Wear Mechanisms of Tib and Tib - Tisi at Fretting Contacts With Steel and Wc-6 WT% Co

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Int. J. Appl. Ceram. Technol.

, 7 [1] 89–103 (2010)


DOI:10.1111/j.1744-7402.2008.02329.x

Ceramic Product Development and Commercialization

Wear Mechanisms of TiB2 and TiB2–TiSi2 at Fretting


Contacts with Steel and WC–6 wt% Co
Golla Brahma Raju and Bikramjit Basu*
Laboratory for Advanced Ceramics, Department of Materials and Metallurgical Engineering, Indian
Institute of Technology, Kanpur, India

Unlubricated fretting wear tests on TiB2 and TiB2–5 wt% TiSi2 ceramics against two different mating materials (bearing
grade steel and WC–6 wt% Co balls) were performed with a view to understand the counterbody-dependent difference in
friction and wear properties. The fretting experiments were conducted systematically by varying load (2–10 N) at an oscillating
frequency of 4 Hz and 100 mm linear stroke, for a duration of 100,000 cycles. Adhesion, abrasion, and three-body wear have
been observed as mechanisms of material damage for both the TiB2/steel and TiB2/WC–Co tribosystems. The third body is
predominantly characterized as tribochemical layer for TiB2/steel and loose wear debris particles for TiB2/WC–Co tribocou-
ple. An explanation on differences in tribological properties has been provided in reference to the counterbody material as well
as microstructure and mechanical properties of flat materials.

Introduction The other damaging aspect of the fretting is fretting


fatigue. These forms of damage can arise in any assem-
Fretting wear refers to the phenomena of small am- bly of engineering components/machinery, if a source
plitude (r300 mm) relative reciprocatory displacement of vibration is present. For example, fretting occurs in
sliding motion. Basically, fretting is a form of adhesive all quasi-static-loaded assemblies, such as keys, bearing
or abrasive wear, where the normal load causes adhe- races, shafts, orthopedic implants, turbine blade roots,
sion between asperities and oscillatory movement causes electrical contacts, nuclear reactor components, and
rupture, resulting in the formation of wear debris.1,2 power plant machinery, etc. Among the various modes
of fretting contacts, mode I fretting wear, that is, under
*bikram@iitk.ac.in tangential displacement conditions has been broadly
r 2008 The American Ceramic Society studied for various engineering materials.1–4
90 International Journal of Applied Ceramic Technology—Brahma Raju and Basu Vol. 7, No. 1, 2010

The use of ceramics for many tribological applica- materials such as steel and WC-based cermets. It also
tions increased considerably because of their unique needs to be mentioned here that till now there are no
combination of properties such as low density, low ther- studies on the tribological behavior of TiB2/cemented
mal expansion, high hardness, and corrosion resistance. carbide tribocouple. In the present study, an attempt has
Titanium diboride (TiB2) belongs to this group of been made to investigate the load (P: 2, 5, 10 N) de-
materials and is well known for its high hardness, high pendent wear mechanisms for hot-pressed TiB2 and
melting point, good electrical and thermal conduc- TiB2–5 wt% TiSi2 ceramics.
tivities.4 The potential applications of TiB2 materials
include cutting tools, impact resistant armor, wear re- Experimental Procedure
sistant parts, and electrodes for electrodischarge ma-
chining. However, densifying TiB2 at lower sintering Materials
temperatures without compromising much on its un-
ique combination of properties remains as a challenge.5 In a recent study,33 we reported that an optimal
Considerable research efforts have been directed toward combination of density, hardness, and toughness can be
obtaining dense TiB2-based materials, possessing better obtained in TiB2–5 wt% TiSi2 and accordingly, the
properties, via various sintering techniques and incor- same ceramic composition is selected for the present
poration of different metallic as well as nonmetallic study. Monolithic TiB2 is chosen as a baseline material
additives.5–10 for comparison. Both TiB2 and TiB2–5 wt% TiSi2 ma-
The tribological performance of the materials de- terials were processed from commercially available
pends on various parameters (fracture toughness, hard- TiB2 (Grade F, H.C. Starck, Goslar, Germany) and
ness, elastic modulus, microstructural characteristics), TiSi2 (Goodfellow Cambridge, Huntingdon, U.K.)
operating parameters (load, test environment, test du- powders. An approximate amount of powder mixtures
ration, etc.) as well as counterbody material. Extensive were hot pressed at a temperature of 16501C for 1 h,
research has been carried out to understand the tribo- with an applied pressure of 30 MPa, in a flowing argon
logical behavior of hard materials like ceramics and cer- atmosphere. The binderless TiB2 could be densified to
mets11–21 with limited tribological work on TiB24,22–26 around 95% theoretical density (rth), while almost full
The fretting wear behavior of TiB2 containing compos- densification (B99% rth) was achieved with the TiB2–
ites under dry sliding conditions against bearing steel is 5 wt% TiSi2 via the liquid-phase sintering. Microstruc-
reported and a tribochemical wear model is formu- tural investigation of the polished surfaces was per-
lated.4 Basu et al.22 studied the mechanisms of material formed by means of a scanning electron microscope
removal of TiB2-based composites after fretting against (SEM) (Quanta FEI 200, Eindhoven, The Nether-
ball bearing steel in lubricating medium. During the lands), attached with energy-dispersive spectroscopy
fretting tests against steel in paraffin oil significant re- (EDS). A representative SEM–EDS image of etched
duction in friction (coefficient of friction [COF] around TiB2–5 wt% TiSi2 is presented in Fig. 1. The micro-
0.08–0.12) could be observed when compared with that structure reveals the presence of equiaxed TiB2 grains
in water for all the investigated materials. Wasche (gray) and intergranular Ti5Si3 (bright) phase; the cor-
et al.26 investigated the influence of relative humidity responding EDS analysis of the phases evidences the
(RH) on friction and wear behavior of pressureless sin- presence of Ti, B, and Si. The average grain size of TiB2
tered TiB2 ceramic against SiC and Al2O3 balls under is about 3.5 mm. Vickers hardness of the dense TiB2
unlubricated conditions at room temperature. TiB2 ex- samples were measured using an indent load of 50 N.
hibited low wear rates (o1  106 mm3/N m), that is, The fracture toughness evaluations were done by crack
high wear resistance. While reviewing the existing liter- length measurement of the radial crack patterns formed
ature, it was noted that the tribological properties of around Vickers indents according to the formulation
TiB2 coatings were also investigated to some extent.27–32 proposed by Anstis et al.34 Elastic modulus (E) was de-
The use of hard TiB2 coating is reported to increase the termined using dynamic elastic properties analyzer,
wear resistance by a factor of 10 than that of bulk TiB2.30 which works on the principle of resonance frequency.
The present investigation is carried out to under- The details of starting powders, microstructural charac-
stand the tribological performance of the newly devel- terization, densification mechanism, and properties of
oped TiB2 ceramics against widely used engineering the materials are described elsewhere.33
www.ceramics.org/ACT Wear Mechanisms of TiB2 and TiB2–TiSi2 91

Fig. 2. Schematic of the fretting test set-up. The testing conditions:


constants-stroke length 100 mm, oscillation frequency 4 Hz and
cycles (100,000); variables—normal load (P: 2, 5, and 10 N).
Fig. 1. Scanning electron microscope image of the polished and
chemically etched (unworn) surface of TiB2–5 wt% TiSi2. The and WC–6 wt% Co balls of 10 mm diameter). The TiB2-
energy-dispersive spectroscopy analysis indicating TiB2 (dark) based ceramics were mechanically held on a translation
phase: Legend 1 and Ti5Si3 (bright) phase: Legend 2. table, which was made to oscillate at a preset relative
linear displacement of constant stroke and frequency.
Wear Tests The flat specimens (TiB2 and TiB2–5 wt% TiSi2)
for fretting tests were cut from the hot-pressed disks and
The fretting experiments were performed using a machined into bar shapes with dimensions of 3 mm 
computer-controlled fretting machine (DUCOM 4 mm  20 mm. The samples were polished to a surface
TR281-M, Bangalore, India), which produces a linear roughness of RaB0.20 mm by standard ceramographic
relative oscillating motion with ball-on-flat configura- techniques. The surface roughness of the samples was
tion. By a stepper motor, the flat sample is made to os- measured by using laser surface profilometer. The me-
cillate with a relative linear displacement of constant chanical properties of all the TiB2-based ceramics and
stroke and frequency. An inductive displacement trans- the counterbodies are presented in Table I. It can be
ducer monitors the displacement of the flat sample and noted that TiB2–5 wt% TiSi2 exhibited superior hardness
a piezoelectric transducer is used to measure the friction (Hv5B25 GPa), indentation toughness (B6 MPa m1/2)
force. The variation in tangential force is recorded and and elastic modulus (B518 GPa), when compared with
the corresponding COF is calculated on-line with the monolithic TiB2. As per the suppliers data, the bearing
help of a computer-based data acquisition system. A fret- grade steel ball (commercial SAE 52100 grade) pos-
ting loop gives the evolution of the tangential force as a sessing a hardness of HvB7 GPa and WC–6 wt% Co
function of the displacement amplitude during each cycle. (cemented carbide) balls are characterized by a hardness
The COF is further evaluated from the average of the two of B18 GPa and toughness of B14 MPa m1/2.
plateau values of the tangential force in the fretting loop. Before the fretting tests, all the flats and balls were
The use of the same testing set-up has been made in our ultrasonically cleaned in acetone. The displacement
research on various ceramics and composite systems.13,24 stroke, normal load, frequency, and number of fretting
A schematic representation of the test configuration is cycles are the external variables associated with the fret-
presented in Fig. 2. The TiB2 flat samples were fretted ting machine. The fretting experiments were performed
against two different counterbodies (bearing grade steel at an oscillating frequency of 4 Hz and 100 mm linear
92 International Journal of Applied Ceramic Technology—Brahma Raju and Basu Vol. 7, No. 1, 2010

Table I. Mechanical Properties and Surface Roughness of the Flat TiB2 Ceramics as Well as the Bearing Grade
Steel and WC–6 wt% Co Counterbody Materials
Vickers hardness, Elastic modulus Fracture toughness Surface roughness
Material HV 5 (GPa) (GPs) (MPa m 1/2 ) (R a, lm)
TiB2 18.3 434.9 3.8 0.20
TiB2–5 wt% TiSi2 25.2 517.9 5.8 0.20
Bearing grade steel ball 7.0 210.0 — 0.02
WC–6 wt% Co ball 18.0 630.0 — 0.02
Data are supplied by the commercial supplier.

stroke, for a duration of 100,000 cycles. The influence very similar behavior. In particular, COF against steel
of varying the load (P: 2, 5, 10 N) on the friction and increases with an increase in load; while less variation in
wear response of all the materials were investigated. It COF is noticed in case of cemented carbide counter-
must be mentioned that the combination of these test- body. The increase in COF with load (up to 5 N) for
ing parameters resulted in the gross slip fretting con- both the TiB2 ceramics against steel can be attributed to
tacts. All experiments were conducted in an ambient the abrasion-induced continuous material removal from
atmosphere at room temperature (301721C) with RH steel counterbody. However, COF at 10 N load is al-
of 4575%. most comparable or even less than that of at 5 N. The
reduction in COF at higher loads of 10 N is due to the
Characterization of the Worn Surfaces establishment of the contact between the tribochemical
layers, instead of the virgin surfaces of the contacting
After the test, detailed characterization of the worn materials. It is also evident that the COF plot consists of
surfaces of the TiB2–TiSi2 flat and counterbody samples many fluctuations during fretting at lower loads up to
were done using SEM. The compositional analysis of 5 N, while it reduces considerably at 10 N load (Fig. 3a).
the worn surfaces and wear debris is carried out using Such characteristic frictional behavior is also noticed
EDS. The wear scar profiles on each sample were obtained with the TiB2–5 wt% TiSi2. However, the COF values
using a computer-controlled laser surface profilometer are relatively higher at all loads when compared with
(Mahr-Perthometer PGK 120, Göttingen, Germany). monolithic TiB2 (see Fig. 3b).
The wear volume was calculated by integrating the The wear volumes of TiB2 ceramics were estimated,
surface area of each 2D profile (extracted from different using laser surface profilometry. A typical 3D profile of
locations on 3D profile) over distance for the flat sample. worn surface of TiB2–5 wt% TiSi2 is shown in Fig. 4a.
From the measured wear volume, the specific wear rates A closer look at Fig. 4a reveals the characteristic peaks
[wear volume/(load  distance)] were calculated. at the worn region, instead of depths, which are com-
posed of various 2D profiles. Such observations impli-
Results cate the negative wear phenomena of the TiB2 samples,
when fretted against steel counterbody (see also Fig. 4b).
Friction and Wear Behavior Against Steel This essentially indicates that on 2D profiles, if taken
Counterbody at various places of worn surface, hills instead of valleys
are commonly observed/recorded for worn surfaces. In
The frictional behavior of both the monolithic contrast, the fretted surfaces on ceramics, after testing
TiB2 and TiB2 reinforced with 5 wt% TiSi2 flat sam- against WC–Co, show regular 3D or 2D profiles, as
ples were investigated during fretting against bearing typically observed for worn material. Therefore, the
steel at varying load from 2 to 10 N. During the fretting wear volumes were calculated for such cases, where
test the other variables (100,000 cycles, 100 mm ampli- 2D profiles show pronounced valley and accordingly,
tude and 4 Hz frequency) are kept constant. The COF the area/maximum depth of profile could be measured.
against number of cycles for different ceramics are plot- Furthermore, the measured wear volumes are normal-
ted in Fig. 3. In general, both the TiB2 ceramics exhibit ized with respect to load and total sliding distance
www.ceramics.org/ACT Wear Mechanisms of TiB2 and TiB2–TiSi2 93

(a) 1.0 (b) 1.0


TiB2 vs. Steel TiB2 -5 wt.% TiSi2 Vs. Steel
0.9 0.9
0.8 0.8
0.7 0.7
0.6 0.6
COF

COF
0.5 0.5
0.4 0.4
0.3 0.3
0.2 2N 0.2 2N
0.1 5N 0.1 5N
10N 10N
0.0 0.0
0 20000 40000 60000 80000 100000 0 20000 40000 60000 80000 100000
No. of cycles No. of cycles

(c) 1.0 (d) 1.0


TiB2 Vs. WC-6 wt.% Co TiB2 -5 wt.% TiSi2 Vs. WC-6 wt.% Co
0.9 0.9
0.8 0.8
0.7 0.7
0.6 0.6
COF
COF

0.5 0.5
0.4 0.4
0.3 0.3
0.2 2N 0.2 2N
0.1 5N 0.1 5N
10N 10N
0.0 0.0
0 20000 40000 60000 80000 100000 0 20000 40000 60000 80000 100000
No. of cycles No. of cycles

Fig. 3. The evolution of coefficient of friction (COF) versus number of cycles for various investigated tribocouples at different loads. Fretting
conditions: 2–10 N load, 4 Hz frequency, 100,000 cycles, and 100 mm stroke length.

(number of cycles  displacement stroke  2) in order to related wear volume and load by the following ex-
obtain normalized specific wear rates. The wear volume pression:
measurements and corresponding wear rates of the TiB2
v ¼ KPx=H ð1Þ
ceramics are shown in Fig. 5. The wear volume increases
with the load for both the materials. The TiB2–5 wt% where v is the wear volume, P is the normal load, x is the
TiSi2 ceramic exhibited the highest wear rate (5.2  sliding distance, H is the hardness of the material under
106 mm3/N m), while the monolithic TiB2 exhibited investigation, and K is the probability of formation of a
lower wear rate (2.8  106 mm3/N m). wear particle during a particular asperity interaction.
It must be mentioned here that wear volume is a Equation (1) can be rewritten in another form
measure of the amount of material removed during in-
K =H ¼ v=xP ð2Þ
teractive relative motion, whereas specific wear rate gives
an idea of the severity of material loss with load and It must be noted that right hand side of Eq. (2)
sliding distance. In other words, such wear rates pro- gives the wear volume per unit sliding distance per unit
vide an idea of the probability of removal of a material load (or the specific wear rate). Hence, the left hand
fragment during a given asperity encounter. Archard35 term (K/H), which is a measure of the probability of
94 International Journal of Applied Ceramic Technology—Brahma Raju and Basu Vol. 7, No. 1, 2010

Fig. 4. Three-dimensional topographical view of the worn surface of TiB2–5 wt% TiSi2/steel at 10 N load (a), the maximum wear peak
profiles obtained at various loading conditions of the TiB2 composite after fretting against steel counterbody (b), three-dimensional
topographical view of worn surface of the TiB2–5 wt% TiSi2/WC–6 wt% Co at 10 N load (c), and the maximum wear depth profiles obtained
at various loading conditions of the TiB2 composite after fretting against WC–6 wt% Co counterbody (d). Fretting conditions: 4 Hz frequency,
100,000 cycles, and 100 mm stroke length.

wear particle formation normalized by hardness, during fretting. A common observation is that the COF
signifies the estimated wear rate. Although in general, values of the TiB2 ceramics are relatively low after fret-
the amount of material removed is expected to increase ting against WC-cemented carbide, when compared
with increase in load, yet depending upon the with the steel counterbody (Table II).
operative wear mechanisms, the probability of further A typical 3D profile of worn surface of TiB2–5 wt%
material removal (specific wear rate) may not increase TiSi2 is shown in Fig. 4c. The wear depth profiles at
accordingly. various loads are presented in Fig. 4d. It can be noted
that maximum wear depth with wider area is observable
Friction and Wear Behavior Against WC–Co at a highest fretting load of 10 N. Figure 5b represents
Counterbody the effect of load on the wear volume and wear rates of
the TiB2 ceramic fretted against cemented carbide. The
Figures 3c and d reveal that the COF of TiB2 ce- wear volume of both the materials increases with in-
ramics varies insignificantly with the load, when fretted creasing load. However, highest wear volume (B3 
against WC–Co-cemented carbide ball. It can be ob- 104 mm3 at a load of 10 N) is measured for the mono-
served that TiB2–5 wt% TiSi2 exhibits better frictional lithic TiB2. Among all the material combinations,
properties than that of monolithic TiB2. The average TiB2–5 wt% TiSi2 exhibited lowest wear rate of 1.2 
COF of the TiB2 ceramics are shown in Table II. 106 mm3/N m at 10 N load. The above observations
Among all the tribocouples, lowest COF of B0.49 is indicate that different frictional behavior is operative,
noticed with the TiB2–5 wt% TiSi2 against cemented when the TiB2-based materials are fretted against the
carbide. It has to be mentioned here that much fluctu- two counterbodies (steel and WC–Co-cemented car-
ations in the frictional response could be observed for bide). The plausible reasons behind such contrasting
both the tribocouples. Such characteristic features can behavior will be critically analyzed in the subsequent
be attributed to the third body (wear debris) presence sections.
www.ceramics.org/ACT Wear Mechanisms of TiB2 and TiB2–TiSi2 95

6.5 6.5
(a) Filled symbols: Wear volume
6.0 Monolithic TiB2 6.0
5.5 TiB2-5 wt.% TiSi2 5.5

Wear rate (–1 x 10–6 mm3 / Nm)


Wear volume (–1 x 10–4 mm3)

5.0 Unfilled symbols: Wear rate 5.0


TiB2
4.5 TiB2-5 wt.% TiSi2 4.5
4.0 4.0
3.5
3.5
3.0
3.0
2.5
2.5
2.0
2.0
1.5
1.5
1.0
1.0
2 4 6 8 10
Load (N)

6.0 6.0
(b) Filled symbols: Wear volume
5.5 5.5
Monolithic TiB2
5.0 TiB2-5 wt.% TiSi2 5.0

Wear rate (x 10–6 mm3 / Nm)


4.5 Unfilled symbols: Wear rate
Wear volume (x 10–4 mm3)

4.5
TiB2
4.0
TiB2-5 wt.% TiSi2 4.0
3.5
3.5
3.0
3.0
2.5
2.5
2.0
1.5 2.0

1.0 1.5

0.5 1.0
0.0 0.5
2 4 6 8 10
Load (N)

Fig. 5. Variation of wear volume and wear rate of the investigated TiB2 ceramics as a function of load after fretting against steel counterbody
(a) and WC–6 wt% Co counterbody (b). Fretting conditions: 2–10 N load, 4 Hz frequency, 100,000 cycles, and 100 mm stroke length.

SEM Observations of Worn Surfaces of 2, 5, and 10 N. The secondary electron images (Figs.
6a, c, and e) reveals the topography of the worn surfaces
In order to understand the wear mechanisms, the covered with adherent tribolayer. The corresponding
topographical observations of the worn surfaces was BSE images (inset in Fig. 6) provide evidences of the
made using SEM, and simultaneously, compositional formation of a thin adherent tribolayer on the surface
analyses of the observed features have been performed even at a lower load of 2 N. However, a thick protective
using EDS, attached with SEM (Figs. 7–10). In the layer formation can be clearly noticed with increasing
following, specific observations are mentioned with ref- load. Similar microstructural features are also observed
erence to two different counterbodies. with TiB2–5 wt% TiSi2/steel tribocouple (not shown).
Figure 6 shows SEM images of monolithic TiB2 A representative wear scar of the TiB2 composite after
after fretting against steel counterbody at various loads fretting at 10 N load is shown in Fig. 7a. The corre-
96 International Journal of Applied Ceramic Technology—Brahma Raju and Basu Vol. 7, No. 1, 2010

Table II. Summary of the Average COF, Wear Volume, and Wear Rate of the Different Investigated
Tribocouples as a Function of the Load
Wear volume of TiB 2 flat Wear rate of TiB2 flat
Average COF materials (mm 3 )  10 4 materials (mm 3 /N m)  10 6

Tribocouple 2N 5N 10 N 2N 5N 10 N 2N 5N 10 N
TiB2 versus steel 0.58 0.69 0.63 1.87 2.98 4.15 4.67 2.98 2.07
TiB2–5 wt% TiSi2 0.62 0.70 0.67 2.09 3.62 5.48 5.22 3.62 2.74
versus steel
TiB2 versus WC–Co 0.56 0.55 0.56 1.64 2.34 3.02 4.10 2.34 1.51
TiB2–5 wt% TiSi2 0.51 0.49 0.56 1.15 2.28 2.53 2.88 2.28 1.26
versus WC–Co
The presented data is an average of at least two test results.
COF, coefficient of friction.

sponding EDS analysis of tribolayer (magnified image) and minute O and no evidence of transferred W can be
reveals strong peaks of Fe and O along with weak Ti and found (inset of Fig. 10b). EDS analysis of wear debris
B elements. Figure 7c shows the worn surface of steel reveals the presence of titanium and oxygen (Figs. 10c
ball at 10 N load. It reveals the presence of tribolayer and d). In case of TiB2–5 wt% TiSi2/cemented carbide
and severe abrasion grooves. The EDS analysis of the tribosystem, similar morphology of worn surfaces, like
worn surface (inset of Fig. 7d) indicated primarily Fe monolithic TiB2, is noticed (see Figs. 11a and b). Figure
and O peaks. In order to further understand the tribo- 11 compares the worn surfaces of both the TiB2–5 wt%
chemistry of the layer, EDS analysis is carried out on the TiSi2 and WC–Co counterbody after fretting at a load
worn surface of TiB2–5 wt% TiSi2 after fretting at 2, 5, of 10 N. The size of wear scar (B451 mm) for both the
and 10 N loads (see Fig. 8). Although the presence of O, TiB2 composite and the WC–Co ball remains almost
Ti, and Fe are detected after testing at 2 N load; much same. SEM observation of the worn surface of WC–Co
pronounced peaks of O and Fe can only be noticed at ball shows the occurrence of very mild abrasion wear,
5 and 10 N load. It is possible that TixOy, FexOy or their with grain pull-out. The corresponding EDS spectrum
mixed oxides form during fretting of the TiB2 ceramics obtained from the abraded area of cemented carbide ball
against steel. indicates the predominant presence of W and minute
The topographical features of the wear scars of the Co and O. No evidence of transferred Ti can be found
monolithic TiB2 after fretting against the cemented car- (inset of Fig. 11d). Further discussion regarding the
bide are shown in Fig. 9. The worn surfaces are much wear mechanisms responsible for different tribological
different from those obtained during fretting against responses of the TiB2-based materials will be presented
steel. The mild abrasive scratches, along with brittle in the following section.
fracture induced pull-outs near the edges of wear are
observed for monolithic TiB2 (see insets of Figs. 9b
and d). The worn surfaces along with the corresponding
EDS of the TiB2 flat after fretting at a load of 10 N is Discussion
shown in Fig. 10. The presence of wear debris near the
edge of the wear scar is observed on TiB2 sample (see Based on the experimental results and SEM–EDS
Fig. 10a). It reveals the presence of very fine debris along investigation of worn surfaces of both the TiB2 ceramics
with agglomerated particles. Mild abrasion along with and counterbodies, we will now make an attempt to
the minute particle pullouts are noticed and the forma- discuss the underlying material removal mechanisms
tion of tribochemical layer on the worn surfaces was not during fretting. It can be realized that variation of
observed. EDS spectrum obtained from the abraded load and counterbody material notably affects the fric-
area of TiB2 indicates the predominant presence of Ti tion and wear behavior of TiB2-based materials.
www.ceramics.org/ACT Wear Mechanisms of TiB2 and TiB2–TiSi2 97

Fig. 6. Overview and details of as-fretted surfaces (scanning electron microscope [SEM] images) of monolithic TiB2 after 2 N (a, b),
5 N (c, d), and 10 N loads (e, f). (a), (c), and (e) represent the topography of worn surfaces (SE SEM images), while (b), (d), and (f) indicate
the magnified view of worn surfaces (BSE SEM images). Fretting conditions: 4 Hz frequency, 100 mm stroke length, and 100,000 cycles:
counterbody: bearing steel (arrow marks indicate the fretting direction, inset diagrams are BSE images of the worn surfaces).

Negative Wear Phenomena of TiB2 Ceramics with increasing load. The direct contact of hard TiB2
Against Steel ceramic with the steel counterbody causes the adhesion
and subsequent fracture of asperities of mating surfaces
It is known that during initial stages of fretting, the and leads to the wear debris formation. At 2 N load,
deformation or fracture of surface asperities of mating fluctuations in COF curve is mainly due to the forma-
bodies result in the formation of wear debris. The fric- tion of wear debris and constant adhesion/abrasion of
tional heat, generated at the sliding contacts, oxidizes the contact surfaces. During the test, the wear debris
the debris particles at ambient testing conditions and particles are compacted to form a tribolayer under load
leads to the tribochemical reactions. Both the TiB2 ce- and atmospheric conditions. When the load increases to
ramics exhibits almost similar trend of increasing COF 5 or 10 N, this takes place more promptly.
98 International Journal of Applied Ceramic Technology—Brahma Raju and Basu Vol. 7, No. 1, 2010

(a) (b)

(c) (d)

Fig. 7. Worn surfaces of TiB2–5 wt% TiSi2 (a), the magnified image of the worn surface along with the energy-dispersive spectroscopy (EDS)
reveals the tribochemical layer (b), the corresponding worn surface of steel ball (c), and the magnified image of the worn surface along with the
EDS (d). Fretting conditions: 10 N load, 4 Hz frequency, 100 mm stroke length, and 100,000 cycles: counterbody: bearing steel (arrow marks
indicate the fretting direction).

From Table I, it is very clear that the TiB2 samples Fe–O, while traces of Ti along with Fe and O at higher
are B2.6–3.5 times harder than the steel ball. Hence, loads (  5 N). It reflects that with increasing load,
during fretting under the application of load, the steel more iron is getting transferred to the counterbody. A
body got abraded by TiB2 and subsequently, the wear detailed study on tribochemistry of TiB2/steel tribocou-
debris of steel got compacted and form tribochemical ple by using Raman spectroscopy and XPS analysis can
layer on the mating contact surfaces. Figure 4b indicates be found in our earlier work.4,23 Such studies revealed
that the tribolayer thickness increases (from 2.5 to that the tribolayer composed of TiO2 and Fe2O3 for the
8 mm) with increasing the load from 2 to 10 N. The TiB2/steel-fretting contact. The present experimental
topography of the worn surfaces of TiB2 samples further results also identify similar tribochemistry of the worn
confirms the increment of the adherent stable tribolayer surfaces. The possible tribochemical reactions involved
with the load (see Fig. 6). Based on the SEM–EDS of are as follows:
worn surfaces of the TiB2 ceramics as well as steel
2TiB2 þ 5O2 ! 2TiO2 þ 2B2 O3 ð3Þ
counterbody during fretting, it is clear that the oxide
debris are formed mainly from steel surface. The worn
Fe þ ð1=2ÞO2 ! FeO ð4Þ
surfaces of steel body clearly evidences abrasive grooves.
However, no such feature can be found on the TiB2
2Fe þ ð3=2ÞO2 ! Fe2 O3 ð5Þ
counterbody (see Fig. 7). A closer look at the EDS
spectra of both the worn surfaces of TiB2–5 wt% TiSi2
3Fe þ 2O2 ! Fe3 O4 ð6Þ
sample and the steel ball at a load of 10 N reveals the
predominant presence of strong O and Fe peaks. These In the present investigation, we observe a new phe-
observations clearly indicate the formation of iron oxide nomenon of negative wear of TiB2 ceramics under the
layer on the worn surfaces. Figure 8 represents that at a present testing conditions. Table II presents that nega-
lower load of 2 N, the tribolayer is composed of Ti– tive wear volume of the samples systematically increases
www.ceramics.org/ACT Wear Mechanisms of TiB2 and TiB2–TiSi2 99

demonstrate that as the TiB2 composite exhibits better


mechanical properties, it wears out the counterbody
more than the monolithic TiB2 and results in higher
values of negative wear. Endo and Maru36 reported that
for cemented carbide/steel-fretting contact, a negative
wear is induced due to the adherence of iron oxide of
the steel body to cemented carbide. It is interesting to
note that wear rate decreases with increasing load (see
Table II). At higher loads, the contact between the
tribolayers of the tribocouples reduces the probability
of further wear of steel counterbody. SEM observations
also reveal the presence of a protective tribolayer at
higher load (10 N).

Wear behavior of TiB2 Ceramics Against WC–Co

Based on the friction and wear data as well as to-


pographical analysis of tribosurfaces, the tribological
behavior TiB2/WC–Co tribocouple under unlubricated
fretting conditions will be discussed in this section. Fur-
ther, the possible wear mechanisms can be understood
in the light of the friction and wear behavior of the
materials.
It has been observed that the frictional behavior of
both the TiB2-based ceramics, during fretting against
the WC–Co-cemented carbide counterbody are nearly
similar. Up to a load of 5 N, the COF curves show
fluctuations during the entire fretting duration. It can
be attributed to the formation of wear debris and their
escape/displacement from the contact zone during fret-
ting. It has to be mentioned here that fretting wear has
the characteristic of forming considerable amounts of
very fine oxide debris and the wear debris particles sub-
sequently govern the wear of the materials.3 Singer37
reported that wear debris can reduce friction by accom-
modating sliding and absorbing the deformation energy
and thereby, also reduce wear. The COF of TiB2/
cemented carbide is relatively lower when compared
with TiB2/steel tribocouple. It is generally recognized
that interactive relative motion of a ceramic against an-
Fig. 8. Energy-dispersive spectroscopy compositional analysis of the
worn surfaces of TiB2–5 wt% TiSi2 after fretting at various
other hard ceramic, as compared with ceramic against
loading conditions for a duration of 100 K cycles: Other fretting metal, leads to lower friction and wear due to formation
parameters include: 4 Hz frequency, 100 mm stroke length, of smaller real contact area. Such a phenomenon is ex-
counterbody: bearing steel. pected as a result of nonconformity and limited elastic/
plastic deformation between the two surfaces.
with the load. However, maximum negative wear vol- Looking at the SEM images, the worn surfaces
ume of 5.4  104 mm3 is measured with the TiB2– are characterized with mild abrasive scratches, grain
5 wt% TiSi2 and is attributed to its high density and pullouts, and polishing of the surfaces (Fig. 9). How-
high hardness. These observations comprehensively ever, at certain regions, more material removal and par-
100 International Journal of Applied Ceramic Technology—Brahma Raju and Basu Vol. 7, No. 1, 2010

Fig. 9. Scanning electron microscope images revealing the topography of worn surfaces of monolithic TiB2 at 2 N (a, b) and 5 N loads
(c, d). (a) and (c) represent the wear scars, while (b) and (d) indicate the magnified images of the worn surfaces. Fretting conditions:
4 Hz frequency, 100 mm stroke length, and 100,000 cycles: counterbody: WC–6 wt% Co (double-headed arrow marks indicate the
fretting direction, inset diagrams are representative of highly worn regions at certain region of wear scars: single-headed arrow marks
from (a) and (c)).

ticle pullouts are observed (see inset Figs. 9b and d). and b). Here it has to be mentioned that the micro-
Even at a higher load of 10 N, the characteristics of the structure of unworn surface reveals the presence of sec-
worn surface remains the same as that of 2 and 5 N. ondary Ti5Si3 phase (black contrast), compacted wear
With increasing load, relatively more abrasive grooves debris and particle pullouts. The worn surface of WC–
can be seen (Fig. 10). EDS analysis of the worn surface Co ball also consists of mild abrasive grooves and EDS
reveals strong Ti peak and traces of O. There is no in- analysis of the worn surface reveals W, Co, and O peaks.
dication of any transfer of material from the cemented Hence, the primary wear mechanism is abrasion and
carbide counterbody. From the micrographs, finer and no tribochemical effects, except mild oxidation can be
nearly spherical wear debris particles can be observed on noticed for the TiB2/WC–Co tribosystem. Maximum
the surfaces at higher loads (Fig. 10). A large quantity of wear volume of 3  104 is measured with TiB2/WC–
wear debris is located around the wear scar and some Co at 10 N load. From Table II, it can be noticed that
amount of compacted debris could be found on the both the COF and wear volumes of TiB2 composites are
worn surface. EDS analysis of wear debris also reveals relatively lower than monolithic TiB2. Hence, a better
the presence of Ti and O only (Fig. 10d). It further wear resistance of TiB2 composites can be attributed to
confirms that except the third body abrasion, no tribo- its higher density and good mechanical properties.
chemical wear takes place for TiB2/cemented carbide Yang et al.25 investigated tribological performance
tribocouples. of TiB2 against different counterbodies (SiC, Al2O3,
A representative worn surface of the TiB2 compos- and mullite). The COF varied between 0.63 and 0.77 and
ite shows mild abrasive grooves and the corresponding the specific wear rates from (7.1 to 68.5)  105 mm3/
EDS reveals the presence of Ti, B, Si, and O (Figs. 11a N m with different tribocouples. For TiB2/alumina fret-
www.ceramics.org/ACT Wear Mechanisms of TiB2 and TiB2–TiSi2 101

(a) (b)

(c) (d)

Fig. 10. Scanning electron microscope images revealing the topography of worn surface of monolithic TiB2 along with the wear debris after
fretting against WC–6 wt% Co at 10 N load (a), the magnified image of worn surface along with the energy-dispersive spectroscopy (EDS) (b),
magnified image of wear debris (c), and the corresponding EDS of wear debris (d) Fretting conditions: 4 Hz frequency, 100 mm stroke length,
and 100,000 cycles (double-headed arrow marks indicate the fretting direction).

ting contacts, a steady-state COF of 0.5 and wear rate of (a) In case of TiB2/steel tribosystem, the COF var-
2  106 mm3/N m is reported.23 In the present inves- ied between 0.58 and 0.70. However, relatively higher
tigation, the COF lies between 0.49 and 0.56 and wear COF against steel is noticed with TiB2–5 wt% TiSi2 at
rate measured to be (1.2–4.1)  106 mm3/N m when all loads, when compared with monolithic TiB2.
the TiB2 ceramics fretted against WC–Co counterbody, (b) Adhesion, abrasion, and tribochemical layer
depending on the loading condition. Our experimental formation are identified as the major wear mechanism
results clearly indicate that the friction and wear proper- for TiB2 against steel. The harder TiB2 abrades the steel
ties of the developed TiB2 ceramics are relatively better or counterbody and the transfer/sticking of tribolayer to
comparable with the earlier reported TiB2 tribosystems. the TiB2 surface contributes to negative wear of TiB2
The results of the present study therefore reveal the po- ceramics. Maximum wear rate of 5.2  106 mm3/
tentiality of the TiB2 materials in contact with the steel N m is measured with the TiB2 composite.
and cemented carbide for wear resistant applications. (c) In case of TiB2/WC–Co tribosystem, the COF
varied between 0.49 and 0.56. The COF values of the
TiB2 composites are relatively lower than monolithic
Conclusions TiB2. The presence of wear debris as third body between
the contacting bodies appears to cause low COF.
The present investigation reports the tribological (d) Adhesion and mild abrasion are major wear
performance of TiB2 and TiB2–5 wt% TiSi2, in fretting mechanisms for the TiB2 ceramics when fretted against
contact with two different counterbodies (steel and cemented carbide. The wear rate varied between (1.2
WC–Co cemented carbide) under varied load condi- and 4.1)  106 mm3/N m for both the TiB2 ceramics
tions. and falls with in the mild wear regime. However, rel-
102 International Journal of Applied Ceramic Technology—Brahma Raju and Basu Vol. 7, No. 1, 2010

(a) (b)

(c) (d)

Fig. 11. Worn surfaces of TiB2–5 wt% TiSi2 (a), the magnified image of the worn surface along with the energy-dispersive spectroscopy
(EDS) (b); the corresponding worn surface of WC–6 wt% Co ball (c), and the magnified image of the worn surface along with the EDS
(d). Fretting conditions: 10 N load, 4 Hz frequency, 100 mm stroke length, and 100,000 cycles (arrow marks indicate the fretting direction).

5. B. Basu, G. B. Raju, and A. K. Suri, ‘‘Processing and Properties of Monolithic


atively lower wear rates are measured with the TiB2 TiB2-Based Materials,’’ Int. Mater. Rev., 51 [6] 1–23 (2006).
composite, when compared with monolithic TiB2. The 6. T. S. R. Ch. Murthy, B. Basu, R. Balasubramaniam, A. K. Suri, C. Subra-
better wear resistance of TiB2 composite is attributed monian, and R. K. Fotedar, ‘‘Processing and Properties of TiB2 with
MoSi2 Sinter-Additive: A First Report,’’ J. Am. Ceram. Soc., 89 [1] 131–
to microstructural difference and better mechanical 139 (2006).
properties. 7. M. K. Ferber, P. F. Becher, and C. B. Finch ‘‘Effect of microstructure on the
properties of TiB2 ceramics,’’ Communications of the American Ceramic
Society, January C-2-3 1983.
8. S. Kang, D. J. Kim, E. S. Kang, and S. S. Baek, ‘‘Pressureless Sintering and
Acknowledgments Properties of Titanium Diboride Ceramics Containing Chromium and
Iron,’’ J. Am. Ceram. Soc., 84 [4] 893–895 (2001).
9. L. H. Li, H. E. Kim, and E. S. Kang, ‘‘Sintering and Mechanical Properties of
G. B. R. thanks Dr. S. J. Cho and his group mem- Titanium Diboride with Aluminum Nitride as a Sintering Aid,’’ J. Eur.
bers at KRISS, Korea for extending co-operation during Ceram. Soc., 22 973–977 (2002).
his stay to hot press some of the ceramic samples, which 10. J. Ho Park, Y. Koh, C. Kim, H. Hwang, and E. Kong, ‘‘Densification and
Mechanical Properties of Titanium Diboride with Silicon Nitride as Sintering
are used in the present investigation. Aid,’’ J. Am. Ceram. Soc., 82 [11] 3037–3042 (1999).
11. Z. Xingzhong, L. Jiajun, Z. Baoliang, O. Jinlin, and X. Qunji, ‘‘Tribological
Properties of TiC-Based Ceramic/High Speed Steel Pairs at High Temper-
ature,’’ Ceram. Int., 24 13–18 (1998).
References 12. A. H. Jones, R. S. Dobedoe, and M. H. Lewis, ‘‘Mechanical Properties and
Tribology of Si3N4–TiB2 Ceramic Composites Produced by Hot Pressing
1. R. B. Waterhouse, ‘‘Fretting Wear,’’ Wear, 100 107–118 (1984). and Hot Isostatic Pressing,’’ J. Eur. Ceram. Soc., 21 969–980 (2001).
2. D. Klaffke, ‘‘Fretting Wear of Ceramics,’’ Tribol. Int., 22 [2] 89–101 (1989). 13. B. Basu, D. Sarkar, and T. Venkateswaran, ‘‘Pressureless Sintering and
3. M. Varenberg, G. Halperin, and I. Etsion, ‘‘Different Aspects of the Role of Tribological Properties of WC–ZrO2 Composites,’’ J. Eur. Ceram. Soc., 25
Wear Debris in Fretting Wear,’’ Wear, 252 902–910 (2002). 1603–1610 (2005).
4. J. Vleugels, B. Basu, K. C. H. Kumar, R. G. Vitchev, and O. Van Der Biest, 14. B. Basu, R. G. Vitchev, J. Vleugels, J. P. Celis, and O. Van Der Biest, ‘‘In-
‘‘Unlubricated Fretting Wear of TiB2 Containing Composites Against Bear- fluence of Humidity on the Fretting Wear of Self-Mated Tetragonal Zirconia
ing Steel,’’ Metall. Mater. Trans. A, 33 [12] 3847–3859 (2002). Ceramics,’’ Acta Mater., 48 2461–2471 (2000).
www.ceramics.org/ACT Wear Mechanisms of TiB2 and TiB2–TiSi2 103

15. W. M. Rainforth, ‘‘The Sliding Wear of Ceramics,’’ Ceram. Int., 22 365–372 26. R. Wasche, D. Klaffke, and T. Troczynski, ‘‘Tribological Performance of SiC
(1996). and TiB2 Against SiC and Al2O3 at Low Sliding Speeds,’’ Wear, 256 695–704
16. J. Mukerji and B. Prakash, ‘‘Wear of Nitrogen Ceramics and Composites in (2004).
Contact with Bearing Steel Under Oscillating Sliding Condition,’’ Ceram. 27. B. Prakash, C. Ftikos, and J. P. Celis, ‘‘Fretting Wear Behavior of PVD TiB2
Int., 24 19–24 (1998). Coatings,’’ Surf. Coat. Technol., 124 253–261 (2000).
17. J. Larsen-Basse, ‘‘Binder Extrusion in Sliding Wear in WC–Co Alloys,’’ Wear, 28. M. Burger and S. Hogmark, ‘‘Tribological Properties of Selected PVD Coat-
105 247–256 (1985). ings when Slid Against Ductile Materials,’’ Wear, 252 557–565 (2002).
18. M. Z. Huq and J. P. Celis, ‘‘Expressing Wear Rate in Sliding Contacts Based 29. M. Burger and S. Hogmark, ‘‘Evaluation of TiB2 Coatings in Sliding Contact
on Dissipated Energy,’’ Wear, 252 375–383 (2002). Against Aluminium,’’ Surf. Coat. Technol., 149 14–20 (2002).
19. J. Pirso, S. Letunovits, and M. Viljus, ‘‘Friction and Wear Behaviour of 30. M. Burger, M. Larsson, and S. Hogmark, ‘‘Evaluation of Magnetron-
Cemented Carbides,’’ Wear, 257 257–265 (2004). Sputtered TiB2 Intended for Tribological Applications,’’ Surf. Coat. Technol.,
20. J. L. Ortiz-Merino and R. I. Todd, ‘‘Relationship between Wear Rate, 124 253–261 (2000).
Surface Pullout and Microstructure During Abrasive Wear of Alumina 31. M. Berger, L. Karlsson, M. Larsson, and S. Hogmark, ‘‘Low Stress Coatings with
and Alumina/SiC Nanocomposites,’’ Acta Mater., 53 3345–3357 Improved Tribological Properties,’’ Thin Solid Films, 401 179–186 (2001).
(2005). 32. Y. Yang, Z. Zheng, X. Wang, X. Liu, J. G. Han, and J. S. Yoon, ‘‘Micro-
21. M. Kalin and J. Vizintin, ‘‘Use of Equations for Wear Volume Determination structure and Tribology of TiB2 and TiB2/TiN Double-Layer Coatings,’’
in Fretting Experiments,’’ Wear, 237 39–48 (2000). Surf. Coat. Technol., 84 404–408 (1996).
22. B. Basu, J. Vleugels, and O. Vanderbiest, ‘‘Fretting Wear Behavior of TiB2- 33. G. B. Raju and B. Basu, ‘‘Densification, Sintering Reactions and Properties of
Based Materials Against Bearing Steel Under Water and Oil Lubrication,’’ Titanium Diboride with Titanium Disilicide as a Sintering Aid,’’ J. Am.
Wear, 250 631–641 (2001). Ceram. Soc., 90 [11] 3415–3423 (2007).
23. B. Basu, J. Vleugels, and O. Vanderbiest, ‘‘Unlubricated Tribological Per- 34. G. R. Anstis, P. Chantukul, B. R. Lawn, and D. B. Marshall, ‘‘A Critical
formance of Advanced Ceramics and Composites at Fretting Contacts with Evaluation of Indentation Techniques for Measuring Fracture Toughness,’’
Alumina,’’ J. Mater. Res., 18 [6] 1314–1324 (2003). J. Am. Ceram. Soc., 64 533–538 (1981).
24. T. S. R. Ch. Murthy, B. Basu, A. Srivastava, R. Balasubramaniam, and A. K. 35. J. F. Archard and W. Hirst, ‘‘The Wear of Metals Under Unlubricated Con-
Suri, ‘‘Tribological Properties of TiB2 and TiB2–MoSi2 Ceramic Compos- ditions,’’ Proc. R. Soc. Lond., A236 397–410 (1956).
ites,’’ J. Eur. Ceram. Soc., 26 1293–1300 (2006). 36. H. Endo and E. Maru, ‘‘Studies on Fretting Wear: Influence of Rubbing
25. Q. Yang, T. Senda, N. Kotani, and A. Hirose, ‘‘Sliding Wear Behavior and Surface Materials and Some Considerations,’’ Wear, 253 795–802 (2002).
Tribofilm Formation of Ceramics at High Temperatures,’’ Surf. Coat. Tech- 37. I. L. Singer, ‘‘How Third-Body Processes Affect Friction and Wear,’’ MRS
nol., 184 270–277 (2004). Bull. 37–40 (1998).

You might also like