Sunteți pe pagina 1din 7

Wear 310 (2014) 8389

Contents lists available at ScienceDirect

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

Wear resistance and tribological features of pure aluminum


and AlAl2O3 composites consolidated by high-pressure torsion
Kaveh Edalati a,b,n, Maki Ashida b, Zenji Horita a,b, Toshiaki Matsui c, Hirotaka Kato c
a

WPI, International Institute for Carbon-Neutral Energy Research (WPI-I2CNER), Kyushu University, Fukuoka 819-0395, Japan
Department of Materials Science and Engineering, Faculty of Engineering, Kyushu University, Fukuoka 819-0395, Japan
c
Department of Mechanical Engineering, Fukui National College of Technology, Fukui 916-8507, Japan
b

art ic l e i nf o

a b s t r a c t

Article history:
Received 27 September 2013
Received in revised form
21 December 2013
Accepted 24 December 2013
Available online 30 December 2013

Ultrane-grained pure Al and Al-based composites with 10 and 20 vol% of Al2O3 were produced by cold
consolidation of powders using high-pressure torsion (HPT). Ball-on-disc wear resistance of Al was improved
by HPT when compared to its coarse-grained counterpart processed with H24 treatment (cold rolling
followed by low-temperature annealing). Wear width decreased but wear depth and wear volume increased
with the addition of Al2O3 to the Al matrix. The wear mechanism was mainly due to adhesion in Al and Al
Al2O3 composites. It was found that the variation of wear width can be represented by a unique function of
the ratio of load/hardness in consistency with the Reye0 s hypothesis and the Holm and Archard relationships.
However, the variation of wear depth was inconsistent with the hardness variations.
& 2013 Elsevier B.V. All rights reserved.

Keywords:
Severe plastic deformation (SPD)
Ultrane-grained materials
Ball-on-disc sliding wear
Adhesive wear
Adhesion

1. Introduction
During last three decades, different severe plastic deformation
(SPD) methods have been utilized to achieve ultrane-grained
(UFG) microstructures in metallic materials [1,2]. The most popular SPD methods are currently Equal-Channel Angular Pressing
(ECAP) [3], High-Pressure Torsion (HPT) [4] and Accumulative RollBonding (ARB) [5]. Superior mechanical properties such as high
hardness, high strength and reasonable plasticity are often
achieved in the SPD-processed metallic materials [16].
Despite numerous papers regarding the microstructural renement and mechanical property improvement of SPD-processed
materials, there have been rather conicting reports on the evolution
of wear resistance in materials processed by ECAP [739], ECAP
followed by cold rolling [40], ECAP followed by HPT [41], HPT [35,42
49] and ARB [3739,5055]. Wear resistance is an important
mechanical property which should be improved for specimens with
mechanical contact and relative motion between their surfaces.
Different studies have reported either increase [730,40,4250],
decrease [3035,38,39,41,48,5153] or no appreciable change [36
39,49,54,55] in the wear resistance after SPD. Several strategies have
been introduced to improve the wear resistance of SPD-processed

n
Corresponding author at: Kyushu University, Department of Materials Science
and Engineering, Faculty of Engineering, 744 Motooka, Nishi-ku, Fukuoka-shi,
Fukuoka 819-0395, Japan. Tel./fax: 81 92 802 2992.
E-mail address: kaveh.edalati@zaiko6.zaiko.kyushu-u.ac.jp (K. Edalati).

0043-1648/$ - see front matter & 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.wear.2013.12.022

materials [56,57]: (i) heat treatment after the SPD processing [25,26],
(ii) coating the surface with ultrahard materials such as TiO2 [27],
diamond-like carbon [45] and TiN [46,47], and (iii) addition of
ceramic reinforcements such as SiC [28,30,49] and Al2O3 [29,30] to
the matrix.
Pure Al is among the most interesting model materials for SPD
processing because of its unique microstructural and mechanical
features (e.g., see a recent report in [58]). Tribological behavior of
pure Al has also been of interest for several decades because of its
ductility and capacity to deform plastically under wear (e.g., see a
report by Kuo and Rigney on ball-on-disc dry sliding of Al [59]).
Several papers investigated the wear resistance of pure Al processed by ECAP [34,35], HPT [35] and ARB [51,53,54] and reported
an increase in the hardness but a reduction in the wear resistance
after the SPD processing. It is an important issue to explore new
strategies to improve the wear resistance of SPD-processed Al.
In this study, the HPT processing is applied to pure Al powders and
AlAl2O3 powder mixtures to consolidate the powders, as attempted
earlier (e.g., see [6063]), and produce bulk forms of Al and AlAl2O3
composites. It is shown that the cold consolidation of powders using
the HPT processing is an effective strategy to improve the wear
resistance.

2. Experimental materials and procedures


Materials used in this study were high purity Al (99.99%)
powders with particle size less than 150 m and pure -Al2O3

84

K. Edalati et al. / Wear 310 (2014) 8389

powders with an average particle size of  30 nm. The morphology of Al and Al2O3 powders has been reported in an earlier paper
[63]. The Al powders were mixed with either 10 or 20 vol% Al2O3
powders using mechanical agitation. The powder mixtures were
then subjected to ball milling in air for 100 min at a rotation speed
of 300 rpm using a planet type ball mill, where the weight ratio of
the balls to powders was 2:1. To prevent any possible contamination during operation of BM, alumina vessels were used with
alumina balls. HPT was conducted at room temperature using the
pure Al powders without any BM processing and using the BMprocessed AlAl2O3 powder mixtures to consolidate to discs with
20 mm diameter and 0.8 mm thickness under a pressure of
P 1.5 GPa and subsequently introduce strain through N 10 turns
with a rotation speed of 1 rpm. It should be noted that a
pressure of 1.5 GPa is high enough to process most Al-based
metallic materials using HPT. As a reference coarse-grained material, commercially pure Al (Al-1050 with 99.5% Al) was received in
a form of sheet with a thickness of 0.95 mm after cold rolling
followed by annealing under the H24 condition. The sheet was cut
to discs with 20 mm diameter using a wire-cutting electric
discharge machine.
The disc samples after HPT as well as after H24 processing were
evaluated in terms of density measurement, Vickers microhardness
measurement, wear test and scanning electron microscopy (SEM).
First of all, both sides of the disc samples were polished to a
mirror-like surface and their density was determined by
Archimedes0 principle using an electronic balance with an accuracy of 0.1 mg.
Second, the Vickers microhardness was measured with an
applied load of 50 g for 15 s along the radii from the center to
edge at two different radial directions with 0.5 mm increments as
depicted by cross marks in Fig. 1.
Third, for wear test, the disc samples were ground with a
#2000 SiC abrasive paper and further washed in acetone. The disc
samples were examined using a ball-on-disc dry sliding method,
as shown schematically in Fig. 2, in pure argon (99.999%) atmosphere. A 6 mm diameter high carbon-chromium steel (SUJ2) ball
with 798 Hv hardness and mirror-like surface was contacted on
the surface of disc at either 2.5 or 7.5 mm from the disc center
under the loads of 2, 5 and 10 N. The disc was concurrently rotated
for 10 min with a rotation speed of 2 rpm at room temperature.
Following the wear test, the disc samples and balls were examined
by a digital microscope. Moreover, cross-sectional proles of the
wear scar along the radii were obtained using a prolometer at
8 different radial directions, as shown schematically in Fig. 1. The
wear depth, the wear width and the wear area were determined
from each cross-sectional prole and the average values of
8 measurements were calculated. The average wear volume within

Fig. 1. Schematic illustration of disc and positions for microhardness measurements, wear test and SEM observation.

Fig. 2. Schematic illustration of wear test using ball-on-disc sliding method.

10 min was calculated by multiplying the sliding length in


wear area.
Fourth, SEM was performed at 15 kV for observation of disc
surface before and after the wear test.

3. Results and discussion


SEM micrographs are shown in Fig. 3 for (a and d) pure Al, (b
and e) Al10% Al2O3 and (c and f) Al20% Al2O3 after processing
with HPT, where (ac) were taken from the center of disc and (df)
were taken from the edge of disc. It is apparent that the pores
which are locally visible at the center of discs, as indicated by
arrows in (ac), are not visible at the edge of discs because of
better consolidation resulting from higher imposed shear strain,
( 2rN=h, r: distance from disc center, N: number of turns, h:
disc thickness). The dark areas in (d) and (f) correspond to Al2O3
particles and it is apparent that there are many dark areas of
which sizes are at the micrometer range. This indicates that many
fractions of the Al2O3 particles are present in an agglomerated
form. Density measurements after HPT show that the relative
density is  100% for the pure Al but decreases to  98% and  96%
for the composite containing 10% and 20% of Al2O3, respectively.
Evolution of microstructure in Al powders and AlAl2O3 powder mixtures were examined in earlier papers using transmission
electron microscopy (TEM), indicating that the average grain size
reaches 500 nm in Al [60] and  480 nm in the AlAl2O3
composites [63]. These grain sizes are smaller than the grain size
of  1900 nm reported for the bulk sample of Al after HPT [58].
This indicates that oxide layers on the surface of Al powders and
the addition of Al2O3 nano-particles to the Al matrix contribute to
an appreciable renement of the microstructure. Other publications also reported that the consolidation of Ni powders [61] and
Ti powders [62] also has a large effect on the grain renement and
the mechanical property improvement due to pining effect of the
oxide particles.
Fig. 4 plots the hardness variation with the distance from the
disc center for pure Al, Al10% Al2O3 and Al20% Al2O3 after
processing with HPT. The hardness level for the H24-treated Al
sample is also included in Fig. 4. The hardness increases with the
distance from the disc center after HPT. This increase becomes
prominent as the Al2O3 fraction increases. Whereas the hardness
saturates quickly to a constant level for the pure Al sample, the
hardness keeps increasing with increasing the distance from the
disc center to 6.5 mm and 8.0 mm for the samples with the 10%
and 20% Al2O3 fractions, respectively.
Fig. 5 shows the appearance of disc samples of pure Al, Al10%
Al2O3 and Al20% Al2O3 after wear test under different loads of

K. Edalati et al. / Wear 310 (2014) 8389

85

Fig. 3. Micrographs of (a and d) Al (99.99%), (b and e) Al10% Al2O3 and (c and f) Al20% Al2O3 after HPT, where (ac) correspond to center of discs and (df) correspond to
edge of discs. Relative density values for each sample were also included.

Fig. 4. Vickers microhardness plotted against distance from disc center for HPTprocessed Al (99.99%), Al10% Al2O3 and Al20% Al2O3 discs, including hardness
level for H24-processed Al (99.5%) sample.

Fig. 5. Appearance of disc samples after wear test under loads of 2, 5 and 10 N.

2, 5 and 10 N. The wear width increases with increasing the load


in consistency with the Reye0 s hypothesis (wear volume is proportional to frictional energy) [64] and the Holm and Archard
relationships [65,66], but this increase becomes less prominent
as the HPT processing is used or the Al2O3 fraction increases. This
trend is more clearly demonstrated in digital micrographs in Fig. 6
(ac), where the decrease in wear width is obvious for the HPT-

processed Al and it is more signicant for the AlAl2O3 composite


consolidated by HPT. The current trends suggest that the harder
materials exhibit lower wear width which is consistent with the
predictions of Holm and Archard relationships [65,66].
Examination of balls using digital microscope after wear test, as
shown in Fig. 6(df), indicates that plastic deformation and
adhesion of materials from disc to the ball occurs in pure Al
samples. The adhesion appears to be more signicant for the
coarse-grained Al sample when compared to the HPT-processed
sample. For the HPT-processed AlAl2O3 composite, Fig. 6(f) shows
buildup material on the ball surface, which was probably produced
by transfer from the disc or adhesion of wear debris. Detailed
examinations using energy dispersive X-ray spectrometry conrmed that the adhesion of Al to the ball occurred. Although a few
studies found that the main wear mechanism in the AlAl2O3
composites with high fractions of Al2O3 at low sliding speeds is
abrasion [67], no evidence for abrasive wear could be found in this
study. For the HPT-processed AlAl2O3 composite, debris, which
stack to the ball, are sharp akes elongated in the wear direction,
indicating that the plastic deformation should have occurred
insignicantly during the wear test. Inspection of the wear surface
using SEM, as shown in Fig. 7(a), also clearly shows the occurrence
of surface plastic deformation and delamination in the HPTprocessed Al, as indicated by A in (a). The occurrence of intensive
plastic deformation in pure Al during wear test has been shown
clearly in earlier papers [59]. However, the wear surface of the
HPT-processed AlAl2O3 composites, as shown in Fig. 7(b), exhibits
a river-like pattern with no signicant surface plastic deformation.
Wear cross-sectional proles are shown in Fig. 8 for three
samples: H24-processed pure Al, pure Al after HPT, and Al20%
Al2O3 composite after HPT. The variation of wear width is
essentially the same as the ones shown in Fig. 6(ac) and the
wear width decreases by the HPT processing and by the addition
of Al2O3 particles to Al. The wear depth for the HPT-processed Al is
also smaller than that for the coarse-grained Al sample, indicating
that the cold consolidation of Al powders using HPT is an effective
strategy to minimize the wear rate of pure Al. However, the wear
depth for the AlAl2O3 composite is unusually higher than those
for the pure Al samples. The high wear depth in the AlAl2O3
composite should be due to the intensive fracture and transfer
(adhesive wear) owing to the lack of plastic deformation of the
composite during wear testing and due to agglomeration of Al2O3
particles, as shown in Fig. 3. It should be noted that the
wear cracks could be detected neither in the Al samples nor in
the AlAl2O3 composite.

86

K. Edalati et al. / Wear 310 (2014) 8389

Fig. 6. Appearance of (ac) wear scar on discs and (df) ball after wear test at 7.5 mm from disc center under a load of 5 N, observed by digital microscope.

Fig. 8. Cross-sectional prole of wear scars on HPT-processed discs of Al (99.99%)


and Al20% Al2O3 including annealed Al (99.5%) disc after wear test at 7.5 mm from
disc center under a load of 5 N.

Fig. 7. Appearance of wear scar on HPT-processed discs of (a) Al (99.99%) and (b)
Al20% Al2O3 after wear test at 7.5 mm from disc center under a load of 10 N,
observed by SEM.

The variation of (a) wear width, (b) wear depth and (c) wear
volume for four different samples are shown in Fig. 9 with respect
to the wear load. Fig. 10 shows the variation of (a) wear width and
(b) wear depth for two different distances from the disc center,
r 2.5 and 7.5 mm. The following ve important points are
obtained from Figs. 9 and 10.
(i) The wear width, the wear depth and wear volume increase
with an increase in the load, which is well consistent with the
Reye0 s hypothesis [64] and the Holm and Archard relationships [65,66].
(ii) Materials with higher hardness exhibit lower wear width.
(iii) The HPT-processed Al exhibits the lowest wear depth and the
lowest wear volume because of fragmentation of surface
oxide layer and distribution of very ne oxide particles,

whereas the wear depth and the wear volume unusually


increases for the AlAl2O3 composites because of agglomeration of Al2O3 nano-particles and lack of plastic deformation.
(iv) For the AlAl2O3 composites, the wear width at r 2.5 mm
and r 7.5 mm (corresponding to a shear strains of  200 and
 590, respectively) are reasonably the same, but the wear
depth signicantly decreases with increasing the distance
from the disc center (i.e., increasing the shear strain). This
suggests that a further increase in the shear strain can
improve the wear resistance of the AlAl2O3 composites
because of an increase in the hardness as well as because of
more homogeneous dispersion of Al2O3 particles. As reported
earlier [63] and shown in Fig. 3, agglomeration of Al2O3
particles still remains even after HPT processing for 10 turns
but, the agglomeration becomes less signicant with increasing the shear strain. The results in Fig. 10 indicate an
important suggestion that the wear resistance should be
enhanced if nano-sized Al2O3 powders are homogeneously
dispersed without agglomeration.
(v) For the HPT-processed Al, the wear depths at r 2.5 mm
(onset of steady state in Fig. 4) and r 7.5 mm (steady state
in Fig. 4) are the same, but the wear width increases with
increasing the distance from the disc center as from
r 2.5 mm to r 7.5 mm. The difference in the wear width
must be due to the difference in the microstructures. It is
known that the dislocation and grain boundary structures
appreciably inuence the wear behavior of soft materials such

K. Edalati et al. / Wear 310 (2014) 8389

87

Fig. 10. Variation of (a) wear width and (b) wear depth for different samples after
wear test at 2.5 and 7.5 mm from disc center under a load of 10 N.

Fig. 9. Variation of (a) wear width, (b) wear depth and (c) wear volume with load
after wear test at 7.5 mm from disc center.

as Al [59] and Cu [68]. It was shown earlier that the


microstructure of Al at the onset of steady state (r 2.5 mm)
contains higher densities of dislocations than those at the
steady sate (r 7.5 mm). The latter microstructure appears to
exhibit a larger wear width than the former one.
Fig. 11 shows the variation of (a) wear width and (b) wear
depth with respect to the wear load normalized by Vickers
microhardness (F/HV). This normalization was attempted because
the load normalized by the ow pressure (here hardness) is a
controlling factor for wear rate in the Holm and Archard relationships [65,66]. It is apparent from Fig. 11(a) that the wear width
increases monotonically with increasing F/HV, as expected
from the Holm and Archard relationships. However, inspection of
Fig. 11(b) shows that the wear depth can not be represented by a
unique function of F/HV for four selected materials. This contradiction arises because the wear mechanism and the wear resistance is inuenced signicantly not only by the hardness but also
by the microstructure, plasticity, friction coefcient and the adhesion between the two contacting materials. Close inspection of
Fig. 11(b), however, shows that the wear depth increases linearly
with increasing F/HV for each selected material. According to the

Fig. 11. Variation of (a) wear width and (b) wear depth with load normalized by
Vickers microhardness after wear test at 2.5 mm and 7.5 mm from disc center
under loads of 2, 5 and 10 N (all data points shown in Figs. 9 and 10).

88

K. Edalati et al. / Wear 310 (2014) 8389

Holm and Arachard relationships, the slopes of the tted lines in


Fig. 11(b) represent the wear resistance (and thus, materials with
smaller slopes have higher wear resistance). It is concluded that
pure Al consolidated with HPT exhibits the highest resistance and
the Al20% Al2O3 composite the lowest. It is probable that this
consequence should be attributed to the contribution of Al2O3
nano-powders on the plasticity and the wearing behavior of the
samples. For the former, a ne homogeneous distribution of Al2O3
particles arising from the fragmentation of the surface layers on
the initial Al powders helps enhance the wear resistance, while for
the latter, a rather incomplete heterogeneous distribution of Al2O3
particles through agglomeration is less effective for the improvement despite the inclusion of the large quantity as 20 vol%.

4. Conclusions
Pure Al and AlAl2O3 composites with ultrane-grained microstructures were consolidated by HPT and their wear properties
were investigated using a ball-on-disc method. The following
conclusions were obtained.
1. The wear width, the wear depth and the wear volume of HPTconsolidated Al with ultrane-grained structure are smaller
when compared to those of coarse-grained Al, indicating that
cold consolidation of powders using HPT is an effective strategy
to improve the wear resistance.
2. Wear width decreases, but the wear depth and the wear
volume unusually increase with the addition of Al2O3 nanoparticles to the Al matrix because of lack of plasticity. However,
poor wear resistance of AlAl2O3 composites can be improved
by increasing shear strain because of increasing hardness and
homogeneous distribution of Al2O3 nano-particles in the Al
matrix.
3. The variation of wear width is represented by a unique function
of wear load normalized by hardness, whereas this is not the
case for the variation of wear depth.
4. The dominant wear mechanism is adhesion in Al and AlAl2O3
composites.

Acknowledgments
One of the authors (KE) thanks the Japan Society for Promotion
of Science (JSPS) for a grant (No. 25889043). This work was
supported in part by the Light Metals Educational Foundation of
Japan and in part by a Grant-in-Aid for Scientic Research from the
MEXT, Japan, in Innovative Areas Bulk Nanostructured Metals
22102004.
References
[1] R.Z. Valiev, R.K. Islamgaliev, I.V. Alexandrov, Bulk nanostructured materials
from severe plastic deformation, Prog. Mater. Sci. 45 (2000) 103189.
[2] R.Z. Valiev, Y. Estrin, Z. Horita, T.G. Langdon, M.J. Zehetbauer, Y.T. Zhu,
Producing bulk ultrane-grained materials by severe plastic deformation,
JOM 58 (4) (2006) 3339.
[3] R.Z. Valiev, T.G. Langdon, Principles of equal-channel angular pressing as a
processing tool for grain renement, Prog. Mater. Sci. 51 (2006) 881981.
[4] A.P. Zhilyaev, T.G. Langdon, Using high-pressure torsion for metal processing:
Fundamentals and applications, Prog. Mater. Sci. 53 (2008) 893979.
[5] Y. Saito, N. Tsuji, H. Utsunomiya, T. Sakai, R.G. Hong, Ultra-ne grained bulk
aluminum produced by accumulative roll-bonding (ARB) process, Scr. Mater.
39 (1998) 12211227.
[6] R. Pippan, S. Scheriau, A. Taylor, M. Hafok, A. Hohenwarter, A. Bachmaier,
Saturation of fragmentation during severe plastic deformation, Annu. Rev.
Mater. Res. 40 (2010) 319343.

[7] M.I.A.E. Aal, N.E. Mahallawy, F.A. Shehata, M.A.E. Hameed, E.Y. Yoon, H.S. Kim,
Wear properties of ECAP-processed ultrane grained AlCu alloys, Mater. Sci.
Eng. A 527 (2010) 37263732.
[8] X. Cheng, Z. Li, G. Xiang, Dry sliding wear behavior of TiNi alloy processed by
equal channel angular extrusion, Mater. Des. 28 (2007) 22182223.
[9] L.L. Gao, X.H. Cheng, Effect of ECAE on microstructure and tribological
properties of Cu10%Al4%Fe alloy, Tribol. Lett. 27 (2007) 221225.
[10] L.L. Gao, X.H. Cheng, Microstructure and dry sliding wear behavior of Cu10%
Al4%Fe alloy produced by equal channel angular extrusion, Wear 265 (2008)
986991.
[11] L.L. Gao, X.H. Cheng, Microstructure, phase transformation and wear behavior
of Cu10%Al4%Fe alloy processed by ECAE, Mater. Sci. Eng. A 473 (2008)
259265.
[12] S.J. Huang, V.I. Semenov, L.S. Shuster, P.C. Lin, Tribological properties of the
low-carbon steels with different micro-structure processed by heat treatment
and severe plastic deformation, Wear 271 (2011) 705711.
[13] Y. Kim, H.S. Yu, D.H. Shin, Tribological characteristics of coarse and ultra-ne
grained ferrite-martensite dual phase steel fabricated by equal channel
angular pressing, Solid State Phenom. 124126 (2007) 13891392.
[14] L.G. Korshunov, N.I. Noskova, A.V. Korznikov, N.L. Chernenko, N.F. Vildanova,
Effect of severe plastic deformation on the microstructure and tribological
properties of a babbit B83, Phys. Metal Metallogr. 108 (2009) 519526.
[15] L.G. Korshunov, V.G. Pushin, N.L. Chernenko, V.V. Makarov, Structural transformations, strengthening, and wear resistance of titanium nickelide upon
abrasive and adhesive wear, Phys. Metal Metallogr. 110 (2010) 91101.
[16] T. Kucukomeroglu, Effect of equal-channel angular extrusion on mechanical
and wear properties of eutectic Al12Si alloy, Mater. Des. 31 (2010) 782789.
[17] P.Q. La, J.Q. Ma, Y.T. Zhu, J. Yang, W.M. Liu, Q.J. Xue, R.Z. Valiev, Dry-sliding
tribological properties of ultrane-grained Ti prepared by severe plastic
deformation, Acta Mater. 53 (2005) 51675173.
[18] A. Moshkovich, V. Perlyev, I. Lapsker, D. Gorni, L. Rapoport, The effect of grain
size on Stribeck curve and microstructure of copper under friction in the
steady friction state, Tribol. Lett. 42 (2011) 8998.
[19] A. Moshkovich, V. Perlyev, D. Gorni, I. Lapsker, L. Rapoport, The effect of Cu
grain size on transition from EHL to BL regime (Stribeck curve), Wear 271
(2011) 17261732.
[20] E. Ortiz-Cuellar, M.A.L. Hernandez-Rodriguez, E. Garcia-Sanchez, Evaluation of
the tribological properties of an AlMgSi alloy processed by severe plastic
deformation, Wear 271 (2011) 18281832.
[21] G. Purcek, O. Saray, T. Kucukomeroglu, M. Haouaoui, I. Karaman, Effect of
equal-channel angular extrusion on the mechanical and tribological properties of as-cast Zn40Al2Cu2Si alloy, Mater. Sci. Eng. A 527 (2010)
34803488.
[22] V.I. Semenov, L.S. Shuster, S.J. Huang, P.C. Lin, Tribological behavior of lowcarbon steel depending on treatment and structural state, J. Frict. Wear 32
(2011) 205211.
[23] N. Thiyaneshwaran, P. Sureshkumar, Microstructure, mechanical and wear
properties of aluminum 5083 alloy processed by equal channel angular
extrusion, Int. J. Eng. Res. Technol. 2 (2013) 1724.
[24] J. Xu, X. Wang, X. Zhu, M. Shirooyeh, J. Wongsa-Ngam, D. Shan, B. Guo,
T.G. Langdon, Dry sliding wear of an AZ31 magnesium alloy processed by
equal-channel angular pressing, J. Mater. Sci. 48 (2013) 41174127.
[25] Z.H. Li, X.H. Cheng, Effects of equal channel angular extrusion and subsequent
heat treatment on tribological properties of TiNi alloy, Surf. Eng. 23 (2007)
434438.
[26] Z.H. Li, X.H. Cheng, Improvement in wear resistance of TiNi alloy processed by
equal channel angular extrusion and annealing treatment, Adv. Tribol. 2010
(2010) 257259.
[27] A. Alsaran, G. Purcek, I. Hacisalihoglu, Y. Vangolu, O. Bayrak, I. Karaman,
A. Celik, Hydroxyapatite production on ultrane-grained pure titanium by
micro-arc oxidation and hydrothermal treatment, Surf. Coat. Technol. 205
(2011) S537S542.
[28] V.I. Semenov, Y.R. Jeng, S.J. Huang, Y.Z. Dao, S.J. Hwang, L.S. Shuster,
S.V. Chertovskikh, P.C. Lin, Tribological properties of the AZ91D magnesium
alloy hardened with silicon carbide and by severe plastic deformation, J. Frict.
Wear 30 (2009) 194198.
[29] S. Bera, Z. Zuberova, R.J. Hellmig, Y. Estrin, I. Manna, Synthesis of copper alloys
with extended solid solubility and nano Al2O3 dispersion by mechanical
alloying and equal channel angular pressing, Philos. Mag. 90 (2010)
14651483.
[30] D. Nickel, G. Alisch, H. Podlesak, M. Hockauf, G. Fritsche, T. Lampke, Microstructure, corrosion and wear behavior of UFG-powder-metallurgical AlCu
alloys, AlC/Al2O3(p) and AlCu/SiC(p) composites, Rev. Adv. Mater. Sci. 25
(2010) 261269.
[31] S. El-Hadad, H. Sato, Y. Watanabe, Anisotropic mechanical properties of equal
channel angular pressed Al5% Zr alloy containing platelet particles, Mater.
Sci. Eng. A 527 (2010) 46744679.
[32] S. El-Hadad, H. Sato, Y. Watanabe, Investigation of wear anisotropy in a
severely deformed AlAl3Ti composite, Metall. Mater. Trans. A 43 (2012)
32493256.
[33] H. Sato, S. Elhadad, O. Sitdikov, Y. Watanabe, Effect of processing routs on wear
property of AlAl3Ti alloys severely deformed by ECAP, Mater. Sci. Forum 584
586 (2008) 971976.
[34] C.T. Wang, N. Gao, R.J.K. Wood, T.G. Langdon, Wear behavior of an aluminum
alloy processed by equal-channel angular pressing, J. Mater. Sci. 46 (2011)
123130.

K. Edalati et al. / Wear 310 (2014) 8389

[35] C.T. Wang, N. Gao, R.J.K. Wood, T.G. Langdon, Wear behavior of Al-1050 alloy
processed by severe plastic deformation, Mater. Sci. Forum 667-669 (2011)
11011106.
[36] G. Purcek, O. Saray, O. Kul, I. Karaman, G.G. Yapici, M. Haouaoui, H.J. Maier,
Mechanical and wear properties of ultrane-grained pure Ti produced by
multi-pass equal-channel angular extrusion, Mater. Sci. Eng. A 517 (2009)
97104.
[37] K.S. Suresh, M. Geetha, C. Richard, J. Landoulsi, H. Ramasawmy, S. Suwas,
R. Asokamani, Effect of equal channel angular extrusion on wear and corrosion
behavior of the orthopedic Ti13Nb13Zr alloy in simulated body uid, Mater.
Sci. Eng. C 32 (2012) 763771.
[38] Y.S. Kim, J.S. Ha, W.J. Kim, Dry sliding wear characteristics of severely
deformed 6061 aluminium and AZ61 magnesium alloys, Mater. Sci. Forum
449-452 (2004) 597600.
[39] Y.S. Kim, H.S. Yu, D.H. Shin, Low sliding-wear resistance of ultrane-grained Al
alloys and steel having undergone severe plastic deformation, Int. J. Mater.
Res. 100 (2009) 871874.
[40] V.V. Stolyarov, L.S. Shuster, M.S. Migranov, R.Z. Valiev, Y.T. Zhu, Reduction of
friction coefcient of ultrane-grained CP titanium, Mater. Sci. Eng. A 371
(2004) 313317.
[41] A.P. Zhilyaev, I. Shakhova, A. Belyakov, R. Kaibyshev, T.G. Langdon, Wear
resistance and electroconductivity in copper processed by severe plastic
deformation, Wear 305 (2013) 8999.
[42] M.I.A.E. Aal, H.S. Kim, Wear properties of high pressure torsion processed
ultrane grained Al-7%Si alloy, Mater. Des. 53 (2014) 373382.
[43] C.T. Wang, N. Gao, M.G. Gee, R.J.K. Wood, T.G. Langdon, Effect of grain size on
the micro-tribological behavior of pure titanium processed by high-pressure
torsion, Wear 280-281 (2012) 2835.
[44] S. Faghihi, D. Li, J.A. Szpunar, Tribocorrosion behaviour of nanostructured
titanium substrates processed by high-pressure torsion, Nanotechnolgy 21
(2010) 485703.
[45] C.T. Wang, A. Escudeiro, T. Polcar, A. Cavaleiro, R.J.K. Wood, N. Gao,
T.G. Langdon, Indentation and scratch testing of DLC-Zr coatings on
ultrane-grained titanium processed by high-pressure torsion, Wear 306
(2013) 304310.
[46] C.T. Wang, N. Gao, M.G. Gee, R.J.K. Wood, T.G. Langdon, Processing of an
ultrane-grained titanium by high-pressure torsion: an evaluation of the wear
properties with and without a TiN coating, J. Mech. Behav. Biomed. 17 (2013)
166175.
[47] C.T. Wang, N. Gao, M.G. Gee, R.J.K. Wood, T.G. Langdon, Tribology testing of
ultrane-grained Ti processed by high-pressure torsion with subsequent
coating, J. Mater. Sci. 48 (2013) 47424748.
[48] H. Kato, Y. Todaka, M. Umemoto, K. Morisako, M. Hashimoto, M. Haga,
Dry sliding wear properties of sub-microcrystalline ultra-low carbon steel
produced by high-pressure torsion straining, Mater. Trans. 53 (2012) 128132.
[49] N.I. Noskova, L.G. Korshunov, A.V. Korznikov, Microstructure and tribological
properties of AlSn, AlSnPb and SnSbCu alloys subjected to severe plastic
deformation, Mater. Sci. Heat Treat. 50 (2008) 593599.

89

[50] E. Darmiani, I. Danaee, M.A. Golozar, M.R. Toroghinejad, A. Ashra, A. Ahmadi,


Reciprocating wear resistance of AlSiC nano-composite fabricated by accumulative roll bonding process, Mater. Des. 50 (2013) 497502.
[51] M. Eizadjou, H. Danesh Manesh, K. Janghorban, Microstructure and mechanical properties of ultra-ne grains (UFGs) aluminum strips produced by ARB
process, J. Alloys Compd. 474 (2009) 406415.
[52] M. Eizadjou, A. Kazemi Talachi, H. Danesh Manesh, K. Janghorban, Sliding
wear behavior of severely deformed 6061 aluminum alloy by accumulative
roll bonding (ARB) process, Mater. Sci. Forum 667669 (2011) 11071112.
[53] A. Kazemi Talachi, M. Eizadjou, H. Danesh Manesh, K. Janghorban, Wear
characteristics of severely deformed aluminum sheets by accumulative roll
bonding (ARB) process, Mater. Charact. 62 (2011) 1221.
[54] Y.S. Kim, T.O. Lee, D.H. Shin, Microstructural evolution and mechanical
properties of ultrane grained commercially pure 1100 aluminum alloy
processed by accumulative roll-bonding (ARB), Mater. Sci. Forum 449-452
(2004) 625628.
[55] Y.S. Kim, J.S. Ha, D.H. Shin, Sliding wear characteristics of ultrane-grained
nan-strain-hardening aluminum-magnesium alloys, Mater. Sci. Forum 475
479 (2005) 401404.
[56] N. Gao, C.T. Wang, R.J.K. Wood, T.G. Langdon, Wear resistance of SPDprocessed alloys, Mater. Sci. Forum 667669 (2011) 10951100.
[57] N. Gao, C.T. Wang, R.J.K. Wood, T.G. Langdon, Tribological properties of
ultrane-grained materials processed by severe plastic deformation, J. Mater.
Sci. 47 (2012) 47794797.
[58] K. Edalati, Z. Horita, T. Furuta, S. Kuramoto, Dynamic recrystallization and
recovery during high-pressure torsion: experimental evidence by torque
measurement using ring specimens, Mater. Sci. Eng. A 559 (2013) 506509.
[59] S.M. Kuo, D.A. Rigney, Sliding behavior of aluminum, Mater. Sci. Eng. A 157
(1992) 131143.
[60] T. Tokunaga, K. Kaneko, Z. Horita, Production of aluminummatrix carbon
nanotube composite using high pressure torsion, Mater. Sci. Eng. A 490 (2008)
300304.
[61] A. Bachmaier, A. Hohenwarter, R. Pippan, New procedure to generate stable
nanocrystallites by severe plastic deformation, Scr. Mater. 61 (2009)
10161019.
[62] K. Edalati, Z. Horita., H. Fujiwara, K. Ameyama, Cold consolidation of ballmilled titanium powders using high-pressure torsion, Metall. Mater. Trans. A
41 (2010) 33083317.
[63] M. Ashida, Z. Horita, Effects of ball milling and high-pressure torsion for
improving mechanical properties of AlAl2O3 nanocomposites, J. Mater. Sci. 47
(2012) 78217827.
[64] T. Reye, Zur theorie der zapfenreibung, Der Civilingenieur 4 (1860) 235255.
[65] R. Holm, Electrical Contacts, H. Gerber Pubblications, Stockholm, 1946.
[66] J.F. Archard, Contact and rubbing of at surfaces, J. Appl. Phys. 24 (1953)
981988.
[67] J.I. Song, K.S. Han, Effect of volume fraction of carbon bers on wear behavior
of Al/Al2O3/C hybrid metal matrix composites, Compos. Struct. 39 (1997)
309318.
[68] D.A. Rigney, M.G.S. Naylor, R. Divakar, Low energy dislocation structures
caused by sliding and by particle impact, Mater. Sci. Eng. 81 (1986) 409425.

S-ar putea să vă placă și