Sie sind auf Seite 1von 7

The mechanical response of Al/Si/Mg/SiC

p
composite: inuence of
porosity
C. Tekmen
a,
*, I. Ozdemir
a,b
, U. Cocen
a
, K. Onel
a
a
Department of Metallurgical and Materials Engineering, Faculty of Engineering, Dokuz Eylul University, Bornova, Izmir 35100, Turkey
b
Material Process Laboratory, Toyota Technological Institute, 2-12-1 Hisakata, Tempaku 468-8511, Japan
Received 10 February 2003; received in revised form 2 June 2003
Abstract
The effect of porosity on the mechanical and fracture behaviour in Al /Si matrix alloy and composites reinforced with SiC
particles of 10 and 20 vol.% in the as-cast state and after extrusion process has been studied. Matrix alloy and composites were
fabricated by compocasting and extrusion. Samples were characterized by optical microscopy, image analyzer, scanning electron
microscopy and tensile tests. The results demonstrate that hot extrusion considerably reduces the porosity, while size and
distribution of the reinforcement particles are also affected. In the point of fracture behaviour, the existence of large porosity is more
effective.
# 2003 Elsevier B.V. All rights reserved.
Keywords: Metal matrix composite; Porosity; Strength; Extrusion
1. Introduction
The use of Al /Si alloys in the manufacture of
automotive engine components, such as cylinder blocks,
cylinder heads, pistons and piston rings, has dramati-
cally increased. The principal usage of these alloys is
their manufacturability, high wear resistance, low ther-
mal expansion coefficient, good corrosion resistance,
low density and improved elevated temperature proper-
ties [1/4]. It is generally known that the addition of
ceramic particulates to these alloys increases the high
temperature strength, strength and wear resistance at
ambient temperature [5,6]. Nevertheless, some defects in
the cast microstructure will occur during the manufac-
ture of particulate metal matrix composites (MMCs)
using the processing techniques, such as melt-stirring or
powder metallurgy. Defects will undermine the perfor-
mance characteristics and casting quality [7].
An important defect is porosity, which tends to cause
a reduction in mechanical and fatigue properties of the
composites [8]. Porosity formation can be attributed to
two factors: (a) shrinkage coupled with a lack of
interdentritic feeding during mushy zone solidification
and (b) evolution of hydrogen gas bubbles due to a
sudden decrease in hydrogen solubility during solidifica-
tion [7/9]. The solubility is a function of temperature,
pressure and alloy composition. After solidification is
complete, these bubbles become micropores. Once
porosity forms, the pores will grow until they have
reached equilibrium between the forces acting on them
to include pressure, hydrogen solubility and interfacial
energy. On the other hand, it has been observed that
reinforcement particles have a tendency to associate
themselves with porosity, thereby giving rise to particle-
porosity clusters [1]. The melt stirring method is
economical, easy to apply and convenient for mass
production. However, in this technique the mixing of
reinforcement with the molten metal has problems, such
as low wettability and particle settling. Increasing the
liquid temperature, coating or oxidizing the reinforce-
ment particles, adding some surface-active elements
such as magnesium and lithium into the matrix [10]
and stirring of molten matrix alloy for an adequate time
period during incorporation are some viable ways used
to eliminate the defects during casting. The porosity in
* Corresponding author. Tel.: /90-232-388-2880; fax: /90-232-
388-7864.
E-mail address: cagri.tekmen@deu.edu.tr (C. Tekmen).
Materials Science and Engineering A360 (2003) 365/371
www.elsevier.com/locate/msea
0921-5093/03/$ - see front matter # 2003 Elsevier B.V. All rights reserved.
doi:10.1016/S0921-5093(03)00461-1
composites that enters the melt along with the reinforce-
ment during stir mixing can be eliminated by applying
pressure during solidification [11] and the application of
plastic forming processes to the composites [12/16]. In
other words, thermomechanical processes, such as
forging, extrusion, and rolling, are important methods
that can be used to improve the mechanical properties of
MMCs and produce standard products having stable
properties. It has been reported that improvements in
strength and ductility are observed with the application
of plastic forming processes to the composites [12,16/
19]. The observed improvement in ductility and proper-
ties is attributed to (a) the decrease in porosity content,
(b) better interfacial bonding between particle and
matrix and (c) the refinement of the matrix structure.
Studies conducted on the effect of reinforcement on
porosity showed that the porosity size and shape was
affected by the presence of reinforcement [7]. In general,
SiC particles tend to restrict the growth of pores. Also,
increasing the reinforcement volume fraction signifi-
cantly increases the pore count [1,7]. This could be
attributed to the longer stirring time that causes a larger
amount of gas dissolved in the molten metal vortex.
The formation of porosity and its effect on mechan-
ical properties of MMCs have been the matter of several
studies [20,21]. It is generally accepted that tensile
properties decrease with an increase in porosity content.
However, the effect of parameters, such as porosity
type, volume fraction, size and distribution, process
variables (e.g. mould and pouring temperature) and
hydrogen content on mechanical properties and fracture
behaviour are still incomplete. In this study, the effect of
porosity on mechanical properties and fracture beha-
viour, in addition to the role of reinforcement on
porosity formation in Al /Si /Mg/SiC composites, have
been investigated using optical microscopy, image
analyzer, scanning electron microscopy and tensile
testing.
2. Experimental
The chemical composition of the matrix alloy is given
in Table 1. The matrix alloy was reinforced with 10 and
20 vol.% of silicon carbide particles (SiC
p
) by using the
compocasting technique. The production of the matrix
alloy and composites used in the present study was
carried out as follows: SiC particles peroxidised at
900 8C were added to the semi-solid matrix alloy at
600 8C. An argon atmosphere was maintained over the
melt to reduce oxidation. The mixture was rapidly
heated to 750 8C and the composite slurry was poured
into a preheated permanent iron die (150 8C). In order
to minimize solidification porosity and obtain better
mechanical properties, the samples were hot extruded.
Extrusion of the ingots was performed at an extrusion
ratio of 10:1.
It has been observed that the produced samples
contain various types of porosity. Composite samples
with high and low content porosity were studied and an
average porosity content of the produced samples in as-
cast and extruded conditions was determined using the
Archimedean method. Microstructural characterization
studies were carried out in order to determine porosity
size distribution, area fraction and distribution of the
porosity by using the optical microscopy and image
analyzer. Tensile samples were machined from the
extruded bars, having a gauge length of 16 mm and
diameter of 6 mm. Tensile tests were conducted at a
cross head speed of 0.5 mm s
1
. The fractured surfaces
were characterized by scanning electron microscope
(SEM) and optical microscope.
3. Results and discussion
3.1. Porosity in as-cast condition
Typical microstructures of the composites in as-cast
condition are shown in Fig. 1. It is evident from Fig. 1
that different types of porosity, such as gas bubbles and
particle-porosity clusters, are present in the structure,
which are inevitable when the composites produced by
casting technique.
In order to determine the overall porosity content,
density measurements were conducted on unreinforced
and composites reinforced with 10 and 20 vol.% SiC
particles and the results are given in Table 2. The
difference between the calculated density (D
c
), which
was obtained by using chemical composition of the
composites and experimental density (D
e
) is due to the
existence of porosity in the structure. The effect of
reinforcement volume fraction on observed porosity of
the composite samples in as-cast condition is illustrated
Table 1
Chemical composition of the matrix alloy used in the present study
Material Composition (wt.%)
Si Mg Cu Mn Ni Zn Fe Al
Matrix alloy 6.62 0.67 0.013 0.027 0.008 0.08 0.298 Bal.
C. Tekmen et al. / Materials Science and Engineering A360 (2003) 365/371 366
in Fig. 2. The size distribution of porosity in the as-cast
samples are shown in Fig. 3. An increase in reinforce-
ment volume fraction increases the porosity content
(Table 2). In contrast, Bindumadhavan et al. [1] and
Samuel et al. [7] evaluated on particle cluster type
porosity and found that the presence of a larger volume
fraction of SiC particles physically restricts the growth
of porosity and thus reduces the overall porosity
content. Difference in results can be attributed to the
fact that measurements in this study were carried out on
both gas bubbles and shrinkage type porosity, thus the
overall porosity size was found greater.
Reinforcement particles have a tendency to associate
themselves with porosity and give rise to particle-
porosity clusters [1,7]. In the present study, compared
to 10 vol.% SiC
p
composite, a greater number of
particle-porosity clusters were found for the 20 vol.%
SiC
p
composite (see Fig. 4). In contrast, Bindumadha-
van et al. [1] studied a A356 matrix alloy reinforced with
SiC particles having an average particle size of 47 mm,
observed that the tendency for formation of particle-
porosity clusters was greater in low volume fraction
composites. They explained that in high volume fraction
composites, the geometric capturing of particles restricts
their movement inside the melt during solidification.
The difference in test results can be attributed to the
following reasons: (i) for the same reinforcement volume
fraction, due to lower reinforcement particle size (12 mm,
in this study) causes an increase in nucleation sites for
porosity at the SiC/matrix interface; and (ii) due to
improper feeding, where fluidity of liquid metal is
insufficient to fill the gaps between adjacent particles.
Fig. 1. Optical micrograph of 20 vol.% SiC
p
composite in as-cast
conditions (a) typical microstructure, (b) particle-porosity clusters and
(c) gas porosity.
Table 2
Density and porosity values of the matrix alloy and composites in as-cast and extruded conditions
Materials Calculated density D
c
(g cm
3
) Experimental density D
e
(g cm
3
) Porosity (%)
As-cast Extruded As-cast Extruded
Matrix alloy 2.6917 2.6558 2.6670 1.33 0.92
Al /10% SiC
p
2.7427 2.6162 2.7135 4.61 1.06
Al /20% SiC
p
2.7942 2.4342 2.7610 12.88 1.19
Fig. 2. The change of porosity content of the samples as a function of
SiC
p
volume fraction.
C. Tekmen et al. / Materials Science and Engineering A360 (2003) 365/371 367
3.2. The effect of extrusion process on porosity
The difference between the as-cast and hot-extruded
composite microstructures is that the SiC
p
clusters
initially present in some areas in the as-cast composite
have disappeared, giving a more uniform distribution of
SiC
p
, as shown in Fig. 5. The microstructures revealed
that the number of resolvable pores is reduced, some
particle fragmentation is noticeable and some particle
orientation in the direction of extrusion has taken place
during hot-extrusion. A detailed microstructural char-
acterization and X-ray diffraction analyses of the matrix
alloy and composites produced in the same manner was
demonstrated in the previous study [18].
Fig. 3. The porosity size distribution of the composite samples in (a, b) as-extruded and (c, d) as-cast states.
Fig. 4. Typical particle-porosity cluster formed in 20 vol.% SiC
p
composite.
Fig. 5. Optical micrograph of typical microstructure of 20 vol.% SiC
p
composite in extruded conditions.
C. Tekmen et al. / Materials Science and Engineering A360 (2003) 365/371 368
Porosity volume fractions were found to vary between
1.33 vol.% for the unreinforced alloy and 12.88 vol.%
for the 20 vol.% SiC
p
composite in the as-cast condition.
Following hot-extrusion, the porosity levels were found
to be 0.92 vol.% for the unreinforced alloy and 1.19
vol.% for the 20 vol.% SiC
p
composite (Table 2).
Consequently, extrusion decreases the porosity content
while also changing the shape of the porosity. Similar
results also obtained from image analyses (Table 3). In
addition, the average porosity size after extrusion is
considerably reduced compared to the as-cast condition
(Fig. 3; Table 3).
Application of the secondary deformation processes
to discontinuously reinforced composites results in to
break up of particle or whisker clusters, reduction or
elimination of porosity and improved bonding charac-
teristics between particle and the matrix. Therefore, the
secondary process is very important in improving the
mechanical properties of MMCs, while it is an essential
step in engineering applications of MMCs for producing
standard products having stable properties. Rozak et al.
reported that the porosity level of A356 alloy based SiC
p
reinforced composites can be reduced by plastic work-
ing. Moreover, if the amount of applied deformation is
around 90%, porosity can be eliminated [12]. Zhang et
al. observed that after applying isothermal hot indirect
extrusion to cast Al /2024/SiC
p
composites, the content
of porosity reduced from 5.56 to 0.56% for an extrusion
ratio of 39 [22]. As reported earlier, Ozdemir et al.
investigated the effect of the hot forging on microstruc-
ture and mechanical properties of SiC
p
reinforced Al /Si
alloy based composites having similar compositions to
those used in the present work. It was pointed out that
the porosity level of the composites could be reduced to
below 1.5 wt.% by application of hot-forging [16]. This
is a higher value than obtained during extrusion of the
composite. Although extrusion decreases the porosity
level, as shown in Fig. 3, it is not effective in eliminating
porosity, which exceeds a critical size. On the other
hand, by increasing the extrusion ratio, it is possible to
eliminate the large-sized porosity but this could ad-
versely affect mechanical properties due to particle
fracture and particle/matrix debonding, which is con-
ducive for void formation. Further, the extrusion
process affects the shape of the porosity by flattening
the porosity along the extrusion direction, as shown in
Fig. 6.
3.3. The effect of porosity on mechanical behaviour
When the composites are fabricated by melt-stirring,
the bonding strength maybe lowered by porosity and
segregation at the interface between the matrix and the
reinforcement. When pores are located at the boundary
of matrix and particles, they cause debonding of
particles from the matrix under low stress and reduce
the ability of load transfer to the particle and minimal
strengthening is achieved. The second type of porosity
located away from the particles tends to reduce the
effective area supporting the load and is detrimental to
strength [21,22]. When porosity or an equivalent defect
is present in a sample, the load bearing area is reduced.
It can be safely assumed that the defective region will
yield first, thereby concentrating the strain. On the other
hand, voids present in the cross section create a multi-
axial stress state and cause local strain concentrations in
their vicinity [8]. It is clear from Fig. 7 that an increase in
porosity content decreases both yield and ultimate
tensile strength values of the produced samples.
Fig. 8 illustrates the micrographs near the fracture tip
of 10 vol.% SiC
p
composite having low porosity content.
As seen from Fig. 8, fracture damage of the 10 vol.%
SiC
p
composite occurred as particle cracking (PC) and
interfacial debonding (ID) for the 20 vol.% SiC
p
composite having high porosity content, the existence
Table 3
The results of porosity measurements of the composite samples in as-extruded and as-cast conditions
Materials Measured area (mm
2
) Mean size (mm) Min. size (mm) Max. size (mm) Porosity area fraction (%)
As-extruded Al /10% SiC
p
3.4/10
7
25.1 1.7 117 1.2
Al /20% SiC
p
3.7/10
7
35.6 10.9 180 2.2
As-cast Al /10% SiC
p
5.8/10
7
30.8 6.2 158 4.5
Al /20% SiC
p
6.1/10
7
65.3 14.9 276 9.6
Fig. 6. The effect of extrusion on porosity shape in 20 vol.% SiC
p
composite.
C. Tekmen et al. / Materials Science and Engineering A360 (2003) 365/371 369
of a large size porosity causes fracture, as shown in Fig.
9. The micrograph clearly shows that almost all the
particles are actually broken, since SiC particles were
observed on the fracture surface and the surface
contains large-sized porosity. Caceres and Selling [8]
carried out studies on Al /Si /Mg alloys and found the
tensile strength to show little or no correlation with bulk
porosity content, while the mechanical performance
decreased with an increase in area fraction of defects
on the fracture surface of the samples. Also, for some
samples it was observed that having low porosity
content the mechanical results are lower compared to
samples having a higher porosity content. This result is
attributed to the extrusion process, which is effective in
eliminating porosity lower than a critical size. For the
case of large-sized porosity, the extrusion process aids in
reducing the size and changing the shape of the pores. It
is possible to draw a conclusion that the average
porosity content is not a reliable parameter to predict
the mechanical properties. Further, large-sized pores,
which cannot be eliminated through extrusion, is an
important factor that promotes fracture while reducing
the strength of the composites.
4. Conclusions
The conclusions obtained are summarized in the
following paragraphs.
The increase in reinforcement volume fraction in-
creases the overall porosity content in both as-cast and
extruded conditions. With the application of the extru-
sion process, the porosity content and size is substan-
tially reduced to low levels. However, the extrusion
process is only effective to eliminate porosity, which is
lower than a certain size.
The increase in porosity content decreases both the
yield and ultimate tensile strength values of the pro-
duced samples. The average porosity content is not a
reliable parameter to predict the mechanical results. In
the point of fracture behaviour, the existence of large
pores is more effective than the overall porosity content.
References
[1] P.N. Bindumadhavan, T.K. Chia, M. Chandrasekaran, H.K.
Wah, L.N. Lam, O. Prabhakar, Mater. Sci. Eng. A315 (2001)
217/226.
[2] O. Mantaux, E. Lacoste, M. Danis, Comp. Sci. Tech. 62 (2002)
1801/1809.
[3] P.D. Lee, A. Chirazi, D. See, J. Light Metals 1 (2001) 15/30.
Fig. 7. The effect porosity content on (a) yield strength and (b)
ultimate tensile strength.
Fig. 8. Cross-section micrograph of tensile tested samples; damage as
particle cracking (PC) and interfacial debonding (ID) in 10 vol.% SiC
p
composite.
Fig. 9. SEM micrograph of 20 vol.% SiC
p
composite after tensile
testing.
C. Tekmen et al. / Materials Science and Engineering A360 (2003) 365/371 370
[4] Y.M. Li, R.D. Li, Sci. Tech. Adv. Mater. 2 (2001) 277/280.
[5] R.A. Siddiqui, H.A. Abdullah, K.R. Al-Belushi, J. Mater.
Process. Tech. 102 (2000) 234.
[6] M. Gu, Y. Jin, Z. Mei, Z. Wu, R. Wu, Mater. Sci. Eng. A252
(1998) 188.
[7] A.M. Samuel, A. Gotmare, F.H. Samuel, Comp. Sci. Tech. 53
(1995) 301/315.
[8] C.H. Caceres, B.I. Selling, Mater. Sci. Eng. A220 (1996) 109/116.
[9] J.P. Anson, J.E. Gruzleski, Mater. Char. 43 (1999) 319/335.
[10] A. Mortensen, I. Jin, Int. Mater. Rev. 37 (1992) 101/128.
[11] M.R. Ghomashchi, A. Vikhrov, J. Mater. Proc. Tech. 101 (2000)
1/9.
[12] G.A. Rozak, J.J. Lewondowski, J.F. Wallace, A. Atmisoglu, J.
Comp. Mater. 26 (1992) 2079/2106.
[13] M.G. McKimpson, T.E. Scott, Mater. Sci. Eng. A107 (1989) 93/
106.
[14] F.M. Hosking, F.F. Portillo, R. Wunderlin, R. Mehrabian, J.
Mater. Sci. 17 (1982) 477/498.
[15] W.C. Harrigan, G. Gaebler, E. Davis, E.J. Levin, in: J.E. Hack,
M.F. Amateau (Eds.), Mechanical Behaviour of Metal Matrix
Composites, Metallurgical Society, Warrendale, PA, 1983, p. 169.
[16] I. Ozdemir, U. Cocen, K. Onel, Comp. Sci. Tech. 60 (2000) 411/
419.
[17] Y.H. Seo, C.G. Kang, Comp. Sci. Technol. 59 (1999) 643/654.
[18] U. Cocen, K. Onel, Mater. Sci. Eng. A221 (1996) 187.
[19] U. Cocen, K. Onel, Comp. Sci. Tech. 62 (2002) 275/282.
[20] S. Ray, J. Mater. Sci. 28 (1993) 5397/5413.
[21] L. Molliex, J.P. Favre, A. Vassel, M. Rabinovitch, J. Mater. Sci.
29 (1994) 6033/6040.
[22] L.B. Zhang, H. Jintao, W. Yanwen, J. Mater. Process. Tech. 84
(1998) 271/273.
C. Tekmen et al. / Materials Science and Engineering A360 (2003) 365/371 371

Das könnte Ihnen auch gefallen