Sei sulla pagina 1di 10

Microstructural evolution in ultrasonically processed AZ91 matrix composites and their

mechanical and wear behaviour

Sachin Vijay Muley* (sachinmul@gmail.com), Satya Prakash Singh*


(satya7.iitr@gmail.com), Piyush Sinha* (sinhaumt@iitr.ernet.in), P.P. Bhingole, G.P.
Chaudhari**

Department of Metallurgical and Materials Engineering, Indian Institute of Technology,


Roorkee 247667, India

* Undergraduate students

** Corresponding author, email: chaudfmt@iitr.ernet.in; Tel.: +91 1332 285524; Fax: +91
1332 285243.

Abstract
Commercial AZ91 magnesium alloy matrix composites (nominal composition Mg–9%Al,
1%Zn, 0.3%Mn, balance Mg in weight percent) reinforced with Si particles were fabricated
by solidification under ultrasonic vibration. Si particles of average 4 µm size and weight
fractions of 3% and 5% were employed. The microstructures of synthesized composites were
investigated by light optical and scanning electron microscopy and XRD. Results indicated
that the size, morphology and distribution of the in-situ Mg2Si particles were greatly
optimized with the assistance of the high-intensity ultrasonic field. The amount of in situ
Mg2Si particles increased, its size was refined and the distributions became uniform. The
Mg2Si particles were observed to be dispersed within the grains. Microstructural studies of
composites revealed enhanced grain refinement. With the addition of silicon particles, grain
size of matrix decreased in 3 and 5 wt. % Si/AZ91 composites. This is attributed to the
combined effect of heterogeneous nucleation and enhanced nucleation caused by water
cooling. The hardness and compressive ultimate stress of the 3 and 5 wt. % Si/AZ91
composites were evidently increased as compared with the as-cast AZ91 magnesium alloy.
The sliding wear behaviour of the composites was improved as compared to that of
monolithic alloy under varying normal loads and constant velocity. Abrasion was observed to
be the dominant wear mechanism at lower sliding velocities.

Keywords: A. light microscopy; A. mechanical characterization; B. magnesium alloys; C.


casting; D. grain refinement; D. nucleation

1. Introduction

Mg and its alloys have a combination of low density, high specific strength, good specific
stiffness, and excellent castability. Owing to these properties, they are promising materials
for use in automotive and aerospace industries and also, in electronic products [1-3].
Unfortunately, the utilization of Mg alloys was limited because they had inferior mechanical
properties at room and elevated temperatures and also poor heat resistance when compared to
other materials [4-7]. Therefore, it is necessary to develop new magnesium alloy matrix
composites because these showed promise for many structural applications, owing to their
improved mechanical properties over unreinforced monolithic metal counterparts. Recently,
magnesium metal matrix composites (MMCs) have been studied extensively because of their
good dimensional stability and high damping capacity, in addition to high specific strength,
low density, and high specific stiffness [8-12].
AZ 91 is one of the common magnesium alloys which are commercially available. At
ambient temperature, the solid solubility of Al in Mg is 2.1 wt.%. When this solubility limit is
exceeded, Mg17Al12 intermetallic phase is formed. Addition of Zn (solubility limit-0.8 at.%
at RT) further improves mechanical properties of the Mg-Al alloys [13-14]. On addition of
silicon to the magnesium based matrix, Mg2Si intermetallic phase is formed as a result of
negligible Si solubility in Mg. This results in improved mechanical properties at elevated
temperatures (up to about 150oC) because Mg2Si has high melting point (1085oC), high
hardness, low thermal expansion coefficient (7.5 × 10-6 K-1), and high Young’s modulus (120
GPa) [15-16].
While there are many manufacturing techniques to fabricate MMCs, stir casting has special
importance. As compared to other manufacturing techniques, stir casting is mostly used to
fabricate MMCs reinforced with micron sized particles as it ensures a good dispersion of
reinforced particles. However, the magnesium based MMCs which are fabricated by this
technique exhibit poor ductility at room temperature which bars them from many potential
uses [17-18]. To overcome this drawback, a new solidification processing technique that uses
ultrasonic vibrations has been developed. Although this can be used to fabricate metal
(including magnesium) matrix composites reinforced with ceramic nanoparticles [19-21],
magnesium based MMCs reinforced with micro-SiC particles were fabricated using
ultrasonic vibration technique, previously. It was observed that reinforcing micro-SiC
particles in the magnesium matrix allowed the ductility of the alloy to be retained or even
improved [22]. It is known that the density of SiC (~3.21 gm/cc) is more than that of Si
(~2.33 gm/cc). Thus it can be expected that micron sized Si powder will get uniformly
dispersed by ultrasonic vibration process.
Particle reinforced MMCs have exhibited improved wear resistance as compared to the
corresponding base metals due to presence of reinforced particles in the matrix, which restrict
continuous material removal at the tribocontact. Their tribological behaviour depends on the
type of matrix and counter-body material, testing conditions, reinforcement volume and its
chemistry. In one of the early works on sliding, abrasion and erosion behavior of alumina
fibre reinforced Mg and Mg–9Al–Zn matrix composites, it has been reported that with an
increase in volume fraction of fibre, an increase in the wear resistance can be obtained.
Whereas it decreased for three-body abrasion and erosion, the wear resistance increased for
two-body abrasion [23]. As compared to base Mg, researchers reported an improved wear
resistance for Mg–30 vol.% SiCp composites during adhesive wear without explicitly
exploring the wear mechanism [24]. The sliding wear behavior of Mg alloy-feldspar
composite against EN24 steel was studied and it was reported that the wear rate reduced with
increasing amount of feldspar particles [25]. It was indicated that the tribological properties
of composites are influenced by the inter-particle distance of SiC particles [26]. Recently, the
wear behavior of Mg–9 wt.% Al/SiCp composite was studied and abrasion, adhesion,
delamination, oxidation, melting and thermal softening were identified as the operating wear
mechanisms at varying loads of 10 and 30 N with sliding velocity of 0.2–5.0 m/s. The wear
rate initially decreased and then slightly increased with sliding speed in this work [27].
In the present work, AZ91 alloy-based magnesium metal matrix composites reinforced with 0
wt.% (no reinforcement), 3 wt.% and 5 wt.% micron sized Si particles were fabricated using
ultrasonic vibration process. The microstructural evolution is studied in order to reveal in situ
formation of new phases. The effect of evolved microstructure on the mechanical and wear
properties of these composites are studied in order to understand the effect of different
quantities of reinforcing particles.
2. Experimental Details

A commercial ingot of AZ91D alloy (nominal composition of Mg–9Al–1 Zn) was chosen as
matrix alloy and Si particles with an average size of 4 µm. First, about 0.5 kg of AZ91D alloy
was melted at 720oC in a steel crucible in an electric muffle furnace. Si particles were added
into the molten alloy. The content of Si particles in the Si/AZ91 composites was 3 wt.% and
5 wt.%. Then the ultrasonic probe was dipped into the melt to about 20 mm depth. The
ultrasonic vibration device (model no.-VCX 1500, Sonics and Materials, USA) consists of a
transducer with a maximum power of 1.5 kW and frequency of about 20 kHz. The melt was
ultrasonically processed at 1.5 kW power level for 3 minutes. The melt was then cooled to
room temperature using water quenching. For comparison, an AZ91 alloy ingot was similarly
ultrasonically processed and cast without reinforcing particles under identical conditions. The
composite ingots were machined to produce test specimens for different tests.
X-ray diffractometer (BRUKER D8) was used to determine the phases in the as-
prepared specimens. A QUANTA 200 FEG field emission scanning electron microscope
(FESEM) was used to analyze the microstructures. Samples for microstructure analysis were
prepared by the conventional mechanical polishing and etching using mixture of 10 ml acetic
acid , 4.2 gm Picric acid ,10 ml distilled water and 70ml of methanol for 10-15 s.
The hardness of polished samples was determined by computerized Vickers hardness
tester using 5-kgf indenting load and a dwell time of 30 s. Hardness test results are reported
as an average of five values at different locations on each sample. Compression tests were
performed on a thermo-mechanical simulator, Gleeble 3800®, at room temperature using
initial strain rate of 0.001 s-1. Dimensions of the cylindrical samples used in compression tests
were 10 mm diameter and 12 mm length. The compression tests are an average of three
values corresponding to each composite processing condition.
For dry sliding wear testing, pin specimens of 5 mm diameter and 10 mm length were
machined from the as-cast composites. The end surfaces of the machined pins were prepared
by polishing sequentially on 320, 800, 1200 and 1500 grit size silicon carbide papers. Sliding
wear tests were carried out using a pin-on disc configuration (Ducom, India). A pin-holder
loaded the stationary pins vertically onto a rotating AISI-O1 tool steel disc with hardness of
RC 63. Experiments were carried out in air without the aid of lubricant. Temperature and
relative humidity were maintained between 24–28oC and 52–65%, respectively. A constant
sliding velocity of 1 m/s was employed and three different loads 0.5, 1.0 and 1.5 kgf were
selected. For each sliding condition, two runs of 600 m and two runs of 900 m (total sliding
distance=3000 m) were performed. At the end of each test, the pins were weighed using an
electronic balance to measure weight loss. Wear of the pin was measured using weight loss
after each run. Using measured densities, the weight loss data were converted to volume loss.
In this study, the wear rate is defined as volume loss per sliding distance. Wear debris was
collected after each run. The worn pins surfaces and wear debris were examined and analyzed
using field emission scanning electron microscope (FESEM) and X-ray diffractometer
(XRD).

3. Results and Discussion


3.1 Constituent phase analysis
Fig. 1a shows the morphology and size distribution of Si particles employed in this work. It
shows particles with angular morphology. The corresponding XRD pattern of the Si powder
is shown in Fig. 1b. Fig. 1c shows the XRD pattern of composite that was fabricated by
ultrasonic vibration process. Diffraction peaks for Mg, Mg17Al12, Mg2Si, and Mg3Al2 are
present. Fig. 2 shows the elemental mapping of the composites. While the occurrence of
Mg17Al12 and Mg2Si is explained in the introductory note, diffraction peaks corresponding to
Mg3Al2 are also noticed.
Fig. 2 shows elemental mapping in the ultrasonically processed composite. Fig 2a shows the
distribution of aluminium in the alloy matrix except at some isolated spots. However, Mg is
uniformly present throughout the matrix (Fig. 2b). In Fig. 2c silicon is observed at some
locations where Al is absent, suggesting the formation a compound of Mg and Si. In Fig. 2d,
aluminium is seen as segregated along the grain boundaries suggesting the formation of
commonly observed Mg17Al12 secondary phase in the AZ91 alloys.
Based on XRD results on formation of various phases, some reactions are observed to take
place in molten magnesium. The in situ formation of Mg2Si results from the following
reaction [28].
2Mg+Si→Mg2Si (1)
The Gibbs’ Free Energy value for this reaction was calculated to predict its thermodynamic
feasibility. For a typical reaction n1A1+n2A2→n3A3+n4A4, the free energy is calculated as
follows [29].
𝑇 𝑇
∆𝐻 = ∆𝐻°+ ∫298 ∆Cp(1) 𝑑𝑇 + ∆𝐻𝑚 + ∫298 ∆𝐶𝑝(2)𝑑𝑇 (2)
𝑇 ∆𝐶𝑝(1) ∆𝐻𝑚 𝑇 ∆𝐶𝑝(2)
∆𝑆 = ∆𝑆°+ ∫298 𝑇 𝑑𝑇 + 𝑇 + ∫298 𝑇 𝑑𝑇 (3)
∆𝐺 = ∆𝐻 − 𝑇∆𝑆 (4)
Here Δ Cp =n3Δ Cp(3)+n4Δ Cp(4) +n1Δ Cp(1) +n2Δ Cp(2) (5)
For simplifying these calculations, above equations can be written as
∆𝐺 = 𝑎 + 𝑏𝑇 (6)
The thermodynamic parameters for the reaction are listed in the table.
Using values from the table, the following expression for free energy for the reaction is
arrived at.
∆𝐺 = −97518.784 − 17323.74𝑇 (7)
When the melt temperature is 993 K, ∆𝐺 = −17300.01 kJ/mol. This means that the reaction
can take place spontaneously at melt conditions used in this work. Thus, the thermodynamic
calculations are in accordance with results of XRD data analysis and the in situ formation of
Mg2Si phase in the melt is confirmed.
3.2 Microstructural analysis
Fig. 3 shows the SEM micrographs of as-cast AZ91 alloy and the composites fabricated by
ultrasonic vibration process. Fig. 3a shows the micrograph of alloy without any
reinforcements. It shows a relatively coarse dendritic structure. Fig. 3b and 3c shows that the
dendritic structure is replaced by relatively fine equiaxed grain structures. Table 2
summarizes the relative grain size of AZ91 alloy and the composites. With the addition of Si
particles, the grain size is refined as seen in 3 and 5 wt.% Si/AZ91 composites.
It can be clearly seen from Fig. 4 that the use of ultrasonic vibration processing has caused
nearly uniform dispersion and distribution of in situ formed Mg2Si particles, and the
Mg17Al12 secondary phase in the magnesium alloy matrix (Fig. 3). The in situ Mg2Si particles
are present in near spherical shape or distinct polygonal-shape with their size in the micron
range. This is likely due to the effects of the high-energy ultrasonic field. As shown in Fig. 5,
it is clear that the in situ reaction is proceeding at the interface of the added silicon particles
and the molten magnesium alloy matrix. As compared to Fig. 1a wherein the Si particles have
sharp angular morphology, the particles appear rounded as a result of reaction and dispersion
of the reaction product throughout the matrix. It is expected that ultrasonic processing for a
prolonged duration may improve the dissolution, dispersion and distribution of the
particulates and also result in an increase in the formation of Mg2Si phase at the expense of
Si.
The effect of ultrasonic vibration processing is now discussed. The sonic pressure gradients
are built from the sonic source primarily due to the collapse of cavitation bubbles and also,
the viscous force of the magnesium melt [30]. Due to action of high pressure pulses,
circulatory flow is set up in the magnesium melt. The flowing speed near transducer is
relatively fast and the flowing direction of the melt is vertically downward. This speed
reduces progressively on moving towards the bottom of the crucible. A spiral vortex is now
formed near the end face of the amplitude lever. These induce the acoustic streaming effects:
stirring and vibration. These effects of the high-energy ultrasonic vibration processing
contribute to the diffusion of the reactants and the products. Thus, the in situ chemical
reaction is speeded and the in site formed and broken Mg2Si and Mg17Al12 phases distribute
nearly uniformly in the magnesium melt.
The reinforcing Si particles act as additional sites for nuclei formation and thus facilitate the
process of heterogeneous nucleation in the magnesium melt. At such preferential sites, the
effective surface energy is lower, which diminishes the free energy barrier and facilitates
nucleation. Surfaces promote nucleation because of wetting (contact angles greater than zero
between phases encourage particles to nucleate) [31]. The nucleus that forms on a ready-
made substrate is so oriented that the atoms on its external face are arranged in the same
manner as those of the contacting face. The growth of a nucleus proceeds by build-up of new
single-atom layers on its surface. For successful nucleation, these nuclei must be of a size not
less than a definite critical value [31]. Also for increased degrees of supercooling, this critical
size of nucleus gets reduced and less work is required for nucleus formation. The larger the
number of nuclei and the slower their growth, the smaller the crystals that will be that grow
out of each nucleus. Therefore, the combined effect of adding reinforcement and water-
quenching is finer grain size in the composite as compared to as-cast AZ91.
3.3 Mechanical properties
The hardness values for the processed composites and the as-cast AZ91 are listed in Table 3.
The listed value for each specimen is an average of 5 hardness values taken at different
locations, separated by required minimal distance so as to avoid the effect of strain-hardening
near the indented spot. With the increase in reinforcements, a gradual increase is observed in
the hardness values. Samples for compressive test samples were prepared and tested at room
temperature under an initial strain rate of 0.001 /s. As is evident from Fig. 6, the ultimate
compressive stress was highest for composite with 5 wt.% silicon content and lowest for the
as-cast alloy. The compressive strength is improved by addition of reinforcing silicon
particles and the resulting Mg2Si phase.
The improvement in the mechanical properties due to reinforcing silicon can be attributed to
various strain hardening mechanisms.
When a single crystal is deformed under tension, it deforms freely on a single slip system and
changes its orientation by lattice rotation while extension is happening. However, in a
polycrystalline material under tension, individual grains are not subjected to a single uniaxial
stress. For the grain boundaries between the deforming crystals to remain intact, continuity
must be maintained in the material. As is expected, considerable differences in the
deformation between neighbouring grains and also within the grains are caused by the
imposed constraints. In spite of a continuous stress across boundaries, there may be distinct
possibility of a steep strain gradient in this region [32]. With simultaneous decrease in grain
size and increase in strain, the deformation assumes a more homogeneous character. Even at
low strains, slip occurs on several systems due to the constraints. In the regions surrounding
the boundaries, slip can occur on non-close-packed planes. This implies that different slip
systems can operate in adjacent regions of the same grain [32]. Complex lattice rotations
result, which leads to formation of deformation bands. For a decrease in grain diameter, the
effects of the grain boundaries will now be felt at the grain center. Thus, the strain hardening
of a fine-grained material is greater than a coarse-grained polycrystalline aggregate.
According to the classic Hall–Petch equation: σy= σ0+Kyd-1/2 (8)
where σy is the yield strength, σ0 and Ky are material constants, and d is the mean grain size,
the value of Ky depends on the number of slip systems. As compared to FCC and BCC
metals, it is higher for HCP metals [33]. Since Mg lattice is HCP, the grain size affects the
yield strength significantly. As shown in Fig.3 and Table 2, the grains in matrix are gradually
refined with increase in the volume fraction of Si particles. As a result, the yield strength of
composites is gradually enhanced. An increase in yield strength is also reflected in higher
hardness values for processed composites.
3.4 Wear characteristics
The volumetric wear rate for monolithic AZ91, AZ91–3 wt% Si and AZ91–5 wt % Si
composites are plotted against constant sliding speed of 1 m/s under varying loads of 5, 10
and 15 N in Fig 7. When we compare the wear behaviour for the same specimen under
different loads, the volumetric wear rate rises with increase in load. It is also clearly evident
that there is a gradual reduction in wear volume with an increase in Si content for the same
normal load. The improved wear resistance can be attributed to two reasons. One is the
presence of finely distributed hard Mg2Si particulates formed by in situ reaction in AZ91
matrix. The other is increase in strength and hardness due to the finer grain size. The
observation agrees with Archard’s hypothesis [34] which states that the materials with
improved hardness will exhibit better resistance to wear.
𝐾.𝑊.𝑆
𝑉= (9)
𝐻
In the beginning (run in period) of the wear tests, adhesive wear is likely to dominate due to
sliding of the cleaner surfaces in contact. As observed in Figs. 7 a), b), c), the volumetric
wear of AZ91 alloy and the composites does not vary much initially. Even with 5 wt % Si
added composite, no significant beneficial effect is seen on wear behavior. This can be
ascribed to better cold weldability of fine grained structure due its tendency to reduce its
surface energy [35]. With the progress of the test, the abrasive wear dominates because of
three body abrasion involving the specimen, the counter face, and the wear debris generated
due to sliding wear. This is supported from the Archard wear coefficient obtained from the
experimental results. The Archard wear coefficient was found to be of the order of 10-3 for all
the three materials indicating abrasive wear [35]. Significant improvement in wear
performance is observed in 5 wt% reinforced composite because the abrasive wear which is
characterised by scouring of the material is resisted by grain size strengthened and dispersion
strengthened composites. This observation could also be attributed to the reduced effective
area fraction of the metal matrix in contact with the counter face due to the presence of the
hard Mg2Si particulates in the matrix. However, the wear rate has decreased with increase in
silicon content and thus, the present study cannot provide evidence for any type of optimum
sliding conditions.
SEM studies of the worn pin surfaces (Fig. 9) of the monolithic alloy and reinforced
composites identified abrasive wear mechanism under different normal loads. The presence
of abrasive wear in the low speed regime has been reported by a number of investigators. In a
study involving pure magnesium reinforced with nano-alumina particles, it was reported that
abrasion was the dominant wear mechanism at 1 m/s [36]. Abrasive wear is generally
characterized by numerous grooves and shallow scratches running parallel to the sliding
direction as shown in Fig 9. Grooves and scratches are clearly evident for the alloy and the
composites. This type of wear is predominant at sliding speed of 1 m/s at 5, 10 and 15 N
loads. Detailed examination of wear track and the pin surface revealed the features associated
with abrasion. These features indicate abrasive wear, which was caused by material being
removed from the pin surface by protuberances or hard particles on the counterface, forcing
against and cutting or ploughing into the surface [35].
The deep grooves observed under low speed and varying normal loads indicate that mode of
abrasion is cutting. This results from the penetration of the hard asperities onto the pin
surface with minimal displacement of material on both sides of the grooves during sliding.
These deep grooves are more severe in the monolithic matrix compared to its composites
because of the relatively brittle hexagonal closed-packed crystals of the magnesium matrix
which favours abrasive wear through micro-cutting [36].

4. Conclusions
AZ91 alloy matrix composites are fabricated by ultrasonic processing and their properties are
studied. The following conclusions can be drawn.
1. With the increase in silicon content, the grain size of the composites decreased
progressively. This is largely attributed to the occurrence of heterogeneous nucleation
and supercooling due to water-quenching. The morphology of the in situ formed
Mg2Si phase was similar to the silicon particulates: some were almost spherical and
others were polygonal. A good distribution of particulates was observed by ultrasonic
vibration process. Better dispersion and distribution of particulates may be observed
by ultrasonic vibration processing for a longer duration.
2. The composites showed improved mechanical properties as compared to monolithic
alloy. Better hardness and compressive strength were observed, owing to reduced
grain size of composites and strain-hardening due to distortion of magnesium matrix
by silicon particulates.
3. Under varying normal loads and low sliding velocity regime, the composites showed
improved wear behaviour as compared to the monolithic alloy. The observed wear
mechanism was abrasion and the wear behaviour complied with Archard’s wear law.

Acknowledgements
Funding provided by SERC, Department of Science and Technology, India is gratefully
acknowledged.

References
[1] E.A. Brandes, G.B. Brook, Smithells Light Metals Handbook, Oxford: Butterworth-
Heinemann, 1998.
[2] I.J. Polmear, Mater. Sci. Technol. 10 (1994) 1–16.
[3] B.L. Mordike, T. Ebert, Mater. Sci. Eng. A 302 (2001) 37–45.
[4] S.W. Xu, N. Matsumoto, K. Yamamoto, S. Kamado, T. Honma, Y. Kojima, Mater.
Sci. Eng. A 509 (2009) 105–110.
[5] A. Mallick, S. Vedantam, L. Lu, Mater. Sci. Eng. A 515 (2009) 14–18.
[6] H. Lianxi, W. Erde, Mater. Sci. Eng. A 278 (2000) 267–271.
[7] Y. El-Saeid Essa, J.L. Perez-Castellanos, J. Mater. Process Tech. 143(144) (2003)
856–859.
[8] X.J. Wang, K. Wu, W.X. Huang, H.F. Zhang, M.Y. Zheng, D.L. Peng, Compos. Sci.
Technol. 67 (2007) 2253–2260.
[9] H.Z.Ye, X.Y. Liu, J. Alloys Compd. 402 (2005) 162–169.
[10] Q.C. Jiang, X.L. Li, H.Y. Wang, Scripta Mater. 48 (2003) 713–717.
[11] A. Bochenek, K.N. Braszczynska, Mater. Sci. Eng. A 290 (2000) 122–127.
[12] Y. Cai, M.J. Tan, G.J. Shen, H.Q. Su, Mater. Sci. Eng. A 282 (2000) 232–239.
[13] G.V. Raynor, The physical metallurgy of magnesium and its alloys, Pergamon
Press, 1959, pp. 33.
[14] H. Okamoto, in: T.B. Massalski (editor), Binary alloy phase diagrams, ASM
International, 1990.
[15] G. Frommeyer, S. Beer, K. Von Oldenburg, Z. Metallkund 85(5) (1994) 372–
377.
[16] A.A. Nayeb-Hashemi, J.B. Clark, in: T.B. Massalski (editor), Binary alloy
phase diagrams, ASM International, 1990.
[17] A. Luo, Mater. Trans. A 26 (1995) 2445–2455.
[18] R.A. Saravanan, M.K. Surappa, Mater. Sci. Eng. A 276 (2000) 108–116.
[19] J. Lan, Y. Yang, X. Li, Mater. Sci. Eng. A 386 (2004) 284–290.
[20] Y. Yang, J. Lan, X. Li, Mater. Sci. Eng. A 380 (2004) 378–383.
[21] G. Cao, H. Konishi, X. Li, Mater. Sci. Eng. A 486 (2009) 357–362.
[22] K.B. Nie, X.J. Wang, K. Wu., L. Xu, M.Y. Zheng, X.S. Hu, J. Mater. Sci. 47
(2012) 138-144.
[23] A. Alahelisten, F. Bergman, M. Olsson, S. Hogmark, Wear 165(2) (1993)
221–226.
[24] R.A. Saravanan, M.K. Surappa, Mater. Sci. Eng. A 276 (2000) 108–116.
[25] S.C. Sharma, B. Anand, M. Krishna, Wear 241 (2000) 33–40.
[26] S.K. Thakur, B.K. Dhindaw, Wear 247 (2001) 191–201.
[27] C.Y.H. Lim, S.C. Lim, M. Gupta, Wear 255 (2003) 629–637.
[28] K. Kondoh, H. Oginuma, R. Tuzuki, T. Aizawa, Mat. Trans. 44(4) (2003) 611-
618.
[29] D.R. Gaskell, Introduction to Thermodynamics of Materials, fourth ed., Taylor
& Francis, 2009.
[30] S. Zhang, Y. Zhao, G. Chen, Nonferrous Met. Soc. China 20 (2010) 2096-
2099.
[31] Y. Lakhtin, Engineering Physical Metallurgy, Mir Publishers Moscow, 1977.
[32] G.E. Dieter, Mechanical Metallurgy, SI Metric Ed., McGraw-Hill Book Co.,
1988.
[33] H.Z. Ye, X.Y. Liu, J. Mater. Sci. 39 (2004) 6153–6171.
[34] J.F. Archard, J. Appl. Phys. 24 (1953) 981–988.
[35] I.M. Hutchings, Tribology: Friction and Wear of Engineering Materials,
Edward Arnold, London, 1992.
[36] C.Y.H. Lim, D.K. Leo, J.J.S. Ang, M. Gupta, Wear 259 (2005) 620–625.

Fig. 1a. SEM image of Si powder employed in this work.


Fig.1b. X-Ray diffraction pattern of the Si powder.
Fig. 1c. X-Ray diffraction pattern of ultrasonically processed AZ91 matrix composite.
Fig.2. Elemental mapping of 3 % Si reinforced ultrasonically processed AZ91 composite
shows distribution of elements obtained from energy dispersive spectroscopy.
Fig. 3. FESEM images of AZ91 alloy (a) without reinforcement, reinforced with (b) 3%, and
(c) 5% (by weight) of Si powder followed by ultrasonic treatment. Decrement in grain size is
clearly observed.

Fig. 4. Elemental mapping of 5% Si reinforced ultrasonically processed AZ91 composite


shows distribution of elements obtained from energy dispersive spectroscopy.
Fig.5. SEM image of a Si particle dispersing into the matrix following the in situ reaction
with magnesium alloy matrix. The edges of the Si particle can be seen to be dissolving into
the matrix.

Fig.6. Compressive stress-strain curve obtained from tests performed using thermo
mechanical simulator-Gleeble 3800®.
Fig. 7. Variation of cumulative wear loss with sliding distance for (a) 500 g,(b) 1000 g, and
(c) 1500 g load, and (d) variation in wear rate with applied load for the AZ91 base alloy and
ultrasonically processed composites.
Fig. 8. SEM images worn surfaces of AZ91 alloy- (a) without reinforcement, reinforced with
(b) 3%, and (c) 5% (by weight) of Si powder followed by ultrasonic treatment. Wider and
larger number of grooves in AZ91 shows enhanced wear as compared to composites.

Table 1: This table gives thermodynamic parameters of elements and compounds of interest
[29].
Substance Cp# ∆𝐻° ∆𝑆° Tm/K ∆𝐻𝑚
3 -5 -1 -1 -1
a b/10 c/10 kJ mol kJ mol K kJ mol-1
Mg(s) 22.39 10.29 -0.433 0 32.802 872 8.988
Mg(l) 32.76 - - - - - -
Mg2Si 73.58 15.04 -8.86 -79.38 64.05 1077 86.1
Si 24.02 2.478 -4.158 0 18.9 1685 50.4
-2
# Cp =a+bT+cT

Table 2: This table denotes gradual decrease in grain size with increase in reinforcement
content.
% Reinforcement 0% 3% 5%
Average Grain Size 73 51 40
(µm)

Table 3: This table denotes gradual increase in hardness values with increase in
reinforcement content.
Reinforcement , in 0 (as-cast) 3 5
weight %
Hardness (VHN) 65.2 ±3 97.4 ±3 108.2 ±4

Potrebbero piacerti anche