You are on page 1of 13

Available online at www.sciencedirect.

com

Wear 264 (2008) 47–59

Synergic effect of reinforcement and heat treatment on the two body abrasive
wear of an Al–Si 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 Al–Si alloy (BS: LM13) under varying loads and abrasive sizes. Silicon carbide particles with size 50–80 ␮m were
reinforced in Al–Si 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 pistons, brake drums, cylinder heads, connecting rods and drive
shafts for automobile sectors and impellers, agitators, turbine
Al-alloy matrix composites (AMCs) containing hard disper- blade, valves, pump inlet, vortex finder for marine and min-
soids are gaining immense industrial importance because of ing sectors [9–11]. In recent years, fabrication of several such
their excellent combination of physical, mechanical and tri- components from AMCs, reaches to the commercial production
bological properties over base alloys [1]. These include high stage [7–12]. Most of the aforesaid components are subjected to
wear and seizure resistance, high specific strength and stiffness, different kinds of wear and tear related failure. In this context, it
improved high temperature strength, controlled thermal expan- is required to characterize Al-composites in terms of wear under
sion coefficient and high damping capacity [1–8]. It is reported different experimental conditions.
that Al-alloy reinforced with 10 wt% SiC particle composite Sliding wear behavior of AMCs has been studied by many
provides comparable mechanical properties but better thermal investigators [13–16], however, limited attempts have been made
conductivity and specific heat than the cast irons [5,7]. As a to study the abrasive wear characteristics of Al-composites
result, frictional heating of these composites are noted to be sig- [17–27]. According to these reports, wear and seizure resistance
nificantly less than that of cast irons [7]. This leads to the use of of AMCs is considerably higher than that of the base alloys. This
these composites in several automobile and engineering com- is primarily attributed to the fact that the hard dispersoids (rein-
ponents where wear, tear and seizure are the major problems in forcing phase) protect the surface from the destructive action
addition to the weight saving. Some of these components are of the abrasives by reducing the depth of penetration of the
abrasives (hard asperities of the counter surface) and the con-
tact between the abrasive and the matrix. On the other hand,
∗ Corresponding author. Tel.: +91 755 2488562; fax: +91 755 2587042. few investigators have reported a transition of abrasive wear
E-mail address: sdas88@hotmail.com (S. Das). behaviour of composites which was dependent on abrasive size

0043-1648/$ – see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.wear.2007.01.039
48 S. Das et al. / Wear 264 (2008) 47–59

and applied load [19,21,24,28]. According to these investiga- ness and toughness of Al–Si alloy and Al–Si alloy reinforced
tors, above a critical value of load and abrasive size, AMCs with hard particle composites by altering the matrix microstruc-
exhibit higher wear rate than the alloys [24–30]. Under such ture through heat treatment [36,37]. Studies on sliding wear
circumstances, the depth and width of wear track become larger characteristics reported that the wear and seizure resistance of
than the size of dispersoid (i.e., h/d and/or w/d is greater than the alloy or composite are improved by heat treatment [36,37].
unity; where h and w are the average depth and width of wear However, it is interesting to examine the abrasive wear behavior
groove, respectively, and d is the average diameter of the rein- of Al–Si alloy reinforced with hard particles after heat treatment.
forcing phase in the composite). This leads to the scooping off In the present study, an attempt has been made to examine
of the dispersoid from the surface of the composites samples the synergic effect of particle content and heat treatment on
[19,21,24]. Furthermore, it is evident from the literature that the abrasive wear behavior of an Al–Si alloy and Al–Si–SiC
the wearing surface and the subsurface undergo plastic defor- composite under varying applied loads and abrasive sizes.
mation, and this deformation becomes more severe when the
abrasive size is coarser and the applied load is higher [20,29]. 2. Experimental
As a result of such plastic deformation, the hard ceramic dis-
persoid gets fractured and fragmented into finer ones, and/or 2.1. Materials
debonded from the matrix [23–29]. In due course, these frag-
mented particles come out from the specimen surface. Thus the The Al–Si (British Standard: LM13) alloy and LM13 alloy
abrasive wear behavior of composite depends on the material containing 10 wt% and 15 wt% SiC particle have been selected
characteristics like shape, size, distribution and volume fraction in the present investigation. The LM13 alloy nominally con-
of the dispersoids and experimental parameters like applied load tains 11.95% Si, 1.0% Mg, 1.5% Ni, 1.0% Cu, 0.8% Fe, 0.6%
and abrasive size [21–31]. Mn and balance is Al. Silicon carbide particles of size 50–80 ␮m
It has been reported that the wear resistance of composite have been incorporated into the Al–Si alloy melt in varying con-
increases with increase in volume fraction and size of the dis- centration (10 wt% and 15 wt% of the matrix alloy) by vortex
persoids [24–29]. One of the prime factors of the improvement technique. The composite melt was solidified in a cast iron die
in wear resistance is increased in hardness of the Al-alloy due to in the form of circular disc of diameter 100 mm and thickness
the addition of hard dispersoids [24–31] and more protection of 5 mm. The Al-alloy melt was also cast in the same die. The cast
the matrix from the destructive action of the abrasive as the mean alloy and composites were solutionized at 495 ◦ C for 6 h and
free path between the SiC particles is reduced with increase in quenched in water. The samples were tempered at 175 ◦ C for
volume fraction of SiC particle [32]. Several investigators have 6 h followed by air-cooling.
also proposed that wear resistance of a material not only depends
on its hardness and strength but also on its ductility and tough- 2.2. Microstructural examination
ness [33,34]. The reinforcement of Al2 O3 particles in aluminum
alloy enhances the abrasive wear of the matrix. The reinforce- For microstructural observations, samples were mechanically
ment of 16 ␮m Al2 O3 particle strengthens the aluminum matrix polished using standard metallographic techniques, etched with
and enhances the wear resistance. However, the reinforcement of Keller’s reagent and observed in a JEOL 5600 scanning electron
coarse particle (66 ␮m) shows higher wear resistance. The oper- microscope (SEM). Prior to SEM examination, the samples were
ating wear mechanism is mainly consists of plastic deformation sputtered with gold.
of the matrix material [35].
The hardness as well as toughness of a composite mate- 2.3. Two-body abrasive wear testing
rial depends significantly on the matrix microstructure, size and
distribution of the dispersoids and the interfacial bonding char- High stress (two-body) abrasive wear tests of Al alloy and
acteristics [32]. The hardness of the composite increases with Al alloy–SiC composites were conducted on 40 mm × 35 mm
increase in the volume fraction of the dispersoid but at the same × 5 mm rectangular specimens using Suga made Abrasion
time its toughness decreases. Additionally, the casting defects Tester (Model: NUSI, Japan). A schematic view of the apparatus
may increase or the possibility of clustering of dispersoid parti- is shown in Fig. 1. Emery paper embedded with SiC parti-
cle may be more as one increases the dispersoid content. These cles (size: 40 ␮m, 60 ␮m and 80 ␮m) was used as the abrasive
may reduce the wear resistance of the composite [30]. Thus medium. These emery papers were cut into sizes and fixed on
the overall improvement in wear resistance of the alloy, due the wheel (50 mm diameter), which rotates against the speci-
to addition of more dispersoid may not be so high especially men surface. The specimen over the wheel was fixed with a
at severe wearing conditions (i.e., at higher applied load and locking arrangement and load on the specimen was applied with
coarser abrasive size). the help of cantilever mechanism. The specimen was subjected
The mechanical and tribological properties of the compos- to reciprocating motion against the wheel on which the abrasive
ite also depend upon the matrix microstructure; hence, the media is rigidly fixed. The wheel makes one complete revolution
properties of composite can be improved by heat treatment. after each 400 cycles (corresponds to 26 m of sliding distance)
During heat treatment, the matrix of composite behaves almost of reciprocating motion of the specimen. The abrasive paper
similar to that of the base alloy and the dispersoid remains on the wheel was changed after an interval of each 400 cycles
unchanged. Attempts have been made to improve strength, hard- so that the specimen surface always makes contact with fresh
S. Das et al. / Wear 264 (2008) 47–59 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 Keller’s reagent and observed
with SEM.

Fig. 1. Schematic view of Suga made abrasion tester (1) specimen stage; (2) 2.5. Mechanical testing
specimen guide; (3) abrasion wheel: (4) specimen press; (5) metal fitting; (6)
double number detector; (7) motor; (8) fixed weight; (9) weight; (10) weight The hardness of the alloy and composites, in as cast and
scale; (11) lock lever.
heat-treated conditions, is measured using a Vicker’s 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 Al–Si (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 Al–Si (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) 47–59

data. Tensile test was conducted in an INSTRON (Universal graph of the composite shows bonding between SiC particle
Testing Machine) at a strain rate of 10−4 s−1 . and the Al–Si alloy matrix (Fig. 3b). The change in matrix
microstructure of composite due to heat treatment is noted to
3. Results be similar to that observed in the case of heat-treated alloy as
shown in (Fig. 2b). Fig. 3c shows the microstructure of cast
3.1. Microstructure Al–Si–15 wt% SiC composite. It also indicates distribution of
SiC particles in Al matrix. The distribution of SiC particle in
Fig. 2a shows a typical microstructure of cast Al–Si (LM13) Al–Si alloy matrix is not affected by the heat treatment. Fig. 3d
alloy which depicts dendrites of aluminium (marked ‘A’) and shows the microstructure of heat-treated Al–SiC composite
needle shaped eutectic silicon (marked ‘B’) in the interdendritic depicts interface bonding between SiC and Al–Si alloy matrix.
regions and around the dendrites. A higher magnification micro-
graph (Fig. 2b) clearly depicts needle shaped eutectic silicon 3.2. Mechanical properties
(marked ‘B’) in Al matrix. The microstructure of heat-treated
LM13 alloy shows fragmentation of needle-shaped eutectic sili- The hardness, ultimate tensile strength and elongation of
con (in cast condition) into more or less spherical one (Fig. 2c). A LM13 alloy and LM13–SiC composites in cast and heat-treated
higher magnification micrograph (Fig. 2d) clearly shows heat- conditions are shown in Table 1. It indicates that hardness and
treated eutectic silicon (marked ‘A’) and intermetallic phases strength of Al–SiC composite are 25–30% higher (depending on
(arrow marked). SiC content and heat treatment) than that of alloy, respectively.
Fig. 3a represents the microstructure of cast Al–Si alloy rein- But the ductility of composite is recorded to be less than that
forced with 10 wt% SiCp composite showing distribution of SiC of the alloy. It is further noted from Table 1 that all the above
particles in Al–Si alloy matrix. A higher magnification micro- mentioned properties are improved significantly due to heat

Fig. 3. (a) Microstructure of Al–10 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: Al–Si alloy; C: eutectic silicon; D: Al/SiC interface). (c) Microstructure
of cast Al–15 wt% SiC composite showing distribution of SiC particle in Al matrix. (d) Microstructure of heat-treated Al–15 wt% SiC composite shows interface
bonding between SiC particle and Al matrix.
S. Das et al. / Wear 264 (2008) 47–59 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 132 180 1.0


Heat-treated 146 210 3.0
LM13–10 wt% SiC As cast 151 210 1.0
Heat-treated 155 220 1.0
LM13–15 wt% SiC As cast 159 220 1.0
Heat-treated 173 230 1.0

treatment. It is further noted that hardness and strength of and leads to very less material removal in composites. Because
composite increased with increase in SiCp content but the of such low wear rate, the exact contribution from SiCp cannot
ductility of composite is reduced considerably. clearly be understood. Fig. 5 clearly indicates the effect of
SiCp content and heat treatment on the abrasive wear behaviour
3.3. Two body abrasive wear behavior as a function of sliding distance of alloy and composites
under higher load (7 N) and coarser abrasives (80 ␮m). It may
The wear rates of LM13 alloy and LM13–SiC composite are be noted from this figure that the composite exhibits higher
plotted as a function of sliding distance when they are tested wear resistance than the alloy and the wear rate of composite
against different abrasive sizes under varying applied loads. reduces with increase in SiCp content irrespective of cast and
Fig. 4 represents variation of the wear rate of alloy and compos- heat-treated conditions. However, it is interesting to note that the
ite as a function of sliding distance, when tested at 1 N load and wear rate of LM13–10 wt% SiCp composite is marginally less
40 ␮m abrasive size. It is noted that wear rate is almost invariant than that of the alloy. The heat-treated alloy exhibits even less
to the sliding distance irrespective of the material. It has further wear rate than the cast composite reinforced with 10 wt% SiC
been noted that the wear rate of cast composites decreases with particles.
increase in SiCp content. However, after heat treatment, the wear It is evident from Figs. 4 and 5 that the wear rate of each of the
rate of alloy and composite were reduced. On the other hand, the material is increasing with increasing in abrasive size and load.
wear rate of LM13–15 wt% SiCp composite is noted to be more However, the wear rate is not following any specific functional
or less same in heat-treated and in cast conditions. It may be relationship with abrasive size irrespective of material. It is noted
noted that the wear rate of LM13–10 wt% SiCp composite and that the wear rate of the alloy is higher than those of composites.
LM13–15 wt% SiC composite at 1 N load is more or less same. The wear rate of composite containing 15 wt% SiCp is noted to
However, at an applied load of 7 N, appreciable decrease in wear be less than that of composite containing 10 wt% SiCp under
rate was observed in LM13–15 wt% SiC than LM13–10 wt% heat-treated condition. Similar type of plots at loads of 3 N, 5 N
SiC composite. The wear rate of heat-treated LM13–15 wt% and 7 N are shown in Figs. 6–8, respectively. It is evident from
SiCp composite is noted to be 7.0 × 10−11 m3 /m, whereas it Fig. 8 that in most of the cases, the wear rate is almost invariant
is found to be 15 × 10−11 m3 /m in LM13–10 wt% SiCp com- to the sliding distance irrespective of the materials when the
posite. Additionally, when the load is significantly low and the abrasive size is 40 ␮m. Fig. 9 shows the effect of abrasive size
abrasive size is finer, the abrasives are unable to penetrate deep on the wear rate of LM13 alloy and LM13–SiC composite at
into the specimen surface especially in the case of composite. In an applied load of 5 N. It is evident from Fig. 9 that in most of
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 Fig. 5. Wear rate of Al alloy and composite as a function of sliding distance
(applied load: 1 N and abrasive size: 40 ␮m). (applied load: 7 N and abrasive size: 80 ␮m).
52 S. Das et al. / Wear 264 (2008) 47–59

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 spe-
cific functional relationship except for cast LM13 alloy and cast
LM13–10 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 demon-
strates 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 Fig. 10. Wear rate of Al alloy and composite as a function of applied load
(applied load: 7 N and abrasive size: 40 ␮m). (abrasive size: 60 ␮m).
S. Das et al. / Wear 264 (2008) 47–59 53

4. Discussion surface [31–38]. The hardness of the composites is noted to be


higher than that of the alloy and the hardness increases with
4.1. Materials and mechanical properties increase in SiCp content (Table 1). Additionally, the hard SiC
particles in composite act like hard protrusion over the spec-
In the heat treatment of Al–Si (LM13) alloy, the needle- imen surface. As a result, less number of abrasives come in
shaped eutectic silicon is fragmented into more or less spherical contact with the matrix alloy and hence their depth of pene-
one during solution treatment and fine precipitation on the tration is reduced. The abrasive particles, which penetrate into
matrix takes place while ageing treatment. This is attributed the composite surface, interact with the SiC particles during
to reduction in surface energy of the needle shaped silicon reciprocating motion. During this process, the abrasive particles
particles while changing into near spherical ones as the surface may either scoop off from the abrasive media, as the bond-
area to volume ratio is the minimum for spherical shape. This ing between SiC and Al matrix is stronger than the bonding
reduction in surface energy acts as the driving force for such of abrasive particles in the cloth or paper, or get degraded by
transformation. Again, the concentration of Si per unit volume the SiC particles. Fig. 11a shows a typical SEM micrograph
is more at the sharp edges of the needle shaped particle, which of emery paper (particle size: ∼80 ␮m) before wear test. It
led to the existence of concentration gradient between the sharp shows SiC particles with sharp edges embedded on the paper.
edges and the wider portion of the Si needle. As a result, the Fig. 11b shows a SEM micrograph of the emery paper (parti-
silicon from the sharp edges is dissolved and diffused to the cle size : ∼80 ␮m) after the wear test ( load: 5 N and distance:
wider region and in due course of time, it turns into near spher- 81 m). It clearly depicts the degradation of SiC particles and
ical shape. Here, the heat energy applied during heat treatment smearing off fine fragmented particles on the emery paper.
acts as the activation energy for such kind of transformation. In However in some instances, scooping off (marked ‘A’) and
addition, because of the same reasons, there is a possibility of breaking (arrow marked) of the SiC particle are also observed
coarsening of spherical silicon particles with aging time. Hence, (Fig. 11c). As a matter of these facts, the cutting efficiency
it is expected that the spherical silicon particles may be coarser or the severity of the destructive action of abrasive media is
when they are aged for longer duration. Further more, during reduced substantially when they move against the composite
aging, precipitation of alloying elements like Cu, Mg, Ni and Fe surface and finally resulting in significantly lower wear rate of
may take place which might have caused further strengthening the composite materials (Figs. 4–9). The number of SiC parti-
of the Al-alloy. The composite exhibits more hardness and cles on the composite increases with increasing SiCp content,
strength than that of alloy. This is attributed to the fact that by and the hardness of composite is also increased due to addi-
addition of SiC particles in the alloy generates more dislocations tion of more SiC. Furthermore, inter-particle distances are also
because of thermal mismatch stress [38,39]. Such stress is reduced due to increase in SiC content. These, in turn, decrease
originated due to differences in thermal expansion coefficient the wear rate of composite with increase in SiC content. The
of the matrix and the SiC particles [38,39] and makes the matrix improvement in wear resistance of the alloy due to addition
plastically constrained and caused higher dislocation strength- of SiC particles is shown in Figs. 4–9. It is clearly noted that
ening of the matrix. The hard SiC particle also acts as barrier the wear resistance of the alloy is improved by 20% due to
to plastic deformation of the matrix alloy, which results in addition of SiC and other experimental conditions. The improve-
triaxial interaction stress at the interface. This interaction stress ment is reduced with increase in applied load and abrasive
also results in higher strength of the composite. Heat treatment size.
results in formation of spherical Si particles and precipitation of It is a well-established phenomenon that during abrasive
other alloying elements (Figs. 2c and 3b) in terms of complex wear a portion of effective stress caused plastic deformation
intermetallic compounds. This leads to higher strength as of the specimen surface [33,34]. The hard ceramic dispersoids
well as higher toughness of the alloy and composite after heat carry the major portion of this stress and hence the area around
treatment. these ceramic particles undergoes more plastic deformation.
As the matrix in composite is plastically constrained and
4.2. Effect of reinforcement on the abrasive wear the SiC particles are brittle in nature, there are possibilities
of either debonding or fracturing of SiC particles at severe
In two-body abrasive wear, the abrasive particles in the abra- wear conditions (higher load and coarser abrasive size). These
sive media make contact with the specimen surface and penetrate particles, as a consequence, scooped off from the specimen
inside the specimen surface when load is applied. The penetrated surface and may lead to increased wear rate. However, there is
abrasive is subjected to reciprocating motion over the speci- a possibility of picking up of finer fragmented SiC particles or
men surface and thus resulting in formation of wear grooves degraded abrasives in the wear groove which may improve wear
on the surface of the test specimen. These grooves are gen- resistance of the matrix so long they are remained intact on the
erated due to flow or removal of materials either by cutting wear grooves. The plastic deformation becomes significantly
or ploughing action. The depth and width of the grooves and high when wear is occurring under the condition of high
their number on the test specimen surface depend on the shape, applied load and coarser abrasive size. Due to significantly high
size and rake angle of the abrasives, specimen surface char- plastic deformation, the lateral and longitudinal cracks on the
acteristics such as surface roughness and the stability of the specimen surface are generated. As a result, at higher applied
asperities or hard protrusions, and the hardness of the specimen load and coarser abrasive size, the possibility of scooping off
54 S. Das et al. / Wear 264 (2008) 47–59

4.3. Effect of heat treatment

The abrasive wear behaviour of a material depends on fac-


tors 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 rela-
tive 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 15–20% after heat treatment and the wear rate is reduced
by 15–70% depending on test conditions. These facts suggest
that, not only the hardness but also other mechanical proper-
ties 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 Al–Si
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 sur-
face. 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 inter-
action between the SiC particles and load applied on the sample.
Fig. 12 shows a SEM micrograph of the subsurface of the Al–Si
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 Fig. 12. SEM micrograph of subsurface of Al–Si alloy depicting fragmentation
size. of eutectic silicon.
S. Das et al. / Wear 264 (2008) 47–59 55

raiser and caused more surface cracks during abrasive wear of the and protect the specimen from destructive action of the abrasive
cast specimens. Furthermore, the inter-silicon particle distance media [35]. The interaction of these facts causes very marginal
may be higher in cast conditions, and hence, either the abra- variation in wear rate with abrasive size up to a size of 60 ␮m
sive particles may penetrate deeper into the specimen surface (Fig. 9).
or more number of abrasives penetrates into the worn surface. It has been reported that the abrasive wear is also associated
These facts lead to considerably higher wear rate of cast alloy with plastic deformation of the surface and subsurface. During
and composites as compared to that in heat-treated condition. such kind of wear, the abrasive particles make wear grooves
It may be noted that the wear resistance of heat-treated alloy either by cutting or ploughing action over the specimen surface.
improved by 40–70% over the as cast condition depending upon It has been reported that the wear rate depends on w/r ratio,
the applied load and abrasive size. The wear resistance of heat- where w is the groove width and r is the radius of the spheri-
treated composite improved by 20–70% over the cast one. It cal tip of the abrasive particle [32]. The radius of the abrasive
shows that heat treatment also significantly reduced the wear tip is varying with increase in abrasive size. But the width of
rate of the materials. However, the contribution to reduce wear the groove increases substantially with increase in abrasive size.
rate due to particle reinforcement and heat treatment is difficult This may be attributed to the fact that, the effective stresses
to quantify as each of these contributions are varying depending on each individual abrasives increases substantially when the
on test conditions and material. In a nutshell, it demonstrates abrasives become coarser in size, and make it more effective to
qualitatively the synergic effect of particle reinforcement and penetrate deeper into the specimen surface. In addition, degra-
heat treatment. dation of abrasive is reduced when the abrasive size becomes
As the silicon particle become spherical, the points of stress coarser. Because of greater penetration, the surface and subsur-
raiser is reduced considerably and the possibility of fracture face is subjected to more plastic deformation, which causes more
and fragmentation of the silicon particle is reduced. The silicon number of cracks on the surfaces. As a result, the wear rate of
particles remain intact within the Al-matrix. During abrasive material increases rapidly when abrasive size is increased from
wear, a fraction of the matrix material is displaced from the wear 60 ␮m to 80 ␮m in majority of the cases. Fig. 13a shows the wear
grooves and spread along the side of wear grooves in the form surface of an Al–Si alloy abraded using emery paper of abrasive
of flakes. These flaky materials remain intact along the wear size 80 ␮m at an applied load of 7 N. It clearly shows deep and
track for a longer duration or get smeared on the wear surface wide wear grooves and formation of cracks in the longitudinal
because of their higher ductility. Higher ductility of heat-treated as well as in transverse directions. Under the same experimen-
Al–Si alloy is primarily attributed to near spherical shape silicon tal conditions, the subsurface examination indicated severely
particles and results in less surface and subsurface cracking. The deformed region with transverse and longitudinal cracks (shown
flakes formed due to frictional heating adhered and deformation by arrow mark in Fig. 13b).
on the wear surface in due course is delaminated in the form of The plastic contact of the abrasive with the specimen sur-
long flakes and plates. As a matter of this fact, the heat-treated face increases with increase in applied load. This leads to more
materials are exhibiting relatively less wear rate. depth of penetration of the abrasives into the specimen surface
and more plastic deformation of the wear surface and subsur-
4.4. Effect of load and abrasive size face. The w/r ratio may be increasing with increase in applied
load as r remains constant for fixed abrasive size. In case of com-
The abrasive wear behaviour of a material is significantly posites, the SiC particles act as barrier against the penetration
influenced by the combined actions of load and abrasive size. and movement of abrasives when the applied load is less. As a
The efficiency of material removal by an abrasive media depends result w/r ratio may not be changing significantly with applied
on the elastic and plastic contact load [20], which varies with load when the applied load is in the lower range (≤3 N). This
applied load and abrasive size. If the applied load is fixed, then leads to either more or less wear rate or slower rate of increase
the effective stress on individual abrasives increases with coarser in wear rate with applied load when the load is kept less than or
abrasive particles, as the load is shared by less number of abra- equal to 3 N (Fig. 10). When the load increases above 3 N, the
sives. When the abrasive particles are finer in size, they make surface and subsurface is subjected to severe plastic deforma-
only elastic contact with the test specimen surface, as the effec- tion in addition to cutting and ploughing of material. Under these
tive stress in individual abrasive is less. As a result, these abrasive circumstances a large number of silicon particles and a few SiC
particles only support the applied load without contributing suf- particles of composite get fractured or fragmented and also the
ficient material removal. However, at higher load regime, the w/r ratio increases significantly. It led to debonding and scoop-
effective stress on each individual abrasive particles reach to a ing off of SiC particles from the matrix, in addition to removal
level where the abrasives make plastic contact with the spec- of matrix material by cutting and ploughing actions. Severe sur-
imen surface and causing more surface damage even at finer face cracking also facilitates more surface damage and reduces
abrasive size. Thus the overall rate of material removal depends the possibility of picking up of detached abrasive particles and
on the extent of plastic contact of the abrasives with the specimen the flaky debris in the wearing surface. These facts as a whole
[30]. Further more, coarser size abrasives generally contain large leads to significantly rapid increase in wear rate especially of
number of flows and hence may break easily [35]. As a result, composite at intermediate load level (3–5 N).
cutting efficiency of the abrasive media is reduced. These bro- At higher applied load, severe plastic deformation of the sur-
ken abrasives are also picked up by the wear surface more easily face and subsurface causes heating of the wear surface which
56 S. Das et al. / Wear 264 (2008) 47–59

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

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
Fig. 13. (a) SEM micrograph shows deep and wide grooves and formation of of the alloy and composite is increasing almost linearly with
longitudinal and transverse cracks (load: 7 N; abrasive size: 80 ␮m). (b) SEM applied load. This may be attributed to the fact, that the coarser
micrograph showing subsurface depicting severe deformation and formation abrasives are stronger than the finer one and they experienced
of cracks in longitudinal and transverse directions (A: deformed region; B: higher effective stress even at lower applied load which has been
undeformed region).
mentioned earlier. As a result, the surface is subjected to higher
plastic deformation even at lower applied load. The wear rate is
makes the alloy or the matrix of composite more ductile. As a primarily depending on the depth and width of the wear grooves
consequence, major portion of the material from the wear sur-
face 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 sub-
jected 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 (espe-
cially 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 experi-
ment conducted at a load of 7 N and sliding distance of 108 m. It
clearly depicts degraded SiC particles. Because of higher plastic-
ity (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 con-
ducted against finer abrasives. At coarser abrasive, the depth of
penetration of the abrasive particle is so high that matrix can-
not hold the SiC particles. Under such conditions these debris Fig. 15. SEM micrograph of emery paper (size: 40 ␮m) shows degraded SiC
may not be spread over the matrix so effectively and removed abrasive (load: 7 N and sliding distance: 108 m).
S. Das et al. / Wear 264 (2008) 47–59 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 7.80(7.71) 18.71(18.71) 25.65(28.2) 32.47(36.25)


Heat-treated 4.50(5.58) 12.29(13.41) 15.60(20.39) 25.15(26.27)
LM13–10 wt% SiC As cast 4.90(5.2) 12.70(12.48) 19.30(18.85) 25.44(24.51)
Heat-treated 2.76(4.05) 8.52(9.02) 19.61(14.71) 18.27(19.25)
LM13–15 wt% SiC As cast 3.70(3.98) 10.7(9.8) 17.40(16.46) 23.80(20.91)
Heat-treated 1.46(3.14) 7.5(7.5) 9.85(11.30) 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 consider-
ing 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 Fig. 16. Effect of abrasive size on the probability of formation of wear debris
(K) of Al alloy and Al–SiC composite.
The above discussion thus suggests that the wear behavior of a
material is governed by several factors which include experimen- relationship with abrasive size; one can express the equation for
tal parameters like load and abrasive size; material properties fixed abrasive size. For example, the abrasive wear rate of alloy
such as hardness, strength and ductility and microstructural or composite at an abrasive size of 80 ␮m can be expressed by
characteristics like shape, size and volume fraction of the hard the following equation:
ceramic dispersoid. Assuming, uniform distribution of SiC par-
ticle in metallic matrix, good interfacial bonding between SiCp WR = 270 × 10−11 H −0.55 T −0.30 P 0.80 (1 − Vf0.53 ) (2)
and matrix and toughness proportional to the product of UTS and Because of the above facts, in few cases significant deviation
elongation; the wear rate (WR ) of the materials may be expressed from the experimental value is noted. However, in most of the
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 elonga-
tion (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 consider-
ing K = 270 × 10−11 are compared with the experimental values
(Table 2).
The value of K and m may be different in different materi-
als. 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 Fig. 17. Effect of relative abrasive size on the formation of wear debris (K) of
wear rate of alloy or composite does not follow any specific Al alloy and Al–SiC composite.
58 S. Das et al. / Wear 264 (2008) 47–59

cases the calculated values are in close approximate with experi- Acknowledgement
mental values. This suggests the reliability of the above equation
for predicting wear rate of Al-alloys and composites from mea- Authors are grateful to Director, RRL, Bhopal, for encour-
sured tensile data, hardness at any applied load and abrasive size. agement and giving permission to publish this paper.
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 References
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 [1] H. Warren, J.R. Hunt, Particulate reinforced MMCs, in: Anthony Kelly, Carl
Zweben (Eds.), Comprehensive Composite Materials, vol. 3, Metal Matrix
size is coarser than 60 ␮m. An attempt is also made to under- Composites, in: T.W. Clyne (Vol. Editor), Pergamon, 2000, pp. 701–715.
stand the effect of relative abrasive size (size of abrasive/size [2] P.K. Rohatgi, R. Asthana, S. Das, Solidification, structure and properties of
of reinforcement) on the formation of wear particle (K value). cast metal–ceramic particle composites, Int. Met. Rev. 31 (1986) 115–139.
Fig. 17 shows the effect of relative abrasive size on the value of [3] M. Hunt, Aerospace composites, Mater. Eng. 108 (6) (1991) 27–30.
K. It indicates that when the relative abrasive size is more than [4] M. Noguch, K. Fukizawa, Alternate materials reduce weight in automo-
biles, Adv. Mater. Process. 143 (6) (1993) 20–26.
0.8, there is an abrupt increase in the value of ‘K’. [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
5. Conclusions Eng. Mater. 127–131 (1997) 985–992.
[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
(1) The wear rate of composite is less than that of the alloy and
size of abrasive, Wear 223 (1998) 131–138.
it decreases with increase in SiC content. [7] T. Zeuner, P. Stojanov, P.R. Sahm, H. Ruppert, A. Engels, Developing trends
(2) The hardness and strength of composite are higher than that in disc brake technology for rail application, Mater. Sci. Technol. 14 (1998)
of alloy and they increase with increase in SiC content. 857–863.
Whereas reverse trend is noted to be true for the ductility [8] H. Ahlatci, E. Candan, H. Çimenoglu, Mechanical properties of Al–60%
SiCp composites alloyed with Mg, Metall. Mater. Trans. A 35 (7) (2004)
of these materials. The hardness and strength of compos-
2127–2141.
ite is noted to be more after heat treatment. But in case of [9] S. Das, Development of aluminium alloy composites for engineering appli-
alloy, the hardness and strength are noted to be more when cations, Trans. IIM 57 (2004) 325–334.
they are aged for 6 h. This may be attributed to the fact that [10] S. Das, D.P. Mondal, O.P. Modi, R. Dasgupta, Influence of experimental
composites are aged faster than the alloy. parameters on the erosive-corrosive wear of Al–SiC particle composite,
Wear 231 (1999) 195–205.
(3) The abrasive wear rate (WR ) of the materials is a function
[11] P.K. Rohatgi, Adv. Mater. Technol. Monitor 17 (1990) 1–47.
of applied load (P), hardness (H), strength and ductility (T) [12] M. Roy, B. Venkataraman, V. Bhanuprasad, Y. Mahajan, G. Sundararajan,
of the materials, volume fraction of the hard dispersoids The effect of particulate reinforcement on the sliding wear behaviour of
(Vf ) and relative size of abrasive with respect to size of the aluminium matrix composites, Metall. Trans. 23A (1992) 2833–2847.
dispersoid. The wear rate can be expressed by the following [13] A.T. Alpas, J. Zhang, Effect of SiC particulate reinforcement on the dry
sliding wear of aluminium–silicon alloys, Wear 155 (1992) 83–104.
type of relations:
[14] S.F. Moustafa, Wear and wear mechanisms of Al-22% Si/Al2 O3f , Wear
 m 185 (1995) 189–195.
d
WR = K H −0.55 T −0.30 P 0.80 (1 − Vf0.53 ) [15] G. Straffelini, F. Bonollo, A. Tizaiani, A. Molinari, Influence of matrix hard-
1+D ness on the dry sliding behaviour of 20 vol.% Al2 O3 -particulate-reinforced
6061 Al metal matrix composite, Wear 211 (1997) 192–197.
[16] I.M. Hutchings, Tribological properties of metal matrix composites, Mater.
This equation clearly demonstrates the effect of size and
Sci. Technol. 10 (1994) 513–517.
volume fraction of reinforcing phase and the size of the [17] K.J. Bhansali, R. Mehrabian, Abrasive wear of aluminium matrix compos-
abrasive particle on the wear rate of Al alloy and compos- ites, J. Met. 34 (1982) 30–34.
ites. It suggests that the wear rate of the composite will be [18] B.K. Prasad, S. Das, A.K. Jha, O.P. Modi, R. Dasgupta, A.H. Yegneswaran,
increasing with increasing in size of reinforcing phase and Factors controlling the abrasive wear response of a zinc-based alloy silicon
carbide particle composite, Composites A 28A (1997) 301–908.
the composite may be suffered from higher wear rate than
[19] J. Larsen Badse, Influence of size on groove formation during sliding
the alloy if the abrasive size is higher than that of reinforcing abrasion, Wear 11 (1968) 213–222.
phase. [20] A.G. Wang, I.M. Hutchings, Wear of alumina fiber aluminium metal matrix
(4) The wear rate of Al alloy and Al composite is invariant to the composites by two-body abrasion, Mater. Sci. Technol. 5 (1989) 71–76.
sliding distance and increases with increase in applied load [21] A.G. Wang, H.J. Rack, Abrasive wear of silicon carbide particulate and
and abrasive size. The effect of abrasive size is noted to be whisker-reinforced 7091 aluminium matrix composites, Wear 146 (1991)
337–348.
insignificant when the abrasive size is less than 60 ␮m. Sud- [22] N. Axen, S. Jacobson, Transitions in the abrasive wear resistance of fibre-
den increase in the wear rate was observed when the abrasive and particle-reinforced aluminium, Wear 178 (1994) 1–7.
size was more than 60 ␮m. But the wear rate increases [23] N. Axen, A. Alahelistan, S. Jacobson, Abrasive wear of alumina fibre-
almost linearly with applied load. reinforced aluminium, Wear 173 (1994) 95–104.
(5) Addition of SiC particle and heat treatment provide compa- [24] S. Das, S. Gupta, D.P. Mondal, B.K. Prasad, Influence of load and abrasive
size on the two body abrasive wear of Al–SiC composites, Aluminum
rable improvement in wear resistance. Thus, their synergic Trans. 2 (2000) 27–36.
effect is noted to be quite significant for improvement the [25] S. Das, D.P. Mondal, G. Dixit, Mechanical properties of pressure die cast
wear resistance of Al–Si alloy. Al hard part composite, Metall. Mater. Trans. 33A (2001) 633–642.
S. Das et al. / Wear 264 (2008) 47–59 59

[26] G.Y. Lee, C.K.H. Dharan, R.O. Ritchie, A physically based abrasive [33] E. Hornbogen, The role of fracture toughness in the wear of metals, Wear
wear model for composite materials, Wear 252 (3–4) (2002) 322– 33 (1975) 251–259.
331. [34] S. Das, The influence of matrix microstructure and particle reinforcement
[27] A.A. Torrance, The effect of grit size and asperity blunting on abrasive on the two-body abrasive wear of cast Al–Si-alloy composites, J. Mater.
wear, Wear 253 (2002) 813–819. Sci. Lett. 16 (1997) 1757–1760.
[28] S. Sawla, S. Das, Combined effect of reinforcement and heat treatment on [35] M. Kok, Abrasive wear of Al2 O3 particle reinforced 2024 aluminium
the two body abrasive wear of Al alloy and Al particle composite, Wear alloy composites fabricated by vortex method, Compos. Part A: Appl. Sci.
257 (5–6) (2004) 555–561. Manuf. 37 (3) (2006) 457–464.
[29] H.-L. Lee, W.-H. Lu, S. Chan, Abrasive wear of powder metallurgy Al [36] R.J. Arsenault, The strengthening of aluminum alloy 6061 by fiber and
alloy 6061–SiC particle composites, Wear 159 (1992) 223–231. platelet silicon carbide, Mater. Sci. Eng. 64 (1984) 171–181.
[30] K.S. Al-Rubaie, H.N. Yoshimura, J.D. Biasoli de Mello, Two-body abrasive [37] R.J. Arsenault, N. Shi, Dislocation generation due to differences between
wear of Al–SiC composites, Wear 233–235 (1999) 444–454. the coefficients of thermal expansion, Mater. Sci. Eng. 81 (1986) 175–187.
[31] I.M. Hutchings, Wear by particulates, Chem. Eng. Sci 42 (40) (1987) [38] T.O. Mulhearn, L.E. Samuels, The abrasion of metals: a model of the
869–878. process, Wear 5 (1962) 478–498.
[32] R.L. Deuis, C. Subramaniun, J.M. Yellup, Abrasive wear of aluminium [39] H. Sin, N. Saka, N.P. Suh, Abrasive wear mechanisms and the grit size
composites—a review, Wear 201 (1996) 132–144. effect, Wear 55 (1979) 163–190.

You might also like