Вы находитесь на странице: 1из 49

CHAPTER - 6

MACROMECHANICAL BEHAVIOUR

6.1 INTRODUCTION
6.2 THREE-DIMENSIONAL MATERIAL ANISOTROPY
6.3 MATERIAL SYMMETRY
6.4 ELASTIC CONSTANTS AND
ENGINEERING CONSTANTS

COMPLIANCES

IN

TERMS

OF

6.5 CYLINDRICAL ORTHOTROPY


6.6 TWO-DIMENSIONAL CASE: PLANE STRESS
6.7 TRANSFORMATION OF ELASTIC CONSTANTS AND COMPLIANCES
6.7.1 Three-Dimensional Case
6.7.2 Two-Dimensional Case
6.8 PARTICULATE AND SHORT FIBRE COMPOSITES
6.9 MULTIDIRECTIONAL FIBRE REINFORCED COMPOSITES
6.10 UNIDIRECTIONAL LAMINA
6.11 BIDIRECTIONAL LAMINA
6.12 GENERAL LAMINATES
6.13 LAMINATE HYGROTHERMAL STRAINS
6.14 STRENGTH CRITERIA FOR ORTHOTROPIC LAMINA
6.15 BIBLIOGTAPHY
6.16 EXERCISES

6.1 INTRODUCTION
The heterogeneity in a composite material is introduced due to not only its biphase or in some cases multi-phase composition, but also laminations. This leads to a
distinctly different stress strain behaviour in the case of laminates. The anisotropy
caused due to fibre orientations and the resulting extension-shear and bending-

twisting coupling as well as the extension-bending coupling developed due to


unsymmetric lamination add to the complexities. A clear understanding of the
constitutive equations of a composite laminate is thus desirable before these are used
in analysis and design of composite structures. In this chapter, we first introduce to the
readers the basic constitutive equations for a general three-dimensional anisotropic
material with and without material symmetry, elastic constants and compliances and
their relations to engineering constants, as well as transformation laws for elastic
constants and compliances for both three and two-dimensional cases. We also discuss
constitutive relations for several composite materialsparticulate and short fibre
composites, multidirectional fibre reinforced composites, unidirectional lamina and
general laminates as well as lamina strength criteria.
6.2 THREE-DIMENSIONAL MATERIAL ANISOTROPY
For a three-dimensional elastic anisotropic body (Fig. 6.1), the generalized
Hook's law is expressed as
(i, j = 1,2,3)

(6.1)

where
and
are the stress and strain tensors, respectively, and
are the elastic
constants. Here the indices i, j, k and l can assume values of 1, 2 and 3. This implies
that there may exist 34 = 81 independent elastic constants. However, it is known from
the theory of elasticity, that both stress tensor
As

and as

Thus,

and strain tensor

are symmetric.

=
,

=
=

(6.2)
=

(6.3)

This results in reduction of possible independent elastic constants to thirty-six.


Further, if there exists a strain energy U such that
(6.4)
with the property that
=

, then
(6.5)

Equation 6.5 in conjunction with Eq. 6.3 finally reduce the total number of
independent elastic constants from thirty-six to twenty-one only. Such an anisotropic
material with twenty-one independent elastic constants is termed as triclinic. Now,
using the following contracted single index notations

(6.6)

the constitutive relations for the general case of material anisotropy are expressed as

(6.7)

or,

Here, [

i, j = 1, 2,.,6

] is the elastic constant matrix.

(6.8)

Conversely, {
(6.9)

} = [Sij ] {

} ;

i, j =1, 2,..,6

where [Sij] is the compliance matrix.


Note that
[Sij] = [Cij ]-1
(6.10)
Also,

] =[

] and [Sij] = [Sji] due to symmetry.

6.3 MATERIAL SYMMETRY


There may exist several situations when the distribution and orientation of
reinforcements may give rise to special cases of material property symmetry. When
there is one plane of material property symmetry (say, the plane of symmetry is x 3 = 0,
i.e., the rotation of 180 degree around the x 3 axis yields an equivalent material), the
elastic constant matrix [

] is modified as

(6.11)

Thus there are thirteen independent elastic constants, and the material is monoclinic.
The compliance matrix [Sij] for a monoclinic material may accordingly be written
from Eq. 6.11 by replacing 'C ' with 'S '.
If there are three mutually orthogonal planes of symmetry, the material
behaviour is orthotropic. The elastic constant matrix

is then expressed as

orthotropic =

(6.12)

Thus there are nine independent elastic constants. Correspondingly there exist nine
independent compliances.
Two special cases of symmetry, square symmetry and hexagonal symmetry,
may arise due to packing of fibres in some regular fashion. This results in further
reduction of independent elastic constant. For instance, if the fibres are packed in a
square array (Fig. 6.2) in the X2X3 plane. Then
[

square array

(6.13)

There exist now six independent elastic constants. Similarly, when the fibres are
packed in hexagonal array (Fig. 6.3),

(6.14)

In the case of hexagonal symmetry, the number of independent elastic constants is


reduced to five only. The material symmetry equivalent to the hexagonal symmetry, is
also achieved, if the fibres are packed in a random fashion (Fig. 6.4) in the
X2X3 plane. This form of symmetry is usually termed as transverse isotropy. The [ ]
matrix due to the transverse isotropy is the same as that given in Eq. 6.14. The
compliance matrices corresponding to Eqs. 6.12 through 6.14 can be accordingly
written down. However,it may be noted that in the case of rectangular array (Fig. 6.5),
C12 C13, C22 C33 and C55 C66 (Eq. 6.13).
Material Isotropy
The material properties remain independent of directional change for an
isotropic material. The elastic constant matrix [
material are expressed as

] for a three dimensional isotropic

(6.15)

The compliance matrix [Sij] for an isotropic material can be accordingly derived.
6.4 ELASTIC CONSTANTS AND COMPLIANCES IN TERMS OF
ENGINEERING CONSTANTS
The elastic constants or compliances are essentially material constants.
Incidentally, the determination of all these elastic constants or compliances is not easy
to accomplish by simple tests. The material constants that are normally determined
through characterization experiments (see chapter 4) are termed as engineering
constants. They can also be evaluated using the micromechanics material models
(chapter 5).
All nine independent compliances and therefore elastic constants listed in Eq.
6.12 are now expressed in terms of nine independent engineering constants. The
stress-strain relations for a three-dimensional orthotropic material, in terms of
engineering constants, can be written as follows:

(6.16)

We know that, in terms of compliances, the stress-strain relations are

(6.17)
Comparing Eqs. 6.16 and 6.17, we can express the compliances in terms of
engineering constants.

(6.18)

The elastic constants can then be derived by inversion of the compliance


matrix i.e. [

] = [Sij]-1 and are given as follows:

(6.19)

where
(6.20)
In terms of engineering constants, the elastic constants and compliances for an
isotropic material are given by

(6.21)
and
6.5 CYLINDRICAL ORTHOTROPY
Consider cylindrical coordinates r, , z as illustrated in Fig. 6.6. Here the zaxis is assumed to coincide with the X 3-axis. The stress and strain components are
represented as

and

(6.22)

The stress-strain relations, in terms of compliances, become

(6.23)

where

(6.24)

The elastic constant matrix [

] is obtained by inversion of the compliance matrix

[Sij] i.e., [ ] = [Sij]-1 or from Eq. 6.19 by replacing the indices 1,2,3 with r, , z
respectively.
6.6 TWO-DIMENSIONAL CASE: PLANE STRESS
For the case of plane stress (Fig. 6.7)
3 = 4 = 5 = 0
(6.25)
The stress-strain relations, with two-dimensional anisotropy, are

(6.26)

or,
(6.27)

i, j = 1,2,6

Where [Qij] are the reduced stiffnesses (elastic constants) for plane stress.
Similarly, in terms of compliances, the stress-strain relations are

(6.28)

or,
(6.29)

i, j =1,2,6

For the case of two-dimensional orthotropy (Fig. 6.8) the stress-strain relations
are

(6.30)
and

(6.31)
with

(6.32)
and
(6.33)
Note that the engineering constants
,
, , (or
to the orthotropic axis system X'1X'2 (i.e., the material axes).

) and G'12 are referred

6.7 TRANSFORMATION OF ELASTIC CONSTANTS AND COMPLIANCES


If the elastic constants and compliances of a material are known with respect
to a given co-ordinate system, then the corresponding values with respect to any other
mutually perpendicular coordinates can be determined using laws of transformation.
These are explained in Appendix A.
6.7.1 Three-Dimensional Case
The transformation of elastic constants from the X' 1X'2X'3 coordinates to
mutually orthogonal X1X2 X3 coordinates (Refer Fig. A.1 and Eq. A. 22) is given as
follows:
(6.34)
where transformation matrix

is given by Eq. A.8.

Note that the elements of


X'1X'2X'3 coordinates, respectively.

and

correspond to the X1X2 X3 and

One can use Eq. 6.34 in the following form


-T

=
=

-1

[T]

[T]T

(6.35)
if the transformation is required from X 1X2 X3 coordinates to theX'1X'2X'3 coordinates.
Note that [T] is defined by Eq. A.13.
Simalarly,
(6.36)
(6.37)
The corresponding elastic constants

and compliances [Sij] due to special

cases of material symmetry and transformation matrices


and [T] due to specific
orientation of axes are to be reduced from the general three dimensional cases, before
transformation is sought from one axis system to the other.
6.7.2 Two-Dimensional Case
If the elements of
and
refer to the X1X2 and X'1X'2 coordinates (see
Eqs. 6.26 and 6.30 and Figs. 6.7 and 6.8), respectively, then transformation laws for
reduced elastic constants are obtained as follows:
(6.38)
(6.39)
where
and
are defined by Eqs. A.18 and A.19. The compliance matrices are
accordingly transformed using Eqs. 6.36 and 6.37.
Accordingly, from Eqs. 6.38 and A.18 it can be shown that

(6.40)

In a similar way from Eqs. 6,36 and A.19 one obtains

(6.41)

Note that m = cos

and n = sin

and

and

are defined by Eqs. 6.32 and

6.33 respectively, in terms of engineering constants


G'12 corresponding to principal material directions.

(or

) and

If transformation is required from one anisotropic material axis system (say


X1X2 X3) to another anisotropic material axis system (say,
6.38 and A. 18 we can

or,

or,

), then from Eqs.

(6.42)

Similarly, using Eqs. 6.36 and A.19 one can write

or,

(6.43)

6.8 PARTICULATE AND SHORT FIBRE COMPOSITES


Particulate composites, where reinforcements are in the form of particles,
platelets and flakes, and short fibre composites may exhibit a wide range of elastic

material behaviour depending on the shapes, sizes, orientations and distributions of


reinforcements in the matrix phase as well as elastic properties of the constituent
materials. The matrix behaviour is normally isotropic. The composition of these
composites are first established by examining their morphology and then proper
stress-strain relations can be obtained from the equations developed in the preceding
sections. It is also to be noted whether the composite body under consideration is
three-dimensional or two dimensional in character.
For example, the behaviour of a three-dimensional composite with a typical
reinforcement packing shown in Fig. 6.9a is anisotropic in nature. Here the
reinforcements are oriented in some regular fashion with respect to the reference axes
X1X2 X3. The stress strain relations

for this type of composites are

given by Eq 6.8 with elements of


listed in Eq. 6.7. When the reinforcements are
arranged parallel to the axes (Fig. 6.9b), the composite behaviour is orthotropic and
Eq. 6.12 defines the corresponding
. If the orientation and distribution of
reinforcements are found to be random in the matrix phase, as shown in Fig. 6.9c, the
composite is assumed to behave like an isotropic material. Consequently, the elastic
constant matrix

is reduced to that given in Eq. 6.15.

For two-dimensional anisotropic, orthotropic and isotropic cases, some


possible reinforcement arrangements are illustrated in Fig. 6.10. The stress-strain
relations, as presented in section 6.6 can be accordingly used for these cases.
If transformations of elastic constants and compliances are required from one
axes system to another, then one can use the transformation rules discussed in section
6.7.
Fig. 6.10
6.9 MULTIDIRECTIONAL FIBRE REINFORCED COMPOSITES
Composites exhibit strong directional properties, when reinforcements are in
the form of continuous fibres. In a multidirectional composite, fibres can be placed in
any desired direction in a three-dimensional space, along which better stiffness (or
strength) is desired. The shear properties can be greatly improved by providing
diagonal reinforcements. Carbon-carbon composites form an important class of
multidirectional composites due to several variations in weave design and perform
construction. Similar multi-directional composite systems can also be designed and
developed with both metal-matrix and ceramic-matrix composites. A typical multidirectional (5D) composite is shown in Fig. 6.11a. There are three bundles of
orthogonal fibres f1, f2, f3 and two bundles of diagonal fibres f4, f5. We consider here
an integrated multidirectional fibre reinforced composite moder which contains n

number of unidirectional fibre composite blocks that are oriented in n arbitrary


directions with respect to a three-dimensional reference axes X 1X2 X3. Each unit block
may have different fobre volume fractions. This arrangement makes n number of
material axis systems, and therefore yield n sets of direction cosines between n
material axis systems and the reference axes X 1X2 X3. For example, Figure 6.11b
represents the orientation of the material axis system for the ith block. The
corresponding transformation matrices
Eqs. A.8 and A.13, respectively.

and

can then be written down using

The material behaviour for each block with respect to its axes is orthotropic.
The elastic constants for the ith block are then given as

(6.44)

The effective elastic constants for the n-directional fibre reinforced composite are then
determined by averaging the transformed properties as follows:
(6.45)
Note that the overall fibre volume fraction is given as

(6.46)
6.10 UNIDIRECTIONAL LAMINA
A unidirectional lamina is a thin layer (ply) of composite and is normally
treated as a two-dimensional problem. It contains parallel, continuous fibres and
provides extremely high directional properties. It is the basic building unit of a
laminate and finds very wide applications in composite structures specially in the

form of laminates. Therefore, the knowledge of its elastic macromechanical behaviour


is of utmost importance to composite structural designers.
Figure 6.12a depicts a unidirectional lamina where parallel, continuous fibres,
are aligned along the X'1 axis (fibre axis or longitudinal direction). The X' 2 axis
(transverse direction) is normal the fibre axis. The axes X' 1X'2 are referred as material
axes. The material axes are oriented counter clockwise by angle with respect to the
reference axes X1X2. The angle (also referred as fibre angle) is considered positive
when measured counterclockwise from the X 1 axis. This type of unidirectional lamina
is termed as off-axis lamina. An off-axis lamina behaves like an anisotropic twodimensional body, and the stress-strain relations, given by Eqs. 6.26 through 6.29, can
be used for the present case.
When the material axes coincide with the reference axes (i.e., =0), as shown
in Fig. 6.12b, the lamina is termed as on-axis lamina and its behaviour is
orthotropic in nature. The stress-strain relations are defined by Eqs. 6.30 and 6.31.
The engineering constants
,
, , (or ) and G'12 are usually known, as
these can be determined either by using micromechanics theories (chapter 4) or by
characterization tests (chapter 5). Using these engineering constants, the reduced
stiffnesses

and compliances

are then determined for an orthotropic lamina

with the help of Eqs. 6.32 and 6.33. The transformed reduced stiffnesses
can now be evaluated employing Eqs. 6.40 and 6.41. The stiffness

and
and

compliances
for three composite systems are computed for various fibre
orientations and are listed in Tables 6.1 and 6.2. Typical variations of transformed
properties
and
with change in the fibre angle are illustrated in Figs. 6.13
and 6.14. Such plots aid to the basic understanding of the stiffness behaviour of an
off-axis lamina with different fibre orientations. Note that the case =0 corresponds
to an on-axis lamina.
6.11 BIDIRECTIONAL LAMINA
A bidirectional lamina is one which contains parallel, continuous fibres aligned along
mutually perpendicular directions, as shown in Fig. 6.15. A lamina reinforced with woven fabrics
that have fibres in the mutually orthogonal warp and fill directions can also be treated as a
bidirectional lamina. The effects of undulation (crimp) and other problems associated with
different weaving patterns are however, neglected. In Fig. 6.15 the X 1' X2' is referred as material
axes. The amount of fibres in both directions need not necessarily be the same. In a hybrid
lamina, even the fibres in two directions may vary, but when the material axes X 1' X2' coincide
with the reference axes X1X2 (Fig. 6.15a), the material behaviour is orthotropic and the lamina
may be termed as on-axis bidirectional lamina. If the X1' X2' plane rotates by an angle

with respect to the X1X2 axes (Fig. 6.15a), then the oriented lamina behaves as an anisotropic
material and it can be identified as an off-axis bidirectional lamina can also be treated as a
two-dimensional problem and its elastic properties can be determined in an usual manner as
discussed in sections 6.6 and 6.10. It may be mentioned that the anisotropy and stiffness
behaviour of a bidirectional lamina can be greatly controlled by varying the types of fibres (say,
carbon fibre along the X1' direction and glass fibre along the X2' direction) and volume fractions
of fibres (Vf) in both directions. When the fibres and Vf are same in both directions, then
E'11 = E'22 and the material behaviour is square symmetric. Note that a square symmetric material
is different from an isotropic material.

6.12 GENERAL LAMINATES


We consider here a general thin laminate of thickness h (Fig. 6.16). The X 3 axis
is replaced here by the z axis for convenience. The laminate consists of n number of
unidirectional and/or bidirectional laminae, where each lamina may be of different
materials and thicknesses and have different fibre orientations ( ). A thin general
laminate is essentially a two-dimensional problem, but cannot be treated as a twodimensional plane stress problem as has been done for a unidirectional lamina. The
existence of extension bending couling causes bending, even if the laminate is
subjected to inplane loads only. Therefore, thin plate bending theories are employed in
derivation of constitutive relations. We assume that Kirchhoff 's assumptions related to
the thin plate bending theory are applicable in the present case.
Let u10, u20 and w are the mid-plane displacements, and w is constant through
the thickness of the lamina. Then the mid-plane strains are given by
(6.47)
and the curvatures, which are constant through the thickness of the laminate, are
(6.48)
The strains at any distance z are then given as
(6.49)
Now from Eq. 6.26, we have at any distance z

(6.50)
The stress and moment resultants (Fig. 6.17) are evaluated per unit length of
the laminate as follows:

and
Thus,

where

where

and

(6.51)

Proceeding in a similar manner, all stress and moment resultants can be expressed as
listed below:

(6.52)

with (Aij, Bij, Dij) =


(6.53)

ij

(1, z, z2) dz;

i, j = 1, 2, 6

Equation 6.52 represents the constitutive relations for a general laminate, and
Aij, Bij, and Dij are the inplane, extension bending coupling and bending stiffnesses,
respectively. Note that all these stiffnesses are derived for a unit length of the
laminate. The elastic properties of each lamina are generally assumed to be constant
through its thickness, as these laminae are considered to be thin. Then A ij, Bij, and
Dij are approximated as

(6.54)

From Eq. 6.52, it is seen that there exist several types of mechanical coupling in a
general laminate. These are grouped together as follows:

Extension Shear

: A16, A26

Extension Bending

: B11, B12, B22

Extension Twisting

: B16 , B26

Shear Bending
Shear Twisting
Bending Twisting

B16 , B26

B66
:

D16 , D26

Biaxial Extension

A12

Biaxial Bending

D12

As stated earlier, the coupling terms B ij occur due to unsymmetry about the middle
surface of a laminate. However, all terms containing suffices '16 ' and '26 ' are resulted
due to anisotropy caused by the fibre orientation other than 00 and 900. Those
containing suffices '12 ' are due to Poisson's effect. Although a heneral unsymmetric
laminate contains all coupling terms, there are several laminates where some of these
may vanish. These are listed in Table 6.3. There are several important points that are
to be noted here. The first two laminates (serial nos. 1 and 2) which are christened as
off-axis laminate and on-axis laminate; respectively are essentially paralles
ply laminates where all laminae in a laminate have the same fibre orientation and
therefore are stacked parallel to each other. These are, in fact, similar to unidirectional
laminae. For a symmetric balanced angle-ply laminate D 16 and D26 do not vanish,
although A16 = A26 = 0. The only coupling effect that appears in an anti-symmetric
cross-ply laminate is the extension-bending coupling due to presence of B 11and
B22 and note that B22 = - B11. But the existence of B 16 and B26 cause an antisymmetric
angle-ply laminate to experience extension-twisting coupling. Note also that
extension-bending coupling is predominant for an unsymmetric cross-ply laminate.
The mechanical coupling, as discussed above, influences the deformation
behaviour of a laminate to a great extent. This can be better understood by examining
the deformed shapes of a couple of laminates as illustrated in Figs. 6.18 through 6.20.
Here the dotted lines represent the undeformed shape and the firm lines, deformed
shapes. Consider first a simple off-axis laminate (or unidirectional lamina), subjected
to an inplane stress resultant N 1 (Fig.6.18a) and an out-of-plane moment resultant
M1 (Fig. 6.18b). We know from Eq. 6.52 and Table 6.1 (Bij=0) that
(6.55)

Thus, as illustrated in Fig. 6.18a, it is noted that a simple tension causes not only
extension and contraction, but also shearing of the laminate. While the extension and
contraction are due to A11 and A12, respectively and the inplane shear deformation is
due to presence of A16. This characteristic behaviour is seen especially in an
anisotropic (off-axis) laminate. The shear deformation vanishes, if A 16 = 0, as in the
case of an orthotropic (on-axis) laminate (serial no.2 of Table 6.1). Similarly, as can be
seen in Fig. 6.18b, a simple bending due to M 1 has resulted not only longitudinal
bending (due to D11) and transverse bending (due to D 12), but also twisting (due to
D16).
Figure 6.19 describes the deformation behaviour of an antisymmetric cross-ply
laminate. The extension-bending coupling due to B 11 and B22 can be clearly observed.
In Fig. 6.19a a simple inplane tension is found to introduce bending in the laminate.
Conversely, a simple bending causes extension of the laminate, a shown in Fig. 6.19b.
Figure 6.20 depicts the deformed shape of an antisymmetric angle-ply
laminate. Here the extension-bending and bending-shear coupling effects due to
B16 and B26 are presented. In a similar manner, the deformation characteristics of other
types of laminates can be illustrated. The most important point that is to be focused
here is that fibre orientation and lamina stacking sequence affect laminate stiffness
properties, which, in turn, control the deformation behaviour of a laminate.
Table 6.4 provides the stiffnesses [A ij], [Bij]and [Dij] for various stacking
sequences of carbon/epoxy composites. The [Q ij] values given in Table 6.1 have been
used to compute the above stiffnesses.
6.13 LAMINATE HYGROTHERMAL STRAINS

The changes in moisture concentration and temperature introduce expansional


strains in each lamina. The stress-strain relation of an off-axis lamina (Eq. 6.28) is
then modified as follows

(6.56)
with

(6.57)
and

and

(6.58)

where the superscripts e, H, T refer to expansion, moisture and temperature,


respectively, C and T are the change in specific moisture concentration and
temperature, respectively, and 's and 's are coefficients of moisture expansion and
thermal expansion respectively.
Note that the spatial distributions of moisture concentration and temperature
are determined from solution of moisture diffusion and heat transfer problems.
Expansional strains transform like mechanical strains (Appendix A)
i.e.,

Inversion of Eq. 6.56 yields (see also Eq. 6.26), at any distance z (Fig.

6.16),
Thus, for a general laminate Eq. 6.52 will be modified as

(6.59)

(6.60)
where the expansional force resultants are

(6.61)
and the expansional moments are

(6.62)
These expansional force resultants and moments may considerably influence the
deformation behaviour of a laminate.

6.14 STRENGTH CRITERIA FOR ORTHOTROPIC LAMINA

Isotropic materials do not have any preferential direction and in most cases
tensile strength and compressive strength are equal. The shear strength is also
dependent on the tensile strength. A strength criterion for an isotropic lamina is,
therefore, based on stress components, 1, 2 and 6for a two-dimensional problem and
a single strength constant i.e., ultimate strength X. An orthotropic lamina (Fig. 6.8), on
the other hand, exhibits five independent strength constants e.g., tensile strength X' 11t a
dcompressive strength X'11c along the X'1 direction; tensile strength X'22t and

compressive strength X'22c along the X'2 direction and inplane shear strength X' 12.
Hence a strength criterion for a two-dimensional orthotropic lamina should involve
the stress components '1, '2 and '6 and strength constants X'11t, X'11c, X'22t X'22c and
X'12. We present here a few important strength criteria that are commonly used to
evaluate the failure of an orthotropic lamina. Maximum Stress Criterian
A lamina is assumed to fail, if any of the following relations is satisfied
, when
,

and

when

are tensile

and

are

compressive.

(6.63)
It is assumed that inplane shear strengths are equal under positive or negative shear
load.
Maximum Strain Criterian
A lamina fails, if any of the following is satisfied
when
when

and
and

are tensile
are compressive.

(6.64)

Note that the addition of suffix 'u' in strain components indicates the corresponding
ultimate strains. The ultimate shear strains are also assumed to be equal under positive
or negative shear load. If a material behaves linearly elastic till failure, the ultimate
strains can be related to ultimate strength constants as follows:

(6.65)
Tsai-Hill Criterion
The general three-dimensional orthotropic strength criterion is given by
(6.66)

Assuming that normal stresses


and substituting
= X'11,
criterion, we obtain

and

= X'22,

an dshear stress
and

act independently

in the above strength

(6.67)

Combining Eqs. 6.67 we get

(6.68)

Assuming transverse symmetry X'22 = X'33 and two-dimensional plane stress case ( 3 =
4 = 5 =0), Eq. 6.66 reduces to
or,
(6.69)
When

or both are tensile or compressive, Eq. 6.69 can be used by substituting

the corresponding tensile or compressive strength constants in it. Thus, if


X'11 = X'11t , and if

is tensile,

is compressive, X'22= X'22c and so on.

Tsai-Hill / Hoffman Criterion


Tsai-Hill/Hoffman criterion accounts for unequal tensile and compressive
strengths. For a three-dimensional state of stress in an orthotropic material, this
criterion is given as
(6.70)
If tensile

acts only and

=X'11t , then from Eq. 6.70

(C2 + C3) X'11t + C4 = 1/X'11t


(6.71)
If compressive

acts only and -

= X'11c, then

(C2 + C3) X'11c C4 = 1/X'11c

(6.72)

From Eqs. 6.71 and 6.72, we obtain


(6.73)
Similarly, consideration of

and

yields
(6.74)
(6.75)

Now, assuming
C2 and C3:

, we derive from the above the following relations for C 1,

(6.76)
(6.77)
Further, applying

only and

yields

(6.78)
Now, considering a two dimensional state of plane stress condition
and substituting the values of C1, C2, C3, C4, C5 and C9 from the above relations, the
strength criterion takes the following form:
(6.79)

Tsai-Wu Quadratic Interaction Criterion


For an orthotropic material under a two-dimensional state of plane stress
condition, this criterion assumes the form
(6.80)
Considering that the positive or negative inplane shear stress
results, the terms F16
to

, F26

and F6

should not affect the

should vanish. Hence Eq. 6.80 reduces


(6.81)

Now applying independently tensile and compressive normal stresses


inplane shear stresses
X'22c and

, and substitution of

=X'11t -

=X'11c,

and

, and

=X'22t , -

=X'12 in Eq. 6.81 yields

(6.82)
Employing the von Mises plane stress analogy, the remaining interaction coefficient
F12 can be defined

(6.83)
Combining Eqs. 6.81-6.83, the Tsai-Wu criterian takes the following form:

(6.84)

It is to be mentioned that the Tsai-Wu criterion (Eq. 6.84) accounts for interaction of
stress components as well as both tensile and compressive strength constants and
shear strength and is considered as a reasonably accurate and consistent representation
of failure of an orthotropic lamina under biaxial stresses. The Tsai-Hill criterion (Eq.
6.69) is also very popular with composite structural designers.

Table 6.1: Stiffnesses

and

for three unidirectional composites (GPa)

Material
Kelvar/Epoxy

91.87

4.03

1.41

2.26

Carbon/Epoxy

133.94

8.32

2.16

3.81

Boron/Polyimide

242.39

14.93

3.88

5.53

Material
(deg
ree)
0

91.
87

4.0
3

1.4
1

2.
26

0.0
0

0.
00

30

54.
15

10.
23

17.
17

18
.0
2

28.
12

9.
92

45

26.
93

26.
93

22.
42

23
.2
7

21.
96

21
.9
6

60

10.
23

54.
15

17.
17

18
.0
2

9.9
2

28
.1
2

90

4.0
3

91.
87

1.4
1

2.
26

0.0
0

0.
00

Kelvar/
Epoxy

Carbon/

13
3.9
4

8.3
2

2.1
6

3.
81

0.0
0

0.
00

30

79.
53

16.
72

25.
17

26
.8
2

40.
48

13
.9
2

45

40.
46

40.
46

32.
84

34
.4
8

31.
40

31
.4
0

60

16.
72

79.
53

25.
17

26
.8
2

13.
92

40
.4
8

90

8.3
2

133
.94

2.1
6

3.
81

0.0
0

0.
00

24
2.3
9

14.
93

3.8
8

5.
53

0.0
0

0.
00

30

14
2.8
8

29.
15

46.
53

48
.1
7

73.
87

24
.6
2

45

71.
80

71.
80

60.
74

62
.3
9

56.
87

56
.8
7

60

29.
15

142
.88

46.
53

48
.1
7

24.
62

73
.8
7

90

14.
93

242
.39

3.8
8

5.
53

0.0
0

0.
00

Epoxy

Boron/
Plyimide

Table 6.2: Compliance

and

for three unidirectional composites (TPa)-1

Material
Kelvar/Epoxy

10.94

249.7

-3.83

443.4

Carbon/Epoxy

7.50

120.6
6

-1.95

262.4
7

Boron/Polyimide

4.14

67.27

-1.08

180.9
6

Material
(deg
ree)
0

10.
94

249
.75

44
3.83 3.
46

0.0
0

0.0
0

30

10
3.4
8

222
.88

31
36.6 2.
6
13

141
.32

65.
50

45

17
4.1
2

174
.12

26
47.6 8. 119. 11
1
35 40 9.4
0

60

22
2.8
8

103
.48

31
36.6 2.
6
13

65.
50

14
1.3
2

90

24
9.7
5

10.
94

44
3.83 3.
46

0.0
0

0.0
0

7.5
0

120
.66

26
1.95 2.
47

0.0
0

0.0
0

30

60.
24

116
.82

16
26.4 4.
0
66

77.
23

20.
76

45

96.
68

96.
68

13
34.5 2.
5
05

56.
58

56.
58

Kelvar/
Epoxy

Carbon/
Epoxy

Boron/
Plyimide

60

11
6.8
2

60.
24

16
26.4 4.
0
66

20.
76

77.
23

90

12
0.6
6

7.5
0

26
1.95 2.
47

0.0
0

0.0
0

4.1
4

67.
27

1.08

18
0.
96

0.0
0

0.0
0

30

40.
06

71.
63

10
21.2 0.
1
42

50.
59

4.0
8

45

62.
56

62.
56

73
27.9 .5
3
7

31.
56

31.
56

60

71.
63

40.
06

10
21.2 0.
1
42

4.0
8

50.
59

90

67.
27

4.1
4

1.08

18
0.
96

0.0
0

0.0
0

Table 6.3 : Stiffnesses for various types of laminates


Case Laminate type

Elastic behaviour

Stiffnesses

1. Off-axis laminate
(all plies oriented
at )

anisotropic
and uncoupled

all Bij=0; Aij= h Qij


Dij = (h3/12) Qij

2. On-axis laminate
(all plies oriented either
00 or 900)

orthotropic
and uncoupled

3. Symmetric cross-ply
(odd number of

specially
orthropic and

I. Symmetric Laminates

all Bij=0; Aij=h Qij


and Dij = h3/12) Qij
with Q16 = Q26 = 0
all Bij=0; A16= A26=
D16= D26=0; rest of

00 / 900 / 00, etc. plies)

uncoupled

4. Symmetric angle-ply
Dij
(odd number of /- / ,
etc. plies)
5.

anisotropic and
uncoupled

Symmetric balanced angle


A26=0 rest
ply ( /- /- / , etc. plies)
finite.

Aij and Dij are finite


all Bij=0; all Aij and
are finite

anisotropic and
uncoupled

all Bij=0; A16=


of Aij and Dij are

II. Unsymmetric Laminates

6. Antisymmetric cross-ply
B26=
(even number of
D26=0
00 / 900 / 00/900, etc. plies)
Dij are

orthotropic and

A16= A26= B16=

partly coupled

B12= B66= D16=


rest of Aij ,Bij and
finite with B22=-B11;
D22=-D11

7. Antisymmetric angle-ply
(even number of
D26=0
( /- / /- , etc. plies)
Dij are

anisotropic and
partly coupled

A16= A26=B11=B22
B12= B66= D16=
rest of Aij, Bij and
finite.

8. Unsymmetric cross-ply
(irregular stacking of
00 or 900 plies)
9.

General unsymmetric
Dij are
laminate

Table 6.4: Stiffneses

orthotropic but
coupled

A16= A26= B16= B26=


D16= D26=0; rest of
Aij, Bij and Dij are
finite.

anisotropic and
strongly coupled

all Aij, Bij and


finite.

for carbon/epoxy composite laminates

Laminate Thickness : 4mm


Units : [Aij], GPa-mm;

[Bij], GPa-mm2;

[Dij], GPa-mm3

1. 00 / 900 / 00 laminates

2. 450 / -450 / 450 laminate

3. 450 /-450 /450 / 450 laminate

4. 00 /900 / 00 / 900 laminate

5. 450 /-450 /450 / -450 laminate

6. 00 /900 /00 / 00 laminate

6.15 BIBLIOGTAPHY
1. S.P. Timoshenko and J.N. Goodier, Theory of Elasticity, McGraw Hill, N.Y.,
1970.
2. Y.C. Fung, Foundations of Solid Mechanics, Englewood Cliffs, N.J., 1965.
3. S.G. Lekhnitskii, Theory of Elasticity of an Anisotropic Body, MIR Publ.
Moscow, 1981.
4. J.C, Halpin, Primer or Composite Materials: Analysis, Technomic Publ. Co.,
Inc. Lancaste, 1984.
5. R.M. Christensen, Mechanics of Composite Materials, Wiley Interscience,
N.Y., 1979.
6. Z. Hashin and C.T. Herakovich (Eds.), Mechanics of Composite MaterialsRecent Advances, Pergamon Press, N.Y.,1983.
7. S.W. Tsai and H.T. Hahn, Introduction to Composite Materials Technomic
Publ. Co., Inc., Lancaster,1980.
8. J.M. Whitney, Structural Analysis of Laminted Composites, Technomic Publ.
Co., Inc.,Lancaster, 1987.
9. J.R. Vinson and R.L. Sierakowski, The Behaviour of Structures Composed of
Composite Materials, Kluwar Academic Publ., MA,1985.
10. S.W. Tsai, J.C. Halpin and N.J. Pangano (Eds.) Composite Materials
Workshop, Technomic Publ. Co., Inc., Lancaster, 1968.
6.16 EXERCISES
1. State the generalized Hooke's law for a three-dimensional elastic anisotropic
material and show that there are twenty-one independent elastic constants for a
triclinic material.
2. Write down the elastic constant matrix for three-dimensional orthtropic, square
symmetric, hexagonal symmetric and isotropic materials.
3. Distinguish between elastic constants and engineering constants.

4. For a two-dimensional orthotropic case, express


engineering constants.
5. Derive expressions for

and

in terms of angle

and

in terms of

and show that

and

6. Assume properties given in Table 4.4 for Kevlar/epoxy and


carbon/epoxy/composites and determine [A ij], [Bij] and [Dij]
a

hybrid laminate (thickness 4 mm).

7. Make a critical assessment of various lamina failure theories.


8. Derive expressions for Tsai-Hill and Tsai-Wu strength criteria.

for

Вам также может понравиться