Sunteți pe pagina 1din 9

Wear 290–291 (2012) 1–9

Contents lists available at SciVerse ScienceDirect

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

Three body abrasion of laser surface alloyed aluminium AA1200


L.A.B Mabhali a,b, N. Sacks a,n, S. Pityana b
a
School of Chemical and Metallurgical Engineering, University of the Witwatersrand, Private Bag 3, Wits 2050, South Africa
b
Council for Scientific and Industrial Research, National Laser Centre, PO Box 375, Pretoria 0001, South Africa

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

Article history: Laser surface alloying of aluminium AA1200 was performed with a 4 kW Nd:YAG laser to improve the
Received 17 June 2011 abrasion wear resistance. Aluminium surfaces reinforced with metal matrix composites and inter-
Received in revised form metallic phases were achieved. The phases present depended on the composition of the alloying
30 May 2012
powder mixture. The wear performance of the alloyed surfaces was characterised using an ASTM G65
Accepted 31 May 2012
three body dry abrasion apparatus. A maximum 82% improvement in the wear resistance of the pure
Available online 9 June 2012
aluminium was achieved with a 40 wt% Ni þ 20 wt% Ti þ40 wt% SiC composition. The three alloys which
Keywords: had the best wear resistance were all produced with a composition of 40 wt% SiC and Ti and Ni powders
Laser surface alloying ranging from 20 to 40 wt%. No direct correlation was observed between hardness and wear resistance.
Aluminium
Microstructural examination showed that the main wear mechanisms were intense plastic deformation
Metal-matrix composites
with micro-fracture of the SiC particles and intermetallic phases. The wear behaviour is mainly
Intermetallics
Three body-body abrasion determined by the response of the different alloy phases, either independently or in combination, to the
action of the abrasive particles and the precise nature of this response is complex and requires
further study.
& 2012 Elsevier B.V. All rights reserved.

1. Introduction several authors [1–22]. Staia et al. [8] alloyed aluminium A356
with 96 wt% WC, 2 wt% Ti and 2 wt% Mg. The wear behaviour of
Aluminium is widely used in industry due to its attractive the alloyed material against AISI52100 steel balls under a load of
attributes such as low density, high strength to weight ratio, high 5 N was investigated. For high laser scanning speeds, large WC
thermal conductivity and good formability [1]. Its drawback is the particles were formed which served as the load carrying particles
poor surface properties such as hardness and wear resistance. and severely abraded the steel. For low laser scanning speeds, the
These surface properties can be improved by depositing hard WC particle size decreased and the wear mechanisms changed as
reinforcement materials on its surface in an attempt to provide a the aluminium matrix participated in the transference process.
composite material which has a tough matrix reinforced with The wear mechanisms of the aluminium A356 was adhesive with
hard particles which should lead to improved wear resistance. high quantities of aluminium transferred to the steel counterface.
This can be achieved by a laser alloying process. Alloying Almeida et al. [9] studied the dry sliding wear mechanisms of
materials are deposited as powders into a melt pool generated Al–Mo deposited on an aluminium substrate by laser surface
on the material surface by a focused laser beam. The beam is alloying. The wear mechanisms were predominantly adhesion
scanned over the part and the deposited material resolidifies followed by material detachment and transfer, oxidation and
resulting in good bonding between the substrate and the alloyed some abrasion, mainly by hard intermetallic compound particles
layer. The surface of the material is modified by changing the on the steel counterbody. Elleuch et al. [10] studied the abrasive
composition and microstructure without affecting the bulk prop- wear of aluminium alloys rubbed against sand. The author
erties of the material. Process parameters such as laser power, reported that the wear rate increased by three times as the
laser beam spot size, laser scanning speed and powder feed rate incident angle of sand was increased from 01 to 451.
have to be controlled to achieve the desired metallurgical bonding S- ahin [11] studied the abrasive wear of aluminium AA2014
and alloyed surface properties. reinforced with SiC particles of different sizes (9 mm, 14 mm and
Research on the wear properties of aluminium alloys rein- 33 mm). The authors observed that the wear rate decreased with
forced with either ceramic or metallic materials has been done by increasing hardness of the MMC while it increased with increas-
ing abrasive particle size of the counter surface, applied load and
sliding distance (up to a maximum value, then either decreased or
n
Corresponding author. remained constant). Miyajima and Iwai [12] reported that SiC and
E-mail address: natasha.sacks@wits.ac.za (N. Sacks). Al2O3 reinforcement caused severe wear of a steel counter surface

0043-1648/$ - see front matter & 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.wear.2012.05.034
2 L.A.B Mabhali et al. / Wear 290–291 (2012) 1–9

during sliding wear of the aluminium matrix composite on a pin- increasing volume fractions of the intermetallic compounds. The
on-disc wear tester. Metal matrix composites with low volume wear resistance of the alloys were tested under dry sliding wear
fractions of embedded ceramics produced severe wear charac- conditions using applied loads of 0.15 and 1 N respectively. The
terised by plastic deformation and large grooves. MMCs with high wear resistance was found to increase with an increase in the
volume fractions of embedded ceramics did not wear severely volume fraction of the intermetallic compounds. However higher
and the surfaces were smooth and flat without large grooves. The wear rates were observed at the lower load. This result was
embedded ceramic acted as inhibitors against plastic flow and attributed to the resulting wear mechanisms which included
adhesion of the matrix metal to the steel counter surface. adhesion, oxidation as well as detachment and transfer of the
Majumdar et al. [13] laser alloyed Ti with Si, Ti with Al, and Ti surface material. Almeida et al. [18] reported that the addition of
with SiþAl and studied the wear behaviour on a ball-on-disc niobium to aluminium provided a good combination of a very
wear testing machine. Surfaces alloyed with Al suffered extensive high hardness and moderate toughness which was attributed to
amounts of abrasive and adhesive wear while surfaces alloyed the volume fraction and dendritic spacing of the Al3Nb phase.
with Si þAl suffered abrasive wear. No significant amount of Several authors have attempted to model the wear behaviour of
adhesive or abrasive wear was observed in samples alloyed only metal matrix composites and general multiphase materials as some
with Si. This decrease in the wear rates was associated with the of the prevailing wear theories do not provide adequate insight into
formation of the Ti5Si3 phase which was found in higher volume multiphase material behaviour under abrasive wear conditions
fractions in the surfaces alloyed only with Si. [16,23–28]. Recently Colaco and Vilar [16] developed a mathema-
Anandkumar et al. [15] conducted three body abrasion wear tical model to provide understanding of the functional dependence
tests on Al–Si/SiC composite coatings using SiC as the abrasive. of wear resistance on the properties of the reinforcement phases in
The coatings were produced using two different specific energies, metal matrix composites. The model was based on the Rabinowicz
namely 26 MJ/m2 and 58 MJ/m2 which yielded hardness values of approach [29] on defining abrasive wear mechanisms of multiphase
120 and 250 VH respectively. The higher hardness was attributed materials and also considered the mechanisms proposed by Hutch-
to the larger proportions of Al4SiC4 and Si precipitates. However ings [28] and Zum-Ghar [24] which reported that material loss
the harder coating had the highest wear rate. The authors occurs by wear of the matrix, wear of the reinforcement particles
concluded that this was due to the significant difference between and extraction of the reinforcement particles from the matrix. The
the SiC abrasive and the composite hardness. The lower wear rate model was tested against experimental data based on three body
achieved by the softer coating was attributed to the presence of abrasion wear of Fe-0.25%C-15%Cr coatings reinforced with different
SiC particles, which although do not provide an improvement in volume fractions of Nb. The authors reported a good correlation
hardness, do resist scratching by the abrasive. The friction and between the model and experimental data and provided detailed
wear behaviour of aluminium laser surface alloyed with SiC and understanding of the functional dependence of the wear resistance
SiCþAl was also studied by Majumdar et al. [21,22]. It was on the proportion of reinforcement particles. The main conclusions
found that some the SiC dissociated to form Al4C3 which led showed that while the average hardness increases as the volume
to an increase in the aluminium hardness from 25 VHN to fraction of reinforcement increases, increased wear rates are possi-
200–250 VHN. The resulting wear resistance was improved up ble when the volume fraction of reinforcement is increased where
to three times and was attributed to the dispersion strengthening cracking and extraction of the particles may increase due to
effect provided by the Al4C3 phase. interaction with the abrasive.
Shipway et al. [14] studied the sliding wear mechanisms of Published work is generally limited to higher aluminium alloys
aluminium reinforced with TiC particles against a carbon-man- (i.e. 2xxx to 7xxx series). To the authors knowledge no published
ganese steel (BS 080A15) counterface and found that the wear work is available on the abrasive wear of Aluminium AA1200
rates increased as the applied load increased and delamination which has been laser alloyed with ceramic and metallic materials
cracks were observed in the MMC layer. The addition of the TiC simultaneously. Therefore the aim of this work was to investigate
particles resulted in counter surface wear as hard particles acted the abrasive wear resistance of Aluminium AA1200 by laser
as abrasives in the sliding process. The wear rates of the steel surface alloying with mixed Ni, Ti and SiC powders. Intermetallic
counter surfaces increased as the volume fraction of the TiC phases are formed when aluminium is laser alloyed with metallic
particles increased. These TiC particles resulted in ploughing materials (e.g. Ni and Ti) while metal matrix composites are
and cutting of the steel counter surfaces in the sliding direction. formed when alloying with ceramics (e.g. SiC).
In dry sliding wear tests of Al–12Si/TiB2 laser clads with an
AISI 440C tool steel counterbody by Anandkumar et al. [17] the
laser clads displayed ultra mild wear. The addition of TiB2 led to 2. Material and methods
an improvement in the hardness and wear resistance of the
Al–12Si alloy. The authors attributed this to the role played Laser alloying of the aluminium AA1200 surface was per-
by the reinforcement phase in protecting the Al-based alloy formed with a Nd:YAG laser. Aluminium AA1200 plates were
against severe plastic deformation. Similar results were found initially sand blasted to enhance the absorption of laser energy by
by Majumdar et al. [20] who studied the abrasion wear resistance the aluminium substrate. The chemical composition of the alu-
of aluminium reinforced with TiB and TiB2 particles where the minium plates was 0.12 wt% Cu, 0.13 wt% Si, 0.59 wt% Fe and the
volume fractions were varied between 7 and 18%. The addition of balance was Al. The laser parameters used were 4 kW of power, a
the TiB improved the hardness threefold using a laser power of beam diameter of 4 mm and a scanning speed of 10 mm/s. These
1.2 kW and a scan speed of 500 mm/min. The wear resistance was parameters ensured that sufficient laser energy was supplied for
improved and decreased with an increase in applied load. This the dissolution of the powders. The alloying powder mixture
was attributed to the increased hardness provided by the TiB consisted of different compositions of Ni, Ti and SiC powder
particles. mixtures as listed in Table 1. The powder feed rate was 2.5–3 g/
Molybdenum has been shown to improve the mechanical min (depending on the powder composition) which ensured
properties of aluminium. In a study done by Almeida et al. [9] sufficient supply of powder during the alloying experiments.
on a range of Al–Mo alloys (14.8–19.1 wt% Mo) the hardness was Argon was used as the carrier and shielding gas to prevent
shown to increase from 85 to 160 VH while Young’s Modulus oxidation during the alloying process. Ten overlapping tracks
increased from 84 to 92 GPa. These increases were attributed to with a track overlap of 50% were made.
L.A.B Mabhali et al. / Wear 290–291 (2012) 1–9 3

Table 1 Table 2
Starting powder compositions. Vickers hardness of the alloyed surfaces.

Sample no. Powder composition Powder composition Hardness (HV1)

Al Untreated aluminium AA1200 Untreated aluminium AA1200 24.03 7 0.27


1 10 wt% Ni þ70 wt% Tiþ 20 wt% SiC 10 wt% Ni þ 70 wt% Ti þ20 wt% SiC 143.697 9.32
2 20 wt% Ni þ30 wt% Tiþ 50 wt% SiC 20 wt% Ni þ 30 wt% Ti þ50 wt% SiC 149.537 10.10
3 20 wt% Ni þ40 wt% Tiþ 40 wt% SiC 20 wt% Ni þ 40 wt% Ti þ40 wt% SiC 172.387 13.63
4 30 wt% Ni þ30 wt% Tiþ 40 wt% SiC 30 wt% Ni þ 30 wt% Ti þ40 wt% SiC 184.487 6.92
5 30 wt% Ni þ40 wt% Tiþ 30 wt% SiC 30 wt% Ni þ 40 wt% Ti þ30 wt% SiC 192.217 8.19
6 33.3 wt% Ni þ 33.3 wt% Tiþ33.3 wt% SiC 33.3 wt% Ni þ33.3 wt% Ti þ33.3 wt% SiC 197.437 7.53
7 40 wt% Ni þ20 wt% Tiþ 40 wt% SiC 40 wt% Ni þ 20 wt% Ti þ40 wt% SiC 198.197 13.49
8 40 wt% Ni þ30 wt% Tiþ 30 wt% SiC 40 wt% Ni þ 30 wt% Ti þ30 wt% SiC 218.90 7 6.41
9 40 wt% Ni þ40 wt% Tiþ 20 wt% SiC 40 wt% Ni þ 40 wt% Ti þ20 wt% SiC 239.687 10.28
10 50 wt% Ni þ20 wt% Tiþ 30 wt% SiC 50 wt% Ni þ 20 wt% Ti þ30 wt% SiC 250.10 7 7.75
11 60 wt% Ni þ30 wt% Tiþ 10 wt% SiC 60 wt% Ni þ 30 wt% Ti þ10 wt% SiC 264.277 9.39
12 70 wt% Ni þ10 wt% Tiþ 20 wt% SiC 70 wt% Ni þ 10 wt% Ti þ20 wt% SiC 290.42 7 11.86
13 70 wt% Ni þ20 wt% Tiþ 10 wt% SiC 70 wt% Ni þ 20 wt% Ti þ10 wt% SiC 295.817 11.18
14 80 wt% Ni þ15 wt% Tiþ 5 wt% SiC 80 wt% Ni þ 15 wt% Tiþ 5 wt% SiC 318.647 9.13

and phases of the alloyed surfaces. Metal matrix composites


reinforced with SiC particles and intermetallic phases were
formed in the alloyed layers. The predominant intermetallic
phases were Al3Ni, Al3Ni2, Al3Ti, TiC and Ti5Si3. The Al3Ni, Al3Ni2
and Al3Ti phases were formed due to the reaction of Al with Ni
and Ti. The dissociation of SiC particles and reactions with Ti
resulted in the formation of TiC and Ti5Si3 phases. In alloys with a
high SiC content, Al4C3, Al4SiC4, Ti3SiC2 and Al3TiC2 intermetallic
phases were also formed from the reaction of the dissociated SiC
with either Al or Ti. A comprehensive study of the microstructures
produced by each powder composition and the methods used for
phase identification (EDS, XRD, phase diagrams, thermodynamic
calculations) can be found in Ref. [31]. The hardness results show
that laser alloying improved the surface hardness for all the
compositions used. Grain refinement, due to the rapid heating
Fig. 1. Wear response of untreated aluminium and laser alloyed surfaces. and cooling rates associated with laser alloying plays a role in
increasing the hardness of the laser alloyed surfaces [32]. The
The hardness was measured using a 1 kgf load and the average highest hardness was obtained when alloying with high Ni
of eleven indentations is reported as the surface hardness. The contents.
microstructures of the alloyed surfaces were studied using scan- Fig. 1 shows the typical wear behaviour of the materials during
ning electron microscopy (SEM) and x-ray diffraction (XRD). the wear test. The wear response of all the alloys was initially
Samples were ground and polished to a 0.04 mm (colloidal silica high during the first 10 min of testing followed by a levelling off
OP-S suspension) surface finish and etched with Keller’s reagent of the wear rate for the remainder of the test. Fig. 2 illustrates the
(3 ml HClþ2 ml HFþ5 ml HNO3 þ190 ml HO2) prior to micro- wear rates of all the alloyed surfaces after 60 min of wear. A
structural characterisation. reduction in wear rate (or improved wear resistance) was found
Three body abrasion wear tests were conducted using an ASTM in all the alloyed surfaces with the maximum reduction in wear of
G65-04 [30] dry sand rubber wheel apparatus shown in Fig. 1. Silica 82% observed when laser alloying with 40 wt% Niþ20 wt%
sand obtained from Rolfes (Pty) Ltd. with a particle size range of Tiþ40 wt% SiC. No correlation was observed between wear
0.3–0.65 mm was used as the abrasive. Prior to wear testing, the silica response and hardness. The 80 wt% Niþ15 wt% Tiþ5 wt% SiC
sand was sieved with a MACSALAB electronic sieve shaker to alloy which had the highest hardness did not show the best wear
determine the particle size and distribution. Majority of the sand resistance. This alloy has the ninth highest wear rate of the fifteen
particles were in the 500–600 mm range with a D50 of 525 mm and materials. The best wear resistance was shown by the alloy which
were generally angular in shape. The sand feed rate was 2.3 g/s. This only had the eighth highest hardness.
feed rate ensured that sufficient sand was introduced into the rubber From the SEM images of the worn surfaces similar features
wheel/specimen contact area in the test chamber. The samples, were found on the wear scars of all the alloys. The degree to
70  20  6 mm in size, were polished to a 1 mm surface finish prior which the different types of mechanisms occurred depended on
to the wear tests. The applied load was kept constant at 9.8 N. The the alloying composition and phases present. Figs. 3–6 illustrate
duration of each wear test was 60 min and the mass of the samples the typical wear scars observed.
was recorded at 10 min intervals. Three tests were conducted to Fig. 3 shows the wear scars on the pure Al. Due to the high
determine the average for each alloy. The wear scars were examined ductility of aluminium, some of the fragmented SiO2 particles
using a JEOL JSM-840 scanning electron microscope. were embedded in the deformed microstructure. During testing
the rubber wheel pushes the sand against the specimen leading to
point indentation of the SiO2 particles on the aluminium surface
3. Results under the applied load. Fig. 3(a) and (b) shows plastic deforma-
tion of the surface which is characterised by non-uniform ridges
Laser alloying was successfully performed with the laser and plastic flow. The extent of plastic deformation on the surface
processing parameters used. Tables 2 and 3 list the hardness is also evident on the cross-sections of the aluminium shown in
4 L.A.B Mabhali et al. / Wear 290–291 (2012) 1–9

Table 3
Phases identified in the alloyed surfaces.

Powder composition Phases observed

10 wt% Ni þ 70 wt% Tiþ 20 wt% SiC a-Al, SiC, TiC, Al3Ni, Al3Ti and Ti5Si3
20 wt% Ni þ 30 wt% Tiþ 50 wt% SiC a-Al, SiC, TiC, Al3Ni, Al3Ni2, Al3Ti, Ti5Si3, Al3TiC2 and Al4SiC4
20 wt% Ni þ 40 wt% Tiþ 40 wt% SiC a-Al, Si, SiC, TiC, Al3Ni, Al3Ti, Ti5Si3, Ti3SiC2 and Al4SiC4
30 wt% Ni þ 30 wt% Tiþ 40 wt% SiC a-Al, Si, SiC, TiC, Al3Ni, Al3Ni2, Al3Ti, Ti5Si3, Ti3SiC2 and Al4SiC4
30 wt% Ni þ 40 wt% Tiþ 30 wt% SiC a-Al, Si, SiC, TiC, Al3Ni, Al3Ni2, Al3Ti, Ti5Si3, Ti3SiC2 and Al4SiC4
33.3 wt% Ni þ 33.3 wt% Tiþ 33.3 wt% SiC a-Al, SiC, TiC, Al3Ni, Al3Ni2, Al3Ti and Ti5Si3
40 wt% Ni þ 20 wt% Tiþ 40 wt% SiC a-Al, SiC, TiC, Al3Ni, Al3Ni2, Al3Ti, Ti5Si3, Ti3SiC2 and Al4SiC4
40 wt% Ni þ 30 wt% Tiþ 30 wt% SiC a-Al, SiC, TiC, Al3Ni, Al3Ni2, Al3Ti, Ti5Si3 and Ti3SiC2
40 wt% Ni þ 40 wt% Tiþ 20 wt% SiC a-Al, SiC, TiC, Al3Ni, Al3Ni2, Al3Ti, Ti5Si3 and Ti3SiC2
50 wt% Ni þ 20 wt% Tiþ 30 wt% SiC a-Al, SiC, TiC, Al3Ni, Al3Ni2, Al3Ti, Ti5Si3 and Al4C3
60 wt% Ni þ 30 wt% Tiþ 10 wt% SiC a-Al, SiC, TiC, Al3Ni, Al3Ni2, Al3Ti and Ti5Si3
70 wt% Ni þ 10 wt% Tiþ 20 wt% SiC a-Al, SiC, TiC, Al3Ni, Al3Ni2, Al3Ti, Ti5Si3 and Al4C3
70 wt% Ni þ 20 wt% Tiþ 10 wt% SiC a-Al, SiC, Al3Ni, Al3Ni2 and Al3Ti
80 wt% Ni þ 15 wt% Ti þ5 wt% SiC a-Al, SiC, Al3Ni, Al3Ni2 and Al3Ti

abrasion wear conditions. A maximum reduction in wear rate of


82% (compared to pure Al) was observed when laser alloying with
a 40 wt% Niþ20 wt% Tiþ40 wt% SiC powder composition. This
enhanced surface abrasive wear performance may be attributed
to a good combination of the tough aluminium matrix reinforced
with hard metal matrix composites and intermetallic and ceramic
phases.
One of the main results of the study was that no direct
correlation was observed between hardness and wear resistance.
Similar trends were found by others [16]. The alloy with the
highest hardness namely, the 80 wt% Niþ 15 wt% Ti þ5 wt% SiC
composition had the ninth highest wear rate of the fifteen
materials tested, while the best wear resistance was shown by
the alloy which only had the eighth highest hardness. The general
Fig. 2. Wear rate of untreated and laser alloyed aluminium surfaces. Numbers increase in hardness of the alloys is attributed to the presence of
correspond to alloy compositions listed in Table 1. the carbide (TiC, Al4SiC4,) and aluminide (Al3Ni, Al3N2) phases. TiC
has a hardness range of 2000–3000 HV [33] and was present in
alloys having both high hardness and high wear resistance, except
Fig. 3(c) and (d) which also highlights the extent and depth to for the two hardest alloys which showed no traces of TiC. The
which the fragmented SiO2 abrasive particles were embedded in aluminides, Al3Ni and Al3Ni2, have hardnesses of 732 HV and
the deformed aluminium. A few micro-cracks close to the surface 1013 HV respectively [34] and were present in most alloys. One of
indicate that material removal occurred by progressive detach- the main phase differences in the five alloys which had the best
ment of the deformed material. wear resistance and the five alloys which had the highest hard-
The worn surfaces of the alloyed layers was characterised by ness was the presence of the Ti3SiC2 and Al4SiC4 phases in the
plastic deformation mechanisms, fracture of the embedded SiC alloys with the best wear resistance; these phases were not
particles and micro-fracture of the intermetallic phases as pre- present in alloys having high hardness values. Ti3SiC2 is known
sented in Figs. 4–6. The plastic deformation observed in Figs. 4 and to have a low hardness, but has a high Young’s modulus and a
5 is characterised by a non-uniform ridge structure where partial high damage tolerance [35] while Al4SiC4 has a hardness of
delamination (Fig. 5(c)) of the deformed surface layers can also be 1200 HV [15]. The presence of Al3Ti in all the alloys also led to
seen. Random paths of smearing are also present; these paths were a strength increase in the alloys [36]. Despite the high hardness
likely caused by the movement of wear debris during the wear test. values of some of the individual phases the overall average
Sub-surface plastic deformation occurred in the laser alloyed hardness of the alloyed layers was in the range of 144–319 HV1.
materials. This was evident in the cross-sections of the worn Detailed analysis and discussion of the phases produced in the
surfaces where distortions in the microstructures were observed, alloyed surfaces and their contribution to hardness can be found
for example uniform changes in grain orientation in the direction in references [31,37]. For the current discussion on the response
of motion of the abrasive. This deformation is caused by the cyclic of the materials to abrasive wear it is sufficient to state that the
action of the abrasive over the surface under the influence of the test results indicate that there is not a linear relationship between
applied load. Lateral cracking seen in the cross-sectional views was alloy hardness and wear rate and therefore cannot be used as a
transgranular and contributed to delamination of the worn surface primary parameter to control the wear resistance of these alloys.
layer as seen in Fig. 6(a) and (b). Fracturing of the SiC particles The wear behaviour is mainly determined by the response of the
(Fig. 6(c–e)) and transgranular cracking of the intermetallic matrix different alloy phases, either independently or in combination, to
(Fig. 6(f)) was also observed. the action of the abrasive particles.
It has been shown that there is a direct correlation between
the microstructural properties, bonding and volume fraction of
4. Discussion the reinforcement particles in metal matrix composites and
intermetallic phases and their relationship to abrasive wear
A reduction in wear rate was observed for the laser alloyed [16,24,33]. The influence of volume fraction of reinforcement
surfaces compared to the pure aluminium metal under three body phase on the abrasive wear has been studied by several authors;
L.A.B Mabhali et al. / Wear 290–291 (2012) 1–9 5

Fig. 3. Worn surfaces of the untreated aluminium in (a) and (b) plan view; (c) and (d) cross-sections showing gross plastic deformation as well as embedded SiO2 abrasive
particles.

Fig. 4. SEM plan view images of wear scars of the laser alloyed surfaces showing typical features observed in all the alloyed layers. Images taken at low magnifications.
6 L.A.B Mabhali et al. / Wear 290–291 (2012) 1–9

Fig. 5. SEM plan view images of wear scars of the laser alloyed surfaces showing typical features observed in all the alloyed layers. Images taken at high magnifications.

however contradictory results have been obtained. Some authors the latter case the wear rates are influenced by the brittle
reported that as the volume fraction increased the wear resis- behaviour of the particles and matrix, the hardness of the
tance increased [23], while others observed the opposite trend reinforcements and the abrasive as well as the volume fraction
[24]. In an attempt to address these contradictions Colaco and of reinforcements present. These factors influence the wear
Vilar [16] developed a model which showed that while the mechanisms which will prevail and determine the resulting
increase in volume fraction of reinforcement phase led to wear rates.
increased hardness, the wear rates were influenced by the action In the current study the worn surfaces did not display parallel
of the abrasive on the reinforcements which can cause cracking grooving which is typically observed under abrasion wear testing.
and detachment of the reinforcements which may subsequently Rather gross plastic deformation as well as brittle fracture of the
lead to increased wear rates. SiC and intermetallic phases was observed. Thus the dimension of
To a certain extent these factors determine if the alloyed damage done by the SiO2 abrasive could not be calculated. To
aluminium surfaces would respond homogenously or heteroge- accurately assess this aspect it should be possible to conduct
neously under the abrasive wear test conditions. The type of scratch tests on the surfaces and to analyse the resulting wear
response expected has been related to a comparison between the behaviour. This was not attempted in the current study. A partial
size of the reinforcement particles and the dimension of damage understanding of the material response can be gained by studying
caused by the abrasive [24,28]. If the sizes are similar then a the worn surfaces and comparing properties between the phases
heterogeneous response is expected and the hardness properties present and the abrasive used. The largest reinforcement particle
of abrasive and reinforcement will determine the wear rates. If size present was that of the retained SiC particles which had a
the size of the reinforcement particles is much smaller than the maximum size of at least 50 mm. The average SiO2 abrasive
dimensions of the damage then the response is homogenous. In particle size was 525 mm. On this basis it is likely that the alloyed
L.A.B Mabhali et al. / Wear 290–291 (2012) 1–9 7

Fig. 6. SEM images of cross-sections of the worn surfaces of the laser alloyed surfaces showing fractured SiC grains and micro-fracture of the intermetallic phases.

layers responded homogenously during wear. The precise nature 1.2 boundary condition and the wear rates are then influenced by
of the response in terms of linking increasing wear rates to the ability of the SiO2 abrasive to plastically deform, fracture or
particle hardness, reinforcement volume fraction and phase cause detachment of the particles from the matrix and would be
brittleness is more complex due to the wide range of reinforce- strongly influenced by the volume fraction of each reinforcement
ment phases present. phase present, the matrix properties and the applied stress during
If the alloyed surfaces were treated as having a ‘pure’ homo- testing. The applied stress was calculated to be 0.02 MPa but the
geneous response and the hardness of the SiO2 abrasive, 800 HV actual pressure is expected to be higher than this as the real
[33], is compared to the composite hardness, which ranged from contact area between individual abrasive particles and the alloyed
144 to 319 HV1, the ratio of the hardness of the abrasive particles surface is lower than the surface area of 420 mm2 used to
to the hardness of the alloyed surfaces (Ha/Hm) is in the range of calculate the nominal stress. For each alloyed layer produced
2.5–5.6. These calculated ratios are greater than 1.2 which sig- the strength properties of the specific phases (matrix and rein-
nifies the boundary condition between soft and hard abrasion as forcement) present, would determine that phase’s response to the
defined by Hutchings [33] and would therefore place the abrasive applied pressure. This would lead to different wear mechanisms
wear of the alloyed surfaces in the hard abrasion category. This being experienced by the different phases and ultimately result in
classification would confer with the extent of hard abrasion a variety of measured wear rates.
damage mechanisms observed on the wear scars of the alloyed The influence of increasing reinforcement volume fraction
surfaces. If the individual reinforcement hardnesses are used, showed a variable wear response and cannot be easily quantified
then the Ha/Hm ratios span a range of values on either side of the as the alloying compositions were not kept constant for a wide
8 L.A.B Mabhali et al. / Wear 290–291 (2012) 1–9

range of compositions. However a few deductions can be made. particles were observed. The Ti containing and the Al–SiC inter-
When the Ni volume fraction was kept constant at 40 wt%, the metallic phases resulted in large Al mean free paths. This
wear rate increased as the Ti volume fraction increased along promoted wear by plastic deformation. In a few instances the
with a simultaneous decrease in the SiC volume fraction. When hard phases, e.g. the equiaxed dendritic Al3Ni2 grains and the SiC
the Ti volume fraction was kept constant at 40 wt%, the wear rate particles, were also seen to interrupt plastic deformation. Fracture
increased with a decrease in SiC content, but no linear effect was of the SiC particles and intermetallic phases can be reduced if
observed as the Ni volume fraction was simultaneously narrower distributions of starting powders were used to ensure
decreased. When the SiC volume fraction was kept constant at either complete dissolution of the SiC particles or to limit
40 wt%, no linear relationships were observed with increasing the retention of larger particles. It has been shown that if the extent
Ti and decreasing the Ni volume fractions simultaneously. It is of cracking or extraction of reinforcement particles is negligible,
noted that the three alloys which had the best wear resistance then the abrasive wear resistance may be improved [16].
were all produced with a composition of 40 wt% SiC and Ti and Ni The hard intermetallic phases probably cracked due to
powders ranging from 20 to 40 wt%. Due to the influence of laser repeated loading and the stresses applied during wear and these
processing conditions on the amount of dissolution and reaction micro-cracks facilitated material removal as wear progressed.
rates, the final proportions of each phase in the alloyed layers are Fracturing of the embedded SiC particles also occurred due to
likely to change. Thus in order to state which composition would cyclic loading but the strong bonding between the particles and
ultimately produce the best wear resistance the final proportions the matrix ensured that decohesion (interfacial debonding and
of each phase in the alloyed layers should be accurately deter- pull-out) of the SiC particles was prevented. A strong interfacial
mined. Due to the complex nature of the alloy system investi- bond between the reinforcements and the matrix is important.
gated and the variety in phase morphologies present this proved When the bond is weak the reinforcements are easily pulled out
to be difficult; even qualitative results by EDS is considered during wear and can act as additional abrasives thereby increas-
misleading. ing the wear rate, as was observed by reference [39]. Due to the
Abrasive wear generally occurs by plastic flow, plastic defor- strong interfacial bonding and prior to fracture, the embedded SiC
mation, brittle fracture or combinations of the different modes particles in the MMCs probably led to lower wear rates in alloys
[18] and in this work all modes were observed. The presence of where these particles were present in high volume fractions. This
intermetallic phases such as Al3Ni and Al3Ti together with the soft strong bond between the matrix and the reinforcements is
a-Al in the inter-dendritic regions constituted a favourable formed by a reaction between the ceramic particles and the metal
combination with these hard phases resisting abrasion and the matrix at the metal/ceramic interface and therefore it is not easy
soft aluminium phase suppressing macro-crack growth in these to detach these particles from the matrix even though they have
regions. The same behaviour was observed by Man et al. [36] who fractured. In wear tests conducted on Al–7 wt% Si laser alloyed
studied the abrasive wear of aluminium AA6061 laser surface with SiC, Anandkumar et al. [15] observed discontinuous wear
alloyed with NiTi and reported that wear generally occurred by grooves confined to the matrix on the worn surfaces and reported
micro-cracking and flaking of the surface. The brittleness of the that the hard SiC reinforcement particles interrupted the abrasive
material has been shown to increase with increasing proportion grooving action.
of reinforcement particles [23]. In a study on the relationship Deeper analysis is required to fully understand the response of
between hardness and brittleness of intermetallic phases Guo the alloyed layers to abrasive wear and to quantify the role of
et al. [38] reported that hardness and brittleness of the inter- each individual phase in contributing to the overall wear perfor-
metallic phases play a role in determining which phase would mance. Existing models based on the abrasive wear of multiphase
have more wear resistance. Phases that have a high hardness and materials are generally limited to the addition of one reinforce-
are less brittle will have better wear resistance but when ductility ment phase in varying proportions to a constant matrix. In the
is low (for brittle phases) fracture occurs leading to high wear current work three reinforcement materials were added to a
rates. This may be the reason for the high wear rates observed by constant matrix which produced a wide variety of reinforcement
the high-Ni alloys which showed the highest hardness, but did phases. Due to the complex nature of the investigated systems a
not have the best wear resistance mainly due to the brittle model to understand the abrasive wear has not been attempted in
fracture of the aluminides. the present work. Existing models may be used to explain the
The role of the mean free path between reinforcements has behaviour of individual phases, e.g. fracture of the SiC and micro-
also been reported to play a role in the wear resistance of cracking of intermetallic particles. However the application of
multiphase materials. Axén and Zum Gahr [39] studied the sliding these models is limited as they do not provide a holistic view of
wear resistance of a tool steel 90MnCrV8 reinforced with TiC both the individual and the combined phase response of the
particles and concluded that the mean free path (defined as the investigated system under abrasive wear conditions.
average spacing between reinforcement particles) also plays an
important role during wear. Matrix protection is improved by a
reduced mean free path between the reinforcing particles. Large 5. Conclusions
reinforcement particles result in an increased mean free path
which ultimately results in reduced wear resistance. The effect of In general laser alloying improved the wear resistance of the
the size of the reinforcements was also found to depend on the aluminium AA1200 surface with an improvement of 19–82% was
matrix structure. Small reinforcement particles are of greater achieved. The three alloys which had the best wear resistance
advantage within a hard matrix than large particles within a soft were all produced with a composition of 40 wt% SiC and Ti and Ni
matrix [39]. It is well known that a high volume fraction of small powders ranging from 20 to 40 wt%. No direct correlation was
particles will result in low mean free paths which would generally observed between hardness and wear resistance. The alloy with
lead to a high wear resistance; this is based on the theory of the highest hardness had the ninth highest wear rate while the
dispersion strengthening [33]. In the current study the alloyed best wear resistance was shown by the alloy which only had the
layers with small Al mean free paths had high wear rates as brittle eighth highest hardness. The predominant wear mechanisms
fracture was promoted by the formation of high densities of the were intense plastic deformation with micro-fracture of the SiC
hard and brittle intermetallic and ceramic phases in which particles and intermetallic phases. The wear behaviour is mainly
cracking of the intermetallic phases and fracturing of the SiC determined by the response of the different alloy phases, either
L.A.B Mabhali et al. / Wear 290–291 (2012) 1–9 9

independently or in combination, to the action of the abrasive [13] J.D. Majumdar, B.L. Mordike, I. Manna, Wear 242 (2000) 18–27.
particles and the precise nature of this response in terms of [14] P.H. Shipway, A.R. Kennedy, A.J. Wilkes, Wear 216 (1998) 160–171.
[15] R. Anandkumar, et al., Surface and Coatings Technology 201 (2007)
linking increasing wear rates to specific phases, particle hardness, 9497–9505.
reinforcement volume fraction and phase brittleness is complex [16] R. Colaco, R. Vilar, Wear 254 (2003) 625–634.
due to the wide range of reinforcement phases present and [17] R. Anandkumar, et al., Surface and Coatings Technology 205 (2011)
3824–3832.
requires further study.
[18] A. Almeida, et al., Materials Science and Engineering A 303 (2001) 273–280.
[19] A. Almeida, R. Vilar, Scripta Materialia 63 (2010) 811–814.
[20] J.D. Majumdar, et al., Surface and Coatings Technology 201 (2006)
Acknowledgements 1236–1242.
[21] J.D. Majumdar, et al., Wear 262 (2007) 641–648.
[22] J.D. Majumdar, et al., Materials Science Engineering A 433 (2006) 241–250.
The authors gratefully acknowledge the financial support of [23] J.W. Liou, L.H. Chen, T.S. Lui, Journal of Materials Science 30 (1995) 258–262.
the Department of Science and Technology and the Council for [24] K.-H. Zum-Ghar, Tribology International 31 (10) (1998) 587–596.
Scientific and Industrial Research in South Africa. The University [25] M.M. Krushkov, Wear 28 (1974) 69–88.
[26] W.M. Garrison, Wear 82 (1982) 213–220.
of the Witwatersrand in South Africa is also acknowledged for [27] N. Axen, I.M. hutchings, S. Jacobson, Tribology International 29 (6) (1996)
technical support. 467–475.
[28] I.M. Hutchings, Materials Science and Technology 10 (1994) 513–517.
[29] E. Rabinowicz, Friction and Wear of Materials, Wiley, New York, 1965.
References [30] American Standard Testing Methods, Standard Test Method for Measuring
Abrasion Using the Dry Sand/Rubber Wheel Apparatus, ASTM G65-04, 2004.
[1] S. Tomida, K. Nakata, Surface and Coatings Technology 174–175 (2003) [31] L.A.B. Mabhali, Laser Surface Alloying and In-Situ Formation of Aluminium
559–563. Metal Composites Reinforced With Ceramics and Intermetallics, Ph.D. Thesis,
[2] L.R. Katipelli, A. Agarwal, N.B. Dahotre, Applied Surface Science 153 (2000) University of the Witwatersrand, South Africa, 2011.
65–78. [32] Z. Brytan, et al., Journal of Achievements in Materials and Manufacturing
[3] H.C. Man, S. Zhang, F.T. Cheng, Materials Letters 61 (2007) 4058–4061. Engineering 40 (1) (2010) 70–78.
[4] A. Almeida, et al., Surface and Coatings Technology 70 (1995) 221–229. [33] I.M Hutchings, Tribology: Friction and Wear of Engineering Materials,
[5] N. Ravi, et al., Materials and Manufacturing Processes 15 (3) (2000) 395–404. Edward Arnold, 2002.
[6] M.H. Staia, M. Cruz, N.B. Dahotre, Wear 251 (2001) 1459–1468. [34] M.A. Zamzam, et al., Materials and Design 23 (2002) 161-138.
[7] M. Kok, Journal of Materials Processing Technology 161 (2005) 381–387. [35] J. Lis, et al., Ceramics International 19 (1993) 219–222.
[8] M.H. Staia, M. Cruz, N.B. Dahotre, Thin Solid Films 377–378 (2000) 665–674. [36] H.C. Man, S. Zhang, F.T. Cheng, Materials Letters 61 (2007) 4058–4061.
[9] A. Almeida, et al., Surface and Coatings Technology 200 (2006) 4782–4790. [37] L.A.B. Mabhali, et al., Molecular Crystals and Liquid Crystals 555 (1) (2012)
[10] K. Elleuch, S. Mezlini, N. Guermazi, Ph. Kapsa, Wear 261 (2006) 1316–1321. 138–148.
[11] Y. S- ahin, Tribology International 43 (2010) 939–943. [38] B. Guo, et al., Surface and Coatings Technology 202 (2008) 4121–4129.
[12] T. Miyajima, Y. Iwai, Wear 255 (2003) 606–616. [39] N. Axén, K.H. Zum Gahr, Wear 157 (1992) 189–201.

S-ar putea să vă placă și