Sunteți pe pagina 1din 10

M A TE RI A L S CH A RACT ER IZ A TI O N 60 ( 20 0 9 ) 3 2 7 –3 3 6

Microstructure, mechanical properties, electrical conductivity


and wear behavior of high volume TiC reinforced
Cu-matrix composites

Farid Akhtar a,b,⁎, Syed Javid Askari a , Khadijah Ali Shah a , Xueli Du a,c , Shiju Guo a
a
Institute of Powder Metallurgy, Materials Science Department, University of Science and Technology Beijing, 100083, China
b
Department of Metallurgical and Materials Engineering, University of Engineering & Technology, Lahore 54890, Pakistan
c
School of Materials Science and Engineering, Tianjin University of Technology, Tianjin 300191, China

AR TIC LE D ATA ABSTR ACT

Article history: This study deals with the processing, microstructure, mechanical properties, electrical
Received 22 February 2007 conductivity and wear behavior of high volume titanium carbide reinforced copper matrix
Received in revised form composites. The microstructural study revealed that the titanium carbide particles were
21 August 2008 distributed uniformly in the matrix phase. No interface debonding and micro-cracks were
Accepted 28 September 2008 observed in the composite. The addition of alloying elements in the copper considerably
increased the sintered density and properties. The composite hardness and strength
Keywords: increased with titanium carbide content and alloying elements in the matrix phase. The
Titanium carbide composites electrical conductivities of the composites were predicted using three point upper bound
Microstructure and two phase self consistent predictive models. The wear resistance of the composites was
Electrical conductivity studied against high speed steel. Wear mechanisms were discussed by means of microscope
Mechanical properties observations on the worn surfaces. The ratio of titanium carbide average grain size to the
Wear mean free path of the binder was introduced as a parameter to determine wear
performance.
© 2008 Elsevier Inc. All rights reserved.

1. Introduction strength above 500 °C due to the structural instability


associated with the coarsening of the precipitate phase [6,7].
Copper is used in industrial products because of its high On the other hand, it is more appropriate to incorporate hard
electrical and thermal conductivities, low cost, ease of phase particles such as carbides, oxides, borides into copper
fabrication and good corrosion resistance. However, the i.e. by developing copper base metal matrix composites.
relative low hardness, low tensile strength and poor wear Metal matrix composites with ceramic reinforcement
resistance limits its applications. There are generally two particles are candidates for structural applications in the
ways to improve the mechanical properties and wear resis- wear industry, primarily because of their superior toughness
tance of copper; either by an age hardening mechanism or by and wear resistance. Alumina/silicon carbide particles are
incorporation of a hard second phase [1–5]. In the case of age often used as dispersoids to reinforce copper, which results in
hardening, addition of small amounts of chromium or superior elevated temperature strength, increased hardness
zirconium to copper can lead to the precipitation of a hard and improved wear resistance [8–12]. Cu-matrix composites
second phase, which is not soluble in the copper matrix at are promising candidates for sliding contact applications,
lower temperatures. These age-hardenable alloys show lack of where high electrical/thermal conductivity and good wear

⁎ Corresponding author. Institute of Powder Metallurgy, Materials Science Department, University of Science and Technology Beijing,
100083, China.
E-mail address: faridmet22@hotmail.com (F. Akhtar).

1044-5803/$ – see front matter © 2008 Elsevier Inc. All rights reserved.
doi:10.1016/j.matchar.2008.09.014
328 MA TE RI A L S CH A R A CT ER IZ A TI O N 60 ( 20 0 9 ) 3 2 7– 3 3 6

Table 1 – Fretting conditions used in the present study matrix powders were wet milled in ethanol for 48 h. The
Parameter magnitude Equipment range milled powders were poured into the die and pressed
uniaxially at 500 MPa pressure. The green compact was
Normal load 100 N–300 N Up to 1200 N
37.5 mm in diameter. The prepared green compacts were
Oscillating amplitude 1 mm 1–15 mm
sintered in a temperature range from 1200 °C to 1350 °C for 1 h
Oscillating frequency 40 Hz 200 Hz
Counter body High speed steel (10 mm) in vacuum. The sintering cycle applied to the samples was as
Specimen temperature Room (22 °C) follows; samples were heated to 900 °C at a rate of 10 °C/min
Relative humidity range 50–70% and held at 900 °C for 15 min, then the samples were heated to
Humidity not controlled the desired sintering temperatures at a rate of 5 °C/min and
they were held at sintering temperature for 1 h. The sintered
density was measured by the Archimedes water immersion
method. Hardness was measured on a micro-indentation
resistance are needed. In the case of carbide reinforced copper hardness tester. Three-point bend tests were conducted at a
matrix composites, several studies have been reported on SiC strain rate of 0.25 mm/min. Samples for bend strength
reinforced copper matrix composites [11–14]. However, the measurement were machined from the sintered samples in
literature on the use of other carbides (titanium carbide/ cuboids and polished on SiC emery papers up to 1000#. At least
tungsten carbide) for reinforced Cu-matrix composites is very four tests for each composition were conducted under the
limited [15,16]. same conditions. For metallographic study, the samples were
Ductile copper generally exhibits poor wear resistance.
Material removal in the form of wear debris is commonly
observed on the copper surface during sliding wear on a hard
steel counterpart [17]. Oxides, carbides, and borides are widely
incorporated in the soft and ductile materials like aluminium,
to reduce the extent of wear deformation in the subsurface
regions. Consequently, the wear resistance of Al is increased
considerably [18–20].
As mentioned above, Cu-base metal matrix composites are
candidates for sliding contacts. It is very practical to improve
wear resistance and to understand wear behavior of Cu-base
composites. In this study, we attempted to fabricate Cu-base
TiC reinforced composites with a high volume fraction of TiC
particles (69 vol.% to 77 vol.%). The titanium carbide particle is
an interesting candidate for reinforcement of copper because
of its high melting point, high hardness and superior wear
resistance.
The copper matrix composites are prepared via a powder
metallurgy (P/M) route. P/M has been used for several years as
a low-cost fabrication method for producing near-net shape
composites; therefore, P/M may be a cost-effective method for
producing TiC–Cu matrix composites. P/M technology was
selected as the processing method, due to the simplicity
associated with the incorporation of a hard phase into a metal
phase through the blending of powders [21]. It is ideal to
prepare copper matrix composites via powder metallurgy
because of its efficient dispersion of the fine reinforcing
particles. The fretting wear behavior of the Cu-matrix
composites is studied under different wear conditions.

2. Experimental

In this study, powder technology is used to prepare the Cu-


matrix TiC reinforced composites. The starting materials were
titanium carbide powders (5 μm), copper powders (N37 μm),
nickel powders (N37 μm), titanium powders (N37 μm), cobalt
powders (N37 μm) and aluminium powders (N37 μm). Five
different compositions were fabricated. Pure Cu, Cu–Ti–Al
(66 at.% Cu) and Cu–Ni–Co (84.5 at.% Cu) were used as the Fig. 1 – SEM micrographs of composites sintered at 1300 °C
binder phase for TiC reinforcements. TiC was incorporated in (a) 77 vol.% TiC–Cu composite (b) 77 vol.% TiC–Cu–Ti–Al
the composite from 69 vol.% to 77 vol.%. TiC and Cu base composite (c) 77 vol.% TiC–Cu–Ni–Co composite.
M A TE RI A L S CH A RACT ER IZ A TI O N 60 ( 20 0 9 ) 3 2 7 –3 3 6 329

prepared according to standard metallographic procedure. phase is homogeneous. Homogeneous distribution of the
The microstructure was characterized in a scanning electron reinforcements ensures the isotropic mechanical properties
microscope. Quantitative metallography was used to deter- and uniform distribution of stresses in the sintered composite
mine the microstructural parameters; average TiC grain size [22–25]. Fig. 1(a) shows the scanning electron micrograph of
and mean free path of the binder phase. At least, three TiC reinforced copper matrix composite containing 77 vol.%
micrographs taken from different areas were used to obtain an TiC particles. A range of reinforcement size is present from
accurate measurement of microstructural parameters. The DC 0.2 μm to 6 μm. The TiC particles have retained their shape.
electrical conductivities of the TiC reinforced copper matrix This is due to the low solubility of TiC particles in a pure
composites were measured at room temperature after grind- copper matrix. Small and large TiC particles do not undergo
ing with successively smoother grades of SiC emery paper solution and re-precipitation process and give a range of
down to 1500#. particle size in the sintered composite.
A sphere-on-flat geometry was used for the low-amplitude Fig. 1(b) and (c) shows the micrographs of 77 vol.% TiC
oscillating dry sliding tests, in ambient conditions. The tests particles embedded in Cu–Ti–Al (66 at.% Cu) and Cu–Co–Ni
were performed under gross slip conditions. An amplitude of (84.5 at.% Cu) matrix respectively. Alloying elements Ti, Al, Co
1 mm and an oscillating frequency of 40 Hz were selected for and Ni are added to the copper matrix in order to increase
the tests in the present study. The applied dead loads were wettability of the matrix phase with the reinforcements. Cu–Ti–Al
150 N and 300 N. The values for the tangential force and the and Cu–Co–Ni binder phase is chosen carefully [26,27]. These two
displacement pass through an analogue/digital card into a binder phases are at the eutectic composition and form a liquid
computer where, at specified intervals, friction force displace- phase at 885 °C and 1040 °C respectively, which is well below the
ment loops were stored. The coefficient of friction (COF) was melting temperature of copper. This serves a dual purpose. First, it
calculated as the dissipated energy per cycle divided by the reduces the sintering temperature. Second, it increases the
normal load and twice the stroke. The friction data of the wettability of the binder phase with the reinforcements. The
fretting tests was presented both by the COF and the dissipated lack of literature on TiC reinforced copper composite is due to the
energy. The room temperature range was 18–22 °C and the fact that it is not easy to sinter TiC with a copper matrix. Fig. 1(b)
relative humidity 40–60%. High speed steel balls of hardness and (c) shows the distribution and morphologies of TiC particles
62.5HRC, with 10 mm diameter, were used as counter bodies, in a copper alloy matrix. The distribution is homogeneous and TiC
loaded with a dead weight against the TiC–Cu matrix reinforcements are found in the particle size range of 0.2 μm to
composites. The wear conditions are summarized in Table 1. 7 μm. The morphologies of TiC reinforcement in the alloy matrix
are different from the pure copper matrix. Some round TiC
particles indicate the dissolution process of TiC particles in the
3. Results and Discussion binder phase. This is also an indication of increased wettability
between the ceramic and metal phase, and is further supported
3.1. Microstructure by increased sintered density, mechanical properties and wear
resistance which are discussed below in this study. Absence of
Three different copper base binders were used to process high interface debonding and microcracks ensures higher mechanical
volume fraction TiC reinforced copper matrix composites. The and wear properties.
micrographs of the composites are shown in Fig. 1. It shows
that TiC particles are surrounded by the matrix phase. A very 3.2. Mechanical Properties
important factor influencing the structure and properties of
the composite is the ceramic metal interface [22,23]. Interface An overview of the mechanical properties of high volume TiC
debonding is not found in the composite. The matrix phase reinforced copper matrix composites is given in Table 2. It
and reinforcement do not show interfacial debonding, indicat- shows that the hardness and strength of the composites
ing the good wettability of TiC particles with the binder phase. increase with increase in density and TiC content. Due to a
For the TiC reinforced copper matrix composites, it is very high volume fraction and higher elastic modulus of TiC than
important to obtain homogeneous reinforcements in the the copper phase, most of the load is applied to the TiC phase,
matrix phase in order to enhance mechanical and electrical when an external load is applied. Therefore, the hardness of
properties. The distribution of TiC particles in the matrix TiC reinforced copper composite, which represents resistance

Table 2 – An overview of the properties of the TiC–Cu base matrix composites


Binder TiC Hardness %Relative density Bending strength Bending strength (MPa) of
(vol.%) (HV) (MPa) TiC-465steel

Cu 77 544 ± 10 93.4 335 ± 20 791 ± 25


Cu–Ti–Al 69 514 ± 10 98.2 748 ± 20 1034 ± 25
Cu–Ti–Al 77 682 ± 10 98.8 787 ± 20 791 ± 25
Cu–Ni–Co 69 518 ± 10 98.9 682 ± 20 1034 ± 25
Cu–Ni–Co 77 719 ± 10 98.6 814 ± 20 791 ± 25

TiC-465 stainless steel composite is with equal volume fraction of TiC particles.
330 MA TE RI A L S CH A R A CT ER IZ A TI O N 60 ( 20 0 9 ) 3 2 7– 3 3 6

of alloying elements in the binder phase. The composite


containing Ni and Co as alloying elements in copper shows the
maximum hardness (719HV). Hardness values of the compo-
sites are given in Table 2 and are suitable to meet the
requirement for sliding contact applications.
The bending strength of the composite shows a similar
trend. With increase in volume percent of TiC particles,
bending strength increases. It is well known that copper has
very low strength. S.C. Tjong et al. incorporated SiC in copper
matrix up to 20 vol.%. They observed a large decrease in the
yield strength and elongation of the composite [4]. Similarly,
there are some other studies dealing with low volume fraction
ceramic reinforced copper matrix composites [3,28]. Recently,
P. K. Deshpanday et al. introduced high volume fraction WC
reinforced copper composite by melt infiltration [16]. They
obtained a maximum hardness of 368HV with 53 vol.% WC.
However, strength of the composite is not mentioned. In our
case, we measured the bending strength of the TiC reinforced
copper composite and, for the purpose of comparison, TiC
reinforced stainless steel matrix composites containing an
equal volume fraction of TiC reinforcements are used, due to
the lack of literature on high volume fraction ceramic
reinforced copper matrix composites. TiC reinforced copper
matrix composites show the least strength due to the lowest
density. Defects such as porosity, macro and microcracks and
interface debonding promote premature failure and ultimate
fracture occurs earlier. In the case of TiC reinforced copper
alloy matrix, a high bending strength is obtained. The TiC
reinforced copper alloy composite shows an increase in
strength with increase in TiC volume fraction, which is
contrary to the previous findings in the Cu–SiC system [4].
Increase of strength with increase in TiC volume fraction
demonstrates the stability of TiC reinforcements in a copper
alloy matrix and is a measure of TiC copper alloy matrix
interfacial bond. The values of strength of TiC reinforced
copper alloy composite are comparable in the case of 77 vol.%
reinforcement content, with the TiC reinforced steel compo-
site prepared by the same processing route and an equal
Fig. 2 – SEM fracture morphologies of composites sintered at
volume fraction of TiC particles.
1300 °C (a) 77 vol.% TiC–Cu composite (b) 77 vol.% TiC–Cu–Ti–Al
The fracture morphologies of the TiC reinforced copper
composite (c) 77 vol.% TiC–Cu–Ni–Co composite.
matrix composites are shown in Fig. 2. Fig. 2(a) shows the
fracture morphologies of TiC reinforced pure copper matrix. It
against plastic deformation, is mainly dependent on the shows the TiC reinforcements, copper matrix and porosity.
volume fraction of TiC particles in the composite. For sliding The fracture mode is mixed. The fracture path passes through
contact applications, hardness of the copper reinforced the TiC particles and the binder phase. No secondary cracks
composite is an important parameter and determines the are observed in the brittle TiC grains indicating premature
wear resistance of the composite. The hardness values of the failure of the composite. The binder phase decohesion is not
composites containing alloying elements are higher than the observed and shows the interfacial bond in strong. Fig. 2(b)
composite containing pure copper as a binder phase. This is and (c) shows the fracture morphologies of the TiC reinforced
due to increase in the density and the solute hardening effect copper alloy (Cu–Ti–Al) and (Cu–Co–Ni) matrix composite

Table 3 – Electrical conductivities of the TiC reinforced copper matrix composites


Composite TiC Binder Electrical conductivity Three point upper bond model Two phase self-consistent
(vol.%) (vol.%) (Exp) %IACS (%IACS) (%IACS)

TiC–Cu 77 23 2.24 19 8
TiC–Cu–Ti–Al 69 31 5.10 27 15
TiC–Cu–Ti–Al 77 23 4.10 19 8
TiC–Cu–Ni–Co 69 31 6.65 27 15
TiC–Cu–Ni–Co 77 23 4.64 19 8
M A TE RI A L S CH A RACT ER IZ A TI O N 60 ( 20 0 9 ) 3 2 7 –3 3 6 331

Porosity is absent, indicating that the addition of alloying


elements to the binder phase at the eutectic composition (with
a liquid phase well below the melting point of pure copper) is
effective in reducing the porosity in the sintered composite.

3.3. Electrical Conductivity

The TiC content in the composites varies from 69 vol.% to 77 vol.


%. As alloying elements are added in the binder phase, the
density increases, because the wettability of binder with the
reinforcements is increased, and the liquid phase appears at
lower sintering temperatures. The electrical conductivity
decreases with increasing TiC content because the volume
fraction of nonconducting particles increases. The electrical
conductivities of TiC reinforced Cu base matrix composite are
predicted using the predictive three point upper bound model
[29], the two phase self-consistent model [30] and are compared
to those of present composites (experimentally measured) and
are given in Table 3. It is found that the three point upper bound
model has large differences in electrical conductivities com-
pared to the experimental values because the model takes into
account microscopically isotropic geometries of either conduct-
ing or nonconducting phases. Also, alloying elements are added
to the matrix phase. The experimental values are closer to the
two phase self-consistent model. The differences are due to the
addition of alloying elements in the copper phase, in the case of
copper alloy matrix, and porosity in the case of a pure copper
matrix. It can be seen from Table 3 that the addition of alloying
elements to the copper base matrix increased the electrical
conductivity. This is due to the elimination of porosity from the
composite. Ideally the alloying elements additions (Ni,Co,Ti,Al)
adversely affect the electrical conductivity. But in our case with
pure copper, porosity present in the composite reduces the
electrical conductivity to a larger extent as shown in Table 3.
With the addition of alloying elements the porosity phase is
reduced/eliminated and the overall increase in the electrical
conductivity value is observed (Table 3). It is also worth
mentioning that when we compare the electrical conductivities
of composites containing Cu–Ti–Al (66 at.% Cu) and Cu–Ni–Co
(84.5 at.% Cu) as binders, Cu–Ni–Co composites show greater
electrical conductivity. This is due to the greater amount of
copper (84.5 at.% Cu) present in the Cu–Co–Ni binder phase.

3.4. Wear Properties

Fig. 3 shows the variation of weight loss with fretting distance of


450 m for the TiC reinforced Cu base matrix composites with
69 vol.% and 77 vol.% TiC content tested under an applied load of
150 N and 300 N at 40 Hz frequency with 1 mm amplitude. Fig. 3(a)
Fig. 3 – (a) Wear loss (mg) of TiC–Cu composite with 77 vol.%
shows the wear loss of TiC–Cu composite with 77 vol.% TiC
TiC content as a function of fretting distance; (b) wear loss
content at 150 N and 300 N load. The wear loss increases with the
(mg) of TiC–Cu–Ti–Al composite with 77 vol.% and 69 vol.%
fretting distance and applied load. At higher load (300 N) the wear
TiC content as a function of fretting distance; (c) wear loss
loss of the composite is very high. The hardness of the counter
(mg) of TiC–Cu–Ni–Co composite with 77 vol.% and 69 vol.%
body is 62.5HRC, which is much higher than the hardness of
TiC content as a function of fretting distance.
copper. The copper matrix is removed by the hard counter body
and this removal is very rapid under a 300 N load.
The wear mechanisms operative at 150 N are plastic
respectively. The fracture path passes through the TiC grains ploughing and grooving of the copper base matrix phase. At
and matrix phase. A few secondary cracks are visible. Binder higher loads, removal of the copper from the surface is very high.
phase decohesion is not observed. The material loss is considered to result from removal of material
332 MA TE RI A L S CH A R A CT ER IZ A TI O N 60 ( 20 0 9 ) 3 2 7– 3 3 6

Fig. 4 – An overview of the hardness and wear loss of the TiC–Cu


base matrix composites at different wear loads.

chips from the specimen, due to the microcutting of the abrasive


particles. Residual porosity plays an important role when the
wear is performed under a high wear load [16]. So, the porosity
effect must be counted. Porosity in the wear surface of the
composite effectively reduces the contacting surface area and
thus increases the net wear load. Other factors that increase the
wear loss due to porosity are the notch effect and the shape of the
porosity. The TiC reinforced pure copper matrix composite shows
a higher wear loss due to the porosity effect. Fig. 3(b) shows the
wear loss vs. fretting distance of TiC reinforced Cu–Ti–Al binder. It
is apparent that the composite experiences a larger weight loss as
the amount of TiC decreases from 77 vol.% to 69 vol.%. Moreover, Fig. 5 – Variation of friction coefficient with the fretting
the wear loss of TiC cermets increases rapidly with increasing distance of TiC–Cu base matrix composites (a) at 150 N load;
fretting distance, indicating that the counter body (high speed (b) at 300 N load.
steel) can deform and remove the material from TiC composites
progressively. However, a much slower increase in wear loss is
strongly depend on TiC loading in the composite and wear loss
observed at a lower load (150 N) for all TiC composites, whereas
has a direct relation with the hardness of the composite. At
the wear loss varies almost linearly with the fretting distance
lower loads and higher TiC content, the matrix deformation
under specific loading conditions. It can be noted that at higher
and removal is protected by the TiC particles. As the hardness
loads, material removal is fast. This implies that the extent of
of metal matrix composites (MMCs) tends to increase with
plastic deformation or material removal from the matrix of TiC
increasing TiC content, thus the fretting wear resistance of TiC
composite decreases with increase in TiC loading in the
cermets improves considerably with increasing TiC content.
composite. TiC reinforcing particles can resist plastic deformation
The variation of friction coefficient of 77 vol.% TiC
and provide protection to the soft copper alloy matrix. TiC
composite under different loading conditions is given in Fig. 5.
composite with 77 vol.% TiC content shows a little variation in
It can be seen that the friction coefficient follows similar trends
wear loss with sliding distance. This is because the TiC composite
to the wear performance. The friction coefficient increases with
containing 77 vol.% TiC exhibits the higher hardness than 69 vol.%
increase in load from 150 N to 300 N, and causes more wear loss
TiC, and more TiC particles resist plastic deformation of the
under higher loads. In the case of 77 vol.% TiC content in the
matrix phase.
Fig. 3(c) shows the wear loss of TiC–Cu–Ni–Co matrix Table 4 – Effect of binder (vol.%) change on the wear loss of
composite. It shows that the composite wear loss strongly the composite
depends on TiC loading in the composite. As the TiC content
Composite Binder Wear loss (mg) Wear loss (mg)
increases from 69 vol.% to 77 vol.%, the wear loss decreases. (vol.%) at 150 N at 300 N
This composite also shows that wear loss increases with the
TiC–Cu–Ti–Al 23 1.0 2.9
applied load.
TiC–Cu–Ti–Al 31 6.2 12.5
The comparison of hardness with wear loss at 150 N and
TiC–Cu–Ni–Co 23 1.0 2.1
300 N is summarized in Fig. 4. Comparing the hardness and TiC–Cu–Ni–Co 31 6.1 14.1
wear loss of composites it can be concluded that the wear loss
M A TE RI A L S CH A RACT ER IZ A TI O N 60 ( 20 0 9 ) 3 2 7 –3 3 6 333

Table 5 – Effect of ratio of average grain size of TiC to mean free path of binder phase (GSTiC/MFPbinder) on the wear
performance of the composite
Composite Binder Mean free path Average grain Ratio of average grain size of Wear loss (mg) Wear loss (mg)
(vol.%) (μm) of binder size of TiC (μm) TiC to mean free path of binder at 150 N at 300 N

TiC–Cu–Ti–Al 23 0.9 3.4 3.77 1.0 2.9


TiC–Cu–Ti–Al 31 1.2 3.7 3.08 6.2 12.5
TiC–Cu–Ni–Co 23 0.9 3.4 3.77 1.0 2.1
TiC–Cu–Ni–Co 31 1.25 3.6 2.88 6.1 14.1

composite, the wear mechanisms operative are polishing wear i.e. 600% increase in wear loss. The increase in binder volume
at lower load and polishing wear with plastic ploughing at fraction rapidly decreases the wear resistance of material at
higher wear loads. lower wear loads. The wear loss is very sensitive to the binder
Tables 4 and 5 summarize the wear loss of the composite volume fraction. Small variations in binder volume fraction
under different fretting conditions with the binder volume have a major effect on the wear rate.
fraction in the composite and the mean free path of the binder Table 5 shows the variation of wear loss with the mean free
phase. Table 4 shows that with an increase in the binder path of the binder phase. As long as the ratio of TiC particle
phase, the wear loss increases. With a binder increase of size to the mean free path is greater than the TiC particle size
31 vol.% the wear loss increases rapidly at lower load of 150 N in the composite, it shows higher wear performance at all

Fig. 6 – SEM micrographs of wear surfaces at 150 N load (a) 77 vol.% TiC–Cu composite; (b) 77 vol.% TiC–Cu–Ti–Al composite;
(c) 77 vol.% TiC–Cu–Ni–Co composite; (d) 69 vol.% TiC–Cu–Ti–Al composite; (e) 69 vol.% TiC–Cu–Ni–Co composite.
334 MA TE RI A L S CH A R A CT ER IZ A TI O N 60 ( 20 0 9 ) 3 2 7– 3 3 6

Fig. 7 – SEM micrographs of wear surfaces at 300 N load (a) 77 vol.% TiC–Cu composite; (b) 77 vol.% TiC–Cu–Ti–Al composite;
(c) 77 vol.% TiC–Cu–Ni–Co composite; (d) 69 vol.% TiC–Cu–Ti–Al composite; (e) 69 vol.% TiC–Cu–Ni–Co composite.

wear loads. When, the ratio of TiC particle size to the mean Fig. 6 shows the worn surfaces of the TiC composites tested
free path is lower than the TiC particle size in the composite, under an applied load of 150 N. The worn surface is
the wear properties of the composite show degradation and characterized by long continuous grooves, which form as the
when this ratio is decreased much the wear loss becomes very counter body ploughs across the surface and eventually
high even at lower wear loads. So, the binder volume fraction removes or pushes material into ridges along the sides of
and the ratio of the average particle size of the reinforcements the grooves. The surface morphologies of these worn surfaces
to the mean free path of the binder phase is the key factor show the ploughing effect on the copper matrix.
controlling the wear mechanisms and wear performance [31] The severity of micro-ploughing is substantially less
(In Tables 4 and 5, TiC reinforced pure copper matrix pronounced on the worn surface of TiC composite containing
composite is not included because it contains a large amount 77 vol.% TiC [Fig. 6(a), (b) and (c)]. Careful examination of the
of porosity). When the ratio of TiC particle size to the mean SEM micrographs reveals that reinforcing particles are
free path is greater than the TiC particle size in the composite exposed on the worn surface of MMC specimens at lower
(GSTiC/MFPbinder N GSTiC) the dominating wear mechanism is loads. It can be seen from these micrographs that TiC
polishing wear and less wear loss is observed. When the ratio reinforcing particles reside on the top surface of TiC cermet
of TiC particle size to the mean free path is lower than the TiC after fretting. No particle cracking and separation of particle
particle size in the composite (GSTiC/MFPbinder b GSTiC) the from the matrix can be observed at 150 N wear loads, shown in
dominating wear mechanisms are plastic ploughing, grooving Fig. 6. Thus, the volume fraction of TiC reinforcing particles is
and microcutting, and high wear loss is observed. very effective in increasing the fretting wear resistance.
M A TE RI A L S CH A RACT ER IZ A TI O N 60 ( 20 0 9 ) 3 2 7 –3 3 6 335

It is considered that the counter body (high speed steel 5. The wear resistance of the composites was studied against
62.5HRC) deforms the alloy matrix of the specimens during the high speed steel. Microploughing of copper base matrix
early stages of fretting wear and causes microploughing and was found to be the dominant wear mechanism. Less
grooving in the surface of copper base matrix. Subsequently, microploughing and wear loss was observed at lower loads
material chips are removed from the grooves formed in the and high vol.% of TiC particles. Heavy microploughing
surface, thereby exposing the hard reinforcing particles observed under a 300 N wear load caused very rapid
[18,32]. In this case, the hard TiC particles come in contact removal of material from the wear surface.
with the counter body. The large size of TiC particles can offer 6. As long as the ratio of TiC average grain size to mean free
protection to Cu/Cu alloy matrix during fretting. Once the path of the binder was higher than the average grain size of
reinforcing TiC particles fracture or loosen from the matrix the TiC, the material showed good wear performance.
phase, they can be removed easily, resulting in a certain When, this ratio was decreased to less than the average TiC
amount of wear loss. The hardness of the steel counter body is grain size, the plastic ploughing, grooving and cutting of
lower than the TiC particles, and microploughing was the the matrix phase was the dominant wear mechanism, and
dominant wear mechanism on 69 vol.% TiC [4,33]. In the case high wear loss was observed.
of 77 vol.% TiC, less binder phase was present and the wear
mechanisms were polishing wear and microploughing at a
lower load. It is apparent from the worn surfaces in Fig. 6 that REFERENCES
TiC cermet with a higher TiC content provides better wear
resistance.
Fig. 7 shows the worn surfaces of the TiC composites tested [1] Morris MA, Morris DG. Microstructures and mechanical
under an applied load of 300 N. The worn surface is properties of rapidly solidified Cu–Cr alloys. Acta Metall
1987;35:2511–22.
characterized by long and deep continuous grooves, which
[2] Lee YF, Lee SL, Huang CH, Lee CK. Effects of Fe additive on
form as the counter body plough across the surface under
properties of Si reinforced copper matrix composites fabricated
higher load and forces material into ridges along sides of the by vacuum infiltration. Powder Metall 2001;44:339.
grooves. The surface morphology of this worn surface shows [3] Tjong SC, Lau KC. Abrasive wear behavior of TiB2 particle-
very heavy ploughing of the copper matrix. At higher loads, reinforced copper matrix composites. Mater Sci Eng A
micro-ploughing is severe and cause fast removal of material 2000;282:183.
from the surface of the composite [Fig. 7(b) to (e)]. Thick [4] Tjong SC, Lau KC. Tribological behaviour of SiC particle-reinforced
copper matrix composites. Mater Lett 2000;43:274.
material chips are present on the surface of the composite as
[5] Dong SR, Tu JP, Zhang XB. An investigation of the sliding wear
shown in Fig. 7 produced, by heavy deformation of the matrix behavior of Cu-matrix composite reinforced by carbon
under high wear loads. The TiC–Cu matrix composite exhibits nanotubes. Mater Sci Eng A 2001;313:83.
the mechanism of removal of Cu matrix chips at 300 N load, as [6] Correia JB, Davies HA, Sellars CM. Strengthening in rapidly
shown in Fig. 7(a). This is due to the notch effect of porosity at solidified age hardened Cu–Cr and Cu–Cr–Zr alloys. Acta
higher wear stress. No particle cracking and separation of Mater 1997;45:177.
[7] Morris MA, Morris DG. Rapid solidification and mechanical
particle from the matrix can be observed under a 300 N wear
alloying techniques applied to Cu–Cr alloys. Mater Sci Eng
load, as shown in Fig. 7. So, the volume fraction of TiC
1988;104:201.
reinforcing particles and the ratio of TiC particle size to the [8] Upadhyaya A, Upadhyaya GS. Sintering of copper–alumina
mean free path are key factors controlling the wear rate of the composites through blending and mechanical alloying powder
TiC–Cu matrix composites at low and higher loads. metallurgy routes. Mater Design 1995;16:41.
[9] Wan YZ, Wang YL, Cheng GX, Tao HM, Cao Y. Effects of
processing parameters, particle characteristics, and metallic
coatings on properties of Al2O3 copper alloy matrix composites.
4. Conclusions
Powder Metall 1998;41:59.
[10] Broyles SE, Anderson KR, Groza JR, Gibeling JC. Creep
1. Three TiC reinforced copper base matrix composites were deformation of dispersion strengthened copper. Metall Mater
fabricated by powder technology. Alloying elements (Al, Ti, Trans A 1996;27:1217.
Ni and Co) were added to enhance the sintering and [11] Chang SY, Lin SJ. Fabrication of SiCw reinforced copper matrix
mechanical properties. composite by electroless copper plating. Scripta Mater
2. The microstructure of the composites revealed that the TiC 1996;35:225.
[12] Ma ZY, Tjong SC. High temperature creep behavior of in-situ
particles were distributed uniformly in the matrix phase.
TiB2 particulate reinforced copper-based composite. Mater Sci
The microstructure of TiC reinforced copper alloy matrix Eng A 2000;284:70.
composites was defect-free. No interface debonding and [13] Sundberg G, Paul P, Sung C, Vasilos T. Fabrication of CuSiC
micro-cracks were observed in the composite. metal matrix composites. J Mater Sci 2006;41:485.
3. The hardness and strength of the composite increased with [14] Wu TF, Qiu ZW, Lee SL, Lee ZG, Lin JC. Effects of graphite on
TiC content and alloying elements in the matrix phase. wear and corrosion behaviour of SiCp-reinforced copper
4. Electrical conductivities of the composites were predicted matrix composites formed by hot pressing. Corr Eng Sci
Technol 2004;39:229.
using three point upper bound model and two phase self-
[15] Sabatello S, Frage N, Dariel MP. Graded TiC-based cermets.
consistent model and compared to the experimentally Mater Sci Eng A 2000;288:12.
measure values. Experimental values were close to a two [16] Deshpande PK, Lin RY. Wear resistance of WC particle
phase self consistent model and seemed to be affected by reinforced copper matrix composites and the effect of
the addition of alloying elements in the copper binder. porosity. Mater Sci Eng A 2006;418:137.
336 MA TE RI A L S CH A R A CT ER IZ A TI O N 60 ( 20 0 9 ) 3 2 7– 3 3 6

[17] Alpas AT, Hu H, Zhang J. Plastic deformation and damage [26] Villars P, Prince A, Okamoto H, editors. Handbook of ternary
accumulation below the worn surfaces. Wear 1993;162–164:188. alloy phase diagrams, vol. 3. ASM International; 1997.
[18] Tjong SC, Lau KC. Properties and abrasive wear of TiB2/Al–4% [27] Villars P, Prince A, Okamoto H, editors. Handbook of ternary
Cu composites produced by hot isostatic pressing. Comp Sci alloy phase diagrams, vol. 6. ASM International; 1997.
Technol 1999;59:2005. [28] Tu JP, Wang NY, Yang YZ, Qi WX, Liu F, Zhang XB, et al.
[19] Xu J, Liu WJ. Wear characteristic of in situ synthetic TiB2 Preparation and properties of TiB2 nanoparticle reinforced
particulate-reinforced Al matrix composite formed by laser copper matrix composites by in situ processing. Mater Lett
cladding. Wear 2006;260:486. 2002;52:448.
[20] Du ZM, Li JP. Study of the preparation of Al2O3·SiCp/Al [29] Thovert JF, kim IC, Torquato S, Acrivos A. Bounds on the
composites and their wear-resisting properties. J Mater effective properties of polydispersed suspensions of spheres:
Process Technol 2004;151:298. an evolution of two relevant morphological parameters. J
[21] Akhtar F, Guo SJ, Yang X, Lian Y. Stainless steel binder for the Appl Phys 1990;67:6088.
development of novel TiC reinforced steel cermets. J Univ Sci [30] Landucer R. The electrical resistance of binary metallic
Technol Beijing 2006;13:546–50. mixtures. Appl Phys 1952;23:779.
[22] Akhtar F, Guo SJ. On the processing, microstructure, [31] Akhtar F, Guo SJ. Microstructure, mechanical and fretting
mechanical and wear properties of cermet/stainless steel wear properties of TiC-stainless steel composites. Mater
layer composites. Acta Mater 2007;55:1467–77. Charact 2008;59:84–90.
[23] Hussainova I. Effect of microstructure on the erosive wear of [32] Tjong SC, Lau KC. Abrasion resistance of stainless-steel
titanium carbide-based cermets. Wear 2003;255:121. composites reinforced with hard TiB2 particles. Comp Sci
[24] Akhtar F, Islam SH, Askari SJ, Tian J, Guo SJ. Effect of WC Technol 2000;60:1141.
particle size on the microstructure, mechanical properties [33] Skolianos S, Kattamis TZ, Chen M, Chembers BV. Cast
and fracture behavior of WC–(W, Ti, Ta) C–6 wt% Co cemented microstructure and tribological properties of particulate
carbides. Int J Ref Met Hardmater 2007;25:405–10. TiC-reinforced Ni-base or stainless steel matrix composites.
[25] Akhtar F, Guo SJ, Feng-e C, Feng PZ, Tao L. TiB2 and TiC Mater Sci Eng A 1994;183:195.
stainless steel matrix composites. Mater Lett 2007;61:189.

S-ar putea să vă placă și