You are on page 1of 9

Tribology International 57 (2013) 6775

Contents lists available at SciVerse ScienceDirect

Tribology International
journal homepage: www.elsevier.com/locate/triboint

Sliding wear behaviour of surface mechanical attrition treated


AISI 304 stainless steel
Y. Sun
Department of Engineering, Faculty of Technology, De Montfort University, Leicester LE1 9BH, UK

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

Article history: In the present work, the sliding wear behaviour of surface mechanical attrition (SMA) treated AISI 304
Received 12 June 2012 stainless steel by spherical shot peening has been studied under both unlubricated and lubricated
Received in revised form conditions under various contact loads. It was found that although SMA treatments could produce a
13 July 2012
hardened layer of several hundred microns thick with much increased hardness from 200 HV0.1 to 480
Accepted 16 July 2012
Available online 25 July 2012
HV0.1, SMA treatments did not have a signicant effect on the unlubricated wear behaviour of the steel.
However, under oil-lubricating conditions, the SMA treated steel showed much better wear resistance
Keywords: than the untreated steel over a wide range of contact loads. The results are discussed in terms of plastic
Stainless steel deformation and property changes induced during the sliding wear process.
Mechanical attrition
& 2012 Elsevier Ltd. All rights reserved.
Strain hardening
Wear

1. Introduction these newly developed processes in enhancing the fatigue life of


metals have been demonstrated by several investigators [1921].
Surface modication of metallic materials by mechanical means It is also logical to conceive that due to the formation of a
such as cold working is not new. Such mechanical techniques as hardened layer as a result of strain hardening and surface nanocrys-
shot peening, deep rolling and burnishing have been widely used by tallization, these SMA treatments would have the potential use for
designers in industry to improve the fatigue life of engineering tribological applications. However, relatively few efforts have been
components [13]. These processes induce severe plastic deforma- reported regarding the tribological behaviour of shot peened and
tion in the near surface region of the component, which leads to SMA treated materials. The experimental results reported so far by
strain hardening, structural changes and the evolution of compres- different investigators are contradicting. Wang et al. [22] reported
sive residual stresses in the deformation zone. The benecial effects that the surface nanocrystallized layer produced on carbon steel by
of these mechanical treatments have been demonstrated by many SMA treatment can improve the wear resistance of the steel under
investigators and by many successful industrial applications [47]. dry sliding conditions. Similar observations have also been made for
More recently, due to the ability of severe plastic deformation in surface nanocrystallized medium carbon steel [23], pure copper [24],
creating nanocryslline structures, intensive efforts have been made CrSi alloy steel [25] and an aluminium alloy [26]. Ma et al. [14] and
to develop various surface nanocrystallization techniques to produce Wang et al. [16] also reported that ne particle bombardment and
a nanocrystalline layer in the surface region with a grain size sandblasting can improve the wear resistance of austenitic stainless
gradient [8,9]. These new developments include ultrasonic shot steel under dry sliding conditions. However, different results have
peening [10,11], laser shock peening [12,13], ne particle bombard- been reported by other investigators. Fridrici et al. [27] studied the
ment [14], grit blasting [15,16] and other surface mechanical fretting wear behaviour of shot peened titanium alloy and found that
attrition (SMA) treatments [17,18]. A common feature of most of shot peening has no benecial effect in improving the fretting wear
these SMAT processes lies in that the surface being treated is resistance of the alloy. Yan et al. [28] reported that the abrasive wear
plastically deformed repeatedly at a high strain rate. Accompanied resistance of Hadeld steel was increased by surface nanocrystalliza-
with structural renement is the hardening of the nanocrystalline tion only when soft abrasives were involved, while under hard
layer in accordance with the HallPetch relation and strain hard- abrasive conditions, the surface nanocrystallized layer showed dete-
ening caused by the cold work, together with the generation of a riorated abrasive wear resistance.
compressive residual stress prole. Indeed, the benecial effects of The aim of the present work was to study the wear behaviour
of SMA treated AISI 304 austenitic stainless steel (304SS) under
both dry and lubricated sliding conditions. 304SS is the most
widely used stainless steel grade and its response to shot peening,
E-mail address: ysun01@dmu.ac.uk ultrasonic peening and other SMA treatments has been widely

0301-679X/$ - see front matter & 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.triboint.2012.07.015
68 Y. Sun / Tribology International 57 (2013) 6775

studied in terms of structural evolution and surface hardening and made contact with the rotating specimen under a controlled
effect [8,10,1317]. This paper reports the experimental wear test contact load. All the tests were conducted at a xed rotation
results and demonstrates that SMA treatment under the present speed of 120 rpm for a total duration of 3600 s, i.e., 7200 sliding
conditions can considerably improve the wear resistance of the cycles. The wear track diameter for all tests was 7 mm, giving a
steel under lubrication conditions, whilst under unlubricated sliding speed of 4.4 cm s  1. Both unlubricated and lubricated
sliding conditions, SMA treatment is not benecial in improving tests were conducted. In unlubricated tests, two different types
wear resistance. of spherical sliders were used, i.e., a stainless steel bearing ball of
6 mm diameter (hardness: 910 HV0.1) and an alumina ball of
8 mm diameter (hardness: 2350 HV0.1). The contact load used in
2. Experimental unlubricated tests ranged from 1 N to 20 N, corresponding to
initial maximum contact pressures from 640 MPa to 1730 MPa for
The material used in the present work was AISI 304 austenitic the steel ball contact and from 590 MPa to 1600 MPa for the
stainless steel with the following nominal chemical compositions alumina ball contact. Lubricated tests were conducted by immer-
(in wt%): 0.07% C, 18.51% Cr, 9.42% Ni, 2.12% Mn and balance Fe. sing the test specimen in mineral oil of 15W40 grade, which is
The material was received in the form of a hot-rolled plate and typical of boundary lubrication. Only the steel ball was used as
was machined into test specimens of 30  30  3 mm3 sizes. The the slider in lubricated tests. All tests were conducted at room
original structure of the steel was austenite with an average grain temperature (22 1C), in ambient atmosphere and triplicated to
size of 45 mm. Apart from some annealing twins, no martensite was ensure reproducibility of test results. It was found that the wear
detected. Before SMA treatment, the specimens were wet ground volume from test to test under the same condition for the same
using a series of SiC grinding papers down to the 1200 grade and specimen varied within 20%. The average results are presented in
nished by 1 mm diamond polishing to achieve a mirror-like surface. this paper, mostly without including the error bars for clarity
Surface mechanical attrition (SMA) treatment was realized using purpose.
an electromagnetic shaker controlled by a signal generator and After each test, the wear volume from each wear track was
powered by an amplier. The vibration frequency and amplitude evaluated by measuring the surface proles across the wear track
could be adjusted to achieve various degrees of mechanical attrition at four locations of 901 apart, using a stylus surface prolometer.
onto the specimen surface. The principle of the system was similar The average cross-sectional area of the wear track was then
to those reported by other investigators [8,17,18,29] and was a estimated from the digitized proles by numerical integration
variant of conventional shot peening. Basically, a cylindrical cham- with reference to the original surface and the wear volume was
ber of 25 mm diameter and 25 mm long was attached to the shaker, obtained by multiplying the cross-sectional area by the circum-
and 20 AISI 440 stainless steel balls of 6 mm diameter (hardness: ferential length of the wear track. The surface morphology of the
910 HV0.1) were placed inside the chamber. The specimen was then tested specimens was examined by optical and scanning electron
secured to the top of the chamber. During the treatment process, the microscopy (SEM).
steel balls inside the chamber vibrated with the shaker and
impacted onto the specimen surface repeatedly at a high speed
and at the controlled frequency, causing indentation and plastic 3. Results
deformation on the impacting surface. Fig. 1 shows some interesting
images of the specimen surface after a short time treatment (3 s) at 3.1. Morphological and hardness evolution by SMA treatment
a frequency of 40 Hz. The image on the left shows an area with a
single indent, typical of normal impact. The one in the middle shows As can be clearly seen from Fig. 1, the impact of the spherical
an area with two overlapping impacts: a ridge was formed between steel shots on to the specimen surface at high speeds resulted in the
two dimples with material pileup. On the right hand side, the image formation of overlapping spherical indents on the surface with
shows the overlapping of three impacts with the formation of two permanent plastic deformation. With sufcient time of SMA treat-
ridges. Obviously, with a sufcient treatment time, the process ment, complete coverage of the surface by the indents was achieved.
would result in complete coverage of the surface with indents and Subsequent impacts would result in many shots on to the ridges
multiple overlapping. In the present work, based on many experi- formed by earlier impacts, leading to further plastic deformation in
mental trials, the specimens for wear testing were treated for the near surface region and changes to the surface morphology.
various times from 20 min to 300 min, at a vibration frequency of Fig. 2 compares the surface morphologies resulting from SMA
40 Hz and amplitude of 7 mm. treatments for various times. Individual dimples and ridges can be
Wear tests were conducted using a pin-on-disk tribometer clearly seen on the 20 min treated surface. As a result of such an
manufactured by Teer Coating Ltd. During the test, the specimen SMA treatment, the surface was signicantly roughened. After SMA
served as the disk rotating at a controlled speed, while a steel ball treatment for a longer time, it became more difcult to distinguish
or a ceramic ball served as the slider pin which was stationary individual dimples. The dimple feature, to a lesser extent, can still be

100m

Fig. 1. Optical micrographs showing the surface morphologies resulted from SMA treatment for 3 s with one 6 mm diameter ball at 40 Hz and 7 mm amplitude.
Y. Sun / Tribology International 57 (2013) 6775 69

80 m 80 m
20 min 90 min

80 m 80 m
180 min 300 min

Fig. 2. Optical micrographs showing the surface morphologies resulted from SMA treatment for various times with 20 balls at 40 Hz and 7 mm amplitude.

Table 1
Summary of surface roughness parameters.

Specimen Ran (mm) Rznn


(mm)

Untreated (polished nish) 0.07 0.31


20 min SMAT 0.81 4.73
90 min SMAT 0.72 3.55
180 min SMAT 0.41 2.40
300 min SMAT 0.33 1.85

n
Central line average.
nn
10 peak/valleys average.

observed after 90 min treatment. However, after SMA treatment for 100m
180 min and longer, the dimple feature disappeared; instead, the
ridges were attened and the surface became smoothened. Table 1
600
summarises the roughness parameter values (Ra and Rz) resulting
from various SMA treatment times. It can be seen that as compared
with the original polished surface nish, the SMA treated surfaces 300 min
500
were much rougher. Such a surface roughening was more signicant
for short time treatments, but with increasing treatment time, the
Microhardness (HV0.1)

surface roughness decreased, obviously due to the attening of the 400 180 min
ridges.
Fig. 3a shows the cross-sectional morphology of the SMA speci-
men treated for 180 min, together with microhardness indentations 300
90 min
made in the near surface region during microhardness prole
measurements. The measured hardness proles for the specimens 20 min
200
SMA treated for various times are given in Fig. 3b. Although it is
outside the scope of the present paper to give a full account of the
structural evolution by SMA treatments, some structural features 100
can be seen from Fig. 3a. First, a severely deformed surface layer
about 200 mm thick with obvious structural modications can be
clearly seen. The originally well-dened grain boundaries became 0
blurred in the severely deformed zone, which was populated with 0 100 200 300 400 500 600
deformation bands, identied as mechanical twins for such auste- Distance From Surface (um)
nitic stainless steel with low stacking fault energy [8,10,16,17]. Close Fig. 3. Microscopic image showing the cross-sectional morphology of the SMA
to the surface, multidirectional twinning within a grain was evident, sample treated for 180 min (a) and microhardness proles measured across the
with two or three operating systems. Such a morphological feature SMA layers treated for various times.
70 Y. Sun / Tribology International 57 (2013) 6775

6 mm diameter steel ball as the slider and under a constant load of


10 N. Typical wear track proles are illustrated in Fig. 5a, from which
it can be seen that the wear tracks produced on all specimens
showed a similar feature of being considerably roughened, but the
width and depth of the wear track varied slightly from specimen to
specimen. The main difference between the untreated specimen and
the SMA treated specimens lies in the material pileup at the two
edges of the wear track on the untreated specimen as can be seen
from the two lips of up to 5 mm height. This provides clear evidence
of severe plastic deformation of the untreated specimen during the
wear process. In Fig. 5b, the measured wear volumes are compared,
from which it can be seen that SMA treatment for 20 min and 90 min
reduced the wear volume of the steel only marginally. Increasing the
SMA treatment time to 180 min seemed to have enhanced the wear
resistance further by 20%. However, further increasing the SMA
treatment time to 300 min, although the hardening effect was
further increased (Fig. 3b), the wear resistance was deteriorated,
which was even worse than the untreated specimen. It thus
appeared that although SMA treatment under certain conditions
was slightly effective in improving wear resistance, other treatment
conditions did not seem to largely affect the wear resistance of the

5
Untreated

0 180 min

20 min
-5
Depth (m)

-10

-15 300 min

-20
10 N, 120 rpm
Fig. 4. Optical (a) and SEM (b) micrographs showing the mechanical twinning Steel ball, dry
structure 10 mm below the surface SMA treated for 90 min.
-25
0.0 0.4 0.8 1.2 1.6
can be more clearly seen in Fig. 4, which was obtained by slightly Horizontal Distance (mm)
grinding and polishing to remove a thin layer of about 10 mm thick
from the SMA treated surface, followed by etching in the 50 vol% 0.3
HCl25 vol% HNO3 25 vol% H2O etchant. The number of operating Unlubricated
systems of mechanical twins decreased to two and then one with Steel ball, 10 N
increasing distance from the surface. Such severe deformation 0.25
behaviour has been shown to be responsible for the structural
renement and nanocrystallization in the near surface region [17].
Wear Volume (mm3)

0.2
Apart from severe plastic deformation and surface nanocrystalli-
zation, another structural modication is the formation of strain-
induced martensite, which is a well-known phenomenon in severely 0.15
deformed austenitic stainless steel [10,17]. It has been shown that
martensites form at intersections between multidirectional
mechanical twins [17]. Thus, the much increased hardness in the 0.1
deformed layer shown in Fig. 3b is the combination results of strain
hardening, structural nanocrystallization and martensite formation.
0.05
From Fig. 3b, it is also evident that the hardness and thickness of the
deformed layer increased with increasing SMA treatment time, but
seem to be saturated after prolonged treatments. 0
Untreated
12345 20 min 90 min 180 min 300 min
3.2. Unlubricated sliding wear
Fig. 5. Surface proles across the wear tracks on selected specimens (a) and
comparison of wear volumes resulted from the untreated specimen and SMA
The rst set of wear tests was conducted on the untreated specimens treated for different times (b). Test conditions: unlubricated, steel ball,
specimen and SMA specimens treated for different times, using a 10 N, 120 rpm, 3600 s.
Y. Sun / Tribology International 57 (2013) 6775 71

1
0.8
Untreated 0.02

SMATed
Untreated
0.8
0.01
0.6

Wear Volume (mm3)


0.6
0
180 min 90 min 1N 2N Untreated
COF

0.4
0.4

0.2
0.2
Unlubricated Unlubricated
steel ball SMATed Steel ball
10 N, 120 rpm 120 rpm, 3600 s
0 0
0 600 1200 1800 2400 3000 3600 0 5 10 15 20 25
Time (s) Contact Load (N)
Fig. 6. Recorded friction curves for selected specimens tested under unlubricated 1
condition.

SMATed

steel under the present testing conditions. The difference in wear 0.8
volume from different specimens was less than 20%, within the
experimental error. The similar tribological behaviour of various Wear Volume (mm3)
specimens was further conrmed by examining the friction curves 0.6
Untreated
recorded during the wear tests, as shown in Fig. 6 for selected
specimens. The untreated and the SMA specimens treated for
different times exhibited similar frictional behaviour during the 0.4
whole test period: the coefcient of friction initially increased with
sliding time during the running-in period and then became stable at
similar values for all specimens. 0.2
Unlubricated
Further unlubricated wear tests were conducted under various Alumina ball
contact loads from 1 N to 20 N on the untreated specimen and the 120 rpm, 3600 s
180 min treated specimen which showed the best wear resistance 0
under the 10 N load as discussed above. The results, obtained 0 5 10 15 20 25
using a 6 mm diameter steel ball as a slider, are presented in Contact Load (N)
Fig. 7a. Except for the test under the 10 N load, the tests under
Fig. 7. Wear volume as a function of applied load under unlubricated condition for
other loads showed that the wear volume from the SMA treated the untreated and 180 min treated specimens against (a) a steel ball and (b) an
specimen was 1020% larger than that from the untreated speci- alumina ball slider.
men. Considering the error involved in such tests (20%), it is safe
to state that the effect of SMA treatment on unlubricated wear typical COF curves for the untreated and SMA treated specimens.
resistance of the investigated steel was insignicant. If there were All COF curves recorded at different loads were similar, charac-
any effect, it was more likely to be negative rather than positive in terised by the initially high COF for the rst 200 s testing and a
many cases. Such a conclusion was further supported by the fact steady state where the COF varied between 0.11 and 0.12, which
that all the data points in Fig. 7a, both from the untreated and is much lower than that measured under unlubricated condition
SMA treated specimens, fall in a general trend line, suggesting (about 0.7). The SMA treated specimen exhibited higher initial
similar wear behaviour. COF than the untreated specimen, obviously due to the rougher
To ascertain the above results, similar unlubricated tests were surface resulting from SMA treatment (see Table 1). In the steady
conducted using an alumina ball of 8 mm diameter as the slider. state, no large difference in COF was found between the SMA
The results summarised in Fig. 7b for the alumina ball slider are treated specimen and the untreated specimen. In Fig. 8b, the
similar to those in Fig. 7a for the steel ball slider. All the data measured wear volume was plotted against contact load for both
points in Fig. 7b also follow a similar trend line, suggesting again specimens. It can be clearly seen that under the present lubrica-
that the wear behaviour of both untreated and SMA treated tion condition, the SMA treated specimen exhibited much better
specimens was similar under the present unlubricated sliding wear resistance than the untreated specimen, in particular at high
conditions. This conclusion was further substantiated by examin- loads, where the wear volume of the steel was reduced by three
ing the friction curves recorded during the wear tests. It was times as a result of SMA treatment. Fig. 9 shows typical wear
observed that under similar loading and counterface conditions, track proles measured by prolometry. All wear tracks produced
the untreated and SMA treated specimens exhibited similar under lubrication showed similar characteristics. Material pileup
frictional behaviour, identical to those shown in Fig. 6. at the two edges of the wear track was evident in both specimens,
but was more signicant in the untreated specimen where the
3.3. Lubricated sliding wear pileup height was as large as the wear depth. From these wear
track proles, it was estimated that the pileup volume was as
In order to generate measurable wear, lubricated wear tests large as two third of the volumetric loss of material from the wear
were conducted at higher loads, from 5 N to 40 N. Fig. 8a shows track on the untreated specimen (with reference to the original
72 Y. Sun / Tribology International 57 (2013) 6775

surface). Thus, material loss from the wear track took two forms:
0.25 in the form of wear debris which formed a third body in the
Oil lubrication tribosystem and in the form of material pileup (pushing up) at the
5N, steel ball
wear track edges due to plastic deformation. The former was
0.2 more dominant in the SMA treated specimen, while the latter was
more dominant in the untreated specimen. It thus seems that the
ability of the SMA treated specimen to reduce plastic deformation
0.15 SMATed was responsible for the observed much reduced wear under the
lubrication condition.
COF

0.1
4. Discussion
Untreated

0.05 The wear test results obtained in this work demonstrated that
although SMA treatment can produce a hardened layer of several
hundred microns thick on the surface of the investigated auste-
nitic stainless steel, the effect of SMA treatment on unlubricated
0
0 600 1200 1800 2400 3000 3600 sliding wear resistance was insignicant. However, SMA treat-
Time (s) ment was very effective in improving wear resistance under
lubrication conditions. These results have obvious implications
0.06 in design regarding the use of SMA and possibly other severe
Mineral oil surface plastic deformation techniques as economic means in
Steel ball practical tribological applications to achieve enhanced compo-
120 rpm, 3600 s
0.05 nent performance and material sustainability.
Untreated The ineffectiveness of SMA treatment in improving unlubricated
sliding wear resistance of AISI 304 stainless steel investigated in this
Wear Volume (mm3)

0.04
work was found with no surprise. Although according to the theory
of adhesive and abrasive wear [30], the much increased hardness in
0.03 the surface layer induced by SMA treatment (see Fig. 3) is expected
to increase wear resistance, it has been known that for many metals,
0.02 hardening by cold working may not lead to pronounced increase in
SMATed friction and wear resistance [31]. Recently, there have been claims
that similar SMA treatments could improve unlubricated sliding
0.01 wear resistance of austenitic stainless steels [14,16] and other
materials [2226]. However, different observations have been made
by other investigators regarding the effect of prestrain [3234], cold
0
0 5 10 15 20 25 30 35 40 45 working [35], surface nanocrystallization [28] and bulk nanocrys-
Contact Load (N) tallization [36,37] on the unlubricated wear behaviour of metallic
materials. The early work reported by Khruschov [32] and Richard-
Fig. 8. Typical friction curves (a) and wear volume as a function of applied normal
son [33] clearly demonstrated that increasing the hardness of metals
load (b) under oil lubricated condition for the untreated and SMA treated
specimens against a steel ball slider.
by work hardening has little effect on abrasive wear resistance. Ray
et al. [34] studied the effect of prestrain by unidirectional tension on
unlubricated sliding wear behaviour of low carbon steel and found
that prestrain deteriorated wear resistance of the steel. Similarly,
Wu et al. [35] found that cold deformation of an AlTi alloy was only
effective in reducing unlubricated wear if oxidative wear was the
9
principal wear mechanism, while in the adhesive wear regime, cold
deformation had a detrimental effect. Regarding nanocrystallization,
6 although its benecial effect on improving tribological properties
have been reported [2226], Lv et al. [36] found that nanocrystalli-
zation of iron ingot by severe rolling deteriorated the unlubricated
3 wear resistance of the ingot. Talachi et al. [37] showed that
SMATed
Depth (um)

nanocrystallization of aluminium sheets by accumulative roll bond-


0 ing did not have a benecial effect on unlubricated wear resistance.
The reported deterioration in unlubricated wear resistance by
severe plastic deformation has been explained by the loss of
-3 plasticity and strain hardening ability due to the existence of
Untreated residual strain in the deformed layer [34,35]. During the wear
-6 process, the strain-hardened or nanocrystallized layer cannot
Oil-lubrication accommodate the shear strain induced by the contact traction,
Steel ball, 20 N instead voids and cracks tend to develop in the near surface
-9 region, leading to the formation of wear debris and accelerated
0.0 0.2 0.4 0.6 0.8 1.0 1.2 material loss through delamination [35].
Horizontal Distance (mm) The contact loads employed in the present wear tests were
Fig. 9. Typical surface proles across the wear tracks on the untreated and SMA sufcient to cause severe plastic deformation to the untreated
treated specimens resulted from oil-lubricated wear testing under 20 N load. surface. Coupled with the large tangential traction arising from a
Y. Sun / Tribology International 57 (2013) 6775 73

high COF of around 0.7 under unlubricated conditions, severe and properties similar to those induced by SMA treatments. It is
shear strain would occur at the surface of the wear track. Thus thus conceivable that after the running-in stage, the structure and
sliding wear is itself a severe surface plastic deformation process properties of the wear track on the untreated specimen should be
and would induce structural and property changes inside the similar to those on the SMA treated specimen. To further conrm
wear track. Indeed, studies by other investigators have clearly this, microhardness measurements were conducted at the centre
shown that sliding wear could cause strain hardening, phase of each wear track after wear testing. The results are summarised
transformation, structural renement and even amorphous phase in Table 2. It can be seen that after unlubricated wear testing, the
formation in stainless steels [38,39]. These are typical structures hardness of the untreated specimen in the wear track was
increased from 204 HV0.1 to 514582 HV0.1, and that of the
SMA treated specimen was increased from 485 HV0.1 to 529
Table 2 586 HV0.1. Clearly, despite the large difference in hardness before
Surface hardness of wear tracks before and after wear testing against a steel ball wear testing, the wear tracks on the untreated specimen and
slider under unlubricated condition.
those on the SMA treated specimen exhibited similar hardness. It
Contact load (N) Hardness (HV0.1) is thus not surprising that the two specimens exhibited similar
steady-state wear behaviour, as observed in Fig. 7.
Untreated 180 min SMAT Fig. 10 shows typical microscopic images of the wear tracks
produced on the two specimens. In consistence with prolometry
Before wear After wear Before wear After wear
measurements (Fig. 5a), the morphology of the wear tracks on the
1 205 514 485 529 SMA treated specimen are similar to that on the untreated
2 205 536 485 545 specimen, except for that two lips were formed at the track edges
5 205 559 485 567 on the untreated specimen due to continuous plastic deformation
10 205 570 485 579
20 205 582 485 586
as more materials were brought into contact (Fig. 10a). On the
other hand, no such lips were observed at the edges of the wear

Adhesion

Oxidation
Pile-up Pile-up

Abrasion

200m

Oxidation

Adhesion

Delamination

200m

Fig. 11. SEM micrographs showing the worn surface of the (a) untreated and
Fig. 10. Optical micrographs showing the wear track on the (a) untreated and (b) SMA treated specimen tested under 20 N load without lubrication. Counter-
(b) SMA treated specimen produced under 10 N load without lubrication. face: steel ball.
74 Y. Sun / Tribology International 57 (2013) 6775

tracks on the SMA treated specimen because the SMA treated induced by the contact traction. Instead, voids and cracks were
surface had been previous deformed and hardened (Fig. 10b). As formed and developed at the surface and subsurface.
can be clearly seen in Fig. 10 and Fig. 11, material removal from Under lubrication conditions, the wear behaviour of both speci-
the wear track was the results of adhesive, abrasive and oxidative mens was different from that observed under unlubricated condi-
wear, as well as delamination. Fig. 12 shows a cross-sectional tions discussed above. The most striking difference is that under
view of a typical wear track on the SMA treated specimen. Plate- lubrication, SMA treatment was very effective in improving the wear
like wear particles of varying sizes were formed through surface resistance of the investigated 304 stainless steel: the wear volume
and subsurface crack formation and propagation which led to was reduced by up to 3 times as a result of SMA treatment (Fig. 8b).
nal fracture and material delamination. This was the result of Similar observations have been made by Lv et al. [35] who found
the fact that as the wear track surface was fully hardened by that nanocrystallization of iron ingot was effective in improving
strain, it was no longer able to accommodate the plastic strain lubricated sliding wear resistance, although it reduced unlubricated
sliding wear resistance. Prolometry measurements (Fig. 9) revealed
that plastic deformation has played an important role in material
removal from the wear track: two third of the material removed
from the wear track on the untreated specimen was in the form of
material pileup at the edges of the track. Material removal in this
form from the SMA treated specimen was less, but still amounted up
to 40%. Clearly, the high hardness of the SMA treated surface
provided more resistance to plastic deformation and thus lowered
the wear rate under the lubrication condition. Material loss in the
form of wear debris formation was signicantly reduced due to the
lubrication effect. Fig. 13 shows typical microscopic images of the
Fig. 12. Microsection showing the worn subsurface of the SMA treated specimen wear tracks. No signs of adhesive wear and delamination were
tested under 20 N load without lubrication. Counterface: steel ball. observed, and the principal wear mechanism seems to be plastic
deformation and mild abrasive wear due to asperity contact with
the harder slider. In addition, the low COF under lubrication (0.11)
determines that the shear stress at the surface was relatively low
which also contributed to the much reduced delamination wear
which requires crack initiation and propagation.

5. Conclusions

(1) The present surface mechanical attrition treatment produces


a hardened layer of several hundred microns thick on the
surface of the stainless steel. In the hardened layer, the
hardness is increased from the base level of about 200 HV
to nearly 500 HV.
(2) The sliding wear behaviour of the SMA treated layer depends
on lubrication condition. Under dry sliding without lubrica-
tion, the SMA treated layer is not benecial in improving wear
resistance; while under oil immersion lubrication, the SMA
treated layer is very effective in enhancing wear resistance by
up to 3 times.
(3) During unlubricated sliding wear, material removal from the
wear track is in the form of debris formation resulting from
adhesion, abrasion, oxidation and delamination. The wear
track on the untreated specimen is markedly hardened during
sliding, such that it exhibits similar hardness to that of the
wear track on the SMA treated specimen. This could explain
the similar wear behaviour of the specimens.
(4) Under lubricated sliding, plastic deformation plays a more
important role in material removal from the wear track, in the
form of material pushing up to the two edges of the wear track
as a result of the ploughing action of the slider. The SMA treated
layer, due to its much higher hardness, provides more resistance
to plastic deformation and thus better wear resistance.
(5) Thus, severe surface plastic deformation via SMA treatments is
recommended for tribological application in lubrication situa-
tions, but is not recommended for unlubricated applications.

References

Fig. 13. Optical micrographs showing the wear track on the (a) untreated and [1] Wagner L. Mechanical surface treatments on titanium, aluminum and
(b) SMA treated specimen produced under 20 N load with oil lubrication. magnesium alloys. Materials Science and Engineering 1999;A263:2106.
Y. Sun / Tribology International 57 (2013) 6775 75

[2] OHara P. Developments in the shot peening process. Mater Design [19] Mordyuk BN, Prokopenko GI. Fatigue life improvement of a-titanium by
1984;5:1616. novel ultrasonically assisted technique. Materials Science and Engineering
[3] Hassan Adel Mahmood. The effects of ball- and roller-burnishing on the 2006;A437:396405.
surface roughness and hardness of some non-ferrous metals. Journal of [20] Roland T, Retraint D, Lu K, Lu J. Fatigue life improvement through surface
Materials Processing Technology 1997;72:38591. nanostructuring of stainless steel by means of surface mechanical attrition
[4] Benedetti M, Fontanari V, Hohn B-R, Oster P, Tobie T. Inuence of shot treatment. Scripta Materialia 2006;54:194954.
peening on bending tooth fatigue limit of case hardened gears. International [21] Dai K, Shaw L. Analysis of fatigue resistance improvements via surface severe
Journal of Fatigue 2002;24:112736. plastic deformation. International Journal of Fatigue 2008;30:1398408.
[5] Juijerm P, Noster U, Altenberger I, Scholtes B. Fatigue of deep rolled [22] Wang ZB, Tao NR, Li S, Wang W, Liu G, Lu J, Lu K. Effect of surface
AlMg4.5Mn (AA5083) in the temperature range 20300 1C. Material Science nanocrystallization on friction and wear properties in low carbon steel.
& Engineering 2004;A379:28692. Materials Science and Engineering 2003;A352:1449.
[6] Liu J, Gou WX, Liu W, Yue ZF. Effect of hammer peening on fatigue life of [23] Guobin Li, Jie Chen, Delin Guan. Friction and wear behaviour of nanocrystal-
aluminum alloy 2A12T4. Mater Design 2009;30:19449. line surface layer of medium carbon steel. Tribology International 2010;43:
[7] Fathallah R, Laamouri A, Sidhom H, Braham C. High cycle fatigue behavior 221621.
prediction of shot-peened parts. International Journal of Fatigue 2004;26: [24] Zhang YS, Han Z, Wang K, Lu K. Friction and wear behaviour of nanocrystal-
105367. line surface layer of pure copper. Wear 2006;260:9428.
[8] Lu K, Lu J. Nanostructured surface layer on metallic materials induced by [25] Ba DM, Ma SN, Meng FJ, Li CQ. Friction and wear behaviour of nanocrystalline
surface mechanical attrition treatment. Materials Science and Engineering surface layer of chrome-silicon alloy steel. Surface and Coatings Technology
2004;A375-377:3845. 2007;202:25460.
[9] Liu ZG, Fecht HJ, Umemoto M. Microstructural evolution and nanocrystal [26] Arun Prakash N, Gnanamoorthy R, Kamaraj M. Friction and wear behaviour
formation during deformation of FeC alloys. Materials Science and Engineering of surface nanocrystallized aluminium alloy under dry sliding condition.
Material Science & Engineering 2010;B168:17681.
2004;A375-377:83943.
[27] Fridrici V, Fouvry S, Kapsa Ph. Effect of shot peening on the fretting wear of
[10] Liu G, Lu J, Lu K. Surface nanocrystallization of 316 L stainless steel induced
Ti6Al4V. Wear 2001;250:6429.
by ultrasonic shot peening. Materials Science and Engineering 2000;A286:
[28] Weilin Yan, Liang Fang, Zhanguang Zheng, Kun Sun, Yunhua Xu. Effect of
915.
surface nanocrystallization on abrasive wear properties in Hadeld steel.
[11] Sanda A, Garcia Navas V, Gonzalo O. Surface state of inconel 718 ultrasonic
Tribology International 2009;42:63441.
peened: effect of processing time, material and quantity of shot balls and
[29] Villegas Juan C, Dai Kun, Shaw Leon L, Liaw Peter K. Nanocrystallization of a
distance from radiating surface to sample. Mater Design 2011;32:221320.
nickel alloy subjected to surface severe plastic deformation. Material Science
[12] Lu JZ, Zhang L, Feng AX, Jiang YF, Cheng GG. Effects of laser shock processing
& Engineering 2005;A410411:25760.
on mechanical properties of FeNi alloy. Mater Design 2009;30:36738.
[30] Archard JF. Contact and rubbing of at surfaces. Journal of Applied Physics
[13] Mordyuk BN, Milman YuV, Iemov MO, Prokopenko GI, Silberschmidt VV,
1953;24:9818.
Danylenko MI, Kotko AV. Characterization of ultrasonically peened and laser-
[31] Hutchings IM. Tribology: friction and wear of engineering materials. Edward
shock peened surface layers of AISI 321 stainless steel. Surface and Coatings Arnold; 1992 p210.
Technology 2008;202:487583. [32] Khruschov MMM. Principles of abrasive wear. Wear 1974;28:6988.
[14] Ma Guo-zheng, Xu Bin-shi, Wang Hai-dou, Si Hong-juan, Yang Da-xiang. [33] Richardson RCD. The wear of metals by hard abrasive. Wear 1967;10:
Effect of surface nanocrystallization on the tribological properties of 291309.
1Cr18Ni9Ti stainless steel. Materials Letters 2011;65:126871. [34] Ray KK, Toppo V, Singh SB. Inuence of pre-strain on the wear resistance of a
[15] Multigner M, Ferreira-Barragans S, Frutos E, Jaafar M, Ibanez J, Marin P, plain carbon steel. Material Science & Engineering 2006;A420:33341.
Perez-Prado MT, Gonzalez-Doncel G, Senjo A, Gonzalez-Carrasco JL. Super- [35] Wu JM, Zheng SL, Li ZZ, Zeng YW. Effects of cold deformation on the low
cial severe plastic deformation of 316 LVM stainless steel through grit speed sliding wear of the RS/PM Al10 wt% Ti alloy against carbon steel.
blasting: effects on its microstructure and subsurface mechanical properties. Wear 1999;232:2530.
Surface and Coatings Technology 2010;205:18307. [36] Lv XR, Wang SG, Liu Y, Long K, Li S, Zhang ZD. Effect of nanocrystallization on
[16] Wang XY, Li DY. Mechanical, electrochemical and tribological properties of tribological behaviours of ingot iron. Wear 2008;264:53541.
nano-crystalline surface of 304 stainless steel. Wear 2003;255:83645. [37] Kazemi Talachi A, Eizadjou M, Danesh Manesh H, Janghorban K. Wear
[17] Zhang HW, Hei ZK, Liu G, Lu J, Lu K. Formation of nanostructured surface characteristics of severely deformed aluminium sheets by accumulative roll
layer on AISI 304 stainless steel by means of surface mechanical attrition bonding (ARB) process. Materials Characterization 2011;62:1221.
treatment. Acta Materialia 2003;51:187181. [38] Zandrahimi Morteza, Raza Bateni M, Poladi A, Szpunar Jerzy A. The formation
[18] Arifvianto B, Suyitno M, Mahardika P, Dewo PT, Iswanto UA. Salim, effect of of martensite during wear of AISI 304-stainless steel. Wear 2007;263:6748.
surface mechanical attrition treatment (SMAT) on microhardness, surface [39] Petrov Yuri N, Gavriljuk Valentin G, Berns Hans, Schmalt Fabian. Surface
roughness and wettability of AISI 316 L. Materials Chemistry and Physics structure of stainless and hadeld steel after impact wear. Wear 2006;260:
2011;125:41826. 68791.

You might also like