Anda di halaman 1dari 10

Wear 303 (2013) 276285

Contents lists available at SciVerse ScienceDirect

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

Reciprocating friction and wear behavior of reaction-formed SiC


ceramic against bearing steel ball
Hongyan Xia a,b, Guanjun Qiao a, Shanlin Zhou a, Jiping Wang a,n
a
b

State Key Laboratory for Mechanical Behavior of Materials, Xian Jiaotong University, Xian 710049, China
State Key Laboratory for Mechanical Structure Strength and Vibration, Xian Jiaotong University, Xian 710049, China

art ic l e i nf o

a b s t r a c t

Article history:
Received 25 October 2012
Received in revised form
8 March 2013
Accepted 14 March 2013
Available online 2 April 2013

The friction and wear behavior of mesocarbon microbeads-based reaction-formed SiC ceramic was
investigated. The SiC, C and Si phase contents varied in the SiC ceramics. The results indicated that when
the hard SiC phase content increased, the local contact area increased and the contact pressure
decreased, resulted in lower friction coefcient and wear rate. SiC ceramics mainly had abrasive,
adhesive and plowing wear during dry sliding. There was Fe-rich not graphite-rich MML formation on
the worn surfaces, which contributed to the reduction of the contact pressure and friction force.
Lubricant effect of C phase and adhesive wear generated by Si phase were not clearly seen. Besides, the
friction and wear behavior were also related to mechanical properties, thermal conductivity, grain size
and humidity-driven tribo-reactions.
& 2013 Elsevier B.V. All rights reserved.

Keywords:
Sliding wear
Engineering ceramics
Bearings
Wear testing

1. Introduction
SiC ceramic has a variety of benecial properties such as high
hardness and strength, excellent thermal stability corrosion resistance, high thermal conductivity and low thermal expansion
coefcient, and has been widely used for tribological applications
like bearings, cylinder liners and mechanical seals. Many investigations have assessed the effects of operating parameters (load,
sliding velocity, humidity, temperature, etc.) and/or material
microstructure and properties (grain size, volume fraction, distribution, hardness, fracture toughness, etc.) on the friction and
wear properties [16]. Based on the results, many different wear
and friction mechanisms were presented. It is generally considered
that the formation of Si(OH)4 or SiOx-rich lm was an important
factor for lowering the friction coefcient of SiC ceramic, resulted
from tribochemical reactions, which was found to be sensitive to
microstructural features (grain size, chemistry, phases, etc.), atmospheric conditions (humidity, temperature) and local contact
stresses. Dong et al. [7] pointed out that at room temperature,
under high loads and high environmental humidity, the tribological behavior was controlled by tribochemical reactions between
the SiC surface and water vapor in the environment. Guha and
Basu [8] considered that tribochemical layer had a limited contribution to the wear in liquid nitrogen. Besides, it is noticed that
mechanically mixed layer (MML) or lm, mainly decided by the

Corresponding author. Tel.: +86 29 82667942; fax: +86 29 82663453.


E-mail addresses: jpwang@xjtu.edu.cn, jipingwang@gmail.com (J. Wang).

0043-1648/$ - see front matter & 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.wear.2013.03.038

microstructural features of the original base materials, could have


a strong inuence on the friction performance and wear resistance
[911]. It is reported a graphite-rich MML facilitated the low
friction and wear whereas the graphite-short MML resulted in
high friction and wear. The detachment and removal of the MML
aggravated the wear. Furthermore, according to Hertzian contactbased models of brittle materials, the wear volume for ball-on-at
contact had a great relationship with the properties (elastic
modulus, hardness and fracture toughness, etc) of samples.
Reaction-forming is an attractive method for fabricating SiC
ceramic due to its low cost and near-net shape fabrication
capability. Compared to liquid phase sintering and hot-pressed
SiC ceramics, reaction-formed SiC ceramic has much more contents of C and Si phases. As said above, many researchers pointed
out C particle in hard ceramic can display self-lubrication effect on
the reduction in friction coefcient and wear rate, yet some
researchers presented that the inuence of carbon on friction
and wear behavior was not observed or no evidence for slidinginduced formation of a benecial graphite lm was obtained
[12,13]. For Si phase, it was generally considered that it can be
softened by sliding, generating adhesive wear [14].
The microstructural characteristics and mechanical properties
of reaction-formed SiC ceramic by liquid inltrating molten Si into
mesocarbon microbeads (MCMBs)-based preform have been
investigated in our previous paper [15]. Less attention is devoted
in the literature to the tribological properties of reaction-formed
SiC while, as known, there is a great demand for structural
materials with low cost exhibiting improved resistance of friction
and wear. The aim of the present work was to investigate the

H. Xia et al. / Wear 303 (2013) 276285

277

reciprocating friction and wear behavior of MCMBs-derived reaction-formed SiC submitted to dry sliding tests against bearing steel
ball under different loads, in order to determine the effects that
the microstructural features and properties can have on friction
and wear behavior, as well as to clarify the wear mechanisms.

linear speed of 0.083 m/s. The sliding time for each test was
normally 20 min. Friction coefcient was measured by an analogto-digital converter and recorded in a computer. The wear rate was
calculated by

2. Materials and experimental methods

where b, h, and l were the width, depth and length of the wear scar
respectively, in which b and h were the mean value of ve different
sites on the wear scar measured by a non-contact 3D surface
proler (VK-9700, keyence, Osaka, Japan); l is the stroke amplitude, 5 mm; L is the total sliding distance.
For each sample, a minimum of two trials were conducted to
ensure repeatability of test data. Because of a sharp deection of
friction coefcient during the running-in stage, the change tendency and values of friction coefcient given in this paper were
that after 20 min running-in stage. After the wear test, the worn
surface, longitudinal section, wear debris of samples and worn
surface of counterfaces were characterized using a scanning
electron microscope (SEM, VEGAII XMU, Tescan, Czech) coupled
with an energy dispersive spectroscopy detector (EDX, INCA,
Oxford Instrument).

2.1. Materials
Reaction-formed SiC ceramics (10RFSC, 20RFSC, 30RFSC and
40RFSC) were fabricated by inltrating molten Si into the carbon
preform derived from a mixture of MCMBs and SiC powders,
whose content ranged from10 to 40 wt%. Here SiC powder limited
the volume shrinkage of MCMBs during sintering and the porosity
of the preform can be adjusted by altering SiC powder content.
With increasing SiC powder content, the morphology of reactionformed SiC changed where the SiC and Si phase contents increased
and C phase content decreased (the estimated values were in
Table 1, the OM images can be seen in [15], SEM images after
etching can be seen in Fig. 1). Meanwhile the reaction-formed SiC
ceramics have wide range of mechanical properties, as seen in
Table 1.

bhl
L

3. Results

2.2. Friction and wear test

3.1. Friction coefcient and wear rate

Dry sliding wear tests were conducted on a reciprocating ballon-at contact tribometer (LFT-1, Zhongke Kaihua, Lanzhou,
China), which moved the specimen stage back and forth under a
xed ball. The ball, bearing steel GCr15 with a diameter of 5 mm,
was positioned on top of horizontal at specimen and loaded with
normal force FN. SiC specimen was xed on the base of the
tribometer that can be actuated to vibrations with a constant
stroke and a frequency of . Prior to wear test, all the contact
surfaces were polished, cleaned in acetone in an ultrasonic cleaner
and nally dried by blowing warm air. All the tests were carried
out at room temperature with a relative humidity of 60%. The
normal loads FN were 20 N, 40 N and 60 N and a constant stroke
was 5 mm at a frequency of 500 rpm, which was equal to the

The representative friction traces of the SiC ceramic worn at


loads of 20 N, 40 N and 60 N are shown in Fig. 2. It can be seen that
after running-in test with 20 min, the friction coefcient of the
samples presented relatively stable behavior at the initial stage
and gradually became nearly constant for all the samples at
different loads. For 10RFSC ceramic, the difference among the
friction coefcients at different loads were very small, indicating
that the loads between 20 N and 60 N played small effect on the
friction coefcient. For 20RFSC, 30RFSC and 40RFSC, the friction
coefcient decreased with increasing load from 20 to 60 N. In the
last 10 min of stable stage, the friction coefcients of 10RFSC,
20RFSC, 30RFSC and 40RFSC basically ranged from 0.40.45,
0.40.6, 0.320.55 and 0.280.55, respectively. It is worth noting

Table 1
Microstructural parameters and properties of SiC ceramic.
Sample

10RFSC
20RFSC
30RFSC
40RFSC
a
b

Density, g cm3

2.59
2.63
2.69
2.81

Phase contenta, vol%


Si

SiC

0
3
6
15

49
60
65
73

51
37
29
12

Open porosity, %

Bending strength, MPa

Fracture toughness MPa m1/2

Hardnessb, GPa

Young's modulusb, GPa

4.18
5.28
5.14
3.90

247
311
330
359

2.6
4.4
3.8
3.4

18
20
22
30

247
239
288
319

Estimated by OM in [15].
Obtained by nanoindentation test using a TI950 Tribo indenter under maximum load 4000 N, other values were obtained from [15].

Fig. 1. SEM morphologies of polished cross-sections of (a) 20RFSC, (b) 40RFSC and (c) high magnication images of SiC phase after etching.

278

H. Xia et al. / Wear 303 (2013) 276285

Fig. 2. Representative friction coefcient curves versus time of the SiC ceramics at different loads: (a) 10RFSC, (b) 20RFSC, (c) 30RFSC, and (d) 40RFSC.

Fig. 3. Wear rates of (a) the four SiC ceramics at 40 N and (b) 20RFSC and 40RFSC at different loads.

that 40RFSC loaded at 60 N had very small friction coefcient and


the mean value was about 0.3. This value is much lower than those
of refractory ceramics under dry conditions [12,16,17], e.g. multiphase SiC based ceramic (40.36, Al2O3 ball, ball-on-at method)
[12], Al/SiC/graphite hybrid composites (40.8, monolithic A356 Al

alloy as counterface, vane-on-disk method) [16], and 70W2B5/C


composites (0.8, GCr15 bearing steel, block-on-ring method) [17].
Fig. 3(a) shows the wear rates of the four specimens tested at
constant load of 40 N. It can be seen that the wear rates of
10RFSC, 20RFSC, 30RFSC and 40RFSC were 0.20, 0.19, 0.15 and

H. Xia et al. / Wear 303 (2013) 276285

0.089  103 mm3/m respectively. The change tendency was similar to ceramic-based W2B5C and Si3N4C composites, which had
increasing tendency of wear rate when C content increased [17,10].
Fig. 3(b) shows the wear rates of 20RFSC and 40RFSC at different
loads. Totally, the wear rates increased with increasing applied load
for the samples. Meanwhile, it is clearly noted that the wear rate of
20RFSC was much sensitive to load than 40RFSC. When applied
load increased from 20 to 60 N, the wear rates of 20RFSC sharply
increased from 0.13 to 0.30  103 mm3/m. For 40RFSC ceramic, the
value only increased from 0.076 to 0.096  103 mm3/m. On the
basis of severity indexes and morphological observations, specic
wear rate k1  1015 m2 N1 or k3  1015 m2 N1 have been
reported as acceptable limits between mild and severe wear
regimes for ceramic materials [18], so the SiC ceramic in our work

279

(the lowest specic wear rate k was 3.8  1015 m2 N1) underwent
serious wear process during sliding.
3.2. Worn surfaces
After wear test, the samples had shiny worn surfaces except
10RFSC, seen by naked eyes. With increasing applied load, worn
tracks became wider, but limited in 11.5 mm. Further investigation on worn surface morphologies of the samples are conducted
by SEM and shown in Fig. 4. For 10RFSC, large amount of debris
covered on the surface, coupled with indentation aw, which
would result in abrasive adhesive wears. For 20RFSC, some narrow
and deep furrows can be seen on the surface. According to the
high magnication SEM micrographs, the soft phase of carbon was

Fig. 4. Micrographs of wear scars of SiC ceramics loaded at 20 N: (a) and (b) 10RFSC, (c) and (d) 20RFSC, (e) and (f) 30RFSC, (g) and (h) 40RFSC.

280

H. Xia et al. / Wear 303 (2013) 276285

peeled off from the SiC matrix and the connecting neck of network
SiC matrix was destroyed, where large amount of debris scattered.
These morphologies indicated that serious micro-cutting happened on the contact surfaces under dry sliding procedure, which
could result in high friction coefcient and wear rate. Besides, the
light gray regions (denoted as A) and dark gray regions (denoted
as B) were investigated by EDX and the results revealed that light
gray regions mainly included C, O, Si, Cr and Fe elements, which
had contents of 8.69, 7.79, 2.34, 1.17 and 80.00 at%, respectively,
and dark gray regions mainly included C and Si elements, which
had contents of 37.83 and 62.17 at%, respectively (Fig. 5). So the
dark gray regions should be the SiC matrix. The light gray regions
should be the MML, which were the debris mainly peeled from
counterpart steel ball and subsequently transferred onto the
sample, underwent complex formation process of fragmentation,
mixing, accumulation, oxidation, and compaction, etc.
Compared to 20RFSC, the worn surface of 30RFSC became
neater and smoother, accompanying with some narrow and

shallow furrows. As seen from high magnication SEM micrograph, network SiC was remained relatively intact, having no great
damage. Besides, it is worthy noted that there had network light
gray regions of MML covered on the network SiC matrix. For
40RFSC, seldom furrows and debris were seen on the worn
surface. However, some cavities left by carbon grains pulled out
from the worn surface were seen on the wear scar. The elemental
distribution map of worn surface of 40RFSC was tested. The results
showed that the content of Fe element was very small; the main
elements were Si and C. The contents of Fe, Si and C were 0.5, 47.3
and 52.2 at%, indicating that the load of 20 N hardly had effect on
serious damage for 40RFSC. Furthermore, the worn surface morphology of 40RFSC loaded at 60 N was conducted by SEM, as
shown in Fig. 6. Besides a large amount of scattered debris,
cavities, and high density of furrows, some thin lms, which
mainly came from counterpart steel ball, were covered on the
worn surface. As seen from the wear scar morphologies, the four
SiC ceramics mainly had abrasive, adhesive and plowing wear

Fig. 5. EDX results of (a) light gray regions (denoted as A) and (b) dark gray regions (denoted as B).

H. Xia et al. / Wear 303 (2013) 276285

281

Fig. 6. Micrographs of wear scar of 40RFSC loaded at 60 N.

Fig. 7. SEM of the longitudinal sections below the worn surfaces of (a) 20RFSC and (b) 40RFSC.

during dry sliding. For 40RFSC, the wear mainly were plowing and
polishing under 20 N and abrasive, adhesive and plowing at 60 N,
this was in agreement with the report to a certain extent [7].
Observation on the longitudinal sections of the two specimens
(20RFSC and 40RFSC) displayed the microstructure features of the
subsurface below the worn surface, as shown in Fig. 7. It can be
seen that partial regions had MML, having a thickness of about
75 m and consisting of ne particles, located on the top of
substrates of 20RFSC. Meanwhile some microcracks in the MML
and downward-sloping microcracks below the MML can be seen
clearly. As known, ceramic materials are generally subjected not
only to brittle fracture under high or impact loads, but also to
fatigue cracking induced mainly at grain boundaries by the high
surface temperatures and cyclic stresses generally attained under
dry sliding conditions. Under the reciprocating sliding process, the
grains beneath the MML were fractured, thereby resulting in the
formation of microcracks beneath the MML. When the microcracks propagated, connected and deected to the worn surface,
the plate-like MML was peeled out from worn surface and serious
wear happened. Under high load, the fracture and removal of the
MML increased; thereby the wear rate increased and was sensitive
to applied load. The above results were in good agreement with
the results of wear rate values.
Different from 20RFSC, the MML on the top of substrate of
40RFSC was very thin, about 1 m, and no cracks appeared. The
line EDX result revealed that the MML of 20RFSC and 40RFSC both
mainly consisted of Fe element, indicating that wear of steel
bearing ball plowed by SiC ceramic at was the main wear
evidence. Meanwhile for 40RFSC the MML with thin thickness
and no cracks formation indicated that the cyclic contact pressure
should be lower than those of 20RFSC.
3.3. Wear debris
Four different types of morphologies were observed in the
debris of samples: mixed particles with large and small size for
10RFSC, plate-like akes coupled with ne accumulated particles

for 20RFSC, ne accumulated particles for 30RFSC, and relatively


large amorphous particles for 40RFSC (Fig. 8). The EDX result
noted the plate-like debris of 20RFSC was abundant in C, O, Si, Cr
and Fe, which had contents of 8.05, 6.82, 12.52, 1.28 and 71.33 at%
respectively (Fig. 9). This result strongly suggested that the platelike debris was in fact detached from the MML, by delamination as
said above. Moreover, the EDX result of the debris of 40RFSC
illustrated that Fe and O element contents (22.07 at% and 2.83 at%,
respectively) was lower than 20RFSC, while C and Si element
contents (26.26 at% and 48.55 at%, respectively) were higher than
20RFSC. Here C element came from C and SiC particles, and Si
element came from Si and SiC particles. It was considered that
ash temperature of the contact point under dry sliding reached
1000 1C for SiC ceramic, making the worn debris oxidation [8].
Wsche and Klaffke reported [12] that oxidation would make the C
element from carbon and SiC particles disappear. According to the
EDX results, it can be seen that serious oxidation occurred for
20RFSC, further indicating that the surface temperatures and
friction force on the 20RFSC was higher than 40RFSC.
3.4. Worn surfaces of steel ball
SEM photos of counterpart steel balls loaded at 60 N are shown
in Fig. 10. Considerable plowing effect was observed on the wear
scars of four samples. On the worn surface of steel ball against
10RFSC, some shallow furrows together with large amount of
debris were seen. For steel ball against 30RFSC, almost no particles
can be seen and wide and smooth grooves were found on the
worn surface. Compared to the steel balls against 10RFSC and
30RFSC, more serious friction and wear happened for steel balls
during sliding against 20RFSC and 40RFSC. For steel ball against
20RFSC, there were large amount of deep furrows together with
cavities and cracks seen on the worn surface. Meanwhile some
white particles, mainly consisted of SiC particles, scattered in the
cavities, which should be attributed to the plowing by the convex
on the surface of 20RFSC. For the counterpart steel ball corresponding to the 40RFSC, the width of furrows was wide and

282

H. Xia et al. / Wear 303 (2013) 276285

Fig. 8. Typical SEM morphologies of the wear debris of (a) 10RFSC, (b) 20RFSC, (c) 30RFSC and (d) 40RFSC.

unsmooth, some white SiC particles scattered on the surface.


Based on the EDX results, Si and C element contents were very
small on the contacts of the four balls; the main elements came
from steel ball itself. This phenomenon indicated that the hard SiC
ceramics, primarily through plowing and abrasive wear, led to
severe wears of the steel balls. Lubricant effect of C phase and
adhesive wear generated by Si phase were not clearly seen.

4. Discussion
4.1. MML formation based on microstructural observations
On the basis of the careful examination and analysis of the
worn surface, longitudinal section of the worn surface and wear
debris, it can be seen that initial surface morphology played an
important role on the friction behavior and wear resistance. In the
initial stage of sliding wear, supercial carbon grains were preferentially peeled by the asperities of the counterface and released
on the worn surface as debris due to its softness and low strength.
However, because of high hardness, the SiC particles cannot be
peeled from the surface at that moment, thus to the direct
mechanical interactions between the convex of hard SiC and
counterpart steel ball, resulting in the friction force increase.
Subsequently, under high friction force, the counterpart steel ball
was plowed and transferred onto the surface of samples. As sliding
wear continued, microcracks would be produced in the samples
and gradually propagated. Meanwhile a drop of SiC debris was
generated due to its fracture under repeated shearing stress. The
debris of C, Si, Fe and SiC were subjected to a complicated process
of fragmentation, mixing, oxidation and compacting under normal
load and frictional force to form MML.
For 10RFSC, the contents of C and SiC grains were the largest
and smallest respectively among the four samples, meanwhile the
homogeneity of phase distribution was the worst [15], which
remarkably increased the local friction and micro-cutting effect,
playing very bad effect on the formation of stable MML. Khurshudov et al. reported [19] that grain detachment from the
ceramic slider surface was the most important mechanism

controlling the life of MML. The SiC grains with relative low
content in 10RFSC were much easier to be detached out from
the body and MML was hardly built up. The detached particles
scattered as debris and the main wear was abrasive wear, resulted
in the largest wear rate. But due to the large C content, the friction
coefcient was relatively low. As the content of C grains decreased
and the homogeneity of phase distribution increased, although
some Fe-rich MML was seen covered on the surface for 20RFSC,
the main contact was still the contact between the convex of hard
SiC and the counterpart steel ball. The friction force was high and
the connecting neck of the network SiC was destroyed. Correspondingly the roughness considerably increased, resulted in high
friction coefcient and wear rate. For 30RFSC, network Fe-rich
MML was covered on the SiC grains. As for 40RFSC, the network
microstructure basically vanished, SiC was the main phase where
small contents of C and Si separately and discontinuously dispersed. At this moment the contact area increased and the contact
pressure (the load divided by the contact area) decreased, the
plowing wear was the main wear mechanism. After a few minutes
of sliding, Fe-rich thin lm formed. As seen above, there were no
graphite-rich lms or MML formation on the surfaces of the four
samples. It was considered that Fe-rich lms or MML formation on
the contact surface contributed to the increase of local contact
area, further reducing the contact pressure and friction force. As a
result, the effect of local high friction force on the friction
coefcient and wear rate became smaller when increasing SiC
powder in raw materials in SiC ceramics.

4.2. Mechanical properties


Besides of the formation of MML on contact surface, the
mechanical properties, such as hardness and fracture toughness,
are also critical in controlling the wear resistance.
Because of small diameter of 5 mm of steel ball, it is expected
that contact pressure was high and here was severe deformation.
As seen in Fig. 7(a), there were lateral cracks due to the residual
stresses caused by indentation. Thus according to Evans and
Marshall's report (assuming sharp indenter), the wear volume

H. Xia et al. / Wear 303 (2013) 276285

283

Fig. 9. EDX results of the wear debris of (a) 20RFSC and (b) 40RFSC.

can be given as [20]


 4=5
P 9=8
E
l1
V 3 1=2 5=8
H
KC H

where P was the load; KC, E and H were the fracture toughness,
elastic modulus and hardness, respectively, of the material; l1 was
the stroke length; and 3 was a material-independent constant. So
the difference in wear volume among four samples can be
explained in consideration of the differences in fracture toughness,
elastic modulus, hardness, and load. Based on the mechanical
properties of SiC ceramics (Table 1), the wear volumes were
calculated and the results are shown in Fig. 11. It is clearly seen
that the change tendency in wear volumes of the theoretical
expectation was basically the same as the experimental wear rate
(the wear volume divided by the total sliding distance, which was
constant for the four samples) in Fig. 3. The wear volume increased
linearly versus applied load and the increase was less intensive to
load with increasing SiC powder in raw materials. So the

mechanical properties of SiC ceramics in the present work played


very important roles in the differences in wear resistance.
Meanwhile the sliding wear behavior of ceramics can be
classied as severe wear with elastic deformation and material
fracture under high contact pressures, and mild wear with plastic
deformation under low contact pressures. Kong and Ashby [21]
proposed an expression for determining the critical load for
initiation and propagation of a pre-existing crack under a spherical
indenter using the Hertzian contact theory, the expression of
critical load (PC) is:
PC

4:33

KC3

R2

1 103 C 0 3=2 E2

where is the coefcient of the friction of the tribosystem, C0 is


the initial crack length and R is the spherical indenter radius. It can
be seen that when spherical indenter radius and initial crack
length were constant, friction coefcient, fracture toughness and
elastic modulus were the key factors to the transition from mild

284

H. Xia et al. / Wear 303 (2013) 276285

Fig. 10. Micrographs of wear scars of steel balls loaded at 60 N against: (a) 10RFSC, (b) 20RFSC, (c) 30RFSC and (d) 40RFSC.

coefcient. When it was stable friction, the maximum value of


the temperature of frictional contact point was [19]
T max

PV
4a1 2

where 1 and 2 was the thermal conductivity of friction material


and counterpart material, a was a dimension-dependent constant,
and V was friction velocity. It can be seen that the high thermal
conductivity of friction material can lower the temperature of
frictional contact point, further reducing the friction coefcient.
The thermal conductivity of SiC ceramics at RT were 35.5, 43.7,
81.5, 103.2 W m1 K1 for 10RFSC, 20RFSC, 30RFSC and 40RFSC
respectively; so the 40RFSC should have the lowest frictional
contact temperature, also the lowest friction coefcient. To some
extent, this was in agreement with the EDX results and change
tendency of measured friction coefcient.
Fig. 11. Calculated wear volume versus applied load of SiC ceramics.

4.4. Grain size


wear to severe wear. Here assuming that R was 2.5 mm (the radius
of steel ball was 2.5 mm), C0 was equal to the size of soft C grains,
10 m, the coefcient of the friction was 0.42, 0.51, 0.47, and 0.42
(the mean values loaded at 40 N) for 10RFSC, 20RFSC, 30RFSC and
40RFSC respectively. According to the values of KC, E in Table 1, PC
was calculated and the values were 0.18, 0.57, 0.31 and 0.24 N for
10RFSC, 20RFSC, 30RFSC and 40RFSC respectively (0.52 N for
Ti3SiC2, 0.54 N for WCZrO2, 0.72 N for TiCN20Ni10WC in
[21]). So the SiC ceramics had very low resistance for the transition
to severe wear. In our work, the applied loads FN were 20 N, 40 N
and 60 N, which were much higher than the critical loads of the
four samples, resulting in the severe wear. This correlated well
with our experimental results, as seen from the wear rates and
morphologies of worn surfaces.
4.3. Thermal conductivity
Based on the Archard model, the wear resistance of materials
had inverse relation with contact ash temperature. The lower the
temperature of frictional contact point, the lower the friction

In general, ceramics with ne grain size can obtain relative


stable smooth friction surface, having a low friction coefcient or
wear [22,23]. However, it was reported that the ceramic with the
matching size of the SiC particles had good friction and wear.
EI-Raghy et al. [24] and Sang et al. [25] considered that the coarse
SiC particles made the formation of the cracks difcult and the ne
particles made the propagation difcult. Meanwhile, the coarse
particles had crack-bridging ability. The crack bridging would result
in the increasing of paths or the closing of cracks in the propagation
processes. In our work, as seen from Fig. 1(b), the SiC phase
consisted of large SiC particles, but from high magnication images
of SiC phase after etching (Fig. 1(c)), the large SiC particles further
consisted of very small (the size was lower than 1 m) SiC particles.
For 40RFSC mainly consisted of SiC phase, there is a possibility that
this microstructure contributed to the improvement of resistance to
the formation and propagation of the cracks, resulted in the
reduction of wear rate. But for other samples, many weak interfaces
and microcracks between SiC and C particles and inside of soft C
particles made the formation and propagation of the cracks easier.

H. Xia et al. / Wear 303 (2013) 276285

4.5. Humidity-driven tribo-reactions


Because the relative humidity of 60% in present work was
relatively high, the humidity-driven tribo-reactions should be
considered. Since the materials were composed of SiC, free silicon
and carbon, partial SiC and Si formed soft layer of hydroxilated
silicon oxide [4,5], partial C vanished as CO, CO2. According to the
EDX results in Figs. 5 and 9, it was deduced that soft layers of
hydroxilated silicon oxide was formed on the wear surfaces of the
four samples. When the contents of soft C and Si phases decreased
for 30RFSC and 40RFSC, the stability of soft layer increased. This
should be one of reasons of friction behavior of reaction-formed
SiC ceramics.
5. Conclusions
The following conclusions can be drawn from this study:
1) Reaction-formed SiC ceramics were fabricated by inltrating
molten Si into the carbon preform derived from a mixture of
MCMBs and SiC powders, whose content ranged from 10 to
40 wt% (10RFSC, 20RFSC, 30RFSC and 40RFSC). In the SiC
ceramics, the SiC and Si phase contents increased and C phase
content decreased on increasing the SiC powder in raw
materials.
2) The friction coefcients of 10RFSC, 20RFSC, 30RFSC and 40RFSC
basically ranged from 0.40.45, 0.40.6, 0.320.55 and 0.28
0.55, respectively, when loaded from 20 to 60 N. These values
were much lower than those of refractory ceramics under dry
conditions. SiC ceramic underwent serious wear process during
sliding. Wear rate increased when C content and applied load
increased. Meanwhile the wear rate of 20RFSC was much
sensitive to load than 40RFSC. For 20RFSC and 40RFSC ceramics, the wear rates increased from 0.13 to 0.30  103 mm3/m
and 0.076 to 0.096  103 mm3/m respectively when loaded
from 20 to 60 N.
3) Based on the observations of the worn surface, longitudinal
section, wear debris of samples and worn surface of counterfaces, the four SiC ceramics mainly had abrasive, adhesive and
plowing wear during dry sliding. When the hard SiC phase
content increased, the local contact area increased and the
contact pressure decreased, resulted in lower friction coefcient and wear rate. Moreover, it is noted that there were
Fe-rich not graphite-rich MML formation on the worn surfaces,
which contributed to the increase of local contact area, further
reducing the contact pressure and friction force. Lubricant
effect of C phase and adhesive wear generated by Si phase
were not clearly seen. Besides, the friction and wear behavior
was also related to mechanical properties, thermal conductivity, grain size and humidity-driven tribo-reactions.

Acknowledgments
The authors are grateful to the National Natural Science Foundation of China (51202181), the China Postdoctoral Science Foundation
(2012M511997), the Program for New Century Excellent Talents in

285

University (NCET-11-0425) and the Natural Science Foundation of


Shaanxi province (2012JM6004) for nancial supports.

References
[1] M.-Y. Chn, B. Bhushan, L.C. DeJonghe, Wear behavior of ceramic sliders in
sliding contact with rigid magnetic thin-lm disks, Tribology Transactions 32
(1992) 603610.
[2] K.-H. Zum Gahr, R. Blattner, D.-H. Hwang, K. Phlmann, Micro- and macrotribological properties of SiC ceramics in sliding contact, Wear 250 (2001)
299310.
[3] E. Ciudad, O. Borrero-Lpez, F. Rodrguez-Rojas, A.L. Ortiz, F. Guiberteau, Effect
of intergranular phase chemistry on the sliding-wear resistance of pressureless liquid-phase-sintered -SiC, Journal of the European Ceramic Society 32
(2012) 511516.
[4] G. Amirthan, M. Balasubramanian, Reciprocating sliding wear studies on Si/SiC
ceramic composites, Wear 271 (2011) 10391049.
[5] R. Dhiman, K. Rana, E. Bengu, P. Morgen, Synthesis, characterization, and wear
and friction properties of variably structured SiC/Si elements made from wood
by molten Si impregnation, Journal of the European Ceramic Society 32 (2012)
11051116.
[6] V.S.R. Murthy, H. Kobayashi, N. Tamari, S. Tsurekawa, T. Watanabe, K. Kato,
Effect of doping elements on the friction and wear properties of SiC in
unlubricated sliding condition, Wear 257 (2004) 8996.
[7] X. Dong, S. Jahanmir, L.K. Ives, Wear transition diagram for silicon carbide,
Tribology International 28 (1995) 559572.
[8] T.K. Guha, B. Basu, Microfracture and limited tribochemical wear of silicon
carbide during high-speed sliding in cryogenic environment, Journal of the
American Ceramic Society 93 (2010) 17641773.
[9] Y.B. Liu, S.C. Lim, S. Ray, P.K. Rohatgi, Friction and wear of aluminiumgraphite
composites: the smearing process of graphite during sliding, Wear 159 (1992)
201205.
[10] A. Gangopadhyay, S. Jahanmir, Friction and wear characteristics of silicon
nitridegraphite and aluminagraphite composites, Tribology Transactions 34
(1991) 257265.
[11] H. Liu, Q. Xue, The tribological properties of TZPgraphite self-lubricating
ceramics, Wear 198 (1996) 143149.
[12] R. Wsche, D. Klaffke, Wear of multiphase SiC based ceramic composites
containing free carbon, Wear 249 (2001) 220228.
[13] P.J. Blau, B. Dumont, D.N. Braski, T. Jenkins, E.S. Zanoria, M.C. Long, Reciprocating friction and wear behavior of a ceramic-matrix graphite composite for
possible use in diesel engine valve guides, Wear 225229 (1999) 13381349.
[14] S. Fan, L. Zhang, L. Cheng, J. Zhang, S. Yang, H. Liu, Wear mechanisms of the
C/SiC brake materials, Tribology International 44 (2011) 2528.
[15] H. Xia, J. Wang, H. Jin, Z. Shi, G. Qiao, Fabrication and properties of reactionformed SiC by inltrating molten Si into mesocarbon microbeads-based
carbon perform, Materials Science and Engineering A 528 (2010) 283287.
[16] M.L. Ted Guo, C.Y.A Tsao, Tribological behavior of self-lubricating aluminium/
SiC/graphite hybrid composites synthesized by the semi-solid powder-densication method, Composites Science and Technology 60 (2000) 6574.
[17] Y. Lv, G. Wen, T.Q. Lei, Tribological behavior of W2B5 particulate reinforced
carbon matrix composites, Materials Letters 60 (2006) 541545.
[18] L. Micelea, G. Palombarini, S. Guicciardi, L. Silvestroni, Tribological behavior
and wear resistance of a SiCMoSi2 composite dry sliding against Al2O3, Wear
269 (2010) 368375.
[19] A.G. Khurshudov, M. Olsson, K. Kato, Tribology of unlubricated sliding contact
of ceramic materials and amorphous carbon, Wear 205 (1997) 101111.
[20] A. Tewari, B. Basu, R.K. Bordia, Model for fretting wear of brittle ceramics, Acta
Materialia 57 (2009) 20802087.
[21] H. Kong, M.F. Ashby, Wear mechanisms in brittle solids, Acta Metallurgica et
Materialia 40 (1992) 29072920.
[22] K.H. Zum Gahr, W. Bundschuh, B. Zimmerlin, Effect of grain size on friction
and sliding wear of oxide ceramics, Wear 162164 (1993) 269279.
[23] R. Singha Roy, H. Guchhait, A. Chanda, D. Basu, M.K. Mitra, Improved sliding
wear-resistance of alumina with sub-micron grain size: a comparison with
coarser grained material, Journal of the European Ceramic Society 27 (2007)
47374743.
[24] T. EI-Raghy, P. Blau, M.W. Barsoum, Effect of grain size on friction and wear
behavior of Ti3SiC2, Wear 238 (2000) 125130.
[25] K. Sang, L. Liu, Z. Jin, Improvements on dry friction and wear properties for
reaction-sintered silicon carbide by the matching size of SiC particles,
Materials and Design 28 (2007) 735738.

Anda mungkin juga menyukai