Sunteți pe pagina 1din 18

Article

Journal of Thermoplastic Composite


Materials
Solid particle erosion 26(6) 777–794
ª The Author(s) 2012
behaviour of glass Reprints and permissions:
sagepub.co.uk/journalsPermissions.nav

mat-based polyester DOI: 10.1177/0892705711429490


jtc.sagepub.com

laminate composite
materials
M Bagci1, H Imrek1 and A Aktas2

Abstract
In this study, solid particle erosion behaviour of glass fibre mat-based polyester laminate
composite materials was investigated by varying the impact velocity and the impingement
angle and sizes of the abrasive particles. Impact velocities used in the tests were 23, 34
and 53 m/s, while the impingement angles were 15 , 30 , 45 , 60 , 75 and 90 . Silicon
dioxide particles with average diameters of 250, 500 and 1000 mm were used as abrasive
particles. Experimental data were taken as reference and values of erosion rates were
obtained as functions of impact velocity, impingement angle and the size of the abrasive
particles. The results obtained exhibited typical ductile behaviour of glass fibre-
reinforced composites, where the maximum erosion rate was recorded at an impinge-
ment angle of 30 , and the erosion rates decreased as the impingement angle increased.
Moreover, the erosion rates increased with the increase in impact velocity and particle
size. Finally, the worn out surfaces of the test specimens were investigated on two dif-
ferent devices, optical microscope and scanning electron microscope.

Keywords
Polyester, glass mat, impact velocity, impingement angle, abrasive particles, erosion rate

1
Department of Mechanical Engineering, Faculty of Engineering and Architecture, Selcuk University, Konya,
Turkey
2
University of Istanbul, Faculty of Engineering, Department of Mechanical Engineering, 34320 Avcilar, Istanbul,
Turkey

Corresponding author:
M Bagci, Department of Mechanical Engineering, Faculty of Engineering and Architecture, Selcuk University,
Alaeddin Campus, 42075 Konya, Turkey.
Email: meh_bagci@yahoo.com

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


778 Journal of Thermoplastic Composite Materials 26(6)

Introduction
Due to their excellent properties, composites are extensively used as structural materials
in various components and engineering parts in automotive, aerospace, marine and
energy applications. Composites in pipeline carrying sand slurries in petroleum refining,
helicopter rotor blades, pump impeller blades, high-speed vehicles and aircraft operating
in desert environments are often exposed to conditions in which they may be subjected to
solid particle erosion. Their mechanical properties such as flexural strength can be
degraded by the presence of localized impact damage after particle erosion. It is also
widely recognized that composites have a poor erosion resistance against operational
requirements of industrial environment. That may be overcome by understanding the
characteristics of the composites.1–4
There has been an increasing trend, recently, towards researches on erosive wear; one
of the types of surface damages where eroding particles of different dimensional sizes
under various impact velocities and impingement angles is striking and affecting the top
surfaces of materials, leading to material losses and functional variations.5–9
Many researchers, therefore, have investigated the erosion behaviour of composites
worn out by solid particles. Patnaik et al.10 carried out erosion test to study the effects of
various operational and material parameters on erosive wear behaviour of polyester
composites reinforced with three different weight fractions of woven E-glass fibre-
reinforced composites in an interacting environment. The findings of the experiments
indicate that erodent size, fibre loading, impingement angle and impact velocity are the
significant factors in a declining sequence affecting the wear rate. Significance of ero-
sion efficiency in identifying the wear mechanism is highlighted. Tewari et al.11 studied
solid particle erosion behaviour of unidirectional carbon fibre- and glass fibre-reinforced
epoxy composites. They evaluated different impingement angles (15 –90 ) at three dif-
ferent fibre orientations (0 , 45 and 90 ). They found that unidirectional carbon fibre-
and glass fibre-reinforced epoxy composites exhibit semi-ductile erosion behaviour,
with a maximum rate of erosion at 60 impingement angle. They also found that the fibre
orientations had a significant influence on erosion. Li and Hutchings12 measured the
rates of erosive wear for a series of eight polyester-based one-component castable poly-
urethane elastomers, with widely varying mechanical properties. Erosion tests were con-
ducted with airborne silica sand, 120 mm in particle size, at an impact velocity of 50 m/s
and impact angles of 30 and 90 . For these materials that showed similar values of
rebound resilience, the erosion rate increased with increasing hardness, tensile modulus
and tensile strength. Patnaik et al.13 reviewed the research on solid particle erosion
behaviour of fibre- and particulate-filled polymer composites. The new aspects in the
experimental studies of erosion of fibre- and particulate-filled polymer composites were
emphasized in this article. Implementation of experimental designs and statistical tech-
niques in analyzing the erosion behaviour of composites was discussed. Recommenda-
tions were given on how to solve some open questions related to the structure–erosion
resistance relationships for polymers and polymer-based hybrid composites. Arjula and
Harsha14 evaluated the erosion efficiency (Z) of polymers and polymeric composites by
collecting the available data from the literature pertaining to solid particle erosion under

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


Bagci et al. 779

normal impact conditions. The objective of this article is to show the usefulness of the Z
parameter in identifying various mechanisms in solid particle erosion. The erosion effi-
ciency map is plotted, which indicates the influence of hardness of various polymers and
polymer composites on their erosion resistance. Tilly and Sage15 investigated the influ-
ences of velocity, impact angle, particle size and weight of eroding particles on nylon,
carbon fibre-reinforced nylon, epoxy resin, polypropylene and glass fibre-reinforced
plastic material. Lindsley and Marder16 found impact velocity (v) to be a critical test
variable in erosion, and that it can easily overshadow changes in other variables, such
as target material, impact angle, and so on. Goretta et al.17 scrutinized solid-particle ero-
sion of portland cement and portland-cement-based concrete and, for comparison, a
nuclear waste glass. Using angular aluminium oxide (Al2O3) erodents of mean diameter
42, 63, 143 and 390 mm impacting the surface at 20 and 90 at 50 m/s, the researchers
found the erosion rate to be a strong function of erodent size. Erosion rates of the three
targets by rounded silicon dioxide (SiO2) sand erodent of mean diameter 450 mm were
generally different from those of the 390 mm Al2O3 erodent. In a study conducted
by Suresh et al.,18 solid particle erosion behaviour of polyetherketone (PEK) reinforced
by short glass fibres with varying fibre content (0–30 wt%) was investigated. Here,
the researchers evaluated different impact angles (15 –90 ) and impact velocities
(25–66 m/s) using silica sand particles as abrasives. PEK and its composites exhibited
maximum erosion rate at 30 impact angle indicating ductile erosion behaviour. The ero-
sion rates of PEK composites increased with increase in the amount of glass fibre. In the
study by Harsha and Jha,19 the researchers investigated erosion resistances of neat epoxy
composites, unidirectional glass fibre-reinforced epoxy composites, unidirectional car-
bon fibre-reinforced epoxy composites and bidirectional E-glass woven-reinforced
epoxy composites. Bidirectional glass fibre-reinforced epoxy composites exhibited
higher erosion resistance than their unidirectional fibre-reinforced counterparts. This
is connected to the fact that double directional composites absorb more impact energy.
Roy et al.20 characterized the solid particle erosion behaviour of four different types of
polymer matrix composites reinforced with glass fibres. The erosion rates of these com-
posites have been evaluated at two impact angles (90 and 30 ) and two impact velocities
(38 and 45 m/s). The erosion response, erosion efficiency and the erosion micromechan-
isms of these composites are presented and discussed in detail and also compared with
the available data in the literature on similar materials. Rattan and Bijwe21 conducted a
research where they evaluated erosion wear behaviour of polyetherimide (PEI) and its
composites using silica sand particles striking at a constant impact velocity but with
varying angles of impingement. Although the mechanical properties of PEI improved
substantially by carbon fabric reinforcement, they found that the erosion resistance of
the material deteriorates by a factor of about four to six times at all angles of impinge-
ment. In spite of the fact that PEI is not a very ductile polymer (elongation to fracture,
60%), they found that it exhibits maximum wear at 15 , which is a characteristic of duc-
tile and semi-ductile mode of failure.
In this study, erosion wear behaviour of glass fibre mat-based polyester laminate com-
posite materials that have high strength, good electrical and thermal conductivities and
are corrosion resistant, the features that make them suitable for long time use without the

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


780 Journal of Thermoplastic Composite Materials 26(6)

Table 1. General properties of fibres and test specimens.

Properties Values

Fibre diameter 12 mm
Fibre density 2.6 g/cm3
Fibre aerial weight 225 g/m2
Fibre tensile strength 2.4 GPa
Tensile strength 80 MPa
E-modules 10 GPa
Bending strength 170 MPa
Specific gravity 1.7 g/cm3
Temperature index 135 TI
Electrical strength (in oil at 90 C) 9.0 kV/mm

need for frequent maintenance services in areas such as cooling towers, ships, platforms
and many other application fields, was investigated, wherein three different particle
impact velocities, six different impingement angles and three different particle sizes
were used in undertaking the tests.

Experimental study
Materials
The mat-reinforced composite materials used in this study were industrially produced at
Pultech FRP, a Turkish Materials Company, where ‘‘hand lay-up’’ technique is used to
fabricate the composites. Glass fibre mat-laminated sheet is produced by combining
unsaturated polyester resin (Neoxil CE 92 N8) with glass fibre mat and inorganic fillers
(alumina trihydrate, kaolin and calcium carbonate) and laminating under heat and pres-
sure. The nominal thickness of the composite plates obtained in this process is 3.2 mm.
These laminates have excellent mechanical, physical and electrical properties and
include special grade flame resistance. These properties are shown in Table 1. DIN
Hm 2471 quality grade was used in our glass fibre mat-laminated sheet composite
materials. Material produced under this grade is self-extinguishing (Flame-retardant
UL94–FVO), arc extinguishing and track resistant with very good electrical and
mechanical properties. Preferentially, these materials are applied in high-voltage and
electrical engineering fields as construction units.

Erosion testing
Solid particle erosion testing device used in this study is given in Figure 1. The device
basically consists of a sand hopper, flow control valves, manometers, particle tank,
nozzle, specimen holder, pressure setting valve and eroded sand collector. Dry com-
pressed air is mixed with the particles, which are fed at a constant rate from the sand
hopper into the pressurized particle tank and then accelerated by a compressor, thereby

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


Bagci et al. 781

Figure 1. (a) Erosion test device: 1: sand hopper; 2,5: spherical valve; 3,10: manometer; 4: pres-
surized particle tank; 6: nozzle; 7: specimen holder; 8: eroded sand collector; 9: valve pressure set-
ting; 11: flow control valves; 12: compressor insert manifold. (b) Details of test region: 13:
specimen; 14: nozzle.

forcing the mixture through a tungsten carbide converging nozzle of 6 mm diameter.


These accelerated particles impact the specimen, which can be held at various angles
with respect to the impacting particles using an adjustable sample holder.
The impact velocity of the particles can be varied by varying the pressure of the
compressed air. In order to determine the velocity of the eroding particles, many
methods22–25 were used previously. In this study, the double disc method26 was used to
determine the velocity of the eroding particles, because this method is simple and eco-
nomic. The particle velocities obtained by double disc method were 23, 34 and 53 m/s. For
the erosion tests, a diamond-impregnated slitting saw was used to cut the samples into a

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


782 Journal of Thermoplastic Composite Materials 26(6)

Figure 2. Optic microscopic views of SiO2 abrasive particles: (a) 250 mm, (b) 500 mm and (c)
1000 mm.

Table 2. Chemical compositions of abrasive particles (% weight).

Content Minimum Maximum

% Humidity 3 8
% Clay 0.1 0.5
% SiO2 98 99
% Fe2O3 0.18 0.4
% Al2O3 0.5 1.2

size of 30  30  3.2 mm3 from the manufactured composite plate. All cut edges were
finished using a fine silicon carbide paper. The standard test procedure was performed
in accordance with ASTM G76-9527 for each erosion test. The samples were weighed
to an accuracy of 0.1 mg using an electronic balance. Then 10 kg of abrasive particles
at a flow rate of 182.5 g/s were spurted on the specimen and then the latter was weighed
again to determine its weight loss.
Optic microscopic views and chemical compositions of these particles are shown in
Figure 2 and Table 2, respectively. The distance from specimen surface to nozzle end
was 10 + 1 mm as described in ASTM G76-95 standards.

Results and discussion


Erosion rate
In this experimental study, glass fibre mat-based polyester laminate materials were sub-
jected to erosive wear, where three different types of abrasive particles having sizes of
250, 500 and 1000 mm were used to strike the materials at velocities of 23, 34 and 53 m/s
and impingement angles of 15 , 30 , 45 , 60 , 75 and 90 . The results obtained were
used to plot the graphs in Figures 3 to 5. These graphs show variations in erosion rates

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


Bagci et al. 783

23 m/s 34 m/s 53 m/s

7
6
Erosion x 10 (g/g) 5
-6

4
3
2
1
0
0 15 30 45 60 75 90
Impingement angle (°)

Figure 3. Variation in erosion rates with impingement angles for three different impact velocities
of 250 mm SiO2 abrasive particles.

23 m/s 34 m/s 53 m/s

12
Erosion x 10 (g/g)

10
8
-6

6
4
2
0
0 15 30 45 60 75 90
Impingement angle (°)

Figure 4. Variation in erosion rates with impingement angles for three different impact velocities
of 500 mm SiO2 abrasive particles.

23 m/s 34 m/s 53 m/s

35
30
Erosion x 10 (g/g)

25
-6

20
15
10
5
0
0 15 30 45 60 75 90
Impingement angle (°)

Figure 5. Variation in erosion rates with impingement angles for three different impact velocities
of 1000 mm SiO2 abrasive particles.

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


784 Journal of Thermoplastic Composite Materials 26(6)

of SiO2 abrasive particles of average diameters of 250, 500 and 1000 mm against
impingement angles of the particles.
The values of erosion rates (g/g) used to plot the graphs were determined by dividing
the weight loss in the test specimens by the total weight of abrasive particles used in each
test. In addition, every value in these graphs was obtained by taking an average value of
three different tests performed around that value.

Comparison with literature


Studies made on erosive behaviour of composites, so far, have confirmed that wear rates
generally depend on abrasive particles’ impingement angle, striking velocity and their
sizes and that the wear rate is effectively affected by striking angle.28,29 Maximum
erosion rate for ductile materials generally occurs at 15 –30 impingement angles, while
materials with brittle and semi-ductile behaviours exhibit maximum erosion rates at 90
and 45 –60 , respectively.30,31
The tests conducted indicate that, at low and moderate particle speeds, the materials
subjected to solid particle erosion experience surface damages prior to undergoing material
losses. Then, as the particles persist keeping their contact with the surface, they cause deep
indentations and grooves on the surface of the material. Figure 6 shows the graphs of var-
iations of erosion with impact angle as obtained by combining studies from related literature.
When the graphs formed as a result of our experimental study are investigated, it is
found that the maximum erosion rate occurs at 30 angle of the abrasive particles. The
erosion rate seems to decrease as the striking angle increases. This is a typical erosion
wear behaviour encountered in ductile materials.33,34
In previous studies conducted on ductile materials, it was found that the value of
maximum wear is between two and three times higher than the values of wear formed on
normal impingement angles.35 In our study, apart from observing similar situation, it was
found that the impact velocities and sizes of abrasive particles have greater effects on
erosion rate too.
The increase in impact velocity led to the increase in the wear rate. The remarkable
increase in the wear rate is correlated to particle sizes used in the tests. It is therefore
concluded that abrasive particles have more effects on erosive wear rate than the impact
velocity.
Figure 7 shows the photographs of test specimens that were subjected to 1000 mm
diameter SiO2 abrasive particle bombardments at impact velocities of 23 and 53 m/s. By
looking closely at the photographs, it is seen that there is an increase in the deformation
of the test specimen surfaces as the impact velocities increase. Moreover, when com-
parison is made in terms of the impingement angles, the specimens that were stricken at
between 15 and 30 seem to exhibit deeper erosion marks on the surfaces. The marks
tend to be mild as the angle increases.
Figure 8 shows the optic microscopic photographs of test specimens stricken by SiO2
abrasive particles with average diameter of 250 and 1000 mm at constant impact velocity
(34 m/s). In this situation also, where the effects of abrasive particles at a constant impact
velocity was investigated, it was found that the surface damage tends to increase as the

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


Bagci et al. 785

Figure 6. Schematic figure showing variation in erosion with impact angle for ductile and brittle
material behaviour.32

particle sizes got bigger. The variations in the impingement angles were also studied here
and deep marks were observed at low angles (15 –30 ). The marks decreased as the
angles increased.
The pictures show that the material removal took place at several steps. Initially,
microcracks were formed on the surface during the particle bombardments and then the
surface layer on the matrix was displaced as the particles continued to strike. The
interface between the matrix layer and the mat had undergone damages as the particle
bombardments persisted. The matrix layer was then separated from the mat as a result of
this damage. The continuous flow of particles led the way to increased fractures on the
mats, which consequently intensified deep cracks and wear debris. From investigation of
the microscopic photos, the cracks and debris seemed to increase in depth as the abrasive
particle sizes increased parallel with the impact velocities.

Steady state erosion


Materials erosion wear behaviour can be classified as ductile and brittle erosion,
although this grouping is not definitive. Thermoplastic matrix composites usually exhibit

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


786 Journal of Thermoplastic Composite Materials 26(6)

Figure 7. Optic microscopic photographs showing test specimens stricken by abrasive particles
with an average diameter of 1000 mm at six different impingement angles (15 –90 ) at (a) 23 m/s
and (b) 53 m/s.

ductile behaviour with their highest erosion rate at around 30 impact angle because
cutting mechanism is dominant in erosion while the thermosetting ones erode in a brittle
manner with the peak erosion occurring at normal impact. However, there is a dispute
about this failure classification as the erosive wear behaviour depends strongly on the

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


Bagci et al. 787

Figure 8. Optic microscopic photographs showing test specimens stricken by abrasive particles at
a velocity of 34 m/s at particle diameters of (a) 250 mm and (b) 1000 mm.

experimental conditions and the composition of the target material.36 To characterize the
erosion morphology as received and eroded surfaces with the mode of material removal,
the eroded samples are observed under scanning electron microscope (SEM). Figure 9(a)
shows the dominance of microchipping and microcracking phenomena. It can be seen
that the multiple cracks originate from the point of impact, intersect one another and

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


788 Journal of Thermoplastic Composite Materials 26(6)

Figure 9. SEM views of test specimens: (a) 30 , (b) 60 and (c) 90 (34 m/s particle impact velocity
and 500 mm average particle diameter).

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


Bagci et al. 789

form wear debris due to ductile fracture in the fibre body. After repetitive impacts, the
debris in platelet form is removed and account for the measured wear loss. The occur-
rence of peak erosion rate at 30 impact is understandable. In this case, both abrasion and
erosion processes play important roles. The sand particles after impacting slide on
the surface and abrade while dropping down. The wear and, subsequently, the damage
are therefore more than that in the case of normal impact. Marks of microploughing on
the ductile polyester matrix region seen in Figure 9(a) support this argument. Figure 9(b)
shows the local removal of resin material from the impacted surface resulting in the
exposure of fibres to the erodent flux. This micrograph also reveals that due to sand
particle impact on fibres, there is formation of transverse cracks that break these fibres.
On comparing this microstructure with that of the same composite eroded at a higher
impingement angle (90 ), it can be seen that the breaking of glass fibres is more
prominent (Figure 9(c)). It appears that cracks have grown on the fibres giving rise to
breaking of the fibres into small fragments. Furthermore, the cracks have been anni-
hilated at the fibre matrix interface and seem not to have penetrated through the matrix.
Change in impact angle from normal to oblique alters the topography of the damaged
surface very significantly.

Morphology of eroded surfaces


Figures 10 and 11 show SEM views of the test specimens. The views show variations in
erosive wear caused by the differences in impact velocities and particle sizes. Figure 10(a)
shows variation on test specimen where abrasive particles having diameters of approxi-
mately 1000 mm strike the surfaces of test specimens at 23 m/s. Scratch marks were
observed on the surfaces of the specimen. This can be explained as being the cause of
specimen deformation as a result of particle bombardments. By keeping all other variables
constant, the velocity was increased from 23 to 53 m/s, and the test was conducted. Then
the SEM views seen in Figure 10(b) were obtained. Apart from the scratches on the
specimen surfaces, the increase in the impact velocity led to the formation of debris on
the surfaces. This caused deeper scratches and hence increased wear, thereby worsening
the condition to a critical state.
The fundamental factor that causes negative effects on erosive wear at an increased
striking speed of particles is the change in the kinetic energy of the particles. That is, at
higher speeds and hence at higher kinetic energies of the particles, the abrasive par-
ticles that strike the surfaces of the composite material fracture the matrix more easily
and thus leading to intensified fracturing of the fibres. From the SEM views given in
Figure 10, the deformation on the surfaces of the test specimens subjected to abrasive
bombardments at striking speed of 53 m/s seems to be more vivid. In other words,
deeper grooves are formed on the surfaces of the test specimens as a result of the
particle strikes and that the fractured fibres are at more noticeable sizes. At particles
striking speed of 23 m/s, the specimen surfaces exhibited only mild variations with
shallow grooves.
Figure 11(a) shows SEM views of test specimens taken after being subjected to
erosive wear tests at 34 m/s impact velocity and 30 impingement angles of abrasive

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


790 Journal of Thermoplastic Composite Materials 26(6)

Figure 10. SEM views of test specimens: (a) 23 m/s and (b) 53 m/s (30 impingement angle and
1000 mm average particle diameter).

particles with average diameter of 250 mm. In Figure 11(b), the SEM views are given
when the velocity and impingement angles are kept constant at 34 m/s and 30 ,
respectively, where the particle diameters are 1000 mm. The two photographs, when
studied, seem to have experienced local gaps as a result of particle strikes. But larger
particles seem to have caused severe negative effects on the test specimens. This can be
seen on the SEM views when looking at the damaged fibres. From these SEM views,

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


Bagci et al. 791

Figure 11. SEM views of test specimens: (a) 250 mm and (b) 1000 mm (30 impingement angle
and 34 m/s particle impact velocity).

which were obtained as a result of bombardments of particles having two different sizes
of 250 and 1000 mm, it is found that the increase in particle sizes leads to a decrease in
wear resistance. That is, as the striking particles get larger they tend to exert more
potential energy onto the specimen surfaces and cause extensive damage on the spe-
cimens. Consequently, the increase in both striking speed and particle sizes causes
substantial changes in wear rates.

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


792 Journal of Thermoplastic Composite Materials 26(6)

Conclusions
The following conclusions have been obtained from this study.

1. Maximum erosion rate is observed at 30 impact angle. However, parallel to the
increase in an impingement angle, the values of erosion rate drop. This is a typical
erosion wear behaviour encountered in ductile materials.
2. The increase in impact velocity leads to an increase in erosion rate.
3. The remarkable increase in the erosion rate is correlated with particle sizes used in
the tests. Moreover, changes in abrasive particle size bring about more effects on
erosive wear rate than changes in impact velocities.
4. Looking closely at the optical microscopic and SEM views, it is seen that there has
been an increase in deformation of the test specimen surfaces as the impact veloci-
ties and abrasive particle sizes increase. In addition, adverse effects can be seen at
small impingement angles (15 –30 ).

Funding
This research received no specific grant from any funding agency in the public, commer-
cial, or not-for-profit sectors.

References
1. Kim A and Kim I. Solid particle erosion of CFRP composite with different laminate orienta-
tions. Wear 2009; 267 (11): 1922–1926.
2. Pool KV, Dharan CKH and Finnie I. Erosive wear of composite materials. Wear 1986; 107
(1): 1–12.
3. Rajesh JJ, Jayashree B, Tewari US and Venkataraman B. Erosive wear behavior of various
polyamides. Wear 2001; 249(8): 702–714.
4. Ahmed TJ, Nino GF, Bersee HEN and Beukers A. Improving erosion resistance of polymer
reinforced composites. J Thermoplast Compos 2009; 22 (6): 703–725.
5. American Society for Testing and Materials. ASTM G0040–02. Metals Test Methods and
Analytical Procedures, Wear and Erosion; Metal Corrosion. In: Annual book of ASTM
standards, 03.02, Terminology Relating to Wear and Erosion Terminology Relating to Wear
and Erosion. West Conshohocken, PA: American Society for Testing and Materials, 2005;
3(2): 1–8.
6. Mills D. Erosive wear. In: Pneumatic conveying design guide, 2nd ed. Great Britain: Elsevier
Butterworth–Heineman, 1992, pp.498–525.
7. Kosel TH. Solid particle erosion. In: ASM handbook: friction, lubrication and wear technol-
ogy, Vol. 18. USA: ASM International, 1992, pp.199–213.
8. Sundararajan G and Roy M. Solid particle erosion behaviour of metallic materials at room and
elevated temperatures. Tribol Int 1997; 30(5): 339–359.
9. Hebbar A, Ouinas D, Lousdad A and Bouiadjra BB. Erosive wear modeling of polymeric com-
posite materials. J Reinf Plast Comp 2010; 29 (12): 1893–1899.
10. Patnaik A, Satapathy A, Mahapatra SS and Dash RR. A taguchi approach for investigation of
erosion of glass fiber – polyester composites. J Reinf Plast Comp 2008; 27 (8): 871–888.
11. Tewari US, Harsha AP, Hager AM and Friedrich K. Solid particle erosion of carbon fibre- and
glass fibre-epoxy composites. Compos Sci Technol 2003; 63(3–4): 549–557.

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


Bagci et al. 793

12. Li J and Hutchings IM. Resistance of cast polyurethane elastomers to solid particle erosion.
Wear 1990; 135: 293–303.
13. Patnaik A, Satapathy A, Chand N, Barkoula NM and Biswas S. Solid particle erosion wear
characteristics of fiber and particulate filled polymercomposites: a review. Wear 2010; 268:
249–263.
14. Arjula S and Harsha AP. Study of erosion efficiency of polymers and polymer composites.
Polym Test 2006; 25: 188–196.
15. Tilly GP and Sage W. The interaction of particle and material behavior in erosion process.
Wear 1970; 16: 447–465.
16. Lindsley BA and Marder AR. The effect of velocity on the solid particle erosion rate of alloys.
Wear 1999; 225–229: 510–516.
17. Goretta KC, Burdt ML, Cuber MM, Perry LA, Singh D, Wagh AS, et al. Solid-particle
erosion of Portland cement and concrete. Wear 1996; 224: 106–112.
18. Suresh A, Harsha AP and Ghosh MK. Erosion studies of short glass fiber-reinforced thermo-
plastic composites and prediction of erosion rate using ANNs. J Reinf Plast Comp 2010; 29
(11): 1641–1652.
19. Harsha AP and Jha SK. Erosive wear studies of epoxy-based composites at normal incidence.
Wear 2008; 265: 1129–1135.
20. Roy M, Vishwanathan B and Sundararajan G. The solid particle erosion of polymer matrix
composites. Wear 1994; 171: 149–161.
21. Rattan R and Bijwe J. Influence of impingement angle on solid particle erosion of carbon
fabric reinforced polyetherimide composite. Wear 2007; 262: 568–574.
22. Finnie I, Wolak J and Kabil Y. Erosion of metals by solid particles. J Mater 1967; 2: 682–700.
23. Barkalow RH, Goebel JA and Pettit FS. Erosion-corrosion of coatings and superalloys in high-
velocity hot gases. In: Adler WF (ed.) Erosion: prevention and useful applications, ASTM STP
664. 1979, pp.163–192.
24. Ninham AJ and Hutchings IM. A computer model for particle velocity calculation in erosion
testing. In: Proceedings of sixth international conference on erosion by liquid and solid
impact. Cambridge, UK: University of Cambridge, 1983, pp.50–51.
25. Stevenson ANJ and Hutchings IM. Scaling laws for particle velocity in the gas-blast erosion
test. Wear 1995; 181–183: 56–62.
26. Ruff AW and Ives LK. Measurement of solid particle velocity in erosive wear. Wear 1975; 35:
195–199.
27. American Society for Testing and Materials. ASTM G0076–04. Metals Test Methods and
Analytical Procedures, Wear and Erosion; Metal Corrosion. In: ASTM annual book of stan-
dards. 03.02, Test Method for Conducting Erosion Tests by Solid Particle Impingement Using
Gas Jets. West Conshohocken, PA: American Society for Testing and Materials, 2000: 1–6.
28. Imrek H, Bagci M and Khalfan OM. Experimental investigation of effects of external loads on
erosive wear. J Achieve Mater Manuf Eng 2009; 32 (1): 18–22.
29. Sari NY, Sinmazcelik T and Yilmaz T. Erosive wear studies of glass fiber- and carbon
fiber-reinforced polyetheretherketone composites at low particle speed. J Thermoplast
Compos 2011; 24 (3): 333–350.
30. Harsha AP and Thakre AA. Investigation on solid particle erosion behaviour of polyetheri-
mide and its composites. Wear 2007; 262: 807–818.
31. Zahavi J and Schmitt GF. Solid particle erosion of reinforced composite materials. Wear 1981;
71: 179–190.
32. Finnie I. The mechanism of erosion of ductile metals. In: Proceedings of third US national
congress of applied mechanics. 1958; 527–532.

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016


794 Journal of Thermoplastic Composite Materials 26(6)

33. Harsha AP, Tewari US and Venkatraman B. Solid particle erosion behaviour of various
polyaryletherketone composites. Wear 2003; 254: 693–712.
34. Jha AK, Mantry S, Satapathy A and Patnaik A. Erosive wear performance analysis of
jute-epoxy-SiC hybrid composites. J Compos Mater 2010; 44 (13): 1623–1641.
35. Hutchings IM. The erosion of ductile metals by solid particles. PhD Dissertation. University
of Cambridge, UK, 1974.
36. Barkoula NM and Karger-Kocsis J. Effects of fibre content and relative fibre-orientation on
the solid particle erosion of GF/PP composites. Wear 2002; 252 (1–2): 80–87.

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 15, 2016

S-ar putea să vă placă și