Você está na página 1de 8

Journal of Alloys and Compounds 429 (2007) 176183

Development of nano-Y2O3 containing magnesium


nanocomposites using solidification processing
S.F. Hassan, M. Gupta
Department of Mechanical Engineering, National University of Singapore, 9 Engineering Drive 1, Singapore 117575, Singapore
Received 6 February 2006; received in revised form 19 April 2006; accepted 19 April 2006
Available online 23 May 2006

Abstract
In the present study, nano-Y2 O3 particulates containing magnesium nanocomposites were synthesized using an innovative disintegrated melt
deposition technique followed by hot extrusion. Microstructural characterization of the composite samples showed retention of reinforcement,
significant grain refinement of magnesium matrix, and the presence of minimal porosity. Mechanical properties characterization revealed that the
presence of nano-Y2 O3 reinforcement lead to significant improvement in hardness, 0.2% YS and UTS of the magnesium matrix. The ductility of
magnesium matrix increases to a peak with 0.22-vol% of nano-Y2 O3 but started to reduce with subsequent increase in the amount of Y2 O3 . The
results further revealed that the combination of 0.2% YS, UTS, and ductility of magnesium matrix with up to 0.66-vol% of nano-Y2 O3 reinforcement
remained much superior even when compared to magnesium reinforced with much higher volume percentage of micrometer size SiC. An attempt is
made in the present study to correlate the effect of nano-Y2 O3 as reinforcement and its increasing amount with the microstructural and mechanical
properties of magnesium.
2006 Elsevier B.V. All rights reserved.
Keywords: Magnesium; Yttria; Solidification; Microstructure; Mechanical properties; Ductility

1. Introduction
Magnesium based materials have emerged as light weight
alternatives to conventional structural materials in aerospace,
automotive, electronics, and consumer products [1]. The growing requirements of high specific mechanical properties with
weight savings have fueled significant research activities in
recent times targeted primarily for further development of magnesium based materials [116]. Among all the reinforcements
used in magnesium based composites only nano-size particulates has shown their potential superiority by simultaneously
improving both the specific mechanical properties and ductility of magnesium with noticeable weight savings [3,6,1215].
It is noticed that the intrinsic limited ductility of conventional
magnesium based materials [4,5,710,17] restricts its solid state
formability which is vital for intricate structural parts production
and remains one of the major concern in its engineering applications. Search of open literature indicates that some research

Corresponding author. Tel.: +65 68746358; fax: +65 67791459.


E-mail address: mpegm@nus.edu.sg (M. Gupta).

0925-8388/$ see front matter 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.jallcom.2006.04.033

work is conducted to study the effect of Al2 O3 , Y2 O3 , and


ZrO2 particulates in nanometer length scale on the properties of magnesium [6,1315]. However, no attempt is made
so far to synthesize the Mg based nanocomposites containing Y2 O3 particulates using solidification process and to study
its effect on the microstructural and mechanical properties of
pure magnesium. Thermally stable oxide ceramics particulates
such as Y2 O3 are expected to be capable of forming chemical bond at the metalreinforcement interface [18,19] in composite processed using solidification technique. Disintegrated
melt deposition (DMD) technique [20] has been chosen in
this study which remains an attractive choice among the solidification processing routes due to its capability of bringing
together the advantages of spray processing and conventional
casting.
Accordingly, the primary aim of the present study was to
synthesize magnesium based composite materials reinforced
with nano-size Y2 O3 particulates using DMD technique. The
composites thus obtained were hot extruded and characterized
for their microstructural and mechanical properties. Particular
emphasis was placed to study the effect of the presence of
nano-size Y2 O3 particulates and its increasing amount on the

S.F. Hassan, M. Gupta / Journal of Alloys and Compounds 429 (2007) 176183

microstructure and mechanical response of commercially pure


magnesium.

177

2.5. Microstructural characterization


Microstructural characterization studies were conducted on metallographically polished extruded nanocomposites samples to investigate reinforcement
distribution, reinforcement/matrix interfacial integrity and morphological characteristics of grains. The Hitachi S4100 field-emission scanning electron microscope (FESEM) and OLYMPUS metallographic microscope were used. Grain
size determination was carried out by image analysis using the Scion system.

2. Experimental procedures
2.1. Materials
In this study, magnesium turnings of more than 99.9% purity (supplied by
ACROS Organics, NJ, USA) were used as the base material and nano-sized
Y2 O3 particulates (supplied by Nanostructured & Amorphous Materials Inc.,
USA) with an average size of 29-nm were used as reinforcement phase.

2.2. Primary processing


Synthesis of monolithic and nano-sized Y2 O3 containing magnesium
nanocomposites (Mg/Y2 O3 ) containing three different volume percentages
of yttria was carried out using DMD technique. Synthesis of the composites involved heating the magnesium turnings with reinforcement particulates
(placed in a multi-layer sandwich form) to 750 C in an inert Ar gas atmosphere in
a graphite crucible using a resistance heating furnace. The crucible was equipped
with an arrangement for bottom pouring. Upon reaching the superheat temperature, the molten slurry was stirred for 2.5 min at 460 rpm using a twin blade (pitch
45 ) mild steel impeller to facilitate the incorporation and uniform distribution
of reinforcement particulates in the metallic matrix. The impeller was coated
with Zirtex 25 (86% ZrO2 , 8.8% Y2 O3 , 3.6% SiO2 , 1.2% K2 O and Na2 O, and
0.3% trace inorganic) to avoid iron contamination of the molten metal. The melt
was then released through a 10 mm diameter orifice at the base of the crucible.
The nanocomposite melt was disintegrated by two jets of argon gas orientated
normal to the melt stream and located 265 mm from the melt pouring point. The
argon gas flow rate was maintained at 25 L min1 . The disintegrated composite
melt slurry was subsequently deposited onto a metallic substrate located 500 mm
from the disintegration point. Preform of 40 mm diameter was obtained following the deposition stage. The synthesis of monolithic magnesium was carried
out using steps similar to those employed for the reinforced materials except
that no reinforcement particulates were added.

2.3. Secondary processing

2.6. Hardness
Microhardness and macrohardness measurements were made on the polished
monolithic and reinforced samples. Vickers microhardness was measured by
Matsuzawa MXT50 automatic digital microhardness tester using 25gf-indenting
load. Rockwell 15T superficial scale was used for macrohardness measurement
in accordance with ASTM E18-94 standard.

2.7. Tensile testing


The smooth bar tensile properties of the extruded monolithic and reinforced
samples were determined in accordance with ASTM test method E8M-01 using
Instron 8516 machine with a crosshead speed set at 0.254 mm/min on round tension test specimens of 5 mm diameter and 25 mm gauge length. Instron 2630-100
Series Clip-On type extensometer was used for strain recording. Fractography
was done on the tensile fractured samples using JEOL JSM-5800 LV scanning
electron microscope (SEM).

3. Results
3.1. Macrostructure
The result of macrostructural characterization conducted on
the as deposited monolithic and reinforced materials did not
reveal any presence of macropores or shrinkage cavity. Following extrusion, there was also no evidence of any macrostructural
defects.

The deposited monolithic and nano-sized Y2 O3 containing magnesium preforms were machined to 36 mm diameter and hot extruded using an extrusion
ratio of 20.25:1 on a 150 tonnes hydraulic press. Extrusion was carried out at
250 C. The preforms were held at 300 C for 90 min in a constant temperature
furnace before extrusion. Colloidal graphite was used as lubricant. Rods of 8 mm
diameter were obtained following extrusion.

The results of the density measurements are shown in Table 1.


Results show that near dense composites are formed using the
fabrication methodology adopted in the present study.

2.4. Density measurements

3.3. Microstructural characterization

Density () measurements were performed in accordance with Archimedes


principle [811] on three randomly selected polished samples of Mg and
Mg/Y2 O3 taken from the extruded rods. Distilled water was used as the immersion fluid. The samples were weighed using an A&D ER-182A electronic
balance, with an accuracy of 0.0001 g.

Microstructural studies conducted on the extruded nanocomposite specimens showed reasonably uniform reinforcement
distribution (see Fig. 1a) with good reinforcementmatrix interfacial integrity, and significant grain refinement (see Fig. 1b

3.2. Density measurements

Table 1
Results of density, porosity, and grain morphology characterization of DMD processed Mg/Y2 O3 nanocomposites
Materials

Mg/0.00Y2 O3
Mg/0.22Y2 O3
Mg/0.66Y2 O3
Mg/1.11Y2 O3

Y2 O3 (wt%)

0.0
0.6
1.9
3.1

Density (gm/cm3 )

Porosity (%)

Theoretical

Experimental

1.7400
1.7472
1.7616
1.7763

1.7397 0.0009
1.7472 0.0029
1.7598 0.0015
1.7730 0.0036

Grains morphology
Size (m)

0.02
0.00
0.10
0.19

49
10
6
6

8
1
1
2

Aspect ratio
1.5
1.4
1.5
1.5

0.4
0.6
0.3
0.3

178

S.F. Hassan, M. Gupta / Journal of Alloys and Compounds 429 (2007) 176183

Table 2
Results of room temperature mechanical properties of DMD processed Mg/Y2 O3 nanocomposites
Materials

Macrohardness (HR15T)

Microhardness (HV)

0.2% YS (MPa)

UTS (MPa)

Ductility (%)

Mg/0.00Y2 O3
Mg/0.22Y2 O3
Mg/0.66Y2 O3
Mg/1.11Y2 O3
Pure Mg [35]
Mg/9.3SiC [8]

37 1
56 1
58 0
49 0

40 0
51 0
56 0
52 1

97 2
218 2
312 4

69105
120 5

173 1
277 5
318 2
205 3
165205
181 6

7.4 0.2
12.7 1.3
6.9 1.6
1.7 0.5
58
4.7 1.3

Fig. 1. Representative micrographs showing: nano-Y2 O3 reinforcement (marked by arrow) distribution in the case of Mg/1.11Y2 O3 nanocomposite (using FESEM)
in (a) and grain morphology for Mg and Mg/0.66Y2 O3 in (b) and (c), respectively.

and c; Table 1). Observations also revealed minimal presence


of porosity in nanocomposite materials.

Ductility of magnesium matrix increased with the addition of


0.22-vol% of nano-Y2 O3 and decreased with further addition.

3.4. Hardness

3.6. Fractography

The results of the macrohardness and microhardness measurements conducted on extruded Mg and Mg/Y2 O3 samples
revealed an increasing trend in matrix hardness (both at macroand microlevel) till a volume percentage of 0.66 and decreasing trend thereafter (see Table 2). The macro- and microharness
of the nanocomposites remained superior when compared to
monolithic magnesium.

Fracture of magnesium matrix in nanocomposite samples


changed from cleavage mode to mixed mode of ductile and inter-

3.5. Tensile characteristics


The results of ambient temperature tensile tests conducted
on extruded monolithic and reinforced samples revealed that
the presence of nano-Y2 O3 particulates reinforcement lead
to an increase in 0.2%YS, UTS, and work of fracture (see
Tables 2 and 3) of magnesium till the addition of 0.66-vol%.

Table 3
Specific strength and work of fracture of DMD processed Mg/Y2 O3
nanocomposites
Materials

Work of fracture
(J/m3 )a

Mg/0.0Y2 O3
Mg/0.22Y2 O3
Mg/0.66Y2 O3
Mg/1.11Y2 O3
Mg/9.3SiC [8]

11.1
29.6
18.2
1.9
8.8

a
b

0.3
3.5
4.7
0.7
2.0

0.2%YS /
(kN m/kg)

UTS / (kN m/kg)

56
125
177

65b

99
159
181
116
98b

Determined from engineering stressstrain diagram using EXCEL software.


Density given was 1.850 gm/cm3 .

S.F. Hassan, M. Gupta / Journal of Alloys and Compounds 429 (2007) 176183

179

Fig. 2. Representative SEM fractographs showing: cleavage steps of Mg in (a), dimples and intergranular crack propagation for Mg/0.22Y2 O3 in (b) and (c), and
intergranular and transgranular microcracks for Mg/1.11Y2 O3 in (d) and (e), respectively.

Fig. 3. Representative SEM fractographs showing: (a) straight lines due to slip in the basal plane in Mg, (b) uneven lines due to combined effect of basal and non-basal
slip in Mg/0.22Y2 O3 , respectively.

granular, dominated by formation, growth, and coalescence of


the microscopic voids with the activation of non-basal slip system (see Figs. 2 and 3) due to the presence of nano-Y2 O3 up
to 0.66-vol%. However, predominantly intergranular fracture
mode for magnesium matrix was observed when nano-Y2 O3
was added to a 1.11-vol%.

4. Discussion
4.1. Synthesis of Mg and Mg/Y2 O3 materials
Synthesis of monolithic and reinforced Mg materials was
successfully accomplished by DMD process followed by hot

180

S.F. Hassan, M. Gupta / Journal of Alloys and Compounds 429 (2007) 176183

extrusion with: (a) minimal oxidation of magnesium and (b)


absence of macropores and blowholes. The inert atmospheric
condition (O2 < 2 ppm) used during melt processing, dispersion,
deposition and solidification was instrumental in the prevention of reaction between air/oxygen and Mg melt. The absence
of macropores, blowholes, and segregation or agglomeration of
reinforcement particulates due to the effect of gravity indicates
the suitability of stirring conditions in the crucible and the realization of good solidification conditions during deposition. The
absence of macropores and blowholes also suggests that the continuous flow of argon during the melting, stirring and deposition
process did not lead to the entrapment of gases. It should be noted
that, no reaction between magnesium melt/nanocomposite slurries with the graphite crucible was detected and it is primarily
due to the inability of magnesium to form stable carbides [4]. The
results, in essence, indicate the feasibility of the DMD process
as a potential fabrication technique for nano-Y2 O3 containing
Mg nanocomposite.
4.2. Microstructural behavior
Microstructural characterization of extruded nanocomposites
samples are discussed in terms of: (a) distribution of reinforcement, (b) reinforcementmatrix interfacial characteristics, (c)
grain size and shape, and (d) amount of porosity.
The reasonably uniform distribution of reinforcement particulates (Fig. 1a) can be attributed to: (a) limited agglomeration of
reinforcement during melting of matrix due to uniform arrangement of raw materials in crucible for melting [911,13], (b) minimal gravity-associated segregation due to judicious selection of
stirring parameters [20], (c) good wetting of reinforcement by
the matrix melt [5,18,19], and (d) disintegration of the composite
slurry by argon jets, and subsequent deposition in metallic mold.
Almost zero standard deviation in density measurement results
also reflect the uniform distribution of the reinforcement in
nanocomposites. Interfacial integrity between matrix and reinforcement as expected was good [5,18,19] and was assessed in
terms of interfacial debonding and nano-voids at the Y2 O3 /Mg
interface.
Metallography of the extruded samples revealed that the
matrix recrystallized completely. The sizes of the near-equiaxed
grains of magnesium matrix in the case of nanocomposite samples were distinctly smaller when compared to that of unreinforced magnesium (see Fig. 1b and c; Table 1). Grain refinement
in the case of composite samples can primarily be attributed
to the coupled effects of: (i) capability of nano-Y2 O3 particulates as reinforcement to nucleate magnesium grains during
recrystallization [21] and (ii) restricted growth of recrystallized
magnesium grains as a result of pinning by nano-Y2 O3 particulates [17]. The fundamental principles behind the ability of fine
inclusions in the metallic-matrix to nucleate recrystallized grains
and to inhibit the grain growth have already been established and
will not be discussed here. The result also revealed that grain
refinement of magnesium matrix enhanced till the addition of
0.66-vol% of nano-Y2 O3 reinforcement, however, beyond that
relatively higher tendency of clustering of nano-size particulates
[22] might be the reason for no further incremental effect.

The presence of minimal porosity in composite materials,


also supported by the experimental density values (see Table 1),
can be attributed to: (a) good compatibility between Mg and
Y2 O3 [5,18,19] and (b) the use of an appropriate extrusion ratio
[4,8].
4.3. Mechanical behavior
The results of hardness measurements revealed that incorporation of nano-Y2 O3 reinforcement lead to an increase in
the hardness (both in micro and bulk level) of the Mg/Y2 O3
nanocomposites (see Table 2) and can be attributed primarily
to: (a) the presence of relatively harder ceramic nanoparticulates in the matrix [6], (b) a higher constraint to the localized
matrix deformation during indentation due to their presence, and
(c) reduced grain size (see Table 1) [23,24]. The hardness results
obtained in the present study are similar to the findings reported
for nano-size ceramic reinforced magnesium matrices [6]. However, the results revealed that the hardness of nanocomposites
peaked at 0.66-vol% of Y2 O3 and beyond that reduced hardness
value (although still remain much higher than the monolithic
magnesium) and it might be the coupled effect of increasing
presence of porosity (see Table 1) and clustering of nano-Y2 O3
in the case of Mg/1.11Y2 O3 nanocomposite.
The results of room temperature tensile testing revealed significant increase in strength (see Fig. 4; Table 2) of pure magnesium due to the presence of nano-size Y2 O3 as reinforcement
and can primarily be attributed to the coupled effect of: (a)
increase in grain boundary area due to grain refinement [17,23],
(b) heavily built multi-directional thermal stress at the particulate/matrix interface at grain boundaries induced due to the large
difference of coefficient of thermal expansion between matrix
and reinforcement (CTE of Mg is 27.1 106 K1 [25] and
Y2 O3 is 7.5 106 K1 [26], respectively), and (c) the effective transfer of applied tensile load to the uniformly distributed
well-bonded strong Y2 O3 particulates (yield strength of ceramic
materials lies much higher than metallic materials [27]) [28]. It
may be noted that, since density of thermally induced dislocation
increases with decrease in reinforcement particulates size, the
extremely finer Y2 O3 with even such low volume percentages

Fig. 4. Representative tensile stressstrain curves of magnesium and its


nanocomposites.

S.F. Hassan, M. Gupta / Journal of Alloys and Compounds 429 (2007) 176183

as used in this study can induce high dislocation density in the


Mg/Y2 O3 system [29]. In general, the yield stress of material is
the stress required to operate dislocation sources and is governed
by the dislocation density and magnitude of all the obstacles
that restricts the motion of dislocation in the matrix. Under the
applied stress, enormous number of reinforcement particulates
and increasing amount of grain boundaries (due to significant
grain refinement) act as obstacles to the dislocation movement
and end up with dislocation pile ups [17,23,30]. Again, multidirectional thermal stress induced during processing easily starts
multi-gliding system [31] under applied stress so that dislocations were found developing and moving in several directions.
Multi-glide planes agglomerate under thermal and/or applied
tensile stress to form grain boundary ledges [23]. As the applied
tensile load increases, these ledges too act as obstacle to dislocation movement resulting in further pile-ups. The coupled
effects of these obstacles lead to the significant increase in the
yield strength of the composites over pure magnesium. During
plastic deformation under tensile load the increasing presence
of the excessive grain boundary ledges made their coalescence
to cavitation to microcracks and ultimately failure easier. In the
case of nanocomposite containing 0.66-vol% of Y2 O3 , marginal
increase in tensile load after yielding (see Table 2) led to the failure of matrix whereas almost instantaneous failure on yielding
of matrix was observed in the case of nanocomposite containing
1.11-vol% of Y2 O3 . Capability of strengthening the magnesium
matrix by Y2 O3 particulates as reinforcement when present in
much higher volume fraction and in compressive mode has also
been reported much earlier [5].
Room temperature tensile test also revealed the capability
of nano-size Y2 O3 particulates reinforcement to increase the
ductility of pure magnesium (see Fig. 4) when present with 0.22vol%. The increment in ductility of magnesium matrix due to
presence of nanoparticulates can primarily be attributed to the
coupled effect of: (a) grain refinement [32], (b) presence of reasonably uniformly distributed reinforcement particulates [30],
and (c) slip on extra non-basal slip system [33]. Grain refinement
particularly benefits hexagonal metals in ductility increment
where limited intergranular fracture arises from intercrystalline
stresses [32] and this can be clearly seen in Fig. 2c. Again, dispersed phases in brittle matrix, where dislocation mobility is
restricted and crack propagation is relatively easy, act as ductility
enhancer, an anomaly to their effect in ductile matrix [30]. Dispersed reinforcement particulates in brittle metal matrix serve
to: (i) provides sites where cleavage cracks may open ahead of
an advancing crack front, (ii) dissipate the stress concentration
which would otherwise exist at the crack front, and (iii) alter
the local effective state of stress from plane strain to one of
plane stress in the neighborhood of the crack tip. In addition, it
has been understood through recent studies that non-basal slip
system activate under axial tensile stress and increases ductility [33] which is also evident in fractography results obtained
in this study (see Fig. 3b). It is interesting to note that metallic
particulate like Ti [33] and nano-size ceramic oxides such as
Al2 O3 , Y2 O3 , and ZrO2 [6,1315] reinforcements seem to be
capable to activate non-basal slip system at room temperature
in pure magnesium matrix and increase the ductility. Effective

181

realization of change in the fracture mode from brittle cleavage


to typical ductile intergranular one could not improve ductility in
Mg/0.66Y2 O3 due to easy and early initiation of grain boundary cavitation and coalescence to formation and propagation
as microcracks to failure. Further addition of Y2 O3 to 1.11vol% in Mg/1.11Y2 O3 nanocomposite led to rapid formation
and propagation of microcracks immediate after matrix yielding
in almost complete intergranular mode (see Fig. 2d) to failure
with drastic reduction in ductility. Unlike Mg/0.22Y2 O3 and
Mg/0.66Y2 O3 nanocomposite dimples corresponding to ductile
fracture of metal were absent in Mg/1.11Y2 O3 nanocomposite.
Improvement in ductility of magnesium alloys exploiting thermally unstable finer precipitation at grain boundaries also has
been reported earlier [3,12].
The work of fracture expresses the ability of each material to
absorb energy up to fracture under tensile load and was computed
using stressstrain diagram [17]. It reveals that nanocomposites reinforced with nano-size Y2 O3 of up to 0.66-vol% are
distinctly superior when compared to monolithic magnesium
(see Table 3) indicating that the presence of nano-Y2 O3 significantly improved the fracture resistance of matrix. In essence,
the results of tensile testing revealed that the nano-Y2 O3 can
be used as the reinforcement (up to 0.66-vol %) in the magnesium for the development of materials for strength based designs
(higher yield strength when compared to magnesium) and damage tolerant designs (higher work of fracture when compared to
magnesium) with good formability.
The results further revealed that strength of these nano-Y2 O3
reinforced nanocomposites remained distinctly superior even
when compared to magnesium reinforced with much higher
volume percentage of micrometer size SiC. Ductility of magnesium matrix increased with the presence of 0.22-vol% of nanoY2 O3 and although beyond that ductility start to reduce it still
remains much higher in the case of Mg/0.66Y2 O3 when compared with Mg/9.3SiC [8]. This also translates into even lighter
weight and further enhanced specific mechanical properties
(see Table 3).
4.4. Fracture behavior
The meticulous study of uniaxially deformed fracture surfaces revealed the microstructural effects on tensile ductility
and fracture properties of nano-Y2 O3 reinforced pure magnesium. Typical brittle fracture surface (Fig. 2a) [17] with the
presence of microscopically rough small steps has been seen in
the case of Mg samples. Fractography of nanocomposite samples containing up to 0.66-vol% of nano-Y2 O3 revealed salient
features such as: (a) ductile fracture of metal with void-sheet
mechanism showing dimples [17] (see Fig. 2b), (b) intergranular crack propagation (typical for hexagonal metal to improve
its ductility [30]) (see Fig. 2c), and (c) increasing amount of
microcracks with increasing volume fraction of reinforcement
(see Fig. 2e). However, intergranular fracture mode for magnesium matrix was observed when nano-Y2 O3 was added to a
1.11-vol%.
Under the influence of far-field tensile load the fine microscopic voids including grain boundary ledges appeared and

182

S.F. Hassan, M. Gupta / Journal of Alloys and Compounds 429 (2007) 176183

propagated by the void-sheet mechanism before coalescing to


form dimples on the fracture surface. This is typical ductile
nature of fracture mode that requires homogeneous deformation of matrix which is naturally restricted in hexagonal crystal
structured magnesium by basal slip system [17]. Activation of
non-basal slip system is necessary to modify typical cleavage
crack propagation to failure with homogeneous plastic deformation to failure in pure magnesium matrix and it is apparent in the
cases of nano-Y2 O3 dispersion strengthened magnesium. Presence of nano-size oxide reinforcement and refined microstructure is believed to initiate the activation of non-basal slip system
in the Mg/0.22Y2 O3 Mg/0.66Y2 O3 nanocomposites as seen in
Fig. 3b. Activation of non-basal slip system in magnesium matrix
due to the presence of reinforcement [33] and grain refinement
[1315,34] has been reported in open literature earlier. However, failure of tensile specimens immediately after yielding
due to rapid intergranular crack propagation (see Fig. 2d) did
not gave the opportunity to materialize the activation of nonbasal slip system in Mg/1.11Y2 O3 nanocomposite matrix. It is
interesting to note that the fracture of the plastically deforming
magnesium was changed from cleavage mode to mixed ductile
and intergranular mode, dominated by formation, growth, and
coalescence of the microscopic voids with the activation of nonbasal slip system when nano-Y2 O3 were present up to 0.66-vol%
and further addition to 1.11-vol% led to complete intergranular
mode.
5. Conclusions
The main conclusions that can be drawn from this study are
as follows:
1. Disintegrated melt deposition technique coupled with hot
extrusion can be used to synthesize nano-Y2 O3 particulates
containing magnesium composites.
2. Microstructural characterization showed reasonably uniform distribution of reinforcement particulates, significantly
increasing grain refinement which stabilizes at 0.66-vol%
of reinforcement, and the presence of minimal porosity in
the nanocomposite microstructure developed using the primary and secondary processing parameters in the present
study.
3. The results of mechanical characterization revealed that the
presence of nano-Y2 O3 up to 0.66-vol% lead to an increase
in the combination of hardness, 0.2%YS, UTS, and work of
fracture of magnesium. Mg/1.11Y2 O3 formulation besides
having superior hardness and strength when compared to
monolithic magnesium remained quite brittle.
4. Fractography of nanocomposites revealed change in fracture mode from cleavage in the case of pure magnesium to
mixed mode of ductile and intergranular, dominated by formation, growth and coalescence of the microscopic voids
with the activation of non-basal slip system in the case of
nanocomposites till the volume percentage of nano-Y2 O3
reinforcement was 0.66. However, predominantly intergranular fracture mode for magnesium matrix was observed when
nano-Y2 O3 was added to a 1.11-vol%.

Acknowledgements
Authors wish to acknowledge NUS RP R-265-000-104-112
and NUSNNI for supporting this research effort. Thanks are also
due to Ms. Zhong Xiang Li for her help in FESEM.

References
[1] R. Fink, in: K.U. Kainer (Ed.), Die-Casting Magnesium, MagnesiumAlloys and Technologies, Wiley-VCH Verlag GmbH & Co., Germany,
2003, pp. 2344.
[2] D. Magers, J. Brussels, in: B.L. Mordike, K.U. Kainer (Eds.), Magnesium Alloys and Their Applications, Werkstoff-Informationsge-Sellschaft,
Wolfsburg, Germany, 1998, pp. 105112.
[3] A.G. Guy, Elements of Physical Metallurgy, second ed., Addison-Wesley,
USA, 1967, p. 233.
[4] D.J. Lloyd, Int. Mater. Rev. 39 (1) (1994) 123.
[5] B.Q. Han, D.C. Dunand, Mater. Sci. Eng. A 277 (2000) 297
304.
[6] R. Unverricht, V. Peitz, W. Riehemann, H. Ferkel, Proceeding of Conference on Magnesium Alloys and Their Applications, Germany, 1998, pp.
327332.
[7] R.A. Saravanan, M.K. Surappa, Mater. Sci. Eng. A 276 (2000) 108
116.
[8] M. Gupta, M.O. Lai, D. Saravanaranganathan, J. Mater. Sci. 35 (2000)
21552165.
[9] S.F. Hassan, M. Gupta, J. Mater. Sci. 37 (2002) 24672474.
[10] S.F. Hassan, M. Gupta, Mater. Sci. Technol. 19 (2003) 253259.
[11] S.F. Hassan, M. Gupta, J. Alloys Compd. 345 (2002) 246251.
[12] Alok Singh, M. Nakamura, M. Watanabe, A. Kato, A.P. Tsai, Scripta Mater.
49 (2003) 417422.
[13] S.F. Hassan, M. Gupta, Mater. Sci. Technol. 20 (2004) 1383
1388.
[14] S.F. Hassan, M. Gupta, Mater. Sci. Eng. A 392 (2005) 163168.
[15] S.F. Hassan, M. Gupta, J. Mater. Sci., in press.
[16] C.Y.H. Lim, D.K. Leo, J.J.S. Ang, M. Gupta, Wear 259 (2005) 620
625.
[17] R.E. Reed-Hill, Physical Metallurgy Principles, second ed., D. Van Nostrand Company, New York, 1964.
[18] N. Eustathopoulos, M.G. Nicholas, B. Drevet, Wettability at High
Temperatures, vol. 3, Pregamon Materials Series, Elsevier, UK,
1999.
[19] J.D. Gilchrist, Extraction Metallurgy, third ed., Pergamon Press, Great
Britain, 1989.
[20] L.M. Tham, M. Gupta, L. Cheng, Mater. Sci. Technol. 15 (1999)
11391146.
[21] P.G. Shewmon, Transformation in Metals, McGraw-Hill Book Co., New
York, 1969, p. 69.
[22] G.N. Hassold, E.A. Holm, D.J. Srolovitz, Scripta Met. Mater. 24 (1990)
101106.
[23] L.E. Murr, Interfacial Phenomena in Metals and Alloys, Addison-Wesley,
Massachusetts, 1975.
[24] J. Naser, W. Riehemann, H. Ferkel, Mater. Sci. Eng. A 234236 (1997)
467469.
[25] M. Bauccio (Ed.), ASM Metal Reference Book, third ed., ASM International, Materials Park, USA, 1993, p. 145.
[26] Y.S. Touloukian, Thermophysical Properties of High Temperature Solid
Materials, Purdue University, USA, 1967.
[27] M.F. Ashby, D.R.H. Jones, Engineering Materials I, ButterworthHeinemann, Boston, 1996.
[28] M. Gupta, T.S. Srivatsan, J. Mater. Eng. Perform. 8 (4) (1999)
473478.
[29] A.L. Geiger, J.A. Walker, JOM 43 (1991) 815.
[30] G.S. Ansell, in: R.W. Cahn (Ed.), Physical Metallurgy, North-Holland Publishing Company, Netherlands, 1970.

S.F. Hassan, M. Gupta / Journal of Alloys and Compounds 429 (2007) 176183
[31] F. Wua, J. Zhua, Y. Chen, G. Zhang, Mater. Sci. Eng. A 277 (2000) 143
147.
[32] W. Yang, W.B. Lee, Mesoplasticity and Its Applications, Materials
Research and Engineering, Springer-Verlag, Germany, 1993.
[33] P. Perez, G. Garces, P. Adv, Compos. Sci. Technol. 64 (2004) 145151.

183

[34] J. Koike, T. Kobayashi, T. Mukai, H. Watanabe, M. Suzuki, K. Maruyama,


K. Higashi, Acta Mater. 51 (2003) 2055.
[35] ASM International Handbook Committee, Properties and Selection: NonFerrous Alloys and Special-Purpose Materials, vol.2. ASM Handbook,
ASM International, Ohio, USA, 1990, p. 1134.

Você também pode gostar