Você está na página 1de 86

N OTES ON

S TRUCTURAL A NALYSIS OF
C OMPOSITE S HELL S TRUCTURES

Jan Stegmann and Erik Lund


2001

Institute of

Mechanical Engineering
Aalborg University

ii

Aalborg University

Institute of Mechanical Engineering

Pontoppidanstraede 101 DK-9220 Aalborg East Denmark

Phone +45 96 35 80 80

Notes on
Structural Analysis of Composite Shell Structures

By
Jan Stegmann? and Erik Lund

Ph.D. Student  js@ime.auc.dk


Assoc. Professor, Ph.D.  el@ime.auc.dk
?

First Edition, September 2001

iv

Contents
1 Introduction

Introduction to Mechanics of Composite Materials

2.1

What is a composite material? . . . . . . . . . . . . . . . . . . . . . . . . . .

2.2

Background on materials selection . . . . . . . . . . . . . . . . . . . . . . . .

2.2.1

Material selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2.3

The composite lamina . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

14

2.4

Micromechanical behavior of a lamina . . . . . . . . . . . . . . . . . . . . . .

18

2.5

Laminate theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

20

2.5.1

Classical laminated plate theory . . . . . . . . . . . . . . . . . . . . .

20

2.5.2

ReissnerMindlin plate theory . . . . . . . . . . . . . . . . . . . . . .

26

2.5.3

Higher order plate theories . . . . . . . . . . . . . . . . . . . . . . . .

29

2.5.4

Concluding remarks on plate theories . . . . . . . . . . . . . . . . . .

30

2.6

Sandwich structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

31

2.7

Glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

32

Continuum-Based Shell Elements

35

3.1

Governing equations Equations of motion . . . . . . . . . . . . . . . . . . .

35

3.1.1

Discrete governing equations Finite element formulation . . . . . . .

37

3.2

The basic shell concept . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

40

3.3

Element formulation be degeneration of continuum element . . . . . . . . . . .

41

3.3.1

First order shear deformation theory a quick review . . . . . . . . . .

42

Geometric and kinematic representation . . . . . . . . . . . . . . . . . . . . .

42

3.4

CONTENTS

vi

3.5

Local coordinate systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

46

3.5.1

Element coordinate system (ECS) . . . . . . . . . . . . . . . . . . . .

46

3.5.2

Material coordinate system . . . . . . . . . . . . . . . . . . . . . . . .

47

3.5.3

User-defined element coordinate system . . . . . . . . . . . . . . . . .

48

Linear element matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

50

3.6.1

Interpolation matrix . . . . . . . . . . . . . . . . . . . . . . . . . . .

50

3.6.2

Strain-displacement matrix . . . . . . . . . . . . . . . . . . . . . . . .

50

Constitutive relation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

52

3.7.1

Single layers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

53

3.7.2

Modelling materials using constitutive relations . . . . . . . . . . . . .

56

3.7.3

Multiple layers Composite laminates and sandwich structures . . . .

56

3.7.4

Other properties of the constitutive relation . . . . . . . . . . . . . . .

58

3.7.5

Transformation of the constitutive matrix . . . . . . . . . . . . . . . .

60

Numerical integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

61

3.8.1

Numerical thickness integration of laminates . . . . . . . . . . . . . .

61

Locking problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

63

3.9.1

Membrane locking . . . . . . . . . . . . . . . . . . . . . . . . . . . .

64

3.9.2

Shear locking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

64

3.9.3

Volumetric and thickness locking . . . . . . . . . . . . . . . . . . . .

65

3.10 Mixed Interpolation of Tensorial Components MITC . . . . . . . . . . . . .

66

3.6

3.7

3.8

3.9

4 Geometrical Non-Linear Analysis of Laminated Shell Structures


4.1

4.2

69

Governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

69

4.1.1

Deformation gradient . . . . . . . . . . . . . . . . . . . . . . . . . . .

69

4.1.2

Governing equations for non-linear analysis . . . . . . . . . . . . . . .

72

Non-linear element matrices . . . . . . . . . . . . . . . . . . . . . . . . . . .

73

4.2.1

75

Tangent stiffness matrix . . . . . . . . . . . . . . . . . . . . . . . . .

Chapter

Introduction

hese notes have been elaborated for the lecture on Structural Analysis of Composite Shell
structures held in connection with the graduate course on Aerodynamics and Structural
Dynamics of Wind Turbines. The course is addressed at 9th semester graduate students at
The Department of Building Technology and Structural Engineering, The Department of Civil
Engineering and The Institute of Mechanical Engineering.

Chapters 3 and 4 of the notes have to a large extend been based on the Masters Thesis by Lars
R. Jensen, Jens M. Rauhe & Jan Stegmann (Stegmann et al., 2001) but have been revised by the
authors to better fulfill the requirements of the course. These notes, example files and necessary
software as well as the full Masters Thesis (Stegmann et al., 2001) are available on-line from
www.ime.auc.dk/js.
Useful literature on the basic concepts of finite element theory can be found in the notes for the
FEM3 course, available at www.ime.auc.dk/el. Further study of numerical methods in engineering, particularly finite element methods, can be made through a number of articles and textbooks. Some of the major contributions over the last years have been made by e.g. Bathe (1996),
Zienkiewicz and Taylor (2000a), Hughes (1987) and Belytschko et al. (2000). Other recent contributions include the work of e.g. McNeal (1998) and Bletzinger et al. (2000). The use of finite
element methods for composite laminates and sandwich structures has been covered extensively
by e.g. Ochoa and Reddy (1992), Klinkel et al. (1999) and Alfano et al. (2001) but numerous
others have published on the subject as well.
This part of the course consists of three lectures and these notes are divided into three chapters,
each pertaining to one of the three lectures. The content is as follows:
Chapter 2 introduces the concepts of laminated composite structures. First, the lamina is
treated and the basic concepts of single plies are introduced. Subsequently, these concepts
are extended to laminates, consisting of multiple plies. Both composite and sandwich
laminates will be addressed.
Chapter 3 deals with the formulation of shell finite elements for the analysis of composite
structures. Geometric representation, kinematics, governing equations, element matrices
and constitutive relations will all be introduced. Numerical solution and problems regarding the solution will be addressed as well.
Chapter 4 extends the formulation derived in Chapter 3 to geometrically non-linear
analysis. This includes derivation of the governing equation and element matrices.

1. Introduction

Some of the concepts used in shell analysis are new to most, and the student is therefore advised
to make him or herself familiar with the nomenclature and terminology before each lecture.
At the end of lecture 3 exercises will be performed using the MUltidisciplinary Synthesis Tool
(MUST) and FEPlot, which are both available for download at www.ime.auc.dk/js. MUST
is a finite element based system for design, analysis and optimization of structural, fluid and
thermal problems. The system has been developed over the last three years by The FluidStructure Interaction Group at The Institute of Mechanical Engineering and the development is
set to continue in the future. MUST is not a stand-alone application but relies on commercial
software typically COSMOS/M for meshing and general preprocessing. FEPlot is a postprocessing program for visualization of results.
Both applications should be downloaded before-hand and preferably familiarized by the student using the simple example available online. This greatly increases the benefit of doing the
exercises. Both an installation instruction and a short users guide are available for download.

Chapter

Introduction to Mechanics of Composite Materials

he purpose of this chapter is to give a short introduction to the concepts for structural design
of a wind turbine blade, see Fig. 2.1, and to give a basic background to the mechanical
behavior of the composite structures used. A glossary of frequently used terms for describing
composite materials can be found at the end of the chapter in Section 2.7.

Beam
Aerodynamic shell

Figure 2.1 A wind turbine blade (Ris, 2001)

The outer contour of a wind turbine blade is prescribed from aerodynamical considerations and
the mechanical objective of the blade design is to render it as light as possible while maintaining
sufficient stiffness and strength.
To accommodate this the designer may vary both structural composition and material usage.
As a consequence, the structural composition is a shell stiffened by a beam-like structure in
the flapwise direction (see Fig. 2.1). The materials used for the blade should have favorable
stiffness, weight, strength, and fatigue properties as well as low material price and manufacturing cost. These demands are hard to meet but composite materials seem to be the obvious
choice, and it is the purpose of this chapter to give a basic description of these materials and
their mechanical behavior.
The use of composite materials in wind turbine blades is illustrated in Fig. 2.2 which shows a
cross-section of a wind turbine blade manufactured by L.M. Glasfiber (printed courtesy of L.M.
Glasfiber).

2. Introduction to Mechanics of Composite Materials

Figure 2.2 Cross section of a wind turbine blade

This represents a typical structural composition of wind turbine blades and can be investigated
by regarding the schematic representation in Fig. 2.3.
Glued joint
Main laminate at the upper part

Main laminate
at the front edge

Sandwich structures at lower part

Main laminate at
the rear edge

Glued joint

Figure 2.3 Elements of a typical cross section of a wind turbine blade (Ris, 2001)

As illustrated the blade is constructed with zones of different composite laminates and in some
areas the laminate is stiffened by adding a sandwich core (the shaded areas in Fig. 2.3).
As can be seen from both Fig. 2.1 and Fig. 2.2 the beam-like structure is made up of gluedtogether webs which form a box-like structure (hollow beam) in the flapwise direction. A
reasonably good representation of the stiffness of the blade can therefore be obtained by a simple
beam model, but it is the goal of these notes to describe more detailed models for analysis of
wind turbine blades; consequently, simplified beam models will not be covered.
The figures of the cross sections introduce the basic structural elements a laminate and a sandwich structure which will be described in detail in this chapter. However, let us start by stating
what makes a material a composite material.

2.1

What is a composite material?

The answer can be stated in short through the following definition from Hansen (1996):
A composite material can be defined as a material that is composed of two or more
distinct phases, usually a reinforcing material supported in a compatible matrix, assembled in prescribed amounts to achieve specific physical and chemical properties.

Notes on Structural Analysis of Composite Shell Structures

In order to classify and characterize composite materials distinction between the following four
types is commonly accepted (Jones, 1999):
1. Fibrous composite materials that consist of fibers in a matrix, e.g.
quasi-isotropic random reinforced materials: chopped fibers in a matrix
orthotropic aligned reinforced materials: stiffeners, wires, fibers, fabrics in a matrix
2. Laminated composite materials that consist of layers of various materials, e.g.
laminated glass, bimetals, clad metals
3. Particulate composite materials that are composed of particles in a matrix, e.g.
quasi-isotropic random reinforced materials: powders or particles in a matrix like
ceramics
4. Combinations of some or all of the first three types, e.g.
laminated fiber-reinforced materials: orthotropic laminae bonded together to form
an anisotropic material
sandwich constructions: face sheets bonded to a lightweight core
In this course the focus will not be on one of the four types alone but on a hybrid class, namely
laminated fiber-reinforced composite materials as they are the basic building element of a wind
turbine blade. These materials are hybrids between type 1 and type 2 since they consist of fibrous composite materials (type 1) but involve lamination techniques (type 2) in their structural
composition. Typically, such materials consist of stacks of bonded-together layers made from
fiber-reinforced material. The layers will often be oriented in different directions to provide
specific and directed strengths and stiffnesses of the laminate. Thus, the strength and stiffnesses
of the laminated fiber-reinforced composite material can be tailored to the specific design requirements of the structural element being built.
Composite materials have many mechanical behavioral characteristics which are different from
those of more conventional engineering materials such as metals. More precisely, composite materials are often both inhomogeneous and non-isotropic (orthotropic or, more generally,
anisotropic). These are key concepts:
An inhomogenous body has nonuniform material properties over the body, i.e., the properties depend on position in the body
An orthotropic body has material properties that are different in three mutually perpendicular directions at a point in the body and, further, has three mutually perpendicular
planes of material property symmetry. Thus, the properties depend on orientation at a
point in the body.
An anisotropic body has material properties that are different in all directions at a point
in the body. No planes of material property symmetry exist. Again, the properties depend
on orientation at a point in the body.

2. Introduction to Mechanics of Composite Materials

These properties of composite materials significantly alter their response to loading compared
to isotropic materials. This is illustrated in Fig. 2.4.

(a)

isotropic

orthotropic

anisotropic

(b)

Figure 2.4 Mechanical behavior of various materials

Because of the inherent heterogeneous nature of composite materials, they can be studied from
a micromechanical or a macromechanical point of view. In this course a macromechanical
point of view is taken, where the material is presumed homogeneous and the effects of the
constituent materials are detected only as averaged apparent macroscopic properties of the
composite material. This approach is generally accepted when modelling gross response of
composite structures.

2.2

Background on materials selection

Before we proceed with a more throughout description of composite materials, let us justify the
use of composite materials in wind turbine blades.
Let us start by asking a simple question. Why do engineers often use aluminum and not steel for
light weight structures like aeroplanes? To answer the question we might consider the properties
of the materials, stated in Table 2.1.
Material
Youngs Modules, E [MPa] Density, [kg/m3 ]
Steel
210 103
7800
3
Aluminum
70 10
2700
Table 2.1 Material properties for steel and aluminum

If we use the ratio of stiffness to weight, E/, as a measure of a materials "suitability" in light
weight design we see that steel and aluminum are almost identical with values of 27 and 26
respectively. Consequently, a thinner structure of steel would show the same structural stiffness
as a thicker structure of aluminum and be equally suitable for light weight design. Or would it?
To better quantify our answer we adopt the excellent description by Ashby (1999) on materials
selection in mechanical design.
2.2.1 Material selection
Any engineering component has one or more functions: to support a load, to contain a pressure,
to transmit heat and so forth. In designing a component, the designer has an objective: to make

Notes on Structural Analysis of Composite Shell Structures

it as cheap as possible, perhaps, or as light, or as safe or perhaps some combination of these,


as in the case of design of wind turbine blades. This objective must be achieved subject to
constraints: that certain parameters are fixed, that the component must carry the given load or
pressure without failure, that it can function in a given range of temperature, and in a given
environment and many more. Examples of objectives and constraints can be seen in Fig. 2.5,
where a material index M is found.

Figure 2.5 The specification of function, objective and constraint leads to a material index M . The
combination in the highlighted boxes leads to the index E 1/2 / (Ashby, 1999)

As an example let us find the material index M to consider in the case of finding the lightest
beam subject to a constraint on its stiffness, see Fig. 2.6.

Figure 2.6 A beam of square section, loaded in bending. Its stiffness is S = F/ where F is the load
and is the deflection (Ashby, 1999)

A beam of cross section b b and length l is considered. The loading F is given, and due to the
constraint on stiffness S, the deflection must not exceed a certain value, i.e., this constraint
requires that S = F/ be greater than some critical value:
S=

C1 EI
F

l3

(2.1)

2. Introduction to Mechanics of Composite Materials

where E is Youngs Modulus, C1 is a constant depending on the distribution of load and I is


the moment of inertia. In the case of a square cross section of the beam, I is given as:
I=

b4
A2
=
12
12

(2.2)

We would like to find the lightest beam when the cross section A = b b is free to vary. We
express the mass as:
m=Al
(2.3)
Combining Equations (2.1), (2.2) and (2.3), the free cross section A can be eliminated and we
obtain:


 
12S
(2.4)
l3
m
C1 l
E 1/2

The brackets in (2.4) represent functional requirements, geometry, and material, respectively.
Thus, the lightest beam is found by selecting the material from the maximum value of the
material index M given by:
E 1/2
M=
(2.5)

In the above case the free cross section A remained square while varied, but in many cases
either the height or the width are allowed to vary freely.
If only the height is free, the material index is found as:
M=

E 1/3

(2.6)

and if only the width is free, the material index is found as:
M=

(2.7)

Depending on the design variable, the design of the lightest beam with a stiffness constraint
thus depends on the relative stiffness E/, E 1/2 /, or E 1/3 / . This explains why aluminum
is often used for lightweight structures instead of steel even though the ratio E/ is equal for
the two materials. When loaded in bending (as in this example) and the height of the beam is
the design variable there is a significant difference in material index M = E 1/3 / , favoring
aluminum, between the two materials, as illustrated in Table 2.2.
Material
Youngs Modules, E [MPa] Density, [kg/m3 ]
Steel
210 103
7800
3
Aluminum
70 10
2700

E/
27
26

E 1/2 / E 1/3 /
1.9
0.7
3.1
1.5

Table 2.2 Stiffness comparisons of steel and aluminum

Table 2.3 shows the material indices to maximize in many different stiffness-limited situations
(Ashby, 1999):

Notes on Structural Analysis of Composite Shell Structures

Functions and constraints

Maximize

Tie (tensile strut)


stiffness, length specified; section area free

E/

Shaft (loaded in torsion)


stiffness, length, shape specified, section area free
stiffness, length, outer radius specified; wall thickness free
stiffness, length, wall-thickness specified; outer radius free

G1/2 /
G/
G1/3 /

Beam (loaded in bending)


stiffness, length, shape specified; section area free
stiffness, length, height specified; width free
stiffness, length, width specified; height free

E 1/2 /
E/
E 1/3 /

Column (compression strut, failure by elastic buckling)


buckling load, length, shape specified; section area free

E 1/2 /

Panel (flat plate, loaded in bending)


stiffness, length, width specified, thickness free

E 1/3 /

Plate (flat plate, compressed in-plane, buckling failure)


collapse load, length and width specified, thickness free

E 1/3 /

Cylinder with internal pressure


elastic distortion, pressure and radius specified; wall thickness free

E/

Spherical shell with internal pressure


elastic distortion, pressure and radius specified, wall thickness free

E/((1 ))

To minimize cost, use the above criteria for minimum weight, replacing density by C m , where Cm is the
material cost per kg. To minimize energy content, use the above criteria for minimum weight replacing density
by q where q is the energy content per kg. To minimize environmental impact, replace density by I e instead,
where Ie , is the eco-indicator value for the material.

E = Youngs modulus for tension, the flexural modulus for bending or buckling; G = shear modulus; =
density, q = energy content/kg; Ie = eco-indicator value/kg.

Table 2.3 Material indices to maximize in stiffness-limited design at minimum mass (cost, energy, environmental impact ), from Ashby (1999).

The following figure is a material selection chart in case of light, stiff structures. The guide
lines show the loci of points for which
1. E/ = C (minimum weight design of stiff ties, minimum deflection in centrifugal
loading, etc.)
2. E 1/2 / = C (minimum weight design of stiff beams, shafts and columns)
3. E 1/3 / = C (minimum weight design of stiff plates)

10

2. Introduction to Mechanics of Composite Materials

The value of the constant C increases as the lines are displaced upwards and to the left. Materials offering the largest stiffness-to-weight ratio lie towards the upper left corner.

Figure 2.7 Youngs modulus, E, against density (Ashby, 1999).

Fig. 2.7 illustrates why composite materials are so well suited for light weight structures. They
can be tailored to have high stiffness in critical directions and low stiffness in others, thus reducing the overall weight while maintaining structural stiffness, resulting in a very favorable
E p / -ratio.
Another important material property to consider is strength, and Table 2.4 shows the material
indices to maximize in many different strength-limited situations (Ashby, 1999):

Notes on Structural Analysis of Composite Shell Structures

11

Functions and constraints

Maximize

Tie (tensile strut)


stiffness, length specified; section area free

f /

Shaft (loaded in torsion)


load, length, shape specified, section area free
load, length, outer radius specified; wall thickness free
load, length, wall-thickness specified; outer radius free

f /
f /
1/2
f /

Beam (loaded in bending)


load, length, shape specified; section area free
load, length, height specified; width free
load, length, width specified; height free

f /
f /
1/2
f /

Column (compression strut)


buckling load, length, shape specified; section area free

f /

Panel (flat plate, loaded in bending)


stiffness, length, width specified, thickness free

f /

Plate (flat plate, compressed in-plane, buckling failure)


collapse load, length and width specified, thickness free

f /

Cylinder with internal pressure


elastic distortion, pressure and radius specified; wall thickness free

f /

Spherical shell with internal pressure


elastic distortion, pressure and radius specified, wall thickness free

f /

Flywheels, rotating discs


maximum energy storage per unit volume; given velocity
maximum energy storage per unit mass; no failure

f /

2/3

2/3

1/2

1/2

To minimize cost, use the above criteria for minimum weight, replacing density by C m , where Cm is the
material cost per kg. To minimize energy content, use the above criteria for minimum weight replacing density
by q where q is the energy content per kg. To minimize environmental impact, replace density by I e instead,
where Ie , is the eco-indicator value for the material.

f = failure strength (the yield strength for metals and ductile polymers, the tensile strength for ceramics,
glasses and brittle polymers loaded in tension; the flexural strength or modulus of rupture for materials loaded in
bending); = density.

For design for infinite fatigue life, replace f by the endurance limit e .

Table 2.4 Material indices to maximize in strength-limited design at minimum mass (cost, energy, environmental impact ), from Ashby (1999).

2. Introduction to Mechanics of Composite Materials

12

The following figure is a material selection chart in case of light, strong structures. The guide
lines show the loci of points for which
1. f / = C (minimum weight design of strong ties, minimum rotational velocity of discs,
etc.)
2/3

2. f / = C (minimum weight design of strong beams and shafts)


1/2

3. f / = C (minimum weight design of strong plates)


The value of the constant C increases as the lines are displaced upwards and to the left. Materials offering the largest strength-to-weight ratio lie towards the upper left corner.

Figure 2.8 Strength, f , against density (Ashby, 1999).

Fig. 2.9 compares various forms of specific composite materials with structural-grade metals.

Notes on Structural Analysis of Composite Shell Structures

13

Figure 2.9 Strength and stiffness of composite materials and metals (Jones, 1999)

The above considerations, among others, lead to the following reasons for selecting polymer
matrix composites like GFRP (glass fiber reinforced plastic) and CFRP (carbon fiber reinforced
plastic) as the main materials for wind turbine blades:
high specific stiffness E/, E 1/2 /, E 1/3 /
2/3

1/2

high specific strength f /, f /, f /


low weight
good fatigue properties
good corrosion resistance
attractive surface finish
competitive material and manufacturing costs
low cost design optimization
directionally tailored properties
aeroelastic tailoring
Thus, having justified the use of composite materials in engineering design we proceed by
describing a composite lamina.

2. Introduction to Mechanics of Composite Materials

14

2.3

The composite lamina

The basic building block of a laminated fiber-reinforced composite material is the composite
lamina. A lamina is a flat (or sometimes curved as in a shell) arrangement of unidirectional or
woven fibers in a supporting matrix. The following demands must be made on fibers used in
reinforced composite materials (Hansen, 1996):
high tensile strength
high modulus of elasticity
lower ultimate elongation than the matrix
good adhesion to the matrix
good resistance to the matrix and its additives
The influence of the matrix on the composite is as follows (Hansen, 1996):
binds the reinforcement and distributes the load
protects the fibers from physical and chemical damage
dominant factor in determining transverse shear and throughthickness properties
dominant factor in determining impact resistance and fracture toughness
dominant factor in determining long time (creep) response
dominant factor in determining service temperature
The purpose of this section is to provide a basic understanding of the macromechanical behaviour of a lamina when averaged apparent mechanical properties are considered. The rest of
this chapter is based on Hansen (1996) and Jones (1999).

Figure 2.10 Basic questions of lamina macromechanics (Jones, 1999)

The basic questions of lamina macromechanics as illustrated by Fig. 2.10 are: 1) what are the
characteristics of a lamina and 2) how does a lamina respond to loading?

Notes on Structural Analysis of Composite Shell Structures

15

The materials are assumed to behave linearly elastic, i.e. the generalized Hookes law is used
for relating stresses to strains. A material coordinate system 1 2 and a Cartesian coordinate
system x y are introduced for the unidirectionally reinforced lamina, see Fig. 2.11.

Figure 2.11 Positive rotation of principal material axes 1-2 from x-y axes (Jones, 1999)

The notation used in the following is described by Table 2.5. The general anisotropic constituTable 2.5 Tensor and contracted notation for stresses and strains

Stresses

Strains

Tensor
notation

Contracted
notation

Tensor
notation

Contracted
notation

11
22
33
23 = 23
31 = 31
12 = 12

1
2
3
4
5
6

11
22
33
223 = 23
231 = 31
212 = 12

1
2
3
4
5
6

tive relation is given by Hookes law:


1
C11

2
C12


C13
3
=
C14

23

C15

31

12
C16

C12
C22
C23
C24
C25
C26

C13
C23
C33
C34
C35
C36

C14
C24
C34
C44
C45
C46

C15
C25
C35
C45
C55
C56

C16
C26
C36
C46
C56
C66

1
2
3
23
31
12

i.e., twenty-one independent material constants are used to describe the material.

(2.8)

2. Introduction to Mechanics of Composite Materials

16

For the composite lamina illustrated in Fig. 2.11, there are two orthogonal planes of material
property symmetry and the material is termed orthotropic. The stress-strain relations in coordinates aligned with principal material directions are given as

C
C
C
0
0
0


11
12
13
1
1

C
C
C
0
0
0


12
22
23
2
2

3
C
C
C
0
0
0

13
23
33
3

(2.9)
=
0
0 C44 0
0
23
23



0
0
0 C55 0

31
31

12
12
0
0
0
0
0 C66
i.e., nine independent material constants characterize the material.

If a plane stress assumption is used for the orthotropic composite lamina, the stress-strain relations are given as

Q11 Q12 0
1
1
2
2
= Q12 Q22 0

12
0
0 Q66
12
where

Q11
Q22
Q12
Q66

= E11 /(1 12 21 ),
= E22 /(1 12 21 ),
= 12 E22 /(1 12 21 ) = 21 E11 /(1 12 21 ),
= G12

(2.10)

The number of independent material constants is now reduced to four.


We would like to express the stress-strain relations for the lamina of arbitrary orientation as
illustrated in Fig. 2.11, and for this please recall the transformation equations for expressing
stresses in a x y coordinate system in terms of stresses in a 1 2 coordinate system:

x
1
y
2
= T1
(2.11)

xy
12

where

cos2
sin2
2 sin cos
T=
sin2
cos2
2 sin cos
sin cos sin cos cos2 sin2

(2.12)

In the transformation of strains it is, of course, important to realize that the above stress transformation is only valid for a second order tensor, i.e., engineering strain vectors cannot be used
immediately. However, introducing the Reuter matrix:

1 0 0
R= 0 1 0
(2.13)
0 0 2
the transformation of strains can be written in compact form using both tensor and engineering
notation for the strain tensor:

x
1
x
1
y
2
y
2
= T1
,
= RT1 R1
(2.14)
1

xy
12
2 xy
2 12

Notes on Structural Analysis of Composite Shell Structures

17

Using the above transformations, the stress-strain relations for arbitrary lamina orientation can
be written as:

Q11 Q12 Q16 x


x
y
y
= Q12 Q22 Q26

xy
xy
(2.15)
Q16 Q26 Q66
where

Q = T1 QRTR1 = T1 QTT

A specially orthotropic composite lamina is one


aligned with the structural axes. For example:


Q11 Q12 Q16
Q12 Q22 Q26 =
Q16 Q26 Q66

for which the principal material axes are

Q11 Q12 0
Q12 Q22 0
0
0 Q66

(2.16)

Finally, a generally orthotropic composite lamina is an orthotropic lamina in which the principal
material axes are not aligned with the structural axes. Thus the Q matrix is uniquely defined
by four material properties even though the matrix is apparently that of an anisotropic material.
That is:

Q11 Q12 Q16


Q11 Q12 0
Q12 Q22 Q26 = T1 Q12 Q22 0 TT
(2.17)
0
0 Q66
Q16 Q26 Q66
Graphite/Epoxy (3M SP-2 88/T300)
22
20

Q11

18

Q22

16

(psi 10 )

14
12
10
8
6

Q12

4
2
0

20

40

60

80

100

120

140

160

180

Lamina Orientation (deg)

Figure 2.12 Results for graphite epoxy (Hansen, 1996)

These stress-strain relations for a composite lamina of arbitrary orientation are best illustrated
by figures, and as examples please consider the components of the matrix Q for a graphite epoxy
in Fig. 2.12 and Fig. 2.13.

2. Introduction to Mechanics of Composite Materials

18

Graphite/Epoxy (3M SP-2 88/T300)


8

Q66

Q16

(psi 10 )

-2

-4

Q26
-6

-8

20

40

60

80

100

120

140

160

180

Lamina Orientation (deg)

Figure 2.13 Results for graphite epoxy - continued (Hansen, 1996)

From Fig. 2.12 and Fig. 2.13 it is obvious that these directional dependencies of stiffness (and
strength) are vital properties when designing with composite materials.

2.4

Micromechanical behavior of a lamina

Throughout these notes a macromechanical viewpoint is taken, i.e., we consider apparent averaged properties of a lamina. However, in order to obtain the apparent averaged material
properties, micromechanical studies must be performed, i.e., the basic questions regarding stiffness and strength illustrated by Fig. 2.14 must be answered.

Figure 2.14 Basic questions of micromechanics (Jones, 1999)

Most of the established work in this area describes the stiffness properties of the lamina whereas

Notes on Structural Analysis of Composite Shell Structures

19

not much work is available regarding micromechanical theories of strength. A short overview
of theories of micromechanical behaviour of a lamina can be found in Jones (1999).
In order to illustrate the mechanics of materials approach to stiffness of a lamina, consider the
representative volume element of a lamina loaded in the 1-direction at Fig. 2.15.

Figure 2.15 Representative volume element loaded in the 1-direction (Jones, 1999)

A basic assumption is that the strains in the fiber direction of a unidirectional fiber-reinforced
composite material are the same in the fibers as in the matrix. Using equilibrium considerations
on the representative element in Fig. 2.15 lead to the rule of mixture for the apparent Youngs
modulus E1 of the composite material in the direction of the fibers:
E1 = Ef Vf + Em Vm = Ef Vf + Em (1 Vf )

(2.18)

Here Ef , Em represent Youngs modulus of the fiber and matrix materials, respectively, and
Vf , Vm represent volume fractions of fibers and matrix. This predicted stiffness is in excellent
agreement with measured stiffnesses, see Fig. 2.16.

Figure 2.16 Predicted versus measured E1 stiffness of unidirectional boron-epoxy composite material
(Jones, 1999)

The rule of mixture approach is used to predict stiffnesses of a given lamina, whereas theories for
strength of a lamina will not be described in these notes. The reader is referred to Jones (1999)
for further details.

2. Introduction to Mechanics of Composite Materials

20

2.5

Laminate theory

A laminate is a series of laminae bonded together to act as an integral structural element. Thus,
a laminate is not a material but instead a structural element with essentiel features of both
material properties and geometry. The stiffnesses and strengths of such a composite material
structural configuration are obtained from the properties of the constituent laminae, and the
macromechanical behaviour of a laminate is the main topic of this section. The lamination
theory described can be considered as a single layer rule of mixtures representation of the
interaction between the multiple laminae in a plate or shell.

Figure 2.17 The basic questions of laminate analysis (Jones, 1999)

As illustrated in Fig. 2.17 the individual laminae may have arbitrary orientations, and the major purpose of lamination is to tailor the directional dependence of stiffness and strength of a
composite material to match the loading environment of the structural element. The layers of
the laminate are usually bonded together by the same matrix material that is used in the individual laminae, i.e., in the following it is assumed that each layer is a fiber-reinforced composite
lamina.
2.5.1 Classical laminated plate theory
Classical laminated plate theory is often also called Classical Lamination Theory (CLT) and can
be regarded as a process of finding effective and reasonably accurate simplifying assumptions
that reduces the three-dimensional elasticity problem to a solvable two-dimensional problem.

Figure 2.18 Geometry of deformation in the x-z plane (Jones, 1999)

Notes on Structural Analysis of Composite Shell Structures

21

The assumptions of classical laminated plate theory are based on the Kirchhoff-Love hypothesis
for plates and shells:
1. The plate is thin. That is the thickness t is small compared to the other physical dimensions.
2. The displacements u(x, y, z), v(x, y, z) and w(x, y, z) are small compared to the plate
thickness.
3. The inplane strains x , y and xy are small compared to unity.
4. The transverse normal strain z is negligible.
5. The transverse shear stresses xz , yz are negligible.

Figure 2.19 Symmetric angle-ply laminate geometry and stresses (Jones, 1999)

The strain-displacement relations in case of linear elasticity are given as:


x =
xy

u
,
x

y =

u v
=
+
,
y x

xz

v
,
y

u w
=
+
,
z
x

z =
yz

w
z

(2.19)

v w
=
+
z
y

The consequences of the assumptions of classical plate theory are:


Assumption #4 z =

w
= 0 which in turn implies that w(x, y, z) = w(x, y)
z

Assumption #5 xz =
z

w(x, y)
x

u w
+
= 0 which in turn implies that u(x, y, z) = u(x, y)
z x

2. Introduction to Mechanics of Composite Materials

22

Similarly assumption #5 v(x, y, z) = v(x, y) z

w(x, y)
y

The plate straindisplacement relations may thereby be written as


2w
u
z 2
x
x
v
2w
=
z 2
y
y
2w
u v
+
2z
=
y x
xy

x =
y
xy

(2.20)

Introducing middlesurface strains x , y , xy


and middlesurface curvatures x , y , xy , the
plate straindisplacement relations may also be expressed as:


x
x
x

y
y
+ z
(2.21)
=

y
xy
xy
xy

where the middlesurface strains and middlesurface curvatures are defined as

2w
u

x
x

x
2
v
w
y
=
,
y
=
y
y 2

u v

2w

2
+
xy
xy
y x
xy

(2.22)

Inserting the strain variation through the thickness given by (2.21) into the stress-strain relation
given by (2.15), the stresses in the k th layer can be expressed in terms of the laminate middlesurface strains and curvatures as:


Q11 Q12 Q16
x
x
x


y
y
= Q12 Q22 Q26
+ z
(2.23)

xy k
Q16 Q26 Q66 k
xy
xy
Next, resultant forces and moments are introduced. Resultant forces can be found by regarding
Fig. 2.20.

Figure 2.20 In-plane forces on a flat laminate (Jones, 1999)

The force resultants acting on the laminate are obtained by integration of the stresses in each
layer or lamina through the laminate thickness, i.e. the force resultants are obtained as:
Nx =

Zt/2

t/2

x dz ,

Ny =

Zt/2

t/2

y dz ,

Nxy =

Zt/2

t/2

xy dz

(2.24)

Notes on Structural Analysis of Composite Shell Structures

23

The moment resultants are obtained in a similar manner by regarding Fig. 2.21.

Figure 2.21 Moments on a flat laminate (Jones, 1999)

Integration over the thickness yields the moment resultants as:


Mx =

Zt/2

t/2

x zdz ,

My =

Zt/2

t/2

y zdz ,

Mxy =

Zt/2

xy zdz

(2.25)

t/2

It should be noted that the force resultants represent force per unit width of the cross section of
the laminate. Similarly, the moments are per unit width.

Figure 2.22 Geometry of an N-layered laminate (Jones, 1999)

The force and moment resultants can be written in matrix form as:

Zt/2
N
1

=
{} dz

M
z
t/2

zk
Z
Zt/2 X
N
N
1
1

X
{}k
dz =
{}k
dz
=

k=1
k=1 z
z
z
t/2
k1

(2.26)

Substituting for stresses in terms of strains yields

zk
Zzk
Z
N
N
1
N X
1

=
[Q]k {} dz =
[Q]k { + z} dz

k=1 z
k=1 z
M
z
z
k1
k1

(2.27)

Here k = 1 corresponds to the bottom ply and the numbering proceeds to k = N at the top ply.
zk and zk1 are the upper and lower z coordinate of the k th ply, see illustration of geometry in
Fig. 2.22.

2. Introduction to Mechanics of Composite Materials

24

After integration through the thickness this becomes:

A | B 
N

= +

B | D

(2.28)

where

Aij =
Bij = 1/2

N
X

k=1
N
X
k=1

Dij = 1/3

N
X
k=1

(Qij )k (zk zk1 ) ,

i, j = 1, 2, 6

(2.29)

2
(Qij )k (zk2 zk1
),

i, j = 1, 2, 6

(2.30)

3
(Qij )k (zk3 zk1
),

i, j = 1, 2, 6

(2.31)

The physical significance of the stiffness terms in (2.28) can be seen in Fig. 2.23.

Figure 2.23 Physical significance of stiffness terms in force and moment resultants (Jones, 1999)

The bending-extension coupling is illustrated by an unsymmetric laminate after curing in Fig. 2.24.
The laminate was flat before curing but the thermally induced residual stresses cause the laminate to become highly curved.

Figure 2.24 An unsymmetric laminate after curing (Jones, 1999)

The effect of bend-twist coupling on plate bending can be seen in Fig. 2.25.

Notes on Structural Analysis of Composite Shell Structures

25

Figure 2.25 Effect of bend-twist coupling on plate bending (Jones, 1999)

An example of advanced design utilizing the bend-twist coupling can be seen on Fig. 2.26 which
illustrates the response of forward-swept wings on the Grumman X-29A.

Figure 2.26 Response of forward-swept wings (Jones, 1999)

Forward-swept wings are subjected to aerodynamic loads that tend to twist the wing about an
axis that is along the wing and off perpendicular to the fuselage. The possible result of these
loads are aerodynamic divergence (gross wing flapping that in the limit tears the wing off).
The wings are designed using composite laminates at various angles to the wing axis which
results in D16 and D26 bend-twist coupling that cause the wing to twist in the opposite sense
to the aerodynamic wing-twisting effect. Such examples of aeroelastic tailoring of composite
structures clearly illustrates the large potential of designing mechanical systems using these
materials.
Next stresses in the laminate can be determined, see Fig. 2.27. The strains are continuous
from lamina to lamina, and because material properties Q ij are discontinuous, the stresses also
becomes discontinuous from lamina to lamina in the classical lamination theory.

2. Introduction to Mechanics of Composite Materials

26

Figure 2.27 Hypothetical variation of strain and stress through the laminate thickness (Jones, 1999)

The stress evaluation in each layer k is given as:

Q11 Q12 Q16


x
x
y
y
= Q12 Q22 Q26

xy k
xy k
Q16 Q26 Q66 k

Q11 Q12 Q16


= Q12 Q22 Q26
Q16 Q26 Q66


x + z x

 + z y
y

xy + z xy k
k

(2.32)

Finally, the following conclusions can be made on classical laminated plate theory:
Advantages
Simple extension of isotropic plate theory.
Computer codes designed for isotropic plates can be easily modified to do composite
analysis.
Gives excellent results for many problems. Particularly true if the results involve
average structural properties.
Disadvantages
Transverse shear stresses are assumed to be zero, see Fig. 2.19, i.e., cannot predict
delamination.
Stress equilibrium is not satisfied across ply interfaces, which is an unrealistic result.
Results are only reliable for thin laminates. Please remark that thin in case of
laminated plates is a combination of geometry and stiffness.
2.5.2 ReissnerMindlin plate theory
Another so-called single layer theory is the ReissnerMindlin plate theory and the following
description is based on Hansen (1996).
The basic assumptions of ReissnerMindlin plate theory are:

Notes on Structural Analysis of Composite Shell Structures

27

1. The plate may be moderately thick.


2. The displacements u(x, y, z), v(x, y, z) and w(x, y, z) are small compared to the plate
thickness.
3. The inplane strains x , y and xy are small compared to unity.
4. The transverse normal strain z is negligible.
5. The transverse shear stresses xz , yz are parabolic in z.
The displacement representation is given by
u(x, y, z) = u(x, y) + zx (x, y)
v(x, y, z) = v(x, y) + zy (x, y)
w(x, y, z) = w(x, y)

(2.33)

where x (x, y) and y (x, y) are rotation like variables. The straindisplacement relations thus
are:
u
x
x =
+z
x
x
v
y
y =
+z
y
x


x y
u v
xy =
+
+z
+
(2.34)
y x
y
x
w
+ x
xz =
x
w
yz =
+ y
y
The constitutive relations in lamina coordinates are given as:

Q11 Q12 0
0
0

1
1

Q12 Q22 0

0
0

0 Q44 0
0
23
4
=
=
0

0
0
0
Q
0

31

5
55

0
0
0
0 Q66
12
6
and in structural coordinates as:

yz
=

zx

xy

Q11 Q12 0
0 Q16
Q12 Q22 0
0 Q26
0
0 Q44 Q45 0
0
0 Q45 Q55 0
Q16 Q26 0
0 Q66

x
y
yz
zx
xy

1
2
23
31
12

(2.35)

(2.36)

2. Introduction to Mechanics of Composite Materials

28

The resultants are given as:


Stress:
t/2
R

Nx =
Ny =

t/2
t/2
R

Moment:
t/2
R
x z dz
Mx =

x dz
y dz

t/2
t/2
R

Nxy =

My =

t/2
t/2
R

xy dz

Qx =

y z dz

t/2
t/2
R

Mxy =

t/2

Shear:
t/2
R

t/2
t/2
R

Qy =

xz dz
yz dz

(2.37)

t/2

xy z dz

t/2

The stress, moment and shear resultants can be written in matrix form as:



A | B | 0
N


+ +

B
|
D
|
0
M

+ +


Q
z
0 | 0 | A

(2.38)

where:

Aij =

N
X

Bij = 1/2

Dij = 1/3

i, j = 1, 2, 6

(2.39)

2
(Qij )k (zk2 zk1
),

i, j = 1, 2, 6

(2.40)

3
),
(Qij )k (zk3 zk1

i, j = 1, 2, 6

(2.41)

i, j = 4, 5

(2.42)

(Qij )k (zk zk1 ) ,

k=1
N
X

k=1
N
X

k=1
N
X

Aij =

k=1

(Qij )k (zk zk1 ) ,

The stress evaluation in each layer k is given as

Q11 Q12 0
0 Q16
x

0 Q26
y
Q12 Q22 0
yz
=
Q
Q
0
0
0
44
45

0
0 Q45 Q55 0

zx
xy k
Q16 Q26 0
0 Q66

Q11 Q12 0
0 Q16
Q12 Q22 0
0 Q26

= 0
0 Q44 Q45 0
0
0 Q45 Q55 0
Q16 Q26 0
0 Q66

x
y
yz
zx
xy

u
y

u
x
+ z
x
x

v
+ z xy
y

w
+ y
y

w
x +
x
v
x
+ z
x
y

The following things should be noted about the stress evaluation:

y
+ x
k
(2.43)

Notes on Structural Analysis of Composite Shell Structures

29

Strains are continuous from lamina to lamina.


Material properties Qij are discontinuous
stresses are discontinuous from lamina to lamina.
Transverse shear strains yz , zx are constant with respect to z.
Transverse shear stresses yz , zx are piecewise constant from lamina to lamina.
yz , zx Dont satisfy equilibrium at lamina interfaces.
yz , zx are not parabolic?
Requires shear correction factor.
The following conclusions can be stated for ReissnerMindlin plate theory :
Advantages
Excellent results for thin to moderately thick plates and shells.
Computer codes are simpler. Lower order derivatives than classical plate theory.
Transverse shears for prediction of throughthickness failures.
Disadvantages
Transverse shear stresses do not satisfy surface equilibrium.
Equilibrium not satisfied at ply interfaces.
Shear correction factors required.

Shear locking for thin plates in finite element analysis.


Special tricks required.

2.5.3 Higher order plate theories


Assumptions of a higher order plate theory can also be used within a single layer formulation:
1. The inplane displacements u(x, y, z), v(x, y, z) are cubic functions of z.
2. The normal displacement w(x, y, z) is independent of z.
3. The transverse shear stresses xz , yz are parabolic in z and vanish at the upper and lower
surfaces of the plate.
The following conclusions can be made for such a higher order plate theory:
Advantages
Gives excellent results for thin to moderately thick plates and shells.
The transverse shears satisfy surface boundary conditions and are quadratic in z. No
shear correction factor necessary.
Disadvantages

2. Introduction to Mechanics of Composite Materials

30

Stress equilibrium is not satisfied across ply interfaces.


Computer codes more complex than ReissnerMindlin.
Five variables u(x, y), v(x, y), w(x, y), x(x, y), y (x, y) to be determined.
2.5.4 Concluding remarks on plate theories
Before the concluding remarks are stated about the described laminated plate theories, a short
summary of requirements and available theories are given (Hansen, 1996):
Requirements
Displacement continuity at ply interfaces (not strain).
Equilibrium at ply interfaces (including plate top and bottom).
Conventional single layer plate theories
Dont work.
Displacements are continuous functions of z.
3D finite elements
Displacement based formulations dont satisfy interelement equilibrium.
Mixed formulation; x , y , z , xy , xz , yz u, v, w
Computationally very intensive.
Layered plate theories
Treat plies as connected plates
Similar to 3D finite elements.
The concluding remarks on the described laminated plate theories are (Hansen, 1996):
1. Lamination theory is a rule of mixtures representation of the lamina interaction in a
plate or shell.
2. Classical plate theory predicts thin plate deflections, natural frequencies, buckling loads
and inplane failure easily and accurately.
3. Classical plate theory cannot predict throughthickness failure, delamination or edge failures. A desire for accurate prediction of these failures has been the driving force behind
the introduction of enhanced plate theories.
4. ReissnerMindlin and higherorder plate theories yield better throughthickness stress
distributions and better results for thick plates and shells.
5. There are no plate theories available which yield reliable, accurate, throughthickness
stress results for arbitrary laminates. More work and perhaps a rethinking of the problem
may be in order.

Notes on Structural Analysis of Composite Shell Structures

31

Using the concepts developed over the preceeding sections it is possible to fully describe composite laminates. However, a special case of laminates must be treated seperately, namely sandwich structures.

2.6

Sandwich structures

Finally, a special kind of laminated composite called a sandwich structure is described.


A structurel sandwich can be defined as a special type of laminated composite comprising of a combination of different materials that are bonded to each other so as to
utilize the properties of each separate component to the structural advantage of the
whole assembly (Kildegaard et al., 1999).
A classical three-layer sandwich structure can be seen in Fig. 2.28.

Figure 2.28 Classical three-layer sandwich (Kildegaard et al., 1999)

The face of the sandwich structure will typically be metal, GFRP (glass fiber reinforced plastic)
or CFRP (carbon fiber reinforced plastic) having the following characteristics:
large stiffness
large strength
large resistance against impact, etc.
nice surface finish
The core may be honeycombs, balsa tree, polymer foam (PVC, PUR, PS), etc., with the following characteristics:
low density
large shear stiffness
large shear strength
large transverse stiffness

2. Introduction to Mechanics of Composite Materials

32

low thermal conduction


The geometrical effect on the bending stiffness and strength of a structural element can be seen
in Fig. 2.29.
Total thickness of
face = t

Weight

Bending
stiffness

Bending
strength

Figure 2.29 Sandwich effect (Kildegaard et al., 1999)

This sandwich effect is used to a high extent in the design of a wind turbine blade, see Fig. 2.1Fig. 2.3.

2.7

Glossary

The following list contains a glossary of frequently used terms for describing composite materials (Kildegaard et al., 1999):
Composite material - A material made by combining two or more materials, macroscopically different in form and composition, and reciprocally insoluble.
Fibrous composite - A composite material made by two components, one of which (the
fiber) in fibrous form, generally destined to give rigidity and strength, and the other (the
matrix), binding together the reinforcement material.
Reinforced plastics - A plastic material to which reinforcing fibers have been added, to
provide strength and rigidity greatly superior to those of the base resin.
Continuous fiber composite - A composite whose reinforcement is made by fibers of
indefinite length.
Discontinuous fiber composite - A composite whose fibers have a limited length, typically 3 to 50 mm.
Oriented fiber composite - A composite whose fibers are aligned according to one or
more preferred orientations.

Notes on Structural Analysis of Composite Shell Structures

33

Unidirectional composite - A composite in which the fibers are aligned according to one
principal orientation. An unidirectional composite is anisotropic.
Hybrid composite - A composite reinforced with fibers of different nature (e.g. glass
and graphite).
Glass fiber reinforced plastic (GFRP) - A polymeric matrix reinforced by glass fibers.
Carbon fiber reinforced plastic (CFRP) - A polymeric matrix reinforced by carbon
fibers.
Kevlar fiber reinforced plastic (KFRP) - A polymeric matrix reinforced by kevlar
fibers.
Strand - A bundle of continuous fibers combined in a single unit without twist. A strand
is typically made of 50 to 200 fibers.
Yarn - An assemblage of twisted bundles to form a continuous length suitable for use in
weaving into textile materials. A yarn is typically made of more than 10000 fibers.
Mat - A fibrous material for reinforced plastics consisting of randomly oriented continuous or, more frequently, chopped filaments held together by a binder.
Fabric - A reinforcing planar structure for reinforced plastics obtained by interlacing
orthogonally oriented yarns in a loom. A glass fabric usually weights no more than 400
g/m2 .
Woven roving - A heavy glass fiber, usually planar, made by weaving roving bundles.
Lamina (ply) - A composite made by a single layer of mat, fabric, woven rowing or
unidirectional fibers held together by a matrix.
Laminate - A material made by bonding together a series of laminae.
Stacking sequence - The order in which laminae are laid up in a laminate.
Symmetric laminate - A composite laminate in which the sequence of laminae below
the laminate midplane is a mirror image of the stacking sequence above the midplane.
Balanced laminate - A symmetric laminate in which, if a lamina oriented at +8 with
respect to a reference axis is present, an equal lamina oriented at 8 is also present.

34

2. Introduction to Mechanics of Composite Materials

Chapter

Continuum-Based Shell Elements

his chapter introduces the concept of a shell structure and further develops the necessary
theory for formulating shell finite elements.

The first step is to derive the governing equations for a general continuum mechanical system
and subsequently, to impose certain restrictions on the system thus arriving at the shell description. We develop the element formulation for a shell and present a thorough treatment of geometric and kinematic representation of a shell. Furthermore, constitutive relations, numerical
integration and potential sources of error (locking) are addressed as well.
We first briefly review the governing equations for a three-dimensional problem.

3.1

Governing equations Equations of motion

The equations of motion are the governing equations for any mechanical problem and hence essential for the understanding of the finite element method. The general equations of motion can
be derived directly by investigating the forces on a mechanical system and enforcing Newtons
law. Consider a system under some loading conditions, as shown in Fig. 3.1.
c

z
y
x

Figure 3.1 Some mechanical system under loading from concentrated load, r c , surface traction, rs , body
forces, rb (not illustrated).

3. Continuum-Based Shell Elements

36

Any small volume fraction of the body having mass, dm, must fulfill the condition of equilibrium of forces:
dr = dm
u
(3.1)
where u
is the acceleration vector and r is the resulting force vector of the small volume.
Considering Fig. 3.1 the relation in (3.1) may be written for the entire body as:
Z
Z
Z
s
b
ri d + ri d =

ui d
(3.2)

where denotes the entire domain and is the surface area on which the force r is is acting. In
(3.2) all concentrated loads are included in the term, rs , which is a reasonable assumption since
forces in effect always act on an area, no matter how small it may be. The term "point load" is
only a conceptual aid when dealing with continuous systems but holds a more natural meaning
when the system is discretized. Consequently, the forces, rc , will from this point be included in
the term rs if nothing else is stated.
All internal forces in (3.2) not acting on a surface cancel out as a consequence of Newtons third
law and hence, using Cauchys formula the tractions can be stated as:
(3.3)

ris = ij nj
where, nj , is the surface normal. The relation in (3.2) may then be written as:
Z
Z
Z
b
ij nj d + ri d =

ui d

(3.4)

To convert the first integral Rfrom surface toR volume integral the Gauss divergence theorem is
employed, which states that Aij d = Aij nj d. By collecting terms under the integral,
(3.4) may be reduced to:
Z

ij,j + rib
ui d = 0
(3.5)

Since (3.5) has been derived for a body of arbitrary shape it seems reasonable to assume that
the condition in (3.5) is satisfied if the integrant is zero. Hence, the general equation of motion
is written as:
ij,j + rib
ui = 0
(3.6)

If there is no motion of the system the acceleration vanishes from the equations of motion in
(3.6) and the equilibrium equations are obtained:
ij,j + rib = 0

(3.7)

The equations of motion, or the equilibrium equations, are the governing equations for any mechanical system. However, in displacement based finite element another form of the equations
is employed. These equations are derived using the principle of virtual displacements to equation (3.6) or (3.7). Consider a small displacement field, ui , which is assumed to fulfill the
essential boundary conditions, i.e. the displacement field causes no displacements or rotations
where these are constrained. This is an fundamental requirement since a displacement field not
fulfilling these conditions would yield a wrong expression for the work. Applying the virtual

Notes on Structural Analysis of Composite Shell Structures

37

displacements to the integral form of (3.6) yields an expression for the virtual work done by the
forces:
Z
Z
Z
b
ij,j ui d + ri ui d
ui
ui d = 0

Z
Z
Z
(3.8)
b
((ij ui ),j ij ui,j ) d + ri ui d
ui
ui d = 0

since it has been exploited that (ij ui ),j = ij,j ui + ij ui,j which is the rule of differentiation
for a product. Rewriting this entity and employing Gauss divergence theorem again to convert
the first integral, the traction forces reappear:
Z
Z
Z
Z
s
b
ri ui d ij ui,j d + ri ui d
ui
ui d = 0
(3.9)

Noting that ui,j = ij and collecting terms we arrive at the work conjugate equilibrium equations for the work done in the system by introducing ij in accordance with (4.10):
Z
Z
Z
Z
b
s
ij ij d =
ri ui d + ri ui d
ui
ui d
(3.10)

|
{z
} |
{z
}
Internal work

External work

The left-hand side of (3.10) constitutes the internal work which is the work done by the internal
stresses as a consequence of the work done by the external forces, described by the right-hand
side of the equation. So far it has been implicitly assumed that the body considered has no prior
deformation history and consequently that the body has no initial stresses or strains. However,
by introducing a more general form of the constitutive relation in (3.40) we may introduce a
work contribution from an initial stress state, denoted by 0 ij , and an initial strain state denoted
by 0 ij . The more general form of the constitutive relation including initial effects may then be
written as:
ij = Cijkl (kl 0 kl ) + 0 ij
(3.11)

which describes the "total stress" after deformation and not only the increment from a reference
state. Introducing the total stress in the energy expression (3.10) yields:
Z
Z
Z
Cijkl kl ij d
Cijkl 0 kl ij d + 0 ij ij d

Z
Z
Z
(3.12)
b
s
=
ri ui d + ri ui d
ui
ui d

which is a more general form of (3.10). The expressions in (3.10) or (3.12) contain the basic equations for the finite element method. The same result could have been obtained using
the principle of stationary potential energy, but the principle of virtual displacements has been
applied as a more intuitive method. To solve the governing equations using the finite element
method they must be discretized which is the subject of the following.
3.1.1 Discrete governing equations Finite element formulation
The first step in discretizing the governing equation is to introduce the concept of a discrete
domain as oppose to the continuous domain of Fig. 3.1. A discrete model of the mechanical

3. Continuum-Based Shell Elements

38

z
y
x

Figure 3.2 A discrete model of the mechanical system in Fig. 3.1 under equivalent loading conditions.

system in Fig. 3.1 under the same loading conditions could look like the one in Fig. 3.2. The
real, physical loads are represented by equivalent nodal forces and similarly, the constraints are
represented by equivalent nodal constraints. The term point-load now seems more appropriate
since also surface tractions and body loads are treated as discrete point loads in the model.
When solving the governing equations for a discrete domain such as the one in Fig. 3.2 the
equations are only solved in a finite number of points. Consequently, a need arises for a way
of describing the intermediate points, corresponding to any point in the continuous domain. To
this end an interpolation, N, is introduced within each element, linking nodal values to any
point in the body as x = Nxa where xa contains coordinates of all nodes, na , in the element.
If the finite element interpolation is isoparametric, the same interpolation is used to evaluate
displacements from nodal values, i.e. u = Nua . It follows immediately that strains, , can be
found from nodal displacements, ua , as:
(3.13)

 = Nua = Bua

where B is the strain-displacement matrix. By using the strain-displacement relation (3.13)


and the interpolation matrix, N, the discrete displacements may be introduced in the governing
equations (3.10) and expressed for any element of volume V and area A:
Z
Z
Z
T T
T
T b
ua B dV =
ua N r dV +
uTa NT rs dA
V

na
X
k=1

uTk rck

uTa NT N
ua

(3.14)

dV

where point loads have been naturally introduced again as nodal loads. Note that matrix and
vector quantities have been transposed and moved in (3.14) in order to fulfill the rules of linear
algebra; this does not in any way alter the meaning of the equation. The global governing
equations are obtained as the sum of (3.14) over all elements, nel , in the discretized domain.
In (3.14) it proves convenient to introduce an internal force vector, , and a mass matrix, M.
When noting that the
are the unknowns the mass matrix of elementRk is defined
R displacements
T
naturally as Mk = V N N dV and the internal element force vector as k = V BT dV .
By further introducing a resulting force vector, r, and collecting terms the expression in (3.14)

Notes on Structural Analysis of Composite Shell Structures

39

can be written for the entire system as a sum over all the elements, nel :
nel
X

k +

nel
X

=
Mk u

(3.15)

rk

k=1

k=1

k=1

nel
X

where uTa has been eliminated through a standard algebraic operation. Note that the subscript,
a, referring to nodal values has been left out for the sake of brevity. In the following it will
therefore be implied that u refers to nodal values. The vector r contains the sum of all external
forces in the system and is often referred to as the nodal load vector. When performing static
analysis the acceleration term disappears and we arrive at the static equations of equilibrium,
which state that internal forces must balance external forces:
nel
X

k =

k=1

nel
X

(3.16)

rk

k=1

The solution of (3.16) depends on the strain tensor. The general Green-Lagrange strain tensor
is given as:


1 u
u
u u
ij =
(3.17)
+
+
2 xj xi xi xj

which contains both linear and non-linear terms. In linear analysis we disregard the non-linear
terms of (3.17) and thus, the strain-displacement matrix, B, and hence the stiffness matrix,
K, become independent of the displacements. In this case we refer to the global system of
equations (3.16) as:
Kd = r
(3.18)
where K = BT CB and d is the global displacement vector which constitutes the unknown in
the expression. The expression in (3.18) relates the response of a body characterized by K to
the external loading conditions, r. It is easily realized that the matrix K is a measure of the
compliance of the body since an increase in numerical value of K will yield a smaller deformation (reduced compliance or increased stiffness) for the same forces and vice versa (increased
compliance or decreased stiffness). This naturally gives way to the term stiffness matrix as
a characteristic measure of the response of a body. The equations in (3.18) are in essence a
linear system of algebraic equations and are solved using standard numerical techniques. The
non-linear terms will be left out until Chapter 4.

So far the force vector, rk , in e.g. (4.15) has just been stated and implicitly regarded as "welldefined" in broad terms. However, for the force vector to fulfill the equation of equilibrium
(3.14) it must be work-equivalent with the distributed forces. This has merely been implied
in the preceding but is an important feature of the element nodal load vector. To illustrate the
property of work-equivalence consider the element nodal load vector, defined as the sum of the
forces given in the right-hand side of (3.14):
Z
Z
Z
T b
T s
c
k dV
(3.19)
rk =
Nk rk dV +
Nk rk dA + rk
Mk u
V

We now consider only the body force term, rbk , and the equivalent nodal force contributions and
apply a small nodal displacement, uk :
Z
Z
T
T
T b
uk rk =
uk Nk rk dV =
uT rbk dV
(3.20)
V

3. Continuum-Based Shell Elements

40

since u = Nuk as stated earlier. From (3.20) it is apparent that the work done by the nodal
forces as a consequence of the nodal displacements is equivalent to that done by the external
forces for equivalent distributed displacements; hence it has been shown tentatively that the
nodal load vector and the real loads are work-equivalent and it is straight-forward to show the
same for the other distributed loads. It can further be shown (as by e.g. Cook et al. (1989)) that
the loads calculated from the nodal load vector are statically equivalent to those found from the
distributed loads, which is an equally important feature.
The choice of interpolation functions, N, is very important for the accuracy of the results obtained from a finite element analysis and some general remarks can be made. It has already
been stated that the use of the same interpolations for the coordinate and displacement mappings is referred to as isoparametric element formulation. If the same interpolation is used in
the expression for the mass matrix, as implied in the previous discussion, the mass matrix is
said to be consistent. Similarly, the nodal load vector is consistent if the displacement interpolations are used in the mapping of distributed forces to nodal values. In general we may state
that the consistency property applies if a quantity (be it rk or Mk ) uses the same interpolation
as used in the calculation of the stiffness matrix, Kk ; this seems reasonable when considering
the governing equations in (4.15).
In the following the formulation of a suitable element for the solution of (3.18) is treated in
depth for the special case of a shell structure. First, we introduce the concept of a shell.

3.2

The basic shell concept

A shell is characterized as a curved or doubly curved structure which carries loads primarily as
membrane stresses. The classical example of a shell is an egg loaded with uniform pressure on
the entire surface. However, this strict definition is often softened since application of external
loads and supports almost always involve localized bending. Consequently, shell structures in
engineering applications will often carry loads as a mixture of membrane, bending and shearing
stresses. Shell theories are as described in Chapter 2 in general formulated in force and moment
resultants whereby stresses (and strains) are related to integral force and moment quantities
as oppose to the direct relation obtained from 3D continuum mechanics. The consequence of
doing so is that explicit integration over the thickness, emanating from the derivation of the
resultants, is introduced directly in the governing equations. Hence, the thickness dependency
is eliminated leading to a 2D representation of the problem. This constitutes a fundamental
characteristic of shell models.
A consequence of the 2D shell representation is the introduction of curvilinear coordinates
whereby the three-dimensional shell is uniquely defined by two parameters r and s; this is
illustrated in Fig. 3.3.
In effect, the shell geometry and kinematics may then be related solely to the shaded surface in
Fig. 3.3, normally referred to as the reference surface or neutral surface; this greatly simplifies
the geometric and kinematic descriptions. In the following the kinematic representation for a
shell will be formulated using the degeneration approach.

Notes on Structural Analysis of Composite Shell Structures

41

v(r,s)
r
z
y
x

Figure 3.3 Three dimensional shell and two dimensional shell representation (shaded area) characterized
by the curvilinear coordinates (r, s) in the reference surface.

3.3

Element formulation be degeneration of continuum element

The degeneration method is a very "inexpensive" way of formulating shell finite elements and
has been the leading method ever since it was introduced in 1970 by Ahmad et al. (1970). Basically, the method starts from a standard 3D continuum solid element (hence continuum based
shell elements) and, by enforcing various constraints on the element behavior, arrives at a 2D
shell element. The procedure of the method is to eliminate nodes in the 3D continuum element
by enforcing different constraints on the behavior of the element. An 8-node degenerated element can for example be constructed from a solid 20 node brick in three steps as shown in
Fig. 3.4.

Figure 3.4 A schematic representation of the degeneration procedure from a 20 node solid element to an
8 node shell element (from McNeal (1998))

In step (1) nodes in the middle plane are removed, corresponding to assuming constant transverse strain, in (2) opposite nodes are linked by assuming equal displacements and assigning
two rotational degrees of freedom to each pair of nodes and in (3) the motion of each straight

3. Continuum-Based Shell Elements

42

line is described by 5 degrees of freedom in 1 node, lying in the reference surface. The procedure corresponds to creating a geometric 2D element (in curvilinear coordinates) from a 3D
continuum element by enforcing the same assumptions as made for the first order shear deformation theory for shells; these assumptions are revised in the following.
3.3.1 First order shear deformation theory a quick review
The first order shear deformation theory (FSDT) is also often referred to as the Reissner-Mindlin
theory of shells in honor of Reissner (1945) and Mindlin (1951) who proposed the theory. The
four basic assumptions of the FSDT are analogues to those used in Timoshenko beam theory
and can be stated as follows: (1) the shell is thin compared to the radius of curvature (t/R  1),
(2) the linear and angular deformations of the shell are small, (3) the transverse normal stress,
33 , is negligible and (4) normals to the reference plane before deformation remain straight and
inextensible after deformation.
The first assumption is the most basic since it implies the other three, i.e. assumption 1 can
not be violated without violating assumptions 24. Assumption 2 in affect means that the state
of the deformed shell can be related directly to the state of the non-deformed shell, and thus it
ensures that the resulting equations become linear. Assumption 3 is reasonable for thin shells
and has no further implications besides simplifying the derivation of the governing equations.
Assumption 4 has two implications; first, the inextensibility assumption implies zero transverse
normal strain (or 33 = 0) and second, it leads to the introduction of two independent rotations,
and , which describe the rotation of the normal.
The FSDT accounts for transverse shearing strains but neglects transverse normal strains. Consequently, the theory is applicable only to thin to moderately thick shells since transverse normal
effects become predominant in thick shells. Finite elements based on these assumptions will
therefore in general have the same limitations. In the following the four basic assumptions
stated above will be used to formulate the geometric and kinematic representations for a shell
element.

3.4

Geometric and kinematic representation

First, let us extend the the curvilinear coordinate system defined in Fig. 3.3 to also include a
thickness coordinate, t. This coordinate is defined as being normal to the tangent plane spanned
by the tangent vectors at any point, (rp , sp ), and is thus also normal to the shell surface at any
point. This leads to a (r, s, t)-coordinate system as shown in Fig. 3.5(a), which uniquely defines
any point in the shell.
However, the curvilinear coordinate system is tedious to handle when developing the governing equations of the shell, and we therefore employ the mapping of coordinates inherent in the
isoparametric formulation of finite elements. In the isoparametric formulation a natural coordinate system is introduced with fixed boundaries [1, 1] in all three coordinates. The mapping
to the natural coordinate system is shown in Fig. 3.5(b) where the natural space becomes a unit
cube.
Recall that the natural coordinate system is used to express the shape functions, N, which for a

Notes on Structural Analysis of Composite Shell Structures

43

t s

z
y

(a)

(b)

Figure 3.5 Mapping of shell element from (a) global space to (b) natural space

solid 8-node brick are given as (see e.g. Cook et al. (1989)):
1
Na (ri ) = (1 r)(1 s)(1 t)
8

(3.21)

P
where a indicates the ath node in the element. The geometric interpolation is thus x =
N a xa
where xa are nodal coordinate values. The shape functions in (3.21) will in the following be
used for deriving a 4-node shell element. The procedure is to describe the entire shell geometry
from the geometry of a reference surface as shown in Fig. 3.6.
t

s
7

4
3

5
1

V3

2
2

(a)

(b)

Figure 3.6 Degeneration of an 8-node solid element (a) into a 4-node shell element (b); a "" represents
a node and a "" indicates a ghost point

The vectors V3a are called node directors since they indicate the direction of the ghost points
with respect to the ath node at the reference surface. The vectors are unit vectors, pointing
from the ath bottom node to the ath top node as shown in Fig. 3.6. Thus, the node directors
can be defined from the nodal coordinates, i.e. we may write V 3a = ({xai }top {xai }bottom )/ha ,
where ha is the thickness of the element at node a. Consequently, any point above or below the
nodes in the thickness direction can be expressed from the coordinate of the reference surface
node and some distance along the node director. This is a fundamental property of shell finite
elements. Adding the "over the element interpolation" using Na any point within the element
may be described using the node directors, the nodal coordinates and the shape functions.
However, the basic shape functions of (3.21) can not be used directly. When deriving the
new interpolations we proceed by dividing the 3D interpolation functions into a plane part,
Na2D = 14 (1r)(1s), and a thickness direction part, Nat = 21 (1t). Using this the interpolation

3. Continuum-Based Shell Elements

44

for an 8-node solid element can be rewritten as:


xi =

n
X
1
a=1

(1 +

t)Na {xai }top

n
X
1
a=1

(1 t)Na {xai }bottom

(3.22)

where n = 4 is now the number of nodes in the degenerated element, i.e. n shell = nsolid /2 in
the following if nothing else is stated. Rearranging the expression in (3.22) the position of all
points is described by the geometry of the reference plane as:
xi =

n
X

Xt
1
1
Na2D ( {xai }top + {xai }bottom ) +
Na2D ({xai }top {xai }bottom )
2
2
2
a=1
a=1

n
X

n
X
t 2D
2D a
Na ha V3a
=
N a xi +
2
a=1
a=1

(3.23)

where xai is the position of the node on the middle surface.


From (3.23) the position of an arbitrary point in the shell is interpolated by use of the coordinates
and the node directors, V3a , of the nodes on the midsurface. Using the geometry representation
in (3.23) the interpolation of the shell kinematics can be determined.
The displacement, ui , of any point is expressed as the change in position of that point from an
undeformed state at time 0 to a deformed state at time t, i.e. ui = t xi 0 xi where the subscripts 0 and t denote the undeformed and the deformed state respectively. Using the geometry
interpolation in (3.23) the displacement is then expressed as:
ui =

n
X
a=1

Na (t xai 0 xai ) +

n
X

Na uai

|a=1 {z

N odal displacement

n
X
t
Na ha (t V3a 0 V3a )
2
a=1

n
X
t
+
Na ha (t V3a 0 V3a )
2
|a=1
{z
}

(3.24)

Relative nodal displacement

where uai is the displacement of the nodes on the midsurface. From (3.24) the displacement of
an arbitrary point is given as the displacement of the node and a displacement relative to the
node, described by the node directors V3a . The main problem in developing an expression for the
displacement interpolation is to express the difference between the deformed and undeformed
vector V3a , i.e. the rotation of 0 V3a about the ath node. There are numerous ways of doing so
but in the following the method used by e.g. Bathe (1996) and Zienkiewicz (1977) is adopted.
We proceed by defining a Cartesian coordinate system in every node of the element the socalled director coordinate system. The coordinate system is set-up with its third axis identical
to the vector 0 V3a itself, and the other two axes are defined through the global unit base vectors,
i, j and k by enforcing that the axes must be perpendicular to each other. The two additional
base vectors are thus defined through the cross product as:
j 0 V3a
=
|j 0 V3a |
a
a
a
0 V2 = 0 V3 0 V1
a
0 V1

(3.25)

Notes on Structural Analysis of Composite Shell Structures

45

It should be noted that if 0 V3a is identical to j the definition in (3.25) fails. However, in this
case 0 V1a is simply set equal to the global base vector k. The director coordinate system is used
to describe the rotation of the vector 0 V3a as the element deforms. By considering Fig. 3.7 the
rotations can be determined.

-a
b
t

z
0V3

V3

V2

x
a

V1

Figure 3.7 Displacement components resulting from rotation of the unit node director 0 V3a in node a.
The rotations are defined in the global (x, y, z)-coordinate system and can be described by the displacements caused by the rotations and .

When the vector 0 V3a is rotated an angle about 0 V1a as shown in Fig. 3.7 the tip of the vector
will be displaced the arc length |0 V3a | sin() in the direction of 0 V2a . If we assume that the
rotation is small, sin() ' whereby the rotation causes a linear displacement of a distance
in the direction of 0 V2a since |0 V3a | = 1. Using the same assumption for we obtain that
when 0 V3a is rotated an angle about 0 V2a , its tip will have a positive displacement along 0 V1a
corresponding to a distance . The third rotation of 0 V3a is the rotation around the vector itself
but it does not contribute to any relative displacement and is therefore disregarded. The above
assumptions in effect state that the vector 0 V3a remains straight after deformation and rotates
relative to the reference surface. This constitutes one of the Mindlin assumptions as stated
earlier. The condition of inextensibility is not enforced through the geometric description but is
introduced via the constitutive relation as described in Section 3.7. It is important to note that
the above assumptions regarding the rotations may introduce errors in the model when the shell
element is used in problems involving large rotations. Consequently, shell elements based on
this description are limited to model moderate rotations.
Now the relative displacement of 0 V3a is expressed through the two rotations and in the
director coordinate system. To obtain the relative displacement in the global coordinate system,
= (, , 0) in the director coordinate system and apply the
we form a displacement vector u
standard transformation law for Cartesian tensors (see e.g. Heinbockel (1996)). Thereby, the
displacement in the global coordinate system, u, can be expressed as:
j ei
ui = u
j e
= a 0 V2ia + a 0 V1ia

(3.26)

j = Vj and ei are the global base vectors i, j and k. Note that 0 V1ia and 0 V2ia are the ith
since e
components of the vectors. The relative displacement term in (3.24) can then be expressed by
(3.26) whereby the total displacement of an arbitrary point in the shell element can be expressed

3. Continuum-Based Shell Elements

46

as:
ui =

n
X
a=1

Na uai

n
X
t
+
ha Na (a 0 V2ia + a 0 V1ia )
2
a=1

(3.27)

The subscript referring to the deformation state of V1 and V2 will in the following derivations
be implied and we simply write V2ia and V1ia .
As can be seen from (3.27) each node in the element has 5 degrees of freedom, (u, v, w, , ).
This corresponds to the displacement field originally proposed by Mindlin for plates and shells.
The disadvantage of using 5 d.o.f. is a loss in generality since moment boundary conditions
cannot be applied around the third local axis. On the other hand, introducing a 6th d.o.f. leads
to a badly conditioned stiffness matrix since no physical stiffness is associated with this extra
rotation. We here maintain the 5 d.o.f. formulation since moment boundary conditions are
relatively uncommon and also because we wish to achieve a robust formulation.
In the preceding discussion only the global and natural coordinate systems have been employed.
In practice however, it proves convenient to define two element specific local coordinate systems; the element coordinate system and the material coordinate system.

3.5

Local coordinate systems

The element coordinate system (or ECS) is used to define the stress and strain locally for each
element. The primary reason for doing so is that the material properties are only unambiguously
expressed in a local coordinate system. The material coordinate system is analogous but defines
the directions of the principle material directions used for setting up the constitutive matrix. A
natural consequence of the introduction of additional coordinate systems is the introduction of
transformation laws. The reason is that the stiffness matrix for the entire system is only defined
in global coordinates and consequently, local strains and stresses must be transformed from
local to global coordinates. In the following the two coordinate systems are defined, starting
with the element local coordinate system.
3.5.1 Element coordinate system (ECS)
The element coordinate system is defined as a covariant base rather than the well-known
Cartesian base. The reason for this is obvious from a general continuum mechanical point
of view since strains are naturally defined in such a system (see e.g. Heinbockel (1996) or
Bathe (1996)). Thus, we define a local coordinate system by the three vectors g i as shown in
Fig. 3.8.
It is important to note that the base vectors gi are in general not orthogonal and furthermore,
need not be of unit length. This is a fundamental difference from cartesian bases in which the
base vectors (often referred to as i, j and k) are always orthogonal and of unit length. The
definition of the new base vectors is as follows.
Consider the vector, v = xi + yj + zk, in Fig. 3.8, which is the position vector of the point
considered. The base vectors, gi , are called the covariant base vectors, and are defined in global
coordinates as:
v
gi =
(3.28)
xi

Notes on Structural Analysis of Composite Shell Structures

47

g3
g1

g2

v
r
z
y
x

Figure 3.8 An arbitrary surface with global Cartesian (x, y, z) coordinate system, local (r, s) curvilinear
coordinate system and local covariant coordinate system spanned by g i

In finite element analysis the local coordinates, xi , of the point will be the natural coordinates,
r, s, t or simply ri . There also exists a reciprocal non-orthogonal basis, called the contravariant
base, also defined in global coordinates as:
gi =

xi
v

(3.29)

Note that in (3.28) and (3.29) the indices indicate a tensor defined in covariant basis and contravariant basis when written respectively as subscripts and superscripts. Also note that although
the base vectors define a local coordinate system they are defined in global coordinates. As a
consequence of the introduction of a local coordinate system it is understood in the following
that all quantities are referred to a local coordinate system. Hence, everywhere on the shell 33
is transverse normal stress at the point considered, 12 is local shearing stress in the (x, y)-plane
spanned by the local x and y axes and so on.
As already mentioned the covariant base is a natural choice for expressing strains, since they
hold a natural meaning in tensor notation. However, the matrix of material constants (the constitutive matrix), C, is only well-defined in a Cartesian base. Consequently, the need arises
for a Cartesian base which is "close" in some sense to the covariant base as described by e.g.
Hughes (1987). This cartesian coordinate system is the material coordinate system.
3.5.2 Material coordinate system
The material coordinate system is defined in order to express the direction of the material properties in the element and to make transformation of the material constants to the global coordinate system possible. The material coordinate system is a Cartesian system having its first and
second axes as tangents to the surface and the third axis as a normal to the surface. Since the
third axis is normal to the surface the third base vector of the system, m 3 , can be constructed
through the expression in (3.25). In order to determine the other two base vectors, m1 and m2 ,
it is necessary to define two auxiliary vectors a and b from the covariant vectors and m 3 as
done by e.g. Hughes (1987):
a=

1
(gr
2
1
k 2 (gr

+ gs )
+ gs )k

b=

m3 a
km3 ak

(3.30)

3. Continuum-Based Shell Elements

48

From the two auxiliary vectors the two remaining base vectors are determined as:

2
2
(a b)
m2 =
(a + b)
m1 =
2
2

(3.31)

Thereby, the three base vectors, mi are uniquely defined and the constitutive matrix can be
stated in this orthogonal coordinate system. For an isotropic material the expression in (3.31) is
enough to uniquely define the directions of the material properties since the only requirement of
the direction is that it must be in the plane of the shell and no concern has to be taken regarding
the specific material directions. For an orthotropic laminate, however, this is not the case since
it is required that the lamina coordinate axes lie in the principle material directions, denoted as
1 and 2.
From the analysis input a material angle, , is stated which defines the direction of the first
principal material direction relative to the first axis of the material coordinate system. The
interpretation of the angle is left to the analysis module and the implemented definition is as
follows. The principal material directions are those obtained by rotating the material coordinate
system the angle about m3 . The in-plane transformation (rotation) about the normal, m3 , is
carried out using a standard transformation matrix (see Section 3.7, (3.65)). Consequently, the
principal material directions are the rotated m1 and m2 axes as shown in Fig. 3.9.
m3
2
m2
1

q
m1

Figure 3.9 Definition of principal material directions, 1 and 2, for orthotropic materials based on the
lamina coordinate system, mi , and the material angle .

These definitions are based on the standard element coordinate system, defined from the covariant base vectors. Depending on the geometry and mesh of the model, these coordinate systems
will in general not be continuous between adjacent elements. Consequently, defining the material directions from the default ECS may lead to an erroneous model. The reason is that the
default element coordinate system is dependent on the Jacobian matrix (because of the covariants) which in turn is dependent on node numbering and element orientation. This is without
consequence for the stresses and strains but the material representation becomes meaningless
as shown in Fig. 3.10.
The laminate we are trying to model is a +45/-45 lay-up. As can be seen the interpretation of
the angles by the analysis module leads to a wrong lay-up. This in turn leads to very erroneous
results. Therefore, it is possible to specify a user-defined element coordinate system.
3.5.3 User-defined element coordinate system
If an element coordinate system has been defined in the analysis input, this definition will
override the material coordinate system described above. Hence, if an ECS has been defined, the
principal material directions will be defined by rotating the defined element coordinate system
the angle .

Notes on Structural Analysis of Composite Shell Structures

49

Figure 3.10 Interpretation of material angle of 45 , relative to the z-axis using standard element coordinate system.

The definition of the element coordinate system has been adopted from the commercial software
package COSMOS/M. It states that when an element coordinate system, (x e , ye , ze ) has been
defined (see Fig. 3.11) the coordinate system is considered as:
If the angle, , between the surface normal, l3 , and xe is smaller than 45 then x is the
projection of xe on the element plane.
Conversely, if the angle is larger or equal to 45 then x is the projection of ye on the
element plane.
The y axis is defined to complete the right-hand rule, i.e. as the cross-product of l 3 and x.
The definition may seem complicated as first glance but it actually provides a well-defined element coordinate system. The definition is illustrated in Fig. 3.11.
x^
l3, z^

l1

y^
x^
l1

l3, z^

ye
y^

l2
xe

f > 45

ze

l2

f 45

Figure 3.11 Definition of local element coordinate system (


x, y, z) from a defined element coordinate
system (xe , ye , ze ).

Using the ECS-definition stated above let us consider the example of Fig. 3.10 again. We proceed by defining an ECS with its first-axis along the global z-axis. This leads to the lay-up
interpretation in Fig. 3.12.

3. Continuum-Based Shell Elements

50

Figure 3.12 Interpretation of material angle of 45 , relative to the z-axis using user-defined element
coordinate system.

Now, all material coordinate systems are aligned at 45 to the global z-axis which is the correct
interpretation. The discussion on user-defined element coordinate systems leads to a very important note. When dealing with laminated structures, particularly complicated geometries,
it is important to define an ECS to ensure well-defined material directions. Particularly so
when using commercial software such as COSMOS/M, Ansys etc. which unlike MUST do not
provide any facilities for visualizing the material directions. Still, using MUST the recommendation stands since severe errors may be introduced when using the standard ECS.
When the geometry is well-defined using the definitions discussed in the preceding section we
may proceed by defining the essential element matrices.

3.6

Linear element matrices

In the following the element matrices i.e. the interpolation matrix, N and the strain-displacement
matrix, B are determined for the 4-node shell element. The starting point is the interpolation
matrix.
3.6.1 Interpolation matrix
The interpolation matrix is the matrix that interpolates the displacement of an arbitrary point
using the nodal displacements i.e. u = Nua . Furthermore, the interpolation matrix is used
when calculating respectively the element nodal load vector as shown in (3.19) and the mass
matrix in (3.14). From the displacement interpolation in (3.27) the interpolation matrix can be
determined by collecting terms and forming a matrix equation as:


t
t
a
a
N a 0
N
h
V
N
h
V
0

a
a
a
a
21
11
2
2




t
t
(3.32)
N = 0 Na 0 2 Na ha V22a 2 Na ha V12a


t
t
a
a
0
0 Na 2 Na ha V23 2 Na ha V13
The interpolation matrix is also used when the strain-displacement matrix is defined.

3.6.2 Strain-displacement matrix

Notes on Structural Analysis of Composite Shell Structures

51

The strain displacement relation is usually written as  = u, where is a differential operator


and the vector, , is defined as  = {11 , 22 , 33 , 212 , 223 , 213 }T = {1 , 2 , 3 , 4 , 5 , 6 }T .
Inserting (3.32) in this relation yields  = Nua whereby the strain-displacement matrix can
be written as B = N. When developing B we first state the strain definition. We use the
linear part of the Green-Lagrange strain tensor:


1 u
u
ij =
+
(3.33)
2 xj xi

By inserting the displacement interpolations in (3.33) it is possible to derive a discretized expression for the linear strains. However, since the displacements are derived with respect to the
global coordinates and the displacement interpolations are functions of the natural coordinates,
a transformation between the natural and global coordinate system is needed. Using the chain
rule of differentiation the derivative of the displacement with respect to the global coordinates
can be expressed as:
ui rk
ui
=
(3.34)
xj
rk xj
The derivatives of the displacement with respect to the natural coordinates (the first right-hand
term of (3.34)) can be determined directly from the displacement interpolation in (3.27) which
yields the following relation:

a
u
t
t
a
a
i
h
N
V
h
N
V
N

ui
a
a,r
a
a,r
a,r
n
2i
1i
r
2
2
ui X

t
t
a
a
(3.35)
Na,s 2 ha Na,s V2i 2 ha Na,s V1i a
s =
1
1
ui
a
a
a=1
h N V
0
2 ha Na V2i
a
2 a a 1i
t

In order to use the transformation in (3.34) the derivatives of the natural coordinates with respect
to global coordinates are also needed (the second right-hand term of (3.34)). They can be
determined by first finding the derivative of the global coordinates with respect to the natural
coordinates and subsequently inverting that expression. Taking the derivative of the position
interpolation in (3.23) yields the expressions for xi /rj to be inverted:

n 
t
xi X
a
a
Na,r xi + ha Na,r V3i
=
r
2
a=1

n 
xi X
t
a
a
=
Na,s xi + ha Na,s V3i
(3.36)
s
2
a=1
n

xi X 1
=
ha Na V3ia
t
2
a=1

By collecting the derivatives of the global positions from (3.36) in a matrix the Jacobian matrix,
J, is formed:
x y z
r

J = x
s
x

r
y
s
y
t

r
z
s
z
t

(3.37)
t
Na,r xa + t ha Na,r V a Na,r ya + t ha Na,r V a Na,r za + t ha Na,r V a
31
32
33


2
2
2


= Na,s xa + 2t ha Na,s V31a Na,s ya + 2t ha Na,s V32a Na,s za + 2t ha Na,s V33a


1
1
1


h N Va
h N Va
h N Va
2 a a 31
2 a a 32
2 a a 33

3. Continuum-Based Shell Elements

52

The last right-hand term of (3.34) may then be determined as the inverse terms of the Jacobian.
Thereby, all quantities in (3.34) are determined and the derivatives of the displacements with
respect to the global coordinate can be found. Practically, this is done by multiplying the inverse
Jacobian matrix with the derivative of the displacements in (3.35). Thereby an expression for
the derivatives of the displacements with respect to the global coordinates is obtained:
a
ui

ui
Na,x 1i a1 2i a1
n
x
X

ui
a
a
(3.38)
y =
Na,y 1i 2 2i 2 a
a
a
ui
a=1 N
a
a,z 1i 3 2i 3
z

1
1
where we for brevity define Na,xi = Ji1
Na,r + Ji2
Na,s , 1 = 12 ha V2a , 2 = 21 ha V1a and
1
1
1
a
i = t(Ji1 Na,r + Ji2 Na,s ) + Ji3 Na . By inserting the displacement derivatives from (3.38)
into the strain definition in (3.33), the linear strain vector can be expressed as e = B0 ua , where
the linear strain-displacement matrix, B0 , is given as :



Na,x
0
0
11 a1
21 a1



0 N
a
a
0
12 2
22 2


a,y


a
a

0
23 3
0
Na,z
13 3

(3.39)
B0 =

a
a
a
a
+

N
N
0

22 1
21 2
12 1
a,x
11 2

a,y


Na,z Na,y 12 a3 + 13 a2 22 a3 + 23 a2
0


Na,z
0 Na,x 13 a1 + 11 a3 23 a1 + 21 a1

As can be seen the expression in (3.39) is independent of the displacements. The straindisplacement matrix is essential for setting up the stiffness matrix of the element. So is the
constitutive matrix, which will be developed in the following.

3.7

Constitutive relation

Stresses are related to strains and hence displacements through a constitutive relation which for
linear material behavior is Hookes general law. In index notation this constitutive relation may
be expressed generally by noting from the outer product rule (see e.g. Heinbockel (1996)) that
two second-order tensors can be related through a fourth-order tensor as:
ij = Cijkl kl

(3.40)

where the elastic coefficient tensor (or constitutive matrix in algebraic notation), C ijkl , is to be
determined. If the material considered is heterogenous the elastic coefficient tensor, C ijkl , is
dependent of the position and the elements of the tensor will be functions of the coordinates,
xi . However, since homogeneity is assumed for all materials under investigation, C ijkl , is independent of the position and the elements of the tensor are scalars. Consequently, since a fourth
order tensor consists of 34 = 81 components, 81 equations in 81 unknowns must generally be
solved to determine the relation in (3.40). However, applying various symmetry considerations
(see e.g. Timoshenko and Goodier (1970)) the relation in (3.40) can be reduced to an equation

Notes on Structural Analysis of Composite Shell Structures

53

in 21 unknowns:

C11 C12 C13
11
22
C22 C23

33
C33
=
12

23
Sym.
13

C14
C24
C34
C44

C15
C25
C35
C45
C55


11
C16
22
C26


C36
33

C46
12
C56 23
13
C66

(3.41)

This is the most general form of the so-called constitutive matrix, C. By assuming e.g. isotropic
or orthotropic material behavior the number of constants in (3.41) can be further reduced.
First, the case of constitutive relations for a single layer is considered.
3.7.1 Single layers
The constitutive matrix presented in (3.41) above is for the general case of an anisotropic material where no symmetry between the material properties exist. As can be seen from the constitutive matrix in (3.41) coupling exist between each stress component and all the strain components. This means that under any prescribed load, the material will deform both by extension
and shear deformation in all planes. However, in general engineering we often restrict ourself
to isotropic or orthotropic materials, which are special cases of the anisotropic material. These
materials exhibit "simpler" response when loaded due to various degrees of symmetry in the
material properties (fewer non-zero values in C).
To investigate the symmetry of material properties we introduce the plane of elastic symmetry
as done by Ochoa and Reddy (1992): if a number of coordinate systems are defined as mirror
images of each other in a certain plane, and the coefficients of C in each of these coordinate
system have the same values, then the plane is called a plane of elastic symmetry. If the material
has one plane of elastic symmetry, the material is called monoclinic or aelotropic. Consider
such a material having the (x y)-plane as an elastic plane of symmetry. For this material it is
possible to define an auxiliary coordinate system (
x, y, z) having the following properties:
x = x

y = y

z = z

(3.42)

The auxiliary coordinate system is a mirror image of the (x, y, z) system in the (x y)-plane.
Since the material is monoclinic, it should be possible to establish Hookes generalized law for
both the (
x, y, z) and (x, y, z) coordinate system with the same constitutive matrix. To establish
Hookes law the strain and stress tensors are first expressed in the (
x, y, z) system and are then
transformed to the (x, y, z) system using a tensor transformation for Cartesian bases.






xi
xj
xi
xj
mn = ij ei
ej

mn = ij ei
ej
(3.43)
xm
xn
xm
xn
Through (3.42) the components of the derivatives in (3.43) can be determined; only the nonzero
components are shown in (3.44):
x
=1
x

y
=1
y

z
= 1
z

(3.44)

3. Continuum-Based Shell Elements

54

By inserting the components found in (3.44) into (3.43), the stress and strain in the auxiliary
coordinate system can be determined.

11

22

33

12

23

13

= 11
= 22
= 33
= 12
= 23
= 13

11
22
33
12
23
13

= 11
= 22
= 33
= 12
= 23
= 13

Inserting the stresses and strains from (3.45) into Hookes law yields:


11
C11 C12 C13 C14 C15 C16
11

22
C22 C23 C24 C25 C26

22

33
C33 C34 C35 C36

33

12 =

C44 C45 C46 12


23
Sym.
C55 C56 23
13
C66
13

(3.45)

(3.46)

If the constitutive matrix in (3.46) has to be identical to the one for the (x, y, z) coordinate
system C15 , C16 , C25 , C26 , C35 , C36 , C45 , C46 needs to be zero. Thereby the constitutive matrix
for a monoclinic material will be of the form:

C11 C12 C13 C14 0


0

C22 C23 C24 0


0

C33 C34 0
0

(3.47)
C=

C
0
0
44

Sym.
C55 C56
C66
From (3.47) it can be seen that the number of independent constants in C have been reduced
from 21 to 13.

If three mutually orthogonal planes of elastic symmetry exist in the material, it is called orthotropic. For the orthotropic material, it is possible to define an additional auxiliary coordinate
system, giving 2 mirrored coordinate systems in all. The new coordinate system can be a mirror
image of the (x, y, z) coordinate system in e.g. the (y z)-plane and in that case given as:
x = x

y=y

z=z

(3.48)

Performing similar manipulations as those done for the monoclinic material earlier the constitutive matrix for an orthotropic material appears in the from:

C11 C12 C13 0


0
0

C22 C23 0
0
0

C
0
0
0
33

(3.49)
C=

C44 0
0

Sym.
C55 0
C66

Notes on Structural Analysis of Composite Shell Structures

55

As can be seen from (3.49) the number of independent constants has now been further reduced
to 9. The coefficients in (3.49) are quite abstract and do not hold any real physical meaning.
Therefore, they are expressed through the engineering material constants E i , ij and Gij which
are well-defined and can be measured through laboratory tests. The relation between the elastic
coefficients and the engineering constants has been shown by e.g. Grdal et al. (1999) to be:
E1 (1 23 32 )

E2 (12 + 13 32 )
=

E3 (13 + 12 23 )
=

= G12

E1 (21 + 23 31 )

E2 (1 13 31 )
=

E3 (23 + 13 21 )
=

= G23

E1 (31 + 21 32 )

E2 (32 + 12 31 )
=

E3 (1 12 21 )
=

= G13
(3.50)

C11 =

C12 =

C13 =

C21

C22

C23

C31
C44

C32
C55

C33
C66

The operator in (3.50) is defined as = 1 12 21 13 31 12 23 31 13 21 32 23 32


where the indices refer to the directions shown in Fig. 3.13. From (3.50) it appears that there
are more than 9 independent constants. However, it can be shown that ij /Ei = ji /Ej , and
thus that the number of independent constants is indeed 9.
When describing materials with additional material symmetries we define an axis of symmetry which is the axis about which the material properties are the same. If a material has two
axes of symmetry and three planes of elastic symmetry the material properties are the same
in all directions and the material is defined as isotropic. In order to develop an expression for
the constitutive matrix of an isotropic material, it is assumed that the x and y axis are axes
of symmetry. Hence, two auxiliary coordinate systems are established as a mirror image of
the (x, y, z) coordinate system around these two axes respectively, leading to two additional
auxiliary coordinates systems:
x =
x
x = z

y = z
y= y

z = y
z=
x

(3.51)

Through similar manipulations as for the monoclinic and orthotropic cases, the constitutive
matrix can be reduced to:

C11 C12 C12 0


0
0

C22 C12 0
0
0

C
0
0
0
33

(3.52)
C=

C44 0
0

Sym.
C44 0
C44

At first glance it appears that the matrix C in (3.52) is dependent on 5 elastic constants. However, it can be shown through thermodynamical considerations as done by Sokolnikoff (1956)
that the constants are mutually dependent and that the number of independent constants is only
2. The elastic coefficients in (3.52) can then be defined as done by e.g. Grdal et al. (1999):
C11 = C22 = C33 =

E(1 )
(1 + )(1 2)

E
(1 + )(1 2)
C11 C12
=
2

(3.53)

C12 =

(3.54)

C44

(3.55)

3. Continuum-Based Shell Elements

56

Now, three types of constitutive relations have been developed for single layers, namely monoclinic, orthotropic and isotropic materials. In the following a short discussion is given of which
relations are suitable for modelling what kind of materials.
3.7.2 Modelling materials using constitutive relations
The concept of lamina or layers is almost always associated with the analysis of composite structures. However, in general such layers could be anything including the traditional engineering
materials such as steel and aluminium. These materials behave isotropic and consequently, the
isotropic constitutive relation in (3.52) is incorporated in the later formulation of shell elements
in order to model such materials.
To access how to model composite lamina we consider the lamina in Fig. 3.13 where the fibers
are assumed to be parallel.
3

Figure 3.13 Fibers in a lamina and the associated material directions.

The 1, 2 and 3 direction in Fig. 3.13 are commonly known as the material directions. The
material properties in the fiber directions will be the same everywhere in the lamina since the
fibers are assumed to be parallel and similarly, the material properties transverse to the fibers
will be the same everywhere. This means that planes of elastic symmetry exist in the (1 2)plane, the (2 3)-plane and the (1 3)-plane respectively, as can be seen from Fig. 3.13. Hence
the lamina can be assumed to be orthotropic in accordance with the definition of orthotropy
stated in Chapter 2.
Composite lamina may also exhibit near-isotropic behavior in some cases. When e.g. using
composite lamina made from chopped mat where the fibers are short and oriented randomly,
isotropic properties are often associated with the lamina. Furthermore, when dealing with sandwich structures it is possible to encounter core materials with isotropic behavior. Both types of
lamina are modelled as being isotropic and hence, the constitutive relation (3.52) is also used
for composite laminates as stated previously.
So far materials has been treated as single layers. In the following we extend the description to
include stacks of lamina.
3.7.3 Multiple layers Composite laminates and sandwich structures
When forming a laminate of single layers it is assumed that the plies are perfectly bonded together. In turn, the laminate will be assumed to deform as a whole with no discontinuities in
the geometry. Two types of stacked lamina will be treated; composite laminates and sandwich
structures. In principle the two are very much alike, but the response of the two is very different and consequently, the modelling of composite and sandwich structures must be separated.

Notes on Structural Analysis of Composite Shell Structures

57

Common to both is the procedure of assembling the global constitutive matrix, as explained in
the following.
As described in the preceding sections orthotropic lamina exert no coupling effects when loaded
in the material directions. However, depending on the orientation of the global coordinate
system coupling effects are introduced, which is especially true for multiple layers as explained
in the following for orthotropic plies.
When the global coordinate system is aligned with the material directions, the material behavior can be described through the orthotropic constitutive matrix (3.49) resulting in no coupling
between the normal stresses and shear strains, the shear stresses and normal strains and the
shear stresses in one plane and the shear strains in other planes. However, if the global coordinate system is not aligned with the material directions, the constitutive matrix needs to be
transformed into the global coordinate system since the stiffness matrix is always expressed
globally. This transformation is carried out through the general tensor transformation rule (see
e.g. Heinbockel (1996)) and consequently, the transformed constitutive matrix will in general
have no zero-valued elements. Thus, the material in the global coordinate system will behave as
an anisotropic material with couplings between all stress and strain components as also shown
in (2.17).
From the preceding discussion it seems reasonable to assume that laminates in general must
have anisotropic behavior since it is unlikely that all layers in the laminate will be aligned with
the global coordinate system. Consequently, the global constitutive matrix for the laminate must
be anisotropic and contain the coupling effects of all lamina in the laminate. This leads to the
conclusion that the global constitutive matrix is dependent on the thickness coordinate, t, so
that C = C(h1 , h2 , h3 ) since the different coupling effects are dependent on the thickness to
a different power (from 0 to 2 in local coordinates). The same conclusion could be drawn by
regarding the A-B-D method described in Chapter 2 (see (2.38)). Each of the matrices A, B
and D are thickness dependent to a different power of h due to the explicit integration carried
out to determine the force and moment resultants. Since the constitutive matrix, C, in a sense
contains both A, B and D it leads to the same conclusion as before. This thickness dependency
involves a number of problems when determining the global constitutive relations since explicit
integration cannot be carried out for C.
The classical way to solve the problem of thickness integration and determine the behavior of a
laminate is to use the A-B-D approach as already mentioned. The methods yields three constitutive matrices for each ply, each of which are well-defined in terms of thickness dependency.
Thus the stiffness matrix can be found globally in one step. An alternative way of determining
the global behavior of the laminate is to use the constitutive matrix, C, for each lamina directly
in the stiffness matrix. This means however, that it is necessary to determine the stiffness matrix for each lamina in the laminate and collect the element stiffness matrix in several steps. We
choose to employ the latter approach since we believe that it constitutes a more general framework than the A-B-D approach. Consequently, exact numerical integration is carried out over
the thickness and the formulation can easily be extended to solid elements.
The procedure described above is sufficient for determining the behavior of composite laminates, but experience shows that special care must be taken when dealing with sandwich structures. The reason is that sandwich structures carry transverse shear loads differently from composites and hence, the composite model yields poor results for sandwich structures. Conse-

3. Continuum-Based Shell Elements

58

quently, an adjustment of the composite model must be made for the analysis of sandwich
structures. To this end we consider the transverse shear load carrying characteristic of the sandwich plate under transverse normal load in Fig. 3.14 (a).

(a)

(b)

Figure 3.14 Load carrying characteristic of (a) a sandwich plate under transverse normal load and (b)
spring analogy for the transverse shear modulus

The load carrying mechanism of the sandwich plate can be evaluated by considering a spring
stiffness analogy. If the stiffness of each ply is modelled by a spring having a stiffness of
Gply hply the total stiffness can be evaluated by regarding the sandwich as a series of springs as
shown in Fig. 3.14 (b). Thereby, the total stiffness is obtained from the reciprocal rule:
1
Gtotal h

1
1
1
+
+
Gskin hskin Gcore hcore Gskin hskin

(3.56)

Considering (3.56) we note that the product Gskin hskin will in general be greater by several
orders of magnitude than Gcore hcore since Gcore  Gskin and hskin < hcore . Consequently,
it is reasonable to assume that the core carries all the transverse shearing stresses. Therefore,
the transverse shear moduli of the skins are set to zero in the constitutive relation, i.e. G 23 =
G13 = 0. This in experience leads to very good results and the approach has been implemented
as a sandwich input option for all shell elements in MUST. However, if the stiffness of the
skins becomes comparable in size to that of the core, the assumption is invalid and would
lead to erroneous results. To avoid such cases a condition has been included in the element
formulation, stating that hcore /h 5/6 when using the sandwich option. This relationship has
been adopted from the commercial software package Ansys. It is not a fool-proof condition
since the assumption is violated if the core/skin stiffness-ratio becomes large (maybe in the
order of 103 or so). In extreme cases the sandwich option can be disabled in the analysis input
whereby all stiffness contributions are included in the constitutive relation as for composite
laminates.
3.7.4 Other properties of the constitutive relation
In degenerated elements the constitutive relation is used not only to describe the material behavior but also for enforcing the Mindlin shell assumption of negligible transverse effects. As
explained briefly earlier one of the Mindlin assumptions states that the midplane normal is inextensible, i.e. there is no transverse normal strain, i.e. 33 = 0. Furthermore, it was assumed
that the transverse normal stress is zero since the shell is thin to moderately thick. However,
using the "full" constitutive relations would lead to transverse normal stresses due to Poisson
effects. This problem can be circumvented by forcing 33 = 0 in the constitutive relation which
is demonstrated in the following.

Notes on Structural Analysis of Composite Shell Structures

59

First, we obtain an expression for the transverse normal stress, 3 , by writing out the product of
the strain vector and the third row of the constitutive matrix in (3.41):
3 = C31 1 + C32 2 + C33 3 + C34 4 + C35 5 + C36 6 = 0

(3.57)

From (3.57) the transverse normal strain can be isolated, yielding an expression for  3 :
3 =

C3i i
C33

(3.58)

where i = 1, 2, 4, 5, 6. Inserting (3.58) in the constitutive relation (3.41) results in the following

expression for the coefficients in the modified constitutive matrix, C:


Ci3 C3j
Cij = Cij
C33

(3.59)

where i = 1, 2, 3, 4, 5, 6 and j = 1, 2, 4, 5, 6. Using (3.59) on the orthotropic constitutive matrix


(3.49) yields the following matrix:

C11 C12 0 0
0
0

C22 0 0
0
0

0 0
0
0

C=
(3.60)

C
0
0
44

Sym.
C55 0
C66
where the coefficients Cij are given as:

E2
21 E1
E1
C22 =
C12 =
1 12 21
1 12 21
1 12 21
= G12
C55 = G23
C66 = G13

C11 =
C44

(3.61)

Similar derivations can be made for the isotropic constitutive relation whereby the coefficients
in the matrix in (3.49) are given as:
C11 = C22 =

E
1 2
E
=
2(1 )

C12 =
C44

E
1 2

(3.62)

The coefficients in (3.61) and (3.62) are those used when formulating the constitutive relations
for the shell elements since they contain the Mindlin assumption.
A further use of the constitutive relations is to correct the transverse shear stresses. If the
transverse shear stresses are computed using three dimensional theory of elasticity, it can be
shown that they will vary quadratically through the thickness. Computing 13 and 23 using the
displacement interpolations in (3.27) results in constant through-thickness stresses. In order to
obtain the same strain energy for the three dimensional solution and the shell element solution,
the transverse shear stresses are corrected with the so-called shear correction factor, k, which
is done by multiplying C55 and C66 with k.

3. Continuum-Based Shell Elements

60

The shear correction factor is not constant but dependent on geometry, the loading situation, the
material properties and the boundary conditions as stated by e.g. Ochoa and Reddy (1992). In
the isotropic case the factor does not change significantly and it is therefore common practice to
use the factor obtained for a rectangular plate as done by e.g. Zienkiewicz and Taylor (2000b),
namely k = 0.8. In the case of laminated structures the factor changes considerably when
one or more of the aforementioned conditions are altered and consequently, it is necessary to
recalculate the factor repeatedly. Several methods have been proposed for determining the correct factor, k, among others can be mentioned the predictor-corrector method developed by
Noor et al. (1990). None of these methods have been implemented in the element formulation
and the discussion will therefore not be extended further. Instead of calculating a shear correction factor for each laminated structure considered, it is chosen to use k = 0.8 for ordinary
laminates and k = 1.0 for sandwich shells. This strategy was suggested by Professor Jorn S.
Hansen, The University of Toronto, who has frequently published on the subject of modelling
sandwich structures (see e.g. Oskooei and Hansen (2000)).
The constitutive relation has now been fully developed but in order to use the matrix for determining K it must be transformed to the global coordinate system.
3.7.5 Transformation of the constitutive matrix
For isotropic materials the constitutive matrix (3.62) is defined in the element plane and is
independent of orientation. Hence, the global constitutive matrix for an element can be obtained
from the material coordinate system to the global C in the (x, y, z)
by transforming the local C
system. The transformation of the constitutive matrix is made using the transformation C =
as done by e.g. Cook et al. (1989). The transformation matrix T is defined from the
TT CT
material base vectors, m1 , m2 and m3 and the global base vectors, i, j and k:

2
a1
b21
c21
a 1 b1
b1 c1
c1 a 1

a22
a 2 b2
b2 c2
c2 a 2
c22
b22

2
2
2

a3
a
b
b
c
c
a
c
b
3
3
3
3
3
3
3
3

(3.63)
T=
2a1 a2 2b1 b2 2c1 c2 a1 b2 + a2 b1 b1 c2 + b2 c1 c1 a2 + c2 a1

2a2 a3 2b2 b3 2c2 c3 a2 b3 + a3 b2 b2 c3 + b3 c2 c2 a3 + c3 a2


2a1 a3 2b1 b3 2c1 c3 a1 b3 + a3 b1 b1 c3 + b3 c1 c1 a3 + c3 a1
where ai , bi and ci are defined as:

a1 = cos(i, m1 ) b1 = cos(j, m1 ) c1 = cos(k, m1 )


a2 = cos(i, m2 ) b2 = cos(j, m2 ) c2 = cos(k, m2 )
a3 = cos(i, m3 ) b3 = cos(j, m3 ) c3 = cos(k, m3 )

(3.64)

For orthotropic materials the constitutive matrix (3.61) is dependent of orientation and by definition expressed in the principal material directions. Consequently, the transformation to the
global coordinate system must be made from the principal material directions and not the lamina
coordinate system. However, since both T and the in-plane rotation of the material coordinate
system are already defined it is convenient to use these transformations. To do so the transfor is rotated from the material directions to the
mation is carried out in two steps: in the first C
material coordinate system and in step two T transforms to the global coordinate system. The

Notes on Structural Analysis of Composite Shell Structures

is carried out using T which is a simplified form of T:


rotation of C
2

a
b2 0
0
0 0
b2
a2 0
0
0 0

0
0 1
0
0 0

T =
2
2

2ab
2ab
0
a

b
0
0

0
0 0
0
a b
b2
a2 0
0
b a

where a and b is given as

a = cos()

b = sin()

61

(3.65)

(3.66)

Using the relations in (3.63) and (3.65) it is possible to find the constitutive matrix in the global
and C = TT CT

(x, y, z)-system. In general, the two successive transformations C = T T CT

will produce a global constitutive matrix with non-zero values in all elements of the matrix.
This elaborates the statement made previously that almost all laminates will exhibit anisotropic
behavior when loaded.
Using the constitutive matrix and the strain-displacement matrix developed earlier we may assemble the element stiffness matrix by employing the expression K = B T CB. For the sake
of brevity the complete stiffness matrix is not stated. In order to assemble the stiffness matrix,
which is an integral expression (see Section (3.1)), we must perform numerical integration.

3.8

Numerical integration

The numerical integration is performed using the Gauss quadrature both in-plane and out-ofplane. First, let us briefly revise the Gauss quadrature.
The Gauss quadrature method approximates an integral as a summation over the evaluation
points:
Z 1Z 1Z 1
X
(Ti (rj ))|J| drdsdt =
(ri , sj , tk )|J|Wi Wj Wk
(3.67)
1

i,j,k

The evaluation points, ri , sj and tk , are located symmetrically around the midpoint of the integration interval, and symmetrically paired evaluation points have the same weight factor, W . It
has been shown by Gauss that the location of the evaluation points is determined as roots in the
Legendre polynomial, (see e.g. Bathe (1996)). Furthermore the expression in (3.67) is obtained
by successive use of a 1-dimensional quadrature in each of the directions r, s and t. Tabulations of Gauss-point locations and Gauss weight factors can be found in e.g. Bathe (1996) and
Cook et al. (1989). When Gauss quadrature is used in two or three dimensions it is most common to use the same order of integration in all directions, but different orders may be used in
different directions.
3.8.1 Numerical thickness integration of laminates
For plane elements the through-the-thickness integration is normally performed explicitly by
multiplying by the thickness of the element which is valid since both the strain-displacement
matrix, B, and the Jacobian matrix, J, are independent of the thickness
R coordinate,R t. Consequently, they integrate as constants with regard to the thickness, i.e. Bdt = B dt = Bh
where h is the shell thickness. In the present element formulation, the through-the-thickness

3. Continuum-Based Shell Elements

62

integration cannot be performed explicitly since both the strain-displacement matrix and the Jacobian matrix are dependent of the thickness coordinate as shown in (3.39) and (3.37). Therefore integration in the thickness direction is performed numerically using the Gauss quadrature method in (3.67). For laminated elements the thickness integration can not be performed
through the application of (3.67) alone as explained in the following.
For laminated elements, where the material has orthotropic properties, different couplings such
as extension-bending, extension-shear, shear-bending and others are introduced when the structure is not loaded in the principal material directions as discussed in Chapter 2. The behavior
resulting from these coupling effects depends on the stacking sequence of the layers, the fiber
orientation in the layers and the loading conditions. If the integration in the transverse direction
is performed as one integration as for isotropic shells the coupling effects will not be correctly
included in the governing equations and thus, the response of the model would not correlate
with the response of the real structure. A more thorough treatment of the different coupling
effects may be found in e.g. Grdal et al. (1999) or Jones (1999).
The new integration scheme is set up as a piecewise through-the-thickness numerical integration, which seems natural when considering a laminated element as shown Fig. 3.15.

t=1
t=0
t = -1

Figure 3.15 A laminated element shown with four layers of equal thickness.

The constitutive matrix will in general be different for each layer and in effect vary as a function of the thickness in global terms as explained in Section 3.7. The different stresses vary
differently with respect to the thickness why we in the previous chapter justified the notation
C = C(h1 , h2 , h3 ) when regarding the entire element.
This property constitutes the fundamental problem of finding the constitutive matrix for an
entire laminate. In classical laminate theory this problem is solved by using resultant formulation and dividing C into several matrices (the A-B-D approach described in Chapter 2). In
the present study the integration of the laminated element is instead performed as a full three
dimensional integration in each layer, meaning integration is performed over r, s and t for all
layers. Consequently, the individual layers are treated as "sub-elements" of the laminated element and thus, the stiffness of the entire element is determined as the sum of the layer stiffness
contributions. This is treated extensively in e.g. Panda and Natarajan (1981). By performing the
integrating of each layer separately in this manner the integration limits for the thickness integration will be dependent on the set-up of the laminate (number of layers and layer thicknesses).
Hence, to use the Gauss quadrature coefficients the integration limits have to be changed since
integrations must be carried out over the interval [1; 1] when using Gauss quadrature. This is
solved by transforming the transverse coordinate t into tl so that tl in the lth layer varies from
1 to 1. This approach is shown in Fig. 3.16.
The transformation of the natural coordinate, t, to a layer coordinate, t l , is performed through

Notes on Structural Analysis of Composite Shell Structures

63

t
tl

n
n-1

l'th layer

hl

2
-1

-1

(a)

(b)

Figure 3.16 Schematic representation of the integration scheme for laminated elements where (a) shows
the entire element and (b) the lth sub-layer where the transformed coordinate t l runs from 1 to 1.

the transformation:
1
t = 1 +
h

l
X
i=1

hi hl (1 tl )

(3.68)

where hl is the thickness of the lth layer and the sum of the first term of the parenthesis constitutes the total thickness of the "preceding sub-elements". Inserting the modified coordinate in
the Gauss integration represents an integration by substitution and consequently, the derivative
of the coordinate transformation in (3.68) must be found and multiplied to the integrand. The
derivative of (3.68) is found as:
dt(tl )
hl
=
dtl
h

dt =

hl
dtl
h

The Gauss quadrature for laminated elements is thus stated as:


Z 1Z 1Z 1
X
hl
(Ti (rl ))|J| drdsdtl =
(ri , sj , tk ) |J|Wi Wj Wk
h
1 1 1
i,j,k

(3.69)

(3.70)

where tk is found from (3.68). It is obvious that the piecewise through-the-thickness integration
of laminated elements makes them computationally expensive and hence, a lot of research has
been made on how to increase the effectiveness of the integration. In e.g. Kumar and Palaninathan (1997)
different improved integration schemes for the piecewise through-the-thickness integration is
discussed but none of these improved schemes have been implemented for the laminated shell
elements presented here.
In principle the element formulation is completely described from the preceding sections. However, the resulting element (a so-called purely displacement-based element) will suffer heavily
from locking, which is described and addressed in the following.

3.9

Locking problems

When displacement based elements are formulated the decisive and crucial choice is the choice
of interpolation functions (as described earlier). If the interpolation chosen can not model the
behavior of the real structure sufficiently accurate, the response of the model may not only be
inaccurate but incorrect. Locking occurs when the finite element model becomes overly stiff in
comparison with the real structure. Locking may occur in various loading situations and several

64

3. Continuum-Based Shell Elements

locking problems may arise simultaneously or single phenomenon may appear isolated. This
renders the identification of locking very difficult for an engineer reviewing the results of a finite
element analysis. Therefore, the need arises for stable, locking free elements.
The problems and solutions of locking are treated in general terms by Cook et al. (1989) and
in depth by e.g. Bathe (1996), Belytschko et al. (2000) or Bischoff (1999). In the following a
brief introduction will be given to three locking phenomena and possible solution strategies will
be presented.
3.9.1 Membrane locking
Membrane locking is only a problem in curved beam and shell elements, (see e.g. Bischoff (1999)).
If considering a shell, the membrane stiffness is very large compared to the bending stiffness
and the shell may therefore bend without stretching. This is called inextensional bending and
membrane locking is caused by the elements inability to represent this type of bending. This
means that when the element is loaded in pure bending parasitic membrane stresses may be
introduced whereby the elements exerts overly stiff behavior. Since membrane locking only
occurs in curved elements (such as 6,- 8- or 9-node elements) it is not necessary to address the
problem when formulating 3- and 4-node elements since these elements are plane. To make
curved elements reliable it is therefore necessary to consider membrane locking when formulating such elements.
The solution to membrane locking is to modify the interpolation of the in-plane stresses. These
modifications can be made in various ways such as the Enhanced Assumed Strain-method
(EAS) or by MITC.
3.9.2 Shear locking
Shear locking is a problem for many types of elements. Shear locking is a result of parasitic
transverse shear stresses introduced when the element is loaded in bending. In a state of pure
bending, resulting from application of a constant moment, M , the shear force is the derivative
of the moment, i.e. V = M/xi . Hence, for constant moment the shear force is zero. Using
a Timoshenko beam analogy (which is reasonable since the Mindlin assumptions for shells
constitute the same restrictions as made by Timoshenko for beams) the shear stress is given
from the shear formula, stated here as 13 = V , where is a geometric quantity. Consequently,
the shearing stresses must be zero since the shear force is zero for constant bending moment.
Shear locking occurs when the element is unable to calculate zero transverse shearing stresses
in pure bending. It may also arise when applying a linearly varying moment to an element if the
element in this case is unable to determine constant transverse shearing stresses. The former is
a problem in first-order elements and the latter in second-order elements; this is explained in
the following.
For the linear elements shear locking arises since the elements are incapable of representing
the deformation caused by pure bending. This is shown in Fig. 3.17 where an edge view of a
deformed linear element is shown in Fig. 3.17a and a correct deformed geometry is shown in
Fig. 3.17b.
If instead considering the quadratic elements they can, because of the mid-side node, properly
represent the deformation caused by pure bending. In the quadratic elements the parasitic transverse shear stresses instead arises when the elements are loaded under linearly varying bending.

Notes on Structural Analysis of Composite Shell Structures

M1

M1

65

M1

M1

(b)

(a)
x

Figure 3.17 Edge view of (a) deformed linear element, (b) correct deformed geometry in pure bending

Since shear locking is a problem in many elements a lot of different methods have been developed to eliminate shear locking. For all the improvement methods shear locking is avoided by
modifying the determination of the transverse shear stresses. This means that the last two rows
in the strain-displacement matrix, B, are modified.
3.9.3 Volumetric and thickness locking
Volumetric locking is different from other locking problems in that volumetric locking is caused
by a material parameter namely the Poisson ratio, . Volumetric locking appears for isotropic
materials when the material becomes/are incompressible or nearly incompressible meaning
0.5. This may be realized from Hookes generalized law rewritten from (3.40) for isotropic
materials:
ij = ij kk + 2ij
(3.71)
where and are Lames constants given as:
=

E
(1 + )(1 2)

E
2(1 + )

(3.72)

If considering it is seen that when 0.5 then . This means that some of the terms
in the constitutive matrix tend towards infinity and in turn, terms in the stiffness matrix tend
towards infinity. If considering the modified constitutive matrix in (3.62) it can be seen that
none of the terms reaches infinity when 0.5. The reason is that the modified constitutive
matrix has been derived for reduced state of stress for which the term (1 2) vanishes.
For orthotropic materials it may be seen from (3.61) that terms in the constitutive matrix may
reach infinite if 1. If considering a single layer (stacks of layers may have entirely different
material properties when viewed as a whole) it seems reasonable to conclude that for common
engineering materials and the case of linearly elastic materials this will never be the case. For
this reason volumetric locking does not represent a problem in the formulated elements and
hence, this locking phenomenon is not treated any further.
Besides the three locking phenomena explained in the preceding two different thickness locking
phenomena exist, namely thickness locking and curvature thickness locking. Both of these
locking effects are caused by a parasitic transverse normal stress and therefore only appear when
thickness changes are taken into account in the shell formulation. This means that for the five
parameter formulation (the Mindlin assumption) used in the implemented shell elements these
two types of locking will not come into effect, (see e.g. Bletzinger et al. (2000)). Consequently,
no strives towards eliminating such locking problems are taken in the formulation of the shell
elements.

3. Continuum-Based Shell Elements

66

It follows from the discussion of locking problems that only shear locking needs to be countered
when implementing plane elements. To this end the MITC approach will be utilized.

3.10

Mixed Interpolation of Tensorial Components MITC

The MITC method is based on a change of strain component interpolations instead of a revised
integration scheme as for selective integration. Consequently, the MITC approach uses full
integration for all strain components but a new set of interpolation expressions. The MITC
method was originally proposed for bilinear elements with a new transverse shear interpolation
(the MITC4 element formulated by Dvorkin and Bathe (1984) for eliminating shear locking),
but has later been extended to higher order elements by e.g. Bucalem and Bathe (1993). This
has lead to the MITCn-family of elements including triangular and quadrilateral elements with
n number of nodes as suggested by e.g. Bathe (1996), in which new interpolations are used
for all strain components in order to simultaneously circumvent the problems of shear and
membrane locking.
In general the "new" assumed strains (AS) will be expressed from the "old" or "directly interpolated" (DI) strains through a new set of interpolation functions as:
AS
ij

p
X

Nkij (r, s)
DI
ij (rp , sp , t)

(3.73)

k=1

where Nkij are the new interpolation functions in r and s corresponding to the ijth strain and the
kth tying point. Note that the directly interpolated strains in (3.73) constitute those calculated
from a standard finite element formulation as the derivatives of the displacements. The new
interpolations must naturally fulfill the relation:
Nkij |l Nkij (rl , sl ) = kl

(3.74)

so that the kth interpolation function assumes the value 1 in the kth tying point and the value
0 in all other tying points. The new interpolations will generally be of the same order as the
standard isoparametric interpolation functions used in the element, i.e. in the case of the bilinear
element we use linear interpolations for the assumed strains.
The expression in (3.73) is the key assumption of the MITC method. The crucial point for the
success of the method is the choice of tying points, (rp , sp , t), from which the strains are interpolated. The tying points are chosen in the reference plane and the t-coordinate thus remains
independent. It follows from (3.73) that the assumed strains and the directly interpolated strains
are equal at the tying points. The location of the tying points in the natural coordinate system
is given by Bathe (1996), Bucalem and Bathe (1993) and Dvorkin and Bathe (1984) for most
types of MITC elements. The justification of the tying point coordinates is not given as a continuum mechanical explanation but is based on a numerical verification and the same approach
will be taken here. In the following the derivation of the MITC interpolation expressions will
be limited to the 4-node bilinear element; for this case the tying point are given by Bathe (1996)
as in Fig. 3.18
The tying points, A and C, in Fig. 3.18 are used for the interpolation of AS
13 and the points, B
AS
and D, are used for 23 and hence, the number of tying points, p, "per strain" is 2. The superscript "AS" will be omitted in the following and it will be understood that the transverse shear
strains are the new, interpolated strains.

Notes on Structural Analysis of Composite Shell Structures

67

s
r

Figure 3.18 Location of tying points for a MITC4 element; A = (0, 1), B = (1, 0), C = (0, 1) and
D = (1, 0) in natural coordinates

To derive the interpolation expressions for the MITC4 element we start by rewriting (3.73) for
a 4-node element by defining a new set of interpolation functions. Using the relation (3.74) a
new set of linear interpolation functions for the four tying points are written tentatively as:
1
NA13 = (1 s)
2
1
NB23 = (1 + r)
2

1
NC13 = (1 + s)
2
1
ND23 = (1 r)
2

(3.75)

Using the interpolations of (3.75) in the general MITC expression (3.73) yields the new assumed
strains:
1
A
13 = (1 s)
13 +
2
1
23 = (1 r)
D
23 +
2

1
(1 + s)
C
13
2
1
(1 + r)
B
23
2

(3.76)

This corresponds to a linear interpolation between the directly interpolated strains at points A
C and D B, respectively. The directly interpolated transverse strains are thus replaced by two
assumed strains expressed in natural coordinates. To form the strain-displacement matrix these
local strains must be transformed to global coordinates as described in e.g. Heinbockel (1996).
This is done using the standard tensor transformation:
ij = kl (gk ei )(gl ej )

(3.77)

where ei are the global base vectors. For the special case of a plate, most of the terms in (3.77)
vanish but for the general case all nine components contribute to the global strain. The new
strain-displacement matrix which incorporates the MITC method can be derived in a straightforward (however tiresome) manner, but the derivation is omitted at present.
The MITC 4-node element has been implemented in MUST and will be employed in the exercises following lecture 3.

68

3. Continuum-Based Shell Elements

Chapter

Geometrical Non-Linear Analysis of Laminated Shell


Structures

his chapter deals with an extension of the theory presented in Chapter 3 which was limited
to geometrically linear structures (small displacements). In the following we develop first
the governing equations for a non-linear problem and subsequently extend the derivation of
element matrices to include the non-linear contributions.

As before we start by reviewing the governing equations.

4.1

Governing equations

The starting point of the previous chapter was a equilibrium consideration resulting in the governing equation 3.16, stated again for reference:
nel
X
k=1

k =

nel
X

rk

(4.1)

k=1

which is the general system of governing equations for both linear and non-linear analysis.
In the previous chapter we proceeded by assuming linear strain but now, we shall include the
non-linear terms as well and develop the governing equation from the full strain tensor:


u
u u
1 u
+
+
(4.2)
ij =
2 xj xi xi xj
which is often written in finite element analysis as a sum of the linear part, e ij , and the nonlinear part, ij . The strain in (4.2) is referenced to the undeformed configuration of the body. In
linear analysis this is sufficient since the deformed and undeformed states are almost identical
but when performing non-linear analysis the need arises for an entity which relates the deformed
state to the undeformed state; this is the deformation gradient.
4.1.1 Deformation gradient
The deformation gradient is very important in relation to the concept of work conjugacy as
described by e.g. Bonet and Wood (1997). The work done by a set of external loads is equal to
the resulting internal work given as the product of stress and strain. The term conjugacy refers
to that stresses and strains must be referenced to the same state for the work to be evaluated

4. Geometrical Non-Linear Analysis of Laminated Shell Structures

70

correctly. When dealing with finite strains the product of the Cauchy stresses, ij and the
Green-Lagrange strain tensor will not yield the correct internal work as is the case in linear
theory of elasticity. To determine how to proceed we consider the body in Fig. 4.1 shown in
undeformed and deformed states.
x ( xi, t)

t i 0

dz

dz

0
0

dy

dx

dy

dx

z
y
x

(a)

(b)

Figure 4.1 Differential volume of a body in (a) the undeformed state and (b) the same volume deformed
in the final state

The mapping, t xi (0 xi , t), between the two states is used to define the deformation gradient. We
use the same procedure as for the Jacobian and define the deformation gradient, t0 Uij , as:
t
0 Uij

t xi
= 0 i t xj
0 xj

(4.3)

where is the gradient operator (hence deformation gradient). The notation t0 Uij indicates that
the deformation gradient describes the change in stretches and rotations between two states at
time 0 and time t. Thus we obtain the differential line segment in the deformed state, t dxi , from
the same line segment in the initial state simply as t dxi = t0 Uij 0 dxj . Writing out the expression
in (4.3) in matrix form and noting that the position at time t is t xj = 0 xj + t uj yields:

t
0U

(0 x+t u1 )
0x
(0 y+t u2 )
x
0

(0 z+t u3 )
0x

(0 x+t u1 )
0y
(0 y+t u2 )
0y
(0 z+t u3 )
0y

(0 x+t u1 )
0z
(0 y+t u2 )
0z
(0 z+t u3 )
0z

t u1
0x
t u2
0x
t u3
0x

1+

t u1
0y
u
1+ t 2
0y
t u3
0y

t u1
0z
t u2
0z
u
1+ t 3
0z

(4.4)

This can conveniently be written in terms of the displacements in tensor form by introducing
Kroneckers delta in the equation :
t
0 Uij

= ij + t ui,j

(4.5)

From the preceding it is obvious that the Jacobian and the deformation gradient are closely
related. Consider again the two bodies in Fig. 4.1. The volume at time 0 is easily calculated
as 0 dV = 0 dx(0 dy 0 dz) but the volume at time t is only for small strain equal to the initial
volume, i.e. in general 0 dV 6= t dV . Since the deformation gradient describes the mapping
of the differential lengths, 0 dxi , it seems natural that the deformation gradient also maps the
differential volume. The volume in the deformed state is written in terms of the initial state

Notes on Structural Analysis of Composite Shell Structures

71

lengths and the deformation gradient as:


0 dV

= 0 dx(0 dy 0 dz)


t xi
t xi t xi

=
t dx1t dx2t dx3
0 x1 0 x2 0 x3
= |J|t dV

(4.6)

from which it follows that |J| = |t0 U|. As a consequence, the deformation gradient (or Jacobian)
also maps the density (which is not assumed constant when dealing with large strain). If no mass
is assumed to disappear between the two states, then the mass dV must remain constant, i.e.
0 0 dV = t t dV whereby 0 = |J|t from (4.6). Considering now that also work per unit mass
must remain constant, we obtain the following relationship:
1
1
im t Dmj = im t Dmj
t
0
1
im =
im
|J|

(4.7)

As can be seen from (4.7) the Kirchhoff stress, im is work conjugate with the rate of deformation change, t Dij , with reference to the initial volume. Note that the rate of deformation change
is analogous to strain velocity, ij , and thus t Dmn = 0t Ui,m 0t Uj,m ij . Therefore (4.7) does not
provide any useful physical information and is thus only used to provide an expression for the
Cauchy stress, ij .
We employ the Green-Lagrange strain definition and therefore need a stress measure which is
referenced to the initial state. This is provided through the second Piola-Kirchhoff stress tensor,
Sij , which is defined from the inverse deformation gradient as Sij = 0t Uim 0t Ujn mn . Using
the relation between ij and ij the definition of the second Piola-Kirchhoff stress tensor is
obtained:
Sij = |J| 0t Uim 0t Ujn t mn
(4.8)
The expression in (4.8) can be rearranged to yield an expression for the Cauchy stresses:
t ij

1 t
Uim t0 Ujn Smn
|J| 0

(4.9)

The work done by a set of external loads is then obtained as the product of the second PiolaKirchhoff stress and the Green-Lagrange strain. However, when performing non-linear analysis
the Cauchy stress must still be determined since the second Piola-Kirchhoff stress is referenced
to the initial state and as such has no real physical interpretation. Note that in linear analysis
we use the product of the Cauchy stress tensor and the Green-Lagrange strain tensor directly.
The validity of this approach can be seen by noting that for small strain 0t Uij = t0 Uij = ij or in
matrix notation 0t U = t0 U = I. And since |I| = 1 we obtain from (4.8) that Sij ' ij .
The internal work in terms of the quantities defined in the preceding may be expressed as:
Z
Z
W =
Sij ij dV =
ij tij dV
(4.10)
V0

Vt

where ij is the Green-Lagrange strain defined in (3.17) and tij is the true strain referenced to
the deformed state. The expression in (4.10) is work conjugate since both stresses and strains

4. Geometrical Non-Linear Analysis of Laminated Shell Structures

72

on the left hand side are referenced to the initial undeformed state and the stresses and strains
on the right hand side are referenced to the deformed state. In non-linear analysis the second
Piola-Kirchhoff stress and not the Cauchy stress is determined through the constitutive relation.
Consequently, the stress will in the following treatment be expressed as S ij .
Using the relations developed in the preceding all quantities of the governing equations are
well-defined and we proceed by addressing the solution strategy for non-linear analysis.
4.1.2 Governing equations for non-linear analysis
The quantities in the governing equations are in non-linear analysis no longer independent of
the displacements. This can be realized when developing the element matrices and is caused
by the non-linear terms of the strain tensor. Consequently, the governing equations can no
longer be solved as a standard system of algebraic equations. In stead we must employ iterative techniques such as the Newton-Raphson method (see e.g. Zienkiewicz and Taylor (2000a),
Bonet and Wood (1997) or Lund and Condra (2001)). The first step in such a solution is to
rewrite (4.1) to its weak form:
nel
X
(t k t r k ) = 0
(4.11)
k=1

where k = V BT S dV in which S is the second Piola-Kirchoff stress introduced to obtain


work conjugancy as discussed previously. The B-matrix is for non-linear analysis dependent on
the displacements due to the non-linear terms of the strain tensor (4.2). Consequently, we write
the strain-displacement matrix as a sum of the linear part (derived in (3.39)) and a non-linear
part, i.e. we write:
B = B0 + BN L (t d)
(4.12)

where t d is the global vector containing all nodal displacements of the model. This vector is
also the solution to (4.11). However, in practice the solution will never satisfy the equation
numerically and hence, the final solution will be the one coming within a specified tolerance.
We assume that the solution to the previous time 0 is known and for convenience write the entire
equation (4.11) as a linear operator in the displacements, i.e. we write (d)k = k rk . The
solution at time t can then be approximated from the solution at time 0 by a first order Taylor
expansion as:
(0 d)
(t d) (0 d) +
d
(4.13)
0d
where d is the global nodal displacement increment. The expression in (4.13) states that the
present solution can be found from the previous iteration and time-step by using (1) the solution
of the previous iteration, (0 d), (2) the gradient from the previous iteration, (0 d)/ 0 d, and
(3) some increment in the displacements from the previous time-step, d. The gradient, or
tangent, term from (4.13) may be found by differentiating the vector (d) while noting that
only k is a function of the displacements. Hence, using the integral expression of k defined

Notes on Structural Analysis of Composite Shell Structures

73

above and differentiating with respect to the displacements we obtain a matrix defined as:
n

el
(0 d) X
=
0d
k=1

=
=

nel Z
X

k=1 V
nel Z
X
k=1

(BT Sk ) dV
uk k

el
X
BTk
Sk dV +
uk

k=1 V
nel Z
X

BTk
Sk dV +
uk
k=1

BTk

Sk k
dV
k uk

(4.14)

BTk Ck Bk dV

since Sk /k = C and k /uk = B. Thus we from (4.14) obtain a definition for the global
tangent stiffness matrix, KT = (0 d)/ 0 d.
The first right-hand term of (4.14) is customarily denoted using the constitutive relation as
K = BT /uCBu. The matrix K is thus naturally referred
to as the global initial stress
R
stiffness matrix. The second term is for now defined as K = V BT CB dV but will be extended
in Section 4.2.
By introducing the definition of the tangent stiffness matrix and the residual in (4.13) the equations of static equilibrium are obtained for the entire system, i.e. all quantities have been summated:
KT d = r
(4.15)
where r and are the global external and internal force vector, respectively. The tangent stiffness matrix, KT is defined as explained above as:
KT = K + K
Z
Z
BT
=
CBua dV +
BT CB dV
u
a
V
V

(4.16)

It is apparent from the preceding discussion that the method of solution requires some initial
values or a "starting guess" in order to solve (4.15). The obvious solution is to provide a zerovalue initial guess, which is also the practice chosen in MUST.
Now, the governing equations for the geometrically non-linear analysis have been defined and
subsequently, the element matrices will be developed.

4.2

Non-linear element matrices

The interpolation matrix is stated as for the linear case (see (3.32)). When determining the
non-linear strain-displacement matrix from the variation we consider the variation of the
non-linear part of the strain vector as given from the strain definition in (4.2), stated again for
reference:
1 uk uk
ij =
(4.17)
2 xj xi

74

4. Geometrical Non-Linear Analysis of Laminated Shell Structures

When expressing the variation of the non-linear strain vector, it proves convenient to define an
auxiliary matrix, A, whereby (4.17) is written as:
1
A
2
T
x
0

0
A =
T
y

(4.18)

Tz

0
Ty
0
Tx
Tz
0

0
0

T
z

T
y
Tx

where Tx = [u,x , v,x , w,x ], Ty = [u,y , v,y , w,y ] and Tz = [u,z , v,z , w,z ]. The vector is given as
T = [ x , y , z ] and for reference 0 = [0, 0, 0]. Taking the variation of in (4.18) yields the
following expression:
1
1
= A + A
2
2

(4.19)

since both A and are functions of the nodal displacements. In order to obtain the variation
of A, an interesting property of the matrix is used. By inserting the expressions for and A in
(4.19) it is easily shown that the variation can be written as A = A. Thereby (4.19) can
be written simply as = A. The matrix A is not suitable as a strain-displacement relation
since must be a direct function of the nodal displacements. Therefore, another expression is
needed and to this end an auxiliary matrix is again introduced. Consequently, the vector is
written as a product of the matrix G and the nodal displacements, i.e. = Gu a . The matrix G
is the auxiliary matrix determined directly from (3.38) due to the definition of :


Na,x
0
0 11 a1 21 a1

0 N
0 12 a1 22 a1
a,x

a
a
0

0 Na,x 13 1 23 1


Na,y
0
0 11 a2 21 a2



0 12 a2 22 a2
G = 0 Na,y
(4.20)


a
a
0
0 Na,y 13 2 23 2

N
0
0 11 a3 21 a3
a,z



a
a
0
Na,z
0 12 3 22 3


0
0
Na,z 13 a3 23 a3

Notes on Structural Analysis of Composite Shell Structures

75

Inserting the expression for in yields = AGua , whereby the non-linear strain-displacement
matrix can be determined as BN L = AG:


Na,x u,x
Na,x v,x
Na,x w,x


Na,y u,y
Na,y v,y
Na,y w,y



Na,z u,z
Na,z v,z
Na,z w,z

BN L =
Na,y u,x + Na,x u,y Na,y v,x + Na,x v,y Na,y w,x + Na,x w,y

N u + N u
a,z ,y
a,y ,z Na,z v,y + Na,y v,z Na,z w,y + Na,y w,z

Na,z u,x + Na,x u,z Na,z v,x + Na,x v,z Na,z w,x + Na,x w,z
(4.21)

a a
a a

11 1
21 1


a a
12 2
a22 a2


a a
a a

13 3
23 3

a12 a1 + a11 a2 a22 a1 + a21 a2

a a
a a
a a
a a
13 2 + 12 3 23 2 + 22 3

a13 a1 + a11 a3 a23 a1 + a21 a3
where

aij

3
X

(4.22)

ila um,j

l=1

Now, both the linear and non-linear strain-displacement matrices have been determined and
the overall strain-displacement matrix can be found as the sum of the linear contribution from
(3.39) and the non-linear contribution from (4.21), i.e. we write B = B0 + BN L . From B the
tangent stiffness matrix can be determined.
4.2.1 Tangent stiffness matrix
The general expression for the element tangent stiffness matrix was defined as (4.14). The
matrix K can be determined in a straight-forward manner by inserting the strain-displacement
relation, B = B0 + BN L , in (4.14):
Z
K=
(BT0 + BTN L )C(B0 + BN L ) dV
Z
ZV
(4.23)
T
BT0 CBN L + BTN L CB0 + BTN L CBN L dV
=
B0 CB0 dV +
|V
{z
} |V
{z
}
K0

KL

The matrix K0 is known as the small deformation stiffness matrix and KL is the initial displacement stiffness matrix also referred to as the large displacement stiffness matrix. The last term
in the expression for the tangent stiffness matrix (4.16) is the initial stress matrix or geometric
stiffness matrix, defined again as:
Z
BT
S dV
(4.24)
K =
V ua
Realizing from (4.12) that only the non-linear part of B is dependent on the displacements,
i.e. B(ua ) = B0 + BN L (ua ) whereby BT /ua = BTN L /ua , and using the definition

4. Geometrical Non-Linear Analysis of Laminated Shell Structures

76

BN L = AG from (4.21) an expression for K is obtained by inserting in (4.24):


Z
AT
K =
GT
S dV
ua
V

(4.25)

Again, exploiting the properties of the matrix A, it can easily be shown that the following
relation is applicable
AT

S=H
= HG
(4.26)
ua
ua
where the matrix H is defined as:

S1 I S 4 I S 6 I
S2 I S5 I
H=
(4.27)
Sym.
S3 I

The components of the stress vector in (4.26) are defined as for the strain vector so that S =
{S1 , S2 , S3 , S4 , S5 , S6 } = {S11 , S22 , S33 , S12 , S23 , S13 }. In (4.27) the matrix I is the 3 3
identity matrix. Inserting (4.26) into (4.25) yields the following expression for the initial stress
matrix:
Z
K =
GT HG dV
(4.28)
V

The tangent stiffness matrix can now be determined using the expressions found in respectively
(4.23) for the small deformation stiffness matrix and for the initial displacement matrix and in
(4.28) for the initial stress matrix.
Z
Z
T
KT = B0 CB0 dV +
BT0 CBN L + BTN L CB0 + BTN L CBN L dV
VZ
V
(4.29)
T
+
G HG dV
V

All quantities in (4.29) are known and the assembly of KT is relatively straight-forward. The
expression in (4.29) is an integral expression and thus, the tangent stiffness matrix is assembled
using the Gauss quadrature described in Section 3.8.

Bibliography
Ahmad, S., B. M. Irons, and O. C. Zienkiewicz (1970). Analysis of thick shell and thin shell structures by curved
finite elements. International Journal for Numerical Methods in Engineering 2, 419459.
Alfano, G., F. Auricchio, L. Rosati, and E. Sacco (2001). Mitc finite elements for laminated composite plates.
International Journal for Numerical Methods in Engineering 50, 707738.
Ashby, M. F. (1999). Materials Selection in Mechanical Design (Second ed.).
ISBN: 0-7506-4357-9.

Butterworth-Heinemann.

Bathe, K.-J. (1996). Finite Element Procedures (second ed.). Prentice-Hall. ISBN: 0-13-301458-4.
Belytschko, T., W. K. Liu, and B. Moran (2000). Nonlinear finite elements for continua and structures (First ed.).
John Wiley & Sons Inc. ISBN: 0-471-98774-3.
Bischoff, M. (1999). Theorie und Numerik einer dreidimensionalen Schalenformulierung. Institut fr Baustatik,
Universitt Stuttgart. Ph.D. thesis.
Bletzinger, K., M. Bischoff, and E. Ramm (2000). A unified approach for shear-locking-free triangular and rectangular shell finite elements. Computers & Structures 75, 321334.
Bonet, J. and R. D. Wood (1997). Nonolinear continuum mechanics for finite element analysis (first ed.). Cambridge University Press. ISBN: 0-521-57272-X.
Bucalem, M. L. and K. J. Bathe (1993). Higher-order mitc general shell elements. International Journal for
Numerical Methods in Engineering 36, 37293754.
Cook, R. D., D. S. Malkus, and M. E. Plesha (1989). Concepts and applications of finite element analysis (Third
ed.). John Wiley & Sons Inc. ISBN: 0-471-84788-7.
Dvorkin, E. N. and K.-J. Bathe (1984). A continuum mechanics based four-node shell element for general nonlinear analysis. Journal of Engineering Computations 1, 7788.
Grdal, Z., R. T. Haftka, and P. Hajela (1999). Design and optimization of laminated composite materials (first
ed.). John Wiley & sons inc. ISBN: 0-471-25276-X.
Hansen, J. S. (1996). Notes on Optimal Design of Composite Structures. Institute of Mechanical Engineering,
Aalborg University. Part 1.
Heinbockel, J. H. (1996). Introduction to Tensor Calculus and Continuum Mechanics (first ed.). Department of
Mathematics and Statistics, Old Dominion University (US). www.math.odu.edu/jhh/counter2.html.
Hughes, T. J. R. (1987). The Finite Element Method (first ed.). Prentice-Hall. ISBN: 0-13-317025-x.
Jones, R. M. (1999). Mechanics of composite materials (Second ed.). Taylor & Francis. ISBN: 1-56032-712-x.
Kildegaard, A., J. S. Hansen, and O. T. Thomsen (1999). Lette Konstruktioner i Komposit- og Sandwichmaterialer.
Institute of Mechanical Engineering, Aalborg University. Livslang Uddannelse.

BIBLIOGRAPHY

78

Klinkel, S., F. Gruttmann, and W. Wagner (1999). A continuum based three-dimensional shell element for laminated structures. Computers & Structures 71, 4362.
Kumar, W. P. P. and R. Palaninathan (1997). Finite element analysis of laminated shells with exact throughthickness integration. Computers & Structures 63, 173180.
Lund, E. and T. J. Condra (2001).
Notes and Exercises for Course on Numerical Methods.
Lecture notes available at www.ime.auc.dk/el.
McNeal, R. H. (1998). Perspective on finite elements for shell analysis. Finite Elements in Analysis and Design 30,
175186.
Mindlin, R. D. (1951). Influence of rotatory inertia and shear on fleural motions of isotropic, elastic plates. Journal
of Applied Mechanics 18, 3138.
Noor, A. K., W. S. Burton, and J. M. Peters (1990). Predictor-corrector procedures for stress and free vibration
analyses of multilayered composite plates and shells. Computer Methods in Applied Mechanics and Engineering 82, 341363.
Ochoa, O. O. and J. N. Reddy (1992). Finite Element Analysis of Composite Laminates (First ed.). Kluwer
Academic Publishers. ISBN: 0-7923-1125-6.
Oskooei, S. and J. S. Hansen (2000). A higher order finite element for sandwich plates. Journal of the American
Institute of Aeronatuics and Astronautics 38, 525533.
Panda, S. and R. Natarajan (1981). Analysis of laminated composite shell structures by finite element method.
Computers & Structures 14, 225230.
Reissner, E. (1945). The effect of transverse shear deformation on the bending of elastic plates. Journal of Applied
Mechanics 12, 6976.
Ris (2001). Ris Nyt Juni 2001. Ris National Laboratory. Used by courtesy of Leif Snderberg Petersen, Ris
National Laboratory.
Sokolnikoff, I. S. (1956). Tensor analysis (Second ed.). John Wiley & Sons, inc.
Stegmann, J., L. R. Jensen, and J. M. Rauhe (2001). Finite Elements for Geometric Non-Linear Analysis of
Composite Laminates and Sandwich Structures. Masters thesis available at www.ime.auc.dk/js.
Timoshenko, S. P. and J. N. Goodier (1970).
ISBN: 0-07-085805-5.

Theory of elasticity (Third ed.).

McGraww-Hill.

Zienkiewicz, O. C. (1977). The Finite Element Method (third ed.). McGraw-Hill. ISBN: 0-07-084072-5.
Zienkiewicz, O. C. and R. L. Taylor (2000a). The Finite Element Method, vol 1 (fifth ed.). Butterworth-Heinemann.
ISBN: 0-7506-5049-4.
Zienkiewicz, O. C. and R. L. Taylor (2000b). The Finite Element Method, vol 2 (fifth ed.). Butterworth-Heinemann.
ISBN: 0-7506-5055-9.

Index
3D finite elements, 30
Anisotropic, 5, 15
Axis of symmetry, 54
Balanced laminate, 33
Bend-twist coupling, 24, 25
Carbon fiber reinforced plastic (CFRP), 13, 31, 33
Cauchys formula, 35
Classical laminated plate theory, 20
Classical lamination theory (CLT), 20
Composite material, 4, 32
Constitutive matrix, 52
Constitutive relation, 51
Constitutive relations
aelotropic, 52
isotropic, 54
modified , 57
monoclinic, 52
orthotropic, 53
transformation, 59
Continuous fiber composite, 32
Contravariant base, 45
Coordinate system
director, 43
Coordinates
curvilinear, 39
Core, 31
Covariant base, 45
Deformation gradient, 67
Density change, 69
Element Input, 74
SHELL4L, 75
SHELL4T, 74
Equations of equilibrium , 35
Equations of motion, 35
discrete form, 36
Fabric, 33
Fibers, 14
Fibrous composite, 5, 32
Force resultants, 22
Gauss divergence theorem, 35
Gauss quadrature, 60
Glass fiber reinforced plastic (GFRP), 13, 31, 33

Higher order plate theory, 29


Hookes general law, 51
Hookes law, 15, 51
Inhomogenous, 5
Initial stress matrix, 71
Internal force vector, 37
Interpolation
displacement, 45
geometry, 43
matrix, 37, 49
Isoparametric
definition, 37
Jacobian
matrix, 50
Kevlar fiber reinforced plastic (KFRP), 33
Kirchhoff-Love hypothesis, 21
Lames constants, 64
Lamina, 14, 33, 52
Laminate, 4, 20, 33, 55
Laminate theory, 20
Laminated composite materials, 5
Laminated fiber-reinforced materials, 5
Layered plate theories, 30
Locking
curvature thickness locking, 64
membrane locking, 62
shear locking, 63
thickness locking, 64
volumetric locking, 64
Macromechanical, 6, 14
Mass matrix
consistent, 39
definition, 37
Mat, 33
Material coordinate system, 15
Materials selection, 6
Matrix, 14
Micromechanical, 18
Mindlinn assumption, 57
MITC
elements, 64
tying points, 65
Mixed Interpolation of Tensorial Components, 64
Moment resultants, 23

INDEX

80

Neutral surface, 39
Nodal load vector, 38
consistent, 39
definition, 38
Notation, 15
Numerical integration
piecewise through-the-thickness, 61
thickness integration, 60
Orthotropic, 5, 16
Plane of elastic symmetry, 52
Plane stress, 16
Real constants
SHELL4L, 76
SHELL4T, 75
Reference surface, 39
Reinforced composite materials, 14
Reinforced plastics, 32
ReissnerMindlin plate theory, 26, 29
Reissner-Mindlin, 41
Reissner-Mindlin assumption, 57
Rule of mixture, 19, 20, 30
Sandwich structure, 4, 31
Shear correction factor, 58
Shell theories
first order shear deformation, 41
Mindlin, 41
Reissner, 41
SHELL4L, 75
SHELL4T, 74
Stacking sequence, 33
Stiffness matrix
geometric, 73
initial displacement, 73
initial stress, 71, 73
large displacement, 73
small deformation, 73
tangent, 71, 73
Stiffness-limited design, 9
Strain
velocity, 69
Strain-displacement matrix, 37, 49
linear, 51
non-linear, 73
Strain-displacement relation, 37
Strand, 33
Strength-limited design, 11
Stress tensor
Cauchy, 69
Kirchhoff, 69
Piola-Kirchhoff, second, 69
Stress-strain relations, 16
Symmetric laminate, 33
Through-the-thickness integration

laminated elements, 60, 61


Transformation, 16
Unidirectional composite, 33
Unsymmetric laminate, 24
Virtual displacements, 35
Virtual work, 36
Volume change, 68
Wind turbine blade, 3
Work equivalence, 38
Woven roving, 33
Yarn, 33

Você também pode gostar