You are on page 1of 12

Wear 271 (2011) 1928–1939

Contents lists available at ScienceDirect

Wear
journal homepage: www.elsevier.com/locate/wear

Friction and wear behaviour of cast Al 6063 based in situ metal matrix
composites
C.S. Ramesh ∗ , Abrar Ahamed
Department of Mechanical Engineering, PES Institute of Technology, Bangalore, India

a r t i c l e i n f o a b s t r a c t

Article history: Al 6063 based in situ composites were manufactured from Al–10%Ti and Al–3%B master alloys by liquid
Received 1 September 2010 metallurgy route. The in situ TiB2 reinforced Al 6063 composites were synthesized through the exother-
Received in revised form mic reaction between Al–10%Ti and Al–3%B master alloys, which were used in the ratio of 1:2 respectively
23 December 2010
in Al 6063 matrix alloy. Tribological properties of both Al 6063 matrix alloy and the developed in situ
Accepted 23 December 2010
composites have been evaluated. Dry sliding friction and wear tests were carried out using a pin on
disc type machine with steel counter disc hardened to HRC60. A load range of 10–50 N with the sliding
velocity varying from 0.209 m/s to 1.256 m/s were adopted. Results have revealed that the developed
Keywords:
In situ composite
in situ composites have lowered coefficient of friction and wear rates when compared with Al 6063
Al 6063 matrix alloy under all the test conditions studied. The excellent wear resistance of the in situ composites
TiB2 results from the formation of fine TiB2 particles uniformly dispersed within the Al 6063 matrix alloy.
The coefficient of friction of both matrix alloy and in situ composites decreased with increase in load,
whereas it increased with increase in sliding velocity. However, wear rates of both matrix alloy and in situ
composites increased with increase in both load and sliding velocity.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction react with aluminium thereby eliminating any reaction products


at the interface between the reinforcement and the matrix [5].
Aluminium alloys are the most widely used non ferrous materi- Conventional practice of developing aluminium based com-
als in engineering applications owing to their attractive properties posites reinforced with hard ceramic particles involves ex situ
such as high strength to weight ratio, good ductility, excel- technique. This ex situ technique involves addition of rein-
lent corrosion resistance, availability and low cost [1]. However, forcement particles into the molten matrix alloy using liquid
their applications have often been restricted because conventional metallurgical route, which could lead to segregation of reinforce-
aluminium alloys are soft and notorious for their poor wear resis- ment particles and poor adhesion at the interface, unless there
tance. This problem is over by reinforcing hard ceramic particles exists good wettability between the matrix and particles or suitably
in aluminium and its alloys to produce a discontinuous rein- modified [6,7].
forced metal matrix composite which possesses nearly isotropic Recently, in situ technique has been developed to fabri-
properties. cate metal matrix composites (MMCs), which exhibits a clean
In addition, these composites exhibit high stiffness coupled matrix/reinforcement interface, which leads to better improve-
with superior strength. This suite of properties makes the par- ment in mechanical properties of the composites [8]. In in situ
ticle reinforced metal matrix composites (MMCs) attractive to a process, ultra fine ceramic particles are formed by the exother-
wide range of applications in automotive, aerospace and transport mic reaction between the elements or their compounds with
industries. Commonly used reinforcement particles in developing molten matrix alloy. These in situ routes provide advantages
MMCs include carbides, nitrides, borides and oxides [2]. Among such as uniform distribution of reinforcement, finer particle size,
these ceramics, TiB2 has emerged as a promising reinforcement clean interface, thermodynamically stable reinforcement phase
as it possesses high melting point, elastic modulus, hardness and and process economy in comparison with the conventional ex
strength at elevated temperatures coupled with good thermal con- situ processes. These advantages will lead to better and improved
ductivity and excellent wear resistance [3,4]. Further, it does not mechanical and tribological properties of the in situ composites
when compared with the matrix alloy [9]. By adopting in situ
techniques, MMCs with a wide range of matrix materials and sec-
∗ Corresponding author. Tel.: +91 80 26721983; fax: +91 80 26720886. ond phase particles can be produced [10–13]. However, meagre
E-mail address: csr gce@yahoo.co.in (C.S. Ramesh). information is available as regards the development of alu-

0043-1648/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.wear.2010.12.048
C.S. Ramesh, A. Ahamed / Wear 271 (2011) 1928–1939 1929

Fig. 1. SEM micrograph of Al 6063 alloy and the developed in situ composites. (a) Al 6063 alloy, (b) Al 6063–2.8 wt%TiB2 and (c) Al 6063–10 wt%TiB2 .

minium based in situ composites by use of master alloys Table 1


Percentage weight of TiB2 in the in situ formed composites.
[14].
In the light of the above, this paper reports on the friction and Al 6063–10 Al 6063–15 Al 6063–20
wear behaviour of Al 6063–TiB2 in situ composites produced by use wt%Al–10%Ti + 5 wt%Al–10%Ti + 7.5 wt%Al–10%Ti + 10
wt%Al–3%B wt%Al–3%B wt%Al–3%B
of master alloys.
2.8 10 6.7
2. Experimental details

Commercially available Al 6063 matrix alloy and the master in our earlier work [14]. A maximum amount of 10 wt%TiB2 was
alloys of Al–10%Ti and Al–3%B are melted in an electric resistance produced in situ within the matrix alloy on use of 15 wt%Al–10%Ti
furnace with a stoichiometric Ti:B ratio of 2:1. The proportion of and 7.5 wt%Al–3%B master alloys as reported in Table 1.
Al–10%Ti was varied from 10 wt% to 20 wt% in steps of 5 with SEM studies were conducted using JEOL scanning electron
the stoichiometric weight ratio being intact. Detailed procedure microscope. Microhardness tests were conducted on the polished
of preparation of the Al 6063 based in situ composites along with samples by applying a load of 100 g for a period of 10 s using Vick-
quantitative analysis of in situ phases of TiB2 and TiAl3 are reported ers microhardness tester. The test was carried out at five different

Fig. 2. Variation of coefficient of friction with sliding distance. Fig. 3. Variation of coefficient of friction with load.
1930 C.S. Ramesh, A. Ahamed / Wear 271 (2011) 1928–1939

all the tests. The loads and sliding velocities were varied from 10 N
to 50 N and 0.209 m/s to 1.256 m/s respectively. Frictional force was
measured using load cell of accuracy 0.1 N, while the wear loss in
terms of height loss was measured using linear variable differential
transducer (LVDT) of accuracy 1 ␮m. The coefficient of friction was
calculated using frictional and normal load data. The wear rates
were calculated from the height loss data in terms of volumetric
wear loss per unit sliding distance.

3. Results and discussion

3.1. SEM studies

SEM micrograph of both the base alloy and developed in situ Al


Fig. 4. Variation of coefficient of friction with sliding velocity.
6063–TiB2 composites are shown in Fig. 1. It is observed that the
in situ synthesized TiB2 particles in the matrix are cubic and spher-
locations in order to contradict the possible effect of indentor rest- ical in form. The size of TiB2 particles ranged from 0.2 ␮m to 2 ␮m,
ing on the harder reinforcement particles. The average of all the which is much lower than that of ex situ particles in discontinu-
five readings was taken as hardness of the samples. ously reinforced composites. Further, there is homogeneity in the
Dry sliding friction and wear tests were performed using pin distribution of in situ TiB2 particles within the matrix alloy with a
on disc apparatus, (Make: Magnum Engineers, Bangalore) as per clear matrix-reinforcement interface.
ASTM G99-95. Cylindrical specimens (both matrix alloy and in situ
composites) of 8 mm diameter and 20 mm height were used as test 3.2. Microhardness
samples. The specimen end surfaces were flat and polished met-
allurgically. Counter disc used is a hardened HRC60 steel disc. The Microhardness of the matrix alloy and the developed in situ
initial surface finish of the steel disc is 1 ␮m. Track radius of 40 mm composites are reported in Table 2. It is observed that there is
has been used for all the experiments. All the tests were conducted significant improvement in the microhardness of the developed
in air at room temperature. Test duration of 30 min was adopted for in situ composites with increased content of TiB2 . An improve-

Fig. 5. SEM of wear tracks of Al 6063 alloy and the developed in situ composites at a load of 10 N and sliding velocity of 0.209 m/s.
C.S. Ramesh, A. Ahamed / Wear 271 (2011) 1928–1939 1931

Fig. 6. EDAX on wear tracks of Fig. 5.

Table 2
Microhardness and tribological properties.

Al 6063 alloy Al 6063–2.8 wt%TiB2 Al 6063–6.7 wt%TiB2 Al 6063–10 wt%TiB2

Microhardness (VHN) 46.56 51.79 56.39 59.25


Coefficient of friction (␮) 0.97 0.8 0.75 0.5
Wear rate (mm3 /m) 0.02204 0.0178 0.01376 0.01175
1932 C.S. Ramesh, A. Ahamed / Wear 271 (2011) 1928–1939

Fig. 7. SEM of wear debris of Al 6063 alloy and the developed in situ composites at load of 10 N and sliding velocity of 0.209 m/s.

ment in microhardness of about 11.23%, 21.11% and 27.25% for Al composites. This increase in coefficient of friction with increase in
6063–2.8 wt%TiB2 , Al 6063–6.7 wt%TiB2 and Al 6063–10 wt%TiB2 sliding distance can be attributed to rise in temperature of rubbing
composites respectively when compared with Al 6063 alloy have surfaces which leads to higher frictional force. Similar trends have
been observed. Further, there is no much variation in microhard- been observed by other researchers also [20]. However, for all slid-
ness in between Al 6063–6.7 wt%TiB2 and Al 6063–10 wt%TiB2 . ing distances studied the coefficient of friction of in situ composites
decreases with increased content of TiB2 .
3.3. Coefficient of friction Fig. 3 shows the effect of normal load on the coefficient of fric-
tion of Al 6063 matrix alloy and the developed in situ composites.
Coefficient of friction values of Al 6063 alloy and the developed It is observed that with an increase in load, there is a decrease
in situ composites are given in Table 2. It is observed that there is in coefficient of friction of both the matrix alloy and the devel-
a significant decrease in the coefficient of friction of in situ com- oped in situ composites. However, the developed in situ composites
posites with increase content of TiB2 . A reduction of 17.52%, 22%, shows lower coefficient of friction values when compared with
and 48.45% in the coefficient of friction is observed for compos- matrix alloy under all the loads studied. It is also observed that there
ites containing 2.8 wt%, 6.7 wt% and 10 wt% TiB2 respectively when is a gradual decrease in coefficient of friction for both the matrix
compared with the matrix alloy. Similar trend has been reported alloy and in situ composites up to 30 N of load and no significant
by Kumar et al. [9]. This can be mainly attributed to the excellent variation of coefficient of friction is observed beyond 30 N, which is
lubricating properties of TiB2 as it possesses a layered structure. due to the presence of secondary phases such as Si in matrix alloy,
Further, it can be also due to smaller particle size [15,16], improved Si and TiB2 in the composites which restricts the flow of metal dur-
dispersion as evidenced in Fig. 1. It is reported that good dispersion ing sliding [9]. Further, fragments of softened debris generated at
and clean interface of the particle in matrix leads to a lower value of higher loads may provide some degree of solid lubrication leading
coefficient of friction of composites [17]. The superior antifrictional to reduction in coefficient of friction at higher loads [21]. In case
behaviour of the in situ composites can also be due to the fact that of composites, increased load leads to higher extent of transfer of
in situ TiB2 particles which are homogeneously distributed do play TiB2 onto both the mating and sliding surfaces. Larger the trans-
the role of load bearer during the sliding process. It is reported that fer, better will be the lubrication at the sliding interface. Further,
the reinforced particles do act as load bearing members [18,19], since TiB2 is produced by in situ reaction, it can accommodate larger
and hence friction should have mainly occurred between TiB2 and loads with less plastic elongation owing to good bond between the
steel disc surface. matrix and the reinforcement. This improved load bearing capacity
Fig. 2 shows the effect of sliding distance on coefficient of friction of the in situ composites is primarily responsible for its excellent
of Al 6063 alloy and its in situ composites. It is observed that with an antifrictional behaviour especially at higher loads.
increase in sliding distance, there is uniform and marginal increase Fig. 4 shows the effect of sliding velocity on coefficient of friction
in coefficient of friction for both the matrix alloy and the in situ of Al 6063 alloy and its in situ composites. It is observed that, with
C.S. Ramesh, A. Ahamed / Wear 271 (2011) 1928–1939 1933

Fig. 8. EDAX on wear debris of Al 6063 alloy and the developed in situ composites at a load of 10 N and sliding velocity of 0.209 m/s.

increase in sliding velocity, there is increase in coefficient of fric- forming a thin adherent solid lubricating film resulting in lower-
tion for both the matrix alloy and the in situ composites. This can ing of coefficient of friction. This phenomenon proceeds at a faster
be attributed to the fact that with an increase in sliding velocity rate with increased sliding velocities. However, at higher sliding
there is an increase in frictional force. The increase in coefficient velocity the existing solid lubricating film gets thickened and may
of friction with increased sliding velocity can also be attributed fragment. Thicker the solid film, the fragmented solid film gets
to the fact that the brittle particulate reinforcements in the com- clogged between the rubbing surfaces resulting in higher friction.
posites get cracked and are squeezed out onto the matrix surfaces Also at higher velocities temperature of mating surfaces [22] leads
1934 C.S. Ramesh, A. Ahamed / Wear 271 (2011) 1928–1939

mation of transfer layers have been observed by several other


researchers especially in case of aluminium based composites
sliding on steel [18,27,28].

Fig. 5 shows the SEM photographs of worn surfaces of Al 6063


matrix alloy and its in situ composites. It is observed that the mor-
phology of worn surfaces of composites is different from that of the
matrix alloy. The matrix alloy shows wider and shallow grooves
on the surface, whereas fine grooves are noticed in case of the
developed in situ composites. This observation supports the lower
wear rates of the composites when compared with Al 6063 matrix
alloy. Worn surfaces also show the presence of some agglomerated
wear debris which is oriented in the direction of sliding. Further,
Fig. 9. Variation of wear rate with sliding distance for both the matrix alloy and the these debris are mixture of matrix material, steel counterface and it
developed in situ composites.
includes fractured TiB2 particles in case of the in situ composites as
evidenced in Fig. 6. Apparently, the worn surfaces are characterized
to more asperity junctions as a consequence of which the coefficient
by long continuous grooves, which form as the counter body plough
of friction increases.
across the surface and eventually removing or pushing material
into ridges along sides of the grooves. The surface morphology of
3.4. Wear studies
the worn surface is similar to that observed in the early stage of the
preparation of metallographic specimens in case of the in situ com-
Effect of reinforcement on wear rates of Al 6063 matrix alloy
posites. And, the severity of microploughing is substantially less
and the developed in situ composites is as reported in Table 2. It is
pronounced on the worn surface of composites. It is considered that
observed that the in situ composites possess lower wear rates than
the counter body (HRC60 steel disc) deforms the alloy matrix of the
that of the base alloy. A reduction of 19.23%, 37.56% and 46.68%
specimens during the early stages of wear and causes microplough-
in wear rates of 2.8 wt%, 6.7 wt% and 10 wt% reinforced TiB2 com-
ing and grooving in the surface of aluminium matrix. Subsequently,
posites respectively when compared with Al 6063 matrix alloy are
material chips are removed from the grooves formed in the surface,
noticed. This considerable improvement in wear resistance of the
thereby exposing the hard reinforcing particles. Since, the hardness
in situ composites can be attributed to the following factors:
of counter body is lower than TiB2 particles so, microploughing
is dominant wear mechanism in case of the composites. It is evi-
• The microhardness of in situ composite increases with increase in
dent from the worn surfaces that the composites with higher TiB2
the amount of TiB2 reinforcement. Increasing hardness reduces content provides better wear protection or resistance.
the wear intensity. It is reported that severe adhesive wear Energy dispersion X-ray analysis pattern of worn surfaces of
depends on material hardness [23]. Further, there is experimen- Al 6063 and the developed in situ composites are shown in
tal support and practical evidence [24] to suggest that the onset Fig. 6. It is observed that the extent of iron on the worn sur-
of adhesive process, such as scuffing and seizure are retarded by faces is more on the in situ composite samples which indicates
increasing hardness of the mating metals. a lower wear rate of the composites. It is reported [19] that
• Existence of excellent bond between the matrix and the particles
the transferred iron layer may get oxidized due to increased
as evidenced by SEM of the in situ composites shown in Fig. 1. The localized heating under the oxidizing atmosphere leading to low-
interfacial bond between the matrix and the particle reinforce- ering of material transfer. Further, once the transferred layer is
ment play a significant role in the wear process [25]. Presence established on the composites the steel counter face can be con-
of good bond minimizes the possibility of decohesion or pull out sidered to be mainly in contact with mixture of iron and its oxide.
of reinforced phase from the matrix alloy there by limiting the These mechanically mixed layers are expected to provide in situ
situation of three body abrasive wear, which otherwise would lubricating effect leading to higher wear resistance of the compos-
have resulted in higher material removal as reported by several ites.
researchers [26]. Fig. 7 shows the SEM micrographs of wear debris of Al 6063
• There exists a formation of inhomogeneous transfer layer, which
alloy and the developed in situ composites. The investigation of
may consist of a mixture of iron debris, aluminium matrix and collected debris from base alloy samples generated under dry con-
fragments of second phases on the sliding surface under steady tacts revealed large, irregular profiles and unequal dimensions. The
state wear as evidenced from EDAX studies. This kind of for- generation of this kind of debris can be attributed to an abrasive
micro-cutting effect. However, the wear debris of Al 6063–TiB2
composites are smaller which can be attributed to decrease in the
severity of micro-cutting process during sliding contacts due to the
enhanced hardness of the composites with increased content of
TiB2 .
Fig. 8 shows the EDAX analysis performed on the wear debris
of both unreinforced alloy and composites. It is observed that,
extent of iron transfer is relatively high in composite than the alloy
under identical test conditions. This might be due to the fact that,
presence of TiB2 particles in composite materials increases the pos-
sibility of abrasive action by ploughing into the steel counterface
and increases the material transfer.
Fig. 9 shows the effect of sliding distance on wear rate of both
Al 6063 alloy and the in situ composites. It is observed that with
Fig. 10. Variation of wear rate with load for both the matrix alloy and the developed increase in sliding distance there is gradual increase in wear rate of
in situ composites. both matrix alloy and the developed in situ composites. A steady
C.S. Ramesh, A. Ahamed / Wear 271 (2011) 1928–1939 1935

Fig. 11. SEM of wear tracks of Al 6063–10 wt%TiB2 at different loads and sliding velocity of 0.209 m/s.

state is reached beyond 628.31 m. This increase in wear rate with that with increase in applied normal load, there is increase in wear
increase in sliding distance is due to the more intimate contact rate of both the matrix alloy and the in situ composites. However,
between the sliding surfaces of the specimen and the rotating disc at all the loads studied in situ composites possessed lower wear
with elapse of time. The increase in wear rate with sliding dis- rates when compared with the matrix alloy.
tance can also be attributed to the fact that with increase in sliding Increase in wear rate with increased load in both matrix alloy
distance the temperature at the specimen disc interface increases and the in situ composites can be attributed to higher extent of plas-
and hence the material gets softened and tends to get into plastic tic deformation. Greater the extent of plastic deformation higher
state. Similar trends have been observed by many other researchers will be the possibility of sub surface cracking which leads to larger
[29,30]. material removal [31]. Further, increased load will result in onset
However, at all slid distances studied, the developed in situ com- delamination leading to higher wear rates for both the matrix alloy
posites do possess lower wear rates when compared with matrix and the composites.
alloy. Fig. 11 shows the SEM micrographs of worn surfaces of Al
The influence of applied normal load on the wear rates of Al 6063 6063–10 wt%TiB2 in situ composite at different loads. The micro-
alloy and the in situ composites are shown in Fig. 10. It is observed graphs clearly indicate that the material have been removed by
1936 C.S. Ramesh, A. Ahamed / Wear 271 (2011) 1928–1939

Fig. 12. SEM of wear debris of Al 6063–10 wt%TiB2 for different loading conditions at sliding velocity of 0.209 m/s.

plastic flow in all the cases. It is observed that the severity of plas-
tic deformation is increasing with increase in load. As the load
increases, the morphology of the worn surface gradually changes
from fine scratches to distinct grooves. Further, damaged spots
in the form of craters can also be observed. At lower loads, the
composite specimens show a mixed abrasion-plastic deforma-
tion mechanism. Abrasion is evident from the scoring grooves
visible on the worn surfaces in the direction of sliding. It is pos-
sible that the scored grooves might have been formed due to the
action of the wear-hardened deposits on the disk track. But at
higher loads the locally damaged and even fractured spots are
observed. These are the indications of severe deformation and
fracture resulting in a high wear rate. Under high loads, the protec-
Fig. 13. Variation of wear rate with sliding velocity for both the matrix alloy and
the developed in situ composites. tive layer of the reinforcing particles can no longer remain stable
C.S. Ramesh, A. Ahamed / Wear 271 (2011) 1928–1939 1937

Fig. 14. SEM of wear tracks of Al 6063–10 wt%TiB2 for different sliding velocities at a load of 10 N.

under the ploughing action and the wear strips formed are more The effect of sliding velocity on wear rate of Al 6063 alloy and
distinct. the in situ composites is shown in Fig. 13. It is clearly observed that
Fig. 12 shows the SEM micrographs of wear debris of Al with increase in sliding velocity, there is increase in wear rate of
6063–10 wt%TiB2 in situ composite at loads varying from 10 N to both matrix alloy and the developed in situ composites. This can be
50 N. It is observed that with increase in load, there is a gradual attributed to high strain rate subsurface deformation as reported
increase in the size of the wear debris collected, which explains the by [32]. The increased rate of subsurface deformation increases the
higher rate of wear with increase in load. This is well supported by contact areas by fracture and fragmentation of asperities leading to
the wear rate results as discussed earlier. It is also observed that enhanced delamination contributing to higher wear loss. However,
at higher loads the debris is mixture of fine and coarse sized par- the in situ composites possesses lower wear rates when compared
ticles. This is because of break down of large sized particles due to with the matrix alloy for all the sliding velocities studied.
increased interaction during sliding. Further, it is reported that [29], Fig. 14 shows the SEM photographs of worn surfaces of Al
at higher loads, since the severity of surface contact is very high, 6063–10 wt%TiB2 composite with varying sliding velocities. It is
the stresses exceeds the fracture stress of the reinforced particles, observed that with increase in sliding velocity there is an increase in
leading to fracture of the reinforced phase. severity of the damage to the worn surfaces. The worn surface with
1938 C.S. Ramesh, A. Ahamed / Wear 271 (2011) 1928–1939

Fig. 15. SEM of wear debris of Al 6063–10 wt%TiB2 in situ composite at different sliding velocities at a load of 10 N.

highest sliding velocity of 1.256 m/s has experienced severe groov- and flaky in shape. This supports the kind of wear mechanism that
ing and plastic flow. It is reported that while sliding a protective occurs with increase in sliding velocity and as discussed earlier and
layer is formed on the surface of the worn surfaces. At lower slid- well supported by the wear tracks of the same composite shown in
ing velocities, this protective layer is found to be stable and protects Fig. 14.
the material by acting as solid lubricant, whereas at higher sliding
velocities it becomes unstable and is easily removed by delam- 4. Conclusions
ination. Such delaminated layers get entrapped between sliding
surfaces resulting in three body abrasive wear. Thus, delaminat- In situ 6063/TiB2 composites were successfully prepared via
ing and abrasive wear together causes excessive wear at higher chemical reaction between Al–10%Ti and Al–3%B master alloys in
velocities. Al 6063 matrix alloy using liquid metallurgy route. A significant
Fig. 15 shows the SEM micrographs of wear debris of Al decrease in coefficient of friction and wear rates with increase
6063–10 wt%TiB2 in situ composite collected at different sliding in TiB2 content in the in situ composites has been observed. A
velocities. It is observed that with increase in sliding velocity there maximum reduction of 48.45% and 46.68% in coefficient of fric-
is an increase in size of the debris particle collected, which explains tion and wear rates respectively for composites containing 10 wt%
the severity of wear rate. It is also observed that at lower slid- TiB2 when compared to the matrix alloy have been observed. For
ing velocities, the wear debris are finer and granular in structure, a maximum load of 50 N 31% and 32.1% reduction in coefficient of
while at higher sliding velocities, the wear debris are larger in size friction and wear rates respectively were observed for 10 wt% TiB2
C.S. Ramesh, A. Ahamed / Wear 271 (2011) 1928–1939 1939

reinforced composites when compared with the matrix alloy. For [13] H.J. Brinkman, J. Duszczyk, L. Katgerman, In situ formation of TiB2 in a P/M
a maximum sliding velocity of 1.256 m/m 36.9% and 50% reduction aluminium matrix, Scripta Metallurgica et Materialia 37 (1997) 293–297.
[14] C.S. Ramesh, B.H. Abrar Ahamed, R. Channabasappa, Keshavamurthy, Devel-
in coefficient of friction and wear rates respectively were observed opment of Al 6063–TiB2 in situ composites, Materials & Design 31 (2010)
for 10 wt% TiB2 reinforced composites when compared with the 2230–2236.
matrix alloy. [15] M. Godet, The third body approach: a mechanical view of wear, Wear 100
(1984) 437–452.
[16] S. Basavarajappa, G. Chandramohan, Arjun Mahadevan, Mukundan Thangavelu,
Acknowledgments R. Subramanian, P. Gopalakrishnan, Influence of sliding speed on the dry slid-
ing wear behavior and the subsurface deformation on hybrid metal matrix
composites, Wear 262 (2007) 1007–1012.
The authors would like to express their deep sense of gratitude [17] Hongzhan Yi, Naihang Ma, Yijie Zhang Ma, Yijie Zhang, Xiangfeng Li, Haowei
to Dr. K.N.B. Murthy Principal and Director, PESIT, Bangalore and Wang, Effective elastic moduli of Al-Si composites reinforced with in situ par-
Prof. D. Jawahar, CEO, PES Group of Institutions. ticles, Scripta Materialia 54 (2006) 1093–1097.
[18] L.F. Zhang, L.C. Zhang, Mai, Y. Woo, Particle effects on friction and wear of alu-
minium matrix composites, Journal of Material Science 30 (1995) 5999–6004.
References [19] S. Wilson, A.T. Alphas, Wear mechanism maps for metal matrix composites,
Wear 212 (1997) 41–49.
[1] Jiang Xu, Wenjin, Liu, Wear characteristic of in situ synthetic TiB2 particulate- [20] Hozumi Goto, Shunji Omori, Friction and characteristics of aluminium alloy
reinforced Al matrix composite formed by laser cladding, Wear 260 (2006) impregnated carbon composites, Transactions of ASME 121 (1999) 294–299.
486–492. [21] M. Kestursatya, J.K. Kim, Rohatgi, Friction and wear behaviour of a centrifugally
[2] S.C. Tjong, S.Q. Wu, H.G. Zhu, Wear behavior of in situ TiB2 .Al2 O3 /Al and cast lead free alloy containing graphite particles, Journal of Metallurgical and
TiB2 .Al2 O3 /Al-Cu composites, Composites Science and Technology 59 (1999) Materials Transactions A 32 (2001) 2115–2125.
1341–1347. [22] F. Rana, D.M. Stefanescu, Friction properties of Al-1.5%Mg/SiC particulate
[3] G.V. Samsonon, I.M. Vinitskii, Handbook of Refractory Compounds, Plenum, metal-matrix composites, Metallurgical and Materials Transactions A 20 (1989)
New York, 1980, p. 40. 1564–1566.
[4] S.C. Tjong, K.C. Lau, Properties and abrasive wear of TiB2 /Al-4%Cu composites [23] G. Levy, R.G. Lingford, L.A. Mitchell, Wear behaviour and mechanical properties:
produced by hot isostatic pressing, Composite Science & Technology 59 (1999) the similarity of seemingly unrelated approaches, Wear 21 (1972) 167–177.
2005–2013. [24] Q.J. Xue, K.C. Ludema, Plastic failure effects in scuffing of soft metals, in: K.C.
[5] K.L. Tee, L. Lu, M.O. Lai, Synthesis of in situ Al-TiB2 composites using stir cast Ludema (Ed.), Proceedings of the Conference on Wear of Materials, ASME, New
route, Composite Structures 47 (1999) 589–593. York, 1985, pp. 499–506.
[6] T.D.P. Rajan, R.M. Pillai, B.C. Pai, Reinforcement coatings and interfaces in [25] C.S. Ramesh, S.K. Seshadri, Tribological characteristics of nickel based compos-
aluminium metal matrix composites, Journal of Material Science 33 (1998) ite coatings, Wear 255 (2003) 893–902.
3491–3503. [26] H.E. Hintermann, Adhesion friction and wear of thin hard coatings, Wear 100
[7] B.S. Murthy, S.K. Thakur, B.K. Dhindaw, On the infiltration behavior of Al, Al-Li (1984) 381–397.
and Mg melts through SiCp bed, Metallurgical and Materials Transactions A 31 [27] S.K. Biswas, B.N. Pramila Bai, Dry wear of Al–graphite particle composites, Wear
(2000) 319–325. 68 (1981) 347–358.
[8] S.C. Tjong, Z.Y. Ma, Microstructural and mechanical characteristics of in situ [28] T.P. Murali, S.V. Prasad, M.K. Surappa, P.K. Rohatgi, K. Gopinath, Friction and
metal matrix composites, Material Science and Engineering R29 (2000) 49–113. wear behavior of aluminum alloy coconut shell char particulate composites,
[9] S. Kumar, M. Chakraborty, V. Subramanya Sarma, B.S. Murthy, Tensile and Wear 80 (1982) 149–158.
wear behaviour of in situ Al-7Si/TiB2 particulate composites, Wear 265 (2008) [29] I.M. Hutchings, Tribological properties of metal matrix composites, Materials
134–142. Science and Technology 10 (1994) 513–517.
[10] A.K. Kuruvilla, K.S. Prasad, V.V. Bhanuprasad, Y.R. Mahajan, Microstructure- [30] Tiejun Ma, Hideki Yamaura, Donald A. Koss, Robert C. Voigt, Dry sliding wear
property correlation in Al/TiB2 (XD) composites, Scripta Metallurgica et behavior of cast SiC-reinforced Al MMCs, Materials Science and Engineering A
Materialia 24 (1990) 873–878. 360 (1–2) (2003) 116–125.
[11] Z.Y. Ma, J.H. Li, M. Luo, X.G. Ning, Y.X. Lu, J. Bi, Y.Z. Zhang, In situ formed Al2 O3 [31] Farid Akhtar, Microstructure evolution and wear properties of in situ synthe-
and TiB2 particulates mixture-reinforced aluminium composite, Scripta Met- sized TiB2 and TiC reinforced steel matrix composites, Journal of Alloys and
allurgica et Materialia 31 (1994) 635–639. Compounds 459 (2008) 491–497.
[12] H. Nakata, T. Choh, N. Kanetake, Fabrication and mechanical properties of in situ [32] A.A. Hamid, P.K. Ghosh, S.C. Jain, S. Ray, The influence of porosity and particles
formed carbide particulate reinforced aluminium composite, Journal of Mate- content on dry sliding wear of cast in situ A(Ti)-Al2 O3 (TiO2 ) composites, Wear
rial Science 30 (1995) 1719–1727. 265 (2008) 14–26.

You might also like