You are on page 1of 13

Available online at www.sciencedirect.

com

Wear 264 (2008) 4759

Synergic effect of reinforcement and heat treatment on the two body abrasive
wear of an AlSi alloy under varying loads and abrasive sizes
S. Das , D.P. Mondal, S. Sawla, N. Ramakrishnan
Regional Research Laboratory (CSIR), Bhopal 462026, India
Received 18 January 2006; received in revised form 19 January 2007; accepted 24 January 2007
Available online 27 February 2007

Abstract
In the present study, an attempt was made to examine the synergic effect of SiC particle reinforcement and heat treatment on the two body
abrasive wear behavior of an AlSi alloy (BS: LM13) under varying loads and abrasive sizes. Silicon carbide particles with size 5080 m were
reinforced in AlSi alloy, in varying concentration (10 wt% and 15 wt%), by solidification process (vortex technique) and the composite melt was
solidified by gravity casting in a cast iron die. The alloy and composites were solution treated at 495 C for 8 h, quenched in water and aged at
175 C for 6 h and cooled in air. Two body abrasive wear behaviour of cast and heat-treated alloy and composite, was examined against abrasives of
different sizes (40 m, 60 m and 80 m), at varying applied loads (1 N, 3 N, 5 N and 7 N), up to a sliding distance of 108 m. It has been noted that
the alloy suffers from higher wear rate than that of composites either in cast or heat-treated conditions, irrespective of applied load and abrasive
size. Further, in most of the cases, the wear rate of composite decreases with increase in SiC particle content. Efforts were made to correlate wear
behavior of Al alloy and composites in terms of mechanical properties, microstructural characteristics, applied load and abrasive size through an
empirical equation.
2007 Elsevier B.V. All rights reserved.
Keywords: Al alloy; Al alloy composites; Abrasive wear; Heat treatment; Abrasive size

1. Introduction
Al-alloy matrix composites (AMCs) containing hard dispersoids are gaining immense industrial importance because of
their excellent combination of physical, mechanical and tribological properties over base alloys [1]. These include high
wear and seizure resistance, high specific strength and stiffness,
improved high temperature strength, controlled thermal expansion coefficient and high damping capacity [18]. It is reported
that Al-alloy reinforced with 10 wt% SiC particle composite
provides comparable mechanical properties but better thermal
conductivity and specific heat than the cast irons [5,7]. As a
result, frictional heating of these composites are noted to be significantly less than that of cast irons [7]. This leads to the use of
these composites in several automobile and engineering components where wear, tear and seizure are the major problems in
addition to the weight saving. Some of these components are

Corresponding author. Tel.: +91 755 2488562; fax: +91 755 2587042.
E-mail address: sdas88@hotmail.com (S. Das).

0043-1648/$ see front matter 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.wear.2007.01.039

pistons, brake drums, cylinder heads, connecting rods and drive


shafts for automobile sectors and impellers, agitators, turbine
blade, valves, pump inlet, vortex finder for marine and mining sectors [911]. In recent years, fabrication of several such
components from AMCs, reaches to the commercial production
stage [712]. Most of the aforesaid components are subjected to
different kinds of wear and tear related failure. In this context, it
is required to characterize Al-composites in terms of wear under
different experimental conditions.
Sliding wear behavior of AMCs has been studied by many
investigators [1316], however, limited attempts have been made
to study the abrasive wear characteristics of Al-composites
[1727]. According to these reports, wear and seizure resistance
of AMCs is considerably higher than that of the base alloys. This
is primarily attributed to the fact that the hard dispersoids (reinforcing phase) protect the surface from the destructive action
of the abrasives by reducing the depth of penetration of the
abrasives (hard asperities of the counter surface) and the contact between the abrasive and the matrix. On the other hand,
few investigators have reported a transition of abrasive wear
behaviour of composites which was dependent on abrasive size

48

S. Das et al. / Wear 264 (2008) 4759

and applied load [19,21,24,28]. According to these investigators, above a critical value of load and abrasive size, AMCs
exhibit higher wear rate than the alloys [2430]. Under such
circumstances, the depth and width of wear track become larger
than the size of dispersoid (i.e., h/d and/or w/d is greater than
unity; where h and w are the average depth and width of wear
groove, respectively, and d is the average diameter of the reinforcing phase in the composite). This leads to the scooping off
of the dispersoid from the surface of the composites samples
[19,21,24]. Furthermore, it is evident from the literature that
the wearing surface and the subsurface undergo plastic deformation, and this deformation becomes more severe when the
abrasive size is coarser and the applied load is higher [20,29].
As a result of such plastic deformation, the hard ceramic dispersoid gets fractured and fragmented into finer ones, and/or
debonded from the matrix [2329]. In due course, these fragmented particles come out from the specimen surface. Thus the
abrasive wear behavior of composite depends on the material
characteristics like shape, size, distribution and volume fraction
of the dispersoids and experimental parameters like applied load
and abrasive size [2131].
It has been reported that the wear resistance of composite
increases with increase in volume fraction and size of the dispersoids [2429]. One of the prime factors of the improvement
in wear resistance is increased in hardness of the Al-alloy due to
the addition of hard dispersoids [2431] and more protection of
the matrix from the destructive action of the abrasive as the mean
free path between the SiC particles is reduced with increase in
volume fraction of SiC particle [32]. Several investigators have
also proposed that wear resistance of a material not only depends
on its hardness and strength but also on its ductility and toughness [33,34]. The reinforcement of Al2 O3 particles in aluminum
alloy enhances the abrasive wear of the matrix. The reinforcement of 16 m Al2 O3 particle strengthens the aluminum matrix
and enhances the wear resistance. However, the reinforcement of
coarse particle (66 m) shows higher wear resistance. The operating wear mechanism is mainly consists of plastic deformation
of the matrix material [35].
The hardness as well as toughness of a composite material depends significantly on the matrix microstructure, size and
distribution of the dispersoids and the interfacial bonding characteristics [32]. The hardness of the composite increases with
increase in the volume fraction of the dispersoid but at the same
time its toughness decreases. Additionally, the casting defects
may increase or the possibility of clustering of dispersoid particle may be more as one increases the dispersoid content. These
may reduce the wear resistance of the composite [30]. Thus
the overall improvement in wear resistance of the alloy, due
to addition of more dispersoid may not be so high especially
at severe wearing conditions (i.e., at higher applied load and
coarser abrasive size).
The mechanical and tribological properties of the composite also depend upon the matrix microstructure; hence, the
properties of composite can be improved by heat treatment.
During heat treatment, the matrix of composite behaves almost
similar to that of the base alloy and the dispersoid remains
unchanged. Attempts have been made to improve strength, hard-

ness and toughness of AlSi alloy and AlSi alloy reinforced


with hard particle composites by altering the matrix microstructure through heat treatment [36,37]. Studies on sliding wear
characteristics reported that the wear and seizure resistance of
the alloy or composite are improved by heat treatment [36,37].
However, it is interesting to examine the abrasive wear behavior
of AlSi alloy reinforced with hard particles after heat treatment.
In the present study, an attempt has been made to examine
the synergic effect of particle content and heat treatment on
the abrasive wear behavior of an AlSi alloy and AlSiSiC
composite under varying applied loads and abrasive sizes.
2. Experimental
2.1. Materials
The AlSi (British Standard: LM13) alloy and LM13 alloy
containing 10 wt% and 15 wt% SiC particle have been selected
in the present investigation. The LM13 alloy nominally contains 11.95% Si, 1.0% Mg, 1.5% Ni, 1.0% Cu, 0.8% Fe, 0.6%
Mn and balance is Al. Silicon carbide particles of size 5080 m
have been incorporated into the AlSi alloy melt in varying concentration (10 wt% and 15 wt% of the matrix alloy) by vortex
technique. The composite melt was solidified in a cast iron die
in the form of circular disc of diameter 100 mm and thickness
5 mm. The Al-alloy melt was also cast in the same die. The cast
alloy and composites were solutionized at 495 C for 6 h and
quenched in water. The samples were tempered at 175 C for
6 h followed by air-cooling.
2.2. Microstructural examination
For microstructural observations, samples were mechanically
polished using standard metallographic techniques, etched with
Kellers reagent and observed in a JEOL 5600 scanning electron
microscope (SEM). Prior to SEM examination, the samples were
sputtered with gold.
2.3. Two-body abrasive wear testing
High stress (two-body) abrasive wear tests of Al alloy and
Al alloySiC composites were conducted on 40 mm 35 mm
5 mm rectangular specimens using Suga made Abrasion
Tester (Model: NUSI, Japan). A schematic view of the apparatus
is shown in Fig. 1. Emery paper embedded with SiC particles (size: 40 m, 60 m and 80 m) was used as the abrasive
medium. These emery papers were cut into sizes and fixed on
the wheel (50 mm diameter), which rotates against the specimen surface. The specimen over the wheel was fixed with a
locking arrangement and load on the specimen was applied with
the help of cantilever mechanism. The specimen was subjected
to reciprocating motion against the wheel on which the abrasive
media is rigidly fixed. The wheel makes one complete revolution
after each 400 cycles (corresponds to 26 m of sliding distance)
of reciprocating motion of the specimen. The abrasive paper
on the wheel was changed after an interval of each 400 cycles
so that the specimen surface always makes contact with fresh

S. Das et al. / Wear 264 (2008) 4759

49

abrasives. The tests were conducted at different loads (i.e., 1 N,


3 N, 5 N and 7 N) and at varying abrasive sizes (40 m, 60 m
and 80 m) up to a sliding distance of 108 m. The wear rate of
specimen was calculated from weight loss measurements. The
specimens were ultrasonically cleaned prior to and after each
interval of the wear tests.
2.4. SEM examination of wear surface and subsurface
In order to understand the mechanism of material removal,
the wear surface and subsurface were examined in SEM. For
subsurface examination, the worn samples were cut transversely,
mounted, polished, etched with Kellers reagent and observed
with SEM.
Fig. 1. Schematic view of Suga made abrasion tester (1) specimen stage; (2)
specimen guide; (3) abrasion wheel: (4) specimen press; (5) metal fitting; (6)
double number detector; (7) motor; (8) fixed weight; (9) weight; (10) weight
scale; (11) lock lever.

2.5. Mechanical testing


The hardness of the alloy and composites, in as cast and
heat-treated conditions, is measured using a Vickers hardness
tester at 2 N load. Before hardness measurements, each sample is
polished and its opposite sides are made perfectly parallel. Yield
strength, ultimate tensile strength and percentage elongation of
Al-alloy and Al-composite were determined from tensile test

Fig. 2. (a) Microstructure of AlSi (LM13) alloy (A: aluminum, B: silicon). (b) A higher magnification micrograph of (a) clearly depicts needle shaped eutectic
Si in Al matrix (B: silicon needle). (c) Typical microstructure of heat-treated AlSi (LM13) alloy showing near spherical Si particle in Al matrix. (d) A higher
magnification micrograph showing the spherical shaped eutectic Si (marked A) and intermetallic phases (arrow marked).

50

S. Das et al. / Wear 264 (2008) 4759

data. Tensile test was conducted in an INSTRON (Universal


Testing Machine) at a strain rate of 104 s1 .
3. Results
3.1. Microstructure
Fig. 2a shows a typical microstructure of cast AlSi (LM13)
alloy which depicts dendrites of aluminium (marked A) and
needle shaped eutectic silicon (marked B) in the interdendritic
regions and around the dendrites. A higher magnification micrograph (Fig. 2b) clearly depicts needle shaped eutectic silicon
(marked B) in Al matrix. The microstructure of heat-treated
LM13 alloy shows fragmentation of needle-shaped eutectic silicon (in cast condition) into more or less spherical one (Fig. 2c). A
higher magnification micrograph (Fig. 2d) clearly shows heattreated eutectic silicon (marked A) and intermetallic phases
(arrow marked).
Fig. 3a represents the microstructure of cast AlSi alloy reinforced with 10 wt% SiCp composite showing distribution of SiC
particles in AlSi alloy matrix. A higher magnification micro-

graph of the composite shows bonding between SiC particle


and the AlSi alloy matrix (Fig. 3b). The change in matrix
microstructure of composite due to heat treatment is noted to
be similar to that observed in the case of heat-treated alloy as
shown in (Fig. 2b). Fig. 3c shows the microstructure of cast
AlSi15 wt% SiC composite. It also indicates distribution of
SiC particles in Al matrix. The distribution of SiC particle in
AlSi alloy matrix is not affected by the heat treatment. Fig. 3d
shows the microstructure of heat-treated AlSiC composite
depicts interface bonding between SiC and AlSi alloy matrix.
3.2. Mechanical properties
The hardness, ultimate tensile strength and elongation of
LM13 alloy and LM13SiC composites in cast and heat-treated
conditions are shown in Table 1. It indicates that hardness and
strength of AlSiC composite are 2530% higher (depending on
SiC content and heat treatment) than that of alloy, respectively.
But the ductility of composite is recorded to be less than that
of the alloy. It is further noted from Table 1 that all the above
mentioned properties are improved significantly due to heat

Fig. 3. (a) Microstructure of Al10 wt% SiC composite in cast condition showing distribution of SiC particle in Al matrix. (b) A higher magnification micrograph
of (a) exhibits interface bonding between SiC particle and Al matrix (A: SiC particle; B: AlSi alloy; C: eutectic silicon; D: Al/SiC interface). (c) Microstructure
of cast Al15 wt% SiC composite showing distribution of SiC particle in Al matrix. (d) Microstructure of heat-treated Al15 wt% SiC composite shows interface
bonding between SiC particle and Al matrix.

S. Das et al. / Wear 264 (2008) 4759

51

Table 1
Mechanical properties of as cast and heat-treated Al alloy and composites
Material

Processing condition

Hardness (MPa)

UTS (MPa)

Elongation (%)

LM13

As cast
Heat-treated

132
146

180
210

1.0
3.0

LM1310 wt% SiC

As cast
Heat-treated

151
155

210
220

1.0
1.0

LM1315 wt% SiC

As cast
Heat-treated

159
173

220
230

1.0
1.0

treatment. It is further noted that hardness and strength of


composite increased with increase in SiCp content but the
ductility of composite is reduced considerably.
3.3. Two body abrasive wear behavior
The wear rates of LM13 alloy and LM13SiC composite are
plotted as a function of sliding distance when they are tested
against different abrasive sizes under varying applied loads.
Fig. 4 represents variation of the wear rate of alloy and composite as a function of sliding distance, when tested at 1 N load and
40 m abrasive size. It is noted that wear rate is almost invariant
to the sliding distance irrespective of the material. It has further
been noted that the wear rate of cast composites decreases with
increase in SiCp content. However, after heat treatment, the wear
rate of alloy and composite were reduced. On the other hand, the
wear rate of LM1315 wt% SiCp composite is noted to be more
or less same in heat-treated and in cast conditions. It may be
noted that the wear rate of LM1310 wt% SiCp composite and
LM1315 wt% SiC composite at 1 N load is more or less same.
However, at an applied load of 7 N, appreciable decrease in wear
rate was observed in LM1315 wt% SiC than LM1310 wt%
SiC composite. The wear rate of heat-treated LM1315 wt%
SiCp composite is noted to be 7.0 1011 m3 /m, whereas it
is found to be 15 1011 m3 /m in LM1310 wt% SiCp composite. Additionally, when the load is significantly low and the
abrasive size is finer, the abrasives are unable to penetrate deep
into the specimen surface especially in the case of composite. In
such case, the abrasive primarily rubs on the specimen surfaces

Fig. 4. Wear rate of Al alloy and composite as a function of sliding distance


(applied load: 1 N and abrasive size: 40 m).

and leads to very less material removal in composites. Because


of such low wear rate, the exact contribution from SiCp cannot
clearly be understood. Fig. 5 clearly indicates the effect of
SiCp content and heat treatment on the abrasive wear behaviour
as a function of sliding distance of alloy and composites
under higher load (7 N) and coarser abrasives (80 m). It may
be noted from this figure that the composite exhibits higher
wear resistance than the alloy and the wear rate of composite
reduces with increase in SiCp content irrespective of cast and
heat-treated conditions. However, it is interesting to note that the
wear rate of LM1310 wt% SiCp composite is marginally less
than that of the alloy. The heat-treated alloy exhibits even less
wear rate than the cast composite reinforced with 10 wt% SiC
particles.
It is evident from Figs. 4 and 5 that the wear rate of each of the
material is increasing with increasing in abrasive size and load.
However, the wear rate is not following any specific functional
relationship with abrasive size irrespective of material. It is noted
that the wear rate of the alloy is higher than those of composites.
The wear rate of composite containing 15 wt% SiCp is noted to
be less than that of composite containing 10 wt% SiCp under
heat-treated condition. Similar type of plots at loads of 3 N, 5 N
and 7 N are shown in Figs. 68, respectively. It is evident from
Fig. 8 that in most of the cases, the wear rate is almost invariant
to the sliding distance irrespective of the materials when the
abrasive size is 40 m. Fig. 9 shows the effect of abrasive size
on the wear rate of LM13 alloy and LM13SiC composite at
an applied load of 5 N. It is evident from Fig. 9 that in most of

Fig. 5. Wear rate of Al alloy and composite as a function of sliding distance


(applied load: 7 N and abrasive size: 80 m).

52

S. Das et al. / Wear 264 (2008) 4759

Fig. 6. Wear rate of Al alloy and composite as a function of sliding distance


(applied load: 3 N and abrasive size: 40 m).

Fig. 9. Wear rate of Al alloy and composite as a function of abrasive size (applied
load: 5 N).

the cases the wear rate is almost invariant to the abrasive size
irrespective of the materials, when the abrasive size is less than
60 m. However, the wear rate increases rapidly if the abrasive
size is coarser than 60 m.
Fig. 10 shows the variation of the wear rate of materials with
applied load, when tested against abrasive of size 60 m. It
is noted that the wear rate of cast alloy and cast composites
is increases with applied load and does not follow any specific functional relationship except for cast LM13 alloy and cast
LM1310 wt% SiC composite which follows a liner relationship
with applied load. It may be noted further that at lower applied
load (3 N), the wear rate of composites varies marginally with
heat treatment. But at the higher applied load (5 N) the wear
rate of composite reduced significantly due to heat treatment and
increase in SiCp content. This figure (Fig. 10) clearly demonstrates that the wear of alloy decreases after heat treatment and
reinforcement of SiC particle.
Fig. 7. Wear rate of Al alloy and composite as a function of sliding distance
(applied load: 5 N and abrasive size: 40 m).

Fig. 8. Wear rate of Al alloy and composite as a function of sliding distance


(applied load: 7 N and abrasive size: 40 m).

Fig. 10. Wear rate of Al alloy and composite as a function of applied load
(abrasive size: 60 m).

S. Das et al. / Wear 264 (2008) 4759

4. Discussion
4.1. Materials and mechanical properties
In the heat treatment of AlSi (LM13) alloy, the needleshaped eutectic silicon is fragmented into more or less spherical
one during solution treatment and fine precipitation on the
matrix takes place while ageing treatment. This is attributed
to reduction in surface energy of the needle shaped silicon
particles while changing into near spherical ones as the surface
area to volume ratio is the minimum for spherical shape. This
reduction in surface energy acts as the driving force for such
transformation. Again, the concentration of Si per unit volume
is more at the sharp edges of the needle shaped particle, which
led to the existence of concentration gradient between the sharp
edges and the wider portion of the Si needle. As a result, the
silicon from the sharp edges is dissolved and diffused to the
wider region and in due course of time, it turns into near spherical shape. Here, the heat energy applied during heat treatment
acts as the activation energy for such kind of transformation. In
addition, because of the same reasons, there is a possibility of
coarsening of spherical silicon particles with aging time. Hence,
it is expected that the spherical silicon particles may be coarser
when they are aged for longer duration. Further more, during
aging, precipitation of alloying elements like Cu, Mg, Ni and Fe
may take place which might have caused further strengthening
of the Al-alloy. The composite exhibits more hardness and
strength than that of alloy. This is attributed to the fact that by
addition of SiC particles in the alloy generates more dislocations
because of thermal mismatch stress [38,39]. Such stress is
originated due to differences in thermal expansion coefficient
of the matrix and the SiC particles [38,39] and makes the matrix
plastically constrained and caused higher dislocation strengthening of the matrix. The hard SiC particle also acts as barrier
to plastic deformation of the matrix alloy, which results in
triaxial interaction stress at the interface. This interaction stress
also results in higher strength of the composite. Heat treatment
results in formation of spherical Si particles and precipitation of
other alloying elements (Figs. 2c and 3b) in terms of complex
intermetallic compounds. This leads to higher strength as
well as higher toughness of the alloy and composite after heat
treatment.
4.2. Effect of reinforcement on the abrasive wear
In two-body abrasive wear, the abrasive particles in the abrasive media make contact with the specimen surface and penetrate
inside the specimen surface when load is applied. The penetrated
abrasive is subjected to reciprocating motion over the specimen surface and thus resulting in formation of wear grooves
on the surface of the test specimen. These grooves are generated due to flow or removal of materials either by cutting
or ploughing action. The depth and width of the grooves and
their number on the test specimen surface depend on the shape,
size and rake angle of the abrasives, specimen surface characteristics such as surface roughness and the stability of the
asperities or hard protrusions, and the hardness of the specimen

53

surface [3138]. The hardness of the composites is noted to be


higher than that of the alloy and the hardness increases with
increase in SiCp content (Table 1). Additionally, the hard SiC
particles in composite act like hard protrusion over the specimen surface. As a result, less number of abrasives come in
contact with the matrix alloy and hence their depth of penetration is reduced. The abrasive particles, which penetrate into
the composite surface, interact with the SiC particles during
reciprocating motion. During this process, the abrasive particles
may either scoop off from the abrasive media, as the bonding between SiC and Al matrix is stronger than the bonding
of abrasive particles in the cloth or paper, or get degraded by
the SiC particles. Fig. 11a shows a typical SEM micrograph
of emery paper (particle size: 80 m) before wear test. It
shows SiC particles with sharp edges embedded on the paper.
Fig. 11b shows a SEM micrograph of the emery paper (particle size : 80 m) after the wear test ( load: 5 N and distance:
81 m). It clearly depicts the degradation of SiC particles and
smearing off fine fragmented particles on the emery paper.
However in some instances, scooping off (marked A) and
breaking (arrow marked) of the SiC particle are also observed
(Fig. 11c). As a matter of these facts, the cutting efficiency
or the severity of the destructive action of abrasive media is
reduced substantially when they move against the composite
surface and finally resulting in significantly lower wear rate of
the composite materials (Figs. 49). The number of SiC particles on the composite increases with increasing SiCp content,
and the hardness of composite is also increased due to addition of more SiC. Furthermore, inter-particle distances are also
reduced due to increase in SiC content. These, in turn, decrease
the wear rate of composite with increase in SiC content. The
improvement in wear resistance of the alloy due to addition
of SiC particles is shown in Figs. 49. It is clearly noted that
the wear resistance of the alloy is improved by 20% due to
addition of SiC and other experimental conditions. The improvement is reduced with increase in applied load and abrasive
size.
It is a well-established phenomenon that during abrasive
wear a portion of effective stress caused plastic deformation
of the specimen surface [33,34]. The hard ceramic dispersoids
carry the major portion of this stress and hence the area around
these ceramic particles undergoes more plastic deformation.
As the matrix in composite is plastically constrained and
the SiC particles are brittle in nature, there are possibilities
of either debonding or fracturing of SiC particles at severe
wear conditions (higher load and coarser abrasive size). These
particles, as a consequence, scooped off from the specimen
surface and may lead to increased wear rate. However, there is
a possibility of picking up of finer fragmented SiC particles or
degraded abrasives in the wear groove which may improve wear
resistance of the matrix so long they are remained intact on the
wear grooves. The plastic deformation becomes significantly
high when wear is occurring under the condition of high
applied load and coarser abrasive size. Due to significantly high
plastic deformation, the lateral and longitudinal cracks on the
specimen surface are generated. As a result, at higher applied
load and coarser abrasive size, the possibility of scooping off

54

S. Das et al. / Wear 264 (2008) 4759

4.3. Effect of heat treatment


The abrasive wear behaviour of a material depends on factors like abrasive size and applied load as the depth of cutting
and over all cutting efficiency of the abrasive media increases
with increase in abrasive size and applied load. As the relative hardness of the specimen surface with respect to abrasive
increases, the depth of penetration of abrasive decreases and
causing less surface damage. This leads to the fact that both
the alloy and composites exhibit minimum wear rate after heat
treatment. Similarly, the hardness of cast composite is improved
by 1520% after heat treatment and the wear rate is reduced
by 1570% depending on test conditions. These facts suggest
that, not only the hardness but also other mechanical properties and microstructural characteristics of alloy and composite
control abrasive wear behaviour depending on the experimental
parameters.
It was mentioned that plastic deformation of the surface and
subsurface is taking place during abrasive wear. In cast AlSi
alloy, needle shaped eutectic silicon is randomly oriented. In few
regions, these silicon particles are parallel to the wear grooves
and in some other regions they are placed perpendicular to the
wear groove direction at which the abrasive scratches the surface. Under this condition, when the abrasive interacts with the
silicon needle oriented perpendicular to the wear direction, the
abrasive movement is restricted until the Si needles are fractured.
The fracturing of the silicon needle is happened due to the interaction between the SiC particles and load applied on the sample.
Fig. 12 shows a SEM micrograph of the subsurface of the AlSi
alloy clearly depicts the highly deformed region (marked A),
cracks (arrow marked) and undeformed region (marked B). It
is observed that the sharp edged silicon needles are fragmented
in to smaller size particles in the subsurface deformed region.
During abrasive wear, the effective stress on the tip of abrasive
may be significantly high and the depth of penetration may be
considerably higher than the diameter of the silicon needles.
This in turn causes fracturing of Si particles and is removed as
micro cutting chips from the specimen surface. Additionally,
the edges of the needle shaped silicon particles may act as stress

Fig. 11. (a) SEM micrograph of as received emery paper (abrasive size: 80 m).
(b) SEM micrograph of emery paper (abrasive size: 80 m) after wear test (load:
5 N, distance: 81 m). (c) SEM micrograph depicting degraded, scooping off
(marked A) and breaking (arrow marked) of SiC abrasive particles in emery
paper.

hard SiC particles becomes higher and the improvement in wear


resistance of the composite over the base alloy is relatively less
as compared to that obtained at lower load and finer abrasive
size.

Fig. 12. SEM micrograph of subsurface of AlSi alloy depicting fragmentation


of eutectic silicon.

S. Das et al. / Wear 264 (2008) 4759

raiser and caused more surface cracks during abrasive wear of the
cast specimens. Furthermore, the inter-silicon particle distance
may be higher in cast conditions, and hence, either the abrasive particles may penetrate deeper into the specimen surface
or more number of abrasives penetrates into the worn surface.
These facts lead to considerably higher wear rate of cast alloy
and composites as compared to that in heat-treated condition.
It may be noted that the wear resistance of heat-treated alloy
improved by 4070% over the as cast condition depending upon
the applied load and abrasive size. The wear resistance of heattreated composite improved by 2070% over the cast one. It
shows that heat treatment also significantly reduced the wear
rate of the materials. However, the contribution to reduce wear
rate due to particle reinforcement and heat treatment is difficult
to quantify as each of these contributions are varying depending
on test conditions and material. In a nutshell, it demonstrates
qualitatively the synergic effect of particle reinforcement and
heat treatment.
As the silicon particle become spherical, the points of stress
raiser is reduced considerably and the possibility of fracture
and fragmentation of the silicon particle is reduced. The silicon
particles remain intact within the Al-matrix. During abrasive
wear, a fraction of the matrix material is displaced from the wear
grooves and spread along the side of wear grooves in the form
of flakes. These flaky materials remain intact along the wear
track for a longer duration or get smeared on the wear surface
because of their higher ductility. Higher ductility of heat-treated
AlSi alloy is primarily attributed to near spherical shape silicon
particles and results in less surface and subsurface cracking. The
flakes formed due to frictional heating adhered and deformation
on the wear surface in due course is delaminated in the form of
long flakes and plates. As a matter of this fact, the heat-treated
materials are exhibiting relatively less wear rate.
4.4. Effect of load and abrasive size
The abrasive wear behaviour of a material is significantly
influenced by the combined actions of load and abrasive size.
The efficiency of material removal by an abrasive media depends
on the elastic and plastic contact load [20], which varies with
applied load and abrasive size. If the applied load is fixed, then
the effective stress on individual abrasives increases with coarser
abrasive particles, as the load is shared by less number of abrasives. When the abrasive particles are finer in size, they make
only elastic contact with the test specimen surface, as the effective stress in individual abrasive is less. As a result, these abrasive
particles only support the applied load without contributing sufficient material removal. However, at higher load regime, the
effective stress on each individual abrasive particles reach to a
level where the abrasives make plastic contact with the specimen surface and causing more surface damage even at finer
abrasive size. Thus the overall rate of material removal depends
on the extent of plastic contact of the abrasives with the specimen
[30]. Further more, coarser size abrasives generally contain large
number of flows and hence may break easily [35]. As a result,
cutting efficiency of the abrasive media is reduced. These broken abrasives are also picked up by the wear surface more easily

55

and protect the specimen from destructive action of the abrasive


media [35]. The interaction of these facts causes very marginal
variation in wear rate with abrasive size up to a size of 60 m
(Fig. 9).
It has been reported that the abrasive wear is also associated
with plastic deformation of the surface and subsurface. During
such kind of wear, the abrasive particles make wear grooves
either by cutting or ploughing action over the specimen surface.
It has been reported that the wear rate depends on w/r ratio,
where w is the groove width and r is the radius of the spherical tip of the abrasive particle [32]. The radius of the abrasive
tip is varying with increase in abrasive size. But the width of
the groove increases substantially with increase in abrasive size.
This may be attributed to the fact that, the effective stresses
on each individual abrasives increases substantially when the
abrasives become coarser in size, and make it more effective to
penetrate deeper into the specimen surface. In addition, degradation of abrasive is reduced when the abrasive size becomes
coarser. Because of greater penetration, the surface and subsurface is subjected to more plastic deformation, which causes more
number of cracks on the surfaces. As a result, the wear rate of
material increases rapidly when abrasive size is increased from
60 m to 80 m in majority of the cases. Fig. 13a shows the wear
surface of an AlSi alloy abraded using emery paper of abrasive
size 80 m at an applied load of 7 N. It clearly shows deep and
wide wear grooves and formation of cracks in the longitudinal
as well as in transverse directions. Under the same experimental conditions, the subsurface examination indicated severely
deformed region with transverse and longitudinal cracks (shown
by arrow mark in Fig. 13b).
The plastic contact of the abrasive with the specimen surface increases with increase in applied load. This leads to more
depth of penetration of the abrasives into the specimen surface
and more plastic deformation of the wear surface and subsurface. The w/r ratio may be increasing with increase in applied
load as r remains constant for fixed abrasive size. In case of composites, the SiC particles act as barrier against the penetration
and movement of abrasives when the applied load is less. As a
result w/r ratio may not be changing significantly with applied
load when the applied load is in the lower range (3 N). This
leads to either more or less wear rate or slower rate of increase
in wear rate with applied load when the load is kept less than or
equal to 3 N (Fig. 10). When the load increases above 3 N, the
surface and subsurface is subjected to severe plastic deformation in addition to cutting and ploughing of material. Under these
circumstances a large number of silicon particles and a few SiC
particles of composite get fractured or fragmented and also the
w/r ratio increases significantly. It led to debonding and scooping off of SiC particles from the matrix, in addition to removal
of matrix material by cutting and ploughing actions. Severe surface cracking also facilitates more surface damage and reduces
the possibility of picking up of detached abrasive particles and
the flaky debris in the wearing surface. These facts as a whole
leads to significantly rapid increase in wear rate especially of
composite at intermediate load level (35 N).
At higher applied load, severe plastic deformation of the surface and subsurface causes heating of the wear surface which

56

S. Das et al. / Wear 264 (2008) 4759

Fig. 14. SEM micrograph showing formation of large size debris and retained
along the wear track.

Fig. 13. (a) SEM micrograph shows deep and wide grooves and formation of
longitudinal and transverse cracks (load: 7 N; abrasive size: 80 m). (b) SEM
micrograph showing subsurface depicting severe deformation and formation
of cracks in longitudinal and transverse directions (A: deformed region; B:
undeformed region).

makes the alloy or the matrix of composite more ductile. As a


consequence, major portion of the material from the wear surface is removed by ploughing action. In this case, large flakes
are generated and retained along the side of the wear track for
a long duration (Fig. 14). Due to heating, these flakes are subjected to severe plastic deformation and spread over the wearing
surface. In due course, these are adhered to the worn surface and
finally removed. Additionally, the abrasives are degraded (especially finer ones) under higher applied load. Thus the cutting
efficiency of such abrasive is reduced. Fig. 15 shows the SEM
micrograph of emery paper (size: 40 m) used in an experiment conducted at a load of 7 N and sliding distance of 108 m. It
clearly depicts degraded SiC particles. Because of higher plasticity (due to frictional heating) under higher applied load (>5 N),
the matrix can also hold SiC particles more effectively. This
fact becomes more dominant especially when the tests are conducted against finer abrasives. At coarser abrasive, the depth of
penetration of the abrasive particle is so high that matrix cannot hold the SiC particles. Under such conditions these debris
may not be spread over the matrix so effectively and removed

easily. All these facts led to slower rate of increase in wear rate
above an applied load of 5 N especially for composite materials
when tested against finer abrasive size, but at coarser abrasive
this effect is less dominating.
In case of alloy, the increase in wear rate may be almost
same for the entire range of applied load. At higher applied load,
because of generation of high heat, the alloy becomes softer than
the composite and during abrasive action; the softer alloy from
the wear groove may also be spreading over the wear surface.
But the depth of cut is so high that significantly larger flakes are
generated and removed from the alloy surface.
When the abrasive size is coarser (60 m), the wear rate
of the alloy and composite is increasing almost linearly with
applied load. This may be attributed to the fact, that the coarser
abrasives are stronger than the finer one and they experienced
higher effective stress even at lower applied load which has been
mentioned earlier. As a result, the surface is subjected to higher
plastic deformation even at lower applied load. The wear rate is
primarily depending on the depth and width of the wear grooves

Fig. 15. SEM micrograph of emery paper (size: 40 m) shows degraded SiC
abrasive (load: 7 N and sliding distance: 108 m).

S. Das et al. / Wear 264 (2008) 4759

57

Table 2
Experimental and calculated (shown in bracket) values of abrasive wear rate of Al alloy and composites at different applied loads
Material

Processing condition

Wear rate of different load


1N

3N

5N

7N

LM13

As cast
Heat-treated

7.80(7.71)
4.50(5.58)

18.71(18.71)
12.29(13.41)

25.65(28.2)
15.60(20.39)

32.47(36.25)
25.15(26.27)

LM1310 wt% SiC

As cast
Heat-treated

4.90(5.2)
2.76(4.05)

12.70(12.48)
8.52(9.02)

19.30(18.85)
19.61(14.71)

25.44(24.51)
18.27(19.25)

LM1315 wt% SiC

As cast
Heat-treated

3.70(3.98)
1.46(3.14)

10.7(9.8)
7.5(7.5)

17.40(16.46)
9.85(11.30)

23.80(20.91)
14.50(14.8)

made by the abrasives and also on the surface and subsurface


cracking. At coarser abrasive size and higher applied load, width
of wear grooves and surface cracking are increased steadily.
Because of relatively higher effective stress and higher level
of frictional heating of the surface, the material removal due
to cutting and ploughing action is increased. While considering the composite materials, it may be found that the fracture
and fragmentation of SiC particles are increased steadily with
applied load. As a result, in majority of the cases the wear rate is
increasing almost linearly with applied load. However, in some
cases composites exhibit comparable wear rate with that of the
alloy.
4.5. Correlation of abrasive wear rate with mechanical
properties, microstructure, applied load and abrasive size
The above discussion thus suggests that the wear behavior of a
material is governed by several factors which include experimental parameters like load and abrasive size; material properties
such as hardness, strength and ductility and microstructural
characteristics like shape, size and volume fraction of the hard
ceramic dispersoid. Assuming, uniform distribution of SiC particle in metallic matrix, good interfacial bonding between SiCp
and matrix and toughness proportional to the product of UTS and
elongation; the wear rate (WR ) of the materials may be expressed
by following empirical equation:

m
d
WR = K
H 0.55 T 0.30 P 0.80 (1 Vf0.53 )
(1)
1+D
where K is a constant which signifies the probability of wear
particle formation, H the hardness of a material (in MPa), P the
applied load (in N), T the product of UTS (in MPa) and elongation (in percentage), Vf the volume fraction of dispersoids, D the
size of reinforcing particle (in m), d the size of the abrasive (in
m) and m is a constant. The calculated values of the wear rate
(shown in the bracket) using the above equation and considering K = 270 1011 are compared with the experimental values
(Table 2).
The value of K and m may be different in different materials. It may also vary with heat treatment, presence of ceramic
dispersoids and the experimental conditions. As in the earlier
discussions, it has been noted that the rate of increase in the
wear rate of alloy or composite does not follow any specific

Fig. 16. Effect of abrasive size on the probability of formation of wear debris
(K) of Al alloy and AlSiC composite.

relationship with abrasive size; one can express the equation for
fixed abrasive size. For example, the abrasive wear rate of alloy
or composite at an abrasive size of 80 m can be expressed by
the following equation:
WR = 270 1011 H 0.55 T 0.30 P 0.80 (1 Vf0.53 )

(2)

Because of the above facts, in few cases significant deviation


from the experimental value is noted. However, in most of the

Fig. 17. Effect of relative abrasive size on the formation of wear debris (K) of
Al alloy and AlSiC composite.

58

S. Das et al. / Wear 264 (2008) 4759

cases the calculated values are in close approximate with experimental values. This suggests the reliability of the above equation
for predicting wear rate of Al-alloys and composites from measured tensile data, hardness at any applied load and abrasive size.
The value of K at different abrasive sizes are calculated using
Eq. (1) and the variation of K as a function of abrasive size is
shown in Fig. 16. It depicts that the formation of wear particle (K
value) is more for larger abrasive size. It may also be noted that
there is sudden increase in the value of K when the abrasive
size is coarser than 60 m. An attempt is also made to understand the effect of relative abrasive size (size of abrasive/size
of reinforcement) on the formation of wear particle (K value).
Fig. 17 shows the effect of relative abrasive size on the value of
K. It indicates that when the relative abrasive size is more than
0.8, there is an abrupt increase in the value of K.
5. Conclusions
(1) The wear rate of composite is less than that of the alloy and
it decreases with increase in SiC content.
(2) The hardness and strength of composite are higher than that
of alloy and they increase with increase in SiC content.
Whereas reverse trend is noted to be true for the ductility
of these materials. The hardness and strength of composite is noted to be more after heat treatment. But in case of
alloy, the hardness and strength are noted to be more when
they are aged for 6 h. This may be attributed to the fact that
composites are aged faster than the alloy.
(3) The abrasive wear rate (WR ) of the materials is a function
of applied load (P), hardness (H), strength and ductility (T)
of the materials, volume fraction of the hard dispersoids
(Vf ) and relative size of abrasive with respect to size of the
dispersoid. The wear rate can be expressed by the following
type of relations:


d
WR = K
1+D

m

H 0.55 T 0.30 P 0.80 (1 Vf0.53 )

This equation clearly demonstrates the effect of size and


volume fraction of reinforcing phase and the size of the
abrasive particle on the wear rate of Al alloy and composites. It suggests that the wear rate of the composite will be
increasing with increasing in size of reinforcing phase and
the composite may be suffered from higher wear rate than
the alloy if the abrasive size is higher than that of reinforcing
phase.
(4) The wear rate of Al alloy and Al composite is invariant to the
sliding distance and increases with increase in applied load
and abrasive size. The effect of abrasive size is noted to be
insignificant when the abrasive size is less than 60 m. Sudden increase in the wear rate was observed when the abrasive
size was more than 60 m. But the wear rate increases
almost linearly with applied load.
(5) Addition of SiC particle and heat treatment provide comparable improvement in wear resistance. Thus, their synergic
effect is noted to be quite significant for improvement the
wear resistance of AlSi alloy.

Acknowledgement
Authors are grateful to Director, RRL, Bhopal, for encouragement and giving permission to publish this paper.
References
[1] H. Warren, J.R. Hunt, Particulate reinforced MMCs, in: Anthony Kelly, Carl
Zweben (Eds.), Comprehensive Composite Materials, vol. 3, Metal Matrix
Composites, in: T.W. Clyne (Vol. Editor), Pergamon, 2000, pp. 701715.
[2] P.K. Rohatgi, R. Asthana, S. Das, Solidification, structure and properties of
cast metalceramic particle composites, Int. Met. Rev. 31 (1986) 115139.
[3] M. Hunt, Aerospace composites, Mater. Eng. 108 (6) (1991) 2730.
[4] M. Noguch, K. Fukizawa, Alternate materials reduce weight in automobiles, Adv. Mater. Process. 143 (6) (1993) 2026.
[5] H.N. Yoshimuro, M. Goncalves, H. Goldenstein, The effect of SiCp clusters
and porosity on the mechanical properties of PM Al matrix composites, Key
Eng. Mater. 127131 (1997) 985992.
[6] D.P. Mondal, S. Das, A.K. Jha, A.H. Yegneswaran, Abrasive wear of Al
alloy-Al2 O3 particle composite: a study on the combined effect of load and
size of abrasive, Wear 223 (1998) 131138.
[7] T. Zeuner, P. Stojanov, P.R. Sahm, H. Ruppert, A. Engels, Developing trends
in disc brake technology for rail application, Mater. Sci. Technol. 14 (1998)
857863.
[8] H. Ahlatci, E. Candan, H. C
imenoglu, Mechanical properties of Al60%
SiCp composites alloyed with Mg, Metall. Mater. Trans. A 35 (7) (2004)
21272141.
[9] S. Das, Development of aluminium alloy composites for engineering applications, Trans. IIM 57 (2004) 325334.
[10] S. Das, D.P. Mondal, O.P. Modi, R. Dasgupta, Influence of experimental
parameters on the erosive-corrosive wear of AlSiC particle composite,
Wear 231 (1999) 195205.
[11] P.K. Rohatgi, Adv. Mater. Technol. Monitor 17 (1990) 147.
[12] M. Roy, B. Venkataraman, V. Bhanuprasad, Y. Mahajan, G. Sundararajan,
The effect of particulate reinforcement on the sliding wear behaviour of
aluminium matrix composites, Metall. Trans. 23A (1992) 28332847.
[13] A.T. Alpas, J. Zhang, Effect of SiC particulate reinforcement on the dry
sliding wear of aluminiumsilicon alloys, Wear 155 (1992) 83104.
[14] S.F. Moustafa, Wear and wear mechanisms of Al-22% Si/Al2 O3f , Wear
185 (1995) 189195.
[15] G. Straffelini, F. Bonollo, A. Tizaiani, A. Molinari, Influence of matrix hardness on the dry sliding behaviour of 20 vol.% Al2 O3 -particulate-reinforced
6061 Al metal matrix composite, Wear 211 (1997) 192197.
[16] I.M. Hutchings, Tribological properties of metal matrix composites, Mater.
Sci. Technol. 10 (1994) 513517.
[17] K.J. Bhansali, R. Mehrabian, Abrasive wear of aluminium matrix composites, J. Met. 34 (1982) 3034.
[18] B.K. Prasad, S. Das, A.K. Jha, O.P. Modi, R. Dasgupta, A.H. Yegneswaran,
Factors controlling the abrasive wear response of a zinc-based alloy silicon
carbide particle composite, Composites A 28A (1997) 301908.
[19] J. Larsen Badse, Influence of size on groove formation during sliding
abrasion, Wear 11 (1968) 213222.
[20] A.G. Wang, I.M. Hutchings, Wear of alumina fiber aluminium metal matrix
composites by two-body abrasion, Mater. Sci. Technol. 5 (1989) 7176.
[21] A.G. Wang, H.J. Rack, Abrasive wear of silicon carbide particulate and
whisker-reinforced 7091 aluminium matrix composites, Wear 146 (1991)
337348.
[22] N. Axen, S. Jacobson, Transitions in the abrasive wear resistance of fibreand particle-reinforced aluminium, Wear 178 (1994) 17.
[23] N. Axen, A. Alahelistan, S. Jacobson, Abrasive wear of alumina fibrereinforced aluminium, Wear 173 (1994) 95104.
[24] S. Das, S. Gupta, D.P. Mondal, B.K. Prasad, Influence of load and abrasive
size on the two body abrasive wear of AlSiC composites, Aluminum
Trans. 2 (2000) 2736.
[25] S. Das, D.P. Mondal, G. Dixit, Mechanical properties of pressure die cast
Al hard part composite, Metall. Mater. Trans. 33A (2001) 633642.

S. Das et al. / Wear 264 (2008) 4759


[26] G.Y. Lee, C.K.H. Dharan, R.O. Ritchie, A physically based abrasive
wear model for composite materials, Wear 252 (34) (2002) 322
331.
[27] A.A. Torrance, The effect of grit size and asperity blunting on abrasive
wear, Wear 253 (2002) 813819.
[28] S. Sawla, S. Das, Combined effect of reinforcement and heat treatment on
the two body abrasive wear of Al alloy and Al particle composite, Wear
257 (56) (2004) 555561.
[29] H.-L. Lee, W.-H. Lu, S. Chan, Abrasive wear of powder metallurgy Al
alloy 6061SiC particle composites, Wear 159 (1992) 223231.
[30] K.S. Al-Rubaie, H.N. Yoshimura, J.D. Biasoli de Mello, Two-body abrasive
wear of AlSiC composites, Wear 233235 (1999) 444454.
[31] I.M. Hutchings, Wear by particulates, Chem. Eng. Sci 42 (40) (1987)
869878.
[32] R.L. Deuis, C. Subramaniun, J.M. Yellup, Abrasive wear of aluminium
compositesa review, Wear 201 (1996) 132144.

59

[33] E. Hornbogen, The role of fracture toughness in the wear of metals, Wear
33 (1975) 251259.
[34] S. Das, The influence of matrix microstructure and particle reinforcement
on the two-body abrasive wear of cast AlSi-alloy composites, J. Mater.
Sci. Lett. 16 (1997) 17571760.
[35] M. Kok, Abrasive wear of Al2 O3 particle reinforced 2024 aluminium
alloy composites fabricated by vortex method, Compos. Part A: Appl. Sci.
Manuf. 37 (3) (2006) 457464.
[36] R.J. Arsenault, The strengthening of aluminum alloy 6061 by fiber and
platelet silicon carbide, Mater. Sci. Eng. 64 (1984) 171181.
[37] R.J. Arsenault, N. Shi, Dislocation generation due to differences between
the coefficients of thermal expansion, Mater. Sci. Eng. 81 (1986) 175187.
[38] T.O. Mulhearn, L.E. Samuels, The abrasion of metals: a model of the
process, Wear 5 (1962) 478498.
[39] H. Sin, N. Saka, N.P. Suh, Abrasive wear mechanisms and the grit size
effect, Wear 55 (1979) 163190.

You might also like