Sei sulla pagina 1di 75

SHEFFIELD HALLAM UNIVERSITY

Department of Engineering & Mathematics


Faculty of Arts, Computing, Engineering
and Sciences








Composite Materials

















Dr Syed T Hasan
Materials Engineering
Room 4124 Sheaf Building
s.hasan@shu.ac.uk
Tel: 0114 225 3407 (direct line)

POLYMERS:

- The word polymer originates from the Greek word polymeros and means
many-membered. In polymer science it refers to molecules held together
by covalent bonds and composed of small units which are repeated many
times to form very large molecules.


ALL POLYMERS


THERMOPLASTICS THERMOSETS


CRYSTALLINE AMORPHOUS



MOLECULAR STRUCTURES:

- Linear Polymers
- Branched Polymers
- Lightly Crosslinked Polymers
- Network and Heavily Crosslinked Polymers

Molecular weight:

Extremely large molecular weights are to be found in polymers with very long
chains. During the polymerisation process in which these large
macromolecules are synthesised from smaller molecules, not all polymer
chains will grow to the same length; this results in a distribution of chain
lengths or molecules weights.

There are several methods of defining average molecular weight.

i) Number-Average Molecular Weight M
n
:

The number average molecular weight M
n
is obtained by dividing the chains
into a series of size ranges and then determining the number fraction of
chains with each size range. This number average molecular weight is
expressed as,
M
n
= Ex
i
M
i

where M
i
represents the mean molecular weight of size range i, and x
i
is the
fraction of the total number of chains within the corresponding size range.





ii) Weight-Average Molecular Weight M
w
:
A weight average molecular weight M
w
is based on the weight fraction of
molecules within the various ranges, expressed as,
M
w
= Ew
i
M
i

where M
i
is the mean molecular weight within a size range, where as w
i

denotes the weight fraction of molecules within the same size interval.


Degree of Polymerisation, n:

An alternative way of expressing average chain size of a polymer is as the
degree of polymerisation which represents the average number of mer units
in a chain. Both number-average and weight-average degrees of
polymerisation can be calculated,
n
n
= M
n
/ m

n
w
= M
w
/ m
where M
n
and M
w
are the number-average and weight-average molecular
weights while m is the mer molecular weight.


Effects of Molar Mass on Properties:

- Properties such as, density, transparency and refractive index vary very
little with molar mass. However, properties such as softening temperature,
melting temperature, melt viscosity, tensile strength, elastic modulus and
toughness vary considerably with molar mass.
- Most of these properties change rapidly with molar mass over the lower
molar mass range but stabilise at higher values.


Polymer Crystallinity:

- The crystalline state may exist in polymeric materials. However, it involves
molecules instead of just atoms or ions, as with metals and ceramics, the
atomic arrangements will be more complex for polymers.
- Polymer crystallinity can be considered as the packing of molecular chains
so as to produce an ordered atomic array.
- The degree of crystallinity may range from completely amorphous to
almost entirely crystalline (95%).
- The density of crystalline polymer will be greater than an amorphous one
of the same material and molecular weight.
- The degree of crystallinity by weight can be determined from accurate
density measurements,

%crystallinity
c
s a
s
c a
=


|
\

|
.
|
|
\

|
.
|


100

where
s
is the density of a specimen for which the percent crystallinity is to
be determined,
a
is the density of the totally amorphous polymer and
c
is
the density of the perfectly crystalline polymer.

- The degree of crystallinity of a polymer depends on the rate of cooling
during solidification as well as on the chain configuration. The molecular
chemistry as well as chain configuration also influence the ability of
polymer to crystallise.


Polymer Crystals:

- A semicrystalline polymer consists of small crystalline regions, having a
precise alignment, which are embedded within the amorphous matrix
composed of randomly oriented molecules.
- These crystals are regularly shaped, thin platelets approximately 10 to 20
nm thick and 10 m long.
- Most of the polymers when crystallise from melt form spherulites. The
spherulite consists of an aggregate of ribbonlike chain-folded crystallites
that radiate from the centre to outward.


Mechanical and Thermomechanical Behaviour of Polymers:

- The mechanical properties of polymers are specified with many of the
same parameters that are used for metals, modulus of elasticity, tensile,
impact and fatigue strengths.

- The mechanical characteristics of polymers are highly sensitive to the rate
of deformation (strain rate), the temperature and the chemical nature of
the environment.

Mechanism of Deformation of Semicrystalline Polymers:

Macroscopic Deformation

Crystallisation:

- The crystallisation of a molten polymer occurs by nucleation and growth
processes. For polymers upon cooling through the melting temperature
nuclei form wherein small regions of the tangled and random molecules
become ordered and aligned in the manner of chain folded layers.


Melting:

- There are several features distinctive to the melting of polymers that are
not normally observed with metal and ceramics.
- Melting of polymers takes place over a range of temperatures.
- The melting behaviour depends on the history of the specimen, in
particular the temperature at which it crystallised.
- The thickness of chain-folded lamellae will depend on crystallisation
temperature, the thicker the lamellea the higher the melting temperature.
- Melting behaviour is a function of the rate of heating, increasing the rate
results in an elevation of the melting temperature.
- Polymeric materials are also responsive to heat treatments which produce
structural and property alterations.


Glass Transition:

- The glass transition occurs in amorphous or glassy polymers when upon
cooling from liquid crystallisation does not take place, that is polymer
chains are not able to rearrange into a three dimensional long range
ordered structure.
- Upon cooling the glass transition corresponds to an increase in viscosity
and the gradual transformation from a liquid to a rubbery material and
finally to a rigid solid.
- The temperature at which the polymer experiences the transition from
rubbery to rigid states is termed the glass transition temperature.

Thermoplastic and Thermosetting Polymers:

Thermoplasts: Thermoplasts soften when heated and harden
when cooled. These materials are normally fabricated by the simultaneous
application of heat and pressure.

Thermosetting: Thermosetting polymers become permanently hard when
heat is applied and do not soften upon subsequent heating.

Fracture of Polymers:




Introduction to Composites:

A composite is considered to be any multiphase material that exhibits a significant
proportion of the properties of both constituents phases.

A composite, in the present context, is a multiphase material that is artificially made,
as opposed to one that occurs or forms naturally. The constituent phases must be
chemically dissimilar and separated by a distinct interface.

Most composites have been created to improve combination of mechanical
characteristics such as stiffness, toughness and ambient and high temperature strength.

The properties of composites are a function of the properties of the constituent phases,
their relative amounts and the geometry of the dispersed phase.

Composite Theory
In its most basic form a composite material is one, which is composed of at least two
elements working together to produce material properties that are different to the
properties of those elements on their own. In practice, most composites consist of a
bulk material (the matrix), and a reinforcement of some kind, added primarily to
increase the strength and stiffness of the matrix. This reinforcement is usually in fibre
form. Today, the most common man-made composites can be divided into three main
groups:
Polymer Matrix Composites (PMCs) These are the most common and will
be discussed here. Also known as FRP - Fibre Reinforced Polymers (or Plastics)
these materials use a polymer-based resin as the matrix, and a variety of fibres such as
glass, carbon and aramid as the reinforcement.
Metal Matrix Composites (MMCs) - Increasingly found in the automotive
industry, these materials use a metal such as aluminium as the matrix, and reinforce it
with fibres such as silicon carbide.
Ceramic Matrix Composites (CMCs) - Used in very high temperature
environments, these materials use a ceramic as the matrix and reinforce it with short
fibres, or whiskers such as those made from silicon carbide and boron nitride.
Polymer Matrix Composites
Polymer matrix composites consists of a polymer resin as the matrix with fibres as the
reinforcement medium. These materials are used in the greatest diversity of composite
applications as well as in the largest quantities in light of their room temperature
properties, ease of fabrication and cost.

Resin systems such as epoxies and polyesters have limited use for the manufacture of
structures on their own, since their mechanical properties are not very high when
compared to, for example, most metals. However, they have desirable properties, most
notably their ability to be easily formed into complex shapes. Materials such as glass,
aramid and boron have extremely high tensile and compressive strength but in solid
form these properties are not readily apparent. This is due to the fact that when
stressed, random surface flaws will cause each material to crack and fail well below
its theoretical breaking point. To overcome this problem, the material is produced in
fibre form, so that, although the same number of random flaws will occur, they will be
restricted to a small number of fibres with the remainder exhibiting the materials
theoretical strength. Therefore a bundle of fibres will reflect more accurately the
optimum performance of the material. However, fibres alone can only exhibit tensile
properties along the fibres length, in the same way as fibres in a rope.
It is when the resin systems are combined with reinforcing fibres such as glass, carbon
and aramid that exceptional properties can be obtained. The resin matrix spreads the
load applied to the composite between each of the individual fibres and also protects
the fibres from damage caused by abrasion and impact. High strengths and stiffnesses,
ease of moulding complex shapes, high environmental resistance all coupled with low
densities, make the resultant composite superior to metals for many applications.
Since PMCs combine a resin system and reinforcing fibres, the properties of the
resulting composite material will combine something of the properties of the resin on
its own with that of the fibres on their own, as surmised in Figure 1.

Figure: 1 Illustrating the combined effect on Modulus of the addition of fibres to a
resin matrix.
Overall, the properties of the composite are determined by:
The properties of the fibre
The properties of the resin
The ratio of fibre to resin in the composite (Fibre Volume Fraction (FVF))
The geometry and orientation of the fibres in the composite
The ratio of the fibre to resin derives largely from the manufacturing process used to
combine resin with fibre. However, it is also influenced by the type of resin system
used, and the form in which the fibres are incorporated. In general, since the
mechanical properties of fibres are much higher than those of resins, the higher the
fibre volume fraction the higher will be the mechanical properties of the resultant
composite. In practice there are limits to this, since the fibres need to be fully coated
in resin to be effective, and there will be an optimum packing of the generally circular
cross-section fibres. In addition, the manufacturing process used to combine fibre with
resin leads to varying amounts of imperfections and air inclusions.
Typically, with a common hand lay-up process as widely used in the boat-building
industry, a limit for FVF is approximately 30-40%. With the higher quality, more
sophisticated and precise processes used in the aerospace industry, FVFs approaching
70% can be successfully obtained.
The geometry of the fibres in a composite is also important since fibres have their
highest mechanical properties along their lengths, rather than across their widths. This
leads to the highly anisotropic properties of composites, where, unlike metals, the
mechanical properties of the composite are likely to be very different when tested in
different directions. This means that it is very important when considering the use of
composites to understand at the design stage, both the magnitude and the direction of
the applied loads. When correctly accounted for, these anisotropic properties can be
very advantageous since it is only necessary to put material where loads will be
applied, and thus redundant material is avoided.
It is also important to note that with metals the material supplier largely determines
the properties of the materials, and the person who fabricates the materials into a
finished structure can do almost nothing to change those in-built properties.
However, a composite material is formed at the same time, as the structure is itself
being fabricated. This means that the person who is making the structure is creating
the properties of the resultant composite material, and so the manufacturing processes
they use have an unusually critical part to play in determining the performance of the
resultant structure.
Loading
There are four main direct loads that any material in a structure has to withstand:
tension, compression, shear and flexure.
Tension
Figure 2 shows a tensile load applied to a composite. The response of a composite to
tensile loads is very dependent on the tensile stiffness and strength properties of the
reinforcement fibres, since these are far higher than the resin system on its own.

Figure 2 Illustrates the tensile load applied to a composite body.
Compression
Figure 3 shows a composite under a compressive load. Here, the adhesive and
stiffness properties of the resin system are crucial, as it is the role of the resin to
maintain the fibres as straight columns and to prevent them from buckling.

Figure 3 - Illustrates the compression load applied to a composite body.
Shear
Figure 4 shows a composite experiencing a shear load. This load is trying to slide
adjacent layers of fibres over each other. Under shear loads the resin plays the major
role, transferring the stresses across the composite. For the composite to perform well
under shear loads the resin element must not only exhibit good mechanical properties
but must also have high adhesion to the reinforcement fibre. The interlaminar shear
strength (ILSS) of a composite is often used to indicate this property in a multiplayer
composite (laminate).

Figure 4 - Illustrates the shear load applied to a composite body.
Flexure
Flexural loads are really a combination of tensile, compression and shear loads. When
loaded as shown (Figure 5), the upper face is put into compression, the lower face into
tension and the central portion of the laminate experiences shear.

Figure 5 - Illustrates the loading due to flexure on a composite body.
Comparison with Other Structural Materials
Due to the factors described above, there is a very large range of mechanical
properties that can be achieved with composite materials. Even when considering one
fibre type on its own, the composite properties can vary by a factor of 10 with the
range of fibre contents and orientations that are commonly achieved. The comparisons
that follow therefore show a range of mechanical properties for the composite
materials. The lowest properties for each material are associated with simple
manufacturing processes and material forms (e.g. spray lay-up glass fibre), and the
higher properties are associated with higher technology manufacture (e.g. autoclave
moulding of unidirectional glass fibre prepreg), such as would be found in the
aerospace industry.
For the other materials shown, a range of strength and stiffness (modulus) figures are
also given to indicate the spread of properties associated with different alloys, for
example.

Figure 6 Tensile Strength of Common Structural Materials

Figure 7 Tensile Modulus of Common Structural Materials
The above Figures (6 and 7) clearly show the range of properties that different
composite materials can display. These properties can best be summed up as high
strengths and stiffnesses combined with low densities. It is these properties that give
rise to the characteristic high strength and stiffness to weight ratios that make
composite structures ideal for so many applications. This is particularly true of
applications, which involve movement, such as cars, trains and aircraft, since lighter
structures in such applications play a significant part in making these applications
more efficient. The strength and stiffness to weight ratio of composite materials can
best be illustrated by the following graphs that plot specific properties. These are
simply the result of dividing the mechanical properties of a material by its density.
Generally, the properties at the higher end of the ranges illustrated in the previous
graphs (Figures 6 and 7) are produced from the highest density variant of the material.
The spread of specific properties shown in the following graphs (Figures 8 and 9)
takes this into account.

Figure 8 Specific Tensile Strength of Common Structural Materials

Figure 9 - Specific Tensile Modulus of Common Structural Materials




Continuous Fibre-Reinforced Composites:

The most important composites are those in which the dispersed phase is in the form
of a fibre. Design goals of fibre reinforced composites often include high strength and
stiffness on a weight basis and these characteristics are expressed in terms of specific
strength and specific modules parameters.

Fibre reinforced composites are sub-classified by fibre length.

Influence of Fibre Length:

The mechanical characteristics of a fibre reinforced composite depend not only on the
properties of the fibre but also on the degree to which an applied load is transmitted to
the fibres by the matrix phase. The extent of the load transmittance is the magnitude
of the interfacial bond between the fibre and matrix phases.

A critical fibre length is necessary for effective strengthening and stiffening of the
composite material.The critical fibre length l
c
is dependent on the fibre diameter d and
its ultimate tensile strength o
f
and on the fibre-matrix bond strength t
c
as,
l
c
f
d
c
=
o
t 2

For a number of glass and carbon fibre-matrix combinations, the critical length is the
order of 1 mm, which is 20 - 150 times the fibre diameter.If the fibre length is (l > 15
l
c
), it is termed as continuous fibre reinforcement.

For discontinuous fibres of lengths significantly less than l
c
the matrix deforms around
the fibre such that there is virtually no stress transference and little reinforcement by
the fibre.

Influence of Fibre Orientation and Concentration:

Longitudinal Loading:

The properties of a composite having its fibres aligned are highly anisotropic.

Let us consider the deformation of a composite in which a stress is applied along the
direction of alignment. The total load sustained by the composite F
c
is equal to the
loads carried by the matrix phase F
m
and the fibre phase F
f
,

F
c
= F
m
+ F
f

or
o
c
A
c
= o
m
A
m
+ o
f
A
f

o
c
= o
m
(A
m
/A
c
) + o
f
(A
f
/A
c
)
where A
m
/A
c
and A
f
/A
c
are the area fractions of the matrix and fibre phases
respectively.

If the composite, matrix and fibre lengths are all equal, A
m
/A
c
is equivalent to the
volume fraction of the matrix V
m
and likewise for the fibres V
f
.
o
c =
o
m
V
m
+ o
f
V
f
Now if we assume an isostrain state,
c
c
= c
m
= c
f
when each term in equation -- is divided by its respective strain and if composite,
matrix and fibre deformations are all elastic then,
E
cl
= E
m
V
m
+ E
f
V
f

where, E
c
= o
c
/c
c
etc. and since composite consists of only matrix and fibre phases,
that is
V
m
+ V
f
= 1 then
E
cl
= E
m
(1+ V
f
)+ E
f
V
f

The ratio of the load carried by the fibres to that carried by the matrix is,
(F
f
/ F
m
) = (E
f
V
f
) / (E
m
V
m
)
Transverse Loading:

A continuous and oriented fibre composite may be loaded in the transverse direction,
the load is applied at a 90
o
angle to the direction of fibre alignment. The stress o to
which the composite as well as both phases are exposed is the same,
o
c
= o
m
= o
f
= o
This is termed as isostress state. The strain of the entire composite c
c
is,
c
c
= c
m
V
m
+ c
f
V
f

but, since c = o/E,
(o/E
ct
) = (o/E
m
)V
m
+ (o/E
f
)V
f
where E
ct
is the modules of elasticity in the transverse direction and dividing by o
yield,
E
ct
E
m
E
f
V
f
E
f
V
f
E
m
=
+
|
\

|
.
|
|
1


The Fibre Phase:

On the basis of diameter and character fibres are grouped into three different
classifications: whiskers, fibres, and wires.

Whiskers are very thin single crystals that have extremely large length to diameter
ratios.

Materials that are classified as fibres are either polycrystalline or amorphous and have
small diameters.

Fine wires have relatively large diameters.

The Matrix Phase:

The matrix phase of fibrous composites may be a metal, polymer or ceramic. In
general metals and polymers are used as matrix materials because some ductility is
desirable. For fibre reinforced composites the matrix phase serves several functions.
First it binds the fibres together and act as the medium by which an externally applied
stress is transmitted and distributed to the fibres. The second function of the matrix is
to protect the individual fibres from surface damage as a result of mechanical abrasion
or chemical reactions with the environment.Finally the matrix separates the fibres and
by virtue of its relative softness and plasticity prevents the propagation of brittle
cracks from fibre to fibre, which could result in catastrophic failure.It is essential that
adhesive bonding forces between fibre and matrix be high to minimise fibre pull out.

The most widely utilised and least expensive polymer resins are the polyesters and
vinyl esters, these matrix materials are used primarily for glass fibre reinforced
composites. The epoxies are more expensive are utilised extensively in aerospace
applications. For high temperature applications polyimide resins are used and their
upper temperature limit is approximately 230
o
C. High temperature thermoplastic
resins offer the potential to be used in future aerospace applications.

Glass fibre reinforced polymer composites:

Fibreglass is simply a composite consisting of glass fibres, either continuous or
discontinuous, contained within a polymer matrix.Fibre diameter normally range
between 3 and 20 m.

Glass is popular as a fibre reinforcement material for several reasons:

1. it is easily drawn into high-strength fibres from the molten state.
2. it is readily available and may be fabricated into a glass reinforced plastic
economically using a wide variety of composite manufacturing techniques.
3. as a fibre, it is relatively strong and when embedded in a plastic matrix it
produces a composite having a very high specific strength.
4. when coupled with the various plastics it possesses a chemical inertness that
renders the composite useful in a variety of corrosive environments.

Carbon fibre reinforced polymer composite:

Carbon is a high performance fibre material that is the most commonly used
reinforcement in advanced polymer matrix composites. The reasons for this are as
follows: Carbon fibres have the highest specific modules and specific strength of all
reinforcing fibre materials. They retain their high tensile modules and high strength at
elevated temperatures, high temperature oxidation may be a problem. At room
temperature carbon fibres are not affected by moisture nor a wide variety of solvents,
acids and bases. These fibres exhibit a diversity of physical and mechanical
characteristics allowing composites incorporating these fibres to have specific
engineered properties. Fibre and composite manufacturing processes have been
developed that are relatively inexpensive and cost effective.

Production and Properties of Commonly Used Fibres:

(i) Graphite/Carbon. Polymer fibres e.g. PAN, stretched, heated to about 220C to
cross-link and stabilise. Further heated to 900C to carbonise and then to 1300C+ to
produce correct graphite structure. High modulus or high strength fibres can be
produced.

(ii) Glass. Homogeneous melt of high purity. Spun from a high temperature, rapidly
chilled and immediately given a protective coating.

(iii) Kevlar. Concentrated solution of an aramid polymer in concentrated sulphuric
acid. Spun into a neutralising bath. Jet shape and degree of stretch are critical
parameters to achieve required properties. Fibre washed dried and heated to 550C in a
nitrogen atmosphere.

(iv) Boron. Core of tungsten or graphite is heated to about 1000-1300C in atmosphere
containing volatile compound of boron e.g. boron chloride. CVP to produce pure
boron with few flaws. Careful cleaning of base fibre in hydrogen is carried out. This
prevents excessive crystal growth which would reduce strength.

PROPERTIES

(i) Very high specific strengths. Strain to failure dependent on flaw population. 2%
strain to failure is possible. Tensile modulus in excess of 250GPa depending on
production route. Stiffness of Strength can be maximised but not at same time. This
situation is improving. Pitch fibres give high Modulus but lower strength. Vice versa
for PAN.

(ii) Glass fibres have significantly lower modulus e.g. 40-120GPa. Good strength of 3-
4.5GPa. Higher density giving considerably lower specific strengths.

(iii)Kevlar fibres have modulus values between 50-130GPa with strengths up to 3.1
GPa High specific properties. High strain to fracture and hence good damage
tolerance.

(iv) Boron fibres. Low density metal, which is weak in bulk. High modulus fibre,
370-410GPa. Good in compression. Strength up to 5GPa. Good specific properties.
Applications in terms of requirements e.g. cost versus weight. High specific properties
e.g. E/ in aerospace. Glass for lower cost land based applications. Kevlar where high
damage tolerance required. Hybrids e.g. Glass/carbon, carbon/Kevlar are common to
produce required property combinations.




Mechanics of Composite Materials
Version 2.1
Bill Clyne, University of Cambridge
Boban Tanovic, MATTER Project
Assumed Pre-knowledge
It is assumed that the student is familiar with simple concepts of mechanical behaviour, such as the broad meanings of
stress and strain. It would be an advantage for the student to understand that these are really tensor quantities,
although this is by no means essential. All of the terms associated with the assumed pre-knowledge are defined in the
glossary, which can be consulted by the student at any time.
Most of the material in this package is based on a recently published book. This is:
"An Introduction to Composite Materials", D.Hull and T.W.Clyne, Cambridge University Press (1996) Order!
This source should be consulted for background to the treatments in this module, particularly mathematical details.
What is a Composite Material?
Most composites have strong, stiff fibres in a matrix which is weaker and less stiff. The objective is usually to make
a component which is strong and stiff, often with a low density. Commercial material commonly has glass or carbon
fibres in matrices based on thermosetting polymers, such as epoxy or polyester resins. Sometimes, thermoplastic
polymers may be preferred, since they are mouldable after initial production. There are further classes of composite
in which the matrix is a metal or a ceramic. For the most part, these are still in a developmental stage, with problems
of high manufacturing costs yet to be overcome. Furthermore, in these composites the reasons for adding the fibres
(or, in some cases, particles) are often rather complex; for example, improvements may be sought in creep, wear,
fracture toughness, thermal stability, etc. This software package covers simple mechanics concepts of stiffness and
strength, which, while applicable to all composites, are often more relevant to fibre-reinforced polymers.
Module Structure
The module comprises three sections:
- Load Transfer
- Composite Laminates
- Fracture Behaviour
Brief descriptions are given below of the contents of these sections, covering both the main concepts involved and the
structure of the software.




Load Transfer
Summary
This section covers basic ideas concerning the manner in an applied mechanical load is shared between the matrix
and the fibres. The treatment starts with the simple case of a composite containing aligned, continuous fibres. This
can be represented by the slab model. For loading parallel to the fibre axis, the equal strain condition is imposed,
leading to the Rule of Mixtures expression for the Young's modulus. This is followed b y the cases of transverse
loading of a continuous fibre composite and axial loading with discontinuous fibres.
What is meant by Load Transfer?
The concept of load sharing between the matrix and the reinforcing constituent (fibre) is central to an understanding
of the mechanical behaviour of a composite. An external load (force) applied to a composite is partly borne by the
matrix and partly by the reinforcement. The load carried by the matrix across a section of the composite is given by
the product of the average stress in the matrix and its sectional area. The load carried by the reinforcement is
determined similarly. Equating the externally imposed load to the sum of these two contributions, and dividing
through by the total sectional area, gives a basic and important equation of composite theory, sometimes termed the
"Rule of Averages".

(1)
which relates the volume-averaged matrix and fibre stresses ( ), in a composite containing a volume (or
sectional area) fraction f of reinforcement, to the applied stress s
A
. Thus, a certain proportion of an imposed load will
be carried by the fibre and the remainder by the matrix. Provided the response of the composite remains elastic, this
proportion will be independent of the applied load and it represents an important characteristic of the material. It
depends on the volume fraction, shape and orientation of the reinforcement and on the elastic properties of both
constituents. The reinforcement may be regarded as acting efficiently if it carries a relatively high proportion of the
externally applied load. This can result in higher strength, as well as greater stiffness, because the reinforcement is
usually stronger, as well as stiffer, than the matrix.
What happens when a Composite is Stressed?

Figure 1
Consider loading a composite parallel to the fibres. Since they are bonded together, both fibre and matrix will stretch
by the same amount in this direction, i.e. they will have equal strains, e (Fig. 1). This means that, since the fibres are
stiffer (have a higher Young modulus, E), they will be carrying a larger stress. This illustrates the concept of load
transfer, or load partitioning between matrix and fibre, which is desirable since the fibres are better suited to bear
high stresses. By putting the sum of the contributions from each phase equal to the overall load, the Young modulus
of the composite is found (diagram). It can be seen that a "Rule of Mixtures" applies. This is sometimes termed the
"equal strain" or "Voigt" case. Page 2 in the section covers derivation of the equation for the axial stiffness of a
composite and page 3 allows the effects on composite stiffness of the fibre/matrix stiffness ratio and the fibre volume
fraction to be explored by inputting selected values.
What about the Transverse Stiffness?
Also of importance is the response of the composite to a load applied transverse to the fibre direction. The stiffness
and strength of the composite are expected to be much lower in this case, since the (weak) matrix is not shielded from
carrying stress to the same degree as for axial loading. Prediction of the transverse stiffness of a composite from the
elastic properties of the constituents is far more difficult than the axial value. The conventional approach is to assume
that the system can again be represented by the "slab model". A lower bound on the stiffness is obtained from the
"equal stress" (or "Reuss") assumption shown in Fig. 2. The value is an underestimate, since in practice there are
parts of the matrix effectively "in parallel" with the fibres (as in the equal strain model), rather than "in series" as is
assumed. Empirical expressions are available which give much better approximations, such as that of Halpin-Tsai.
There are again two pages in the section covering this topic, the first (page 4) outlining derivation of the equal stress
equation for stiffness and the second (page 5) allowing this to be evaluated for different cases. For purposes of
comparison, a graph is plotted of equal strain, equal stress and Halpin-Tsai predictions. The Halpin-Tsai expression
for transverse stiffness (which is not given in the module, although it is available in the glossary) is:

(2)
in which



Figure 2
The value of x may be taken as an adjustable parameter, but its magnitude is generally of the order of unity. The
expression gives the correct values in the limits of f=0 and f=1 and in general gives good agreement with experiment
over the complete range of fibre content. A general conclusion is that the transverse stiffness (and strength) of an
aligned composite are poor; this problem is usually countered by making a laminate (see section on "composite
laminates").
How is Strength Determined?
There are several possible approaches to prediction of the strength of a composite. If the stresses in the two
constituents are known, as for the long fibre case under axial loading, then these values can be compared with the
corresponding strengths to determine whether either will fail. Page 6 in the section briefly covers this concept. (More
details about strength are given in the section on "Fracture Behaviour".) The treatment is a logical development from
the analysis of axial stiffness, with the additional input variable of the ratio between the strengths of fibre and matrix.
Such predictions are in practice complicated by uncertainties about in situ strengths, interfacial properties,
residual stresses etc. Instead of relying on predictions such as those outlined above, it is often necessary to measure
the strength of the composite, usually by loading parallel, transverse and in shear with respect to the fibres. This
provides a basis for prediction of whether a component will fail when a given set of stresses is generated (see section
on "Fracture Behaviour"), although in reality other factors such as environmental degradation or the effect of
failure mode on toughness, may require attention.
What happens with Short Fibres?
Short fibres can offer advantages of economy and ease of processing. When the fibres are not long, the equal strain
condition no longer holds under axial loading, since the stress in the fibres tends to fall off towards their ends (see
Fig. 3). This means that the average stress in the matrix must be higher than for the long fibre case. The effect is
illustrated pictorially in pages 7 and 8 of the section.

Figure 3
This lower stress in the fibre, and correspondingly higher average stress in the matrix (compared with the long fibre
case) will depress both the stiffness and strength of the composite, since the matrix is both weaker and less stiff than
the fibres. There is therefore interest in quantifying the change in stress distribution as the fibres are shortened.
Several models are in common use, ranging from fairly simple analytical methods to complex numerical packages.
The simplest is the so-called "shear lag" model. This is based on the assumption that all of the load transfer from
matrix to fibre occurs via shear stresses acting on the cylindrical interface between the two constituents. The build-up
of tensile stress in the fibre is related to these shear stresses by applying a force balance to an incremental section of
the fibre. This is depicted in page 9 of the section. It leads to an expression relating the rate of change of the stress in
the fibre to the interfacial shear stress at that point and the fibre radius, r.

(3)
which may be regarded as the basic shear lag relationship. The stress distribution in the fibre is determined by
relating shear strains in the matrix around the fibre to the macroscopic strain of the composite. Some mathematical
manipulation leads to a solution for the distribution of stress at a distance x from the mid-point of the fibre which
involves hyperbolic trig functions:

(4)
where e
1
is the composite strain, s is the fibre aspect ratio (length/diameter) and n is a dimensionless constant given
by:

(5)
in which n
m
is the Poisson ratio of the matrix. The variation of interfacial shear stress along the fibre length is
derived, according to Eq.(3), by differentiating this equation, to give:

(6)
The equation for the stress in the fibre, together with the assumption of a average tensile strain in the matrix equal to
that imposed on the composite, can be used to evaluate the composite stiffness. This leads to:

(7)
The expression in square brackets is the composite stiffness. In page 10 of the section, there is an opportunity to
examine the predicted stiffness as a function of fibre aspect ratio, fibre/matrix stiffness ratio and fibre volume
fraction. The other point to note about the shear lag model is that it can be used to examine inelastic behaviour. For
example, interfacial sliding (when the interfacial shear stress reaches a critical value) or fibre fracture (when the
tensile stress in the fibre becomes high enough) can be predicted. As the strain imposed on the composite is
increased, sliding spreads along the length of the fibre, with the interfacial shear stress unable to rise above some
critical value, t
i*
. If the interfacial shear stress becomes uniform at t
i*
along the length of the fibre, then a critical
aspect ratio, s
*
, can be identified, below which the fibre cannot undergo fracture. This corresponds to the peak
(central) fibre stress just attaining its ultimate strength sf*, so that, by integrating Eq.(3) along the fibre half-length:

(8)
It follows from this that a distribution of aspect ratios between s
*
and s
*
/2 is expected, if the composite is subjected to
a large strain. The value of s
*
ranges from over 100, for a polymer composite with poor interfacial bonding, to about
2-3 for a strong metallic matrix. In page 10, the effects of changing various parameters on the distributions of
interfacial shear stress and fibre tensile stress can be explored and predictions made about whether fibres of the
specified aspect ratio can be loaded up enough to cause them to fracture.
Conclusion
After completing this section, the student should:
- Appreciate that the key issue, controlling both stiffness and strength, is the way in which an applied load is
shared between fibres and matrix.
- Understand how the slab model is used to obtain axial and transverse stiffnesses for long fibre composites.
- Realise why the slab model (equal stress) expression for transverse stiffness is an underestimate and be able
to obtain a more accurate estimate by using the Halpin-Tsai equation.
- Understand broadly why the axial stiffness is lower when the fibres are discontinuous and appreciate the
general nature of the stress field under load in this case.
- Be able to use the shear lag model to predict axial stiffness and to establish whether fibres of a given aspect
ratio can be fractured by an applied load.
- Note that the treatments employed neglect thermal residual stresses, which can in practice be significant in
some cases.





Composite Laminates
Summary
This section covers the advantages of lamination, the factors affecting choice of laminate structure and the approach
to prediction of laminate properties. It is first confirmed that, while unidirectional plies can have high axial stiffness
and strength, these properties are markedly anisotropic. With a laminate, there is scope for tailoring the properties in
different directions within a plane to the requirements of the component. Both elastic and strength properties can be
predicted once the stresses on the individual plies have been established. This is done by first studying how the
stiffness of a ply depends on the angle between the loading direction and then imposing the condition that all the
individual plies in a laminate must exhibit the same strain. The methodology for prediction of the properties of any
laminate is thus outlined, although most of the mathematical details are kept in the background.
What is a Laminate?
High stiffness and strength usually require a high proportion of fibres in the composite. This is achieved by aligning a
set of long fibres in a thin sheet (a lamina or ply). However, such material is highly anisotropic, generally being
weak and compliant (having a low stiffness) in the transverse direction. Commonly, high strength and stiffness are
required in various directions within a plane. The solution is to stack and weld together a number of sheets, each
having the fibres oriented in different directions. Such a stack is termed a laminate. An example is shown in the
diagram. The concept of a laminate, and a pictorial illustration of the way that the stiffness becomes more isotropic
as a single ply is made into a cross-ply laminate, are presented in page 1 of this section.
What are the Stresses within a Crossply Laminate?
The stiffness of a single ply, in either axial or transverse directions, can easily be calculated. (See the section on Load
Transfer). From these values, the stresses in a crossply laminate, when loaded parallel to the fibre direction in one of
the plies, can readily be calculated. For example, the slab model can be applied to the two plies in exactly the same
way as it was applied in the last section to fibres and matrix. This allows the stiffness of the laminate to be calculated.
This gives the strain (experienced by both plies) in the loading direction, and hence the average stress in each ply, for
a given applied stress. The stresses in fibre and matrix within each ply can also be found from these average stresses
and a knowledge of how the load is shared. In page 2 of this section, by inputting values for the fibre/matrix stiffness
ratio and fibre content, the stresses in both plies, and in their constituents, can be found. Note that, particularly with
high stiffness ratios, most of the applied load is borne by the fibres in the "parallel" ply (the one with the fibre axis
parallel to the loading axis).
What is the Off-Axis Stiffness of a Ply?
For a general laminate, however, or a crossply loaded in some arbitrary direction, a more systematic approach is
needed in order to predict the stiffness and the stress distribution. Firstly, it is necessary to establish the stiffness of a
ply oriented so the fibres lie at some arbitrary angle to the stress axis. Secondly, further calculation is needed to find
the stiffness of a given stack. Consider first a single ply. The stiffness for any loading angle is evaluated as follows,
considering only stresses in the plane of the ply The applied stress is first transformed to give the components parallel
and perpendicular to the fibres. The strains generated in these directions can be calculated from the (known) stiffness
of the ply when referred to these axes. Finally, these strains are transformed to values relative to the loading direction,
giving the stiffness.

Figure 4
These three operations can be expressed mathematically in tensor equations. Since we are only concerned with
stresses and strains within the plane of the ply, only 3 of each (two normal and one shear) are involved. The first step
of resolving the applied stresses, s
x
, s
y
and t
xy
, into components parallel and normal to the fibre axis, s
1
, s
2
and t
12
(see
Fig. 4), depends on the angle, f between the loading direction (x) and the fibre axis (1)

(9)
where the transformation matrix is given by:

(10)
in which c = cosf and s = sinf. For example, the value of s
1
would be obtained from:

(11)
Now, the elastic response of the ply to stresses parallel and normal to the fibre axis is easy to analyse. For example,
the axial and transverse Youngs moduli (E
1
and E
2
) could be obtained using the slab model or Halpin-Tsai
expressions (see Load Transfer section). Other elastic constants, such as the shear modulus (G
12
) and Poissons ratios,
are readily calculated in a similar way. The relationship between stresses and resultant strains dictated by these elastic
constants is neatly expressed by an equation involving the compliance tensor, S, which for our composite ply, has
the form:

(12)
in which, by inspection of the individual equations, it can be seen that





Application of Eq.(12), using the stresses established from Eq.(9), now allows the strains to be established, relative to
the 1 and 2 directions. There is a minor complication in applying the final stage of converting these strains so that
they refer to the direction of loading (x and y axes). Because engineering and tensorial shear strains are not quite the
same, a slightly different transformation matrix is applicable from that used for stresses

(13)
in which,

and the inverse of this matrix is used for conversion in the reverse direction,

(14)
in which,

The final expression relating applied stresses and resultant strains can therefore be written,

(15)
The elements of | |, the transformed compliance tensor, are obtained by concatenation (the equivalent of
multiplication) of the matrices | T '|
-1
, | S | and | T |. The following expressions are obtained





(16)


Figure 5
The final result of this rather tedious derivation is therefore quite straightforward. Eq.(16), together with the elastic
constants of the composite when loaded parallel and normal to the fibre axis, allows the elastic deformation of the ply
to be predicted for loading at any angle to the fibre axis. This is conveniently done using a simple computer program.
The results of such calculations can be explored using pages 4 and 5 in this section. As an example, Fig. 5 shows the
Young's modulus for the an polyester-50% glass fibre ply as the angle, f between fibre axis and loading direction
rises from 0 to 90. A sharp fall is seen as f exceeds about 5-10.
How is the Stiffness of a Laminate obtained?
Once the elastic response of a single ply loaded at an arbitrary angle has been established, that of a stack bonded
together (i.e. a laminate) is quite easy to predict. For example, the Young's modulus in the loading direction is given
by an applied normal stress over the resultant normal strain in that direction. This same strain will be experienced by
all of the component plies of the laminate. Since every ply now has a known Young's modulus in the loading
direction (dependent on its fibre direction), the stress in each one can be expressed in terms of this universal strain.
Furthermore, the force (stress times sectional area) represented by the applied stress can also be expressed as the sum
of the forces being carried by each ply. This allows the overall Young's modulus of the laminate to be calculated. The
results of such calculations, for any selected stacking sequence, can be explored using pages 4 and 5.
Are Other Elastic Constants Important?
There are several points of interest about how a ply changes shape in response to an applied load. For example, the
lateral contraction (Poisson ratio, n) behaviour may be important, since in a laminate such contraction may be
resisted by other plies, setting up stresses transverse to the applied load. Another point with fibre composites under
off-axis loading is that shear strains can arise from tensile stresses (and vice versa). This corresponds to the elements
of S which are zero in Eq.(12) becoming non-zero for an arbitrary loading angle (Eq.(16)). These so-called "tensile-
shear interactions" can be troublesome, since they can set up stresses between individual plies and can cause the
laminate to become distorted. The value of , for example, represents the ratio between g
12
and s
1
. Its value can be
obtained for any specified laminate by using page 6 of this section. It will be seen that, depending on the stacking
sequence, relatively high distortions of this type can arise. On the other hand, a stacking sequence with a high degree
of rotational symmetry can show no tensile-shear interactions. When the tensile-shear interaction terms contributed
by the individual laminae all cancel each other out in this way, the laminate is said to be "balanced". Simple crossply
and angle-ply laminates are not balanced for a general loading angle, although both will be balanced when loaded at
f=0 (i.e. parallel to one of the plies for a cross-ply or equally inclined to the +q and -q plies for the angle-ply case). If
the plies vary in thickness, or in the volume fractions or type of fibres they contain, then even a laminate in which the
stacking sequence does exhibit the necessary rotational symmetry is prone to tensile-shear distortions and
computation is necessary to determine the lay-up sequence required to construct a balanced laminate. The stacking
order in which the plies are assembled does not enter into these calculations.
Conclusion
After completing this section, the student should:
- Appreciate that, while individual plies are highly anisotropic, they can be assembled into laminates having a
selected set of in-plane properties.
- Understand broadly how the elastic properties of a laminate, and the partitioning of an applied load between
the constituent plies, can be predicted.
- Be able to use the software package to predict the characteristics of specified laminate structures.
- Understand the meaning of a "balanced" laminate.
















Fracture Behaviour
Summary
This section covers simple approaches to prediction of the failure of composites from properties of matrix and fibre
and from interfacial characteristics. The axial strength of a continuous fibre composite can be predicted from
properties of fibre and matrix when tested in isolation. Failures when loaded transversely or in shear relative to the
fibre direction, on the other hand, tends to be sensitive to the interfacial strength and must therefore be measured
experimentally. An outline is given of how these measured strengths can be used to predict failure of various laminate
structures made from the composite concerned. Finally, a brief description is given of what is meant by the toughness
(fracture energy) of a material. In composites the most significant contribution to the fracture energy usually comes
from fibre pullout. A simple model is presented for prediction of the fracture energy from fibre pullout, depending on
fibre aspect ratio, fibre radius and interfacial shear strength.
How do Composites Fracture?

Figure 6
Fracture of long fibre composites tends to occur either normal or parallel to the fibre axis. This is illustrated on page
1 of this section - see Fig. 6. Large tensile stresses parallel to the fibres, s
1
, lead to fibre and matrix fracture, with the
fracture path normal to the fibre direction. The strength is much lower in the transverse tension and shear modes and
the composite fractures on surfaces parallel to the fibre direction when appropriate s
2
or t
12
stresses are applied. In
these cases, fracture may occur entirely within the matrix, at the fibre/matrix interface or primarily within the fibre.
To predict the strength of a lamina or laminate, values of the failure stresses s
1*
, s
2*
and t
12*
have to be determined.
Can the Axial Strength be Predicted?
Understanding of failure under an applied tensile stress parallel to the fibres is relatively simple, provided that both
constituents behave elastically and fail in a brittle manner. They then experience the same axial strain and hence
sustain stresses in the same ratio as their Young's moduli. Two cases can be identified, depending on whether matrix
or fibre has the lower strain to failure. These cases are treated in pages 2 and 3 respectively.

Figure 7
Consider first the situation when the matrix fails first (e
m*
<e
f*
). For strains up to e
m*
, the composite stress is given by
the simple rule of mixtures:

(17)
Above this strain, however, the matrix starts to undergo microcracking and this corresponds with the appearance of a
"knee" in the stress-strain curve. The composite subsequently extends with little further increase in the applied stress.
As matrix cracking continues, the load is transferred progressively to the fibres. If the strain does not reach e
f*
during
this stage, further extension causes the composite stress to rise and the load is now carried entirely by the fibres. Final
fracture occurs when the strain reaches e
f*
, so that the composite failure stress s
1*
is given by f s
f*
. A case like this is
illustrated in Fig. 7, which refers to steel rods in a concrete matrix.
[FB2, RHS, real system data, mild steel fibres, concrete matrix, fibre fraction 40%, "strength v. fraction of fibres"
clicked]

Figure 8
Alternatively, if the fibres break before matrix cracking has become sufficiently extensive to transfer all the load to
them, then the strength of the composite is given by:

(18)
where s
fm*
is the fibre stress at the onset of matrix cracking (e
1
=e
m*
). The composite failure stress depends therefore
on the fibre volume fraction in the manner shown in Fig. 8. The fibre volume fraction above which the fibres can
sustain a fully transferred load is obtained by setting the expression in Eq.(18) equal to f s
f*
, leading to:

(19)
If the fibres have the smaller failure strain (page 3), continued straining causes the fibres to break up into
progressively shorter lengths and the load to be transferred to the matrix. This continues until all the fibres have
aspect ratios below the critical value (see Eq.(8)). It is often assumed in simple treatments that only the matrix is
bearing any load by the time that break-up of fibres is complete. Subsequent failure then occurs at an applied stress of
(1-f) s
m*
. If matrix fracture takes place while the fibres are still bearing some load, then the composite failure stress is:

(20)
where s
mf
is the matrix stress at the onset of fibre cracking. In principle, this implies that the presence of a small
volume fraction of fibres reduces the composite failure stress below that of the unreinforced matrix. This occurs up to
a limiting value f ' given by setting the right hand side of Eq.(20) equal to (1-f) s
m*
.

(21)
The values of these parameters can be explored for various systems using pages 2 and 3. Prediction of the values of
s
2*
and t
12*
from properties of the fibre and matrix is virtually impossible, since they are so sensitive to the nature of
the fibre-matrix interface. In practice, these strengths have to be measured directly on the composite material
concerned.
How do Plies Fail under Off-axis Loads?
Failure of plies subjected to arbitrary (in-plane) stress states can be understood in terms of the three failure
mechanisms (with defined values of s
1*
, s
2*
and t
12*
) which were depicted on page 1. A number of failure criteria
have been proposed. The main issue is whether or not the critical stress to trigger one mechanism is affected by the
stresses tending to cause the others - i.e. whether there is any interaction between the modes of failure. In the simple
maximum stress criterion, it is assumed that failure occurs when a stress parallel or normal to the fibre axis reaches
the appropriate critical value, that is when one of the following is satisfied:



(22)

For any stress system (s
x
, s
y
and t
xy
) applied to the ply, evaluation of these stresses can be carried out as described in
the section on Composite Laminates (Eqs.(9) and (10)).

Figure 9
Monitoring of s
1
, s
2
and t
12
as the applied stress is increased allows the onset of failure to be identified as the point
when one of the inequalities in Eq.(22) is satisfied. Noting the form of | T | (Eq.(10)), and considering applied
uniaxial tension, the magnitude of s
x
necessary to cause failure can be plotted as a function of angle f between stress
axis and fibre axis, for each of the three failure modes.

(23)

(24)

(25)
The applied stress levels at which these conditions become satisfied can be explored using page 5. As an example, the
three curves corresponding to Eqs.(23)-(25) are plotted in Fig. 9, using typical values of s
1*
, s
2*
and t
12*
. Typically,
axial failure is expected only for very small loading angles, but the predicted transition from shear to transverse
failure may occur anywhere between 20 and 50, depending on the exact values of t
12*
and s
2*
.
In practice, there is likely to be some interaction between the failure modes. For example, shear failure is expected to
occur more easily if, in addition to the shear stress, there is also a normal tensile stress acting on the shear plane. The
most commonly used model taking account of this effect is the Tsai-Hill criterion. This can be expressed
mathematically as

(26)
This defines an envelope in stress space: if the stress state (s
1
, s
2
and t
12
) lies outside of this envelope, i.e. if the sum
of the terms on the left hand side is equal to or greater than unity, then failure is predicted. The failure mechanism is
not specifically identified, although inspection of the relative magnitudes of the terms in Eq.(26) gives an indication
of the likely contribution of the three modes. Under uniaxial loading, the Tsai-Hill criterion tends to give rather
similar predictions to the Maximum Stress criterion for the strength as a function of loading angle. The predicted
values tend to be somewhat lower with the Tsai-Hill criterion, particularly in the mixed mode regimes where both
normal and shear stresses are significant. This can be explored on page 6.
What is the Failure Strength of a Laminate?
The strength of laminates can be predicted by an extension of the above treatment, taking account of the stress
distributions in laminates, which were covered in the preceding section. Once these stresses are known (in terms of
the applied load), an appropriate failure criterion can be applied and the onset and nature of the failure predicted.

Figure 10
However, failure of an individual ply within a laminate does not necessarily mean that the component is no longer
usable, as other plies may be capable of withstanding considerably greater loads without catastrophic failure.
Analysis of the behaviour beyond the initial, fully elastic stage is complicated by uncertainties as to the degree to
which the damaged plies continue to bear some load. Nevertheless, useful calculations can be made in this regime
(although the major interest may be in the avoidance of any damage to the component).In page 7, a crossply (0/90)
laminate is loaded in tension along one of the fibre directions. The stresses acting in each ply, relative to the fibre
directions, are monitored as the applied stress is increased. Only transverse or axial tensile failure is possible in either
ply, since no shear stresses act on the planes parallel to the fibre directions. The software allows the onset of failure to
be predicted for any given composite with specified strength values. Although the parallel ply takes most of the load,
it is commonly the transverse ply which fails first, since its strength is usually very low.

In page 8, any specified laminate can be subjected to an imposed stress state and the onset of failure predicted. An
example of such a calculation is shown in Fig. 10.
What is the Toughness (Fracture Energy) of a Composite?
The fracture energy, G
c
, of a material is the energy absorbed within it when a crack advances through the section of a
specimen by unit area. Potentially the most significant source of fracture work for most fibre composites is interfacial
frictional sliding. Depending on the interfacial roughness, contact pressure and sliding distance, this process can
absorb large quantities of energy. The case of most interest is pull-out of fibres from their sockets in the matrix. This
process is illustrated schematically in page 9.
The work done as a crack opens up and fibres are pulled out of their sockets can be calculated in the following way.
A simple shear lag approach is used. Provided the fibre aspect ratio, s (=L/r), is less than the critical value, s*
(=s
f*
/2t
i*
), see page 10 of the Load Transfer section, all of the fibres intersected by the crack debond and are
subsequently pulled out of their sockets in the matrix (rather than fracturing). Consider a fibre with a remaining
embedded length of x being pulled out an increment of distance dx. The associated work is given by the product of
the force acting on the fibre and the distance it moves
dU = (2prxt
i*
) dx (27)

where t
i*
is the interfacial shear stress, taken here as constant along the length of the fibre. The work done in pulling
this fibre out completely is therefore given by

(28)
where x
0
is the embedded length of the fibre concerned on the side of the crack where debonding occurs (x
0
= L). The
next step is an integration over all of the fibres. If there are N fibres per m
2
, then there will be (N dx
0
/ L) per m
2
with
an embedded length between x
0
and (x
0
+ dx
0
). This allows an expression to be derived for the pull-out work of
fracture, G
c


(29)
The value of N is related to the fibre volume fraction, f, and the fibre radius, r
N =
(30)
Eq.(29) therefore simplifies to

(31)
This contribution to the overall fracture energy can be large. For example, taking f=0.5, s=50, r=10 m and
t
i*
=20 MPa gives a value of about 80 kJ m
-2
. This is greater than the fracture energy of many metals. Since s
f*
would
typically be about 3 GPa, the critical aspect ratio, s* (=s
f*
/2t
i*
), for this value of t
i*
, would be about 75. Since this is
greater than the actual aspect ratio, pull-out is expected to occur (rather than fibre fracture), so the calculation should
be valid. The pull-out energy is greater when the fibres have a larger diameter, assuming that the fibre aspect ratio is
the same. In page 10, the cumulative fracture energy is plotted as the crack opens up and fibres are pulled out of their
sockets. The end result for a particular case is shown in Fig. 11.

Figure 11
Conclusion
After completing this section, the student should:
- Appreciate that a unidirectional composite tends to fracture axially, transversely or in shear relative to the
fibre direction.
- Be able to use simple expressions for axial composite strength, based on fibre and matrix fracturing similarly
in the composite and in isolation.
- Understand what is meant by "mixed mode" failure and be able to use Maximum Stress or Tsai-Hill criteria
to predict how a unidirectional composite will fail under multi-axial loading.
- Be able to use measured strength values for a unidirectional composite to predict how ply damage will
develop in a laminate.
- Understand the concept of the fracture energy of a composite and be able to use the software package to
predict the contribution to this from fibre pull-out.
Bibliography
The student is referred to the following resources in this module:
Chavla, K.K., Ceramic Matrix Composites, Chapman and Hall,1993
Clyne, T.W., and Withers, P.J., An Introduction to Metal Matrix Composites, Cambridge University Press,
1993
Hull, D. and Clyne, T.W., An Introduction to Composite Materials, Cambridge University Press, 1996
Piggott, M.R. Load Bearing Fibre Composites, Pergamon Press, 1980
Chou, T.W., Microstructural Design of Fibre Composites, Cambridge University Press, 1992
Harris, B., Engineering Composite Materials, Institute of Metals, 1986
Kelly, A.(Ed), Concise Encyclopaedia of Composite Materials, Pergamon Press, 1994



Content:


- Introduction to Metal Matrix Composites (MMC)


- Theoretical Analysis of MMCs


- Production of Reinforcements


- Production and Applications of MMCs
Metal Matrix Composite:

- Metal matrix composites materials have been active subjects of scientific
investigation and applied research for more than three decades.

- A significant volume fraction of a stiff non-metallic phase in a ductile metal matrix
results in phenomena that are specific to reinforced metals.

- In the past few years, we have seen significant advances in our understanding of
these materials and of phenomena specific to their fabrication and behaviour.

- Scientific investigation have addressed the governing principles of their
processing and general laws have been identified for the influence exerted by the
reinforcement on the microstructural evolution of the matrix.

- Advances in computational mechanics have brought to light practically important
micromechanical phenomena that were often ignored in analytic treatments.

A composite material can be described as a mixture of component materials
designed to meet a specific engineering role by exploiting the desirable properties of
the components, whilst minimising the harmful effects of their less desirable
properties.

- It must be man made

- It must be a combination of at least two chemically distinct materials

- It should be created to obtain the properties which were not otherwise be
achieved by any of the individual constituents

- Metal matrix composites consist of at least two components, one is the matrix
(usually a metal or an alloy) and the second a reinforcement

- The distinction of metal matrix composites from other two or more phase
alloys comes from the processing of the composite

Types of Metal Matrix Composites:

- Dispersion Strengthened: Microstructure consisting of an elemental matrix
within which fine particles are uniformly dispersed. The particle diameter ranges
from 0.01 to 0.1 m and the volume fraction 1 to 15%.

Reinforcing particles are much smaller having diameters between 0.01 and 0.1
m (10 and 100 nm). The mechanism of strengthening is similar to that for
precipitation hardening, the matrix bears the major portion of applied load and the
small dispersed particles hinder or impede the motion of dislocations. Which
results in restricted plastic deformation and leads to improved yield, tensile
strength and as well as hardness of the composite.


- Particle Reinforced: Dispersed particles of greater than 1.0 m diameter
with a vol. frac. of 5 to 40%. There are further two subdivisions of particle
reinforced composites as large particle and dispersion strengthen composites.

The distinction between these is based upon reinforcement or strengthening
mechanism. The hard and stiffer particles tend to restrain movement of the matrix
phase in the vicinity of each particle. The matrix (softer phase) transfers some of
the applied stress to the particles which bear a fraction of the load.The degree of
reinforcement or improvement of mechanical behaviour depends on strong
bonding at the matrix-particle interface.


- Fibre Reinforced: The reinforcing phase in fibre composite materials span the
entire size range and spans the entire range of volume concentrations.

The distinguishing micro-structural feature of fibre-reinforced materials is that the
reinforcing fibre has one long dimension, whereas the reinforcing particles of the
other two types do not.


Basic Mechanical Behaviour:

A stress-strain curve of a metal matrix composite can conveniently represent its
basic mechanical behaviour: stiffness, yield flow and fracture stresses.

The stress-strain curve of a typical continuous fibre metal matrix composite consists
at most of three stages.

First stage both matrix and fibre deform elastically,

Second stage the matrix deform plastically while the fibre remains elastic,

and the third stage both matrix and fibre deform plastically.



The existence of the second and third stages of the stress-strain curve of a metal
matrix composite depends on the types of the matrix metal and fibre.


Short Fibre Metal Matrix Composites:

The stress-strain curve of these compositions can be divided into two classes.

First there is a distinct linear region in the initial portion of the stress-strain curve
corresponding to stage I in the continuous fibre composites followed by a parabolic
stress-strain curve corresponding to stage III in the continuous fibre composites.


Second the stress-strain curve does not have the linear region, the stress-strain
curve are parabolic in shape over the entire range.






Deformation Characteristics of MMC Containing Second Phase Particles:

- The strengthening of a pure metal is carried out by alloying and supersaturating to
a extent that on suitable heat treatment the excess alloying addition precipitate
out (generally known as ageing process).

- To study the deformation behaviour of precipitate hardened alloy or particulate
reinforced metal matrix composites,

the interaction of dislocation with reinforcing particles is much more
dependent on particle size, spacing and density than on the composition.


- When a particle is introduced in a matrix, an additional barrier to the movement of
dislocation is created and the dislocation must behave one of two ways,

- cut through the particles or

- take a path around the obstacles


Mott and Nabarro Theory:

Mott and Nabarro considered an alloy with spherical solute atoms or groups of solute
atoms, which by virtue of their different atomic size from that of the solvent atoms,
caused internal stresses in the matrix.



If the atomic radius of the solvent atom is R
s

then the atomic radius of the solute atom is R
s
(1+u) where u is the misfit parameter
defined as

1
a
da
dc
. (1)


a being the lattice parameter and c the atomic concentration of solute. The same
concept can be applied to groups of solute atoms.



The internal stress fields caused by the elastic strain between the group of particles
and the matrix are an average distance A apart which is the wavelength of the
internal stress field.


From elasticity theory, the shear strain c in the internal stress field at distance l from
the centre of a spherical particle of radius r
o
when l>>r
o
is given as:

c
u
=
r
l
o
3
3
(2)

They then calculated a mean shear strain c
M
from equation 2 by assuming that l
could be defined as half the distance between particles:

l N =

1
2
1
3
(3)

where N is the number of particles per unit volume, so

c u u
M
o o
r N r N = = ( )( )
3
1
3
3 3
2 8 (4)

The concentration of solute C r N
o
o
=
4
3
3
t . The critical shear stress of the
system is defined in terms of the mean elastic strain as,

t c u u
o M
o
o
G G r N G C = = ~ 8 2
3
(5)

This expression is independent of the particle spacing, the yield stress depending
only on the mismatch function and the solute concentration.

The experimental results shows that for incoherent particles the yield stress is
related inversely to the particle spacing.

In the above analysis a rigid dislocation line was assumed but later modification to
the theory introduced the idea of flexible dislocation movement.

The movement would then be dependent on the distance between internal stress
centres.

A dislocation line always tends to reduce its energy by shortening, i-e it tries to
straighten itself. So we can introduce the concept of tension T along the line.

Mott and Nabarro have shown that T~ Gb
2
.

So if we have a curved dislocation, it can only be in equilibrium if acted on by a
stress. Let us assume that t
o
is the stress needed to maintain the dislocation in a
curvature of radius r.









If we consider a small arc of os of a dislocation of strength b. The angle subtended
by the arc at the centre of curvature O is o| = os/r.

There is an outward force along OA due to the applied stress equal to t
o
bos, and an
opposing inward force due to the line tension of


2
2
T T sin
o|
o| ~ (6)

In equilibrium

T b s
o
o| t o = . (7)

so,
t
o|
o
0
= = =
T
b s
T
br
Gb
r
.
(8)

Thus the radius of curvature of a dislocation is inversely proportional to the applied
stress.




Production of Metal Matrix Composites:


Matrices and Reinforcements for MMCs.


Matrices:

- Aluminium alloys

- Copper alloys

- Iron and steel alloys

- Magnesium alloys

- Nickel based alloys

- Titanium alloys

- Zinc alloys

Reinforcements:

Non-metallic

Alumina Boron
Boron carbide Boron nitride
Graphite Niobium carbide
Silica Silicon nitride
Tantalum carbide Titanium boride
Titanium carbide Titanium aluminide
Tungsten carbide Vanadium carbide
Zircon Zirconia
Zirconium

Metal Wires

Molybdenum
Stainless steel
Titanium
Tungsten

Production of a component from MMCs involves:

- bulk production of the composite from its component materials

- secondary working of the composite material into some form

- joining of the composite materials, leading to final fabrication of the engineered
component


Essential requirements for any production route:

- the reinforcement must be distributed in a controlled manner in the metal matrix, i-e either
uniformly distributed throughout or placed in designated locations of the component

- minimal porosity and full density should result in the final component

- typically, volume fractions of 0.1 - 0.6 of reinforcement need to be incorporated in the
matrix

- reactions at the reinforcement/matrix interface should be controlled to promote optimum
bond strength and avoid reinforcement degradation

- the reinforcement should be incorporated into the matrix without breakage. This is a
particularly important factor when processing continuous fibre and whisker reinforced
MMCs
- during composite joining and forming, minimal reinforcement degradation of either
chemical or physical means should result. Reinforcement alignment and distribution
should be maintained

- the route should be as flexible as possible in terms of matrices and reinforcements to
which it can be applied

- subsequent post-fabrication heat treatments should be allowed for

- the route should be capable of producing components with a high degree of
reproducibility at minimum product variability at minimum cost and maximum
productivity

- flexibility in the range of shapes capable of being produced is highly desirable

- any proposed process route should be amenable to scale-up.

Processing of metal matrix composites can be broadly divided into two categories of
fabrication techniques:

Solid state including powder metallurgy and diffusion bonding and

Liquid state A majority of the commercially viable applications are now
produced by liquid state processing technique over solid-state techniques.

Liquid-State Processing:

Liquid-state processing technologies can be divided into four major categories:

- Infiltration

- Dispersion

- Spraying and

- In-situ fabrication
Infiltration Processes:

Infiltration processes involve holding a porous body of reinforcing phase within a mould and
infiltrating it with molten metal that flows through interstices to fill the pores and produce a
composite.

The mechanical properties of any given reinforcement-matrix system are highly dependent on
the matrix microstructure. The rules developed for microstructural control in the
solidification of unreinforced metal are generally not directly applicable to metal matrix
composites because the reinforcing phase influences matrix solidification.

The reinforcement act as a barrier to mass transfer during solidification. This causes an
alloyed matrix to grow in avoidance of the fibres because the fibres present a barrier to solute
diffusion. Thus in most cases the last phase to solidify is found in the vicinity of the
reinforcement surface.

If the interfibre spacing is smaller or comparable to the size of the solidification structure
found in the unreinforced metal, the transitions from plane front to cellular growth and from
cellular to dendritic growth are shifted.

When solidification is dendritic in the unreinforced alloy, ripening of dendrite arms ceases in
the composite when the arm spacing reaches the fibre spacing, further coarsening of the
structure occurs by coalescence.
Matrix microsegregation is also strongly affected by the reinforcement, which places an
upper limit on diffusion distance in the solid phase during solidification.

Thus microsegregation can be reduced by solid-state diffusion during solidification to a much
greater extent than that which occurs in unreinforced metals.

If high pressure is applied during solidification, the matrix exhibits a finer grain than would
be obtained at atmospheric pressure.

Dispersion Processes:

In dispersion processes the reinforcement is incorporated in loose form into the metal matrix.
Due to poor wetting characteristic of metal -reinforcement systems, mechanical force is
required to combine these phases.

The simplest dispersion process is the Vortex method, which consists of vigorous stirring of
the liquid metal and the addition of particles in the vortex.

Mixing of particles and metal can also be achieved while the alloyed metal is kept between
solidus and liquidus temperature. This process is known as Compocasting.

The advantage of using semi-solid metal is the increase in the apparent viscosity of the slurry.

Another method to mix semi-solid metal and particulates is the Thixomoulding process ,
whereby metal pallets and particles are extruded through an injection moulding apparatus.

Critical to the success of dispersion processes is the control over generally undesirable
features such as porosity resulting from gas entrapment during mixing, oxide inclusions,
reaction between reinforcement and metal matrix, particle migration and clustering during
and after mixing.

When solidification takes place, particle migration caused by solidification effects competes
with migration caused by gravity.

Spray Processes:

In these processes droplets of molten metal are sprayed together with the reinforcing phase
and collected on a substrate where metal solidification is completed.

Alternatively the reinforcement may be placed on the substrate and molten metal may be
sprayed onto it.

The critical parameters in spray processing are the initial temperature, size distribution and
velocity of the metal drops, the velocity, temperature and feeding rate of the reinforcement.

Most spray deposition processes use gases to atomise the molten metal into fine droplets.

The particles can be injected within the droplet stream or between the liquid stream and the
atomising gas.

The advantage of spray-deposition techniques resides in the resulting matrix microstructure
that features fine grain size and low segregation.

One of the drawback of the process is the amount of residual porosity and the resulting need
to further process the materials.

Osprey Process:

The osprey process is a rapid solidification spray process developed by Osprey Metals for the
production of both unreinforced and particulate reinforced.

The process involved;

the production of controlled stream of molten metal

the conversion of the stream to spray of molten droplets by means of inert gas atomisation.

impacting of droplets on to a collecting surface and re-coalescence.

injection of reinforcement particles directly into the atomised spray permits MMC
production.

In-Situ Processes:

In-situ processes are the ideal techniques to produce MMC directly in a single step without
the need for intermediate formation of the reinforcement and also a capability to produce near
net shape components.

The reinforcements are formed in-situ in the metal matrix and the surfaces of reinforcements
generated tend to remain clean i-e free from gas absorption oxidation or other detrimental
surface reactions and the matrix-reinforcement interface bond tends to be stronger.


Optimisation of Liquid-Metal Processing:

Reinforcement:

A number of problems associated with liquid-state processes arise from the reinforcement.

In infiltration processes a preform of the reinforcing phase is often prepared prior to
infiltration, by dispersing the reinforcement in a solution containing an inorganic binder,
pressing to the desired volume fraction and drying the cake composed of fibres and residual
binder material.

When short fibres are used, fibre alignment is difficult to achieve, unless the composite is
extruded, which generally damages the fibres.
Solid-State Processing:

- Solid-state processes are generally used to obtain the highest mechanical properties in
MMCs, particularly in discontinuous MMCs.

- Because segregation effects and brittle reaction product formation are at a minimum for
these processes, especially when compared with liquid-state processes.

Powder Consolidation:

- Powder metallurgy is the most common method for fabricating metal-ceramic and metal-
metal composites.

- With the advent of rapid solidification technology, the matrix alloy is produced in a
prealloyed powder form rather than starting from elemental blends.

- After blending the powder with ceramic reinforcement particulates, cold isostatic pressing
is utilised to obtain a green compact that is then thoroughly degassed and forged or
extruded.

- In some cases hot isostatic pressing of the powder blend is required, prior to which
complete degassing is essential.

- Consolidation of matrix powder with ceramic fibres has also been achieved but the
difficulties encountered when attempting to maintain uniform fibre spacing.

- Powder based routes for MMC production tend to be more expensive than liquid-based
routes and therefore generally occupy the more specialist high cost markets for MMCs.

Diffusion Bonding:

- Diffusion bonding is probably the most widely used technique that is applicable to
preforms and filaments. The reinforcements are placed between foils of the matrix metal
and are then subjected to a pressure-temperature treatment. The bond between the matrix
and the reinforcement is made good by inter-diffusion between the two.

- In principle this is a solid-state creep deformation process.

- The pressure-time-temperature requirements for consolidation can be determined from
matrix flow stress.

- To avoid reinforcement degradation care must be exercised to maintain low pressure
during consolidation.

- Thermal expansion mismatch between the fibre and matrix can often causes tensile
stresses and matrix cracking upon cooling from the diffusion bonding temperature.

- Hot roll bonding: metal matrix foils and fibre arrays or preforms are continuously fed
between rollers which apply both heat and pressure. The advantage of the technique is a
short reaction time, which limits the fibre degradation problems.

- Deposition processes for fibre composites generally applies a monotape that can be
stacked in layers and diffusion bonded into a composite.

Some degree of porosity is usually present in the monotape but that can be eliminated during
diffusion bonding or hot isostatic pressing.

Deformation Processing of Metal Matrix Composites:

Secondary processing of the discontinuously reinforced composites leads to break up of
particles, reduction or elimination of porosity and improved particle to particle bonding, all
of which tend to improve the mechanical properties of these materials.

The common methods of secondary working are,

- Extrusion

- Rolling

- Thermomechanical processing

- Superplastic deformation and

- Forging

Diffusion Bonding Processing Route:

Squeeze Casting:

Pressure Casting:

Hot Isostatic Pressing:

Reocasting:

Superplastic Forming:

Metal Matrix Composite Applications

- Metal matrix composites have emerged as a class of materials capable of advanced
structural, aerospace, automotive, electronic, thermal management and wear applications.

- The performance of metal matrix composites is superior in terms of improved physical,
mechanical and thermal properties, although substantial technical and infrastructural
challenges remain.

- The performance advantage of metal matrix composites is their tailored mechanical,
physical and thermal properties that include low density, high specific strength, high
specific modules, high thermal conductivity, good fatigue response, control of thermal
expansion and high abrasion and wear resistance.

- The ability of a metal matrix composite for a cost effective application involves several
factors, including a large material production capacity, reliable static and dynamic
properties, cost effective processing and a change in design philosophy based on
experience and durability.

- The structural potential, particularly for aluminium MMC materials has been
demonstrated, although the result has rarely proven to be cost effective.

- In application-driven environments, MMCs are gaining rapid prominence in the
automotive, electronic and aerospace sectors.

- Higher operating temperatures of materials and wear resistance are mandated by the
increasingly stringent performance requirements of the aerospace and automotive
industry.

- Several industrial efforts have been made to provide a low cost, cast metal matrix
composite, e.g.:

- Development of reinforced pistons and engine blocks using preform infiltration
technology on a mass production basis by Toyota and Honda respectively.

- Spray deposition processes for producing MMC ingots and products by Alcan.

- Production by Lanxide via direct metal oxidation and pressureless infiltration of wear
resistance composites and electronic packaging components.

- Development of large-scale production capacity for commercial particulate composites by
both foundry and powder routes.

- By selecting appropriate reinforcing constituents for matrix material, it is possible to
design alloys that will enhance strength and stiffness.

- Metal matrix composites incorporate a wide variety of metal systems using several types
of reinforcements in the form of whiskers, monofilament, continuous fibres and
particulates.




Applications in the Automotive Industry:


- The automobile industry is currently facing substantial technical challenges as it seeks to
improve fuel economy, reduce vehicle emissions, increase styling options and enhance
performance.

- The increased use of lightweight metals has been considered for high volume production
applications.

- The enhancement of the power to weight ratio demonstrates the viability of aluminium
base engines and MMC cylinder blocks.

- Replacing the grey cast iron liners with an aluminium MMC enables a weight savings of
4.5 kg, with an increased in engine displacement and increased cooling efficiency.

Engine block with MMC liners:

MMC drive shaft:

MMC brake disc:

Forged MMC connecting rod:

MMC pistons:

Aerospace Applications:

- Motivation for metal matrix composites has been driven by defence applications.

- Aerospace applications require higher performance, weight savings, high operating
temperature and competitive life cycle costs.

- For aerospace engine applications include exhaust nozzles, links, blades, cases, shafts.

- Structural aerospace applications include fuselage structures, landing and arresting gears,
stiffeners, drag brakes and compression and torque tubes.


Typical Tutorial: Axial and Transverse Strength of Unidirectional
reinforced composite.

Q.1:

A continuous and aligned glass reinforced composite consists of 40
vol% of glass fibres having a modules of elasticity 69 GPa and 60 vol%
of a polyester resin that, when hardened, displays a modules of 3.4
GPa.

a) Calculate the modules of elasticity of this composite in the
longitudinal direction.
b) If the cross section area is 250 mm
2
and a stress of 50 MPa is
applied in longitudinal direction, calculate the magnitude of the
load carried by each of the fibre and matrix phases.
c) Determine the strain that is sustained by each phase when the
stress calculated in part b is applied.
d) Assuming tensile strengths of 3.5 GPa and 69 MPa for glass
fibres and polyester resin respectively, determine the
longitudinal tensile strength of the composite.
e) If the stress is applied perpendicular to the direction of fibre
alignment, calculate the elastic modules of the composite
described above.


Q.2:

It is desired to produce an aligned and continuous fibre reinforced
epoxy composite having a maximum of 50 vol% fibres. In addition, a
minimum longitudinal modulus of elasticity of 50GPa is required, as
well as minimum tensile strength of 1300 MPa .

Of E-glass, carbon and aramid fibre materials which are possible
candidates and Why?

The epoxy has a modulus of elasticity of 3.1 GPa and tensile strength
of 75 MPa. In addition assume the following stress levels on the epoxy
matrix at fibre failure: E-glass--------- 70MPa; Carbon ----------- 30 MPa ;
and aramid ---------- 50 MPa.

Q.3:

A metal matrix composite is made from a boron fibre reinforced aluminium
alloy. To form the boron fibre, a tungsten wire (r = 10 m) is coated with
boron, giving a final radius of 75 m. The aluminium alloy is then bonded
around the boron fibres giving a volume fraction of 0.65 for the aluminium
alloy. Assuming that the rule of binary mixture applies also to ternary
mixtures, calculate the effective tensile elastic modules of the composite
material under isostrain conditions.

Data:

E
W
= 410 GPa
E
B
= 379 GPa
E
Al
= 68.9 GPa

Q.4:

A metal matrix composite is made with 80% by volume of aluminium alloy
2124-T6 and 20% by volume of SiC whiskers. The density of the 2124-T6
alloy is 2.77g/cm
3
and that of the whiskers is 3.10 g/cm
3
. Calculate the
average density of the composite material.

(Basis is 1 m
3
of material, thus we have 0.80 m
3
of 2124 alloy and 0.20 m
3
of
SiC fibre in 1 m
3
of material).

Q.5:

A metal matrix composite is made of a 2024-Al alloy matrix and continuous
boron fibres. The boron fibres are produced with a 11.5 m diameter tungsten
wire core which is coated with boron to make a final 105 m diameter fibre. A
unidirectional composite is made with 51 vol% of the boron fibres in the
Aluminium matrix. Assuming the law of mixtures applies to isostrain
conditions, calculate the tensile modules of the composite in the direction of
the fibres.

Data:

E
W
= 410 GPa
E
B
= 379 GPa
E
Al
= 72.4 GPa

Q.6:

A new developmental metal matrix composite is made for the national
Aerospace plane with a matrix of the intermetallic compound titanium
aluminide (Ti
3
Al) and continuous silicon carbide fibres. A unidirectional
composite is made with the SiC continuous fibres all in one direction. If the
modules of the composite is 210 GPa and assuming isostrain conditions,
what must be the volume percent of SiC in the composite if

E(SiC) = 390 GPa and E(Ti
3
Al) = 145 GPa.

Q.7:

A metal matrix composite is made with a matrix of 6061-Al alloy and 45 vol %
SiC continuous fibres in one direction. If isostrain conditions prevail, what is
the tensile modules of the composite in the direction of the fibres?

Data:

E(SiC) = 390 GPa and E(6061 Al) = 68.9 GPa.

Q.8:

An MMC is made with a 2024 Al alloy with 15 vol% SiC whiskers. If the
density of the composite is 2.95 g/cm
3
and that of the SiC fibres is 3.10 g/cm
3
,
what will be the density of the 2024 Al alloy?

Q.9:

A ceramic matrix composite is made with continuous SiC fibres embedded in
a glass ceramic matrix.

a) calculate the tensile modules of the composite under isostrain
conditions and

b) calculate the stress o at which the cracks start to grow.

Data:

Glass ceramic matrix:

E = 94 GPA
K
IC
= 2.4 MPa \m
Largest pre-existing flaw is 10 m in diameter.

SiC fibres:

E = 350 GPa
K
IC
= 4.8 MPa \m
Largest surface notches are 5 m deep.

Typical Example of Examination Questions

Q.1:

a) Explain the role of the reinforcement in continuous fibre reinforced
composites. [6 Marks]
b) Briefly discuss the performance of composite materials compared to
conventional engineering materials in terms of operating temperature
and specific strength.
[10 Marks]
c) Describe the filament-winding process. What are the distinct
advantages for this process from an engineering design standpoint?
[9 Marks]
Total 25 Marks


Q.2:

a) Differentiate between a laminate and sandwich panel. In terms of their
construction and performances.
[3 Marks]
b) The rule of mixtures predicts the longitudinal and transverse modulus of the
unidirectional composites.

Explain how the rule of mixtures can be applied to calculate stresses within a
cross ply laminate composite and also comment on the accuracy of the
predictions made by such method of analysis.
[10 Marks]
c) By using the Halpin Tsai relationship estimate the % by which
theequal stress (Reuss) expression from the slab model underestimates the
transverse stiffness of the following materials:

a) Epoxy 40% carbon (HS)
b) Polyester 30% carbon (HS)
c) Aluminium 20% carbon (HS)

Table 2.1
Materials Elastic modulus, GPa
Carbon (HS) 230
Epoxy 3.1
Polyester 3.4
Aluminium 70

[12 Marks]
Total 25 Marks





Q 3:
a) A composite material consists of a continuous glass fibre reinforced epoxy
resin, produced by using 60% (by volume) E-glass fibres and a hardened
epoxy resin.

Calculate
i) The modulus of elasticity, E
c

ii) The tensile strength,
c
and
iii) The fraction of the load carried by fibre, F
f

Data:
Table 3.1
Materials Elastic modulus, GPa Tensile strength, MPa
E-glass 76 345
Epoxy 3.1 75

[12 Marks]

b) Explain how the tensile modulus of the composite, described above, will be
affected by the presence of;

i) aligned short fibres, and [3 Marks]
ii) a range of fibre orientations. [3 Marks]

c) Describe and discuss the role played by a continuous matrix in a fibre
reinforced composite.
[7 Marks]
Total 25 Marks


Q.4:

(a) Briefly discuss the factors that affect the fracture toughness of MMCs.
[10 Marks]

(b) Explain the differences between dispersion strengthened, particle
reinforced and fibre reinforced MMCs.
[9 Marks]

(c) An MMC is made with a 2024 Al alloy with 15 vol% SiC whiskers. If the
density of the composite is 2.95 g/cm
3
and that of the SiC fibre is 3.10
g/cm
3
, what will be the density of the 2024 Al alloys?
[6 Marks]
Total 25 Marks


Q.5:

(a) Explain, why the deformation behaviour of a precipitate hardened alloy
or a precipitate reinforced metal matrix composites depend on
particulate size, shape, spacing and density rather than on composition
of the alloy.
[15 Marks]

(b) A new developmental metal matrix composite is made for the national
Aerospace aeroplane with a matrix of the intermetallic compound
titanium aluminide (Ti
3
Al) and continuous silicon carbide fibres. A
unidirectional composite is made with the SiC continuous fibres all in
one direction. If the modules of the composite is 210 GPa and
assuming isostrain conditions prevail, what must be the volume
percent of SiC in the composite if

E
(SiC)
= 390 GPa and E
(Ti3Al)
= 145 GPa.
[10 Marks]
Total 25 Marks


Q.6:

a) Discuss the fracture toughness of conventional ceramics and explain
the microstructural mechanisms by which this property is improved by
the introduction of reinforcement as ceramic reinforced composites.
[15 Marks]

b) Draw a typical stress strain curve for a continuous fibre and discuss its
shape in terms of deformation behaviour.
[10 Marks]
Total 25 Marks]


Additional Information

Composite Formulas


VOIGT: E
c
= V
f
E
f
+ (1-V
f
) E
m


REUSS: 1/E
c
= (V
f
/E
f
) + ((1-V
f
) / E
m
)

HALPIN - TSAI: E
c
= E
m
(1+nV
f
) / (1-nV
f
)

where;

n = ((E
f
/E
m
) -1) / ((E
f
/E
m
) + 1)

Potrebbero piacerti anche