Vous êtes sur la page 1sur 8

Wear 390391 (2017) 4148

Contents lists available at ScienceDirect

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

Identication of parameters inuencing the glaze layer formation and MARK


stability at high temperature for a Waspaloy/Ren125 contact under fretting
wear conditions

F. Alkelae , S. Fouvry
Laboratoire de Tribologie et Dynamique des Systmes, UMR 5513, Ecole Centrale de Lyon, 36 avenue Guy de Collongue, 69134 Ecully cedex, France

A R T I C L E I N F O A B S T R A C T

Keywords: In the last decades, industrial investigations were mostly interested in developing new coatings to improve
Fretting wear materials endurance at high temperature. In this study, our interest will be focused on knowing what happens to
High temperature basic materials. Using ball-on-at conguration, with Waspaloy as a counterbody and Ren125 as the sample,
Nickel based alloys results show that, when submitted to fretting wear solicitations, a glaze layer is developed within the contact,
Glaze layer
which allows an important decrease simultaneously in the friction and the wear rate.
Oxides formation
To better understand this glaze layer behavior, many parameters were studied, such as test lifetime, sliding
amplitude, temperature, load and frequency. In the aim to determine the eciency domain of the glaze layer, a
mapping based on normalized sliding velocity is proposed. This map will gather all operating parameters that
control wear behavior of our tribosystem, it plots wear rate versus sliding velocity, it distinguishes formation/
degradation limits of this glaze layer.
Increasing the test lifetime does not aect outstandingly the progress of the glaze layer, which is reected in
the friction coecient and the wear evolutions, but it is worthy to notice that this layer is formed and stabilized
once the temperature is higher than 400 C , and it is governed by sliding amplitude, frequency as well as load at
high temperature. A synthesis curve is presented indicating bilinear trend of wear rate depending on operating
parameters, highlighting for each case dissipated energy needed to reach the stable regime of glaze layer de-
velopment.

1. Introduction structure. The spinel ferrous oxide Fe3 O4 which can be easily sheared
o becomes more thermodynamically stable above 200 C than the hard
As known at high temperature working, many destructive phe- rhombohedral Fe2 O3, which has an abrasive character. Stott and co-
nomena may be encountered like tribocorrosion and oxidative wear. workers [1214] related this reduction on friction and wear to the
The former occurs especially when working in lubricated conditions. It formation of protective oxide layers at high temperature, called the
is detected for applications like automotive [1], oil and gas industries glaze layer. Jiang et al. [15,16] found that strong adhesion occurs
[2], for pressurized water reactors [3] and also when using ionic liquids between ne debris especially at high temperature, which enables fast
[4]. The later leads to premature failure of materials and reduces their establishment of protective oxide layers. To this, we add investigations
lifetime endurance, as reported by [5] for cutting tools, or by Water- conducted by Kato and Komei [17,18] about the diusion character of
house for electrical contacts [6]. That is why it is important to develop articially oxide particles supplied, if these oxides are highly diusive,
coatings more resistant to high temperature wear damages. These are transition from severe to mild wear is observed. Conversely, no tran-
aspects of detrimental oxidative wear eect, in this study, benecial sition is observed when using poorly diusive oxides. Rybiak et al. [19]
oxidative wear will be tackled. The friction and wear responses of steel investigated Jethete M152 stainless steel behavior fretted against aus-
[7], titanium [8] and especially nickel alloys [9] under high tempera- tenitic steel at high temperature, whose results are in good agreement
ture and dry conditions, in many assemblies like turbine engine struc- with Kato's theory [18]. At a threshold temperature (220 C), a sig-
tures, have been widely investigated. While using stainless steel mate- nicant decrease on both wear and friction is noticed due to formation
rials, a sharp decrease in wear and friction is noted after 200 C [10,11]. of a stabilized glaze layer formed at the interface. The objective of
According to Hurricks, this behavior is due to transformation in oxide this research work is to investigate the stability of the glaze layers


Corresponding author.
E-mail address: alkelae.fathia@gmail.com (F. Alkelae).

http://dx.doi.org/10.1016/j.wear.2017.07.008
Received 5 April 2017; Received in revised form 4 July 2017; Accepted 7 July 2017
Available online 08 July 2017
0043-1648/ 2017 Elsevier B.V. All rights reserved.
F. Alkelae, S. Fouvry Wear 390391 (2017) 4148

Table 1 800 C , and the displacement range was increased so that the dis-
Chemical composition of alloys used in this study. placement amplitude can be imposed from 5 m to 200 m. Test
apparatus was optimized to investigate very high fretting frequency.
Co% Cr% W% C% Mo% Ni% Fe% Ti% Al% Ta%
Initially limited at 13 Hz, the test system used in this investigation
Waspaloy 13.5 19 4.3 56.7 3 1.5 showed that the apparatus allows high frequencies up to 150 Hz
Ren 125 10 8.5 8 0.11 2 59 0 2.5 4.8 3.8 without inducing signicant dynamical perturbation in the fretting
cycle. This investigation has been done monitoring the sliding ampli-
tude g dened value Q* = 0, and therefore avoiding the test tan-
formed with Waspaloy (ball) /Ren 125 (at) (Nickel based alloys)
gential accommodation. (i.e g = C.Q = 0 if Q = 0). In this study, the
under fretting wear solicitations at 700 C, as a function of the applied
eect of temperature, sliding amplitude, applied load and frequency
sliding amplitude, frequency, applied load and test duration. Results
will be investigated. To conduct these experiments, a conguration of
obtained will be used to establish an eciency domain mapping of
ball-on-at is used, with a ball radius of 12.7 mm.
those materials in service conditions, followed by a general summary of
the glaze layer stability versus the energy dissipated at the interface.
3. Tribological response of Waspaloy/Ren 125 contact as a
function of test conditions
2. Experimental procedure
3.1. Temperature eect
2.1. Materials
The eect of temperature was investigated on Waspaloy/Ren 125
Nickel based superalloys are known to have an exceptional combi-
tribosystem. A range of temperature from 25 to 700 C was applied
nation of high temperature strength, toughness, and resistance to de-
keeping all other operating parameters constant (P = 67 N, f = 25 Hz,
gradation in corrosive or oxidizing environments [20], they are widely
Nc = 500 000 cycles, g = 100 m). Tests were conducted using
used in aircraft engines, and hot end components of various types of gas
ball-on-at conguration. After cleaning both counterbodies in an ul-
turbines.
trasonic bath with ethanol for 15 min, in order to remove debris
For the Crankcase/Turbine Low Pressure contact, materials used
trapped in the wear scars, 3D surface proles were performed using a
are: Waspaloy which represents the ball counterbody, with surface
prolometer to quantify the wear volume. For both friction coecient
roughness (Ra) of 1.1 m, against the at specimen of Ren 125 with
and total wear volume evolutions (Fig. 2), we can note the presence of
surface roughness (Ra) of 1.04 m. Both are rich in chromium and
three wear regimes:
cobalt as shown in Table 1. Alloying elements play a potent role in
dening materials behavior under given conditions. For nickel alloys,
I. From room temperature up to 200 C, severe wear regime char-
elements such as Ni, Co, Cr, Mo, W form the FCC phase (named ), and
acterized by high wear rate and high friction coecients,
thereby stabilize it. Ti, Ta, and Al promote the formation of ordered
II. From 200 C to 400 C, Transition from severe to mild wear regime
phases named that reside in the phase. For Ren 125, the presence
characterized by a sharp decrease of wear rate and medium friction
of carbon may promote the formation of carbides by combing carbon
coecient,
with reactive elements including Ta, Ti, Al, W, Mo and Cr, which have
III. From 400 C up to 700 C, a stabilized friction coecient around 0,4
an eective role on the rupture strength by inhibiting the grain-
is observed with almost no wear, which characterizes mild wear
boundary sliding [21].
regime.

2.2. Experimental layout Investigations on steel and iron based materials [22,23] attributed
this decrease of friction and wear with increasing temperature to
The fretting wear devicer used is composed of a mobile part acti- transformation of oxides formed, whereas below 200 C, Fe2 O3 is the
vated by an electromagnetic shaker (Fig. 1a), and a xed part on which main constituent in the wear debris. Above this temperature, the
is clamped the plane specimen. During the test, the normal force P is amount of Fe3 O4 becomes more signicant. Person and Shipway [24]
kept constant while the tangential force Q and displacement are re- investigated the eect of temperature on a high strength steel (Super-
corded, which enables online plotting of the fretting loop Q - (Fig. 1b). CMV), they found that increasing temperature gave dierent micro-
More details about the characteristics of this equipment at room tem- structures to the wear scar, starting with 24 C, where debris exhibited
perature can be found in previous studies [19]. roughened appearance with large acks of oxide, at 85 C, a total
The fretting setup was improved by Rybiak et al. investigations in coverage of the wear scar with compacted oxides smoother in appear-
many aspects. Hence the temperature capacity was increased up to ance, simultaneously, in other areas delamination and breakdown still

Fig. 1. Contact conguration.

42
F. Alkelae, S. Fouvry Wear 390391 (2017) 4148

Fig. 2. Tribological parameters versus temperature.

occurred, at 120 C the smoothness of the wear scar improved. In- 3.2. Test duration eect
creasing the temperature to 150 C , the surface appeared glassy and
burnished giving rise to the glaze layer establishment and evolution To follow the glaze layer behavior as a function of test duration,
with increasing temperature. Thereby reducing wear rate and friction tests were conducted keeping same operating parameters with in-
under fretting conditions. Investigations on nickel based alloys (Ni- creasing progressively number of cycles for every test [P = 67 N, g
monic 75, Nimonic C263, Nimonic 108 and Incoloy 901) [25], showed = 100 m, f = 25 Hz, T = 700 C ].
that there is a temperature value above which reduction in friction and The evolution of friction coecient (Fig. 3c) indicates that once the
wear takes place. This threshold temperature is 600 C and it whitnesses contact is activated, a high friction value is reached ( = 0.9), and after
a change of oxides nature formed. Oxides hence formed contain almost 3000 cycles, an abrupt decrease is observed, which can be due to the
all the alloying elements of the contacting bodies. formation of a lubricant oxide lm at the interface. For all tests, the
During sliding at 400 C and 800 C , tests conducted versus tem- friction was stabilized at = 0.4.
perature and versus chromium content on nickel - chromium alloys (in Fig. 3a shows that while the wear rate of Waspaloy increases,
the range 540% Cr) [26], have shown that, for a given alloy, an ef- Ren125 wear rate decreases, that means that wear by transfer occurs.
fective glaze layer is formed at higher temperatures, and for a given Debris generated from Waspaloy are the basis of the glaze layer formed
temperature, it is formed for a lower chromium-content alloy. on Ren 125, prohibiting the metal-metal contact and leading to a sharp
Bill [27] performed fretting wear tests on pure Ni and Ni - Cr - Al decrease on friction and wear.
alloys (A, B, C and D) at a range of temperatures up to 800 C , with: The total wear rate follows a zigzag evolution versus test duration. A
progressive rise up to 1.5. 106 cycles, followed by linear decrease until
A: Ni - 20% Cr - 5% Al 3.5. 106 cycles then a new linear increase up to 5. 106 cycles. The rst
B: Ni - 20% Cr - 2% Al (similar to Waspaloy) maximum wear volume corresponds to almost no wear on Ren
C: Ni - 10% Cr - 5% Al (similar to Ren 125) 125 + 4. 103 mm3 wear of Waspaloy, the minimal wear volume ob-
D: Ni - 10% Cr - 2% Al tained corresponds to a maximum transfer to the Ren 125 surface
2. 103 mm3 + 3. 103 mm3 wear of Waspaloy, and the second max-
Results showed that from room temperature up to 540 C , increasing imum wear volume corresponds to a maximum wear volume of
chromium content from 10% to 20% is eective in reducing fretting Waspaloy 5. 103 mm3 + an important transfer to Ren 125 surface
wear damage, and increasing aluminium content from 2% to 5% ef- 1.8. 103 mm 3 .
fectively reduced the fretting wear at temperatures up to 816 C. A and These results show that the responsible for the tribosystem wear is
B alloys showed more fretting wear at 650 C than at 540 C , conversely, the Waspaloy counterbody. The second and third tests show that debris
C and D alloys showed less fretting wear at same temperature values. generated from wear of Waspaloy are not retained in the interface. This
These results assume that higher chromium concentration accelerates can be interpreted by gross particles production that could not be
surface fatigue damage by chromium diusion to crack sites [27]. Be- fragmented, comminuted and compacted in order to form a protective
sides, formation of aluminium oxides such as Al2 O3 or NiAl2 O4 in- oxide layer, since just ne particles are able to be trapped in the in-
creases the materials fretting wear resistance. These oxides can form at terface. These gross particles are ejected leading to a maximum wear
high temperatures with 10% Cr 5% Al content, but not with higher rate.
chromium content (20% Cr), which explains in our case, the higher
wear resistance of Ren 125 in comparison to Waspaloy.
In our case, at 200 C, unstable thermodynamically oxides start to 3.3. Applied load eect
form (NiCr2 O4 , Ni2 O3 , CoO), with increase of temperature, especially
after 400 C, these oxides tend to transform to more stable oxides The real area of contact in sliding wear, is a very small fraction
(NiO, Cr2 O3 , Co3 O4 ) [28], which allows formation of protective lu- compared to the apparent area. Under load eect, contact occurs at the
bricant oxide layer, thus, decreasing wear and friction. level of asperities that break down forming wear debris, these frag-
mented asperities, under cyclic solicitations, will undergo plastic

43
F. Alkelae, S. Fouvry Wear 390391 (2017) 4148

Fig. 3. Evolution of wear volume as a function of test duration.

deformation as described by Suh [29] in the theory of delamination.


In the aim to elucidate the load eect on wear behavior of materials,
a series of tests was conducted with varying the applied load from 30 to
80 N as shown in Fig. 4. Linear evolution for both materials was ob-
tained. Wear of Waspaloy is twice more important than that of Ren
125.
Fig. 4b indicates an increase of total wear volume with applied load.
Once the glaze layer is established in the interface, it acts as a stick-slip
zone, thus reducing friction and wear. When load increases, wear scar
radius increases while maintaining same slip regime (Fig. 5).
Abhishek [30] has evocated the glaze layer stability depending on
the wear behavior response, using non oxide ceramic and non oxide
particulate reinforced composite against Steel. For oxidative wear, with
a stable oxide layer, the wear volume increases exponentially as load
increases, while for tribomechanical wear, it increases linearly with the
load, which give rise to unstable glaze layer in the interface. In our case,
the wear volume increases linearly with increasing load, which means
that we have interminably a metal-metal contact regime, resulting on
the fracture of oxide layers already formed, which puts us into the
fracture wear regime. Fig. 5. Evolution of wear scar radius with applied load.

Fig. 4. Evolution of wear volume versus applied load.

44
F. Alkelae, S. Fouvry Wear 390391 (2017) 4148

Fig. 6. Evolution of wear volume as a function of sliding amplitude.

3.4. Sliding amplitude eect keeping other operating parameters constant [P = 67 N, g


= 100 m, Nc = 106 cycles, T = 700 C]. Fig. 9 plots the wear rate
The eect of variation of sliding amplitude on wear behavior is evolution of Waspaloy and Ren 125 versus the frequency. Progressive
plotted in Fig. 6, keeping all other operating parameters constant [P = increase of Waspaloy wear rate occurs with increasing the frequency.
67 N, Nc = 106 cycles, f = 100 Hz, T = 700 C]. A change in slope is On the other hand, Ren 125 follows no tendency, that is interpreted by
observed for both materials at g = 175 m, which indicates tran- continuous process of formation/degradation of the glaze layer. Wear
sition from mild to severe wear. If we refer to Figs. 7 and 8, we note an rate is 4 times more signicant at 100 Hz than at 25 Hz, which is related
increase of sliding amplitude accompanied with extension of the worn as Stott evocated [31], to an eect of the glaze layer thickness. While a
area, resulting of formation and ejection of progressively bigger parti- thickness threshold is reached, the glaze layer is destructed and wear
cles that can not be retained in the interface anymore. We can assume debris generated are fastly ejected as abrasive particles due to lack of
that only ne particles (g = 75 m to 120 m) can easily be time to be fragmented and compacted. Metal-metal contact is almost
sintered, oxidized and compacted leading to formation of lubricant always activated, prohibiting the glaze layer formation and leading to
oxide layers [18]. more wear and high friction.
To this conclusion, we add the glaze layer thickness eect shown in
Fig. 8, until 120 m of sliding amplitude, we note progressive growth of 4. Mapping of the Glaze layer eciency domain
the glaze layer thickness on the surface of Ren 125, due to wear by
transfer, which we can note also on Fig. 6a (indicated by negative wear A mapping of the glaze layer eciency is proposed, based on results
rate on Ren 125 surface), after this threshold, we note an eective above and considering the normalized sliding velocity parameter.
wear on both surfaces (Ren 125 and Waspaloy).
Debris ejected from Waspaloy are mostly deposited on Ren 125 V
Vnorm =
surface until glaze layer thickness of 35 m, beyond this value, the Vref . (1)
glaze layer is destructed by fatigue eect leading to ejection of gross
particles that can not be trapped in the interface, metal-metal contact is Vref = 4. fref . gref . =40 mm / s (2)
activated, leading to widening of the contact area, hence, more wear
This approach considers operating parameters such as the sliding
with increasing the sliding amplitude.
amplitude, the frequency, and other sets of experiments with 3 dierent
applied load values (34 N, 67 N and 100 N). All these tests were con-
3.5. Frequency eect ducted at 700 C (service temperature) and 106 cycles. the change of
slope is observed at Vthreshold = 72 mm/ , which is believed to be the limit
In order to investigate the frequency eect on fretting wear beha- of the glaze layer stability domain. At this value of normalized sliding
vior of our materials, tests were conducted at several frequencies while velocity, the total wear volume is 0.05 mm3 (mild wear regime), which

Fig. 7. Sliding amplitude eect on particles size ejected.

45
F. Alkelae, S. Fouvry Wear 390391 (2017) 4148

Fig. 8. Glaze layer thickness as a function of sliding amplitude.

Fig. 9. Evolution of wear volume versus the frequency.

indicates the presence of a stable glaze layer in the interface. Exceeding and sintered, allowing a stable glaze layer regime establishment.
this threshold value, a sharp increase of wear is observed, leading to According to this result, and referring to sliding amplitude section,
transition to severe wear, we can assume that the glaze layer wear re- we conrm that the responsible for wear transition from wild to severe
sistance is weakened, the highly loading conditions inhibit the forma- oxidational regime, is the fragmentation-compaction eect, assumed by
tion of an eective glaze layer, due to formation of gross wear particles, high loads and high sliding amplitudes imposed. Tu et al. [32] in-
that cannot be compacted and sintered to form a protective layer in the vestigated temperature eect on improving wear resistance of 5
interface, hence ejected, leading to abrasive action. CrNiMo steel against 40 MnB steel, under unlubricated highly loaded
Fig. 10 shows that tests conducted under high loads (100 N), kept sliding wear condition. It was shown that at low temperature, there are
materials behavior under mild oxidational wear regime. This suggests no oxide layers transferred on the steel, severe abrasive wear occurred
that fragmentation of gross particles is assumed due to high compaction indicated by grooves on the wear scars. Conversely, at high tempera-
action. Debris generated were relatively easier to be destroyed, oxidized ture, a compact oxide transfer layer was formed on the worn surface,
improving considerably friction and wear resistance of materials. In our
study, all tests were performed at 700 C, which means that compaction
eect is only due to high load and sliding amplitude parameters.
At low sliding velocity, surfaces have enough time to interact with
environment, which enables enough oxidation to contacting surfaces,
this interaction time is reduced at high sliding velocity [33], which
means that oxygen diusion to the interface is limited, this time, me-
tallic debris that could not be oxidized are also present in the interface,
thus a metal-metal contact is activated, leading to increase of wear rate.

5. Discussion

In this section, the impact of the energy dissipated at the interface


on wear rate behavior will be highlighted, based on results above.
Fig. 11 plots total wear volume versus test duration of three sets of
experiments (all at 700 C and 67 N). The rst set is conducted con-
sidering g = 100 m/f = 25 Hz , the second g =
100 m/f = 100 Hz , and the third g = 150 m/f = 100 Hz . VGL is
debris volume needed to form an eective stable glaze layer under
given conditions, its stabilization is function of operating parameters.
Fig. 10. Mapping of the Glaze layer eciency domain. The gure compares their energy dissipated at the interface:

46
F. Alkelae, S. Fouvry Wear 390391 (2017) 4148

required (VGL ) to get to the mild wear regime (Fig. 11). Once reached, if
operating parameters are always the same, mild wear persists, other-
wise, if sliding conditions become aggressive by increasing one of op-
erating parameters, transition to severe wear occurs. The process of
producing more wear debris until reaching the new VGL corresponding
to new operating parameters is activated, as schematized on Fig. 12,
then, another parameter takes control on wear behavior, as highlighted
by [31], if the glaze layer reaches its critical thickness, it breaks down
into wear debris, this process is repeated on the whole contact area,
leading to continuous formation/degradation of the oxide lm layers.

6. Conclusion

In the light of this study, the impact of the glaze layer development
and stabilization on wear behavior of Waspaloy/Ren125 is in-
vestigated, and the following conclusions can be drawn:
Fig. 11. Relation between dissipated energy and wear volume needed for a stabilized
glaze layer.
1. The fretting wear behavior of nickel based alloys (Waspaloy/
Ren125) depends greatly on temperature. Sharp decrease starts at
200 C and almost no wear beyond 400 C, indicating a stable glaze
layer development at the interface.
2. Increasing operating parameters at 700 C (load, sliding amplitude,
frequency and test duration) increases wear rate, with major inu-
ence of sliding amplitude and load, while remaining in the mild
wear regime.
3. The glaze layer stability domain may be demarcated by applying the
normalized sliding velocity approach. A stable glaze layer is main-
tained in the interface until exceeding Vthreshold = 72 mm/s , beyond
this threshold value, it breaks down leading to increase in wear rate.
4. The energy dissipated at the interface shows a strong dependence of
the volume needed to form a stable glaze layer (VGL ) on operating
parameters. The higher the loading conditions, the larger the debris
volume needed to reach the mild wear regime is.
5. The key parameter of monitoring the glaze layer once stabilized is
its thickness, even if operating parameters are stable, once the glaze
layer exceeds a threshold thickness, its breaks down. This process
lasts as much as the mild wear regime is activated.

Appendix A. Supplementary data


Fig. 12. Operating parameters eect on GL behavior.
Supplementary data associated with this article can be found in the
Ed = 4. P. g . . N = 4. P. g . . f . t (3) online version at http://dx.doi.org/10.1016/j.wear.2017.07.008.

With Ed the energy dissipated at the interface (Fig. 1b) References


: the coecient of friction, t: test duration in seconds
It shows a bilinear wear trend for the three conditions (A and B). [1] F. Bratu, L. Benea, J.-P. Celis, Tribocorrosion behaviour of Ni-SiC composite coat-
The evolution A characteristic of the severe oxidational wear, shows ings under lubricated conditions, Surf. Coat. Technol. 201 (16) (2007) 69406946.
[2] M. Stack, Mapping tribo-corrosion processes in dry and in aqueous conditions: some
for every given condition, that wear increases linearly with test dura-
new directions for the new millennium, Tribol. Int. 35 (10) (2002) 681689.
tion, metal-metal contact is activated with continuous ejection of debris [3] E. Lemaire, M. Le Calvar, Evidence of tribocorrosion wear in pressurized water
generated, their ejection rate exceeds their formation rate, which in- reactors, Wear 249 (5) (2001) 338344.
[4] A. Jim'enez, M. Bermudez, Ionic liquids as lubricants of titaniumsteel contact. part
hibits the development of a stable glaze layer at this stage. The B
2: friction, wear and surface interactions at high temperature, Tribol. Lett. 37 (2)
evolution takes place, indicating transition to mild oxidational wear, (2010) 431443.
volume of debris needed to form a stable glaze layer is reached, and the [5] N. Dhar, M. Kamruzzaman, M. Ahmed, Eect of minimum quantity lubrication
energy dissipated needed to compact and sinter debris is aorded by (mql) on tool wear and surface roughness in turning aisi-4340 steel, J. Mater.
Process. Technol. 172 (2) (2006) 299304.
test parameters and ash temperature (heat generated by frictional [6] R. Waterhouse, Fretting wear, Wear 100 (13) (1984) 107118.
energy by sliding motion of the two surfaces, it reaches its maximum at [7] J. Hardell, B. Prakash, High-temperature friction and wear behaviour of dierent
the contacting asperities points). Once the glaze layer is formed, wear tool steels during sliding against Al-Si-coated high-strength steel, Tribol. Int. 41 (7)
(2008) 663671.
rate is stabilized. [8] T. Kitagawa, A. Kubo, K. Maekawa, Temperature and wear of cutting tools in high-
ne size debris are able to be trapped between the contacting sur- speed machining of inconel 718 and Ti-6Al-6V-2Sn, Wear 202 (2) (1997) 142148.
faces, and integrated directly in the development of the compact layers [9] F. Stott, D. Lin, G. Wood, The structure and mechanism of formation of the glaze
oxide layers produced on nickel-based alloys during wear at high temperatures,
on contacting surfaces [34,35]. As test conditions become tougher, Corros. Sci. 13 (6) (1973) 449IN3455454IN6469.
large-sized wear particles are formed, leading to increase in wear rate, [10] P. Hurricks, The fretting wear of mild steel from room temperature to 200 C , Wear
and time to reach mild wear regime is more important, due to time 19 (2) (1972) 207229.
[11] P. Hurricks, The fretting wear of mild steel from 200 to 500 C , Wear 30 (2) (1974)
needed to their fragmentation and compaction. This concept is illu-
189212.
strated in Fig. 12. [12] F. Stott, J. Glascott, G. Wood, The sliding wear of commercial Fe-12% Cr alloys at
As much as sliding conditions are severe, more wear volume is high temperature, Wear 101 (4) (1985) 311324.

47
F. Alkelae, S. Fouvry Wear 390391 (2017) 4148

[13] J. Glascott, F. Stott, G. Wood, The transition from severe to mild sliding wear for Fe- [24] S. Pearson, P. Shipway, J. Abere, R. Hewitt, The eect of temperature on wear and
12% Cr-base alloys at low temperatures, Wear 97 (2) (1984) 155178. friction of a high strength steel in fretting, Wear 303 (1) (2013) 622631.
[14] J. Glascott, F. Stott, G. Wood, The eectiveness of oxides in reducing sliding wear of [25] F. Stott, D. Lin, G. Wood, Glazes' produced on nickel-base alloys during high
alloys, Oxid. Met. 24 (3) (1985) 99114. temperature wear, Nature 242 (118) (1973) 7577.
[15] J. Jiang, F. Stott, M. Stack, The role of triboparticulates in dry sliding wear, Tribol. [26] F. Stott, D. Lin, G. Wood, C. Stevenson, The tribological behaviour of nickel and
Int. 31 (5) (1998) 245256. nickel-chromium alloys at temperatures from 20 to 800 C , Wear 36 (2) (1976)
[16] J. Jiang, F. Stott, M. Stack, Some frictional features associated with the sliding wear 147174.
of the nickel-base alloy N80A at temperatures to 250 C , Wear 176 (2) (1994) [27] R.C. Bill, Fretting of Nickel-chromium-aluminum Alloys at Temperatures to 816 C .
185194. [28] G. Wallwork, The oxidation of alloys, Rep. Prog. Phys. 39 (5) (1976) 401.
[17] H. Kato, K. Komai, Tribolm formation and mild wear by tribo-sintering of nan- [29] N.P. Suh, An overview of the delamination theory of wear, Wear 44 (1) (1977)
ometer-sized oxide particles on rubbing steel surfaces, Wear 262 (1) (2007) 3641. 116.
[18] H. Kato, Severe-mild wear transition by supply of oxide particles on sliding surface, [30] A. Tewari, Load dependence of oxidative wear in metal/ceramic tribocouples in
Wear 255 (1) (2003) 426429. fretting environment, Wear 289 (2012) 95103.
[19] R. Rybiak, S. Fouvry, B. Bonnet, Fretting wear of stainless steels under variable [31] F. Stott, The role of oxidation in the wear of alloys, Tribol. Int. 31 (1) (1998) 6171.
temperature conditions: introduction of a composite wear law, Wear 268 (3) [32] J. Tu, X. Jie, Z. Mao, M. Matsumura, The eect of temperature on the unlubricated
(2010) 413423. sliding wear of 5 CrNiMo steel against 40 MnB steel in the range 400 600 C ,
[20] T.M. Pollock, S. Tin, Nickel-based superalloys for advanced turbine engines: Tribol. Int. 31 (7) (1998) 347353.
chemistry, microstructure and properties, J. Propuls. Power 22 (2) (2006) 361374. [33] A. Pauschitz, M. Roy, F. Franek, Mechanisms of sliding wear of metals and alloys at
[21] R.C. Reed, The Superalloys: Fundamentals and Applications, Cambridge University elevated temperatures, Tribol. Int. 41 (7) (2008) 584602.
Press, 2008. [34] H. Kato, Eects of supply of ne oxide particles onto rubbing steel surfaces on
[22] T. Kayaba, A. Iwabuchi, The fretting wear of 0.45% C steel and austenitic stainless severe-mild wear transition and oxide lm formation, Tribol. Int. 41 (8) (2008)
steel from 20 to 650 C in air, Wear 74 (2) (1981) 229245. 735742.
[23] S. Hernandez, J. Hardell, C. Courbon, H. Winkelmann, B. Prakash, High tempera- [35] J. Jiang, F. Stott, M. Stack, A generic model for dry sliding wear of metals at ele-
ture friction and wear mechanism map for tool steel and boron steel tribopair, vated temperatures, Wear 256 (9) (2004) 973985.
Tribol.-Mater. Surf. Interfaces 8 (2) (2014) 7484.

48

Vous aimerez peut-être aussi