Вы находитесь на странице: 1из 7

Wear 257 (2004) 555561

Combined effect of reinforcement and heat treatment on the two body


abrasive wear of aluminum alloy and aluminum particle composites
S. Sawla, S. Das
Regional Research Laboratory (CSIR), Near Habibganj Naka, Bhopal 462-026, MP, India
Received 16 October 2003; received in revised form 14 January 2004; accepted 17 February 2004

Abstract
An attempt has been made to study the two body abrasive wear behavior of LM13 alloy and LM1315 wt.% SiC composite, in cast and
heat-treated conditions, as a function of applied load. The wear constant (K) was calculated based on the wear rate data, which signifies
the probability of formation of wear particles during abrasive wear process. It was observed that the wear constant decreases with load. In
the case of cast alloy the value of wear constant was higher than that of the heat-treated alloy and composite. The wear surface and the
subsurface were studied using scanning electron microscope (SEM). The wear surface and subsurface studies indicated that at low load
regime (1 N) the fragmentation of the wear surface is more by cracking, however, at high load regime plastic deformation is dominated.
During the wear process, the cracks are mainly nucleated at the Al/Si and Al/SiC interfaces and joining the cracks forming the wear
debris. Heat-treated alloy and composite showed better strength and hardness, resulted in less propensity for crack nucleation and showed
enhancement in wear resistance. The subsurface studies showed plastic deformation and formation of mechanically mixed layer consisting
of SiC particle, silicon and deformed Al. The subsurface deformation clearly depicts the propagation of cracks in longitudinal as well as in
transverse directions in low load condition. However at high load, the SiC particles are seen embedded in the plastically deformed matrix.
Pullout of SiC particles from the Al matrix was not observed.
2004 Published by Elsevier B.V.
Keywords: AlSi alloy; Abrasive wear; AlSiC composite; Effect of load; Subsurface studies; Wear constant

1. Introduction
Aluminum matrix composites (AMCs) are being used as a
suitable lightweight and high abrasion resistance material in
place of cast iron and steel in automobile, mining and mineral sectors [14]. Because of lightweight, low coefficient
of thermal expansion and high abrasion resistance properties, AMCs are found in few marked applications [5,6].
Additionally, aluminum matrix composites have excellent
combination of properties such as high hardness, strength
and stiffness as compared to the base alloy [7,8]. Abrasive
wear behavior of various Al alloys, such as AlMg [9,10],
AlCu [11,12] and AlZnMg [13,14], reinforced with hard
particle has been studied and found that the wear rates of
these hard particle composites are significantly lower than
the wear rates of corresponding base alloys. Prasad et al.
[1517] have studied zinc alloy reinforced with SiC particle
composite and stated that subsurface work hardening, microcracking tendency, fragmentation/removal of dispersed
Corresponding author. Fax: +91-755-587-042.
E-mail address: sdas@rrlbpl.org (S. Das).

0043-1648/$ see front matter 2004 Published by Elsevier B.V.


doi:10.1016/j.wear.2004.02.001

phase, clogging, shelling and attrition controlled the wear


of material.
In general, the reinforced hard particles protect the matrix surface against destructive action of the abrasive during
wear process [1821]. The abrasive wear behavior of composites influenced by several factors like shape and content
of reinforcement [22,23], size [22,2426] and rack angle of
the abrasive [27]. It was reported that reinforced particles
generate high stress concentration around the particles that
influenced the flow stress and wear rate of material [28].
It was also noted that the highest wear resistance was obtained when the microstructure was associated with fine,
well-dispersed semi-coherent particles in metallic matrix.
The influence of carbide on the wear resistance depended on
their hardness relative to the matrix hardness [28]. Banerjee et al. [29] showed that addition of zircon particle (size:
100 m) in a series of Al alloy improves the high stress abrasive wear resistance and abrasive wear resistance increases
linearly with the volume fraction of zircon. This study indicated that the abrasion resistance increased possibly due to
blunting of alumina particle of the abrasive cloth. It was also
reported that the low stress abrasive wear rate decreased by

556

S. Sawla, S. Das / Wear 257 (2004) 555561

a factor of 5 due to the addition of 0.35 volume fraction of


zircon particle [29].
Composite containing a high volume fraction of hard particles exhibited high wear resistance [11,30]. Turenne [31]
was reported that the wear resistance of composite increases
with decreased abrasive particle size. It was also reported
that in case of Al-6061SiC composites, containing 20 volume fraction of reinforcement, the transition in the wear
behavior was observed when the size of abrasive is than
20 m [9]. It was mentioned that the wear mechanism is
changed from chipping and removal of reinforcement to
more a smearing of aluminum over the SiC particle. It was
concluded that at lower size abrasives, the wear behavior
is considered as similar to low stress abrasion resulting in
an improved abrasion resistance [10]. Further, Wang and
Hutching [10] reported that coarse abrasive particles and
high volume fraction of reinforcement resulting in decreased
wear resistance. This is attributed to breakage and pull out
of reinforcement during high stress abrasive wear. The possibility of removal of reinforcing phase from the composite
surface depends on the relative size of abrasive with respect
to the reinforcement size. It was also observed in case of
AlSi30 wt.% zircon particles composites (size: 5090 m)
that the wear rate increases with the sliding distance which
is caused mainly due to formation of microcracks and fracturing tendency of reinforcement [32]. Axen and Zum Gahr
[33] reported that the wear resistance of TiC reinforced steel
is governed by the size of the reinforcement, size of the abrasive and matrix hardness. Further, Wang and Rack [14] proposed a term relative penetration depth which was defined
as the ratio of penetration depth of the abrasive (h) to size
of the reinforcing phase (d) in SiC reinforced Al composites. It was reported that the ratio h/d was a critical parameter to control the wear resistance of composites. If the ratio
decreases below unity, the wear resistance increases significantly but above unity the wear resistance is independent
of the penetration depth. If the hardness of abrasive is more
than reinforcing phase then there would be an increase in
the penetration depth and the wear of material is determined
by the removal and fracture of reinforcing phase [34]. The
shape of abrasive particles is also influenced the wear rate
of materials; the angular-shaped abrasive particle enhances
the wear rate [34].
It was mentioned that the particle-reinforced composites
offered better wear resistance than fiber-reinforced one [35].
This happens mainly because of poor fibermatrix interfacial bonding that dictates the crack nucleation and propagation at the interface. Silicon carbide particle-reinforced
metal composites exhibit less wear rate due to good interfacial bonding between reinforcement and metallic matrix
causes less propensity for nucleation of microcracks. Several investigators have studied the effect of heat treatment
on the abrasive wear behavior of Al alloy hard particle composites [10,13,14,36]. In general, it is stated that the heat
treatment improves the mechanical properties and enhances
the wear resistance. However, Lin and Liu [13] studied the

abrasion resistance of AlZnMgSiC composite and suggested that ageing temperature and duration are the dominant factors controlling the abrasion rate rather than ageing
hardness. Overageing of composite results in achieving the
best abrasion resistance, which explained by the anchoring
and contraction mechanisms [13].
In the present investigation, an attempt has been made to
study the combined effect of heat treatment and reinforcement of SiC particles on the two body abrasive wear behavior
of AlSi (LM13) alloy. The heat treatment of AlSi alloy alters the morphology of eutectic Si from coarse plate-shaped
to more or less spherical. The spherical shape eutectic silicon
improves the mechanical properties of AlSi alloy in addition to the strengthening effect due to precipitation. The wear
surface and subsurface deformation of cast and heat-treated
alloy and composite were studied in order to ascertain the
wear mechanism during two body abrasive wear. Further,
the wear mechanism is correlated with probability of formation of wear particles.
2. Experimental
2.1. Materials
AlSi (BS: LM13) alloy was used as the matrix material. The alloy contains 11.00 wt.% Si, 1.00 wt.% Mg,
1.50 wt.% Ni, 1.00 wt.% Cu, 0.80 wt.% Fe, 0.50 wt.% Mn
and balance was Al. LM1315 wt.% SiCp composite was
prepared by dispersing hard particles in aluminum matrix
using stir-casting technique. The steps involved in preparing
the composite were melting the alloy, stirring the melt using
a mechanical stirrer, dispersing the silicon carbide particles
in the vortex of the melt and solidifying the composite melt
in a predetermined cast iron mold in the form of circular
disc (diameter: 100 mm; thickness: 5 mm). The silicon carbide particles (size: 5080 m) were used as reinforcement
for synthesizing LM1315 wt.% SiC composite. AlSi alloy
melt was also cast in the same mold.
2.2. Heat treatment
The cast alloy and composite were heat-treated following a cycle consisting of solutionizing at 490 C for 8 h.
quenched in water (at 35 C) followed by tempering at
170 C for 6 h and finally cooled in ambient air.
2.3. Microstructure
For microstructural investigation, the cast and heat-treated
alloy and composite were mechanically polished using standard metallographic practices, etched with Kellers reagent
and observed in scanning electron microscope (SEM). For
wear surface studies, the specimens were cut from the
sample and examined in SEM. For subsurface deformation
studies, samples were sectioned perpendicular to the wear

S. Sawla, S. Das / Wear 257 (2004) 555561

surface and the cross-section was metallographically polished, etched and observed in SEM. All these samples were
gold sputtered prior to SEM examination.
2.4. Abrasive wear test
Two body abrasive wear tests were conducted on
40 mm 35 mm 5 mm rectangular specimens using a
Suga Abrasion Tester (Model: NUSI, Japan) at different applied loads (1 7 N). Emery papers, containing SiC particles
(size: 80 m) were cut into exact dimensions and fixed on a
wheel (diameter: 50 mm; thickness: 12 mm), to serve as the
abrasive medium. The specimen was fixed with a locking
arrangement against the abrasive medium. The load over the
specimen was applied through cantilever mechanism. The
wear rates were calculated by weight loss methods. The loss
of weight of the specimen before and after each test was
measured using a microbalance. The details of the test procedure are given elsewhere [37]. It is observed that the abrasive
particles, in the emery paper, blunt due to abrasive action.
2.5. Hardness
The hardness of cast and heat-treated alloy and composite was measured in HV scale using a Vickers hardness
tester at an applied load of 2.5 kg. The specimens were metallographically polished and opposite sides were made perfectly parallel before the hardness measurement. Ten hardness readings were taken on each sample and the average
value was reported.

557

Table 1
Hardness of LM13 alloy and LM1315 wt.% SiCp Composite in as-cast
and heat-treated conditions
Sl. no.

Material

Hardness
value

1
2
3
4

LM13 alloy (as-cast)


LM13 alloy (heat-treated)
LM1315 wt.% SiCp composite (as-cast)
LM1315 wt.% SiC composite (heat-treated)

130
146
159
173

of alloy is improved by 22% due to reinforcement of 15 wt.%


SiC particle in aluminum matrix. An improvement of hardness by 33% is observed due to combined effect of heat
treatment and reinforcement of 15 wt.% of SiC particles.
3.3. Wear constant as a function of load
It is known from the basic wear law [38] that the volume
loss of the material (V), during wear process, is directly
proportional to the sliding distance (D) and applied load (L)
and inversely proportional to the hardness of the material
(Hv ). Hence the wear relationship can be written as
V

DL
,
Hv

V =K

DL
Hv

(1)

where K is the wear constant. The physical significance of


K, in this relation, is the probability of formation of wear
particles:
K=

V
Hv
DL

since

V
= W (wear rate)
D

Therefore,
3. Results
3.1. Microstructure
In general, the microstructure of alloy, in cast condition, shows aluminum dendrites and eutectic silicon (sharp
edges plate-shaped) in the inter-dendritic regions and around
the dendrites. The microstructure of alloy, in heat-treated
condition, depicts smooth edges near-spherical-shaped silicon in aluminum matrix. During the heat treatment, the
sharp edges plate-shaped silicon is changed to smooth edges
near-spherical-shaped silicon.
The micrograph of composite, in cast condition, depicts
uniform distribution of silicon carbide particles in Al matrix.
The interfacial investigation shows good interface bonding
between the aluminum matrix and silicon carbide particles
and eutectic silicon is observed at the matrix/particle interface. The heat-treated micrograph also shows near spherical
eutectic silicon at the SiC/Al interface.
3.2. Hardness
The hardness values of alloy and composite, in cast and
heat-treated conditions, are shown in Table 1. The hardness

K=

W
Hv
L

Hence K is defined as the wear rate per unit load multiplied by hardness of the material. Fig. 1 shows the wear
constant (K) of alloy and composite as a function of load
at a sliding distance of 108 m. It shows that the wear constant (K) of cast alloy decreased drastically from a value
of 12.0 104 to 8.5 104 when the applied load is increased from 1 to 3 N. Further increasing the load up to 7 N,
the value of K decreases monotonically. For instance, at an
applied load of 5 N the value of K is found to be around
6.8 104 and it is reduced to 6.2 104 when the load
is increased to 7 N. However, in the case of heat-treated alloy and cast composite, the wear constant (K) is found to
be around 8.0 104 at an applied load of 1 N and it decreased to around 6.2 104 when the load is increased to
3 N. Further, increases the load to 7 N the wear constant (K)
is found to be around 5.8 104 . In the case of heat-treated
composite, the value of K is found to be around 6.4 104
and is reduced to 5.0 104 when the load is increased
from 1 to 3 N. Further, increasing the load to 5 and 7 N, the
values of K are reduced to 4.0 104 and 3.5 104 , receptively. This figure clearly indicates that the probability

558

S. Sawla, S. Das / Wear 257 (2004) 555561

of formation of wear particles is higher at low load regime


and it reduces as the applied load increases. The probability
of formation of wear particles in heat-treated composite is
found to be less amongst all.
3.4. Wear surface
The wear surface of cast alloy and composite was studied using SEM. The wear surface shows continuous groove

Fig. 1. The wear constants (K) of cast and heat-treated LM13 alloy and
LM1315 wt.% SiCp composite as a function of load, after a sliding
distance of 108 m (abrasive size: 80 m).

Fig. 2. Scanning electron micrograph of wear surface (a) cast alloy at an


applied load of 1 N shows propagation of cracks along the longitudinal
and transverse directions and formation of wear debris by joining of
cracks; (b) cast alloy at an applied load of 7 N shows continuous grooves
and plastically deformed region.

Fig. 3. Scanning electron micrographs of subsurface deformation (a) cast


alloy at an applied load of 1 N depicts fragmentation of eutectic silicon
in highly deformed region and propagation of cracks; (b) cast composite
at an applied load of 1 N depicting a SiC particle entrapped in subsurface
region; (c) cast composite at an applied load of 7 N shows a SiC particle
embedded in plastically deformed layer.

S. Sawla, S. Das / Wear 257 (2004) 555561

formation, damaged regions and cracks propagation along


the longitudinal and transverse directions. Fig. 2(a) shows
a typical wear surface of cast alloy, at an applied load of
1 N. It clearly shows wear grooves and the cracks propagate
along the longitudinal and transverse directions. It is noted
that longitudinal and transverse cracks join together (shown
by arrow marks) to form wear particles. Fig. 2(b) shows the
wear surface of alloy at an applied load of 7 N. It shows continuous grooves and plastically deformed regions. It is seen
that the cracking tendency is less (in Fig. 2(b)) as compared
to Fig. 2(a). SEM observation of cast and heat-treated composites is also revealed similar surface features as observed
in Fig. 2(a) and (b). It stems from the above wear surface
observation that at low load regime (1 N) the tendencies
of cracking and fracturing of wear surface are more. However, at high load regime the wear surface is prone to have
more of plastically deformed rather than cracking.
3.5. Subsurface
A typical subsurface of cast alloy, at an applied load of
1 N, is shown in Fig 3(a). It shows deformed region and the
depth of deformation is measured to be around 50 m. The
deformed region also shows fragmentation of eutectic silicon (arrow marked) and cracks propagating along the longitudinal as well as in the transverse directions. During the
sliding action, the cracks propagate and join together to form
the wear debris. While increasing the applied load to 7 N,
the depth of deformation is measured to be around 70 m. It
also shows cracks and fragmentation of silicon particles in
the deformed region. Fig. 3(b) shows a typical subsurface of
cast composite, at an applied load of 1 N. It shows a SiC particle entrapped in subsurface deformed region. The cracks
propagating along the longitudinal and transverse directions
are also seen. Fig. 3(c) shows subsurface of cast composite at an applied load of 7 N. It is noted that the degree of
cracking is less in Fig. 3(c) as compared to Fig. 3(b). It also
shows silicon carbide particle embedded in the plastically
deformed aluminum matrix.

4. Discussion
The phases present in cast alloy are aluminum in dendritic morphology and coarse faceted eutectic silicon. The
eutectic silicon is seen in the inter-dendritic regions and
around the dendrites. The morphology of eutectic silicon is
sharp edges plate-shaped and it is changed to smooth edges
near-spherical-shaped due to heat treatment. In the case of
composite, apart from aluminum and silicon, uniform distribution of SiC particle is observed in AlSi alloy matrix.
During solidification of AlSiSiC composite melt, the SiC
particles are pushed by the aluminum dendrites into the last
freezing eutectic liquid and thus SiC particles are seen surrounded by the eutectic silicon. It is revealed good interfacial bonding between aluminum matrix and SiC particles.

559

In an AlSi alloy system, hard Si phase is distributed in


soft Al matrix. AlSi eutectic, in Al melt, not only imparts
the fluidity but it also enhances the wear properties. During
abrasive wear process, the wear of material depends primarily upon nucleation and propagation of cracks on the surface and subsurface regions. In general, the available site
for crack nucleation in AlSi alloy is the Al/Si interface. In
the initial stage of abrasive wear process, silicon particle resists against destructive action of the abrasive and protects
the surface. In cast alloy, the sharp edges eutectic silicon
(sites for high stress concentration) act as sites for crack nucleation. During wear process, these cracks are propagated
and later joined together to form flake-shaped wear debris.
Additionally, the hard silicon particles tend to fragment into
small pieces and separated out from the matrix. Moreover,
the fragmentation of silicon particles is taking place when
the stress applied on the Si particle is more than the fracture
strength of silicon. Heat-treated alloy showed lower value of
wear constant, K (i.e. less probability of formation of wear
particle) than that of the cast alloy. This happens mainly due
to less propensities for crack nucleation, at the Al/Si interface, in heat-treated alloy as compared to the cast alloy. In
heat-treated alloy, the wear surface is covered with smaller
size near-spherical silicon particles, which in turn protect
the surface from wear. In abrasive wear of heat-treated alloy, the material removes from the wear surface mainly due
to ploughing action by the abrasive particles. During the
ploughing action, material below the abrasives is plastically
deformed and then accumulated on both sides of the groove
and the material is removed from the wear surface as fine debris and causes lower wear constant (K) than that of the cast
alloy. Moreover, heat-treated alloy forms a mating surface of
silicon particles over the aluminum matrix and protects the
matrix most effectively from severe contact of abrasive and
results in shallower and finer grooves in the wear surface.
However, in the case of cast alloy, the material is subjected
to cutting type of wear. The material accumulated around
the groove, deformed plastically and subsequent wear process, it detached from the wear surface by nucleation and
propagation of the cracks. It stems from the above facts that
the AlSi alloy exhibits two types of abrasive wear depending upon the morphology of silicon; one is by ploughing
action (heat-treated alloy) and the other one is by the combination of cutting and ploughing action (cast alloy). The
transition from ploughing type of wear (heat-treated alloy)
to cutting type (cast alloy) depends on silicon morphology.
Heat-treated AlSi alloy shows some ductility (5%), wears
out essentially by ploughing action and the cast alloy which
is brittle in nature (ductility <1%) undergoes wear mainly
by cutting action (crack nucleation and propagation).
It is reported that the tensile strength, hardness, modulus,
etc. of Al alloy can be improved by the reinforcement of hard
particles [39]. The wear behavior of hard particle-reinforced
composite depends primarily on the type of interfacial bonding between the Al matrix and the reinforcement. When the
silicon carbide particles are strongly bonded with Al matrix,

560

S. Sawla, S. Das / Wear 257 (2004) 555561

they protect the surface against severe destructive action of


the abrasives. This is because of the strong interfacial bond,
which plays a critical role in transferring loads from the matrix to the hard particles, results in less wear of material. In
the case of poor interfacial bonding, the interface offers site
for crack nucleation and tends to pull out the particle from
the wear surface [40]. The subsurface of composite shows
less depth of deformation and less cracking tendency. But
in the case of alloy, the depth of deformation is more than
composite and cracking tendency at the interface is more. At
low load condition, the SiC particle protruded and protects
the wear surface from loss of material. This clearly indicated
by lower value of K of the composite than the alloy.
It has been observed that the value of wear constant K of
the heat-treated alloy and cast composite are more or less
same irrespective of load. This is explained based on the
fact that the propensity for crack nucleation in the case of
heat-treated alloy is less due to spherical nature of silicon
particle. On the other hand, in the case of cast composite,
the SiC particle protects the surface and stop severe cracking
of the metallic matrix. This explains that probability of formation of wear particle (K) is same in both the cases. It was
observed that the probability of formation of wear particles
is more at low load regime because of less depth of penetration of abrasives. In case of low load regime, the formation
of wear particle is dictated mainly by nucleation and propagation of microcracks. In the case of cast alloy possibility
for cracking is more because of sharp edges silicon particles in aluminum matrix. However, at high load regime, the
depth of penetration of the abrasives is more and the metallic matrix is deformed plastically. Further, nucleation and
joining of cracks leads to formation of few coarse debris.
Because of this reason, the probability of formation of wear
particle is less at high load regime than that of at low load.
In the later stages of abrasive wear, sharp edges eutectic
silicon gets fragmented and the cracks are joined together
with the SiC particles to form wear debris. The propensity
for crack nucleation is more in cast composite than that of
the heat-treated one. During severe wear process (usually at
high load and coarse abrasive size), the SiC reinforcements
are getting fractured and fractured particles tend to come out
from the wear surface. In the case of heat-treated composites, cracking tendency of the matrix is less as compared to
the cast composites due to matrix strengthening and higher
matrix hardness. Moreover, the cracking tendency would be
less in heat-treated composites due to near-spherical nature
of silicon phase. In the case of heat-treated alloy, the effective stress applied on the composite surface during wear
process is less due to higher strength and ductility of the Al
matrix. This resulted in less cracking tendency of the composite surface as compared to the cast alloy.
It is also to be noted that the fragmentation of eutectic silicon and cracking tendency of the composite
surface are dominating in cast composites, however, in
heat-treated composites only fragmentation of SiC particles
are dominated. The subsurface of composite shows less

depth of deformation as compared to cast composite. The


fragmented silicon and SiC particles are entrapped in the
mechanically mixed layer of cast composite.
The present result shows the superior abrasive wear resistance in case of heat-treated composite over the cast composite. It is also noted that the wear surface and subsurface
deformation of heat-treated composite depicted less damaged, crack propagation and less depth of deformation as
compared to cast composite. This is achieved due to the
combined effects of the reinforcement of SiC and heat treatment of alloy, which causes improvement in hardness and
better wear resistance of composite.

5. Conclusions
Following conclusions are made from the present study:
1. The probability of formation of wear particle (K) decreases with increase in applied load.
2. In the case of cast alloy, at low load regime (1 N), the
wear constant (K) is found higher value and decreased
drastically to lower value with increase in applied load
(3 N).
3. In the case of heat-treated alloy and composites, the wear
constant (K) decreases monotonically with load.
4. SEM observation of wear surface and subsurface suggested that at low load regime, the wear of material is
controlled by nucleation and propagation of crack, however, at high load regime, the material removal is dominated by plastic deformation.At high load condition, the
probability of formation of wear particles is less, which is
mainly because of higher depth of penetration and plastic
deformation of the wear surface and led to the formation
in coarse wear debris.

Acknowledgements
The authors thank Dr. N. Ramakrishanan, Director, RRL,
Bhopal, for his encouragement and for permission to publish
this article.

References
[1] A.I. Nussbaum, Light Met. Age 55 (2) (1997) 5458.
[2] I.M. Hutching, S. Wilson, A.T. Alpas, Comprehensive composites
materials, in: A. Kelly (Ed.), Metal Matrix Composites, vol. 3, 2000,
pp. 501519.
[3] D.H. Aylor, P.J. Maron, J. Electrochem. Soc. 132 (1985) 12711281.
[4] P.K. Rohatgi, R. Asthana, S. Das, Int. Met. Rev. 31 (1986) 115139.
[5] T. Zeuner, P. Stojanov, P.R. Sahn, H. Ruppert, A. Enels, Mater. Sci.
Technol. 14 (910) (1998) 857863.
[6] J.A. Hooker, J.K. Doorbar, Mater. Sci. Technol. 16 (78) (2000)
725731.
[7] A.A. Das, M.M. Yacaub, Z. Zantout, A.I. Cleg, Cast Met. 1 (1988)
6080.

S. Sawla, S. Das / Wear 257 (2004) 555561


[8] R.L. Deuis, C. Subramanium, J.M. Yellup, Wear 201 (1996) 132
144.
[9] S. Wilson, A. Ball, in: P.K. Rohatgi, P.J. Blaue, C.S. Yust (Eds.),
Tribology of Composites Materials, ASM International, Novelty, OH,
1990, pp. 103112.
[10] A.G. Wang, I.M. Hutching, Mater. Sci. Technol. 5 (10) (1989) 7176.
[11] K.J. Bansali, R. Mehrabian, J. Met. 34 (9) (1982) 3034.
[12] A.T. Alpas, J.D. Embury, in: K.C. Ludema, R.G. Bayer (Eds.),
Proceedings of the Conference on Wear of Materials, ASME, New
York, 1991, pp. 159166.
[13] S.-J. Lin, S.K. Liu, Wear 121 (1988) 114.
[14] A. Wang, H.J. Rack, Wear 146 (1991) 337348.
[15] B.K. Prasad, A.J. Jha, O.P. Modi, S. Das, A.H. Yegneswaran, Mater.
Trans. JIM 36 (1995) 10481057.
[16] B.K. Prasad, O.P. Modi, A.K. Jha, J. Tribol. Int. 27 (1994) 153
158.
[17] B.K. Prasad, S. Das, R. Dasgupta, O.P. Modi, A.K. Jha, A.H.
Yegneswaran, J. Mater. Sci. Lett. 17 (1998) 901903.
[18] A. Sato, R. Mehrabian, Metall. Trans. 7B (1976) 443451.
[19] P.K. Rohotgi, Adv. Mater. Technol. Monitor 17 (1990) 147.
[20] A.T. Alpas, J.D. Embury, Scripta Metall. 24 (1990) 931935.
[21] B.C. Pai, P.K. Rohatgi, S. Venkatesh, Wear 30 (1974) 117125.
[22] T. Kulik, T.H. Kosel, V. Xu, in: K.C. Ludema (Ed.), Proceedings of
the International Conference on Wear of Material, Elsevier, 1989,
pp. 2330.

561

[23] T. Jain-Main, S. Ye-Yug, Z. Hug-Ji, Z. Chingan, K. Xianwa, Tribol.


Int. 18 (1985) 101109.
[24] T. Hisakado, H. Suda, T. Trukui, Wear 155 (1992) 297307.
[25] B.R. Lawn, T. Jansan, A. Arora, J. Mater. Sci. 11 (1978) 573580.
[26] S.W. Date, S. Malkin, Wear 40 (1976) 223235.
[27] R.T. Spurr, Wear 65 (1981) 315328.
[28] K.H. Zum Gahr, Met. Prog. 116 (1979) 4652.
[29] A. Banerjee, S.V. Prasad, M.K. Surappa, P.K. Rohatgi, Wear 82
(1982) 141151.
[30] J. Fohl, T. Weissenberg, J. Wiedemeyer, Wear 130 (1989) 275288.
[31] S. Turenne, S. Caron, O. Weiss, J. Masounave, Fabrication
of Particulate Reinforced Metal Composites, ASM International,
Materials Park, OH, 1990, pp. 271276.
[32] S. Das, J. Mater. Sci. Lett. 16 (1997) 17571760.
[33] N. Axen, K.H. Zum Gahr, Wear 157 (89) (1992) 189201.
[34] I.M. Hutching, Chem. Eng. Sci. 42 (40) (1987) 869.
[35] B.K. Prasad, S.V. Prasad, A.A. Das, J. Mater. Sci. 27 (1997) 4489
4494.
[36] W.O. Song, P. Kraualis, A.P. Mourtiz, S. Bandopadhyay, Wear 185
(1995) 125130.
[37] S. Das, D.P. Mondal, S. Sawla, S. Dixit, Metall. Mater. Trans. 33A
(2002) 30313044.
[38] J.F. Archard, J. Appl. Phys. 24 (1953) 981988.
[39] D.J. Lloyd, Int. Met. Rev. 39 (1984) 123.
[40] S.V. Prasad, P.K. Rohatgi, J. Met. 39 (1987) 2226.

Вам также может понравиться