Sunteți pe pagina 1din 9

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/256913852

Tribological behaviour of shot peened CuNi


austempered ductile iron
ARTICLE in WEAR APRIL 2013
Impact Factor: 1.91 DOI: 10.1016/j.wear.2012.12.027

CITATIONS

READS

75

5 AUTHORS, INCLUDING:
Ann Zammit

Stephen C Abela

University of Malta

University of Malta

2 PUBLICATIONS 7 CITATIONS

12 PUBLICATIONS 14 CITATIONS

SEE PROFILE

SEE PROFILE

Lothar Wagner

Mansour Mhaede

Technische Universitt Clausthal

Technische Universitt Clausthal

130 PUBLICATIONS 961 CITATIONS

58 PUBLICATIONS 128 CITATIONS

SEE PROFILE

SEE PROFILE

Available from: Mansour Mhaede


Retrieved on: 19 December 2015

Wear 302 (2013) 829836

Contents lists available at SciVerse ScienceDirect

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

Tribological behaviour of shot peened CuNi austempered ductile iron


A. Zammit a,n, S. Abela a, L. Wagner b, M. Mhaede b, M. Grech a
a
b

Department of Metallurgy and Materials Engineering, Faculty of Engineering, University of Malta, Msida MSD 2080, Malta
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:
Received 6 August 2012
Received in revised form
10 December 2012
Accepted 13 December 2012
Available online 23 December 2012

Wear and fatigue properties of power transmission components are usually improved by various
surface engineering processes. One surface modication process is shot peening which is generally
carried out to improve bending fatigue. However there are contrasting studies meant to investigating
whether shot peening actually increases the sliding wear resistance of austempered ductile iron (ADI).
Unlubricated wear tests were conducted on ground ADI and shot peened ADI pins. Hardness
measurements of the worn ADI surfaces showed a 19% increase in hardness after testing at low loads,
possibly due to strain hardening and frictional heating. Metallography of the worn surfaces showed
a distorted microstructure at the surface, indicative of surface ow. On the other hand, samples tested
at high loads showed a 73% increase in hardness. A white non-etchable layer which was identied
as untempered martensite formed upon cooling of wear test samples. Calculation of the wear factors
and friction coefcients 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.

Keywords:
Austempered ductile iron
Sliding wear
Shot peening
Phase transformation

1. Introduction
Careful selection of austempering heat treatment parameters
applied to ductile iron results in a variety of microstructures and a
correspondingly wide range of bulk mechanical properties. This
renders austempered ductile iron (ADI) a potential alternative to
steel, having comparable strength and toughness, lower density
and greater damping capacity, combined with excellent castability.
ADI is in fact suitable for automotive components such as crank
shafts, connecting rods and transmission gears [1,2].
It has been reported [36] that the unique wear behaviour of
ADI is affected by the presence of surface graphite nodules as well
as the ability of the retained austenite, which is metastable at
room temperature, to transform to martensite when loaded.
Straffelini et al. [4] show that ADI exhibited a lower coefcient
of friction and wear coefcient than that of nitrided steel during
dry rolling-sliding wear testing. The authors attribute this to the
smearing of graphite on to the surface which in turn served as
a solid lubricant between the two wear surfaces. Straffelini et al. [4]
determined the wear mechanism occurring during sliding of
austempered ductile iron by using the wear-mechanism maps
described by Lim and Ashby [7]. The latter authors gave graphical

Corresponding author. Tel.: 356 2340 2066; fax: 356 21343577.


E-mail address: ann.t.triccas@um.edu.mt (A. Zammit).

0043-1648/$ - see front matter & 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.wear.2012.12.027

presentations of wear phenomena showing the wear rate and the


wear mechanism dominance in steel/steel tribocontacts over a
wide range of loads and sliding speeds.
Surface engineering techniques are usually applied to improve
the surface properties namely by changing the microstructure or
the composition of the surface. This can be achieved by thermal,
chemical, thermochemical or mechanical treatments. Tan et al. [8]
reported that after laser hardening, the surface hardness and
abrasive wear resistance of nodular cast irons could be considerably improved due to a predominantly martensitic structure
produced in the hardened zone. Lu and Zhang [9] obtained
relatively high sliding wear resistance for both austempered and
laser-hardened CuMo ADI specimens when compared to ductile
iron specimens. This was attributable to the strain-induced
martensite transformation of the retained austenite occurring
during the wear process and the martensite formed during laser
processing. In addition, Xue et al. [10] reported that ADI specimens with and without laser hardening showed a higher contact
fatigue resistance than that of induction hardened steel.
Shot peening (SP) is a conventional mechanical surface treatment that may be used to improve the fatigue strength of
automotive components subjected to fatigue loading. In this
process, the surface of a material is bombarded with a ow of
spherical media, creating a layer of compressive residual stress
and inducing high dislocation densities [11]. The compressive
layer induced by shot peening increases the resistance to crack

830

A. Zammit et al. / Wear 302 (2013) 829836

initiation and propagation which in turn prolongs the lifetime of


components. However, as the spherical shots hit the surface,
dimples are formed which roughen the surface. The surface nish
of mating surfaces affects the wear resistance of components. This
combination of a high friction coefcient at the beginning of the
test and a work hardened surface was also observed by Ohba et al.
[12] when studying rolling contact fatigue properties of ADI.
Similar ndings were also reported by Ho et al. [13] who showed
that shot peening did not improve the sliding wear resistance of
annealed 1018 steel, but it did decrease the wear rate of hardened
4340 steel. On the other hand, work by Vaxevanidis et al. [14]
showed that shot peening had a benecial effect on the tribological behaviour of steel.
A large number of studies have shown that shot peening
improves the bending fatigue strength [1519]. However, very
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
wear behaviour of ground ADI with shot peened ADI specimens
tested under dry conditions.

of standard deviation to the mean) of the measurements was in


all cases below 5%.
Phase analysis before and after shot peening was carried out
using the X-ray diffraction method and a Bruker D8 Advance
X-Ray diffractometer (Mo-Ka radiation). The scanning step was
0.011, the dwell time 0.2 s and 2y values between 15 and 401. The
tube acceleration voltage and current used were 45 kV and
35 mA, respectively. The XRD patterns obtained were subjected
to the SavitzkyGolay smoothing lter which performs a local
polynomial regression to the raw data reducing the signal noise
[21]. A 3rd order regression was found to preserve features of the
pattern including relative maxima and width of peaks. The
retained austenite content (gret) in the ADI was measured with
the X-ray diffractometer using the simplied method described
by Miller [22].

2.4. Test equipment and conditions

After heat treatment of both pins and disks, the surfaces were
ground up to a mean surface roughness Ra of 0.2 mm. After
grinding, half of the pins were shot peened. Shot peening was
carried out using S330 shots, with an Almen intensity of
0.38 mmA up to 100% coverage. The stand-off distance was
90 mm while the angle of impingement was set at 901. The
surface roughness Ra of the shot peened pins was measured to
be 3.7 mm.

Dry sliding wear tests were carried out using a conventional


pin-on-disk tribometer capable of maintaining a constant unidirectional, sliding velocity between the pin and disk. The machine
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
disk. In this study, a cantilever loaded at ended cylindrical ADI
pin was made to slide over a rotating hardened steel disk. The pin
was xed to one end of the cantilever arm, and the other end was
attached to displacement and force transducers. The tribometer
was connected to a computer which monitored and recorded the
displacement of the pin and disk and the frictional force.
Tests followed ASTM G99-05 (Standard test method for wear
testing with a pin-on-disk apparatus) procedures and were carried
out in ambient air held at room temperature. Tests were
performed at two values of pressure acting between the cylindrical surface of the pin and the horizontal rotating disk namely at
2.5 and 10 MPa, while the sliding distance, sliding velocity and
radius at which the pin acted upon the disk were kept constant at
3.6 km, 4 m/s and 26 mm, respectively.
Before and after tribological tests, both the pins and the disks
were cleaned for 10 min in an acetone ultrasonic bath, and then
rinsed in isopropanol and dried in a jet of hot air. For each
experiment, a new pin and new counter body were used. The
mass of both the pins and the disks was measured before and
after each test using a digital balance having an accuracy of
70.1 mg. The mass lost was converted into wear volume W,
taking the density r of the ADI material as 6890 kg/m3. The wear
factor K was then calculated using the relation K W/Fs, where W
represents the wear volume in mm3, F is the applied load in N and
s is the sliding distance in m [23]. Two surface conditions of ADI
were tested namely: ground ADI (G) and shot peened ADI (SP)
(Table 3). At least ve sliding tests were carried out for each
condition and applied load, and the average data is reported.

2.3. Characterisation

Table 2
Chemical composition of the steel reference disks.

2. Experimental procedure
2.1. Material and processing
Test pins of 5 mm diameter used for the pin-on-disk wear
testing were machined from ductile iron keel blocks, having the
composition shown in Table 1. After machining, samples were
austenitised at 900 72 1C, and then rapidly quenched in a salt
bath at 36075 1C and held for 1 h. The samples were then air
cooled to room temperature. These austempering parameters
were optimised in a previous study [20]. Samples were coated
using a dedicated paint (SEMCO Zir H) to prevent decarburisation
during the austempering process.
Test disks of 90 mm diameter were made out of D2 tool steel
of chemical composition shown in Table 2 and heat treated to a
hardness of 61 HRC. The heat treatment cycle consisted of preheating, followed by austenitising at a temperature of 1025 1C
and then quenching using nitrogen at a pressure of 5 bar. After
that, the disks were tempered at a temperature of 190 1C for 3 h.
2.2. Surface treatment

The microhardness measurements were taken using a


Mitutoyo MVK-H12 microhardness tester. Three measurements
for the hardness were taken. The coefcient of variation (the ratio

Element

Cr

Mo

Mn

Si

Fe

Composition (wt%)

11.8

1.55

0.8

0.8

0.4

0.3

Bal

Table 1
Chemical composition of the keel blocks.
Element

Si

Cu

Ni

Mn

Mg

Al

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) 829836

831

Table 3
Tribological test conditions.
Specimen
identication

Surface
condition

Nominal applied
pressure (MPa)

G2.5
SP2.5
G10
SP10

Ground ADI
Shot peened ADI
Ground ADI
Shot peened ADI

2.5
10

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
EDX were used to obtain information regarding the morphology
and chemical composition of the wear debris.

3. Results and discussion


3.1. ADI microstructure
The microstructure of the austempered material (Fig. 1) contains 200 graphite nodules per mm2 with graphite nodularity
greater than 95% in a matrix consisting of acicular ferrite and
carbon enriched austenite. The macrohardness of the ADI structure is 336715 HV, while the microhardness of the ausferritic
structure is 370710 HV.
3.2. Characterisation of shot peened specimens
Shot peening of austempered ductile iron results in a straininduced phase transformation. In fact, the face centred cubic (FCC)
retained austenite present in the ausferrite matrix transforms to body
centred tetragonal (BCT) martensite by cold-working caused by shots
impinging on the surface of the component. It is evident from the
micrograph in Fig. 2 that the microstructure at the surface is distinct
from that of the base matrix.

Fig. 3. X-ray diffraction patterns of as-treated and shot peened ADI.

Optical micrograph observations are supported by the X-ray


diffraction patterns shown in Fig. 3. The amount of retained austenite
in as-austempered polished specimens was calculated to be about
44%. In comparison, none of the austenite peaks (g111, g200, g220, g311)
were observed in the shot peened ADI specimen, but only ferrite and
martensite peaks. This indicates that the austenite has transformed to
martensite as a result of shot peening.
The work hardening induced by shot peening and the phase
transformation to martensite results in an increase of the surface
hardness from 370 to 535 HV. The microhardness depth-prole
shown in Fig. 4 also indicates that the hardness falls continuously
as the distance from the surface increases.
The residual compressive stress of shot peened specimens was
measured in a previous study [15]. The maximum occurred at the
surface and was about 975 MPa, 67% higher than the yield point of
the material.
3.3. Study of the surfaces of wear test samples
In tribological tests, the wear process changes as the properties of the surface alter. At a low applied pressure of 2.5 MPa, the
microhardness of the worn surface is about 19% higher than that

832

A. Zammit et al. / Wear 302 (2013) 829836

Fig. 4. Microhardness prole of specimens before wear tests.

Fig. 5. Microhardness proles of a cross section of the worn surface.

of the bulk; 440 and 370 HV, respectively (Fig. 5) for tested G
specimens. The thickness of this hardened layer is around
100 mm. This was also observed by Refaey et al. and Islam et al.
[24,25] who state that this increase in hardness is due to strain
hardening of the ausferritic matrix at the surface region which
predominates over any frictional heating effect. As a result of this
plastic deformation, the material is stronger and causes surface
ow and the microstructure to distort. Fig. 6 shows the distorted
microstructure of a section taken across the worn surface after
testing with the lower load.
It is noted that the hardness of SP specimens decreases after
wear testing, from 535 to 450 HV. This is probably due to the
removal of part of the shot peened layer during the wear test, or
tempering of the martensite which was formed during the shot
peening process (Fig. 3).
On the other hand, the microhardness at the surface of specimens tested at the higher load is over 600 HV (Fig. 5). This
indicates a phase transformation to a high hardness phase.
Micrographs show that a white non-etchable phase is present at
the surface of the specimens tested with the higher load (Fig. 7).
When the two surfaces slide over each other, most of the work
done against friction is converted into heat, causing a general rise

Fig. 6. Microstructure just below the worn surface shows distortion and surface
ow.

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


10 MPa.

in temperature, as well as localised temperature spikes where an


asperity makes contact with the mating surface. The resulting rise
in temperature may modify the mechanical and metallurgical
properties of the sliding surfaces, causing them to oxidise, or
possibly melt. This high temperature transforms the ausferrite to
austenite and can result in carbon diffusion into the austenite,
making it stable and hence increasing the hardenability of the pin.
As a consequence, the critical cooling rate is lowered, resulting in
the formation of untempered martensite at a slow cooling rate
upon cooling of the pin and disk after the test is stopped. Another
plausible reason of martensite formation is that the austenite
being produced due to the high temperatures being formed at the
asperities is rapidly cooled due to heat being conducted into the
underlying bulk material.
This was also observed by Fordyce et al. [26] who reported a
white non-etching layer formed during the unlubricated sliding
wear of austempered spheroidal cast iron. However, these white
layers were not found on the worn surfaces of the as-cast
spheroidal iron. Straffelini et al. [4] explained how the wear rate
of ADI at high sliding speeds (1.52.6 m/s) was dominated by the
formation and cracking of this white layer formed on the sliding
surface. Sharma [27] has also shown that high loads applied

A. Zammit et al. / Wear 302 (2013) 829836

during wear testing may transform the metastable austenite to


martensite.
3.4. Wear rates
In all tests, the disk wear was negligible. This can be explained
by the higher hardness of the steel disks compared to that of the
ADI pins, and the contact area on the disk which is several times
greater than that of the area of the pin.
The wear displacement, which is the progressive movement of
the pin and disk [28], can be seen in Fig. 8. The graphs for the
higher load are divided in two regions, a short running-in stage
(about 300 m) and the longer steady-state region. The displacement is much larger during the running-in stage, while during the
steady-state region, the wear rate increases more or less linearly
with the sliding distance. The graphs showing wear displacement
for tribological tests carried out at the lower load show a constant
wear rate throughout the entire test, without a running-in region.
Fig. 9 shows the wear factor, K, of ground and shot peened
specimens as a function of sliding distance. Even though there is a
large scatter of wear factor values, it is noted that the ground
surfaces show a lower wear rate than the shot peened surfaces.
These results may suggest that a higher original hardness does
not necessarily result in better wear resistance. It may be argued

833

however that the potential improvement resulting from an


increase in hardness is being counterbalanced by the increased
surface roughness caused by shot peening.
The wear factor was higher for specimens which were tested
at the higher load. The austenite to martensite transformation
increases the strength at the surface, rendering it able to withstand the high contact loads. However, martensite is known to be
brittle and is susceptible to cracking. Analysis of the worn
surfaces showed cracks at the surface of specimens (Fig. 10).
Cracks could also be a result of micro-adhesive interactions
between the pin and disk. Propagation of these cracks led to
pitting of the surface (Figs. 10 and 11), resulting in loss of material
and increased wear rate.
The coefcients of variation (COV) which is the standard
deviation divided by the mean were calculated in order to give
an indication of the data dispersion and therefore reliability of the
results. The dispersion of the wear data was between 29 and 79%.
The possible causes for the disparity in wear test results include
the variation of the applied load caused by vibrations generated
during testing, the difference in composition and microstructure
due to the multiple phases of the materials being tested, and the
accuracy of material loss measurements.
The presence of graphite nodules has a major inuence on the
wear rate of ADI as supercial 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 graphitematrix interface, this being the path of least
resistance. On the other hand, a nodule may arrest crack propagation, 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 inuence the
propagation path.

3.5. Coefcient of friction (CoF)

Fig. 8. Wear displacement as a function of sliding distance for G and SP specimens.

Fig. 9. Wear factors of G and SP samples.

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 3.7 mm) as compared to the initial surface roughness of the
ground specimens (Ra 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 roughness of the dimpled surface resulting from the shot peening
treatment. On the contrary, Mokhtar claimed that heat treatment,

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

834

A. Zammit et al. / Wear 302 (2013) 829836

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


Fig. 14. Friction coefcient evolution.

Fig. 12. Crack propagation through graphite nodules (G10).

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 benecial inuence of any phase transformation (Figs. 2 and 3), and compressive residual stresses present at
the surface of SP specimens [15]. Analysis of the surface roughness 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
of the surfaces and the frictional resistance. Also, martensite has
better thermal properties, providing better heat dissipation leading 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

Fig. 13. Crack arrested by the graphite nodule.

alloying or shot peening, increased surface hardness resulting in


lower frictional resistance [29]. He attributed this to the fact that
hardening and shot peening introduce residual stresses and phase
transformations which lower the strength of the cold-welded
junctions. This, he maintains, lowers adhesion of the surfaces and

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
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 lm produced
and the generation of frictional heat which raises the temperature
of the sliding surfaces. This weakens the bonding strength
between the oxide lm and the substrate, resulting in delamination of the oxide layer. The worn surfaces and the wear debris
collected after wear tests had a redbrown tinge and EDX analysis

A. Zammit et al. / Wear 302 (2013) 829836

identied large amounts of oxygen (Fig. 16). This conrms the


oxidational wear as the main wear mechanism which occurred
during the tests. The other identied peaks in Fig. 16 shows the
presence of iron, carbon, silicon and copper, all of which are
elements present in the material being studied.
SEM observation of wear debris generated by sliding showed
that small particles were agglomerated to combine into larger
wear particles (Fig. 17). This was also observed by Stachowiak
where the wear particles were seen to agglomerate into clusters
during the unlubricated sliding wear of steels and cast iron [36].

835

Fig. 18 shows the wear scar of a pin where sliding marks


parallel to the direction of sliding were observed. The ne grooves
show that several plateaus are formed at the beginning of the
wear process when the contacting surfaces achieved conformity
[23]. During sliding, these plateaus become unstable, they break
up to form debris, and other wear grooves tracks are formed. The
generation of wear particles can change the wear mechanism to
three-body abrasive wear, which leads to microploughing of the
surface. The dimpled surface of the SP specimens can trap these
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 CuNi 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:

Fig. 15. Oxide layer on worn surface of G2.5 pin.

(1) Shot peening of ADI results in an increase in surface roughness 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
ow. An increase in surface hardness of 19% was noted on
these surfaces after wear tests. On the other hand, untempered 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) 829836

tested at the higher load. This was attributed either to the


high ash temperatures and/or to the high stresses reached
during sliding.
(3) The main wear mechanism was mild oxidative wear where
a reddish-brown layer was seen on the worn surfaces. In
addition, material loss was due to the result of three-body
abrasive wear.
(4) No improvement was noticed on the wear factor after shot
peening, despite the increase in surface microhardness, the
introduction of residual compressive stresses, work hardening
of the surface and stress-induced austenite to martensite
transformation. The roughened surface induced by shot peening,
and possibly the entrapment of wear particles between the
dimpled surface of the SP specimens and the disk may have had
an inuence on the wear rate. In addition, no differences were
noticed on the microstructures just below the worn surfaces,
debris analysis and friction coefcients of ground and shot
peened samples.
(5) This work shows that shot peening does not result in an
improvement in the sliding wear resistance. This indicates
that shot peening can be applied to components which
require a higher bending fatigue resistance, without lowering
signicantly the wear resistance.

Acknowledgements
The authors would like to thank Mr. Mark Joseph Zerafa
(B.Eng.(Hons.)) for his contribution in machining and testing of
specimens.
In addition, the authors would like to acknowledge the
positive impact of ERDF funding and the purchase of the testing
equipment through the project: Developing an Interdisciplinary
Material Testing and Rapid Prototyping R&D Facility (Ref. no. 012).
References
[1] R.C. Voigt, C.R. Loper, Austempered ductile ironprocess control and quality
assurance, Journal of Heat Treating 3 (1984) 291309.
[2] J. Race, L. Stott, Practical experience in the austempering of ductile iron, Heat
Treatment of Metals 4 (1991) 105109.
[3] U.R. Kamari, R.P. Rao, Study of wear behaviour of austempered ductile iron,
Journal of Materials Science 44 (2009) 10821093.
[4] G. Straffelini, M. Pellizzari, L. Maines, Effect of sliding speed and contact
pressure on the oxidative wear of austempered ductile iron, Wear 270 (2011)
714719.
[5] T. Nasir, D.O. Northwood, J. Han, Q. Zou, G. Barber, X. Sun, P. Seaton, Heat
treatment-microstructure-mechanical/tribological property relationships in
austempered ductile iron, in: Surface Effects and Contacts Mechanics X, 2011.
[6] B. Bosnjak, B. Verlinden, B. Radulovic, Dry sliding wear of low alloyed
austempered ductile iron, Materials Science and Technology 19 (7) (2003)
650656.
[7] S.C. Lim, M.F. Ashby, Wear-mechanism maps, Acta Metallurgica 35 (1987)
124.
[8] Y.H. Tan, S.I. Yu, J.L. Doong, J.R. Wang, Abrasive wear property of bainitic
nodular cast iron in laser processing, Journal of Materials Science 25 (1990)
41334139.

[9] G.X. Lu, H. Zhang, The structure and sliding-contact wear resistance of a
laser-hardened austempered ductile iron, Wear 138 (1990) 112.
[10] L. Xue, M.U. Islam, G. McGregor, Dot matrix hardening of steels using a ber
optic coupled pulsed Nd:YAG laser, Materials and Manufacturing Processes
14 (1999) 5365 1999.
[11] V. Schulze, Proceedings of the Eighth International Conference on Shot
Peening, Garmisch-Partenkirchen, Wiley, Germany, 2002.
[12] H. Ohba, S. Matsuyama, T. Yamamoto, Effect of shot peening treatment on
rolling contact fatigue properties of austempered ductile iron, Tribology
Transactions 45 (2012) 576582.
[13] J.W. Ho, C. Noyan, J.B. Cohen, V.D. Khanna, Z. Eliezer, Residual stresses and
sliding wear, Wear 84 (1983) 183202.
[14] N.M. Vaxevanidis, D.E. Manolakos, A. Koutsomichalis, G. Petropoulos,
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,
Parma, Italy, 2006.
[15] A. Zammit, M. Mhaede, M. Grech, S. Abela, L. Wagner, Inuence of shot
peening on the fatigue life of CuNi austempered ductile iron, Materials
Science and Engineering A 545, pp. 7885.
[16] V. Schulze, Modern Mechanical Surface Treatment, WILEY-VCH, 2006.
[17] Y. Ochi, K. Masaki, T. Matsumura, T. Sekino, Effect of shot-peening treatment
on high cycle fatigue property of ductile cast iron, International Journal of
Fatigue 23 (2001) 441448.
[18] A. Ebenau, D. Lohe, O. Vohringer, E. Macherauch, Inuence of Shot Peening on
the Microstructure and the Bending Fatigue Strength of Bainitic-Austenitic
Nodular Cast Iron in ICSP-4, 1990, pp. 389398.
[19] M.H. Mhaede, K.M. Ibrahim, M. Wollmann, L. Wagner, Enhancing Fatigue
Performance of Ductile Iron by Austempering and Mechanical Surface
Treatments, in: Arabcast 2008, 2008.
[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.
[21] A. Savitzky, M.J.E. Golay, Smoothing and differentiation of data by simplied
least squares procedures, Analytical Chemistry 36 (1964) 16271639.
[22] R.L. Miller, A rapid X-ray measurement for the determination of retained
austenite, ASM Transactions 57 (1964) 892899.
[23] A. International, Friction, Lubrication and Wear Technology Vol. 18, 1992.
[24] A. Refaey, N. Fatahalla, Effect of microstructure on properties of ADI and low
alloyed ductile iron, Journal of Materials Science 38 (2003) 351362.
[25] M.A. Islam, A.S.M.A. Haseeb, A.S.W. Kurny, Study of wear of as-cast and heattreated spheroidal graphite cast iron under dry sliding conditions, Wear 188
(1995) 6165.
[26] E.P. Fordyce, The dry sliding wear behaviour of an austempered spheroidal
cast iron, Wear 135 (1990) 265278.
[27] V.K. Sharma, Roller Contact Fatigue Study of Austempered Ductile Iron, vol. 3,
ASM, 1984 326334.
[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,
Wear 79 (1982) 297304.
[30] M.A. Mokhtar, M.A.E. Radwan, The inuence of quenching techniques on
frictional behaviour of carbon steels, in: Semin on Heat and Mass Transfer,
Dubrovnik, 1979.
[31] M.O.A. Mokhtar, M. Zaki, G.S.A. Shawki, Effect of mechanical properties on
frictional behaviour of metals, Tribology International (1979) 165267.
[32] A. Gural, Inuence of martensite particle size on dry sliding wear behaviour
of low carbon dual phase powder metallurgy steel, Metallic Materials 48
(2010) 2531.
[33] S. Sendooran, P. Raja, Metallurgical investigation on cryogenic treated HSS
tool, International Journal of Engineering Science and Technology 3 (2011)
39923996.
[34] F.H. Stott, The role of oxidation in the wear of alloys, Tribology International
31 (1998) 6171.
[35] S.C. Lim, The relevance of wear-mechanism maps to mild-oxidational wear,
Tribology International 35 (2002) 717723.
[36] G.W. Stachowiak, G.B. Stachowiak, Unlubricated friction and wear behaviour
of toughened zirconia ceramics, Wear 132 (1989) 151171.
[37] G.W. Stachowiak, WearMaterials, Mechanisms, and Practice, John, Wiley &
Sons, Ltd, 2005.

S-ar putea să vă placă și