) 1000 m. & (b) 250 m
It is evident from the SEM photographs (comparing Fig. 5b with Fig. 6b) that the 10 % Al2O3 filled G-E is
showing lesser degree of worn surface features compared to unfilled G-E sample at 36 N load applications. In the
case of samples subjected to 36 N load, one can see less number of broken fibers with less debris formation in the
Al2O3 filled G-E sample (Fig. 6b), whereas in the unfilled G-E sample (Fig. 5b), de-bonding of the fiber with
cleavage type of fracture is seen. Now, coming to the samples subjected to higher abrading distance (1000 m),
masking of fibers are noticed in the Al2O3 filled sample (Fig. 6a). On the other hand, the SEM features of unfilled GE sample (Fig. 5a) show large number of broken fibers with lot of distortion in the matrix and also higher degree of
debris formation.

ISBN 978-93-82338-04-8 | 2012 Bonfring

International Conference on Challenges and Opportunities in Mechanical Engineering, Industrial Engineering and Management Studies
(ICCOMIM - 2012), 11-13 July, 2012

IV.

364

CONCLUSIONS

In this study, an experimental investigation has been conducted to evaluate the mechanical properties and threebody abrasive wear behaviour of glass-fiber reinforced epoxy matrix filled with different proportions of very fine
Al2O3 particles. The following main conclusions can be drawn from this study.
1.
2.

3.

4.

5.

The present work shows that incorporation of Al2O3 filler into G-E composites modifies the tensile and
tribological properties of the composites.
Compared with unfilled G-E composite, the tensile modulus of Al2O3 filled G-E improves a little but the
tensile strength improves obviously. However, glass fiber- Al2O3 exhibits synergistic effects on wear
resistance and reinforcing the epoxy simultaneously, because of its special spherical-cylindrical rod
structure and the interaction between Al2O3 and glass fiber. The higher percentage (10 wt%) of Al2O3
content in G-E composite results to higher tensile strength, hardness and a less percentage of elongation at
break.
The maximum wear volume and specific wear rate is observed for unfilled G-E composites. The specific
wear rate increased with applied load at lower abrading distance and decreased with increasing abrading
distance.
The worn surface morphology for unfilled and Al 2O3 filled G-E composites show that the plowing action is
predominating, brittle behaviour and the cracking mechanism under three-body abrasive wear conditions.
During wear process the important SEM features observed fiber fracture, tearing of matrix, cracking wear
mechanism etc.
For Al2O3 filled G-E, increase in the filler loading causes a decrease in wear, due to ridge formation.

REFERENCES
[1] Kuljanin, J., Vuckovic, M.,. Comor,M.I., Bibic, N., Djokovic, V., Nedeljkovic, J.M., Influence of CdS-filler on
the thermal properties of polystyrene, European Polymer Journal,2002, 38(8), pp 16591662 .
[2] Weidenfeller, B., Hofer, M., Schilling, F., Thermal conductivity, thermal diffusivity, and specific heat capacity
of particle filled polypropylene, Composites Part A: Applied Science and Manufacturing, 2004, 35 (4), pp 423
429.
[3] Ferreira, J.M., Pires, J.T.B., Costa, J.D., Zhang, Z.Y., Errajha ,O.A., and Richardson, M., Fatigue Damage
Analysis of Aluminized Glass Fiber Composites, Materials Science and Engineering A, 2005, 407, pp 16
[4] Amuzu, J.K.A., Briscoe, B.J.and Tabor, D., Polymers as Bearing and Lubricants; Aspects and Fundamental
Research: Advances in Tribology, 1978, pp 59-62.
[5] Kaushal ,S.,and Kishore, Analysis of Deformation Behaviour and Fracture Features in Glass-epoxy Composites
Toughened by Rubber and Carbon Additions, Journal of Materials Science Letters, 1992, 11, pp 8688.
[6] Bahadur, S.,and Zheng, Y., Mechanical and Tribological Behaviour of Polyester Reinforced with Short Glass
Fibers, Wear, 1990,137, pp 251-266.
[7] Zhang, Y., Pickels, C.A., and Cameron, J., The Production and Mechanical Properties of Silicon Carbide and
Alumina Whisker-Reinforced Epoxy Composites, Journal of Reinforced Plastics and Composites, 1992, 11, pp
11761186.
[8] Cao, Y.,and Cameron, J. ,Flexural and Shear Properties of Silica Particle Modified Glass Fiber Reinforced
Epoxy Composite, Journal of Reinforced Plastics and Composites, 2006, 25, pp 347359.
[9] Srivastava, V.K., Pathak, J.P., and Tahzibi, K.,Wear and Friction Characteristics of Mica-filled Fiber Reinforced
Epoxy Resin Composites, Wear, 1992,152, pp 343350.
[10] D-H kim Composite structures for civil and architectural engineering.
[11] unal ,H., and mimargolu, A., Influence of filler addition on the mechanical properties of nylon-6 polymer,
journal of reinforced plastics and composites, 2004, 23, pp 5461-69.
[12] Osmani, Mechanical Properties of Glass-Fiber Reinforced Epoxy Composites Filled with Al2O3 Particles.
Journal of Reinforced Plastics and Composites, 28, 2861-2867 (2009).
[13] Suresha, B., Siddharamaiah, Kishore, Seetharamu, S., Sampath kumaran, P., Investigations on the influence of
graphite filler on dry sliding wear and abrasive wear behaviour of carbon fabric reinforced epoxy composites,
Wear, 2009, pp 267
[14] Wetzel, B., Haupert, F., Zhang, M.Q., Epoxy nanocomposites with high mechanical and tribological
performance, Composite Science and Technology, 2003, 63, pp 2055-2067.
[15] Chand, N., Naik, A.M, and Neogi, S., Three-Body Abrasive Wear of Short Glass Fiber Polyester Composite,
Wear, 2000, 242(1-2), pp 38-46.

ISBN 978-93-82338-04-8 | 2012 Bonfring

International Conference on Challenges and Opportunities in Mechanical Engineering, Industrial Engineering and Management Studies
(ICCOMIM - 2012), 11-13 July, 2012

365

[16] Chand N., and Gautam, K.K.S., Influence of Load on Abrasion of Fly Ash-glass Fiber reinforced Composites,
Journal of Materials Science Letters, 1994, 13(4), pp 227-300.
[17] Suresha, B., Chandramohan, G., Siddaramaiah, Sampathkumaran, P. , and Seetharamu, S., Three-Body
Abrasive Wear Behavior of Carbon and Glass Fiber Reinforced Epoxy Composites, Mater. Sci. Eng. A, 2007,
443(1-2), pp 285-292.
[18] Suresha, B., and Chandramohan, G., Three-Body Abrasive Wear Behavior of Particulate- Filled Glass-Vinyl
Ester Composites, J. Mater. Process. Tech., 2008, 200(1), pp 306-311.
[19] Suresha, B., Chandramohan, G., Kishore, Sampathkumaran, P., and Seetharamu, S., Mechanical and ThreeBody Abrasive Wear Behavior of SiC Filled Glass-Epoxy Composites, Polym. Compos., 2008, 29(9), pp 10201025 .
[20] Visconti, I.C. , Langella, A., and Durante, M., The Wear Behaviour of Composite Materials with Epoxy Matrix
Filled with Hard Powder, Applied Composite Materials, 2001, 8(3), pp 179-189.
[21] Vaziri, M., Spurr , R.T., and Stott, F.H., An Investigation of the Wear of Polymeric Materials, Wear, 1988,
122(3), pp 329-342.
[22] Vishwanath, B., Verma A.P. , and Rao, C.V.S.K., Wear Study of Glass Woven Roving Composite, Wear, 1989,
131(2), pp 197-205.
[23] Suresha, B., Chandramohan, G., Siddaramaiah and Jayaraju, T., Three-body Abrasive Wear Behavior of E-glass
Fabric Reinforced/Graphite-filled Epoxy Composites, Polymer Composites, 2008, 29(6), pp 631-637.
[24] Suresha, B., Chandramohan , G.and Mohanram, P.V. , Role of fillers on three-body abrasive wear behavior of
glass fabric reinforced epoxy composites, Polymer Composites, 2009, Vol. 30(8), pp. 1106-1113.
[25] Bartenov ,G.M., and Lavrentov, V.V. , Lee L.H., and Ludema, K.C., Friction and Wear of Polymers, Tribology
Series 6, Elsevier, Amsterdam ,1981.
[26] Jayshree Bijwe, John Rajesh, J., Jeyakumar, A., Gosh A., and. Tewari, U.S , Tribology International, 2000, 33,
6971.

ISBN 978-93-82338-04-8 | 2012 Bonfring

You might also like