Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Wear 258 (2005) 1348–1356

Friction and wear of titanium alloys sliding against metal,


polymer, and ceramic counterfaces夽
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 (Ti–6Al–4V and Ti–6Al–2Sn–4Zr–2Mo). 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 for oxygen results in the formation of an adherent surface ox-


ide, but sub-stoichiometric TiO2 can act as a solid lubricant.
In comparison to light weight alloys based on aluminum A great deal is known about the physical metallurgy,
and magnesium, titanium alloys present interesting possibil- heat treatment, and mechanical properties of titanium alloys,
ities as tribomaterials, but they have not been widely inves- thanks to extensive aerospace-related research and develop-
tigated as bearing materials. They are harder and stiffer than ment. Tribological concerns for Ti in aerospace components
Mg and Al alloys, and they resist exposure to heat and aque- have focused mainly on their fretting behavior, leading to re-
ous corrosion much better. Like Al and Mg, their high affinity search on surface treatments like ion implantation and solid
film lubrication [1,2]. Needs in the chemical process industry
夽 Research sponsored by the U.S. Department of Energy, Assistant Secre- motivated a 1991 study of the galling and sliding wear be-
tary for Energy Efficiency and Renewable Energy, Office of FreedomCAR havior of commercial-purity Ti and alloy Ti–6Al–4V [3]. In
and Vehicle Technologies, as part of the High Strength Weight Reduction that investigation, the best wear and friction results for Ti al-
Materials Program, under contract DE-AC05-00OR22725 with UT-Battelle, loys were obtained for anodized counter-surfaces coated with
LLC.
∗ Corresponding author. Tel.: +1 865 574 4560; fax: +1 865 574 6918. MoS2 solid-film or with polytetrafluoroethylene (PTFE), but
E-mail address: qujn@ornl.gov (J. Qu). the abrasion resistance was poor. Relatively few additional

0043-1648/$ – see front matter © 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.wear.2004.09.062
J. Qu et al. / Wear 258 (2005) 1348–1356 1349

Table 1
Compositions and characteristics of Ti–6Al–4V and Ti–6Al–2Sn–4Zr–2Mo
Ti alloy Compositions (wt%)* (balance Ti) Microindentation hardness HV (GPa) Tensile strength UTS (MPa)
Ti64 6.53 Al, 3.89 V, 0.13 Fe 3.36 ± 0.17 954.9
Ti6242 5.85 Al, 1.98 Sn, 4.22 Zr, 1.95 Mo 3.31 ± 0.10 957.0
∗ Analyses supplied by Titanium Metals Corporation.

studies have been conducted on sliding wear mechanisms of of these alloys, provided by the supplier for these heats of ma-
Ti alloys. Molinari et al. highlighted the mechanisms respon- terial, are listed in Table 1 and their microstructures are shown
sible for the wear resistance under different load and sliding in Fig. 1(a) and (b), respectively. The typical ␣, ␣/␤-phase
speed conditions in self-mated Ti–6Al–4V disk-on-disk slid- grain structure can readily be identified on the etched cross-
ing tests [4], and Dong and Bell [5] reported unexpectedly section of Ti64. Ti6242 has finer grain size and is dominated
high wear rates for alumina sliding against Ti–6Al–4V (pin- by ␣-phase. The two alloys have similar microindentation
on-disk tests). hardness and tensile strength (see Table 1).
Recent developments in Ti processing forecast the avail- Friction and wear tests were conducted using a pin-on-disk
ability of lower-cost Ti and that has prompted further interest apparatus. The diameter of the fixed ball sliders was 9.53 mm.
in exploring the tribological behavior of Ti alloys as bearing As shown in Table 2, 440C stainless steel, silicon nitride, alu-
materials [6]. Focus by the U.S. Department of Energy on im- mina, and polytetrafluoroethylene (PTFE), were selected to
proved brake materials for fuel-efficient heavy trucks, has led represent metallic, ceramic, and polymeric bearing materials.
to the consideration of Ti for disc brake rotors as well. In fact, The Ti alloy disks were 63.5 mm diameter and 12.7 mm thick.
coated Ti brake discs are already showing promise in auto rac- The disk surfaces were polished by 600 grit wet SiC abrasive
ing [7]. This renewed interest in the friction and wear of Ti paper and the pre-test arithmetic average surface roughness
alloys has prompted the current laboratory study of the behav- (Ra ), measured with a Taylor Hobson TalysurfTM 10 stylus
ior of two commercially-available Ti alloys sliding against profilometer with a 2.5 ␮m tip radius, was 0.11 ± 0.02 ␮m.
model metallic, ceramic, and polymeric counterfaces. One Concentric wear tracks ranging from 18 to 52 mm in diameter
of the two alloys (Ti–6Al–4V) has had more tribological at- were used, and the disk rotation rate was adjusted accordingly
tention than the other, but the other (Ti–6Al–2Sn–4Zr–2Mo) to provide either 0.3 or 1.0 m/s sliding speed. A 10 N normal
has attractive elevated temperature properties and was felt load was applied and the test was run for 500 m sliding dis-
to be of interest as well. There are very few studies on the tance. In follow-on experiments, and to provide a comparison
tribological properties of Ti–6Al–2Sn–4Zr–2Mo in the liter- to the results for the Ti alloy disks, a 304 stainless steel disk
ature. In this study, it is intended to establish baseline data for (63.5 mm diameter and 6.35 mm thick) was tested against
these alloys with which to compare the tribological behavior 440C stainless steel and ceramic (Si3 N4 and Al2 O3 ) slid-
of new surface treatments or coatings in future work. ers 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.
2. Materials and testing procedure The friction force was monitored by a load cell-based
force measurement system. The wear volumes of the sliders
Two titanium alloys, Ti–6Al–4V and Ti–6Al–2Sn– and disks were determined by weight change measurements
4Zr–2Mo, were tested in this study. These alloys shall subse- with an accuracy of 0.1 mg. The wear factor is defined as the
quently be referred to as Ti64 and Ti6242. The compositions 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) 1348–1356

Table 2
Characteristics of slider materials
Sliders Supplier Specification Vickers hardness (GPa) Fracture toughness (MPa m1/2 )
440C stainless steel McMaster-Carr Grade 100 Hardened 12.60a 23.7b
Silicon nitride Cerbec East Granby, CT NBD200 Grade 5 19.37a 5.2c
Alumina Southern Bearing Service AFBMA Grade 25 24.75a 3–4c
PTFE W.M. Berg, Inc. East Rockaway, NY – N/A N/A
a The Vickers hardness was measured using a 100 g load.
b The Izod impact strength of hardened 440C stainless steel is 4 ft lb [8]. The fracture toughness here was estimated based on Barsom–Rolfe’s 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.

distance of the pin. All the tests were conducted in ambient large fluctuation of the friction coefficient was thought to
air conditions with temperature and humidity in the range be caused by formation and periodic, localized fracture of
of 18–22 ◦ C and 52–62%, respectively. At least two dupli- a transfer layer. Titanium alloy commonly transfers to the
cates were run at each test condition. Good repeatability was counterface when rubbing against other metals or ceramics
obtained in both friction and wear results. [3–5]. 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 stain-
3. Results less steel ball (see Fig. 3(a)). The energy dispersive spec-
troscopy (EDS) analysis detected Ti and/or Al on the worn
Results are summarized in Table 3(a) and (b) for Ti64 and surfaces of the metal and ceramic balls, as shown in Fig. 4.
Ti6242, respectively. Friction and wear results are presented More discussion on surface analysis can be found on Section
separately. 4.2.
For metal and ceramic balls, lower friction coefficient and
3.1. Friction 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
Table 3(a) and (b) present the average friction coeffi- sliding speed, the contact area had higher temperature, which
cient and its fluctuation at steady-state for each test con- generally reduced the shear strength and led to lower friction
dition. Selected friction traces of the four different sliders forces.
against Ti64 disks are shown in Fig. 2. The PTFE slider gen-
erated a fairly smooth friction trace (Fig. 2(d)) due to its 3.2. Wear
self-lubricating nature. The metal and ceramic sliders pro-
duced friction coefficient in the range of 0.34–0.50 with rel- As shown in Table 3(a) and (b), up to five times higher wear
atively large fluctuation, as illustrated in Fig. 2(a)–(c). The 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 1.0 m/s

Friction coefficient Wear factor (mm3 /N m) Friction coefficient Wear factor (mm3 /N m)

Ball Disk Ball Disk


(a) Ti–6Al–4V disks
440C stainless steel 0.50 ± 0.05 6.9 × 10−6 1.7 × 10−4 0.35 ± 0.05 1.6 × 10−6 1.5 × 10−4
Silicon Nitride 0.47 ± 0.07 3.8 × 10−5 3.5 × 10−4 0.36 ± 0.07 6.2 × 10−6 1.3 × 10−4
Alumina 0.49 ± 0.07 5.7 × 10−5 5.7 × 10−4 0.44 ± 0.07 1.6 × 10−5 2.0 × 10−4
PTFE 0.28 ± 0.001 8.4 × 10−4 N/Ma 0.29 ± 0.001 6.1 × 10−4 N/Ma
(b) Ti–6Al–2Sn–4Zr–2Mo disks
440C stainless steel 0.48 ± 0.05 5.1 × 10−6 1.3 × 10−4 0.34 ± 0.04 1.22 × 10−6 1.1 × 10−4
Silicon nitride 0.47 ± 0.08 4.4 × 10-5 3.5 × 10−4 0.37 ± 0.02 9.40 × 10-6 1.1 × 10−4
Alumina 0.49 ± 0.08 1.2 × 10−4 3.4 × 10−4 0.42 ± 0.04 2.33 × 10−5 2.2 × 10−4
PTFE 0.27 ± 0.001 9.9 × 10−4 N/Ma 0.29 ± 0.001 7.92 × 10−4 N/Ma
a N/M, not measurable.
J. Qu et al. / Wear 258 (2005) 1348–1356 1351

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

than those at 1.0 m/s. Similar sliding speed dependency was tected by the polymeric layer transferred from the PTFE
also reported by other researchers [5]. ball and had almost no surface damage except a few shallow
The Ti disks suffered high wear rates, in the order of circular groves ground by third body particles, probably
10−4 mm3 /N m, against the metal and ceramic balls, and some metal debris, embedded in the PTFE ball.
harder sliders generated relatively more (or at least compara- It has been seen that these two Ti alloys showed simi-
ble) wear on the Ti disks. Although harder balls were expected lar friction and wear behavior. Therefore, discussion will be
to have higher wear resistance, the results in Table 3(a) and focused on Ti64 only.
(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 4. Discussion
were at least five times higher than those of the steel balls.
Dong and Bell also reported a higher wear rate of an alumina The most unusual finding of this study was the observa-
ball than that of a steel ball when sliding against a Ti64 disk tion that the relatively hard ceramic sliders wore considerably
[5]. More analysis and discussion are presented in Section 4. more severely than the softer stainless steel slider. Mechan-
Fig. 3 shows the wear scars on the metal and ceramic ical and chemical analyses have been conducted to try to
balls with features of abrasive wear, adhesive wear, and plas- explain these results.
tic deformation. Abrasive wear seemed to dominate the wear
process at 0.3 m/s. Those wear scars were larger and flatter, 4.1. Fracture toughness
corresponding to their higher wear factors. The wear scars
generated at 1.0 m/s were smaller but much rougher with The wear resistance of 440C stainless steel, silicon car-
larger patches of transferred material implying more severe bide, and alumina pins is in the reverse order as their relative
adhesive wear, possibly due to higher temperature at the con- Vickers hardness numbers, but in the same relative order as
tact area. EDS analysis also showed higher Ti and/or Al con- their fracture toughness.
centration on the wear scars at 1.0 m/s. Recognizing that the current work was on sliding wear, the
The PTFE slider had the highest wear factor abrasive wear of ceramics has been proposed to be a function
(10−3 mm3 /N m). Its counterface (Ti disk) was pro- of both hardness and fracture toughness [10]. Further studies
1352 J. Qu et al. / Wear 258 (2005) 1348–1356

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

on ceramic wear mechanisms showed that fracture toughness sliding and impact, since vibrations may occur if the disk
may play a dominating role in wear resistance. For example, surface is not perfectly normal to the axis of rotation. The
Fischer [11] has demonstrated that, in the case of yttria sta- tendency for impact to occur for small errors in alignment
bilized zirconia ceramics, the wear resistance increases with depends also on the sliding speed and how far the contact
the fourth power of fracture toughness. is from the center of rotation. Unlike sliding that usually
The pin-on-disk apparatus is mainly intended to evaluate causes plastic shearing in materials, impact may introduce
sliding wear, but in practice, the load history can consist of catastrophic failures, such as cracking and crushing of the
J. Qu et al. / Wear 258 (2005) 1348–1356 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 0.55 ± 0.05 <1 × 10−6 2.52 × 10−4
Silicon nitride 0.68 ± 0.08 1.15 × 10−6 2.01 × 10−4
Alumina 0.57 ± 0.03 9.80 × 10−6 5.41 × 10−4

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

It is known that mechanically deformed surfaces usu-


ally have different chemical reactivity than purely thermally
stressed solids [14]. Tribochemical reactions may signifi-
cantly 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).
Fig. 4. EDS analysis of the wear scars on the sliders. X-ray diffraction was then used to further analyze the wear
scars on the ceramic sliders. A four-axis goniometer [15]
contact surfaces, leading to faster material removal and the was employed for the grazing incidence (2◦ ) X-ray diffrac-
production of sharp ceramic debris fragments that can in turn tion measurements using Cu K␣ radiation and parallel beam
cause three-body abrasion. Brittle materials, like ceramics, optics. That technique eliminates the sample surface dis-
are more sensitive to such repeated impact effects than are placement errors due to the spherical shape. Fig. 5 shows
tougher metals. Rice et al. [12,13] have studied the wear rate the XRD patterns of the wear scar and debris generated by
and mechanism of compound impact (impact and sliding) silicon nitride against Ti64. Fig. 5(a) reveals that the wear
on metals and superalloys. The material with lower fracture scar on the silicon nitride ball contains silicon nitride (Si3 N4 )
toughness had a higher wear rate and gave strong evidence in both ␣- and ␤-phase and silicon oxide nitride (Si2 N2 O).
for subsurface damage. Due to the peak superposition, silicon aluminum oxide nitride
To test the dependency of the wear rate on fracture tough- (Si5 AlON7 ) cannot be distinguished from Si3 N4 . However,
ness, comparative tests were conducted on a 304 stainless the presence of Al detected by EDS supports this possibility.
steel disk sliding against the 440C stainless steel, silicon ni- As shown in Fig. 5(b), the XRD analysis on the wear debris
tride, and alumina sliders, under 10 N load and at 0.3 m/s found titanium, silicon nitride, and titanium nitride, but no
speed for 500 m. Friction and wear results for the 304 stain- indication of titanium oxides. This was a little surprising be-
less steel disk are shown in Table 4. The alumina, silicon cause titanium oxides have the lower Gibbs free energy of
nitride, and 440C stainless steel balls had the wear rate from formation than the titanium nitride in the ambient environ-
high to low. This confirmed the fracture toughness effect. ment. One possible explanation is that the titanium oxides
However, it has been noticed that the sliders against the steel were amorphous due to severe plastic deformation and were
1354 J. Qu et al. / Wear 258 (2005) 1348–1356

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 sur- on the XRD analysis of the wear debris produced by alumina
face on the alumina ball sliding against a Ti64 disk is shown sliding against Ti64.
in Fig. 6. The observed spinel (MgAl2 O4 ) was probably a The high wear rates of alumina and silicon nitride slid-
sintering aid. The XRD pattern may suggest some possibil- ers may be attributed to the formation of chemical reac-
ity of forming titanium–aluminum intermetallic compounds tion products between them and the Ti and/or Al transferred
(Al3 Ti, Al2 Ti), but there was no strong evidence. Titanium from the Ti64 disks. Such tribochemical reactions were also
aluminides were also suspected by Dong and Bell [5] based aided by the lower thermal conductivity of Ti that promotes
J. Qu et al. / Wear 258 (2005) 1348–1356 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 prod- less steel disk supported the fracture toughness depen-
ucts bonded to the ceramic contact surfaces may deteriorate dency of the wear rate.
their mechanical properties and result in micro fractures lead- (6) EDS and XRD analyses confirmed material transfer from
ing to high wear. The wear process continuously developed the Ti alloy disks to their counterfaces and suggested
“fresh surfaces” and in turn accelerated the tribochemical re- possible tribochemical reactions.
actions.

Acknowledgements
5. Summary
The authors with to acknowledge with appreciation Y.
The tribological behavior and responsible wear mech- Kosaka of Titanium Metals Corporation, USA, for supply-
anisms for titanium alloys Ti–6Al–4V and Ti–6Al–2Sn– ing alloy billets along with their chemical analyses. Support
4Zr–2Mo, sliding against 440C stainless steel, silicon nitride, for this research was provided by the U.S. Department of
alumina, and PTFE were investigated. The following obser- Energy, Assistant Secretary for Energy Efficiency and Re-
vations and conclusions were obtained: newable Energy, Office of FreedomCAR and Vehicle Tech-
(1) The two Ti alloys had similar friction and wear perfor- nologies, as part of the High Strength Weight Reduction Ma-
mance, although their grain structures and compositions terials Program, under contract DE-AC05-00OR22725 with
are different. UT-Battelle, LLC. J. Qu and N. Kulkarni were supported in
(2) Large frictional fluctuations occurred when metal and part by appointments to the ORNL Postdoctoral Research As-
ceramic balls slid against Ti alloy disks, probably caused sociates Program administered jointly by ORNL and ORISE.
by formation and periodic, localized fracture of a transfer
layer.
(3) Higher friction coefficient with larger fluctuation and References
higher wear rate were observed at the lower sliding speed.
(4) Despite their higher hardness, ceramic sliders experi- [1] F.M. Kustas, M.S. Misra, Friction and wear of titanium alloys, in:
enced much higher wear and created more wear on the P.J. Blau (Ed.), ASM Handbook, Friction, Lubrication, and Wear
counterfaces than did the stainless steel sliders. Technology, 18, ASM International, 1992, pp. 778–784.
[2] R.B. Waterhouse, A. Iwabuchi, The effect of ion implantation on the
(5) Fracture toughness and tribochemical reactions have fretting wear of four titanium alloys at temperatures up to 600 ◦ C, in:
been proposed to explain the unexpected high wear rates Proceedings of the International Conference on Wear of Materials,
on the ceramic sliders. Comparative tests on a 304 stain- ASME, New York, 1985, pp. 471–484.
1356 J. Qu et al. / Wear 258 (2005) 1348–1356

[3] K.G. Budinski, Tribological properties of titanium alloys, Wear 151 Testing of Materials STP 466, ASTM, Philadelphia, 1979, pp. 281–
(1991) 203–217. 302.
[4] A. Molinari, T.B. Straffelini, T. Bacci, Dry sliding wear mechanisms [10] A.G. Evans, D.B. Marshall, Fundamentals of Friction and Wear of
of the Ti6Al4V alloy, Wear 208 (1997) 105–112. Materials, ASM, 1980, p. 439.
[5] H. Dong, T. Bell, Tribological behavior of alumina sliding against [11] T.E. Fischer, Friction and wear of ceramics, Scripta Metall. Mater.
Ti6Al4V in unlubricated contact, Wear 225–229 (1999) 874–884. 24 (1990) 833–838.
[6] EHKT Technologies, Opportunities for low cost titanium in re- [12] S.L. Rice, The role of microstructure in the impact wear of two
duced fuel consumption, improved emissions, and enhanced dura- aluminum alloys, ASME Proc. Wear Mater. (1979) 27–34.
bility heavy-duty vehicles, Oak Ridge National Laboratory Report, [13] S.L. Rice, H. Nowotny, S.F. Wayne, Characteristics of metallic sub-
ORNL/Sub/4000013062/1, Oak Ridge, Tennessee, 2002, p. 59. surface zones in sliding and impact wear, ASME Proc. Wear Mater.
[7] Ultra-Lite Brakes and Components, Literature, Red Devil Brakes, (1981) 47–52.
Inc., Mt. Pleasant, Pennsylvania, not dated. [14] G. Heinicke, Tribochemistry, Carl Hanser Verlag Munchen Wien,
[8] S. Lampman, Fatigue and fracture properties of stainless steel, in: Berlin, 1984.
S.R. Lampman (Ed.), ASM Handbook, Fatigue and Fracture, 19, [15] H. Krause, A. Haase, X-Ray Diffraction System PTS for Powder,
ASM International, 1996, pp. 712–732. Texture and Stress Analysis, in: H.J. Bunge (Ed.), Experimental
[9] J.M. Barsom, S.T. Rolfe, Correlations between KIC and Charpy V- Techniques of Texture Analysis, vol. 405–408, DGM Informations-
notch test results in the transition temperature range, in: Impact gesellschaft, Verlag, 1986.

You might also like