You are on page 1of 44

3 MICROMECHANICS OF FRP COMPOSITES

3.1 INTRODUCTION
Most of composite structures made of fibrous composites consist of
several distinct unidirectional laminae.
A lamina is a flat or curved arrangement of unidirectional or woven
fibres in a support matrix. The unidirectional lamina (Figure 3.1a) is the basic
building block in a laminated FRP composite material.
A unidirectional composite consists of parallel fibres embedded in a
matrix. The direction parallel to the fibres is called the longitudinal direction
(axis 1 or L) and the direction perpendicular to the fibres in the 1-2 plane is
called the transverse direction. Any direction in the 2-3 plane is also a
transverse direction. These axes are also referred to as the material axes of the
lamina. A similar system of axis can be attached to a FRP reinforcing bar,
Figure 3.1b.
Figure 3.1a A unidirectionally fibre reinforced lamina
72
Transverse direction
Longitudinal direction
(3)
(2) T
(1) L
3(T)
2(T)
1(L)
Figure 3.1b A unidirectionally fibre reinforced round bar
The in-plane elastic behaviour of a unidirectional lamina may be fully
described in terms of four basic lamina properties:
longitudinal modulus (E
1
=E
L
)
transverse modulus (E
2
=E
T
)
shear modulus (G
12
=G
LT
)
the major Poissons ratio (
12
=
LT
)
Many failure theories for composite materials have been proposed up to
now. All theories can be expressed in terms of the basic strength parameters
referred to the principal material axes of the unidirectional lamina (Figure 3.2):
longitudinal tensile strength (f
Lt
)
longitudinal compressive strength (f
Lc
)
transverse tensile strength (f
Tt
)
transverse compressive strength (f
Tc
)
in-plane shear strength (f
LTs
)
The properties of a composite material depend on the properties of its
constituents and their distribution and physical and chemical interactions. These
properties can be determined by experimental measurements but one set of
experimental measurements determines the properties of a fibre-matrix system
produced by a single fabrication process. When any change in the system
variables occur, additional measurements are required. These experiments may
become time consuming and cost prohibitive, therefore a variety of methods
have been used to predict properties of composite materials (Agarwal et al, 2006
Daniel and Ishai, 2006). The mechanics of materials approach is based on
micromechanics.
In most of the composites literature micromechanics means the analysis
of the effective composite properties in terms of constituent material properties.
A complete definition of micromechanics (Lee 1989) is given below:
Micromechanics is a set of concepts, math-models and detailed studies
used to predict composite properties from constituent material properties,
stresses and strains, geometric configuration and fabrication process variables.
The unidirectional composite shows different properties in the material
axes directions. Thus, this type of composites are orthotropic with their axes
1,2,3 as axes of symmetry (Figure 3.1a,b). These unidirectionally fibre
73
reinforced composite elements have the strongest properties in the longitudinal
direction; material behaviour in the other two directions (2,3) is nearly identical
because of the random fibre distribution in the cross section. Therefore, a
unidirectional composite can be considered to be transversely isotropic, that is,
it is isotropic in the 2-3 plane.
3.2 VOLUME AND MASS FRACTIONS
A key element in micromechanical analysis is the characterization of the
relative volume and/or weight content of the various constituent materials. The
mass fractions are easier to obtain during fabrication or using one of the
experimental methods after fabrication; the volume fractions are used in
micromechanics of composites. Therefore it is desirable to determine these
fractions and the relationships between the mass fractions and volume fractions.
Consider a volume v
c
of a composite material which consists of volume v
f
of fibres and volume v
m
of the matrix material. The subscripts c,f and m
represent the composite material, fibres, and the matrix material respectively.
74
Figure 3.2 Lamina loading schemes for basic strength parameters:
a) longitudinal tensile stress (f
Lt
); b) longitudinal compressive stress (f
Lc
);
c) transverse tensile stress (f
Tt
); d) transverse compressive stress (f
Tc
);
e) in-plane shear stress (f
LTs
).
a. b.

1
=
L

1
=
L

1
=
L

1
=
L
c. d. e.

2
=
T

2
=
T

2
=
T

2
=
T

12
=
LT

12
=
LT

2
1
=

T
L

2
1
=

T
L
Let us also consider m
c
, m
f
and m
m
the corresponding masses of the composite,
fibres and the matrix material respectively. Let the volume fraction and the mass
fractions be denoted by V and M respectively. Assuming that no voids are
present in the composite the volume fractions and the mass fractions are defined
as follows:
m f c
v v v +
(3.1a)
c
f
f
v
v
V
and
c
m
m
v
v
V
(3.1b)

m f c
m m m +

(3.2a)
c
f
f
m
m
M
and
c
m
m
m
m
M

(3.2b)
The density
c
of the composite can be obtained in terms of the densities
of the constituents (
f
and
m
) and their volume fractions or mass fractions. The
mass of a composite can be written as:
m m f f c c
v v v +

(3.3)
Dividing both sides of Equation (3.3) by v
c
and using the definition for the
volume fractions, the following equation can be derived for the composite
material density:
m m f f c
V V +

(3.4)
Equation (3.4) for a composite material with two constituents can be
generalized for a composite with n constituents:

n
i
i i c
V
1

(3.5)
The density of composite materials in terms of mass fractions can be
obtained as:
m m f f
c
M M

1
(3.6)
which can also be generalized for n constituents:
75

n
i
i i
c
M
1
1

(3.7)
Considering the definition of mass fractions and replacing the mass by the
product of density and volume, the conversion between the mass fractions and
volume fractions can be obtained:
m
c
m
m f
c
f
f
V M V M


(3.8)
or, in general
i
c
i
i
V M

(3.9)
The reverse relations are
m
m
c
m f
f
c
f
M V M V


(3.10)
and, in general
i
i
c
i
M V

(3.11)
The composite density calculated theoretically from the mass fractions
may not agree with the experimentally determined density. Assuming that the
theoretically calculated density is
ct
and the experimentally determined density
is
ce
the volume fraction of voids V
v
is given by:
ct
ce ct
v
V

(3.12)
The void content may significantly influence some mechanical properties of a
composite material. A good composite must have less than 1% voids, whereas
a poorly made composite can have up to 5% void content (Agarwal and
Broutman, 1990). Higher void contents lead to increased scatter in strength
properties, to lower fatigue resistance and greater susceptibility to water
penetration.
For any number of constituent materials, n, the sum of the constituent
volume fractions must be unity:

n
i
i
V
1
1
(3.13)
and, when the composite material consists of fibres, matrix and voids:
76
1 + +
g m f
V V V
(3.14)
The range of constituent volume fractions that may be expected in fibre
reinforced composites can be determined using representative area elements for
idealized fibre-packing geometries such as the square and triangular arrays
shown in Figure 3.3. If it is assumed that the fibre spacing, s, and the fibre
diameter, d, do not change along the fibre length, then, the area fractions must
be equal to the volume fractions. The fibre volume fraction for the square array
is found by dividing the area of the fibre enclosed in the square by the total area
of square:
2
2
2
4
1
4

,
_


s
d
s
d
V
f

(3.15)
The maximum theoretical fibre volume fraction occurs when s=d. In this case:
785 . 0
4
max

f
V
(3.16)
In case of a triangular array
2
3 2

,
_

s
d
V
f

(3.17)
and, when s=d, the maximum fibre volume fraction is:
907 . 0
3 2
max


f
V

(3.18)
These theoretical limits are not generally achievable in practice. In most
continuous fibre composites the fibre volume fractions range from 0.5 to 0.8.
77
Figure 3.3 Representative area elements for idealized fibre-
packing geometries: a) square; b) triangular
a.
b.
3.3 LONGITUDINAL PROPERTIES OF UNIDIDIRECTIONAL
FRP COMPOSITES
3.3.1 Initial linear elastic behaviour
A unidirectional fibre reinforced composite material subjected to an axial
load parallel to fibres deforms in the following stages (Fig. 3.4) depending on
the relative ductility or brittleness of the fibres and the matrix (Fig. 3.5).
78
*
*
Stage 1 Stage 2
2
Stage 3 Stage 4

?
matrix
Composite
mmmamaterial
fibre
Figure 3.4 Deformation stages of a fibre reinforced
composite material (Jones 1999)












s
t
r
e
s
s
Fibre
Composite,V
f
=50%
Composite,V
f
=25%
Matrix
strain
brittle fibre
ductile fibre
failure
yield stress
Figure 3.5 Stress-strain diagrams for hypothetical fibre
reinforced composites (Agarwal and Broutman 1990)
The four stages that can be identified during the composite loading are:
1-Both the fibres and the matrix deform in a linear elastic fashion.
2-The fibres continue to deform elastically but the matrix now deform
plastically.
3-Both the fibres and the matrix deform plastically.
4-The fibres fracture followed by the composite failure.
Elementary mechanics of materials models have been adopted in the
elastic range, based on the following assumptions:
- A unidirectional composite may be modelled by assuming fibres to be
uniform in properties and diameter, continuous, and parallel throughout the
composite.
- It may be assumed that a perfect bonding exists at the interface, so that
no slip occurs between fibre and matrix materials.
- The fibre and matrix materials are assumed to be homogeneous and
linearly elastic. The matrix is assumed to be isotropic, but the fibre can be either
isotropic or orthotropic.
- Since it is assumed that the fibres remain parallel and that the
dimensions do not change along the length of the element, the area fractions
must equal the volume fractions.
Let us consider the model of the unidirectional composite shown in Figure 3.6.
Since no slippage occurs at the interface and the strains of fibre, matrix and
composite are equal we can write:
1 f

= 1 m

= 1 c

(3.19)
in which subscripts f, m and c refer to fibre, matrix and composite, respectively
and the second subscript refers to the direction.
For the model shown in Figure 3.6 the load (P
c
=
L
A
c
) is shared between
the fibres (P
f
=
f1
A
f
) and the matrix (P
m
=
m1
A
m
).
79
Figure 3.6 Model of FRP composite for predicting longitudinal behaviour
fibre
matrix
l
c

c
L
(2) T
(1) L
Static equilibrium requires that the total force on the lamina cross section
must equal the sum of the forces acting on the fibre and matrix:

L
A
c
=
A A A
m f f c c 1 1 1
+
m
(3.20)
Since the area fractions are equal to the corresponding volume fractions,
Equation 3.20 can be rearranged to give an expression for the composite
longitudinal stress:
m m f f c L
V V +
1 (3.21)
Equation (3.21) can be differentiated with respect to strain, which is the same
for the composite, fibres and matrix:
m
m
m
f
f
f
L
c
c
V
d
d
V
d
d
d
d

,
_

,
_

,
_

(3.22)
where (d/d) represents the slope of the corresponding stress-strain diagrams at
the given strain. If the stress-strain curves of the materials are linear, the slopes
(d/d) are constants and they can be replaced by the corresponding elastic
modulus in Equation (3.22). Thus:
m m f f c L
V E V E E E +
1
1 (3.23)
Relationships (3.21) and (3.23) are known under the name rule of mixtures
indicating that the contributions of the fibres and the matrix to the composite
stress and elastic modulus respectively are proportional to their volume
fractions.
In Equation (3.23) it is assumed that the fibre can be anisotropic with
different properties in the longitudinal and transverse directions and that the
matrix is isotropic. For example aramid and carbon fibres are anisotropic
whereas glass is practically isotropic. Carbon and aramid fibres are orthotropic
and they have very different values of longitudinal modulus (E
fl
) and transverse
modulus, E
ft
. The ratio E
fl
/E
ft
=24 for Kevlar, 15.3 for high strength carbon and
65 for high modulus carbon (Gay et al 2007). The matrix modulus does not need
a second subscript. The rule of mixtures predictions for the longitudinal elastic
modulus is very close to the experimental results.
Equations (3.21) and (3.23) can be generalized for n constituents as:
( )

n
i
i i c
V
1

(3.24)
80
and

n
i
i i L
V E E
1
(3.25)
Equation (3.23) can be also written as:
) 1 (
1
f m f f L
V E V E E +
(3.23a)
and it predicts a linear variation of the longitudinal modulus E
L
with fibre
volume fraction as shown in Figure 3.7
3.3.2 Behaviour beyond initial deformation
The rule of mixtures accurately predicts the stress-strain behaviour of a
unidirectional composite subjected to longitudinal loads provided that Equation
(3.21) is used for the stress and Equation (3.22) for the slope of the stress-strain
curve. However, the replacement of slopes by the elastic moduli is possible only
when both constituents deform elastically. This may constitute only a small
portion of the stress-strain behaviour and is primarily applicable for glass fibre
reinforced thermosetting plastics. In general, the deformation of a fibre
reinforced composite may proceed in the four stages shown in Figure 3.4. Stage
2 usually occupies the largest portion of the composite stress-strain curve, and in
this period of development the matrix stress-strain curve is no longer linear, so
that the composite modulus must be predicted at each strain level by:
81
Figure 3.7 Variation of predicted E
L
with
fibre volume fraction
E
L
E
m
E
f
0 0.25 0.5 0.75 1.0
V
f
m
m
m
f f L
V
d
d
V E E

,
_

1
(3.26)
where (d
m
/d
m
) is the slope of stress-strain curve of the matrix at the
corresponding strain of the composite. The stress-strain diagrams for
hypothetical composites with ductile and brittle fibres are shown in Figure 3.5. It
can be observed that the stress-strain curves of a composite fall between those of
the fibre and the matrix. However if the fibres are capable of deforming
plastically within the matrix, the fracture strain of fibres in the composite may
be larger than the fracture strain of the fibres working separately. Thus the
fracture strain of the composite may exceed that of the fibres, (Figure 3.5).
3.3.3 Longitudinal tensile strength
When a fibre reinforced composite is subjected to longitudinal tension the
constituent with the lower ultimate strain will fail first. Under assumption of
uniform strengths, two cases are distinguished depending on the relative
magnitudes of the ultimate strains of fibres and matrix. When the ultimate
tensile strain of the fibre is lower than that of the matrix (
fu
<
mu
) the composite
will fail when its longitudinal strain reaches the ultimate strain in the fibre,
(Figure 3.8a). Then, the longitudinal tensile strength of the composite can be
calculated with:
) 1 (
f f ft Lt
V V f f
m
+
(3.27a)
where
f
Lt
= longitudinal composite tensile strength
f
ft
=longitudinal fibre tensile strength
m

= average matrix stress at the fibre fracture strain (Fig. 3.8a)


V
f
= fibre volume fraction
Expressing
m
by the product of the matrix elastic modulus times the
corresponding strain the following formula can be utilized to determine the
composite longitudinal tensile strength:
m
Lt ft f f
fl
E
f f [V (1 V )]
E
+
(3.27b)
If the fibre volume fraction is below the so called V
min
the matrix is able to
support the entire composite load when all the fibres break. After that, the
matrix takes additional load with increasing strain. It is assumed that the fibres
do not support any load at composite strains higher than the fibre fracture strain.
The composite eventually fails when the matrix reaches its ultimate tensile
82
strength (f
mt
). Thus the ultimate strength of a composite with the fibre volume
fraction less than V
min
is given by:
) 1 (
f mt Lt
V f f
(3.28)
By equating right-hand sides of Equations (3.27) and (3.28) the value V
min
can be
determined:
By equating right-hand sides of Equations (3.27a) and (3.28) the value V
min
can
be determined as :
m mt ft
m mt
f
f f
f
V V


min
(3.29a)
or utilising Equation 3.27b :
( )
( )
1
mt ft m fl
min
mt ft m fl
f f E / E
V
f f E / E

1
+
]
(3.29b)
Figure 3.9 illustrates the longitudinal composite strengths plotted against fibre
volume fraction. The solid portions of the lines represent the range of their
applicability and their intersection defines V
min
. It can be noticed that Equation
(3.28) predicts composite strength that is always less than the strength of
83

fu

mu
Figure 3.8 Longitudinal stress-strain curves for composite and constituents
a. fibre dominated strength (
fu
<
mu
)
b. matrix dominated strength (
mu
<
fu
)
a. b.
f
ft
f
mt

f
fibre
composite
matrix
f
ft
f
Lt
f
mt

m

fu

mu
fibre
composite
matrix
stress
strain strain
stress
f
Lt
nonreinforced matrix, whereas Equation 3.27 predicts composite strength that
can be higher or lower than the matrix strength depending on the fibre volume
fraction. In Figure 3.9 the fibre volume fraction can be utmost 0.785 in case of
square fibre array and up to 0.907 for triangular fibre packing.
A critical fibre volume fraction, V
crit
, which must be exceeded for
strengthening can be defined as follows:
mt f m f ft Lt
f V V f f + ) 1 (
(3.30)
m ft
m mt
crit f
f
f
V V


(3.31a)
or
( )
1
( / )
/
mt ft m fl
cr
ft m fl
f f E E
V
f E E

]
(3.31b)
The reinforcing fibres begin to act effectively only when V
f
exceeds V
crit
,
a condition that is important to understand when composites are designed,
(Gerdeen and Rorrer, 2012).
In polymeric composites V
crit
and V
min
are very small because most polymers
exhibit only a limited amount of plastic flow and strain hardening. For example
in case of a glass-fibre reinforced epoxy composite V
min
would range between
0.25% and 1% a fibre volume fraction much lower than usual fibre content
(Agarwal et al, 2006).
When the ultimate matrix tensile strain is lower than that of the fibre
(
mu
<
fu
) the composite fails when its longitudinal strain reaches the fracture
strain of the matrix (Fig. 3.8b). Then, the longitudinal tensile strength of the
composite can be calculated with:
) 1 (
f mt f f Lt
V f V f +
(3.32a)
Expressing
f
by the product of the longitudinal fibre elastic modulus times the
corresponding strain, (
mu
), the following formula can be established to
determine the composite longitudinal tensile strength:
1 1
f
Lt mt f
m
E
f f V
E
1 _
+
1
,
]
(3.32b)
The previous analysis is based on the assumption that all fibres are continuous,
have the same strengths, and are all fractured at the same longitudinal position.
The results above, Equations (3.27) and (3.32), do not consider the statistical
distribution of the constituent strength. In the case of fibre-dominated strength,
fibre strength varies from point to point and fibre to fibre, and a certain fibre
84
may break at a weak point, generating a nonuniform state of stress around the
fibre break (Figure 3.10). The effect of the fibre break on adjacent fibres is an
increase in fibre stress and in interfacial shear stress in the failure region. The
fibre stress increases from zero at the fibre end to the nominal far field value
after a characteristic value (l
1
) from the break. Thus the broken fibre becomes
ineffective on a length 2l
1
. Different failure mechanisms may develop,
depending on the properties of the constituents:
-transverse matrix cracking when composites have a brittle matrix and
strong interface (Figure 3.11a);
-fibre-matrix debonding in the case of weak interface and/or relatively
high ultimate fibre strain (Figure 3.11b);
-conical shear failures in matrix in the case of a relatively ductile matrix
and strong interface (Figure 3.11c).
As the longitudinal tension force increases, the fibre breaks increase in density,
these localized failure interact eventually leading to catastrophic fracture,
(Figure 3.12). In the case of brittle matrix composites, the ultimate strain of the
matrix is lower than that of the fibres and failure starts with the development of
multiple matrix cracks. These cracks produce local stress distributions and high
interfacial shear stresses, initiating fibre/matrix debonding and fibre breaks. The
failure mechanisms consist of transverse matrix cracks, fibre breaks and fibre
pull-out.
85
Figure 3.9 Variation of longitudinal tensile strength with
fibre volume fraction
V
f
V
crit
V
min
f
mt
f
Lt

f
mt
f
ft
fu

f
ft
) 1 (
f m f ft Lt
V V f f +
) 1 (
f mt Lt
V f f
Tensile
longitudinal
strength
Fibre volume fraction
Figure 3.10 Stress distribution around a fibre break
Figure 3.11 Tensile failure mechanisms in a unidirectional composite
loaded in tension
3.3.4 Longitudinal compression
86
When fibre reinforced composite materials are loaded in longitudinal
compression the models for tensile longitudinal strength cannot be used since
the failure of the composite is, in many cases, associated with microbuckling or
kinking of the fibre within the restraint of the matrix material.
Accurate experimental values for the compressive strength are difficult to
be obtained and they are highly dependent on specimen geometry and the testing
method. There are three main longitudinal compression failure modes (Jones
1999):
-microbuckling of fibres in either extensional or shear mode
-shear failure of fibres without buckling
-transverse tensile fracture due to Poisson strain
A fibre buckling in the matrix can be thought as a column on an elastic
foundation. For such a model the buckle wave length can be shown to be
proportional to the fibre diameter. Fibre buckling can also be caused by
shrinkage stresses developed during curing of the composite material. These
shrinkage stresses result from the matrix having a higher coefficient of thermal
expansion (CTE) than the fibres. Two modes of fibre buckling are possible in
the representative volume element, Figure 3.13. In the first mode the fibres can
buckle out of phase relative to one-another, to give the so called extensional
buckling mode (Figure 3.13) in which the matrix is extended or compressed in
the y direction.
87
In the second mode, called the shear mode, the fibres buckle in phase with one
another, Figure 3.13c, and the matrix shears in the x-y plane with all the shear
being relative to x axis. The fibres can be considered much stiffer than the
matrix (G
f
>>G
m
) and the fibre shear deformations may be neglected. Two-
dimensional models were used, with the fibres represented as plates separated by
matrix blocks.
The buckling load of a fibre surrounded by a supporting matrix material is
higher than if there is no matrix material around the fibre, since the lateral
support of the continuous matrix material has an effect similar to increasing the
number of discrete lateral supports for an Euler column.
This effect leads to a buckling load:
EI
l
m
N
cr 2
2 2

(3.33)
in which E is the elastic modulus of the fibre material, I is the fibre moment of
inertia, l is the fibre length, and m is the number of half sine waves (Figure
3.14). The buckling load depends on the number of lateral supports (m-1), and
the buckling load is much larger than if m=1 (column without lateral support).
88
Figure 3.12 Failure evolution in unidirectional composite
under longitudinal tension

L
To find the fibre buckling load in each buckling mode the energy method can be
utilised (Timoshenko and Gere 1961, Jones 1999). Using the energy method, the
work done by external forces (W) is equal to the corresponding change in strain
energy of fibres ( U
f
) plus the change in the strain energy of the matrix (
U
m
):
W U U
m f
+
(3.34)
In the energy method the buckling loads are calculated with Equation (3.34)
using deflection configurations approximated for the various buckle modes.
89
Figure 3.14 Buckling of an Euler column with discrete supports
a. b.

L
N
cr
N
cr
m=1
m=2 m=3
Figure 3.13 Modes of fibre buckling
a-representative volume element; b-extension mode; c-shear mode
a.
b. c.

L

L

L

L

L

L
2c d
L
x
y
Usually the unknown buckling transverse displacement can be represented by
the Fourier sine series. In case of extension buckling mode a sinusoidal buckled
shape is assumed and the following formula can be developed for the fibre
critical stress:
) 1 ( 3
2
f
f m f
fcr
V
E E V


(3.35)
in which V
f
is the fibre volume fraction.
The maximum compressive stress in the composite material is:
) 1 ( 3
2
max
f
f m f
f Lc c
V
E E V
V f


(3.36)
In Equation (3.35) it is assumed that the matrix is unstressed in the x-
direction. The strain at buckling in the x-direction can be calculated using the
expression of
cr
from Equation (3.35):
f
m
f
f
fcr
E
E
V
V
) 1 ( 3
2


(3.37)
Assuming that the matrix has the same strain in the fibre direction as the fibre
(
m
=
fcr
) the following equations can be written:
fcr m m
E
(3.38)
fcr m f fcr f c
E V V ) 1 (
max
+
(3.39)
( )
fcr
f
m
f f c
E
E
V V
1
1
]
1

+ 1
max (3.40)
( )
) 1 ( 3
1 2
max
f
f m f
f
m
f f c
V
E E V
E
E
V V

1
]
1

+
(3.41)
For high ratios E
f
/E
m
the difference between Equations (3.36) and (3.41) are
insignificant. When the shear buckling mode occurs (Figure 3.13c) the fibre
displacements are equal and in phase with one another. The matrix material is
alternately sheared in one direction and then the other. It can be assumed that the
changes in deformations in the y direction are negligible and the shear strains
can be considered to be only a function of the fibre direction coordinate. The
90
change in the strain energy can be expressed in terms of matrix shear properties
and the following formula for the fibre buckling stress is determined:
( )
f f
m
fcr
V V
G

(3.42)
and the maximum composite stress is:
f
m
Lc c
V
G
f


1
max

(3.43)
while the fibre strain at buckling is determined with:
1
]
1

f
m
f f
cr
E
G
V V ) 1 (
1

(3.44)
Equations (3.36) and (3.43)
( )
f
f m f
f Lc
V
E E V
V f

1 3
2
and
f
m
Lc
V
G
f

1
are plotted in Figure 3.15 for a glass-epoxy composite material. It can be noticed
that the shear mode has the minimum strength for the composite over a wide
range of fibre volume fraction. It can also be observed that the extensional mode
governs the compressive strength for low fibre volume fractions (V
f
=0.10.2)
and is not important for practical composites. The predicted strength should be
below the curve labelled elastic shear mode in Figure 3.15. The inelastic
shear mode curve is obtained by replacing the elastic matrix shear modulus in
Equation. (3.43) by a shear modulus that varies linearly from the elastic value at
1% strain to a zero value at 5% strain (Figure 3.16) but the predictions are still
too high. Reasonable predictions of compressive strength for graphite/epoxy
composites have been obtained on a model including the effects of material
nonlinearity and the effects of initial fibre curvature.
The predicted compressive strengths are now closer to the actual values
but they are still higher than the real ones. An explanation may be that the
analysis of the buckling problem has been performed two dimensionally instead
of the actual three-dimensional buckling problem. The influence of the matrix
shear modulus reduction due to inelastic deformation is illustrated in Figure
3.17. In this figure the composite material strain at buckling versus fibre volume
fraction is plotted.
91
The results are calculated from Equation (3.44) for two values of the ratio
E
f
/G
m
and the matrix Poissons ratio 0.25. Again the shear mode governs the
material behaviour for the most range of the fibre volume fractions.
Experimental work carried out on boron/epoxy composite materials have shown
that theory agrees quite well with experimental data if the matrix moduli in
Equations (3.36) and (3.43) are multiplied by 0.63.
Another possible failure mode under longitudinal compression is the
failure of fibres in direct shear due to maximum shear stress. This occurs at an
angle =45
o
to the loading axis, Figure 3.18. At the highest values of V
f
for well
aligned fibres pure compressive failure, which can be related to shear failure of
the fibres, may be encountered. In case of the shear mode governed by the shear
strength of the fibre, the predicted strength is:
92
Figure 3.16 Variation of shear
modulus with shear strain
1% 5% 0
G
me
Matrix G
m
shear
modulus
Figure 3.15 Compressive strength of glass-epoxy composite materials
a-extension mode; b-shear mode
0.2 0.4 0.6 0.8 1.0
Lc
f

L
a
b
elastic
inelastic
V
f
] ) 1 ( [ 2
f
m
f f fs Lc
E
E
V V f f +
(3.45)
in which f
fs
is the shear strength of the fibres.
93
Figure 3.17 Compressive strain
at microbuckling for fibre
reinforced composite materials
0 0.2 0.4 0.6 0.8 1.0
1.0
0.5
0.1
0.05
50
m
f
G
E
100
m
f
G
E
extension
shear

fcr
V
f


Compressive
strain at
microbuckling

m
=0.25
Good agreement with experimental data has been reported for graphite/epoxy
composite when the maximum shear stress is given by a rule of mixtures, so that
the compressive strength is:
) ( 2
m LTsm f LTsf Lc
V f V f f +

(3.46)
where f
LTsf
and f
LTsm
are the shear strengths of fibre and matrix respectively in the
plane LT.
A model of failure under longitudinal compressive loading is based on
the transverse tensile fracture due to Poisson strains (Figure 3.19).
The longitudinal compressive stress
L
produces the longitudinal strain:
L
L
L
E


(3.47)
where E
L
is the longitudinal modulus of the composite in compression.
Under the compressive longitudinal stress, the transverse Poisson strain is:

L LT T

L
L
LT
E

(3.48)
Compressive failure of a unidirectional fibre reinforced composite loaded
in the fibre direction may be caused by transverse splitting of the material.
94
The transverse tensile strain resulting from the Poisson ratio effect can exceed
the ultimate strain capability of the composite resulting in longitudinal cracks at
the interface. At failure L is the ultimate compressive strength (f
Lc
) such that
LTL
equals the ultimate transverse tensile strain (
Tu
) of the composite:
Tu
L
L
LT
E


(3.49)
Tu
LT
L
Lc
E
f

(3.50)
The ultimate transverse strain of the composite can be calculated from the
ultimate tensile strain (Agarwal and Broutman 2006) of the matrix (
mu
):
) 1 (
3 / 1
f mu Tu
V
(3.51)
and the longitudinal compressive strength of the composite is:
) 1 (
) 1 )]( 1 ( [
3 / 1
f m f f
mu f f m f f
Lc
V V
V V E V E
f
+
+

(3.52)
Experimental results are in better agreement with predictions of Equation
(3.52) than with the predictions based on microbuckling of fibres. The
95
Figure 3.18 Shear failure without fibre
buckling of a unidirectional composite
Figure 3.19 Transverse tensile rupture
due to Poisson strains

L
L

agreement is particularly good when


mu
=0.5% which is a reasonable value for
the epoxy resin.
A number of other factors have been shown to affect longitudinal
compressive strength, among them the fibre/matrix interfacial strength, in case
of transverse tensile rupture due to Poisson strain. Other experiments have
shown that the compressive strength of graphite/epoxy is strongly related to the
interfacial shear strength, which also depends on the fibre surface treatments.
3.4 TRANSVERSE STIFFNESS AND STRENGTH OF
UNIDIRECTIONAL COMPOSITES
3.4.1 Transverse modulus
The transverse modulus is a matrix-dominated property being sensitive to
the local state of stress. In the case of transverse loading the state of stress in the
matrix surrounding the fibres is complex and more affected by interaction from
neighbouring fibres. Approaches leading to the mechanics of materials
predictions are based on simplified stress distributions and do not yield accurate
results. Let us consider a simple mathematical model shown in Figure 3.20. As
in the previous section the fibres are assumed to be uniform in properties and
diameter, continuous and parallel throughout the composite. The composite is
represented by a series model of matrix and fibre elements, and the main
assumption is that the stress is the same in the fibre and matrix. Both
constituents are assumed to be linear-elastic materials and the fibre-matrix
bond is perfect. Considering the model made up of layers representing fibres and
matrix materials it is clear from Figure 3.20 that each layer has the same area on
which load acts, experiencing the same stress. The cumulative thicknesses of the
fibre and matrix layers are proportional to their respective volume fractions
because each layer is assumed to be uniform in thickness. In this case, the
composite transverse elongation ( cT

) is the sum of the fibre ( f

) and matrix (
m

) elongation respectively. The elongation of each constituent can be written


as the product of the strain and its cumulative thickness:
m m f f c cT
m m mT f f fT c cT cT
mT fT cT
l l l
l l l


+

+
; ;
(3.53)
96
Since the dimensions of the representative volume element do not change
along the longitudinal direction, the length fractions must be equal to the volume
fractions:
t l
t l
V
c
f
f

t l
t l
V
c
m
m

(3.54)
Assuming the fibres and matrix to deform elastically and the stress is the same
in the fibre, matrix and composite, in the transverse direction, we can write:
( )
m
m
m
f
f
f
T
T c
V
E
V
E E


+
(3.55)
and:

T
E
1
m
m
f
f
E
V
E
V
+
(3.56)
which is the inverse rule of mixtures for the transverse modulus.
Equation (3.56) can be also written as:
m f f m
m f
T
V E V E
E E
E
+

(3.57)
where E
f
is the transverse modulus of the fibres.
For a composite made from n different constituents Equation (3.56) can be
generalised as :
97
fibre
Figure 3.20 Model of a unidirectional composite under transverse normal stress
matrix
l
c

m
t
l
f
l
m

T
( )

n
i
i i
T
E V
E
1
1
(3.58)
The transverse modulus of a unidirectional composite as predicted by
Equation (3.57) is plotted in Figure 3.21 as a function of fibre volume fraction.
The longitudinal modulus as predicted by the rule of mixtures it is also shown in
the same figure.
It can be noticed that the fibres do not contribute much to the transverse
modulus unless the fibre volume fraction is very high. This is in sharp contrast
to the effect of fibres on the longitudinal modulus. Theoretically, the transverse
modulus can be raised to seven times the matrix modulus by providing 90%
fibres, which is not practical.
The model utilised to determine the transverse modulus is not
mathematically rigorous. In a real composite the parallel fibres are dispersed in
the matrix material in a random fashion; generally both constituents will be
present at any section perpendicular to the load, especially at the higher volume
fraction. Thus the load is shared between the fibres and the matrix and the
assumption that the stresses and the matrix are equal is inaccurate and the
mechanics of materials prediction underestimates the transverse modulus,(Kaw,
2006).
Halpin and Tsai developed (Halpin and Tsai 1967) semiempirical equations to
match the results of more exact micromechanics analyses.
98
m
T L
E
E or E
28
24
20
16
12
8
4
E
m
V
f
0 0.25 0.5 0.75 1.0
E
L
E
T
E
f
=30Em
Figure 3.21 Variation of E
L
and E
T
as a function of fibre volume fraction
f
f
m T
V
V
E E
1
1 1
1
1


(3.59)
where:
( )
( )
1
1
1

m f
m f
E E
E E
(3.60)
and
1
is the reinforcing efficiency factor for transverse loading. Its value
depends on the fibre geometry, packing geometry and loading conditions.
The prediction above tends to agree with experimental results for values of

1
=1.0 to 2.0. For usual case of circular-section fibres, satisfactory results are
obtained by taking
1
=2. Predictions of the Halpin-Tsai equations for transverse
modulus of a unidirectional composite are shown as a function of fibre volume
fraction in Figure 3.22 for different constituent modulus ratios.
The transverse modulus of a unidirectional composite is much smaller than its
longitudinal modulus. An increase in fibre volume fraction results in the
increase of transverse modulus similar to the longitudinal modulus, while an
increase in fibre modulus does not have a significant influence on the transverse
modulus. When =0, the Halpin-Tsai equation reduces to the inverse rule of
mixtures, whereas a value of =

yields the rule of mixtures.


3.4.2 Transverse tensile strength
The transverse tensile loading is the most critical loading of a
unidirectional composite. Many factors influence the transverse tensile strength
and the most important are: the matrix strength, the fibre-matrix interface
properties, and defects in matrix such as microcraks and voids. In case of
99
12
8
4
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
m
T
E
E
V
f
Figure 3.22 Predictions of Halpin-Tsai equation for transverse
modulus of a unidirectional composite
a
b
c
d
a.
b.
c.
d.
transverse loading, the high-modulus fibres act as effective constraints on the
deformation of the matrix, causing stress and strain concentrations in this
constituent and at the fibre-matrix interface, where the critical stresses and
strains usually occur (Gibson 1994). The maximum stress in the matrix for a
square fibre array is the axial stress at the interface along the loading direction,
Figure 3.23.
The stress concentration factor (k

) is defined as the ratio of the maximum


internal stress to the applied average stress. Its value depends on the relative
properties of the constituents and their volume fractions:
] / 1 [ ) / 4 ( 1
] / 1 [ 1
2 / 1
f m f
f m f
E E V
E E V
k

(3.61)
In Figure 3.24 the variation of the stress concentration factor for two
polymeric composites is illustrated. Knowing the value of k

the composite
transverse strength, f
Tt
, can be predicted dividing the tensile matrix strength, f
mt
,
to the stress concentration factor, k

.
100
x
y

2
=
T

2
=
T

y
=
r

x
=

x
=

Figure 3.23 Local stresses in transversely loaded


unidirectional composites
Another characteristic quantity for transversely loaded composites is the
strain concentration factor (k

) (Daniel and Ishai 2006) defined as the ratio of


the maximum strain (
max
) to the average strain (
2
). A formula to calculate k

is
given below:
( ) ( )
( )
m
m m
m
E
E
k k

,
_


1
2 1 1
2
2
max
(3.62)
where max

and
2

are the maximum and average strains, respectively and


m

the matrix Poissons ratio. In the formula above it was assumed there is a
perfect bond between the constituents and the fibres are much stiffer than the
matrix. The strain concentration factor, also termed strain magnification factor
(Agarwal and Broutman, 2006) can also be determined with:
] / 1 [ ) / 4 ( 1
1
2 / 1
f m f
E E V
k

(3.63a)
A mechanics of materials approach based on the model shown in Figure 3.25
leads to another formula (Gibson 1994) for the strain concentration factor:
101
Figure 3.24 Stress concentration in matrix of unidirectional composites
with square fibre array under transverse tensile loading
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
3.5
3.0
2.5
2.0
1.5
1.0
0.5
V
f

















S
t
r
e
s
s

c
o
n
c
e
n
t
r
a
t
i
o
n

f
a
c
t
o
r
,

k

glass/epoxy
carbon/epoxy

2
=
T

2
=
T
1 1
1
+
1
1
]
1

f
m
E
E
s
d
k

(3.63b)
where d is the fibre diameter and s the distance between the fibre centres.
In Figure 3.26 the variation of the strain concentration factor of a unidirectional
composite is illustrated. It can be seen that k

increases sharply for fibre volume


fractions exceeding 0.5.
102
s
d

2
Figure 3.25 Mechanics of materials model for strain
concentration factor

2

The value of the transverse failure strain (
cT
) due to the strain
concentration in the matrix around the fibres is:

k
mT
cT

(3.64)
where
mT
=
m
is the matrix tensile failure strain (matrix is assumed
isotropic). If a linear behaviour to failure can be assumed, the corresponding
composite transverse strength can be determined with:
103
Figure 3.26 Variation of strain concentration factors (k

and k

) of a
unidirectional composite with fibre volume fraction
s
d

2
0 0.2 0.4 0.6 0.8
20
16
12
8
4
Fibre volume fraction, V
f





















S
t
r
a
i
n

c
o
n
c
e
n
t
r
a
t
i
o
n

c
o
e
f
f
i
c
i
e
n
t
,

k



o
r

k

k E
f E
E f
m
mt T
T cT Tt


(3.65)
An empirical approach (Nielsen 1974) for the prediction of transverse tensile
strength of fibrous composites is given below:
) 1 (
3 / 1
f
m
mt T
Tt
V
E
f E
f

(3.66)
The preceding equations above assume perfect adhesion between phases and
thus failure occurs by matrix fracture at or near the interface. A reduction
coefficient (C
v
) to account for voids can be used to modify Equation (3.66). C
v
can be determined with:
) 1 (
4
1
f
v
v
V
V
C

(3.67)
where V
v
is the void volume fraction.
Another empirical formula (Barbero 1999) can also account for voids:
1
1
]
1

,
_

+
f
m
f f v mt Tt
E
E
V V C f f 1 ) ( 1
(3.68)
The effect of voids is very detrimental to the transverse strength and this is
reflected by both empirical formulas. Although the results provided by these
formulas can be used for preliminary design, experimental data are usually
required if transverse strength is the controlling mode of failure of the
component. Failure of a transversely loaded composite is strongly influenced by
the residual stresses and strains caused by matrix curing or thermal stresses and
strains due to thermal expansion mismatch (Daniel and Ishai 2006). Assuming a
linear elastic behaviour of the matrix to failure, the maximum tensile stress or
strain failure criterion can be used to predict the tensile strength for a
unidirectional composite:
a)
[ ]
rm mt Tt
f
k
f


1
(3.70)
for the maximum tensile stress criterion and
104
b)
) (
) 2 1 )( 1 (
1
m rm mt
m m
m
Tt
E f
k
f

(3.71)
for the maximum tensile strain criterion, where
rm
and
rm
are the maximum
residual stress and residual strain respectively.
Failure of unidirectional composites subjected to transverse tensile loads occurs,
in most cases, because of matrix or interface tensile failure. In some cases they
may fail by fibre transverse tensile failure if the fibres are highly oriented and
weak in the transverse direction.
Therefore the unidirectional composite failure modes under transverse tensile
loads may be described as: matrix tensile failure, constituent debonding and
fibre splitting. Failure takes the form of isolated interfacial microcracks
increasing in number as load increases and finally coalescing into a catastrophic
macrocrack, Figure 3.27.
105
Figure 2.27 Failure stages of a unidirectional composite
under transverse tension

2

2

2

2

2

2

2

2
3.4.3 Transverse compression
A unidirectional fibre reinforced composite subjected to transverse
compressive load, Figure 3.28, may fail by shear failure of the matrix or matrix
shear failure with constituent debonding and/or fibre crushing. The failure
modes are illustrated in Figure 3.29, where some portions of the failure surface
are created by constituent debonding.
The transverse compressive strength is lower than the longitudinal
compressive strength, but the comments made in the previous case are also
valid. For the most frequent failure mechanism the predicted transverse
compressive composite strength, f
Tc
, is:
106
Figure 3.29 Failure of unidirectional
composite under compressive load

2
Figure 3.28 Unidirectional composite
under transverse compression

k
f
f
mc
Tc

(3.72)
where f
mc
is the compressive strength of the matrix and k

is the stress
concentration factor discussed in the previous section. When the maximum
residual stresses at the interface,
rm
, are taken into account, Equation (3.72)
becomes:

k
f
f
rm mc
Tc
+

(3.73)
Transverse compressive strength values are higher than tensile strength values
for both matrix and composite. Also the transverse compressive strength
increases with increase in the fibre volume fraction. This is explained by the
additional constraints placed on the matrix, preventing its deformation in the
direction perpendicular to the plane of load-fibre axes.
3.5 SHEAR STIFFNESS AND STRENGTH OF UNIDIRECTIONAL
COMPOSITES
3.5.1 In-plane shear modulus
The behaviour of unidirectional composites under in-plane shear loading
is dominated by the matrix properties and the local stress distributions. The
mechanics of materials approach uses a series model under uniform shear stress,
Figure 3.30 to determine the shear modulus. Using the notations shown in the
figure, the total shear deformation of the composite,
c
, is the sum of the shear
deformations of the fibre,
f
, and the matrix,
m
; each shear deformation can be
then expressed as the product of the corresponding shear strain (
c
,
f
,
m
) and the
cumulative widths of the material(l
c
, l
f
, l
m
):
107
Figure 3.30 a) Model of unidirectional composite for prediction of shear
modulus; b) shear deformations for constituents and for the model
t
l
f
l
m
m
f

LT

TL

TL

LT

a.
b.
L
T
l
c

m
m f c
+
(3.74)
m m f f c c
l l l +
(3.75)
Dividing both sides of Equation (3.75) by l
c
and recognising that the width
fraction is proportional to volume fractions, yields:
m m f f c
V V +
(3.76)
Assuming linear shear stress-shear strain behaviour of fibres and matrix, the
shear strains can be replaced by the ratios of shear stress and the corresponding
shear modulus:
m
m
m
f
f
f
c
LT
LT
l
G
l
G
l
G

+
(3.77)
where G
LT
is the in-plane shear modulus of the composite, G
f
is the shear
modulus of fibres and G
m
the shear modulus of matrix. But the shear stresses are
equal on composite, fibres and matrix and from Equation (3.77) we obtain:
m
m
f LT
G
V
G
V
G
f
+
1
(3.78)
or
m f f m
m f
LT
V G V G
G G
G
+


(3.79)
Equation (3.79) can be rewritten as:
) / ( ) 1 (
f m f f
m
LT
G G V V
G
G
+


(3.79a)
and if the fibres are much stiffer than the matrix (G
f
>>G
m
) the in-plane shear
modulus can be approximated as:
f
m
LT
V
G
G

1
(3.79b)
As in the case of transverse modulus Equation (3.79) underestimates the values
of the in-plane shear modulus, and the Halpin-Tsai equations can be used to give
better predictions:
108
f
f
m LT
V
V
G G
2
2 2
1
1

(3.80)
where:
( )
( )
2
2
1

m f
m f
G G
G G
(3.81)
and
2
is the reinforcing efficiency factor for in-plane shear. The best agreement
with experimental results has been found for
2
=1. Shear modulus as predicted
by Equation (3.80) is shown as a function of the constituent property ratios and
fibre volume fraction in Figure 3.31. Assuming
2
=1, Equation (3.80) becomes:
) ( ) (
) ( ) (
m f f m f
m f f m f
m LT
G G V G G
G G V G G
G G
+
+ +


(3.82)
Equation (3.82) is identical to the formula derived by the self-consistent field
model and to the lower bond of the variational approach. It may be underlined
that G
m
has a significant influence on G
LT
, similar to E
m
on E
T
.
In this section, the matrix and the fibres have been assumed to be isotropic; the
shear modulus of the constituents can be computed from the elastic modulus, E,
and Poissons ratio, using the following formula:
) 1 ( 2 +

E
G
(3.83)
When the reinforcing fibres are anisotropic, the corresponding shear modulus
(G
12
) should be utilised.
109
Figure 3.31 Variation of shear modulus according to Halpin-Tsai equations
10
20
50 100
m
f
G
G
V
f
m
LT
G
G
7
6
5
4
3
2
1
0 0.2 0.4 0.6 0.8
3.5.2 In-plane shear strength
Under in-plane shear, Figure 3.32, high shear stress concentration may
develop at the fibre-matrix interface. The failure could occur by matrix failure,
constituent debonding or a combination of the two. Shear failure may also occur
when off-axis unidirectional composite elements are loaded in axial tension.
Based on matrix shear failure, the following formula can be used to determine
the in-plane shear strength, f
LTs
, of the composite:

k
f
f
ms
LTs

(3.84)
where f
ms
is the matrix shear strength and k

is the shear concentration factor.


When the shear
strain
concentration factor, k

, is utilised, f
LTs
can be estimated using the procedure
outlined in section 3.4.2 for transverse tensile strength. The representative
shaded rectangular element shown in Figure 3.33 loaded by in-plane shear stress
gives the following formula (Gibson 1994) to determine the shear strain
concentration factor, k

:
1 1
1
+
1
1
]
1

f
m
G
G
s
d
k

(3.85)
where d and s are shown in Figure 3.33. The shear strain concentration factor
causes the composite shear failure strain to be less than the matrix failure strain.
If a linear behaviour to failure can be assumed the corresponding shear strength
can be determined with:
110

12
=
LT
Figure 3.32 In-plane shear failure of unidirectional composite

12
=

LT

T
L
=

2
1
f
ms
f
ms
Failure surface

2
1
=

T
L
L
T

k G
f G
f
m
ms LT
LTs

(3.86)
As in case of transverse tensile strength, the matrix shear strength can be used as
an upper bound on the composite shear strength. For a preliminary design, the
in-plane shear strength may be evaluated using a formula (Barbero 1999) similar
to (3.68) replacing the matrix tensile strength with the shear strength of the
matrix as follows:
1
1
]
1

,
_

+
f
m
f f v ms LTs
G
G
V V C f f 1 ) ( 1
(3.87)
where C
v
is the reduction coefficient (described in section 3.4.2) to account for
voids.
Again, in this section the matrix and the fibres have been assumed to be
isotropic; when the reinforcing fibres are anisotropic, the corresponding shear
modulus (G
12
) should be utilised.
3.6 PREDICTION OF POISSONS RATIO
111
d
s

21
=
TL

21
=
TL

12
=
LT

12
=
LT
d
matrix
fibre
Figure 3.33 Mechanics of materials model for strain concentration
factor under in plane shear loading
The Poisson coefficients (
ij
) are important engineering constants of
composite materials. The first subscript, i, refers to the direction of the applied
stress, and the second one, j, corresponds to the direction of the associated
transverse strain.
i
j
ij


(3.88)
for
i
0, ij, and i,j=1,2. In case of a unidirectional composites
ij

ji.
Two Poisson ratios are considered for in-plane loading of a unidirectional
fibre reinforced unidirectional composite. Using the axis system shown in
Figure 3.34 the first Poisson ratio,
LT
, relates the longitudinal stress,
L
, to the
transverse strain,
T
, and is normally referred to as the major Poisson ratio:
L
T
LT


(3.89)
where
L
is the longitudinal strain and the loading scheme is:
L
0,
T
=0 and

LT
=0. The second one called the minor Poisson ratio,
TL
, relates the transverse
stress,
T
, to the longitudinal strain,
L
:
T
L
TL


(3.90)
when
T
0,
L
=0 and
LT
=0.
A model similar to that used to predict E
T
can be used to determine
LT
;
however, the load is applied parallel to the fibres, Figure 3.34. The deformation
pattern illustrated in this figure, for cumulative thicknesses of layers is utilised
to express the transverse strains in the composite and constituents (fibre and
matrix) in terms of longitudinal strains and the Poisson ratio.
112
Figure 3.34 Model of unidirectional composite for
prediction of Poissons ratio
l
f
l
m

L
l
c
Deformed composite
Undeformed composite
f
m
The total transverse deformation of the composite,
c
, is the sum of the
constituent transverse deformations,
f
and
m
. Transverse deformations can be
written as the product of strain and cumulative thickness:
( ) ( ) ( )
c T c c m T m m f
T
f f m f c
l l l + ; ; ;
(3.91)
Transverse strains in the composite, (
c
)
T
, fibre (
f
)
T
and matrix, (
m
)
T
, can
be expressed in terms of the corresponding longitudinal strains, (
c
)
L
, (
f
)
L
, (
m
)
L
and the Poisson ratios as follows:
( ) ( ) ( ) ( ) ( ) ( )
L
f f
T
f L m m T m L c LT T c
v v v ; ;
(3.92)
Equation (3.91) now becomes:
( ) ( ) ( )
m L m m f f f c L c LT
l v l v l v
(3.93)
Assuming that no slippage occurs at the interface and the strains experienced by
the composite, fibre and matrix are equal:
( ) ( ) ( )
L m
L
f L c

and that the widths
are proportional to the volume fractions the following formula is obtained for
the major Poisson ratio:
m m f f LT
V v V v v +
(3.94)
Equation (3.94) is the rule of mixtures for the major Poisson ratio of a
unidirectional composite. The plot of
LT
with respect to the fibre volume
fraction is similar to the plot of the longitudinal modulus (Taranu and Isopescu
1996), Figure 3.35.
The following functional relationship (presented in macromechanics of
composites) exists between engineering constants:
L TL T LT
E E (3.95)
Thus the minor Poisson ratio can be obtained from the already known
engineering constants E
L
, E
T
and
LT
:
L
T
LT TL
E
E

(3.96)
113
or in the extended form:
( ) [ ]
( ) ( ) [ ]
( )
f m f f
f m f f m f
f m f f TL
V E V E
V E V E E E
V V
+
+
+
1
1 /
1

(3.97)
REFERENCES
1. Agarwal, B.D., Broutman, L.J., Chandrashekhara, K, (2006), Analysis and
performance of fiber composites. Third edition. Wiley-Interscience, New-York.
2. Barbero, E. J, (2011), Introduction to composite materials design. Second edition,
CRC Press, Taylor & Francis, Boca Raton.
3. Daniel I., Ishai O, (2006), Engineering mechanics of composite materials. Second
Edition, University Press, Oxford.
4. Gay, D., Hoa, S.V, (2007), Composite materials. Design and applications. CRC Press,
Boca Raton.
5. Gerdeen, J.C., Rorrer, R.A.L, (2012), Engineering design with polymers and
composites. Second edition, CRC Press, Boca Raton
6. Gibson, R. F. (2012), Principles of composite material mechanics. Third edition, CRC
Press Boca Raton.
7. Jones, R. M. (1999), Mechanics of composite materials. Taylor & Francis,
Philadelphia.
8. Kaw, A.K., (2006), Mechanics of composite materials, Second edition, , CRC Press
Boca Raton.
9. Lee, S. M. (1989), Dictionary of composite materials technology. Technomic,
Lancaster.
V
f

LT

m
0 0.2 0.4 0.6 0.8
114
Figure 3.35 Poisson ratio v
LT
as a function of
fibre volume fraction (
m
>
f
)
10. Nielsen, L. E. (1974), Mechanical properties of polymers and composites. Vol. 2,
Marcel Dekker, New York.
11. Taranu N., Isopescu D. (1996) Structures made of composite materials. Vesper, Iasi.
12. Timoshenko, S.P, Gere, J. M. (1961), Theory of elastic stability. McGraw Hill, New
York.
115

You might also like