Está en la página 1de 12

Proceedings of the Institution of Mechanical

Engineers, Part L: Journal of Materials Design


andhttp://pil.sagepub.com/
Applications
Optimization of mechanical properties of Al2O3SiC nanocomposite body

A Asadi, S A Sadough Vanini and A Farrokhi


Proceedings of the Institution of Mechanical Engineers, Part L: Journal of Materials Design and Applications 2011 225:
266 originally published online 25 August 2011
DOI: 10.1177/1464420711418429
The online version of this article can be found at:
http://pil.sagepub.com/content/225/4/266

Published by:
http://www.sagepublications.com

On behalf of:

Institution of Mechanical Engineers

Additional services and information for Proceedings of the Institution of Mechanical Engineers, Part L: Journal of Materials Design and
Applications can be found at:
Email Alerts: http://pil.sagepub.com/cgi/alerts
Subscriptions: http://pil.sagepub.com/subscriptions
Reprints: http://www.sagepub.com/journalsReprints.nav
Permissions: http://www.sagepub.com/journalsPermissions.nav
Citations: http://pil.sagepub.com/content/225/4/266.refs.html

>> Version of Record - Oct 13, 2011


Proof - Aug 25, 2011
What is This?

Downloaded from pil.sagepub.com by Narayanasamy P on October 18, 2011

266

Optimization of mechanical properties


of Al2O3SiC nanocomposite body
A Asadi*, S A Sadough Vanini, and A Farrokhi
Mechanical Engineering Department, Amirkabir University of Technology, Tehran, Iran
The manuscript was received on 19 August 2010 and was accepted after revision for publication on 8 July 2011.
DOI: 10.1177/1464420711418429

Abstract: Experimental procedure of Al2O3SiC nanocomposite body with addition of small


amount of MgO for hot-press sintering process is considered. For improving the densification
behaviour and mechanical properties of this nanocomposite body, experimental variables such as
sintering temperature, sintering duration, SiC vol%, and MgO amount were optimized. For optimization, many tests were conducted on the nanocomposite materials with different produce
conditions to achieve the best physical and mechanical properties. Obviously, achieving the
best condition for all the physical and mechanical properties at one time is unreachable matter;
so concern is on the properties which are important at high-speed impacts. It was observed that
specimen with 10 vol% SiC, 500 ppm MgO, sintering temperature of 1650  C, sintering duration of
90 min, and sintering pressure of 30 MPa, had the optimum physical and mechanical properties.
Keywords:

Al2O3SiC nanocomposite, MgO, hot-press sintering, microstructure, optimization

INTRODUCTION

Because of the low density, high strength, and high


hardness of nanocomposite materials such as
ceramics with nanoscale particles, many research
studies have been reported on the subject of these
materials [1, 2]. To overcome the brittleness of
ceramics, new technique concepts for creating the
ceramic nanocomposites have been developed [3].
One of the techniques to improve the strength of
ceramics is particle dispersion of different phases in
an alumina matrix [4]. New material design proposed
by Niihara [3] significantly improved strength of
the material by dispersing second phase nanosize
particles within the matrix grains and on the grain
boundaries. SiC particles were mainly located
within Al2O3 matrix grains because of the transformation of alumina from the g phase to the a phase unlike
common composites of which second phase particles
were mainly located on the boundaries [5]. Thermal
expansion difference between ceramic and SiC
particles produces significant improvement in
*Corresponding author: Mechanical Engineering Department,
Amirkabir University of Technology (Tehran Polytechnic), 424
Hafez Avenue, 15914 Tehran, Iran
email: taban1293@aut.ac.ir; taban1293@yahoo.com

mechanical properties of nanocomposite, especially


fracture toughness of the material by producing
nanoscale cracks in material. These nanoscale
cracks cause the growth of the frontal process zone
(FPZ) size ahead of a crack tip and improve the fracture toughness of the material [6]. A theoretical model
for investigating thermal expansions anisotropy of
the Al2O3 matrix and SiC particles has been presented
by which it is showed that high strength could be
achieved because of the residual microstresses,
despite such microstresses are distributed in the lattice rather than being localized at the grain boundaries [7]. It has been shown by Medvedovski [8] for
aluminamullite ceramics that satisfactory highspeed impact performance of the aluminamullite
ceramics can be explained by their high impact
energy dissipation ability. Therefore, energy dispersion (D-criterion) is one of the important parameters
in the high-speed impact and penetration for which
this study was calculated in accordance with the formula proposed by Neshpor et al. [9]
D

0:36Hv Ec
KIc2

where Hv is the Vickers hardness, E the Youngs modulus, c the sonic velocity, and KIc the fracture

Proc. IMechE Vol. 225 Part L: J. Materials: Design and Applications


Downloaded from pil.sagepub.com by Narayanasamy P on October 18, 2011

Optimization of mechanical properties of Al2O3SiC nanocomposite body

toughness. Sonic velocity was calculated using the


following formula [10]
s
E 1  v
2
c
1 v1  2v
where  is the density and  the Poissons ratio for
which the nanocomposite of this study was assumed
as 0.2. Since hardness, Youngs modulus, fracture
toughness, and density have direct impact on the
energy dissipation value, consequently tests for
these properties have been considered as the main
test of this study.
Also, small amount of MgO had a significant effect
on improving the densification and mechanical properties of alumina nanocomposite with decreasing the
grain size of the alumina [1113]. Effects of the MgO
on Al2O35 vol% SiC had been investigated by Wang
et al. [11], who showed that both sintering density
and mechanical properties of the Al2O3SiC ceramic
material were improved by adding a small amount of
the MgO. It was seen in their study that with higher
amounts of the MgO, the effects rate of MgO on both
density and hardness were decreased. MgO, as the
second phase, in Al2O3 is able to reduce sintering
temperature and also eliminate the abnormal grain
growth [12]. This study showed that hardness was
increased as a result of lower porosity because of
MgO additive. MgO increases densification rate by
increasing lattice and boundary diffusions and
reduction of pores [14]. It also had been reported
that when great quality alumina pellet is placed
between two compacts of Analar MgO, MgAl2O4
would be forming on the end layers of alumina
pellet, which is helping the densification of alumina
to its theoretical value [15]. It was shown that in
alumina sintering, MgO solid solution is responsible
for the improving effect of MgO addition to alumina,
and excess addition of MgO beyond its solid solubility
limit exists would cause non-stoichiometric spinel in
grain boundaries of sintered Al2O3 [16, 17].
In all the previous studies, produced specimens
had small dimensions and because of the small
scale of test specimens, dispersion of the second
phase in the material was homogenous, and since
perfect dispersion of the second phase nanoparticles
had a significant impact on the mechanical properties of the material, the high mechanical properties
like 1 GPa bending strength in small specimens by
Niihara [3] was achievable. Collection of these small
specimens has application in construction of segment panels; on the other hand, the larger scale
bodies have widespread applications at high-speed
impact problems as tile panels. In this study, specimens are produced in larger scales and test samples

267

were cut and polished from the sintered specimens.


This study is aimed to investigate the effects of the
experimental variables of hot-press sintering on
physical and mechanical properties of the large
nanocomposite ceramic body and finding the optimum condition to improve the properties of the
material.

2 EXPERIMENTAL PROCEDURE
Commercially available a-Al2O3 (A-11, Aldrich,
USA), b-SiC (B-1, Ibiden Co, Japan), and MgO
(Haghighatshimi Co, Iran) powders were used in
this study in which the alumina powder has the
purity of 99.6 per cent and average grain size of 3
mm; also, the SiC and MgO powders have average
grain size of 80 and 70 nm, respectively. The processing steps for hot-press sintering method were
outlined in Fig. 1. Six different mixtures of starting
powders Al2O3 and SiC, containing 0, 2.5, 5, 7.5, 10,
and 15 vol% SiC, were mixed with 0, 500, 1000, and
1500 ppm MgO. The mixtures are milled in a milling
machine (Fritsch, Pulverisette 5) with 17 tungsten
carbide grinding balls and the diameter of 2 cm, in
isopropyl alcohol environment at 250 r/min for 3 h.
Milled mixture initially dries in magnetic mixer at
100  C and at last, it dries in an oven at 130  C for
3 h. Since SiC particles are fine and they easily aggregate, it is very important to disperse SiC particles
homogeneously in Al2O3 [18]; therefore, after the
drying process, the prepared mixture was dry milled
for 24 h. Then, the mixture was sieved and analysed
with scanning electron microscope (SEM; Philips,
XL30 at 20 kV) apparatus but the powder size distribution was not measured.

Fig. 1

Process flow chart for the preparation of


nanocomposites

Proc. IMechE Vol. 225 Part L: J. Materials: Design and Applications


Downloaded from pil.sagepub.com by Narayanasamy P on October 18, 2011

268

A Asadi, S A Sadough Vanini, and A Farrokhi

The mixed powder (630 g) is compressed and


shaped to size of 120  120  12 mm3 at a pressure
of 20 bar for 30 s. Sintering of the powder compacts
was carried out by a uniaxial hot-press machine and
high-temperature induction furnace with argon
atmosphere at a pressure of 30 MPa and at various
temperatures including 1600  C, 1650  C, 1700  C,
and 1750  C for three sintering durations of 60, 90,
and 120 min, when heating rate was controlled at
10  C/min from room temperature to different end
temperatures; then, the furnace was turned off
and the specimens were slowly cooled to environment
temperature in the furnace. Because of the sintering in
induction furnace with graphite mould, a layer of
graphite will cover surfaces of the specimens. For
removing that layer, samples were heated at 900  C
for 4 h. Produced samples were cut and polished for
the microscopic analysis.
The densities of the sintered nanocomposites were
measured by immersing in distilled water using
Archimedes method based on ASTM B311 standard
[19]. Bending strength of the samples was measured
using three-point bending test based on ASTM C1161
standard [20]. For this test, sintered samples were cut
into rectangular bar specimens (3  4  45 mm3) with
a span length of 40 mm and test speed of 0.5 mm/
min. Hardness of the samples was measured using
Vickers method based on the ASTM C1327 standard
with indentation load of 10 kg for 30 s [21]. For measuring of the bending strength and hardness, three
tests were conducted and the mean value was
reported. Fracture toughness of material was calculated from previous measurements using the following equation [22]

KIC a

E
Hv

0:5 

P
d 1:5


3

where a is a constant coefficient independent of


material and equal to 0.16  0.04, as proposed by
Anstis et al. [23] in which 0.016 was used, d the
radial/median crack length (mm), E the Youngs
modulus (GPa), H the Vickers hardness, and P the
indentation load (N). For estimation of the Youngs
modulus of material, the slope of the stressstrain
measurements from the three-point bending strength
test was used based on the ASTM D790 standard. For
estimation of properties, always three samples were
used.
For analysis of physical and mechanical properties
from different locations of the nanocomposite
ceramic body, the body with the optimized properties
is divided into nine equal parts (40  40  12 mm3)
and the properties are measured independently in
every section for comparison.

3 RESULTS AND DISCUSSION


This section is divided to three parts. (1) First part
has focused on the physical and mechanical
optimization of the nanocomposite based on the
MgO addition amount. For achieving this task, a
set of four specimens were produced with MgO
amounts of 0, 500, 1000, and 1500 ppm at sintering
temperature of 1650  C, sintering pressure of 30 MPa,
and SiC vol% of 10 per cent. (2) The second part has
focused on the optimization of physical and mechanical properties of SiC vol% and sintering temperature,
which was conducted at the sintering pressure of
30 MPa and with the MgO addition, which has
attained from the first part. (3) The third part has
focused on the physical and mechanical optimization
of the sintering duration, which was conducted at
different sintering durations of 60, 90, and 120 min
in the optimum conditions, which had attained
from the previous parts.
3.1 Optimization of MgO addition
Since the purpose is to find the nanocomposite
ceramic with optimized properties in addition of
SiC particles and small amount of MgO simultaneously, the tests for different MgO amounts for this
part were carried out with addition of some SiC
particles from which 10 vol% SiC was chosen for
this part based on the results obtained in literature
[11]. Physical and mechanical properties of the
material versus different MgO additions at sintering
temperature of 1650  C and pressure of 30 MPa
for sintering duration of 120 min were shown in
Table 1. It was seen that relative density had an
ascending characteristic with relation to addition of
MgO. The rate of the increase is lower at higher
amount of MgO and relative density for the best
case reached 99.34 per cent with addition of
1000 ppm MgO.
The energy dissipation criterion has been presented in Table 1 for different MgO additions.
Nonetheless that maximum mechanical properties
have obtained from the specimen with 500 ppm
MgO, mechanical properties of specimen with MgO
amount greater than 500 ppm have improved in comparison with the specimen without MgO. With
increasing the amount of MgO more than 500 ppm,
the bending strength, Youngs modulus, energy
dissipation, and hardness were decreased unlike the
fracture toughness that had an ascending characteristic with addition of the MgO. Increase of the fracture
toughness with addition of the MgO is related to
the densification of the material and reduction
of the pores. Also, in Table 1, deviation of the

Proc. IMechE Vol. 225 Part L: J. Materials: Design and Applications


Downloaded from pil.sagepub.com by Narayanasamy P on October 18, 2011

Optimization of mechanical properties of Al2O3SiC nanocomposite body

Table 1

269

Physical and mechanical properties of nanocomposite with 10 vol% SiC and sintering
duration time of 120 min for different MgO addition amounts at sintering temperature
of 1650  C

MgO amount
(ppm)

Relative
density
(vol%)

Bending
strength
(MPa)

Vickers
hardness

Youngs
modulus
(GPa)

Fracture
toughness,
KIc (MPa m1/2)

Energy
dissipation,
D (1012 s1)

Energy
dissipation
deviation (%)

0
500
1000
1500

98.05
99.13
99.34
99.28

321.6
391.3
377.8
356.4

1650
1930
1880
1850

209.5
307.8
293.2
288.5

3.01
3.23
3.45
3.55

1.07
1.92
1.52
1.38

0
79.4
42.05
28.97

Fig. 2

Relative density of Al2O3SiC as a function


of sintering temperature and the volume fraction of SiC, with 500 ppm MgO and sintering
duration of 120 min

energy dissipation is listed, which is calculated as the


percentage difference between the energy dissipation
of ceramic without MgO and other specimens with
different volume percents of MgO particles. It was
seen that the specimen with 500 ppm had the maximum energy dissipation which with considering the
energy dissipation deviation, its value is nearly
1.8 times of the ceramic without MgO. This indicates
that the developed ceramic with addition of small
amount of MgO may have a benefit in comparison
with the ceramic without the addition of the MgO
and the MgO amount of the 500 ppm was chosen
for the rest of the analysis.
3.2 Sintering temperature and SiC vol%
optimization
Relative density values of the prepared nanocomposites versus sintering temperature were shown for
various volume percents of nanoSiC particles in
Fig. 2. Because of the existence of 500 ppm MgO
which was attained from Section 3.1, relative densities
for 0 vol% SiC are so close to the theoretical relative
density of the nanocomposite for all sintering temperatures. Actually, relative densities, for specimens with
sintering temperature of 1750  C and for all SiC

volume percents, were measured higher than 99 per


cent. A small amount of MgO in alumina ceramics, in
addition to the high temperature and pressure during sintering, caused the grain-boundary mobility
to decrease. Therefore, the abnormal grain growth
would be reducing. On the other hand, it would
increase the surface diffusivity and therefore, increasing the pore mobility which would decrease the pores
and consequently, it would cause the densification of
nanocomposites.
It was showed that with increase of SiC vol%, relative
density values for all sintering temperatures were
decreased and this reduction is more significant at
lower sintering temperatures. At lower sintering temperatures, most of the SiC particles have dispersed on
the grain boundaries rather than to disperse in alumina
grains; consequently, the grain-boundary diffusion path
becomes longer in proportion to the size of the SiC particles. Also, it was observed that higher sintering temperatures would improve the diffusivity of the grain
boundary and lattice in which its rate has decreased by
addition of SiC nanoparticles. Consequently, for reaching higher relative densities with addition of SiC particles, higher sintering temperature would be needed.
SEM pictures of nanocomposite specimens from a
surface area with thermal etching at 10 vol% SiC for sintering temperature of 1650  C were shown in Fig. 3. The
SiC particles were shown as dispersed white spots which
some of them were indicated with arrows. It was shown
that grain size of the alumina matrix was significantly
reduced by dispersion of SiC particles, because the SiC
grain particles had decreased the grain-boundary
mobility. It was seen for Al2O3 without MgO in
Fig. 3(a) that alumina matrices have abnormal grain
growth while grain growth has been decreased by addition of 500 ppm MgO particles in Fig. 3(b). Decrease of
grain growth is because of the SiC particles, which have
placed on grain boundary of alumina matrix. It was
observed that effects of SiC particles on grain growth
were trivial at low SiC vol% and high sintering temperatures, which for these conditions most of the nanoSiC
particles were placed in the alumina matrix.
The results for bending strength tests versus sintering temperatures for various SiC vol% were shown in
Proc. IMechE Vol. 225 Part L: J. Materials: Design and Applications

Downloaded from pil.sagepub.com by Narayanasamy P on October 18, 2011

270

A Asadi, S A Sadough Vanini, and A Farrokhi

Fig. 4

Fig. 3

Surface area SEM picture of a hot-press sintering specimen in sintering temperature of


1650  C, sintering duration time of 90 min, and
the pressure of 30 MPa with 10 vol% SiC:
(a) without MgO and (b) with 500 ppm MgO

Fig. 4. Maximum bending strengths were attributed


to specimens with 10 vol% SiC in sintering temperatures of 1650  C and higher than that. It shows
that bending strength has risen from 251 MPa, for
specimen without SiC particles, to 391 MPa for specimen with 10 vol% SiC and 500 ppm MgO at sintering
temperature of 1650  C. In this specimen, bending
strength of Al2O3 ceramic was increased by 55
per cent. It was shown that addition of SiC nanoparticles to 10 vol% has increased the bending strength of
the nanocomposite specimen. Addition of SiC nanoparticles to 10 vol% has increased the bending
strength of the nanocomposite, which is for the fact
that grain size of alumina matrix is decreased by addition of the SiC particles. Also, thermal expansion

Bending strength of Al2O3SiC as a function of


sintering temperature and the volume fraction
of SiC, with 500 ppm MgO and sintering duration of 120 min

mismatch between SiC particles and Al2O3 matrix has


produced residual tensile stress in Al2O3 that helps to
increase the toughness of nanocomposite. Thermal
expansion of Al2O3 is higher than that of SiC particles;
so, during sintering cooling process, most of the SiC
particles were located in the Al2O3 grains and because
of the mismatch between the thermal expansions, SiC
particles would have less twitch in comparison to
alumina matrix. Therefore; residual tensile stress
would produce in Al2O3 grains which would have
effect on the toughness of nanocomposite [7].
It was conducted from the results in Fig. 4, for
the specimen with 15 vol% SiC that with increase of
SiC vol% after 10 vol%, the density was decreased.
On the other hand, tensile residual stress caused by
the thermal expansion mismatch was amplified
with addition of SiC particles, and nanocracks were
produced in alumina matrix, thereby reducing the
toughness. Figure 4 shows that with increase of
sintering temperature for specimens with 0 and
2.5 vol% SiC, bending strength decreased constantly.
Increase in grain growth with rise in the sintering
temperature caused the reduction in bending
strength of both 0 and 2.5 vol% SiC specimen.
For specimens with 510 vol% SiC particles, the bending strength would increase with the rising of the sintering temperature from 1600  C to 1650  C, because
of the higher melting temperature of SiC (2970  C) in
comparison to alumina (2322  C), due to perfect sintering of the nanocomposite at higher temperature.
Average grain size versus SiC vol% for different
sintering temperatures was shown in Fig. 5. The average grain size, G, was calculated using the relation
G

1:5L
MN

Proc. IMechE Vol. 225 Part L: J. Materials: Design and Applications


Downloaded from pil.sagepub.com by Narayanasamy P on October 18, 2011

Optimization of mechanical properties of Al2O3SiC nanocomposite body

where 1.5 is a geometry-dependent proportionality


constant, L the total test line length, M the magnification, and N the total number of intercepts. It can
be seen from Fig. 5 that the grain size is decreased
with addition of the SiC particles and increase of the
sintering temperature but at higher amounts of SiC,
effect of the sintering temperature on grain size are
very low.
Youngs modulus versus SiC vol% for sintering
temperatures from 1600  C to 1750  C was shown in
Fig. 6. It was observed that Youngs modulus had
maximum value between 7.5 and 10 vol% of SiC for
every sintering temperature. The SiC particles that
were dispersed on the grain boundaries locked the
alumina matrices together and have increased the
Youngs modulus of the material, but with addition

Fig. 5

Average grain growth of Al2O3SiC as a function


of sintering temperature and the volume fraction of SiC, with 500 ppm MgO and sintering
duration of 120 min

Fig. 6

Youngs modulus of Al2O3SiC as a function of


sintering temperature and the volume fraction
of SiC, with 500 ppm MgO and sintering duration of 120 min

271

of the SiC to 15 vol% Youngs modulus had a significant decrease because of the reduction of density and increase of pores. The Youngs modulus
were decreased with rising of the sintering temperature from 1650 to 1750  C, which shows the
effects of the sintering temperature on the dispersion of the SiC particles and reduction of the
density.
Hardness versus SiC vol% for sintering temperatures from 1600  C to 1750  C was shown in Fig. 7.
As it was seen, sintering temperature 1600  C had
lowest hardness which with addition of SiC had
reduced. Formation of SiC particle lattice and sintering properties quality reduction had decreased the
hardness; so, sintering temperature of 1600  C was
not the proper temperature. For sintering temperature increasing from 1650 to 1750  C, hardness was
decreased because of the grain growth. The grain
growth with increasing temperature was shown in
Fig. 5. Fracture toughness against SiC vol% for
sintering temperatures from 1600  C to 1750  C was
shown in Fig. 8. Because of proper sintering at higher
temperature, toughness was increased with temperature increase. At volume fractions of SiC over 5 vol%
by increasing the amount of SiC particles, a considerable fraction of SiC particles would be trapped at
grain boundaries. Therefore, with the increase of
SiC vol%, toughness has been extremely reduced
because of the thermal expansion mismatch effect
of SiC particles at the boundaries and intersection
of boundaries. Consequently, it would produce tensile residual stress in the material. Fracture toughness
of grain boundaries is usually lower than that
observed within the grains; therefore, addition of
SiC particles would change the fracture mode of
nanocomposite
from
intergranular
to

Fig. 7

Vickers hardness of Al2O3SiC as a function of


sintering temperature and the volume fraction
of SiC, with 500 ppm MgO and sintering duration of 120 min

Proc. IMechE Vol. 225 Part L: J. Materials: Design and Applications


Downloaded from pil.sagepub.com by Narayanasamy P on October 18, 2011

272

A Asadi, S A Sadough Vanini, and A Farrokhi

transgranular fracture. The transgranular fracture


mode significantly helps the strength of grain boundaries and resists against crack propagation.
For better understanding of the magnitude of the
residual stresses caused by dispersion of SiC particles
in the alumina grains and its boundaries, these two
types of dispersion have been analysed here. The
residual stress caused by a SiC particle placed in the
intersection of the boundaries as is shown in Fig. 9(a)
can be calculated by [7]
w

E
m  d T
21  v 2

where w is the opening stress as the result of the


thermal expansion mismatch between Al2O3 matrix
m , SiC dispersed d , and T is the range of temperature over which diffusional relaxation at grain
boundaries is not allowed in polycrystalline Al2O3.
Considering
E 395 GPa, v 0:25, m  d

4:0  106 C1 , and T 1180  C [24], the tensile


residual stress caused by this phenomenon would
be 0.99 Gpa at the intersection of the boundaries.
On the other hand, if the SiC particle be placed in
the middle of the Al2O3 matrix as it is shown in Fig.
9(b), a compressive stress would be created in the
Al2O3 matrix as the result of the thermal expansion
mismatch of the SiC particles and Al2O3 at the time of
sintering. For estimation of this compressive residual
stress, first it has been considered that the grain is
deformed as the result of the compressive stress produced in the grain. Then, forces have been applied on
the boundaries of the grain to reshape it to be able to
fill the cavity of the matrix. For the stress continuity
on the matrix and grain boundaries, forces equivalent
in magnitude and opposite in direction have been
applied on the boundaries of the matrix. Therefore,
the hydrostatic stress field generated in the matrix is
given by [7]
r Vf

E
m  d T
21  v 2

where Vf is the volume fraction of the SiC particles


in alumina. In Fig. 9(b), the intersection of the boundaries and the hydrostatic stress field acting on them is
depicted. The stress field generated at any arbitrary
point (x,z ) by a stress, p , distributed homogeneously
along a grain-boundary line at any arbitrary angle,
, can be expressed as [7]
p cos 
4
z  sin 
2
 2z sin   x cos  x 2 z 2

xx x, z
l
0

Fig. 8

Fracture toughness of Al2O3SiC as a function of


sintering temperature and the volume fraction
of SiC, with 500 ppm MgO and sintering duration of 120 min

(a)

Fig. 9


1v


21 v cos   x2
d
2 2z sin   x cos  x 2 z 2

(b)

Grain boundary with (a) the tensile residual wedge opening microstress caused by a single
SiC particle placed at boundary intersection and (b) the stress field caused by the thermal
expansion mismatch by SiC particle placed in the middle of the grains

Proc. IMechE Vol. 225 Part L: J. Materials: Design and Applications


Downloaded from pil.sagepub.com by Narayanasamy P on October 18, 2011

Optimization of mechanical properties of Al2O3SiC nanocomposite body

p cos 
4
z  sin 
2
 2z sin   x cos  x 2 z 2

zz x, z
l
0


3v


21 v cos   x2
d
2 2z sin   x cos  x 2 z 2
8

p cos 
4
 cos   x
2 2z sin   x cos  x 2 z 2

xz x, z
l
0


1v 


21 v cos   x2
d
2 2z sin   x cos  x 2 z 2
9

where  is the direction taken along the selected


grain boundary of length l For calculation of the
stress component xx , which arises from the thermal
expansion mismatch between Al2O3 and SiC phases
along boundary AB, the residual stresses caused by
the boundary AB itself and its neighbouring boundaries CA, DA, BE, and BF must be included. With integration of equations (7) to (9), the stress distribution
along boundary AB has been computed. The residual
stresses for 10 vol% SiC, which has the optimum
bending strength in Fig. 4 along the boundary AB,
are presented in Fig. 10. The value of stress at intersection of the boundaries has singularity, and also, it
can be seen that the stress generated by a SiC particle
that placed in the middle of alumina grain would
cause a compressive stress on the boundaries.
Detail description is described elsewhere [7].
The energy dissipation D versus SiC vol% at different sintering temperatures has been shown in Fig. 11.
It was seen that sintering temperature of 1650  C

had the maximum energy dissipation criterion in


comparison with the other sintering temperatures.
Also, the energy dissipation had an ascending
characteristic from 0 to 10 vol% SiC for every sintering
temperature and for all the sintering temperatures,
the energy dissipation criterion has decreased for
15 vol% SiC nanoparticles which with considering
the energy dissipation formula of equation (1) and
Figs 6 to 8, it concludes that significant decrease of
Youngs modulus of E for every sintering temperature
has caused the decrease of the energy dissipation
criterion at 15 vol% SiC. Considering the results from
physical and mechanical properties such as density,
bending strength, hardness and toughness, and
energy dissipation criterion D from Fig. 11, specimens
containing 10 vol% SiC and sintering temperature of
1650  C were chosen for optimum condition for problems with high-speed impacts. For a few mechanical
and physical properties such as fracture toughness
and relative density, selected conditions did not
have the best properties between all tested specimens
but the optimum condition chosen considering with
all the physical and mechanical properties as a whole,
also for better fracture and density properties, a
SiC vol% between 7.5 and 10 is proposed.
3.3 Sintering duration time optimization
Physical and mechanical properties for three sintering duration times 60, 90, and 120 min in sintering
temperature of 1650  C were outlined in Table 2. It
was seen that at higher sintering duration times
because of the abnormal grain growth of Al2O3,
the relative density has increased. Considering
the mechanical properties, it was concluded that
sintering duration of 60 min in sintering temperature
of 1650  C was not enough for proper boundary

Fig. 11
Fig. 10

Residual stress caused by the thermal expansion mismatch along boundary AB for the SiC
particle placed in the middle of grains

273

Energy dissipation of Al2O3SiC as a function


of sintering temperature and the volume fraction of SiC, with 500 ppm MgO and sintering
duration of 120 min

Proc. IMechE Vol. 225 Part L: J. Materials: Design and Applications


Downloaded from pil.sagepub.com by Narayanasamy P on October 18, 2011

274

A Asadi, S A Sadough Vanini, and A Farrokhi

Table 2

Physical and mechanical properties for different sintering duration times with 10 vol%
SiC and 500 ppm MgO at 1650  C

Sintering
duration
time (min)

Relative
density
(vol%)

Bending
strength
(MPa)

Vickers
hardness

Youngs
modulus
(GPa)

Fracture
toughness
KIc (MPa m1/2)

Energy
dissipation,
D (1012 s1)

Energy
dissipation
deviation (%)

60
90
120

98.82
99.02
99.13

358.3
410.2
391.3

1900
1950
1930

312.1
321.2
307.8

3.19
3.22
3.23

1.98
2.08
1.92

0.03
0.08
0

Table 3

Fig. 12

Mapping of the different locations on the nanocomposite body (dimensions per millimetre)

surface diffusivity; consequently, low mechanical


properties were attainted for 60-min sintering duration time at sintering temperature of 1650  C.
Fracture toughness had ascending characteristics
with relation to the increasing of the sintering duration time. Higher sintering duration time has caused
the abnormal grain growth; therefore, more of the SiC
particles were located in the Al2O3 matrix and causing
the reduction in Youngs modulus of material. Also,
SiC particles which have located in the alumina
matrix would expand the FPZ size and improve the
fracture toughness of the materials. Because of the
abnormal grain growth, Vickers hardness and bending strength at sintering duration time of 120 min
were decreased. Energy dissipation deviation is calculated as the percentage difference between the
energy dissipation results of different sintering duration times and the duration time of 120 min. It can be
seen that the effect of sintering duration on energy
dissipation is more trifling in comparison to the
MgO addition effects. Considering with the above
conclusions, 90 min was chosen for the optimum sintering duration time.
3.4 Post-analysis of nanocomposite
ceramic body
For analysis of physical and mechanical properties
from different locations on the nanocomposite

Physical and mechanical properties of


nanocomposites in different locations on the
ceramic body produced in optimized
conditions

Location
number of the
measurements

Relative
density
(vol%)

Vickers
hardness

Fracture
toughness,
KIc (MPa m1/2)

Average
grain size
(mm)

1
2
3
4
5
6
7
8
9

99.559
99.521
99.545
99.523
99.512
99.526
99.553
99.529
99.541

1936
1953
1944
1958
1961
1950
1941
1957
1939

3.281
3.221
3.291
3.212
3.208
3.225
3.275
3.231
3.275

8.11
7.75
8.10
7.71
7.58
7.69
8.12
7.73
8.14

ceramic body, mapping of the ceramic body with


optimized properties is done as it can be seen in
Fig. 12. The body was divided into nine equal pieces
with the size of 40  40  12 mm3. In Table 3, physical
and mechanical properties of every piece are listed
and it can be seen that the piece number 5 has the
lowest mean grain size, which shows better dispersion of SiC and MgO in that area and as a result of
the decrease in grain size; consequently, density of
the body has its lowest value in the piece number 5.
Also, hardness has its higher value in the piece
number 5.
High-speed impact properties of the nanocomposite ceramic with optimized parameters of the
manufacturing procedure, which have been obtained
in this study, were examined for 7.62 projectile with
the speed of 820  15 m/s. The results in comparison
to the pure alumina show 30 per cent decrease in
surface density of targets. Detail discussions are
described elsewhere [25, 26].
4 CONCLUSIONS
1. All mechanical properties improved by addition
of the small amount of MgO in comparison to the
one without MgO, but the maximum mechanical
properties attained from the specimen with
500 ppm MgO.
2. Youngs modulus, hardness, bending strength, and
energy dissipation had decreased with addition of

Proc. IMechE Vol. 225 Part L: J. Materials: Design and Applications


Downloaded from pil.sagepub.com by Narayanasamy P on October 18, 2011

Optimization of mechanical properties of Al2O3SiC nanocomposite body

3.

4.

5.

6.

7.

more than 500 ppm MgO. However, relative density and fracture toughness have increased continuously with addition of MgO.
The relative density has been increased with addition of MgO, but it has descending characteristic
with addition of SiC particles.
Additions of 10 vol% of SiC and 500 ppm MgO
increased the bending strength of the ceramic
by 55 per cent in comparison to the pure Al2O3
ceramic.
Relative density, fracture toughness, and energy
dissipation criterion had an ascending characteristic with relation to the increase of the sintering
duration time, unlike the hardness, bending
strength, and Youngs modulus which had had
optimum property at sintering duration time of
90 min.
The optimum physical and mechanical properties
attained from the specimen with 10 vol% SiC,
500 ppm MgO at sintering temperature of
1650  C, sintering duration of 90 min, and sintering
pressure of 30 MPa.
From the measurement of properties in different
locations on the nanocomposite ceramic body, it
has been concluded that the area near intersection
of the diagonals of the body (location number 5)
has lower relative density and higher hardness in
comparison to other locations in the ceramic body.

Authors 2011

REFERENCES
1 Niihara, K. and Nakahira, A. Strengthening of oxide
ceramics by SiC and Si3Ni4 dispersion. In Proceedings
of the 3rd International Symposium on Ceramic materials and components for engines 3rd Int Symposium,
Las Vegas, 2730 November 1988, pp. 919926.
2 Niihara, K., Nakahira, A., Sasaki, G., and
Hirabayashi, M. Development of strong Al2O3/SiC
composites. In Proceedings of the 1st MRS
International Meeting on Advanced Materials,
Tokyo, Japan, 23 June 1988, pp. 129134.
3 Niihara, K. New design concept of structural
ceramicsceramic nanocomposites. J. Ceram. Soc.
Jpn, 1991, 99, 974982.
4 Davidge, R. W. Effect of microstructure on the
mechanical properties of ceramics. Fracture mechanics of ceramics, Vol. 2 (Eds R. C. Bradt, D. P.
H. Hasselman, and F. F. Lange), 1973, pp. 447468
(Plenum Press, New York).
5 Wang, H. Z., Gao, L., and Guo, J. K. The effect of nanoscale SiC particles on the microstructure of Al2O3
ceramics. Ceram. Int., 2000, 26, 391396.

275

6 Seong-Min,
Choi.
and
Hideo,
Awaji.
Nanocompositesa new material design concept.
Sci. Technol. Adv. Mater., 2005, 6, 210.
7 Pezzoti, G. and Muller, W. H. Strengthening mechanisms in Al2O3/SiC nanocomposites. Comput.
Mater. Sci., 2001, 22, 155168.
8 Medvedovski, E. Aluminamullite ceramics for structural applications. Ceram. Int, 2006, 32, 369375.
9 Neshpor, V. C., Zaitsev, G. P., Dovgal, E. J., et al.
Armour ceramics ballistic efficiency evaluation. In
Proceedings of the 8th CIMTEC on Ceramics: charting the future (Ed. P. Vincenzini), Florence, Italy, 28
June4 July1994, pp. 23952401 (TechnaS.r.l).
10 Kinsler, L. E., Frey, A. R., Coppens, A. B., and
Sanders, J. V. Fundamentals of acoustics, ed.4, 2000
(John Wiley and Sons Inc, New York, USA).
11 Wang, J., Lim, S. Y., Ng, S. C., Chew, C. H., and Gan,
L. M. Dramatic effect of a small amount of MgO
addition on the sintering of Al2O3-5 vol% SIC nanocomposite. Mater. Lett., 1998, 33, 273277.
12 Rittidech, A., Portia, L., and Bongkarn, T. The relationship between microstructure and mechanical
properties of Al2O3MgO ceramics. Mater. Sci.
Engng A, 2006, 438440395398.
13 Takayasu, Ikegami., Nobuo, Iyi., and Isao,
Sakaguchi. Influence of magnesia on sintering
stress of alumina. Ceram. Int., 2010, 36, 11431146.
14 Bennison, S. J. and Harmer, M. P. Grain-growth
kinetics for alumina in the absence of a liquid
phase. J. Am. Ceram. Soc., 1985, 68(1), C-22C-24.
15 Warman, M. O. and Budworth, D. W. Residual gas
effects on the sintering of alumina to theoretical
density in vacuum. Trans. Br. Ceram. Soc., 1967, 66,
265271.
16 Johnson, W. C. and Coble, R. L. A test of the secondphase and impurity-segregation models for MgOenhanced densification of sintered alumina. J. Am.
Ceram. Soc., 1978, 61(34), 110114.
17 Taylor, R. I., Coad, J. P., and Brook, R. J. Grain
boundary segregation in Al2O3. J. Am. Ceram. Soc,
1974, 57(12), 539540.
18 Wang, H. Z., Gao, L., Gui, L. H., and Guo, J. K.
Preparation and properties of intragranular Al2O3SiC nanocomposites. Nanostruct. Mater., 1998,
10(6), 947953.
19 ASTM B311 - 93(2002)el - Test method for density
determination for powder metallurgy (P/M) materials containing less than two percent porosity,
developed by Subcommittee, B09.11, Book of
Standards Volume: 02.05, 2002.
20 ASTM C1161 - 02c (2008) el - Standard test method
for flexural strength of advanced ceramics at ambient temperature, developed by Subcommittee,
C28.01, Book of Standards Volume: 15.01, 2008.
21 ASTM C1327 - 08 - Standard test method for Vickers
indentation hardness of advanced ceramics, developed by Subcommittee, C28.01, Book of Standards
Volume: 15.01, 2008.
22 ASTM C790 - 07el - Standard test method for flexural
properties of unreinforced and reinforced plastics
and electrical insulating materials, developed by

Proc. IMechE Vol. 225 Part L: J. Materials: Design and Applications


Downloaded from pil.sagepub.com by Narayanasamy P on October 18, 2011

276

A Asadi, S A Sadough Vanini, and A Farrokhi

Subcommittee, C20.10, Book of Standards Volume:


08.01, 2007.
23 Anatis, G. R., Chantikul, P., Lawn, B. R., and
Marshall, D. B. A critical evaluation of indention
techniques for measuring fracture toughness: I,
Direct crack measurements. J. Am. Ceram. Soc,
1981, 64(9), 533538.
24 Sergo, V., Wang, X. L., Clarke, D. R., and Becher, P.
F. Residual stresses in alumina/ceria stabilized zirconia composite. J. Am. Ceram. Soc, 1995, 78(8),
22132214.

25 Asadi, A. and Sadough, A. Investigation of the


impact behavior of Al2O3-SiC-MgO nanoceramic/
metal laminated composite. In Proceedings of
the 11th Iranian Conference on Manufacturing engineering, Tabriz, Iran, 1921 October 2010,
pp. 292923.
26 Asadi, A., Sadough, A., and Jabbari, A. Investigation
of the impact behavior of Al2O3-SiC-MgO nanoceramic/metal laminated composite. J. Mech. Sci.
Technol., 2011, 25(9), 16.

Proc. IMechE Vol. 225 Part L: J. Materials: Design and Applications


Downloaded from pil.sagepub.com by Narayanasamy P on October 18, 2011

También podría gustarte