You are on page 1of 8

Wear 302 (2013) 829–836

Contents lists available at SciVerse ScienceDirect

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

Tribological behaviour of shot peened Cu–Ni austempered ductile iron


A. Zammit a,n, S. Abela a, L. Wagner b, M. Mhaede b, M. Grech a
a
Department of Metallurgy and Materials Engineering, Faculty of Engineering, University of Malta, Msida MSD 2080, Malta
b
Institute of Materials Science and Engineering, Clausthal University of Technology, Agricolastrasse 6, D-38678 Clausthal-Zellerfeld, Germany

a r t i c l e i n f o abstract

Article history: Wear and fatigue properties of power transmission components are usually improved by various
Received 6 August 2012 surface engineering processes. One surface modification process is shot peening which is generally
Received in revised form carried out to improve bending fatigue. However there are contrasting studies meant to investigating
10 December 2012
whether shot peening actually increases the sliding wear resistance of austempered ductile iron (ADI).
Accepted 13 December 2012
Unlubricated wear tests were conducted on ground ADI and shot peened ADI pins. Hardness
Available online 23 December 2012
measurements of the worn ADI surfaces showed a 19% increase in hardness after testing at low loads,
Keywords: possibly due to strain hardening and frictional heating. Metallography of the worn surfaces showed
Austempered ductile iron a distorted microstructure at the surface, indicative of surface flow. On the other hand, samples tested
Sliding wear
at high loads showed a 73% increase in hardness. A white non-etchable layer which was identified
Shot peening
as untempered martensite formed upon cooling of wear test samples. Calculation of the wear factors
Phase transformation
and friction coefficients showed that shot peening does not improve the wear resistance. This has been
attributed to the fact that the potential advantages resulting from the higher hardness at the surface,
stress-induced austenite to martensite transformation and the residual compressive stresses of the shot
peened specimens are counteracted by the induced surface roughness.
& 2012 Elsevier B.V. All rights reserved.

1. Introduction presentations of wear phenomena showing the wear rate and the
wear mechanism dominance in steel/steel tribocontacts over a
Careful selection of austempering heat treatment parameters wide range of loads and sliding speeds.
applied to ductile iron results in a variety of microstructures and a Surface engineering techniques are usually applied to improve
correspondingly wide range of bulk mechanical properties. This the surface properties namely by changing the microstructure or
renders austempered ductile iron (ADI) a potential alternative to the composition of the surface. This can be achieved by thermal,
steel, having comparable strength and toughness, lower density chemical, thermochemical or mechanical treatments. Tan et al. [8]
and greater damping capacity, combined with excellent castability. reported that after laser hardening, the surface hardness and
ADI is in fact suitable for automotive components such as crank abrasive wear resistance of nodular cast irons could be consider-
shafts, connecting rods and transmission gears [1,2]. ably improved due to a predominantly martensitic structure
It has been reported [3–6] that the unique wear behaviour of produced in the hardened zone. Lu and Zhang [9] obtained
ADI is affected by the presence of surface graphite nodules as well relatively high sliding wear resistance for both austempered and
as the ability of the retained austenite, which is metastable at laser-hardened Cu–Mo ADI specimens when compared to ductile
room temperature, to transform to martensite when loaded. iron specimens. This was attributable to the strain-induced
Straffelini et al. [4] show that ADI exhibited a lower coefficient martensite transformation of the retained austenite occurring
of friction and wear coefficient than that of nitrided steel during during the wear process and the martensite formed during laser
dry rolling-sliding wear testing. The authors attribute this to the processing. In addition, Xue et al. [10] reported that ADI speci-
smearing of graphite on to the surface which in turn served as mens with and without laser hardening showed a higher contact
a solid lubricant between the two wear surfaces. Straffelini et al. [4] fatigue resistance than that of induction hardened steel.
determined the wear mechanism occurring during sliding of Shot peening (SP) is a conventional mechanical surface treat-
austempered ductile iron by using the wear-mechanism maps ment that may be used to improve the fatigue strength of
described by Lim and Ashby [7]. The latter authors gave graphical automotive components subjected to fatigue loading. In this
process, the surface of a material is bombarded with a flow of
spherical media, creating a layer of compressive residual stress
n
Corresponding author. Tel.: þ356 2340 2066; fax: þ 356 21343577. and inducing high dislocation densities [11]. The compressive
E-mail address: ann.t.triccas@um.edu.mt (A. Zammit). layer induced by shot peening increases the resistance to crack

0043-1648/$ - see front matter & 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.wear.2012.12.027
830 A. Zammit et al. / Wear 302 (2013) 829–836

initiation and propagation which in turn prolongs the lifetime of of standard deviation to the mean) of the measurements was in
components. However, as the spherical shots hit the surface, all cases below 5%.
dimples are formed which roughen the surface. The surface finish Phase analysis before and after shot peening was carried out
of mating surfaces affects the wear resistance of components. This using the X-ray diffraction method and a Bruker D8 Advance
combination of a high friction coefficient at the beginning of the X-Ray diffractometer (Mo-Ka radiation). The scanning step was
test and a work hardened surface was also observed by Ohba et al. 0.011, the dwell time 0.2 s and 2y values between 15 and 401. The
[12] when studying rolling contact fatigue properties of ADI. tube acceleration voltage and current used were 45 kV and
Similar findings were also reported by Ho et al. [13] who showed 35 mA, respectively. The XRD patterns obtained were subjected
that shot peening did not improve the sliding wear resistance of to the Savitzky–Golay smoothing filter which performs a local
annealed 1018 steel, but it did decrease the wear rate of hardened polynomial regression to the raw data reducing the signal noise
4340 steel. On the other hand, work by Vaxevanidis et al. [14] [21]. A 3rd order regression was found to preserve features of the
showed that shot peening had a beneficial effect on the tribolo- pattern including relative maxima and width of peaks. The
gical behaviour of steel. retained austenite content (gret) in the ADI was measured with
A large number of studies have shown that shot peening the X-ray diffractometer using the simplified method described
improves the bending fatigue strength [15–19]. However, very by Miller [22].
few works have been conducted on the tribological behaviour
after shot peening, and it is thus not clear whether it actually
improves the wear resistance. This paper compares the sliding 2.4. Test equipment and conditions
wear behaviour of ground ADI with shot peened ADI specimens
tested under dry conditions. Dry sliding wear tests were carried out using a conventional
pin-on-disk tribometer capable of maintaining a constant unidir-
ectional, sliding velocity between the pin and disk. The machine
2. Experimental procedure used was an Italdesign TR-20 which allowed control of the load,
velocity, duration of test and radius at which the pin acts on the
2.1. Material and processing disk. In this study, a cantilever loaded flat ended cylindrical ADI
pin was made to slide over a rotating hardened steel disk. The pin
Test pins of 5 mm diameter used for the pin-on-disk wear was fixed to one end of the cantilever arm, and the other end was
testing were machined from ductile iron keel blocks, having the attached to displacement and force transducers. The tribometer
composition shown in Table 1. After machining, samples were was connected to a computer which monitored and recorded the
austenitised at 900 72 1C, and then rapidly quenched in a salt displacement of the pin and disk and the frictional force.
bath at 36075 1C and held for 1 h. The samples were then air Tests followed ASTM G99-05 (Standard test method for wear
cooled to room temperature. These austempering parameters testing with a pin-on-disk apparatus) procedures and were carried
were optimised in a previous study [20]. Samples were coated out in ambient air held at room temperature. Tests were
using a dedicated paint (SEMCO Zir H) to prevent decarburisation performed at two values of pressure acting between the cylind-
during the austempering process. rical surface of the pin and the horizontal rotating disk namely at
Test disks of 90 mm diameter were made out of D2 tool steel 2.5 and 10 MPa, while the sliding distance, sliding velocity and
of chemical composition shown in Table 2 and heat treated to a radius at which the pin acted upon the disk were kept constant at
hardness of 61 HRC. The heat treatment cycle consisted of pre- 3.6 km, 4 m/s and 26 mm, respectively.
heating, followed by austenitising at a temperature of 1025 1C Before and after tribological tests, both the pins and the disks
and then quenching using nitrogen at a pressure of 5 bar. After were cleaned for 10 min in an acetone ultrasonic bath, and then
that, the disks were tempered at a temperature of 190 1C for 3 h. rinsed in isopropanol and dried in a jet of hot air. For each
experiment, a new pin and new counter body were used. The
2.2. Surface treatment mass of both the pins and the disks was measured before and
after each test using a digital balance having an accuracy of
After heat treatment of both pins and disks, the surfaces were 70.1 mg. The mass lost was converted into wear volume W,
ground up to a mean surface roughness Ra of 0.2 mm. After taking the density r of the ADI material as 6890 kg/m3. The wear
grinding, half of the pins were shot peened. Shot peening was factor K was then calculated using the relation K ¼W/Fs, where W
carried out using S330 shots, with an Almen intensity of represents the wear volume in mm3, F is the applied load in N and
0.38 mmA up to 100% coverage. The stand-off distance was s is the sliding distance in m [23]. Two surface conditions of ADI
90 mm while the angle of impingement was set at 901. The were tested namely: ground ADI (G) and shot peened ADI (SP)
surface roughness Ra of the shot peened pins was measured to (Table 3). At least five sliding tests were carried out for each
be 3.7 mm. condition and applied load, and the average data is reported.

Table 2
2.3. Characterisation Chemical composition of the steel reference disks.

The microhardness measurements were taken using a Element Cr C Mo V Mn Si Fe

Mitutoyo MVK-H12 microhardness tester. Three measurements Composition (wt%) 11.8 1.55 0.8 0.8 0.4 0.3 Bal
for the hardness were taken. The coefficient of variation (the ratio

Table 1
Chemical composition of the keel blocks.

Element C Si Cu Ni Mn P Mg Al S Fe

Composition (wt%) 3.26 2.36 1.63 1.58 0.24 0.011 0.057 0.024 0.006 Bal
A. Zammit et al. / Wear 302 (2013) 829–836 831

Table 3
Tribological test conditions.

Specimen Surface Nominal applied


identification condition pressure (MPa)

G2.5 Ground ADI 2.5


SP2.5 Shot peened ADI
G10 Ground ADI 10
SP10 Shot peened ADI

Fig. 2. SP microstructure.

Fig. 1. Austempered ductile iron microstructure.

The tests were carried out at an ambient temperature of 1873 1C


and relative humidity of 4275%.

2.5. Post-wear characterisation

The worn surfaces of the pins were examined using optical


microscopy and scanning electron microscopy. Changes in the
subsurface region were also investigated by using metallographic
techniques and by taking microhardness measurements on the
cross-section wear test samples.
After the tests, the debris was collected in order to investigate
the products generated during sliding of the specimens. SEM and
Fig. 3. X-ray diffraction patterns of as-treated and shot peened ADI.
EDX were used to obtain information regarding the morphology
and chemical composition of the wear debris.
Optical micrograph observations are supported by the X-ray
diffraction patterns shown in Fig. 3. The amount of retained austenite
3. Results and discussion in as-austempered polished specimens was calculated to be about
44%. In comparison, none of the austenite peaks (g111, g200, g220, g311)
3.1. ADI microstructure were observed in the shot peened ADI specimen, but only ferrite and
martensite peaks. This indicates that the austenite has transformed to
The microstructure of the austempered material (Fig. 1) con- martensite as a result of shot peening.
tains 200 graphite nodules per mm2 with graphite nodularity The work hardening induced by shot peening and the phase
greater than 95% in a matrix consisting of acicular ferrite and transformation to martensite results in an increase of the surface
carbon enriched austenite. The macrohardness of the ADI struc- hardness from 370 to 535 HV. The microhardness depth-profile
ture is 336715 HV, while the microhardness of the ausferritic shown in Fig. 4 also indicates that the hardness falls continuously
structure is 370710 HV. as the distance from the surface increases.
The residual compressive stress of shot peened specimens was
3.2. Characterisation of shot peened specimens measured in a previous study [15]. The maximum occurred at the
surface and was about 975 MPa, 67% higher than the yield point of
Shot peening of austempered ductile iron results in a strain- the material.
induced phase transformation. In fact, the face centred cubic (FCC)
retained austenite present in the ausferrite matrix transforms to body 3.3. Study of the surfaces of wear test samples
centred tetragonal (BCT) martensite by cold-working caused by shots
impinging on the surface of the component. It is evident from the In tribological tests, the wear process changes as the proper-
micrograph in Fig. 2 that the microstructure at the surface is distinct ties of the surface alter. At a low applied pressure of 2.5 MPa, the
from that of the base matrix. microhardness of the worn surface is about 19% higher than that
832 A. Zammit et al. / Wear 302 (2013) 829–836

Fig. 6. Microstructure just below the worn surface shows distortion and surface
flow.
Fig. 4. Microhardness profile of specimens before wear tests.

Fig. 7. Martensite formed on worn surfaces, tested at an applied pressure of


10 MPa.
Fig. 5. Microhardness profiles of a cross section of the worn surface.

of the bulk; 440 and 370 HV, respectively (Fig. 5) for tested G in temperature, as well as localised temperature spikes where an
specimens. The thickness of this hardened layer is around asperity makes contact with the mating surface. The resulting rise
100 mm. This was also observed by Refaey et al. and Islam et al. in temperature may modify the mechanical and metallurgical
[24,25] who state that this increase in hardness is due to strain properties of the sliding surfaces, causing them to oxidise, or
hardening of the ausferritic matrix at the surface region which possibly melt. This high temperature transforms the ausferrite to
predominates over any frictional heating effect. As a result of this austenite and can result in carbon diffusion into the austenite,
plastic deformation, the material is stronger and causes surface making it stable and hence increasing the hardenability of the pin.
flow and the microstructure to distort. Fig. 6 shows the distorted As a consequence, the critical cooling rate is lowered, resulting in
microstructure of a section taken across the worn surface after the formation of untempered martensite at a slow cooling rate
testing with the lower load. upon cooling of the pin and disk after the test is stopped. Another
It is noted that the hardness of SP specimens decreases after plausible reason of martensite formation is that the austenite
wear testing, from 535 to 450 HV. This is probably due to the being produced due to the high temperatures being formed at the
removal of part of the shot peened layer during the wear test, or asperities is rapidly cooled due to heat being conducted into the
tempering of the martensite which was formed during the shot underlying bulk material.
peening process (Fig. 3). This was also observed by Fordyce et al. [26] who reported a
On the other hand, the microhardness at the surface of speci- white non-etching layer formed during the unlubricated sliding
mens tested at the higher load is over 600 HV (Fig. 5). This wear of austempered spheroidal cast iron. However, these white
indicates a phase transformation to a high hardness phase. layers were not found on the worn surfaces of the as-cast
Micrographs show that a white non-etchable phase is present at spheroidal iron. Straffelini et al. [4] explained how the wear rate
the surface of the specimens tested with the higher load (Fig. 7). of ADI at high sliding speeds (1.5–2.6 m/s) was dominated by the
When the two surfaces slide over each other, most of the work formation and cracking of this white layer formed on the sliding
done against friction is converted into heat, causing a general rise surface. Sharma [27] has also shown that high loads applied
A. Zammit et al. / Wear 302 (2013) 829–836 833

during wear testing may transform the metastable austenite to however that the potential improvement resulting from an
martensite. increase in hardness is being counterbalanced by the increased
surface roughness caused by shot peening.
3.4. Wear rates The wear factor was higher for specimens which were tested
at the higher load. The austenite to martensite transformation
In all tests, the disk wear was negligible. This can be explained increases the strength at the surface, rendering it able to with-
by the higher hardness of the steel disks compared to that of the stand the high contact loads. However, martensite is known to be
ADI pins, and the contact area on the disk which is several times brittle and is susceptible to cracking. Analysis of the worn
greater than that of the area of the pin. surfaces showed cracks at the surface of specimens (Fig. 10).
The wear displacement, which is the progressive movement of Cracks could also be a result of micro-adhesive interactions
the pin and disk [28], can be seen in Fig. 8. The graphs for the between the pin and disk. Propagation of these cracks led to
higher load are divided in two regions, a short running-in stage pitting of the surface (Figs. 10 and 11), resulting in loss of material
(about 300 m) and the longer steady-state region. The displace- and increased wear rate.
ment is much larger during the running-in stage, while during the The coefficients of variation (COV) which is the standard
steady-state region, the wear rate increases more or less linearly deviation divided by the mean were calculated in order to give
with the sliding distance. The graphs showing wear displacement an indication of the data dispersion and therefore reliability of the
for tribological tests carried out at the lower load show a constant results. The dispersion of the wear data was between 29 and 79%.
wear rate throughout the entire test, without a running-in region. The possible causes for the disparity in wear test results include
Fig. 9 shows the wear factor, K, of ground and shot peened the variation of the applied load caused by vibrations generated
specimens as a function of sliding distance. Even though there is a during testing, the difference in composition and microstructure
large scatter of wear factor values, it is noted that the ground due to the multiple phases of the materials being tested, and the
surfaces show a lower wear rate than the shot peened surfaces. accuracy of material loss measurements.
These results may suggest that a higher original hardness does The presence of graphite nodules has a major influence on the
not necessarily result in better wear resistance. It may be argued wear rate of ADI as superficial graphite is smeared over the
surface and aids in lubricating the surfaces in sliding contact. Also,
as can be seen in Fig. 12, cracks have a tendency of passing
through the graphite–matrix interface, this being the path of least
resistance. On the other hand, a nodule may arrest crack propaga-
tion, Fig. 13. Whether or not a crack is arrested or assisted to
propagate as it reaches a graphite nodule would depend on the
angle of approach. It follows that graphite nodules influence the
propagation path.

3.5. Coefficient of friction (CoF)

Frictional heat is generated when the pin slides on the disk.


It can be seen from CoF data shown in Fig. 14 that there is only
a minimal difference in the values pertaining to ground and shot
peened specimens. This is despite the initial rougher SP surface
(Ra ¼3.7 mm) as compared to the initial surface roughness of the
ground specimens (Ra ¼ 0.2 mm). Results suggest that the higher
wear rate of the SP specimens shown in Fig. 9, cannot be
attributed to localised heating caused by the increase in rough-
Fig. 8. Wear displacement as a function of sliding distance for G and SP specimens. ness of the dimpled surface resulting from the shot peening
treatment. On the contrary, Mokhtar claimed that heat treatment,

Fig. 9. Wear factors of G and SP samples. Fig. 10. Pitting of worn G surfaces tested at an applied pressure of 10 MPa.
834 A. Zammit et al. / Wear 302 (2013) 829–836

Fig. 11. Pitting of worn SP surfaces tested at an applied pressure of 10 MPa.


Fig. 14. Friction coefficient evolution.

hence lowers the frictional resistance [30]. This contrasts with the
results shown in the present work. One may argue that the
negative impact of the rougher surface, was in this study more
dominant than the beneficial influence of any phase transforma-
tion (Figs. 2 and 3), and compressive residual stresses present at
the surface of SP specimens [15]. Analysis of the surface rough-
ness of the shot peened wear test samples used in the two studies
may throw more light on this apparent inconsistency.
Fig. 14 also shows that a lower CoF was recorded when testing
specimens at a higher applied load. High applied pressure
promotes the phase transformation of retained austenite to
martensite, a harder and load bearing phase. It is known that
friction properties are generally improved when the hardness of
the surface is increased. As explained by Mokhtar et al. [31], cold
weld junctions formed when hard phases like martensite are
present are relatively easy to break, hence lowering the adhesion
Fig. 12. Crack propagation through graphite nodules (G10). of the surfaces and the frictional resistance. Also, martensite has
better thermal properties, providing better heat dissipation lead-
ing to a reduction in the CoF [32,33].
On the other hand, under the action of low applied loads, no
martensite formation is observed. The asperities of the harder
disk indent into the softer ausferritic structure of the pin causing
plastic deformation and strong cold-welded junctions are formed.
Frictional sliding resistance to motion is thus higher due to the
larger force required to shear these welded junctions [31].

3.6. Wear mechanism

Different wear mechanisms are believed to have occurred


during the tests. As observed by Straffelini et al. [4], the wear
factors shown in Fig. 9 are typical of mild oxidational wear [7,34].
During dry sliding, the surfaces interact with the atmosphere
resulting in a mild form of corrosive wear, known as oxidational
wear, which primarily occurs during unlubricated conditions of
sliding [23]. Due to the high frictional heating during sliding, fast
oxidation occurs. The oxide layers formed usually appear as
Fig. 13. Crack arrested by the graphite nodule. islands on the sliding surfaces (Fig. 15). The separation of the
surfaces due to these layers results in mild wear [35].
Debris is formed due to the fatigue of the oxide film produced
alloying or shot peening, increased surface hardness resulting in and the generation of frictional heat which raises the temperature
lower frictional resistance [29]. He attributed this to the fact that of the sliding surfaces. This weakens the bonding strength
hardening and shot peening introduce residual stresses and phase between the oxide film and the substrate, resulting in delamina-
transformations which lower the strength of the cold-welded tion of the oxide layer. The worn surfaces and the wear debris
junctions. This, he maintains, lowers adhesion of the surfaces and collected after wear tests had a red–brown tinge and EDX analysis
A. Zammit et al. / Wear 302 (2013) 829–836 835

identified large amounts of oxygen (Fig. 16). This confirms the Fig. 18 shows the wear scar of a pin where sliding marks
oxidational wear as the main wear mechanism which occurred parallel to the direction of sliding were observed. The fine grooves
during the tests. The other identified peaks in Fig. 16 shows the show that several plateaus are formed at the beginning of the
presence of iron, carbon, silicon and copper, all of which are wear process when the contacting surfaces achieved conformity
elements present in the material being studied. [23]. During sliding, these plateaus become unstable, they break
SEM observation of wear debris generated by sliding showed up to form debris, and other wear grooves tracks are formed. The
that small particles were agglomerated to combine into larger generation of wear particles can change the wear mechanism to
wear particles (Fig. 17). This was also observed by Stachowiak three-body abrasive wear, which leads to microploughing of the
where the wear particles were seen to agglomerate into clusters surface. The dimpled surface of the SP specimens can trap these
during the unlubricated sliding wear of steels and cast iron [36]. wear particles, thus increasing the wear rate [37].

4. Conclusions

In this study, unlubricated sliding wear tests were carried out


to determine the effect of shot peening (SP) on the tribological
behaviour of Cu–Ni alloyed austempered ductile iron. Pin-on-disk
tests were conducted on ground ADI, and shot peened ADI
specimens using two nominal applied loads. The results of the
present work are summarised as follows:

(1) Shot peening of ADI results in an increase in surface rough-


ness and hardness, together with austenite to martensite
transformation.
(2) Specimens tested at the lower load showed a distorted
microstructure just below the wear scar indicative of surface
flow. An increase in surface hardness of 19% was noted on
these surfaces after wear tests. On the other hand, untem-
Fig. 15. Oxide layer on worn surface of G2.5 pin. pered martensite was observed on the surface of specimens

Fig. 16. EDX analysis of the wear debris. Fig. 18. Ploughing marks on the pin surface.

Fig. 17. SEM micrographs of debris collected from wear tests; (a) ADI 10 MPa, (b) SP 10 MPa.
836 A. Zammit et al. / Wear 302 (2013) 829–836

tested at the higher load. This was attributed either to the [9] G.X. Lu, H. Zhang, The structure and sliding-contact wear resistance of a
high flash temperatures and/or to the high stresses reached laser-hardened austempered ductile iron, Wear 138 (1990) 1–12.
[10] L. Xue, M.U. Islam, G. McGregor, Dot matrix hardening of steels using a fiber
during sliding. optic coupled pulsed Nd:YAG laser, Materials and Manufacturing Processes
(3) The main wear mechanism was mild oxidative wear where 14 (1999) 53–65 1999.
a reddish-brown layer was seen on the worn surfaces. In [11] V. Schulze, Proceedings of the Eighth International Conference on Shot
Peening, Garmisch-Partenkirchen, Wiley, Germany, 2002.
addition, material loss was due to the result of three-body [12] H. Ohba, S. Matsuyama, T. Yamamoto, Effect of shot peening treatment on
abrasive wear. rolling contact fatigue properties of austempered ductile iron, Tribology
(4) No improvement was noticed on the wear factor after shot Transactions 45 (2012) 576–582.
[13] J.W. Ho, C. Noyan, J.B. Cohen, V.D. Khanna, Z. Eliezer, Residual stresses and
peening, despite the increase in surface microhardness, the sliding wear, Wear 84 (1983) 183–202.
introduction of residual compressive stresses, work hardening [14] N.M. Vaxevanidis, D.E. Manolakos, A. Koutsomichalis, G. Petropoulos,
of the surface and stress-induced austenite to martensite A. Panagotas, I. Sideris, A. Mourlas, S.S. Antoniou, The effect of shot peening
on surface integrity and tribological behaviour of tool steels in AITC-AIT,
transformation. The roughened surface induced by shot peening,
Parma, Italy, 2006.
and possibly the entrapment of wear particles between the [15] A. Zammit, M. Mhaede, M. Grech, S. Abela, L. Wagner, Influence of shot
dimpled surface of the SP specimens and the disk may have had peening on the fatigue life of Cu–Ni austempered ductile iron, Materials
an influence on the wear rate. In addition, no differences were Science and Engineering A 545, pp. 78–85.
[16] V. Schulze, Modern Mechanical Surface Treatment, WILEY-VCH, 2006.
noticed on the microstructures just below the worn surfaces, [17] Y. Ochi, K. Masaki, T. Matsumura, T. Sekino, Effect of shot-peening treatment
debris analysis and friction coefficients of ground and shot on high cycle fatigue property of ductile cast iron, International Journal of
peened samples. Fatigue 23 (2001) 441–448.
[18] A. Ebenau, D. Lohe, O. Vohringer, E. Macherauch, Influence of Shot Peening on
(5) This work shows that shot peening does not result in an the Microstructure and the Bending Fatigue Strength of Bainitic-Austenitic
improvement in the sliding wear resistance. This indicates Nodular Cast Iron in ICSP-4, 1990, pp. 389–398.
that shot peening can be applied to components which [19] M.H. Mhaede, K.M. Ibrahim, M. Wollmann, L. Wagner, Enhancing Fatigue
Performance of Ductile Iron by Austempering and Mechanical Surface
require a higher bending fatigue resistance, without lowering Treatments, in: Arabcast 2008, 2008.
significantly the wear resistance. [20] A. Zammit, L. Hopkins, J.C. Betts, M. Grech, Austenite transformation in
austempered ductile iron, in: Materials Science and Engineering (MSE 2008),
Nuremberg, Germany, 2008.
Acknowledgements [21] A. Savitzky, M.J.E. Golay, Smoothing and differentiation of data by simplified
least squares procedures, Analytical Chemistry 36 (1964) 1627–1639.
[22] R.L. Miller, A rapid X-ray measurement for the determination of retained
The authors would like to thank Mr. Mark Joseph Zerafa austenite, ASM Transactions 57 (1964) 892–899.
[23] A. International, Friction, Lubrication and Wear Technology Vol. 18, 1992.
(B.Eng.(Hons.)) for his contribution in machining and testing of
[24] A. Refaey, N. Fatahalla, Effect of microstructure on properties of ADI and low
specimens. alloyed ductile iron, Journal of Materials Science 38 (2003) 351–362.
In addition, the authors would like to acknowledge the [25] M.A. Islam, A.S.M.A. Haseeb, A.S.W. Kurny, Study of wear of as-cast and heat-
positive impact of ERDF funding and the purchase of the testing treated spheroidal graphite cast iron under dry sliding conditions, Wear 188
(1995) 61–65.
equipment through the project: ‘‘Developing an Interdisciplinary [26] E.P. Fordyce, The dry sliding wear behaviour of an austempered spheroidal
Material Testing and Rapid Prototyping R&D Facility (Ref. no. 012)’’. cast iron, Wear 135 (1990) 265–278.
[27] V.K. Sharma, Roller Contact Fatigue Study of Austempered Ductile Iron, vol. 3,
ASM, 1984 326–334.
References [28] M.J. Neale, M. Gee, Wear Problems and Testing for Industry, William Andrew
Publishing, 2001.
[29] M.O.A. Mokhtar, The effect of hardness on the frictional behaviour of metals,
[1] R.C. Voigt, C.R. Loper, Austempered ductile iron—process control and quality Wear 79 (1982) 297–304.
assurance, Journal of Heat Treating 3 (1984) 291–309. [30] M.A. Mokhtar, M.A.E. Radwan, The influence of quenching techniques on
[2] J. Race, L. Stott, Practical experience in the austempering of ductile iron, Heat frictional behaviour of carbon steels, in: Semin on Heat and Mass Transfer,
Treatment of Metals 4 (1991) 105–109. Dubrovnik, 1979.
[3] U.R. Kamari, R.P. Rao, Study of wear behaviour of austempered ductile iron, [31] M.O.A. Mokhtar, M. Zaki, G.S.A. Shawki, Effect of mechanical properties on
Journal of Materials Science 44 (2009) 1082–1093. frictional behaviour of metals, Tribology International (1979) 165–267.
[4] G. Straffelini, M. Pellizzari, L. Maines, Effect of sliding speed and contact [32] A. Gural, Influence of martensite particle size on dry sliding wear behaviour
pressure on the oxidative wear of austempered ductile iron, Wear 270 (2011) of low carbon dual phase powder metallurgy steel, Metallic Materials 48
714–719. (2010) 25–31.
[5] T. Nasir, D.O. Northwood, J. Han, Q. Zou, G. Barber, X. Sun, P. Seaton, Heat [33] S. Sendooran, P. Raja, Metallurgical investigation on cryogenic treated HSS
treatment-microstructure-mechanical/tribological property relationships in tool, International Journal of Engineering Science and Technology 3 (2011)
austempered ductile iron, in: Surface Effects and Contacts Mechanics X, 2011. 3992–3996.
[6] B. Bosnjak, B. Verlinden, B. Radulovic, Dry sliding wear of low alloyed [34] F.H. Stott, The role of oxidation in the wear of alloys, Tribology International
austempered ductile iron, Materials Science and Technology 19 (7) (2003) 31 (1998) 61–71.
650–656. [35] S.C. Lim, The relevance of wear-mechanism maps to mild-oxidational wear,
[7] S.C. Lim, M.F. Ashby, Wear-mechanism maps, Acta Metallurgica 35 (1987) Tribology International 35 (2002) 717–723.
1–24. [36] G.W. Stachowiak, G.B. Stachowiak, Unlubricated friction and wear behaviour
[8] Y.H. Tan, S.I. Yu, J.L. Doong, J.R. Wang, Abrasive wear property of bainitic of toughened zirconia ceramics, Wear 132 (1989) 151–171.
nodular cast iron in laser processing, Journal of Materials Science 25 (1990) [37] G.W. Stachowiak, Wear—Materials, Mechanisms, and Practice, John, Wiley &
4133–4139. Sons, Ltd, 2005.

You might also like