You are on page 1of 12

Materials and Design 32 (2011) 37183729

Contents lists available at ScienceDirect

Materials and Design


journal homepage: www.elsevier.com/locate/matdes

Investigation of microstructure, hardness and wear properties of Al4.5 wt.% CuTiC


nanocomposites produced by mechanical milling
N. Nemati a, R. Khosroshahi a, M. Emamy b,, A. Zolriasatein c
a

Materials Engineering Department of Sahand University of Technology (SUT), Tabriz, Iran


Center of Excellence for High Performance Materials, School of Metallurgy and Materials, University of Tehran, Tehran, Iran
c
Faculty of Mechanical Engineering, K.N. Toosi University of Technology, Tehran, Iran
b

a r t i c l e

i n f o

Article history:
Received 5 February 2011
Accepted 24 March 2011
Available online 30 March 2011
Keywords:
A. Metal matrix composite
C. Mechanical alloying
E. Wear

a b s t r a c t
The present work deals with studies on the manufacturing and investigation of mechanical and wear
behavior of aluminum alloy matrix composites (AAMCs), produced using powder metallurgy technique
of ball milled mixing in a high energy attritor and using a blendpresssinter methodology. Matrix of
pre-mechanical alloyed Al4.5 wt.% Cu was used to which different fractions of nano and micron size
TiC reinforcing particles (ranging from 0 to 10 wt.%) were added. The powders were mixed using a planetary ball mill. Consolidation was conducted by uniaxial pressing at 650 MPa. Sintering procedure was
done at 400 C for 90 min. The results indicated that as TiC particle size is reduced to nanometre scale
and the TiC content is increased up to optimum levels, the hardness and wear resistance of the composite
increase signicantly, whereas relative density, grain size and distribution homogeneity decrease. Using
micron size reinforcing particulates from 5% to 10 wt.%, results in a signicant hardness reduction of the
composite from 174 to 98 HVN. Microstructural characterization of the as-pressed samples revealed reasonably uniform distribution of TiC reinforcing particulates and presence of minimal porosity. The wear
test disclosed that the wear resistance of all specimens increases with the addition of nano and micron
size TiC particles (up to 5 wt.%). Scanning electron microscopic observation of the worn surfaces was conducted and the dominant wear mechanism was recognized as abrasive wear accompanied by some
delamination wear mechanism.
2011 Elsevier Ltd. All rights reserved.

1. Introduction
Aluminum-based alloys are widely used as aerospace and
automotive components, because of their high specic strength,
stiffness and formability. However, both pure Al and Al alloys possess poor wear resistance. On the other hand, Al alloy matrix
composites are known to offer better wear resistance and bulk
mechanical properties. These composites are synthesized by liquid
or powder metallurgy routes. The wear performance of these composites is a subject of strong interest, especially for their potential
application in automobile components including cylinder block,
piston and brake disks [1]. The wear test is usually conducted
under sliding wear conditions using pin-on-disk or block-on-ring
tests. Though high volume fractions of hard reinforcements are
favored for wear resistance, the wear rate of the counter-body is
found to be greatly enhanced by the abrasive action of the reinforcements [2]. Carbides, oxides, nitrides and different intermetallics compounds have been used extensively as reinforcing
particulates for AMCs [3]. In specic, SiC and Al2O3 were the
Corresponding author. Tel.: +98 21 82084083; fax: +98 21 61114083.
E-mail address: emamy@ut.ac.ir (M. Emamy).
0261-3069/$ - see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.matdes.2011.03.056

commonly used ceramic reinforcements and recently NiAl, Ni3Al


and intermetallics composite have been shown to improve the
wear resistance of aluminum and magnesium alloys to a level similar to that of SiC reinforced composite, whilst reducing counterface wear rates [46]. TiC was not investigated enough as a
ceramic reinforcement to Al alloy matrix nanocomposites, however it has been receiving much attention lately for its high melting
temperature (3160 C), low thermal coefcient of expansion,
extraordinary hardness, excellent wear and abrasion resistance
[7,8]. A decrease of the reinforcement particle size from micrometric to nanometric scale, brings a superior increase in mechanical
strength of the composite, but the tendency of particle clustering
and agglomeration also increases [911]. It is important to note
that a homogeneous distribution of the reinforcing particles is
essential for achieving the improved properties [12]. Mechanical
alloying (M/A) via ball milling has been successfully employed to
improve particle distribution throughout the matrix [1315].
Mechanical alloying is solid-state synthesis process that
consists of repeated cold-welding, fracturing, dynamic recrystallization and mechanically activated inter diffusion among the powder particles. A high energy ball mill offers indeed supplementary
degrees of freedom in the choice of possible routes for synthesizing

3719

N. Nemati et al. / Materials and Design 32 (2011) 37183729

new materials and appears further as an attractive method of synthesis in view of its potential for large scale production [16]. To obtain an absolute nanocrystalline alloy powder, materials from
initial mixtures of elemental powders in a grain size range of
micrometers with appropriate weight proportions are charged into
a high energy ball mill container. By applying the proper milling
condition such as suitable process control agent (PCA), ball to powder ratio (BPR) and time of milling, the specic alloy will be
achieved in solid state. Recently, a number of Al65Cu20TM15 or
Al50TM40Si10 (TM = early transition metals like Ti, Nb and Zr) alloys
have been developed by mechanical alloying with amorphous and/
or nanocrystalline matrix and in situ nano-intermetallic phase/
compound dispersion [6,17,18]. These composites have exhibited
nearly 23 times greater wear resistance and hardness than the
conventional microcrystalline and age hardenable Al alloys [19
21]. The general characteristics and mechanisms of mechanical
milling and alloying have been studied, but the inuence of milling
parameters on mechanical properties of composites is not yet fully
known [22]. Wear behavior of these novel mechanically alloyed Al
alloys or composites has not been investigated fully [2].
The present study has focused on the manufacturing of MMCs
using a nanocrystalline alloying matrix for the desired composite
with the composition of Al4.5 wt.% Cu, which according to the
previous study was prepared through mechanical alloying using
a high energy attritor ball mill to produce the nanostructure matrix
powder of the composites [23]. Different concentrations of nano
and micron size TiC particles as reinforcement were added to the
matrix to improve the hardness and wear resistance of the MMCs,
proper distribution of the reinforcing particles in the matrix is very
important in blending process which is measured through precise
EDX analysis [4,6], then the effect of different weight fractions of
reinforcement on the wear resistance, hardness, relative density
and microstructure of a new generation nanocrystalline matrix
composites are investigated. An effort has also been made to
understand the operative wear mechanism.
2. Experimental procedure
Pre-alloyed powder of the composite matrix produced by
mechanical alloying technique with nominal stoichiometry of
Al4.5 wt.% Cu (M) and particle size in range of 1020 lm with a
ake like morphology, typical of mechanical alloyed particles were
used. The chemical composition of the matrix is shown in Table 1,
according to previous study to produce the alloying matrix, pure Al
and Cu elemental powders with average particle size of 150 and
40 lm respectively were subjected to mechanical alloying in an
attrition ball mill. Milling was operated for 70 h at the rotational
speed of 270 rpm with 12:1 ball to powder weight ratio (BPR)
and 1.5 wt.% stearic acid was used as the milling process control
agent (PCA), using stainless steel vial (AISI 316) and hardened
chromium steel balls (6 and 10 mm diameter with the ratio of
1:2). The powder in the vial was protected with high purity argon
gas and the milling tank was cooled with a x ow of cold water
through a rotating system. Samples for analysis were removed by
interrupting the ball mill at various intervals (each 10 h). Alloying
behavior, solution and crystallite size evolution during milling
were determined by X-ray diffraction (XRD) analysis using a Philips (30 kV and 25 mA) diffractometer with Cu Ka radiation (k =
1.5405 ). All XRD experiments were done with a step size of

0.02 and a time per step of 0.1 s. Crystallite size of powders was
evaluated using the WilliamsonHall and Sherrer methods
[23,24]. Since both Al and Cu are ductile materials, the dominant
mechanism that occurs through mechanical alloying of these elements is indeed ductileductile which results in formation of
lamellar structure of repeated attened layers of aluminum and
copper in each grain of the powder particles [24].
In order to study the effect of reinforcing particulate amount
and size on the hardness, sliding wear behavior and the microstructure of the composites, different weight percent of nano-size
(0%, 0.5%, 1%, 3%, 5% and 7 wt.%) and micron size (0%, 5%, 7% and
10 wt.%) TiC particulates were added to the prepared alloy matrix
powder and blending was performed for 2 h in a planetary ball mill
at nominal room temperature and at a rotation speed (cup speed)
of 400 rpm with 12:1 ball to powder weight ratio using tungsten
carbide (WC) coated vials and hardened chromium steel balls in
four different sizes (20, 15, 10, 6 mm diameter) under Ar atmosphere and applying 1.5 wt.% stearic acid as the milling PCA [24].
The microstructure and distribution of the milled TiC powders
and the compacts were studied by CamScan MV3200 scanning
electron microscope and supplemental EDX.
The same equipment was employed to characterize the milled
powder morphology and the topology of the surfaces of compact
samples. The as-received and milled powders were consolidated
into compacts using cold and hot unidirectional pressing. First,
4 g of powders were measured and poured into in a cylindrical tool
steel die and were cold pressed. Fig. 1a shows the scheme of the die
with an inner diameter of 10 mm. Powder mixture was sustained
for 10 min in the pressure of 650 MPa then according to Fig. 1b
by adjusting an auto-controlled power of the heating system, the
compacted samples were hot pressed at 400 C under pressure of
500 MPa for 80 min. Then they are allowed to cool to room temperature and stripped off the die for characterization and mechanical
tests. Density values of compact samples were measured applying
three different methods of geometrical, Archimedes and theoretical approaches. Vickers microhardness method was used to determine the hardness of consolidates using a Mitutoyo microhardness
tester. The VHN measurements were conducted at 50 gf for 10 s of
dwell time. A minimum of ve indentations were measured per
condition.
The dry wear tests were performed according to ASTM standard
G99 [25] on a unidirectional pin-on-disk aperture wear tester with
a counter face of AISI 52100 steel which was hardened to 62 HRC at
ambient temperature, the cylindrical shape pins with the dimensions of 10 mm in diameter and 15 mm in length were prepared
by machining the consolidated samples for the tests. Prior to wear
tests, the pins and disks were polished/grinded with 800 grit SiC
paper to a 1 lm nish. Samples were cleaned thoroughly using
an ultrasonic device with acetone and dried prior and after the
tests and then weighed to determine the wear loss using an electronic balance (GR200-AND) with an accuracy of 0.1 mg. All data
on wear were measured from the pins. The volume loss of samples,
according to Eq. (1), was calculated through the mass loss divided
by the Archimedes density of the materials and the friction coefcient (l) values were measured by a load cell equipped with the
pin-on-disk apparatus and the wear rate values of the samples
were calculated for all samples according to Eq. (2).
The morphology of the worn surfaces was analyzed using similar SEM, the same test parameters were used for all samples: 10

Table 1
Chemical composition of the alloying matrix.
Materials

Si

Fe

Cu

Mn

Mg

Cr

Ni

Zn

Ti

Al

Chemical composition (wt.%)

0.01

0.012

4.542

0.003

0.002

0.02

0.001

0.01

0.004

Base

3720

N. Nemati et al. / Materials and Design 32 (2011) 37183729

Fig. 1. (a) Schematic of cylindrical tool steel die and its dimensions used for cold and hot press and (b) auto-controlled power of the heating system used for compacts to be
hot pressed.

and 20 N for the normal load, 200 rpm for the sliding speed,
1500 m for the sliding distance and 25 C for the ambient
temperature.

Volume loss mm3

Weight loss g
 1000
Density g=mm3

Volume loss mm3


Sliding distancekm

Wear Rate mm3 =km

3. Results and discussion


3.1. Microstructural characterization
3.1.1. Milled Al + Cu powder mixtures
Fig. 2a and b shows the SEM micrographs of Al and Al4.5 wt.%
Cu (M) powder mixtures before and after 10 h of milling respectively. It can be seen that the morphology of as-received aluminum

powders (after milling the mixture for 10 h) changes from rather


spherical to plate-like, due to milling process of ductile powders
which leads to squeezing the powder particles. Although the milling time rst was set to be 70 h to observe the satisfactory condition for alloying [24], after 40 h milling Cu atoms were eventually
solvated in the Al crystal lattice, as shown in Fig. 3. As the time of
milling is the most important parameter, after each 10 h a proportion of powder was unloaded and structural changes were investigated and detailed studies on the microstructure of the powder
grain size and lattice strain accumulated were done applying Sherrer and WilliamsonHall methods [23,24]. According to Fig. 3a, the
XRD patterns of the powder produced at different stages of the
milling (Cu/Al powder mixture) show that Cu peaks are gradually
disappeared and Al peaks are broadened and shifted to greater degrees through milling process which is a strong probability to
prove the occurrence of alloying. Fig. 3b depicts the overlap peak
of the most intensive Al peaks gained in different intervals of milling. The arrows on the image indicate broadening and the shift of
the position of Al peaks toward greater degrees as a proof of

Fig. 2. SEM micrographs of: (a) as-received Al powder and (b) Al4.5 wt.% Cu (M) powders mixture after 10 h milling time.

N. Nemati et al. / Materials and Design 32 (2011) 37183729

3721

Fig. 3. XRD patterns of: (a) Cu/Al powder mixture produced and (b) most intensive Al peak positions produced at different intervals of milling.

diffusion of Cu atoms into Al lattice which caused alteration of the


lattice parameters of Al. A detailed report of the applied method
and results of synthesizing the alloying matrix with (Al4.5 wt.%
Cu composition as matrix alloy) is presented in previous studies
[23,24].
It can also be assumed that the intermetallics compounds are
dispersed into the Al as ultrane isolated particles which are virtually rather undetectable by XRD [25]. Other probabilities for elimination of Cu peaks are as below:
Continuous size reduction of the particles and strain induced
due to the sever deformation occurring to the particles and last
but not least, diffusion of Cu into the lattice of Al, which can be a
reasonable explanation for noticeable reduction and eventually
elimination of Cu peaks in the presented XRD patterns (Fig. 3a).
This mentioned procedure will be continued in further milling
times until 40 h which after this time, no sign of Cu peaks are observed in the pattern. In the nal milling time sequences (4070 h)
the geometry alteration of the peaks are solely due to the particle
size reduction and strain induced by M/A and that is because of the
absence of solute peaks (sufcient Cu atoms as they were consumed in the previous times of alloying). Work hardening during
milling resulted in continuous renement of matrix powder particles so that the average powder particle size was 0.5 to less than
10 lm after 40 h of ball milling.
3.1.2. Structural evolution in compacts
After exerting uniaxial presssinter methodology on the asmilled powders, the microstructure of the compact samples has
been studied applying optical and scanning electron microscopy.
According to Fig. 4, observations revealed that by increasing the
amount of reinforcing TiC nano-particles from 0% to 7 wt.%, the
grain size of the compacts decreases and the quality of grain junctions are weakened, while the TiC hard reinforcing nano-particles
can act as an obstacle against the grain boundary movement at
the time of the sintering and hence retard grain growth. Also, the
particle distribution in the sample containing 7 wt.% TiC decreases
because of higher tendency for clustering and agglomeration at
high levels of the reinforcing particles [26]. Fig. 4 shows the microstructure of the samples containing different weight percents of
TiC nano-particles (0%, 3%, 5%, 7 wt.%) with average particle size
of 30 nm. Fig. 4 depicts that as reinforcement weight percent in-

creases, the grain size decreases to some extent. The point to bear
in mind is the homogeneity of the composites is reduced in specimens containing more than 5 wt.% TiC nano and micron size particles. The reason for this phenomenon can be due to their higher
specic surface of ner particles in comparison with coarse particulates [27]. An increase in the specic surface leads to higher interparticle friction and thus leading to decreased particle distribution
and so weakening of grain junctions occurs [28]. It should be stated
that the addition of hard reinforcing particles affects both compressibility and sinterability of compacts. So, the relative density
of the compacts decreases and porosities are formed [26]. The
experimental, theoretical and relative density values of the compacts are shown in Table 2.
The relative density of the consolidated samples, measured by
Archimedes water immersion method, is demonstrated in Table
2. By increasing nano-size particulates portion, in comparison with
same portion of micron sized particulates, density rst increases
slightly and then decreases (10 wt.% TiCl). The reinforcing particles
which are smaller than the matrix particles situate among the matrix particles, lling pores. This causes an increase in the density of
the nano-particle reinforced composites compared with the matrix
and micron size particle reinforced composites. However, the increase in agglomeration of the particles by increasing the nanosized TiC particulates may lead to the decrease of the density, in
case of further TiC addition (>7 wt.%). It would be worth mentioning that nano-particles are prone to agglomerate and pores retained in the agglomerated zones, which may result in
decreasing the density [29].
Moreover, according to Fig. 5, XRD patterns reveal that no
detectable interaction layer is observed between TiC and Al alloy
matrix during different mixing intervals. Unequivocally, formation
of interaction layer between TiC and the matrix is very likely to occur especially after long mixing periods, the products (different
carbide or intermetallics) are so small, which is undetectable by
X-ray diffraction technique. Besides the production of this interaction will make mechanical properties to decline signicantly, so
the shortest time of mixing (2 h with planetary ball mill) was selected to ensure the minimum interaction and well distribution
to avoid the drawbacks.
In order to investigate the morphology and analysis of TiC nanoparticles, SEM image with high resolution and high magnication

3722

N. Nemati et al. / Materials and Design 32 (2011) 37183729

Fig. 4. Optical micrographs of the microstructure of the compacts with different contents of TiC nano-particles: (a) Al + 4.5 wt.% Cu(M), (b) M + 3 wt.% TiC, (c) M + 5 wt.% TiC
and (d) M + 7 wt.% TiC.

sion), is presented in Fig. 8ae. The EDX spectra clearly


demonstrate the rather uniform distribution of the TiC nano-particles through the matrix.

Table 2
Experimental and theoretical density values of consolidated samples.
Sample
name

Theoretical
density
(0.001 g/cm3)

Geometrical
density
(0.001 g/cm3)

Archimedes
density
(0.001 g/cm3)

Relative
density

Pure Al
Al + 4.5 wt.%
Cu (M)
M + 0.5% TiC
M + 1% TiC
M + 3% TiC
M + 5% TiC
M + 7% TiC
M + 5% TiCl
M + 7% TiCl
M + 10%
TiCl

2.701
2.982

2.601
2.752

2.601
2.833

0.97
0.95

2.991
3.001
3.041
3.081
3.121
3.081
3.121
3.176

2.706
2.680
2.672
2.671
2.692
2.758
2.743
2.688

2.841
2.880
2.888
2.895
2.803
2.772
2.746
2.763

0.95
0.96
0.95
0.94
0.90
0.90
0.88
0.87

3.2. Microhardness measurement

and corresponding EDX point analysis is provided in Fig. 6ac.


Cloudy shape of nanopowders is due to the agglomeration of
nano-particles for the great specic surface of the powders.
The presence of minimal porosities is expected from calculation
of relative density values, which is conrmed with experimental
results. Fig. 7a and b shows the presence of porosity on the surfaces
of M + 7 wt.% TiC and M + 3 wt.% TiC composites which are natural
results of consolidation process. However, passing optimum levels,
the more the reinforcing weight percent, the more the porosities
and the less the relative density.
In order to investigate the distribution homogeneity of reinforcing particulates, EDX analyses for the composite (5 wt.% TiC disper-

Fig. 9 shows the results of microhardness measurement of the


bulk aluminum, Al4.5 wt.% Cu alloy and composite samples.
According to the data population (ranging from 40 to 174 HV)
and standard deviation values obtained from the experiments
(Fig. 9), it can be seen that data is less scattered indicating that
the consolidated material has rather uniform and consistent properties through the sample. The average hardness values of pure
aluminum, (M) and (M) + 7% TiC nanocomposite were measured
to be 40 HV, 130 HV and 174 HV respectively. From Fig. 9, it is also
clear that the hardness of the matrix is remarkably improved in
comparison with bulk pure aluminum which is truly the result of
mechanical alloying. As expected, the natural ageing of the composition that occurs through hot press process may affect the hardness results [29]. In other words the main reason for this
enhancement can be due to the formation of Al (Cu) solid solution
and further more heat treatment process that naturally occurs during hot press. Increased dislocation density and microstrain and
also grain size renement due to the milling process which leads
to work hardening of the milled materials can be considered as
other reasons for that phenomenon [16,3032].
Furthermore, it is understood that increasing the weight percentage of the hard phase in the composite in case of nanoscale
TiC particulates results in a great, gradual increase in hardness
values (Fig. 9). This increase is expected because hard particulate

N. Nemati et al. / Materials and Design 32 (2011) 37183729

3723

effect of Orowan strengthening mechanism decreases [33]. According to the literature, the optimum levels of Al2O3 and SiC also had
been detected to be effective in enhancing mechanical properties
of composites [2,5]. In case of micron size TiC particulates, by
increasing its content from 5% to 10 wt.% the hardness decreases
slightly.
3.3. Investigation of wear behavior

Fig. 5. XRD patterns of TiC particulates and alloying matrix (M) mixture in different
intervals of mixing via ball milling.

reinforcements act as a barrier to the dislocation movement within


the matrix and exhibit greater resistance to the indentation of the
hardness tester. The reason for that is mainly due to the decreased
inter-particle distance in case of using nanometric reinforcement
(at constant vol.% in comparison with micrometric particles),
which results in more effectiveness action of Orowan mechanism
[5,32,33]. In addition, depending on the particle size after a certain
amount of added particle, they can easily agglomerate to form
clusters. The agglomeration phenomena and poor distribution of
hard particles cause a signicant reduction in mechanical properties of the composite specimens. In this case, the actual inter-particle distance becomes larger than the expected distance, therefore,

To investigate the size effect of reinforcing particulates a comparison is made between the composites reinforced with nano
and micron size TiC. Figs. 10 and 11 show the weight loss values
for all specimens under 10 and 20 N applied loads. It is generally
believed that contribution of hard particles to aluminum alloys results in an improvement of the wear resistance of the base alloy to
a great extent [34,35]. Based on the results, under both loads the
amount of weight loss of the all nano-particle reinforced composites is less than that of micron size TiC reinforced condition. The
main reason for this fact is clearly related to the hardness enhancement of nano-size TiC reinforced samples and according to Archad
equation:

Q K

W
H

In this equation, Q is wear rate (mm3/km), the volume of worn


material per distance, K is a constant called wear coefcient and H
is the hardness of the specimen in Vickers scale (kg/mm2) [35].
From Archad equation, it is clear that the wear resistance of the
materials is directly related to their hardness. In this work, the
maximum wear resistance enhancement due to the reinforcing
particle inuence, is for M + 5 wt.% TiC nano-particle composite.
According to Fig. 12, it is evident that the wear rate of the composites decreases with increasing TiC nano-particle volume fraction and the improvement trend gradually decreases with

Fig. 6. (a) SEM image of TiC nano powders, (b) high magnication SEM image of TiC nano-particles indicating the formation of agglomerated powders and (c) corresponding
EDX point analysis of TiC powders coated with Au.

3724

N. Nemati et al. / Materials and Design 32 (2011) 37183729

Fig. 7. SEM micrographs of the surfaces of consolidated: (a) M + 7 wt.% TiC and (b) M + 3 wt.% TiC nanocomposites.

Fig. 8. (a) SEM image, (b) corresponding line scan EDX spectra and EDX elemental maps of (c) Al, (d) Cu and (e) Ti of the surface of M + 5 wt.% TiC nanocomposite.

N. Nemati et al. / Materials and Design 32 (2011) 37183729

3725

Fig. 9. Microhardness test results of bulk aluminum, Al4.5 wt.% Cu (M) and
composites along the diameter of consolidated samples.

Fig. 11. Weight loss values under 20 N applied load for: (a) bulk Al, alloying matrix
(M) and TiC nano-particle reinforced specimens and (b) micron size TiC reinforced
specimens.

Fig. 10. Weight loss values under 10 N applied load for: (a) bulk Al, alloying matrix
(M) and TiC nano-particle reinforced specimens and (b) micron size TiC reinforced
specimens.

increasing reinforcement content. As expected, the weight loss and


wear rate for all samples under 20 N applied load is more for 10 N
load. It can be explained that at higher load the separation of surface asperities and hence the number of contact asperities increase
and consequently the worn surface material increases [36,37]. It is
generally believed that the incorporation of hard particles to aluminum alloys contributes to the improvement of the wear resistance of the base alloy to a great extent [35,38]. As seen in
Fig. 13, adding TiC improves the sliding wear resistance of the matrix alloy signicantly. TiC nano-size particles in the composite
may act as a load-supporting element and this hard particles resist
against destruction action of the abrasive and protect the surface,
especially at optimum level (5 wt.%). With increasing its content,
the wear resistance enhances but it seems this enhancement continues until the reinforcement can improve the mechanical properties such as hardness. As mentioned before, in case of nano-sized
particle clustering, the trend of wear resistance improvement
may be affected by the particle agglomeration.
In case of micron size reinforcements, it is evident that the coarser the TiC particles, the higher the abrasion and wear rates, as
shown in Fig. 12. It is interesting to note that similar behavior is
seen between wear rate results of nano and micron size TiC particles at higher contents (>5 wt.%). Although the high probability of
particle agglomeration and lack of even homogeneity in the
structure of the consolidated composite (with higher levels of

3726

N. Nemati et al. / Materials and Design 32 (2011) 37183729

Fig. 12. Variation of wear rate values of the composites, Al and matrix under 10 and
20 N applied loads.

Fig. 14. Variation of the friction coefcient (l) of: (a) M + 5 wt.% TiC and (b) M
samples via sliding distance under 20 N applied load.

Fig. 13. Variation of the friction coefcient (l) of: (a) M + 5 wt.% TiC and (b) M
samples via sliding distance under 10 N applied load.

nano-size TiC particles) is the main reason for the increased wear
rates, the fragmentation of coarse and brittle TiC particles during
wear test may contribute to the increased wear rates of the composite containing micron size TiC particles. As a result lower hardness and wear resistance will be achieved.

As expected, the weight loss and wear rate values for all samples under 20 N applied load is more for 10 N load. It can be explained that at higher load the separation of surface asperities
and hence the number of contact asperities increase and consequently the worn surface material increases [37]. Based on the
weight loss data, specimens with 5 wt.% TiC and without reinforcement have the highest wear resistance among the all tested samples, also unreinforced pure aluminum give the lowest wear
resistance.
Figs. 13 and 14 show the variation of the friction coefcient (l)
of samples via sliding distance under the both applied loads. The
average friction coefcient of Al4.5 wt.% Cu + 5 wt.% TiC sample
was 0.31 under 10 N applied load which is much smaller than that
for the matrix sample (0.575) in the same condition. This is mainly
due to the high hardness of the nanostructured composite. The
reinforcement particles support the load, so as to lessen the touch
area between the pin and counter disk surface, decrease the friction coefcient and can prevent the scratch and cut from the surface [38]. Also, a comparison between Figs. 13 and 14 reveals
that increasing the normal load from 10 to 20 N has no signicant
effect on the variation trend of l values, but in case of matrix specimen the friction coefcient is higher in 20 N applied load. Similar
results for aluminum alloy matrix composites have been reported
by Bermudez et al. [39]. For all specimens, initially friction coefcients have a quick increase with sliding distance to an approximately steady range after about 50 m. This increase in l

N. Nemati et al. / Materials and Design 32 (2011) 37183729

3727

Fig. 15. SEM micrographs of worn surfaces of: (a) bulk pure Al, (b) alloying matrix (M), (c) M + 3 wt.% TiC, (d) M + 5 wt.% TiC and (e) M + 7 wt.%TiC nanocomposites under
10 N applied load.

corresponds to an increase in apparent contact area, until full contact occurs across the full diameter of the pin.
Results of SEM revealed complex wear mechanisms on the sliding surface of the composites and the aluminum alloy matrix, as
seen in Figs. 15 and 16. Three different wear mechanisms were observed to coexist, namely abrasion, delamination and adhesion.
Surface coverage of impression for each of the mechanisms depends on the load applied normal to the worn surface. Identication of the type of wear mechanisms involved depends on
surface morphology of the worn specimen. SEM results were performed after completion of the wear experiments. Fig. 15 shows
the wear surfaces of samples under 10 N applied load which reveals great details of the contributed wear mechanisms. As it is
clear, the worn surface of the aluminum consists of large delaminated areas and no sign of grooves are observed (solid arrows indicate abrasive wear and bald arrows indicate delaminating or
adhesive ones in these gures), in contrast as can be seen in other
worn surfaces micrographs, as TiC nano-particle reinforcement
fraction is increased through the matrix the surface coverage

impression alters gradually and grooves parallel to the sliding vector are observed.
Distinct ner grooves and the abrasive zones have become
more, especially for the unreinforced alloying matrix. The worn
surface observation suggests that the dominant wear mechanism
for samples containing different levels of TiC nano-particles is
abrasive wear accompanied by some delaminating wear mechanism under both loads, regarding to the fact that for the composite,
because of the hard reinforcement presence and higher hardness,
the portion of adhesive mechanism is less or ignorable, the worn
surface is smoother and grooves are ner and also the total depth
of deformation is smaller than for the aluminum alloy matrix. The
depth of deformation should be linearly related to the true contact
area and the asperity contact radius. In case of micron size TiC reinforced samples, as can be seen in Fig. 16 there is no sign of grooves
and mainly adhesive wear mechanism is present in the surface of
the samples with 7% and 10 wt.% TiC. This is due to lower hardness
of these samples and also excessive delaminated areas are formed
due to particle decohesion during wear test, as hard micron size

3728

N. Nemati et al. / Materials and Design 32 (2011) 37183729

Fig. 16. SEM micrographs of worn surfaces of micron size TiC reinforced composites: (a) M + 7 wt.% TiC and (b) M + 10 wt.% TiC, under 10 N applied load.

particulates have not been distributed evenly through the matrix


the agglomerates are decohered. It was observed that under 20 N
applied load, for most of the specimens and composites the specic
wear mechanism is delaminating. It has been also reported that
when the applied load is more than 10 N, the wear mechanism alters and delamination wear is predominates [40].
4. Conclusions
The overall results of this research are as following:
1. The hardness of nano-particle reinforced composites was higher
than that of micron size TiC reinforced ones. With increasing
the volume fraction of nano-size TiC up to 5 wt.%, the hardness
of the composites continuously increased and after that level,
particle agglomeration reduced the amount of effective particulate available and the particle strengthening effect diminished.
2. The wear resistance of all samples reinforced with nano-size TiC
was higher than micron size TiC reinforced composites and the
matrix alloy. Increasing TiC content up to 5 wt.% resulted in
increased wear weight loss and wear rates of the specimens
with changing the applied load from 10 to 20 N.
3. Worn surface observation suggested that the dominant wear
mechanism for non reinforced pure Al specimen has been delaminating wear accompanied by some adhesive wear mechanism regarding to that, worn surfaces of the nano-particle
reinforced composites were smoother and the total depth of
deformations were smaller, grooves were ner than the unreinforced aluminum alloy matrix specimens.

Acknowledgements
The authors would like to thank University of Tehran and Sahand University of Technology for nancial supports of this
research.
References
[1] Lee HS, Yeo JS, Hong SH, Yoon DJ, Na KH. The fabrication process and
Mechanical properties of SiCp/AlSi metal matrix composites for automobile
air-conditioner compressor pistons. J Mater Process Technol 2001;113:2028.
[2] Kang YC, Chan SL. Tensile properties of nanometric Al2O3 particulatereinforced aluminum matrix composites. Mater Chem Phys 2004;85:43843.
[3] Miyajima T, Iwai Y. Effects of reinforcements on sliding wear behavior of
aluminum matrix composites. J Wear 2003;255:60616.

[4] DaCosta CE, Zapata WC, Velasco F, Ruiz-Prieto JM, Torralba JM. Wear behavior
of aluminum reinforced with nickel aluminide MMCs. J Mater Process Technol
1999;9293:6670.
[5] Zhang Z, Chen DL. Contribution of Orowan strengthening effect in particulate
reinforced metal matrix nanocomposites. J Mater Sci Eng A 2008;483
484:14852.
[6] Manna I, Nandi P, Bandyopadhyay B, Ghoshray K, Ghoshray A. Microstructural
and nuclear magnetic resonance studies of solid-state amorphization in AlTi
Si composites prepared by mechanical alloying. J Acta Mater
2004;52:413342.
[7] Wang Y, Rainforth WM, Jones H, Lieblich M. Dry wear behaviour and its
relation to microstructure of novel 6092 aluminium alloyNi3Al powder
metallurgy composite. J Wear 2001;251:142132.
[8] Salem H, El-Eskandarany SH, Kandil A, Abdul Fattah H. Bulk behavior of ball
milled AA2124 nanostructured powders reinforced with TiC. J Nanomater.
Hindawi Publishing Corporation; 2009. 12 p. [Article ID 479185].
[9] Hanada K, Murakoshi Y, Negishi H, Sano T. Microstructures and mechanical
properties of AlLi/SiCp composite produced by extrusion processing. J Mater
Process Technol 1997;63:40510.
[10] Tan MJ, Zhang X. Powder metal matrix composites: selection and processing. J
Mater Sci Eng A 1998;244:805.
[11] Fogagnolo JB, Robert MH, Torralba JM. Mechanically alloyed AlN particlereinforced Al-6061 matrix composites, powder processing, consolidation and
mechanical strength and hardness of the as-extruded materials. J Mater Sci
Eng A 2006;426:8594.
[12] Khakbiz M, Akhlaghi F. Synthesis and structural characterization of AlB4C
nanocomposite powders by mechanical alloying. J Alloy Compd
2009;479:33441.
[13] Mohammad Shari E, Karimzadeh F, Enayati MH. Mechanochemically
synthesized Al2O3TiC nanocomposite. J Alloy Compd 2009:102667.
[14] Zakeri M, Yazdani-Rad R, Enayati MH, Rahimipoor MR. Synthesis of MoSi2
Al2O3 nanocomposite by mechanical alloying. J Mater Sci Eng A
2006;430:1858.
[15] Razavi Hesabi Z, Hazpour HR, Simchi A. An investigation on the
compressibility of aluminum/nano-alumina composite powder prepared by
blending and mechanical milling. J Mater Sci Eng A 2007;454455:8998.
[16] Suryanarayan C. Mechanical alloying and milling. J Progrss Mater Sci
2001;46:1184.
[17] Manna I, Chattopadhyay PP, Banhart F, Fecht HJ. Development of amorphous
and nanocrystalline Al65Cu35xZrx alloys by mechanical alloying. J Mater Sci.
Eng A 2004;379:3605.
[18] Nandi P, Chattopadhyay PP, Pabi SK, Manna I. Solid state synthesis of Al-based
Amorphous and nanocrystalline AlCuNb alloys. J Mater Sci Eng A
2003;359:117.
[19] Roy D, Kumari S, Mitra R, Manna I. Microstructure and mechanical properties
of mechanically alloyed and spark plasma sintered amorphousnanocrystalline Al65Cu20Ti15 intermetallic matrix composite reinforced with
TiO2 nanoparticles. J Intermetall 2007;15:1595605.
[20] Meyers MA, Mishra A, Benson DJ. Mechanical properties of nanocrystalline
materials. J Progr Mater Sci 2006;51:427556.
[21] Subramaniam C. Some consideration towards the design of a wear resistance
Al alloy. J Wear 1992;155:153544.
[22] Bermudez MD, Carrion FJ, Iglesias P, Nicolas GM, Herrera EJ, Rodriguez JA.
Inuence of milling conditions on the wear resistance of mechanically alloyed
aluminum. J Wear 2005;258:90614.
[23] Amadora D, Ruiz-Navas EM, Torralba JM, Fogagnolo JB. Solid solution in
Al4.5 wt% Cu produced by mechanical alloying. J Mater Sci Eng A 2006;433:
4549.

N. Nemati et al. / Materials and Design 32 (2011) 37183729


[24] Nemati N, Zolriasatein A, Emamy M, Khosroshahi RA. Fabrication and alloying
behavior of nanostructured Al4.5wt%Cu alloy in solid state by mechanical
alloying. The 2nd national nano materials and nano technology conference.
Isfahan, Iran; May 45, 2010.
[25] ASTM. Standard test method for wear testing with a pin-on-disc apparatus.
ASTM G99-95. Philadelphia: PA; 1995.
[26] Tabandeh Khorshid M, Jenabali Jahromi SA, Moshksar MM. Mechanical
properties of tri-modal Al matrix composites reinforced by nano- and
submicron-sized Al2O3 particulates attrition milling and hot extrusion
developed by wet. J Mater Des 2010;31:38804.
[27] Rahimian M, Parvin N, Ehsani N. Investigation of particle size and amount of
alumina on microstructure and mechanical properties of Al matrix composite
made by powder metallurgy. J Mater Sci Eng A 2010;527:10318.
[28] Slipenyuk A, Kuprin V, Milman Y, Goncharuk V. Eckert J. J Acta Mater
2006;54:15766.
[29] Jafari M, Enayati MH, Abbasi MH, Karimzadeh F. Compressive and
wear behaviors of bulk nanostructured Al2024 alloy. J Mater Des 2010;31:
663669.
[30] Suryanarayana C, Koch CC. Nanocrystalline materialscurrent research and
future directions. J Hyperne Interact 2000;130:544.
[31] El-Eskandarany SHMS. Mechanical alloying for fabrication of advanced
engineering materials. New York: Noyes Publications; 2001.

3729

[32] Roy D, Singh SS, Basu B, Lojkowski W, Mitra R, Manna I. Studies on wear
behavior of nano-intermetallic reinforced Al-base amorphous/nanocrystalline
matrix in situ composite. J Wear 2009;266:11138.
[33] Ma ZY, Li YL, Liang Y, Zheng F, Bi J, Tjong SC. Nanometric Si3N4 particulatereinforced aluminum composite. J Mater Sci Eng A 1996;219:22931.
[34] Lashgari HR, Suzadeh AR, Emamy MT. He effect of strontium on the
microstructure and wear properties of A35610%B4C cast composites. J
Mater Des 2010;31:218795.
[35] Hosseini N, Karimzadeh F, Abbasi MH, Enayati MH. Tribological properties of
Al6061Al2O3 nanocomposite prepared by milling and hot pressing. J Mater
Des 2010;31:477785.
[36] Miracle DB. Metal matrix composites from science to technological
signicance. J Compos Sci Technol 2005;65:252640.
[37] Abouelmagd G. Hot deformation and wear resistance of P/M aluminum metal
matrix composites. J Mater Process Technol 2004;155156:1395401.
[38] Zhiqiang S, Di Zh, Guobin L. Evaluation of dry sliding wear behavior of silicon
particles reinforced aluminum matrix composites. J Mater Des 2005;26:4548.
[39] Bermudez MD, Martinez-Nicolas G, Carrion FJ, Martinez-Mateo I, Rodriguez JA,
Herrera EJ. Dry and lubricated wear resistance of mechanically-alloyed
aluminum-base sintered composites. J Wear 2001;248:17886.
[40] Alpas AT, Zhang J. Effect of SiC particulate reinforcement on the dry sliding
wear of aluminiumsilicon alloys (A356). J Wear 1992;155:83104.

You might also like