Vous êtes sur la page 1sur 78

Chapter 3

COMPOSITE LAMINATES
3.1

Macromechanical behavior of lamina

In this chapter, we discuss some fundamental problems concerning fiberreinforced composite laminates; i.e. the classical part of the general theory of
composite materials.
The basic results existing in this field can be found, for instance, in the
monographies due to Ashton and Whitney [3.1], Jones [3.2], Christensen [3.3],
Tsai and Hahn [3.4], Cristescu [3.5], Whitney [3.6] and Gibson [3.7].
The fiber-reinforced composite laminates are made of fiber-reinforced laminae. The fibers considered here are long and continuous. A lamina is a plane
arrangement of unidirectional fibers strongly bounded in a matrix. In Figure 3.1
is shown a typical lamina together with its material symmetry axis, named also
principal material axes or directions.

Figure 3.1: Lamina with unidirectional fibers.

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

112

CHAPTER 3. COMPOSITE LAMINATES

Axis 1 is parallel to the fibers, axis 2 is perpendicular to the fibers in the


plane of the lamina and axis 3 is perpendicular to the plane of lamina. The fibers
or filaments are the main reinforcing or load-carrying elements. They are generally
strong and stiff. The matrix can be organic, ceramic, or metallic. The function of
the matrix is to support and protect fibers and to provide a means of distributing
and transmitting load among fibers. The fibers generally exhibit linear elastic
behavior. Fiber-reinforced composites, such as boron-epoxy and graphite-epoxy are
usually considered to be linear elastic materials since the fibers provide most of
the stiffness.
A laminate is a stack of laminae with various orientations of the principal
material directions with respect to the laminae as shown in Figure 3.2.

Figure 3.2: Exploded view of laminate structure.


Generally the fiber orientation of the layers cannot be symmetric about the
middle surface of the laminate. The layers of a laminate are usually firmly bounded
together by the same matrix material that is used in laminae. Laminates can be
composed of plates of different materials, or layers of fiber-reinforced laminae, as
shown in Figure 3.2. Also, various laminae can have various thicknesses.
A major purpose of lamination is to determine the directional dependence
of stiffness of a material in accordance with the given loading environment of the
structural element. Laminates are suited to this objective since the principal material directions of each layer can be oriented according to the need. For example, six
layers of a ten-layer laminate could be oriented in one direction and the other four
at 90 with respect to that direction. The resulting laminate has an extensional
stiffness roughly 50 percent higher in one direction than in the other one.
The fiber-reinforced lamina is the basic building block in a laminated fiberreinforced composite or laminate. Thus, the knowledge of the mechanical behavior
of a laminae is essential to the understanding of laminated fiber-reinforced structures. Analyzing the macro and micromechanical behavior of a laminae, we assume
that the matrix, the fibers and the lamina itself have linear elastic behavior. Also

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.1. MACROMECHANICAL BEHAVIOR OF A LAMINA

113

we suppose that the fibers and the matrix are firmly bounded together. The same
assumption will be made concerning the laminae forming a laminate.
At the macro-mechanical level, the fiber-reinforced lamina will be assumed
to be an orthotropic linearly elastic material. The symmetry axis are parallel and
perpendicular to the fibers direction as shown in Figure 3.2. The most advantageous description of the stress-strain relation involves the (macro-mechanical or
effective or equivalent or overall ) technical or engineering constants of the lamina, considered as a homogeneous body. These constants are particulary helpful in
describing material behavior since they are determined by obvious and relatively
simple mechanical tests.
In the following, our attention will be focused on stress-strain relation for
orthotropic materials in a plane stress state, the most common condition satisfied
by a loaded composite lamina. The constitutive relations, initially formulated using
the material symmetry axes, will be expressed later by using coordinate systems
that are not aligned along the principal material directions. Such a change is
necessary in order to describe the global behavior of various laminates, composed
of laminae with various orientations of the reinforcing fibers.
Let us consider now a lamina in the 1-2 plane as shown in Figure 3.1. Here
the axes 1, 2, 3 are the principal material directions of the laminae, assumed to
be (macroscopically) orthotropic.
As usual, we say that the lamina is in a plane stress state relative to its
symmetry plane 1-2 if the components of the stress tensor satisfy the following
relations:
31 = 32 = 33 = 0.
(3.1.1)
Since the material is orthotropic, according to the constitutive equation
(2.2.70), from the above relation, it follows that the components of the strain
tensor satisfy the equations
31 = 32 = 0 , 33 = S13 11 + S23 22 ,
and, thus, the stress-strain relation (2.2.70) reduces to


1
1
S11 S12 0
2 = S12 S22 0 2 .
0
0
S66
6
6

(3.1.2)

We recall that in the above matrix form of the remaining constitutive equation,
we have used the Voigts convention; i.e.
1 = 11 , 2 = 22 , 6 = 212 , 1 = 11 , 2 = 22 , 6 = 12 .
Also, we note again that the axis 1, 2, 3 are the principal material directions
of the lamina, axis 1 being parallel to the fibers, axis 2 being perpendicular to the
fibers and situated in the plane of the lamina and axis 3 being perpendicular to
this plane.

Copyright 2004 by Chapman & Hall/CRC

114

CHAPTER 3. COMPOSITE LAMINATES

The general relation (2.2.71) shows that the involved components S11 , S12 , S22
and S66 of the compliance matrix [S] can be expressed in terms of the technical
constants of the orthotropic lamina by the following equations:

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

S11 =

12
21
1
,
=
, S12 =
E2
E1
E1
1
1
.
, S66 =
S22 =
G12
E2

(3.1.3)

Since the matrix [S] is positive definite, the relation (3.1.2) can be inverted
to obtain the inverse stress-strain relations

1
Q11 Q12 0
1
1
2 = Q12 Q22 0
2 = [Q] 2 .
(3.1.4)
6
0
0
Q66
6
6

The quantities Q11 , Q12 , Q22 and Q66 are named reduced stiffnesses. They
have the following expressions:
Q11 =

S11
S12
S22
2
,
, Q66 = G12 , S = S11 S22 S12
, Q22 =
, Q12 =
S
S
S

(3.1.5)

or, in terms of the engineering constants

E2
21 E1
12 E2
E1
, Q66 = G12 .
, Q22 =
=
, Q12 =
1 12 21
1 12 21
1 12 21
1 12 21
(3.1.6)
The reduced constitutive equations (3.1.4) represent the basis for the analysis
of the behavior of an individual lamina subjected to forces acting in its own plane.
For such special loading, the orthotropic lamina is indeed in a plane stress state.
We stress again that E1 is Youngs modulus in the fibers direction, E2 is
Youngs modulus in the direction perpendicular to the fibers and situated in the
lamina plane, 12 and 21 are Poissons ratios in the same plane, and G12 is the
shear modulus in the lamina plane.
We now present some numerical values of the involved material parameters for
laminae frequently used in applications. The values are taken from the monograph
[3.4] by Tsai and Hahn (see pp. 19 and 20). The material constants having physical
dimensions (such as E1 , E2 , G12 , S11 , ..., S66 , Q11 , ..., Q66 ) are expressed in GP a =
109 N m2 . Obviously, if E1 , E2 , 12 and G12 are known from experimental data
S11 , ..., S66 and Q11 , ..., Q66 can be calculated using the relation (3.1.3) and (3.1.6).
The data given in Tables 3.1, 3.2 and 3.3 show that for fiber-reinforced laminae, generally
E2 << E1 and G12 << E1

Q11 =

and
Q22 << Q11

Copyright 2004 by Chapman & Hall/CRC

and

Q66 << Q11 .

115

3.1. MACROMECHANICAL BEHAVIOR OF A LAMINA

Type
T300/5208
B(4)/5505
AS/3501

Material
Graphite/Epoxy
Boron/Epoxy
Graphite/Epoxy

E1
181
204
138

E2
10.3
18.5
8.96

12
0.28
0.23
0.30

G12
7.17
5.59
7.1

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

Table 3.1: Engineering constants of typical fiber-reinforced laminae.

Type
T300/5208
B(4)/5505
AS/3501

S11
5.525
4.902
7.246

S22
97.09
54.05
111.6

S12
-1.547
-1.128
-2.174

S66
139.5
172.7
140.8

Table 3.2: Compliance components of typical fiber-reinforced laminae.

Type
T300/5208
B(4)/5505
AS/3501

Q11
181.8
205.0
138.8

Q22
10.34
18.58
9.013

Q12
2.897
4.275
2.704

Q66
7.17
5.75
7.1

Table 3.3: Reduced stiffnesses of typical fiber-reinforced laminae.

We shall see in the Section 5, that the above large differences between the
magnitudes of the different rigidity moduli of a fiber-reinforced composite material have essential implications on the stability behavior of these bodies, having
obviously an internal structure.
We recall that the reduced constitutive relations (3.1.4) are expressed using
the stress and strain components corresponding to the material symmetry direction
of the lamina. These special directions often do not coincide with the coordinate
direction which are geometrically related to a given problem. Hence, we must
be able to express the reduced stress-strain relations using arbitrary systems of
coordinates x1 = x, x2 = y, x3 = z. For our needs, we assume that the principal
material direction 3 and the direction of the axis x3 = z coincide. Also, we
suppose that the planes x, y and 1, 2 coincide, and the principal directions 1, 2 are
obtained by rotating the axes x, y with an angle about the axis z, as shown in
Figure 3.3.
In the above mentioned case, the orthogonal matrix [qkr ], present in the
general lows (1.1.14) characterizing the connections between the components of a
tensor in the old and new axes have, according to the relations (1.1.8), the following

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

116

CHAPTER 3. COMPOSITE LAMINATES

Figure 3.3: Positive rotation of principal material axes 1, 2 from arbitrary axes
x, y.

form:

cos
[qkr ] = sin
0

sin
cos
0

0
0
1

(3.1.7)

For simplicity, we shall denote by x , y , xy the components 11 , 22 , 12 of


the stress tensor in the coordinate system (x, y, z), and by x , y , xy = xy /2
the components 11 , 22 , 12 of the strain in the same coordinate system (x, y, z).
Taking into account (3.1.7) and the general transformation law (1.1.14) or
its special form (1.1.16), we get

1
x
1
x
2 = [T ()] y , 2
= [T ()] y ,
(3.1.8)
6
xy
6 /2
xy
where the 3 3 square matrix [T ()]

cos2

[T ()] = sin2
sin cos
1

is given by the equation


sin2
cos2
sin cos

2 sin cos
2 sin cos .
cos2 sin2

Denoting by [T ()] the inverse matrix of [T ()] from

x
1
x
y = [T ()]1 2 , y = [T ()]1
xy
6
xy

(3.1.9)

(3.1.8), we get

2
(3.1.10)
6 /2

Taking into account the geometrical significance of the transformation matrix


[T ()] , or by direct computations, it is easy to see that

cos2
sin2
2 sin cos
1
, (3.1.11)
[T ()] = [T ()] = sin2
cos2
2 sin cos
2
2
sin cos sin cos cos sin

Copyright 2004 by Chapman & Hall/CRC

117

3.1. MACROMECHANICAL BEHAVIOR OF A LAMINA

Consequently, (3.1.10) can be expressed in the following equivalent form:

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

x
1
y = [T ()] 2 ,
xy
6
Introducing the Reuters

1 0
[R] = 0 1
0 0
we have

1
1
2 = [R] 2
,
6
6 /2

x
1
y = [T ()] 2
.
xy
6 /2

matrices

0
0 ,
2

[R]

1
= 0
0

0
1
0

0
1/2

x
x
y = [R]1 y , since
xy
xy

(3.1.12)

(3.1.13)

xy = xy /2.

Now, returning to the primary stress-strain relation (3.1.4) and using the
above equations, we successively get

x
1
1
1
y = [T ()] 2 = [T ()] [Q] 2 = [T ()] [Q] [R] 2

xy
6
6
6 /2

x
x
1
= [T ()] [Q] [R] [T ()] y = [T ()] [Q] [R] [T ()] [R] y .
xy
xy

Using (3.1.9), (3.1.11) and (3.1.13), it is easy to see that

sin cos
= [T ()]T .
[R] [T ()] [R]
sin cos
cos2 sin2
(3.1.14)
Consequently, the needed stress-strain relation becomes
1

cos2

= sin2
2 sin cos

sin2
cos2
2 sin cos

Q11
x
 x

y = Q() y = Q12
xy
xy
Q16

with

Q12
Q22
Q26

Q16
x
Q26 y ,
xy
Q66


T
Q() = [T ()] [Q] [T ()]

(3.1.15)

(3.1.16)

Finally, using the relations (3.1.11), (3.1.14) and the last equation, after long,

but elementary computations, we get for the components of the matrix Q() the

Copyright 2004 by Chapman & Hall/CRC

118

CHAPTER 3. COMPOSITE LAMINATES

following expressions:

Q11 () = Q11 cos4 + 2(Q12 + 2Q66 ) sin2 cos2 + Q22 sin4 ,

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

Q12 () = (Q11 + Q22 4Q66 ) sin2 cos2 + Q12 (sin4 + cos4 ),

Q22 () = Q11 sin4 + 2(Q12 + 2Q66 ) sin2 cos2 + Q22 cos4 ,

Q16 () = (Q11 Q12 2Q66 ) sin cos3 + (Q12 Q22 + 2Q66 ) sin3 cos ,

Q26 () = (Q11 Q12 2Q66 ) sin3 cos + (Q12 Q22 + 2Q66 ) sin cos3 ,

Q66 () = (Q11 + Q22 2Q12 2Q66 ) sin2 cos2 + Q66 (sin4 + cos4 ).
(3.1.17)


The matrix Q() is named the transformed reduced stiffness matrix, and its
components Q11 (), ..., Q66 () are the transformed reduced stiffness of the fiberreinforced lamina.
Note that the transformed reduced stiffness matrix has non-vanishing coefficients in all nine positions in contrast to the zeros existing in the primary
reduced stiffness matrix [Q]. However, there are still only four independent material constants since the lamina is orthotropic and it is in a plane stress state. The
stress-strain relation (3.1.15) shows that in general, with arbitrary x, y axis, there
is coupling between normal stresses and shear strains and between shear stresses
and normal strains. Thus, in the coordinates x, y, named in the following body
coordinates, even an orthotropic lamina behaves as would a general anisotropic.
That is the reason why such a lamina is called general orthotropic lamina, even if
it is actually orthotropic.
We observe now that, as an alternative to the foregoing procedure, we can
express in the body coordinates the strains in terms of stresses, by inverting the
relation (3.1.15) and by using the property

[Q]

S11
[S] = S12
0

S12
S22
0

0
0 .
S66

Thus, by using also (3.1.16) and the equation [T ()] = [T ()]

S 11
x
 x

y = S() y = S 12
xy
xy
S 16

with

Copyright 2004 by Chapman & Hall/CRC

S 12
S 22
S 26

S 16
x
S 26 y ,
xy
S 66


T
S() = [T ()] [S] [T ()] .

(3.1.18)

, we obtain

(3.1.19)

(3.1.20)

3.1. MACROMECHANICAL BEHAVIOR OF A LAMINA

119

Now, using (3.1.19), (3.1.18) and the last equation, we obtain

S 11 () = S11 cos4 + (2S12 + S66 ) sin2 cos2 + S22 sin4 ,

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

S 12 () = S11 (sin4 + cos4 ) + (S11 + S22 S66 ) sin2 cos2 ,

S 22 () = S11 sin4 + (2S12 + S66 ) sin2 cos2 + S22 cos4 ,

S 16 () = (2S11 2S12 S66 ) sin cos3 (2S22 2S12 S66 ) sin3 cos ,

S 26 () = (2S11 2S12 S66 ) sin3 cos (2S22 2S12 S66 ) sin cos3 ,

S 66 () = 2(2S11 + 2S22 4S12 S66 ) sin2 cos2 + S66 (sin4 + cos4 ).


(3.1.21)
Note that because of the presence of Q16 , Q26 in (3.1.15), and of S 16 , S 26
in (3.1.19), there is no difference between the behavior of the general orthotropic
lamina and the actually anisotropic lamina in plane stress-state. As for anisotropic
lamina, the coefficients S 11 , .., S 66 of the generally orthotropic lamina can be expressed in terms of the apparent technical or engineering coefficients, introduced
in the following way (see for instance Jones [3.2] Chapter 2 or Lekhnitski [3.8]
Chapter 2):

1
1
yx
xy
1
,
, S 66 =
, S 22 =
=
, S 12 =
Gxy
Ey
Ey
Ex
Ex
xy,y
x,xy
y,xy
xy,x
, S 26 =
=
=
=
.
(3.1.22)
Gxy
Ex
Ey
Gxy

S 11 =

S 16

The mechanical significance of the apparent Young moduli Ex , Ey , the Poisson ratios xy , yx and the shear modulus Gxy is the same as in the case of an
orthotropic material. Obviously, their usual significance must be related to the
coordinate axes x and y.
As can be seen, we have also introduced new engineering coefficients xy,x ,
x,xy , xy,y and y,xy . These material constants are named by Lekhnitski coefficients of mutual influence and are defined as:
i,ij = coefficient of mutual influence of the first kind which characterizes
the stretching in the idirection caused by shear in the ij plane, that is i,ij =
ii /2ij , for ij = , all other stresses being zero, and i 6= j;
ij,i = coefficient of mutual influence of the second kind which characterizes
the shearing in the ij plane caused by a normal stress in the idirection, that
is ij,i = ij /i , for ii = , all other stresses being zero, and i 6= j.
Obviously, the apparent technical moduli depend on the angle by which the
principal mutual directions were rotated.
Using the relation (3.1.22) and the equations (3.1.21), the apparent moduli
can be expressed in terms of the primary engineering moduli of the lamina and

Copyright 2004 by Chapman & Hall/CRC

120

CHAPTER 3. COMPOSITE LAMINATES

the angle . Elementary computations give




212
1
1
1
4
sin2 cos2 +

cos

+
=
E1
G12
E1
Ex

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

xy = Ex
1
Ey

1
Gxy

1
E1

=2

 
sin4 + cos4 E11 +

12
E1

sin4 +

2
E1

xy,x = Ex

xy,y = Ey

n

n

2
E2

1
G12

212
E1

412
E1

212
E1

1
G12

2
E1

212
E1

1
G12

sin2 cos2 +

1
G12

2
E1

1
E2

1
E2

1
G12

1
E2

sin2 cos2 +

sin cos3
sin3 cos

sin4 ,

o
sin2 cos2 ,

cos4 ,
1
G12


sin4 + cos4 ,

2
E2

212
E1

1
G12

2
E2

212
E1

1
G12

o
sin3 cos ,

o
sin cos3 .

(3.1.23)
An important consequence of the presence of the coefficients xy,x and xy,y
is that traction tests in non principal material directions result, not only in axial
extensions and lateral contractions, but also in shear deformations.
Following Jones (see [3.2], Chapter 2), values typical for a glass/epoxy composite (E1 = 3E2 , E2 = 8.27GP a, G12 = 0.5E2 , 12 = 0.25) are plotted in Figure
3.4. In Figure 3.4, Ex is divided by E2 and Gxy by G12 . This normalization permits
an easier analysis of the behavior of the apparent technical moduli as a function
of .
3,0

3,0

2,5

E1
=3
E2
G12
=0.5
E

EX
E2

EX
E2

2,0

xy

=0.25

1,5

2,0

12

xy

Gxy
G12

xy,x

2,5

1,5

G xy
G12
1,0

1,0

0,5

0,5

xy,x
0

15

30

45

60

75

0
90

Figure 3.4: Normalized moduli for glass/epoxy.


The figure shows that xy,x is vanishing at = 00 and = 900 , as is to

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.1. MACROMECHANICAL BEHAVIOR OF A LAMINA

121

be expected, since the laminae actually is orthotropic. Also it can be seen that
at intermediate angles, this coefficient of mutual influence achieves large values as
compared to the apparent Poisson ratio xy . Also, as the first two equations (3.1.23)
show, the transverse axial modulus Ey behaves essentially like the longitudinal one
Ex , with the exception that Ey is small for near 00 and large when is near 900 .
Similar comments can be made for yx and xy,y .
We observe that the behavior presented in the Figure 3.4 is not always typical for all composites, fiber-reinforced laminae. For the considered glass/epoxy
composite, the maximal value of Ex is just E1 . There exist cases where Ex can
actually exceed both E1 and E2 , or can be smaller than both E1 and E2 , for some
orthotropic laminae and some intermediate values of the angle (see P.3.8).
The reduced stiffnesses given in relation (3.1.17) are relatively complicated
functions of the four primary material characteristics E1 , E2 , 12 , G12 , as well as of
the angle of rotation . There exists an ingenious recasting of the stiffness transformations equations that enables a more clear understanding of the consequences of
rotating a lamina in a laminate (see Jones [3.2], Chapter 2). By using elementary
trigonometric identities, the transformed reduced stiffnesses can be expressed in
the following way:

Q11 = U1 + U2 cos 2 + U3 cos 4,


Q12 = U4 U3 cos 4,
Q22 = U1 U2 cos 2 + U3 cos 4,
1
Q16 = U2 sin 2 U3 sin 4,
2
1
Q26 = U2 sin 2 + U3 sin 4,
2
Q66 = U5 U3 cos 4,

(3.1.24)

where
U1 =

U2 =

U3 =

U4 =

U5 =

1
(3Q11 + 3Q22 + 2Q12 + 4Q66 ) ,
8
1
(Q11 Q22 ) ,
2
1
(Q11 + Q22 2Q12 4Q66 ) ,
8
1
(Q11 + Q22 + 6Q12 4Q66 ) ,
8
1
(Q11 + Q22 2Q12 + 4Q66 ) .
8

(3.1.25)

The advantage of writing the expressions of the reduced stiffnesses in the


above form is that these relations show just those parts of Q11 , .., Q66 which rest
invariant under rotation of the lamina. This concept of invariance is useful when
examining the prospect of orienting a lamina at various angles to achieve a certain

Copyright 2004 by Chapman & Hall/CRC

122

CHAPTER 3. COMPOSITE LAMINATES

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

stiffness property. For example, the first equation (3.1.24) shows that the value
of Q11 is determined by a fixed constant, U1 , plus a quantity of low frequency
variation with , U2 cos 2, plus a third quantity, U3 cos 4, of higher frequency
variation with . Hence, U1 is an effective measure of lamina stiffness in a design
application, and it is not being affected by the orientation of the lamina.

3.2

Strength of materials approach

In the Section 3.1, our approach was macromechanical or macroscopic considering the overall properties of a lamina. That is, a large enough piece of the
lamina has been considered as being (macroscopically) homogeneous. The fact
that the lamina is piece-wise homogeneous, being made of two constituent materials (the matrix and the fibers) was neglected. In this sense, we were able to
say that a boron/epoxy composite lamina with unidirectional boron fibers has
certain elasticities and stiffnesses which were experimentally determined. In this
homogenized situation, the following question cannot be asked and cannot be
answered: how can the (effective, equivalent, overall) stiffness of the composite be
varied by changing the amount of boron fibers in the lamina? Because there must
be some rationales (reasons) for selecting a particular stiffness for a particular design application, there must also exist a rationale for determining how to find the
best procedure to achieve that stiffness for a fiber-reinforced lamina. That is, how
can the percentage or the concentration or the volume fraction of the constituent
materials be varied so as to arrive at the desired (overall, macroscopic, equivalent)
stiffness?
There are two methods to answer the above questions which can be characterized as being either micromechanical or macromechanical. In micromechanics,
the composite material behavior is studied taking into account the interaction
of the constituent materials, that is the composite is analyzed as being a (piecewise) heterogeneous body. In macromechanics, the composite material behavior
is analyzed assuming the body as being homogeneous, and the effects of the actual non-homogeneities are taken into account only as averaged apparent, overall,
equivalent properties of the composite.
When using micromechanical methods, the properties of a lamina can be
mathematically derived on the basis of the properties of the constituent materials.
When using macromechanical methods, the properties of a lamina can be experimentally determined is the as mate state. That is, we can predict the lamina
properties by the procedures of micromechanics and we can measure the lamina
properties by mechanical experiments and use the properties obtained by one of
the above methods in a macroscopic analysis of the structure.
Knowledge of how to predict properties is essential in order to construct
composites that must have certain apparent, overall, equivalent or macroscopical
properties. Consequently, micromechanics is a natural approach beside macromechanics when viewed from a design rather than an analysis point of view. Obvi-

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.2. STRENGTH OF MATERIALS APPROACH

123

ously, the real design efficiency is evidenced when the micromechanical predictions
of the properties of the composite agree with the measured properties. Unfortunately, the micromechanical approach has inherent limitation. For example, a
perfect bound between fibers and matrix is a usual analysis restriction that might
not be satisfied by some composites. Thus, the micromechanical predictions must
be validated by careful experimental work.
Nowadays there exist two basic approaches in the micromechanics of composite materials: (i) mechanics (strength) of materials; (ii) elasticity.
The mechanics of materials approach contains simplifying assumptions concerning the hypothesized behavior of the mechanical system.
The elasticity approach is actually: (i) bounding principles; (ii) exact solutions; (iii) approximate solutions. Some of these approaches will be discussed in
detail, for some important cases, in Section 4 devoted to macroscopically homogeneous composites. We shall present bounds for the overall moduli, obtained by Hill,
Hashin and Shtrikman for macroscopically isotropic and transversally isotropic
composites. Exact solutions will also be presented due to Hill and one, derived by
Budiansky and Hill. Also we shall discuss briefly some results obtained by taking
into account various geometrical models of different composite materials.
The final objective of all micromechanical approaches is to determine the
overall (equivalent, macroscopic, effective) elastic moduli or stiffness of a composite material in terms of the elastic moduli and concentrations of the constituent
bij of
materials or phases. For example, the overall elastic moduli, designed by C
a fiber-reinforced composite lamina must be expressed in terms of the fibers and
matrix moduli and their concentrations
bij = C
bij (Em , m , Ef , f , cm , cf ) ,
C

where Em , m and Ef , f are Youngs moduli and Poissons ratios of the matrix
and of the fibers, respectively, and
cm = vm /v , cf = vf /v
represent the concentration or volume fractions of the matrix and of the fibers,
respectively, v, vm , vf being the volumes occupied by the lamina, the matrix and
the fibers, respectively.
As we shall see, the above problem generally cannot be solved without introducing unrealistic assumptions, used in the strengths of materials. The overall
properties obtained in this way, generally do not agree with the measured ones.
This is the main reason why the much powerful approach formulated on the base of
elasticity and on the theory of macroscopically homogeneous composite materials
must be involved. In this way, generally, we can derive lower and upper bounds
for the overall moduli, and if these bounds are close, the obtained results can be
used in the design.
According to the micromechanical approach used, we must impose some basic restrictions on the composite material that can be treated, using the methods

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

124

CHAPTER 3. COMPOSITE LAMINATES

of strength of materials or those of elasticity theory. For instance, in the case of


a fiber-reinforced lamina we assume that: (i) the matrix is linearly elastic, homogeneous and isotropic or transversally isotropic; (ii) the fibers are linearly elastic, homogeneous and isotropic or transversally isotropic, and perfectly aligned;
(iii) the lamina is macroscopically linearly elastic, homogeneous and transversally
isotropic or orthotropic. We suppose also that no voids can exist in the fibers or
in the matrix or between them, and the fibers and matrix are firmly bounded
together.
Basic in the discussion of micro and macromechanics of a macroscopically homogeneous composite is its representative volume element (RVE). Roughly speaking the RVE is the smallest region or piece of composite material over which
the stresses and strains are macroscopically uniform. However, it is obvious, that
microscopically the stresses and strains are nonuniform in the RVE, due to the
heterogeneity of the composite material. Thus, the scale of the RVE is very important. Other concepts concerning the characteristics of the RVE, if they exist,
will be presented and discussed in Chapter 4 concerning the elasticity approach
for macroscopically homogeneous composites.
Here we shall present and discuss briefly only the mechanics of material approach to the micromechanics for the overall material stiffnesses. In this way, we
shall obtain very simple, but generally unrealistic approximations, to the effective
engineering constants of the fiber-reinforced lamina, assumed to be macroscopically orthotropic. For simplicity, the matrix and the fibers are supposed to be
homogeneous and isotropic. In this Section, the mechanical and geometrical characteristics of the matrix will be designed by m, and those of the fibers by f .
As we already know, the key feature of the mechanics of material approach is
that certain simplifying assumption are made regarding the mechanical behavior of
a composite material. Using this procedure we can derive the mechanics of material
expression for the overall orthotropic moduli of the unidirectionally reinforced
fibrous composite material.
It is assumed that the RVE contains only one fiber.
b1
Determination of E
The first overall modulus to be determined is that of the composite in the
fiber direction. We suppose that the axial strain 1 in the fiber direction is the
same in the matrix and in the fiber. Such a hypothesis was first made by Voigt in
1910. From the Figure 3.5, we get
1 =

4L
,
L

where 1 is the axial strains for both the fibers and the matrix, according to the
basic Voigt type assumption. Then, the axial stresses m and f in the matrix and
in the fiber are
m = E m 1 , f = E f 1 .

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.2. STRENGTH OF MATERIALS APPROACH

125

Figure 3.5: Representative volume element loaded in the 1-direction.

The average axial stress 1 acts on the cross sectional area S, m acts on the cross
sectional area Sm of the matrix, and f acts on the cross sectional area Sf of the
fiber. Thus, the resultant axial force F on the RVE is
F = 1 S = m Sm + f Sf .
Using the obtained results and taking into account that according to the
b1 , we have
definition of the overall axial moduli E
we get

b1 1 ,
1 = E

b1 = Sm Em + Sf Ef .
E
S
S
But the concentrations or volume fractions cm = vm /v and cf = vf /v of the
matrix and of the fibers can be expressed as

cm =

Sf
Sm
,
, cf =
S
S

where v, vm and vf are the volumes occupied by the RVE, by the matrix and by
the fiber, respectively. In this way, finally we get
b1 = cm Em + cf Ef .
E

(3.2.1)

This expression for the overall (apparent, equivalent, macroscopical) Young modulus in direction of the fibers is known as the rule of mixture or as the Voigt type
estimate. This rule leads to a simple linear variation of the overall Young modulus
b1 from Em to Ef as the fibers concentration cf varies from 0 to 1.
E
We stress the fact that, according to its definition, the overall axial modulus
b1 connects the mean stress and the mean strain, evaluated on the RVE of the
E

Copyright 2004 by Chapman & Hall/CRC

126

CHAPTER 3. COMPOSITE LAMINATES

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

composite material. In the elasticity approach of the problem, the overall moduli
will be introduced in the same way!
b2
Determination of E
b2 , in the direction transverse
We now consider the overall Young modulus E
to the fibers. In the mechanics of the material approach, the same transverse stress
2 is assumed to be applied to both the matrix and the fiber, as shown in Figure
3.6. Such kind of hypotheses was first made by Reuss in 1929.

Figure 3.6: Representative volume element loaded in 2-direction.


The transverse strains m and f , in the matrix and in the fiber, respectively,
are therefore
2
2
.
, f =
m =
Ef
Em

The transverse direction over which on the average m acts is approximately cm W ,


whereas f acts on cf W . Thus, the total transverse deformation is
2 W = c m W m + c f W f ;
Hence, the mean transverse deformation 2 becomes
2 = c m m + c f f .
Introducing here the stress-strains relations, we get
2 = c m

2
2
.
+ cf
Ef
Em

b2 must satisfy
Recognizing that according to its definition the overall moduli E
the material law
1
,
2 =
b2 2
E
Copyright 2004 by Chapman & Hall/CRC

3.2. STRENGTH OF MATERIALS APPROACH

127

finally we get

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

1
1
1
.
+ cf
= cm
b
E
E
f
m
E2

(3.2.2)

b2 in the transverse direction


This expression for the overall Young modulus E
of fibers is known as the Reuss type estimate.
Obviously, we have
cm + cf = 1.
Hence if cm = 1, that is cf = 0, according to the above rule, the overall
modulus predicted is that of the matrix; if cf = 1, hence cm = 0, the modulus
predicted is that of the fibers. However, now the rule does not represent a linear
b2 as cf goes from 0 to 1. Let us observe
variation of the overall Young modulus E
also that according to the Reuss type estimate, more than 50 percent by volume of
b2 to twice the matrix modulus,
fibers is required to raise the transverse modulus E
even if Ef = 10Em . That is, if 2 is a tensile test, the fibers cannot contribute
much to the overall transverse modulus unless their percentage is very high, and,
obviously the bound between the fibers and the matrix is perfect. In exchange, no
such bound is needed if 2 is a compression test!
Obviously, the assumptions involved in the foregoing derivation are not entirely consistent, since the transverse stresses in the matrix and in the fibers are
not the same. Indeed, if the Poisson ratios of the matrix and the fibers are not
the same, then longitudinal stresses are introduced in the matrix and fiber, with
accompanying shear stresses at the matrix-fiber boundary. Such shearing stresses
lead to a stress state much more complicated than that assumed in our derivation.
The consequence of such inconsistent assumptions can be measured only by comparison with experimental results.
Determination of b12
The overall Poisson ratio b12 can be determined using the assumption made
b1 ; that is, supposing that the axial strains in the matrix and the fiber
to obtain E
are the same, 1 . Denoting by 2 the (mean) transverse strain of the RVE, b12 is
defined by
2
b12 = ,
1

for the stress state 1 6= 0 and all over stresses are zero.
According to the Figure 3.7, the transverse deformation 2 is
2 =

W
= b
12 1 .
W

We also have
W = Wm + Wf ,

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

128

CHAPTER 3. COMPOSITE LAMINATES

Figure 3.7: Representative volume element loaded in 1-direction.

Wm and Wf being the transverse displacements of the matrix and of the fiber,
respectively. Consequently

Wf
Wm
= b12 1 .
+
W
W

Following the same procedure as in analysis for the overall transverse Young
b2 , we assume that the transverse displacements Wm and Wf are
modulus E
approximately
Wm = W cm m 1 , Wf = W cf f 1 ,

m 1 and f 1 being the average transverse deformations of the matrix and


of the fiber, respectively. Combining the last equations, we get
b12 = cm m + cf f .

(3.2.3)

The strength of materials rule leads to the mixture rule or to the Voigt type
estimate of the overall Poisson ratio b12 .

b 12
Determination of G
b 12 of a lamina is estimated in the meThe overall in-plane shear modulus G
chanics of materials approach by assuming a Reuss type hypothesis. It is supposed
that the same shear stress acts in the matrix and in the fiber. Denoting by m
and f , the shear strains in the matrix and fiber, respectively, we get
m =

1
1
.
, f =
f
m

The loading is shown in Figure 3.8 and the deformations on microscopic scale in
Figure 3.9. The total (mean) shear deformation is
=

Copyright 2004 by Chapman & Hall/CRC

,
W

129

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.2. STRENGTH OF MATERIALS APPROACH

Figure 3.8: Representative volume element loaded in shear.


FIBER
MATRIX

m /2

f
MATRIX

Figure 3.9: Shear deformation of a representative volume element.


and we have
= m + f ,
m and f being the horizontal displacements of the matrix and of the fiber,
respectively. Denoting by m and f the shear strains in the matrix and fiber,
respectively, we approximately get
m = c m W m f = c f W f .
Hence,
= c m m + c f f .
b 12 , connecting the mean strain and
The overall in-plane shear modulus G
the mean stress , is defined by
=

1
.
b 12
G

Thus, using the above formulas, we obtain

1
1
1
+ cf .
= cm
b 12
f
m
G
Copyright 2004 by Chapman & Hall/CRC

(3.2.4)

130

CHAPTER 3. COMPOSITE LAMINATES

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

Material
Ef
f
Em
m
cf
Methods
Experimental
Mixture rules

Material
Ef
f
Em
m
cf
Methods
Experimental
Mixture rules

b1
E

49.40
47.16

Glass/Epoxy
70
0.17
2.85
0.33
0.66
b2
b 12 b12
E
G
18.00
7.77

7.80
2.95

Boron/Epoxy
413
0.2
4.10
0.35
0.70
b 12
b2
b1
G
E
E
237.8
290

13.3
26.7

5.5
12.2

0.22
0.224

b1
E

Carbon/Epoxy
234
0.2
3.8
0.33
0.6
b2 G
b 12 b12
E

151
141.8

9.3
9.2

6.2
3.5

0.32
0.25

b12

0.245

Table 3.4: Experimental and calculated values of the overall elastic coefficients.

The strength of materials approach leads to a Reuss type estimation for the
b 12 .
overall in-plane shear modulus G
b2 , only for a fiber volume greater than 50 percent of the
As in the case of E
b 12 rise to above twice m even if f = 10m .
total volume does G
Using the data given by Barran and Loroze [3.9], we present in Table 3.4 the
mechanical characteristics of three fiber-reinforced composite materials, giving also
the fiber concentrations. We also give the overall elastic coefficients experimentally
determined and the values of the overall moduli calculated using the mixture rules
b1 and
obtained by the strength of materials approach. The axial Young modulus E
the transverse Poisson ratio b12 are evaluated taking into account the Voigt type
b2 and the inmixture rules (3.2.1), (3.2.2), and the transverse Young modulus E
b
plane transverse shear modulus G12 are obtained using the Reuss type mixture
rules (2.3.3), (3.2.4). The axial and transverse Young moduli, as well as the shear
modulus are expressed in GP a = 109 P a.
Examining the above data, we can see that the calculated values of the overall
b1 and those of the overall transverse Poisson ratio b12 are
axial Young modulus E
acceptable as first approximations. However, the calculated values of the overall
b2 and those of the overall transverse shear modulus
transverse Young modulus E
b
G12 are not acceptable, and cannot be used as a first approximation. Generally,
we can say that much more powerful methods are necessary to evaluate and/or
to bound the overall moduli as those obtained with the strength of materials
approach.

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.3. GLOBAL CONSTITUTIVE EQUATIONS

131

The above estimations are only examples of the type of mechanics of materials
approaches that can be used to obtain approximate expression for the overall
moduli. Other assumptions of mechanical behavior lead to different estimations
for the overall elastic moduli of the lamina.
The true significance of the Voigt and Reuss type estimates can be clarified
only by using the elasticity approach to get the overall stiffnesses. As we shall
see in Section 4.1, the Voigt and Reuss type estimations give universal-bounds for
the overall moduli. Generally, these estimates are the worst bounds that can be
derived by the elastic approach.
We end this Section with some words concerning the approach named netting
analysis (see Jones [3.2], Chapter 3, Section 3.3.1). The basic assumption in netting
analysis is that the fibers provide all the longitudinal stiffness and the matrix
provide all the transverse and shear stiffness as well as the Poisson effect. Even
on the base of the above results furnished by the mechanics of material, we can
see that the assumptions made by the netting analysis must generally be rejected.
In turn, the results due to the strength of materials approach must be carefully
analyzed in light of the elasticity approach. Some important results of this analysis
will be presented in Chapter 4.

3.3

Global constitutive equations

As we have seen, a laminate is composed of two or more laminae bounded


together to act as a structural element. The constituent laminae are oriented to
produce a structural element capable of resisting load in several directions. The
stiffness of such a composite body results from the properties of the constituent
laminae, as well as from their relative orientations. In the following, we present the
basic formulation of the classical lamination theory. The major difference between
this theory and the classical theory of homogeneous isotropic plates is in the form
of the stress strain relationships of the lamina. Other elements of the theory such as
the deformation hypothesis, the equilibrium equation and the strain displacement
relationships are the same as those used in the classical plate theory.
Although the laminate is made up of multiple laminae, it is assumed, that
the individual laminae are perfectly bounded together so as to behave as a unitary,
nonhomogeneous, anisotropic plate. Interfacial slip is not allowed and the interfacial bounds are not allowed to deform in shear, which mean that the displacement
across laminae interfaces are assumed to be continuous. The assumptions imply
that deformation hypothesis from the classical homogeneous plate theory can be
used for the laminated composite plate.
Figure 3.10 shows the coordinate system to be used in developing the laminated plate analysis. The x1 , x2 , x3 coordinate system is assumed to have its origin
on the middle surface of the plate, so that the x1 x2 planes lie in the middle plane.
The components of the displacement u are u1 , u2 , u3 and they depend on x1 , x2 , x3 .
Frequently, x1 , x2 , x3 are denoted by x, y, z and u1 , u2 , u3 by u, v, w.

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

132

CHAPTER 3. COMPOSITE LAMINATES

Figure 3.10: Coordinate system for laminated plate.

The basic assumptions made in the frame-work of the classical composite


laminate theory are the following:
(1) The plate consists of orthotropic laminae bounded together, with the
principal material axes of the orthotropic laminae oriented at arbitrary direction
with respect to the x1 , x2 axes.
(2) The thickness h of the plate is much smaller than the length along the
plate edges a and b.
(3) The displacements u1 , u2 , u3 are small compared with the plate thickness
h.
(4) The in-plane strains 11 , 22 , 12 are small compared with unity.
(5) Transverse shear strains 13 and 23 are negligible.
(6) The transverse normal strain 33 is negligible.
(7) The normal stress 33 is small in comparison with the other stress components.
(8) The transverse shear stresses 13 and 23 vanish on the plate surfaces
x3 = h2 .
(9) Each lamina obeys the reduced stress-strain relation corresponding to
plane stress state.
The assumption (2) stresses the fact that we develop here the classical thin
lamination theory. The assumptions (3) and (4) show that the theory refers to
small deformations, that is it is geometrically linear. The assumptions (5) and (6)
express the classical Love-Kirchhoff hypothesis, known also as the hypothesis of
plane sections: any normal to the middle surface remains straight and normal to
the deformed middle surface, and at the same time, its magnitude rests constant
during the deformation. The assumptions (7), (8) and (9) express the fact the
stresses 13 , 23 and 33 are assumed to be small in comparison with the stresses
11 , 22 and 12 . That is, as is stated in the assumption (9), the stresses 11 , 22 and
12 and the strains 11 , 22 and 12 can be related using the reduced stress-strain
relatives corresponding to the plane stress state of the laminae. Since the assumed
hypotheses are similar to those used in the classical Love-Kirckhhoff theory of homogeneous isotropic thin plate, the classical lamination theory of composite laminates inherits all internal contradictions and inconsistencies of the Love-Kirckhhoff

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.3. GLOBAL CONSTITUTIVE EQUATIONS

133

theory. This observation concerns all of the internal contradiction existing between
the assumptions (7), (8) and (9): though 13 , 23 and 33 are not vanishing, we
use the reduced stress-strain relations corresponding to vanishing 13 , 23 and 33 .
Obviously, the seriousness and consequences of these inconveniences can be established only by studying the implication of the theory based on the assumptions
(1)-(9). For this purpose, we must first develop the classical lamination theory,
using the supposed hypothesis.
From assumption (6), we obtain
33 (x1 , x2 , x3 ) =

u3
(x1 , x2 , x3 ) = 0.
x3

Consequently, the normal displacement u3 (x1 , x2 , x3 ) depends only on x1 and x2 ;


i.e.
u3 = U3 (x1 , x2 ).
(3.3.1)
From the assumption (5), we get
213 (x1 , x2 , x3 ) =

u3
u1
(x1 , x2 , x3 ) = 0,
(x1 , x2 , x3 ) +
x1
x3

223 (x1 , x2 , x3 ) =

u3
u2
(x1 , x2 , x3 ) = 0.
(x1 , x2 , x3 ) +
x2
x3

From here, according to (3.3.1), the displacements u1 (x1 , x2 , x3 ) and


u2 (x1 , x2 , x3 ) depend linearly on x3 ; i.e.
u1 = U1 (x1 , x2 ) x3

U3 (x1 , x2 )
U3 (x1 , x2 )
.
, u2 = U2 (x1 , x2 ) x3
x2
x1

(3.3.2)

In the above relations, U3 (x1 , x2 ) is the normal displacement of the middle surface,
and U1 (x1 , x2 ), U2 (x1 , x2 ) characterize the tangential displacement of the same
surface.
From (3.3.2), we obtain the following expressions for the non-vanishing strain
components 11 , 22 , 33 :
= e + x3 k , , = 1, 2,
where
e = e (x1 , x2 ) =

1
U
1 U
) = (U, + U, )
+
(
2
x
2 x

(3.3.3)

(3.3.4)

describe the deformation of the middle surface x3 = 0, and


k (x1 , x2 ) = k (x1 , x2 ) =

are the curvatures of the same surface.

Copyright 2004 by Chapman & Hall/CRC

U3
= U3,
x x

(3.3.5)

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

134

CHAPTER 3. COMPOSITE LAMINATES


In the following, the Greek indices take the values 1 and 2.
For later use, we introduce the following matrix notations:

x
11
x
11
[] = y = 22 , [] = y = 22 ,
xy
12
xy
212

e11
[e] = e22 ,
2e12

Thus from (3.3.3) we get

k11
[k] = k22
2k12

[] = [e] + x3 [k] .

(3.3.6)

(3.3.7)

In Figure 3.11, we present the geometry of an N-layered laminate, clarifying in


this way the relations which will be used in what follows. Occasionally, to simplify
some formulas, we shall use the notation x3 = z.

Figure 3.11: Geometry of an N -layered laminate.


The k-th lamina occupies the domain defined by
zk1 < z < zk , k = 1, ..., N with x3 = z.
Obviously, z0 = h2 and zN = h2 .
We return now to assumption (9). According to this hypothesis, in each
lamina the reduced and transformed stress-strain relation (3.1.15) is valid. Hence,
we have

Q11 Q12 Q16


x
x
y = Q12 Q22 Q26 y for k = 1, .., N.
(3.3.8)

xy k
Q16 Q26 Q66 k
xy
k

Using the simplified matrix notation (3.3.6), we get


 
[]k = Q k [k ] for zk1 < x3 = z < zk and k = 1, .., N.

Copyright 2004 by Chapman & Hall/CRC

(3.3.9)

135

3.3. GLOBAL CONSTITUTIVE EQUATIONS

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

Since according to (3.3.7), e and k depend on x1, x2 only, the last equation becomes
 
 
[]k = Q k [e] + x3 Q k [k] for zk1 < x3 = z < zk and k = 1, .., N. (3.3.10)

The last equation expresses the plane stress 11 , 22 , 12 in the k-th layer, in
terms of the laminate middle surface strains and curvatures.
Expanded, the equation (3.3.10) becomes

k11
Q11 Q12 Q16
e11
Q11 Q12 Q16
11
22 = Q12 Q22 Q26 e22 +x3 Q12 Q22 Q26 k22
12 k
Q16 Q26 Q66 k 2k12
Q16 Q26 Q66 k 2e12
(3.3.11)
for zk1 < x3 = z < zk and k = 1, .., N.
In the above equations, k denotes the k-th lamina, ( )k , , = 1, 2 are the
stress in the k-th lamina, (Qij )k , i, j = 1, 2, 6 are the transformed reduced stiffness
of the k-th lamina, zk1 and zk are the distances from the middle surface to the
inner and to the outer surfaces of the k-th lamina, respectively, and N is the total
number of the laminae.
We recall that the reduced stiffness Qij , i, j = 1, 2, 6 depend on , the angle
made by the fibers with axis Ox1 , and we have

(Qij )k = Qij (k ) for zk1 < x3 = z < zk and k = 1, ...N,

(3.3.12)

k representing the angle made by the fibers in the k-th lamina and the body axis
x1 .
Since (Qij )k can be different for each lamina of the laminate, the stress
variation through the thickness is not necessarily linear, even though the strain
variation is linear, as can be seen by examining equation (3.3.3).
In the laminated plate analysis, it is convenient to use the forces N and
the moments M per unit length, defined by the following relations:

N = N =

h
2

h
2

dx3 , M = M =

h
2

h
2

x3 dx3 , , = 1, 2. (3.3.13)

Let us observe that usually N11 , N12 = N21 , N22 are denoted by Nxx , Nxy =
Nyx and Nyy , respectively, and also M11 , M12 = M21 , M22 are denoted by Mxx , Mxy
= Myx , Myy , respectively.
According the relation (3.3.13)1 N11 , N12 , N22 are forces per unit length of
the cross-section. The mechanical meaning of these force resultants are shown in
the Figure 3.12.
Similarly, equation (3.3.13)2 shows that M11 , M12 , M22 are moments per unit
length of the cross-section. In Figure 3.13 is shown the mechanical meaning of
these moment resultants.

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

136

CHAPTER 3. COMPOSITE LAMINATES

Figure 3.12: In-plane forces on a flat laminate.

Figure 3.13: Moments on a flat laminate.

The relations (3.3.13) show that these force and moments resultants do not
depend on x3 , but are functions of x1 and x2 , the in the plane coordinates of the
laminate middle surface.
In more detail, the defining equations (3.3.13) can be written as

N Z zk
N11
11
X
22 dz;
[N ] = N22 =
z
k1
N12
12 k
k=1

Z
N
M11
11
zk
X
[M ] = M22 =
z 22 dz.
z
k1
M12
12 k
k=1

(3.3.14)

The integrations indicated in these equations can be rearranged to take advantage of the fact that the stiffness matrix for a lamina is constant within each
lamina. Thus, substituting the stress-strain relations (3.3.11) and taking into ac-

Copyright 2004 by Chapman & Hall/CRC

137

3.3. GLOBAL CONSTITUTIVE EQUATIONS

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

count the fact that e and k do not depend on x3 = z, we get


)
)
(N
(N
Z zk
Z zk
X
X
zdz [k],
dz [e] +
[Q]k
[N ] =
[Q]k

[M ] =

zk1

k=1

N
X

[Q]k

k=1

zk

zdz
zk1

[e] +

zk1

k=1

N
X

k=1

[Q]k

zk

z dz
zk1

Finally, these equations can be written as


N11
A11 A12 A16
e11
B11
[N ] = N22 = A21 A22 A26 e22 + B21
N12
A61 A62 A66
2e12
B61


M11
B11
[M ] = M22 = B21
M12
B61

B12
B22
B62

B12
B22
B62


B16
e11
D11
B26 e22 + D21
B66
2e12
D61

D12
D22
D62

[k].

B16
k11
B26 k22 ,
B66
2k12

D16
k11
D26 k22 ,
D66
2k12
(3.3.15)

where the coefficients Aij , Bij , Dij , i, j = 1, 2, 6 are defined by


N
P

(Qij )k (zk zk1 ),


k=1
N
P
2
),
(Qij )k (zk2 zk1
Bij = Bji = 12
k=1
N
P
3
).
(Qij )k (zk3 zk1
Dij = Dji = 13
k=1

Aij = Aji =

Introducing the symmetric 3 3 matrices

A11 A12 A16


B11 B12 B16
D11
[A] = A12 A22 A26 , [B] = B12 B22 B26 , [D] = D12
A61 A26 A66
B61 B26 B66
D61

(3.3.16)

D12
D22
D26

D16
D26
D66
(3.3.17)

the equations (3.1.15) can be expressed in a concentrated matrix form


[N ] = [A][e] + [B][k], [M ] = [B][e] + [D][k]

(3.3.18)

the 3 1 matrixes [e] and [k] being defined by equation (3.3.6)3,4 .


Also, the system (3.3.18) can be replaced by the following matrix equation:

N11
A11 A12 A16 B11 B12 B16
e11
N22 A21 A22 A26 B21 B22 B26 e22

N12 A61 A62 A66 B61 B62 B66 2e12

M11 B11 B12 B16 D11 D12 D16 k11 . (3.3.19)

M22 B21 B22 B26 D21 D22 D26 k22


M12
B61 B62 B66 D61 D62 D66
2k12

Copyright 2004 by Chapman & Hall/CRC

138

CHAPTER 3. COMPOSITE LAMINATES

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

This equation describing the global behavior of the laminate, can be expressed
in the concentrated form

 




N
A B
e
e
=
= [E]
.
(3.3.20)
M
B D
k
k
The 6 6 symmetric matrix [E] is the global laminate stiffness matrix.
The coefficients Aij are called extensional stiffnesses, the coefficients Bij describe the coupling stiffness, and the coefficients Dij are called bending stiffnesses.
The presence of the coefficients Bij implies coupling between bending and extension
of a laminate. That is, it is impossible to pull on a laminate that has non-vanishing
Bij terms, without at the same time bending and/or twisting the laminate. Thus,
an extensional force results not only in extensional deformation of the middle surface, but also in twisting and/or bending of the laminate. Also, such a laminate
cannot be subjected to a moment without at the same time being subjected to
an extension of the middle surface. The experiments made with laminates confirm
these theoretical predictions. In spite of this fact, in the stability analysis of laminates, this coupling is generally neglected and we shall discuss this question in
Chapter 7.
It is easy to see that the matrix equation (3.3.19) can be written in the
following tensorial or component form, very useful in many problems
N = A e + B k ,
M = B e + D k , , , , = 1, 2.

(3.3.21)

The coefficients of these equations can be expressed simply and obviously by


using the quantities A11 , ..., D66 . For instance, we have
A1111 = A11 , A1122 = A2211 = A12 ,
A1112 = A1121 = A1211 = A2111 = A16 ,
A1212 = A1221 = A2112 = A2121 = A66 , ...,
B1111 = B11 , B1122 = B2211 = B12 ,
B1112 = B1121 = B1211 = B2111 = B16 ,
B1212 = B1221 = B2112 = B2121 = B66 , ...,
D1111 = D11 , D1122 = D2211 = D12 ,
D1112 = D1121 = D1211 = D2111 = D16 ,
D1212 = D1221 = D2112 = D2121 = D66 , ....

(3.3.22)

It is also clear that the following symmetry relations take place


A = A = A = A ,
B = B = B = B ,
D = D = D = D .

Copyright 2004 by Chapman & Hall/CRC

(3.3.23)

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.3. GLOBAL CONSTITUTIVE EQUATIONS

139

We stress the fact that the constitutive coefficients A , B , D


have the same symmetries as the elasticities of a linearly elastic material. As we
shall see later, this property will have important consequences.
We assume now that the global constitutive equation (3.3.20) is invertible. In
Chapter 7, we shall see that this property is always true if the initial configuration
of the laminate is (locally) stable.
In order to express [e] and [k] in terms of [N ] and [M ], we rewrite (3.3.20) as
[N ] = [A][e] + [B][k], [M ] = [B][e] + [D][k].

(3.3.24)

From the first equation we get


[e] = [A]1 [N ] [A]1 [B][k].

(3.3.25)

Substitution of (3.3.25) in (3.3.24)2 gives


[M ] = [B][A]1 [N ] [B][A]1 [B][k] + [D][k].

(3.3.26)

Equation (3.3.25) and (3.3.26) give a partially inverted form of the equation
(3.3.20)

 


e
A B
N
=
,
(3.3.27)
M
C D
k

with

[A ] = [A]

, [B ] = [A]

[B] , [C ] = [B][A]1 , [D ] = [D] [B][A]1 [B].


(3.3.28)
Now, using (3.3.27) for [k], we get
[k] = [D ]1 [M ] [D ]1 [C ][N ].
Introducing (3.3.29) in (3.3.25), we obtain


[e] = [A ] [B ][D ]1 [C ] [N ] + [B ][D ]1 [M ].

(3.3.29)

(3.3.30)

Finally, (3.3.29) and (3.3.30) lead to the following inverted global constitutive
equation.

  0




e
A B0
N
N
1
=
=
[E]
,
(3.3.31)
k
C 0 D0
M
M

where

[A0 ] = [A ] [B ][D ]1 [C ],
[B 0 ] = [B ][D ]1 ,
[C 0 ] = [D ]1 [C ] = [B 0 ]| = [B 0 ],
[D0 ] = [D ]1 .

(3.3.32)

The last results show that the 6 6 global compliance matrix [E]1 is symmetric. This is an obvious result, since if the global stiffness matrix [E] being
symmetric, its inverse, if it exists, must also be symmetric.

Copyright 2004 by Chapman & Hall/CRC

140

CHAPTER 3. COMPOSITE LAMINATES

3.4

Special classes of laminates

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

This Section considers special classes of laminates for which the stiffnesses
can easily be calculated. The special classes will be presented in increasing order
of complexity.
1. Single - layered configurations
For a single isotropic layer with material properties E, and thickness h,
equations (3.3.16) give
A11 = A22 =

1
Eh
A , Bij = 0 ,
A, A12 = A , A16 = A26 = 0 , A66 =
2
2
1

1
Eh3
D.
D , D12 = D , D16 = D26 = 0 , D66 =
2
2
12(1 )
(3.4.1)
In order to obtain the above relations, we must use the equations (3.1.6),
supposing an isotropic material.
From (3.4.1), we can conclude that the resultant forces depend only on the
in-plane strains of the laminate middle surface, and the resultant moments depend
only on the curvatures of the middle surface. There is no coupling. The constitutive
equations become


A A
0
e11
N11
e22 ,
N22 = A A
0
2e12
N12
0
0 (1 ) A2
(3.4.2)


D D
0
k11
M11
k22 .
M22 = D D
0
D
2k12
M12
0
0 (1 ) 2
D11 = D22 =

In particular, we have

h2
A.
(3.4.3)
12
For a simple specially orthotropic layer of thickness h the lamina stiffnesses
are given by equation (3.1.6). Hence, according to (3.3.10), the laminate stiffnesses
are
D=

A11 = hQ11 , A12 = hQ12 , A22 = hQ22 , A16 = A26 = 0 , A66 = hQ66 ,
Bij = 0,
D11 =

h3
h3
h3
h3
Q66 .
Q22 , D16 = D26 = 0 , D66 =
Q12 , D22 =
Q11 , D12 =
12
12
12
12
(3.4.4)

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.4. SPECIAL CLASSES OF LAMINATES

141

Again, the resultant forces depend only on the in-plane strains, and the resultant moments depend only on the curvatures. There is no coupling. The constitutive equation becomes


e11
N11
A11 A12
0
N22 = A12 A22
0 e22 ,
0
0
A66
2e12
N12
(3.4.5)

M11
D11 D12
0
k11
M22 = D12 D22
0 k22 .
0
0
D66
M12
2k12
2. Symmetric laminate
For laminates that are symmetric in both geometry and material properties
about the middle surface, the general stiffness equations (3.3.16) simplify considerably. Because of the symmetry of the transformed stiffnesses (Qij )k and of the
thicknesses hk , it can be shown that all coupling stiffness Bij of the laminate are
zero. There is no coupling. Obviously such laminates are much easier to analyze
than laminates with coupling. Consequently, symmetric laminates are commonly
used unless special circumstances require an unsymmetrical laminate possessing
the coupling property.
The constitutive equations for a symmetric laminate are

N11
A11 A12 A16
e11
N22 = A12 A22 A26 e22 ,
N12
A16 A26 A66
2e12
(3.4.6)

M11
D11 D12 D16
k11
M22 = D12 D22 D26 k22 .
M12
D16 D26 D66
2k12

In the following, we shall present some special cases of symmetric laminates,


determining the stiffness Aij and Dij in each case.
For symmetric laminates with multiple isotropic layers, multiple isotropic
laminae of various thicknesses are arranged symmetrically about the middle surface
from both a geometric and a material property standpoint. The resulting laminate
does not exhibit coupling between extension and bending. The extensional and
bending stiffnesses are calculated from the equations (3.3.16), where, according to
(3.1.6), and assuming an isotropic material, for the k-th layer, we get

Ek
, (Q16 )k = (Q26 )k = 0 ,
1 k2
Ek
k Ek
.
, (Q66 )k =
(Q12 )k =
2(1 + k2 )
1 k2

(Q11 )k = (Q22 )k =

Copyright 2004 by Chapman & Hall/CRC

(3.4.7)

142

CHAPTER 3. COMPOSITE LAMINATES

In these equations, Ek and k are the Youngs modulus and the Poissons
ratio for the k-th lamina.
It is easy to see that

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

A11 = A22 , A16 = A26 = 0 , D11 = D22 , D16 = D26 = 0.


Hence, the constitutive equations become

N11
A11 A12
0
e11
N22 = A12 A11
0 e22 ,
0
0
A66
N12
2e12

D11
M11
M22 = D12
0
M12

D12
D11
0

k11
0
0 k22 .
2k12
D66

(3.4.8)

(3.4.9)

A symmetric laminate with multiple specially orthotropic layers is made of


orthotropic layers that have their principal material directions aligned with the
laminate axes, and the layers (laminae) having various thicknesses, are arranged
symmetrically about the middle surface both from a geometric and a material
property standpoint. The stiffnesses of the laminate are calculated from the general
equations (3.3.16), whereas, according to (3.1.6) for the k-th lamina
(Q11 )k =

k
E2k
21
E1k
E1k
,
, (Q22 )k =
, (Q12 )k =
k k
k
k
k
k
1 12
1 12 21
1 12 21
21

(Q66 )k = Gk12 , (Q16 )k = (Q26 )k = 0 ,

(3.4.10)

k
k
E1k , E2k , 12
, 21
and Gk12 being the engineering material constants of the k-th specially orthotropic lamina.
Because (Q16 )k and (Q26 )k are zero, it is easy to see that A16 , A26 , D16 and
D26 vanish; i.e.
A16 = A26 = 0 , D16 = D26 = 0.
(3.4.11)

Also, because of symmetry, the coupling stiffnesses Bij are all zero; i.e.
Bij = 0.

(3.4.12)

Hence, the constitutive equation for the laminate takes the form

N11
A11 A12
0
e11
N22 = A12 A22
0 e22 ,
N12
0
0
A66
2e12


M11
D11
M22 = D12
0
M12

Copyright 2004 by Chapman & Hall/CRC

D12
D22
0

k11
0
0 k22 .
D66
2k12

(3.4.13)

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.4. SPECIAL CLASSES OF LAMINATES

143

Taking into account the above equations, this type of laminate could be called
specially orthotropic laminate in analogy to a special orthotopic lamina.
A regular symmetric cross-ply laminate represents a very common special case
of symmetric laminates with multiple specially orthotropic laminae (layers). The
regular symmetric cross-ply laminate occurs when the laminae are all of the same
thickness and material properties, and their major principal material direction
(that is, the fiber directions) alternate at 00 or 900 with respect to the laminate
(body) axes, for examples (00 /900 /00 ) as in Figure 3.14.

Figure 3.14: Exploded (unbounded) view of a three-layered regular symmetric


cross-ply laminate.
The regular symmetric cross-ply laminate must have an odd number of layers
if we wish to satisfy the symmetric requirement by which coupling between bending
and extension is eliminated. Cross-ply laminates with an even number of layers
are not symmetric and will be discussed a little later on.
Before analyzing other special classes of laminates, let us say a few words
about the logic to establish various stiffnesses.
For simplicity, we denote x3 by z; i.e. x3 = z.
Let us consider the extensional stiffnesess
Aij =

N
X

k=1

Qij

(zk zk1 ) ,

given by equation (3.3.16)1 .


Since zk zk1 > 0 for k = 1, ..., N , the above
equation shows that the only

way to have an Aij zero is either for all Qij k to be zero, or for some of the

Qij k to be a negative and some positive, so that the sum of their products with
their respective thicknesses be zero. From equation (3.1.17) giving the transformed
reduced stiffnesses Qij , follows that Q11 , Q12 , Q22 and Q66 are positive, since all

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

144

CHAPTER 3. COMPOSITE LAMINATES

trigonometrical functions are involved with even powers and Q11 , Q12 , Q22 , Q66
are positive. Thus, A11 , A12 , A22 and A66 are all positive since the thicknesses
of the laminae are obviously positive. However, (Q16 )k and (Q26 )k are zero for
lamina orientation of 00 or 900 to the laminate axes. Thus, A16 and A26 are zero
for laminates made of orthotropic laminae oriented at either 00 or 900 to the
laminate axes.
Next, we consider the coupling stiffnesses
N

Bij =


1X
2
,
(Qij )k zk2 zk1
2
k=1

given by equation (3.3.16)2 .


It is easy to see that if the cross-ply laminate is symmetric about the middle
surface, then all the Bij vanish.
Finally, we consider the bending stiffnesses
N

Dij =


1X
3
,
(Qij )k zk3 zk1
3
k=1

given by equation (3.3.16)3 .






3
Since zk3 zk1
> 0 and Q11 k , Q12 k , Q22 k , Q66 k > 0, it results that


D11 , D12 , D22 and D66 are positive. Also Q16 k and Q26 k are zero for laminae
having principal material property orientation of 00 or 900 with respect to the
laminate coordinates. Thus, D16 and D26 also vanish.
Summing up, we can say that the status of the extensional and bending
stiffnesses is the same.
As we have seen, a laminate of multiple generally orthotropic layers that are
symmetrically disposed about the middle surface exhibits no coupling between
bending and extension; that is the Bij are zero. Therefore, the force and moments
resultants are given by equation (3.4.6) There, all the Aij and Dij are required
because of forces and shearing strains, shearing force and normal strains, normal
moments and twist, and twisting moment and normal curvatures coupling between
normal forces N11 , N12 and shearing strains e12 , shearing force N12 and normal
strains e11 , e22 , normal moments M11 , M22 and twist k12 and twisting moment M12
and normal curvatures k11 , k22 . Such coupling is evidenced by the A16 , A26 , D16
and D26 stiffnesses.
A special subclass of this class of symmetric laminates is the regular symmetric angle-ply laminate. Such laminates have orthotropic laminae of equal thicknesses. The adjacent laminae have opposite signs of angle of orientation of the
principal material properties with respect to the laminates axes, for example
+/ / + as in Figure 3.15.
For symmetry, there must be an odd number of layers.
The aforementioned coupling that involves A16 , A26 , D16 and D26 takes a
special form for symmetric angle-ply laminates. Those stiffnesses can be shown to

Copyright 2004 by Chapman & Hall/CRC

145

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.4. SPECIAL CLASSES OF LAMINATES

Figure 3.15: Exploded (unbounded) view of a three-layered regular symmetric


angle-ply laminate.

be largest when N = 3 (the lowest N for which this class of laminates exists) and
decrease in proportion to 1/N as N increases
A16 =

N
X

k=1

(Q16 )k (zk zk1 ) and D16 =


1X
3
,
(Q16 )k zk3 zk1
3
k=1

obviously, A16 and D16 are sums of terms of alternating signs since
(Q16 )+ = (Q16 ) .

(3.4.14)

Consequently, for many layered symmetric angle-ply laminates, the values of A 16 ,


A26 , D16 and D26 can be quite small when compared to the other Aij and Dij ,
respectively.
3. Antisymmetric laminates
As we have seen, symmetry of a laminate about a middle surface is generally
desired to avoid coupling between bending and extension. However, many engineering applications of laminated composite require antisymmetric laminates to
achieve design requirements.
The general class of antisymmetric laminates has an even numbers of layers
if adjacent laminae have alternating signs of the principal material directions with
respect to the laminate axes. In addition, each pair of laminae must have the same
thickness.
It can be shown that in the case of an antisymmetric laminate
A16 = A26 = D16 = D26

(3.4.15)

(Q16 ) = (Q16 ) , (Q26 ) = (Q26 ) ,

(3.4.16)

since

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

146

CHAPTER 3. COMPOSITE LAMINATES

where the above symbols represent the corresponding transformed reduced stiffnesses in the laminae with the orientation and of the fibers.
For general antisymmetric laminates, all coupling stiffnesses are non-vanishing.
Hence, the constitutive equation (3.3.15) become

N11
A11 A12
0
e11
B11 B12 B16
k11
N22 = A12 A22
0 e22 + B12 B22 B26 k22 ,
N12
0
0
A66
2e12
B16 B26 B66
2k12


M11
B11
M22 = B12
M12
B16

k11
0
0 k22 .
D66
2k12
(3.4.17)
We discuss now two important subclasses of antisymmetric laminates.
An antisymmetric cross-ply laminate consists of an even number of orthotropic
laminae laid down on each other with principal material directions alternating at
00 and 900 to the laminate axes, as in the example given in Figure 3.16.
B12
B22
B26


B16
e11
D11
B26 e22 + D12
0
B66
2e12

D12
D22
0

Figure 3.16: Exploded (unbounded) view of a two-layered regular antisymmetric


cross-ply laminate.
The antisymmetric cross-ply laminates do not have non-zero A16 , A26 , D16 ,
D26 , but do have coupling between extension and bending. More precisely
A16 = A26 = 0 , D16 = D26 = 0 , B11 = B22 , B12 = B16 = B26 = B66 = 0.
(3.4.18)
Hence, the constitutive equations of the antisymmetric cross-ply laminates become



k11
e11
B11 0
0
N11
A11 A12
0
N22 = A12 A22
B11 0 k22 ,
0 e22 + 0
0
0
0
0
0
A66
2k12
2e12
N12


M11
B11
M22 = 0
M12
0

Copyright 2004 by Chapman & Hall/CRC

0
B11
0


0
e11
D11
0 e22 + D12
0
2e12
0

D12
D22
0

0
k11
0 k22 .
D66
2k12
(3.4.19)

147

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.4. SPECIAL CLASSES OF LAMINATES

A regular antisymmetric cross-ply laminate is a special case, which has laminae of equal thickness. This type of laminate is common because of simplicity of
fabrication.
It can be shown that the coupling stiffness B11 of an antisymmetric cross-ply
laminate approaches zero as the number of layers increases, for a fixed laminate
thickness.
An antisymmetric angle-ply laminate has laminae oriented at + degrees to
the laminate coordinate axes on one side of the middle surface and the corresponding equal thickness laminae on the other side is oriented at degrees. A simple
example is given in Figure 3.17.

y = x2

x = x1

Figure 3.17: Exploded (unbounded) view of a two-layered regular antisymmetric


angle-ply laminate.
A regular antisymmetric angle-ply laminate has all laminae of the same thicknesses.
It can be shown that for an antisymmetric angle-ply laminate, the following
stiffnesses are vanishing:
A16 = A26 = 0 , B11 = B12 = B22 = 0 , D16 = D26 = 0.

(3.4.20)

Consequently, the global constitutive equations become

N11
A11 A12
0
e11
0
0
B16
k11
N22 = A12 A22
0 e22 + 0
0
B26 k22 ,
0
0
A66
N12
2e12
B16 B26 0
2k12


M11
0
M22 = 0
M12
B16


B16
e11
D11
B26 e22 + D12
0
2e12
0

0
k11
0 k22 .
D66
2k12
(3.4.21)
It can be shown for a fixed laminate thickness that the coupling stiffnesses
B16 and B26 tend towards zero as the number of layers in the laminate increases.
Summing up the presented results, concerning some special classes of fiberreinforced composite laminates, we can say the following:

Copyright 2004 by Chapman & Hall/CRC

0
0
B26

D12
D22
0

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

148

CHAPTER 3. COMPOSITE LAMINATES

(1) Simple layer laminates with a reference surface, at the middle surface do
not exhibit coupling between extension and bending.
(2) Multilayered laminates, in general, develop coupling between extension
and bending.
(3) The coupling is influenced by the geometrical as well as by the material
properties of the laminae.
(4) There exist combinations of the material properties and geometrical characteristics for which there is no coupling between extension and bending.
(5) The elastic symmetry of the laminae (transversally isotropy, orthotropy)
is not necessarily conserved in the laminate.
(6) Moreover, the symmetries can be increased, decreased or remain the same.
(7) The symmetry properties of the three stiffness matrices [A], [B] and [D]
need not be the same, and are generally different.
We stress the fact that the fundamental concept of coupling between extension and bending must be well understood, because there exists many applications
of the composite laminates where the neglect of the coupling can be catastrophic.
Considering coupling is the key to the correct analysis of eccentrically stiffened
plates.
The procedure to describe a laminate by use of individual layer thicknesses,
principal material property orientations, and overall sequence can be quite involved. However, all pertinent parameters can be represented in a simple way if
one uses the following stacking sequence terminology.
For regular (equal thickness layers) laminates, a listing of layers and their orientation suffices, for example, [00 /900 /450 ]. Let us observe that only the principal
material direction orientations need to be given.
For irregular (when the layers do not have the same thickness) laminates, a
notation of layers thickness must be added to the previous notation, for example,
[00 /h1 /900 /h2 /450 /h3 ].
For symmetric laminates, the simplest representation of the laminate [0 0 /900 /
0
45 /450 /900 /00 ] is [00 /900 /450 ] sym.
We shall discuss now in greater details the relation by which the cross-ply
laminate stiffnesses can be expressed.
We recall that a cross-ply laminate has N unidirectionally reinforced orthotropic layers with the principal material directions alternatingly oriented at
00 and 900 with respect to the laminate coordinate axes. The fiber direction of
the odd -numbered layers is the x1 direction of the laminate. The fiber direction
of the even-numbered layers is the x2 direction of the laminate. We assume that
all odd-numbered layers have the same thickness, all even-numbered layers have
also equal thicknesses but the odd and even numbered layers do not necessarily
have the same thickness. For this special but important case for applications, two
geometrical parameters are important:
(1) N, the total number of layers,
and

Copyright 2004 by Chapman & Hall/CRC

149

3.4. SPECIAL CLASSES OF LAMINATES

(2) m, the ratio of the total thicknesses of odd numbered layers to the total
thickness of the even-numbered layers, called cross-ply ratio. Hence,

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

m=

hk

k=odd

hk

(3.4.22)

k=even

For instance, for a five-layered cross-ply laminate, which has a lamination or stacking sequence [00 /h1 /900 /2h1 /00 /2h1 /900 /2h1 /00 /h1 ], we get
m=

h1 + h 1 + h 1
= 3/4.
2h1 + 2h1

Let us observe that the cross-ply ratio m has a specific meaning only when
the laminae (layers) have alternating 00 and 900 orientation!
Tsai (see Jones [3.2], Chapter 4) has shown that the laminate stiffnesses
Aij , Bij and Dij for cross-ply laminates with odd or even numbers of layers, can
be expressed in terms of m and N . In addition, Tsai uses the stiffness ratio
f = Q22 /Q11 = E1 /E2 6 1

(3.4.23)

of principal lamina stiffnesses.


Tsai has obtained the following expressions for the stiffnesses:
Symmetric cross-ply laminates (N odd ).

1 + mf
1 + mf
m+f
A11 ,
hQ11 =
hQ11 , A12 = hQ12 , A22 =
m+f
1+m
1+m
A16 = A26 = 0 , A66 = hQ66 , Bij = 0,
h3
h2
1+m
(f 1)p + 1 3
Q12 ,
{(f 1)p + 1} A11 , D12 =
h Q11 =
D11 =
12
12
m+f
12
h2
1+m
(1 f )p + f 3
{(1 f )p + f } A11 ,
h Q11 =
D22 =
12
m+f
12
h3
Q66 ,
(3.4.24)
D16 = D26 = 0 , D66 =
12

A11 =

where
p=

m(N 3){m(N 1) + 2(N + 1)]


1
.
+
(N 2 1)(1 + m)3
(1 + m)3

Copyright 2004 by Chapman & Hall/CRC

150

CHAPTER 3. COMPOSITE LAMINATES


Antisymmetric cross-ply laminates (N even)

1 + mf
1 + mf
m+f
A11 ,
hQ11 =
hQ11 , A12 = hQ12 , A22 =
m+f
1+m
1+m
A16 = A26 = 0, A66 = hQ66 ,
m(f 1)
m(f 1) 2
hA11 , B22 = B11 ,
h Q11 =
B11 =
N (1 + m)(f + m)
N (1 + m)2
B12 = B16 = B26 = B66 = 0,
h3
h2
1+m
(f 1)r + 1 3
Q12 ,
{(f 1)r + 1} A11 , D12 =
h Q11 =
D11 =
12
12
m+f
12
h2
1+m
(1 f )r + f 3
{(1 f )r + f } A11 ,
h Q11 =
D22 =
12
m+f
12
3
h
Q66 ,
(3.4.25)
D16 = D26 = 0 , D66 =
12

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

A11 =

where
r=

8m(m 1)
1
.
+
1 + m N 2 (1 + m)3

Concerning the above formulas, we can make the following observations:


(1) For both symmetric and antisymmetric cross-ply laminates, the extensional stiffnesses Aij are independent of N , the number of layers.
(2) However, the stiffness ratios A11 and A22 depend on the cross-ply ratio
m, as well as on the stiffness ratio f .
(3) The stiffnesses A12 and A66 are independent on m and f .
(4) The remaining stiffnesses A16 and A26 are zero for all cross-ply laminates.
(5) All coupling stiffnesses Bij are zero for the symmetric cross-ply laminates.
(6) For the antisymmetric cross-ply laminates only, the bending stiffnesses
B11 and B22 are not vanishing.
(7) The values of B11 and B22 decrease as N , number of laminae, increases,
the thickness h of the laminate being fixed. Since N must be even to get any coupling, N = 2 corresponds to the largest coupling between extension and bending.
(8) The bending stiffnesses D11 and D22 are involved functions on the numbers of layers, N , the cross-ply ratio, m, and the stiffness ratio, f .
(9) The value of D11 approaches A11 h2 /12 and D22 approaches A22 h2 /12 as
N gets large or as m gets large, or as f approaches 1.
Regular antisymmetric cross-ply laminates (N even) In this case, all layers
have the same thickness h/N , hence, the cross-ply ratio m is one; i.e. m = 1.

Copyright 2004 by Chapman & Hall/CRC

151

3.4. SPECIAL CLASSES OF LAMINATES

Consequently, Tsais formulas (3.4.25) are considerably simplified and we get:


1+f
hQ11 , A12 = hQ12 , A16 = A26 = 0 , A66 = hQ66 ,
2
1 f 1
f 1 2
hA11 , B12 = B16 = B26 = B66 = 0,
h Q11 =
B11 = B22 =
2N f + 1
4N
h3
h2
f +1 3
Q12 ,
A11 , D12 =
h Q11 =
D11 = D22 =
12
12
24
h3
Q66 .
(3.4.26)
D16 = D26 = 0 , D66 =
12

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

A11 = A22 =

Since E2 6 E1 , we have f 6 1; thus,


B11 < 0

and

B22 > 0.

(3.4.27)

In the following, we shall indicate the way in which Tsais formulas (3.4.26)
for a regular antisymmetric cross-ply laminate can be obtained. Such a laminate
is shown in the Figure 3.18.

Figure 3.18: Regular cross-ply laminate having N (even) layers.


Since all laminae have the same thickness h/N , we have

h
h
for k = 0, 1, ..., N.
(3.4.28)
+k
N
2
According to the definition of a regular antisymmetric cross-ply laminate, we have
also
k = 00 if k is odd and k = 900 if k is even .
(3.4.29)
zk =

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

152

CHAPTER 3. COMPOSITE LAMINATES

Conceptually, from the general equations (3.1.17), giving the transformed


reduced stiffnesses Qij , we get
(Q11 )k = Q11 if k is odd and (Q11 )k = Q22 if k is even,
(Q12 )k = Q12 for any k,
(Q22 )k = Q22 if k is odd and (Q22 )k = Q11 if k is even,
(Q66 )k = Q66 for any k,
(Q16 )k = (Q26 )k = 0 for any k.
In order to prove Tsais relations (3.4.26), the above properties must be used.
We begin with A11 , which, according to (3.3.16)1 , is given by the equation
(x3 = z!)
N
X
A11 =
(Q11 )k (zk zk1 ).
k=1

Since all layers have the same thickness h/N , we obviously have
zk zk1 = h/N for any k.

(3.4.30)

Hence,
N
h X
(Q11 )k .
N

A11 =

k=1

According to the above results, the reduced transformed thickness (Q11 )k


takes N/2-times the value Q11 and N/2-times the value Q22 . Hence,

A11 =

h
(Q11 + Q22 ).
2

Using the stiffness ratio f = Q22 /Q11 , we get:


A11 =

1+f
hQ11 .
2

Hence, we get the first Tsai formula (3.4.26), giving the extensional stiffness A 11 .
In the same way all Tsai formulas can be deduced concerning the extensional
stiffnesses Aij .
Let us analyze now the coupling stiffnesses. According to (3.3.16)2 , we have
(x3 = z)
N
1X
2
).
(Qij )k (zk2 zk1
Bij =
2
k=1

At the same time from (3.4.28) and (3.4.30), we get


2
zk2 zk1
=

Copyright 2004 by Chapman & Hall/CRC

h2
{(1 + N ) + 2k}.
N2

153

3.4. SPECIAL CLASSES OF LAMINATES


Hence, the formula giving B11 becomes

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

B11 =

k=1

k=1

X
X
h2
k(Q11 )k }.
(Q11 )k + 2
{(1 + N )
2N

We recall again that (Q11 )k = Q11 if k is odd, and (Q11 )k = Q22 if k is even.
Thus we obtain

B11 =

X
X
N (1 + N )
h2
(Q11 + Q22 ) + 2(Q11
k + Q22
k)}.
{
2
2N
k=odd

k=even

Now let us use the well known relation


n
X

l=

l=0

n(n + 1)
.
2

(3.4.31)

Thus, after some elementary computations, we get


X

k=

N (N + 2)
N2 X
.
,
k=
4
4

(3.4.32)

k=even

k=odd

Introducing these values in the last expression of B11 , after some algebra, it results
B11 =

h2
(Q11 + Q22 ).
4N

But Q22 = f Q11 , and the above equation becomes


B11 =

f 1 2
h Q11 .
4N

Therefore, we have obtained the first Tsai formula (3.4.26), giving the coupling
stiffness B11 .
In the same way, all Tsai formulas can be deduced (3.4.26) giving the coupling
stiffness Bij.
According to (3.3.16)3 , for the bending stiffness D11 , we have (x3 = z)
N

D11 =

1X
3
).
(Q11 )k (zk3 zk1
3
k=1

Furthermore, from (3.4.28) and (3.4.30), it results


3
=
zk3 zk1

Copyright 2004 by Chapman & Hall/CRC

h3
N3


3N 2 + 6N + 4
3(N + 1)k + 3k 2 .
4

154

CHAPTER 3. COMPOSITE LAMINATES

Also, we know that (Q11 )k takes N/2 times the value Q11 , if k is odd, and
N/2-times the value Q22 , if k is even. Hence, the equation giving D11 becomes

D11 =

X
X
h3 3N 2 + 6N + 4 N
(Q
+
Q
)

3(N
+
1)(Q
k
+
Q
k)
{
11
22
11
22
2
N
3N 3

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

k=odd

+3(Q11

k 2 + Q22

k=odd

k=even

k 2 )}.

k=even

We use again the relation (3.4.31) and the well known formula
n
X

l2 =

l=o

n(n + 1)(2n + 1)
.
6

(3.4.33)

Thus, after some algebra, from (3.4.31) and (3.4.33), we get


X

k=odd

k2 =

N (N + 1)(N + 2)
N (N 2 1) X 2
.
,
k =
6
6

(3.4.34)

k=even

Introducing (3.4.32) and (3.4.34) in the last expression of D11 , elementary computations leads to
D11 =

1+f 3
h2
h Q11 ,
(Q11 + Q22 ) =
24
24

since Q22 = f Q11 , according to the definition of the rigidity ratio f.


In this way we have derived the Tsai formula (3.4.26) giving the bending
stiffness D11 .
In the same way can be deduced all Tsai formulas concerning the bending
stiffnesses Dij .

3.5

Equilibrium equations and boundary


conditions

In the theory of thin plate, the Cauchys equilibrium equations are replaced
by global equilibrium conditions satisfied by the force resultants N , moment
resultants M and by the resultant shear forces Q defined by the relations
Q =

h/2

3 dx3 , = 1, 2.

(3.5.1)

h/2

To get these equations, we start with the Cauchys (local) equilibrium conditions, assuming absence of body forces
, + 3,3 = 0 , 3, + 33,3 = 0, , = 1, 2.

Copyright 2004 by Chapman & Hall/CRC

(3.5.2)

3.5. EQUILIBRIUM EQUATIONS AND BOUNDARY CONDITIONS

155

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

We integrate these equations with respect to x3 from h/2 to h/2. At the


same time, we take into account that integration with respect to x3 , and differentiation with respect to x1 and x2 are invertible. Thus, from (3.5.2), and using
equations (3.3.13)1 and (3.5.1), which are defining N and Q , respectively, we
get
N + 3 (h/2) 3 (h/2) = 0 , Q, + 33 (h/2) 33 (h/2) = 0 , , = 1, 2.
We recall now the assumption (8) made at the beginning of Section 3.3. According
to the made assumption, 3 are vanishing on the faces x3 = h/2 of the laminate.
Hence,
3 (h/2) = 0.
Also, we interpret the function
q(x1 , x2 ) = 33 (x1 , x2 , h/2) 33 (x1 , x2 , h/2)
as representing a given external surface force distribution, acting normal to the
face x3 = h/2 of the laminate.
Taking into account the above observations, we obtain the first three (global)
equilibrium equation that must be satisfied by the resultants in-plane forces N
and by the resultant shear forces Q , in the plane domain D occupied by the
middle surface of the laminate
N, = 0 , Q, + q = 0 in D, , = 1, 2.

(3.5.3)

In order to obtain the (global) equilibrium equations satisfied by the moment


resultants M , we multiply the first two Cauchy equations (3.5.2) by x3 and
integrate the obtained results with respect to x3 from h/2 to h/2. In this way,
using equation (3.3.13)2 defining M , and the invertibility of differentiation with
respect to x1 ,x2 and of integration with respect to x3 , we get
M, +

h/2

x3 3,3 dx3 = 0.
h/2

We have
x3 3,3 = (x3 3,3 ) 3 .
Consequently,
Z

h/2

x3 3,3 dx3 =
h/2

h
h
3 (h/2) 3 (h/2)
2
2

h/2

3 dx3 .
h/2

We recall that 3 (h/2) = 0 and we will use the definition (3.5.1) of the resultant
shear forces Q .

Copyright 2004 by Chapman & Hall/CRC

156

CHAPTER 3. COMPOSITE LAMINATES

Thus, we finally get the equilibrium condition which must be satisfied by the
shear forces resultants Q and moment resultants M , in the plane domain D
occupied by the middle surface of the laminate

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

M, Q = 0 in D, , = 1, 2.

(3.5.4)

From (3.5.3)2 and (3.5.4), we can now see that the resultant moments M
must satisfy the following second order (global) equilibrium equation
M, + q = 0 in D , = 1, 2.

(3.5.5)

It remains to formulate the boundary conditions which can be prescribed and


must be satisfied by the displacement U , U3 , by the force resultants N and by
the moments resultants M on the boundary D of the domain D occupied by
the middle surface of the laminate in the x1 , x2 plane.
In order to obtain the possible boundary condition (assuming uniqueness of
the solution), we recall the geometrical equations (3.3.4) and (3.3.5) expressing the
plane strains e and the curvatures k in terms of the displacements U , U3 , the
constitutive relations in tensorial form (3.3.21) expressing the force and moments
resultants N , M in terms of e , k , and the just established equilibrium
conditions (3.5.3), (3.5.4) and (3.5.5). We have
Geometrical equations:
e =

1
(U, + U, ), k = U3, ;
2

(3.5.6)

Constitutive equations:
N = A e + B k , M = B e + D k ;

(3.5.7)

Equilibrium equations:
N, = 0 , Q, + q = 0 , M, Q = 0 , M, + q = 0.

(3.5.8)

We recall also that according to (3.3.23), the constitutive coefficients have the
following symmetry properties :
A = A = A = A ,
B = B = B = B ,

(3.5.9)

D = D = D = D .
Obviously the Greek indices take the values 1, 2 and the Einsteins summation
convention applies.

Copyright 2004 by Chapman & Hall/CRC

3.5. EQUILIBRIUM EQUATIONS AND BOUNDARY CONDITIONS

157

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

In the following we shall prove a theorem of work and energy, appropriate


to the Love-Kirchhoff type of plate theory developed above.
As before, we denote by D the plane domain occupied by the middle surface
of the plate, D will be the boundary curve of D. Let n be the unit outward
normal to D, and let be the unit tangent vector to D, as shown in Figure
3.19.

Figure 3.19: The plane domain occupied by the middle surface of the laminate.
Denoting by n and the (plane) components of n and , we get
n1 = 2 , n2 = 1 .

(3.5.10)

We introduce now the two-dimensional vectors N n and M n defined on D,


by the following equations, giving their components Nn and Mn :
Nn = N n , Mn = M n .

(3.5.11)

Obviously, Nn and Mn are corresponding to the Cauchys stress vector, and


represent the resultant force vector and the resultant moment vector, respectively,
acting on the boundary D of the laminate middle surface.
The normal and tangential components Nnn , Nn and Mnn , Mn of these
vectors have the following expression:
Nnn = n Nn , Nn = Nn , Mnn = n Mn , Mn = Mn .

(3.5.12)

In the same way, the normal and tangential components Un and U of the
in-plane displacement on the boundary D can be obtained using the equations
Un = n U , U = U .

(3.5.13)

Also, for a later use, we introduce on the boundary D the normal and
tangential derivatives U3,n and U3, of the normal displacement U3 . According to

Copyright 2004 by Chapman & Hall/CRC

158

CHAPTER 3. COMPOSITE LAMINATES

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

the general definition of the directional derivative of a scalar field, given in the
Section 1.2, we have
U3,n = n U3, , U3, = U3, .
(3.5.14)
Using these equations, we can express the partial derivatives U3, through
the normal and tangential derivatives U3,n , U3, . From (3.5.10) and (3.5.14), we
get
U3,1 = 2 U3,n + 1 U3, , U3,2 = 1 U3,n + 2 U3, .
(3.5.15)

These equations will be used a little later on.


Multiplying the equilibrium equation (3.5.8)1 by U , integrating the obtained
result on D, using Greens theorem, the symmetry of N , the geometrical equations (3.5.6)1 and the Cauchys type relations (3.5.11)1 , we obtain
Z
Z
N e da =
U Nn ds.
(3.5.16)
D

The two-dimensional scalar product U Nn of the two-dimensional vectors (U1 , U2 )


and (Nn1 , Nn2 ) can be expressed in an equivalent form using the normal and tangential components of the involved vectors; i.e.
U Nn = Un Nnn + U Nn .
Thus, equation (3.5.16) becomes
Z
Z
N e da =
D

(Un Nnn + U Nn )ds.

(3.5.17)

(3.5.18)

We multiply now the second equilibrium equation (3.5.8)2 by U3 and integrate


the obtained result on D. In this way, Greens theorem leads to the following
relation:
Z
Z
Z
Q U3, da.
(3.5.19)
U3 Q n ds =
qU3 da +
D

In order to transform the right hand side term of this equation, we now use
the equilibrium condition (3.5.8)3 . We get
Z
Z
Q U3, da =
U3, M, da.
(3.5.20)
D

Using again the Greens theorem, the geometrical equations (3.5.6)2 and the
Cauchys type relations (3.5.11), from (3.5.20) we obtain
Z
Z
Z
Q U3, da =
U3, Mn ds +
M k da.
(3.5.21)
D

Consequently, from (3.5.19) and (3.5.21), it results


Z
Z
Z
Z
qU3 da +
U3 Qn ds =
U3, Mn ds +
M k da,
D

Copyright 2004 by Chapman & Hall/CRC

(3.5.22)

3.5. EQUILIBRIUM EQUATIONS AND BOUNDARY CONDITIONS

159

where

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

Qn = Q n

(3.5.23)

is the normal component of the resultant shear forces (Q1 , Q2 ).


In order to obtain the final results, we still have to express in an adequate
form the first term from the right hand side of equation (3.5.22). To do this, we
use equation (3.5.15). In this way we successively get
U3, Mn

=
=
=

U3,1 Mn1 + U3,2 Mn2


(2 U3,n + 1 U3, )Mn1 + (1 U3,n + 2 U3, )Mn2
(1 Mn1 + 2 Mn2 )U3, + (2 Mn1 + 1 Mn2 )U3,n .

We use now equations (3.5.10) and (3.5.12)3,4 to get


U3, Mn

(1 Mn1 + 2 Mn2 )U3, + (n1 Mn1 + n2 Mn2 )U3,n

Mn U3, + Mnn U3,n .

(3.5.24)

Transforming adequately the first term in the right hand side of (3.5.24), we obtain
U3, Mn = (U3 Mn ), U3 Mn, + U3,n Mnn .
The tangential derivatives involved in this equation have the following expressions:
(U3 Mn ), = (U3 Mn ),1 1 + (U3 Mn ),2 2 ,
Mn, = Mn,1 1 + Mn,2 2 .
We assume that U3 and Mn are uniform function on D. Hence,
Z
(U3 Mn ), ds = 0.

(3.5.25)

since D is a closed curve. Thus from (3.5.24) and (3.5.25) we finally get
Z
Z
U3, Mn ds =
(U3 Mn, + U3,n Mnn )ds.
D

Using this equation in (3.5.22), we obtain


Z
Z
M k da.
{U3 (Qn + Mn, ) U3,n Mnn }ds =
qU3 da +

(3.5.26)

Now, equations (3.5.18) and (3.5.26) lead to the following work relation:
Z
Z
{Nnn Un + Nn U + (Qn + Mn, )U3 U3,n Mnn }ds +
qU3 da
D
D
Z
= 2
wda,
(3.5.27)
D

Copyright 2004 by Chapman & Hall/CRC

160

CHAPTER 3. COMPOSITE LAMINATES

where the quadratic form w is defined by the following equation:

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

2w = N e + M k .

(3.5.28)

The work relation (3.5.27) is a direct consequence of the geometrical equation


(3.5.6) and equilibrium condition (3.5.8).
Using the constitutive relations (3.5.7) and the symmetry properties (3.5.9),
w can be expressed in the following form:
w=

1
1
e A e + e B k + k D k .
2
2

(3.5.29)

As in the usual elasticity theory, the left-hand side of the work relation (3.5.27) represents the total work of the external force acting on the laminate. Consequently,
w is the specific elastic energy stored in the deformed laminate. Hence, the total
elastic energy W stored in the laminate is
Z
W =
wda.
(3.5.30)
D

The above interpretation and equation (3.5.27) represents the content of the
announced and proved theorem of work and energy.
This theorem tells us what kind of boundary conditions can be given
on the boundary D, in order to assume the uniqueness of the solution of various
boundary value problems. The structure of the left-hand side of the work relation
(3.5.27) shows that the following fields can be prescribed on the boundary D:
Un or Nnn , U or Nn , U3 or Qn + Mn, and U3,n or Mnn .

(3.5.31)

The same work relations show that the following theorem takes place.
Uniqueness theorem. If the specific elastic energy w is a positive definite
quadratic form, the various boundary value problems can have no more than one
regular solution, modulo a rigid displacement.

3.6

Variational and extreme principles

We observe that due to the symmetry relations (3.5.9), the differential system
(3.5.6)(3.5.8) is self-adjoint. Due to this fact we can establish various variational
and extreme principles, corresponding to various boundary value problem.
In order to do this, we introduce first the three-dimensional vectors U and
U having the components (U , U3 ) and (U , U3 ), respectively. Also we consider
the energy functional
Z
W (U) =
w(U)da
(3.6.1)
D

where w is the quadratic form given by the equation (3.5.28) and (3.5.29).

Copyright 2004 by Chapman & Hall/CRC

161

3.6. VARIATIONAL AND EXTREME PRINCIPLES

For brevity we denote simply by W the first variation of W in U, in the


direction U. According to the usual definition, we have

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

W =

d
W (U + U)|=0 , being a real variable.
d

(3.6.2)

In order to evaluate the above variation, we take into account the geometric
relation (3.5.6), the constitutive equations (3.5.7), the symmetry properties (3.5.8)
and the expression (3.5.29) of the specific elastic (strain) energy w. Thus, after
some elementary computations, we get
Z
W =
(N U, M U3, )da.
(3.6.3)
D

Using Greens formula, we obtain


Z
W = (N, U + M, U3 )da
D

(U N n + U3 M, n U3, M n )ds.

(3.6.4)

To obtain this equation, we have used the following relations:


M U3, = (M U3, ), M, U3,
= (M U3, ), (M, U3 ), + M, U3 .
Using the Cauchys type relation (3.5.11)1 , we get
U N n = U Nn = Un Nnn + U Nn ,

(3.6.5)

Un and U being the normal and tangential components of the in-plane displacement field (U1 , U2 ).
Similarly, using the equilibrium condition (3.5.8)4 and the equation (3.5.23),
we obtain
U3 M, n = U3 Q n = U3 Qn ,
(3.6.6)
Qn = Q n being the normal component of the resultant shear force (Q1 , Q2 ).
Using a relation analogous to (3.5.27), we get
U3, M n = U3, Mn = (U3 Mn ), U3 Mn, + U3,n Mnn .

(3.6.7)

Introducing (3.6.5), (3.6.6), (3.6.7) in the line integral of equation (3.6.4), and
observing that
Z
(U3 Mn ), ds = 0,
D

Copyright 2004 by Chapman & Hall/CRC

162

CHAPTER 3. COMPOSITE LAMINATES

we obtain the expression of the variation W


Z
W = (N, U + M, U3 )da

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

{Un Nnn + U Nn + U3 (Mn, + Qn ) U3,n Mnn } ds.

(3.6.8)

The last equation can be used to obtain variational principles, corresponding to


different boundary value problems.We shall illustrate the procedure by analyzing
two possibilities.
In order to do this we shall interpret the third equilibrium condition (3.5.8) 3 ,
that is the relations
Q = M,
(3.6.9)
as representing supplementary constitutive equations, expressing the resultant shear
forces Q in terms of the in-plane deformations e and of the curvatures k .The
status of the relations (3.6.5) is similar to that of the constitutive equations (3.5.7)
expressing the resultant in-plane forces N and the resultant bending moments
M in terms of the above mentioned kinematical fields.
Also, we suppose that the involved displacement fields U and their variations
U are of class C 2 on D and of class C 1 on D = D D, since we take into account
only regular solutions.
First let us assume that on the boundary D are given Nnn , Nn , Mn, +Qn
and Mnn ; i.e.

Nnn = , Nn = , Mn, + Qn = , Mnn = on D,

(3.6.10)

where , , , are given continuous functions on D. We suppose also that q is


a given continuous function on D.
In this case, we introduce the functional I(U) defined by the equation
Z
Z
qda.
(3.6.11)
(Un + U + U3 U3,n )ds
I(U) = W (U)
D

Taking into account equation (3.6.8), for the variation I of I in U, in the direction
U, we get
Z
I =
{N, U + (M, + q)U3 } da
D
Z
+
{(Nnn )Un + (Nn )U
D

+(Mn, + Qn )U3 (Mnn )U3,n }da.

(3.6.12)

Using the above equation, we can prove the following.


First variational principle. If U is a regular solution of the boundary value
problem (3.6.10), the first variation I of I is vanishing in U for any direction

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.6. VARIATIONAL AND EXTREME PRINCIPLES

163

U. Conversely, if I is vanishing in U for any direction U, then U is a regular


solution of the boundary value problem (3.6.10).
In order to prove this theorem, let us assume that U is a regular solution of
the considered boundary value problem. Hence, the equilibrium equations (3.5.8) 1 ,
(3.5.8)4 and the boundary conditions (3.6.10) are satisfied. Hence, according to
(3.6.12), I = 0 in U for any U.
In order to prove the inverse implication, we assume that I = 0 in U for
any U. In this case the relation (3.6.12) shows that the equilibrium equations
(3.5.8)2 and (3.5.8)4 , as well as the boundary conditions (3.6.10) are satisfied.
From the assumed supplementary constitutive relations (3.6.5), it follows that the
remaining equilibrium conditions (3.5.8)2 and (3.5.8)3 are also verified and the
proof is complete.
Let us observe that in the above variational principle, the variations U are
not submitted to any restrictions on the boundary D. In this sense we consider the
boundary conditions (3.6.10) as being natural boundary conditions. The situation is
entirely analogous to that encountered in the usual elasticity theory if the traction
is prescribed on the boundary of the body.
Let us suppose now that, on the boundary D, are given Un , U , U3 and
U3,n ;
Un = , U = , U3 = and U3,n = on D,
(3.6.13)
where , , , are continuous functions given on D. We suppose also that q is
a continuous function given on D.
In this case we shall introduce the functional J(U) defined by the equation
Z
J(U) = W (U)
qU3 da.
(3.6.14)
D

We shall calculate the variation J of J in U, in a direction U that satisfies


homogeneous boundary conditions; i.e.
Un = U = U3 = U3,n = 0 on D.
Using again (3.6.8) and taking into account (3.6.15), we get
Z
{N, U + (M, + q)U3 } da.
J =

(3.6.15)

(3.6.16)

Based on the last equation, we can prove the following.


Second variational principle. If U is a regular solution of the boundary
value problem (3.6.13), then the variation J of J in U is vanishing for any direction (variation) U satisfying the homogeneous boundary conditions (3.6.16).
Conversely, if U is a solution of the given boundary conditions (3.6.13) and if the
variation J of J is vanishing in U for any variation (direction) U, which satisfies the homogeneous boundary conditions (3.6.15), then U is a regular solution
of the boundary value problem (3.6.13).

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

164

CHAPTER 3. COMPOSITE LAMINATES

The proof of the second variational principle is similar to that given for the
first one, and is entirely based on equation (3.6.16).
Let us observe that in the second variational principle, U and U are not
arbitrary on the boundary D; U must satisfy the given boundary conditions and
U must satisfy homogeneous boundary conditions. In this sense, we consider the
boundary conditions (3.6.13) as being essential boundary conditions. The situation is entirely analogous to that existing in the usual elasticity theory where the
displacement is prescribed on the boundary of the body.
The above boundary conditions are not typical either in the classical plate
theory or in the composite laminate theory. In these domains, the most frequently
encountered boundary value problems have a mixed character. That is, on boundary D are given some data concerning the displacements and complementary information concerning the force and moment resultants. Jones shows ([3.2], Chapter
5) that the most frequently used boundary conditions in laminate theory are formulated as a choice between simply supported, clamped or free edges. The situation
for laminate plates (laminates) is complex because there are actually four types
of boundary conditions that can be called simply supported edges and four types
of boundary conditions that can be called clamped edges. According to Jones,
the eight types of simply supported (prefix S) and clamped (prefix C) boundary
conditions are commonly classified as
S1 : U3 = 0 , Mnn = 0 , Un = , U = ,
S2 : U3 = 0 , Mnn = 0 , Nnn = , U = ,
S3 : U3 = 0 , Mnn = 0 , Un = , Nn = ,
S4 : U3 = 0 , Mnn = 0 , Nnn = , Nn = ,

(3.6.17)

and
C1 : U3 = 0 , U3,n = 0 , Un = , U = ,
C2 : U3 = 0 , U3,n = 0 , Nnn = , U = ,
C3 : U3 = 0 , U3,n = 0 , Un = , Nn = ,
C4 : U3 = 0 , U3,n = 0 , Nnn = , Nn = , on D.

(3.6.18)

In these relations, , , and are given functions on the boundary curve D.


The functionals and variational principles corresponding to these eight boundary conditions can be obtained as before, examining the structure of equation
(3.6.8) together with the structure of the boundary conditions taken into account.
Let us observe that the variational principle corresponding to the boundary condition C1 represents a particular case of the second variational principle
discussed above.
In the following we shall prove a relation which can be used to obtain minimum principles. Let us consider two displacement fields U and U0 , and as denoted
0
by e , k and e0 , k
, let the in-plane strains and the curvatures correspond

Copyright 2004 by Chapman & Hall/CRC

3.6. VARIATIONAL AND EXTREME PRINCIPLES

165

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

to U and U0 , respectively. Let us evaluate the functional W corresponding to


U + U0 . According to equations (3.5.29) and (3.5.30), we have
Z
1
0
{ (e + e0 )A (e + e0 )+
W (U + U ) =
D 2

1
0
0
0
)}da.
)D (k + k
) + (k + k
+(e + e0 )B (k + k
2
Taking into account the symmetry relations (3.5.9) and the constitutive equations
(3.5.7), we get
Z
0
M )da,
W (U + U0 ) = W (U) + W (U0 ) +
(e0 N + k
D

where N and M correspond to U. Now we recall the geometric relations (3.5.6)


and use Greens theorem to transform the surface integral into a line integral.
Taking into account supplementary constitutive equations (3.6.5) and using the
relations of type (3.5.24) on (3.5.25), we obtain
Z
W (U + U0 ) = W (U) + W (U0 )
(U0 N, + U30 M, )da
D
Z
0
Mnn }ds.
+
{Un0 Nnn + U0 Nn + U30 (Qn + Mn, ) U3,n
D

(3.6.19)

In order to obtain minimum principles, this equation plays the same role as
that played by equation (3.6.8) used to obtain variational principles.
Let us consider first the boundary value problem (3.6.10) and the functional
I(U) defined by equation (3.6.11). As in the usual elasticity theory, we call I(U)
the potential energy of the laminate, corresponding to the traction boundary
value problem (3.6.10). We shall denote by A the set of all regular displacement
fields U.
Using equation (3.6.19), we can formulate and prove a principle of minimum
potential energy appropriate to the composite laminate theory and to the boundary
value problem (3.6.10).
In order to prove the minimum principles, we assume that the specific elastic
(strain) energy w of the laminate, defined by equation (3.5.30), is a positive definite
quadratic form. We have:
The first principle of minimum potential energy. Let A denote the set
of all regular displacement field U and let I(U) be the functional defined on A by
equation (3.6.11). Let U be the solution of the traction boundary value problem
(3.6.10). Then
e
I(U) I(U)
(3.6.20)
e A, and equality holds only if U
e = U modulo a rigid displacement.
for every U

Copyright 2004 by Chapman & Hall/CRC

166

CHAPTER 3. COMPOSITE LAMINATES

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

e U A and define U0 =U
e U. Using the definition (3.6.11) of the
Let U,
functional I(U), the equation (3.6.19), the equilibrium equation (3.5.8) 1,4 and the
fact that U is a solution of the traction boundary value problem (3.6.10), after
some simple computations, we get
e = I(U) + W (U0 ).
I(U)

(3.6.21)

e
J(U) J(U)

(3.6.22)

Since w is positive definite, W (U0 ) 0, hence (3.6.20) takes place. If I(U) =


e
I(U), then W (U0 ) = 0. Again, since w is positive definite, we get W (U0 ) = 0.
Thus, using again the positive definiteness of w, according to equation (3.5.29),
0
we must have e0 = 0 and k
= 0 in D. Accordingly, U0 is a rigid displacement
field of the laminate and the demonstration is complete.
We observe again that the admissible displacement fields for which the functional I(U) was defined do satisfy any restriction on the boundary line D. This
is so, since the traction boundary conditions (3.6.10) are natural boundary conditions.
Next, we consider the displacement boundary value problem (3.6.13) and
the functional J(U) defined by equation (3.6.14). We call J(U) the potential energy of the laminate corresponding to the displacement boundary value problem
(3.6.13).
We shall denote by B the set of all regular displacement fields U satisfying
the given displacement boundary conditions (3.6.13). We have:
The second principle of minimum potential energy. Let B denote the
set of all regular displacement field U that satisfy the boundary conditions (3.6.13),
and let J(U) be the functional defined on B by equation (3.6.14). Let U be the
solution of the displacement boundary value problem (3.6.13). Then

e B,and the equality holds only if U


e = U on D.
for every U

e U B and let us define U0 =U


e U. Then U0 satisfies homogeneous
Let U,
0
boundary conditions on D; i.e. Un = U0 = U03 = U03,n = 0 on D. Using
the definition (3.6.14) of the functional J, equation (3.6.19), the facts that U is
a solution of the boundary value problem (3.6.13) and U0 satisfies homogeneous
boundary conditions, we obtain
e = J(U) + W (U0 ).
J(U)

(3.6.23)

Since w is positive definite, from (3.6.23) we can conclude that (3.6.22) is true.
0
e
If J(U) = J(U), we obtain e0 = k
= 0 on D, hence U0 is a rigid displacement
0
field. Moreover, U must satisfy the homogeneous boundary conditions on D.
Hence U0 = 0 on D and the proof is complete.
We note again that now the admissible displacement fields for which the
functional J is defined must satisfy the given displacement boundary conditions

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.6. VARIATIONAL AND EXTREME PRINCIPLES

167

on the boundary line D. This is due to the fact that the displacement boundary
conditions (3.6.13) are essential boundary conditions.
We recall that the most frequent boundary conditions in the laminate theory
correspond to simple supported or clamped edges and are of mixed type. Using
equation (3.6.19), appropriate potential energies can be found, and can be formulated and proved appropriate minimum principles, corresponding to various
boundary conditions listed in the relations (3.6.17) and (3.6.18).
As in the usual elasticity theory, we can prove the conversers of the given
minimum principles.
Converse of the first principle of minimum potential energy. Let
U A and suppose that
e
I(U) I(U)
(3.6.24)

e A. Then U is a regular solution of the traction boundary value


for every U
problem (3.6.10).
e = U0 +U A.
Let U0 be an arbitrary vector field of class C on D. Then U
Using the equation (3.6.19) and the assumption (3.6.29), it is easy to see that
Z
{U0 N, + U30 (M, + q)}da
W (U0 )
D
Z
0
{Un (Nnn ) + U0 (Nn ) + U30 (Qn
+
D

0
+Mn, ) U3,n
(Mnn )}ds 0.

Obviously, this inequality must hold when U0 is replaced by U0 , being


an arbitrary real number. Hence we must have
Z
{U0 N, + U30 (M, + q)}da
ZD
+
{Un0 (Nnn ) + U0 (Nn ) + U30 (Qn
D

0
(Mnn )}ds = 0,
+Mn, ) U3,n

since W (U0 ) = 0.
Since U0 is an arbitrary field, from the above equation and from the supplementary constitutive relation (3.6.5), it follows that the equilibrium equations
(3.5.8) and the boundary conditions (3.6.10) are satisfied. Hence, U is a regular
solution of the traction boundary value problem (3.6.10). Analogously, we have
the following.
Converse of the second principle of minimum potential energy. Let
U B. Suppose that
e
J(U) J(U)
(3.6.25)
e B. Then U is a solution of the displacement boundary value
for every U
problem (3.6.13).

Copyright 2004 by Chapman & Hall/CRC

168

CHAPTER 3. COMPOSITE LAMINATES

Let U0 be an arbitrary vector field of class C on D, and suppose that U0


e = U0 +U B. Using equation (3.6.19), the assumption
vanishes on D. Then U
0
(3.6.25) and the fact that Un0 = U0 = U30 = U3,n
= 0 on D, we get
Z
{U0 N, + U30 (M, + q)}da 0.
W (U0 )

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

Since this inequality must remain true if U0 is replaced by U0 , we get


Z
{U0 N, + U30 (M, + q)}da = 0.
D

Since U is arbitrary in D, from the last equation and from the supplementary
constitutive relations (3.6.5), it follows that the equilibrium equations (3.5.8) are
verified by U. Moreover, since U B, U satisfies also the boundary conditions
(3.6.13). Hence, U is a regular solution of the displacement boundary value
problem (3.6.13) and the demonstration is complete.
In a similar way, the converses of principles of minimum potential energies
corresponding to various, possible boundary value problems can be proved.
We note also that, as in the usual elasticity theory, the principles of minimum potential energy can be used to prove uniqueness theorems for the involved
boundary value problems.
As we have seen, in order to prove uniqueness theorems and minimum principles, we have assumed that the specific deformation (strain) energy w is a positive
definite quadratic form.
We recall also that from the beginning we have supposed that the elasticity
tensor c of any linearly elastic body is positive definite. Using this hypotheses, in
the Section 2.2 we have derived various restrictions which must be satisfied by the
elasticities in order to ensure the positiveness of c.
In a natural way, the following question is raised. If the elasticities ck , k =
1, ..., N of the laminate are positive definite, does the specific strain energy w of
the laminate have the same property?
Since the reduced transformed stiffnesses (Qij )k of the laminae and the laminate stiffnesses Aij , Bij , Dij are complicated functions of the laminae mechanical
and geometrical characteristics, it is difficult to find an answer for the above question for a laminate having arbitrary structure.
This is the reason why we shall analyze the problem only for the particular,
but important case of the regular antisymmetric cross-ply laminates, for which
the extensional, coupling and bending stiffnesses of the composite laminate can
be expressed by relatively simple relations in terms of the primary mechanical
characteristics of the laminae, using Tsais formulas (3.4.26).
We recall that from the positive definiteness of the elasticity c of an orthotropic elastic material, it follows that the technical constants of the body must
satisfy the restrictions (2.2.74)(2.2.79). In particular, we have

E1 , E2 , G12 > 0, 12 , 21 < 1,

Copyright 2004 by Chapman & Hall/CRC

(3.6.26)

169

3.6. VARIATIONAL AND EXTREME PRINCIPLES


|12 | <

E1
, |21 | <
E2

E2
.
E1

(3.6.27)

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

If we assume in addition that a traction tensor stress acting in a principal


material direction produces extension in that direction and contractions in the
other two principal material directions, we can conclude that
12 , 21 > 0.

(3.6.28)

Hence in this case, the Poissons ratios 12 and 21 satisfy the restrictions
0 < 12 <

E1
, 0 < 21 <
E2

E2
.
E1

(3.6.29)

We recall that E1 is the Youngs modulus in the fibers direction and E2 is


the Youngs modulus in the perpendicular, in-plane direction. Consequently, we
assume that
E1
< 1.
(3.6.30)
0<f
E2

Using the Tsais stiffness ratio f , we can express the inequalities (3.6.29) satisfied
by Poissons ratios 12 and 21 in the equivalent form
0 < 12 <

p
1
, 0 < 21 < f .
f

(3.6.31)

We return to equations (3.1.6) giving the (primary) reduced stiffness Q11 , Q12 ,
Q22 and Q66 in terms of the engineering constants of the lamina; we have

21 E1
E1
,
, Q12 =
1 12 21
1 12 21
E2
and Q66 = G12 .
=
1 12 21

Q11 =

Q22

(3.6.32)

Thus, from (3.6.26) and (3.6.28) we can conclude that all these stiffnesses are
positive; i.e.
Q11 , Q12 , Q22 , Q66 > 0.

(3.6.33)

The positive definiteness of the elasticity c of the lamina assures only the
positivity of the reduced stiffnesses Q11 , Q22 and Q66 . For the positivity of the
stiffness Q12 some supplementary assumption, leading to the inequalities (3.6.28),
must be also used.
We recall now Tsais formulas (3.4.26) giving the extensional, coupling and
bending stiffnesses of a regular antisymmetric cross-ply laminate. Using also the

Copyright 2004 by Chapman & Hall/CRC

170

CHAPTER 3. COMPOSITE LAMINATES

relation (3.6.32), we get

221
E1
1+f
A11 , A66 = hG12 ,
, A12 =
h
1+f
1 12 21
2
1 f 1
hA11 ,
B11 =
2N f + 1
h2
h2 212
h2
A66 .
(3.6.34)
A11 , D66 =
A11 , D12 =
D11 =
12
12 1 + f
12

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

A11 =

From (3.6.26), (3.6.28), (3.6.30) and (3.6.34), we can conclude that


A11 , A12 , A66 > 0,
B11 < 0,
D11 , D12 , D66 > 0.

(3.6.35)

We observe now that the positive definiteness of the specific strain energy w of
a laminate, defined by equation (3.5.29), is equivalent with the positive definiteness
of the global stiffness matrix [E] of the laminate, introduced by equation (3.3.20).
According to Tsais relations (3.4.26), the stiffness matrix [E] of a regular
antisymmetric cross-ply laminate has the following expression:

A11 A12
0
B11 0
0

A12 A11
0
0
B11 0

0
0
A66 0
0
0
.
(3.6.36)
[E] =

B11 0
0
D11 D12
0

0
B11 0
D12 D11
0
0
0
0
0
0
D66

According to the Sylvesters criterion, [E] is positive definite


following six determinants are positive:

A11 A12 0



A11 A12


, 3 = A12 A11 0
1 = |A11 | , 2 =
A12 A11
0
0
A66




4 =

A11
A12
0
B11

A12
A11
0
0

0
0
A66
0

6 = det [E] .

B11
0
0
D11



A11



A12


, 5 = 0



B11


0

A12
A11
0
0
B11

0
0
A66
0
0

if and only if the





,

B11
0
0
D11
D12

0
B11
0
D12
D11

(3.6.37)
Long, but elementary computations lead to the following expression of the
above determinants:

Copyright 2004 by Chapman & Hall/CRC

171

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.6. VARIATIONAL AND EXTREME PRINCIPLES



2
421
A211 ,
1 = A11 , 2 = 1
(f + 1)2
( 

2
421
1
h2
3

1
A66 A11
4 =
(f + 1)2
3
2
( 

2
421
1
h4 4

1
A11 A66
5 =
(f + 1)2
3
16

3 = 2 A66 ,
2 )

f 1
1
,
N2 f + 1
2 )

f 1
1
,
N2 f + 1

6 = 5 D66 .

(3.6.38)

From the (3.6.35) and the first relation (3.6.38), we get


1 > 0.

(3.6.39)

We use now the restriction (3.6.31)2 satisfied by 21 . Thus we obtain


2

2
f 1
4f
421
> 0.
=
>1
1
f +1
(f + 1)2
(f + 1)2

Consequently, from the second relation (3.6.38), we get


2 > 0.

(3.6.40)

The third relation (3.6.38) and the above equality show that
3 > 0,

(3.6.41)

since, according to (3.6.35), A66 > 0.


Using again the restriction (3.6.31)2 , we successively get
2

2

2



2
f 1
1
1 f 1
f 1
1
421
1
2
>
2
1
f +1
N
3 f +1
f +1
N
(f + 1)2
3
2


f 1
1
1
.

=
f +1
3 N

We recall that the analyzed structure is a regular antisymmetric cross-ply


laminate. Hence N can take only the even values 2, 4, 6, .... Consequently, N 2.
Hence 31 N12 > 31 41 > 0. Thus, from the above relations, we can conclude that
1
3

2
421
1
(f + 1)2

1
2
N

f 1
f +1

2

> 0.

Now the fourth equation (3.6.38) leads to


4 > 0,

Copyright 2004 by Chapman & Hall/CRC

(3.6.42)

172

CHAPTER 3. COMPOSITE LAMINATES

Hence, according to the last two equations (3.6.38), we have also

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

5 > 0 , 6 > 0

(3.6.43)

since the inequalities (3.6.35) take place.


Examining the inequalities (3.6.39)(3.6.43), we can see that all Sylvesters
determinants (3.6.37) are positive. Hence, we have the following.
Theorem of positive definiteness. The specific strain (deformation) energy w of a regular antisymmetric cross-ply laminate is positive definite, if the
specific strain (deformation) energies of the composing laminae are positive definite and if tensile tests applied in the fibers direction produce, in each laminae,
extensions in that direction and contractions in the perpendicular in-plane direction.
As the above example shows, the analysis of the positive definiteness of the
specific strain energy of a composite laminate can generally be a difficult task.
However, such analysis must be done since questions concerning uniqueness and
extreme properties are strongly connected with positive definiteness. Moreover, as
we shall see in the Chapter 7, the stability of a fiber-reinforced composite laminate
can be lost if its specific strain energy ceases to be positive definite.

3.7

Rectangular laminates

Introducing the global constitutive equations (3.5.7) and the geometrical


equations (3.5.6) into the equilibrium equations (3.5.8)1 and (3.5.8)4 , we shall obtain the differential system which must be satisfied by the components U 1 , U2 , U3 ,
of the displacement U in the plane domain D occupied by the middle surface of the
laminate. Elementary, but long computations lead to the following displacement
equilibrium equations:
A11 U1,11 + 2A16 U1,12 + A66 U1,22 + A16 U2,11 + (A12 + A66 )U2,12 + A26 U2,22
B11 U3,111 3B16 U3,221 (B12 + 2B66 )U3,122 B26 U3,222 = 0,
A16 U1,11 + (A12 + A66 )U1,12 + A26 U1,22 + A66 U2,11 + 2A26 U2,12 + A22 U2,22
B16 U3,111 (B12 + 2B66 )U3,112 3B26 U3,122 B22 U3,111 = 0,
D11 U3,1111 + 4D16 U3,1112 + 2(D12 + 2D66 )U3,1122 + 4D26 U3,1222 + D22 U3,2222
B11 U1,111 3B16 U1,112 (B12 + 2B66 )U1,122 B26 U1,222
B16 U2,111 (B12 + 2B66 )U2,112 3B26 U2,122 B22 U2,222 = q.
(3.7.1)

Copyright 2004 by Chapman & Hall/CRC

173

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.7. RECTANGULAR LAMINATES

Generally, the relation (3.7.1) is a complicated and coupled differential system. In-plane deformation and distortions of the laminate appear together, in an
inseparable way. But, obvious and sometimes important simplifications result
when the laminate is symmetric about the middle surface (Bij = 0), specially
orthotropic (all the terms with 16 and 26 indices vanish in addition to the B ij ),
2
homogeneous (Bij = 0 and Dij = Aij h12 ), or isotropic. In all these cases, equations
(3.7.1)1,2 are uncoupled from the equations (3.7.1)3 . In such situations, the first
two equations contain only the in-plane displacements U1 , U2 , and the third one
contains only the normal displacement U3 . Accordingly, equation (3.7.1)3 must
be solved to obtain the deflections of a plate and the system (3.7.1)1,2 must be
integrated to obtain the in-plane deformations of the laminate. Unfortunately, the
more general case of nonsymmetric laminates (there exist nonvanishing coupling
stiffnesses Bij ) requires the simultaneous integration of the coupled system (3.7.1).
Currently, many methods exist to solve the equilibrium equations (3.7.1).
The methods range from exact solutions to approximate numerical integration,
using finite element or finite difference approaches and various approximate energy methods of Rayliegh-Ritz or Galerkin type, based on the given minimum or
variational principles. In what follows, we shall consider only the case of a rectangular laminate, taking into account a small number of special examples.
Let us consider first the general class of laminated rectangular plates, as
shown in Figure 3.20, that are simply supported along edges x1 = 0, x1 = a1 , x2 =
0, x2 = a2 and subjected to a given normal load q = q(x1 , x2 ).
We suppose that the given normal load q = q(x1 , x2 ) can be expanded in a
double Fourier series; i.e.

q(x1 , x2 ) =

m=1 n=1

qmn sin

nx2
mx1
.
sin
a2
a1

(3.7.2)

Figure 3.20: Simply supported laminated rectangular plate under distributed normal load.
In what follows, the various types of possible laminates, such as specially orthotropic, symmetric angle-ply, antisymmetric cross-ply and antisymmetric angleply will be analyzed for the loading q = q(x1 , x2 ) given by equation (3.7.2).

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

174

CHAPTER 3. COMPOSITE LAMINATES

As we know, a specially orthotropic laminate has either a single layer of specially orthotropic material or multiple specially orthotropic layers that are symmetrically placed about the laminate middle surface. For this special case considered, the nonvanishing laminate stiffnesses are A11 , A12 , A22 , A66 , D11 , D12 , D22 ,
D66 . In other words, neither shear or twist coupling, nor bending-extension coupling exists. Also, the in-plane deformations and bending and twisting are decoupled.Thus, for the laminate problem, the vertical deflection are described only by
one differential equation of equilibrium, resulting from the third equation (3.7.1),
D11 U3,1111 + 2(D12 + 2D66 )U3,1122 + D22 U3,2222 = q.

(3.7.3)

The boundary conditions corresponding to simply supported edges become


U3 = 0 , M11 = D11 U3,11 D12 U3,22 = 0 for x1 = 0 and x1 = a1 ,

(3.7.4)

U3 = 0 , M22 = D12 U3,11 D22 U3,22 = 0 for x2 = 0 and x2 = a2 .


The solution of the boundary value problem can be determined using the
separation of variables technique, as in the case of isotropic rectangular plates. In
this way, it is easy to check that the solution satisfying the boundary conditions
(3.7.4) must have the following relatively simple form:
U3 (x1 , x2 ) =

m=1 n=1

Amn sin

nx2
mx1
.
sin
a2
a1

(3.7.5)

This normal displacement field satisfies the equilibrium equation (3.7.3) only
if the Fourier coefficients amn are given by the relation
Amn =

qmn
1
.
m 4
4
D11 ( a1 ) + 2(D12 + 2D66 )( am1 )2 ( an2 )2 + D22 ( an2 )4

(3.7.6)

Once the normal displacement or deflection U3 is known, all force and moment
resultants can be obtained using the corresponding constitutive equations.
The case of a symmetric angle-ply laminate is much more complicated, even
if, for these structures, the coupling stiffnesses Bij are also vanishing. But now the
shear coupling stiffnesses A16 , A26 and the twist coupling stiffnesses D16 , D26 are
not zero. The equilibrium equation (3.7.1)3 , describing the deflection of the plate,
is decoupled and takes the form
D11 U3,1111 + 4D16 U3,1112 + 2(D12 + 2D66 )U3,1122 + 4D26 U3,1222 + D22 U3,2222 = q.
(3.7.7)
The simply supported edge condition becomes
U3 = 0, M11 = D11 U3,11 D12 U3,22 2D16 U3,12 = 0 for x1 = 0 and x1 = a1 ,
U3 = 0, M22 = D12 U3,11 D22 U3,22 2D26 U3,12 = 0 for x2 = 0 and x2 = a2 .
(3.7.8)

Copyright 2004 by Chapman & Hall/CRC

175

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.7. RECTANGULAR LAMINATES

Now the solution of the equilibrium equation (3.7.7) is not as simple as before,
because of the presence of D16 and D26 . Due to the terms involving these nonvanishing coefficients, the method using the separation of variables cannot be applied and a Fourier type expression (3.7.5) does not satisfy the governing equation
(3.7.7). Moreover, the expansion (3.7.5) also does not satisfy the boundary conditions (3.7.8), again since the terms involving D16 and D26 are present. Actually,
the variables x1 and x2 cannot be separated in the expression of the normal displacement U3 . This is the reason why Ashton (see Jones [3.2], Chapter 5) has
solved the problem using the second principle of the minimum potential energy.
The problem being decoupled, the involved functional J(U), given in the equation
(3.6.14), takes the following simplified form:
1
2

2
2
+ 2D12 U3,11 U3,22 + D22 U3,22
(D11 U3,11
Z
2
qU3 da.
+4D66 U3,12 + 4D16 U3,11 U3,12 + 4D26 U3,22 U3,12 )da

J = J(U3 ) =

(3.7.9)

Ashton has approximated the deflection U3 by a finite number of terms of the


Fourier expansion (3.7.5). Such an expression satisfies the displacement boundary
conditions (3.7.8), hence is an admissible displacement field. The boundary conditions (3.7.8) concerning the bending moments must not be a` priori satisfied, since
they are natural boundary conditions. Ashtons approach is obviously a RayleighRitz method, applied to our particular case. Minimizing the functional J(U 3 ),
taking into account a finite number of terms in the expansion (3.7.5), we are lead
to a finite set of simultaneous linear algebraic equations for the chosen unknowns
amn . Such a system can be solved using a computer. Let us observe again that
only the displacement boundary conditions are exactly satisfied, but the natural
boundary conditions will be only approximately satisfied. The convergence of the
method may be slow, just because the natural boundary condition are not satisfied
exactly. For instance, Ashton has used 49 terms (up to m = 7 and n = 7) to obtain
approximatively the deflection U3 . Thus, for a uniformly loaded (q(x1 , x2 ) = 1)
(D12 +2D66 )
D26
16
22
= 1, D
square laminate (a1 = a2 = a) with D
D11 = D11 = 0.5,
D11
D11 = 1,
the maximum deflection at the center of the plate, found by Ashton, is

U3 max = U3 (0, 0) =

0.00425a4 q
.
D11

If D16 and D26 are ignored, that is the symmetric angle-ply laminate is approximated by a special orthotropic laminate having

(D12 + 2D66 )
D11
= 1 , D16 = D26 = 0,
= 1,
D11
D22

the maximum deflection found by Ashton is


U3 max = U3 (0, 0) =

Copyright 2004 by Chapman & Hall/CRC

0.0032a4 q
.
D11

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

176

CHAPTER 3. COMPOSITE LAMINATES

Hence, the error in neglecting the twist coupling terms D16 and D26 , is about
24 percent, which represents a no negligible error. Thus, generally, a specially
orthotropic laminate is an unacceptable approximation for a symmetric angle-ply
laminate.
Let us consider now a regular antisymmetric cross-ply laminate. Such a
laminate has nonvanishing extensional stiffnesses A11 , A12 , A11 = A22 and A66 ,
bending-extensional coupling stiffnesses B11 and B22 = B11 , and bending stiffnesses D11 , D12 , D22 = D11 and D16 . Since B11 and B22 are not vanishing, the
displacement equilibrium equations are coupled. The general system (3.7.1) takes
the following simplified form:
A11 U1,11 + A66 U1,22 + (A12 + A66 )U2,12 B11 U3,111 = 0,
(A12 + A66 )U1,12 + A66 U2,11 + A11 U2,22 + B11 U3,222 = 0,

(3.7.10)

D11 (U3,1111 + U3,2222 ) + 2(D12 + 2D66 )U3,1122 B11 (U1,111 U2,222 ) = 0.


Whitney and Leissa (see Jones [3.2], Chapter 5) have solved the problem for
simply supported edges taking into account boundary conditions of type S 2 (see
Equations (3.6.17))
U3 = 0, M11 = B11 U1,1 D11 U3,11 D12 U3,22 = 0, for x1 = 0 and x1 = a1,
U2 = 0, N11 = A11 U1,1 + A12 U2,2 B11 U3,11 = 0,
U3 = 0, M22 = B11 U2,2 D12 U3,22 D11 U3,22 = 0, for x2 = 0 and x2 = a2 ,
U1 = 0, N22 = A12 U1,1 + A11 U2,2 + B11 U3,22 = q.
(3.7.11)
These authors have observed that in this case the variables x1 and x2 can
be separated and the displacements U1 , U2 , U3 can be obtained using Fouriers
method; i.e.
U1 =

U2 =

U3 =

m=1 n=1
X

m=1 n=1
X

m=1 n=1

mn cos

nx2
mx1
,
sin
a2
a1

mn sin

nx2
mx1
,
cos
a2
a1

mn sin

nx2
mx1
.
sin
a2
a1

(3.7.12)

It is easy to see that the boundary conditions (3.7.11) are satisfied. Obviously,
the Fourier coefficients mn , mn and mn can be obtained taking into account the
equilibrium equation (3.7.10).
The results obtained by Whitney and Leissa are replotted in Figure 3.21 for
the special situation in which the normal load is one term of the Fourier series;

Copyright 2004 by Chapman & Hall/CRC

177

3.7. RECTANGULAR LAMINATES


i.e.

x2
x1
, q0 = const.
sin
a2
a1

Figure 3.21 gives the normalized maximum deflexion for a rectangular (a = b)


regular antisymmetric cross-ply graphite/epoxy laminated plate, for 2,4,6 and an
infinite number of layers.
20
2

15

U3,MAX E 2 h3 3
10
q 0 a24

E1
G
=40 E12=0.5
E2
2

=0.25

12

NUMBER OF LAYERS

10
4
6

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

q = q(x1 , x2 ) = q0 sin

x
q(x1,x2)=q0sin ax1 sin a22
1

2
a
ASPECT RATIO, a 1
2

Figure 3.21: Maximum deflection of a rectangular regular antisymmetric cross ply


laminate under sinusoidal normal load.
Let us observe that the infinite number of layers corresponds to the specially
orthotropic laminate, for which the coupling between bending and extension does
not exist. As the results show, for a two-layered laminate, the neglect of the coupling results in an under-prediction of the deflection by 64 percent. That is the
actual prediction is approximately three times bigger than the prediction corresponding to the orthotropic approximation. However, the effect of coupling on the
deflection dies out rapidly as the number of layers increases, independent of the
plate aspect ratio ab . We can say that when more than six layers exist, the coupling can be neglected without important error and the orthotropic approximation
becomes acceptable.
As Whitney and Leissa have shown, for composite laminates, the effect of
coupling between bending and extension on the plate deflection depends essentially
G12
1
on the stiffness ratio E
E2 , whereas the influence of E2 and 12 is relatively small.

Copyright 2004 by Chapman & Hall/CRC

178

CHAPTER 3. COMPOSITE LAMINATES

In Figure 3.22 are given the maximum deflection of a square regulate antisymmetric cross-ply laminate, under sinusoidal transverse load, in terms of the
E1
1
orthotropic modulus ratio E
E2 . At E2 = 1, the effect of coupling is nonexistent. As
E1
E2 increases, the effect of coupling between bending and extension increases.

25
3

U 3,MAX E 2 h 2
10
q0a 4

G12
=0.5
E2

20

=0.25

12

15

2
10
4

NUMBER
OF
LAYERS

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

30

x
q(x1,x2)=q sin ax1 sin a 2
0

10

20

30

MODULUS RATIO,

E1
E2

40

50

Figure 3.22: Maximum deflection of a square regulate antisymmetric cross-ply


laminated plate under sinusoidal transverse load.
Summing up the conclusions resulting from the above presented examples,
we can say the following:
(1) In few special situations, the variables can be separated and the exact
solution can be obtained, but only as a Fourier series expansion.
(2) Frequently the Fourier series are lent convergent and many terms must
be taken into account to obtain a satisfactory solution.
(3) Generally the variables cant be separated. In these situations RayleighRitz and /or Galerkin type numerical methods can, and must be used, taking into
account the corresponding extreme and variational principles, and using computers to solve the resulting linear algebraic systems. The displacement fields taken
into account must satisfy exactly all given displacement boundary conditions.
(4) The presence in a composite laminate of coupling between bending and
extension generally increases deflections. Hence, coupling decreases the effective
stiffness of a composite laminate.
(5) For laminates exhibiting twist-curvature coupling, the deflections are increased.

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.8. PROBLEMS

179

(6) In the case of bending-extension coupling and twist-curvature coupling,


the effect on deflection decreases rapidly as the number of layers increases.
(7) The approximation of a general laminate by a specially orthotropic one
can lead to bigger errors by a factor of 3. Thus, the use of such an approximation
must be carefully proven to be justified for the case under consideration.
(8) The first general rule that can be established analyzing the presented
results tells us that coupling should be included in every analysis of composite
laminate behavior, unless coupling is proven to be insignificant.
(9) The first general rule tells us that for general laminates, specific investigations are necessary to obtain accurate information concerning their behavior.

3.8

Problems

P3.1 Show that for an isotropic lamina and in plane stress state, the strainstress relation are


1
1
S11 S12 0
2
2 = S12 S11 0
0
0
2(S11 S12 )
6
6

with S11 = E1 , S12 = E , E and being Youngs modulus and the Poissons
ratio of the laminae, respectively.
P3.2 Show that for an isotropic lamina and in plane stress state, the stressstrain relations are

1
Q11 Q12 0
1
2 = Q12 Q11 0
2
0
0
Q66
6
6

with

Q11 =

E
E
E
= = G.
, Q66 =
, Q12 =
2
2
2(1 + )
(1 )
(1 )

P3.3 The engineering constants of a scotchply1002 glass/epoxy fiber-reinforced


lamina have the following values:
E1 = 38.6GP a , E2 = 8.27GP a , 12 = 0.26 , G12 = 4.14GP a.
Find the compliance components and the reduced stiffnesses of this lamina.
P3.4 The engineering constants of a Kevlar 49/E epoxy type aramid/epoxy
lamina are
E1 = 76GP a , E2 = 5.5GP a , 12 = 0.34 , G12 = 2.3GP a.
Find the compliance components and the reduced stiffnesses of this lamina.

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

180

CHAPTER 3. COMPOSITE LAMINATES

P3.5 Prove by direct calculations that equations (3.1.14) are true.


P3.6 Obtain by direct calculations the relations (3.1.19)(3.1.21).
P3.7 Prove that the equations (3.1.23), giving the apparent technical moduli
Ex , xy , Ey , Gxy , xy,x and xy,y , as functions of the angle [00 , 900 ], are true.
0
0
x Gxy
P3.8 Plot E
E2 , G12 , xy and xy,x as functions on [0 , 90 ] for a boron/epoxy composite, an orthotropic material with

E1 = 10E2 , G12 =

1
E2 , 12 = 0.3 , E2 = 18.5GP a.
3

Analyze the obtained results.


P3.9 Show that the apparent axial or longitudinal modulus Ex of a fiberreinforced orthotropic lamina can be written as function of [00 , 900 ] in the
following form:
E1
= (1 + a 4b) cos4 + 2 (2b a) cos2 + a
Ex

with
E1
and b =
a=
Ex

E1
G12

212

P3.10 Use the above expression of Ex to find its maxima and minima. Show
that if
E1
,
G12 >
2 (1 + 12 )

Ex is greater than both E1 and E2 for some value of , and if


G12 <

E1
E1
E2

+ 12

,

Ex is smaller than both E1 and E2 for some value of .


P3.11 Prove the validity of the equations (3.1.24) and (3.1.25).
P3.12 Prove that the combinations Q11 + Q22 + 2Q12 and Q66 Q12 are
invariant under rotations about the x3 = z axis and these invariants have the
constant values given by the equations

Q11 + Q22 + 2Q12 = Q11 + Q22 + 2Q12 = 2 (U1 + U4 ) ,


Q66 Q12 = Q66 Q12 = U5 U6 ,

where U1 , U4 , U5 and U6 are given by equations (3.1.25).


P3.13 Show that the reduced stiffnesses as a function of the angle can be
expressed in the following matrix form:

Copyright 2004 by Chapman & Hall/CRC

181

3.8. PROBLEMS

Q11
U1 U2
0
U3
0
1

Q22 U1 U2 0
U3
0
cos 2

Q12 U4 0
0
U
0
3
sin 2 ,
=

U5 0
Q
0
4U3 0
cos 4
66

2Q 0
0
U2 0
2U3
16
sin 4
0
0
U2 0
2U3
2Q26
where U1 , U2 , U3 , U4 , U5 are given by the equations (3.1.25).
P3.14 The engineering constants of a T300/5208 graphyte/epoxy fiber-reinforced lamina are given in Table 3.1. Find the coefficients U1 , U2 , U3 , U4 and U5
for this material.
P3.15 A boron/epoxy fiber-reinforced composite lamina has the following
engineering constants:
E1 = 206.85GP a , E2 = 20.68GP a , 12 = 0.3 , G12 = 6.86GP a.
(a) Plot the reduced transformed
stiffnesses Q11 (), Q22 (), Q12 () and Q66 ()

for this lamina for 00 , 900 .
(b) Plot the reduced
transformed
stiffnesses Q16 () and Q26 () for the same


lamina and for 00 , 900 .
(c) Analyze the results obtained in (a) and (b).
P3.16 Show that the functions S 11 (), ..., S 66 (), given by equations (3.1.21)
can be expressed in the following matrix form:

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

S 11
S 22
S 12
S 66
S 16
S 26

V1
V1
V4
V5
0
0

V2
V2
0
0
0
0

0
0
0
0
V2
V2

V3
V3
V3
4V3
0
0

0
0
0
0
2V3
2V3

1
cos 2
sin 2
cos 4
sin 4

where
V1 =

V2 =

V3 =

V4 =

V5 =

1
(3S11 + 3S22 + 2S12 + S66 ) ,
8
1
(S11 S22 ) ,
2
1
(S11 + S22 2S12 S66 ) ,
8
1
(S11 + S22 + 6S12 S66 ) ,
8
1
(S11 + S22 2S12 + S66 ) .
2

b1 and the variation


P3.17 Plot the variation of the overall elastic modulus E
of the overall Poissons ratio b12 with the fiber volume fraction cf [0, 1], using
Voigts mixture rule.

Copyright 2004 by Chapman & Hall/CRC

182

CHAPTER 3. COMPOSITE LAMINATES


b

P3.18 Plot the variations of the ratios EEm2 and


cf [0, 1], using Reuss mixture rule, assuming that

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

(a)

Ef
= 10;
Em

(b)

b12
G
Gm

with the volume fraction

Ef
= 100.
Em

P3.19 What conclusion do you get if you assume for the determination of
b2 equal strains in both the fiber and the matrix,
the overall transverse modulus E
instead of equal stresses.
P3.20 Show that if the curvatures k (x1 , x2 ), , = 1, 2 are vanishing, the
middle surface of a composite laminate rests plane after the deformation of the
laminate. In other words, show that the curvatures characterizes the bending and
the twisting of the laminate.
P3.21 Let us assume that the in-plane deformations e (x1 , x2 ) and the curvatures k (x1 , x2 ), , = 1, 2 of a laminate are vanishing.
Find in that case the components uk (x1 , x2 , x3 ), k = 1, 2, 3 of the displacement field and give the geometrical meaning of the obtained result.
P3.22 Assuming that a laminate is submitted to a rigid displacement, find
the components of the in-plane deformation and the components of the curvature.
P3.23 Show that the extensional stiffnesses Aij , i, j = 1, 2, 6 of a composite
laminate can be expressed by the following relations:
A11 = U1 V0A + U2 V1A + U3 V3A ,
A22 = U1 V0A U2 V1A + U3 V3A ,
A12 = U4 V0A U3 V3A ,
A66 = U5 V0A U3 V3A ,
1
A16 = U2 V2A U3 V4A ,
2
1
A26 = U2 V2A + U3 V4A ,
2

where
V0A = h,
N
X
V1A =
(zk zk1 ) cos 2k ,
k=1

V2A =

N
X

k=1

V3A =

N
X

k=1

V4A =

N
X

k=1

Copyright 2004 by Chapman & Hall/CRC

(zk zk1 ) sin 2k ,


(zk zk1 ) cos 4k ,
(zk zk1 ) sin 4k .

183

3.8. PROBLEMS

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

P3.24 Show that the coupling stiffnesses of a composite laminate can be


expressed by the following relations:
B11 = U2 V1B + U3 V3B ,
B22 = U2 V1B + U3 V3B ,

B12 = U3 V3B ,
B66 = U3 V3B ,
1
B16 = U2 V2B U3 V4B ,
2
1
B26 = U2 V2B + U3 V4B ,
2

where
V0B = 0,
N

V1B

1X 2
2
(zk zk1
) cos 2k ,
=
2
k=1

V2B =

1X 2
2
(zk zk1
) sin 2k ,
2
k=1

V3B =

1X 2
2
) cos 4k ,
(zk zk1
2
k=1

V4B =

1X 2
2
) sin 4k .
(zk zk1
2
k=1

P3.25: Show that the bending stiffnesses of a composite laminate can be


expressed by the following relations:
D11 = U1 V0D + U2 V1D + U3 V3D ,
D22 = U1 V0D U2 V1D + U3 V3D ,
D12 = U4 V0D U3 V3D ,

D66 = U5 V0D U3 V3D ,


1
D16 = U2 V2D U3 V4D ,
2
1
D26 = U2 V2D + U3 V4D ,
2

Copyright 2004 by Chapman & Hall/CRC

184

CHAPTER 3. COMPOSITE LAMINATES

where
V0D =

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

V1D =

h3
,
12
N
1X

k=1

3
) cos 2k ,
(zk3 zk1

V2D =

1X 3
3
) sin 2k ,
(zk zk1
3
k=1

V3D =

1X 3
3
) cos 4k ,
(zk zk1
3
k=1

V4D

1X 3
3
) sin 4k .
(zk zk1
=
3
k=1

P3.26 Prove that the relations given in P3.24-P3.26 can be expressed


following concentrated matrix form:

U1 U2
0
U3
0
[A11 , B11 , D11 ]
V0[A,B,D]

[A22 , B22 , D22 ] U1 U2 0


U
0
3
V1[A,B,D]

[A12 , B12 , D12 ] U4 0


0
U3 0
V2[A,B,D]

[A66 , B66 , D66 ] = U5 0


0
U
0
3
V3[A,B,D]

2 [A16 , B16 , D16 ] 0


0
U2 0
2U3
V4[A,B,D]
0
0
U2 0
2U3
2 [A26 , B26 , D26 ]

in the

P3.27 Let us consider a composite laminate submitted to the constant force

resultants N 11 , N 22 and N 12 . Find the appropriate constant moment resultants

M 11 , M 22 and M 12 which must be applied to the laminate to obtain zero curvatures.


(a) Solve first the problem using the global constitutive equations expressing
[N ] and [M ] in terms of [e] and [k].
(b) Next, solve the problem using the inversed global constitutive equations
expressing [e] and [k] in terms of [N ] and [M ].
(c) Compare the results obtained in (a) and (b).
(d) Analyze the case in which the coupling stiffnesses are vanishing.
P3.28 Find for a single layer isotropic laminate with Young modulus E, Poisson ratio and thickness h, the extensional, coupling and bending stiffnesses.
P3.29 Find for a single layer specially orthotropic laminate with technical
constants E1 , E2 , 12 , G12 and thickness h, the components of the global stiffness
matrix [E].
P3.30 Consider two orthotropic laminae with principal material directions at
+ and with respect to the laminate reference axis. Prove that




Q16 + = Q16 and Q26 + = Q26 .

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

3.8. PROBLEMS

185

P3.31 Show that if a cross-ply laminate is symmetric about its middle surface,
all of its coupling stiffnesses are vanishing.
P3.32 Starting with Tsais formulas (3.4.24) and using the definition of a
regular symmetric cross-ply laminate, express the stiffnesses of the laminate in
2
terms of the reduced stiffness Q11 , the thickness h and the stiffness ratio f = E
E1 .
P3.33 Give a direct proof to the results obtained in P3.32.
P3.34 Prove that the global stiffness matrix [E] of a regular symmetric crossply laminate is positive definite if the elasticity tensor of the composing laminae
are positive definite, and the Poissons ratio satisfy the inequalities ij > 0 for
i, j = 1, 2, 3.
P3.35 An expanded view of a [+45/45/45/+45] regular angle-ply laminate
consisting of 0.25-mm thick unidirectional AS/3501 graphite/epoxy laminae is
shown in Figure 3.23. Determine the stiffness matrix [E] of this laminate.The
engineering constants of the laminae are

E1 = 138GP a , E2 = 9GP a , 12 = 0.3 , G12 = 6.9GP a , 21 = 12

E2
= 0.0196.
E1

Figure 3.23: Exploded view of a [+45/45/45/+45] regular angle-ply laminate.


P3.36 An exploded view of a [45/+45/45/+45] regular angle-ply laminate
consisting of the same 0.25-mm thick unidimensional A.S/3501 graphite/epoxy
laminae as used in P3.35 is given in Figure 3.24. Determine the global stiffness
matrix [E] of the laminate.
P3.37 The angle-ply laminate described in P3.35 is subjected to a single
uniaxial force per unit length N11 = 50M P a mm1 .
(a) Determine the resulting in-plane deformations and curvatures associated
with the x1 and x2 axes.

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

186

CHAPTER 3. COMPOSITE LAMINATES

Figure 3.24: Exploded view of a [45/+45/45/+45] regular angle-ply laminate.

(b) Determine the resulting deformations and displacements associated with


the x1 , x2 and x3 axes in each laminae.
(c) Determine the resulting stresses associated with the x1 and x2 axes in
each laminae.
P3.38 The angle-ply laminate described in P3.36 is subjected to a simple
uniaxial force per unit length N11 = 50M P a mm1 .
(a) Determine the resulting in-plane deformations and curvatures associated
with the x1 and x2 axes.
(b) Determine the resulting deformations and displacements associated with
the x1 , x2 and x3 axes in each laminae.
(c) Determine the resulting stresses associated with the x1 and x2 axes in
each laminae.
P3.39 A regular angle-ply laminate has N unidirectionally reinforced orthotropic layers having the same thickness and with principal material directions
alternatingly oriented at + and to the laminate coordinate axes. The oddnumbered plies are at and the even-numbered plies at +. Show that the
transformed reduced stiffnesses of the laminate satisfy the following equations:





Q11 + = Q11 , Q12 + = Q12 ,




Q22 + = Q22 , Q66 + = Q66 ,




Q16 + = Q16 , Q26 + = Q26 .

P3.40 Show that for a symmetric (N odd!) regular angle-ply laminate, the

Copyright 2004 by Chapman & Hall/CRC

187

3.8. PROBLEMS

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

laminate stiffnesses are given by the following equations:


(A11 , A12 , A22 , A66 ) = h Q11 , Q12 , Q22 , Q66 ,

h
Q16 , Q26 , Bij = 0, i, j = 1, 2, 6,
(A16 , A26 ) =
N

h3
Q11 , Q12 , Q22 , Q66 ,
(D11 , D12 , D22 , D66 ) =
12

h3 2N 2 2
Q16 , Q26 .
(D16 , D26 ) =
3
12 N

In these equations in Qij , i, j = 1, 2, 6, and is accounted for.


P3.41 Show that for an antisymmetric (N even!) regular angle-ply laminate,
the laminate stiffnesses are given by the following equations:

(A11 , A12 , A22 , A66 ) = h Q11 , Q12 , Q22 , Q66 , (A16 , A26 ) = 0,

h2
Q16 , Q26 ,
(B11 , B12 , B22 , B66 ) = 0, (B16 , B26 ) =
2N

h3
(D11 , D12 , D22 , D66 ) =
Q11 , Q12 , Q22 , Q66 , (D16 , D26 ) = 0.
12

P3.42 Show that A16 , A26 and D16 , D26 for a symmetric regular angle-ply
laminate approach zero as the number of layers increases, while the total thickness
is held constant. What happens if equal thickness layers are added so that the
total laminate thickness increases, too?
P3.43 Show that B16 and B26 for an antisymmetric regular angle-ply laminate
approach zero as the total number of layers increases, while the total thickness is
held constant. What happens if equal thickness layers are added so that the total
laminate thickness increases, too?
P3.44 In what conditions are the stiffnesses A16 , A26 , D16 , D26 of a symmetric regular angle-ply laminate vanishing? In what conditions are the stiffnesses
B16 , B26 of an antisymmetric regular angle-ply laminate vanishing?
P3.45 Using the notations introduced in the Section 3.5 shows that
U Nn = U1 Nn1 + U2 Nn2 = Un Nnn + U Nn .
P3.46 Show that the specific strain energy w of a composite laminate can be
expressed in the following matrix form:
w=

1 T
[d] [E] [d] ,
2

where

[d] = [e11 , e22 , 2e12 , k11 , k21 , 2k12 ] .

P3.47 Using the work theorem for composite laminates and assuming the positive definiteness of the specific elastic energy w, formulate and prove the uniqueness
theorem corresponding to the classical plate theory of composite laminates. More
exactly, prove that if Un or Nnn , U or Mn , U3 or Qn + Mn, and U3,n or Mnn

Copyright 2004 by Chapman & Hall/CRC

Downloaded by [Hong Kong University of Science and Technology] at 01:44 17 November 2014

188

CHAPTER 3. COMPOSITE LAMINATES

are vanishing on the boundary D of the plane domain D occupied by the middle
surface of the laminate, then e and k are vanishing on D, assuming usual regularity conditions. Hence, the solution is vanishing, modulo a rigid displacement
of the plate. Give conditions in which this displacement is also vanishing!
P3.48 Formulate and prove a variational principle corresponding to the simply
supported edge boundary conditions S1.
P3.49 Formulate and prove a variational principle corresponding to the clamped edge boundary conditions C4.
P3.50 Find the appropriate potential energy and prove its minimum property
for a simply supported laminate, submitted to the boundary conditions S2.
P3.51 Find the appropriate potential energy and prove its minimum property
for a clamped laminate, submitted to the boundary conditions C2.
P3.52 Using the corresponding principles of minimum potential energy prove
the uniqueness theorems corresponding to the boundary value problems S2 and
C2, respectively.
P3.53 Formulate and prove the converses of the principles of minimum potential energy corresponding to the boundary value problems S2 and C2, respectively.

Bibliography
[3.1] Ashton J.E., Whitney J.M., Theory of laminated plates, Progress in Material Science Series, Vol. IV, Technomic Publishing Co., Stanford, 1970.
[3.2] Jones, R.M., Mechanics of composite materials, Hemisphere Publishing Co.,
New York, 1975.
[3.3] Christensen, R.M., Mechanics of composite materials, John Wiley and Sons,
1979.
[3.4] Tsai, W., Hahn, M.T., Introduction to composite materials, Technomic Publishing Co., Westport, Conneticut, 1980.
[3.5] Cristescu, N., Mechanics of composite materials, University of Bucharest,
Bucharest, 1983 (in Romanian).
[3.6] Whitney, J.M., Analysis of laminated anisotropic plates, Technomic, Lancaster, PA, 1987.
[3.7] Gibson, R.F., Principles of composite material mechanics, McGraw-Hill Inc.,
New York, 1994.
[3.8] Lekhnitski, S.G., Theory of elasticity of an anisotropic elastic body, HoldenDay, San Francisco, 1963.

[3.9] Barran, J.J., Laroze, S., Calcul des structures en materiaux composite, Ecole
National Superieur de lAeronautique et de lespace, Dept. structures, materiaux, technologie, 1987.

Copyright 2004 by Chapman & Hall/CRC

Vous aimerez peut-être aussi