You are on page 1of 9

Wear 258 (2005) 13481356

Friction and wear of titanium alloys sliding against metal,


polymer, and ceramic counterfaces
Jun Qua, , Peter J. Blaua , Thomas R. Watkinsa , Odis B. Cavinb , Nagraj S. Kulkarnia
a

Metals and Ceramics Division, Oak Ridge National Laboratory, P. O. Box 2008, MS 6063, Oak Ridge, USA
b University of Tennessee, Knoxville, USA
Received 11 July 2003; received in revised form 21 September 2004; accepted 23 September 2004
Available online 11 November 2004

Abstract
Recent advances in lower-cost processing of titanium, coupled with its potential use as a light weight material in engines and brakes has
renewed interest in the tribological behavior of titanium alloys. To help establish a baseline for further studies on the tribology of titanium
against various classes of counterface materials, pin-on-disk sliding friction and wear experiments were conducted on two different titanium
alloys (Ti6Al4V and Ti6Al2Sn4Zr2Mo). Disks of these alloys were slid against fixed bearing balls composed of 440C stainless steel,
silicon nitride, alumina, and polytetrafluoroethylene (PTFE) at two speeds: 0.3 and 1.0 m/s. The friction coefficient and wear rate were lower at
the higher sliding speed. Ceramic sliders suffered unexpectedly higher wear than the steel slider. The wear rates, ranked from the highest to the
lowest, were alumina, silicon nitride, and steel, respectively. This trend is inversely related to their hardness, but corresponds to their relative
fracture toughness. Comparative tests on a Type 304 stainless steel disk supported the fracture toughness dependency. Energy dispersive
spectroscopy (EDS) and X-ray diffraction (XRD) analyses confirmed the tendency of Ti alloys to transfer material to their counterfaces and
suggested possible tribochemical reactions between the ceramic sliders and Ti alloy disks. These reaction products, which adhere to the
ceramic sliders, may degrade the mechanical properties of the contact areas and result in high wear. The tribochemical reactions along with
the fracture toughness dependency helped explain the high wear on the ceramic sliders.
2004 Elsevier B.V. All rights reserved.
Keywords: Titanium; Ceramics; Material transfer; Tribochemical reaction

1. Introduction
In comparison to light weight alloys based on aluminum
and magnesium, titanium alloys present interesting possibilities as tribomaterials, but they have not been widely investigated as bearing materials. They are harder and stiffer than
Mg and Al alloys, and they resist exposure to heat and aqueous corrosion much better. Like Al and Mg, their high affinity
Research sponsored by the U.S. Department of Energy, Assistant Secre-

tary for Energy Efficiency and Renewable Energy, Office of FreedomCAR


and Vehicle Technologies, as part of the High Strength Weight Reduction
Materials Program, under contract DE-AC05-00OR22725 with UT-Battelle,
LLC.
Corresponding author. Tel.: +1 865 574 4560; fax: +1 865 574 6918.
E-mail address: qujn@ornl.gov (J. Qu).
0043-1648/$ see front matter 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.wear.2004.09.062

for oxygen results in the formation of an adherent surface oxide, but sub-stoichiometric TiO2 can act as a solid lubricant.
A great deal is known about the physical metallurgy,
heat treatment, and mechanical properties of titanium alloys,
thanks to extensive aerospace-related research and development. Tribological concerns for Ti in aerospace components
have focused mainly on their fretting behavior, leading to research on surface treatments like ion implantation and solid
film lubrication [1,2]. Needs in the chemical process industry
motivated a 1991 study of the galling and sliding wear behavior of commercial-purity Ti and alloy Ti6Al4V [3]. In
that investigation, the best wear and friction results for Ti alloys were obtained for anodized counter-surfaces coated with
MoS2 solid-film or with polytetrafluoroethylene (PTFE), but
the abrasion resistance was poor. Relatively few additional

J. Qu et al. / Wear 258 (2005) 13481356

1349

Table 1
Compositions and characteristics of Ti6Al4V and Ti6Al2Sn4Zr2Mo
Ti alloy

Compositions (wt%)* (balance Ti)

Microindentation hardness HV (GPa)

Tensile strength UTS (MPa)

Ti64
Ti6242

6.53 Al, 3.89 V, 0.13 Fe


5.85 Al, 1.98 Sn, 4.22 Zr, 1.95 Mo

3.36 0.17
3.31 0.10

954.9
957.0

Analyses supplied by Titanium Metals Corporation.

studies have been conducted on sliding wear mechanisms of


Ti alloys. Molinari et al. highlighted the mechanisms responsible for the wear resistance under different load and sliding
speed conditions in self-mated Ti6Al4V disk-on-disk sliding tests [4], and Dong and Bell [5] reported unexpectedly
high wear rates for alumina sliding against Ti6Al4V (pinon-disk tests).
Recent developments in Ti processing forecast the availability of lower-cost Ti and that has prompted further interest
in exploring the tribological behavior of Ti alloys as bearing
materials [6]. Focus by the U.S. Department of Energy on improved brake materials for fuel-efficient heavy trucks, has led
to the consideration of Ti for disc brake rotors as well. In fact,
coated Ti brake discs are already showing promise in auto racing [7]. This renewed interest in the friction and wear of Ti
alloys has prompted the current laboratory study of the behavior of two commercially-available Ti alloys sliding against
model metallic, ceramic, and polymeric counterfaces. One
of the two alloys (Ti6Al4V) has had more tribological attention than the other, but the other (Ti6Al2Sn4Zr2Mo)
has attractive elevated temperature properties and was felt
to be of interest as well. There are very few studies on the
tribological properties of Ti6Al2Sn4Zr2Mo in the literature. In this study, it is intended to establish baseline data for
these alloys with which to compare the tribological behavior
of new surface treatments or coatings in future work.

2. Materials and testing procedure


Two titanium alloys, Ti6Al4V and Ti6Al2Sn
4Zr2Mo, were tested in this study. These alloys shall subsequently be referred to as Ti64 and Ti6242. The compositions

of these alloys, provided by the supplier for these heats of material, are listed in Table 1 and their microstructures are shown
in Fig. 1(a) and (b), respectively. The typical , /-phase
grain structure can readily be identified on the etched crosssection of Ti64. Ti6242 has finer grain size and is dominated
by -phase. The two alloys have similar microindentation
hardness and tensile strength (see Table 1).
Friction and wear tests were conducted using a pin-on-disk
apparatus. The diameter of the fixed ball sliders was 9.53 mm.
As shown in Table 2, 440C stainless steel, silicon nitride, alumina, and polytetrafluoroethylene (PTFE), were selected to
represent metallic, ceramic, and polymeric bearing materials.
The Ti alloy disks were 63.5 mm diameter and 12.7 mm thick.
The disk surfaces were polished by 600 grit wet SiC abrasive
paper and the pre-test arithmetic average surface roughness
(Ra ), measured with a Taylor Hobson TalysurfTM 10 stylus
profilometer with a 2.5 m tip radius, was 0.11 0.02 m.
Concentric wear tracks ranging from 18 to 52 mm in diameter
were used, and the disk rotation rate was adjusted accordingly
to provide either 0.3 or 1.0 m/s sliding speed. A 10 N normal
load was applied and the test was run for 500 m sliding distance. In follow-on experiments, and to provide a comparison
to the results for the Ti alloy disks, a 304 stainless steel disk
(63.5 mm diameter and 6.35 mm thick) was tested against
440C stainless steel and ceramic (Si3 N4 and Al2 O3 ) sliders under similar sliding conditions. The microindentation
Vickers hardness of the 304 stainless steel disk was about
3.16 GPa, slightly lower than that of the Ti alloy disks.
The friction force was monitored by a load cell-based
force measurement system. The wear volumes of the sliders
and disks were determined by weight change measurements
with an accuracy of 0.1 mg. The wear factor is defined as the
wear volume normalized by the applied load and the sliding

Fig. 1. Microstructures of two Ti alloys.

1350

J. Qu et al. / Wear 258 (2005) 13481356

Table 2
Characteristics of slider materials
Sliders

Supplier

Specification

Vickers hardness (GPa)

Fracture toughness (MPa m1/2 )

440C stainless steel


Silicon nitride
Alumina
PTFE

McMaster-Carr
Cerbec East Granby, CT
Southern Bearing Service
W.M. Berg, Inc. East Rockaway, NY

Grade 100 Hardened


NBD200 Grade 5
AFBMA Grade 25

12.60a
19.37a
24.75a
N/A

23.7b
5.2c
34c
N/A

The Vickers hardness was measured using a 100 g load.


The Izod impact strength of hardened 440C stainless steel is 4 ft lb [8]. The fracture toughness here was estimated based on BarsomRolfes empirical
formula [9], with the assumption that the Izod impact strength is close to the Charpy V-notch impact strength.
c The fracture toughness values were provided by the suppliers.
b

distance of the pin. All the tests were conducted in ambient


air conditions with temperature and humidity in the range
of 1822 C and 5262%, respectively. At least two duplicates were run at each test condition. Good repeatability was
obtained in both friction and wear results.

large fluctuation of the friction coefficient was thought to


be caused by formation and periodic, localized fracture of
a transfer layer. Titanium alloy commonly transfers to the
counterface when rubbing against other metals or ceramics
[35]. In this study, surface morphology examination and
surface analysis confirmed this tendency. A transfer layer
was easily identified on the wear scar of the 440C stainless steel ball (see Fig. 3(a)). The energy dispersive spectroscopy (EDS) analysis detected Ti and/or Al on the worn
surfaces of the metal and ceramic balls, as shown in Fig. 4.
More discussion on surface analysis can be found on Section
4.2.
For metal and ceramic balls, lower friction coefficient and
smaller instantaneous fluctuation were observed at 1.0 m/s
compared to those at 0.3 m/s, as shown in Table 3. At higher
sliding speed, the contact area had higher temperature, which
generally reduced the shear strength and led to lower friction
forces.

3. Results
Results are summarized in Table 3(a) and (b) for Ti64 and
Ti6242, respectively. Friction and wear results are presented
separately.
3.1. Friction
Table 3(a) and (b) present the average friction coefficient and its fluctuation at steady-state for each test condition. Selected friction traces of the four different sliders
against Ti64 disks are shown in Fig. 2. The PTFE slider generated a fairly smooth friction trace (Fig. 2(d)) due to its
self-lubricating nature. The metal and ceramic sliders produced friction coefficient in the range of 0.340.50 with relatively large fluctuation, as illustrated in Fig. 2(a)(c). The

3.2. Wear
As shown in Table 3(a) and (b), up to five times higher wear
factors were obtained on both the slider and disk at 0.3 m/s

Table 3
Friction and wear results
Slider material

Sliding speed
0.3 m/s
Friction coefficient

1.0 m/s
Wear factor

(mm3 /N m)

Ball

Disk

0.50 0.05
0.47 0.07
0.49 0.07
0.28 0.001

6.9 106
3.8 105
5.7 105
8.4 104

1.7 104
3.5 104
5.7 104
N/Ma

(b) Ti6Al2Sn4Zr2Mo disks


440C stainless steel
0.48 0.05
Silicon nitride
0.47 0.08
Alumina
0.49 0.08
PTFE
0.27 0.001

5.1 106
4.4 10-5
1.2 104
9.9 104

1.3 104
3.5 104
3.4 104
N/Ma

(a) Ti6Al4V disks


440C stainless steel
Silicon Nitride
Alumina
PTFE

N/M, not measurable.

Friction coefficient

Wear factor (mm3 /N m)


Ball

Disk

0.35 0.05
0.36 0.07
0.44 0.07
0.29 0.001

1.6 106
6.2 106
1.6 105
6.1 104

1.5 104
1.3 104
2.0 104
N/Ma

0.34 0.04
0.37 0.02
0.42 0.04
0.29 0.001

1.22 106
9.40 10-6
2.33 105
7.92 104

1.1 104
1.1 104
2.2 104
N/Ma

J. Qu et al. / Wear 258 (2005) 13481356

1351

Fig. 2. Frictional traces of different sliders against Ti64 disks.

than those at 1.0 m/s. Similar sliding speed dependency was


also reported by other researchers [5].
The Ti disks suffered high wear rates, in the order of
104 mm3 /N m, against the metal and ceramic balls, and
harder sliders generated relatively more (or at least comparable) wear on the Ti disks. Although harder balls were expected
to have higher wear resistance, the results in Table 3(a) and
(b) show a reverse order: the alumina ball wore more than the
silicon nitride ball, which in turn wore more than the stainless
steel ball. Remarkably, the wear factors of the ceramic balls
were at least five times higher than those of the steel balls.
Dong and Bell also reported a higher wear rate of an alumina
ball than that of a steel ball when sliding against a Ti64 disk
[5]. More analysis and discussion are presented in Section 4.
Fig. 3 shows the wear scars on the metal and ceramic
balls with features of abrasive wear, adhesive wear, and plastic deformation. Abrasive wear seemed to dominate the wear
process at 0.3 m/s. Those wear scars were larger and flatter,
corresponding to their higher wear factors. The wear scars
generated at 1.0 m/s were smaller but much rougher with
larger patches of transferred material implying more severe
adhesive wear, possibly due to higher temperature at the contact area. EDS analysis also showed higher Ti and/or Al concentration on the wear scars at 1.0 m/s.
The PTFE slider had the highest wear factor
(103 mm3 /N m). Its counterface (Ti disk) was pro-

tected by the polymeric layer transferred from the PTFE


ball and had almost no surface damage except a few shallow
circular groves ground by third body particles, probably
some metal debris, embedded in the PTFE ball.
It has been seen that these two Ti alloys showed similar friction and wear behavior. Therefore, discussion will be
focused on Ti64 only.

4. Discussion
The most unusual finding of this study was the observation that the relatively hard ceramic sliders wore considerably
more severely than the softer stainless steel slider. Mechanical and chemical analyses have been conducted to try to
explain these results.
4.1. Fracture toughness
The wear resistance of 440C stainless steel, silicon carbide, and alumina pins is in the reverse order as their relative
Vickers hardness numbers, but in the same relative order as
their fracture toughness.
Recognizing that the current work was on sliding wear, the
abrasive wear of ceramics has been proposed to be a function
of both hardness and fracture toughness [10]. Further studies

1352

J. Qu et al. / Wear 258 (2005) 13481356

Fig. 3. SEM images of wear scars on the balls sliding against Ti64 disks.

on ceramic wear mechanisms showed that fracture toughness


may play a dominating role in wear resistance. For example,
Fischer [11] has demonstrated that, in the case of yttria stabilized zirconia ceramics, the wear resistance increases with
the fourth power of fracture toughness.
The pin-on-disk apparatus is mainly intended to evaluate
sliding wear, but in practice, the load history can consist of

sliding and impact, since vibrations may occur if the disk


surface is not perfectly normal to the axis of rotation. The
tendency for impact to occur for small errors in alignment
depends also on the sliding speed and how far the contact
is from the center of rotation. Unlike sliding that usually
causes plastic shearing in materials, impact may introduce
catastrophic failures, such as cracking and crushing of the

J. Qu et al. / Wear 258 (2005) 13481356

1353

Table 4
Friction and wear results for 304 stainless steel disks against metal and
ceramic sliders
Slider material

Friction coefficient

Wear factor (mm3 /N m)


Ball

Disk

440C stainless steel


Silicon nitride
Alumina

0.55 0.05
0.68 0.08
0.57 0.03

<1 106
1.15 106
9.80 106

2.52 104
2.01 104
5.41 104

disk had much lower wear factors than those against the Ti
disks at the same testing condition (see Tables 3 and 4), while
the 304 stainless steel and Ti64 disks had similar hardness and
comparable wear factors. This suggests that there might be
other sources influencing the wear rate.
4.2. Tribochemical reactions

Fig. 4. EDS analysis of the wear scars on the sliders.

contact surfaces, leading to faster material removal and the


production of sharp ceramic debris fragments that can in turn
cause three-body abrasion. Brittle materials, like ceramics,
are more sensitive to such repeated impact effects than are
tougher metals. Rice et al. [12,13] have studied the wear rate
and mechanism of compound impact (impact and sliding)
on metals and superalloys. The material with lower fracture
toughness had a higher wear rate and gave strong evidence
for subsurface damage.
To test the dependency of the wear rate on fracture toughness, comparative tests were conducted on a 304 stainless
steel disk sliding against the 440C stainless steel, silicon nitride, and alumina sliders, under 10 N load and at 0.3 m/s
speed for 500 m. Friction and wear results for the 304 stainless steel disk are shown in Table 4. The alumina, silicon
nitride, and 440C stainless steel balls had the wear rate from
high to low. This confirmed the fracture toughness effect.
However, it has been noticed that the sliders against the steel

It is known that mechanically deformed surfaces usually have different chemical reactivity than purely thermally
stressed solids [14]. Tribochemical reactions may significantly accelerate the wear process. Surface analyses (EDS
and XRD) were conducted on the contact surfaces to explore
the possibility of tribochemical reactions that may induce
the unexpected high wear rates on the ceramic sliders. Fig. 4
shows the EDS spectra of the wear scars on the balls that slid
against the Ti64 disks. The EDS analyses indicated Ti and
Al on the worn surfaces of the steel and alumina balls (see
Fig. 4(a) and (c)), which indicate material transfer from the
Ti64 disk to the sliders. It is interesting to notice that only Al
but no Ti was found on the wear scar of the silicon nitride ball
(see Fig. 4(b)). No Al was observed on the unworn region of
this ball. This may imply that the detected Al was probably
not in metallic form (otherwise Ti should be present too), but
had chemical compounds with other elements, such as Si, O,
and/or N present in Fig. 4(b).
X-ray diffraction was then used to further analyze the wear
scars on the ceramic sliders. A four-axis goniometer [15]
was employed for the grazing incidence (2 ) X-ray diffraction measurements using Cu K radiation and parallel beam
optics. That technique eliminates the sample surface displacement errors due to the spherical shape. Fig. 5 shows
the XRD patterns of the wear scar and debris generated by
silicon nitride against Ti64. Fig. 5(a) reveals that the wear
scar on the silicon nitride ball contains silicon nitride (Si3 N4 )
in both - and -phase and silicon oxide nitride (Si2 N2 O).
Due to the peak superposition, silicon aluminum oxide nitride
(Si5 AlON7 ) cannot be distinguished from Si3 N4 . However,
the presence of Al detected by EDS supports this possibility.
As shown in Fig. 5(b), the XRD analysis on the wear debris
found titanium, silicon nitride, and titanium nitride, but no
indication of titanium oxides. This was a little surprising because titanium oxides have the lower Gibbs free energy of
formation than the titanium nitride in the ambient environment. One possible explanation is that the titanium oxides
were amorphous due to severe plastic deformation and were

1354

J. Qu et al. / Wear 258 (2005) 13481356

Fig. 5. X-ray diffraction analysis of the wear scar and debris for silicon nitride sliding against Ti64.

not detected by XRD. The XRD pattern of the worn surface on the alumina ball sliding against a Ti64 disk is shown
in Fig. 6. The observed spinel (MgAl2 O4 ) was probably a
sintering aid. The XRD pattern may suggest some possibility of forming titaniumaluminum intermetallic compounds
(Al3 Ti, Al2 Ti), but there was no strong evidence. Titanium
aluminides were also suspected by Dong and Bell [5] based

on the XRD analysis of the wear debris produced by alumina


sliding against Ti64.
The high wear rates of alumina and silicon nitride sliders may be attributed to the formation of chemical reaction products between them and the Ti and/or Al transferred
from the Ti64 disks. Such tribochemical reactions were also
aided by the lower thermal conductivity of Ti that promotes

J. Qu et al. / Wear 258 (2005) 13481356

1355

Fig. 6. X-ray diffraction pattern of the wear scar on the alumina slider against the Ti64 disk.

a higher temperature near the interface. These reaction products bonded to the ceramic contact surfaces may deteriorate
their mechanical properties and result in micro fractures leading to high wear. The wear process continuously developed
fresh surfaces and in turn accelerated the tribochemical reactions.

5. Summary
The tribological behavior and responsible wear mechanisms for titanium alloys Ti6Al4V and Ti6Al2Sn
4Zr2Mo, sliding against 440C stainless steel, silicon nitride,
alumina, and PTFE were investigated. The following observations and conclusions were obtained:
(1) The two Ti alloys had similar friction and wear performance, although their grain structures and compositions
are different.
(2) Large frictional fluctuations occurred when metal and
ceramic balls slid against Ti alloy disks, probably caused
by formation and periodic, localized fracture of a transfer
layer.
(3) Higher friction coefficient with larger fluctuation and
higher wear rate were observed at the lower sliding speed.
(4) Despite their higher hardness, ceramic sliders experienced much higher wear and created more wear on the
counterfaces than did the stainless steel sliders.
(5) Fracture toughness and tribochemical reactions have
been proposed to explain the unexpected high wear rates
on the ceramic sliders. Comparative tests on a 304 stain-

less steel disk supported the fracture toughness dependency of the wear rate.
(6) EDS and XRD analyses confirmed material transfer from
the Ti alloy disks to their counterfaces and suggested
possible tribochemical reactions.

Acknowledgements
The authors with to acknowledge with appreciation Y.
Kosaka of Titanium Metals Corporation, USA, for supplying alloy billets along with their chemical analyses. Support
for this research was provided by the U.S. Department of
Energy, Assistant Secretary for Energy Efficiency and Renewable Energy, Office of FreedomCAR and Vehicle Technologies, as part of the High Strength Weight Reduction Materials Program, under contract DE-AC05-00OR22725 with
UT-Battelle, LLC. J. Qu and N. Kulkarni were supported in
part by appointments to the ORNL Postdoctoral Research Associates Program administered jointly by ORNL and ORISE.

References
[1] F.M. Kustas, M.S. Misra, Friction and wear of titanium alloys, in:
P.J. Blau (Ed.), ASM Handbook, Friction, Lubrication, and Wear
Technology, 18, ASM International, 1992, pp. 778784.
[2] R.B. Waterhouse, A. Iwabuchi, The effect of ion implantation on the
fretting wear of four titanium alloys at temperatures up to 600 C, in:
Proceedings of the International Conference on Wear of Materials,
ASME, New York, 1985, pp. 471484.

1356

J. Qu et al. / Wear 258 (2005) 13481356

[3] K.G. Budinski, Tribological properties of titanium alloys, Wear 151


(1991) 203217.
[4] A. Molinari, T.B. Straffelini, T. Bacci, Dry sliding wear mechanisms
of the Ti6Al4V alloy, Wear 208 (1997) 105112.
[5] H. Dong, T. Bell, Tribological behavior of alumina sliding against
Ti6Al4V in unlubricated contact, Wear 225229 (1999) 874884.
[6] EHKT Technologies, Opportunities for low cost titanium in reduced fuel consumption, improved emissions, and enhanced durability heavy-duty vehicles, Oak Ridge National Laboratory Report,
ORNL/Sub/4000013062/1, Oak Ridge, Tennessee, 2002, p. 59.
[7] Ultra-Lite Brakes and Components, Literature, Red Devil Brakes,
Inc., Mt. Pleasant, Pennsylvania, not dated.
[8] S. Lampman, Fatigue and fracture properties of stainless steel, in:
S.R. Lampman (Ed.), ASM Handbook, Fatigue and Fracture, 19,
ASM International, 1996, pp. 712732.
[9] J.M. Barsom, S.T. Rolfe, Correlations between KIC and Charpy Vnotch test results in the transition temperature range, in: Impact

[10]
[11]
[12]
[13]

[14]
[15]

Testing of Materials STP 466, ASTM, Philadelphia, 1979, pp. 281


302.
A.G. Evans, D.B. Marshall, Fundamentals of Friction and Wear of
Materials, ASM, 1980, p. 439.
T.E. Fischer, Friction and wear of ceramics, Scripta Metall. Mater.
24 (1990) 833838.
S.L. Rice, The role of microstructure in the impact wear of two
aluminum alloys, ASME Proc. Wear Mater. (1979) 2734.
S.L. Rice, H. Nowotny, S.F. Wayne, Characteristics of metallic subsurface zones in sliding and impact wear, ASME Proc. Wear Mater.
(1981) 4752.
G. Heinicke, Tribochemistry, Carl Hanser Verlag Munchen Wien,
Berlin, 1984.
H. Krause, A. Haase, X-Ray Diffraction System PTS for Powder,
Texture and Stress Analysis, in: H.J. Bunge (Ed.), Experimental
Techniques of Texture Analysis, vol. 405408, DGM Informationsgesellschaft, Verlag, 1986.

You might also like