Sei sulla pagina 1di 29

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/286379081

Metal–Matrix Composites

Article · October 2015


DOI: 10.1016/B978-0-12-803581-8.03950-3

CITATIONS READS

14 14,859

1 author:

Meysam Haghshenas
University of Toledo
75 PUBLICATIONS   620 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

both of project View project

Depth-sensing creep behavior of additively manufactured Ti-6Al-4V View project

All content following this page was uploaded by Meysam Haghshenas on 03 January 2018.

The user has requested enhancement of the downloaded file.


Metal–Matrix Composites$
M Haghshenas, University of Waterloo, Waterloo, ON, Canada
r 2016 Elsevier Inc. All rights reserved.

Introduction

The possibility of mixing different material systems, i.e., metals, ceramics, and polymers, provides the opportunity for infinite
variations and functioning properties. Human beings have been making these combinations for thousands of years. One early
example is mud bricks. Mud can be dried out into a brick shape to give a building material. It provides good compressive strength
but it breaks quite easily under bending due to poor tensile strength. Straw seems very strong in stretching, but it can be crumpled
up easily. By mixing mud and straw together it is possible to make bricks that are resistant to both squeezing and tearing and make
excellent building blocks. Concrete is another example of an ancient compound. It is a mixture of aggregate cement and sand.
Concrete provides good compressive strength (being resistance to squashing). In recent times, by adding metal rods or wires to the
concrete, its tensile (and bending) strength increases (reinforced concrete).
Beside human beings, Mother Nature has developed natural compounds in both animals and plants. Almost all natural
materials which must bear load, i.e., wood, bone, muscle, are indeed compound materials. Wood is a natural compound, made
from long cellulose fibers (a polymer) held together by a much weaker substance called lignin. Cellulose is found in cotton, but
without the lignin to bind it together it is much weaker. The two weak substances, lignin and cellulose, together form a much
stronger structure. The bone in human and animals' body is also a natural compound (Figure 1). It is made from a hard but brittle
material called hydroxyapatite (which is mainly calcium phosphate) and a soft and flexible material called collagen (which is a
protein). Collagen is found in hair and finger nails as well. On its own it would not be much use in the skeleton but it can combine
with hydroxyapatite to give bone the properties that are needed to support the body.

Figure 1 The structure of bone in human and animals' body as a natural composite (www.classes.mst.edu).

A compound (composite) material is a macro-physical combination of, physically and/or chemically, distinct materials
(phases) where the aim is to combine beneficial properties of the constituent materials (Figure 2). Hence, very clear and abstract
knowledge of properties of the constituent materials are needed in order to select compatible materials for composite fabrication.
The end product is a unique and new material which possesses superior characteristics than those of each of the individual
components. The properties of the newly born materials are basically dependent upon the properties of their single components.


Change History: July 2015. M. Haghshenas added an ‘Abstract’ and a detailed ‘Introduction’ including Figures 1–3. ‘Types of MMCs’ were extended and
updated including classification based on matrix material, Figures 5–7, and reinforcement material, Figures 12–15. ‘Properties of MMCs’ extended to physical
and mechanical properties of MMCs and new Figures 21–28 and Table 1 were added in this Section. Two new Sections, ‘Strengthening methods in MMCs’ and
‘Fabrication of MMCs’ were added including Figure 29. Also, MMCs were classified based on matrix materials and most commonly used matrix materials
including Al-, Mg, Ti-, Cu-, and Ni-supper alloy matrixes were discussed in detail and Figures 6–11 were added here. Beside these other new sections including
‘Developing trends of MMCs,’ ‘Microstructure of MMCs,’ and ‘Design consideration’ containing Figures 16–20 were added. ‘Application of MMCs’ was
updated with novel applications of MMCs in different industries, i.e., aerospace, automotive, trains, Electronic and sport. New figures, Figures 30–40 and new
Tables 4–6 were added as well in this Section.

Reference Module in Materials Science and Materials Engineering doi:10.1016/B978-0-12-803581-8.03950-3 1


2 Metal–Matrix Composites

Figure 2 The family of composites among metals/ceramics/polymers.

This distinguishes a composite from other multiphase materials which are produced by bulk processes where one or more phases
result from phase transformation.
Composite materials are, indeed, like sandwiches. A good sandwich contains a variety of ingredients to yield a taste that no
single ingredient could provide by itself. Similarly, composite materials are those which are formed from two or more materials
producing properties or characteristics that could not be obtained from any one material. Having said this, many common
materials like metals, alloys, and processed polymers are not considered as composites even though various amounts of dispersed
phases exist in their structure. Why? The reason is that their properties are similar (or very close) to those of their base constituents.
Many ‘in situ composites.’ such as directionally solidified eutectics (i.e., Al–Si alloys) or alloys from which a ceramic phase is
precipitated, are thus considered to be alloys and not composites.
Why use composites? The biggest advantage of the composite materials is that they are light as well as strong (Figure 3). By
choosing an appropriate combination of matrix and reinforcement material, a composite can be made that exactly meets the
requirements of a particular application. Composites also provide design flexibility because many of them can be molded into
complex shapes. The most applicable properties of composite materials include, Kainer (2006):

• High stiffness and high strength


• High Young's modulus
• High fatigue strength, especially at elevated temperatures
• High corrosion resistance
• Low density
• High thermal and electrical conductivity
• High wear resistance

Figure 3 Physical and mechanical properties of composites as compared with two most commonly used alloys, i.e., steel and aluminum.

Figure 3 shows common physical and mechanical properties of the composites as compared with aluminum and steel. The
composite materials have been shown in purple color.

Matrixes and Reinforcements

The terms of matrix and reinforcement are very often used when talking about composites. Matrix is a relatively ‘soft’ phase with
specific physical and mechanical properties, i.e., ductility, formability and thermal conductivity (Figure 4). In the matrix, ‘hard’
reinforcements with high strength, high stiffness, and low thermal expansions are embedded (Figure 4). Reinforcement phase is,
Metal–Matrix Composites 3

Figure 4 Schematic and actual illustration of matrix and reinforcement within a composite structure, Kainer (2006).

therefore, a second phase (or phases) which is usually stronger and stiffer than the matrix and mainly carries the applied load to
the composite. That is, composites are demonstrated by two or more phases, one of which is stronger and stiffer that the others
and serves as the primary load carrying component. For example, in some composites, we take the advantage of combining
toughness of metals with, in most cases, high intrinsic strength of reinforcing phases (i.e., ceramics). Therefore, matrix is the
primary phase with a continuous character which holds the reinforcement phase and shares applied load with it.
Matrix and reinforcement dictate the performance of a composite depending on:

• Physical, mechanical, and chemical properties of matrix and reinforcement


• Size and distribution of constituents
• Morphology of constituents
• Interface between constituents

Fiberglass is an example of a composite that clearly transfer the concept of matrix and reinforcement phases. Nowadays, it is widely
used for boat hulls, sports equipment, building panels and many car bodies. The matrix is a plastic and the reinforcement is glass
that has been made into fine threads and often woven into a sort of cloth. Glass, by itself, is strong but brittle and will break if bent
sharply. The plastic matrix holds the glass fibers together and also protects them from damage by sharing out the forces acting
on them.
Some advanced composites are now made using carbon fibers instead of glass. These materials are lighter and stronger than
fiberglass but more expensive to produce. They are used in aircraft structures and expensive sports equipment such as golf clubs.
Carbon nanotubes and graphene have also been used successfully to make new class of composites. The potential of graphene
composites includes medical implants, engineering materials for aerospace and renewables and much more. These are even lighter
and stronger than composites made with ordinary carbon fibers but they are still very expensive. They do, however, offer
possibilities for making lighter cars and aircraft (which will use less fuel than the heavier vehicles we have now). For instance, the
new Airbus A380, the world's largest passenger airliner, makes use of modern composites in its design. More than 50% of the A380
is made of composite materials, mainly plastic reinforced with carbon fibers. This design is the first large-scale use of glass-fiber-
reinforced aluminum, a new composite that is 25% stronger than conventional airframe aluminum but 20% lighter.

Matrix
The purpose of the matrix is to bind the reinforcements together by virtue of its cohesive and adhesive characteristics, to transfer
load to and between reinforcements, and to protect the reinforcements from environments and handling. The matrix also provides
a solid form to the composite, which aids handling during manufacture and is typically required in a finished part. This is
particularly necessary in discontinuously reinforced composites, as the reinforcements are not of sufficient length to provide a
handle-able form. Because the reinforcements are typically stronger and stiffer, the matrix is often the ‘weak link’ in the composite,
from a structural perspective. As a continuous phase, the matrix therefore controls the transverse properties, inter-laminar strength,
and elevated-temperature strength of the composite. However, the matrix allows the strength of the reinforcements to be used to
their full potential by providing effective load transfer from external forces to the reinforcement.
On the basis of matrix, composites are classified to:

• Polymer Matrix Composites (PMC)


PMCs are composed of a thermoset or thermoplastic matrix embedded dispersed reinforcement phases, i.e., carbon, glass,
metal, Kevlar fibers.
• Ceramic Matrix Composites (CMC)
CMCs are composed of a ceramic matrix (i.e., SiC, Al2O3, SiN) and embedded fibers of other ceramic materials.
• Metal–matrix composites (MMC)
MMCs are the most widely used composites in the industrial scale. The main advantages of MMCs over PMCs and CMCs are
(Clyne, 2001; Matthews and Rawlings, 1996):
4 Metal–Matrix Composites

J Higher temperature capabilities


J Better radiation resistance
J Higher transverse stiffness and strength
J Less moisture absorption
J Higher thermal and electrical conductivity
J Better process-ability
J Less out gassing
J Higher fire resistance

However, in some aspects, MMCs have some drawbacks as compared with metals and PMCs:
J Higher end cost
J Relatively immature technology
J Complex fabrication methods

Table 1 presents representative properties of selected materials in each category of matrixes. The properties of the four types vary
noticeably causing profound effects on the properties of the composites using them.

Table 1 Properties of selected matrix materials

Material Class Density (g/ Modulus Tensile strength (MPa) Tensile failure strain (%) Thermal conductivity (W/ CTE (ppm/
cm3) (GPa) m K) K)

Epoxy Polymer 1.8 3.5 70 3 0.1 60


Al 6061 Metal 2.7 69 300 10 180 23
Ti (6Al-4V) Metal 4.4 105 1100 10 16 9.5
SiC Ceramic 2.9 520 – o0.1 81 4.9
Al2O3 Ceramic 3.9 380 – o0.1 20 6.7
Borosilicate Ceramic 2.2 63 – o0.1 2 5
Carbon Carbon 1.8 20 – o0.1 5–90 2

Source: Zweben (2006).

In this Chapter, the main focus will be on characteristics of MMCs. The MMCs are composed of a continuous metallic matrix
(i.e., aluminum, magnesium, copper, titanium) and a reinforcement phase which can be either dispersed ceramics (i.e., oxides,
carbides) or metallic phases (i.e., tungsten, molybdenum, lead) that constitutes a few percent to around 50% of the total volume
of the material. The objective of developing MMCs is to combine desirable properties of metallic materials and second phase
(reinforcement). The prime reason for using MMCs in structural applications is their improved specific strength and specific
stiffness relative to unreinforced alloys. Most structural metals (i.e., steel, aluminum, titanium, magnesium) have strength/density
ratios of 26–27 MN m kg1. Reinforcing aluminum with 25% of particulate silicon carbide increases this to nearly 40 MN m kg1;
reinforcing steel with 25% titanium di-boride gives an increase to around 35 MN m kg1. Increases in the specific properties can be
exploited through weight savings and/or improved fatigue resistance, Akhtar (2014).
The main benefits exhibited by MMCs, i.e., lower density, increased specific strength and stiffness, increased high-temperature
performance limits, and improved wear-abrasion resistance, are dependent on the properties of the matrix alloy and of the
reinforcing phase. The selection of the matrix is empirical based, using readily available alloys, and the major consideration is the
nature of the reinforcing phase.
When selecting materials for a MMC, it is critical that the materials selected for matrix and reinforcement are evaluated on their
individual characteristics, as well as on their interactivity with each other during the forming process. The interaction between
reinforcement material and matrix material, quite often, imposes restrictions on the type of materials that can be intermixed
together to produce a MMC.

Classification of MMCs based on the matrix material


MMCs are classified into different categories depending upon their matrix materials. Some examples of most commonly used
metallic matrix configurations are:

• Aluminum-based composites; aluminum as matrix can be either cast alloy or wrought alloy (i.e., AlMgSi, AlMg, AlCuSiMn,
AlZnMgCu, AlCu, AlSiCuMg)
• Magnesium-based composites
• Titanium-based composites
• Copper-based composites
• Super alloy-based composites
Metal–Matrix Composites 5

Figure 5 shows the usage volume of different matrix materials in MMCs. As seen, aluminum is the most commonly sued matrix
material in MMCs.

Figure 5 Usage of matrix materials in MMCs, Adebisi et al. (2011).

Aluminum–matrix composites are most commonly studied MMC as they are widely used in the automotive and aerospace
industries. Reinforcement compounds such as SiC, Al2O3, and B4C can be mixed easily and effectively in molten aluminum.
Magnesium–matrix composites have similar advantages, but due to limitations in fabrication and lower thermal conductivity, they
are not widely used as compared with aluminum-based MMCs. Magnesium–matrix composites have been developed for the space
industry thank to the low density of magnesium and its alloys. Titanium alloys are used as matrix material in fabricating MMCs
due to their good strength at elevated temperatures and excellent corrosion resistance. Compared with aluminum, titanium alloys
hold their strength at higher temperature, which is advantageous in manufacturing aircraft and missile structures, whose operating
speeds are very high. However, their main problem lies with processing of highly reactive titanium with reinforcement materials.
Fiber-based titanium composites are widely used in developing aircraft structures. In terms of thermal conductivity and high-
temperature strength properties, copper–matrix composites are superior compared with other MMCs. Super alloys are commonly
used as candidate materials for manufacturing gas turbine blades, where they operate at higher temperatures and speeds.

Aluminum-based MMCs
Aluminum-based MMCs are the most commonly used MMC in the automotive and aerospace applications. This is mainly due to its
unique properties like greater strength, improved stiffness, reduced density, improved temperature properties, controlled thermal
expansion and improved wear resistance. In aluminum-based MMCs, one of the constituent is an aluminum/aluminum alloy (i.e.,
Al–Si, Al–Cu, Al–Si–Mg alloys), which forms percolating network and is termed as matrix phase. The other constituent is embedded
in this aluminum/aluminum alloy matrix and serves as reinforcement, which is usually non-metallic and common ceramic such as
SiC, Al2O3, C, B, B4C, AlN, and BN. Figure 6 shows the microstructure of a SiC-particle reinforced aluminum–matrix composite
material. Properties of aluminum-based MMCs can be tailored by varying the nature of constituents and their volume fraction.

Figure 6 SiC-particle reinforced aluminum-matrix composite material (SiC-volume content: 70%), Sorensen (2011).

Aluminum-based MMCs are intended to substitute monolithic materials including aluminum alloys, ferrous alloys, titanium
alloys and polymer based composites in several applications. The major advantages of aluminum-based MMCs compared to
unreinforced materials are as follows:
6 Metal–Matrix Composites

• Greater strength
• Improved stiffness
• Reduced density (weight)
• Improved high-temperature properties
• Controlled thermal expansion coefficient
• Thermal/heat management
• Enhanced and tailored electrical performance
• Improved abrasion and wear resistance
• Control of mass (especially in reciprocating applications)
• Improved damping capabilities.

These advantages can be quantified for better appreciation. For example, elastic modulus of pure aluminum can be enhanced from
70 GPa to 240 GPa by reinforcing with 60 vol% continuous aluminum fiber. On the other hand, incorporation of 60 vol%
alumina fiber in pure aluminum leads to decrease in the coefficient of thermal expansion from 24 ppm/1C to 7 ppm/1C. Similarly
it is possible to process Al–9% Si–20 vol% SiCp composites having wear resistance equivalent or better than that of gray cast iron.
All these examples illustrate that it is possible to alter several technological properties of aluminum/aluminum alloy by more than
two to three orders of magnitude by incorporating appropriate reinforcement in suitable volume fraction.
Figure 7 superimposes curves for the three primary aluminum-based MMCs, ceramic particulate, boron filament, and graphite
tow reinforcements, on the curve for conventional 6061-T6. Although these curves are somewhat idealized, the two uniaxially
reinforced continuous fiber composites are markedly stiffer and stronger but also show very low ductility-to-fracture. The DWA1
20, ceramic particulate-reinforced composite (25 vol% 6061-T6) shows a more conventional stress–strain response (exhibiting
ductility), but shows mechanical properties intermediate between those of the unreinforced material and those of the continuous
fiber composites. The exact positions and shapes of the composite stress–strain curves can be varied by matrix heat treatment,
reinforcement volume percent (vol%), or reinforcement orientation effects. It must also be kept in mind that all continuously
reinforced composite material exhibits anisotropic characteristics. In Figure 7, if the transverse direction stress–strain curves for the
boron–aluminum material were to be included, it would be relegated to the extreme lower left-hand corner of the chart: strength
152 MPa, modulus B145 GPa, erB0.4%. Similarly, the graphite–aluminum composite transverse stress–strain response shows
even lower mechanical characteristics.

Figure 7 Comparative stress–strain curves of a 6061 aluminum alloy and the aluminum-based composites, Chawla (2012).

Over the years, aluminum–matrix composites have been used in numerous structural, non-structural and functional appli-
cations in different engineering sectors. Driving force for the utilization of aluminum–matrix composites in these sectors include
performance, economic and environmental benefits. The key benefits of aluminum-based MMCs composites in transportation
Metal–Matrix Composites 7

sector are lower fuel consumption, less noise and lower airborne emissions. With increasing strict environmental regulations and
emphasis on improved fuel economy, use of aluminum-based MMCs in transportation sector will be inevitable and desirable in
the coming years, Capel et al. (2000). The aluminum-based MMCs are very attractive for their isotropic mechanical properties
(higher than their unreinforced alloys) and their low costs (cheap processing routes and low prices of some of the discontinuous
reinforcement such as SiC particles or Al2O3 short fibers).

Magnesium-based MMCs
The increasing demand for lightweight and high-performance materials is likely to increase the need for magnesium-based MMCs.
The MMCs based on magnesium alloys, in particular Mg–Al system, are excellent candidates for engineering light structure
materials, and have great potential in civic, military and aerospace applications. The potential applications of magnesium–matrix
composites in the automotive industry include their use in disk rotors, piston ring grooves, gears, gearbox bearings, connecting
rods, and shift forks. One of the drawbacks of magnesium MMC is higher production costs due to more complicated manu-
facturing processes. To address this, the use of low cost materials, alloys and reinforcements can provide room for this class of low
density materials into the market.
The development of a wide range of reinforcing materials and new processing techniques are among top interests in high-
performance magnesium materials. The magnesium-based MMCs uni-directionally reinforced with continuous carbon fiber can
readily show a bending strength of 1000 MPa with a density as low as 1.8 g cm3 (Capel et al., 2000; Ottinger et al., 1995;
Hausmann et al., 1998). The superior mechanical property can be retained at elevated temperatures of up to 350–400 1C (Kagawa
and Nakata, 1992; Diwanji and Hall, 1992; Ottinger et al., 1997). In some magnesium alloys, formation of a composite may be the
only effective approach to strengthening. Mg–Li binary alloys with eutectic composition, for instance, are composed of HCP (a)
and BCC (b) solid solution phases. The dissolution of Li into Mg causes a partial solid solution strengthening effect without the
formation of any Mg–Li precipitates during the cooling process, Ye and Liu (2004). Therefore, heat treatment is not an effective
way to improve mechanical properties of these alloys. Considering this, the incorporation of thermally stable reinforcements into
composite materials makes them applicable for elevated-temperature applications.
Mg–Al alloys such as AM60 and AZ91 are presently the most prevalent magnesium alloys utilized in the automotive industry.
They are also the most widely studied matrix for magnesium-based composites. Other magnesium materials, such as pure
magnesium, Mg–Li alloy, and Mg–Ag–Re (QE22) alloys, have also been employed as a matrix material, though less frequently.
Ceramic particles are the most widely studied reinforcement for magnesium–matrix composites. Some common properties of
ceramic materials make them desirable for reinforcements. These properties include low density and high levels of hardness,
strength, elastic modulus, and thermal stability. However, they also have some common limitations such as low wettability, low
ductility, and low compatibility with a magnesium matrix. Among the various ceramic reinforcements, SiC is the most popular
one because of its relatively high wettability and its stability in magnesium melt, as compared to other ceramics.
Figures 8 and 9 show optical image of the as-cast AZ91D-based composite with 50 vol% Mg2B2O5 whiskers and scanning
electron micrograph of AZ91 based composite reinforced with SiC particles, respectively.

Figure 8 Optical image of the as-cast AZ91D-based composite with 50 vol% Mg2B2O5 whiskers, Chen et al. (2010).

Copper-based MMCs
When Al2O3 particles are dispersed in copper matrix, unique characteristics can be achieved (i.e., high thermal and electrical
conductivity, as well as high strength and excellent resistance to annealing), Motta et al. (2001). The applications encompass
resistance welding electrodes, lead frames and electrical connectors, Nishi et al. (1998).
The materials for electronic packaging and thermal management applications should have compatible coefficients of thermal
expansion (CTE) with those of semiconductors or ceramic substrates, high thermal conduction and excellent mechanical prop-
erties. Due to the high thermal conductivity of copper and low CTE of SiC, CuSiC MMC can be made to serve as a good solution
for thermal management. Semiconductors and ceramics have CTE in the range of 2–7 ppm/1C. Traditional low-CTE materials like
copper/tungsten (Cu/W), copper/molybdenum (Cu/Mo), copper–Invar–copper (Cu/I/Cu) and copper–molybdenum–copper
(Cu/Mo/Cu) have high densities and thermal conductivities that are little or no better than that of aluminum; Bukhari et al.
(2011). Density, thermal conductivity and CTE concerns mentioned above can be copped by using copper silicon carbide (CuSiC)
8 Metal–Matrix Composites

Figure 9 Scanning electron micrograph of AZ91 metal matrix composite reinforced with SiC particles, Trojanová et al. (2011).

based MMCs. Copper silicon carbide composites provide a good compromise between thermo-mechanical properties and high
conductivity. They have lower density than copper, very good thermal conductivity, low CTE and good machinability. A CuSiC
based MMC heat spreader will offer high thermal conductivity between 250 W/m K and 325 W/m K and corresponding adjustable
CTE between 8.0 ppm/1C and 12.5 ppm/1C; Zweben and Schmidt (1989). However, the primary challenge of CuSiC manu-
facturing is to prevent reaction between copper and silicon carbide during high-temperature densification, which dramatically
degraded the thermal conductivity.

Titanium-based MMCs
Ti-MMCs reinforced by continuous SiC fiber are being developed for aerospace applications in several countries, including the
USA, UK, France, and China, Peng (2005). Ti-MMCs provide outstanding mixture of stiffness, specific strength, fatigue and creep
resistance at elevated temperatures, Hooker and Doorbar (2003). Owning to the active nature of Ti, several solid-state processing
techniques have been developed to date for manufacturing Ti-MMCs, Guo and Derby (1995); Guo (1998), including the foil-fiber-
foil (FFF) method, matrix-coated mono-tape (MCM) method and the matrix-coated fiber (MCF) method. A maximum fiber
volume fraction up to 80% has been achieved with very uniform fiber distribution in MCF. The research on the MCF method has
been mainly concentrated on the consolidation behavior of MCFs with the aim to optimize the processing parameters. Table 2
compares physical and mechanical properties of a Ti alloy and a Ti-MMC. Figure 10 shows the microstructure of an SiC reinforced
titanium matrix composite.

Table 2 Physical and mechanical properties of a Ti alloy and a Ti-MMC

Property Titanium (Ti-6-2-4-2) Ti-MMC

Density, gr/cm3 4.54 3.93


Tensile strength (longitudinal), MPa 931 1689
Young's modulus (longitudinal), GPa 117 200
Tensile strength (transverse), MPa 931 400
Young's modulus (transverse), GPa 117 145
Compression strength, MPa 931 44481
High cycle fatigue (longitudinal), R ¼0.1 and 30 Hz 107 cycles @ 482 MPa 107 cycles @ 531 MPa
Tensile strength at 315.5 1C, MPa 552 1379

Super alloy-based MMCs


The primary application for super alloy matrix composites is gas turbine blades. By enhancing the material operating temperatures
and stresses of turbine blades, an increase in the performance and a reduction in the operating cost can be achieved. Several
composites have been chosen for developments of the next generation of turbine blade materials. Refractory alloy wire reinforced
super alloys have been investigated, for instance, for elevated-temperature applications. Solid state diffusion bonding and liquid
phase infiltration techniques have been used to produce composite samples. Mono-layers tape has been produced using tech-
niques and equipment similar to that used for Ti–matrix composites.
Metal–Matrix Composites 9

Figure 10 Microstructures of a SiC reinforced titanium matrix composite, Poletti et al. (2008).

In the field of super alloy matrix composites, there have been several recent extensive reviews of progress in the development of
Nb-silicide-based composites (Figure 11 shows the microstructure of a Nb-silicide-based composites). In the elevated-temperature
applications, for the next generation of jet engines, Nb- and Mo-silicide-based composites are among the best candidates, Bewlay
et al. (2003). The melting points of the silicide containing composites based on these systems are in excess of 1750 1C. Densities of
the Nb-silicide-based composites are in the range of 6.6–7.2 g/cc The ambient-temperature fracture toughness of Nb-silicide-based
composite systems has been reported as being above 20 MPa m1/2 (Bewlay et al., 2003) while values for the preliminary creep and
oxidation properties indicate that Nb-silicide-based composites could, with further development, be integrated into blade designs
with substantial payoffs in weight and cooling-air savings, relative to fourth generation super alloy blade designs.

Figure 11 Backscattered electron image of a Nb-silicide composite (a two-phase composite of Nb and Nb3Si), Bewlay et al. (2003).

Oxide (ceramic)-fiber-reinforced super alloys are another important class of composites being studied for high-temperature
applications. Single crystal aluminum-oxide super alloys based composites are produced commercially for research activity pur-
poses. Productions rates are relatively slow and fiber unit cost is quite high and composite properties suitable for industrial
applications have yet to be demonstrated.

Reinforcement
Reinforcement phase is stronger than the matrix, and it is why it is called reinforcing phase. Reinforcement (reinforcing phase) is
generally supposed to possess following characteristics (Chawala, 2012):

• Low density
• Good mechanical and chemical compatibility
• Good thermal stability
• High Young's modulus
• High compression and tensile strength
• Good process ability
• Economic efficiency

Generally, non-metal inorganics components (i.e., ceramic particles, carbon fibers) can provide acceptable combinations of
above properties. Continuous, aligned fibers are the most efficient reinforcement form and are widely used, especially in high-
performance applications. However, for ease of fabrication and to achieve specific properties, such as improved through-thickness
strength, continuous fibers are converted into a wide variety of reinforcement forms using textile technology. Most of the fibrous
reinforcements are made of brittle ceramics or carbon, i.e., glass fiber, carbon (graphite) fiber, boron fibers, silicon carbide fibers,
10 Metal–Matrix Composites

alumina fiber, aramid fibers, and high density polyethylene fibers. Figure 12 shows most commonly used reinforced materials and
their usage volume in MMC.

Figure 12 Main reinforced materials and their usage volume in metal–matrix composites, Adebisi et al. (2011).

On the basis of reinforcement, composites are classified to (Figure 13):

Figure 13 Classification of composites based on reinforcement phases, Singh et al. (2014).

• Particulate Composites
Particulate Composites consist of a matrix reinforced by a dispersed phase in form of particles.
(i) Composites with random orientation of particles.
(ii) Composites with preferred orientation of particles; dispersed phase of these materials consists of two-dimensional flat
platelets (flakes), laid parallel to each other.
• Fibrous composites
(i) Short-fiber reinforced composites; these composites consist of a matrix reinforced by a dispersed phase in form of
discontinuous fibers (length o 100  diameter). Short-fiber reinforced composites are further classified to randomly
oriented fibers and preferably oriented fibers.
(ii) Long-fiber reinforced composites; these composites consist of a matrix reinforced by a dispersed phase in form of
continuous fibers. Long-fiber reinforced composites are further classified to unidirectional oriented fibers and bidirectional
oriented of fibers (i.e., woven).
• laminated Composites
Metal–Matrix Composites 11

In laminar composites the layers of reinforcement are stacked in a specific pattern to obtain required properties in the resulting
composite piece. These layers are called plies or laminates. Laminates can be composed of reinforcement material which may be
non-woven, braided, fiber reinforced, matt, 2D-woven, 3D-woven, and unidirectional fibers (Figure 14).

Figure 14 Various configurations of fiber reinforcements, Singha and Singha (2012).

MMCs are generally distinguished by characteristics of the reinforcement. The reinforcing phase can be fibrous, plate like, or
equiaxed (having equal dimensions in all directions), and its size can also vary widely, from about 0.1 mm to more than 100 mm.
In particular, reinforcements used MMCs fall in five categories (Figure 15): continuous fibers, short fibers, whiskers, equiaxed
particles, and interconnected networks. Table 3 displays typical reinforcements used in MMCs.

Figure 15 Types of reinforcement fibers used in MMCs (A) Continuous fibers (B) discontinuous fibers (C) whiskers (D) particulates.

Table 3 Typical reinforcements used in metal–matrix composites

Type Aspect ratio Diameter Examples

Particle 1–4 1–25 mm Sic, Al2O3, BN, WC


Short fiber (whisker) 10–10000 1–5 mm C, SiC, Al2O3, SiO2 þ Al2O3
Continuous fiber 41000 3–150 mm SiC, Al2O3, C, B, W, Nb þ Ti, Nb3Sn
Nanoparticle 1–4 o100 nm C, Al2O3, SiC
Nanotube 41000 o100 nm C

Source: Chawla, 2012.

Developing Trends and Design Consideration of MMCs


The need for developing MMCs for use in high-performance structural and functional applications, i.e., aerospace and defense
industries, automotive and transportation sectors have significantly raised recently. This trend is expected to increase steadily in the
coming years (Figure 16). The need to develop new materials with combinations of low density, improved stiffness and high
12 Metal–Matrix Composites

strength, in order to overcome the limitations of existing alloys, improves the design procedures and results in improvements in
overall efficiency, reliability and performance. This can be achieved by reducing either absolute weight or increases in strength-to-
weight ratio. For this purpose, materials of high ultimate tensile strength, high stiffness, high yield strength and low density are
required.

Figure 16 The global MMC market (USDOC, USCAR, Composite World, Composite UK, Transparency Market Research), Researchmoz Global Pvt.
Ltd. All the products and end-user segments have been analyzed based on present and future trends and the market has been estimated from
2012 to 2019. Regional segmentation includes the current and forecast demand for North America, Europe, Asia Pacific and Rest of the world.

By proper selection of MMCs, as multiphase materials, it becomes possible to meet the challenging needs of specific designs.
Within broad limits, for instance, it is possible to specify mechanical characteristics, i.e., strength and stiffness in one direction, and
physical properties, i.e., CTE in another direction (this is hardly achievable in monolithic (isotropic) materials). Among different
classes of MMCs, particulate-reinforced MMCs tend to be isotropic (like metals). However, their ductility and fracture toughness
are significantly less than monolithic materials due to presence of brittle reinforcements and oxides. The properties of composites
reinforced with whiskers depend strongly on their orientation. Randomly oriented whiskers produce an isotropic material.
However, processes like extrusion will re-orient whiskers resulting in anisotropic properties. Whiskers also reduce ductility and
fracture toughness.
MMCs reinforced with aligned fibers have anisotropic properties. They are stronger and stiffer in the direction of the fibers than
perpendicular to them. The transverse strength and stiffness of unidirectional MMCs, however, are frequently great enough for use
in components such as stiffeners and struts. This is one of the major advantages of MMCs over PMCs, which can rarely be used
without transverse reinforcement. As the modulus and strength of metal matrixes are significant with respect to those of most
reinforcing fibers, their contribution to composite behavior is important. The stress–strain curves of MMCs often show significant
nonlinearity resulting from yielding of the matrix.
Another factor that has a significant effect on the behavior of fiber-reinforced metals is the frequently large difference in CTE
between the two constituents. This can cause large residual stresses in composites when they are subjected to significant tem-
perature changes. In fact, during cool down from processing temperatures, matrix thermal stresses are often severe enough to cause
yielding. Large residual stresses can also be produced by mechanical loading.
Although fibrous MMCs may have stress–strain curves displaying some nonlinearity, they are essentially brittle materials, as are
PMCs. In the absence of ductility to reduce stress concentrations, joint design becomes a critical design consideration. Numerous
methods of joining MMCs have been developed, including metallurgical and polymeric bonding and mechanical fasteners.

Microstructure of MMCs
The microstructure of a MMC comprises the structure of matrix and reinforcement. The key features in the microstructure of a
composite material, resulting from the interaction between the matrix and the reinforcement. This usually includes type, size, and
distribution of secondary reinforcing phases, matrix grain size, matrix and secondary phase interfacial characteristics, and
microstructural defects. The mechanical properties of the composite materials are also strongly influenced by these factors.
Metal–Matrix Composites 13

Figure 17 shows typical microstructures of different particle reinforced aluminum-based composite materials. As seen,
depending on the processing methods, reinforced particle distributions vary. The option of combining particles and fibers to make
a hybrid-reinforced composite with the effects of both reinforcement components is shown in Figure 18.
In multi-filament-strengthened composite materials (Figure 19) the fiber/fiber contacts and non-reinforced areas are recog-
nizable as a result of the infiltration of fiber bunch preforms. Structural defects, like fiber/fiber contacts, pores and non-reinforced
areas are visible, which have a substantial influence on the composite characteristics. Figure 20 shows the optical structure of a SiC
monofilament/Ti composite material.

Figure 17 Arrangement of typical structures of different particle reinforced composites, Kainer (2006); (a) SiC-particle reinforced aluminum (mold
cast), (b) SiC-particle reinforced aluminum (die cast), (c) SiC-particle reinforced aluminum (extruded powder mixture), (d) SiC-particle reinforced
aluminum (cast and extruded).

Figure 18 Structure of formation of hybrid reinforced light metal composite materials with C short fibers and Mg2Si particles, Kainer (2006).

Physical and Mechanical Properties


Composite properties depend first and foremost on the nature of the composite. Composites are classified as complex, hetero-
geneous, and often anisotropic material systems. Their properties are affected by many variables including reinforcement form,
volume fraction, geometry, distribution, matrix/reinforcement interface, void content, and manufacturing process. Hereby,
representative mechanical and physical properties of MMCs at ambient temperature (i.e., room temperature) are presented for a
broad range of mechanical engineering applications.

Mechanical properties of MMCs


By reinforcing metals and alloys with continuous/discontinuous fibers, whiskers, and particles, higher strength and stiffness as well
as better wear resistance can be achieved.
Figure 21 compares the tensile curves for a typical reinforced material, and a conceptual fiber-reinforced MMC as well as a
matrix material (i.e., a metal or alloy). Adding hard, brittle reinforcement particle to a ductile metal matrix can strengthen the
14 Metal–Matrix Composites

Figure 19 Structure of a unidirectional endless fiber reinforced aluminum composite material (transverse grinding): matrix: AA 1085, 52 vol% 15-
mm Altex-fiber (Al2O3), Kainer (2006).

Figure 20 Structure of a titan matrix composite material of SiC monofilaments, World of MMC asses, http://mmc-assess.tuwien.ac.at/mmc.

Figure 21 Schematic representation of tensile curves of matrix and reinforcement as compared with composite material, Ashby (2005).

metal significantly. When tensile stress is applied parallel to the fiber direction in a continuous fiber MMC, both fiber and matrix
deform elastically. By keep applying higher stresses, the matrix undergoes plastic deformation while the fibers are still in the elastic
regime. If the fibers possess some ductility, both matrix and fibers deform plastically; however, fibers break without experiencing
plastic deformation. As the fibers fracture, the load is transferred back to the weaker matrix, and the composite fails immediately.
Metal–Matrix Composites 15

The MMCs exhibit higher strength because they are able to transfer much of the loads to the strong reinforcement particles,
reducing the stress carried by the matrix. The fiber–matrix interphase properties must be carefully tailored and maintained over the
life of the composite to obtain the desirable behaviors. The strength of MMCs depends upon a much more complex manner on
composite microstructure.
In this section the main mechanical properties of MMCs including modulus of elasticity, elongation, strength, fatigue, and
creep are briefly discussed.

• Modulus of elasticity

The stiffness, which is proportional to the elastic modulus, of a MMC increases with introducing of reinforcement. Type and
volume fraction of reinforcement are the main factors contributing the elastic behavior of a composite. For instance, in a
unidirectional reinforced continuous fiber MMC, the longitudinal Young's modulus increases linearly as a function of the fiber
volume fraction according to:

Ec ¼ Vr Er þ Vm Em

where E is the elastic modulus, V is the volume fraction and c, m, and r are indexes for the composite, matrix, and reinforcement,
respectively.
Figure 22 shows an example of modulus increase as a function of fiber volume fraction for an alumina fiber-reinforced
aluminum  lithium alloy matrix composite, Champion et al. (1978). The increase in the longitudinal Young's modulus is in
agreement with the rule-of-mixtures value, whereas the modulus increase in a direction transverse to the fibers is much lower.
Particle reinforcement also results in an increase in the modulus of the composite; the increase, however, is much less than that
predicted by the rule-of-mixtures. This is understandable inasmuch as the rule of mixtures is valid only for continuous fiber
reinforcement. The relatively high cost of many continuous reinforcing fibers used in MMCs has limited the application of these
materials.

Figure 22 Modulus increase as a function of fiber volume fraction Vf for alumina fiber-reinforced aluminum  lithium alloy matrix for (a) E
(elastic modulus), and (b) smax, Champion et al. (1978).

The most widely used MMCs are reinforced with discontinuous fibers or particles. For discontinuously reinforced MMCs:

Em ð1 þ 2sqVr Þ
Ec ¼
1  qVr

where

ðEr =Em  1Þ

ðEr =Em Þ þ 2s

where s is the particle aspect ratio.


Figure 23 shows materials property (Ashby) plot of Young's modulus vs. density for different classes of materials including
composites.

• Strength
The introduction of reinforcements to matrix alloy generally enhances both yield and ultimate tensile strength. The strength
of a fiber-reinforced composite is determined by fracture processes, themselves governed by a combination of microstructural
phenomena and features. These include plastic deformation of the matrix, the presence of brittle phases in the matrix, the
16 Metal–Matrix Composites

Figure 23 Young's modulus versus density for different classes of materials, Ashby (2005).

strength of the interface, the distribution of flaws in the reinforcement, and the distribution of the reinforcement within the
composite. Consequently, predicting the strength of the composite from that of its constituent phases is generally difficult. In
some cases, the improvements are dramatic. The greatest increases in strength and modulus are achieved with continuous
fibers. Figure 24 shows Ashby plot, strength vs. density for different classes of materials including composites.
• Elongation
Elongation of the MMCs is generally less than matrix alloy. For instance, the elongation of a 6061-T6 aluminum alloy is
about 12%; by adding 15 vol% Al2O3 particles to this alloy, the elongation drops to 5.4%. This is mainly due to the fact that
the composite failure is associated with particle cracking and void formation in the matrix within the particle clusters. Also, in
MMCs, the strain in the matrix is constrained with surrounding reinforcement phases (i.e., particles, fibers). Size, type,
distribution of particles, matrix microstructure, and nature of matrix-reinforcement interface are the main contributing factors
in MMCs' ductility.
• Fatigue
Fatigue life of composites is significantly increased by adding the reinforcements. The range of the stress intensity factor, DK,
for Al–SiCp is between 2 and 4 MPa m1/2 which is twice that of unreinforced aluminum (1–2 MPa m1/2). The fatigue resistance
of long-fiber reinforced MMCs is larger than that of unreinforced metals when loaded in tension along the fiber axis. Crack
deflection and reduction in slip band formation are the main contributing factors to the superior fatigue properties of MMCs.
In terms of general S  N curve behavior, a composite such as a silicon carbide particle-reinforced aluminum-based MMC
shows an improved fatigue behavior compared to the reinforced alloy, Figure 25; Chawla et al. (1998). Such an improvement in
stress controlled cyclic loading or high cyclic fatigue is attributed to the higher stiffness of the composite. However, the fatigue
behavior of the composite, evaluated in terms of strain amplitude vs. cycles or low cycle fatigue, was found to be inferior to that
of the unreinforced alloy; Bonnen et al. (1990). This was attributed to the lower ductility of the composite compared to the
unreinforced alloy. At elevated temperatures, particularly close to the aging temperature for age hardenable alloys, the matrix
strength decreases resulting in a decrease in fatigue strength, Figure 26; Chawla et al. (1999).
• Fracture toughness
Fracture toughness of particulate composites generally decreases with increase in particle size at a given volume fraction and
inter-particle spacing. Fracture toughness decreases with increasing tensile strength for cast composites, whereas it increases by
70–100% for forged ones over that of cast ones. This increase in toughness with increasing amount of particulates is related to
blunting of the original crack tip caused by the void nucleated over the particles. While the crack tip radius increases, the stress
Metal–Matrix Composites 17

Figure 24 Strength versus density for different classes of materials, Ashby (2005).

Figure 25 Stress versus cycles (S–N) behavior of a particle reinforced MMC. With increasing volume fraction of particles the fatigue strength of
the composite increases, Chawla et al. (1998).

concentration at crack tip decreases resulting in combined increase in facture toughness values. Unidirectional fibers
reinforcement can lead to easy crack initiation and propagation vis-à-vis the unreinforced alloy matrix. Braiding of fibers can
make the crack propagation toughness increase largely due to extensive matrix deformation, fiber bundle deboning, and
pull out.
Figure 27 shows Ashby plot of fracture toughness vs. Young's modulus for different classes of materials including composites.
• Creep
It has been reported that the creep resistance of Al9Si3Cu/25 vol% Al2O3 composites at 400 1C increased by 100% compared
to that of unreinforced alloys. The dependence of creep rate on both stress level and temperature is more in the composites
compared to the matrix metal/alloy. The whisker-reinforced composites are more creep resistance than particulate composites.
18 Metal–Matrix Composites

Figure 26 Elevated-temperature fatigue behavior of a particle reinforced MMC. With increasing temperature, the matrix strength decreases
resulting in decreased fatigue resistance (aging temperature was 175 1C), Chawla et al. (1999).

Figure 27 Fracture toughness vs. Young's modulus for different materials including composites, Ashby (2005).

The addition of short-fiber reinforcement enhances the creep strength due to effective load transfer to the fibers; Bhatnagar and
Srivatsan (2009).

Whisker or particle reinforcement also results in significant creep strengthening over the unreinforced alloy, Figure 28; Nardone
and Strife (1987). Anomalously high values of the activation energy, Q, and stress exponent, n, have been reported in MMCs;
Chawla and Chawla (2013); Nardone and Strife (1987). Nardone and Strife (1987) rationalized this by proposing the concept of a
threshold stress sr, for creep deformation, originally used to explain the high values for Q and n in dispersion-strengthened alloys.
The physical explanation for the threshold stress in the discontinuously reinforced composite system can be attributed to a variety
of reasons, Russell and Lee (2005): (1) Orowan bowing between particles, (2) back-stress associated with dislocation climb, (3)
Metal–Matrix Composites 19

Figure 28 Steady-state creep rate as a function of applied stress for SiC-particle and SiC whisker-reinforced Al–matrix composites, Nardone and
Strife (1987).

attractive force between dislocations and particles, resulting from relaxation of the strain field of dislocations at the particle/matrix
interface.

Physical Properties of MMCs


Density, CTE, and thermal conductivity are among most commonly discussed physical properties of composites. Composites are
considered multi-functional materials; among the family of composites, there are an increasing number of material systems than
combine superior physical properties including high thermal conductivity along with low density and adaptable CTE as well as
excellent mechanical properties.
Physical properties of matrix can be significantly altered by addition of a reinforcement and chiefly depend on the reinfor-
cement distribution. A good example is aluminum–silicon carbide composites, for which the presence of the ceramic increases,
substantially, the elastic modulus of the metal without greatly affecting its density. Elastic moduli for 6061 aluminum–matrix
composites reinforced with discrete silicon carbide particles or whiskers have been calculated by using the rule of mixtures for the
same matrix reinforced with two types of commercial continuous silicon carbide fibers. As a result, several general facts become
apparent; first, modulus improvements are significant, even with equiaxed silicon carbide particles, which are far less expensive
than fibers or whiskers. However, the level of improvement depends on the shape and alignment of the silicon carbide. Also, it
depends on the processing of the reinforcement: for the same reinforcement shape (continuous fibers), microcrystalline
polycarbosilane-derived SiC fibers yield much lower improvements than do crystalline polycarbosilane SiC fibers. These features,
which influence reinforcement shape, orientation, and processing of modules, are quite general; they are also observed, for
example, in MMC reinforced with aluminum oxide or carbon.
Table 1 includes physical properties of a variety of unidirectional composites reinforced with carbon fibers, along with those of
monolithic copper and 6063 aluminum alloy for comparison. As seen, the specific axial thermal conductivities of the composites
are significantly greater than those of aluminum and copper.

Strengthening Methods in MMCs


Dislocation strengthening
Dislocations' density is higher in the composite matrix than in unreinforced metal with the same history. In a composite, dislocations
can be generated by (1) straining in response to an applied load, (2) straining to relax residual thermal stresses caused by CTE
mismatches between matrix and reinforcement. The increase in dislocations' density can be written as, Russell and Lee (2005):

DaDTf
Ds ¼ 12
bd

where Δr is the increase in dislocation density, Δa is the CTE mismatch, ΔT is the temperature difference, b is Burgers vector, f is the
reinforcement volume fraction, and d is particle size. From this equation, the change in matrix yield strength can be estimated:

pffiffiffiffiffiffi
Ds ¼ Gb Dr
20 Metal–Matrix Composites

Therefore the dislocation density and hence matrix strengthening increase with decreasing reinforcement particle size and increasing
reinforcement volume fraction.

Orowan strengthening
This mechanism strengthens the composite by making subsequent dislocation motion more difficult. Although important in
dispersion-strengthened systems, Orowan mechanism does not effectively strengthen other MMCs because in other MMCs the
reinforcement phases (particles) are too large and to too far to be considered as effective obstacle to the motion of dislocations.
However, Orowan strengthening mechanism can be significant in discontinuously reinforced aluminum composites where the
particles are finer and closer together and causes increase in the strength of the MMC.

Grain refinement strengthening


The MMCs usually possess much finer grain size than that of unreinforced matrix metal. Hall–Petch relationship implies that the
beginning of plastic flow in the matrix depends on the local magnification of stress at grain boundaries resulting from dislocation
pile-ups. A small grain size results in fewer dislocations in the pile-ups, and hence the applied stress must be higher to cause
yielding.

Fabrication of MMCs
For a given MMC application, the processing (fabrication) method is a key factor that determines both cost and properties. The
main challenges in fabricating MMCs include (1) finding a cost effective method to distribute the reinforcement phase in the
reinforcement phase in the desired configuration, (2) achieving a strong bond between the matrix and the reinforcement to allow
an effective load transfer between phases without any failure. The MMCs can be produced by a variety of fabrication techniques
including solid state processes, liquid-state processes and deposition processes.
In solid-state processes, the most spread method is powder metallurgy; it is usually used for high melting point matrixes and
avoids segregation effects and brittle reaction product formation prone to occur in liquid state processes. This method permits to
obtain discontinuously particle reinforced aluminum–matrix composites with the highest mechanical properties. The aluminum–
matrix composites are used for military applications but remain limited for large-scale productions. In liquid-state processes, one
can distinguish the infiltration processes where the reinforcements form a preform which is infiltrated by the alloy melt (1) with
pressure applied by a piston (squeeze-casting) or by an inert gas (gas pressure infiltration) and (2) without pressure. MMC
fabrication under no pressure can be subdivide to (1) reactive infiltration, in which the wetting between reinforcement and melt is
obtained by reactive atmosphere, elevated temperature, alloy modification or reinforcement coating and (2) the dispersion
processes, such as stir-casting, where the reinforcements are particles stirred into the liquid alloy. Process parameters and alloys are
to be adjusted to avoid reaction with particles. In deposition processes, droplets of molten metal are sprayed together with the
reinforcing phase and collected on a substrate where the metal solidification is completed. This technique has the main advantage
that the matrix microstructure exhibits very fine grain sizes and low segregation. This technique however has several drawbacks: the
technique can only be used with discontinuous reinforcements, the costs are high, and the products are limited to the simple
shapes that by obtained by extrusion, rolling or forging.
Another widely used fabrication method comprises spray processes. In this case, a molten metal stream is fragmented by means
of a high-speed cold inert-gas jet passing through a spray gun, and dispersoid powders are simultaneously injected. A stream of
molten droplets and dispersoid powders is directed toward a collector substrate where droplets recombine and solidify to form a
high-density deposit.
Depending on the process, the desired microstructure, and the desired part, MMCs can be produced to net or near-net shape; or
alternatively they can be produced as billet or ingot material for secondary shaping and processing. Figure 29, schematically, shows
different techniques for making MMCs.

Application of MMCs
The combined attributes of MMCs, together with the costs of fabrication, vary widely with the nature of the material, the
processing methods, and the quality of the product. In engineering, the type of composite used and its application vary sig-
nificantly, as do the attributes that drive the choice of MMCs in design. For example, high specific modulus, low cost, and high
weld-ability of extruded aluminum oxide particle-reinforced aluminum are the properties desirable for bicycle frames. High wear
resistance, low weight, low cost, improved high-temperature properties, and the possibility for incorporation in a larger part of
unreinforced aluminum are the considerations for design of diesel engine pistons. MMCs are being used in various applications
including, Chawla and Chawla (2013):

• Aerospace
• Transportation
• Electronics
Metal–Matrix Composites 21

Figure 29 Methods used to make MMCs (www.what-when-how.com).

• Electric power transmission


• Recreational products and sporting goods
• Wear-resistance materials
Figure 30 shows the usage of MMCs per year, broken down by market segment (i.e., automotive, aerospace, consumer
products). Note that the ground transportation industry is, by far, the major market segment for MMCs. Applications in this
segment include drive shafts, engine components, and brake components. The main constraint in the transportation sector is that
it is an extremely cost-sensitive sector. Thus, reducing the manufacturing costs of the MMC components will greatly aid in the
replacement of conventional parts.
The application of MMCs in aerospace industries is due to the fact that materials with enhanced specific stiffness and strength
can significantly increase the performance of the aircraft. The MMCs have been used largely in military and commercial aircrafts. In
the F-16 aircraft, for instance, aluminum access doors have been replaced with SiC-particle reinforced MMCs resulted in increased
fatigue life. Continuous fiber-reinforced MMCs have also been used in military applications, due to high specific strength, stiffness,
and fatigue resistance. SiC monofilament reinforced Ti–matrix composites have been used as nozzle actuator controls for the F119
engine in F-16. The MMC replaced a heavier Inconel 718 in the actuator links and stainless steel in the piston rods.
MMCs are being used in commercial aircraft as well. Figure 31 shows an application on the MMCs in the fan-exit guide vane of
a Pratt & Whitney engine on a Boing 777. The MMC replaced a Carbon/Epoxy composite that had problems with foreign object
damage (FOD). Boeing 787 is the first commercial jet transport to be manufactured out of predominantly composite materials.
Traditional metallic materials are characterized by the isotropic nature of their material properties (no preferred directions in terms
of tensile strength) and the fact that the predominant manufacturing method is based on material removal (milling). Composite
22 Metal–Matrix Composites

Figure 30 The use of MMCs in different market sectors, Business Communications Company (2006).

Figure 31 Application of a SiC-particle reinforced aluminum-based MMC in the fan-exit guide vane of a Pratt & Whitney engine on a Boeing 777,
Chawla and Chawla (2006).

materials, on the other hand, frequently are fibrous and anisotropic. Moreover, the predominant manufacturing methods are
based on material deposition rather than removal.
Boeing 787 makes greater use of composite materials in its airframe and primary structure than any previous Boeing com-
mercial airplane (Figure 32). Undertaking the design process without preconceived ideas enabled Boeing engineers to specify the
optimum material for specific applications throughout the airframe. The result is an airframe comprising nearly half carbon fiber

Figure 32 The percentages of materials used in Boing 787 aircraft, Holley (2013).
Metal–Matrix Composites 23

reinforced plastic and other composites. This approach offers weight savings on average of 20% compared to more conventional
aluminum designs.
MMCs have been used in a variety of automotive applications as well. An early MMC application in an automotive engine was
a hybrid particulate-reinforced aluminum–matrix composite used as a cylinder liner in the Honda Prelude (Figure 33). The
composite consisted of an Al–Si matrix with 12% Al2O3 for wear resistance and 9% carbon lubricity. The composite was integrally
cast with the engine block, had improved cooling efficiency, and exhibited improved wear and a 50% weight savings over cast iron,
without increasing the engine package size.

Figure 33 Hybrid particulate-reinforced aluminum-matrix composite used as a cylinder liner in the Honda Prelude; (a) Prelude engine block, (b)
magnified view of cylinder liner, and (c) microstructure of composite showing carbon short fibers (black) and Al2O3 fibers (dark gray), Chawla and
Chawla (2006).

The automotive market is a high volume and a high technology market, but costs should be as low as possible. However, there
are still a lot of reasons to consider the use of light aluminum composites:

• Reduction of the weight of engine parts;


• Increase of the operation temperature of engines;
• Improvement of the tribological properties of moving and contacting;
• Components (wear resistance, lubrication);
• Increase of stiffness and strength,
• Matching coefficient of thermal expansion (i.e., steel or cast iron in connection with aluminum alloys);
• The use of related manufacturing techniques (especially for discontinuously reinforced aluminum alloys).

A list of typical MMCs, in particular aluminum–matrix composites, for car applications is given in Table 4.
Other current applications are piston parts, cylinder liners and connecting rods. The first high volume application is the
successful aluminum Toyota-piston ring (Figure 34), reinforced with short Saffil fibers and produced by squeeze casting. Both

Table 4 Aluminum–matrix composites for car applications, Campbell (2012)

Material Application Improved Property Feature Manufacturer

Al-shore fiber Piston ring Abrasion resistance High-temperature engine Toyota


performance, lower cost
Al-shore fiber Piston combustion bowl High-temperature performance Improved durability Consortium
Al-shore fiber Selective reinforcement of High-temperature performance Improved durability Peugeot
motor block
Al-shore fiber Cylinder liner Improve stiffness, wear Improved durability Honda
resistance, better heat
conducting, closer tolerances
Al-SiC particles Connecting rod Specific strength, specific Higher performance Nissan
stiffness
Al-Al2O3 fiber Connecting rod Specific strength, specific Higher performance Chrysler
stiffness
24 Metal–Matrix Composites

weight saving and increasing wear resistance are the main reasons for the success. An important potential replacement of steel by
SiC-particle-reinforced aluminum–matrix composites is in the connecting rod (Figure 35). Near-net-shape sinter-forging was used
to fabricate MMCs connecting rods with tensile and fatigue properties comparable to those of extruded materials. Table 5 compares
the weight of the MMC connecting rod versus that of the steel rod. A 57% weight savings was achieved with the MMC rod, with a
moderate increase in cost.

Figure 34 Selectively Reinforced Automotive Piston.

Figure 35 Cross-section of a passenger car engine showing the location of the connecting rod, Chawla and Chawla (2006).

Figure 36 shows particulate MMCs for use in brake drums and brake rotors, as a replacement for cast iron. The high wear
resistance and thermal conductivity coupled with 50–60% weight savings make the MMCs quite attractive for this application.
Figure 37 shows another example of the MMCs in automotive applications which is a 6061/Al2O3/20 p composite used as
driveshaft in the Corvette. The composite exhibits a 36% increase in specific modulus over steel.
Metal–Matrix Composites 25

Table 5 Mass comparison of connecting rod for steel and an MMC, Chawla and Chawla
(2006)

2080/SiC/20 Steel

Pine weight (gr) 65.2 144.7


Crank weight (gr) 184.0 437.7
Total weight (gr) 249.2 582.4

Figure 36 Brake drums and brake rotors made of MMC, Chawla and Chawla (2013).

Figure 37 MMC made driveshaft in the Corvette with 36% increase in specific modulus over steel, Chawla and Chawla (2013).

Railroad brakes
The driving force for lightweight railway vehicles has also prompted the use of high-performance MMCs, Chawla and Chawla
(2013). A conventional brake system for a railway vehicle consists of four brake disks, calipers, hand brake, and electromagnetic
track brake. This makes up about 20% of the total weight of the bogey. A particulate-reinforced aluminum–matrix composite
(AlSi7Mg þ SiC particulates, supplied by Duralcan) is fabricated by a multi-pouring process, where alternating layers of the
unreinforced alloy and MMC is cast in successive layers (Figure 38). This contributed to reduced cost by using less MMC and
placing the composite in the strategically important region (i.e., in contact with the wearing surface). In this application, the weight
is reduced from 115 kg for a spheroidal graphite iron disk to 65 kg for the MMC disk, a weight savings of 43%. The steel brake
shows significant cracking, while the MMC brake is in relatively good condition.

Electronic and communication applications


New generation advanced integrated circuits are generating more heat than previous types. Therefore, the dissipation of heat
becomes a major concern in electrical applications. Indeed, thermal fatigue may occur due to a small mismatch of the CTE
between the silicon substrate and the heat sink (normally molybdenum). This problem can be solved by using the MMCs with
exactly matching coefficients (i.e., aluminum with boron or graphite fibers and aluminum with SiC particles). Besides a low CTE
and a high thermal conductivity, the aluminum-based MMCs also have a low density and a high elastic modulus. Figure 39 shows
cross-section of an electrical conductor for power transmission including alumina fiber (Nextel 610)/Al core strands, Al–Zr alloy
outer strands. The core consists of individual wires made from a continuously reinforced aluminum MMC produced by 3 M. The
MMC core supports the load for the 54 aluminum wires and also carries a significant current, unlike competing steel cores.
Hermetic package materials are developed to protect electronic circuits from moisture and other environmental hazards. These
packages have often glass-to-metal seals. Therefore, materials with an ‘adjustable’ CTE are required. Al-based MMCs are fulfilling
this condition, as the CTE is depending upon the volume fraction of the fibers or particles.
26 Metal–Matrix Composites

Figure 38 Brake rotors for German high speed train ICE-2 made from a particulate-reinforced aluminum alloy (AlSi7Mg þ SiC particulates,
supplied by Duralcan), developed by Knorr Bremse AG. Compared to conventional parts made out of cast iron with 115 kg/piece the 65 kg of the
MMC rotor offers an attractive weight saving potential: each wagon has 8 brake rotors and in combination with the reduction of un-damped
masses a highly efficient component is realized with MMC, Chawla and Chawla (2013); Chawla and Chawla (2006).

Figure 39 Aluminum conductor composite reinforced (ACCR), Campbell (2012).

Sports and leisure market applications


The already well known advantages of Al-based composites are leading to several applications in various leisure and sporting
goods. Typical applications are fishing rods, tennis and squash rackets, bicycle frames, golf club heads, track and field shoes, and
cheetah flex-foot (Figure 40).
Metal–Matrix Composites 27

Figure 40 MMCs in track shoes and Cheetah Flex-Foot Carbon Fiber Reinforced composite.

Table 6 summarizes potential and realistic technical applications of MMC in various industries along with their advantages.

Table 6 Potential and realistic technical applications of metal–matrix composites

Composite Components Advantages

Aluminum–silicon carbide (particle) Piston Reduced weight, high strength and wear resistance
Brake rotor, caliper, liner High wear resistance, reduced weight
Propeller shaft Reduced weight, high specific stiffness
Aluminum–silicon carbide (whiskers) Connecting rod Reduced reciprocating mass, high specific strength
and stiffness, low coefficient of thermal expansion
Sprockets, pulleys, and covers Reduced weight, high strength and stiffness
Aluminum–aluminum oxide (short Piston ring Wear resistance, high running temperature
fibers) Piston crown (combustion bowl) Reduced reciprocating mass, high creep and fatigue
resistance
Aluminum–aluminum oxide (long Connecting rod Reduced reciprocating mass, improved strength and
fibers) stiffness
Copper–graphite Electrical contact strips, electronics Low friction and wear, low coefficient of thermal
packaging, bearings expansion
Aluminum–graphite Cylinder, liner platon, bearings Call resistance, reduced friction, wear and weight
Aluminum–titanium carbide (particle) Piston, connecting rod Reduced weight and wear
Aluminum–fiber flax Piston Reduced weight and wear
Aluminum–aluminum oxide fibers– Engine block Reduced weight, improved strength and wear
carbon fibers resistance
Supper alloy-based composite (Ni– Turbine blades Fatigue resistance, impact strength, temperature
Ni3Nb) resistance
Cu–Nb, Cu–Nb3Sn Super conductor Superconducting, mechanical strength, ductility
Cu–W Spot welding electrodes Burn-up resistance

References

Adebisi, A.A., Maleque, M.A., Rahman, M.M., 2011. Metal matrix composite brake rotor: Historical development and product life cycle analysis. International Journal of
Automotive and Mechanical Engineering 4, 471–480.
Akhtar, F., 2014. Ceramic reinforced high modulus steel composites: Processing, microstructure and properties. Canadian Metallurgical Quarterly 53, 253–263.
Ashby, M.F., 2005. Materials Selection in Mechanical Design, third ed. Cleveland, OH: Elsevier.
Bewlay, B.P., Jackson, M.R., Zhao, J.C., Subramanian, P.R., 2003. A review of very-high-temperature Nb-Silicide-based composites. Metallurgical and Materials Transactions
34A, 2043–2052.
Bhatnagar, N., Srivatsan, T.S. (Eds.), 2009. Processing and fabrication of advanced materials-XVII. New Delhi: I.K. International Publishing House Ltd.
Bonnen, J.J., You, C.P., Allison, J.E., Jones, J.W., 1990. Proceedings of the International Conference on Fatigue. New York: Pergamon Press, pp. 887  892.
Bukhari, M.Z., Brabazon, D., Hashmi, M.S.J., 2011. Assessment of suitable thermally enhanced materials for electronics packaging application. In 1st International Malaysia−
Ireland Joint Symposium on Engineering, Science and Business, Athlone, Ireland.
Business Communications Company, RGB-108N Metal−matrix composites in the 21st Century: Markets and Opportunities, 2006.
Campbell, F.C. (Ed.), 2012. Lightweight Materials, Understanding the Basics. ASM International.
Capel, H., Harris, S.J., Schulz, P., Kaufmann, H., 2000. Correlation between manufacturing conditions and properties of carbon fibre reinforced Mg. Materials Science and
Technology 16, 765–768.
Champion, A.R., Krueger, W.H., Hartman, H.S., Dhingra, A.K., 1978. In Proceedings of the 2nd International Conference on Composite Materials (ICCM/2), TMS−AIME, New
York.
28 Metal–Matrix Composites

Chawla, K.K., 2012. Composite Materials: Science and Engineering. New York: Springer Science Media.
Chawla, N., Andres, C., Jones, J.W., Allison, J.E., 1998. Metallurgical and Materials Transactions 29A, 2843–2854.
Chawla, N., Chawla, K.K., 2006. Metal−matrix composites in ground transportation. JOM. 67–70.
Chawla, N., Chawla, K.K., 2013. Metal−Matrix Composites, second ed. New York: Springer Science.
Chawla, N., Jones, J.W., Allison., J.E., 1999. Elevated temperature tensile and fatigue behavior of particle reinforced metal−matrix composites. Fatigue '99. In: Wu, X.R., Wang,
Z.G. (Eds.), EMAS/HEP.
Chen, S.H., Jin, P.P., Schumacher, G., Wanderka, N., 2010. Microstructure and interface characterization of a cast Mg2B2O5 whisker reinforced AZ91D magnesium alloy
composite. Composite Science and Technology 70, 123–129.
Clyne, T.W., 2001. Metal−matrix composites: Matrices and processing. In: Jürgen Buschow, K.H., Cahn, R.W., Flemings, M.C., Ilschner, B. (Eds.), Encyclopedia of Materials:
Science and Technology. Amsterdam: Elsevier, pp. 1382–1391.
Diwanji, A., Hall, I.W., 1992. Fiber and fiber-surface treatment effects in carbon/aluminum metal−matrix composites. Journal of Materials Science 27, 2093–2100.
Guo, Z.X., 1998. Towards cost effective manufacturing of Ti/SiC fibre composites and components. Materials Science and Technology 14, 864–872.
Guo, Z.X., Derby, B., 1995. Solid-state fabrication and interfaces of fibre reinforced metal−matrix composites. Progress in Materials Science 39, 411–495.
Hausmann, C., Cayron, C., Beffort, O., Long, S., 1998. In 9th International Conference on Modern Materials and Technologies. CIMTEC'98, Florence, Italy.
Holley, R., 2013. The Great Metal Tube in the Sky. Material Strategies, Innovative Applications in Architecture, https://arch5541.wordpress.com.
Hooker, J., Doorbar, P., 2003. Metal and ceramic matrix composites. In: Cantor, B., Dunne, F.P.E., Stone, I.C. (Eds.), Series in Materials Science and Engineering. IoP
Publication.
Kagawa, Y., Nakata, E., 1992. Some mechanical properties of carbon fibre-reinforced magnesium-matrix composite fabricated by squeeze casting. Journal of Materials Letters
11, 176–178.
Kainer, K.U., 2006. Metal−matrix composites. Custom-Made Materials for Automotive and Aerospace Engineering. Weinheim: Wiley-VCH Verlag GmbH & Co. KGaA.
Matthews, F.L., Rawlings, R.D., 1996. Composite Materials: Engineering and Design. London: Chapman & Hall Publication.
Motta, M.S., Jena, P.K., Brocchi, E.A., Solorzano, I.G., 2001. Materials Science and Engineering C 15, 175–177.
Nardone, V.C., Strife, J.R., 1987. Analysis of the creep behavior of silicon carbide whisker reinforced 2124 Al (T4). Metallurgical Transactions 18A, 109–114.
Nishi, H., Araki, T., Eto, M., 1998. Diffusion bonding of alumina dispersion-strengthened copper to 316 stainless steel with interlayer metals. Fusion Engineering and Design 39
(40), 505–511.
Ottinger, O., Grau, C., Winter, R., et al., 1995. In Tenth International Conference on Composite Materials, ICCM-10 Cambridge: Woodhead Publishing.
Ottinger, O., Schaff, W., Hausmann, C., Heyne, T., Singer, R.F., 1997. In Proceedings of the ICCM−11, ICCM−11, Gold Coast, Austraila.
Peng, H.X., 2005. Manufacturing titanium metal−matrix composites by consolidating matrix coated fibers. Journal of Materials Science and Technology 21, 647–651.
Poletti, C.M., Balog, T., Schubert, V.L., Edtmaier, C., 2008. Production of titanium matrix composites reinforced with SiC particles. Composites Science and Technology 68,
2171–2177.
Russell, A.M., Lee, K.L., 2005. Structure−property relations in nonferrous metals. New Jersey: John Wiley & Sons, Inc.
Singh, N.B., Rai, S., Agarwal, S., 2014. Polymer nanocomposites and Cr(VI) removal from water. Nanoscience and Technology 1 (1), 10.
Singha, M., Singha, K., 2012. Applications of textiles in marine products. Marine Science 2 (6), 110–119.
Sorensen, J., 2011. Great Potential (Still) for Metal Matrix Composite Structures, CPS Technologies Blog, http://www.alsic.com
Trojanová, Z., Száraz, Z., Palček, P., Chalupová, M., 2011. Magnesium Alloys − Design, Processing and Properties. Czerwinski, F. (Ed.), Tech Rijeka, pp. 501–526.
Ye, H., Liu, X., 2004. Review of recent studies in magnesium matrix composites. Journal of Materials Science 39, 6153–6171.
Zweben, C., 2006. composite materials. In mechanical engineers'. In: Myer Kutz (Ed.), Handbook: Materials and Mechanical Design, third ed John Wiley & Sons, Inc.
Chapter 10.
Zweben, C., Schmidt, K.A., 1989. Advanced Composite Packaging Materials. In Electronic Materials Handbook. Materials Park, OH: ASM International. vol. 1.

Further Reading

Chawla, N., Shen, Y.L., 2001. Mechanical behavior of particle reinforced metal−matrix composites. Advanced Engineering Materials 3, 357–370.
Surappa, M.K., 2003. Aluminum matrix composites: Challenges and opportunities. Sadhana 28, 319–334.

Relevant Websites

http://classes.mst.edu/civeng120/lessons/composite/materials/index.html
Introduction to Composite Materials.
www.what-when-how.com
Metal−matrix composites.
www.researchmoz.us
ResearchmozGlobal Pvt. Ltd.
http://siags.siam.org/siaggd/problems/grandine/
Siags.
www.classes.mst.edu
The human body.
http://mmc-assess.tuwien.ac.at/mmc
Worldof MMC asses.

View publication stats

Potrebbero piacerti anche