You are on page 1of 23

Particle reinforced aluminium and magnesium

matrix composites
D. J. Lloyd

Particle reinforced metal matrix composites are


now being produced commerically, and in this
paper the current status of these materials is
reviewed. The different types of reinforcement
being used, together with the alternative
processing methods, are discussed. Depending on
the initial processing method, different factors
have to be taken into consideration to produce a
high quality billet. With powder metallurgy
processing, the composition of the matrix and the
type of reinforcement are independent of one
another. However, in molten metal processing they
are intimately linked in terms of the different
reactivities which occur between reinforcement
and matrix in the molten state. The factors
controlling the distribution of reinforcement are
also dependent on the initial processing method.
Secondary fabrication methods, such as extrusion
and rolling, are essential in processing composites
produced by powder metallurgy, since they are
required to consolidate the composite fully. Other
methods, such as spray casting, molten metal
infiltration, and molten metal mixing give an
essentially fully consolidated product directly, but
extrusion, etc., can improve the properties by
modifying the reinforcement distribution. The
mechanical properties obtained in metal matrix
composites are dependent on a wide range of
factors, and the present understanding, and areas
requiring further study, are discussed. The
successful commercial production of metal matrix
composites will finally depend on their cost
effectiveness for different applications. This
requires optimum methods of processing,
machining, and recycling, and the routes being
developed to achieve this are considered.
IMR/255

1994 The Institute of Materials and ASM International.


The author is with Alcan International
Ltd, Kingston
Research and Development
Centre, Kingston,
ON,
Canada.

application as selective reinforcement in the ring land


area of diesel pistons," and whisker reinforcement is
under development for aerospace applications.
Particle reinforced light metals, with their potential
as low cost, high modulus and strength, high wear
resistance, and easily fabricated material, are just
reaching the commercial production stage. .Understanding the factors that influence the physical and
, mechanical properties of these materials presents quite
a challenge, because they are sensitive to the type of
reinforcement,
the mode of manufacture,
and the
details of any fabrication processing of the composite
after initial manufacture. While there are still many
areas which are poorly understood, the work of the
past five years or so has identified many of the key
factors which have to be considered to achieve optimum properties.
The objective of the present paper is to give an
overview of the current understanding
of particle
reinforced light metal composites, identify some of
the key factors which need to be controlled - which
may be specific to the mode of manufacture
or
processing - and consider some of the main problem
areas which need to be overcome if these materials
are to reach their full commerical potential.
In a field which typically has about 20 publications
a month, it is inevitable that a review such as this
has to be highly selective. It will only consider Aland Mg-based matrices, even though Ti-based systems have considerable potential, particularly for high
temperature applications. Most of the published work
has considered AI-based composites, with their attractions of low density, wide alloy range, heat treatment
capability, and processing flexibility. Many of these
features are also exhibited by Mg-based systems, and
with its lower elastic modulus Mg often achieves a
larger property improvement with reinforcement than
does Al, Also, many of the composite fabrication
processes are common to both AI- and Mg-based
systems.

Introduction

The reinforcement

The development of metal matrix composites (MMCs)


has been one of the major innovations in materials
in the past 20 years. Many of the social and technological factors influencing this development have been
reviewed by Kelly.1 The early work on composites
considered continuous fibre reinforcement, and while
work in this area continues.' it was soon apparent
that the cost of continuous fibres, complex fabrication
routes, and limited fabricability, would restrict their
use to those applications requiring the ultimate in
performance. This led to the development of discontinuously reinforced composites, particularly
short
staple Al, 03 fibre" and SiC whisker" reinforced composites. Discontinuous fibres have found commercial

A recent overview of the short fibres and whiskers


available for metal matrix composites has been given
by Stacey." Since most ceramics are available as
particles, there is a wide range of potential reinforcements for particle reinforced composites. Rohatgi and
co-workers have used mica," alumina, silicon carbide,
and clay," zircon," and graphite, 10 and other
reinforcements, such as boron carbide!' and titanium
diboridc.F have also been examined. However, the
choice of reinforcement is not as arbitrary as this
list of composites might suggest, but is dictated by
several factors:
1. The application: if the composite is to be used
in a structural application, the modulus, strength, and
International

Materials Reviews

1994

Vol. 39

No.1

Lloyd

Particle reinforced AI and Mg matrix composites

density of the composite will be important, which


requires a high modulus, low density reinforcement.
Particle shape may be important,
since angular
particles can act as local stress raisers, reducing
ductility. If the composite is to be used in thermal
management applications, the coefficient of thermal
expansion and thermal conductivity are important.
The coefficient of thermal expansion is generally
important because it influences the strength of the
composite, as discussed in the section 'Strength' below.
2. The method of composite
manufacture:
as
considered in the next section, there are two generic
methods for composite manufacture, powder metallurgy and methods involving molten metal. In the
case of powder metallurgy, the matrix alloy powder
is blended with particles of the reinforcement
to
achieve a homogeneous mixture. To achieve this, the
sizes of the metal and ceramic powders need to be
carefully chosen so that agglomerates are not left after
blending, and carryover into the final product. The
appropriate
size ratio will depend on the blending
process used, but in one case a SiC/AI particle size
ratio of 07: 1 gave a more uniform reinforcement
distribution than a ratio of O'3 : 1 (Ref. 13). In powder
metallurgy processing, the brittle ceramic particles
are also susceptible to particle fracture, which is
dependent on the particle aspect ratio and flaw density. Typically, the atomised aluminium
powder
particle size is in the range 20-40 urn, and reinforcement particle sizes are 3-20 urn with aspect ratios
of < 5: 1.
For composites processed in the molten state, there
are different considerations.
In some of these processes, the ceramic particles can spend considerable
time in contact with the molten alloy matrix, and this
can result in reaction between the twO.14 For example,
SiC is thermodynamically
unstable in most molten
Al alloys, reacting to form aluminium carbide, Al4 C3,
whereas it is stable in many molten Mg alloys. On
the other hand, Al2 03 is stable in most Mg-free Al
alloys, but unstable in Mg ailoys, reacting to form
spinel, Al2 MgO 4' Reaction of the reinforcement can
severely degrade the properties of the composite, so
the reinforcement has to be chosen after considering
the matrix alloy, and the processing time and temperature. The reinforcement particle size is also important
because, while it is generally easier to incorporate
coarser particles into the melt, large particles are
more susceptible to gravity settling and can result
in a heavily segregated casting.P However, finer
particles increase the viscosity of the melt, making
processing difficult. Most molten metal processes use
ceramic particles in the 10-20 urn size range.
3. Cost: a major reason for using particles is to
reduce the cost of the composite, so the reinforcement
has to be readily available in the quantities, size, and
shape required at low cost, i.e, about US$5/kg.
Table 1

Cold Isostatic

Compaction

~Degassing

~~
Evaluation

Powder metallurgy

manufacturing

route

With these considerations


in mind, the <, two
reinforcements receiving the most attention are SiC
and Al2 3, and some of the properties of these two
reinforcements are given in Table 1.

Composite manufacture
While the details of the different manufacturing routes
are often proprietary, the important features are quite
well established.
Powder metallurgy
Because of the difficulty in wetting ceramic particles'
with molten metal, the powder metallurgy route was
the first method developed. The essential features of
the powder route are shown in Fig. 1. The metal
powder is usually prealloyed atomised powder, in the
20-40 urn size range, but it can also be a blend of
elemental powders, or rapidly solidified chopped
ribbon or flake, as in the Allied-Signal process.l" As
noted previously, the blending step is important since
this establishes the initial homogeneity of the composite. Because the metal powder particles have a
hydrated oxide film, it is necessary to remove the
associated water molecules before consolidation
if
subsequent gas porosity is to be avoided. The vacuum
hot pressing consolidates the composite to over 95%

Properties of SiC and AI203 "reinforcement


Elastic
modulus,

Particle

GN m-z

Density,
9 cm-3

SiC

420-450

32

Alz03

380-450

396

International

Materials Reviews

1994

Coefficient of
thermal expansion,
K-1

Specific
heat,
J kg-1 K-1

Thermal
conductivity,
Wm-1 K-1

Poisson's
ratio

4'3x10-6
70 x 10-6

840
1050

10-40,

at 11 OOC

017

at 1000C

025

Vol. 39

No.1

5-10,

Lloyd

dense, and can be carried out below the solidus of


the alloy, or in the liquid-solid region. The kinetics
of densification are higher using liquid phase sintering,
but it has the disadvantage that reaction can more
readily occur between the reinforcement
and the
liquid phase to form undesirable intermetallics at the
ceramic particle interface. Liquid phase sintering will
also degrade the microstructure of the rapidly solidified powder particles, and generate coarse eutectic
intermetallic phases in the melted regions. Completely
solid state consolidation
enables advantage to be
taken of the supersaturated metastable alloy compositions that can be obtained by the rapid solidification
atomisation and ribbon casting processes.
The final wrought product is obtained by extrusion,
with an extrusion ratio of about 20: 1 or higher. A
high extrusion ratio is required to disrupt the oxide
film between metal powder particles, allowing metal
to metal contact and the development of a good bond
between the metal particles. A high extrusion ratio
also improves the distribution of reinforcement, since
the plastic flow associated with extrusion tends to
disperse any clusters of reinforcing particles. However,
the extrusion ratio and temperature must be carefully
controlled to avoid reinforcement particle cracking,
and degradation
of the matrix if rapidly solidified
powders are being used.
The powder metallurgy route has several attractive features:
1. It allows essentially any alloy to be used as
the matrix.
2. It also allows any type of reinforcement to be
used because reaction
between the matrix and
reinforcement can be minimised by using solid state
processing.
.
3. Non-equilibrium
alloys can be used for the
matrix by using rapidly solidified material. This is
particularly important where the composite is to be
used for high temperature applications, and rapidly
solidified alloys have much better elevated temperature strength than conventional alloys."?
4. High volume fractions of reinforcement are possible, thus maximising the modulus and minimising
the coefficient of thermal expansion of the composite.
The powder route also has some major disadvantages: it involves handling large quantities of highly
reactive, potentially explosive powders and the manufacturing route is relatively complex and limited in
the initial product forms it can produce. As a result,
the product is expensive in comparison with conventional wrought aluminium alloys, and currently costs
over US$100/kg, but this should decrease with future
scale-up.
Molten metal methods
Early attempts to incorporate ceramic particles into
metallic melts had limited success because most metals
do not wet ceramic particles, and this results in
rejection of the particles from the melt.18,19 The basic
thermodynamics
associated
with incorporating
a
single particle into a melt have been considered in
some detail,20-23 and demonstrate that the contact
angle between molten aluminium and the ceramic
particle must be less than 90 for successful incorpor-

Particle reinforced AI and Mg matrix composites

Microstructure
of
Duralcan
606120 vol.-%AI203 composite after extrusion

ation. The wetting angle is usually measured by the


sessile drop method, and is influenced by several
variables including the heat of formation, stoichiometry, valence electron concentration,
interfacial
reactions, temperature, and time.i" In general, however, molten aluminium does not wet most ceramic
particles at typical casting temperatures, i.e. < 800C,
and the molten metal methods attempt to improve
this wetting behaviour.
Mixing methods
The early mixing method of Surappa and Rohatgi "
introduced ceramic particles through the sides of a
vortex created in the melt with a mechanical impeller.
This method is helped by the addition to the melt of
surface active elements, such as magnesium to aluminium melts,26,27 or metal coated particles." However,
the process is limited to coarse ceramic particles,
> 50 um, and low volume fractions, < 10 vol._%.
The major breakthrough in mixing processes came
with the development of the Duralcan (Duralcan
owned by Alcan Aluminum Corp.) process by Skibo
and Schuster.'? This process is applicable to conventional aluminium
alloys, using uncoated ceramic
particles of about 10 urn and larger, and produces
reinforcement levels of up to 25 vol._%. The typical
microstructure of a 6061/AI203/20p*
Duralcan composite after casting and extrusion is shown in Fig. 2.
The Duralcan material is now being commercially
produced in batches of up to 6800 kg, and cast as pig
or direct chill (DC) ingot.
Two other aluminium companies have announced
the development of a molten mixing method. Hydro
Aluminium AS have discussed composites which
appears to be comparable with Duralcan material. 30
Comalco has recently introduced
Comral, which
is 6061 reinforced with a spherical, mixed oxide,
AI203-Si02,
reinforcement." The extruded microstructure of this composite is shown in Fig. 3.
The molten metal mixing method is attractive
because, in principle, it allows all the conventional
*The suggested Aluminum Association designation for composites:
matrix/reinforcement/volume fraction and reinforcement type
(p refers to particle reinforcement) is used throughout.
International

Materials Reviews

1994

Vol. 39

No.1

Lloyd

Particle reinforced AI and Mg matrix composites


30-.----------------

C\J
I

E
z
V)

25
20

>t::: 15
(j)

u
(j)

10
5

O-+---.----r-~~~~~~~~E....J

o
3

Microstructure
of Comalco 6061-20 vol.-%
spherical oxide after extrusion

10 20 30 40 50 60 70 80 90 100
SHEAR RATE 5-1
I

Shear

rate

dependence

of

viscosity

of

A356-15 vol.-%SiC
metal processing routes to be used, and hence minimises the cost. At the present time, composite material
is available from Duralcan for about US$6/kg. It
does, however, have some problems in terms of reactivity between the reinforcement and the melt, and
particle segregation effects, both of which are considered in later sections. The volume fraction of reinforcement is also limited in the mixing method because
the viscosity of the melt increases with particle content
and becomes non-Newtonian.
The viscosity of unreinforced aluminium is about 10-3Nsm-2,'but
as
seen from Fig. 4, the viscosity of molten composites
is much larger, and shear rate dependent.F The
rheology of molten composites is poorly understood,
but appears to be history dependent." and increases
with increasing
volume fraction and decreasing
particle size. As a result, the power requirements
necessary for mixing limits the amount of reinforcement that can be used.
Semisolid casting
The microstructures developed by stirring a semisolid
melt were investigated in the early 1970s.34 The
partially
solidified,
non-dendritic
microstructure
developed has high viscosity which inhibits ceramic
particle settling and floating, and can be used to
retain particles in the melt. 35,36 This method has been
developed for magnesium alloys by the Dow Chemical
CO.37 The process is restricted to longer freezing
range alloys and, other than this, has the same
limitations as the fully liquid mixing methods.
Pressure infiltration
In this process, a packed bed of ceramic particles is
evacuated and then infiltrated by a pressurised melt
to form a 'master composite hardener' containing
about 50 vol-"
of reinforcement. This hardener can
then be diluted by adding it to an unreinforced melt,
and the reinforcement
redispersed
by submerged
mixing.Pr" However, this dispersion is quite difficult
because of the' high viscosity of the master hardener,
and it is difficult to obtain dispersion without incorporation of gas and oxide into the melt. Recently, a
bottom mixing process has been suggested, where an
International

Materials Reviews

1994

Vol. 39

NO.1

evacuated packed bed in the bottom of a crucible is


covered with a melt, and the stirrer shears the interface
between the particles and the melt, resulting in incorporation.P" Combining these two processes, by replacing the ceramic bed with the pressure infiltrated
composite hardener, may provide a useful way of
diluting the hardener."?
The pressure infiltration route has the advantage
that it is a means of producing composites with a
high volume fraction of reinforcement, and since the
system is evacuated, the master composite has very
low porosity. With the additional dilution stage, the
same factors come into playas with the direct mixing
route, i.e. reactivity, viscosity, etc. No costs for this
process are presently available.
Pressu reless i nfi Itration
A recent molten metal process is the Lanxide Corp.
Primex pressureless infiltration process.?' ,42 In this
process a packed bed of ceramic powder is infiltrated
by an AI- Mg alloy, without any applied pressure, in
a nitrogen atmosphere.
The resulting composite,
which has a packed bed density of about 55 vol.- %,
can then be diluted in the appropriate matrix alloy,
if desired. Ceramic particles of SiC and A1203, with
particle sizes as fine as about 1 urn have been infiltrated in this way, and at infiltration rates of up to
the order of centimetres per minute under some
processing
conditions.
Processing
details of the
Primex route are proprietary, but it would appear
to be a very competitive process for higher volume
fraction composites.
Spray deposition
The spray deposition process for unreinforced alloys
was developed by Singer." and put into commercial
use by Osprey Metals.44,45 It involves atomising a
melt and collecting the semisolid droplets on a substrate. As a result, the process is a hybrid rapid
solidification process. The 'Spray Co-deposition Process' is a variant of the basic process, where ceramic
particles are introduced into the spray and codeposited with the alloy droplets/" The deposition rate

Lloyd

Particle reinforced AI and Mg matrix composites

65G-r--------------------,

PVC Exterior
Claddll)Q

64

v
(/)~ 63

::>

Pressure Relief
Venls \

:5
a
:J

,~ ToAlr

62

61

600-+--.......----,----r---.----r---,---r---,--'
o 0.5
2
3
4
TIME AT 675C, h

ReInforced
Wall

6
5

Spray codeposition

system

is 6-10 kg min - 1, and Alcan is presently developing


this route commercially, producing 200 kg ingots."?
A diagrammatical representation of the A1can system
is shown in Fig. 5, where the 'solid deposit' is built
up to a billet for subsequent fabrication. The process
has the advantage that the contact time between the
melt and the reinforcing particles is brief, so reaction
between the two is limited and a wider range of
reinforcements are possible provided the as sprayed
billets are not remelted. The initial billets are typically
95-98% dense, and require a secondary fabrication
step to achieve full density. As long as the alloy has
a sufficient freezing range to achieve atomisation at
moderate superheats, any matrix composition can be
used, including the advanced aerospace alloys, such
as the AI-Li 8090 alloy.r" The cost of composites
produced by spray codeposition
should be intermediate between powder processed composites, and
material made by the mixing method.
XD process
This is a rather different approach to composite
manufacture than the previous molten metal methods.
The XD process is a patented composite manufacturing method developed by Martin Marietta Corp.,
in which ceramic particles are produced in situ in a
melt.'" The process consists of adding to a solvent
metal, such as aluminium, compounds
which will
react exothermally to produce the required reinforcing
particles. A wide range of ceramic compounds can be
formed by this process, 50 but the two which have
received most attention are TiB2 and TiC, which can
be formed by the following reactions
2B

+ Ti + Al ~

TiB2

+ Al

(1)

(2)

and
C

Decrease of melting point with time at 675C


for 7075-15 vol.-J'oSiC (after Ref. 57)

+ Ti + AI~TiC + Al

The particles are typically single crystal, and should


have clean, unoxidised interfaces because they are
formed in situ. By varying the process parameters,
such as reaction temperature, the reinforcement size
can be varied from ~ 02 to 10 urn, though the
material reported in the literature has particles in the
025-15)lm
range." The initial composite is then

used as a hardener with a pure aluminium matrix,


which is then diluted in a melt of the desired matrix
alloy.
An alternative in situ method for producing TiC
particles is to inject a carbon-carrying
gas into an
AI- Ti melt at a sufficient temperature for the exothermic reaction to TiC to occur. 52
To date, there is little detailed information in the
open literature on the processing aspects of in situ
composites, so it is difficult to assess the cost of this
material. It has the attraction of producing particles
which should be inherently wetted by the matrix, and
therefore provide high interfacial strength. However,
the fine particle size is expected to produce highly
viscous melts, which may make handling
and
dilution difficult.
There is, therefore, a range of composite production
routes available in varying degrees of commercial
development, and the appropriate choice will depend
on the application and acceptable cost.

Matrix-reinforcement

reactivity

A recurring theme in the composite literature is the


reactivity between the matrix and the reinforcement
since it can have a significant effect on the interfacial
strength,53,54 and hence the deformation and fracture
of the composite. However, reaction between the
reinforcement
and the matrix can also result in
changes in the matrix alloy metallurgy which will, in
turn, influence such basic properties as the melting
point of the alloy'" and its strength.
The thermodynamic
stability, in aluminium and
magnesium alloys, of most of the reinforcements of
interest has recently been reviewed.!" In silicon free
alloys, SiC is thermodynamically
unstable above the
melting point, reacting to form aluminium carbide,
Al4 C3 (Ref. 56)
4AI

+ 3SiC~

Al4 C3

+ 3Si

. (3)

with an increase in the silicon level of the matrix. As


the reaction proceeds the silicon level increases, and
the melting point of the composite decreases with
time, as shown in Fig. 6 for an AI-Zn- Mg matrix
composite,7075jSiCj15p.57
The aluminium carbide reaction can be avoided by
using high silicon alloys for the matrix, as shown in
International

Materials Reviews

1994

Vol. 39

No.1

Lloyd

Particle reinforced AI and Mg matrix composites


2.0 +----+---+---+--t--+----+--+---+---I---t

14

-t-

o'!-

13

~
~

12

1.5

11

0.08

"3

0.06

o
~

MgO

C\I

-.g

0.10

<(

CJ)

I
-ri

10

co

1.0

C\I

0.04 ~

CJ)

CJ)

.9

tf)

0.02

CJ)

2 0.5

C\I

.9
CJ)

2
L---'--4-----+--+--f----t--+---+--+--+
0.00
920 940 960 980 1000 1020 1040 1060 1080 1100 1120

TEMPERATURE, K

~oo
7

700
800
TEMPERATURE,oC

Silicon
level required
peratures
to
prevent
reaction (after Ref. 58)

at different
temaluminium
carbide

Fig. 7.58 So for processing routes involving long contact times between the reinforcement and the melt,
high silicon aluminium alloys are p~efe~red. In p~~der
processing using solid state consolidation, alum~n.tum
carbide formation is not a factor because silicon
carbide is stable below the solidus; however, if liquid
phase sintering is involved there is th~ potential for
reaction. It is the molten metal processmg routes that
are particularly prone to reaction, since the p~rticleliquid metal contact times can be long, partt.cularly
when large scale casting processes, or remelting are
involved. The spray codeposition method is the least
susceptible, since in this case the contact time is only
in the order of seconds.
Other carbides, such as boron carbide and titanium
carbide are also thermodynamically
unstable in
molten aluminium, but often react in a more complex
manner. 14
Magnesium has no stable carbide, so ceramic carbides are stable in pure magnesium. However, many
of the magnesium alloys of interest contain alloying
elements such as aluminium, which will form carbides, and in these magnesium alloys reaction may
occur if the contact times are sufficiently long.
Aluminium oxide, Al2 3, is stable in pure aluminium, but reacts with magnesium in AI- Mg alloys

3Mg+AI203~3MgO+2AI

Thermodynamic stability of AI-Mg


AI-Mg alloys (after Ref. 59)

oxides in

shows the spinel crystals formed on the surface of


Al2 03 particles after reaction above the melting point
in an AI- Mg alloy.
Obviously,
from equation
(5), Al203 will be
unstable to some extent in magnesium alloys, but
other oxides, such as MgO and Y 203 will be stable.
While these thermodynamic
considerations
show
the tendency for reaction to occur, it is the kine~ics
and the extent of reaction which are of practical
importance. There are three stages of the reaction
which may be attributed to:
.
1. Nucleation of reaction product at preferred SItes
on the reinforcement.
2. Continued dissolution of reinforcement in direct
contact with liquid aluminium.
3. Dissolution of the reinforcement separated from
the liquid aluminium by reaction product and, in
some cases, a solute enriched region.
Gabryel and McLeod 59 have recently con~idered
the reactivity of Al203 in AI- Mg alloys USIng an
approach which may be applicable to. other reinfo~cements. Expressing the extent of reaction, a(t),at ume
t as
Wo -

a(t)=---

w(t)

. (6)

WO-We

where

Wo

is the initial Mg concentration

at t

0, w(t)

(4)

and
3Mg+4AI203~3MgAI204+2AI

. (5)

The magnesium in equilibrium for equations (4) and


(5) has been calculated, and is shown in Fig. 8.59 At
high magnesium levels, and lower temperatures, MgO
may form, while the spinel will form down to very
low magnesium levels. It is not surprising, therefore,
that Al203 is not thermodynamically
stable in most
aluminium alloys. Other oxides, such as MgO, are
expected to be stable. It should also be ,noted t?a~,
unlike SiC which is stable below the sohdus, this IS
not the case for Al2 3, and reaction can continue in
the solid state. So solid state processing may still
result in reinforcement reaction in this case. Figure 9

International

Materials Reviews

1994

Vol. 39

No.1

Spinel crystals on surface of extracted


crystals

AI203

Lloyd

948 K
0973 K
0998 K

Z"0.8

1023K

1048 K

~ 1.0

+I

Q
t<{

1073K
0.6

0::
t-

~ 0.4
U
Z
00.2
U
CJ)

20.0
10

50

250

100
150
200
TIME, min

300

Experimental results for variation in Mg


concentration
in solution after different
reaction times compared with predictions of
kinetic
model
(continuous
lines)
(after

Ref. 59)

the Mg concentration at time t, and We the equilibrium


Mg concentration,
and applying Mampel's model
for deceleratory nucleation caused by overlapping
nucleations.?" they obtain
In(I - a) = C - 4nkNorZt

. (7)

where C is a constant, k a rate constant, No the initial


surface concentration
of nucleation sites, and r the
particle radius. For fused Al, 03 they found that the
weight per cent of magnesium remaining in solid
solution after reacting for a time t (s), at a temperature
T (K), is given by

w = 003

+ (wo

- 0'03)

x exp{50[exp(-I03

OOO/RT)](to

- t)}

(8)

where the equilibrium concentration


of magnesium
Wo, is taken as 003 wt_%. Figure 10 shows that the
model fits the data well within the temperature-time
regimes where MgO does not form. The results demonstrate the importance of minimising the processing
temperature, and in the case of the spinel reaction,
keeping the magnesium content of the alloy as low
as possible, commensurate with achieving the desired
strength.
The kinetics of particle-matrix
reaction can be
influenced by several methods. Undesirable Al4 C3
can be a voided by appropriate choice of the silicon
content of the alloy, which, while restricting the choice
of matrix alloy, is the most convenient approach for
those SiC reinforced composites involved in molten
metal processing. Another approach is to oxidise the
surface of the SiC, forming an outer layer of SiOz. In
this case the early stages of the reaction involve
reducing the SiOz, rather than dissolving the SiC.
The reaction product will depend on the matrix alloy,
being Alz 03 for pure aluminium, and MgO and
AlzMg04
for Mg-containing
alloy.?' This is a not
very satisfactory approach, since there is still an
interface reaction product, and magnesium is lost
from the matrix,
reducing
the age hardening
response." For Al203 reinforcement, the spinel reaction can be minimised by using a low Mg-content
matrix alloy. There is also some recent information

Particle reinforced AI and Mg matrix composites

that using a mixed oxide, mullite, in the crystallised,


fully dense condition greatly reduces reactivity;" but
no detailed kinetic data have been published.
. Interface reaction can have several undesirable
effects. Al4 C3 dissolves in water, degrading the corrosion behaviour of the composite, and its formation
involves release of Si, modifying the matrix composition. Al; Mg04 is not expected to affect the corrosion
behaviour directly, but will modify matrix composition. Matrix composition modifications can in principal be allowed for, but the interface reaction products
may also modify the mechanical properties of the
interface.l'v'" However, for those composites undergoing melt processing either in primary processing as
in the melt mixing process, or in secondary processing
such as shape casting, it is the effect that reaction has
on the viscosity of the melt, and hence on casting
fluidity, that is particularly important.
Casting fluidity is usually assessed by a test, such
as the spiral test, which measures the distance the
melt will flow before solidification occurs. So the test
involves both rheological factors, such as viscosity,
and solidification factors, such as latent heat, surface
tension, superheat, cooling rate, and the freezing range
of the alloy. For an unreinforced metal, the distance
a melt will flow along a channel before solidification,
Lis given by63
.

(9)

where
C = heat capacity
d = diameter of channel
h = heat transfer coefficient
H = latent heat
Ps = density of solid, taken to be equal to density
of liquid
Tm = melt temperature
To = mould temperature
I1T= melt superheat
v = velocity of liquid.
For composites several terms will be modified. The
density, Ps, will be replaced by the density of the
composite, Pc
Pc

Pm(1-

J!;,)

+ Pp J!;,

(10)

where
Pm = density of matrix
Pp = density of particle
J!;, = volume fraction of particle.
The latent heat involved in solidification will be
reduced because the particles are not involved in the
solidification process; the effective latent heat, He' is
given by

He = H(I-

Vp).

(11)

Similarly, the effective specific heat of the composite,


c.. is given by
Ce = Cm(1 - ~)

+ Cp Wp

(12)

where
Cm = specific heat of matrix
Cp = specific heat of particle
~ = weight fraction of particle.
International

Materials Reviews

1994

Vol. 39

No.1

Lloyd

Particle reinforced AI and Mg matrix composites

70 ~------------..,

70
A

60-

<9 40Z

t-

Eu 50-

..

:r:
t-

W
--J

a::

30-

(a)
50

100

150

200

250

a
300

a 750C; b 800C
Spiral fluidities

of A356-15 vol.-%Sie after different

Substituting the appropriate values for the composite


into equation (9) will result in a decrease in the
expected fluidity length with increasing
particle
content.
However, of more importance is the melt velocity
term, v, under the applied metallostatic head. As seen
from Fig. 4, the viscosity increases with increasing
particle level, and this will reduce the melt velocity.
For
non-metallic
liquids
containing
spherical
particles, the viscosity is given by64
nm(1

+ 2'5Vp + 10'05V~)

where
n; = viscosity of composite
nm = viscosity of unreinforced

. (13)

matrix.

The non-Newtonian
behaviour of molten composites
makes comparison with equation (13) difficult, but it
is expected to be of the right order at very high shear
rates, when Newtonian behaviour is approached. At
low shear rates, or when reaction occurs between the
reinforcement and the melt, it greatly underestimates
the viscosity. If extensive reaction occurs the viscosity
can, in effect, increase to infinity, and the melt will
not flow into the mould at all.As a result, rheological
factors dominate the casting fluidity under these
circumstances. This effect is demonstrated in Fig. 11,
where the spiral fluidities after different holding times
at 750 and 800C are shown. For 750C the spiral
fluidity remains about constant with holding time,
because the extent of aluminium carbide formation is
limited at this temperature
in a 7 wt-%Si alloy.
However at 800C, the amount of aluminium carbide
increases rapidly with time resulting in a marked
decrease in fluidity until, after 250 min, the composite
will not flow into the mould.
As noted previously, the rheological behaviour of
composite melts is poorly understood,
but a few
general statements can be made:
1. The viscosity is non-Newtonian, decreasing with
increasing shear rate.
International

Materials Reviews

1994

Vol. 39

No.1

(b)
I

50

HOLDING TIME, min

10-

i
20-

(/)

10-

n;

a:

20-

(/)

11

--J

30-

<9 40Z

--J

a:
a::

E sau

--J

60-

100

1
,

...

200
250
HOLDING TIME, min
150

300

holding times at different temperatures

2. The viscosity increases with increasing volume


fraction of particle.
3. The viscosity increases with decreasing particle
size.
4. The viscosity is dependent on the history of the
melt, in terms of temperature, time, and shear rate.
5. The viscosity increases with increasing reaction
product at the interface.
The most extensive rheological study of unreacted
composites has been carried out by Moon.i" who
found that they are also thixotropic.
In terms of equation (13), interface reaction will
often result in an increase in the volume fraction of
solid in the melt. For example, from the stoichiometry
of the SiC reaction to Al4 C3, and the lower density
of Al4 C3 to SiC (2'36 cf. 32 g ern - 3), 1 g of SiC
reacted to Al4 C3 will result in a particle volume
increase of 60%. In addition, as pointed out by
Surappa and Rohatgi,66 the viscosity may be dependen t on particle surface area, and this will increase
with interface reaction because the reaction products
tend to be in the form of fine crystals.!" The nonNewtonian and thixotropic nature of composite melts
indicate that the composite melts have a structure
associated with them, presumably reflecting particle
clustering in the melt. Particle clusters could occlude
liquid within the clusters, effectively raising the solid
fraction of the melt. 67
It is apparent that reinforcement reactivity can
influence the behaviour of the composite from a wide
range of viewpoints, and needs to be considered both
in terms of composite processing and composite use.

Microstructures
The most important aspect of the microstructure
is
the distribution of the reinforcing particles, and this
depends on the processing and fabrication routes
involved. However, particles can modify other aspects
of the matrix microstructure.

Lloyd

Reinforcement distribution
In powder processed material, the reinforcement distribution will depend on the blending and consolidation procedures, as well as the relative size of the
matrix and reinforcing particles. If the matrix powder
is large relative to the reinforcement, the reinforcing
particles will agglomerate in the intersticies of the
coarse particles, and be very inhomogeneously distributed in the final product. With the blending and
consolidation route used in one study.t" an AI/SiC
particle size ratio of 07: 1 gave a more uniform
distribution than a 024: 1 ratio. Any secondary processing will also tend to homogenise the particle
distribution.?" (It should be noted that quantitative
determination
of particle inhomogeneity
and clustering, considering the variable shape and size of
reinforcing particles, is not a trivial task. Initial
approaches
have used the Dirichlet
tesselation
method. 69. 70)
In composites processed by molten metal mixing
methods, the situation is somewhat more complicated
because the reinforcement distribution is influenced
'by several factors:
1. Distribution
in the liquid as a result of the
mixing.
2. Distribution in the liquid after mixing but before
solidification.
3. Redistribution as a result of solidification.
The distribution
during mixing will obviously
depend on the mixing process used, and it is essential
to produce as uniform a distribution
as possible
without any gas entrapment, since any gas bubbles
will be lined with reinforcing particles. After mixing
and before solidification, the particles will segregate
due to gravity.P With the relatively high volume
fraction of particles and a range of particle size, the
settling will be hindered 71
Jt::

Vo(l- C)P

where
C = particle concentration
d = particle diameter
D = container diameter
p = 465 + 195d/D for R, < 02
=(435+ 175d/D)R;0,03 for 02<Re<
R; = Reynolds number
Jt:: = particle velocity
Vo = Stokes velocity.

. (14)

The settling rate will be a function of the particle


density and size, and there is also the possibility that
particle shape will playa role."? Particles of different
size and shape will settle at different rates producing
agglomeration.
The third factor which influences reinforcement
distribution
is the solidification
process
itself.
Reinforcing particles do not generally nucleate the
primary solidifying phase, 56 though solidification
nucleation may occur in some hypereutectic
systems.P If solidification nucleation does not occur the
reinforcing particles are rejected at the solid/liquid
interface, and segregate to the interdendritic regions
which solidify last. Particle entrapment or rejection
has been extensively studied, and there are several
recent reviews. 22. 74, 75 The entrapment
models are

Particle reinforced AI and Mg matrix composites

a slow solidification rate investment


rate pressure die casting
12

casting;

Influence of solidification
particle distribution

b high solidification

conditions

on

generally applicable to planar front solidification of


single phase systems, while in reality it is dendritic
solidification of multiphase systems which is relevant.
In these commercial systems the experiments indicate
that, whenever particles are captured by a solid/liquid
interface, the particles act as heterogeneous nuclei for
the solids." Because this does not generally occur,
the particle distribution is influenced by the solidification rate, as shown in Fig. 12. Secondary fabrication
processing, such as extrusion or rolling, can homogenise the structure to some extent, as seen from the
previous extrusion microstructures,
but minimising
reinforcement inhomogeneity during initial processing
is important
for achieving optimum
properties.
Figure 13 shows how the tensile elongation increases
in (AI-Si) A356/SiC/15p with increasing degrees of
extrusion.
At present, quantification of particle inhomogeneity
remains a problem, but recently some authors have
begun to develop methods for addressing it.13, 76, 77
Grain structure
Wrought alloy composites are solution treated and
aged after fabrication, and recrystallisation will usually occur during this heat treatment. Since particles
of diameter larger than about 1 urn will develop an
associated deformation
zone sufficient to generate
International

Materials Reviews

1994

Vol. 39

No.1

10

Lloyd

Particle reinforced AI and Mg matrix composites

15
14

13

12
~11
z'10

0 9

(9 7
Z
6
0
.--J

b.

3
2

T4-5.7 em DIA. BILLET


T4-19.5 em DIA. BILLET
T4-PM CAST TENSILE

b.

1
0
10

EXTRUSION

13

100

1000

RATIO

Variation in tensile elongation with extrusion


ratio in A356-15 vol.-%Sie, T4 condition

recrystallised nuclei;" the reinforcing particles should


produce a high density of nuclei. However, if the
particles are very closely spaced, the subgrain growth
necessary for recrystallisation
nucleation is impeded
and recrystallisation may not occur. Recrystallisation
is impeded when t;,/ d > 01 urn, whereas most commercial composites have t;,/d less than this value,
where t;, is the volume fraction of particles and d the
particle diameter.
The grain size of the recrystallised composite can
be estimated by assuming that each reinforcing
particle of diameter d acts as a nucleus for a spherical
grain D, which is given by 79
D

d[(1 -

t;,)/t;,J 1/3

(15)

Any subsequent grain growth will be limited by the


Zener pinning of the particles on the grain boundaries,
giving a limiting grain size of 2d/3 t;,. For 20 vol-"
of 10 urn particles equation (15) gives D = '" 15 urn,
and the limited grain growth gives D = 33 urn,
Recrystallisation
studies
on Duralcan
(AI-Cu)
2014/AI203/20p
demonstrated
that
the Al203
particles stimulate the nucleation, accelerating recrystallisation, and decreasing the recrystallisation
temperature.P'' The resulting equiaxed grain size was
",15 urn, and after holding the composite for 150 h
at 500C the grains had only grown to 17 urn. These
results show that equation (15) is appropriate
for
predicting both the recrystallised and limiting grain
size, and the high volume fraction of reinforcing
particles is very effective in stabilising the grain size.
In powder processed composites reasonable agreement was obtained with the theory except for very
coarse particles (D = 40 urn), where the grain size was
much finer than expected because of multiple
nucleation occurring at each particle. 79
Since the reinforcing particles are very effective in
nucleating new grains, and stabilising the resulting
fine grain size, MMCs are potentially superplastic
materials. Tensile elongations of 300-8300/0 have been
obtained in powder processed PM-64/SiC/20p,81
6061/Si3N4/20p,82
and (AI-Cu) 2124/Si3N4/20p,83
spray processed 2014/SiCj15p and 7475/SiC/15p,84
and mechanically alloyed IN9021/SiC/15p.85
A high
International

Materials Reviews

1994

Vol. 39

strain rate sensitivity m in the strain rate equation


= kern, is required for superplasticity,
and m-values
in the range 0'4-0'6, which are comparable with those
of superplastic unreinforced alloys, were obtained. An
interesting feature of superplasticity in MMCs is that
the high m-values are often obtained at high strain
rates. In conventional superplastic alloys the highest
m-value, and optimum superplastic behaviour, is usually obtained at low strain rate, about I x 10-4 to
5 X 10-3 S-1. However, in MMCs the largest mvalues, and greatest elongations
are obtained at
greater than 10- 2 S -1 (Ref. 82) and even as high as
10 and 100 s - 1 in mechanically alloyed composites."
In all cases the elongation is limited by cavitation,
and particles increase the extent of cavitation by
limiting grain boundary sliding and providing stress
concentrations
at the interface. (There is also the
possibility of pre-existing cavities at the interface.P")
The extent of cavitation increases with increasing
volume fraction of reinforcement,
but it can be
reduced by using a superimposed
hydrostatic presure."! Using as fine a reinforcement
as possible
should also help to decrease cavitation, and this is in
agreement with the recent work on 2124/Si3N4/20p
where the particles were less than I urn in size and
elongations of 850% were obtained at a strain rate
of 4 x 10 - 2 S- 1 and 515C, in as extruded material.
The yield strength is dependent on grain size
through the Hall- Petch relationship, and the finer
grain size present in composites could contribute
considerably to the strength. Aluminium alloys have
a low Hall- Petch slope, and grain sizes of 10 urn and
less will be required to influence the strength significantly, but in reinforced wrought Mg alloys grain size
strengthening could be very significant, because the
Hall- Petch slope is large.I"
(J

No.1

Aging effects
Another aspect of the microstructure
which may be
modified in composites is that of precipitation. Many
studies of the age hardening response of metal matrix
composites have been reported87-93 and, in general,
an enhancement
of the aging kinetics has been
observed. However, some of the studies demonstrate
that at low temperatures the aging kinetics are slowed
down, or unaffected by the reinforcement,89,91,94 The
more rapid aging kinetics were initially attributed to
easy precipitation nucleation on dislocations punched
out from the reinforcement particle interface as a
result of the coefficient of thermal expansion mismatch
on quenching from the solution treatment temperature. Enhanced precipitation in the neighbourhood
of reinforcement
particles undoubtedly
occurs for
precipitates, such as 0' and S', which are susceptible
to dislocation nucleation." but there are many other
factors which may influence aging. In material processed by powder metallurgy, there are also fine oxide
particles in the matrix which are known from work
on SAP-type alloys to influence the aging kinetics."
The presence of fine oxide particles, together with a
fine grain size is likely to reduce the vacancy concentration by providing a high vacancy sink density, as
do the matrix/reinforcement
interfaces. Vacancies may
also be swept up by moving dislocations associated

Lloyd

Particle reinforced AI and Mg matrix composites

11

130,....,----------------,

120
(\J

IE
zllO

o
~ 100
-.J

::J

o
o

2
u
~

s = 1.5h

/./:,./

90

/./
/.

/..

= 1.0

80

V>

Ljj

70
- -----

60

14

Precipitation at SiC interface in (AI-Zn-Mg)


7091 alloy (courtesy of J. J. Lewandowski)

Rule of Mixtures
Equivalent Inclusion
Tsai-Halpin Equation

50
+------,r-----r----.---~---1
o
5
10
15
20

25

VOLUME, 0/0

with the plastic relaxation of misfit stresses generated


. the vacancy concenon quenc himg. 91 Ch anges In
tration will particularly affect G P zone formation and
the low temperature aging response, consistent with
many of the differential scanning calorimetry studies.90,91,93,94
Solute segregation
associated
with
interfaces may also be present .in the composite,"?
and the solute distribution may also be affected by
any residual stresses present in the matrix. The higher
dislocation density in composites could also increase
the solute diffusivity, and enhanced diffusivity of Mg
atoms has been suggested as the reason for the higher
growth rate of {3' precipitates in 6061-A12 03 composites.?" At higher aging temperatures, or longer times,
precipitation may occur at the particle interface in
some alloys. An example of this in a 7091 alloy matrix
is shown in Fig. 14.
There is very little information in the literature on
the aging kinetics of Mg matrix composites, but work
on SiC reinforced Mg-6Zn
indicated comparable
aging behaviour to that of unreinforced material. 99
From all these considerations
it is clear that the
aging response will depend on a range of factors
including t~e particular matrix, the processing history:
and the agmg temperature. However, it should also
be appreciated that in spite of any modification of
agi.ng kinetics in the composites, the peak aged propertres are usually obtained within the normal commercial aging practice for unreinforced alloys, particularly
for melt processed composites.

Mechanical properties
Elastic modulus
The on~ mechanical property which is always significantly Increased by the addition of reinforcement is
the elastic modulus. The quantitative
value of the
elastic modulus is somewhat
dependent
on the
method of measurement,
with dynamic measuring
methods tending to give larger values than static
measurements
obtained from the elastic portion of
the tensile stress-strain curve. Static values may also
depend on whether the measurements
are made in
tension or compression.l''?
Most of these difficulties

15

Variation of elastic modulus with volume


fraction of SiC particles: 5 is particle aspect
ratio

result from the presence of thermal residual stresses


caused by differences in the coefficient of thermal
expansion
between the matrix and the ceramic
particles. In the case of SiC and Al203 particle
reinforced AI, the matrix is in tension. This means
that when the composite is loaded, plastic flow occurs
earlier in tension than in compression, and the total
strain will consist of both elastic and plastic components. The situation is further complicated
by
inhomogeneity
in the reinforcement
distribution
which can also result in local plasticity. Comparison
with theoretical expectations is also somewhat difficult due to uncertainty in the appropriate value for
the modulus of the particle reinforcement.
Figure 15 shows the increase in Young's modulus
with volume fraction of reinforcement for a variety of
AI-SiC composites (these values were taken from the
commercial literature for wrought 2000, 6000, and
7000 alloys). The rule of mixtures expression

e, = VpEp + VmEm
~here s; Em' s, are ~he elastic

(16)

moduli of the composite, matrix, and p~rtIcle, respectively? and Vm and ~


the ~olume fractions of the matrix and particle,
considerably overestimates the elastic modulus.
The rule of mixtures expression is most appropriate
for continuous reinforcement and it has been modified
for discontinuous
reinforcement in the Halpin- Tsai
equation 101

=
e

Em(J + 2sqJt;,)
l-qJt;,

. (17)

where

(Ep/Em - 1)
q = (Ep/Em) + 2s

. (18)

and s is the particle aspect ratio. As seen from


the figure, the Halpin- Tsai equation gives a good
representation of the results.
International

Materials Reviews

1994

Vol. 39

No.1

12

Lloyd

Particle reinforced AI and Mg matrix composites


400.------------------------,

100
C\I

IE

Z 90-

(/)'

:3

80-

A356
A356/SiC120p

2: 70(/)

360-

Z.250
(/)
(/)

100

l-

(/) 150

200

300

400

50
O-+---+----+--+---l'-+---+----+--+---l'-+---+----+--+--lf---l

0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14


TRUE STRAIN

500

600

TEMPERATURE,oC
16

Temperature dependence of elastic modulus


in A356 and A356/SiC/20p

The elastic modulus can also be calculated using


the Eshelby equivalent inclusion method.l'" and this
approach is also in good agreement with the data.
The dominant
factor in controlling
the elastic
modulus is the volume fraction of reinforcement, and
it is relatively insensitive to the particle distribution,
while variations in the type and shape of the reinforcement can be accounted for by the different expressions.
The improvement in Young's modulus is retained
at higher temperatures, as shown in Fig. 16, where
results for unreinforced matrix and composite are
shown up to 500C. The temperature dependence of
the composite elastic modulus reflects that of the
unreinforced matrix. The higher elastic modulus at
temperature
is important
for higher temperature
applications.
Strength
The first extensive study of the strength of discontinuously reinforced Al alloys was carried out by
Mcfranels.l'" who investigated SiC whisker and
particle reinforcement
in several different alloy
matrices. He reported up to a 60% increase in yield
and ultimate tensile strengths, depending on the
volume fraction of reinforcement, the type of alloy,
and the matrix alloy temper. Subsequent work has
generally confirmed these findings, but the reported
experimental results show an extremely large degree
of scatter, presumably reflecting the quality of the
material and differences in processing. When considering the yield stress there is a more fundamental
difficulty, as pointed out by Humphreys.l'" The yield
strength is usually quoted as the 0.2% proof stress,
and since composites work harden extremely rapidly
at low strains, this may not be equivalent to a
conventional yield stress. However, composites are
now commercially
available which will meet a
strength specification, and Table 2 lists the properties
of some of the composites now available from different
producers. For comparison some typical properties
of unreinforced alloys are given in Table 3.
The strengthening mechanisms which may operate
in particle reinforced MMCs have been considered
in several recent publications,
and the behaviour
has also been extensively
modelled
mathematically.77,104-108 The continuum shear lag models for
International Materials Reviews

1994

100 :

I-

-t----.------.-----.----.,.----.II'I---j
I
I

W200
0::

0
>50

350

IE 300

::::>

C\I

Vol. 39

No.1

17

Stress-strain curves for extruded A356 and


A356/SiC/15p with different
particle sizes
and tempers

reinforcement with an aspect ratio below that necessary for full reinforcement loading, gives the composite
strength as109
. (19)
where (J m is the yield stress of the matrix, and s the
aspect ratio.
For the aspect ratio typically used in particle
MMCs, which are in the 1-5: 1 range, equation (19)
underestimates
the strength,
but Nardone
and
Prewo '?" have suggested that better agreement is
obtained if the equation is modified to allow for end
loading effects. The difficulty with this continuum
approach is that it ignores the influence of particles
on the micromechanics
of deformation, such as the
very high work hardening at low strains, and modifications in microstructure,
such as grain size and
dislocation density.
In the micromechanics
approach,
the possible
strengthening mechanisms are:
1. Orowan strengthening.
2. Grain and substructure strengthening.
3. Quench hardening resulting from the dislocations generated to accommodate the difference in
coefficient of thermal expansion between the reinforcing particles and the matrix.
4. Work hardening, due to the strain misfit between
the elastic reinforcing particles and the plastic matrix.
The extent to which the different mechanisms operate will depend on the microstructure and processing
of the particular composite. In powder processed
composites, the grain size can be extremely small and
significantly contribute to the strength, whereas in
melt processed MMCs the grain size more closely
approaches unreinforced alloys. At typical volume
fractions, Orowan strengthening is not a major factor
with the 5 urn and larger particles usually used, but
particles of this size can result in quench hardening
and enhanced work hardening because of elastic misfit
back stress hardening.111-113 From these considerations, the strength of particle reinforced composites
is most strongly dependent on the volume fraction of
reinforcement with a somewhat weaker dependence
on particle size. lOS Normal matrix strengthening due
to solution and precipitation
hardening will give
additional strength to the composite.

Lloyd
Table 2

Typical properties of some commercially

Particle reinforced AI and Mg matrix composites

13

available metal matrix composites


Elastic

Composite*

YS,t
MN m-2

UTS,
MN m-2

Wrought
AI-M92Si
6061 /AI203/1 Op (T6)
6061/AI203/15p
(T6)
6061/AI203/20p
(T6)
6061/AI203/20p
(T6)
6061 /SiC/15p
(T6)
6061 /SiC/15p
(T4)
6061/SiC/20p
(T4)

296
317
359
305
342
405
420

338
359
379

Duralcan, Alcan
Duralcan, Alcan
Duralcan, Alcan
Comral 85, Comalco
Cospray, Alcan

70
50

98
105

DWAt
DWAt

40

115

DWAt

84
92

Duralcan, Alcan
Duralcan. Alcan

101
100
98
100

Duralcan, Alcan
Cospray, Alcan
Cospray, Alcan

630

33
23
10
20
30
5-7
2-4

116

BP~
DWAt

517
503
503
493
532
610

483
466
460
400
490

Supplier

81
87
98
85
91

460
500
515

483
476

modulus,
GN m-2

75
54
21
34
32

330
364

430

6061/SiC/25p
(T4)
AI-Cu
2014/AI203/10p
(T6)
2014/AI203/15p
(T6)
2014/ AI2 03 /20p (T6)
2014/SiC/15p
(T6)
2618/SiC/12p
(T6)
2124/SiC/178p
(T4)
2124/SiC/25p
(T4)

Elongation,

BP~

2124/SiC/20p
AI-Zn-Mg

(T4)

405

560

105

7075/SiC/15p
7049/SiC/15p

(T651)
(T6)

601
643

(T6)

735

3
2
2

95
90
105

Cospray, Alcan
Cospray, Alcan

7090/SiC/20p
AI-Li

556
598
665

8090/SiC/13p
8090/SiC/13p
8090/SiC/17p
8090/SiC/17p

(T4)
(T6)
(T4)
(T6)

455
499
310
450

520
547
460
540

4
3
4-7
3-4

101
101

Cospray, Alcan
Cospray, Alcan

103
103

BP~
BP~

20

105

XD, Martin

06

82
91
98
95
114

Duralcan,
Duralcan,
Duralcan,
Duralcan,
Duralcan,

DWAt

Cast
AI-Cu
201/TiC/20p
AI-Si

(T7)

420

356/SiC/10p
356/SiC/1 5p
356/SiC/20p
380/SiC/10p
380/SiC/20p
Mg-AI-Zn

(T61)
(T61 )
(T61)
(F)
(F)

287
329

AZ91/SiC/94p
AZ91 /SiC/151
AZ61/SiC/20p

336
245
308

308
336
357
332
356

191
208
260

236
236
328

* Composite designation: matrix/reinforcement/volume


t 02% offset yield stress.
t Composite Specialities lnc., Chatsworth, CA.

~I British

03
04
10
04

fraction

Alcan
Alcan
Alcan
Alcan
Alcan

Dow
Dow
Dow

of particles.

Petroleum.

Figure 17 shows the stress-strain curves for A356


and A356/SiC/15p, for two SiC particles sizes and
two tempers, all materials being extruded bar. The
apparent work hardening rate at low strains is higher
in the composites, and also increases with decreasing

Table 3

Alloy

Typical unreinforced
YS,*
MN m-2

UTS,
MN m-2
310
524
470
470

6061
2014
2124
2618

(T6)
(T6)
(T6)
(T6)

275
476

7075
8090
A356
A380

(T6)
(T6)
(T6)
(F)

505
415
205

AZ61
AZ91

475
54
80

2
1
25

Marietta

325
370

160
157
168

* 02% offset yield strength.

570
485
280
320
198
311

alloy properties
Elongation,
%
20
13
12
9
10
7
6
35
30
21

Elastic
modulus,
GN m-2
69
73
72
74
72
80
76
72
38
49

SiC particle size. At a 0'2% offset strain the composite


work hardening rate in the T4 condition is around
50 GN m -2, which is greater than the shear modulus
of the matrix. Therefore, the higher initial work
hardening rate may, to some extent, reflect differences
in the elastic-plastic
transition between composites
and unreinforced material. This is not too surprising,
considering the presence of a relatively inhomogeneous distribution of elastic particles. After strains
of about 3 % the stress-strain curves of the composite
and unreinforced alloy are essentially parallel, so all
the strengthening
is associated with the first few
percent strain. It is also apparent from Fig. 17 that
the work hardening rate increases with a decrease in
particle size from 16 urn (400 grit) to 78 urn (800 grit).
This is consistent with current models for. coefficient
of thermal expansion and back stress strengthening.112,114 Particle shape, in terms of aspect ratio, will
influence composite strength, but for the typical aspect
ratio range of up to 2: 1 it is not expected to be a
major factor.l!"
International

Materials Reviews

1994

Vol. 39

No.1

14

Lloyd

Particle reinforced AI and Mg matrix composites

601lJ-r---------------------,

20,u,--------------------,

C\I

18

IE 50

z
2

40

IE

16

14

,12

(/)
(/)

~ 30

~1O
0::
~ 8

(/)

..J

A356
A3561l5SiC

(/)

---

C\I

20

........- 606l1SiClfT6

(/)

--+- 6061/SiCfT4
----&- 2124/SiCfT4

W
>=10

a
..J

>=

10

15

20

25.

30

35

VOLUME FRACTION, 010

18

Variation in yield stress with volume fraction


of reinforcement
for powder
processed
composites

The predominant strengthening factor is the volume


fraction of reinforcement, and this is shown for
powder processed composites in Fig. 18. Powder
processed material tends to give somewhat higher
strengths than melt processed composites, probably
because of additional strengthening from oxide dispersoids, and the somewhat finer grain size. Note that
the strength increase appears to be reduced at the
very high volume fractions of reinforcement. This
probably reflects mixing difficulties in obtaining a
uniform particle distribution at high volume fractions.
The extent of strengthening is also dependent on
the matrix alloy microstructure, being lowest in peak
aged material. Age hardened alloys tend to be more
susceptible to strain localisation, and the work hardening rate decreases very rapidly with strain. This
effect will also occur in composites, and it means that
the strengthening observed is alloy and heat treatment
dependent. While dislocation density and elastic misfit
strengthening models are in agreement with the general trends of the strengthening results, they are not
sufficiently detailed to account for particle size and
distribution variations, and how the different microstructural factors interact.
The ability to achieve any strengthening in a composite .is depend ent on the ability to transfer stress
from the matrix to the stronger reinforcing particles.
This, in turn, is dependent on achieving a strong
interfacial bond between the matrix and the reinforcement. If the interfacial bond is weak the interface will
fail before any effective stress transfer to the particle
can occur, and no strengthening is achieved. The
composite may, in fact, appear to be weaker than the
unreinforced matrix because of the reduced effective
area supporting the load. As noted previously in the
discussion of elastic modulus, the presence of thermal
residual stresses, and inhomogeneity of reinforcement
distribution, result in the early onset of plasticity and
no strengthening at very low strains, i.e, close to the
proportional limit of the stress-strain curve.
A final aspect of strength, which is of commercial
importance, is the behaviour at elevated temperatures.
While the presence of particles improves the modulus at higher temperatures (Fig. 16) they do not add
significantly to the high temperature strength. The
reason for this is that the strengthening mechanisms
operating in composites at lower temperatures are
relaxed at higher temperatures, and the composite
International

Materials Reviews

1994

Vol. 39

No.1

50

100

150

200

250

300

350

TEST TEMPERATURE. DC

19

Variation in yield strength with temperature


after 100 h at temperature
for A356 and
A356/SiC/15p

strength is primarily controlled by the higher temperature strength of the matrix. However, a small
improvement in strength over the unreinforced alloy
is retained at higher temperatures, as shown in Fig. 19.
This figure compares the strength at temperature of
unreinforced A356 with A356/SiC/15p after a 100 h
soak at temperature. The composite has about a
10 MN m -2 higher strength than the matrix at 300C,
and the decrease in composite strength with temperature follows that of the matrix. By using powder
processing of rapidly solidified powders, improved
high temperature strengths can be achieved..i i s
Another alternative is to use a very high volume
fraction of reinforcement in a normally solidified
matrix. By going to a high volume fraction of particles
the Orowan strengthening component becomes significant, and this is essentially independent of temperature other than the temperature dependence of the
modulus. The Lanxide Corp. Primex pressureless
metal infiltration technology allows high volume fractions to be obtained in near net shapes, and Fig. 20
shows the temperature dependence of Lanxide NX,
which is essentially 3360 alloy (AI-Si) containing
60 vol._%SiC.1l6
The high temperature strength
advantage of the composite is very significant, but
there is, of course, a sacrifice in ductility.
Elongation

The major limitation in the mechanical properties of


composites is the rather limited ductility, as seen from
Table 2. The tensile elongation decreases rapidly with
the addition of reinforcing particles, Fig. 21, and it
also decreases with increased aging time in the heat
treatable alloys. Recent work has demonstrated that
composite failure is associated with particle cracking
and void formation in the matrix within clusters of
particles.68,llS-12o Particle fracture is more prevalent
in coarser particles, and this would be expected from
two points of view. The larger the particle the more
it will be loaded by conventional fibre loading and
end loading mechanisms. In addition, the coarser
ceramic particles will have a higher probability of
containing fracture initiating defects. Failure associated with particle clusters can be attributed to the
higher stress triaxiality generated in these regions.
Drucker'?' noted that matrix deformation between

Lloyd

Particle reinforced AI and Mg matrix composites

15

20

25u,-------------------,

~ 20

-.-

336/60SiC

18

-.-

ALLOY336

16

cf!.

E
z

215

o Iti um
D

7.5Jlrn

14

[]

612

(!)

~1O

0
...J
6
w

0:::
l(/)5

~1O

I-

0
D

4
2
o

100

200

300

400

500

0+----+--+---+--+--t---f--t--+---+--4--1

600

-200 -150 -100 -50

TEST TEMPERATURE,oC

20

Temperature dependence of strength


with and without 60 vol.-%SiC

of 336

closely spaced elastic particles would be highly constrained, resulting in local stress levels many times
the matrix flow stress. This behaviour has been confirmed by recent continuum modelling.106,122Since
ductile failure mechanisms by void nucleation and
coalescence are very sensitive to the triaxial stress
level,123fracture nucleation within particle clusters is
expected.
Many models have been developed for void
nucleation. Particle cracking by catastrophic propagation of an interal defect is given by the Griffith
equation
(Jf

(~cy)1/2
,'-

(20)

where (J'fis the stress on the particle, y the fracture


surface energy, E the Young's modulus of the particle,
and C the internal crack length.
For coarser particles there is a higher probability
of their containing a defect of length C sufficient to
give particle fracture. It is worth noting that the
particles are being loaded through the particle/
matrix interface, so a high degree of particle fracture
is indicative of a high interfacial strength, particularly
when cracking is occurring in finer particles,
<10 urn,
A criterion for particle-matrix decohesion appropriate for coarse particles has been developed by
Argon et al.
The critical stress for interfacial void

124

2
18

-+-

6061/Sicrr6

-+- 60611A~03

/rs

~1

12

~1
~

0
...J
W
2
0

10

15

VOLUME

21

20

25

50

100

150 200 250 300

TEST TEMPERATURE,

30

35

40

FRACTION, 010

Variation in tensile elongation of 6061 with


volume fraction of reinforcement

22

350

Influence of testing temperature on tensile


elongation for extruded A356/SiC/15p with
400 and 800 grit sizes

nucleation, (J'e,is given by


. . . . . . . (21)
where a.; is the equivalent stress and (J'rnthe mean
stress ((J'rn= (J'ii/3).Void nucleation will therefore be a
function of the matrix flow stress, which will be
influenced by the volume fraction of particles, heat
treatment, etc. In the Argon et al. model the critical
stress is independent of particle size, but particle size
effects would come into the picture if the particle
distribution is a function of particle size, and also
through the particle size dependence for particle
cracking, equation (20).
In unreinforced material, voids grow after being
nucleated until they are of a sufficient size to coalesce,
resulting in final fracture. However, with the high
volume fraction of particles in composite, extensive
void growth does not occur. The limit load failure
model of Thomason 123 predicts the condition for
spontaneous ductile fracture at the void nucleation
strain as
03[(n/4 Jtf)1/2 1 - Jtf

IJ + 06

(J'rn

2 + 2k

. (22)

where Jtf is the volume fraction of nucleated voids,


(J'rnthe mean stress, and k the matrix shear flow
stress.
If the mean stress, (J'rn= 2k, this expression predicts
that a void volume fraction greater than 0055 would
result in spontaneous fracture without extensive void
growth. In composites with typical particle volume
fractions greater than 0'1, the fracture process would
be nucleation controlled, consistent with the limited
void growth seen experimentally.
Since particle fracture is readily observed in most
failed composite, it is often suggested that this is the
dominant failure process. However, particle fracture
occurs from very early strains, as demonstrated by a
detectable decrease in modulus with increasing
strain.119,120,125It is also apparent, in melt processed
composites, that very different particle fracture behaviour is not reflected in differences in the tensile
elongation. Figure 22 shows the variation in tensile
elongation with test temperature for A356/SiC/15p
with two grit sizes. There is very little difference in
the elongation between the two grit sizes, but the
International

Materials Reviews

1994

Vol. 39

No.1

16

Lloyd

Particle reinforced AI and Mg matrix composites

4. Control of particle shape .


5. A ductile matrix.
Composite fabrication and processing will influence.
the degree of flexibility available to meet these
requirements.

0.20,---------------------.

..J

U 0.18

F=

lti urn
7.5Jlrn

~ 0.16

n;

0.14

W 0.12

~ 0.10
50.08
~ 0.06
Z 0.04

o
o

~ 0.02

DOD

....
~ 0.00 +-0----G~
_ _+___+___+-l____+__+___l:jl--_+__+__+-+___+__j
60 80 100 120 140 160 180 200 220 240 260 280 300 320 340
UTS, MN m-2

23

Fraction of cracked particles in A356/SiC/15p


composites
fractured
at different
temperatures, and with grit sizes of 400 and 800

fraction of cracked particles is much larger in the


coarser grit size, as shown in Fig. 23. The minimum
in tensile elongation at 200C is due to precipitation
hardening occurring during the test. It is interesting
to note that the minimum in tensile elongation is not
reflected in a maximum in the extent of particle
cracking, which demonstrates the importance of the
matrix in composite fracture. At temperatures above
250C general cavitation becomes the predominant
failure mechanism, and very little particle fracture
occurs.
Obviously coarse particles, whether reinforcing
particles or intermetallics, should be avoided to minimise particle fracture. A high interfacial strength is
needed to minimise cracking along the interface and
to load the particles effectively. However, the microstructure of the matrix is also important, because
local matrix failure appears ultimately to control the
fracture. In addition, the particle distribution will also
influence the composite ductility. The fracture models,
discussed previously, assumed a uniform particle distribution, and the mean stress in equations (21) and
(22) is a far field stress. With the inhomogeneous
distribution of particles in real composites, intrinsic
local stress triaxiality is generated in the clustered
regions which may dominate any far field stresses,
and make it difficult to compare the models with
experiment. Recent work on notched tensile bars
shows that the extent of far field triaxiality has
only a small influence on the tensile elongation in
composites, compared with the major influence in
unreinforced alloys."? The importance of particle distribution is also demonstrated
by the increase in
tensile elongation with extrusion ratio, as shown
in Fig. 13.
Considering all these different factors it is clear that
the fracture process in particle composites is quite
complex, and a quantitative understanding is lacking.
It is also apparent that different composites may be
dominated
by different fracture processes, but to
maximise ductility for, a particular volume fraction,
the composite should have:
1. Uniform particle distribution.
2. A fine 10 urn), uniform
particle
size
distribution.
3. A high interfacial strength.
International Materials Reviews

1994

Vol. 39

No.1

Fracture toughness
The fracture toughness of particle reinforced composites has been reviewed
in two recent publications.126.127 While the toughness mirrors to some
extent the tensile elongation, decreasing with increasing reinforcement, the decrease is most significant
from zero to 100/0 reinforcement, with only a slight
decrease for higher reinforcement loadings. The fracture toughness is also much less sensitive to the degree
of aging than is the tensile elongation, and there
are some data suggesting that coarser particles are
beneficial. With typical' toughness values in the
15-20 MN m -3/2 range, it is apparent that particle
composites do provide a reasonable degree of fracture
toughness.
In the continuum mechanics approach to fracture,
the fracture toughness of a' material is normally
assessed in terms of some crack tip parameter at the
initiation of crack growth. Specifically, fracture will
occur when the crack tip strain, et, is exceeded over
some microstructurally
significant characteristic distance, 10, ahead of the crack tip. Under conditions of
small scale yielding

et = co]

. . . (23)

where b is the crack tip opening displacement, x the


distance ahead of the crack, and c a constant of the
order of 1.
At fracture initiation, et = ef over a distance x = 10,
when b = be and, for small scale yielding

be

CK2 / Eay

(24)

where K is the stress intensity factor, E Young's


modulus, ay the yield stress, and C a numerical
constant depending on the work hardening exponent
n, and the stress state, typically '" 0'5-0,6.
The problem is then reduced to relating be to some
microstructural feature in the composite. One suggestion is that be scales with some average interparticle
spacing, but for 2000 series alloys reinforced with
Al203 particles the agreement is poor, especially for
particle sizes typical of most commercial composites.
Hahn and Rosenfield 128assumed that crack extension occurs when the extent of the heavily deformed
region ahead of the crack tip is comparable with the
width of the unbroken ligaments separating cracked
particles. They related the critical stress intensity
factor for fracture to the volume fraction of reinforcement by the expression

K1C

[2ayE(n/6)1/3d]1/2j;1/6

. (25)

where K1C is the critical stress intensity factor (i.e.


fracture toughness), a y' the yield stress, E Young's
modulus, d the particle diameter, and fp the volume
fraction of particles.
A problem with this model is that it predicts an
increase in toughness with increasing strength; a
prediction which is not generally valid in unreinforced

Lloyd

alloys. In Al203 reinforced 6061 and 2014 alloys


Klimowicz and Vecchio V? found that the fracture
toughness decreased with aging time, and hence
increasing strength, and also continued to decrease
even in the overaged condition where the strength
was
decreasing.
However,
Manoharan
and
Lewandowski r''' obtained reasonable agreement with
this model for underaged AI-Zn- Mg-Cu matrix composites containing 15 and 20 vol.-oiO SiC particles of
13 urn average diameter. The equation also predicts
an increase in fracture toughness with increasing
particle size, which is also not in agreement with
some experiments on unreinforced alloys.P! Some
studies in composites have reported an increase in
toughness with increasing particle size, 126,132 while
other studies report no affect of particle size. 11 7 In
general, the fracture toughness does not show a strong
dependence on reinforcement particle size.P?
Kraffr'P" also proposed that fracture was controlled
by failure of ligaments ahead of the crack tip and
suggested that ligament failure occurred at the instability strain observed in a smooth bar tensile test.
This model gives the expression
K1C

En(2nA)1/2

(26)

where E is Young's modulus, n the work hardening


exponent, and A the process zone size.
Recently, Topur134 has compared equation (26)
with fracture toughness data from Al203 reinforced
2014 (AI-Cu) and obtained a process zone size
A ~ I urn. She concluded that this was too small for
the 15% of 9 urn diameter particles studied, and
suggested a 'brittle' variant of the model by replacing
the strain hardening exponent with the fracture strain
of the composite. This approach gives A 'in the range
10-100 urn, which is comparable with the spacing
between
particle
clusters
in
the
composite
in ves tiga ted.
The Krafft model is a limit load failure model
which has been further developed by Thomason 123,135
to represent the strain distribution at the crack tip
more accurately. For fracture which is nucleation
controlled the fracture toughness is given by
K1C

258(pG'yEBe)1/2

Particle reinforced AI and Mg matrix composites

Bf,
the fracture strain, may be a reasonable
assumption.
Table 4 compares the fracture toughness predictions of equations (25) and (27) with the experimental
results of Klimowicz
and Vecchio 129 for 6061/
A1203/15p and (AI-Cu) 2014/AI203/15p.
In comparing the data the particles were assumed to be
homogeneously distributed, with a 10 urn particle size,
and the crack radius was taken as 50 urn,
As expected the Hahn and Rosenfield equation
predicts an increase in fracture toughness
with
increased aging which is contrary to experiment. The
limit load approach of Thomason correctly predicts
the general trend with aging time, but tends to
underestimate the fracture toughness, particularly for
the 2014 matrix composite. However, the extent of
agreement
does suggest that models based on
nucleation controlled fracture may be appropriate for
particle composites. In this case control of particle
volume fraction and distribution will be important
for optimum fracture toughness.
There is not much information on the influence
of temperature
on fracture
toughness,
but in
2009/SiC/20p there is a small increase between 25
and 200C, and a decline above 220C.136 A micromechanical model based on a critical strain concept
was used to predict the temperature dependence, with
the local fracture strain being inferred from tensile
data.
It is apparent from this discussion that the present
models of fracture toughness need extensive development to explain the toughness of MMCs. It is difficult
to take account of reinforcement
inhomogeneity,
particle size distribution, and any residual stresses
present in the composite, all of which are likely to
influence the fracture toughness.
Finally, it should be noted that obtaining valid
plane strain fracture toughness values for composites
can be a problem and many of the results in the
Be ~

Table 4

. (27)

where p is the crack tip radius, typically '" 50 urn, G'y


the yield stress, E Young's modulus, and Be the
microvoid nucleation strain, which is dependent on
the mean normal stress, the volume fraction of void
nucleating particles, and the particle-matrix
interfacial strength.
Notice that in equation (27) the decrease in ductility'
with increasing yield stress will counter the predicted
increase in toughness with increasing stress, unlike
the situation
in the Hahn- Rosenfield
model.
However, equation (27) requires an expression for the
critical strain for void nucleation in terms of matrix
strength, particle size, distribution, and stress state
before the fracture toughness can be related to the
microstructural
features of the composite. As noted
previously, the fracture process is not sufficiently
understood for this to be done at present. However,
recent experiments on notched tensile bars showed
that the fracture strain was relatively insensitive to
the degree of far field triaxiality, suggesting that

17

Comparison between theoretical Kth and


experimental
Kexp fracture
toughness
values
cry,
MN m-2

Composite
Thomason
AI-Cu

Elongation,

Kexp,
MN m-3/2

MN m-3/2

242
217
195

254
120
124

190
181

117
156

~h'

equation

2014/AI203/15p
3 h, 160C
7 h, 160C
16 h, 160C
40 h, 160C
48 h, 160C
AI-M92Si
6061/AI203/15p
1 h, 17rC
3 h, 17rC
10 h, 17rC
25 h, 17rC
100 h, 177C

331
441
469
420
372

6
2
1
1

221
317
345
331
276

247

5
3
3
4

232
227
215
212

215
218
176
172
180

331

242

311

221

247

248

Hahn and Rosenfield


AI-Cu
2014/AI203/15p
3 h, 160C
AI-Mg2Si
6061/AI203/15p
1 h, 177C

International

equation

Materials Reviews

1994

Vol. 39

No.1

18

Lloyd

Particle reinforced AI and Mg matrix composites

280-r---=-----------------,

1O-4.-r-----------------~

260-

(\J

IE

Z 240-

D 6061/15A1203,T6

6061~T6

10-5
Q)

u
~

, 220-

(j)
(j)

~ 10-6

~ 200-

EU

t(j)

~
~
X

:J

180-

~ 10-7

160-

0% A~03

ro

"0

(/~~/

140-

15% AIz0,

10-8

~ 120-

103

1b4

105

CYCLES

24

I
I

107

1~6

1O-9-t-----r---.----,---,-r--r-.--.-----.--~"T""""""T"-_._____._J
3
5
10
20
STRESS INTENSITY
RANGE (~K), MN m-3/2

TO FAILURE

S-N curve for 6061- T6 and 6061/AI203/15p,

T6 at R= -1

25

Fatigue crack growth rate for 6061- T6 and


T6 at R=-1

6061/AI203/15p,

literature are acknowledged to be invalid according


to standard testing practice, e.g. ASTM E399 and
BS 5447. Residual stresses and inhomogeneity of
particle distribution often result in excessive curvature
in the initial fatigue crack which invalidates the test.
Various aspects of fracture toughness testing of
particle metal matrix composites have been recently
considered by Roebuck and Lord.137
Fatigue

Many of the potential applications for composites


require a resistance to cyclic loading, and several
fatigue studies in MMCs have been reported - see
for example, Refs. 138-141. Intuitively, one would
expect the low cycle fatigue behaviour of MMCs to
be somewhat worse than unreinforced alloys because
of the lower ductility in composites, whereas the high
cycle performance should be improved because of the
higher modulus. While there is considerable scatter
in the published data, reflecting variations in processing and material quality, these expectations are
generally fulfilled. Figure 24 shows the 8- N curve for
extruded unreinforced 6061- T6 and 6061/Al2 03/
15p- T6 at R = -1. However, several studies141,142
have shown that the improvement in fatigue life
evident in stress life data is eliminated when compared on a strain life basis. Under constant strain
amplitude conditions the MMC is inferior in the
low cycle regime where plastic strains dominate, and
in the high cycle regime the composite is little different
to unreinforced material. The improvement observed
in constant stress amplitude tests reflects the fact that
with the higher Young's modulus of the composite,
the strains in the composite are lower than those in
the unreinforced material at the same stress level.
Constant stress amplitude tests involve both fatigue
crack initiation and crack propagation. Crack
initiation has been associated with defects in the
composite, intermetallic particles, and large reinforcement particle or particle clusters. It has been suggested
that fatigue cracks initiate late in the life of composites.143 Fatigue crack propagation is generally characterised in terms of fracture mechanics, and is investigated by crack growth experiments. Crack growth
experiments on composites show the typical sigmoidal
International

Materials Reviews

1994

Vol. 39

No.1

shaped curve of crack growth rate against stress


intensity range exhibited by unreinforced alloys. The
curve is bounded at the low crack growth end by the
threshold stress intensity range for initiating growth
(AKth) and at the high growth rate end by the stress
intensity for unstable crack growth, i.e. Kmax
approaches the fracture toughness K1C' The data
generally shows that the threshold stress intensity
range is higher in MMCs than in unreinforced material, and an example of this is shown in Fig. 25, again
for 6061/AI203/15p, T6. This higher threshold is
attributed to crack closure effects. You and Allison 144
found that in 2124/SiC/20p, particle-matrix decohesion was prevalent at near threshold, while there
was increasing evidence of particle cracking at higher
AKs. This study also found that the fatigue crack
growth rate was relatively insensitive to aging.
However, in 7XXX/SiC/20p the situation was found
to be more complex, with the behaviour depending
on particle size and the stress intensity range.l'"
Since the crack growth rate plot is bounded by
AKth and Kmax = K1C, and since AKth is higher and
K1C lower than for unreinforced alloys, the stress
intensity range for fatigue crack growth in composites
will be smaller than for unreinforced alloys. Other
than this, the fatigue crack growth rate curve for
composites has the same general shape as for unreinforced alloys, and does exhibit a linear stage II
regime obeying the Paris law
da/dN = BAKs

. . . . . . . . . . (28)

where da/dN is the crack growth rate, AK the stress


intensity range, and Band 8 are constants.
"Davidson 139 found that for a wide range of different
aluminium composites, which included mechanically
alloyed, powder metallurgy, and molten metal processed composites, the parameters Band
8 were
related by the expression
In B

= -16'1-

228 . . . . .

(29)

This relationship was developed for R = 0'1, and


compares with In B = -0,143 - 03758 obtained for
unreinforced Al alloys.l''" Since the correlation
between Band
8 is very sensitive to R ratio in

Lloyd

aluminium alloys, equation (29) is also likely to vary


with R. However, such an approach may prove useful
in estimating fatigue performance.
The results to date show that the fatigue performance of the composite is different to that of the
unreinforced matrix, but whether the fatigue behaviour is improved or not depends on the mode of
testing. Most of the studies indicate that the fatigue
performance can be comparable with or better than
the unreinforced matrix, except at high stress or strain
amplitudes where the reduced ductility of the composite influences the behaviour. There are insufficient
data to assess the influence of particle size, shape,
etc., and fatigue crack growth models are not sufficiently well developed to make any predictions.
Creep

In previous sections it was pointed out that particle


reinforcement provides an increase in the elastic
modulus at elevated temperatures, and only a relatively small increase in the flow stress. for many high
temperature applications the creep properties of the
material are important and .several recent creep studies have been reported, primarily on powder processed
composites.
The steady state creep rate can be expressed as
8 = AO"" exp(-Q/RT)

. . . . . . . . (30)

where 8 is the creep rate, A a constant, 0" the creep


stress, n the stress exponent, Q the activation energy,
R the gas constant, and T the absolute temperature.
Applying this equation to the tensile creep of
composites gives high n-values, 9,5-20,5, and
high activation energies, 390-400 kJ mol-l (Refs.
145-149). These values compare with n = 5 and
Q = 142 kJ mol-l for the self-diffusion of aluminium,
A high stress exponent is reminiscent of oxide dispersion strengthened (ODS) alloys. A recent study
of ODS Al and AI-Mg alloys showed values of
n = 15-25, and apparent
activation energies of
500-540 kJ mol-l (Ref. 150). To explain this behaviour it is usual to introduce the concept of a threshold
stress, 0"0' and replace the creep stress, 0", by (0" - 0"0)
in equation (30). Park et al.151 used this approach in
analysing the creep in shear of 6061/SiC/30p. Their
measurements extended over seven orders of magnitude of strain rate and exhibited two regimes of
behaviour. At low stress and strain rates, n is high
and increases with decreasing stress, indicative of a
threshold stress of 81 MN m - 2. At high stresses and
strain rates n tends towards a constant value of 74.
Substituting the threshold stress of 81 MN m - 2 into
the creep equation gives a value for n of 5, consistent
with unreinforced alloys. Park et ale suggested that
the threshold stress was due to dislocation interaction,
not with the SiC reinforcing interaction, but with fine
oxide particles incorporated into the composite
during powder processing.
One problem with these tensile creep studies is that
the extent of steady state creep is quite limited, raising
questions regarding the validity of some of the data.
To overcome this difficulty Pandey et al.152 have
tested AI/SiC/lOp, 20p, and 30p in compression, and
with two particle sizes, 17 and 145 urn. They found

Particle reinforced AI and Mg matrix composites

19

that the creep rate decreased with increasing volume


fraction of reinforcement, and the threshold stress
also increased with volume fraction, from
145 MN m-2 for 10% (1'7 urn) to 328 MN m-2 for
300~, tested at 350C. This suggests that the creep
rate is not controlled by oxide dispersoids. They
analysed their results in terms of the structure
invariant model proposed by Sherby et al.153
s

3
= A'DL
A
b
5

(0" - 0"0)8

(31)

where A' is a constant, DL the lattice diffusivity, ).,the


subgrain size, b the Burgers vector, and E the Young's
modulus.
/ "This equation describes the creep of metals with a
constant substructure deforming by lattice diffusion
controlled creep. They suggest that the subgrain size
is controlled by the average spacing between reinforcing particles, and also stabilised by the reinforcing
phase. For composites with fine, 17 urn reinforcing
particles, they found good agreement with this equation, and about an order of magnitude improvement
in creep rate over unreinforced alloys. However,
for coarse reinforcement, 145 and 459 urn, while
the threshold of 158 MN m - 2 is equal to that for the
finer reinforcement, the creep rate in terms of the
effective stress, (0" - 0"0) showed no improvement over
unreinforced alloys. This behaviour can be explained
if it is assumed that the interparticle spacing in the
case of the coarser reinforcement is too large to
influence the substructure. However, there is presently
no satisfactory explanation for the value of the threshold stress observed.
From the creep experiments to date it would appear
that a high volume fraction of fine particle reinforcement can provide significant improvement in creep
resistance up to about 350C. At higher temperatures
the reinforcing particles are unlikely to exercise much
constraint on the plasticity, and the creep behaviour
will approach that of the matrix.

Commercial aspects
As noted in the introduction, particle MMCs are now
at the commercial production stage, and a whole
range of factors have to be addressed to produce a
cost competitive component. Table 5 lists the factors
involved in determining the finished component costs.
Considering that MMCs contain hard ceramic

Table 5 Finished component cost factors for an


. extruded wrought product
Extrusion billet cost
Material
Sawing
Extrusion cost
Extrusion speed
Die cost
Die wear
Recoveries
Cutting
Machining
Bending and forming
Welding
Surface finishing - anodising
Transportation
Recycling

International

or painting

Materials Reviews

1994

Vol. 39

No.1

20

Lloyd

Particle reinforced AI and Mg matrix composites

particles, have high wear resistance and limited ductility, even the most straightforward operation, such
as sawing, can require modification of the conventional equipment. Extensive sawing and machining
studies have been carried out, and viable commercial
practices established.P"
Most applications require some degree of machining, such as sawing, milling, drilling, reaming, and
tapping. The ceramic particles present in composites
have hardnesses approaching that of the machine tool
material, which means that abrasive wear of the
tooling is the main issue. To minimise abrasive wear
the area and contact time between the tool and the
workpiece should be kept as low as possible during
the machining operation. This means that MMCs
should be machined at higher feed rates and depth of
cut than traditionally used with unreinforced Al
alloys. As an example, face milling of unreinforced Al
is typically carried out at a feed rate of around
0013 ern per tooth. Increasing this by a factor of 4
for the composite not only increases the removal rate,
but also increases tool life by a factor of 3. Low
cutting speeds, which minimise the temperature rise
during machining, also increases tool life.
The hardness of the tooling is an important factor
in controlling tool life. High speed steel tools are
dulled in seconds, while conventional and coated
carbide tooling last only a few minutes. By far, the
most cost effective tooling is polycrystalline diamond
(PCD), provided it is of high quality. Unfortunately,
PCD tooling is not yet available for taps or very
small diameter drills and reamers, so carbide tools
must be used for these operations.
The composition of the MMC, in terms of type,
shape, size, and volume fraction of reinforcement will
affect machine ability. A matrix with a soft reinforcement, such as graphite, will be easier to machine than
a matrix reinforced with Al2 3, The shape of the
reinforcement may influence the ease with which the
ceramic particles are sheared, but there is no information available on this. Increasing the reinforcement
particle size and volume fraction will increase the
abrasiveness of the composite, but the uniformity of
the reinforcement distribution will also be important.
Large clusters of ceramic particles will act as large
inclusions with regard to tool damage.
Many factors, therefore, affect the machining of
MMCs, but guidelines are being developed which will
enable MMCs to be commercially machined, see for
example, Ref. 155.
Many of the early composite extrusions used lubricated conical dies to facilitate extrusion, but this is
not commercially viable except for very simple, symmetrical shapes. Fortunately, the conventional direct
extrusion process with unlubricated shear faced dies
is feasible with appropriate die design and extrusion
practice, though speeds are lower and extrusion pressures higher.156.157Excessive die wear can be minimised by using high wear resistant inserts. However,
appropriate extrusion die material is still an area
requiring development, particularly for complex
shapes.
Directly casting to final shape, and other near net
shape processing reduces some of these production
difficulties, but a degree of final machining is necessary

International Materials Reviews

1994

Vol. 39

No.1

....J

2
ffi <D
~~

25.----------------------,

AI/SiC

20

I.L X. 15

OZ

I-Q
Z(j)

~zlO

~~
I.LX

I.LW
W

u
01+-r-,__r_r_1""""T""T'""T""T""T'"
.........
"""T""'T""'1"""T"""T"""............-~..,...,_,._r_T"""T__r_r_.....,....,_T'""'T""T'"
.........
""T"""r'T_r_T""'~
10
20
30
40
50
60
70
80
90
100

26

VOLUME FRACTION, 0/0

Variation of coefficient of thermal expansion


with volume fraction of SiC in variety of AI
alloy matrices

in the majority of components. In addition, a key


cost parameter in all processing routes will be the
ability to recycle the scrap and/or the final component
when its useful life is over. This is particularly important for aluminium matrix composites, because the
whole Al business philosophy is based on an ability
to recycle.
There are two strategies which can be used to
handle scrap MMCs:
1. Recycle: for reuse as composite.
2. Reclamation: reclaim the individual components
of the composite, i.e. separate the composite into
its two constituents parts, the matrix alloy and the
ceramic particles.
Both strategies have to be cost effective and
environmentally safe.
To date, no information has been reported on
recycling powder metallurgy material, but it will
presumably involve melting. Composites initially produced by the molten metal route lend themselves to
recycling by remelting, since most of the issues
involved, such as reactivity between the particle and
the molten metal, have already been considered in
the original processing. However, melt cleanness and
degassing are more difficult problems when dealing
with scrap, and scrap sorting is very important.
Fluxing and degassing technologies have been developed by Duralcan, and presumably other composite
producers will develop similar technologies.
The reclamation route is an alternative approach
which is also available to composites which can be
remelted. Using combined argon and salt fluxing, SiC
particles can be removed from the melt, and 85-90%
of the Al can be recovered.

Some general comments


This review has primarily been concerned with the
factors influencing the microstructural-mechanical
properties relationship of composites. It should be
appreciated however, that in some applications Jor
composite the physical properties are of paramount
importance. Addition of SiC particles to aluminium
alloys can reduce the coefficient of thermal expansion
of the alloys, while still maintaining a thermal conductivity equivalent to the matrix alloy. Figure 26

Lloyd

shows the influence of SiC on the coefficient of


thermal expansion of a variety of Al alloys, and the
reduction is close to rule of mixture behaviour. Low
coefficient of thermal expansion and high thermal
conductivity is an attractive combination for applications requiring dimensional stability.
While wear resistance is a systems property rather
than a materials property, there are many situations
where the wear resistance is much higher in the
composite than in the unreinforced alloy. iSS This
makes MMCs attractive for bearings, bushings, cylinder liners, and break rotors. In some cases this
property is only required at the surface, and generating a surface composite layer by spray deposition, or
some other route, may be the appropriate means
of use.
In many applications the corrosion behaviour of
the composite is important. Since many ceramic
particles, such as Al2 3, are insulators, they would
not be expected to affect corrosion behaviour directly.
SiC, however, is a conductor and changes in corrosion
behaviour need to be assessed. Even in the case of
insulating reinforcing particles, the composite microstructure is different to that of the matrix as a result
of the processing, e.g. finer grain size, coarse intermetallics, reaction products, and this can modify the
corrosion response. There is very little published work
on the corrosion behaviour of composite, and no
doubt it will depend critically on the matrix-reinforcement combination and the particular environment.
With the tendency for grain refinement in MMCs, no
increase in susceptibility to stress corrosion is
expected in recrystallised microstructures. However,
it may be more difficult to prevent recrystallisation
in those alloy matrices where an unrecrystallised
microstructure is required for maximum stress corrosion resistance. For those alloy-environment combinations where pitting and intergranular corrosion
are a consideration, a somewhat inferior performance
in the composite is expected. In the studies conducted
to date, corrosion does not appear to be a major
concern provided the appropriate matrix is chosen,
and a suitable microstructure developed.
In summary, a wide range of particle MMCs are
now available in semicommercial and full commercial
quantities. Depending on the initial fabrication route,
there is almost an unlimited range of matrices,
particles, and loadings available, which can provide
a broad range of mechanical and physical properties.
However, the properties will inevitably be inferior to
those obtainable with continuous reinforcement, so
they provide a compromise combination of properties
between unreinforced and continuous fibre reinforced
composite. With this compromise comes the ability
to use conventional fabrication methods and lower
cost. For commercial success large tonnage applications need to be developed, and these are most
likely to be in the transportation industry. Some
commercial application has occurred in the sporting
and bicycle industry, and small quantities of MMCs
have been used in space and aircraft. However, major
applications in the automobile industry, such as brake
rotors and drive shafts, are now in the late stages of
development, and other engineering applications are
also well advanced. Reproducibility of properties in

Particle reinforced AI and Mg matrix composites

21

commercial component runs, and meeting specifications with' high yields and at acceptable cost, now
become the key factors in successful exploitation. The
technologies are now available which should be able
to meet these requirements, and the next 5 years or
so will demonstrate whether particle MMCs will
become a general use engineering material, or be
limited to niche markets, which is the role they have
played to date.

Acknowledgments
The author is. grateful to Alcan International Ltd for
permission to publish, and to his colleagues at the
Alcan Research Centres in Kingston, Banbury, and
Arvida, and at Duralcan USA, San Diego and Dubuc
plants, for their contributions to the ideas and information in this article.

References
1.
2.

A. KELLY: Compos. Sci. Technol.,


T. W.

CHOU, A. KELLY,

and

1985,23,171-199.

A. OKURA: Composites,

1985,

16, 187-206.
3.

and w. R. SYMES:
in Proc. 5th Int. Conf. on 'Composite materials - ICCM-V',
(ed. W. C. Harrigan, Jr et al.), 671-685; 1985, Warrendale,
PA, The Metallurgical Society of AIME.
4. s. V. NAIR, J. K. TIEN, and R. C. BATES: Int. Met. Rev., 1985,
30, 275-290.
5. T. DONOMOTO, K. FUNATANI, N. MIURA, and N. MIYAKE: SAE
Technical Paper 830252, Detroit, MI, 1983.
6. M. H. STACEY: Mater. Sci. Techno!., 1988, 4, 227-230.
7. DEONATH and P. K. ROHATGI: J. Mater. Sci., 1980, 15,
2777-2784.
8. M. K. SURAPPA and P. K. ROHATGI: J. Mater. Sci., 1981,
16, 983-993.
9. A. BANERJI, M. K. SURAPPA, and P. K. ROHATGI: Metall. Trans.,
1983, 14B, 273-283.
10. B. P. KRISHNAN and P. K. ROHATGI: Met. Technol., 1984,
11, 41-44.
11. W. KAI, J. M. YANG, and w. C. HARRIGAN, Jr: Scr. Metall., 1989,
23, 1277-1281.
12. B. ROEBUCK and A. E. J. FORNO: in 'Modern developments in
powder metallurgy', Vol. 20, 451-465; 1988, Princeton, NJ,
Metal Powder Industries Federation, American Powder
Metallurgy Institute.
13. J. J. LEWANDOWSKI, C. LIU, and w. H. HUNT, Jr: in 'Processing
and properties for powder metallurgy composites', (ed.
P. Kumar et a!.), 117-137; 1987, Warrendale, PA, The Metallurgical Society of AIME.
14. D. J. LLOYD, H. P. LAGACE, and A. D. McLEOD: in 'Controlled
interphases in composite materials - ICCM-III', (ed.
R. Ishida), 359-376; 1990, London, Elsevier Applied Science.
15. D. J. LLOYD and B. CHAMBERLAIN: in 'Cast reinforced metal
composites', (ed. S. G. Fishman and A. K. Dhingra), 263-269;
1988, Metals Park, OR, ASM International.
16. c. M. ADAMS and R. E. LEWIS: in 'Rapidly solidified crystalline
alloys', (ed. S. K. Das et al.), 157-183; 1985, Warrendale, PA,
The Metallurgical Society of AIME.
17. M. S. ZEDALIS, J. M. PELTIER, and P. S. GILMAN: in 'Light-weight
alloys for aerospace applications', (ed. E. W. Lee et a!.),
323-334; 1989, Warrendale, PA, The Metallurgical Society
of AIME.
18. F. A. BADIA and P. K. ROHATGI: Trans. AFS, 1969, 76, 402-406.
19. A. M. PATTON: J. Inst. Met., 1972, 100, 197-201.
20. A. W. NEUMANN, C. J. van OSS, and J. SZEKELY: Kolloid-Z.u.Z.
Polym., 1973, 251, 415-423.
21. P. K. ROHATGI: in 'Interfaces in metal-matrix composites', (ed.
A. K. Dhingra and S. G. Fishman), 185-202; 1986, Warrendale, PA, The Metallurgical Society of AIME.
22. K. C. RUSSELL, J. A. CORNIE, and S.-Y. OH: in 'Interfaces in
metal-matrix composites', (ed. A. K. Dhingra and S. G.
Fishman), 61-91; 1986, Warrendale, PA, The Metallurgical
Society of AIME.
J. DINWOODIE, E. MOORE, C. A. J. LANGMAN,

International

Materials Reviews

1994

Vol. 39

No.1

22
23.

Lloyd

Particle reinforced AI and Mg matrix composites

and R. ASTHANA: in 'Cast reinforced metal


composites', (ed. S. G. Fishman and A. K. Dhingra), 61-66;
1988, Metals Park, OH, ASM International.
24. F. DELANNAY, L. FROYEN, and A. DERUYTERRE: J. Mater. Sci;
1987, 22, 1-16.
25. M. K. SURAPPA and P. K. ROHATGI: Met. Technol., 1978,
5, 358-361.
26. v. G. GORBUNOV, u. D. PARSHIN, and v. v. PANIN: Russ. Cast.
Prod., Aug. 1974, 348.
27. J. TAFTO, K. KRISTIANSEN, H. WESTENGEN, A. NYGARD, J. B.
BORRADAILE, and D. O. KARLSEN: in 'Cast reinforced metal
composites', (ed. S. G. Fishman and A. K. Dhingra), 71-75;
1988, Metals Park, OH, ASM International.
28. F. A. BADIA: Trans AFS, 1971, 79, 347-350.
29. Alcan Aluminum Corp.: US Pat. 4786467, 1988.
30. J. B. BORRADAILE, S. SKJERVOLD, and w. RUCH: in Extended
Abstracts of Conf. 'Metal matrix composites: property
optimisation and applications', London, 1989, The Institute
of Metals, Paper 10.1.
31. T. SRITHARAN, K. XIA, H. HEATHCOCK, and J. MIHELICH: in
'Metal and ceramic matrix composites: Processing, modeling
and mechanical behavior', (ed. R. B. Bhagat et al.), 13-22;
1990, Warrendale, PA, The Metallurgical Society of AIME.
32. D. J. LLOYD: in 'High performance composites for the 1990s',
(ed. S. K. Das), 33-45; 1990, Warrendale, PA, The
Metallurgical Society of AIME.
33. J. A. CORNIE, H.-K. MOON, and M. C. FLEMINGS: in 'Fabrication
of particulates reinforced metal composites', (ed. J. Masounave
and F. G. Hamel), 63-78; 1990, Materials Park, OH, ASM
International.
34. R. MEHRABIAN, R. G. RIEK, and M. C. FLEMINGS: Metall. Trans.,
1974, 5, 1899-1905.
35. R. MEHRABIAN, A. SATO, and M. C. FLEMINGS: Light Met., 1975,
38, 177-193.
36. P. R. GIBSON, A. J. CLEGG, and A. A. DAS: Foundry Trade J.,
Feb. 1982, 253-263.
37. Dow Chemical Co.: US Pat. 4432936.
38. E. M. KLIER: SM thesis, Massachusetts Institute of
Technology, 1988.
39. S. CARON and J. MASOUNAVE: in 'Fabrication of particulates
reinforced metal composites', (ed. J. Masounave and
F. G. Hamel), 107-113; 1990, Materials Park, OH, ASM
International.
40. 1. A. CORNIE: Personal communication, Kingston, ON, 1990.
41. A. W. URQUHART: Mater. Sci. Eng., 1991, AI44, 75.
42. M. K. AGHAJANIAN, M. A. ROCAZELLA, J. T. BURKE, and s. D.
KECK: J. Mater. Sci., 1991, 26, 447.
43. A. R. E. SINGER: US Pat. 4515864, 1985.
44. Osprey Metals: UK Pat. 1 379261, 1975.
45. Osprey Metals: UK Pat. 1 472939, 1977.
46. United Kingdom Atomic Energy Authority: US Pat.
4928 745, 1990.
47. T. C. WILLIS: Met. Mater., 1988, 4, (8), 485-488.
48. J. WHITE, I. R. HUGHES, T. C. WILLIS, and R. M. JORDAN: in
'Proc. 4th Int. Conf. on aluminium-lithium alloys', J. Phys.,
1987, 48, (9), C3, 347-353.
49. Martin Marietta Corp.: US Pat. 4 915908, 1990.
50. Martin Marietta Corp.: US Pat. 4 710348, 1987.
51. R. M. AIKIN, Jr: 'The mechanism of dispersion strengthening
and fracture in Al-based XD alloys', Contract Report 4276,
National Aeronautics and Space Administration, Feb.
1990.
52. US Pat. 4 808 372.
53. A. G. METCALFE: 'Composite materials, Vol. 1 - Interfaces in
metal matrix composites'; 1974, New York, Academic Press.
54. A. MORTENSEN: in 'Mechanical and physical behaviour of
metallic and ceramic composites', 9th Rise Int. Symp. on
Metallurgy and Materials Science, Rise National Laboratory,
Roskilde, Denmark, (ed. S. I. Anderson et al.), 141-155.
55. D. J. LLOYD and I. JIN: Meta 11. Trans., 1988, 19A, 3107-3109.
56. R. WARREN and c. H. ANDERSSON: Composites, 1984, 15,
101-111.
57. D. J. LLOYD and E. DEWING: in Proc. Int. Symp. on 'Advanced
structural materials', (ed. D. S. Wilkinson), 71-77; 1988,
Montreal, PQ, Pergamon Press.
58. D. J. LLOYD: Compos. Sci. Technol., 1989,35, 159-179.
59. c. M. GABRYEL and A. D. McLEOD: Metall. Trans., 1992, 23A,
1279-1283.
60. K. L. MAMPEL: Z. Phys. Chern., 1940, A187, 43-57; 235-249.
P. K. ROHATGI

International

Materials Reviews

1994

Vol. 39

No.1

61.

L. SALVO, and M. SuERY: in


'Fabrication of particulates reinforced metal composites', (ed.
J. Masounave and F. G. Hamel), 31-39; 1990, Materials Park,
OH, ASM International.
62. H. RIBES and M. SuERY: Scr. Metall., 1989, 23, 705-709.
63. M. C. FLEMINGS: 'Solidification processing'; 1974, New York,
McGraw-Hill.
64. D. G. THOMAS: J. Colloid Sci., 1965, 20, 267-277.
65. H.-K. MOON: PhD' thesis, Massachusetts Institute of
Technology, 1990.
66. M. K. SURAPPA and P. K. ROHATGI: Metall. Trans., 1981,
12B, 327-332.
67. J. A. CORNIE: Personal communication, Boston, MA, 1991.
68. w. H. HUNT, Jr, D. RICHMOND, and R. D. YOUNG: in Proc. 6th
Int. Conf. on 'Composite materials - ICCM VI', (ed.
F. L. Matthews et al.), 2.209-2.203; 1987, London, Elsevier
Applied Science.
69. T. M. OSMAN, J. J. LEWANDOWSKI, and w. H. HUNT, rr: in
'Fabrication of particulates reinforced metal composites', (ed.
J. Masounave and F. G. Hamel), 209-216; 1990, Materials
Park, OH, ASM International.
,
70. c. A. ROGERS: 'Packing and covering', Cambridge
Mathematical Tracts, Vol. 54; 1964, Cambridge University
Press.
71. J. F. RICHARDSON and w. N. ZAKI: Chern. Eng. Sci., 1954, 3, 65.
72. G. H. GEIGER and D. R. POIRIER: 'Transport phenomena in
metallurgy'; 1973, New York, Addison-Wesley.
73. I. JIN and D. J. LLOYD: in 'Fabrication of particulates reinforced
metal composites', (ed. J. Masounave and F. G. Hamel),
47-52; 1990, Materials Park, OH, ASM International.
74. P. K. ROHATGI, R. ASTHANA, and s. DAS: Int. Met. Rev., 1986,
31, 115-139.
75. D. M. STEFANESCU and B. K. DHINDAW: in 'Metals handbook',
9 edn, Vol. 15, 142-147; 1988, Metals Park, OH, ASM
International.
76. M. TAYA, K. MURAMATSU, D. J. LLOYD, and R. WATANABE:
JSME lnt. J., Ser. 1, 1991, 34, 198-206.
77. D. J. LLOYD: in 'Metal matrix composites - processing, microstructure and properties', 12th Rise Int. Symp. on Metallurgy
and Materials Science, Rise National Laboratory, Roskilde,
Denmark, 1991, (N. Hansen et al.), 81-99.
78. F. J. HUMPHREYS: Acta Metall., 1977,25, 1323-1344.
79. w. S. MILLER and F. J. HUMPHREYS: in 'Fundamental relationships between microstructures and mechanical properties of
metal matrix composites', (ed. M. N. Gungor and P. K. Liaw),
517-541; 1989, Warrendale, PA, The Metallurgical Society
of AIME.
80. M. FERRY, P. MUNROE, A. CROSKY, and T. CHANDRA: in 'Metal
matrix composites - processing, microstructure and prop-.
erties', 12th Rise Int. Symp. on Metallurgy and Materials
Science, Rise National Laboratory, Roskilde, Denmark, 1991,
(N. Hansen et al.), 337-342.
81. M. W. MAHONEY and A. K. GHOSH: Metall. Trans., 1987,
18A, 653-661.
82. M. MABUCHI, K. HIGASHI, Y. OKADA, S. TANIMURA, T. IMAf, and
K. KUBO: Scr. Metall. Mater., 1991,25, 2003-2006.
83. M. MABUCHI, K. HIGASHI, S. WADA, and s. TANIMURA: Scr.
Metall. Mater., 1992,26, 1269-1274.
84. J. PILLING: Scr. Metall., 1989,23, 1375-1380.
85. K. HIGASHI, T. OKADA, T. MUKAI, S. TANIMURA, T. G. NIEH, and
J. WADSWORTH: se-. Metall. Mater., 1992,26, 185-190.
86. G. NUSSBAUM, P. SAINFORT, G. REGAZZONI, and H. GJESTLAND:
Scr. Metall., 1989, 23, 1079-1084.
87. T. G. NIEH and R. F. KARLEK: Scr. Metall., 1984, 18,25:....28.
88. s. NUTT and R. W. CARPENTER: Mater. Sci. Eng., 1985, 75,
169-177.
89. H. J. RACK: in Proc. 6th Int. Conf. on 'Composite materials ICCM-VI', (ed. F. L. Matthews et al.), 2.382-2.389; 1987,
London, Elsevier Applied Science.
90. J. M. PAPAZIAN: Metall. Trans., 1988, 19A, 2945-2953.
91. P. B. PRAGNELL and w. M. STOBBS: in Proc. 7th Int. Symp. on
'Composite materials - ICCM-VII', (ed. Yunshu et al.),
573-578; 1989, Oxford, Pergamon.
92. s. SURESH, T. CHRISTMAN, and Y. SUGIMURA: Scr. Metall., 1989,
23, 1599-1602.
93. K. K. CHAWLA, A. H. ESMAELI, A. K. DATYE, and A. K. VASUDEVAN:
Scr. Metall. Mater., 1991,25, 1315-1319.
94. E. HUNT, P. D. PITCHER, and P. J. GREGSON: Scr. Metall., 1990,
24, 937-941.
J. G. LEGOUX, G. L'ESPERANCE,

.>

Lloyd
95.

P. B. PRANGNELL and w. M. STOBBS: in 'Metal matrix composites - processing; microstructure and properties', 12th Rise
Int. Symp. on Metallurgy and Materials Science, Rise
National Laboratory, Roskilde, Denmark, 1991, (N. Hansen

96.
97.

S. CERESARA and P. FIORINI: Powder Metall., 1979,22, (1), 1-4.


M. STRANGWOOD, C. A. HIPPSLEY, and J. J. LEWANDOWSKI: Scr.

98.

I. DULTA, S. M. ALLEN,

et al.), 603-610.

Metall., 1990, 24, 1483-1487.

and J. L. HAFLEY: Metall. Trans., 1991,


22A, 2553-2563.
99. H. J. RACK and P. K. CHAUDHURY: in 'Light-weight alloys for
aerospace applications', (ed. E. W. Lee et al.), 345-357; 1989, .
Warrendale, PA, The Minerals, Metals and Materials Society.
100. R. J. ARSENAULT and s. B. wu: Mater. Sci. Eng., 1987,96, 77-88.
101. J. C. HALPIN: 'Primer on composite materials: Analysis', Rev.
edn, 130; 1984, Lancaster, PA, Technomic Publ.
102. M. TAYA and R. J. ARSENAULT: Scr. Metall., 1987, 21, 349-354.
103. D. L. McDANELS: Metall. Trans., 1985, 16A, 1105-1115.
104. F. J. HUMPHREYS: in 'Mechanical and physical behaviour of
metallic and ceramic composites', 9th Rise Int. Symp. on
Metallurgy and Materials Science, Rise National Laboratory,
Roskilde, Denmark, (ed. S. I. Anderson et al.), 51-74.
105. F. J. HUMPHREYS, H. BASU, and M. R. DJAZEB: in 'Metal matrix
composites - processing, microstructure and properties', 12th
Rise Int. Symp. on Metallurgy and Materials Science, Rise
National Laboratory, Roskilde, Denmark, 1991, (N. Hansen
et al.), 51-66.

106.

A. NEEDLEMAN, and s. SURESH: Acta Metall.,


198~3~ 3029-305Q
107. v. TVERGAARD: in 'Metal matrix composites - processing,
microstructure and properties'. 12th Rise Int. Symp. on
Metallurgy and Materials Science, Rise National Laboratory,
Roskilde, Denmark, 1991, (N. Hansen et al.), 173-188.
108. A. LEVY and J. M. PAPAZIAN: in 'Metal matrix composites processing, microstructure and properties', 12th Rise Int.
Symp. on Metallurgy and Materials Science, Rise National
Laboratory, Roskilde, Denmark, 1991, (N. Hansen et al.),
475-482.
109. M. R. PIGGOTT: 'Load bearing fibre composites'; 1980, Oxford,
Pergamon Press.
110. v. C. NARDONE and K. M. PREWO: Scr. Metall., 1986,20,43-48.
111. R. J. ARSENAULT and N. SHI: Mater. Sci. Eng., 1986,81, 175-187.
112. M. TAYA, K. E. LULAY, and D. J. LLOYD: Acta Metall. Mater.,
1991, 39, 73-87.
113. P. J. WITHERS, W. M. STOBBS, and o. B. PEDERSEN: Acta Metall.,
1989,37,3061-3084.
114. M. TAYA, K. E. LULAY, and D. J. LLOYD: in 'Morris E. Fine
Symp.', (ed. P. K. Liaw et al.), 153-164; 1990, Warrendale,
PA, The Metallurgical Society of AIME.
115. M. S. ZEDALIS, P. S. GILMAN, and s. K. DAS: in 'High performance
composites for the 1990s', (ed. S. K. Das et al.), 61-81; 1990,
Warrendale, PA, The Metallurgical Society of AIME.
116. R. DWlVEDI, G. ALTLAND, P. BARRON-ANATOLIN, J. LEIGHTON,
and F. G. HAMEL: in 'Int. off highway and powerplant cong.',
Milwaukee, WI, 1991, SAE Technical Paper 911770.
117. Y. FLOM: PhD thesis, University of Maryland, 1987.
118. Y. FLOM and R. J. ARSENAULT: in Proc. 6th Int. Conf. on
'Composite materials - ICCM-VI', (ed. F. L. Matthews et al.),
2.189-2.208; 1987, London, Elsevier Applied Science.
119. D. J. LLOYD: Acta Metall. Mater., 1991,39,59-71.
120. D. J. LLOYD, P. L. MORRIS, and E. NEHME: in 'Fabrication of
particulates reinforced metal composites', (ed. J. Masounave
and F. G. Hamel), 235-243; 1990, Materials Park, OH, ASM
International.
121. D. C. DRUCKER: in 'High strength materials', (ed. V. Zackay),
795-830; 1965, New York, Wiley.
122. G. BAO, J. W. HUTCHINSON, and R. M. McMEEKING: Acta Metall.
Mater., 1991,39, 1871-1882.
123. P. F. THOMASON: 'Ductile fracture of metals'; 1990, Oxford,
Pergamon Press.
124. A. S. ARGON, J. 1M, and R. SAFOGLU: Metall. Trans., 1975,
6A, 825-837.
125. T. MOCHIDA, M. TAYA, and D. J. LLOYD: J. Jpn Inst. Met., 1991,
32, (10) 931.
126. A. MORTENSEN: in 'Fabrication of particulates reinforced metal
composites', (ed. J. Masounave and F. G. Hamel), 235-243;
1990, Materials Park, OH, ASM International.
T. 'CHRISTMAN,

Particle reinforced AI and Mg matrix composites

23

127.

D. L. DAVIDSON: in 'Metal matrix composites: Mechanisms and


properties', (ed. R. K. Everett and R. J. Arsenault), 217-234;
1991, Boston, MA, Academic Press.
128. G. T. HAHN and A. R. ROSENFIELD: Metall. Trans., 1975,
6A, 653-668.
129. T. F. KLIMOWICZ and K. S. VECCHIO: in 'Fundamental relationships between microstructures and mechanical properties',
(ed. M. N. Gungor and P. K. Liaw), 255-267; 1989,
Warrendale, PA, The Metallurgical Society of AIME.
130. M. MANOHARAN and J. J. LEWANDOWSKI: Acta Metall. Mater.,
1990, 38, 489-496.
131. T. B. cox and J. R. LOW: NASA Tech. Rep. No.3, NGR
38-087-003, 1972.
132. J. J. STEPHENS, J. P. LUCAS, and F. M. HOSKINS: Scr. Metall.,
1988,22, 1307-1312.
133. G. T. KRAFFT: Appl. Mater. Res., 1964,1,88-101.
134. E. E. TOPUR: MSc thesis, Lehigh University, Bethlehem, PA,
1990.
135. P. F. THOMASON: Acta Metall. Mater., 1992,40, 241-249.
136. B. P. SOMERDAY, Y. LENG, F. E. WAWNER, and R. P. GANGLOFF:
in 'Advanced metal matrix composites for elevated temperature', (ed. M. N. Gungor), 167-182; 1992, Materials Park,
OH, ASM International.
137. B. ROEBUCK and J. D. LORD: Mater. Sci. Technol., 1990, 6,
1199-1209.
138. w. A. LOGSDON and P. K. LIAW: Eng. Fract. Mech., 1986,
24, 737-751.
139. D. L. DAVIDSON: Eng. Fract. Mech., 1989, 33, 965-977.
140. D. L. DAVIDSON: Metall. Trans., 1991, 22A, 97-112.
141. J. K. SHANG and R. O. RITCHIE: in 'Metal matrix composites:
Mechanisms and properties', (ed. R. K. Everett and
R. J. Arsenault), 255-285; 1991, Boston, MA, Academic Press.
142. J. J. BONNEN, J. E. ALLISON, and J. E. JONES: Metall. Trans.,
1991, 22A, 1007-1019.
143. D. R. WILLIAMS and M. E. FINE: in Proc. 5th Int. Conf. on
'Composite materials - ICCM-V', (ed. W. C. Harrigan, Jr
et al.), 639-670; 1985, Warrendale, PA, The Metallurgical
Society of AIME.
144. c. P. YOU and J. E. ALLISON: in Proc. 7th Int. Conf. on
'Fracture', 3005-3012; 1989, Pergamon Press.
145. J. K. SHANG, W. YU, and R. O. RITCHIE: Mater. Sci Eng. A,
1988, 102, 181-192.
146. J. P. BAILON, J. MASOUNAVE, and c. BATHIAS: Scr. Metall., 1977,
11, 1101-1106.
147. T. G. NIEH, K. XIA, and T. G. LANGDON: J. Eng. Mater. Technol.
(Trans. ASME), 1988, 110, 77-82.
148. K. XIA, T. G. NIEH, J. WADSWORTH, and T. G. LANGDON: in
'Fundamental relationship between microstructure and
mechanical properties of metal matrix composites', (ed.
M. N. Gungor and P. K. Liaw), 543-556; 1989, Warrendale,
PA, The Metallurgical Society of AIME.
149. S. PICKARD and B. DERBY: in 'Developments in the science and
technology of composite materials (ECCM3)', (ed.
A. A. Bunsell et al.), 199-204; 1989, Amsterdam, The
Netherlands, Elsevier.
150. w. C. OLIVER and w. D. NIX: Acta Metall., 1982,30, 1335-1347.
151. K.-T. PARK, E. J. LAVERNIA, and F. A. MOHAMED: Acta Metall.
Mater., 1990, 38, 2149-2159.
152. A. B. PANDEY, R. S. MISRA, and Y. R. MAHAJAN: Acta Metall.
Mater., 1992,40, 2045-2052.
153. o. D. SHERBY, R. H. KLUNDT, and A. K. MILLER: Metall. Trans.,
1977, 8A, 843-850.
154. c. T. LANE: in 'Fabrication of particulates reinforced metal
composites', (ed. J. Masounave and F. G. Hamel), 195-201;
1990, Materials Park, OH, ASM International.
155. 'Machining guidelines', 1992, Duralcan Composites, San
Diego, CA.
156. S. BRUSETHANG, O. REISO, and w. RUCH: in 'Fabrication of
particulates reinforced metal composites', (ed. J. Masounave
and F. G. Hamel), 173-179; 1990, Materials Park, OH, ASM
International.
157. P. W. JEFFREY and s. HOLCOMB: in 'Fabrication of particulates
reinforced metal composites'. (ed. J. Masounave and
F. G. Hamel), 181-186; 1990, Materials Park, OH, ASM
International.
158. s. V. PRASAD and P. K. ROHATGI: J. Met., 1987, 39, (1), 22-26.

International

Materials Reviews

1994

Vol. 39

No.1

You might also like