You are on page 1of 14

This article was downloaded by: [Duke University Libraries]

On: 10 October 2013, At: 01:17


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Tribology Transactions
Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/utrb20

A Study on Tribological Behavior of Alumina-Filled


Glass–Epoxy Composites Using Taguchi Experimental
Design
a a
Sandhyarani Biswas & Alok Satapathy
a
Department of Mechanical Engineering , National Institute of Technology , Rourkela,
769008, India
Published online: 24 Jun 2010.

To cite this article: Sandhyarani Biswas & Alok Satapathy (2010) A Study on Tribological Behavior of Alumina-Filled
Glass–Epoxy Composites Using Taguchi Experimental Design, Tribology Transactions, 53:4, 520-532

To link to this article: http://dx.doi.org/10.1080/10402000903491309

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Tribology Transactions, 53: 520-532, 2010
Copyright C Society of Tribologists and Lubrication Engineers

ISSN: 1040-2004 print / 1547-397X online


DOI: 10.1080/10402000903491309

A Study on Tribological Behavior of Alumina-Filled


Glass–Epoxy Composites Using Taguchi Experimental
Design
SANDHYARANI BISWAS and ALOK SATAPATHY
Department of Mechanical Engineering, National Institute of Technology, Rourkela 769008, India

This article describes the development of alumina-filled rotor blades, pump impeller blades, high-speed vehicles, water
Downloaded by [Duke University Libraries] at 01:17 10 October 2013

glass fiber–reinforced epoxy matrix composites and compared turbines, aircraft engine blades and engineering/structural com-
the experimental results obtained from Taguchi experimental ponents operating in desert environments, etc. (Tewari, et al. (1),
design with the reported theoretical erosion model. An exhaus- (2)).
tive literature survey indicated that six control factors, viz. im- Solid particle erosion manifests itself in thinning of com-
ponents, surface roughening, surface degradation, macroscopic
pact velocity, filler content, erodent temperature, impingement
scooping appearance, and reduction in functional life of the struc-
angle, stand-off distance, and erodent size, predominantly af-
ture. Hence, it has been considered as a serious problem re-
fect on erosion behavior of polymer composites. This system-
sponsible for many failures in engineering applications. Erosion
atic experimentation has led to determination of significant pro- resistance has become an important material property, particu-
cess parameters and material variables that predominantly in- larly while selecting alternative materials such as polymer-based
fluence the wear rate of different particulates-filled glass fiber– composites. Ongoing research efforts in this field are mainly
reinforced composites. The comparative study indicates that focused on studying the influence of experimental and target-
although the alumina-filled composites exhibit relatively infe- related parameters. This has been reviewed earlier by Barkoula
rior mechanical properties, their erosion wear performance is and Karger-Kocsis (3) and the work done by the previous inves-
better than that of the published literature red mud–filled glass tigators has been summarized subsequently by Harsha, et al. (4).
fiber–reinforced composites (S. Biswas and A. Satapathy, Ma- Recently Biswas, et al. (5) have reported exhaustively on the solid
terials and Design, vol. 30, 2009). The glass fiber–reinforced particle erosion behavior of polymer matrix composites. Studies
of erosive wear by solid particle impingement involve the mea-
epoxy matrix composite filled with alumina content showed
surement of several significant experimental parameters such as
semiductile erosion behavior with peak erosion rate occurring
particle velocity, particle size (distribution), angle of incidence,
at 60◦ impingement angle. The morphology of eroded surfaces
and others. Though a number of research papers on the erosion
is examined by using scanning electron microscopy (SEM) and wear exist in the literature, most are confined to fiber-reinforced
possible erosion mechanisms are discussed. composites and only limited reports are available on particulate-
filled polymer composites.
KEY WORDS The inclusion of inorganic fillers into polymers for commer-
Epoxy Resin; E-Glass Fiber; Taguchi Method; ANOVA; Con- cial applications is primarily aimed at the cost reduction and stiff-
firmation Experiment ness improvement (Rothon (6); Bonner (7). It is worth noting
that the inclusion of micrometer-sized particulates into polymers,
INTRODUCTION a high filler content typically greater than 20 vol% is generally
Polymer composites have been used for a long time with required to bring the above-stated positive effects into play. This
an increased demand in various engineering fields, particularly would detrimentally affect some important properties of the ma-
in aerospace applications, due to their high specific mechanical trix polymers such as processability, appearance, density, and ag-
properties compared to the other conventional materials. These ing performance (Attwood, et al. (8). On the other hand, short
composites are also finding various other applications in hos- carbon fiber–reinforced polyetheretherketone (PEEK) compos-
tile environments where these are subjected to external attacks ites possess reasonably high tensile modulus and strength with
such as solid particle erosion. Examples of such applications are 30 vol% fiber and are also characterized by their high fracture
pipeline carrying sand slurries in petroleum refining, helicopter toughness. The key similarity in the above research has been that
the investigations involved the wear of polymers filled with mi-
croscale particles. With this scale of filler particles, the mechan-
Manuscript received August 25, 2009
ical behavior of fillers tends to emulate the behavior of the bulk
Manuscript accepted November 16, 2009
Review led by David Burris filler material (Li, et al. (9); Voss and Friedrich (10). The work

520
Solid Particle Erosion of Polymer Composites 521

NOMENCLATURE α = Angle of impingement, degree


θ = Erodent temperature, ◦ C
Er = Erosion rate, kg/kg θ0 = Room temperature, ◦ C
Erth = Theoretical erosion wear rate, kg/kg η = Erosion efficiency
H = Hardness, N/m2 ηnormal = Erosion efficiency with normal impact
S = Specific heat of silica sand, J/Kg K ρc = Density of composite, kg/m3
U = Impact velocity, m/s ρ = Density of erodent, kg/m3

reported here is carried out with a different combination of and tangle together when mixed. The cast of each composite was
particulate filler (Al2 O3 ) and polymer materials (glass fiber– cured under a load of about 50 kg for 24 h before it was removed
reinforced epoxy) and investigated the reasons for modification from the mold. Then this cast was postcured in the air for another
of the tribological behavior, which included the aspects that have 24 h after removing out of the mold. Specimens of suitable dimen-
not been studied yet. sion were cut using a diamond cutter for physical characterization
To obtain the desired properties from a hybrid composite and mechanical testing.
system, reinforcement and fillers both are added to the poly-
Downloaded by [Duke University Libraries] at 01:17 10 October 2013

mers. The additional improvements in mechanical and tribologi- Test Apparatus


cal properties are in many cases attained through the incorpora-
The solid particle erosion experiments were carried out as
tion of glass or carbon fiber reinforcement and through the filling
per ASTM G76 on the erosion test rig shown schematically in
of particulate matters. However, tribo properties are not intrinsic
Fig. 1. The test rig consists of an air compressor, an air drying
material properties but strongly depend upon the system in which
unit, a conveyor belt–type particle feeder, and an air–particle
material functions (Harsha and Tewari (11). So the influence of
mixing and accelerating chamber. The dried and compressed air
fillers and fibers on the tribo-behavior of composites cannot be
was then mixed with the silica sand (300–600 µm size), which was
predicted a priori and has to be tested in the laboratory. In many
fed constantly by a conveyor belt feeder into the mixing chamber
industrial applications of composites, an understanding of the tri-
and then accelerated by passing the mixture through a conver-
bological behavior is also necessary along with an understanding
gent brass nozzle of 3 mm internal diameter. The setup is capa-
of the mechanical properties.
ble of creating reproducible erosive situations for assessing ero-
Hence, the primary concern here is to study how the alumina
sion wear resistance of the composite samples. The erodent par-
particle–filled glass fiber–reinforced epoxy composites respond to
ticles impact the specimen, which can be held at different angles
the impact of erodent under different operating conditions, to as-
with respect to the direction of erodent flow using a swivel and
sess the damage due to wear and finally to determine the signif-
an adjustable sample clip. The velocity of the eroding particles
icant factor setting for minimization of erosion rate. In addition
was determined using a standard double disc method (Ruff and
to a comprehensive review of the status of research in the field
Ives (12). The apparatus was equipped with a heater, which reg-
of solid particle erosion of polymer composites, it also aims to
ulates and maintains the erodent temperature at any predeter-
compare the theoretical wear rates with the values obtained from
mined fixed value during an erosion trial. In the present study,
experiments carried out as per Taguchi’s design.
dry silica sand (assumed to be square pyramidal shaped) of differ-
ent particle sizes (300, 450, and 600 µm) was used as erodent. The
EXPERIMENTAL DETAILS samples were cleaned in acetone, dried, and weighed to an accu-
Composite Fabrication racy of ±0.1 mg before and after the erosion trials using a pre-
cision electronic balance. The weight loss was recorded for sub-
E-glass fibers (360 roving taken from Saint Gobian Ltd.,
sequent calculation of erosion rate. The erosion process was re-
India) are reinforced in alumina-filled Epoxy LY 556 (NICE Ltd.,
peated until the erosion rate attained a constant value, called the
Kolkata, India), which was used as the matrix material. Its com-
steady-state erosion rate. The surfaces of the specimens were ex-
mon name is bisphenol-A-diglycidyl-ether and it chemically be-
amined directly by a (SEM) scanning electron microscope JEOL
longs to the epoxide family. The epoxy resin and the hardener
JSM-6480LV Japan. The eroded samples were mounted on stubs
were supplied by Hindustan Ciba Geigy Mumbai, India Ltd. E-
with silver paste. To enhance the conductivity of the eroded sam-
glass fiber and epoxy resins have modulus of 72.5 and 3.42 GPa,
ples, a thin film of platinum was vacuum-evaporated onto them
respectively, and possess density of 2590 and 1100 kg/m3 , respec-
before the photomicrographs were taken.
tively. The filler material alumina (density 3.89 g/cm3 ) was pro-
vided by NICE Ltd., Kolkata, India and was sieved to obtain par-
ticle size in the range 70–90 µm. Composites of three different Experimental Design
compositions (0, 10, and 20 wt% alumina filling) were made and Taguchi experimental design is a powerful analysis tool for
the fiber loading (weight fraction of glass fiber in the composite) modeling and analyzing the influence of control factors on per-
was kept at 50% for all the samples. The low-temperature-curing formance output. This method achieves the integration of design
epoxy resin and corresponding hardener (HY951) were mixed in of experiments (DOE) with the parametric optimization of the
a ratio of 10:1 by weight as recommended. The mix was stirred process yielding the desired results. The orthogonal array (OA)
manually to disperse the fibers in the matrix. Care was taken to requires a set of well-balanced (minimum experimental runs) ex-
ensure a uniform sample because fibers have a tendency to clump periments. Taguchi’s method uses a statistical measure of perfor-
522 S. BISWAS AND A. SATAPATHY
Downloaded by [Duke University Libraries] at 01:17 10 October 2013

Fig. 1—A schematic diagram of an erosion test rig.

mance called signal-to-noise ratio (S/N), which is a logarithmic review on erosion behavior of polymer composites reveals that
function of desired output to serve as objective functions for op- factors such as impact velocity, filler content, erodent temper-
timization. The S/N ratio takes both the mean and the variabil- ature, impingement angle, stand-off distance, erodent size, etc.,
ity into account. It is defined as the ratio of the mean (signal) to largely influence the erosion rate of polymer composites (Pat-
the standard deviation (noise). The ratio depends on the quality naik, et al. (14), (15). Hence, in the present work the impact of
characteristics of the product/process to be optimized. The three these six parameters is studied using an L27 (313 ) orthogonal de-
categories of S/N ratios are used: lower-the-better (LB), higher- sign. In a conventional full-factorial experiment design, it would
the-better (HB), and nominal-the-best (NB). The experimental require 36 = 729 runs to study six factors each at three levels,
observations are transformed into an S/N ratio. The S/N ratio for whereas Taguchi’s factorial experiment approach reduces it to
minimum erosion rate coming under the LB characteristic can be only 27 runs, offering a great advantage in terms of experimen-
calculated as logarithmic transformation of the loss function as tal time and cost. The fixed and variable parameters chosen for
shown below: the test are given in Table 1. The selected levels of the six control
S 1  2  parameters are listed in Table 2.
= − 10 log y [1]
N n The plan of the experiments is as follows: the 1st column is
where n is the number of observations, and y is the observed data. assigned to impact velocity (A), the 2nd column to alumina con-
The standard linear graph, as shown in Fig. 2, is used to assign tent (B), the 5th column to erodent temperature (C), the 9th col-
the factors and interactions to various columns of the orthogonal umn to impingement angle (D), the 10th column to stand-off dis-
array (Glen (13)). tance (E), and the 12th column to erodent size (F); the 3rd and
The most important stage in the design of experiment lies
in the selection of the control factors. An exhaustive literature
TABLE 1—PARAMETERS SETTING
B(2) Control Factors Symbols Fixed Parameters

Velocity of impact Factor A Erodent Silica sand


Fiber loading Factor B Erodent feed rate 10.0 ±1.0
(3,4) (6,7) (g/min)
Erodent temperature Factor C Nozzle diameter 3
D(9) E(10) F(12) (13) (mm)
Impingement angle Factor D Length of nozzle 80
A(1) (8,11) C(5) (mm)
Stand-off distance Factor E
Erodent size Factor F
Fig. 2—Linear graphs for L27 array.
Solid Particle Erosion of Polymer Composites 523

TABLE 2—LEVELS FOR VARIOUS CONTROL FACTORS in the composites increases with the filler content. But when this
Level glass-reinforced epoxy is filled with microsized red mud particles,
the density of the resulting hybrid composites assume higher val-
Control factor I II III Units
ues; that is, 1.65 g/cm3 (void fraction of 3.225%) and 1.752 g/cm3
A: Velocity of impact 43 54 65 m/s (void fraction of 7.894%) for composites with red mud contents
B: Filler content 0 10 20 % of 10 and 20 wt%, respectively. Consequently, the void fraction
C: Erodent temperature 40 50 60 ◦C
in the composites increases with the filler content (Biswas and
D: Impingement angle 30 60 90 ◦
Satapathy (16).
E: Stand-off distance 65 75 85 mm
Table 3 presents the tensile and flexural strengths of the com-
F: Erodent size 300 450 600 µm
posites with and without alumina filling. Similarly, Table 3 illus-
trates the variation of tensile modulus with the weight fraction of
alumina content in the composites. It can be seen that the tensile
4th column are assigned to (A × B)1 and (A × B)2 , respectively, properties have become distinctly poorer with the incorporation
to estimate interaction between impact velocity (A) and alumina of alumina particles in the matrix. But the same author reported
content (B); the 6th and 7th column are assigned to (B × C)1 for other filler material like red mud–filled glass fiber–reinforced
and (B × C)2 , respectively, to estimate interaction between alu-
Downloaded by [Duke University Libraries] at 01:17 10 October 2013

epoxy composite shows better tensile strength, flexural strength,


mina content (B) and erodent temperature (C); the 8th and 11th and tensile modulus (Biswas and Satapathy (16)). This is possi-
column are assigned to (A × C)1 and (A × C)2 , respectively, to bly not occurring better in tensile strength, flexural strength and
estimate interaction between the impact velocity (A) and erodent tensile modulus in the present study with the presence of alumina
temperature (C); and the remaining columns are used to estimate particle as the filler material in the epoxy composites instead of
experimental errors. red mud particulates. This might have influenced some other me-
chanical properties of the composites, such as inter laminar shear
RESULTS AND DISCUSSION
strength (ILSS) or impact strength. This phenomenon might re-
Mechanical Properties sult from the comparatively poorer dispersion of microsized par-
The present investigation reveals that the presence of alumina ticles (alumina filler) and a greater possibility of the existence of
has varied effect on the glass–epoxy composites in terms of me- voids in composites. The decreases in the ultimate mechanical
chanical properties. As is seen in Table 3, the composite micro- strengths of the composites are probably caused by an incompat-
hardness has significantly improved with the addition of alumina. ibility of the alumina particles and the epoxy matrix, leading to
The density has also increased with alumina content. The neat poor interfacial bonding. However, it also depends on other fac-
epoxy taken for this study possesses a density of 1.1 g/cm3 , which tors such as the size and shape of the filler in the composites.
increases to 1.53 g/cm3 (with a void fraction of 0.9%) with the A short beam shear test was carried out on these laminates to
reinforcement of 50 wt% of glass fiber in it. But when this glass determine the ILSS of composites prepared with different filler
fiber–reinforced epoxy is filled with microsized alumina particles, content. Apparent ILSS of glass fiber–reinforced epoxy resin is
the density of the resulting hybrid composites assumes higher val- measured as 99.75 MPa. With the addition of alumina, ILSS of
ues; that is, 1.62 g/cm3 (void fraction of 5.241%) and 1.80 g/cm3 laminates decreases drastically, as is seen in the case of the com-
(void fraction of 6.880%) for composites with alumina contents posite with filler up to 20 wt% (Table 3). Similarly, with the ad-
of 10 and 20 wt%, respectively. Consequently, the void fraction dition of red mud, ILSS of laminates decreases drastically, as is

TABLE 3—MECHANICAL AND PHYSICAL PROPERTIES OF THE COMPOSITES

Tensile Tensile Flexural Impact Measured Theoretical


Hardness Strength Modulus Strength ILSS Energy Density Density
Volume Fraction
Composites (Hv) (MPa) (GPa) (MPa) (MPa) (J) (g/cm3 ) (g/cm3 ) of Voids (%)

Epoxy + 50 wt% glass 24.8 516 8.77 393.1 99.75 0.976 1.53 1.544 0.906
fiber
Epoxy + 50 wt% glass 37.05 494 8.24 275.9 59.44 1.301 1.65 1.705 3.225
fiber + 10 wt% Red
mud
Epoxy + 50 wt% glass 43.05 468 8.32 401.81 83.97 0.958 1.752 1.9 7.894
fiber + 20 wt% Red
mud
Epoxy + 50 wt% glass 39.92 427 7.57 378.02 77.86 1.198 1.627 1.717 5.241
fiber + 10 wt% alumina
Epoxy + 50 wt% glass 47.46 385 6.95 258.19 57.83 1.311 1.8 1.933 6.88
fiber + 20 wt% alumina
524 S. BISWAS AND A. SATAPATHY

seen in case of the composite with filler up to 10 wt% (Table 3). upper surface. The distribution of filler as seen in the micrograph
The reduction may be related with the formation of voids in the is reasonably uniform, although at places the particles are seen
matrix, which are generally located at the interlaminar region of to have formed small and large clusters. The SEM image of the
composites. However, for the composite with 20 wt% filler, a rel- eroded surface at impingement angle of 30◦ is seen in Fig. 4(b).
atively higher value of ILSS has been recorded. Compared with The erodent particles at an elevated temperature of 60◦ C hit the
red mud particulates, alumina particulates show better ILSS up to composite surface with an impact speed of 43 m/s. As expected,
10 wt%, after which on further addition of both particulates with due to the parallel hits of the erodent at a low angle, not much
the same composites up to 20 wt% red mud–filled composites fiber cracking is seen to have occurred. Because of the small nor-
show better ILSS (Table 3). For both flexural and interlaminar mal impact component of the erodent, the fibers resist cracking.
shear strength the test is repeated six times and the mean value is As seen in Figs. 4(c) and 4(d) a large number of cracks is not no-
reported. ticed on the fibers (vertical to their length). On the other hand,
In case impact tests the machine is adjusted such that the blade there is a matrix removal between the fibers, which weakens the
on the free-hanging pendulum just barely contacts the specimen fiber-supporting mechanism. Insufficiently supported fibers are
(zero position). Because there are practically no losses due to broken and this may lead to displacement of the fibers from their
bearing friction, etc. (<0.3%), the testing conditions may be re- original positions.
Downloaded by [Duke University Libraries] at 01:17 10 October 2013

garded as ideal. The specimens are clamped in a square support Figures 5(a)–5(d) show SEM micrographs of alumina-filled
and are struck at their central point by a hemispherical bolt of glass fiber–reinforced epoxy matrix composites eroded at a 60◦
diameter 5 mm. The respective values of impact energy of dif- impingement angle where maximum wear rate is recorded. SEM
ferent specimens are recorded directly from the dial indicator. investigations reported previously reveal that the erosion of the
Impact strength is the ability of a material to resist the fracture polymer composites in general and thermoset composites in
under stress applied at high speed. The impact properties of com- particular is a complex process involving matrix microcracking,
posite materials are directly related to their overall toughness. fiber matrix debonding, fiber breakage, and material removal
Composite fracture toughness is affected by interlaminar and in- (Friedrich (17); Pool, et al. (18); Hager, et al. (19); Zahavi and
terfacial strength parameters. As can be seen from Table 3, the Schmitt ((20)). The observed anisotropy of the erosion of these
impact strength of the composites is significantly improved by the materials can be attributed to the following mechanisms. It is well
addition of alumina and red mud particulate compared with un- known that the fibers in composites, subjected to particle flow,
filled glass epoxy composite. The increased filler content (up to break in bending (Hager, et al. (19)). In case of an impact hav-
approximately 20 wt.) leads to a higher impact strength due to the ing a parallel component of velocity with respect to the fiber ori-
interfacial reaction and provides an effective barrier for pinning entation (Figs. 5(c) and 5(d)), bending requires particle inden-
and bifurcation of the advancing cracks. The mechanical prop- tation into the composite. The indentation involves compressive
erty tests were repeated six times and the mean value is reported stresses, and resistance to microbending is very high. Thus, there
in this study. is local removal of resin material from the impacted surface; this
results in the exposure of the fibers. For transverse particle im-
Surface Morphology pact (Figs. 5(a) and 5(b)), the resistance to lateral component of
Figure 3 presents the scanning electron micrographs (SEMs) the bending moment is lower and bundles of fibers are bent and
of the unfilled glass–epoxy composite surfaces eroded under var- break easily. This caused more erosive wear in both composites.
ious test conditions. In Fig. 3(a) no cracks or craters are seen on Also, in case of transverse erosion, high interfacial tensile stresses
the composite surface after erosion due to impact of dry silica are generated by particle impacts. This causes intensive debond-
sand particles (temperature 40◦ C) of smallest grit size (300 µm) ing and breakage of the fibers, which are not supported by the
with a lower impact velocity (43 m/s) at a low impingement angle matrix. The continuous impact of sand particles on the compos-
of 30◦ . But as the erosion tests were carried out with higher ero- ite surface breaks the fibers because of the formation of cracks
dent temperature (60◦ C) and grit size (450 µm), the morphology along their length. The bending of fibers is possible because the
of the eroded surface became different as in Fig. 3(b). The matrix surrounding matrix and supporting fibers have been removed.
was chipped off and the glass fibers are clearly visible beneath As the impact angle of the particles shifts toward larger
the matrix layer. The micrographs with higher magnifications pre- values, the effects of normal force become dominant. Figures
sented in Figs. 3(c), 3(d), and 3(e) distinctly illustrate the craters 6(a)–6(d) illustrate the wear morphology of the material after
formed due to material loss and the arrays of broken and semi- normal erosion (at 90◦ ) at high particle speed of 65 m/s. At this
broken glass fibers within. Due to repeated impact of hard and impingement angle there is no parallel component of particle
high-temperature sand particles there is initiation of cracks on the speed and parallel force. Hence, no abrasive wear is expected.
fibers and as erosion progresses, these cracks subsequently prop- Each erodent particle has a totally normal speed and exerts
agate on the fiber bodies both in a transverse as well as a longi- totally low-energy repeated impact loading for the material
tudinal manner. Such cracks are clearly noticed in Figs. 3(f), 3(g), surface. Many fibers are cracked into small fragments and they
and 3(h), respectively. It is also evident from the microstructures are removed from their places partly with the surrounding
of these eroded surfaces that for higher erodent temperature and matrix like spalled fragments. When comparing two surface mor-
impact velocity, the damage to the surface is relatively greater. phologies, it is clear that the wear morphology in Fig. 6 is quite
Figure 4(a) shows a portion of the composite surface before different compared to Fig. 5(d). On the other hand, the effects of
erosion occurred. Scattered alumina particles are observed on the the particle impacts on the subsurface of the composites should
Solid Particle Erosion of Polymer Composites 525
Downloaded by [Duke University Libraries] at 01:17 10 October 2013

Fig. 3—SEM micrographs of the eroded surfaces of the unfilled glass–epoxy composites.

be noted. At acute angles, due to the smaller normal constituent quently occur. Some of these features are reflected in the micro-
of the particle impingement, the effect of impact phenomena is graph (Fig. 7(d)). The fibers are broken by means of shearing ac-
smaller. Greater part of the particle’s energy spent for removing tion, which can be seen from the micrograph due to impingement
the matrix between the fibers and fracture them to limited extent. of particles at higher impact velocity (Fig. 7(d)).
The eroded surfaces of glass–epoxy composites with alumina
filling (20 wt%) subjected to normal impingement angle at high Steady-State Erosion
impact velocity are shown in Figs. 7(a)–7(d). The worn surfaces The erosion wear rates of glass–epoxy–alumina composites as
exhibit signs of plastic deformation in the matrix regime, which a function of impingement angle are shown in Fig. 8. Steady-state
indicates the initiation of surface damage (Figs. 7(a), (b)). Subse- wear rates are reached during the erosion periods for each
quently, there is removal of matrix material from the surface, re- impingement angle. No incubation period is observed at any
sulting in exposure of fibers to an erosive environment, which can impingement angle or particle velocity. It is observed that the
be clearly seen (Fig. 7(c)). In this micrograph, the fibers are still rate of mass loss of the eroded samples vary with the erosion
held firmly in place by the matrix surrounding them. Repeated time at each impingement angle (15, 30, 45, 60, 75, and 90◦ ). As is
impacts gradually form larger craters and fiber matrix debond- well known, the impingement angle has a great influence on the
ing, brittle fracture of matrix, and pulverization of fibers conse- particle erosion. In order to monitor the effect of impingement
526 S. BISWAS AND A. SATAPATHY
Downloaded by [Duke University Libraries] at 01:17 10 October 2013

Fig. 4—Scanning electron micrograph of alumina-filled glass fiber–epoxy matrix composite eroded surfaces (impact velocity 43 m/s, impingement angle
30◦ , filler content 10 wt%, erodent temperature 60◦ C, erodent size 450 µm, and stand of distance 65 mm.)

angles, experimental results are illustrated in Fig. 8. The erosion between the target material and the trajectory of the erodents.
wear behavior of polymer composites can be grouped into ductile Dependence of erosion rate on the impact angle is largely deter-
and brittle categories, although this grouping is not definitive mined by the nature of target material. In the present study, the
because the erosion characteristics depend on the experimental variation of erosion wear rate of the composites with impinge-
conditions as much as on composition of the target material. It is ment angle is studied by conducting experiments under specified
well known that for ductile materials the peak erosion normally operating conditions. The results are presented in Fig. 8, which
occurs at 15–20◦ angle, whereas for brittle materials the erosion shows the peak erosion taking place at an impingement angle
damage is maximum usually at normal impact; that is, at 90◦ of 60◦ for the unfilled as well as the alumina-filled glass–epoxy
impingement angle. Impingement angle is defined as the angle composites. This clearly indicates that these composites respond

Fig. 5—Scanning electron micrograph of alumina-filled glass fiber–epoxy matrix composite surfaces erodent at impact velocity 43 m/s, impingement
angle 60◦ , filler content 20 wt%, erodent temperature 60◦ C, erodent size 300 µm, and S.O.D 75 mm.
Solid Particle Erosion of Polymer Composites 527
Downloaded by [Duke University Libraries] at 01:17 10 October 2013

Fig. 6—Scanning electron micrograph of alumina-filled glass fiber–epoxy matrix composite surfaces erodent at impact velocity 54 m/s, impingement
angle 90◦ , filler content 20 wt%, erodent temperature 50◦ C, erodent size 300 µm, and S.O.D 85 mm.

to solid particle impact neither in a purely ductile nor in a purely was made using the popular software specifically used for design
brittle manner. This behavior can be termed as semiductile in of experiment applications known as MINITAB 14. Before any
nature, which may be attributed to the incorporation of glass attempt is made to use this simple model as a predictor for the
fibers and alumina particles within the epoxy body. measure of performance, the possible interactions between the
control factors must be considered. Thus, factorial design incor-
Analysis of Experimental Results porates a simple means of testing for the presence of the inter-
In Table 4, the erosion rates of different composites for all 27 action effects. Analysis of the result leads to the conclusion that
test runs and their corresponding S/N ratios are given. Each value a factor combination of A3 , B2 , C1 , D3 , E1 , and F1 gives a min-
is in fact the average of two replications. The overall mean for the imum erosion rate (Fig. 9). The interaction graphs are shown in
S/N ratio of the erosion rate is found to be −47.47 db. The analysis Figs. 10(a)–10(c).

Fig. 7—Scanning electron micrograph of alumina-filled glass fiber–epoxy matrix composite surfaces erodent at impact velocity 65 m/s, impingement
angle 90◦ , filler content 20 wt%, erodent temperature 50◦ C, erodent size 450 µm, and S.O.D 75 mm.
528 S. BISWAS AND A. SATAPATHY

275 defined by Sundararajan, et al. (21) is given as


0 wt% Alumina
2ErHv
250 10 wt% Alumina ηnormal = [2]
ρ V2
Erosion rate (mg/kg)

20 wt% Alumina
But considering impact of erodent at any angle α to the surface,
225
the actual erosion efficiency can be obtained by modifying Eq. [2]
200 as
2ErH
η= [3]
175 ρ U2 Sin2 α
According to the theoretical erosion wear model proposed by
150 Biswas and Satapathy (16), the nondimensional erosion wear rate
15 30 45 60 75 90 of a composite material is given by
Impingement angle (degree) η.ρC  2 2 
Er = U Sin α + 2.S(θ − θ0 ) [4]
3H
Fig. 8—Effect of impingement angle on the erosion wear rate of the com-
The mathematical expression in Eq. [4] can be used for predictive
Downloaded by [Duke University Libraries] at 01:17 10 October 2013

posites.
purpose to make an approximate assessment of the erosion dam-
age from the composite surface. When the erodent temperature
is same as room temperature, Eq. [4] reduces to:
Erosion Efficiency η.ρc  2 2 
Er = U Sin α [5]
The hardness alone is unable to provide sufficient correlation 3H
with erosion rate, largely because it determines only the volume Here in Eq. [5], the role of thermal energy transfer from erodent
displaced by each impact and not really the volume eroded. Thus, to target material in causing erosion is absent and thus the ex-
a parameter that will reflect the efficiency with which the volume pression is similar to the one in the theoretical model proposed
that is displaced is removed should be combined with hardness earlier by Patnaik, et al. (15).
to obtain a better correlation. The erosion efficiency is obviously The evaluation of erosion efficiency can be made only on the
one such parameter. In case of a stream of particles impacting a basis of experimental data. Hence, the values of erosion efficien-
surface normally (i.e., at α = 90◦ ), the erosion efficiency (ηnormal ) cies of these composites calculated using Eq. [3] are summarized

TABLE 4—EXPERIMENTAL DESIGN USING L27 ORTHOGONAL ARRAY

Sl. No. A (m/s) B (%) C (◦ C) D (◦ ) E (mm) F (µm) Er1 (mg/kg) Er2 (mg/kg) Er = ( Er1 + Er2 )/2 (mg/kg) S/N Ratio (db)

1 43 0 30 30 65 300 202.3478 206.3482 204.348 −46.207


2 43 0 50 60 75 450 338.029 346.029 342.029 −50.681
3 43 0 70 90 85 600 410.7203 416.7197 413.720 −52.334
4 43 10 30 60 75 600 219.2472 221.3328 220.290 −46.86
5 43 10 50 90 85 300 173.913 192.753 183.333 −45.265
6 43 10 70 30 65 450 184.058 210.210 197.134 −45.895
7 43 20 30 90 85 450 263.768 315.942 289.855 −49.244
8 43 20 50 30 65 600 179.1923 168.6337 173.913 −44.807
9 43 20 70 60 75 300 192.3188 223.4632 207.891 −46.357
10 54 0 30 60 85 450 223.6232 228.5508 226.087 −47.086
11 54 0 50 90 65 600 342.1473 365.0987 353.623 −50.971
12 54 0 70 30 75 300 376.087 388.207 382.147 −51.645
13 54 10 30 90 65 300 147.2464 154.2036 150.725 −43.564
14 54 10 50 30 75 450 286.5217 293.1883 289.855 −49.244
15 54 10 70 60 85 600 300.1972 302.1208 301.159 −49.576
16 54 20 30 30 75 600 218.3712 233.8028 226.087 −47.086
17 54 20 50 60 85 300 374.4928 379.1312 376.812 −51.523
18 54 20 70 90 65 450 274.3478 267.7582 271.053 −48.661
19 65 0 30 90 75 600 163.7123 163.8237 163.768 −44.285
20 65 0 50 30 85 300 363.7681 355.0719 359.420 −51.112
21 65 0 70 60 65 450 449.4203 438.0037 443.712 −52.942
22 65 10 30 30 85 450 197.1014 208.6966 202.899 −46.146
23 65 10 50 60 65 600 158.8903 164.5877 161.739 −44.176
24 65 10 70 90 75 300 136.7391 149.2069 142.973 −43.105
25 65 20 30 60 65 300 142.8986 146.9574 144.928 −43.223
26 65 20 50 90 75 450 284.9275 283.1885 284.058 −49.068
27 65 20 70 30 85 600 287.8922 298.4238 293.158 −49.342
Solid Particle Erosion of Polymer Composites 529

Main Effects Plot (data means) for SN ratios


A B C
-46

-47

-48
Mean of SN ratios
-49

-50
43 54 65 0 10 20 30 50 70
D E F
-46

-47

-48

-49
Downloaded by [Duke University Libraries] at 01:17 10 October 2013

-50
30 60 90 65 75 85 300 450 600

Signal-to-noise: Smaller is better

Fig. 9—Effect of control factors on erosion rate.

in Table 5 along with their hardness values and operating condi- efficiencies of these composites under normal impact (ηnormal )
tions. This clearly shows that erosion efficiency is not exclusively vary from 3 to 6%, 6 to 9%, and 9 to 12% for impact velocities
a material property but also depends on other operational vari- 65, 54, and 43 m/s, respectively. The value of η for a particular
ables such as impingement angle and impact velocity. The erosion impact velocity under oblique impact can be obtained simply by

TABLE 5—EROSION EFFICIENCY OF GLASS FIBER–REINFORCED ALUMINA–EPOXY RESIN COMPOSITE


Expt. Impact Velocity Density of Eroding Hardness of Eroding Erosion Rate Erosion
No. (V) m/s Material (ρ) kg/m3 Material (Hv) MPa (Er) mg/kg Efficiency (η)

1 43 1530 24.80 204.348 14.05465


2 43 1530 24.80 342.029 7.841816
3 43 1530 24.80 413.720 7.113708
4 43 1627 39.92 220.290 7.645242
5 43 1627 39.92 183.333 4.771697
6 43 1627 39.92 197.134 20.52361
7 43 1800 47.46 289.855 8.107095
8 43 1800 47.46 173.913 19.45703
9 43 1800 47.46 207.891 7.753261
10 54 1530 24.80 226.087 3.286841
11 54 1530 24.80 353.623 3.855488
12 54 1530 24.80 382.147 16.66592
13 54 1627 39.92 150.725 2.487522
14 54 1627 39.92 289.855 19.13473
15 54 1627 39.92 301.159 6.627377
16 54 1800 47.46 226.087 16.03871
17 54 1800 47.46 376.812 8.910926
18 54 1800 47.46 271.053 4.807155
19 65 1530 24.80 163.768 1.232335
20 65 1530 24.80 359.420 10.81837
21 65 1530 24.80 443.712 4.452102
22 65 1627 39.92 202.899 6.419776
23 65 1627 39.92 161.739 2.456526
24 65 1627 39.92 142.973 1.628533
25 65 1800 47.46 144.928 2.365436
26 65 1800 47.46 284.058 3.476976
27 65 1800 47.46 293.158 14.35345
530 S. BISWAS AND A. SATAPATHY

Interaction Plot (data means) for SN ratios


-44 B
0
10
-45 20

-46

SN ratios
-47

-48

-49

-50
43 54 65
A
Downloaded by [Duke University Libraries] at 01:17 10 October 2013

Signal-to-noise: Smaller is better

(a)

Interaction Plot (data means) for SN ratios


-44 C
30
50
-45
70

-46
SN ratios

-47

-48

-49

-50

-51
43 54 65
A
Signal-to-noise: Smaller is better

(b)

Interaction Plot (data means) for SN ratios


-45 C
30
-46 50
70
-47

-48
SN ratios

-49

-50

-51

-52

-53
0 10 20
B
Signal-to-noise: Smaller is better

(c)

Fig. 10a—Interaction graph between A × B for erosion rate (b). Interaction graph between A × C for erosion rate (c). Interaction graph between B × C for
erosion rate.
Solid Particle Erosion of Polymer Composites 531

multiplying a factor 1/Sin2 α with ηnormal . A similar observation on TABLE 6—COMPARISON OF THEORETICAL AND EXPERI-
velocity dependence of erosion efficiency has previously been re- MENTAL RESULTS

ported by investigators (Roy, et al. (22). Expt. No. Erth (mg/kg) Erexpt. (mg/kg) Error (%)
The theoretical erosion wear rate (Erth ) of the alumina-filled
1 231.712 204.348 11.8096
glass–epoxy composites is calculated using Eq. [4]. These val-
2 387.851 342.029 11.8144
ues are compared with those obtained from experiments (Erexpt )
3 479.506 413.720 13.7195
conducted under similar operating conditions. Table 6 presents a 4 198.058 220.290 −11.2248
comparison among the theoretical and experimental results and 5 168.967 183.333 −8.5016
the corresponding associated error percentage. The errors in ex- 6 185.092 197.134 −6.5059
perimental results with respect to the theoretical ones lie in the 7 296.660 289.855 2.2939
range 0–10%. The magnitude of η can be used to characterize 8 168.416 173.913 −3.2634
the nature and mechanism of erosion. For example, ideal mi- 9 183.665 207.891 −13.1903
croploughing involving just the displacement of the material from 10 206.075 226.087 −9.7109
the crater without any fracture (and hence no erosion) will result 11 333.054 353.623 −6.1757
12 378.521 382.147 −0.9579
in η = 0. In contrast, if the material removal is by ideal microcut-
13 135.548 150.725 −11.1966
Downloaded by [Duke University Libraries] at 01:17 10 October 2013

ting, η = 1.0 or 100%. If erosion occurs by lip or platelet forma-


14 277.781 289.855 −4.3465
tion and their fracture by repeated impact, as is usually the case
15 339.767 301.159 11.3632
with ductile materials, the magnitude of η will be very low; that 16 208.071 226.087 −8.6583
is, η ≤ 100%. In the case of brittle materials, erosion occurs usu- 17 380.134 376.812 0.87398
ally by spalling and removal of large chunks of materials resulting 18 247.356 271.053 −9.5801
from the interlinking of lateral or radial cracks and thus η can be 19 160.040 163.768 −2.3292
expected to be even greater than 100% (Arjula and Harsha (23). 20 327.574 359.420 −9.7215
The erosion efficiencies of the composites under the present study 21 395.002 443.712 −12.331
indicate that at low impact speed the erosion response is semiduc- 22 179.336 202.899 −13.138
tile (η = 10–100%). On the other hand, at relatively higher impact 23 186.329 161.739 13.1971
24 153.003 142.973 6.5553
velocity the composites exhibit ductile (η < 10%) erosion behav-
25 132.226 144.928 −9.6057
ior (Roy, et al. (22).
26 268.441 284.058 −5.8174
27 304.157 293.158 3.6163
ANOVA and the Effects of Factors
In order to determine the statistical significance of various fac-
tors like impact velocity (A), alumina content (B), erodent tem- has least effect on performance characteristics. Factors A and C
perature (C), impingement angle (D), stand-off distance (E), and have a significant effect on erosion rate but their interaction has
erodent size (F) on erosion rate, analysis of variance (ANOVA) the least effect on erosion rate (Table 7). Therefore, factor F and
was performed on experimental data. Table 7 shows the results interaction A × C can be omitted for further prediction. The es-
of the ANOVA with the erosion rate. This analysis is undertaken timated S/N ratio for erosion rate can be calculated with the help
for a level of confidence of significance of 5%. The last column of of following prediction equation:
the table indicates that the main effects are highly significant (all
η1 = T + (A2 − T) + (B3 − T) + [(A2 B3
have very small p-values; Montgomery (24).
From Table 7, one can observe that alumina content (p = − T) − (A2 − T) − (B3 − T)] + (C2 − T) + [(B3 C2 − T)
0.044), erodent temperature (p = 0.059), stand-off distance (p =  
− (B3 − T) − (C2 − T)] + (E3 − T ) + F 2 − T [6]
0.102), erodent size (p = 0.152), and impact velocity (p = 0.159)
have a great influence on erosion rate. The interaction of Alu- where is predicted average, T is overall experimental average,
mina Content × Erodent Temperature (p = 0.163) and Impact and A2 , B3 , C2 , E3 and F 2 are mean response for factors and in-
Velocity × Erodent Temperature (p = 0.205) show significant teractions at designated levels.
contribution on the erosion rate but the remaining factor and in- By combining like terms, the equation reduces to
teractions make a relatively less significant contribution to ero-
sion rate. η1 = A2 B3 + B3 C2 − B2 + E3 + F 2 − 2T [7]

A new combination of factor levels A2 , B3 , C2 , E3 , and F2 is used


Confirmation Experiment to predict deposition rate through the prediction equation and it
The optimal combination of control factors has been deter- is found to be η̄1 = − 52.7645 dB.
mined in the previous analysis. However, the final step in any For each performance measure, an experiment is conducted
design of experiment approach is to predict and verify improve- for a different factor combination and compared with the result
ments in observed values through the use of the optimal com- obtained from the predictive equation as shown in Table 8. The
bination level of control factors. The confirmation experiment resulting model seems to be capable of predicting erosion rate
is performed by taking an arbitrary set of factor combination to a reasonable accuracy. An error of 6.63% for the S/N ratio
A2 B3 C2 E3 F2 , but factor D has been omitted and factor A also of erosion rate is observed. However, the error can be further
532 S. BISWAS AND A. SATAPATHY

TABLE 7—ANOVA TABLE FOR EROSION RATE (ALUMINA- impingement angle for all composite samples under various
FILLED COMPOSITES) experimental conditions.
Source DF Seq SS Adj SS Adj MS F P Rank
6. The confirmation tests showed that the error associated be-
tween experimental results and the predictive model for ero-
A 2 15.170 15.170 7.585 5.27 0.159 5 sion rate is 6.63%. However, the error can be further reduced
B 2 62.209 62.209 31.104 21.62 0.044 1 if the number of experiments is increased.
C 2 45.522 45.522 22.761 15.82 0.059 2
D 2 2.255 2.255 1.127 0.78 0.561 6 REFERENCES
E 2 25.464 25.464 12.732 8.85 0.102 3
(1) Tewari, U. S., Harsha, A. P., Hager, A. M., and Friedrich, K. (2003), “Solid
F 2 16.073 16.073 8.036 5.59 0.152 4 Particle Erosion of Carbon Fibre and Glass Fibre Epoxy Composites,”
A*B 4 7.449 7.449 1.862 1.29 0.480 3 Composites Science and Technology, 63, pp 549–557.
A*C 4 23.668 23.668 5.917 4.11 0.205 2 (2) Tewari, U. S., Harsha, A. P., Hager, A. M., and Friedrich, K. (2002),
B*C 4 31.027 31.027 7.757 5.39 0.163 1 “Solid Particle Erosion of Unidirectional Carbon Fibre Reinforced
Polyetheretherketone Composites,” Wear, 252, pp 992–1000.
Error 2 2.877 2.877 1.438 (3) Barkoula, N. M. and Karger-Kocsis, J. (2002), “Review—Processes and
Total 26 231.713 Influencing Parameters of the Solid Particle Erosion of Polymers and Their
Composites,” Journal of Materials Science, 37, pp 3807–3820.
(4) Harsha, A. P., Tewari, U. S., and Venkataraman, B. (2003), “Solid Particle
Downloaded by [Duke University Libraries] at 01:17 10 October 2013

TABLE 8—RESULTS OF THE CONFIRMATION EXPERIMENTS FOR Erosion Behaviour of Various Polyaryletherketone Composites,” Wear,
254, pp 693–712.
EROSION RATE
(5) Biswas, S., Satapathy, A., and Patnaik, A. (2008), “Erosion Wear Be-
Optimal Control Parameters haviour of Polymer Composites: A Review,” Journal of Reinforced Plastics
and Composites. doi: 10.1177/0731684408097786.
Prediction Experimental (6) Rothon, R. N. (1997), “Mineral Fillers in Thermoplastics: Filler Manufac-
ture,” Journal of Adhesion, 64, pp 87–109.
Level A2 B3 C2 E3 F2 A2 B3 C2 E3 F2 (7) Bonner, W. H. (1962), “Aromatic polyketones and preparation thereof,”
S/N ratio for erosion rate (db) −52.7645 −49.2661 U.S. Patent Specification 3065205.
(8) Attwood, T. E., Dawson, P. C., Freeman, J. L., Hoy, L. R. J., Rose, J. B.,
and Staniland, P. A. (1981), “Synthesis and Properties of Polyaryletherke-
tones,” Polymer, 22(8), pp 1096–1103.
(9) Li, T. Q., Zhang, M. Q., Zeng, K., and Zeng, H. M. (2000), “The Depen-
reduced if the number of measurements is increased. This vali- dence of the Fracture Toughness of Thermoplastic Composite Laminates
dates the development of the mathematical model for predicting on Interfacial Interaction,” Composites Science and Technology, 60(3),
the measures of performance based on knowledge of the input pp 465–476.
(10) Voss, H. and Friedrich, K. (1986), “Wear Performance of Bulk Liquid
parameters. Crystal Polymer and Its Short Fibre Composites,” Tribology International,
19, pp 145–156.
CONCLUSIONS (11) Harsha, A. P. and Tewari, U. S. (2002), “Tribo Performance of Pol-
yaryletherketone Composites,” Polymer Testing, 21, pp 697–709.
Based on the research presented in this article the following (12) Ruff, A. W. and Ives, L. K. (1975), “Measurement of Solid Particle Veloc-
conclusions are drawn: ity in Erosive Wear,” Wear, 35(1), pp 195–199.
(13) Glen, S. P. (1993), Taguchi Methods: A Hands on Approach, Addison-
Wesley: New York.
1. These composites suitable for applications in highly erosive (14) Patnaik, A., Satapathy, A., Mahapatra, S. S., and Dash, R. R. (2008),
environments can be prepared by reinforcement of glass “A Modeling Approach for Prediction of Erosion Behavior of Glass
Fiber-Polyester Composites,” Journal of Polymer Research, 15(2), pp 147–
fibers and filling of microsized alumina particles in epoxy
160.
resin. (15) Patnaik, A., Satapathy, A., Mahapatra, S. S., and Dash, R. R. (2009),
2. Erosion characteristics of these composites can be successfully “Tribo-Performance of Polyester Hybrid Composites: Damage Assess-
ment and Parameter Optimization Using Taguchi Design,” Materials and
analyzed using a Taguchi experimental design scheme. The
Design, 30, pp 57–67.
erosion wear performance of these composites improves quite (16) Biswas, S. and Satapathy, A. (2009), “Tribo-Performance Analysis of Red
significantly by addition of alumina filler. Mud Filled Glass-Epoxy Composites Using Taguchi Experimental De-
sign,” Materials and Design, 30, pp 2841–2853.
3. The factors like alumina content (p = 0.044), erodent temper-
(17) Friedrich, K. (1986), “Erosive Wear of Polymer Surfaces by Steel Blast-
ature (p = 0.059), stand-off distance (p = 0.102), erodent size ing,” Journal of Material Science, 21, pp 3317–3332.
(p = 0.152), impact velocity (p = 0.159), and interactions Alu- (18) Pool, K. V., Dharan, C. K. H., and Finnie, I. (1986), “Erosive Wear of
Composite Materials,” Wear, 107, pp 1–12.
mina Content × Erodent Temperature (p = 0.163) and Impact
(19) Hager, A., Friedrich, K., Dzenis, Y. A., and Paipetis, S. A. (1995), “Study
Velocity × Erodent Temperature (p = 0.205) in order of pri- of Erosion Wear of Advanced Polymer Composites,” in ICCM-10 Con-
ority show significant contributions to the erosion rate. ference Proceedings, Street, K. (Ed.), Woodhead PublishingL Cambridge,
UK, pp 155–162.
4. The erosion efficiency (η), in general, characterizes the wear
(20) Zahavi, J. and Schmitt, G. F. (1981), “Solid Particle Erosion of Reinforced
mechanism of composites. The alumina-filled glass fiber– Composite Material,” Wear, 71, pp 179–190.
epoxy composites exhibit semiductile erosion response (η = (21) Sundararajan, G., Roy, M., and Venkataraman, B. (1990), “Erosion
Efficiency—A New Parameter to Characterize the Dominant Erosion
10–60%) for low impact velocities and ductile erosion re-
Micro-Mechanism,” Wear, 140, pp 369–381.
sponse (η < 10%) for relatively high impact velocity. (22) Roy, M., Vishwanathan, B., and Sundararajan, G. (1994) “Solid Particle
5. A study of influence of impingement angle on erosion rate Erosion of Polymer Matrix Composites,” Wear, 171, pp 149–161.
(23) Arjula, S. and Harsha, A. P. (2006), “Study of Erosion Efficiency of Poly-
of the composites filled with different weight percentage of
mers and Polymer Composites,” Polymer Testing, 25, pp 188–196.
alumina reveals their semiductile nature with respect to ero- (24) Montgomery, D. C. (2005), Design and Analysis of Experiments, 5th ed.,
sion wear. The peak erosion rate is found to occur at 60◦ John Wiley & Sons, New York.

You might also like